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

Index Calculations

Yuan Yao

Computing Embedded Contact Homology in Morse-Bott Settings

Yuan Yao
Abstract

Given a contact three manifold YY with a nondegenerate contact form λ\lambda, and an almost complex structure JJ compatible with λ\lambda, its embedded contact homology ECH(Y,λ)ECH(Y,\lambda) is defined ([Hut14]) and only depends on the contact structure. In this paper we explain how to compute ECH for Morse-Bott contact forms whose Reeb orbits appear in S1S^{1} families, assuming the almost complex structure JJ can be chosen to satisfy certain transversality conditions (this is the case for instance for boundaries of concave or convex toric domains, or if all the curves of ECH index one have genus zero). We define the ECH chain complex for a Morse-Bott contact form via an enumeration of ECH index one cascades. We prove using gluing results from [Yao22] that this chain complex computes the ECH of the contact manifold. This paper and [Yao22] fill in some technical foundations for previous calculations in the literature ([Cho16], [HS06]).

1 Introduction

1.1 Embedded contact homology

In this article we develop some tools to compute the embedded contact homology (ECH) of contact 3-manifolds in Morse-Bott settings.

ECH is a Floer theory defined for a pair (Y,λ)(Y,\lambda), where YY is a three dimensional contact manifold with nondegenerate contact form λ\lambda (for an introduction see [Hut14]). The ECH chain complex is generated by orbit sets of the form α={(γi,mi)}\alpha=\{(\gamma_{i},m_{i})\}. Here γi\gamma_{i} are distinct simply covered Reeb orbits of λ\lambda; and the mim_{i} is a positive integer which we call the multiplicity of γi\gamma_{i}. To describe the differential, consider the symplectization (×Y,d(esλ))(\mathbb{R}\times Y,d(e^{s}\lambda)) of YY with almost complex structure JJ. Here ss denotes the variable in the \mathbb{R} direction; and JJ is a generic λ\lambda-compatible almost complex structure (see Definition 2.2). The differential of ECH, which we write as \partial, is defined by counting holomorphic currents of ECH index I=1I=1 in the symplectization. More precisely, the coefficient α,β\langle\partial\alpha,\beta\rangle is defined by counts of JJ-holomorphic currents that approach α\alpha as ss\rightarrow\infty and β\beta as ss\rightarrow-\infty, where convergence to α,β\alpha,\beta is in the sense of currents. The resulting homology, which we write as ECH(Y,ξ)ECH(Y,\xi), is an invariant of the contact structure ξ=kerλ\xi=\textup{ker}\lambda. See Section 2 below for a more precise review of ECH and the ECH index.

In part due to its gauge theoretic origin, ECH has had spectacular applications to understanding symplectic problems and dynamics in low dimensions; for instance sharp symplectic embedding obstructions of four dimensional symplectic ellipsoids ([MS12]), closing lemmas for Reeb flows on contact 3-manifolds ([Iri15]), the Arnold chord conjecture ([HT11, HT13]), and quantitative refinements of the Weinstein conjecture [CH16]. Several computations (e.g. [HS06, Cho16, Leb07]) and applications (e.g. [Hut16]) of ECH have assumed results from its Morse-Bott version, which we develop in detail in this paper.

1.2 Morse-Bott theory

The original definition of ECH requires we use non-degenerate contact forms. However, in practice many contact forms we encounter carry Morse-Bott degeneracies, for which the Reeb orbits are no longer isolated but instead show up in families with weaker non-degeneracy conditions imposed (for a more precise description, see Definition 3.2 in [OW18]). Although all Morse-Bott contact forms can be perturbed to non-degenerate ones, it is often useful to be able to compute ECH directly in the Morse-Bott setting, where often the enumeration of JJ-holomorphic curves is easier.

For ECH, since we only consider 3-manifolds, the two Morse-Bott cases are either when the Reeb orbits come in a two dimensional family, or come in one dimensional families. For the first case it then follows that the entire contact manifold is foliated by periodic Reeb orbits. ECH with this kind of Morse-Bott degeneracy has been computed in many cases by [NW20], see also [Far11].

The other case is when Reeb orbits show up in one dimensional S1S^{1} families, i.e. we see tori foliated by Reeb orbits. We shall call these tori Morse-Bott tori. It is with this case we concern ourselves in this paper (for a description of what the contact form looks like, see Proposition 3.2). Examples of this include boundaries of toric domains, and torus bundles over the circle see [Her98, Cho+14, Cri19, Leb07].

For now we consider (Y3,λ)(Y^{3},\lambda) a contact 3-manifold where λ\lambda is a Morse-Bott contact form all of whose Reeb orbits appear in S1S^{1} families. Later for the case of boundary of convex or concave toric domains (Sections 9,10) we allow the case of both nondegenerate Reeb orbits and S1S^{1} families of Reeb orbits. We consider the symplectization with a generic λ\lambda compatible almost complex structure JJ (see Definition 2.2)

(×Y3,d(esλ)).(\mathbb{R}\times Y^{3},d(e^{s}\lambda)).

Following the recipe described in [Bou02], to compute ECH in the Morse-Bott setting we shall count holomorphic cascades of ECH index one. The philosophy behind this is as follows: given λ\lambda, a Morse-Bott contact form with Reeb orbits in Morse-Bott tori, we can perturb

λλδ\lambda\longrightarrow\lambda_{\delta}

where λδ\lambda_{\delta} with δ>0\delta>0 is a nondegenerate contact form up to a certain action level L>>0L>>0. This perturbation requires the following information. For each circle of orbits parameterized by S1S^{1}, choose a Morse function ff on S1S^{1} with two critical points. The effect of this perturbation is so that each Morse-Bott torus splits into two nondegenerate Reeb orbits (corresponding to the critical points of ff): one is an elliptic orbit and the other is a hyperbolic orbit. We also need to perturb the λ\lambda-compatible almost complex structure on the symplectization into a λδ\lambda_{\delta} compatible almost complex structure, JδJ_{\delta}. Since λδ\lambda_{\delta} is nondegenerate up to action LL, we can define the ECH chain complex up to action LL in this case by counting ECH index one JδJ_{\delta}-holomorphic curves. The idea is to take δ0\delta\rightarrow 0 and see what these ECH index one holomorphic curves degenerate into.

By a compactness theorem in [Bou+03] (see also [Bou02, Yao22]), such JδJ_{\delta}-holomorphic curves degenerate into JJ-holomorphic cascades. For a definition of JJ-holomorphic cascade, see [Yao22]. Roughly speaking, a JJ-holomorphic cascade, which we shall write as uu^{\text{\Lightning}}, consists of a sequence of JJ-holomorphic curves {u1,..,un}\{u^{1},..,u^{n}\} that have ends on Morse-Bott tori. We think of the curves uiu^{i} as living on different levels, with uiu^{i} one level above ui+1u^{i+1}. Between adjacent levels there is the data of a single number Ti[0,]T_{i}\in[0,\infty] described as follows. Suppose a positive end of ui+1u^{i+1} is asymptotic to a simply covered Reeb orbit γ\gamma with multiplicity nn. This γ\gamma corresponds to a point on S1S^{1} (the S1S^{1} that parameterizes the family of Morse-Bott Reeb orbits). Then if we follow the upwards gradient flow of ff for time TiT_{i} starting at the point corresponding to the Reeb orbit γ\gamma, we arrive at a Reeb orbit γ~\tilde{\gamma}, and a negative end of uiu^{i} is asymptotic to γ~\tilde{\gamma} with the same multiplicity nn. We assume all positive ends of ui+1u^{i+1} and negative ends of uiu^{i} are matched up in this way. For an illustration of a cascade111This figure and the accompanying explanations are taken from Figure 1 in [Yao22]., see Figure 1.

Refer to caption
Figure 1: A schematic picture of a cascade: the cascade uu^{\text{\Lightning}} consists of two levels, uu and vv. Horizontal lines correspond to Morse-Bott tori. Moving in the horizontal direction along these horizontal lines corresponds to moving to different Reeb orbits in the same S1S^{1} family. Arrows correspond to gradient flows, and diamonds correspond to critical points of Morse functions on S1S^{1} families of Reeb orbits. Between the holomorphic curves uu and vv, there is a single parameter TT that tells us how long positive ends of vv must follow the gradient flow to meet a negative end of uu.

1.3 Main results

The Morse-Bott ECH chain complex which we write as (CMB,MB)(C_{*}^{MB},\partial_{MB}) (see section 7) can be described as follows. Its generators are collections of Morse-Bott tori, equipped with a multiplicity and additional data, which we write as α={(𝒯j,±,mj)}\alpha=\{(\mathcal{T}_{j},\pm,m_{j})\}. Here 𝒯j\mathcal{T}_{j} denotes a Morse-Bott torus; we call mjm_{j} the multiplicity; and a choice of ++ or -. See Section 5.3 for a description. Suppose we can choose a λ\lambda compatible almost complex structure JJ which is “good” (see definition 4.3), meaning certain transversality conditions (Definition 4.5) are satisfied. The differential in the Morse-Bott chain complex MB\partial_{MB} counts ECH index one cascades between Morse-Bott ECH generators. The ECH index of a cascade is described in Section 5. We describe what it means for an cascade to be asymptotic to a Morse-Bott ECH generator in Section 5.3. For a description of what ECH index one cascades look like, see Corollary 5.29, Prop. 5.33. We prove that

Theorem 1.1.

Let λ\lambda be a Morse-Bott contact form on the contact 3-manifold YY whose Reeb orbits all appear in S1S^{1} families. Assuming the almost complex structure JJ is good (see Definition 4.3), the homology of the Morse-Bott ECH chain complex computes the ECH of the contact manifold ECH(Y,ξ)ECH(Y,\xi).

A slightly more precise version of this theorem that we prove is Theorem 7.1.

We next find some instances there is enough transversality to compute ECH using the Morse-Bott chain complex.

Theorem 1.2.

Let λ\lambda be a Morse-Bott contact form on the contact 3-manifold YY whose Reeb orbits all appear in S1S^{1} families. We can choose a generic JJ so that

  • Every reduced cascade (See Definition 3.13) of 3\leq 3 levels is transversely cut out (see Definition 4.5).

  • Every reduced cascade where all of the (nontrivial) JJ-holomorphic components of the reduced cascade (in all of its levels) are distinct up to translation in the symplectization direction is transversely cut out (see Definition 4.5).

If we can show through some other means that we can choose a small perturbation of JJ to JδJ_{\delta} satisfying conditions of Theorem 7.3 so that for small enough δ\delta, all ECH index one curves degenerate into cascades whose reduced version satisfy either of the above conditions, then consider the Morse-Bott ECH chain complex (CMB,MB)(C_{*}^{MB},\partial_{MB}) as described more precisely in Section 7. For the differential MB\partial_{MB}, if we restrict to “good” cascades (see Sections 5, 7 for the notion of “good”) of ECH index one whose reduced versions are of the above form, the differential is well defined and the chain complex (CMB,MB)(C_{*}^{MB},\partial_{MB}) computes ECH(Y,ξ)ECH(Y,\xi).

For a discussion how these conditions arise and a proof of this theorem, see the Appendix. This list is by no means exhaustive. We expect there are many other situations where transversality can be achieved; the particulars will depend on the specific details of the contact manifold for which we are computing the ECH chain complex. In particular, for the case relevant for boundaries of convex and concave toric domains, we have the following:

Theorem 1.3.

Let λ\lambda be a contact form on the contact 3-manifold YY whose Reeb orbits apppear either in Morse-Bott S1S^{1} families or are non-degenerate. Let δ>0\delta>0, and λδ\lambda_{\delta} be the nondegenerate perturbation of λ\lambda that perturbs each S1S^{1} family of Reeb orbits into two nondegenerate ones. If for δ>0\delta>0 small enough, all JδJ_{\delta} holomorphic curves of ECH index one in ×Y3\mathbb{R}\times Y^{3} have genus zero, then the embedded contact homology of YY can be computed from the Morse-Bott chain complex (CMB,tree,MBtree)(C_{*}^{MB,tree},\partial_{MB}^{tree}) (see Section 8) using an enumeration of tree like cascades.

To be more precise, for the above theorem we need to use a slightly different description of cascades which we call “tree like” cascades, which is explained in Sections 8, 9, 10. Consequently, we can prove

Theorem 1.4.

For boundaries of concave toric domains or convex toric domains, in the nondegenerate case after a choice of generic almost complex structure all curves of ECH index one have genus zero. Therefore the ECH of boundaries of concave/convex toric domains can be computed using the Morse-Bott ECH chain complex (CMB,tree,MBtree)(C_{*}^{MB,tree},\partial_{MB}^{tree}), via counts of tree-like ECH index one cascades.

For a definition of convex and concave toric domains, see Sections 9, 10.

We mention some previous computations of ECH that have assumed Morse-Bott theory of the flavour we develop in this paper, notably in [HS06] for the case of T3T^{3}, and [Cho16] for certain toric contact 3-manifolds, and [Leb07] for the case of T2T^{2} bundles over S1S^{1}. This paper and the gluing paper [Yao22] fill in the foundations for these results.

Remark 1.5.

The above theorems say for genus zero curves we have all the transversality we need by simply restricting to cascades of ECH index one and choosing a generic JJ; however this result is not strict, there could well be other scenarios where transversality can be achieved. For instance we expect with some more care we can show the moduli space of cascades of ECH index one and genus one can be shown to be transverse. For discussion of general difficulties see the Appendix.

1.4 Some technical details

For ECH in the nondegenerate setting (see [Hut14]), as we review in Section 2, the Fredholm index of a somewhere injective curve is bounded from above by its ECH index. Further, the ECH index is superadditive under unions of JJ-holomorphic curves in symplectizations. Using the fact that after choosing a generic almost complex structure, all somewhere injective curves are transversely cut out, it follows that by restricting to only ECH index one curves we do not need to consider multiply covered nontrivial curves. With this, one defines the ECH differential in the nondegenerate setting via counts of ECH index one JJ-holomorphic curves.

Parts of the above story continue to hold in the case of cascades if we assume can choose JJ to be good (Definition 4.3), as we explain below.

We first note that the notion of an ECH index continues to make sense for cascades, as we explain in Section 5. The case of cascades, however, is more complicated, in two directions.

  • During the degeneration process for λδ\lambda_{\delta} as δ0\delta\rightarrow 0, simple curves may degenerate into cascades that have multiply covered components;

  • For generic JJ, and even if we restrict to cascades all of whose curves are somewhere injective, the cascade need not be transversely cut out.

The second bullet point is the most problematic. This happens because by requiring there is a single parameter between adjacent levels, we are imposing restrictions on the evaluation maps on the ends of the curves in a cascade. Hence a cascade lives in a fiber product, which need not be transversely cut out even if we restrict to only somewhere injective curves. For an explanation of this, see the Appendix.

However, if we take as an assumption that JJ is good (which isn’t always possible, it will depend on the specific contact manifold), then all cascades built out of somewhere injective curves that we consider are transversely cut out. Then we can address the first bullet point by using a version of the ECH index inequality for cascades .

To explain the ECH index inequality for cascades, consider the following. Given a cascade, we can pass to a reduced cascade, which means we replace all multiply covered curves with the underlying simple curves. See Section 3 for a precise description of this process. The reduced cascade also lives in a fiber product because of the conditions we imposed on its ends. By the assumption that JJ is good (and consequently transversality assumptions in Definition 4.5 are satisfied), the reduced cascade is transversely cut out. To each reduced cascade we can associate to it a virtual dimension, which is the dimension of the moduli space of curves that lies in the same configuration as the reduced cascade. We prove that the ECH index of the cascade bounds the Fredholm index of the reduced cascade from above; and that equality holds only if the original cascade had no multiply covered components (and is well behaved in various ways, see Section 5).

In [Yao22], we proved a correspondence theorem between certain cascades and JJ-holomorphic curves.

Theorem 1.6 ([Yao22]).

Given a “transverse and rigid” (see Definition 3.4 in [Yao22]) height one JJ-holomorphic cascade uu^{\text{\Lightning}} , it can be glued to a rigid JδJ_{\delta}-holomorphic curve uδu_{\delta} for δ>0\delta>0 sufficiently small. The construction is unique in the following sense: if {δn}\{\delta_{n}\} is a sequence of numbers that converge to zero as nn\rightarrow\infty, and {uδn}\{u^{\prime}_{\delta_{n}}\} is sequence of JδnJ_{\delta_{n}}-holomorphic curves converging to uu^{\text{\Lightning}}, then for large enough nn, the curves uδnu_{\delta_{n}}^{\prime} agree with uδnu_{\delta_{n}} up to translation in the symplectization direction.

In this paper, using index calculations, we show that if JJ is good (some instances of which are described in Theorems 1.2), then essentially all ECH index one cascades are transverse and rigid222Technically we need to restrict ourselves to good ECH index one cascades. This is a fairly minor point, but see Proposition 5.32 and surrounding discussion.. Thus the gluing theorem above is then used to show the Morse-Bott chain complex computes ECH(Y,λ)ECH(Y,\lambda). In the cases where we use “tree like” cascades, for instance for boundaries of convex or concave toric domains, the definitions are slightly different, but essentially the same story holds true and we can always choose a generic JJ so that the Morse-Bott chain complex computes ECH(Y,λ)ECH(Y,\lambda).

Finally in the Appendix we explain why the usual techniques for achieving transversality fails for cascades.

Acknowledgements I would like to thank my advisor Michael Hutchings for his consistent help and support throughout this project. I would like to acknowledge the support of the Natural Sciences and Engineering Research Council of Canada (NSERC), PGSD3-532405-2019. Cette recherche a été financée par le Conseil de recherches en sciences naturelles et en génie du Canada (CRSNG), PGSD3-532405-2019.

2 ECH review

For a thorough introduction to ECH see [Hut14]. We will summarize much of the material from [Hut14] and [Hut02] for convenience of the reader.

Let (Y3,λ)(Y^{3},\lambda) be a contact 3 manifold with nondegenerate contact form λ\lambda. The generator of ECH are collections Θ\Theta, where each Θ\Theta is a set of Reeb orbits with multiplicities

Θ:={(γi,mi)|γiare pairwise distinct simply covered Reeb orbits,mi+}.\Theta:=\{(\gamma_{i},m_{i})|\gamma_{i}\,\text{are pairwise distinct simply covered Reeb orbits},\,m_{i}\in\mathbb{Z}_{+}\}.

We require mi=1m_{i}=1 if γi\gamma_{i} is a hyperbolic orbit. Then the chain for ECH are just

C(λ):=Θi2Θi.C_{*}(\lambda^{\prime}):=\bigoplus_{\Theta_{i}}\mathbb{Z}_{2}\langle\Theta_{i}\rangle.
Remark 2.1.

There is a decomposition of ECH according to homology class of Θi\Theta_{i} in H1(Y)H_{1}(Y). ECH can also be defined using \mathbb{Z} coefficients. We will not address these issues here.

Let α,β\alpha,\beta be ECH generators. Consider the symplectization of YY, defined as the symplectic manifold (×Y,ω:=d(eaλ))(\mathbb{R}\times Y,\omega:=d(e^{a}\lambda)), where aa denotes the \mathbb{R} coordinate. Equip it with a generic λ\lambda compatible almost complex structure JJ. By compatible we mean the following

Definition 2.2.

Let λ\lambda be a contact form (not necessarily nondegenerate) on a contact 3-manifold. Let JJ be a almost complex structure on the symplectization (×Y,ω:=d(eaλ))(\mathbb{R}\times Y,\omega:=d(e^{a}\lambda)). We say JJ is compatible with λ\lambda if

  1. a.

    JJ is invariant in the \mathbb{R} direction;

  2. b.

    Let RR denote the Reeb vector field, then Js=RJ\partial_{s}=R;

  3. c.

    Let ξ\xi denote the contact structure, then Jξ=ξJ\xi=\xi and dλ(,J)d\lambda(\cdot,J\cdot) defines a metric on ξ\xi.

Then the coefficient α,β\langle\partial\alpha,\beta\rangle is defined by

α,β:={2 count of holomorphic currents 𝒞 of ECH index I=1,so that as s+,𝒞 approaches α as a current, and as s,𝒞 approaches β as a current.}\langle\partial\alpha,\beta\rangle:=\left\{\begin{tabular}[]{@{}l@{}}$\mathbb{Z}_{2}$\, {count of holomorphic currents}\, $\mathcal{C}$\, {of ECH index} \,$I=1$,\\ {so that as} $s\rightarrow+\infty,\,\mathcal{C}$ \,\text{approaches}\, $\alpha$ {as a current, and as} $s\rightarrow-\infty$,\\ $\mathcal{C}$\, \text{approaches} $\beta$ \,\text{as a current}.\end{tabular}\right\} (1)

A holomorphic current 𝒞\mathcal{C} is by definition a collection {(Ci,mi)}\{(C_{i},m_{i})\} where each CiC_{i} is a somewhere injective JJ holomorphic curve and mi>0m_{i}\in\mathbb{Z}_{>0} accounts for the multiplicity of this curve. The ECH index II of a holomorphic curve CC (or more generally a relative 2 homology class in H2(α,β,Y)H_{2}(\alpha,\beta,Y), see section below for a definition) is defined by

I(C):=Qτ(C)+cτ(C)+CZI(C)I(C):=Q_{\tau}(C)+c_{\tau}(C)+CZ^{I}(C) (2)

where Qτ(C)Q_{\tau}(C) is the relative intersection number, cτ(C)c_{\tau}(C) is the relative Chern class, and CZCZ is a sum of Conley Zehnder indices used in ECH. We will review these terms in the upcoming subsections.

2.1 Relative first Chern class

Let α,β\alpha,\beta be orbit sets. We define the relative homology group H2(α,β,Y)H_{2}(\alpha,\beta,Y) to be the set of 2-chains Σ\Sigma with

Σ=αβ\partial\Sigma=\alpha-\beta

modulo boundary of 3 chains. This is an affine space over H2(Y)H_{2}(Y), and each JJ holomorphic curve defines a relative homology class.

We fix trivializations τ\tau of the contact structure ξ\xi over each Reeb orbit in YY. We then define the relative first Chern class cτc_{\tau} with respect to this choice of trivialization. For a given homology class in H2(α,β,Y)H_{2}(\alpha,\beta,Y), choose a representative ZH2(α,β,Y)Z\in H_{2}(\alpha,\beta,Y) that is embedded near its boundaries α,β\alpha,\beta. We assume ZZ is a smooth surface. Let ι:ZY\iota:Z\rightarrow Y be the inclusion. Then consider the bundle ιξ\iota^{*}\xi over ZZ. Let ψ\psi be a section of this bundle that is constant with respect to the trivialization τ\tau near each of the Reeb orbits, and perturb ψ\psi so that all of its zeroes are transverse. Then cτ(Z)c_{\tau}(Z) is defined to be the algebraic count of zeroes of ψ\psi. See [Hut14] for a more thorough explanation and that this is well defined.

2.2 Writhe

Let CC be a somewhere injective JJ holomorphic curve in the symplectization of YY, (×Y,d(eaλ))(\mathbb{R}\times Y,d(e^{a}\lambda)) (with generic λ\lambda-compatible complex structure JJ) that is asymptotic to α\alpha as s+s\rightarrow+\infty and β\beta as ss\rightarrow-\infty. For simplicity we focus on s+s\rightarrow+\infty end. It is known (see for example [Sie]) that for ss sufficiently large, C{s}×YC\cap\{s\}\times Y is a union of embedded curves near each orbit of α\alpha. For each orbit γi\gamma_{i} of α\alpha, the curves C{s}×YC\cap\{s\}\times Y forms a braid ξi+\xi_{i}^{+}. We use the trivialization τ\tau to identify the braids ξi+\xi_{i}^{+} with braids in S1×D2S^{1}\times D^{2}. We can define the writhe of ξi+\xi_{i}^{+} by identifying S1×D2S^{1}\times D^{2} with an annulus times an interval, projecting ξi+\xi_{i}^{+} to the annulus, and counting crossings with signs. The same sign convention is clearly explained in [Hut09].

Then given a somewhere injective JJ-holomorphic curve CC that is not the trivial cylinder, with braids ζi+\zeta_{i}^{+} associated to the ii-th Reeb orbit it approaches as s+s\rightarrow+\infty and braids ζj\zeta_{j}^{-} associated to the jjth Reeb orbit it approaches as ss\rightarrow-\infty we define its writhe to be

wτ(C):=iwτ(ζi+)jwτ(ζj).w_{\tau}(C):=\sum_{i}w_{\tau}(\zeta_{i}^{+})-\sum_{j}w_{\tau}(\zeta_{j}^{-}).

We also recall the writhe of the braid ζi+\zeta_{i}^{+} can be bounded by expressions in terms of the Conley-Zehnder indices.

Proposition 2.3.

Let CC be a somewhere injective holomorphic curve that is not a trivial cylinder which is asymptotic to γi\gamma_{i} with total multiplicity nin_{i}. Suppose there are kik_{i} distinct ends of CC that are asymptotic to γi\gamma_{i}, with covering multiplicities qijq_{i}^{j}. Then the writhe associated to the braid ζi+\zeta_{i}^{+} corresponding to Reeb orbit γi\gamma_{i} is bounded above by

wτ(ζi+)jniCZ(γij)jkiCZ(γiqij)w_{\tau}(\zeta_{i}^{+})\leq\sum_{j}^{n_{i}}CZ(\gamma_{i}^{j})-\sum_{j}^{k_{i}}CZ(\gamma_{i}^{q_{i}^{j}}) (3)

A similar bound holds for braids at ss\rightarrow-\infty with signs reversed.

We will derive an analogue of this bound for the Morse-Bott case. For now we recall another definition:

Definition 2.4.

Let CC be a somewhere injective JJ-holomorphic curve that is not a trivial cylinder. For each γi\gamma_{i} that CC is asymptotic to as s+s\rightarrow+\infty, form the sum CZI(γi):=j=1niCZ(γij)CZ^{I}(\gamma_{i}):=\sum_{j=1}^{n_{i}}CZ(\gamma_{i}^{j}) as above, and for each γi\gamma_{i}^{\prime} that CC is asymptotic to as ss\rightarrow-\infty, we form an analogous sum, then we define

CZI(C):=γi,Cis asymptotic toγi,as s+CZI(γi)γi,Cis asymptotic toγi,as sCZI(γi).CZ^{I}(C):=\sum_{\begin{subarray}{c}\gamma_{i},\\ C\,\text{is asymptotic to}\,\gamma_{i},\\ \text{as }\,s\rightarrow+\infty\end{subarray}}CZ^{I}(\gamma_{i})-\sum_{\begin{subarray}{c}\gamma_{i}^{\prime},\\ C\,\text{is asymptotic to}\,\gamma_{i}^{\prime},\\ \text{as }\,s\rightarrow-\infty\end{subarray}}CZ^{I}(\gamma_{i}^{\prime}). (4)

This is the Conley-Zehnder index term that appears in the definition of ECH index.

2.3 Relative adjunction formula

In this section we review the relative adjunction formula (see [Hut14, Hut02]). We first review the notion of relative intersection pairing, which is a map depending on the trivialization τ\tau:

Qτ:H2(α,β,Y)×H2(α,β,Y)Q_{\tau}:H_{2}(\alpha,\beta,Y)\times H_{2}(\alpha,\beta,Y)\rightarrow\mathbb{Z}

as follows. Let SS and SS^{\prime} be surfaces representing relative homology classes in H2(α,β,Y)H_{2}(\alpha,\beta,Y). If we identify ×Y\mathbb{R}\times Y with (1,1)×Y[1,1]×Y(-1,1)\times Y\subset[-1,1]\times Y, then we have by definition

S=S=imi{1}×αiini{1}×βi\partial S=\partial S^{\prime}=\sum_{i}m_{i}\{1\}\times\alpha_{i}-\sum_{i}n_{i}\{-1\}\times\beta_{i}

We make the following requirements on the representatives SS and SS^{\prime}:

  1. a.

    The projections to YY of the intersections of SS and SS^{\prime} with (1ϵ,1]×Y(1-\epsilon,1]\times Y and [0,ϵ)×Y[0,\epsilon)\times Y are embeddings.

  2. b.

    Each end of SS or SS^{\prime} covers Reeb orbits αi\alpha_{i} (resp βi\beta_{i}) with multiplicity 11.

  3. c.

    The image of SS (after projecting to YY in a neighborhood S1×D2S^{1}\times D^{2} of αi\alpha_{i} determined by the trivialization τ\tau) do not intersect, and do not rotate with respect to the chosen trivialization τ\tau as one goes around αi\alpha_{i}. Further, the image of different ends of SS approaching αi\alpha_{i} lie on distinct rays in a neighborhood of αi\alpha_{i}. More concretely using trivialization τ\tau to identify a neighborhood of αi\alpha_{i} with S1×2S^{1}\times\mathbb{R}^{2}, ends of SS approach αi\alpha_{i} along different rays in 2\mathbb{R}^{2}. We make a similar requirement for βi\beta_{i}. We make a similar requirement for SS^{\prime}.

  4. d.

    All interior intersections between SS and SS^{\prime} are transverse.

Representatives satisfying all of the above conditions are called τ\tau -representatives in [Hut02], which is a definition we will adopt. Then given τ\tau representatives as listed above, Qτ(S,S)Q_{\tau}(S,S^{\prime}) is defined to be the algebraic count of intersections between SS and SS^{\prime}.

We are now ready to state the relative adjunction formula, see also [Hut02].

Proposition 2.5.

If CC is a somewhere injective JJ holomorphic curve,

cτ(C)=χ(C)+Qτ(C)+wτ(C)2δ(C)c_{\tau}(C)=\chi(C)+Q_{\tau}(C)+w_{\tau}(C)-2\delta(C) (5)

where δ(C)0\delta(C)\geq 0 is defined to be an algebraic count of singularities of CC. Each singularity is positive due to the fact CC is JJ-holomorphic.

2.4 ECH index inequality

We have now defined all of the terms that appear in the ECH index inequality. We compare this with the Fredholm index. Let CC be a somewhere injective JJ-holomorphic curve, let Ind(C)Ind(C) denote the Fredhom index of CC, which in this case is given by

χ(C)+2cτ(C)+CZInd(C).-\chi(C)+2c_{\tau}(C)+CZ^{Ind}(C).

Here CZInd(C)CZ^{Ind}(C) is defined as follows. If CC is positively asymptotic to γ\gamma with kk ends, each of multiplicity qkq_{k}, then the contribution to CZInd(C)CZ^{Ind}(C) from γ\gamma is given by kCZ(γqk)\sum_{k}CZ(\gamma^{q_{k}}). Similarly if CC is asymptotic to γ\gamma at the negative ends, then its contribution to CZInd(C)CZ^{Ind}(C) is kCZ(γqk)-\sum_{k}CZ(\gamma^{q_{k}}).

Theorem 2.6.

Let CC denote a somewhere injective JJ-holomorphic curve as above, then we have the following inequality

Ind(C)I(C)2δ(C).Ind(C)\leq I(C)-2\delta(C). (6)

An immediate corollary of the above is

Corollary 2.7.

Let 𝒞\mathcal{C} be a JJ-holomorphic current of I(𝒞)=1I(\mathcal{C})=1. Then for generic JJ, the current 𝒞\mathcal{C} must satisfy

  1. a.

    It contains an unique connected embedded curve CC of multiplicity one that is not a trivial cylinder. The ends of CC approach Reeb orbits according to partition conditions. (See [Hut14, Section 3] for a discussion of partition conditions). We will review the relevant partition conditions in the Morse-Bott setting later).

  2. b.

    The other components of 𝒞\mathcal{C} are trivial cylinders with multiplicities.

Convention 2.8.

In this paper we describe a correspondence between ECH index 1 currents in the nondegenerate setting and ECH index 1 cascades in the Morse-Bott setting. We will only care about the nontrivial part of the ECH index 1 current, as the trivial cylinders correspond trivially in the non-degenerate and Morse-Bott situations. Hence when we say cascade, or a sequences of ECH index one curves/currents degenerating into a cascade, unless stated otherwise, we will always be considering what happens to the nontrivial part of the ECH index one current, and what cascade it corresponds to.

2.5 J0J_{0} index and finiteness

We recall (without proof) the following proposition (see [Hut02],[Hut14]):

Proposition 2.9.

Let α,β\alpha,\beta be ECH generators. We choose a generic JJ, and let I=1(α,β)/\mathcal{M}^{I=1}(\alpha,\beta)/\mathbb{R} denote the moduli space of ECH index =1=1 currents from α\alpha to β\beta modulo the action of \mathbb{R}. Then I=1(α,β)/\mathcal{M}^{I=1}(\alpha,\beta)/\mathbb{R} is a finite collection of points.

