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

A contact McKay correspondence for links of simple singularities

Leo Digiosia and Jo Nelson
Abstract

We compute the cylindrical contact homology of the links of the simple singularities. These manifolds are contactomorphic to S3/GS^{3}/G for finite subgroups GSU(2)G\subset\text{SU}(2). We perturb the degenerate contact form on S3/GS^{3}/G with a Morse function, which is invariant under the corresponding HSO(3)H\subset\text{SO}(3) action on S2S^{2}, to achieve nondegeneracy up to an action threshold. The cylindrical contact homology is recovered by taking a direct limit of the action filtered homology groups. The ranks of this homology are given in terms of |Conj(G)||\text{Conj}(G)|, demonstrating a Floer theoretic McKay correspondence.

1 Introduction

A simple singularity is modeled by the isolated singular point of the variety 2/G{\mathbb{C}}^{2}/G, for a finite nontrivial subgroup GSU(2)G\subset\text{SU}(2). The action of GG on [u,v]{\mathbb{C}}[u,v] admits an invariant subring, generated by three monomials, mi(u,v)m_{i}(u,v) for i=1,2,3i=1,2,3, that satisfy a minimal polynomial relation,

fG(m1(u,v),m2(u,v),m3(u,v))=0,f_{G}(m_{1}(u,v),m_{2}(u,v),m_{3}(u,v))=0,

for some nonzero fG[z1,z2,z3]f_{G}\in{\mathbb{C}}[z_{1},z_{2},z_{3}]. These weighted polynomials fGf_{G} provide an alternative perspective of the simple singularities as hypersurface singularities in 3{\mathbb{C}}^{3}. Specifically, the map

2/GVG:=fG1(0),[(u,v)](m1(u,v),m2(u,v),m3(u,v)){\mathbb{C}}^{2}/G\to V_{G}:=f^{-1}_{G}(0),\,\,\,\,\,[(u,v)]\mapsto(m_{1}(u,v),m_{2}(u,v),m_{3}(u,v))

defines an isomorphism of complex varieties, 2/GVG{\mathbb{C}}^{2}/G\simeq V_{G}, and produces a hypersurface singularity given any finite nontrivial GSU(2)G\subset\text{SU}(2). The following table summarizing the relationship of GG to fGf_{G}. The integer triple (p,q,r)(p,q,r) corresponds to the lengths of the 3 branches of the associated Dynkin diagram denoted by Γ(G)\Gamma(G). In the AnA_{n} case, (k,l)(k,l) is an arbitrary pair of positive integers satisfying k+l=n+1k+l=n+1.

Group GG Graph Γ(G)\Gamma(G)      fG(z1,z2,z3)f_{G}(z_{1},z_{2},z_{3}) branches (p,q,r)(p,q,r)
n+1{\mathbb{Z}}_{n+1} AnA_{n} z1n+1+z22+z32z_{1}^{n+1}+z_{2}^{2}+z_{3}^{2} (1,k,l)(1,k,l)
𝔻2n4{\mathbb{D}}^{*}_{2n-4} DnD_{n} z12z2+z1n1+z32z_{1}^{2}z_{2}+z_{1}^{n-1}+z_{3}^{2} (2,2,n2)(2,2,n-2)
𝕋{\mathbb{T}}^{*} E6E_{6} z14+z23+z32z_{1}^{4}+z_{2}^{3}+z_{3}^{2} (2,3,3)(2,3,3)
𝕆\mathbb{O}^{*} E7E_{7} z13z1+z23+z32z_{1}^{3}z_{1}+z_{2}^{3}+z_{3}^{2} (2,3,4)(2,3,4)
𝕀{\mathbb{I}}^{*} E8E_{8} z15+z23+z32z_{1}^{5}+z_{2}^{3}+z_{3}^{2} (2,3,5)(2,3,5)
Table 1: Polynomial relation fGf_{G} for finite subgroups GSU(2)G\subset\text{SU}(2).

One can recover the conjugacy class of GG from VGV_{G} by studying the Dynkin diagram associated to the minimal resolution X~G\widetilde{X}_{G} of a simple singularity 𝟎\mathbf{0} of VGV_{G}, using the McKay correspondence [McK80] summarized below. The Dynkin diagram associated to (the minimal resolution of) (VG,𝟎)(V_{G},\mathbf{0}) is the finite graph whose vertex viv_{i} is labeled by the exceptional holomorphic sphere ZiZ_{i} of self-intersection -2, and viv_{i} is adjacent to vjv_{j} if and only if ZiZ_{i} transversely intersects with ZjZ_{j}. In this way, we associate to any simple singularity (VG,𝟎)(V_{G},\mathbf{0}) the graph Γ(VG,𝟎)\Gamma(V_{G},\mathbf{0}). It is a classical fact that Γ(VG,𝟎)\Gamma(V_{G},\mathbf{0}) is isomorphic to one of the AnA_{n}, DnD_{n}, or the E6E_{6}, E7E_{7}, or E8E_{8} graphs (see [Sl80, §6]), depicted in Figure 1.1.

Refer to caption
Figure 1.1: The Dykin diagrams; AnA_{n} and DnD_{n} feature nn nodes.

The Dynkin diagrams also simultaneously classify the types of conjugacy classes of finite subgroups GG of SU(2)\text{SU}(2). Any finite subgroup GSU(2)G\subset\text{SU}(2) must be either cyclic, conjugate to 𝔻2n{\mathbb{D}}_{2n}^{*}, or is a binary polyhedral group, cf. [Za58, §1.6]. Associated to each type of finite subgroup GSU(2)G\subset\text{SU}(2) is a finite graph, Γ(G)\Gamma(G). The vertices of Γ(G)\Gamma(G) are in correspondence with the nontrivial irreducible representations ViV_{i} of GG, of which there are |Conj(G)|1|\text{Conj}(G)|-1, where Conj(G)\text{Conj}(G) denotes the set of conjugacy classes of a group GG. The McKay correspondence states that Γ(G)\Gamma(G) is isomorphic to one of the AnA_{n}, DnD_{n}, or the E6E_{6}, E7E_{7}, or E8E_{8} graphs, as enumerated in Table 2. The adjacency matrix AijA_{ij} of the Dynkin diagram determines the tensor products 2ViAijVj{\mathbb{C}}^{2}\otimes V_{i}\cong\oplus A_{ij}V_{j} with the canonical representation, cf. [St85]. We also note that the dimension of the cohomology of the minimal resolution is precisely the number of irreducible representations.

GSU(2)G\subset\text{SU}(2) |Conj(G)|1|\text{Conj}(G)|-1 Γ(G)\Gamma(G)
n{\mathbb{Z}}_{n} n1n-1 An1A_{n-1}
𝔻2n{\mathbb{D}}_{2n}^{*} n+2n+2 Dn+2D_{n+2}
𝕋{\mathbb{T}}^{*} 66 E6E_{6}
𝕆{\mathbb{O}}^{*} 77 E7E_{7}
𝕀{\mathbb{I}}^{*} 88 E8E_{8}
Table 2: Dykin diagrams associated to finite subgroups GSU(2)G\subset\text{SU}(2).

We adapt a method of computing the cylindrical contact homology of (S3/G,ξG)(S^{3}/G,\xi_{G}) as a direct limit of action filtered homology groups, described by Nelson in [Ne20]. This process uses a (lift of a) Morse function, which is invariant under the corresponding symmetry group in SO(3)\text{SO}(3), to perturb the standard degenerate contact form. In order to define the exact symplectic cobordism maps necessary to take direct limits, a detailed analysis of the homotopy classes of Reeb orbits is needed due to the presence of contractible and torsion Reeb orbits.

Our computation realizes a contact Floer theoretic McKay correspondence result, namely that the ranks of the cylindrical contact homology of the links111 Recall that the link of a hypersurface singularity in 3{\mathbb{C}}^{3} is the 3-dimensional contact manifold L:=Sϵ5(0){f1(0)}L:=S^{5}_{\epsilon}(0)\cap\{f^{-1}(0)\}, with contact structure ξL:=TLJ3(TL)\xi_{L}:=TL\cap J_{{\mathbb{C}}^{3}}(TL), where J3J_{{\mathbb{C}}^{3}} is the standard integrable complex structure on 3{\mathbb{C}}^{3}, and ϵ>0\epsilon>0 is small. There is a contactomorphism (S3/G,ξG)(L,ξL)(S^{3}/G,\xi_{G})\simeq(L,\xi_{L}), where ξG\xi_{G} on S3/GS^{3}/G is the descent of the standard contact structure ξ\xi on S3S^{3} to the quotient by the GG-action. of simple singularities are given in terms of the number of conjugacy classes of the group GG. It additionally recovers the presentation of the manifold as a Seifert fiber space and, in this sense, provides a natural basis for the cylindrical contact homology in terms of the Reeb orbits realizing the different conjugacy classes of GG, cf. Remark 1.3.

We expect that our explicit description of the cylindrical chain complexes will enable computations of embedded contact homology and its associated spectral invariants after an appropriate adaption of arguments from [NW23, NW2]. Such results will be of interest in the context of gauge theory as well as have applications to the study of symplectic embeddings and fillings. Our computations realize McLean and Ritter’s work, which computes the positive S1S^{1}-equivariant symplectic cohomology of the crepant resolution YY of n/G{\mathbb{C}}^{n}/G in terms of the number of conjugacy classes of the finite GSU(n)G\subset\text{SU}(n), [MR, Theorem 1.10, Corollary 2.13], without needing to know the cohomology of the minimal resolution.

1.1 Definitions and overview of cylindrical contact homology

First we recall some basic definitions. Let (Y,ξ)(Y,\xi) be a closed contact three manifold with defining contact form λ\lambda. This contact form determines a smooth vector field, RλR_{\lambda}, called the Reeb vector field, which uniquely satisfies λ(Rλ)=1\lambda(R_{\lambda})=1 and dλ(Rλ,)=0d\lambda(R_{\lambda},\cdot)=0. A Reeb orbit γ\gamma is a map /TM{\mathbb{R}}/T{\mathbb{Z}}\to M, considered up to reparametrization, with γ˙(t)=Rλ(γ(t))\dot{\gamma}(t)=R_{\lambda}(\gamma(t)). Let 𝒫(λ)\mathcal{P}(\lambda) denote the set of Reeb orbits of λ\lambda. If γ𝒫(λ)\gamma\in\mathcal{P}(\lambda) and kk\in{\mathbb{N}}, then the kk-fold iterate of γ\gamma, denoted γk\gamma^{k}, is the precomposition of γ\gamma with /kT/T{\mathbb{R}}/kT{\mathbb{Z}}\to{\mathbb{R}}/T{\mathbb{Z}}. The orbit γ\gamma is embedded when /TY{\mathbb{R}}/T{\mathbb{Z}}\to Y is injective. If γ\gamma is the mm-fold iterate of an embedded Reeb orbit, then m(γ):=mm(\gamma):=m is the multiplicity of γ\gamma.

For a Reeb orbit γ\gamma as above, the time TT linearized Reeb flow defines a symplectic linear map

Pγ:(ξγ(0),dλ)(ξγ(0),dλ),P_{\gamma}:\left(\xi_{\gamma(0)},d\lambda\right)\to\left(\xi_{\gamma(0)},d\lambda\right),

after making a choice of trivialization, which we also denote by PγP_{\gamma}. We say γ\gamma is nondegenerate if PγP_{\gamma} does not have 1 as an eigenvalue. The contact form λ\lambda is called nondegenerate if all γ𝒫(λ)\gamma\in\mathcal{P}(\lambda) are nondegenerate. A nondegenerate Reeb orbit is said to be elliptic if PγP_{\gamma} has its eigenvalues on the unit circle and hyperbolic if PγP_{\gamma} has real eigenvalues. (If both real eigenvalues are positive then γ\gamma is a positive hyperbolic orbit and if both real eigenvalues are negative then γ\gamma is a negative hyperbolic orbit.)