We will mention two results that go into this proof, for we will need analogous constructions in the Morse-Bott context.

Definition 2.10.

Let α={(αi,mi)},β={(βi,ni)}\alpha=\{(\alpha_{i},m_{i})\},\beta=\{(\beta_{i},n_{i})\} be ECH generators, let ZH2(α,β,Y)Z\in H_{2}(\alpha,\beta,Y) be a relative homology class. We define:

J0(α,β,Z)=cτ(Z)+Qτ(Z)+CZJ0(α,β)J_{0}(\alpha,\beta,Z)=-c_{\tau}(Z)+Q_{\tau}(Z)+CZ^{J_{0}}(\alpha,\beta) (7)

where

CZJ0(α,β):=ik=1mi1CZ(αik)ik=1ni1CZ(βik)CZ^{J_{0}}(\alpha,\beta):=\sum_{i}\sum_{k=1}^{m_{i}-1}CZ(\alpha_{i}^{k})-\sum_{i}\sum_{k=1}^{n_{i}-1}CZ(\beta_{i}^{k}) (8)

We have the following proposition bounding the topological complexity of holomorphic curves counted by ECH index 1 conditions:

Proposition 2.11.

Let 𝒞I=1(α,β)\mathcal{C}\in\mathcal{M}^{I=1}(\alpha,\beta), which decomposes as 𝒞=C0C\mathcal{C}=C_{0}\cup C where C0C_{0} is a union of trivial cylinders, and CC is somewhere injective and nontrivial. Let ni+n_{i}^{+} denote the number of positive ends CC has at αi\alpha_{i}, plus 1 if C0C_{0} includes cylinders of the form ×αi\mathbb{R}\times\alpha_{i}, define njn_{j}^{-} analogously for β\beta and negative ends of CC then

χ(C)+i(ni+1)+j(nj1)J0(C).-\chi(C)+\sum_{i}(n_{i}^{+}-1)+\sum_{j}(n_{j}^{-}-1)\leq J_{0}(C). (9)

Finally we state the version of Gromov compactness for currents. Let α,β\alpha,\beta be orbit sets, we define a broken holomorphic current from α,β\alpha,\beta to be a finite sequence of JJ nontrivial holomorphic currents (𝒞0,..,𝒞k)(\mathcal{C}^{0},..,\mathcal{C}^{k}) in ×Y\mathbb{R}\times Y such that there exists orbit sets α=γ0,γ1,..,γk+1=β\alpha=\gamma^{0},\gamma^{1},..,\gamma^{k+1}=\beta so that 𝒞i(γi,γi+1)\mathcal{C}^{i}\in\mathcal{M}(\gamma^{i},\gamma^{i+1}) (this notation means 𝒞i\mathcal{C}^{i} is a current from the orbit set γi\gamma^{i} to γi+1\gamma^{i+1}). By nontrivial we mean a current is not entirely composed of unions of trivial cylinders. We say a sequence of holomorphic currents {𝒞v1}(α,β)\{\mathcal{C}_{v\geq 1}\}\in\mathcal{M}(\alpha,\beta) converges to (𝒞0,..,𝒞k)(\mathcal{C}^{0},..,\mathcal{C}^{k}) if for each i=0,..,ki=0,..,k there are representatives 𝒞νi\mathcal{C}_{\nu}^{i} of 𝒞ν(α,β)/\mathcal{C}_{\nu}\in\mathcal{M}(\alpha,\beta)/\mathbb{R} such that the sequence {𝒞v1}\{\mathcal{C}_{v\geq 1}\} converges as a current and as a point set on compact sets to 𝒞i\mathcal{C}^{i}.

Proposition 2.12.

([Hut14], [Tau98] Prop 3.3 ) Any sequence {𝒞v}\{\mathcal{C}_{v}\} of holomorphic currents in (α,β)/\mathcal{M}(\alpha,\beta)/\mathbb{R} has a subsequence which converges to a broken holomorphic current (𝒞0,..,𝒞k)(\mathcal{C}^{0},..,\mathcal{C}^{k}). Further if we denote {𝒞v}\{\mathcal{C}_{v}\} the convergent subsequence, we have the equality

[𝒞v]=i=0k[𝒞i]H2(α,β,Y)[\mathcal{C}_{v}]=\sum_{i=0}^{k}[\mathcal{C}^{i}]\in H_{2}(\alpha,\beta,Y) (10)

3 Morse-Bott setup and SFT type compactness

Let (Y,λ)(Y,\lambda) be a contact 3 manifold with Morse-Bott contact form λ\lambda. Throughout we assume the Morse-Bott orbits come in families of tori.

Convention 3.1.

Throughout this paper we fix action level L>0L>0 and only consider ECH generators of action level up to LL. This is implicit in all of our constructions and will not be mentioned further. We construct Morse-Bott ECH up to action level LL, and the full ECH is recovered by taking LL\rightarrow\infty.

The following theorem, which is a special case of a more general result in [OW18], gives a characterization of the neighborhood of Morse-Bott Tori. Let λ0\lambda_{0} denote the standard contact form on (z,x,y)S1×S1×(z,x,y)\in S^{1}\times S^{1}\times\mathbb{R} of the form

λ0=dzydx.\lambda_{0}=dz-ydx.
Proposition 3.2.

[OW18] Let (Y,λ)(Y,\lambda) be a contact 3 manifold with Morse-Bott contact form λ\lambda. We assume the Morse-Bott orbits come in families of tori 𝒯i\mathcal{T}_{i} with minimal period TiT_{i}. Then we can choose coordinates around each Morse-Bott torus so that a neighborhood of 𝒯i\mathcal{T}_{i} is described by S1×S1×(ϵ,ϵ)S^{1}\times S^{1}\times(-\epsilon,\epsilon), and the contact form λ\lambda in this coordinate system looks like:

λ=h(x,y,z)λ0\lambda=h(x,y,z)\lambda_{0}

where h(x,y,z)h(x,y,z) satisfies:

h(x,0,z)=1,dh(x,0,z)=0h(x,0,z)=1,dh(x,0,z)=0

Here we identify zS1/2πTiz\in S^{1}\sim\mathbb{R}/2\pi T_{i}\mathbb{Z}

See [Yao22] Theorem Proposition 2.2 for a sketch of the proof. By the Morse-Bott assumption there are only finitely many such tori up to fixed action LL. We assume we have chosen such neighborhoods around all Morse Bott Tori 𝒯i{\mathcal{T}_{i}}. Next we shall perturb them to nondegenerate Reeb orbits by perturbing the contact form in a neighborhood of each torus as described below. This is the same perturbation as in [Yao22].

Let δ>0\delta>0, let f:x/f:x\in\mathbb{R}/\mathbb{Z}\rightarrow\mathbb{R} be a smooth Morse function with maximum at x=1/2x=1/2 and minimum x=0x=0. Let g(y):g(y):\mathbb{R}\rightarrow\mathbb{R} be a bump function that is equal to 11 on [ϵ𝒯i,ϵ𝒯i][-\epsilon_{\mathcal{T}_{i}},\epsilon_{\mathcal{T}_{i}}] and zero outside [2ϵ𝒯i,2ϵ𝒯i][-2\epsilon_{\mathcal{T}_{i}},2\epsilon_{\mathcal{T}_{i}}]. Here ϵ𝒯i\epsilon_{\mathcal{T}_{i}} is a small number chosen for each 𝒯i\mathcal{T}_{i} small enough so that the normal form in the above theorem applies to all Morse-Bott tori of action <L<L, and that all such chosen neighborhoods these Morse-Bott tori are disjoint. Then in neighborhood of the Morse-Bott tori 𝒯i\mathcal{T}_{i} we perturb the contact form as

λλδ:=eδgfλ.\lambda\longrightarrow\lambda_{\delta}:=e^{\delta gf}\lambda.

We can describe the change in Reeb dynamics as follows:

Proposition 3.3.

For fixed action level L>0L>0 there exists δ>0\delta>0 small enough so that the Reeb dynamics of λδ\lambda_{\delta} can be described as follows. In the trivialization specified by Proposition 17, each Morse-bott torus splits into two non-degenerate Reeb orbits corresponding to the two critical points of ff. One of them is hyperbolic of index 0, the other is elliptic with rotation angle |θ|<Cδ<<1|\theta|<C\delta<<1 and hence its Conley-Zehnder index is ±1\pm 1. There are no additional Reeb orbits of action <L<L.

For proof see [Bou02].

Remark 3.4.

Later when we define various terms in the ECH index, they will depend on the choice of trivializations of the contact structure on the Reeb orbits. We will always choose the trivialization specified by Proposition 3.2. For convenience of notation we will call this trivialization τ\tau and write for example cτc_{\tau} or QτQ_{\tau} for the definition of relative Chern class or intersection form with respect to this trivialization.

We also observe that after iterating the Reeb orbit in the Morse-Bott tori, their Robbin-Salamon index stays the same ([Gut14]). So up to action LL, in the nondegenerate picture, we will only see Reeb orbits of Conley-Zehnder index 1,0,1-1,0,1.

Definition 3.5.

We say a Morse Bott torus is positive if the elliptic Reeb orbit has Conley-Zehnder index 1 after perturbation; otherwise we say it is negative Morse Bott torus. This condition is intrinsic to the Morse-Bott torus itself, and is independent of trivializations or our choice of perturbations.

We recall our goal is to define the ECH chain complex up to filtration LL, and then take LL\rightarrow\infty to recover the entire ECH chain complex. Hence, let us consider for small δ>0\delta>0 the symplectization

(M4,ωδ):=(×Y3,d(esλδ))(M^{4},\omega_{\delta}):=(\mathbb{R}\times Y^{3},d(e^{s}\lambda_{\delta}))

We equip (M,ωδ)(M,\omega_{\delta}) with a λδ\lambda_{\delta} compatible almost complex structure JδJ_{\delta}, and (M,ω):=(×Y3,d(esλδ))(M,\omega):=(\mathbb{R}\times Y^{3},d(e^{s}\lambda_{\delta})) with λ\lambda-compatible almost complex structure JJ. Both JJ and JδJ_{\delta} should be chosen generically, with genericity condition specified in Definition 4.5 and Theorem 7.3. In particular JδJ_{\delta} should be a small perturbation of JJ, i.e. the CC^{\infty} norm difference between JδJ_{\delta} and JJ should be bounded above by CδC\delta. For fixed LL and small enough and generic choice of δ\delta, the ECH of (Y3,λδ)(Y^{3},\lambda_{\delta}) is defined for generators of action less than LL via counts of embedded J-holomorphic curves of ECH index 1. To motivate our construction, we next take δ0\delta\rightarrow 0 to see what kinds of objects these JJ holomorphic curves degenerate into. By a theorem of that first appeared in Bourgeois’ thesis [Bou02] and also stated in [Bou+03] (for a proof see the Appendix of [Yao22]), they degenerate into JJ-holomorphic cascades. (For a more careful definition of cascades see the appendix of [Yao22] that takes into account of stability of domain and marked points, but the definition here suffices for our purposes).

Definition 3.6 ( [Bou02], See also definition 2.7 in [Yao22]).

Let Σ\Sigma be a punctured (nodal) Riemann surface, potentially with multiple connected components. A cascade of height 1, which we will denote by uu^{\text{\Lightning}}, in (×Y3,d(esλ)(\mathbb{R}\times Y^{3},d(e^{s}\lambda) consists of the following data :

  • A labeling of the connected components of Σ=Σ{nodes}\Sigma^{*}=\Sigma\setminus\{\text{nodes}\} by integers in {1,,l}\{1,...,l\}, called sublevels, such that two components sharing a node have sublevels differing by at most 1. We denote by Σi\Sigma_{i} the union of connected components of sublevel ii, which might itself be a nodal Riemann surface.

  • Ti[0,)T_{i}\in[0,\infty) for i=1,,l1i=1,...,l-1.

  • JJ-holomorphic maps ui:(Σi,j)(×Y3,J)u^{i}:(\Sigma_{i},j)\rightarrow(\mathbb{R}\times Y^{3},J) with E(ui)<E(u_{i})<\infty for i=1,,li=1,...,l, such that:

    • Each node shared by Σi\Sigma_{i} and Σi+1\Sigma_{i+1}, is a negative puncture for uiu^{i} and is a positive puncture for ui+1u^{i+1}. Suppose this negative puncture of uiu^{i} is asymptotic to some Reeb orbit γi𝒯\gamma_{i}\in\mathcal{T}, where 𝒯\mathcal{T} is a Morse-Bott torus, and this positive puncture of ui+1u^{i+1} is asymptotic to some Reeb orbit γi+1𝒯\gamma_{i+1}\in\mathcal{T}, then we have that ϕfTi(γi+1)=γi\phi^{T_{i}}_{f}(\gamma_{i+1})=\gamma_{i}. Here ϕfTi\phi^{T_{i}}_{f} is the upwards gradient flow of ff for time TiT_{i} lifted to the Morse-Bott torus 𝒯\mathcal{T}. It is defined by solving the ODE

      ddsϕf(s)=f(ϕf(s)).\frac{d}{ds}\phi_{f}(s)=f^{\prime}(\phi_{f}(s)).
    • uiu^{i} extends continuously across nodes within Σi\Sigma_{i}.

    • No level consists purely of trivial cylinders. However we will allow levels that consist of branched covers of trivial cylinders.

Convention 3.7.

We fix our conventions as in [Yao22].

  • We say the punctures of a JJ-holomorphic curve that approach Reeb orbits as ss\rightarrow\infty are positive punctures, and the punctures that approach Reeb orbits as ss\rightarrow-\infty are negative punctures. We will fix cylindrical neighborhoods around each puncture of our JJ-holomorphic curves, so we will use “positive/negative ends” and “positive/negative punctures” interchangeably. By our conventions, we think of u1u^{1} as being a level above u2u^{2} and so on.

  • We refer to the Morse-Bott tori 𝒯j\mathcal{T}_{j} that appear between adjacent levels of the cascade {ui,ui+1}\{u^{i},u^{i+1}\} as above, where negative punctures of uiu^{i} are asymptotic to Reeb orbits that agree with positive punctures from ui+1u^{i+1} up to a gradient flow, intermediate cascade levels.

  • We say that the positive asymptotics of uu^{\text{\Lightning}} are the Reeb orbits we reach by applying ϕf\phi_{f}^{\infty} to the Reeb orbits hit by the positive punctures of u1u^{1}. Similarly, the negative asymptotics of uu^{\text{\Lightning}} are the Reeb orbits we reach by applying ϕf\phi_{f}^{-\infty} to the Reeb orbits hit by the negative punctures of ulu^{l}. They are always Reeb orbits that correspond to critical points of ff. We note if a positive puncture (resp. negative puncture) of u1u^{1} (resp. ulu^{l}) is asymptotic to a Reeb orbit corresponding to a critical point of ff, then applying ϕf+\phi^{+\infty}_{f} (resp. ϕf\phi_{f}^{-\infty}) to this Reeb orbit does nothing.

Definition 3.8 ([Bou02], Chapter 4, See also definition 2.9 in [Yao22] ).

A cascade of height kk consists of kk height 1 cascades, uk={u1,,uk}u^{\text{\Lightning}}_{k}=\{u^{1\text{\Lightning}},...,u^{k\text{\Lightning}}\} with matching asymptotics concatenated together.

By matching asymptotics we mean the following. Consider adjacent height one cascades, uiu^{i\text{\Lightning}} and ui+1u^{i+1\text{\Lightning}}. Suppose a positive end of the top level of ui+1u^{i+1\text{\Lightning}} is asymptotic to the Reeb orbit γ\gamma (not necessarily simply covered). Then if we apply the upwards gradient flow of ff for infinite time we arrive at a Reeb orbit reached by a negative end of the bottom level of uiu^{i\text{\Lightning}}. We allow the case where γ\gamma is at a critical point of ff, and the flow for infinite time is stationary at γ\gamma. We also allow the case where γ\gamma is at the minimum of ff, and the negative end of the bottom level of uiu^{i\text{\Lightning}} is reached by following an entire (upwards) gradient trajectory connecting from the minimum of ff to its maximum. If all ends between adjacent height one cascades are matched up this way, then we say they have matching asymptotics.

We will use the notation uku^{\text{\Lightning}}_{k} to denote a cascade of height kk. We will mostly be concerned with cascades of height 1 in this article, so for those we will drop the subscript kk and write u={u1,,ul}u^{\text{\Lightning}}=\{u^{1},...,u^{l}\}.

Remark 3.9.

As mentioned in [Yao22], we can also think of a cascade of height kk as a cascade of height 1 where k1k-1 of the intermediate flow times are infinite.

We now state a SFT style compactness theorem relating non-degenerate JδJ_{\delta} holomorphic curves to cascades. However, the precise statement is rather technical and requires us to take up Deligne-Mumford compactifications of the moduli space of Riemann surfaces. The full version is stated in [Bou+03] (see also the Appendix of [Yao22], where we also sketch a proof). For our purposes it will be sufficient to state the theorem informally as below.

Theorem 3.10.

(See [Bou+03]) Let uδnu_{\delta_{n}} be a sequence of JδnJ_{\delta_{n}}-holomorphic curves with uniform upper bound on genus and energy, then a subsequence of uδnu_{\delta_{n}} converges to a cascade of JJ- holomorphic curves (which can be apriori of arbitrary height).

Since ECH is really a theory of holomorphic currents, we find it also useful to define a cascade of holomorphic currents, which is what we shall primarily work with.

Definition 3.11.

A height 1 holomorphic cascade of currents 𝐮={u1,..,un}\mathbf{u}^{\text{\Lightning}}=\{u^{1},..,u^{n}\} consists of the following data:

  • Each uiu^{i} consists of holomorphic currents of the form (Cji,dji)(C^{i}_{j},d_{j}^{i}). Each CjiC^{i}_{j} is a somewhere injective holomorphic curve with E(Cji)<E(C^{i}_{j})<\infty. The positive integer djid^{i}_{j} is then the multiplicity.

  • Numbers Ti[0,),i=1,..,n1T_{i}\in[0,\infty),i=1,..,n-1

  • Let γi\gamma_{i} be a simply covered Reeb orbit that is approached by the negative end of some component of uiu^{i}, say the components Cj1i,,CjkiC^{i}_{j_{1}},...,C^{i}_{j_{k}} (such curves have associated multiplicity dj1i,,djkid^{i}_{j_{1}},...,d^{i}_{j_{k}}). Each CjiC^{i}_{j_{*}} approaches γi\gamma_{i} with a covering multiplicity njn_{j_{*}}, which is how many times γi\gamma_{i} is covered by CjiC^{i}_{j_{*}} as currents. Then the total multiplicity of γi\gamma_{i} as covered by uiu^{i} is given by =1,..kdjinj\sum_{*=1,..k}d^{i}_{j_{*}}n_{j_{*}}. Then consider ϕfTi(γi+1):=γi\phi_{f}^{T_{i}}(\gamma_{i+1}):=\gamma_{i}. Then ui+1u^{i+1} is asymptotic to γi+1\gamma^{i+1} in its positive end with total multiplicity =1,..kdjinj\sum_{*=1,..k}d^{i}_{j_{*}}n_{j_{*}} also.

  • No level consists of purely of trivial cylinders (even if they have higher multiplicities).

We define the positive asymptotics of 𝐮:={u1,..,un}\mathbf{u}^{\text{\Lightning}}:=\{u^{1},..,u^{n}\} as before, except we only care about Reeb orbits up to multiplicity. Then we can similarly define a cascade of currents of height kk by stacking together cascades of currents of height 11.

We will refer to ordinary cascade a “cascade of curves” when we wish to distinguish it from a cascade of currents. Then given a cascade of curves, we can pass it to a cascade of currents by using the following procedure:

Procedure 3.12.
  • Replace every multiple covered non-trivial curve with a current of the form (C,m)(C,m) where CC is a somewhere injective curve, and we translate all mm copies along \mathbb{R} to make the entire collection somewhere injective.

  • If we see a multiply covered trivial cylinder we replace it with (C,m)(C,m) where mm is the multiplicity and CC is a trivial cylinder.

  • If we see a nodal curve in one of the levels, we separate the node and apply the above process to each of the separated components of the nodal curve.

  • We remove all levels that only have currents made out of trivial cylinders. Suppose uiu^{i} is a level only consisting of trivial cylinders to be removed, and suppose the s+s\rightarrow+\infty end is a intermediate cascade level with flow time Ti1T_{i-1}, and the ss\rightarrow-\infty end of uiu^{i} has associated flow time TiT_{i}, after the removal of uiu^{i} level, the newly adjacent levels ui1u^{i-1} and ui+1u^{i+1} have flow time between them equal to Ti+Ti1T_{i}+T_{i-1}.

In passing from cascades of curves to currents we have lost some information, but we shall see currents are the natural settings to talk about ECH index.

We later wish to make sense of the Fredholm index of a cascade of currents. To this end we make the definition of reduced cascade of currents.

Definition 3.13.

Given a cascade of currents 𝐮\mathbf{u}^{\text{\Lightning}}, for components within it of the form (C,m)(C,m) where m>1m>1 and CC is a nontrivial holomorphic curve, we then replace (C,m)(C,m) with just (C,1)(C,1). After we perform this operation we obtain another cascade of currents, which we label 𝐮~\tilde{\mathbf{u}}^{\text{\Lightning}}, which we call the reduced cascade of currents.

4 Index calculations and transversality

The heart of the calculation that underlies ECH is this: the ECH index bounds from the above the Fredholm index, and if there are curves of ECH index one with bad behaviour (singularities, multiply covers), this would imply the existence of somewhere injective curves of Fredholm index less than 1, which cannot happen for generic JJ. In this section we take up the issue of establishing Fredholm index for JJ holomorphic cascades, and explain the transversality issue we encounter.

Given a reduced cascades of currents, 𝐮~={u~1,,u~n}\mathbf{\tilde{u}}^{\text{\Lightning}}=\{\tilde{u}^{1},...,\tilde{u}^{n}\}, we would like to assign to it a Fredholm index. Ideally this Fredholm index measures geometrically the dimension of the moduli space this particular cascade lives in. We note that by passing to the reduced cascade the multiplicities associated to ends of adjacent levels, u~i\tilde{u}^{i} and u~i+1\tilde{u}^{i+1} do not necessarily match up, but by imposing there is a single flow time parameter TiT_{i} between adjacent levels still means we can think of 𝐮\mathbf{u}^{\text{\Lightning}} as living in a fiber product with virtual dimension.

To this end we first recall some conventions when it comes to JJ-holomorphic curves with ends on Morse-Bott critical submanifolds (in this case, tori). Consider u~i\tilde{u}^{i}, for simplicity suppose its domain Σ˙i\dot{\Sigma}_{i} is a punctured Riemann surface that is connected. Let pj±p_{j}^{\pm} label the positive/negative punctures, and the map u~i\tilde{u}^{i} is asymptotic to Reeb orbits (of some multiplicity) on Morse-Bott tori at each of its punctures. We wish to associate to u~i\tilde{u}^{i} a moduli space of curves that contain u~i\tilde{u}^{i} as an element and contains curves that are “close” to u~i\tilde{u}^{i}. To this end we recall some conventions.

To each puncture pj±p_{j}^{\pm} of u~i\tilde{u}^{i}, we can designated it as “fixed” or “free”, and each choice of these designations leads to a different moduli space. The designation “free” means we consider JJ-holomorphic maps from Σ˙i\dot{\Sigma}_{i} so that pj±p_{j}^{\pm} can land on any Reeb orbit with the same multiplicity on the same Morse-Bott torus at the corresponding end of u~i\tilde{u}^{i}. For a puncture to be considered “fixed”, we consider moduli space of JJ-holomorphic curves from Σ˙i\dot{\Sigma}_{i} so that pj±p_{j}^{\pm} lands on a fixed Reeb orbits on a Morse-Bott torus with fixed multiplicity (the same Reeb orbit as u~i\tilde{u}^{i}). Given a designation of “fixed” or “free” on punctures of u~i\tilde{u}^{i}, we can then consider the moduli space of JJ holomorphic curves from Σ˙i\dot{\Sigma}_{i} into ×Y\mathbb{R}\times Y with the same asymptotic constraints as u~i\tilde{u}^{i} and living in the same relative homology class. We shall denote this moduli space as 𝐜(u~i)\mathcal{M}_{\mathbf{c}}(\tilde{u}^{i}), using 𝐜\mathbf{c} to denote our choice of fixed/free ends. This moduli space has virtual dimension given by:

Ind(u~i):=χ(u~i)+2c1(u~i)+pj+μ(γqpj+)pjμ(γqpj)+12#free ends12#fixed endsInd(\tilde{u}^{i}):=-\chi(\tilde{u}^{i})+2c_{1}(\tilde{u}^{i})+\sum_{p_{j}^{+}}\mu(\gamma^{q_{p_{j}^{+}}})-\sum_{p_{j}^{-}}\mu(\gamma^{q_{p_{j}^{-}}})+\frac{1}{2}\#\text{free ends}-\frac{1}{2}\#\text{fixed ends} (11)

where χ\chi is the Euler characteristic, c1c_{1} the relative first Chern class, μ()\mu(-) is the Robbin Salamon index for path of symplectic matrices with degeneracies defined in [Gut14]. We use the symbol γ\gamma to denote the Reeb orbit the end pj±p_{j}^{\pm} is asymptotic to, with multiplicity qpj±q_{p_{j}^{\pm}}.

Given a reduced cascade of currents, 𝐮~\mathbf{\tilde{u}}^{\text{\Lightning}}, let α\alpha denote the designation of “free”/“fixed” ends of u~1\tilde{u}^{1} at the s+s\rightarrow+\infty end, and let β\beta denote the “fixed”/“free” designation of u~n\tilde{u}^{n} at the ss\rightarrow-\infty end. Later we will see we can replace α\alpha and β\beta with Morse-Bott ECH generators. In order to define the Fredholm index we need to assign free/fixed ends to the rest of the ends.

Convention 4.1.

If a non trivial curve uiu^{i} has an end landing on a critical point of ff, then we consider that end to be fixed. If a trivial cylinder has one end on critical point of ff, the other end must also land on the same critical point. We allow trivial cylinders with both ends free. If the trivial cylinder is at a critical point of ff, we take the convention we can only designate one of its ends as fixed.

Definition 4.2.

Let 𝐮~={u1,..,un1}\tilde{\mathbf{u}}^{\text{\Lightning}}=\{u^{1},..,u^{n-1}\} denote a reduced cascade of currents of height 1. Let ind(ui)ind(u^{i}) denote the Fredholm index of each of uiu^{i}. Note this makes sense since we have assigned free/fixed ends to all ends of uiu^{i} by our conventions above.

Suppose there are R2,,Rn1R_{2},...,R_{n-1}\in\mathbb{Z} distinct Reeb orbits approached by free ends as ss\rightarrow-\infty at each intermediate cascade level. Let us denote kik_{i} and kik_{i}^{\prime} the number of free ends in each intermediate cascade level. e.g. elements in u1u^{1} has k2k_{2} free ends as ss\rightarrow-\infty, and u2u^{2} has k2k_{2}^{\prime} free ends as s+s\rightarrow+\infty. Both counts of kik_{i} and kik_{i}^{\prime}, as well as RiR_{i} ignores “free” ends of fixed trivial cylinders, as such “free” ends are artificial to our convention. Now we define the cascade dimension

Ind(𝐮~):=\displaystyle Ind(\tilde{\mathbf{u}}^{\text{\Lightning}}):= Ind(u1)+..+Ind(un1)\displaystyle Ind(u^{1})+..+Ind(u^{n-1})
[k2+kn1][k2++kn1]+[R2+..+Rn1]+(n2)(n1)L\displaystyle-[k_{2}^{\prime}...+k_{n-1}^{\prime}]-[k_{2}+...+k_{n-1}]+[R_{2}+..+R_{n-1}]+(n-2)-(n-1)-L

where LL is the number of intermediate cascade levels without free ends plus the number of intermediate cascade levels whose flow time is zero. Again in the count of LL we ignore “free” ends coming from fixed trivial cylinders.

Observe for (reduced) cascades of height 11, we always have kiRik_{i}\geq R_{i} and kiRik_{i}^{\prime}\geq R_{i}.

We next explain how to define/compute the dimension of height kk cascades. Let 𝐮~={u1,..,un1}\tilde{\mathbf{u}}^{\text{\Lightning}}=\{u^{1},..,u^{n-1}\} denote a reduced cascade of currents of height NN. We recall the difference between height one and height NN cascade is that between cascade levels uiu^{i} and ui+1u^{i+1} we allow flow times Ti=T_{i}=\infty. We assign the free/fixed ends to uiu^{i} depending on whether they land on critical points of ff as before. We can split a height NN cascade into NN height 1 cascades by partitioning the levels where the flow times are infinite. In particular we write 𝐮~={𝐯𝟏~,,𝐯𝐍~}\tilde{\mathbf{u}}^{\text{\Lightning}}=\left\{\tilde{\mathbf{v^{1}}}^{\text{\Lightning}},...,\tilde{\mathbf{v^{N}}}^{\text{\Lightning}}\right\}. Then the index of 𝐮~\tilde{\mathbf{u}}^{\text{\Lightning}} is given by the sum of the indices of 𝐯𝐢~\tilde{\mathbf{v^{i}}}^{\text{\Lightning}}.

Here we come to the key transversality assumption of this paper. We first make sense of the notion of transversality.

Definition 4.3.

Let λ\lambda be a Morse-Bott contact form, whose Reeb orbits come in S1S^{1} families. We say a λ\lambda compatible almost complex structure JJ is good if all reduced cascades of height one are tranversely cut out, which is defined below.

Remark 4.4.

We note the transversality conditions needed to count cascades given below are quite natural. However, since cascades have many parts the notation is bit complicated.

Definition 4.5.

Let 𝐮~={u1,..,un1}\tilde{\mathbf{u}}^{\text{\Lightning}}=\{u^{1},..,u^{n-1}\} denote a reduced cascade of currents of height 1.

We say 𝐮~={u1,..,un1}\tilde{\mathbf{u}}^{\text{\Lightning}}=\{u^{1},..,u^{n-1}\} is transversely cut out if the conditions below are met.

  • Each moduli space c(ui)\mathcal{M}_{c}(u^{i}) is transversely cut out with dimension given by the Fredholm index formula. Here the subscript cc implicitly denotes the assignments of fixed and free ends we assigned to each end of uiu^{i} according to Convention 4.1. Note fixed trivial cylinders are assigned index zero.

Suppose there are R2,,Rn1R_{2},...,R_{n-1}\in\mathbb{Z} distinct Reeb orbits reached by free ends at each intermediate cascade level. We label them by γ(i,j)\gamma(i,j) where j=1,,Rij=1,...,R_{i}, and ii indexes which level we are referring to. For each γ(i,j)\gamma(i,j), we choose a negative puncture of ui1u^{i-1} that is asymptotic to γ(i,j)\gamma(i,j). We call this puncture p(i1,j)p^{-}(i-1,j). The other negative ends of ui1u^{i-1} that are asymptotic to γ(i,j)\gamma(i,j) are labelled p(i1,j,c,l)p^{-}(i-1,j,c,l), where l=1,2..,n(γ(i,j),)l=1,2..,n(\gamma(i,j),-). Next consider ϕTi1(γ(i,j)))\phi^{-T_{i-1}}(\gamma(i,j))). They are approached by positive punctures of uiu^{i}. For each ϕTi1(γ(i,j)))\phi^{-T_{i-1}}(\gamma(i,j))), we pick out a special free puncture p+(i,j)p^{+}(i,j). The remaining free positive ends of uiu^{i} that are asymptotic to ϕTi1(γ(i,j)))\phi^{-T_{i-1}}(\gamma(i,j))) are labelled p+(i,j,c,l)p^{+}(i,j,c,l) for l=1,,n(γ(i,j),+)l=1,...,n(\gamma(i,j),+).