If τ\tau is a homotopy class of trivializations of ξ|γ\xi|_{\gamma}, then the Conley Zehnder index, μCZτ(γ)\mu_{\operatorname{CZ}}^{\tau}(\gamma)\in{\mathbb{Z}} is defined and related to the rotation of the Reeb flow along γ\gamma. The parity of the Conley-Zehnder index does not depend on the choice of trivialization and is even when γ\gamma is positive hyperbolic and odd when γ\gamma is elliptic. If γ\gamma is an embedded negative hyperbolic orbit then the parity of the Conley-Zehnder index is odd for all odd iterates and even for all even iterates, with respect to any homotopy class of trivializations. An orbit γ𝒫(λ)\gamma\in\mathcal{P}(\lambda) is said to be bad if it is an even iterate of a negative hyperbolic orbit, otherwise, γ\gamma is said to be good. Let 𝒫good(λ)𝒫(λ)\mathcal{P}_{\text{good}}(\lambda)\subset\mathcal{P}(\lambda) denote the set of good Reeb orbits.

If c1(ξ),π2(Y)=0\langle c_{1}(\xi),\pi_{2}(Y)\rangle=0 and if μCZτ(γ)3\mu_{\operatorname{CZ}}^{\tau}(\gamma)\geq 3 for all contractible γ𝒫(λ)\gamma\in\mathcal{P}(\lambda) with any τ\tau extendable over a disc, we say the nondegenerate contact form λ\lambda is dynamically convex. The symplectic vector bundle (ξ,dλ)(\xi,d\lambda) admits a global trivialization if c1(ξ)=0c_{1}(\xi)=0, which is unique up to homotopy if rank H1(Y)=0\mbox{rank }H_{1}(Y)=0. In this case, the integral grading |γ||\gamma| of the generator γ\gamma is defined to be μCZτ(γ)1\mu_{\operatorname{CZ}}^{\tau}(\gamma)-1 for any τ\tau induced by a global trivialization of ξ\xi.

Definition 1.1.

We say that an almost complex structure JJ on ×Y{\mathbb{R}}\times Y is λ\lambda-compatible if

  • J(ξ)=ξJ(\xi)=\xi;

  • dλ(v,Jv)>0d\lambda(v,Jv)>0 for nonzero vξv\in\xi;

  • JJ is invariant under translation of the {\mathbb{R}} factor;

  • J(s)=RλJ(\partial_{s})=R_{\lambda}, where ss denotes the {\mathbb{R}} coordinate.

We denote the set of all λ\lambda-compatible JJ by 𝒥(Y,λ)\mathcal{J}(Y,\lambda).

Fix such a λ\lambda-compatible JJ. If γ+\gamma_{+} and γ\gamma_{-} are Reeb orbits, we consider JJ-holomorphic cylinders interpolating between them, which are smooth maps u:×S1×Yu:{\mathbb{R}}\times S^{1}\to{\mathbb{R}}\times Y such that the nonlinear Cauchy-Riemann equation holds

su+Jtu=0,\partial_{s}u+J\partial_{t}u=0,

lims±πu(s,t)=±\lim_{s\to\pm\infty}\pi_{{\mathbb{R}}}\circ u(s,t)=\pm\infty, and lims±πYu(s,)\lim_{s\to\pm\infty}\pi_{Y}u(s,\cdot) is a parametrization of γ±\gamma_{\pm}. Here π\pi_{\mathbb{R}} and πY\pi_{Y} are the respective projections from ×Y{\mathbb{R}}\times Y to {\mathbb{R}} and YY. We say that uu is positively asymptotic to γ+\gamma_{+} and negatively asymptotic to γ\gamma_{-}. We declare two maps to be equivalent if they differ by translation and rotation of the domain ×S1{\mathbb{R}}\times S^{1}, and denote the set of equivalence classes by J(γ+,γ)\mathcal{M}^{J}(\gamma_{+},\gamma_{-}). There is an additional {\mathbb{R}} action J(γ+,γ)\mathcal{M}^{J}(\gamma_{+},\gamma_{-}) by translation of the {\mathbb{R}} factor on the target ×Y{\mathbb{R}}\times Y.

We define the Fredholm index of a cylinder uJ(γ+,γ)u\in\mathcal{M}^{J}(\gamma_{+},\gamma_{-}) by

ind(u)=μCZτ(γ+)μCZτ(γ)+2c1(uξ,τ),\text{ind}(u)=\mu_{\operatorname{CZ}}^{\tau}(\gamma_{+})-\mu_{\operatorname{CZ}}^{\tau}(\gamma_{-})+2c_{1}(u^{*}\xi,\tau),

after fixing a trivialization τ\tau of ξ\xi over γ+\gamma_{+} and γ\gamma_{-}. The relative first Chern class c1(uξ,τ)c_{1}(u^{*}\xi,\tau) vanishes when τ\tau extends to a trivialization of uξu^{*}\xi. For kk\in{\mathbb{Z}}, kJ(γ+,γ)\mathcal{M}_{k}^{J}(\gamma_{+},\gamma_{-}) denotes those cylinders with ind(u)=k\text{ind}(u)=k. The significance of the Fredholm index is that if JJ is generic and ukJ(γ+,γ)u\in\mathcal{M}_{k}^{J}(\gamma_{+},\gamma_{-}) is somewhere injective, then kJ(γ+,γ)\mathcal{M}^{J}_{k}(\gamma_{+},\gamma_{-}) is naturally a manifold near uu of dimension kk.

For a nondegenerate contact form λ\lambda, and under favorable transversality conditions, we define the cylindrical contact homology chain complex CC(Y,λ,J)CC_{*}(Y,\lambda,J) over {\mathbb{Q}} as follows. (The original definition is due to Eliashberg-Givental-Hofer [EGH00] and we are using notation from [HN22], but suppressing some decorations as we only consider one cylindrical flavor of contact homology in this paper.) As a module, CC(Y,λ,J)CC_{*}(Y,\lambda,J) is noncanonically isomorphic to the vector space over {\mathbb{Q}} generateed by good Reeb orbits; an isomorphism is fixed after a choice of coherent orientations, which is used to define a {\mathbb{Z}}-module 𝒪γ\mathcal{O}_{\gamma} that is noncanonically isomorphic to {\mathbb{Z}}, cf. [HN22, A.3]. We then define

CC(Y,λ,J)=γ𝒫good(λ)𝒪γ.CC_{*}(Y,\lambda,J)=\bigoplus_{\gamma\in\mathcal{P}_{\text{good}}(\lambda)}\mathcal{O}_{\gamma}\otimes_{\mathbb{Z}}{\mathbb{Q}}.

The choice of a generator of 𝒪γ\mathcal{O}_{\gamma} for each good Reeb orbit specifies an isomorphism

CC(Y,λ,J)𝒫good(λ).CC_{*}(Y,\lambda,J)\simeq{\mathbb{Q}}\langle\mathcal{P}_{\text{good}}(\lambda)\rangle.

This chain complex admits a canonical /2{\mathbb{Z}}/2-grading determined by the mod 2 Conley-Zehnder index, which can be upgraded to a relative or absolute {\mathbb{Z}} grading in certain circumstances. In the setting of this paper, we have an absolute {\mathbb{Z}} grading given by

|γ|=μCZτ(γ)1,|\gamma|=\mu_{\operatorname{CZ}}^{\tau}(\gamma)-1,

where τ\tau is any homotopy class of the global unitary trivialization constructed in (2.2) and Remark 2.12.

To define the differential, we first define the following operator assuming that all moduli spaces kJ(α,β)\mathcal{M}_{k}^{J}(\alpha,\beta) with Fredholm index k1k\leq 1 are cut out transversely:

δ:CC(Y,λ,J)CC1(Y,λ,J),\delta:CC_{*}(Y,\lambda,J)\to CC_{*-1}(Y,\lambda,J),

given by

δα=β𝒫good(λ)u1J(α,β)/ϵ(u)d(u)β.\delta\alpha=\sum_{\beta\in\mathcal{P}_{\text{good}}(\lambda)}\sum_{u\in\mathcal{M}_{1}^{J}(\alpha,\beta)/{\mathbb{R}}}\frac{\epsilon(u)}{d(u)}\beta.

Here ϵ(u)\epsilon(u) is an element of {±1}\{\pm 1\} after generators of 𝒪α\mathcal{O}_{\alpha} and 𝒪β\mathcal{O}_{\beta} have been chosen, cf. [HN22, Def. A.26], and d(u)>0d(u)\in{\mathbb{Z}}_{>0} is the covering multiplicity of uu, which is 1 if and only if uu is somewhere injective.

Next we define an operator

κ:CC(Y,λ,J)CC(Y,λ,J)\kappa:CC_{*}(Y,\lambda,J)\to CC_{*}(Y,\lambda,J)

by

κ(α)=d(α)α.\kappa(\alpha)=d(\alpha)\alpha.

Under suitable transversality assumptions for 2J(α,β)\mathcal{M}_{2}^{J}(\alpha,\beta), then counting their ends yields

δκδ=0.\delta\kappa\delta=0. (1.1)

This was proven in the dynamically convex case in [HN16] and recovered in arbitrary odd dimensions in the absence of contractible Reeb orbits in [HN22]. As a result of (1.1), we obtain that

:=δκ\partial:=\delta\kappa

is a differential on CC(Y,λ,J)CC_{*}(Y,\lambda,J). The differential preserves the free homotopy class of Reeb orbits because they count cylinders which project to homotopies in YY between Reeb orbits.

Under additional hypotheses, this homology is independent of contact form λ\lambda defining ξ\xi and generic JJ (for example, if λ\lambda admits no contractible Reeb orbits, [HN22, Corollary 1.10]), and is denoted CH(Y,ξ)CH_{*}(Y,\xi). This is the cylindrical contact homology of (Y,ξ)(Y,\xi). Upcoming work of Hutchings and Nelson will show that CH(Y,ξ)=CH(Y,λ,J)CH_{*}(Y,\xi)=CH_{*}(Y,\lambda,J) is independent of dynamically convex λ\lambda and generic JJ.

1.2 Main result and connections to other work

The link of the AnA_{n} singularity is shown to be contactomorphic to the lens space L(n+1,n)L(n+1,n) in [AHNS17, Theorem 1.8]. More generally, the links of simple singularities (L,ξL)(L,\xi_{L}) are shown to be contactomorphic to quotients (S3/G,ξG)(S^{3}/G,\xi_{G}) in [Ne1, Theorem 5.3]. Theorem 1.2 computes the cylindrical contact homology of (S3/G,ξG)(S^{3}/G,\xi_{G}) as a direct limit of filtered homology groups.

Theorem 1.2.

Let GSU(2)G\subset\mbox{\em SU}(2) be a finite nontrivial group, and let m=|Conj(G)|m=|\mbox{\em Conj}(G)|\in{\mathbb{N}} be the number of conjugacy classes of GG. The cylindrical contact homology of (S3/G,ξG)(S^{3}/G,\xi_{G}) is

CH(S3/G,λG,J):=limNCHLN(S3/G,λN,JN)i0m2[2i]i0H(S2;)[2i].CH_{*}(S^{3}/G,\lambda_{G},J):=\varinjlim_{N}CH_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N})\cong\bigoplus_{i\geq 0}{\mathbb{Q}}^{m-2}[2i]\oplus\bigoplus_{i\geq 0}H_{*}(S^{2};{\mathbb{Q}})[2i].

The directed system of filtered cylindrical contact homology groups CHLN(S3/G,λN,JN)CH_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N}) is described in Section 1.3. Upcoming work of Hutchings and Nelson will show that this direct limit is an invariant of (S3/G,ξG)(S^{3}/G,\xi_{G}), in the sense that it is isomorphic to CH(S3/G,λ,J)CH_{*}(S^{3}/G,\lambda,J) where λ\lambda is any dynamically convex contact form on S3/GS^{3}/G with kernel ξG\xi_{G}, and J𝒥(λ)J\in\mathcal{J}(\lambda) is generic.

The brackets in Theorem 1.2 describe the degree of the grading.222For example, 8[5]H(S2;)[3]{\mathbb{Q}}^{8}[5]\oplus H_{*}(S^{2};{\mathbb{Q}})[3] is a ten dimensional space with nine dimensions in degree 5, and one dimension in degree 3. By the classification of finite subgroups GG of SU(2)\text{SU}(2), the following enumerates the possible values of m=|Conj(G)|m=|\text{Conj}(G)|:

  1. (i)

    If GG is cyclic of order nn, then m=nm=n.

  2. (ii)

    If GG is binary dihedral, G𝔻2nG\cong{\mathbb{D}}^{*}_{2n} for some nn, then m=n+3m=n+3.