We next consider the evaluation maps. Given the collection of flow times T1,,Tn1T_{1},...,T_{n-1}. Let {1,..,n1}\mathfrak{I}\subset\{1,..,n-1\} denote the subset for which Ti>0T_{i}>0, we consider the evaluation map

EV:(u1)×(u2)××(un2)(S1)R2×(S1)R3××(S1)Rn1\displaystyle EV^{-}:\mathcal{M}(u^{1})\times\mathcal{M}(u^{2})\times...\times\mathcal{M}(u^{n-2})\rightarrow(S^{1})^{R_{2}}\times(S^{1})^{R_{3}}\times...\times(S^{1})^{R_{n-1}} (12)

given by

(u1,,un2)(ev1(u1),ev2(u2),,evn2(un2))\displaystyle(u^{\prime 1},...,u^{\prime n-2})\rightarrow(ev_{1}^{-}(u^{\prime 1}),ev_{2}^{-}(u^{\prime 2}),...,ev_{n-2}^{-}(u^{\prime n-2})) (13)

Here the evaluation is at the p(i1,j)p^{-}(i-1,j) puncture of ui1u^{i-1}. We also consider the map

EV+:(u2)×(u3)×(un1)(S1)R2×(S1)R3××(S1)Rn1\displaystyle EV^{+}:\mathcal{M}(u^{2})\times\mathcal{M}(u^{3})...\times\mathcal{M}(u^{n-1})\rightarrow(S^{1})^{R_{2}}\times(S^{1})^{R_{3}}\times...\times(S^{1})^{R_{n-1}} (14)

given by:

(u2,,un1)(ev2+(u2),,evn1+(un1))\displaystyle(u^{\prime 2},...,u^{\prime n-1})\rightarrow(ev_{2}^{+}(u^{\prime 2}),...,ev_{n-1}^{+}(u^{\prime n-1})) (15)

where the evaluation is at p+(i,j)p^{+}(i,j) of uiu^{i}. We consider the flow map

Φf:(S1)R2××..×(S1)Rn1×(S1)R2×(S1)R3××(S1)Rn1.\Phi_{f}:(S^{1})^{R_{2}}\times\mathbb{R}^{*}\times..\times(S^{1})^{R_{n-1}}\times\mathbb{R}^{*}\rightarrow(S^{1})^{R_{2}}\times(S^{1})^{R_{3}}\times...\times(S^{1})^{R_{n-1}}.

The notation \mathbb{R}^{*} means the following: if ii\in\mathfrak{I} then we include a factor of \mathbb{R} in the above product, otherwise we omit the factor. For xiS1x_{i}\in S^{1} (i.e. a copy of S1S^{1} among the product (S1)Ri(S^{1})^{R_{i}}), if ii\in\mathfrak{I} then the image of xix_{i} under Φf\Phi_{f} is given by ϕfTi(xi)\phi_{f}^{T_{i}^{\prime}}(x_{i}). If the index ii is not in \mathfrak{I}, then the image under Φf\Phi_{f} is xix_{i}. We use the notation ΦfEV+\Phi_{f}\circ EV^{+} to denote the composition of the two maps, with domain (u2)××(u2)×(un1)×\mathcal{M}(u^{2})\times\mathbb{R}^{*}\times\mathcal{M}(u^{2})...\times\mathcal{M}(u^{n-1})\times\mathbb{R}^{*} and image (S1)R2×(S1)R3××(S1)Rn1(S^{1})^{R_{2}}\times(S^{1})^{R_{3}}\times...\times(S^{1})^{R_{n-1}}.

Let 𝒦\mathcal{K}_{-} denote the subset of (u1)×(u2)××(un2)\mathcal{M}(u^{1})\times\mathcal{M}(u^{2})\times...\times\mathcal{M}(u^{n-2}) so that the ends p(i,j)p^{-}(i,j) and p(i,j,c,l)p^{-}(i,j,c,l) approach the same Reeb orbit. Let 𝒦+\mathcal{K}_{+} denote the subset of (u2)×(u3)×(un1)\mathcal{M}(u^{2})\times\mathcal{M}(u^{3})...\times\mathcal{M}(u^{n-1}) where p+(i,j)p^{+}(i,j) and p+(i,j,c,l)p^{+}(i,j,c,l) are asymptotic to the same Reeb orbit. Then

  • Near 𝐮~\mathbf{\tilde{u}}^{\text{\Lightning}}, both 𝒦±\mathcal{K}_{\pm} are transversly cut out submanifolds.

Then we can restrict EV±EV^{\pm} to 𝒦±\mathcal{K}_{\pm}, in particular the map ΦfEV+\Phi_{f}\circ EV^{+} admits a natural restriction to 𝒦×||\mathcal{K}_{-}\times\mathbb{R}^{|\mathfrak{I}|}, our final condition is:

  • ΦfEV+\Phi_{f}\circ EV^{+} and EVEV^{-}, when restricted to 𝒦+×()||\mathcal{K}_{+}\times(\mathbb{R})^{|\mathfrak{I}|} and 𝒦\mathcal{K}_{-} respectively are transverse at 𝐮~={u1,..,un1}\tilde{\mathbf{u}}^{\text{\Lightning}}=\{u^{1},..,u^{n-1}\}

Assumption 4.6.

We assume we can choose JJ to be good so that all reduced cascades of current we encounter satisfy the transversality condition above.

In particular, this implies all reduced cascades of currents live in a moduli space whose dimension is given by the index formula, and if such index is less than zero, then such cascades cannot exist.

We note that in general the transversality assumption is not automatic. In a reduced cascades of currents, all our curves are somewhere injective, but this is not enough. The issue lies in the fact that the fiber product that defines cascade can fail to have enough transversality. This is because all different levels of the cascade have the same JJ, and this JJ cannot be perturbed independently in each level. When the cascade is complicated enough, the same curve can appear multiple times in different levels, and this causes difficulty with the evaluation map. Consequently when there is not enough transversality for the naive definition of the universal moduli space of reduced cascades to be a Banach manifold, one usually needs some additional arguments.

However in simple enough cases we can still achieve the above transversality condition. This is the content of Theorem 1.2, which is proved in the Appendix.

5 ECH Index of Cascades

In this section we develop the analogue of ECH index one condition for cascades of currents. We shall see this will impose severe limits on currents that can appear in a cascade, provided transversality can be achieved.

To start the definition, we first consider one-level cascades, i.e. holomorphic curves from Morse-Bott tori to Morse-Bott tori. We want to define an index II so that for somewhere injective curves:

I(C)dim(C)+2δ(C)I(C)\geq dim\mathcal{M}(C)+2\delta(C)

where (C)\mathcal{M}(C) denotes the moduli space of holomorphic curves CC belongs in. Note the definition of dimdim\mathcal{M} is ambiguous, because we need to specify which ends are “fixed” and which are “free”. Our definition of II will depend on the type of end conditions imposed on our curve. The key to our construction will be the relative adjunction formula.

5.1 Relative adjunction formula in the Morse-Bott setting

Here we clarify what we mean by the intersection form QQ. We first provide a provisional definition that is very much similar to regular ECH, then we show this definition descends to a more natural definition adapted to the Morse-Bott setting.

Let α,β\alpha,\beta be orbit sets. Observe here this means that they pick out discrete Reeb orbits (potentially with multiplicity) among the S1S^{1} family of Reeb orbits. Then we can define the relative intersection formula as:

Definition 5.1.

We fix trivializations of Morse-Bott tori as we have specified, and denote it by τ\tau. Given α,β\alpha,\beta orbit sets, given Z,ZH2(α,β,Y)Z,Z^{\prime}\in H_{2}(\alpha,\beta,Y) we choose τ\tau representatives SS SS^{\prime} as before, then Qτ(Z,Z)Q_{\tau}(Z,Z^{\prime}) is defined as before as the algebraic count of intersections between SS and SS^{\prime}.

Because τ\tau here provides a global trivialization of all Reeb orbits in a given Morse-Bott torus, the intersection QQ doesn’t depend on which specific Reeb orbit α\alpha or β\beta picks out in a given Morse-Bott torus. We state the phenomenon in terms of a proposition:

Proposition 5.2.

Given orbit sets α,β\alpha,\beta and relative homology classes Z,ZH2(α,β)Z,Z^{\prime}\in H_{2}(\alpha,\beta). For definiteness let γ\gamma be a Reeb orbit in the s+s\rightarrow+\infty end of α\alpha, let γ\gamma^{\prime} be any translation of γ\gamma in its Morse-Bott torus, then using γ\gamma^{\prime} to replace γ\gamma defines another orbit set α\alpha^{\prime}. There exists corresponding relative homology classes Z^,Z^H2(α,β,Y)\hat{Z},\hat{Z}^{\prime}\in H_{2}(\alpha^{\prime},\beta,Y) obtained by attaching a cylinder that connects between γ\gamma to γ\gamma^{\prime} to ends of SS and SS^{\prime} that are asymptotic to γ\gamma, then

Qτ(Z,Z)=Qτ(Z^,Z^)Q_{\tau}(Z,Z^{\prime})=Q_{\tau}(\hat{Z},\hat{Z}^{\prime})
Proof.

Choose τ\tau representatives for Z,ZZ,Z^{\prime} which we write as SS, SS^{\prime}, then attach a cylinder connecting between γ\gamma to γ\gamma^{\prime} to SS and SS^{\prime}. In our trivialization the resulting surfaces are still τ\tau representatives, and this process does not introduce additional intersections. ∎

The above proposition suggests QτQ_{\tau} in the Morse-Bott case descends to a intersection number whose input is not H2(α,β,Y)H_{2}(\alpha,\beta,Y) but a more general relative homology group adapted to the Morse-Bott setting.

Definition 5.3.

We define the relative homology classes 2(α,β,Y)\mathcal{H}_{2}(\alpha,\beta,Y). Here α,β\alpha,\beta are collections of Morse-Bott tori, and multiplicities. For instance we can write α:={(𝒯i,mi)|mi0}\alpha:=\{(\mathcal{T}_{i},m_{i})|m_{i}\in\mathbb{Z}_{\geq 0}\} where 𝒯i\mathcal{T}_{i} are Morse-Bott tori, and mim_{i} are multiplicities. A element Z2(α,β,Y)Z\in\mathcal{H}_{2}(\alpha,\beta,Y) is a 2-chain in YY so that

Z=αβ.\partial Z=\alpha-\beta.

The above equality means the boundary (which includes orientation) of ZZ consists of Reeb orbits on Morse-Bott tori {𝒯i}\{\mathcal{T}_{i}\}, and each 𝒯iα\mathcal{T}_{i}\in\alpha has a total of mim_{i} Reeb orbits (counted with multiplicity) to which the ends of ZZ are asymptotic. Likewise for β\beta. We define a equivalence relation on 2(α,β,Y)\mathcal{H}_{2}(\alpha,\beta,Y), which we write as ZZZ\sim Z^{\prime} as follows: ZZ and ZZ^{\prime} are equivalent if there is a 3-chain WW whose boundary takes the following form:

W=ZZ+{I×S1}\partial W=Z-Z^{\prime}+\{I\times S^{1}\}

where the collection {I×S1}\{I\times S^{1}\} consists of 2 chains on Morse-Bott tori that appear in either α\alpha or β\beta. We think of these 2-chains as an Reeb orbit (which we think of S1S^{1}) times an interval, II.

The idea is we consider 2-chains but allow their ends to slide along the Reeb orbits in the Morse-Bott family. The next proposition proves the relative intersection QQ remains well defined.

Proposition 5.4.

QτQ_{\tau} as defined above descends into a intersection form:

Qτ:2(α,β,Y)×2(α,β,Y).Q_{\tau}:\mathcal{H}_{2}(\alpha,\beta,Y)\times\mathcal{H}_{2}(\alpha,\beta,Y)\rightarrow\mathbb{Z}.
Proof.

For clarity we use Q^τ\hat{Q}_{\tau} to denote the intersection form defined in Definition 5.1. Suppose Z,Z2(α,β,Y)Z,Z^{\prime}\in\mathcal{H}_{2}(\alpha,\beta,Y), and suppose Z′′ZZ^{\prime\prime}\sim Z. We pick a distinguished Reeb orbit γi\gamma_{i} for each Morse-Bott torus that appears in α,β\alpha,\beta, and chosen so that γi\gamma_{i} does not appear as a Reeb orbit in Z,ZZ,Z^{\prime} and Z′′Z^{\prime\prime}. We connect Reeb orbits in ZZ, ZZ^{\prime} and Z′′Z^{\prime\prime} to {γi}\{\gamma_{i}\} counted with multiplicities using cyliners along each Morse-Bott tori to obtain Z^,Z^,Z^′′\hat{Z},\hat{Z}^{\prime},\hat{Z}^{\prime\prime}. We then define

Qτ(Z,Z):=Q^τ(Z^,Z′′^).Q_{\tau}(Z,Z^{\prime}):=\hat{Q}_{\tau}(\hat{Z},\hat{Z^{\prime\prime}}).

Observe in the above Q^τ\hat{Q}_{\tau} is an intersection form defined on H2(α,β,Y)H_{2}(\alpha^{\prime},\beta^{\prime},Y) where α\alpha^{\prime} and β\beta^{\prime} are collections of Reeb orbits of the form {(γi,ni)}\{(\gamma_{i},n_{i})\}. It suffices to prove Qτ(Z′′,Z)=Qτ(Z,Z′′)Q_{\tau}(Z^{\prime\prime},Z^{\prime})=Q_{\tau}(Z,Z^{\prime\prime}). To do this note the fact ZZ′′Z\sim Z^{\prime\prime} in 2(α,β,Y)\mathcal{H}_{2}(\alpha,\beta,Y) extends to an equivalence of Z^Z′′^\hat{Z}\sim\hat{Z^{\prime\prime}} in H2(α,β,Y)H_{2}(\alpha^{\prime},\beta^{\prime},Y), hence Q^τ(Z^′′,Z^)=Q^τ(Z^,Z^)\hat{Q}_{\tau}(\hat{Z}^{\prime\prime},\hat{Z}^{\prime})=\hat{Q}_{\tau}(\hat{Z},\hat{Z}^{\prime}), and hence the proof. ∎

We observe using the above reasoning the relative Chern class also descends to 2(α,β,Y)\mathcal{H}_{2}(\alpha,\beta,Y). We state this in the form of a definition:

Definition 5.5.

Given Z2(α,β,Y)Z\in\mathcal{H}_{2}(\alpha,\beta,Y), we define the relative Chern class cτ(Z)c_{\tau}(Z) the same way as before: choose a representative SS of ZZ that is embedded near the boundary. Let ι:ZY\iota:Z\rightarrow Y be the inclusion, then consider the pullback of the contact structure ιξ\iota^{*}\xi to ZZ, pick a section ψ\psi of ιξ\iota^{*}\xi that does not rotate with respect to τ\tau near the end points and has transverse zeroes, then cτ(Z)c_{\tau}(Z) is the signed count of zeroes of ψ\psi.

Finally we define writhe the same way as before:

Definition 5.6.

Let CC be a somewhere injective curve that is not a trivial cylinder. We assume at s+s\rightarrow+\infty (resp. -\infty) it is asymptotic to orbit set α\alpha (resp. β\beta). The trivialization specified in Theorem 17 gives an identification a neighborhood of each Reeb γα,β\gamma\in\alpha,\beta with S1×2S^{1}\times\mathbb{R}^{2}, then using this we can define writhe of CC as we had before in section 2.

Remark 5.7.

The definition of writhe depends crucially on the fact CC is a holomorphic curve, and does not admit constructions as before where we can slide the Reeb orbits of α,β\alpha,\beta around and obtains a surface with same relative intersection number/Chern class.

Hence we are ready to state the relative adjunction formula.

Theorem 5.8.

If CC is a simple JJ-holomorphic curve, then

cτ(C)=χ(C)+Qτ(C)+wτ(C)2δ(C)c_{\tau}(C)=\chi(C)+Q_{\tau}(C)+w_{\tau}(C)-2\delta(C)

with the definition of relative chern class, relative intersection number, and writhe given above.

Proof.

This is a purely topological formula. The same proof as in [Hut02] follows through. ∎

Hence we would like to define a version of ECH index by applying the relative adjunction formula to the Fredholm index formula of holomorphic curves as in [Hut02]. Recall then the proof of index inequality boils down to bounding the writhe of the JJ holomorphic curve in terms of various algebraic expressions involving the Conley Zehnder indices that the curve is asymptotic to. We turn to this writhe bound in the next subsection.

5.2 Writhe Bound

We recall the Fredholm index of a somewhere injective curve uu depends on which end is free and which end is fixed. Hence we anticipate that the ECH index we assign to a holomorphic curve uu will depend on which end is fixed and which end is free. The writhe inequality we prove shall take into account of the assignment of free and fixed ends. We note that this assignment of an index to a curve that depends on which end is free/fixed is somewhat artificial, but it will be less artificial once we use this index to define the ECH index of an entire cascade.

First we fix some conventions on Conley Zehnder indices. For a given Morse-Bott Torus 𝒯\mathcal{T} assume the JJ holomorphic curve has NN ends that are positively (resp. negatively) asymptotic to Reeb orbits on this torus. They are asymptotic to the individual Reeb orbits labelled R1,..,RnR_{1},..,R_{n}. Writhe bound is a local computation so we only consider a particular Reeb orbit, called R1R_{1}. Assume kk ends of CC are asymptotic to R1R_{1}. They have multiplicity q1,..,qkq_{1},..,q_{k}. We adopt the following convention on Conley Zehnder indices.

Convention 5.9.

Recall for positive Morse-Bott torus μ=1/2\mu=1/2. We declare μ+=1\mu_{+}=1, μ=0\mu_{-}=0. For negative Morse-Bott torus we declare μ+=0\mu_{+}=0, μ=1\mu_{-}=-1.

This has the following significance: for a curve with free end as s+s\rightarrow+\infty landing in a Morse-Bott torus (regardless of whether it is positive or negative torus), the Conley Zehnder index term in the Fredholm index formula associated to this end is μ+\mu_{+} (the specific value depends on the positive/negative Morse-Bott torus as above), and the Conley Zehnder index term assigned to fixed end is μ\mu_{-}. Conversely, at the ss\rightarrow-\infty end we assign μ\mu_{-} to free ends and μ+\mu_{+} to fixed ends.

Using the above conventions given a somewhere injective holomorhic curve uu, we assign its total Conley-Zehnder index denoted by CZInd(u)CZ^{Ind}(u) according to the convention above. The goal of the writhe inequality is to come up with another Conley-Zehnder index term CZECH(u)CZ^{ECH}(u) so that the total writhe of uu is bounded above by

wrτ(u)CZECH(u)CZInd(u)wr_{\tau}(u)\leq CZ^{ECH}(u)-CZ^{Ind}(u) (16)

By way of convention we will use CZ(R1,±)CZ^{*}(R_{1},\pm\infty) where =Ind,ECH*=Ind,ECH to denote the Conley Zehnder index we should assign to the free/fixed ends approaching RR as s±s\rightarrow\pm\infty

5.2.1 Positive Morse-Bott tori

Theorem 5.10.

In the case of positive Morse-Bott torus, ss\rightarrow-\infty, if ξi\xi_{i} is an end of uu with covering multiplicity qiq_{i} and uu is not the trivial cylinder, we have the following inequality

η(ξi)1(single end winding number).\eta(\xi_{i})\geq 1\quad\text{(single end winding number)}.

For single end writhe, we have:

w(ξi)η(ξi)(qi1).w(\xi_{i})\geq\eta(\xi_{i})(q_{i}-1).\quad

Note this holds true for trivial cylinders (as long as it’s somewhere injective).

Let ξ1\xi_{1} and ξ2\xi_{2} be two braids that correspond to two distinct ends of uu that approach the same Reeb orbit, with multiplicities qiq_{i} and winding numbers ηi\eta_{i}, then:

l(ξ1,ξ2)min(q1η2,q2η1)l(\xi_{1},\xi_{2})\geq min(q_{1}\eta_{2},q_{2}\eta_{1})

Note this holds if one of the ends ξi\xi_{i} came from a trivial cylinder.

And finally to calculate the writhe of all ends approach the same Reeb orbit, w(ξ)w(\xi), let ξ\xi denote the total braid and ξi\xi_{i} the various components coming from incoming ends of uu (this holds for both s=±s=\pm\infty):

w(ξ)=iw(ξi)+ijl(ξi,ξj)w(\xi)=\sum_{i}w(\xi_{i})+\sum_{i\neq j}l(\xi_{i},\xi_{j})

In the case of s+s\rightarrow+\infty, using the exactly the same notation, we have the following inequalities:

η(ξi)0\eta(\xi_{i})\leq 0
w(ξi)η(ξi)(qi1)for single end writhew(\xi_{i})\leq\eta(\xi_{i})(q_{i}-1)\,\,\,\,\text{for single end writhe}
l(ξ1,ξ2)max(q1η2,q2η1)l(\xi_{1},\xi_{2})\leq max(q_{1}\eta_{2},q_{2}\eta_{1})
Proof.

(Sketch) The proof constitutes an amalgamation of existing results in the literature. The key result is an description of asymptotics of ends of holomoprhic curves on Morse-Bott torus [Sie]. Namely, near the s+s\rightarrow+\infty end of uu, the ss constant slice of {s}×Y\{s\}\times Y of uu can be described as follows. We can choose a neighborhood of trivial cylinder ×γ\mathbb{R}\times\gamma as ×S1×2\mathbb{R}\times S^{1}\times\mathbb{R}^{2} where ss is the symplectization direction, tt is the variable along the Reeb orbit and {0}×2\{0\}\times\mathbb{R}^{2} is the contact structure along the Reeb orbit, then we can write an end ξi\xi_{i} of uu as

u(s,t)=(qs,qt,i=1neλisei(t))u(s,t)=(qs,qt,\sum_{i=1}^{n}e^{\lambda_{i}s}e_{i}(t)) (17)

where λi\lambda_{i} and eie_{i} are respectively the (negative) eigenvalues and corresponding eigenfunctions of the operator A(t):L2(S1,2)L2(S1,2)A(t):L^{2}(S^{1},\mathbb{R}^{2})\rightarrow L^{2}(S^{1},\mathbb{R}^{2}) coming from the linearization of the Cauchy Riemann operator, which can be written as

A(t)=JtSA(t)=-J\partial_{t}-S

With this normal form, the winding number bound comes from combining the results in [Gut14] about the meaning of Robbin-Salamon index and results in [HWZ96] relating Conley-Zehnder indices to crossing of eigenvalues. The relations on writhe and linking number come from direct modifications from the proofs in [Hut02], once we realize that locally the braids can be described by Equation 17. ∎

Next we move to use these relations to prove writhe bound. As in the case of ECH, equality of the writhe bound implies certain partition conditions, which we will carefully state.

Proposition 5.11 (link, -\infty, positive Morse Bott torus).

Consider the JJ holomorphic curve uu with negative ends on a Reeb orbit γ\gamma. We have kfreek_{free} free ends of multiplicity qifreeq_{i}^{free}, and kfixedk_{fixed} fixed ends with multiplicity qifixedq_{i}^{fixed} and of total multiplicity NfixedN_{fixed}. The writhe bound reads

w(ξ)i=1kfree+kfixedηi+i,jkfree+kfixedmin(ηiqj,ηjqi)(Nfree1+Nfixed)(kfixed)w(\xi)\geq-\sum_{i=1}^{k_{free}+k_{fixed}}\eta_{i}+\sum_{i,j}^{k_{free}+k_{fixed}}min(\eta_{i}q_{j},\eta_{j}q_{i})\geq(N_{free}-1+N_{fixed})-(k_{fixed})

with equality holding implying there can be only free/fixed ends at this Reeb orbit. If there are only fixed ends the partition conditions is (n)(n), and if there are only free ends the partition condition is (n)(n) or (1,n1)(1,n-1).

Proof.

We have the respective bounds

kfree+ikfreemin(ηiqj,ηjqi)Nfree1-k_{free}+\sum_{i}^{k_{free}}min(\eta_{i}q_{j},\eta_{j}q_{i})\geq N_{free}-1

and

kfix+i,jkfixmin(ηiqj,ηjqi)Nfixkfixed-k_{fix}+\sum_{i,j}^{k_{fix}}min(\eta_{i}q_{j},\eta_{j}q_{i})\geq N_{fix}-k_{fixed}

and cross terms will imply strict inequality, hence only free or fixed term appears. In the case of only fixed points, we see the only way equality can hold is with partition condition (n)(n). Similar considerations produces the partition conditions for free ends. ∎

Proposition 5.12 (link, \infty, positive Morse Bott Torus).

In the s+s\rightarrow+\infty end, consider the JJ holomorphic curve uu with ends on a Reeb orbit γ\gamma. We have kfreek_{free} free ends of multiplicity qifreeq_{i}^{free}, and kfixedk_{fixed} fixed ends qifixedq_{i}^{fixed} of total multiplicity NfixedN_{fixed}:

w(ξ)i=1kfree+kfixηi+i,jkfree+kfixmax(qjηi,qi,ηj)Nfree(kfree).w(\xi)\leq-\sum_{i=1}^{k_{free}+k_{fix}}\eta_{i}+\sum_{i,j}^{k_{free}+k_{fix}}max(q_{j}\eta_{i},q_{i},\eta_{j})\leq N_{free}-(k_{free}).

The partition condition implies (1,,1)(1,...,1) on the free ends.

Proof.

We see that lhs0lhs\leq 0, and RHS=0RHS=0 iff the free end satisfies partition conditions (1,1)(1,...1); there are no requirements on fixed ends. ∎

5.2.2 Negative Morse-Bott tori

In this subsection we take up the analogous writhe bounds for negative Morse-Bott tori.

Theorem 5.13.

In the case of negative Morse Bott torus, ss\rightarrow-\infty, we have the following inequalities:

If ξi\xi_{i} is an end of uu and uu is not the trivial cylinder, we have the following inequality:

η(ξi)0\eta(\xi_{i})\geq 0

For writhe of a single end, with covering multiplicity qiq_{i}, we have:

w(ξi)η(ξi)(qi1)w(\xi_{i})\geq\eta(\xi_{i})(q_{i}-1)

Note this holds for the case of a trivial cylinder.

Let ξ1\xi_{1} and ξ2\xi_{2} be two braids that correspond to two distinct ends of uu that approach the same Reeb orbit, with multiplicities qiq_{i} and winding numbers ηi\eta_{i}, then:

l(ξ1,ξ2)min(q1η2,q2η1)l(\xi_{1},\xi_{2})\geq min(q_{1}\eta_{2},q_{2}\eta_{1})

Note this holds if one of the ends ξi\xi_{i} came from a trivial cylinder.

And finally to calculate the writhe of all ends approach the same Reeb orbit, w(ξ)w(\xi), let ξ\xi denote the total braid, and ξi\xi_{i} the various components coming from incoming ends of uu (this holds for both s=±s=\pm\infty):

w(ξ)=w(ξi)+ijl(ξi,ξj)w(\xi)=w(\xi_{i})+\sum_{i\neq j}l(\xi_{i},\xi_{j})

In the case of s+s\rightarrow+\infty, we have the following inequalities

η(ξi)1\eta(\xi_{i})\leq-1
w(ξi)η(ξi)(qi1)for single end writhew(\xi_{i})\leq\eta(\xi_{i})(q_{i}-1)\,\,\,\,\text{for single end writhe}
l(ξ1,ξ2)max(q1η2,q2η1)l(\xi_{1},\xi_{2})\leq max(q_{1}\eta_{2},q_{2}\eta_{1})
Proof.

The exact same proof for the positive Morse-Bott torus except we use Robbin-Salamon index μ=1/2\mu=-1/2. ∎

Proposition 5.14 (link, -\infty,negative Morse Bott torus).

Let uu have ends asymptotic to γ\gamma on a negative Morse-Bott torus as ss\rightarrow-\infty, suppose there are kfreek_{free} free ends of multiplicity qifreeq_{i}^{free}, of total multiplicity NfreeN_{free}; suppose there are kfixk_{fix} fixed ends each of multiplicity qfixq_{fix}, of total multiplicity NfixN_{fix}. Then we have the writhe bound:

w(ξ)ikfix+kfreeηi+i,jkfix+kfreemin(ηiqj,ηjqi)Nfree(kfree)w(\xi)\geq-\sum_{i}^{k_{fix}+k_{free}}\eta_{i}+\sum_{i,j}^{k_{fix}+k_{free}}min(\eta_{i}q_{j},\eta_{j}q_{i})\geq-N_{free}-(-k_{free})

with equality enforcing partition condition (1,..,1)(1,..,1) on free ends and no partition condition on fixed ends.

Proof.

η0\eta\geq 0 so lhs0lhs\geq 0, rhs=kfreeNfreerhs=k_{free}-N_{free} so inequality holds, and equality if free ends has partition conditions (1,..,1)(1,..,1), no restrictions on fixed ends. ∎

Proposition 5.15 (link, ++\infty,negative Morse-Bott torus).

Let uu have ends asymptotic to γ\gamma on a negative Morse-Bott torus as s+s\rightarrow+\infty, suppose there are kfreek_{free} free ends of multiplicity qifreeq_{i}^{free}, of total multiplicity NfreeN_{free}; and suppose there are kfixk_{fix} fixed ends each of multiplicity qfixq_{fix}, of total multiplicity NfixN_{fix}.

w(ξ)ikfix+kfreeηi+i,jkfix+kfreemax(ηiqj,ηjqi)NfixNfree+1+kfixw(\xi)\leq-\sum_{i}^{k_{fix}+k_{free}}\eta_{i}+\sum_{i,j}^{k_{fix}+k_{free}}max(\eta_{i}q_{j},\eta_{j}q_{i})\leq-N_{fix}-N_{free}+1+k_{fix}

with equality enforcing only free or fixed ends. In the case of only fixed ends the partition condition is (n)(n), and in the case of only free ends the partition condition is either (n)(n) or (n1,1)(n-1,1).

Proof.

We can split the sum into:

ikfreeηi+i,jkfreemin(ηiqj,ηjqi)1Nfree-\sum_{i}^{k_{free}}\eta_{i}+\sum_{i,j}^{k_{free}}min(\eta_{i}q_{j},\eta_{j}q_{i})\leq 1-N_{free}

and

ikfixedηi+i,jkfixedmin(ηiqj,ηjqi)kfixNfix.-\sum_{i}^{k_{fixed}}\eta_{i}+\sum_{i,j}^{k_{fixed}}min(\eta_{i}q_{j},\eta_{j}q_{i})\leq k_{fix}-N_{fix}.

Each of the above inequalities hold individually, and when there are both free and fixed ends, there are cross terms that make the inequality strict. As before, we can deduce the partition conditions directly from imposing the equality condition. ∎

5.3 Morse-Bott tori as ECH generators

Recall that for ECH of nondegenerate contact forms, the generators of the chain complex are orbit sets satisfying the condition that if an orbit is hyperbolic then it can only have multiplicity 11. There are analogues of this in Morse Bott tori. In Morse-Bott ECH, we think of the generators of the chain complex as collections of Morse-Bott tori with additional data, written schematically as:

α={(𝒯j,±,mj)}\alpha=\{(\mathcal{T}_{j},\pm,m_{j})\}

and the differential as counting ECH index one height one JJ holomorphic cascades connecting between chain complex generators as above (which we will also call orbit sets). In the above definition mjm_{j} is the total multiplicity, which we think of as total multiplicity of Reeb orbits on 𝒯j\mathcal{T}_{j} hit by the JJ holomorphic curves that have ends on this Morse-Bott torus on the top (resp. bottom) level of a (height 1) cascade. ±\pm is additional information, which specifies how many ends of the JJ holomorphic curve landing on 𝒯j\mathcal{T}_{j} are free/fixed. We see that this also depends on whether α\alpha appears as the top or bottom level of a JJ holomorphic cascade, and in context of our correspondence theorem free/fixed ends correspond to elliptic/hyperbolic orbits in the non-degenerate case. We state this explicitly in the next definition in which we also describe the expected correspondence between Morse-Bott ECH generators and nondegenerate ECH generators after the perturbation.

Definition 5.16.

We consider the case of positive Morse Bott tori. In the nondegenerate case we let γ\gamma_{-} denote the hyperbolic Reeb orbit that arises from perturbation with Conley Zehnder index 0, and γ+\gamma_{+} the elliptic orbit that arose out of the perturbation with Conley Zehnder index 1. Then the description of our Morse-Bott generator, say (𝒯,±,m)(\mathcal{T},\pm,m) (this is just one Morse-Bott torus, in general α\alpha will consist of a collection of such tori, we focus on an example for the sake of brevity) and its correspondence with ECH generators in the perturbed non-degenerate case is given by:

  1. a.

    positive side ss\rightarrow\infty,

    1. (i)

      The Morse-Bott generator (𝒯,+,m)(\mathcal{T},+,m) is defined to require all ends on 𝒯\mathcal{T} are free, with total multiplicity on the torus being mm. In the perturbed nondegenerate case, this corresponds to ECH orbit set (γ+,m)(\gamma_{+},m). We observe the nondeg partition (θ\theta positive) condition is (1,..,1)(1,..,1), and the Morse-Bott partition condition from the writhe bound is (1,..1)(1,..1).