  3. (iii)

    If GG is binary polyhedral, G𝕋,𝕆G\cong{\mathbb{T}}^{*},{\mathbb{O}}^{*}, or 𝕀{\mathbb{I}}^{*}, then m=7,8m=7,8, or 99, respectively.

Remark 1.3.

The cylindrical contact homology in Theorem 1.2 recovers the presentation of the manifold S3/GS^{3}/G as a Seifert fiber space, whose S1S^{1}-action agrees with the Reeb flow of a contact form defining ξG\xi_{G}. Viewing the manifold S3/GS^{3}/G as an S1S^{1}-bundle over an orbifold surface333Namely S2/HS^{2}/H where H=P(G)SO(3)H=P(G)\subset\text{SO}(3) and P:SU(2)SO(3)P:\text{SU}(2)\to\text{SO}(3), cf. Section 1.3. homeomorphic to S2S^{2}, the copies of H(S2;)H_{*}(S^{2};{\mathbb{Q}}) appearing in Theorem 1.2 may be understood as the orbifold Morse homology of this base. Each orbifold point pp with isotropy order npn_{p} corresponds to an exceptional fiber, γp\gamma_{p}, in S3/GS^{3}/G, which may be realized as an embedded Reeb orbit. The generators of the m2[0]{\mathbb{Q}}^{m-2}[0] term are the iterates γpk\gamma_{p}^{k} for k=1,2,,np1k=1,2,\dots,n_{p}-1 so that the dimension m2=p(np1)m-2=\sum_{p}(n_{p}-1) of this summand can be regarded as a kind of total isotropy of the base.

Remark 1.4.

Theorem 1.2 can alternatively be expressed as

CH(S3/G,λG,J){m1=0,m2and even0else.CH_{*}(S^{3}/G,\lambda_{G},J)\cong\begin{cases}{\mathbb{Q}}^{m-1}&*=0,\\ {\mathbb{Q}}^{m}&*\geq 2\,\,\mbox{and even}\\ 0&\mbox{else.}\end{cases}

In this form, we realize the expected isomorphism [BO17] between cylindrical contact homology and the positive S1S^{1}-equivariant symplectic cohomology with coefficients in {\mathbb{Q}} of the crepant resolutions YY of the singularities 2/G{\mathbb{C}}^{2}/G, as computed by McLean and Ritter. Their work shows that these groups with {\mathbb{Q}}-coefficients are free [u]{\mathbb{Q}}[u]-modules of rank equal to m=|Conj(G)|m=|\text{Conj}(G)|, where GSU(n)G\subset\text{SU}(n) and uu has degree 2 [MR, Corollary 2.13].

Remark 1.5.

Recent work of Haney and Mark computes the cylindrical contact homology in [HM22] of a family of hyperbolic Brieskorn manifolds Σ(p,q,r)\Sigma(p,q,r), for pp, qq, rr relatively prime positive integers satisfying 1p+1q+1r<1\frac{1}{p}+\frac{1}{q}+\frac{1}{r}<1, using methods from [Ne20]. Their work uses a family of hypertight contact forms, whose Reeb orbits are non-contractible. These manifolds are also Seifert fiber spaces, whose cylindrical contact homology features summands arising from copies of the homology of the orbit space, as well as summands from the total isotropy of the orbifold.

1.3 Structure of proof of main theorem

We now outline the proof of Theorem 1.2. Section 2 explains the process of perturbing a degenerate contact form λG\lambda_{G} on S3/GS^{3}/G using an orbifold Morse function. Given a finite, nontrivial subgroup GSU(2)G\subset\text{SU}(2), HH denotes the image of GG under the double cover of Lie groups P:SU(2)Spin(3)SO(3)P:\text{SU}(2)\cong\mbox{Spin}(3)\to\text{SO}(3). By Lemma 2.1, the quotient by the S1S^{1}-action on the Seifert fiber space S3/GS^{3}/G may be identified with a map 𝔭:S3/GS2/H\mathfrak{p}:S^{3}/G\to S^{2}/H. This 𝔭\mathfrak{p} fits into a commuting square of topological spaces (2.11) involving the Hopf fibration 𝔓:S3S2\mathfrak{P}:S^{3}\to S^{2}.

An HH-invariant Morse-Smale function on (S2,ωFS(,j))(S^{2},\omega_{\text{FS}}(\cdot,j\cdot)), constructed in Section 2.3, descends to an orbifold Morse function, fHf_{H}, on S2/HS^{2}/H. Here, ωFS\omega_{\text{FS}} is the Fubini-Study form on S2P1S^{2}\cong{\mathbb{C}}P^{1}, and jj is the standard integrable complex structure. By Lemma 2.4, the Reeb vector field of the perturbed contact form on S3/GS^{3}/G

λG,ε:=(1+ε𝔭fH)λG\lambda_{G,\varepsilon}:=(1+\varepsilon\mathfrak{p}^{*}f_{H})\lambda_{G}

is the descent of the vector field

Rλε:=Rλ1+ε𝔓fεXf~(1+ε𝔓f)2R_{\lambda_{\varepsilon}}:=\frac{R_{\lambda}}{1+\varepsilon\mathfrak{P}^{*}f}-\varepsilon\frac{\widetilde{X_{f}}}{(1+\varepsilon\mathfrak{P}^{*}f)^{2}}

to S3/GS^{3}/G. Here, Xf~\widetilde{X_{f}} is a horizontal lift to S3S^{3} of the Hamiltonian vector field XfX_{f} of ff on S2S^{2}, computed with respect to ωFS\omega_{\text{FS}}, and we use the convention that ιXfωFS=df\iota_{X_{f}}\omega_{\text{FS}}=-df. Thus, the Xf~\widetilde{X_{f}} term vanishes along exceptional fibers γp\gamma_{p} of S3/GS^{3}/G projecting to orbifold critical points pS2/Hp\in S^{2}/H of fHf_{H}, implying that these parametrized circles and their iterates γpk\gamma_{p}^{k} are Reeb orbits of λG,ε\lambda_{G,\varepsilon}. Lemma 2.15 computes the Conley-Zehnder index μCZτ(γpk)\mu_{\operatorname{CZ}}^{\tau}(\gamma_{p}^{k}) in terms of kk and the Morse index of fHf_{H} at pp with respect to a global unitary trivialization.