      By the Conley-Zehnder index convention the ECH conley Zehnder index assigned to (𝒯,+,m)(\mathcal{T},+,m) is given by: CZECH((𝒯,+,+,m))=mCZ^{ECH}((\mathcal{T},+\infty,+,m))=m

    2. (ii)

      The Morse-Bott generator (𝒯,,m)(\mathcal{T},-,m) there is one end on 𝒯\mathcal{T} that is fixed with multiplicity 1, on the critical point of ff that corresponds to the hyperbolic orbit. The rest of the ends are free, and the total multiplicity of orbits on 𝒯\mathcal{T} is mm. This corresponds to the orbit set {(γ,1),(γ+,m1)}\{(\gamma_{-},1),(\gamma_{+},m-1)\} in the nondegenerate case. Note the partition conditions between nondegenerate case and Morse-Bott case agree.

      We also have CZECH((𝒯,+,,m))=m1CZ^{ECH}((\mathcal{T},+\infty,-,m))=m-1.

  2. b.

    In the case of negative ends, ss\rightarrow-\infty,

    1. (i)

      The Morse-Bott generator (𝒯,+,m)(\mathcal{T},+,m) is defined to require all ends are fixed and asymptotic to the critical point of ff corresponding to the elliptic orbits, and the total multiplicity is mm. In the nondegenerate case this correspond to the orbit set (γ+,m)(\gamma_{+},m). We observe Morse-Bott and nondegenerate partition conditions agree, both being (m)(m). By our conventions, CZECH(𝒯,+,m)=mCZ^{ECH}(\mathcal{T},+,m)=m

    2. (ii)

      The Morse-Bott generator (𝒯,,m)(\mathcal{T},-,m) requires there is a multiplicity 1 free end landing on 𝒯\mathcal{T}, the remaining ends are fixed and are also required to land on the critical point corresponding to elliptic Reeb orbit. This corresponds to the orbit set {(γ+,m1),(γ,1)}\{(\gamma_{+},m-1),(\gamma_{-},1)\} in the nondegenerate case, and we have analogous partition conditions for both Morse-Bott and nondegenerate case. CZECH(𝒯,,m)=m1CZ^{ECH}(\mathcal{T},-,m)=m-1

We observe (𝒯,±,m)(\mathcal{T},\pm,m) imposes different free/fixed end conditions, depending whether it appears as s±s\rightarrow\pm\infty ends, however we should think of it as being the same generator in the chain complex, as is evidenced by the fact that it is identified to the same nondegenerate orbit set regardless of whether it appears at ++\infty or -\infty end.

We also briefly summarize the analogous result for negative Morse-Bott torus.

Definition 5.17.

In the case of negative Morse Bott tori, we use γ\gamma_{-} to denote the elliptic Reeb orbit after perturbation of Conley Zehnder index -1, and let γ+\gamma_{+} denote the hyperbolic orbit after perturbation of Conley Zehnder index 0. Let (𝒯,±,m)(\mathcal{T},\pm,m) denote a Morse-Bott generator.

  1. a.

    At the positive end as ss\rightarrow\infty,

    1. (i)

      (𝒯,,m)(\mathcal{T},-,m) requires all ends fixed at the critical point of ff corresponding to γ\gamma_{-}, corresponds to (γ,m)(\gamma_{-},m) in nondegenerate case, both degenerate and nondegenerate case has partition conditions (m)(m). CZECH((𝒯,,m))=mCZ^{ECH}((\mathcal{T},-,m))=-m

    2. (ii)

      (𝒯,+,m)(\mathcal{T},+,m) requires one end free with multiplicity 1, the rest have multiplicity m1m-1 fixed at the critical point of ff corresponding to γ\gamma_{-}. The generator corresponds to {(γ+,1),(γ,m1)}\{(\gamma_{+},1),(\gamma_{-},m-1)\}. CZECH((𝒯,+,m))=m+1CZ^{ECH}((\mathcal{T},+,m))=-m+1. Partition conditions match.

  2. b.

    Negative end, as ss\rightarrow-\infty,

    1. (i)

      (𝒯,,m)(\mathcal{T},-,m) has all ends free, of total multiplicity mm. This corresponds to (γ,m)(\gamma_{-},m) in the nondegenerate case. Partition conditions match. CZECH((𝒯,,m))=mCZ^{ECH}((\mathcal{T},-,m))=-m.

    2. (ii)

      (𝒯,+,m)(\mathcal{T},+,m) has one fixed end corresponding to the critical point of ff at γ+\gamma_{+} of multiplicity one; the rest are free and of multiplicity m1m-1. This corresponds to the orbit set {(γ+,1),(γ,m1)\{(\gamma_{+},1),(\gamma_{-},m-1). The partition conditions correspond, and CZECH((𝒯,+,m))=m+1CZ^{ECH}((\mathcal{T},+,m))=-m+1.

We would also like a more general notion of ECH Conley Zehnder index for when there are more free/fixed ends than allowed by ECH generator conditions are above. To keep track of the more refined intersection theory information, we need to make our definition depend slightly on the behaviour of the JJ-holomorphic curve as its ends approach Reeb orbits on Morse-Bott tori. We consider a nontrivial somewhere injective holomorphic curve u:Σ×Y3u:\Sigma\rightarrow\mathbb{R}\times Y^{3}. We isolate this into the following definition.

Definition 5.18.

Let u:Σ×Y3u:\Sigma\rightarrow\mathbb{R}\times Y^{3} be a nontrivial somewhere injective holomorphic curve. Let γ\gamma be a simple Reeb orbit on a positive Morse-Bott torus.

  1. a.

    At the ss\rightarrow\infty end, suppose kfreek_{free} ends approach γ\gamma with total multiplicity NfreeN_{free}, and kfixedk_{fixed} ends approach γ\gamma with total multiplcity NfixedN_{fixed}, then CZECH(γ):=NfreeCZ^{ECH}(\gamma):=N_{free}.

  2. b.

    At the ss\rightarrow-\infty end, suppose kfreek_{free} ends approach γ\gamma with total multiplicity NfreeN_{free}, and kfixedk_{fixed} ends approach γ\gamma with total multiplcity NfixedN_{fixed}, then CZECH(γ):=Nfree+Nfixed1CZ^{ECH}(\gamma):=N_{free}+N_{fixed}-1.

Similarly if γ\gamma is a simply covered Reeb orbit on a negative Morse-Bott torus.

  1. a.

    At the ss\rightarrow\infty end, suppose kfreek_{free} ends approach γ\gamma with total multiplicity NfreeN_{free}, and kfixedk_{fixed} ends approach γ\gamma with total multiplcity NfixedN_{fixed}, then CZECH(γ):=NfixNfree+1CZ^{ECH}(\gamma):=-N_{fix}-N_{free}+1.

  2. b.

    At the ss\rightarrow-\infty end, suppose kfreek_{free} ends approach γ\gamma with total multiplicity NfreeN_{free}, and kfixedk_{fixed} ends approach γ\gamma with total multiplcity NfixedN_{fixed}, then CZECH(γ):=NfreeCZ^{ECH}(\gamma):=-N_{free}.

Note the above definition agrees with that of the ECH Morse-Bott generator. Then let uu be a somewhere injective JJ holomorphic curve with no trivial cylinder components, and we have chosen which ends of uu are fixed/free. Then we define its ECH index using the above notion of ECH Conley-Zehnder index:

Definition 5.19.

We define the ECH index of uu as:

I(u):=cτ(u)+Qτ(u)+CZECH(u)I(u):=c_{\tau}(u)+Q_{\tau}(u)+CZ^{ECH}(u) (18)

Note the above definition not only depends on the relative homology class of uu, it also depends on how the ends of uu are distributed among the Reeb orbits (for information of free/fixed beyond that of the Morse-Bott ECH generators)- in particular we have to keep the information of not only how many free/fix ends land on a Morse-Bott torus, we also need to retain the information which ends are asymptotic to which Reeb orbit.

By using the writhe bound we recover directly

Proposition 5.20.

Let uu be a JJ-holomorphic map satisfying the conditions above,

Ind(u)I(u)2δ(u).Ind(u)\leq I(u)-2\delta(u). (19)

with equality enforcing partition conditions described in the writhe bound section.

We next include the case of trivial cylinders in our definition of ECH Conley-Zehnder index.

Definition 5.21.

Let γ\gamma be a simply covered Reeb orbit on a positive Morse-Bott torus. Let u:Σ×Yu:\Sigma\rightarrow\mathbb{R}\times Y be a JJ-holomorphic curve with potentially disconnected domain. When we say trivial cylinders below, we allow trivial cylinders with higher multiplicities.

  1. a.

    At the ss\rightarrow\infty end, suppose kfreek_{free} ends approach γ\gamma with total multiplicity NfreeN_{free}, and kfixedk_{fixed} ends approach γ\gamma with total multiplcity NfixedN_{fixed}, then CZECH(γ):=NfreeCZ^{ECH}(\gamma):=N_{free}. Here we allow holomorphic curves to be trivial cylinders.

  2. b.

    At the ss\rightarrow-\infty end, suppose kfreek_{free} ends approach γ\gamma with total multiplicity NfreeN_{free}, and kfixedk_{fixed} ends approach γ\gamma with total multiplcity NfixedN_{fixed}. If at least one of the approaching ends is not that of a trivial cylinder, then CZECH(γ):=Nfree+Nfixed1CZ^{ECH}(\gamma):=N_{free}+N_{fixed}-1. If all the approaching ends are trivial cylinders, then CZECH:=NfixedCZ^{ECH}:=N_{fixed}.

Next let γ\gamma be a simply covered Reeb orbit on a negative Morse-Bott torus.

  1. a.

    At the ss\rightarrow\infty end, suppose kfreek_{free} ends approach γ\gamma with total multiplicity NfreeN_{free}, and kfixedk_{fixed} ends approach γ\gamma with total multiplcity NfixedN_{fixed}, If at least one of the approaching ends is not that of a trivial cylinder, then CZECH(γ):=1NfreeNfixCZ^{ECH}(\gamma):=1-N_{free}-N_{fix}. If there are only trivial cylinders, then CZECH=NfixedCZ^{ECH}=-N_{fixed}.

  2. b.

    At the ss\rightarrow-\infty end, suppose kfreek_{free} ends approach γ\gamma with total multiplicity NfreeN_{free}, and kfixedk_{fixed} ends approach γ\gamma with total multiplcity NfixedN_{fixed}. Then we set CZECH(γ):=NfreeCZ^{ECH}(\gamma):=-N_{free}. This includes the case of trivial cylinders.

Proposition 5.22.

Let CC be a JJ holomorphic current which can contain trivial cylinders. Each end in CC is implicitly assigned “free” or “fixed”, and recall the convention that we can at most designate one end of a trivial cylinder as fixed. With CZECHCZ^{ECH} as defined above, we have the inequality:

Ind(C)I(C)2δ(C)Ind(C)\leq I(C)-2\delta(C)
Proof.

Let CC be a JJ-holomorphic current of the form {(Ci,mi}\{(C_{i},m_{i}\} where CiC_{i} are pairwise distinct. If CiC_{i} is nontrivial, and mi>1m_{i}>1, then as in [Hut02], we can consider mim_{i} copies of CiC_{i} translated by mim_{i} distinct factors in the symplectization direction. Then we can represent (Ci,mi)(C_{i},m_{i}) as mim_{i} distinct somewhere injective JJ-holomorphic curves. We do this for all nontrivial components of CC. Each resulting end of CiC_{i} receives an assignment of “free/fixed”, hence both sides of the inequality above are defined. (One can make all the copies of CiC_{i} coming from (Ci,mi)(C_{i},m_{i}) have the same free/fixed assignments at their corresponding ends, but this won’t be necessary.)

As before this boils down to writhe bounds at s=+s=+\infty and s=s=-\infty. We first consider γ\gamma a Reeb orbit on a positive Morse-Bott torus. We first consider the s=+s=+\infty case. Here for trivial cylinders qi=1q_{i}=1 and the linking number is zero, so the same proof as before produces the writhe bound.

In the case ss\rightarrow-\infty, let NtrivialN_{trivial} denote the multiplicity of trivial ends. Let NtrivialN_{trivial} denote the total multiplicity of trivial ends, fixed or free. First assume there is at least one nontrivial end. The apriori bound on writhe is:

w(ξ)#nontrivial ends+i,jnontrivial endsmin(qi,qj)+Ntrivial(#nontrivial ends).w(\xi)\geq-\#\textup{nontrivial ends}+\sum_{i,j\textup{nontrivial ends}}min(q_{i},q_{j})+N_{trivial}\cdot(\#\textup{nontrivial ends}).

With our new definition of CZECHCZ^{ECH}, we need to establish the writhe bound that

#nontrivial ends+i,jnontrivial endsmin(qi,qj)+Ntrivial(#nontrivial ends)Nfree+Nfixed1(kfixed)-\#\textup{nontrivial ends}+\sum_{i,j\textup{nontrivial ends}}min(q_{i},q_{j})+N_{trivial}\cdot(\#\textup{nontrivial ends})\geq N_{free}+N_{fixed}-1-(k_{fixed})

We use the superscript T and NT to distinguish whether the multiplicity is coming from trivial ends or nontrivial ends. But the writhe bound already established implies

#nontrivial ends+i,jnontrivial endsmin(qi,qj)NfreeNT+NfixedNT1kfixedNT-\#\textup{nontrivial ends}+\sum_{i,j\textup{nontrivial ends}}min(q_{i},q_{j})\geq N^{NT}_{free}+N^{NT}_{fixed}-1-k_{fixed}^{NT}

Then it suffices to establish that

Ntrivial(#nontrivial ends)NfreeT+NfixedTkfixedTN_{trivial}\cdot(\#\textup{nontrivial ends})\geq N_{free}^{T}+N_{fixed}^{T}-k_{fixed}^{T}

which always holds, hence the writhe bound continues to hold.

When there are only trivial cylinders, the writhe is automatically zero, likewise the writhe bound is trivially satisfied.

We next consider the case γ\gamma a Reeb orbit on a negative Morse-Bott torus. We first consider the ss\rightarrow-\infty case. Since the winding number η\eta in this case is bounded below by zero, the writhe bound continues to hold even in the presence of trivial cylinders.

In the case of s+s\rightarrow+\infty, the computation is very much similar to the -\infty end of a positive Morse-Bott torus. Assuming there is at least one nontrivial end

w(ξ)+#nontrivial ends+i,jnontrivial endsmax(ηiqj,ηjqi)Ntrivial#nontrivial endsNfixNfree+1+kfix.w(\xi)\leq+\#\textup{nontrivial ends}+\sum_{i,j\textup{nontrivial ends}}max(\eta_{i}q_{j},\eta_{j}q_{i})-N_{trivial}\cdot\#\textup{nontrivial ends}\leq-N_{fix}-N_{free}+1+k_{fix}.

With the previous writhe bound we have already proven

#nontrivial ends+i,jnontrivial endsmax(ηiqj,ηjqi)NfixNTNfreeNT+1+kfixNT\#\textup{nontrivial ends}+\sum_{i,j\textup{nontrivial ends}}max(\eta_{i}q_{j},\eta_{j}q_{i})\leq-N_{fix}^{NT}-N_{free}^{NT}+1+k_{fix}^{NT}

hence suffices to prove

Ntrivial#nontrivial endsNfixTNfreeT+kfixT-N_{trivial}\cdot\#\textup{nontrivial ends}\leq-N_{fix}^{T}-N_{free}^{T}+k_{fix}^{T}

but this follows directly from our assumptions.

In the case there are only trivial ends the total writhe is zero, and the writhe bound is achieved. ∎

We next establish the subadditivity property of the ECH index.

Proposition 5.23.

Let 𝒞1={(Ca,ma)}\mathcal{C}_{1}=\{(C_{a},m_{a})\} and 𝒞2={(Cb,mb)}\mathcal{C}_{2}=\{(C_{b},m_{b})\} denote two JJ-holomorphic currents, and CaC_{a} is never the same as CbC_{b} unless they are both trivial cylinders (they can be \mathbb{R} translates of each other). Then their ECH indices satisfy

I(𝒞1𝒞2)I(𝒞1)+I(𝒞2)+2𝒞1𝒞2.I(\mathcal{C}_{1}\cup\mathcal{C}_{2})\geq I(\mathcal{C}_{1})+I(\mathcal{C}_{2})+2\mathcal{C}_{1}\cap\mathcal{C}_{2}. (20)

In the above 𝒞1𝒞2\mathcal{C}_{1}\cap\mathcal{C}_{2} counts the intersection with multiplicity of CaC_{a} with CbC_{b}. Note by intersection positivity each multiplicity is positive. Further by construction the intersection between trivial cylinders is zero.

Proof.

We again apply the translation in the symplectization trick to represent nontrival currents (Ca,ma)(C_{a},m_{a}) (resp. (Cb,mb)(C_{b},m_{b})) by mam_{a} (rep. mbm_{b}) distinct somewhere injective curves. After relabelling we can also denote them by CaC_{a} (resp. CbC_{b}). We apply the adjunction inequality as in [Hut02, Hut09] to obtain

I(𝒞1𝒞2)I(𝒞1)I(𝒞2)2#𝒞1𝒞2=CZECH(𝒞1𝒞2)CZECH(𝒞1)CZECH(𝒞2)2a,blτ(Ca,Cb)I(\mathcal{C}_{1}\cup\mathcal{C}_{2})-I(\mathcal{C}_{1})-I(\mathcal{C}_{2})-2\#\mathcal{C}_{1}\cdot\mathcal{C}_{2}=CZ^{ECH}(\mathcal{C}_{1}\cup\mathcal{C}_{2})-CZ^{ECH}(\mathcal{C}_{1})-CZ^{ECH}(\mathcal{C}_{2})-2\sum_{a,b}l_{\tau}(C_{a},C_{b}) (21)

Then this reduces to a local computation relating linking number and our choice of Conley-Zehnder indices. We take this up case by case. First consider γ\gamma a Reeb orbit on a positive Morse-Bott torus, consider the ss\rightarrow\infty end. In this case we have CZECH(𝒞1𝒞2)CZECH(𝒞1)CZECH(𝒞2)=0CZ^{ECH}(\mathcal{C}_{1}\cup\mathcal{C}_{2})-CZ^{ECH}(\mathcal{C}_{1})-CZ^{ECH}(\mathcal{C}_{2})=0 and lτ(Ca,Cb)0l_{\tau}(C_{a},C_{b})\leq 0. Hence all the contributions from this end is 0\geq 0.

We next consider γ\gamma on a positive Morse-Bott torus at ss\rightarrow-\infty ends. Because how we assigned Conley-Zehnder indices depends on whether all the ends are trivial, we split into cases. In the case where all ends of 𝒞1\mathcal{C}_{1} and 𝒞2\mathcal{C}_{2} asymptotic to γ\gamma as ss\rightarrow-\infty are trivial, we have again CZECH(𝒞1𝒞2)CZECH(𝒞1)CZECH(𝒞2)=0CZ^{ECH}(\mathcal{C}_{1}\cup\mathcal{C}_{2})-CZ^{ECH}(\mathcal{C}_{1})-CZ^{ECH}(\mathcal{C}_{2})=0 and the linking number vanishes. If one of them has non-trivial ends approaching γ\gamma (WLOG take this to be 𝒞1\mathcal{C}_{1} and take 𝒞2\mathcal{C}_{2} consists purely of trivial ends), then we have the Conley Zehnder contribution being

Nfree1+Nfixed11+Nfixed2(Nfree1+Nfree2+Nfixed1+Nfixed21)=Nfree2N_{free}^{1}+N_{fixed}^{1}-1+N_{fixed}^{2}-(N_{free}^{1}+N_{free}^{2}+N_{fixed}^{1}+N_{fixed}^{2}-1)=-N_{free}^{2}

where we write Nfree1N_{free}^{1} to denote the free ends coming from 𝒞1\mathcal{C}_{1} etc. The linking number contribution is bounded below by 2(Nfixed2+Nfree2)2(N_{fixed}^{2}+N_{free}^{2}), hence the overall contribution is non-negative. The case where both 𝒞1\mathcal{C}_{1} and 𝒞2\mathcal{C}_{2} contains nontrivial ends at γ\gamma as ss\rightarrow-\infty, then the Conley-Zehnder difference term is just 1-1, and the linking number term 2lτ(Ca,Cb)22l_{\tau}(C_{a},C_{b})\geq 2, hence once again the overall contribution is non-negative.

We next consider the case γ\gamma is a Reeb orbit in a negative Morse-Bott torus. This will be largely analogous to the positive Morse-Bott torus case. For ss\rightarrow-\infty, we have the Coneley-Zehnder indices contribute zero, and lτ(Ca,Cb)0l_{\tau}(C_{a},C_{b})\geq 0 as ss\rightarrow\infty, hence the overall contribution is non-negative. We next consider γ\gamma as s+s\rightarrow+\infty. Again we break into cases because of trivial cylinders. In the case where all ends approaching γ\gamma from 𝒞1\mathcal{C}_{1} and 𝒞2\mathcal{C}_{2} are trivial cylinders, the Conley-Zehnder index contribution as well as the linking number is zero. Then in the case 𝒞1\mathcal{C}_{1} has nontrivial ends but 𝒞2\mathcal{C}_{2} has all ends trivial, then the Conley-Zehnder index contribution is given by Nfree2-N_{free}^{2}, and the linking number 2lτ(Ca,Cb)2(Nfree2+Nfixed2)\sum 2l_{\tau}(C_{a},C_{b})\leq-2(N_{free}^{2}+N_{fixed}^{2}), hence the overall contribution is nonnegative. Similarly in the case where both 𝒞1\mathcal{C}_{1} and 𝒞2\mathcal{C}_{2} have nontrivial ends, the difference of Conley-Zehnder index contribution is 1-1, whereas the linking number 2lτ(Ca,Cb)22l_{\tau}(C_{a},C_{b})\leq-2, hence the overall contribution is positive. Hence combining all of the above local inequalities we obtain the overall ECH index inequality. ∎

5.4 Multiple level cascades and ECH index

In this subsection we describe ECH index one cascades. We recall ECH index one cascades should come from degenerations of ECH index one curves, and in particular should respect partition conditions on the end points. In particular we should always keep in mind that ECH index one cascades should flow from a generator Morse-Bott ECH α1\alpha_{1} to another αn\alpha_{n}, which includes the information of multiplicities of free/fixed ends that land on Morse-Bott tori.

Given any cascade uu^{\text{\Lightning}} as given in our previous definition, we first turn it into a “cascade of currents”: 𝐮={u1,..,un1}\mathbf{u}^{\text{\Lightning}}=\{u^{1},..,u^{n-1}\}. Then we can proceed to define the ECH index of 𝐮\mathbf{u}^{\text{\Lightning}}. The following is half definition half theorem, as in if this cascade is transverse and rigid and we glued it into a JJ holomorphic curve the ECH index of its homology class is given by the following calculation. Conversely, if uu came from a cascade of curves that came from a degeneration of I=1I=1 holomorphic curve in the λδ\lambda_{\delta} setting, then our definition of II for the cascade of current will also be one.

Definition 5.24.

Let 𝐮={u1,,un1}\mathbf{u}^{\text{\Lightning}}=\{u^{1},...,u^{n-1}\} be a height 1 cascade of currents. Let its positive asymptotics be denoted by α1\alpha_{1} and negative asymptotics be denoted by αn\alpha_{n}, both Morse-Bott ECH generators. We can then define the ECH index for the cascade of currents as:

I(𝐮)=c1(𝐮)+Qτ(𝐮)+CZECH(𝐮).I(\mathbf{u}^{\text{\Lightning}})=c_{1}(\mathbf{u}^{\text{\Lightning}})+Q_{\tau}(\mathbf{u}^{\text{\Lightning}})+CZ^{ECH}(\mathbf{u}^{\text{\Lightning}}). (22)

The CZECHCZ^{ECH} index term for cascade is just the ECH index terms of α1\alpha_{1} and αn\alpha_{n}, which corresponds to the nondegenerate ECH Conley Zehnder index once we have identified free/fixed ends with elliptic/hyperbolic orbits. The cascade Chern class and relative intersection terms are just the sum of the Chern class of each of the levels, i.e.

c1(𝐮):=c1(u1)++c1(un1)c_{1}(\mathbf{u}^{\text{\Lightning}}):=c_{1}(u^{1})+...+c_{1}(u^{n-1})

and

Qτ(𝐮):=Qτ(u1)++Qτ(un1)Q_{\tau}(\mathbf{u}^{\text{\Lightning}}):=Q_{\tau}(u^{1})+...+Q_{\tau}(u^{n-1})

We would like to compare the ECH index of cascade to the Fredholm index of the reduced version, because then with enough transversality we would be able to rule out certain configurations of cascade of ECH index one by index reasons. To this end, we decompose the ECH index of a cascade into ECH index of its constituents, as follows:

Proposition 5.25.

We assume all ends of u2,..,un2u^{2},..,u^{n-2} are free, and all ends of u1u^{1} and un1u^{n-1} are considered free except those mandated by α1\alpha_{1} and αn\alpha_{n}, and we recall our conventions on trivial cylinders with only one fixed end. Then let Rpos,i+1R_{pos,i+1}^{\prime} denote the number of distinct Reeb orbits on positive Morse-Bott tori approached by nontrivial ends of uiu^{i} as ss\rightarrow-\infty, and let Vpos,i+1V_{pos,i+1}^{\prime} denote the total multiplicity of Reeb orbits on positive Morse-Bott tori approached by uiu^{i} at the ss\rightarrow-\infty, so that at these Reeb orbits there are only trivial ends as ss\rightarrow-\infty. Similarly we let Rneg,iR_{neg,i}^{\prime} denote the number of distinct Reeb orbits on negative Morse-Bott tori approached by nontrivial ends of uiu^{i} as s+s\rightarrow+\infty, and let Vneg,iV_{neg,i}^{\prime} denote the total multiplicity of Reeb orbits on negative Morse-Bott tori approached by uiu^{i} at the s+s\rightarrow+\infty, so that at these Reeb orbits there are only trivial ends as s+s\rightarrow+\infty. Then we have

I(𝐮)=\displaystyle I(\mathbf{u}^{\text{\Lightning}})= I(u1)+I(un1)\displaystyle I(u^{1})...+I(u^{n-1})
Rpos,2Rpos,n1Vpos,2Vpos,n1Rneg,2..Rneg,n1Vneg,2..Vneg,n1\displaystyle-R_{pos,2}^{\prime}-...-R_{pos,n-1}^{\prime}-V_{pos,2}^{\prime}-...-V_{pos,n-1}^{\prime}-R_{neg,2}^{\prime}-..-R_{neg,n-1}^{\prime}-V_{neg,2}^{\prime}-..-V_{neg,n-1}^{\prime}
Proof.

Follows directly from definition of ECH Conley Zehnder index. ∎

Remark 5.26.

Note the assignment of free/fixed end points for calculation of ECH index purposes is different from when we defined free/fixed punctures in the calculation of the Fredholm index.

Remark 5.27.

We remark the above formula makes sense in the case our cascade consists purely of a chain of cylinder at a critical point. If it started at the minimum of ff, the trick is to notice by our convention all trivial cylinders below it are considered free.

In order to compare I(𝐮)I(\mathbf{u}^{\text{\Lightning}}) and Ind(𝐮~)Ind(\tilde{\mathbf{u}}^{\text{\Lightning}}), we first define

I(𝐮~)\displaystyle I(\tilde{\mathbf{u}}^{\text{\Lightning}}) :=I(u1~)+I(un1~)\displaystyle:=I(\tilde{u^{1}})...+I(\tilde{u^{n-1}})
Rpos,2Rpos,n1Vpos,2Vpos,n1Rneg,2..Rneg,n1Vneg,2..Vneg,n1\displaystyle-R_{pos,2}^{\prime}-...-R_{pos,n-1}^{\prime}-V_{pos,2}^{\prime}-...-V_{pos,n-1}^{\prime}-R_{neg,2}^{\prime}-..-R_{neg,n-1}^{\prime}-V_{neg,2}^{\prime}-..-V_{neg,n-1}^{\prime}

by removing all multiple covers of nontrivial curves. Note we have

I(𝐮~)I(𝐮)I(\tilde{\mathbf{u}}^{\text{\Lightning}})\leq I(\mathbf{u}^{\text{\Lightning}}) (23)

with equality holding only if 𝐮\mathbf{u}^{\text{\Lightning}} is already reduced. Next we compare Ind(𝐮~)Ind(\tilde{\mathbf{u}}^{\text{\Lightning}}) and I(𝐮~)I(\tilde{\mathbf{u}}^{\text{\Lightning}}).

Proposition 5.28.

Ind(𝐮~)I(𝐮~)2δ(𝐮~)1Ind(\tilde{\mathbf{u}}^{\text{\Lightning}})\leq I(\tilde{\mathbf{u}}^{\text{\Lightning}})-2\delta(\tilde{\mathbf{u}}^{\text{\Lightning}})-1

Proof.

We make a term-wise comparison, e.g. we compare

Ind(u~i)ki+1ki+1+Ri+1Ind(\tilde{u}^{i})-k_{i+1}^{\prime}-k_{i+1}+R_{i+1} (24)

and

I(u~i)2δ(u~i)Rpos,i+1Vpos,i+1Rneg,i+1Vneg,i+1.I(\tilde{u}^{i})-2\delta(\tilde{u}^{i})-R_{pos,i+1}^{\prime}-V_{pos,i+1}^{\prime}-R_{neg,i+1}^{\prime}-V_{neg,i+1}^{\prime}. (25)

Note there are two different conventions by which we assigned “free” and “fixed” ends to ends of curves appearing in the cascade, we will refer to them respectively as the Fredholm convention and the ECH convention.

We further refine our notation to kpos,i+1,kneg,i+1,kpos,i+1,kneg,i+1k_{pos,i+1},k_{neg,i+1},k_{pos,i+1}^{\prime},k_{neg,i+1}^{\prime} to denote the number of ends among the kik_{i} and ki+1k_{i+1} ends that land on positive/negative Morse-Bott tori, i.e. we have ki=kpos,i+kneg,ik_{i}=k_{pos,i}+k_{neg,i}.

We first restrict to 1<i<n11<i<n-1 To compare these two terms, we first decompose ui~=CiTfree,iTfixed,i\tilde{u^{i}}=C_{i}\cup T_{free,i}\cup T_{fixed,i}, where CiC_{i} is a collection of nontrivial somewhere injective curves, Tfree,iT_{free,i} is a collection of free trivial cylinders according to Fredholm convention, and Tfixed,iT_{fixed,i} is a collection of fixed cylinder according to the Fredholm index convention. Assume CiC_{i} has lfree,il_{free,i} free ends, and lfixed,il_{fixed,i} ends according to Fredholm convention, then we have:

Ind(CiTfree,iTfixed,i)+lfixed,iI(CiTfree,iTfixed,i)2δ(CiTfree,iTfixed,i)|Tfixed,i|Ind(C_{i}\cup T_{free,i}\cup T_{fixed,i})+l_{fixed,i}\leq I(C_{i}\cup T_{free,i}\cup T_{fixed,i})-2\delta(C_{i}\cup T_{free,i}\cup T_{fixed,i})-|T_{fixed,i}|

We may at later points further refine the notation to lfixed,pos/neg,±,il_{fixed,pos/neg,\pm,i} to indicate fixed ends at positive/negative Morse-Bott tori, at positive/negative ends. Note Tfixed,iT_{fixed,i} is regarded as free cylinders when we measure its ECH index. |Tfixed,i||T_{fixed,i}| denotes the total number of fixed trivial cylinders that appear in this level.

We will also later refine our notation to distinguish Tfixed/free,pos/neg,iT_{fixed/free,pos/neg,i} for trivial cylinders on positive/negative Morse-Bott tori.

We next consider the case for i=1i=1. We can decompose as before u~1=C1Tfree,1Tfixed,1Tfixed,1\tilde{u}^{1}=C_{1}\cup T_{free,1}\cup T_{fixed,1}\cup T_{fixed,1}^{\prime}. We explain the notation. C1C_{1} is a collection of nontrivial somewhere injective holomorphic curves. The information of Morse-Bott generator α1\alpha_{1} tells us which of C1C_{1} should already be considered as fixed as ss\rightarrow\infty. There are additionally lfixedl_{fixed} ends of CC that we count as fixed when we compute its Fredholm index because they land on critical points of ff. Tfree,1T_{free,1} is a collection of free cylinders. Tfix,1T_{fix,1} is a collection of fixed trivial cylinders that come from requirements of α1\alpha_{1}. Each positive Morse Bott torus can only have one of these, and they must all be multiplicity 1. Tfixed,1T_{fixed,1}^{\prime} is a collection of trivial cylinders that don’t come from requirements of α1\alpha_{1} but also happen to land on a critical point of ff. The index inequality we have gives:

Ind(C1Tfixed,1Tfree,1Tfixed,1)+lfixed,1I(C1Tfixed,1Tfree,1Tfixed,1)2δ(u~1)|Tfixed,1|Ind(C_{1}\cup T_{fixed,1}\cup T_{free,1}\cup T_{fixed,1}^{\prime})+l_{fixed,1}\leq I(C_{1}\cup T_{fixed,1}\cup T_{free,1}\cup T_{fixed,1}^{\prime})-2\delta(\tilde{u}^{1})-|T_{fixed,1}^{\prime}|

where for the purpose of computing ECH index we have counted elements of Tfixed,1T_{fixed,1}^{\prime} as free cylinders.

Similarly for the i=n1i=n-1 level. As before we can decompose u~n1=Cn1Tfree,n1Tfixed,n1Tfixed,n1\tilde{u}^{n-1}=C_{n-1}\cup T_{free,n-1}\cup T_{fixed,n-1}\cup T_{fixed,n-1}^{\prime} with the same convention as before. Here we only need to prove:

Ind(Cn1Tfree,n1Tfixed,n1)+lfixed,n1I(Cn1Tfree,n1Tfixed,n1)2δ(u~n1)|Tfixed,n1|Ind(C_{n-1}\cup T_{free,n-1}\cup T_{fixed,n-1})+l_{fixed,n-1}\leq I(C_{n-1}\cup T_{free,n-1}\cup T_{fixed,n-1})-2\delta(\tilde{u}^{n-1})-|T^{\prime}_{fixed,n-1}|

which holds by the one-level ECH index inequality. When we take the difference between I(u~)I(\tilde{u^{\text{\Lightning}}}) and Ind(u~)Ind(\tilde{u^{\text{\Lightning}}}), we can break down their difference into the following form:

I(𝐮~)=\displaystyle I(\tilde{\mathbf{u}}^{\text{\Lightning}})= I(u1~)+I(u~n1)\displaystyle I(\tilde{u^{1}})...+I(\tilde{u}^{n-1})
Rpos,2Rpos,n1Vpos,2Vpos,n1Rneg,2..Rneg,n1Vneg,2..Vneg,n1\displaystyle-R_{pos,2}^{\prime}-...-R_{pos,n-1}^{\prime}-V_{pos,2}^{\prime}-...-V_{pos,n-1}^{\prime}-R_{neg,2}^{\prime}-..-R_{neg,n-1}^{\prime}-V_{neg,2}^{\prime}-..-V_{neg,n-1}^{\prime}

and the index term can be re written as

Ind=iind(u~i)i=2,,n1(kpos,i+kpos,iRpos,i)i=2,,n1(kneg,i+kneg,iRneg,i)1LInd=\sum_{i}ind(\tilde{u}^{i})-\sum_{i=2,...,n-1}(k_{pos,i}+k_{pos,i}^{\prime}-R_{pos,i})-\sum_{i=2,...,n-1}(k_{neg,i}+k_{neg,i}^{\prime}-R_{neg,i})-1-L

If we take their difference, and take advantage of the inequalities we proved in the previous paragraphs, we get:

IInd=\displaystyle I-Ind= I(u~i)ind(u~i)+i=2,,n1((kpos,i+kpos,iRpos,iRpos,iVpos,i)\displaystyle\sum I(\tilde{u}^{i})-ind(\tilde{u}^{i})+\sum_{i=2,...,n-1}((k_{pos,i}+k_{pos,i}^{\prime}-R_{pos,i}-R_{pos,i}^{\prime}-V_{pos,i}^{\prime})
+i=2,,n1(kneg,i+kneg,iRneg,iRneg,iVneg,i)+L+1\displaystyle+\sum_{i=2,...,n-1}(k_{neg,i}+k_{neg,i}^{\prime}-R_{neg,i}-R_{neg,i}^{\prime}-V_{neg,i}^{\prime})+L+1
\displaystyle\geq 2δ(𝐮~)+i=2,..,n2(lfixed,i+|Tfixed,i|)+lfixed,1+lfixed,n1+|Tfixed,1|+|Tfixed,n1|\displaystyle 2\delta(\tilde{\mathbf{u}}^{\text{\Lightning}})+\sum_{i=2,..,n-2}(l_{fixed,i}+|T_{fixed,i}|)+l_{fixed,1}+l_{fixed,n-1}+|T^{\prime}_{fixed,1}|+|T^{\prime}_{fixed,n-1}|
+i=2,,n1((kpos,i+kpos,iRpos,iRpos,iVpos,i)\displaystyle+\sum_{i=2,...,n-1}((k_{pos,i}+k_{pos,i}^{\prime}-R_{pos,i}-R_{pos,i}^{\prime}-V_{pos,i}^{\prime})
+i=2,,n1(kneg,i+kneg,iRneg,iRneg,iVneg,i)+L+1\displaystyle+\sum_{i=2,...,n-1}(k_{neg,i}+k_{neg,i}^{\prime}-R_{neg,i}-R_{neg,i}^{\prime}-V_{neg,i}^{\prime})+L+1

It suffices to prove the above expression is bounded below by one. It suffices to prove

i=2,,n1Rpos,i+Rpos,i+Vpos,i+i=2,,n1Rneg,i+Rneg,i+Vpos,i\displaystyle\sum_{i=2,...,n-1}R_{pos,i}+R_{pos,i}^{\prime}+V_{pos,i}^{\prime}+\sum_{i=2,...,n-1}R_{neg,i}+R_{neg,i}^{\prime}+V_{pos,i}^{\prime}
i=2,..,n2(lfixed,i+|Tfixed,i|)+lfixed,1+lfixed,n1+|Tfixed,1|+|Tfixed,n1|\displaystyle\leq\sum_{i=2,..,n-2}(l_{fixed,i}+|T_{fixed,i}|)+l_{fixed,1}+l_{fixed,n-1}+|T^{\prime}_{fixed,1}|+|T^{\prime}_{fixed,n-1}|
+i=2,,n1(kpos,i+kpos,i)+i=2,,n1(kneg,i+kneg,i)+L\displaystyle+\sum_{i=2,...,n-1}(k_{pos,i}+k_{pos,i}^{\prime})+\sum_{i=2,...,n-1}(k_{neg,i}+k_{neg,i}^{\prime})+L

We break down the above inequality into several components. We first observe for i=2,..,n2i=2,..,n-2 we have

Rpos,i+1+Vpos,i+1lfixed,pos,,i+lfixed,pos,+,i+|Tfixed,i|+kpos,i+1+kpos,i+1Rpos,i+1R_{pos,i+1}^{\prime}+V_{pos,i+1}^{\prime}\leq l_{fixed,pos,-\infty,i}+l_{fixed,pos,+\infty,i}+|T_{fixed,i}|+k_{pos,i+1}+k^{\prime}_{pos,i+1}-R_{pos,i+1}

We first observe the multiplicities counted by Rpos,i+1R_{pos,i+1}^{\prime} and Vpos,i+1V_{pos,i+1}^{\prime} are disjoint - if a Reeb orbit appear in considerations of Rpos,i+1R_{pos,i+1}^{\prime} then it is not considered for Vpos,i+1V_{pos,i+1}^{\prime} and vice versa. Multiplicities counted by Vpos,i+1V_{pos,i+1}^{\prime} are contained in kpos,,i+1k_{pos,-\infty,i+1} and |Tfixed,i+1||T_{fixed,i+1}|, and the Reeb orbits counted by Rpos,i+1R_{pos,i+1}^{\prime} are contained in the ends counted by lfixed,pos,,i+1l_{fixed,pos,-\infty,i+1} and kpos,,i+1k_{pos,-\infty,i+1}. We observe for this range of ii, we only needed to use the fixed ends of CiC_{i} in lfixed,il_{fixed,i} as ss\rightarrow-\infty to achieve this inequality, and the prescence of lfixed,i,pos,+l_{fixed,i,pos,+\infty} will make this inequality strict by that factor. Finally we observe kpos,i+1Rpos,i+10k^{\prime}_{pos,i+1}-R_{pos,i+1}\geq 0. This concludes this inequality.

We next consider the case for i=1i=1 for positive Morse-Bott tori, i.e. we consider the inequality

Rpos,2+Vpos,2+Rpos,2lfixed,pos,1+lfixed,pos,+1+|Tfixed,pos,1|+kpos,2+kpos,2R^{\prime}_{pos,2}+V^{\prime}_{pos,2}+R_{pos,2}\leq l_{fixed,pos,-\infty 1}+l_{fixed,pos,+\infty 1}+|T^{\prime}_{fixed,pos,1}|+k_{pos,2}+k_{pos,2}^{\prime}

This inequality does not hold in general. We first observe kpos,2R2,pos0k_{pos,2}^{\prime}-R_{2,pos}\geq 0, and the Reeb orbits counted by Rpos,2R_{pos,2}^{\prime} are included in kpos,2k_{pos,2} and lfixed,pos,,1l_{fixed,pos,-\infty,1}. The issue for Vpos,2V_{pos,2}^{\prime} is slightly more subtle, because each positive Morse-Bott torus can contain one fixed trivial cylinder that is not included in |Tfixed,pos,1||T_{fixed,pos,1}^{\prime}|, hence a Reeb orbit counted by Vpos,2V_{pos,2}^{\prime} that does not necessarily appear on the right hand side. If we follow this trivial cylinder downwards, if we encounter an end of a non-trivial JJ-holomorphic curve that approaches this Reeb orbit at ss\rightarrow\infty, then it will contribute to lfixed,pos,+,il_{fixed,pos,+\infty,i} terms in one of the lower levels. And this lfixed,pos,+,il_{fixed,pos,+\infty,i} term was not used in our previous computations, so after we add up all the terms in the inequality, the overall inequality will still hold.

If we go downwards and do not see a nontrivial end, then there must be a trivial cylinder at the bottom level of the cascade making a contribution to Tfixed,,pos,n1T_{fixed,,pos,n-1}^{\prime} located at this specific Reeb orbit on this positive Morse-Bott torus. This cylinder counted by Tfixed,pos,n1T_{fixed,pos,n-1}^{\prime} is not used anywhere else in any of our other inequalities, so makes up for the deficit coming from the i=1i=1 inequality.

Finally we consider the terms on the last level concerning the positive Morse-Bott tori contributing to our inequality. This is just

|Tfixed,pos,n1|0|T_{fixed,pos,n-1}^{\prime}|\geq 0

which holds trivially. |Tfixed,pos,n1||T_{fixed,pos,n-1}^{\prime}| being nonzero does not necessarily mean our inequality is strict, as some of these may be borrowed to make the inequality hold on the i=1i=1 level as per above.

We now repeat the analogous series of inequalities concerning negative Morse-Bott tori. We first prove the inequalities

Rneg,i+Rneg,i+Vneg,ikneg,i+kneg,i+lfixed,neg,+,i+|Tfixed,neg,i|R_{neg,i}+R^{\prime}_{neg,i}+V^{\prime}_{neg,i}\leq k_{neg,i}+k^{\prime}_{neg,i}+l_{fixed,neg,+\infty,i}+|T_{fixed,neg,i}|

for ii in range 2,,n22,...,n-2. We have as before that Rneg,ikneg,iR_{neg,i}\leq k_{neg,i}. Similarly the count of orbits in Rneg,iR_{neg,i}^{\prime} is included kneg,ik^{\prime}_{neg,i} and lfixed,neg,+,il_{fixed,neg,+\infty,i}, and the count of Vneg,iV_{neg,i}^{\prime} is included among Tfixed,neg,iT_{fixed,neg,i} and kneg,ik^{\prime}_{neg,i}. This concludes the proof of this inequality.

Next we focus on the i=n1i=n-1 case. We consider the inequality

Rneg,n1+Vneg,n1+Rneg,n1lfixed,neg,+,n1+|Tfixed,neg,n1|+kneg,n1+kneg,n1R^{\prime}_{neg,n-1}+V^{\prime}_{neg,n-1}+R_{neg,n-1}\leq l_{fixed,neg,+\infty,n-1}+|T^{\prime}_{fixed,neg,n-1}|+k_{neg,n-1}+k_{neg,n-1}^{\prime}

This does not always hold, as before we first observe kpos,n1Rneg,n10k_{pos,n-1}-R_{neg,n-1}\geq 0, and Rneg,n1R^{\prime}_{neg,n-1} is included in lfixed,neg,+,n1l_{fixed,neg,+\infty,n-1} and kneg,n1k_{neg,n-1}. However each negative Morse-Bott torus can contain one fixed trivial cylinder not included in Tfixed,neg,n1T^{\prime}_{fixed,neg,n-1}. If we follow this trivial cylinder upwards, if we encounter an end of a non-trivial JJ-holomorphic curve that approaches this Reeb orbit at ss\rightarrow-\infty, then it will contribute to lfixed,neg,,il_{fixed,neg,-\infty,i} terms in one of the upper levels. And this lfixed,,neg,,il_{fixed,,neg,-\infty,i} term was not used in our previous computations, so after we add up the terms in the inequality, the overall inequality will still hold.

If we go upwards and do not see a nontrivial end, then there must be a trivial cylinder contributing to Tfixed,neg,1T^{\prime}_{fixed,neg,1} appearing at the very same Reeb oribt. This cylinder’s contribution is not used up by any of our previous inequalities, so makes up for the deficit in the above inequality.

The i=1i=1 level terms for negative Morse-Bott tori is simply |Tfixed,neg,1|0|T_{fixed,neg,1}^{\prime}|\geq 0 which holds trivially. This inequality being strict does not necessarily imply the overall inequality is strict, by the mechanism discussed above.

Adding up the above inequalities we get the inequality in the proposition. ∎

We now state some consequences of the ECH index one condition, assuming transversality can be satisfied.

Corollary 5.29.

Assuming JJ can be chosen to be good, and we have a height one cascade uu^{\text{\Lightning}}. Then we pass to cascade of currents 𝐮\mathbf{u}^{\text{\Lightning}}, the ECH index being one imposes the following conditions:

  1. a.

    𝐮\mathbf{u}^{\text{\Lightning}} is reduced.

  2. b.

    All flow times are strictly positive.

  3. c.

    All curves are embedded. Curves on the same level are disjoint.

  4. d.

    Each level only has one nontrivial curve, the rest are trivial cylinders.

  5. e.

    With the above choice of fixed/free ends, all curves obey partition conditions of free ends for ends that do not land on critical points. They obey the partition conditions for fixed ends for those that land on critical points of ff.

  6. f.

    For any nontrivial curve CC appearing in the cascade of currents 𝐮\mathbf{u^{\text{\Lightning}}}:

    • If CC appears in either u1u^{1} or un1u^{n-1}, then its ends can appear on critical points of ff only as mandated by α1\alpha_{1} or α2\alpha_{2}. All other ends must avoid critical points of ff.

    • If CC appears in a level between u1u^{1} and un1u^{n-1}, its ends can only end on a critical point of ff if this end is then connected by a fixed chain of trivial cylinders to fixed points mandated by α1\alpha_{1} or αn\alpha_{n}. All other ends avoid critical points of ff, and hence are free.

    • Further, if we see a chain of fixed trivial cylinders connecting a positive or negative end of CC to a critical point of ff, suppose the fixed Reeb orbit is called γ\gamma. Then no nontrivial end may land on γ\gamma on any of the levels of the components of the chain of trivial cylinders in either s+s\rightarrow+\infty or ss\rightarrow-\infty. On the level where CC is asymptotic to γ\gamma as ss\rightarrow\infty or ss\rightarrow-\infty, the end of CC is the only end that is asymptotic to γ\gamma as s+s\rightarrow+\infty and ss\rightarrow-\infty respectively.

  7. g.

    In particular, if CC is a nontrivial curve in the cascade, and an end of CC is asymptotic to γ\gamma, a Reeb orbit in the s+s\rightarrow+\infty (resp. -\infty) end, then no other curve (or other ends of CC) in the same level may be asymptotic to γ\gamma as s+s\rightarrow+\infty (resp. )-\infty).

  8. h.

    If an end of a nontrivial curve CC is asymptotic to γ\gamma with multiplicity >1>1, as ss\rightarrow\infty, and if we follow γ\gamma upwards, e.g. we consider CC^{\prime} in the level above which is asymptotic to γ\gamma as ss\rightarrow-\infty. If all curves above CC that are asymptotic to γ\gamma are trivial cylinders, then we cannot draw any conclusions aside from partition conditions of CC. However, if after some chain of gradient flow lines a nontrivial curve C′′C^{\prime\prime} above CC is asymptotic to ϕTf(γ)\phi_{T}^{f}(\gamma) as ss\rightarrow-\infty and is connected to the positive end of CC at γ\gamma via a gradient flow, then by partition conditions both CC and C′′C^{\prime\prime} can only be asymptotic to γ\gamma with multiplicity 1.

Proof.

All statements in the above proposition comes from taking all the inequalities in the previous proposition to be equalities. (a)(a) comes I(𝐮)=I(𝐮~)I(\mathbf{u^{\text{\Lightning}}})=I(\mathbf{\tilde{u}^{\text{\Lightning}}}). (b)(b) comes from L=0L=0. (c)(c) comes from δ(𝐮)=0\delta(\mathbf{u^{\text{\Lightning}}})=0. (d)(d) comes from Ind=0Ind=0, otherwise the cascade lives in a moduli space of dimension greater than zero. (e)(e) comes from the fact that violations of partition conditions for nontrivial curves would make the inequalities comparing Fredholm index to ECH index strict.

Next consider (f)(f), for the nontrivial curves appearing in u1u^{1} or un1u^{n-1}. We first consider the case of u1u^{1}. We observe all contributions to lfixed,+,1l_{fixed,+\infty,1} from the s+s\rightarrow+\infty must be zero for equality in 5.28 to hold. Similarly we observe that for un1u^{n-1} all contributions to lfixed,,n1l_{fixed,-\infty,n-1} from the ss\rightarrow-\infty must be zero for equality to hold.

If CC is a nontrivial curve between u1u^{1} and un1u^{n-1}, we have to separate this into cases. We first assume it has a negative end landing on a critical point of ff on a positive Morse-Bott torus. Then this end makes a contribution to lfixed,pos,l_{fixed,pos,-\infty}, and was used in our computation of inequality. Call this Reeb orbit γ\gamma, and consider levels below CC that have nontrivial ends asymptotic to γ\gamma as s+s\rightarrow+\infty. Say this occurs on level ii. If there are such curves, and if γ\gamma does not appear as a fixed end assigned by α1\alpha_{1} and connected to a trivial cylinder in u1u^{1}, then it is a appearance of lfixed,pos,+,il_{fixed,pos,+\infty,i} that was not used in our proof of inequality in 5.28, hence the inequality is strict.

The case where γ\gamma appears in α1\alpha_{1} as a fixed end of a trivial cylinder is handled as follows. In the case there is a contribution to Tfixed,pos,n1T_{fixed,pos,n-1}^{\prime} on the un1u^{n-1} level from a trivial cylinder at γ\gamma, then we can use the additional lfixed,pos,+,il_{fixed,pos,+\infty,i} at γ\gamma to make the inequality strict. In the case Tfixed,n1T_{fixed,n-1}^{\prime} does not have a trivial cylinder at γ\gamma, then for multiplicity reasons the total multiplicity of nontrivial ends asymptotic to γ\gamma as s+s\rightarrow+\infty in the entire cascade must be greater than equal to two. If they come from two different ends (potentially at different levels), then their contribution to lfixed,pos,+,l_{fixed,pos,+\infty,*} (of various levels) is at least two, which makes the inequality in proposition 5.28 strict. If we only see a single nontrivial end approach γ\gamma as s+s\rightarrow+\infty below u1u^{1} level, then this end must have multiplicity 2\geq 2, and this violation of writhe inequality also ensures the index inequality is strict.

If no nontrivial curves below CC that are positively asymptotic to γ\gamma exist, then with the negative puncture of CC landing at γ\gamma, the negative puncture is connected to the last level un1u^{n-1} at γ\gamma via a chain of fixed trivial cylinders. If γ\gamma is a minimum of ff, then this is a contribution to |Tfixed,pos,n1||T_{fixed,pos,n-1}^{\prime}| that was not considered in the proof of inequality. This will make the overall inequality strict if γ\gamma did not appear as a fixed end connected to a trivial cylinder in u1u^{1}. If γ\gamma did appear (as a fixed end mandated by α1\alpha_{1}), then again for multiplicity reasons there is either an additional lfixed,pos,+il_{fixed,pos,+\infty i} contribution from s+s\rightarrow+\infty ending on γ\gamma on one of the middle levels, or |Tfixed,n1||T_{fixed,n-1}^{\prime}| at γ\gamma has multiplicity greater than or equal to two. Either case makes the index inequality strict.

However if γ\gamma is at a maximum of ff, the inequality is not violated if this is a chain of trivial cylinders connecting to a fixed end mandated by αn\alpha_{n}. If αn\alpha_{n} assigns free ends to this chain of cylinders, then we have extra contributions to Tfixed,pos,n1T_{fixed,pos,n-1}^{\prime} which make the index inequality strict (in this case α1\alpha_{1} cannot assign γ\gamma as a fixed end). Finally if this is indeed a chain of fixed trivial cylinders connecting to a fixed orbit mandated by αn\alpha_{n}, then on the level where CC appears no other nontrivial end may be asymptotic to γ\gamma as ss\rightarrow-\infty, this is because if this is true, then we consider the inequality for CC’s level

Rpos,i+1+Vpos,i+1lfixed,pos,,i+|Tfixed,pos,i|+kpos,i+1+kpos,i+1Rpos,i+1R_{pos,i+1}^{\prime}+V_{pos,i+1}^{\prime}\leq l_{fixed,pos,-\infty,i}+|T_{fixed,pos,i}|+k_{pos,i+1}+k^{\prime}_{pos,i+1}-R_{pos,i+1}

Both nontrivial ends at γ\gamma are counted once by Rpos,i+1R_{pos,i+1}^{\prime}, but twice by lfixed,pos,,il_{fixed,pos,-\infty,i}, which makes this inequality strict. This automatically imposes the partition condition (n)(n) on this particular negative end of CC. Further, down this chain of fixed trivial cylinders, all the way to αn\alpha_{n}, no further lower levels may have non-trivial curves whose ends are asymptotic to γ\gamma as \rightarrow-\infty. This is clear for the lowest level un1u^{n-1}. We already argued lfixed,pos,,n1=0l_{fixed,pos,-\infty,n-1}=0, then all fixed ends landing on γ\gamma must be fixed ends assigned by αn\alpha_{n}, then the partition conditions imposed by ECH index implies we cannot have both trivial and nontrivial ends at γ\gamma. On levels above the lowest level and below the level of CC, this follows from the inequality

Rpos,i+1+Vpos,i+1lfixed,pos,,i+|Tfixed,pos,i|+kpos,i+1+kpos,i+1Rpos,i+1.R_{pos,i+1}^{\prime}+V_{pos,i+1}^{\prime}\leq l_{fixed,pos,-\infty,i}+|T_{fixed,pos,i}|+k_{pos,i+1}+k^{\prime}_{pos,i+1}-R_{pos,i+1}.

If we have both a trivial cylinder and an nontrivial end asymptotic to γ\gamma in the negative end, they make an overall contribution of 11 to the left hand side, but make a overall contribution of 2 to the right hand side by increasing lfixed,pos,,il_{fixed,pos,-\infty,i} and |Tfixed,pos,i||T_{fixed,pos,i}|, hence making this inequality strict.

We next consider CC has a positive end ending on a critical point of ff. Call this Reeb orbit of γ\gamma. If γ\gamma is not a fixed Reeb orbit mandated by α1\alpha_{1}, then this already makes a contribution to lfixed,pos,+,il_{fixed,pos,+\infty,i} we did not use in the index inequality, which makes the overall inequality strict. If γ\gamma indeed appears in α1\alpha_{1} and is in fact connected to a trivial cylinder, then either this end of CC is connected upwards to γ\gamma via a sequence of trivial cylinders, or there are more nontrivial ends above CC that ends on γ\gamma as s+s\rightarrow+\infty, but this makes the index inequality strict due to multiplicity reasons (α1\alpha_{1} can only require a fixed end of multiplicity 1 at γ\gamma). Hence it must be the case CC is connected to γ\gamma on the top level via sequence of fixed trivial cylinders, and no level above CC have nontrivial ends approaching γ\gamma as s+s\rightarrow+\infty. If a curve above CC has a negative end approaching γ\gamma, we are back to the previous case and this also makes the index inequality strict.

The case of negative Morse-Bott tori is similar to positive Morse-Bott tori but with the signs reversed, so we will not repeat it. We remark the proof of Negative Morse-Bott tori is independent of the proof of positive Morse-Bott tori because when we compute |Tfixed,i||T_{fixed,i}^{\prime}| the trivial cylinders at negative and positive Morse-Bott tori are independent of each other.

To prove (g)(g) and (h)(h). We already took care of the case a non-trivial curve that is asymptotic to a Reeb orbit corresponding to a critical point of ff. We next consider the case of free ends. Let our curve be CC in some level of the cascade and consider its ++\infty free ends asymptotic to positive Morse-Bott tori. We have kpos,i+1=Rpos,i+1k_{pos,i+1}^{\prime}=R_{pos,i+1}, this implies each free Reeb orbit as s+s\rightarrow+\infty is approached by a unique positive end of CC. The ECH index also imposes partition conditions of (1,..,1)(1,..,1), hence this end is simply covered. Recalling 𝐮\mathbf{u}^{\text{\Lightning}} is reduced, any ss\rightarrow-\infty free end of curves above CC arrived at by following the gradient flow is also simply covered. This proves (g)(g) and (h)(h) for positive Morse-Bott tori. The result for negative Morse-Bott tori holds by considering the negative free ends of CC. ∎

We would also like a way to prove that provided our transversality conditions hold (i.e. JJ is good), JδJ_{\delta}-holomorphic curves of ECH index one degenerate into cascades of height one, as opposed to cascades of greater height. To do this we need a slight strengthening of the above index inequality where we allow fixed trivial cylinders with higher multiplicities.

Proposition 5.30.

Let α1\alpha_{1} and αn\alpha_{n} be ECH Morse-Bott generators, except we relax the condition on multiplicities of fixed/free ends - they are allowed to be arbitrary. Let uu^{\text{\Lightning}} be a cascade of height one connecting from α1\alpha_{1} to αn\alpha_{n}. Then we have the inequality

Ind(𝐮~)I(𝐮)2δ(𝐮)1Ind(\mathbf{\tilde{u}}^{\text{\Lightning}})\leq I(\mathbf{u}^{\text{\Lightning}})-2\delta(\mathbf{u}^{\text{\Lightning}})-1
Proof.

We repeat the proof of index inequality in Proposition 5.28 and observe the inequalities concerning the intermediate level curves continue to hold. The issue is in allowing fixed trivial cylinders of high multiplicities allowed by α1\alpha_{1} and αn\alpha_{n} at the top and bottom levels. We first focus on what happens near positive Morse-Bott tori. For simplicity we fix γ\gamma a Reeb orbit corresponding to the hyperbolic orbit in a positive Morse-Bott torus and consider what happens to ends of holomorphic curves with fixed ends at γ\gamma. As we have seen above the problematic term comes from the inequality

Rpos,2+Vpos,2kpos,2+kpos,2Rpos,2+lfixed,pos,,1+lfixed,pos,+,1+|Tfixed,1|,R_{pos,2}^{\prime}+V_{pos,2}^{\prime}\leq k_{pos,2}+k_{pos,2}^{\prime}-R_{pos,2}+l_{fixed,pos,-\infty,1}+l_{fixed,pos,+\infty,1}+|T_{fixed,1}^{\prime}|,

where Vpos,2V_{pos,2}^{\prime} can contain fixed trivial cylinders mandated by α1\alpha_{1} that appear in V2,posV_{2,pos}^{\prime} but does not appear in |Tfixed,1||T_{fixed,1}^{\prime}|. For simplicity we consider Tγ,fixedT_{\gamma,fixed} appearing at γ\gamma of multiplicity NN. In order for this to make a contribution to V2,posV_{2,pos}^{\prime} instead of R2,posR_{2,pos}^{\prime}, we assume that u1u^{1} has no nontrivial end that are asymptotic to γ\gamma as ss\rightarrow-\infty. We recall we would like to prove an inequality of the form

I(u)1Ind(𝐮~)+2δ(u)I(u^{\text{\Lightning}})-1\geq Ind(\mathbf{\tilde{u}}^{\text{\Lightning}})+2\delta(u^{\text{\Lightning}})

Consider for i=2,,n1i=2,...,n-1, the nontrivial currents (Ci,j,mi,j)ui(C_{i,j},m_{i,j})\subset u^{i}, where we think of mi,jm_{i,j} as the multiplicity of Ci,jC_{i,j} (since we are working in the nonreduced case). We assume each Ci,jC_{i,j} has li,jl_{i,j} ends asymptotic to γ\gamma as ss\rightarrow\infty, and suppose Ci,jC_{i,j} has total multiplicity ni,jn_{i,j} asymptotic to γ\gamma as ss\rightarrow\infty. Finally let Tfixed,n1,γT_{fixed,n-1,\gamma} denote the number of trivial cylinders at the last level un1u^{n-1} at γ\gamma. We have the inequality

Ni,jmi,jni,j|Tfixed,n1,γ|.N-\sum_{i,j}m_{i,j}n_{i,j}\leq|T^{\prime}_{fixed,n-1,\gamma}|.

Let’s consider I(Ci,j)I(C_{i,j}), by virtue of it being nontrivial and the writhe inequality, jI(Ci,j)j(ni,j+1)\sum_{j}I(C_{i,j})\geq\sum_{j}(n_{i,j}+1). This is coming from the fact in order for the Ci,jC_{i,j} to exist its Fredholm index must be greater or equal to one, and at the ends of γ\gamma the ECH index is treated as free ends whereas the Fredholm index is treated as fixed ends. So in passing from uiu^{i} to u~i\tilde{u}^{i} we decreased the ECH index by at least i,j(mi,j1)(ni,j+1)\sum_{i,j}(m_{i,j}-1)(n_{i,j}+1).

We next compare the ECH index of reduced cascade with its Fredholm index, in particular we consider the inequalities

I(ui~)Ind(ui~)+Rpos,i+1+Vpos,i+1[lfixed,i+|Tfixed,i|+kpos,i+1+kpos,i+1Rpos,i+1]0I(\tilde{u^{i}})-Ind(\tilde{u^{i}})+R_{pos,i+1}^{\prime}+V_{pos,i+1}^{\prime}-[l_{fixed,i}+|T_{fixed,i}|+k_{pos,i+1}+k^{\prime}_{pos,i+1}-R_{pos,i+1}]\geq 0

for i=2,..,n2i=2,..,n-2. We have that by virtue of the writhe inequality occurring at γ\gamma across these levels, the γ\gamma orbit’s contribution is that the left hand side is at least jni,jli,j\sum_{j}n_{i,j}-l_{i,j} bigger than the right hand side.

Finally, on the un1u^{n-1} level, we originally had the inequality

|Tfixed,n1|0|T^{\prime}_{fixed,n-1}|\geq 0

In the above inequality we have included the |Tfixed,n1,γ||T^{\prime}_{fixed,n-1,\gamma}| term coming from the last level in our cascade contributed by γ\gamma, and the writhe bound for this level also implies this there is also an excess of the index inequality of size jnn1,jln1,j\sum_{j}n_{n-1,j}-l_{n-1,j}.

Hence we can think of proving the index inequality as follows: there is a deficit of NN at the top level contributed purely by γ\gamma, and by making the inequalities of the lower levels strict, we can make up for it. In passing from nonreduced to reduced curve, the “excess” of ECH index is bounded below by i,j(mi,j1)(ni,j+1)\sum_{i,j}(m_{i,j}-1)(n_{i,j}+1). The excess of comparing ECH index of reduced curves Ci,jC_{i,j} to their Fredholm index coming from writhe inequality is given by i,jni,jli,j\sum_{i,j}n_{i,j}-l_{i,j}, and the excess in the index inequality of various levels due to contributions to lfixed,pos,+,il_{fixed,pos,+\infty,i} coming from γ\gamma is precisely i,jli,j\sum_{i,j}l_{i,j}. And on the last level the excess is given simply by |Tfixed,n1,γ||T^{\prime}_{fixed,n-1,\gamma}| Hence the excess due to γ\gamma is bounded below by

i,j(mi,j1)(ni,j+1)+i,jni,jli,j+i,jli,j+|Tfixed,n1,γ|\sum_{i,j}(m_{i,j}-1)(n_{i,j}+1)+\sum_{i,j}n_{i,j}-l_{i,j}+\sum_{i,j}l_{i,j}+|T^{\prime}_{fixed,n-1,\gamma}|

Using the fact Ni,jmi,jni,j|Tfixed,n1,γ|N-\sum_{i,j}m_{i,j}n_{i,j}\leq|T^{\prime}_{fixed,n-1,\gamma}|, we see the excess outweighs the deficit at the top level, so fixed trivial cylinders at γ\gamma will keep the overall index inequality intact. We can apply the same reasoning for every γ\gamma at positive Morse-Bott tori.