Next we outline our procedure of taking direct limits in Section 4 of action filtered cylindrical contact homology in Section 3. Given a contact manifold (Y,λ)(Y,\lambda), the action of a Reeb orbit γ:/TY\gamma:{\mathbb{R}}/T{\mathbb{Z}}\to Y is the positive quantity

𝒜(γ):=γλ=T.\mathcal{A}(\gamma):=\int_{\gamma}\lambda=T.

For L>0L>0, we let 𝒫L(λ)𝒫(λ)\mathcal{P}^{L}(\lambda)\subset\mathcal{P}(\lambda) denote the set of orbits γ\gamma with 𝒜(γ)<L\mathcal{A}(\gamma)<L. A contact form λ\lambda is LL-nondegenerate when all γ𝒫L(λ)\gamma\in\mathcal{P}^{L}(\lambda) are nondegenerate. If c1(ξ),π2(Y)=0\langle c_{1}(\xi),\pi_{2}(Y)\rangle=0 and μCZ(γ)3\mu_{\operatorname{CZ}}(\gamma)\geq 3 for all contractible γ𝒫L(λ)\gamma\in\mathcal{P}^{L}(\lambda), we say that the LL-nondegenerate contact form λ\lambda is LL-dynamically convex. By Lemma 2.15, given L>0L>0, all γ𝒫L(λG,ε)\gamma\in\mathcal{P}^{L}(\lambda_{G,\varepsilon}) are nondegenerate and project to critical points of fHf_{H} under 𝔭\mathfrak{p}, when ε\varepsilon is sufficiently small.

This lemma allows for the computation in Section 3 of the action filtered cylindrical contact homology. After fixing L>0L>0, \partial restricts to a differential, L\partial^{L}, on the subcomplex generated by γ𝒫goodL(λ)\gamma\in\mathcal{P}_{\text{good}}^{L}(\lambda), denoted CCL(Y,λ,J)CC_{*}^{L}(Y,\lambda,J), whose homology is denoted CHL(Y,λ,J)CH_{*}^{L}(Y,\lambda,J). This is because the differential decreases action: if 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}) then J(γ+,γ)\mathcal{M}^{J}(\gamma_{+},\gamma_{-}) is empty because by Stokes’ theorem, action decreases along holomorphic cylinders in a symplectization.

In Section 3, we use Lemma 2.15 to produce a sequence (LN,λN,JN)N=1(L_{N},\lambda_{N},J_{N})_{N=1}^{\infty}, where LNL_{N}\nearrow\infty in {\mathbb{R}}, λN\lambda_{N} is an LNL_{N}-dynamically convex contact form for ξG\xi_{G}, and JN𝒥(λN)J_{N}\in\mathcal{J}(\lambda_{N}) is generic. By Lemmas 3.1 and 3.3, every orbit γ𝒫goodLN(λN)\gamma\in\mathcal{P}^{L_{N}}_{\text{good}}(\lambda_{N}) is of even degree, and so LN=0\partial^{L_{N}}=0, providing

CHLN(S3/G,λN,JN)𝒫goodLN(λN)i=02N1m2[2i]i=02N2H(S2;)[2i].CH_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N})\cong{\mathbb{Q}}\langle\,\mathcal{P}_{\text{good}}^{L_{N}}(\lambda_{N})\,\rangle\cong\bigoplus_{i=0}^{2N-1}{\mathbb{Q}}^{m-2}[2i]\oplus\bigoplus_{i=0}^{2N-2}H_{*}(S^{2};{\mathbb{Q}})[2i]. (1.2)

Finally, we prove Theorem 4.4 in Section 4, which states that a completed symplectic cobordism (X,λ,J)(X,\lambda,J) from (λN,JN)(\lambda_{N},J_{N}) to (λM,JM)(\lambda_{M},J_{M}), for NMN\leq M, induces a homomorphism,

Ψ:CHLN(S3/G,λN,JN)CHLM(S3/G,λM,JM)\Psi:CH_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N})\to CH_{*}^{L_{M}}(S^{3}/G,\lambda_{M},J_{M})

which takes the form of the standard inclusion when making the identification (1.2). The proof of Theorem 4.4 comes in two steps. First, the moduli spaces 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) are finite by Proposition 4.7 and Corollary 4.8, implying that the map Ψ\Psi is well defined. Second, the identification of Ψ\Psi with a standard inclusion is made precise in the following manner. Given γ+𝒫goodLN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}_{\text{good}}(\lambda_{N}), there is a unique γ𝒫goodLM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}_{\text{good}}(\lambda_{M}) which

  1. (i)

    projects to the same critical point of fHf_{H} as γ+\gamma_{+} under 𝔭\mathfrak{p},

  2. (ii)

    satisfies m(γ+)=m(γ)m(\gamma_{+})=m(\gamma_{-}).

When (i) and (ii) hold, we write γ+γ\gamma_{+}\sim\gamma_{-}. We argue in Section 4 that Ψ\Psi takes the form Ψ([γ+])=[γ]\Psi([\gamma_{+}])=[\gamma_{-}], when γ+γ\gamma_{+}\sim\gamma_{-}.

Theorem 4.4 now implies that the system of filtered contact homology groups is identified with a sequence of inclusions of vector spaces, providing isomorphic direct limits:

limNCHLN(S3/G,λN,JN)i0m2[2i]i0H(S2;).\varinjlim_{N}CH_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N})\cong\bigoplus_{i\geq 0}{\mathbb{Q}}^{m-2}[2i]\oplus\bigoplus_{i\geq 0}H_{*}(S^{2};{\mathbb{Q}}).

1.4 Connections to orbifold Morse homology

Using the construction of the orbifold Morse-Smale-Witten complex as in Cho and Hong in [CH14] we can draw the following parallels between orbifold Morse homology and cylindrical contact homology. Given an orbifold Morse function ff on an orbifold XX, the chain group CM(X,f)CM_{*}(X,f) is generated by the orientable critical points of ff. The differential, M\partial^{M}, is given as a weighted count of the negative gradient flow lines between orientable critical points in XX. There are two notable similarities between the chain complexes (CC,)(CC_{*},\partial) and (CM,M)(CM_{*},\partial^{M}), exemplified by our computations.

(1) Bad Reeb orbits are analogous to non-orientable critical points.
Bad Reeb orbits are excluded as generators of CCCC_{*} for the same reasons that non-orientable critical points are excluded as generators of CMCM_{*}. A critical point pp of ff on an orbifold is non-orientable if the action of its isotropy group Γp\Gamma_{p} on a choice of unstable manifold is not orientation preserving. Analogously, a Reeb orbit γ\gamma is bad if the action of its cyclic deck group Δγ\Delta_{\gamma} on an asymptotic operator is not orientation preserving.

Our Seifert projections 𝔭:S3/GS2/H\mathfrak{p}:S^{3}/G\to S^{2}/H geometrically realize this analogy: if γ\gamma is a bad Reeb orbit associated to (1+ε𝔭fH)λG(1+\varepsilon\mathfrak{p}^{*}f_{H})\lambda_{G} in S3/GS^{3}/G that projects to orbifold critical point pp of fHf_{H}, then pp is non-orientable. Conversely, if pS2/Hp\in S^{2}/H is a non-orientable critical point of fHf_{H}, then there is a bad Reeb orbit γ\gamma associated to (1+ε𝔭fH)λG(1+\varepsilon\mathfrak{p}^{*}f_{H})\lambda_{G} in S3/GS^{3}/G projecting to pp. This interplay can be realized through the pairs (γ,p)=(h2,ph)(\gamma,p)=(h^{2},p_{h}) for the binary dihedral group in Sections 3.2 and (γ,p)=(2,𝔢)(\gamma,p)=(\mathcal{E}^{2},\mathfrak{e}) for the binary polyhedral group in Section 3.3.

(2) The differentials are structurally identical.
Take good Reeb orbits α\alpha and β\beta with μCZ(α)μCZ(β)=1\mu_{\operatorname{CZ}}(\alpha)-\mu_{\operatorname{CZ}}(\beta)=1, and take orientable critical points pp and qq of orbifold Morse ff with indf(p)indf(q)=1\text{ind}_{f}(p)-\text{ind}_{f}(q)=1. Now compare

α,β=u1J(α,β)/ϵ(u)d(α)d(u),Mp,q=x(p,q)/ϵ(x)|Γp||Γx|.\langle\partial\alpha,\beta\rangle=\sum_{u\in\mathcal{M}_{1}^{J}(\alpha,\beta)/{\mathbb{R}}}\dfrac{\epsilon(u)d(\alpha)}{d(u)},\,\,\,\,\,\,\,\,\,\langle\partial^{M}p,q\rangle=\sum_{x\in\mathcal{M}(p,q)/{\mathbb{R}}}\dfrac{\epsilon(x)|\Gamma_{p}|}{|\Gamma_{x}|}.

Here, (p,q)\mathcal{M}(p,q) is the space of negative gradient paths xx from pp to qq, and Γx\Gamma_{x} is the local isotropy group at any point on the path xx, whose order divides |Γp||\Gamma_{p}|. Both ϵ{±1}\epsilon\in\{\pm 1\} quantities come from choices of orientations in each setting and are well-defined because α\alpha and β\beta are good, and because pp and qq are orientable.

The similarities of both boundary operators as weighted counts of moduli spaces reflect the parallels between the breaking and gluing of the two theories: a single broken gradient path or building may serve as the limit of multiple ends of a 1-dimensional moduli space in either setting. For a thorough treatment of why these signed counts generally produce a differential that squares to zero, see [CH14, Theroem 5.1] (in the orbifold case) and [HN16, §4.3] (in the contact case).

In Sections 2.4 and 2.5, we explain the analogies between the contact data of S3/GS^{3}/G and the orbifold Morse data of S2/HS^{2}/H in further detail.

Acknowledgements

Leo Digiosia thanks his advisor, Jo Nelson, for her exceptional guidance and discussions. Leo Digiosia was supported by NSF grants DMS-1745670, DMS-1840723, and DMS-2104411. Jo Nelson is supported by NSF grants DMS-2104411 and CAREER DMS-2142694. We thank the referee for their thoughtful reading and helpful comments on this paper, especially with respect to the cobordism maps in Section 4. (Section 4 also benefited from a comment from Chris Wendl and his book [We20].)

2 Geometric setup and dynamics

In this section we first review the process of perturbing degenerate contact forms on S3S^{3} and S3/GS^{3}/G using a Morse function to achieve nondegeneracy up to an action threshold, following [Ne20, §1.5]. We then identify the associated Reeb orbits of S3/GS^{3}/G and compute their Conley Zehnder indices in Lemma 2.15. In Section 2.3 we construct the HH-invariant Morse functions we use to perturb the contact forms on S3/GS^{3}/G. In Sections 2.4 and 2.5 we elucidate how the Reeb dynamics realize the Morse orbifold data associated to S2/HS^{2}/H.

2.1 Spherical geometry and associated Reeb dynamics

The diffeomorphism between S32S^{3}\subset{\mathbb{C}}^{2} and SU(2)\text{SU}(2) provides S3S^{3} with the structure of a Lie group:

(α,β)S3(αβ¯βα¯)SU(2)(\alpha,\beta)\in S^{3}\mapsto\begin{pmatrix}\alpha&-\overline{\beta}\,\,\\ \beta&\overline{\alpha}\end{pmatrix}\in\text{SU}(2) (2.1)

and we see that e=(1,0)S3e=(1,0)\in S^{3} is the identity element. The round contact form on S3S^{3}, denoted λ\lambda, is defined as the restriction of the 1-form ιVω0Ω1(2)\iota_{V}\omega_{0}\in\Omega^{1}({\mathbb{C}}^{2}) to S3S^{3}, where

ω0:=i2k=12dzkdzk¯andV:=12k=12zkzk+zk¯zk¯,ιVω0=i4k=12zkdzk¯zk¯dzk.\omega_{0}:=\frac{i}{2}\sum_{k=1}^{2}dz_{k}\wedge d\overline{z_{k}}\,\,\,\,\,\text{and}\,\,\,\,\,V:=\frac{1}{2}\sum_{k=1}^{2}z_{k}\partial_{z_{k}}+\overline{z_{k}}\partial_{\overline{z_{k}}},\,\,\implies\,\,\,\iota_{V}\omega_{0}=\frac{i}{4}\sum_{k=1}^{2}z_{k}\wedge d\overline{z_{k}}-\overline{z_{k}}dz_{k}.