We next consider negative Morse-Bott tori. We assume γ\gamma is Reeb orbit on a negative Morse-Bott torus, and αn1\alpha_{n-1} assigns a fixed end of multiplicity NN to γ\gamma. We consider the overall inequality and show it still holds after we factor in the contributions from other terms. Let |Tfixed,1,γ||T_{fixed,1,\gamma}^{\prime}| denote the number of free trivial cylinders located at γ\gamma at the u1u^{1} level. For i=1,..,n2i=1,..,n-2 we consider (Ci,j,mi,j)ui(C_{i,j},m_{i,j})\subset u^{i} nontrivial curves that asymptote to γ\gamma as ss\rightarrow-\infty. We let li,jl_{i,j} denote the number of such ends at each level and ni,jn_{i,j} denote the multiplicity. Then the same proof as before will show the inequality continues to hold. ∎

In fact we have equality of ECH index to Fredholm index also enforces that the cascade is simple.

We now take care of the case of height kk cascades.

Proposition 5.31.

Consider a sequence of JδnJ_{\delta_{n}}-holomorphic ECH index one curves unu_{n} of bounded energy from α1\alpha_{1} to αn\alpha_{n} (as nondegenerate ECH generators) converging to a cascade uu^{\text{\Lightning}} from α1\alpha_{1} and αn\alpha_{n} viewed as Morse-Bott ECH generators, then uu^{\text{\Lightning}} has height one.

Proof.

Suppose uu^{\text{\Lightning}} is a height kk cascade, then it can be written as kk height 11 cascades, which we write as v1,,vkv_{1}^{\text{\Lightning}},...,v_{k}^{\text{\Lightning}}. We recall that between cascades viv_{i}^{\text{\Lightning}} and vi+1v_{i+1}^{\text{\Lightning}} their end asymptotics are connected by either infinite or semi-infinte gradient flows. We pass each to a cascade of currents, and to each cascade 𝐯i\mathbf{v}_{i}^{\text{\Lightning}} we assign to it two generalized ECH generators at its topmost and bottom-most level, which we write as αi\alpha_{i} and αi+1\alpha_{i+1}^{\prime}. For αi\alpha_{i} we assign all the ends approaching the minimum of ff as fixed, and all others are free. For αi+1\alpha_{i+1}^{\prime} we consider all ends approaching the maximum of ff are fixed, and the rest are free. The exception to this rule is α1\alpha_{1} and αk+1\alpha_{k+1}^{\prime} which we assign Morse-Bott ECH generators corresponding to the degenerating JδJ_{\delta}-holomorphic curve. With this we can assign an ECH index to each cascade I(vi)I(v^{\text{\Lightning}}_{i}). We can also assign a relative ECH index between the general ECH generators αi\alpha_{i} and αi\alpha_{i}^{\prime}, which we write as I(αi,αi)I(\alpha_{i}^{\prime},\alpha_{i}). This number is always 0\geq 0, and we illustrate it as follows. Let 𝒯\mathcal{T} be a Morse-Bott torus, and suppose coming from αi\alpha_{i} there is multiplicity n1n_{1} at the minimum of ff and n2n_{2} away from minimum of ff. From αi\alpha_{i}^{\prime} there is n1n_{1}^{\prime} multiplicity at the maximum of ff, and n2n_{2}^{\prime} away from the maximum of ff. Then we have the inequalities

n1n2n_{1}^{\prime}\geq n_{2}

and

n2n1.n_{2}^{\prime}\leq n_{1}.

Then we say contribution to I(αi,αi)I(\alpha_{i}^{\prime},\alpha_{i}) from this Morse-Bott torus is (n1n2)=n1n20(n_{1}-n_{2}^{\prime})=n_{1}^{\prime}-n_{2}\geq 0. Then we add up this term for each Morse-Bott torus that appears in αi\alpha_{i}. Geometrically this is the total mulitplicity of complete gradient trajectories flowing between viv_{i}^{\text{\Lightning}} and vi1v_{i-1}^{\text{\Lightning}} and has potentially nonzero contributions to the ECH index. Then the fact that the cascade came from a ECH index one curve implies

I(v1)+I(α2,α2)++I(vk)=1I(v^{\text{\Lightning}}_{1})+I(\alpha_{2}^{\prime},\alpha_{2})+...+I(v^{\text{\Lightning}}_{k})=1

And by previous proposition each I(vi)0I(v_{i})\geq 0,with equality only if it consisted entirely of fixed trivial cylinders. Hence there is a unique viv_{i}^{\text{\Lightning}} with ECH index 1, the rest have ECH index zero, and all I(αi,αi)=0I(\alpha_{i}^{\prime},\alpha_{i})=0. This means there can only be fixed trivial cylinders above and below viv_{i}^{\text{\Lightning}} and cannot be infinite gradient flows. This is equivalent to saying the cascade of currents is height one. ∎

The above gives a description of what ECH index one cascades look like from the perspective of currents, we now reverse the process, and use the above to understand all cascades of curves of ECH index one. We need to add back in the information that was lost from passing from curves to currents. We only care about the cascades of curves that resulted from degeneration of a nondegenerate connected ECH index one curve. Call this curve CδC_{\delta}. We observe the Fredholm index of CδC_{\delta}, which we denote by Fred Ind(Cδ)\text{Fred Ind}(C_{\delta}), is equal to one. We assume as δ0\delta\rightarrow 0, CδC_{\delta} degenerates into a cascade of curves uu^{\text{\Lightning}}, and denote 𝐮\mathbf{u}^{\text{\Lightning}} the resulting cascade of holomorphic currents. From the above we know 𝐮\mathbf{u}^{\text{\Lightning}} is a cascade of currents of height one, however uu^{\text{\Lightning}} could apriori be of arbitrary height, and the levels that are removed from uu^{\text{\Lightning}}to form 𝐮\mathbf{u}^{\text{\Lightning}} must all be branched covers of trivial cylinders occurring at critical points of ff.

The first case we need to consider is if 𝐮\mathbf{u}^{\text{\Lightning}} is empty, then this implies that uu^{\text{\Lightning}} consists purely of branched covers of trivial cylinders. To be precise uu^{\text{\Lightning}} may contain many levels that consists of branched covers of trivial cylinders, and levels that begin and end on critical point of ff, however it may also contain levels where the trivial cylinders (branched covered or not) are away from critical points of ff. Here we allow levels where there is only a single unbranched cylinder away from critical points of ff. We assume CδC_{\delta} is connected. If at level ii a trivial cylinder is at the critical point of ff corresponding to elliptic Reeb orbit (hyperbolic for negative Morse-Bott torus) then all levels above ii the trivial cylinders that connected to the original cylinder will be at the same Reeb orbit. Similarly if at level ii a trivial cylinder is at the hyperbolic orbit (resp elliptic orbit for negative Morse-Bott torus) then all the trivial cylinders below this level connecting to this original (potentially branched cover of) cylinder will also be at the same Reeb orbits.

If all the levels of uu^{\text{\Lightning}} are at the same Reeb orbit which is also a critical point, then uu came from a branched cover of trivial cylinder in the nondegenerate case. If this is not the case, then remove the top most and bottom most levels until none of the trivial cylinders in uu^{\text{\Lightning}} begin/end on critical point of ff. Then as currents we don’t care where the branched points are, so we can think of uu^{\prime} as a cascade of currents with only 1 level. Then the ECH index of 𝐮\mathbf{u}^{\text{\Lightning}} is equal to one, which implies 𝐮\mathbf{u}^{\text{\Lightning}} consists of a free trivial cylinder with multiplicity one. Hence the same must be true of uu^{\text{\Lightning}} and there are no top/bottom branch covers.

We now turn our attention to the case where 𝐮\mathbf{u}^{\text{\Lightning}} is nonempty. We shall use the fact the Fredholm index of CδC_{\delta} is one to rule out configurations of height >1>1. We observe the trivial cylinders on levels above/below 𝐮\mathbf{u}^{\text{\Lightning}} admit the following description:

Proposition 5.32.
  1. a.

    Let 𝒯\mathcal{T} denote a positive Morse Bott torus contained in the top level of 𝐮\mathbf{u}^{\text{\Lightning}}. For curves on the top level of 𝐮\mathbf{u}^{\text{\Lightning}}, as s+s\rightarrow+\infty all free ends have multiplicity one, and avoid critical point of ff. The fixed end can only have multiplicity one. Hence all branched covers of trivial cylinders above this level can only happen at the critical point of ff corresponding to the elliptic orbit. Moreover, because CδC_{\delta} obeys partition conditions, the top most level in uu^{\text{\Lightning}} of the stack of branched trivial cylinders has partition conditions (1,..,1)(1,..,1).

  2. b.

    Let 𝒯\mathcal{T} denote a negative Morse Bott torus contained in the top level of 𝐮\mathbf{u}^{\text{\Lightning}}, as s+s\rightarrow+\infty. The positive free end of the top level of 𝐮\mathbf{u}^{\text{\Lightning}} has multiplicity 1, so there cannot be branched cover of trivial cylinder at the critical point of ff corresponding to the hyperbolic orbit. The fixed end at the critical point of ff corresponding to the elliptic orbit can have a stack of branched cover of trivial cylinders on top of it on height levels above uu^{\prime}, and again by partition conditions on CδC_{\delta} the top most level is hit by partition condition (n)(n).

  3. c.

    Let 𝒯\mathcal{T} denote a positive Morse Bott torus contained in the bottom level of 𝐮\mathbf{u}^{\text{\Lightning}}. The free end has multiplicity one, so there cannot be branched covers of trivial cylinders at the critical point of ff corresponding to the hyperbolic orbit. The fixed end at critical point of ff corresponding to elliptic end can have a stack of branched cover of trivial cylinders below it on height levels below uu^{\prime}, and again by partition conditions on CδC_{\delta} the top most level is hit by partition condition (n)(n).

  4. d.

    Let 𝒯\mathcal{T} denote a negative Morse Bott torus contained in the bottom level of 𝐮\mathbf{u}^{\text{\Lightning}}. As ss\rightarrow-\infty all free ends have multiplicity one, and avoid the critical points of ff. The fixed end can only have multiplicity one. Hence all branched covers of trivial cylinders above this level can only happen at the critical point of ff corresponding to the elliptic orbit. Moreover, because CδC_{\delta} obeys partition conditions, the bottom most level (in terms of height) of the stack of branched trivial cylinders has partition conditions (1,..,1)(1,..,1).

In light of the above, we can compute the topological Fredholm index of CδC_{\delta} via the following procedure:

First consider the height level corresponding to 𝐮\mathbf{u}^{\text{\Lightning}}, we know all trivial cylinders connecting between nontrivial curves are simply covered, so all the possible branched covers that appear on this height level are chains of trivial branched covers of cylinders that connect to the top and bottom levels of 𝐮\mathbf{u}^{\text{\Lightning}}. We then create two additional height levels, one above 𝐮\mathbf{u}^{\text{\Lightning}}, denoted by 𝐮¯\overline{\mathbf{u}^{\text{\Lightning}}} and one below 𝐮\mathbf{u}^{\text{\Lightning}}, denoted by 𝐮¯\underline{\mathbf{u}^{\text{\Lightning}}}, and push all branch points of trivial cylinders that appear in 𝐮\mathbf{u}^{\text{\Lightning}} onto these 2 levels 𝐮¯\overline{\mathbf{u}^{\text{\Lightning}}},𝐮¯\underline{\mathbf{u}^{\text{\Lightning}}}, so that all trivial cylinders that appear in 𝐮\mathbf{u}^{\text{\Lightning}} have no branch point (though they may be multiply covered), and hence are transversely cut out. We recall we assign Ind(𝐮)Ind(\mathbf{u}^{\text{\Lightning}}) as the dimension of moduli space of 𝐮\mathbf{u}^{\text{\Lightning}} lives in, viewed as a cascade of currents

Then the Fredholm index of CδC_{\delta} is computed as:

Ind(Cδ)=\displaystyle Ind(C_{\delta})=
Ind(𝐮)+1χ(𝐮¯)χ(𝐮¯)\displaystyle Ind(\mathbf{u}^{\text{\Lightning}})+1-\chi(\overline{\mathbf{u}^{\text{\Lightning}}})-\chi(\underline{\mathbf{u}^{\text{\Lightning}}})

Note by the ECH index assumption Ind(𝐮)=0Ind(\mathbf{u}^{\text{\Lightning}})=0, so it will enforce no branched cover of trivial cylinders appear. Hence we have the proved the following proposition:

Proposition 5.33.

Suppose JJ is chosen to be good, if CδC_{\delta} is a sequence of connected nontrivial ECH index one curves of bounded energy that converges to a cascade of curves, uu^{\text{\Lightning}}, then either

  • uu^{\text{\Lightning}} is a free cylinder of multiplicity one

  • uu^{\text{\Lightning}} is the same as a height one cascade of currents of ECH index one, described above, and all trivial cylinders that appear in levels of uu^{\text{\Lightning}} either unbranched chains of fixed trivial cylinders, or trivial cylinders over a Reeb orbit of multiplicity one.

In the latter case, uu^{\text{\Lightning}} does not contain a sequence of fixed trivial cylinders that do not connect to any nontrivial JJ holomorphic curve. See Convention 2.8.

We call cascades of curves of ECH index one of the form stated in the above theorem good cascades of ECH index 1.

Then this is more or less a complete characterization of ECH index one cascades we should count in the Morse-Bott case provided we can achieve enough transversality. Assuming transversality conditions, we quote a theorem from [Yao22] to show ECH index one cascades can be glued uniquely (up to translation) to ECH index one curves.

Theorem 5.34 (Theorem 3.5 in [Yao22]).

Assuming transversality conditions 4.6, any given ECH index one cascades can be glued uniquely to ECH index one JδJ_{\delta}-holomorphic curves for sufficiently small values of δ>0\delta>0 up to translation in the symplectization direction.

The key is to note ECH index one and transversality implies all of the cascades above are transverse and rigid, as in Definition 3.4 of [Yao22] and hence can be glued. The final ingredient we need is to show that assuming JJ is good, the set of good ECH index one cascades is finite. To do this we need the notion of J0J_{0} index for cascades.

6 Finiteness

In order to prove the differential in Morse-Bott ECH is well defined we need to prove the for given generators α,β\alpha,\beta the set of good ECH index one cascades from α\alpha to β\beta is finite. For JJ-chosen to be good, we already know this set is a zero dimensional space, hence it suffices to prove that it is compact. To this end we develop the analogue of J0J_{0} index in the Morse-Bott world. We start with 1-level cascades then build upwards to nn level cascades. In this section we assume JJ is good throughout.

6.1 Level 1 cascades

Consider an level 1 cascade of ECH index 1 from generator α\alpha to β\beta. In anticipation of multiple level ECH index 1 cascades, here we relax some (but not all) of the conditions on α,β\alpha,\beta to remove conditions that require certain free/fixed ends (depending on whether we are on a positive/negative Morse-Bott torus) to only have multiplicity 1. This corresponds to relaxing the condition in the nondegenerate case to only allow hyperbolic orbits of multiplicity one (see Theorem 5.16). We recall the consequences of generic JJ:

  1. a.

    For positive Morse-Bott tori, as ss\rightarrow\infty, all free ends are disjoint and are asymptotic to Reeb orbits in the torus with multiplicity 1. Let n+pos,freen^{pos,free}_{+} denote the number of such orbits.

  2. b.

    For positive Morse-Bott tori, the fixed ends at ss\rightarrow\infty are disjoint from the free ends. They are hit with partition condition (1)(1). Suppose there are N+pos,fixN^{pos,fix}_{+} such ends.

  3. c.

    For positive Morse-Bott tori, as ss\rightarrow-\infty all free ends are disjoint and cover the Reeb orbits in the torus with multiplicity 1. Let npos,freen^{pos,free}_{-} denote the number of such orbits.

  4. d.

    For positive Morse-Bott tori, as ss\rightarrow-\infty, all fixed ends have partition conditions (n)(n). Suppose there are Npos,fixN^{pos,fix}_{-} such ends, each with multiplicity n,jpos,fixn^{pos,fix}_{-,j}

  5. e.

    For negative Morse-Bott tori, as ss\rightarrow\infty, all free ends are disjoint and cover the Reeb orbits in the torus with multiplicity 1. Let n+neg,freen^{neg,free}_{+} denote the number of such orbits.

  6. f.

    For negative Morse-Bott tori, the fixed ends at ss\rightarrow\infty are disjoint from the free ends. They are hit with partition conditions (n)(n). Suppose there are N+neg,fixN^{neg,fix}_{+} such ends with multiplicity n+,jneg,fixn^{neg,fix}_{+,j}

  7. g.

    For negative Morse-Bott tori, as ss\rightarrow-\infty all free ends are disjoint and cover the Reeb orbits in the torus with multiplicity 1. Let Nneg,freeN^{neg,free}_{-} denote the number of such orbits.

  8. h.

    For negative Morse-Bott tori, as ss\rightarrow-\infty there is only 1 fixed end for each Morse-Bott tori, and has partition conditions (1)(1). Let there be Nneg,fixN^{neg,fix}_{-} such ends total

Definition 6.1.

For a level 1 good ECH index 1 cascade CC connecting generator α\alpha to β\beta, we define:

J0(C,α,β):=cτ(C)+Qτ(C,C)[j(n,jpos,fix1)][j(n+,jneg,fix1)]J_{0}(C,\alpha,\beta):=-c_{\tau}(C)+Q_{\tau}(C,C)-[\sum_{j}(n^{pos,fix}_{-,j}-1)]-[\sum_{j}(n_{+,j}^{neg,fix}-1)] (26)

We observe that J0(C,α,β)J_{0}(C,\alpha,\beta) can be computed from the knowledge of α,β\alpha,\beta and the relative homology class of CC alone. We also remark that the J0J_{0} index can be similarly be defined for nontrivial curves of higher ECH index, as long as they satisfy the long list of partition conditions we listed above, and the same genus bounds below holds. We shall have need for this fact for the proof of finiteness below.

Then we have the following genus bound:

Proposition 6.2.

Let gg denote the genus of a holomorphic curve CC. Then we have the upper bound

χ(C)J0(C,α,β).-\chi(C)\leq J_{0}(C,\alpha,\beta). (27)
Proof.

We recall the adjunction formula in our case says

cτ(C)=χ(C)+Qτ(C)+wτ(C)2δ(C)c_{\tau}(C)=\chi(C)+Q_{\tau}(C)+w_{\tau}(C)-2\delta(C)

plugging this into J0J_{0} yields

J0(C,α,β)=χ(C)wτ(C)[(npos,fix)1][n+neg,fix1]+2δ(C)J_{0}(C,\alpha,\beta)=-\chi(C)-w_{\tau}(C)-[\sum(n^{pos,fix}_{-})-1]-[\sum n_{+}^{neg,fix}-1]+2\delta(C)

hence it suffices to prove

wτ[(npos,fix)1][n+neg,fix1]0.-w_{\tau}-[\sum(n^{pos,fix}_{-})-1]-[\sum n_{+}^{neg,fix}-1]\geq 0.

We break this into cases. If CC is a trivial cylinder, then this is trivial. If CC has a nontrivial component along with fixed trivial cylinders, we only consider the nontrivial component, also denoted by CC. All of the computations below follow from the computations of the writhe bound:

  • At a positive Morse-Bott torus

    • ss\rightarrow\infty, free end. wτ0-w_{\tau}\geq 0 because the multiplicity is one.

    • ss\rightarrow\infty, fixed end wτ0-w_{\tau}\geq 0 because multiplicity is one.

    • ss\rightarrow-\infty, free end. wτ0w_{\tau}\geq 0 by multiplicity.

    • ss\rightarrow-\infty, for given fixed end jj, the writhe at this end satisfies wτnpos,fix1w_{\tau}\geq n_{-}^{pos,fix}-1.

  • At a negative Morse-Bott torus

    • ss\rightarrow\infty, free end. wτ0-w_{\tau}\geq 0 due to multiplicity constraints.

    • ss\rightarrow\infty, for a single fixed end jj, the writhe satisfies wτn+neg,fix1-w_{\tau}\geq n_{+}^{neg,fix}-1.

    • ss\rightarrow-\infty, free end. wτ0w_{\tau}\geq 0 due to multiplicity constraints.

    • ss\rightarrow-\infty, fixed end. wτ0w_{\tau}\geq 0 by multiplicity.

combining all of the above we conclude our inequality.

6.2 Multiple level cascades

We now explain how to generalize the definition of J0(C,α,β)J_{0}(C,\alpha,\beta) to good ECH index one cascades of arbitrary number of levels. Consider a nn level cascade u={u1,..,un}u^{\text{\Lightning}}=\{u^{1},..,u^{n}\} of ECH index one with input α\alpha and output β\beta. Recall we have so called fixed chains of trivial cylinders, i.e. chain of trivial cylinders that all begin/end on a fixed end orbit of either α\alpha or β\beta until this chain of trivial cylinders meet an nontrivial holomorphic curve in one of the intermediate levels (which has an fixed end at said Reeb orbit). We remove all of these kinds of trivial cylinders, then the number J0J_{0} is defined for each of the intermediate cascade levels, which we denote by J0(ui)J_{0}(u^{i}), then we define the J0J_{0} of the entire cascade as

Definition 6.3.
J0(u):=J0(ui)J_{0}(u^{\text{\Lightning}}):=\sum J_{0}(u^{i}) (28)

We observe this definition also only dependents on the relative homology class and α,β\alpha,\beta. Recall the Euler characterisitc of the cascade χ(u)\chi(u^{\text{\Lightning}}) is the Euler characterstic of the surface obtained if we glued a cylinder between each matching end of uiu^{i} and ui+1u^{i+1}, clearly then the Euler characteristic of the cascade is the sum of the Euler characteristic of each of its components. Applying the proposition for level one cascades we get

Proposition 6.4.
χ(u)J0(u).-\chi(u^{\text{\Lightning}})\leq J_{0}(u^{\text{\Lightning}}).

6.3 Finiteness

We finally prove

Theorem 6.5.

Given generators α,β\alpha,\beta, the moduli space of good ECH index 1 cascades from α\alpha to β\beta is compact.

Proof.

Let {um}\{u^{\text{\Lightning}}_{m}\} be a sequence of good ECH index one cascades from α\alpha to β\beta. Each umu^{\text{\Lightning}}_{m} is a cascade of the form {umn}n\{u_{m}^{n}\}_{n}. We show {um}\{u^{\text{\Lightning}}_{m}\} has a convergent subsequence. From the Morse-Bott assumption there is an upper bound to how many cascade levels there are, so we pass to a subsequence where they all have NN levels. For each n=1,..,Nn=1,..,N, we apply the compactness for holomorphic current from [Hut14] to each of unmu^{m}_{n}. To see this, note for fixed nn, the energy constraint of {um}\{u^{\text{\Lightning}}_{m}\} and Morse-Bott condition implies there are only finitely many possible choices for the positive and negative asymptotics of unmu^{m}_{n}, so we pick a subsequence (also denoted by unmu^{m}_{n}) where the positive and negative asymptotics of unmu^{m}_{n} is independent of mm. Here, by positive and negative asymptotics of unmu^{m}_{n} we simply mean the Morse-Bott tori 𝒯\mathcal{T} that unmu^{m}_{n} are asymptotic to at its positive/negative ends, and the total multiplicity of Reeb orbits at each such Morse-Bott tori.

Then using the Gromov compactness for currents (see [Hut14]) applied to {unm}\{u^{m}_{n}\} we conclude we can refine a further subsequence of {unm}\{u^{m}_{n}\} (for all n=1,..,Nn=1,..,N) with the same relative homology class (our notion of relative homology class here is in 2(,,Y)\mathcal{H}_{2}(-,-,Y))). Now for each unmu^{m}_{n} simply the knowledge of its asymptotics (which we can read off directly: by virtue of being part of ECH index one cascade all the ends that avoid the critical points of ff are free, and those at critical points of ff is fixed) and its relative homology class provides an upper bound on its J0J_{0} index. This upper bound on J0J_{0} then provides a bound the genus of each unm,n=1,,Nu_{n}^{m},n=1,...,N.

With the genus bound we can apply SFT compactness: for fixed nn, we observe unmu^{m}_{n} cannot break into a building, for that would yield (if we view umu^{\text{\Lightning}}_{m} as cascade of currents) an ECH index 1 cascade of currents with Ti=0T_{i}=0, which does not exist by genericity conditions. Similarly ruled out by genericity conditions are overlapping free ends and free ends migrating to fixed ends. The umnu^{n}_{m} also cannot converge to a multiple cover of nontrivial curve, for that would yield an ECH cascade of current of index 1 with multiple covers of nontrivial curve, which is ruled out by genericity. Hence we conclude that {um}\{u^{\text{\Lightning}}_{m}\} has a subsequence that converges to a ECH index 1 cascade, and hence we have compactness. ∎

7 Computing ECH in the Morse-Bott setting using cascades

We now define the Morse-Bott ECH chain complex (over 2\mathbb{Z}_{2}). We write the chain complex as

CMB(λ,J):=Θi2Θi.C_{*}^{MB}(\lambda,J):=\bigoplus_{\Theta_{i}}\mathbb{Z}_{2}\langle\Theta_{i}\rangle.

Here Θi={(𝒯j,±,mj)}\Theta_{i}=\{(\mathcal{T}_{j},\pm,m_{j})\} denotes a collections of Morse-Bott ECH generators. Suppose we can choose our JJ to be good, the differential, which we write as MB\partial_{MB} is defined as

MBΘ1,Θ2:={2 count of J-holomorphic cascades 𝒞 of ECH index I=1,so that as s+,𝒞 approaches Θ1 and as s,𝒞 approaches Θ2.}\langle\partial_{MB}\Theta_{1},\Theta_{2}\rangle:=\left\{\begin{tabular}[]{@{}l@{}}$\mathbb{Z}_{2}$\, {count of J-holomorphic cascades}\, $\mathcal{C}$\, {of ECH index} \,$I=1$,\\ {so that as} $s\rightarrow+\infty,\,\mathcal{C}$ \,\text{approaches}\, $\Theta_{1}$ {and as} $s\rightarrow-\infty$,\\ $\mathcal{C}$\, \text{approaches} $\Theta_{2}$.\end{tabular}\right\} (29)

We clarify that in the above definition the cascade 𝒞\mathcal{C} must be decomposable into 𝒞0𝒞1\mathcal{C}_{0}\sqcup\mathcal{C}_{1}, where 𝒞0\mathcal{C}_{0} is a (potentially empty) collection of fixed trivial cylinders with multiplicity, and 𝒞1\mathcal{C}_{1} is a good ECH index one cascade. We note if (T,n)(T,n) is an element of 𝒞0\mathcal{C}_{0}, if it is positively asymptotic to Morse-Bott ECH generator (𝒯,n,±)(\mathcal{T},n,\pm), it is also negatively asymptotic to the Morse-Bott ECH generator (𝒯,n,±)(\mathcal{T},n,\pm) (thus far we only considered nontrivial cascades when we talked about their asymptotics).

We note by Theorem 6.5 the operator MB\partial_{MB} is well defined.

Theorem 7.1.

Assuming JJ is good, the chain complex (CMB,MB)(C_{*}^{MB},\partial_{MB}) computes ECH(Y,ξ)ECH(Y,\xi).

Before we prove this theorem we choose a generic family of almost complex structures JδJ_{\delta}.

Recall that the traditional definition of ECH requires choosing a generic JJ from a residual subset of almost complex structures. For fixed δ>0\delta>0, we say JδJ_{\delta} is ECH adapted if it is an almost complex structure with which the ECH chain complex is well defined.

Definition 7.2.

Consider δ(0,δ0]\delta\in(0,\delta_{0}], we say a path of almost complex structures JδJ_{\delta}, each compatible with λδ\lambda_{\delta} for any δ(0,δ0]\delta\in(0,\delta_{0}], is generic if for any collection of Reeb orbits α,β\alpha,\beta, the moduli space

(α,β,δ):={(u,δ)|¯Jδu=0,usomewhere injective,lims+uconverges toα,limsuconverges toβ}\mathcal{M}(\alpha,\beta,\delta):=\{(u,\delta)|\bar{\partial}_{J_{\delta}}u=0,u\,\textup{somewhere injective},\lim_{s\rightarrow+\infty}u\,\textup{converges to}\,\,\alpha,\lim_{s\rightarrow-\infty}u\,\textup{converges to}\,\,\beta\} (30)

is cut out transversely.

Theorem 7.3.

There is a small enough δ0>0\delta_{0}>0 so that there is a generic path of almost complex structures JδJ_{\delta}, δ(0,δ0]\delta\in(0,\delta_{0}] so that:

  • Jδ0J_{\delta_{0}} is ECH adapted.

  • limδ0Jδ=J\lim_{\delta\rightarrow 0}J_{\delta}=J, where JJ is a generic almost complex structure we have chosen above to count ECH index one cascades.

  • |JJδ|Cδ|J-J_{\delta}|\leq C\delta in CkC^{k} norm, k>100k>100, and JδJ_{\delta} take the prescribed form near small fixed neighborhood of Morse-Bott torus described in Section 3.

  • For a residual subset S(0,δ0]S\subset(0,\delta_{0}], for all δS\delta\in S, JδJ_{\delta} is ECH adapted.

Proof.

This is standard application of Sard-Smale theorem. ∎

Proof of theorem 7.1.

We observe for fixed L>0L>0, there are only finitely many ECH index 1 cascades of energy <L<L. We fix δ0\delta_{0} small enough so that for all δ(0,δ0]\delta\in(0,\delta_{0}] the cascades can be glued (uniquely in our sense specified) to ECH index 1 curves.

We assume δ0\delta_{0} is such that Jδ0J_{\delta_{0}} is ECH adapted. We recall we have chosen a generic family Jδ,δ[0,δ0]J_{\delta},\delta\in[0,\delta_{0}] so that the space:

{(u,Jδ)|δ(0,δ0]uJδholomorphic, somewhere injective ECH index 1}\{(u,J_{\delta})\,|\,\delta\in(0,\delta_{0}]\,\,u\,J_{\delta}\,\text{holomorphic, somewhere injective ECH index 1}\}

modulo translation is a 1-manifold (not necessarily compact). A SFT compactness theorem ([Yao22, Bou02, Bou+03]) tells us the δ=0\delta=0 ends of this manifold are precisely the good ECH index one cascades.

We recall there is a residual set A(0,δ0]A\subset(0,\delta_{0}] so that for all δA\delta\in A, JδJ_{\delta} is ECH adapted and the ECH homology can be computed by counting ECH index one JδJ_{\delta} holomorphic curves for δA\delta\in A.

We make the following observation: if uδu_{\delta} and vδv_{\delta} are JδJ_{\delta}-holomorphic curves of ECH index one that converge to the same cascade as δ0\delta\rightarrow 0, by the gluing theorem, for small enough δ\delta uδu_{\delta} and vδv_{\delta} are in fact the same curve up to \mathbb{R} translation.

Then we claim we can find small enough δA\delta^{\prime}\in A so that the corbordism from δ=0\delta=0 to δ\delta^{\prime} built by {(u,Jδ)|δ(0,δ]uJδholomoprhic, somewhere injective ECH index 1}\{(u,J_{\delta})\,|\,\delta\in(0,\delta^{\prime}]\,\,u\,J_{\delta}\,\text{holomoprhic, somewhere injective ECH index 1}\} is the trivial cobordism. Suppose not, then for arbitrarily small δ\delta we can find uδu_{\delta} a ECH index one somewhere injective curve that does not come from gluing, take δ0\delta\rightarrow 0 and after taking a subsequence, uδu_{\delta} degenerates into a good ECH index one cascade, but by our observation must have come from a curve obtained by gluing together an ECH index one cascade, contradiction. ∎

8 ECH index one curves of genus zero

We showed in the previous section that when there is enough transversality for cascades, the cascades of ECH index one take a particularly nice form. However this is not always achievable, except in special circumstances. In this section and the next we outline some special circumstances in which transversality can always be achieved. Here we consider the case where all ECH index one curves in the perturbed picture must have genus zero. This is the case for T3T^{3} and some toric domains.

We shall use a slightly different description of cascades that do not allow for the presence of trivial cylinders. We will call this description “tree-like” cascades and will be described below. The reason we can use this description is that if the curve has genus zero, we can do the gluing without requiring that between each adjacent cascade levels there is a single flow time parameter; instead we can assign a different flow time between each pair of adjacent nontrivial curves.

We use the following convention to represent our holomorphic curves. We use a vertex to represent a JJ holomorphic curve of genus zero, and use directed edges to denote the positive and negative punctures of the curve. Edges directed away from the JJ-holomorphic curve correspond to positive punctures, and edges directed towards the vertex correspond to negative punctures. The figure below illustrates how we go from JJ-holomorphic curve to vertex with directed edges.

Refer to caption
Figure 2: Passing from genus zero curve to vertex with edges

Then a height one cascade with tree-like compactifications from Morse-Bott ECH generator consists of the following data:

  1. a.

    A collection of vertices {v1,..,vn}\{v_{1},..,v_{n}\} each equipped with the data of inward and outward pointing edges. Each vertex has at least one outgoing edge. Each edge is also equipped with the information of which Reeb orbit it lands on.

  2. b.

    Given two vertices viv_{i} and vjv_{j}, if we can find a Morse-Bott torus 𝒯\mathcal{T} so that a positive puncture of viv_{i} lands on γ\gamma, and if we follow the gradient flow for time Ti,j[0,)T_{i,j}\in[0,\infty) along γ\gamma we arrive at a negative puncture of vjv_{j} landing on the corresponding orbit, then we say it is possible to connect viv_{i} and vjv_{j} via the given pair of edges. The data of a height one cascade in this compactification consists of choices of connections between the vertices of {v1,..,vn}\{v_{1},..,v_{n}\}, so that after we connect the edges, we obtain a connected tree. See figure below for an example. We call these connections internal connections.

  3. c.

    The positive punctures of {v1,..,vn}\{v_{1},..,v_{n}\} that are not assigned internal connections are assigned free/fixed as per required by ECH generator α1\alpha_{1}, and likewise for negative punctures and αn\alpha_{n}.

    Refer to caption
    Figure 3: Cascade with tree like compactification. The green arrow denote finite gradient flow lines.

For genus zero JδJ_{\delta}-holomorphic curves degenerating into a cascade with our previous compactification, we can easily pass to a tree like compactification by removing all the trivial cylinders.

Given a cascade of height one with tree like compactification, which we write as u={v1,..,vn}u^{\text{\Lightning}}=\{v_{1},..,v_{n}\}. We can compute its ECH index as follows: we treat all edges participating in internal connections as free, then the ECH index is simply given by

I(u)=I(v1)+.+I(vn)n+1.I(u^{\text{\Lightning}})=I(v_{1})+....+I(v_{n})-n+1.

In order to talk about Fredholm index we also need to pass to the reduced cascade u~\tilde{u}^{\text{\Lightning}} consisting of curves {v1~,..,vn~}\{\tilde{v_{1}},..,\tilde{v_{n}}\}. If in our tree like compactification all free ends assigned by α1\alpha_{1} and αn\alpha_{n} as well as all internal connections avoided critical points of ff, then the reduced cascade lives in a transversely cut out moduli space of dimension

iInd(vi~)1\sum_{i}Ind(\tilde{v_{i}})-1

since being tree like removes the condition of needing to have the same flow time between adjacent cascade levels.

Hence to achieve the necessary transversality conditions to count ECH index one cascades, we choose a generic JJ so that

  1. a.

    For any punctured sphere that is the domain of a JJ-holomorphic curve, we endow it with an assignment of incoming and outgoing punctures, and for each end we assign a free/fixed end; and if an end is assigned fixed it must land on a Reeb orbit corresponding to a critical point of ff under the JJ-holomorphic map; and if an end is free it must avoid critical points of ff. Then all moduli spaces of somewhere injective JJ holomorphic curves with the above information are transversely cut out with dimension given by the index formula.

  2. b.

    For any two curves v1v_{1} and v2v_{2} satisfying the above condition and both rigid, if their free ends land on the same Morse-Bott torus from opposite sides (one as a positive puncture the other as a negative puncture), then they do not land on the same Reeb orbit in the Morse-Bott family (we only care about where they land on the Morse-Bott torus and ignore information of multiplicity, i.e. even if they cover the same Reeb orbit of different multiplicity on their free ends, this is prohibited).

The above conditions are easily achieved by choosing a generic JJ by classical transversality methods. We next consider cascades of height one. We observe we have the inequality (if we treat all internal connections as free for both ECH index and Fredholm index)

I(u)nInd(vi~)10I(u^{\text{\Lightning}})-n\geq\sum Ind(\tilde{v_{i}})-1\geq 0

since each v~i\tilde{v}_{i}, by virtue of it existing and transversality conditions, must have Fredholm index 0\geq 0. ECH index one implies Ind(vi~)=1Ind(\tilde{v_{i}})=1, hence all these curves are rigid, and embedded. By the above genericity of JJ all flow times are nonzero, and the cascade itself is already reduced. All free ends and ends coming from internal connections avoid critical points of ff. Also observe that by partition conditions derived previous sections that between internal connections, the participating edges can only over Reeb orbits with multiplicity one.

Then suppose a sequence of genus zero ECH index one JδJ_{\delta} holomorphic curves from α1\alpha_{1} to αn\alpha_{n} degenerates into a cascade with tree like compactification for arbitrary height. This just means we allow internal connections adjoint to each other with semi-infinite or infinite gradient trajectories. Then for each internal connection whose flow time is infinite, we separate them into two different cascades. Then we get a collection of height one cascades each of which is tree like. We write them as u1,,uku^{\text{\Lightning}}_{1},...,u^{\text{\Lightning}}_{k}. Then we can assign generalized ECH generators to ends of uiu^{\text{\Lightning}}_{i} as before, and the ECH index one condition imposes

I(u1)+I(u2)++I(uk)+relative difference between ECH generators=1I(u^{\text{\Lightning}}_{1})+I(u_{2}^{\text{\Lightning}})+\dots+I(u_{k}^{\text{\Lightning}})+\textup{relative difference between ECH generators}=1

By relative difference between ECH generators we mean the same construction as proposition 5.31. We have for all height one cascades that

I(ui)1Ind(ui~)0I(u_{i}^{\text{\Lightning}})-1\geq Ind(\tilde{u_{i}}^{\text{\Lightning}})\geq 0

Hence there is either a unique cascade uiu_{i}^{\text{\Lightning}} of index zero, or the entire cascade is just one gradient flow line. By considerations of topological Fredholm index we also rule out additional branched cover of trivial cylinders at the top/bottom level of the cascade with tree- like compactifications. Hence using the above description we have the following proposition.

Proposition 8.1.

In the nondegenerate case, ECH index one curves of genus zero degenerate into ECH index one tree like cascades that are reduced and transversely cut out.

We call the type of cascades of the above proposition “good ECH index one tree like cascades”, because we eliminated branched covers of trivial cylinders via topological Fredholm index.

As in the previous section we choose JδJ_{\delta} to be a generic family of almost complex structures satisfying the same conditions as Theorem 7.3.

We then quote a gluing theorem from [Yao22].

Theorem 8.2.

Let uu^{\text{\Lightning}} be a good ECH index one cascade of genus zero as per above, then for small enough δ>0\delta>0 there exists a unique (up to translation) JδJ_{\delta}-holomorphic curve in an ϵ\epsilon neighborhood of this cascade.

Proof.

The main difference is that because the whole curve is genus zero, we no longer need to make sure the pregluing is well defined by restricting our choice of asymptotic vectors to Δ^\hat{\Delta}, as in proposition 8.28 in [Yao22]. ∎

We define a chain complex as before. We We write the chain complex as

CMB,tree(λ,J):=Θi2Θi.C_{*}^{MB,tree}(\lambda,J):=\bigoplus_{\Theta_{i}}\mathbb{Z}_{2}\langle\Theta_{i}\rangle.

We use the superscript “tree” to denote the fact we are counting tree like cascades. As before Θi={(𝒯j,±,mj)}\Theta_{i}=\{(\mathcal{T}_{j},\pm,m_{j})\} denotes a collections of Morse-Bott ECH generators. After we choose a generic JJ, all good tree like cascades are transversely cut out. Then we define the differential MBTree\partial_{MB}^{Tree} to be

MBtreeΘ1,Θ2:={2 count of tree like J-holomorphic cascades 𝒞 of ECH index I=1,so that as s+,𝒞 approaches Θ1 and as s,𝒞 approaches Θ2.}\langle\partial^{tree}_{MB}\Theta_{1},\Theta_{2}\rangle:=\left\{\begin{tabular}[]{@{}l@{}}$\mathbb{Z}_{2}$\, {count of tree like J-holomorphic cascades}\, $\mathcal{C}$\, {of ECH index} \,$I=1$,\\ {so that as} $s\rightarrow+\infty,\,\mathcal{C}$ \,\text{approaches}\, $\Theta_{1}$ {and as} $s\rightarrow-\infty$,\\ $\mathcal{C}$\, \text{approaches} $\Theta_{2}$.\end{tabular}\right\} (31)

As before, we clarify in the cascade 𝒞\mathcal{C} must be decomposable into 𝒞0𝒞1\mathcal{C}_{0}\sqcup\mathcal{C}_{1}, where 𝒞0\mathcal{C}_{0} is a (potentially empty) collection of fixed trivial cylinders with multiplicity, and 𝒞1\mathcal{C}_{1} is a good ECH index one tree like cascade.

Theorem 8.3.

Suppose JJ is chosen to be generic so that all ECH index one good tree like cascades are transversely cut out, and we can choose a generic family of perturbations to JJ, which we write as JδJ_{\delta} that meets the conditions of Theorem 7.3. We further for small enough δ>0\delta>0, all JδJ_{\delta}-holomorphic curves of ECH index one are genus zero. Then the chain complex (CMB,tree,MBTree)(C_{*}^{MB,tree},\partial_{MB}^{Tree}) computes ECH(Y,ξ)ECH(Y,\xi).

Proof of Theorem 8.3.

The same proof as in Theorem 7.1 works. ∎

9 Applications to concave toric domains

As an application of our methods we show that for concave toric domains, ECH can be computed via enumeration of ECH index one cascades. By what we proved above, it suffices to show all ECH index one holomorphic curves after the Morse-Bott perturbation have genus zero.

We recall the definition of a concave toric domain. Consider 2\mathbb{C}^{2} equipped with the standard symplectic product symplectic form. Consider the diagonal S1S^{1} action on 2\mathbb{C}^{2}, and the associated moment map μ:22\mu:\mathbb{C}^{2}\rightarrow\mathbb{R}^{2} given by

μ(z1,z2)=(π|z1|2,π|z2|2).\mu(z_{1},z_{2})=(\pi|z_{1}|^{2},\pi|z_{2}|^{2}).

Let Ω2\Omega\subset\mathbb{R}^{2} be a domain in the first quadrant of 2\mathbb{R}^{2}, we define the domain XΩX_{\Omega} to be

XΩ:={(z1,z2)|μ(z1,z2)Ω}.X_{\Omega}:=\{(z_{1},z_{2})|\mu(z_{1},z_{2})\in\Omega\}.

Suppose Ω\Omega is a domain bounded by the horizontal segment from (0,0)(0,0) to (a,0)(a,0), the vertical segment from (0,0)(0,0) to (0,b)(0,b) and the graph of a convex function f:[0,a][0,b]f:[0,a]\rightarrow[0,b] so that f(0)=bf(0)=b and f(a)=0f(a)=0. We further assume ff is smooth, f(0)f^{\prime}(0) and f(a)f^{\prime}(a) are irrational, f(x)f^{\prime}(x) is constant near 0 and aa, and f′′(x)>0f^{\prime\prime}(x)>0 whenever f(x)f^{\prime}(x) is rational, then we say XΩX_{\Omega} is a concave toric domain. Note our definition is slightly more restrictive than that of [Cho+14], because we are not interested in capacities; we need the boundary of XΩX_{\Omega} to be well behaved enough to define ECH.

For a concave toric domain XΩX_{\Omega}, its boundary XΩ\partial X_{\Omega} is a contact 3-manifold diffeomorphic to S3S^{3}. We now describe the Reeb orbits that appear in XΩ\partial X_{\Omega}. We also note their Conley Zehnder indices, having chosen the same trivializations as in [Cho+14].

  1. a.

    γ1={(z1,0)XΩ}\gamma_{1}=\{(z_{1},0)\in\partial X_{\Omega}\}. The orbit γ1\gamma_{1} is elliptic with rotation angle 1/f(a)-1/f^{\prime}(a), hence CZ(γ1k)=2k/f(a)+1CZ(\gamma_{1}^{k})=2\lfloor-k/f^{\prime}(a)\rfloor+1

  2. b.

    γ2={(0,z2)XΩ}\gamma_{2}=\{(0,z_{2})\in\partial X_{\Omega}\}. The orbit γ2\gamma_{2} has rotation angle f(0)-f^{\prime}(0), hence CZ(γ2k)=2kf(0)+1CZ(\gamma_{2}^{k})=2\lfloor-kf^{\prime}(0)\rfloor+1.

  3. c.

    Let x(0,a)x\in(0,a) be such that f(x)f^{\prime}(x) is rational. Then the torus described by {(z1,z2)|μ(z1,z2)=(x,f(x))}\{(z_{1},z_{2})|\mu(z_{1},z_{2})=(x,f(x))\} is a (negative) Morse-Bott torus. Each Reeb orbit has Robbin-Salamon index 1/2-1/2.

We say a bit more about the Reeb dynamics for the third case. Consider the point (x,f(x))(x,f(x)) so that f(x)f^{\prime}(x) is rational. We set f(x)=tan(ϕ),ϕ(π/2,0)f^{\prime}(x)=tan(\phi),\phi\in(-\pi/2,0). Then the Reeb vector field is given by (see [Mun20])

R=2πxsin(ϕ)+f(x)cos(ϕ)(sinϕθ1+cos(ϕ)θ2).R=\frac{2\pi}{-x\sin(\phi)+f(x)\cos(\phi)}(-\sin\phi\partial_{\theta_{1}}+\cos(\phi)\partial_{\theta_{2}}).

For large action L>0L>0, we perturb each Morse-Bott torus to a pair of orbits, one elliptic, the other hyperbolic. Then an ECH generator α={αi,mi}\alpha=\{\alpha_{i},m_{i}\} is a collection of nondegenerate Reeb orbits with multiplicities. We associate to each ECH generator a combinatorial generator.

Definition 9.1.

(see [Cho+14]) A combinatorial generator is a quadruple Λ~=(Λ,ρ,m,n)\tilde{\Lambda}=(\Lambda,\rho,m,n) where

  1. a.

    Λ\Lambda is a concave integral path from (0,B)(0,B) to (A,0)(A,0) such that the slope of each edge is in the interval [f(0),f(a)][f^{\prime}(0),f^{\prime}(a)].

  2. b.

    ρ\rho is a labeling of each edge of Λ\Lambda by ee or hh.

  3. c.

    mm and nn are nonnegative integers.

Let Λm,n\Lambda_{m,n} denote the concatenation of the following sequence of paths:

  1. a.

    The highest polygonal path with vertices at lattice points from (0,B+n+mf(0))(0,B+n+\lfloor-mf^{\prime}(0)\rfloor) to (m,B+n)(m,B+n) which is below the line through (m,B+n)(m,B+n) with slope f(0)f^{\prime}(0).

  2. b.

    The image of Λ\Lambda under the translation (x,y)(x+m,y+n)(x,y)\mapsto(x+m,y+n).

  3. c.

    The highest polygonal path with vertices at lattice points from (A+m,n)(A+m,n) to (A+m+n/f(a),0)(A+m+\lfloor-n/f^{\prime}(a)\rfloor,0) which is below the line through (A+m,n)(A+m,n) with slope f(a)f^{\prime}(a).

Let (Λm,n)\mathcal{L}(\Lambda_{m,n}) denote the number of lattice points bounded by the axes and Λm,n\Lambda_{m,n}, not including the lattice points on the image of Λ\Lambda under the translation (x,y)(x+m,y+n)(x,y)\mapsto(x+m,y+n). We then define

Icomb(Λm,n)=2(Λm,n)+h(Λ)I^{comb}(\Lambda_{m,n})=2\mathcal{L}(\Lambda_{m,n})+h(\Lambda)

where h(Λ)h(\Lambda) is the number of edges in Λ\Lambda labelled by hh. To each ECH generator α={(αi,mi)}\alpha=\{(\alpha_{i},m_{i})\} we associate a combinatorial ECH generator (Λ,m,n)(\Lambda,m,n) as follows. The number mm is the multiplicity of γ2\gamma_{2} as it appears in α\alpha, and the integer nn is the multiplicity of γ1\gamma_{1} as it appears in α\alpha. For other (nondegenerate) Reeb orbits of α\alpha, they all come from small perturbations of Morse-Bott tori. If γα\gamma\in\alpha is a Reeb orbit that comes from breaking the degeneracy of a Morse-Bott torus at (x,f(x))(x,f(x)), then let v1v_{1} be the smallest positive integer so that v2=f(x)v1v_{2}=f^{\prime}(x)v_{1}\in\mathbb{Z}. Let vv denote the vector v=(v1,v2)v=(v_{1},v_{2}). The path is obtained by taking each Reeb orbit γ\gamma in α\alpha that come from Morse-Bott tori, associating to it the vector that is vv multiplied by the multiplicity of γ\gamma as it appears in α\alpha, and concatenating these vectors in order of increasing slope. The labelling ρ\rho is obtained by labelling the vector associated to γ\gamma the letter hh if γ\gamma is hyperbolic, and ee if γ\gamma is elliptic.

Proposition 9.2.

([Cho+14]) If CC is a current from α\alpha to β\beta, its ECH index is given by Icomb(α)Icomb(β)I^{comb}(\alpha)-I^{comb}(\beta).

For future usage, we also record how the Chern class is computed (see [Cho+14]). Let α\alpha denote a ECH generator, we associate to it the combinatorial generator (Λ,ρ,m,n)(\Lambda,\rho,m,n), then we take

cτ(α)=A+B+m+n.c_{\tau}(\alpha)=A+B+m+n.

Then if we have a JJ-holomorphic curves from ECH generator α\alpha to β\beta, then its relative first Chern class is calculated by cτ(α)cτ(β)c_{\tau}(\alpha)-c_{\tau}(\beta).

We need a version of the local energy inequality, which we take up presently. Versions of this inequality have appeared in [HS05, YZ22, CHS20, Cho16]. Consider the boundary of Ω\Omega with its intersections with the two coordinate axes removed, then its preimage under the moment map is an interval times a two torus. We write the two torus as (x1,x2)S11×S21(x_{1},x_{2})\in S^{1}_{1}\times S^{1}_{2}, where the first S11S^{1}_{1} is the S1S^{1} coming from rotation in the first complex plane \mathbb{C}, and the second S1S^{1} comes from the second copy of \mathbb{C}. We use \mathbb{Z}\oplus\mathbb{Z} to denote the lattice of first homology with \mathbb{Z} coefficients. Consider a Morse-Bott torus at (x,f(x))(x,f(x)) with f(x)=v2/v1f^{\prime}(x)=v_{2}/v_{1} as before, then the homology class of the Reeb orbit is given by the pair (v2,v1)2(-v_{2},v_{1})\in\mathbb{Z}^{2} (this is true before or after the Morse-Bott perturbation).

Consider F[x0,x1]F_{[x_{0},x_{1}]}, by which we denote the preimage of the graph {(x,f(x))|x[x0,x1]}\{(x,f(x))|x\in[x_{0},x_{1}]\} under the moment map. We similarly consider FxF_{x}, which is the preimage of (x,f(x))(x,f(x)) under the moment map. Let CC be a somewhere injective JJ holomorphic curve, we consider CFx0C\cap F_{x_{0}} (we choose x0x_{0} generically so this intersection is transverse). We orient this intersection using the boundary orientation of CF[x0ϵ,x0]C\cap F_{[x_{0}-\epsilon,x_{0}]}. Its homology class in 2\mathbb{Z}^{2} we write as [Fx][F_{x}].

Proposition 9.3.

Let (p,q)2(p,q)\in\mathbb{Z}^{2} denote the homology of CFx0C\cap F_{x_{0}}, then we have the inequality

p+f(x)q0.p+f^{\prime}(x)q\geq 0.

We further observe equality holds only if CC is a trivial cylinder.

Proof.

We consider CF[x1,x2]C\cap F_{[x_{1},x_{2}]}, and observe with our conventions (CF[x0,x1])=CFx1CFx0\partial(C\cap F_{[x_{0},x_{1}]})=C\cap F_{x_{1}}-C\cap F_{x_{0}}. We next consider

CF[x1,x2]𝑑λ\displaystyle\int_{C\cap F_{[x_{1},x_{2}]}}d\lambda =CFx1λCFx0λ\displaystyle=\int_{C\cap F_{x_{1}}}\lambda-\int_{C\cap F_{x_{0}}}\lambda
=CFx1r1𝑑θ1+r2dθ2CFx0r1𝑑θ1+r2dθ2\displaystyle=\int_{C\cap F_{x_{1}}}r_{1}d\theta_{1}+r_{2}d\theta_{2}-\int_{C\cap F_{x_{0}}}r_{1}d\theta_{1}+r_{2}d\theta_{2}
=(x1x0)p+(f(x1)f(x0))q0.\displaystyle=(x_{1}-x_{0})p+(f(x_{1})-f(x_{0}))q\geq 0.

By taking the limit x0x1x_{0}\rightarrow x_{1}, we conclude the proof. ∎

Suppose the JJ-holomorphic CC current connects from α+\alpha_{+} to α\alpha_{-} and has ECH index one. Suppose CC does not contain trivial cylinder components, hence it is embedded. Let α+\alpha_{+} contain γ1\gamma_{1} with multiplicity n+n_{+}, the orbit γ2\gamma_{2} with multiplicity m+m_{+}, and contains e+e_{+} distinct elliptic orbits and h+h_{+} hyperbolic orbits. Suppose further CC has km+k_{m}^{+} ends at γ2\gamma_{2}, with multiplicities m+im_{+}^{i}, and CC has kn+k_{n}^{+} ends at γ1\gamma_{1} with multiplicities n+in_{+}^{i} Likewise we use m,n,e,hm_{-},n_{-},e_{-},h_{-} and km,mi,kn,nik_{m}^{-},m_{-}^{i},k_{n}^{-},n_{-}^{i} to denote the respective quantities in α\alpha_{-}, except here ee_{-} denotes the number of elliptic Reeb orbits counted with multiplicity. Then the key is the following proposition (similar proofs have appeared in [CHS20, HS05, Cho16])

Proposition 9.4.

For the case of concave toric domains, after a small perturbation away from the Morse-Bott degeneracies, all ECH index one curves have genus zero.

Proof.

Step 1 We know that the integers m±im_{\pm}^{i} and n±in_{\pm}^{i} satisfy partition conditions because CC has ECH index one. Recall that for an elliptic Reeb orbit of rotational angle θ\theta, suppose CC is asymptotic to this Reeb orbit at its positive ends with multiplicity mm. Consider the line y=θxy=\theta x on the xyx-y plane, then draw the maximal concave polygonal path connecting lattice points beneath y=θxy=\theta x. This polygonal path 𝒫\mathcal{P} starts at the origin and connects to (m,mθ)(m,\lfloor m\theta\rfloor). The horizontal displacements of the edges in this path we will write as (mi)(m_{i}) and take the convention that if i<ji<j, then mim_{i} is the segment before mjm_{j} if we count starting from the origin. This gives an integer partition of mm, which is the partition conditions for positive ends of CC that are asymptotic to this Reeb orbit.

We observe that imiθ=θm\sum_{i}\lfloor m_{i}\theta\rfloor=\lfloor\theta m\rfloor. To see this, first it follows from the properties of the floor function that

imiθmθ.\sum_{i}\lfloor m_{i}\theta\rfloor\leq\lfloor m\theta\rfloor.

For the converse inequality, consider the polygonal path 𝒫\mathcal{P} with vertices at (ikmi,ikmiθ)(\sum_{i}^{k}m_{i},\lfloor\sum_{i}^{k}m_{i}\theta\rfloor). It suffices to show

mkθikmiθik1miθ.\lfloor m_{k}\theta\rfloor\geq\lfloor\sum_{i}^{k}m_{i}\theta\rfloor-\lfloor\sum_{i}^{k-1}m_{i}\theta\rfloor.

This follows from the fact that

θikmiθik1miθmk\theta\geq\frac{\lfloor\sum_{i}^{k}m_{i}\theta\rfloor-\lfloor\sum_{i}^{k-1}m_{i}\theta\rfloor}{m_{k}}

which is a consequence of the fact that 𝒫\mathcal{P} is maximally concave.

We next recall the partition conditions for negative ends of CC asymptotic to the Reeb orbit with rotation angle θ\theta. Consider the line y=θxy=\theta x, and the minimal convex path above y=θxy=\theta x that connects between (0,0)(0,0) and (m,mθ(m,\lceil m\theta\rceil through lattice points. The horizontal displacements of the edges of of this path are labelled (in order) mim_{i}, and form the partition conditions for ends of CC. Using a very similar proof as before, we can show

miθ=mθ.\sum\lceil m_{i}\theta\rceil=\lceil m\theta\rceil.

Then we can compute the Fredholm index of CC as

Ind(C)=\displaystyle Ind(C)= 2g2+(e++h++km++kn+)+(e+h+km+kn)\displaystyle 2g-2+(e_{+}+h_{+}+k_{m}^{+}+k_{n}^{+})+(e_{-}+h_{-}+k_{m}^{-}+k_{n}^{-})
+2(A++B++m++n+ABmn)\displaystyle+2(A_{+}+B_{+}+m_{+}+n_{+}-A_{-}-B_{-}-m_{-}-n_{-})
e++e\displaystyle-e_{+}+e_{-}
+(kn++km++km+kn)\displaystyle+(k_{n}^{+}+k_{m}^{+}+k_{m}^{-}+k_{n}^{-})
+i=1kn+2n+i/f(a)+i=1km+2m+if(0)i=1kn2ni/f(a)i=1km2mif(0).\displaystyle+\sum_{i=1}^{k_{n}^{+}}2\lfloor-n_{+}^{i}/f^{\prime}(a)\rfloor+\sum_{i=1}^{k_{m}^{+}}2\lfloor-m_{+}^{i}f^{\prime}(0)\rfloor-\sum_{i=1}^{k_{n}^{-}}2\lceil-n_{-}^{i}/f^{\prime}(a)\rceil-\sum_{i=1}^{k_{m}^{-}}2\lceil-m_{-}^{i}f^{\prime}(0)\rceil.

Step 2 To analyze the above equation further, we first note that

A++n++i=1km+m+if(0)Ani=1kmmif(0)0A_{+}+n_{+}+\sum_{i=1}^{k_{m}^{+}}\lfloor-m_{+}^{i}f^{\prime}(0)\rfloor-A_{-}-n_{-}-\sum_{i=1}^{k_{m}^{-}}\lceil-m_{-}^{i}f^{\prime}(0)\rceil\geq 0 (32)

This is accomplished by considering the interior intersections of CC with γ2×\gamma_{2}\times\mathbb{R}. All such intersection points are positive, by positivity of intersections. The count of interior intersections is given by (see [Hut16a])

l+(C,γ2)l(C,γ2)l_{+}(C,\gamma_{2})-l_{-}(C,\gamma_{2})

where l+l_{+} denotes the linking number of positive ends of CC with γ2\gamma_{2}, and ll_{-} is the linking of negative ends of CC with γ2\gamma_{2}. We note the linking numbers in a concave toric domain are calculated as follows ([Cho+14]):

lk(γ1,γ2)=1,lk(γ1,ov)=v2,lk(γ2,ov)=v1,lk(ov,ow)=min{v1w2,v2w1}.lk(\gamma_{1},\gamma_{2})=1,\quad lk(\gamma_{1},o_{v})=-v_{2},\quad lk(\gamma_{2},o_{v})=v_{1},\quad lk(o_{v},o_{w})=\min\{-v_{1}w_{2},-v_{2}w_{1}\}.

Here we use ovo_{v} to denote nondegenerate orbits that come from perturbing a Morse-Bott torus at (x,f(x))(x,f(x)), with f(x)=v2/v1f^{\prime}(x)=v_{2}/v_{1}.

From this we see that lk+=A++n++i=1km+m+if(0)lk_{+}=A_{+}+n_{+}+\sum_{i=1}^{k_{m}^{+}}\lfloor-m_{+}^{i}f^{\prime}(0)\rfloor, and lk=A+n+i=1kmmif(0)lk_{-}=A_{-}+n_{-}+\sum_{i=1}^{k_{m}^{-}}\lceil-m_{-}^{i}f^{\prime}(0)\rceil. The A±A_{\pm} terms come from ends of CC asymptotic to ovo_{v}, the n±n_{\pm} term comes from ends of CC asymptotic to γ1\gamma_{1}, and the floor and ceiling terms come from ends of CC asymptotic to γ2\gamma_{2} and the fact that CC has ECH index one. From the partition conditions we see that i=1km+m+if(0)=m+f(0)\sum_{i=1}^{k_{m}^{+}}\lfloor-m_{+}^{i}f^{\prime}(0)\rfloor=\lfloor-m_{+}f^{\prime}(0)\rfloor. Likewise we can show

B++m++i=1kn+2n+i/f(a)Bmi=1kn2ni/f(a)0B_{+}+m_{+}+\sum_{i=1}^{k_{n}^{+}}2\lfloor-n_{+}^{i}/f^{\prime}(a)\rfloor-B_{-}-m_{-}-\sum_{i=1}^{k_{n}^{-}}2\lfloor-n_{-}^{i}/f^{\prime}(a)\rfloor\geq 0

Hence we conclude from the Fredholm index formula that if CC has ends at γ+\gamma_{+} or γ\gamma_{-}, then it must have genus 0.

Step 3 Next we consider the case where CC has no ends at γ+\gamma_{+} or γ\gamma_{-}. We assume CC has genus one. Then A+=A,B+=BA_{+}=A_{-},B_{+}=B_{-} from Fredholm index considerations. Let Λ±\Lambda_{\pm} denote the polygonal paths associated to generators α±\alpha_{\pm}. We first show Λ+\Lambda_{+} lies outside Λ\Lambda_{-}. By the above we already know they agree at end points.

As a preamble, we consider the homology classes FxCF_{x}\cap C. First for xx very close to zero, say equal to ϵ>0\epsilon>0, let [Fϵ]=(p,q)[F_{\epsilon}]=(p,q). Then we have p+f(0)q0p+f^{\prime}(0)q\geq 0. Similarly consider [F1ϵ]=(p,q)[F_{1-\epsilon}]=(-p,-q). We have pf(a)q0-p-f^{\prime}(a)q\geq 0. Adding these inequalities to get (f(0)f(a))q0(f^{\prime}(0)-f^{\prime}(a))q\geq 0 from which we deduce q0q\leq 0. Then we have f(a)qpf(0)q-f^{\prime}(a)q\geq p\geq-f^{\prime}(0)q, which implies p=q=0p=q=0. Incidentally this implies a kind of maximal principle for holomorphic curves. Note p+f(x)q=0p+f^{\prime}(x)q=0 only if the curve is a branched cover of a trivial cylinder. This implies for our curves they are confined to have x(0,1)x\in(0,1).

Next we compute [Fx][F_{x}] for any xx irrational and ϵ>0\epsilon>0 sufficiently small. We have

[Fx][Fϵ]+homology class of Reeb orbits in [ϵ,x] approached by positive ends of C\displaystyle[F_{x}]-[F_{\epsilon}]+\text{homology class of Reeb orbits in $[\epsilon,x]$ approached by positive ends of $C$}
homology class of Reeb orbits in [ϵ,x] approached by negative ends of C=0.\displaystyle-\text{homology class of Reeb orbits in $[\epsilon,x]$ approached by negative ends of $C$}=0.

Next we consider the no crossing of polygonal paths.

Suppose the no crossing result does not hold, since we know Λ±\Lambda_{\pm} have the same beginning and end points, there must exists two intersection points which we call (a,b)(a,b) and (c,d)(c,d), with a<ca<c. Then on the interval (a,c)(a,c) the path Λ\Lambda_{-} is strictly above Λ+\Lambda_{+} except at end points where they overlap. Form the line connecting (a,c)(a,c) and (b,d)(b,d), we can find x0(a,c)x_{0}\in(a,c) such that f(x0)=dbcaf^{\prime}(x_{0})=\frac{d-b}{c-a}. We compute [Fx0ϵ][F_{x_{0}-\epsilon}] and apply the local energy inequality to it. We use x0ϵx_{0}-\epsilon to avoid the case where x0x_{0} is the xx coordinate of lattice points in Λ±\Lambda_{\pm}, practically this will not make a difference.

Let the lattice point (p,q)(p,q) have the following property: it is a vertex on Λ+\Lambda_{+}, the edge to the left of this lattice point has slope less than f(x0)f^{\prime}(x_{0}), and the edge to the right of this vertex has slope greater than equal to f(x0)f^{\prime}(x_{0}). Then the contribution to [Fx0ϵ][F_{x_{0}-\epsilon}] from Λ+\Lambda_{+} is simply ((Bq),p)(-(B-q),-p). We also consider the contribution of Fx0ϵF_{x_{0}-\epsilon} from Λ\Lambda_{-}, which takes the form (Bq,p)(B-q^{\prime},p^{\prime}). The lattice point (p,q)(p^{\prime},q^{\prime}) on Λ\Lambda_{-} is chosen the same way as (p,q)(p,q). If no such vertex exists, then Λ\Lambda_{-} must overlap with the line segment connecting (a,b)(a,b) and (c,d)(c,d). Then the point (p,q)(p^{\prime},q^{\prime}) is still the lattice point on Λ\Lambda_{-} which corresponds to the left most end point of where Λ\Lambda_{-} overlaps with the line connecting (a,b)(a,b) to (c,d)(c,d). In either case the local energy inequality says that

(qq)+dbca(pp)0(q-q^{\prime})+\frac{d-b}{c-a}(p^{\prime}-p)\geq 0

We first assume (p,q)(p^{\prime},q^{\prime}) is not on the line connecting (a,b)(a,b) and (c,d)(c,d), then this means that the point (p,q)(p,q) is further away from the line connecting (a,b)(a,b) to (c,d)(c,d) than (p,q)(p^{\prime},q^{\prime}). Geometrically this is described by

(bd)(pp)+(ca)(qq)<0.(b-d)(p-p^{\prime})+(c-a)(q-q^{\prime})<0.

which is impossible. Now assume (p,q)(p^{\prime},q^{\prime}) is on the line connecting (a,b)(a,b) to (c,d)(c,d), then since we have chosen [Fx0ϵ][F_{x_{0}-\epsilon}], we must have p<pp^{\prime}<p. The energy inequality implies

qqpp>dbca\frac{q-q^{\prime}}{p-p^{\prime}}>\frac{d-b}{c-a}

contradicting the geometric picture.

Step 4. After we proved no-crossing in the previous step, we show there cannot be a genus one curve satisfying the assumptions of the previous step. The Fredholm index formula tells us that (recall we are assuming g=1g=1)

1=h++h+2e1=h_{+}+h_{-}+2e_{-}

which means e=0e_{-}=0 and at most one of h+h_{+} and hh_{-} is one. If h+=1h_{+}=1, and h=0h_{-}=0, then α=\alpha_{-}=\emptyset. By inspection CC cannot have ECH index one.

On the other hand, if h+=1h_{+}=1 and h=1h_{-}=1, then Λ\Lambda_{-} consists of a single line segment. Λ+\Lambda_{+} has the same end points as Λ\Lambda_{-} and is concave, hence must also agree with Λ\Lambda_{-} as polygonal paths. One checks easily that in this case the ECH index cannot be one.

This concludes the proof that all ECH index one curves have genus zero. ∎

After we have proved all ECH index one curves have genus zero, we can then use the tree like compatification to describe the moduli space of cascades. However there is the complication that there are two nondegenerate orbits, γ+\gamma_{+} and γ\gamma_{-}. So in the tree like compactification, we allow the ends of JJ-holomorphic curves to land on nondegenerate orbits. Furthermore, connecting between two nontrivial curves, instead of a gradient trajectory, it could be that adjacent ends of JJ-holomorphic curves land on the same non-degenerate orbits and no gradient trajectories connect between them. See figure 4.

Refer to caption
Figure 4: Cascade with tree like compactification for concave toric domains. The unconnected ends of holomorphic curves can land on either Morse-Bott tori or nondegenerate Reeb orbits. The green arrow denotes a finite gradient flow line connecting between two adjacent ends that land on Morse-Bott tori. The dashed line is used to indicate the adjacent ends land on non-degenerate Reeb orbits, and there is no need for gradient trajectories to connect between them.

Given such a cascade of ECH index one, we can cut it into subtrees along each matching pair of nondegenerate orbits, see figure 5.

Refer to caption
Figure 5: We cut along the red dashed lines to sub trees of cascades. For this figure each subtree is circled by dashed blue lines. The ECH index is additive along concatenation of such sub trees.

The ECH index is additive with respect to concatenation of sub-trees. So the ECH index one conditions implies there are no matching along nondegenerate orbits, and we can use the correspondence theorem 8.3 as before.

10 Convex Toric Domains

In this section we show we can compute the ECH chain complex of convex toric domains via enumeration of JJ-holomorphic cascades. As there are many similarities with the case of concave toric domains, we will be brief in its treatment.

Suppose Ω\Omega is a domain bounded by the horizontal segment from (0,0)(0,0) to (a,0)(a,0), the vertical segment from (0,0)(0,0) to (0,b)(0,b) and the graph of a concave function f:[0,a][0,b]f:[0,a]\rightarrow[0,b] so that f(0)=bf(0)=b and f(a)=0f(a)=0. We further assume ff is smooth, f(0)f^{\prime}(0) and f(a)f^{\prime}(a) are irrational, f(x)f^{\prime}(x) is constant near 0 and aa, and f′′(x)<0f^{\prime\prime}(x)<0 whenever f(x)f^{\prime}(x) is rational, then we say XΩX_{\Omega} is a convex toric domain.

As in the case of a concave toric domain, the boundary of XΩX_{\Omega}, written as XΩ\partial X_{\Omega}, is a contact 3-manifold diffeomorphic to S3S^{3}. We now describe the Reeb orbits that appear in XΩ\partial X_{\Omega}. We also note their Conley Zehnder indices, having chosen the same trivializations as in [Hut16]

  1. a.

    γ1={(z1,0)XΩ}\gamma_{1}=\{(z_{1},0)\in\partial X_{\Omega}\}. The orbit γ1\gamma_{1} is elliptic with rotation angle 1/f(a)-1/f^{\prime}(a), hence CZ(γ1k)=2k/f(a)+1CZ(\gamma_{1}^{k})=2\lfloor-k/f^{\prime}(a)\rfloor+1

  2. b.

    γ2={(0,z2)XΩ}\gamma_{2}=\{(0,z_{2})\in\partial X_{\Omega}\}. The orbit γ2\gamma_{2} has rotation angle f(0)-f^{\prime}(0), hence CZ(γ2k)=2kf(0)+1CZ(\gamma_{2}^{k})=2\lfloor-kf^{\prime}(0)\rfloor+1.

  3. c.

    Let x(0,a)x\in(0,a) be such that f(x)f^{\prime}(x) is rational. Then the torus described by {(z1,z2)|μ(z1,z2)=(x,f(x))}\{(z_{1},z_{2})|\mu(z_{1},z_{2})=(x,f(x))\} is a (positive) Morse-Bott torus. Each Reeb orbit has Robbin-Salamon index +1/2+1/2.

Definition 10.1.

A combinatorial generator is a quadruple Λ~=(Λ,ρ,m,n)\tilde{\Lambda}=(\Lambda,\rho,m,n) where

  1. a.

    Λ\Lambda is a convex integral path from (0,B)(0,B) to (A,0)(A,0) such that the slope of each edge is in the interval [f(0),f(a)][f^{\prime}(0),f^{\prime}(a)].

  2. b.

    ρ\rho is a labeling of each edge of Λ\Lambda by ee or hh.

  3. c.

    mm and nn are nonnegative integers.

Let Λm,n\Lambda_{m,n} denote the concatenation of the following sequence of paths:

  1. a.

    The highest polygonal path with vertices at lattice points from (0,B+n+mf(0))(0,B+n+\lfloor-mf^{\prime}(0)\rfloor) to (m,B+n)(m,B+n) which is below the line through (m,B+n)(m,B+n) with slope f(0)f^{\prime}(0).

  2. b.

    The image of Λ\Lambda under the translation (x,y)(x+m,y+n)(x,y)\mapsto(x+m,y+n).

  3. c.

    The highest polygonal path with vertices at lattice points from (A+m,n)(A+m,n) to (A+m+n/f(a),0)(A+m+\lfloor-n/f^{\prime}(a)\rfloor,0) which is below the line through (A+m,n)(A+m,n) with slope f(a)f^{\prime}(a).

Let (Λm,n)\mathcal{L}(\Lambda_{m,n}) denote the number of lattice points bounded by the axes and Λm,n\Lambda_{m,n}, including the lattice points on the edges of Λm,n\Lambda_{m,n}. We then define

Icomb(Λm,n)=2((Λm,n)1)h(Λ)I^{comb}(\Lambda_{m,n})=2(\mathcal{L}(\Lambda_{m,n})-1)-h(\Lambda)

And the Chern class of Λm,n\Lambda_{m,n} is given by

cτ(Λm,n)=A+B+m+n.c_{\tau}(\Lambda_{m,n})=A+B+m+n.
Theorem 10.2.

The ECH index of a holomorphic curve between two ECH generators is the difference of the IcombI^{comb} we associate to their corresponding combinatorial ECH generators.

Proof.

The proof is a generalization of the computation in [Hut16, Cho+14]. We briefly summarize this below. Let α\alpha denote a ECH orbit set. We consider I(α,,Z)I(\alpha,\emptyset,Z) where ZZ is the unique relative homology class that is represented by discs with boundary α\alpha. Let m,nm,n denote the multiplicity of γ2,γ1\gamma_{2},\gamma_{1} respectively in α\alpha, and let Λ\Lambda be the resulting convex integral path defined by associating Reeb orbit sets to integral paths as in [Hut16]. Then it suffices to show I(α,,Z)=Icomb(Λm,n)I(\alpha,\emptyset,Z)=I^{comb}(\Lambda_{m,n}). The computation is the same as the one in [Hut16], except the Conley-Zehnder index terms arising from γ1\gamma_{1} and γ2\gamma_{2} may not just be 11 due to the fact their rotation angles θ\theta need not be very close to zero. This is accounted for by the polygonal paths we append to image of Λ\Lambda under the translation (x,y)(x+m,y+n)(x,y)\mapsto(x+m,y+n).

Theorem 10.3.

A nontrival JδJ_{\delta}-holomorphic curve in a convex toric domain of ECH index one has genus zero. Here we use JδJ_{\delta} to mean we have perturbed away all Morse-Bott degeneracies.

Proof.

We borrow the notation of the previous section, except here e+e_{+} denotes the total multiplicity of elliptic Reeb orbits in α+\alpha_{+} arising from Morse-Bott tori and ee_{-} denotes the total number of distinct elliptic Reeb orbits in α\alpha_{-} arising from perturbations of Morse-Bott tori. The Fredholm index of a connected JJ-holomorphic curve CC between two orbit sets α+\alpha_{+} and α\alpha_{-} is given by

Ind(C)=\displaystyle Ind(C)= 2g2+(e++h++km++kn+)+(e+h+km+kn)\displaystyle 2g-2+(e_{+}+h_{+}+k_{m}^{+}+k_{n}^{+})+(e_{-}+h_{-}+k_{m}^{-}+k_{n}^{-})
+2(A++B++m++n+ABmn)\displaystyle+2(A_{+}+B_{+}+m_{+}+n_{+}-A_{-}-B_{-}-m_{-}-n_{-})
+e+e\displaystyle+e_{+}-e_{-}
+(kn++km++km+kn)\displaystyle+(k_{n}^{+}+k_{m}^{+}+k_{m}^{-}+k_{n}^{-})
+i=1kn+2n+i/f(a)+i=1km+2m+if(0)i=1kn2ni/f(a)i=1km2mif(0).\displaystyle+\sum_{i=1}^{k_{n}^{+}}2\lfloor-n_{+}^{i}/f^{\prime}(a)\rfloor+\sum_{i=1}^{k_{m}^{+}}2\lfloor-m_{+}^{i}f^{\prime}(0)\rfloor-\sum_{i=1}^{k_{n}^{-}}2\lceil-n_{-}^{i}/f^{\prime}(a)\rceil-\sum_{i=1}^{k_{m}^{-}}2\lceil-m_{-}^{i}f^{\prime}(0)\rceil.

The same linking number relations as in 9.4 holds in the case of convex toric domains; so similarly by considering the intersections of CC with the trivial cylinders at γ1\gamma_{1} and γ2\gamma_{2}, we conclude

A++n++i=1km+m+if(0)Ani=1kmmif(0)0A_{+}+n_{+}+\sum_{i=1}^{k_{m}^{+}}\lfloor-m_{+}^{i}f^{\prime}(0)\rfloor-A_{-}-n_{-}-\sum_{i=1}^{k_{m}^{-}}\lceil-m_{-}^{i}f^{\prime}(0)\rceil\geq 0

and

B++m++i=1kn+2n+i/f(a)Bmi=1kn2ni/f(a)0.B_{+}+m_{+}+\sum_{i=1}^{k_{n}^{+}}2\lfloor-n_{+}^{i}/f^{\prime}(a)\rfloor-B_{-}-m_{-}-\sum_{i=1}^{k_{n}^{-}}2\lfloor-n_{-}^{i}/f^{\prime}(a)\rfloor\geq 0.

Hence for CC to have genus nonzero it must not have any ends at γ1\gamma_{1} and γ2\gamma_{2}.

The local energy inequality holds as before, to prove the no-crossing lemma, we can associate two polygonal paths Λ+\Lambda_{+} and Λ\Lambda_{-} to ECH generators α+\alpha_{+} and α\alpha_{-} respectively. As before from index considerations the xx and yy intercepts of Λ+\Lambda_{+} and Λ\Lambda_{-} agree. Hence as before we can choose points (a,b)(a,b) and (c,d)(c,d) where Λ+\Lambda_{+} and Λ\Lambda_{-} intersect, and between these two points Λ\Lambda_{-} is strictly above Λ+\Lambda_{+}. As before we may choose x0(a,c)x_{0}\in(a,c) so that f(x0)=dbcaf^{\prime}(x_{0})=\frac{d-b}{c-a}. Let the lattice point (p,q)(p^{\prime},q^{\prime}) have the following property: it is a vertex on Λ\Lambda_{-}, the edge to the left of this lattice point has slope greater than or equal to f(x0)f^{\prime}(x_{0}), and the edge to the right of this vertex has slope less than f(x0)f^{\prime}(x_{0}). Let (p,q)(p,q) denote a vertex of Λ+\Lambda_{+} with the same property. We assume such a vertex (p,q)(p,q) exists and leave the case where such a vertex does not exist to later. Then consider [Fx0+ϵ]=(qq,pp)[F_{x_{0}+\epsilon}]=(q-q^{\prime},p^{\prime}-p). Now again the energy inequality says

(qq)+dbca(pp)0(q-q^{\prime})+\frac{d-b}{c-a}(p^{\prime}-p)\geq 0

In this case, the point (p,q)(p,q) is closer to the line connecting (a,b)(a,b) and (c,d)(c,d) than (p,q)(p^{\prime},q^{\prime}), but this time on the other side of the line. This means that

(pp)(bd)+(ca)(qq)<0(p-p^{\prime})(b-d)+(c-a)(q-q^{\prime})<0

Comparing with the energy inequality we see a contradiction. Now if (p,q)(p,q) is in fact on the line connecting (a,b)(a,b) and (c,d)(c,d), then since we are computing [Fx0+ϵ][F_{x_{0}+\epsilon}], we must have p>pp>p^{\prime}, from which we have

dbca>qqpp\frac{d-b}{c-a}>\frac{q-q^{\prime}}{p-p^{\prime}}

which is a contradiction.

With the no-crossing result at hand, we turn to the index formula. If CC had genus one, then

1=2e++h++h.1=2e_{+}+h_{+}+h_{-}.

As before we break this into cases. We must have e+=0e_{+}=0. If h+=1h_{+}=1 then Λ+\Lambda_{+} consists of a single edge, by no-crossing Λ\Lambda_{-} is either an identical edge or empty. We check either case cannot produce an ECH index 1 curve. h+h_{+} cannot equal zero because then Λ+=\Lambda_{+}=\emptyset.

Hence we concluded all ECH index one curves are index zero, a similar description of tree-like cascades shows we can use them to compute the ECH chain complex.

Appendix A Appendix: Transversality Issues

In this Appendix we describe some the transversality difficulties in the moduli space of cascades, even if all the appearing curves are somewhere injective. Note we are not claiming transversality is impossible, we are simply saying there are issues with the standard universal moduli space approach of transversality. We give some simple examples below to illustrate this.

Consider the universal moduli space of somewhere injective cascades, written as

:={(u,J)|uis aJ-holomorphic cascade, and that all curves appearing in u are simple}.\mathcal{B}:=\{(u^{\text{\Lightning}},J)|\,u^{\text{\Lightning}}\,\textup{is a}\,J\textup{-holomorphic cascade, and that all curves appearing in $u^{\text{\Lightning}}$ are simple}\}.

We explain why the standard proof that \mathcal{B} is a Banach manifold does not necessarily work. Given a cascade uu^{\text{\Lightning}}\in\mathcal{B}, there are two evaluation maps EV+EV^{+} and EVEV^{-} that map into a product of S1S^{1}, as in Definition 4.5. The usual procedure to show that \mathcal{B} is a Banach manifold is to show the maps EV±EV^{\pm} are transverse to each other. However in complicated enough cascades, the same curve can appear in multiple different levels. An illustration is given in the figure below. Here we have a cascade of 5 levels. The red curve is a map u:Σ×Y3u:\Sigma\rightarrow\mathbb{R}\times Y^{3}, and the blue curve is a map v:Σ×Y3v:\Sigma^{\prime}\rightarrow\mathbb{R}\times Y^{3}. Green horizontal arrows denote the upwards gradient flow, and the black horizontal lines denote Morse-Bott tori. Diamonds denote the critical points of ff on the Morse-Bott tori. For instance, one of the positive ends of the black curve ends on a critical point of ff, and there is a chain of fixed trivial cylinders atop this end.

Refer to caption
Figure 6: Cascade with 5 levels

This is an illustration of how the same curves can happen in the same cascade. To illustrate the transversality issue, we assume that the configuration consisting the red and blue curves (which we labelled uu and vv) in figure 7 happens nn times in a cascade uu^{\text{\Lightning}}.

Refer to caption
Figure 7: A repetitive pattern that can appear multiple times in a cascade.

We assume both uu and vv are rigid (we are allowed since we are working in the universal moduli space, in general more complicated things can still happen but the principle is the same). We label the nn identical copies of uu and vv as ui,viu_{i},v_{i} with i=1,,ni=1,...,n. The two negative ends of uiu_{i} and the two positive ends of viv_{i} are labelled by 1,21,2, as shown in the figure. The remaining end of uiu_{i} and viv_{i} is labelled 33. We denote their evaluation maps by ev(ui,k)ev(u_{i},k) and ev(vi,k)ev(v_{i},k) where k=1,2,3k=1,2,3. As a necessary condition for the EV+EV^{+} and EVEV^{-} to be transverse, we must have

(dev(ui,1)+dev(vi,1)+ti,dev(ui,2)+dev(vi,1)+ti):T𝒲uT𝒲vi=1,..,ni=1,,n(TS1TS1)\bigoplus(dev(u_{i},1)+dev(v_{i},1)+t_{i},dev(u_{i},2)+dev(v_{i},1)+t_{i}):T\mathcal{W}_{u}\oplus T\mathcal{W}_{v}\bigoplus_{i=1,..,n}\mathbb{R}\longrightarrow\bigoplus_{i=1,...,n}(TS^{1}\oplus TS^{1}) (33)

is surjective. Note (t1,,tn)i=1,..,n(t_{1},...,t_{n})\in\bigoplus_{i=1,..,n}\mathbb{R}. The vector space T𝒲uT\mathcal{W}_{u} has the following description. Recall a neighborhood of (not necessarily JJ holomorphic) curves near uu can be represented by W2,p,d(uTM)T𝒥V1V2V3W^{2,p,d}(u^{*}TM)\oplus T\mathcal{J}\oplus V_{1}\oplus V_{2}\oplus V_{3}. Here W2,p,d(uTM)W^{2,p,d}(u^{*}TM) is the Sobolev space of vector fields on uu with exponential weight ed|s|e^{d|s|} near the cylindrical ends. T𝒥T\mathcal{J} is a finite dimensional Teichmuller slice, and the vector spaces ViV_{i} consist of asymptotically constant vectors near each of the cylindrical ends, which we labelled 1,2,31,2,3 (see [Yao22, Wen10]). Recalling the coordinate choices of Section 3 near Morse-Bott tori, the ViV_{i} is spanned by vector fields of the form

βz,βa,βx.\beta\partial_{z},\quad\beta\partial_{a},\quad\beta\partial_{x}.

β\beta here is a cutoff function that is one near a cylindrical neighborhood of a puncture and zero elsewhere. We denote a triple of these vector fields in ViV_{i} as (r,a,p)i(r,a,p)_{i}.

Then the vector space 𝒲u\mathcal{W}_{u} is given by

{(ξ,(r,a,p)i,Y)W2,p,d(uTM)T𝒥V1V2V3T|D¯J(ξ+i(r,a,p)i)+YTuj=0}\{(\xi,(r,a,p)_{i},Y)\in W^{2,p,d}(u^{*}TM)\oplus T\mathcal{J}\oplus V_{1}\oplus V_{2}\oplus V_{3}\oplus T\mathcal{I}|D\bar{\partial}_{J}(\xi+\sum_{i}(r,a,p)_{i})+Y\circ Tu\circ j=0\}

D¯JD\bar{\partial}_{J} is the linearization of Cauchy Riemann operator along uu that includes deformation of the domain complex structure of uu. Here TT\mathcal{I} denotes the Sobolev space that is the tangent space of all λ\lambda compatible almost complex structures (we should choose a Sobolev space for this but that is unimportant for now). A similar expression holds for T𝒲vT\mathcal{W}_{v}. We note the same YTY\in T\mathcal{I} appears in the definition of T𝒲vT\mathcal{W}_{v} as well. Now since uu is rigid for given YY there exists a unique tuple (ξ,(r,a,p)i)(\xi,(r,a,p)_{i}) (up to translation in the symplectization direction) so that (ξ,(r,a,p)i,Y)T𝒲u(\xi,(r,a,p)_{i},Y)\in T\mathcal{W}_{u}. A similar statement holds for 𝒲v\mathcal{W}_{v}. Conversely, given two tuples (p1(u),p2(u),p3(u))(p_{1}(u),p_{2}(u),p_{3}(u)) and (p1(v),p2(v),p3(v))(p_{1}(v),p_{2}(v),p_{3}(v)) (we use brackets to denote whether the vector field is living over uu or vv, we can find YTY\in T\mathcal{I} and (ξ(u),(r(u),a(u))i)(\xi(u),(r(u),a(u))_{i}) and (ξ(v),(r(v),a(v))i)(\xi(v),(r(v),a(v))_{i}) so that the tuples (ξ(u),(r(u),a(u),p(u))i,Y)T𝒲u(\xi(u),(r(u),a(u),p(u))_{i},Y)\in T\mathcal{W}_{u}, and similarly for T𝒲vT\mathcal{W}_{v}. Hence we can think of the map described in Equation 33 as the following. Its imagine is spanned by vector fields of the form

i(x1+y1+ti,x2+y2+ti)\bigoplus_{i}(x_{1}+y_{1}+t_{i},x_{2}+y_{2}+t_{i})

where (x1,y1)(x_{1},y_{1}) and (x2,y2)(x_{2},y_{2}) are arbitrary real numbers. We think of x1x_{1} as p1(u)p_{1}(u), x2x_{2} corresponding to p2(u)p_{2}(u), and likewise for yy and p(v)p(v). For given nn the domain has 2+n2+n independent variables, but the target is 2n2n dimensional. Hence for large values of nn this space cannot be transverse.

Proof of Theorem 1.2.

We note if the above situation does not happen, then the usual proof that \mathcal{B} is a Banach manifold follows through. To be precise, if we let B~\tilde{B} denote the universal moduli space so that

in addition, either all nontrivial curvesare distinct, or the cascade has less than or equal to 3 levels}\tilde{B}:=\left\{(u^{\text{\Lightning}},J)\;\middle|\;\begin{tabular}[]{@{}l@{}}$u^{\text{\Lightning}}$ \, {is a reduced} $J$ {-holomorphic cascade as in Definition \ref{def:transversality conditions}};\\ {in addition, either all nontrivial curves}\\ { are distinct, or the cascade has less than or equal to 3 levels}\end{tabular}\right\}
B~:={(u,J)| u   is a reduced J -holomorphic cascade as in Definition 4.5; (34)

Then B~\tilde{B} is a Banach manifold, and for generic JJ, cascades satisfying the extra hypothesis of B~\tilde{B} are transversely cut out living in moduli spaces given by the virtual dimension. In particular if we take as assumption after we perturb away the Morse-Bott degeneracy, all ECH index one curves degenerate (as reduced cascades) to reduced cascades of the form specified in B~\tilde{B}, then we can choose a JJ so that the conditions 4.5 are satisfied for these cascades. A straightforward modification of the proofs in Sections 6, 7 shows the Morse-Bott chain complex (CMB,MB)(C_{*}^{MB},\partial_{MB}) when we further restrict the differential to only consider cascades whose reduced versions can appear in B~\tilde{B} is well defined and computes ECH(Y,ξ)ECH(Y,\xi). The only different part is showing the cascades counted by MB\partial_{MB} is finite. Consider the following. Suppose unu^{\text{\Lightning}}_{n} is a sequence of cascades of the form allowed in B~\tilde{B} and unuu^{\text{\Lightning}}_{n}\rightarrow u^{\text{\Lightning}}. Then for each unu^{\text{\Lightning}}_{n} there is a sequence of JδnmJ_{\delta_{n}^{m}}-holomorphic curves vnmv_{n}^{m} of ECH index one that converges to unu_{n}^{\text{\Lightning}} as mm\rightarrow\infty. We pass to a diagonal subsequence, which we denote by vnv_{n}, of ECH index one JδnJ_{\delta_{n}}-holomorphic curves that degenerate into uu. By assumption, then the reduced version of uu^{\text{\Lightning}} must be of the form allowed in B~\tilde{B}, and this concludes the proof of finiteness. ∎

References

  • [Bou02] Frederic Bourgeois “A Morse-Bott approach to contact homology”, 2002
  • [Bou+03] Frederic Bourgeois et al. “Compactness results in symplectic field theory” In Geometry & Topology 7.2, 2003, pp. 799–888
  • [Cho16] Keon Choi “Combinatorial embedded contact homology for toric contact manifolds”, 2016 arXiv:1608.07988 [math.SG]
  • [Cho+14] Keon Choi et al. “Symplectic embeddings into four-dimensional concave toric domains” In J. Topol. 7.4, 2014, pp. 1054–1076
  • [Cri19] Dan Cristofaro-Gardiner “Symplectic embeddings from concave toric domains into convex ones” With an appendix by Cristofaro-Gardiner and Keon Choi In J. Differential Geom. 112.2, 2019, pp. 199–232
  • [CHS20] Dan Cristofaro-Gardiner, Vincent Humilière and Sobhan Seyfaddini “Proof of the simplicity conjecture” arXiv, 2020 URL: https://arxiv.org/abs/2001.01792
  • [CH16] Daniel Cristofaro-Gardiner and Michael Hutchings “From one Reeb orbit to two” In J. Differential Geom. 102.1, 2016, pp. 25–36
  • [Far11] David Farris “The embedded contact homology of nontrivial circle bundles over Riemann surfaces”, 2011
  • [Gut14] Jean Gutt “Generalized Conley-Zehnder index” In Annales de la faculté des sciences de Toulouse Mathématiques 23.4 Université Paul Sabatier, Toulouse, 2014, pp. 907–932
  • [Her98] David Hermann “Non-equivalence of Symplectic Capacities for Open Sets with Restricted Contact Type Boundary”, 1998 URL: www.math.u-psud.fr/~biblio/pub/1998/abs/ppo1998_32.html
  • [HWZ96] Helmut Hofer, K. Wysocki and E. Zehnder “Properties of pseudoholomorphic curves in symplectisations. I : asymptotics” In Annales de l’I.H.P. Analyse non linéaire 13.3 Gauthier-Villars, 1996, pp. 337–379
  • [Hut02] Michael Hutchings “An index inequality for embedded pseudoholomorphic curves in symplectizations” In Journal of the European Mathematical Society 004.4 European Mathematical Society Publishing House, 2002, pp. 313–361 URL: http://eudml.org/doc/277717
  • [Hut09] Michael Hutchings “The embedded contact homology index revisited” In New perspectives and challenges in symplectic field theory 49, CRM Proc. Lecture Notes Amer. Math. Soc., Providence, RI, 2009, pp. 263–297
  • [Hut14] Michael Hutchings “Lecture notes on embedded contact homology” In Contact and symplectic topology 26, Bolyai Soc. Math. Stud. János Bolyai Math. Soc., Budapest, 2014, pp. 389–484
  • [Hut16] Michael Hutchings “Beyond ECH capacities” In Geom. Topol. 20.2, 2016, pp. 1085–1126
  • [Hut16a] Michael Hutchings “Mean action and the Calabi invariant” In J. Mod. Dyn. 10, 2016, pp. 511–539
  • [HS05] Michael Hutchings and Michael Sullivan “The periodic Floer homology of a Dehn twist” In Algebr. Geom. Topol. 5, 2005, pp. 301–354
  • [HS06] Michael Hutchings and Michael Sullivan “Rounding corners of polygons and the embedded contact homology of T3T^{3} In Geom. Topol. 10, 2006, pp. 169–266
  • [HT11] Michael Hutchings and Clifford Henry Taubes “Proof of the Arnold chord conjecture in three dimensions 1” In Math. Res. Lett. 18.2, 2011, pp. 295–313
  • [HT13] Michael Hutchings and Clifford Henry Taubes “Proof of the Arnold chord conjecture in three dimensions, II” In Geom. Topol. 17.5, 2013, pp. 2601–2688
  • [Iri15] Kei Irie “Dense existence of periodic Reeb orbits and ECH spectral invariants” In J. Mod. Dyn. 9, 2015, pp. 357–363
  • [Leb07] Eli Lebow “Embedded Contact Homology of 2-Torus Bundles over the Circle”, 2007
  • [MS12] Dusa McDuff and Felix Schlenk “The embedding capacity of 4-dimensional symplectic ellipsoids” In Ann. of Math. (2) 175.3, 2012, pp. 1191–1282
  • [Mun20] Mihai Munteanu “Noncontractible loops of symplectic embeddings between convex toric domains” In J. Symplectic Geom. 18.4, 2020, pp. 1169–1196
  • [NW20] Jo Nelson and Morgan Weiler “Embedded contact homology of prequantization bundles”, 2020 arXiv:2007.13883 [math.SG]
  • [OW18] Yong-Geun Oh and Rui Wang “Analysis of contact Cauchy–Rieman maps II: canonical neighborhoods and exponential convergence for the Morse-Bott case” In Nagoya Mathematical Journal 231 Cambridge University Press, 2018, pp. 128–223
  • [Sie] Richard Siefring “Relative asymptotic behavior of pseudoholomorphic half-cylinders” In Communications on Pure and Applied Mathematics 61.12, pp. 1631–1684
  • [Tau98] Clifford Henry Taubes “The structure of pseudo-holomorphic subvarieties for a degenerate almost complex structure and symplectic form on S1×B3S^{1}\times B^{3} In Geom. Topol. 2, 1998, pp. 221–332
  • [Wen10] Chris Wendl “Automatic transversality and orbifolds of punctured holomorphic curves in dimension four” In Commentarii Mathematici Helvetici, 2010, pp. 347–407
  • [Yao22] Yuan Yao “From Cascades to JJ-holomorphic Curves and Back” arXiv, 2022 DOI: 10.48550/ARXIV.2206.04334
  • [YZ22] Yuan Yao and Ziwen Zhao “Product and Coproduct on Fixed Point Floer Homology of Positive Dehn Twists” arXiv, 2022 URL: https://arxiv.org/abs/2205.14516