The SU(2)\text{SU}(2)-action on 2{\mathbb{C}}^{2} preserves ω0\omega_{0} and VV, and so the SU(2)\text{SU}(2)-action on S3S^{3} preserves λ\lambda.

There is a natural Lie algebra isomorphism between the tangent space of the identity element of a Lie group and its collection of left-invariant vector fields. The contact plane ξe=ker(λe)\xi_{e}=\text{ker}(\lambda_{e}) at the identity element e=(1,0)S3e=(1,0)\in S^{3} is spanned by the tangent vectors x2|e\partial_{x_{2}}|_{e} and y2|e\partial_{y_{2}}|_{e}, where we are viewing

ξeTe2Te4=Span{xi|e,yi|e}.\xi_{e}\subset T_{e}{\mathbb{C}}^{2}\cong T_{e}{\mathbb{R}}^{4}=\text{Span}_{{\mathbb{R}}}\{\partial_{x_{i}}|_{e},\partial_{y_{i}}|_{e}\}.

Let V1V_{1} and V2V_{2} be the left-invariant vector fields corresponding to x2|e=0,0,1,0\partial_{x_{2}}|_{e}=\langle 0,0,1,0\rangle and y2|e=0,0,0,1\partial_{y_{2}}|_{e}=\langle 0,0,0,1\rangle, respectively. Because S3S^{3} acts on itself by contactomorphisms, V1V_{1} and V2V_{2} are sections of ξ\xi and provide a global unitary trivialization of (ξ,dλ|ξ,J2)(\xi,d\lambda|_{\xi},J_{{\mathbb{C}}^{2}}), denoted τ\tau:

τ:S3×2ξ,(p,η1,η2)η1V1(p)+η2V2(p)ξp.\tau:S^{3}\times{\mathbb{R}}^{2}\to\xi,\,\,\,(p,\eta_{1},\eta_{2})\mapsto\eta_{1}V_{1}(p)+\eta_{2}V_{2}(p)\in\xi_{p}. (2.2)

Here, J2J_{{\mathbb{C}}^{2}} is the standard integrable complex structure on 2{\mathbb{C}}^{2}. Note that J2(V1)=V2J_{{\mathbb{C}}^{2}}(V_{1})=V_{2} everywhere. Given a Reeb orbit γ\gamma of any contact form on S3S^{3}, let the symbol μCZ(γ)\mu_{\operatorname{CZ}}(\gamma) denote the Conley Zehnder index of γ\gamma with respect to this τ\tau. If (α,β)S3(\alpha,\beta)\in S^{3}, write α=a+ib\alpha=a+ib and β=c+id\beta=c+id. Then, with respect to the ordered basis (x1,y1,x2,y2)(\partial_{x_{1}},\partial_{y_{1}},\partial_{x_{2}},\partial_{y_{2}}) of T(α,β)4T_{(\alpha,\beta)}{\mathbb{R}}^{4}, we have the following expressions

V1(α,β)=c,d,a,b,V2(α,β)=d,c,b,a.V_{1}(\alpha,\beta)=\langle-c,d,a,-b\rangle,\,\,\,\,V_{2}(\alpha,\beta)=\langle-d,-c,b,a\rangle. (2.3)

Consider the double cover of Lie groups, P:SU(2)SO(3)P:\text{SU}(2)\to\text{SO}(3). The kernel of PP has order 2 and is generated by IdSU(2)-\text{Id}\in\text{SU}(2), the only element of SU(2)\text{SU}(2) of order 2.

(αβ¯βα¯)SU(2)𝑃(|α|2|β|22Im(αβ)2Re(αβ)2Im(α¯β)Re(α2+β2)Im(α2+β2)2Re(α¯β)Im(α2β2)Re(α2β2))SO(3)\begin{pmatrix}\alpha&-\overline{\beta}\,\,\\ \beta&\overline{\alpha}\end{pmatrix}\in\text{SU}(2)\xmapsto{P}\begin{pmatrix}|\alpha|^{2}-|\beta|^{2}&2\text{Im}(\alpha\beta)&2\text{Re}(\alpha\beta)\\ -2\text{Im}(\overline{\alpha}\beta)&\text{Re}(\alpha^{2}+\beta^{2})&-\text{Im}(\alpha^{2}+\beta^{2})\\ -2\text{Re}(\overline{\alpha}\beta)&\text{Im}(\alpha^{2}-\beta^{2})&\text{Re}(\alpha^{2}-\beta^{2})\end{pmatrix}\in\text{SO}(3) (2.4)

A diffeomorphism P1S23{\mathbb{C}}P^{1}\to S^{2}\subset{\mathbb{R}}^{3} is given in homogeneous coordinates (|α|2+|β|2=1|\alpha|^{2}+|\beta|^{2}=1) by

(α:β)P1(|α|2|β|2,2Im(α¯β),2Re(α¯β))S2.(\alpha:\beta)\in{\mathbb{C}}P^{1}\mapsto(|\alpha|^{2}-|\beta|^{2},-2\text{Im}(\bar{\alpha}\beta),-2\text{Re}(\bar{\alpha}\beta))\in S^{2}. (2.5)

We have an SO(3)\text{SO}(3)-action on P1{\mathbb{C}}P^{1}, pulled back from the SO(3)\text{SO}(3)-action on S2S^{2} by (2.5). Lemma 2.1 illustrates how the action of SU(2)\text{SU}(2) on S3S^{3} is related to the action of SO(3)\text{SO}(3) on P1S2{\mathbb{C}}P^{1}\cong S^{2} via P:SU(2)SO(3).P:\text{SU}(2)\to\text{SO}(3).

Lemma 2.1.

For a point zz in S3S^{3}, let [z]P1[z]\in{\mathbb{C}}P^{1} denote the corresponding point under the quotient of the S1S^{1}-action on S3S^{3}. Then for all zS3z\in S^{3}, and all matrices ASU(2)A\in\mbox{\em SU}(2), we have

[Az]=P(A)[z]P1S2.[A\cdot z]=P(A)\cdot[z]\in{\mathbb{C}}P^{1}\cong S^{2}.
Proof.

First, note that the result holds for the case z=e=(1,0)S3z=e=(1,0)\in S^{3}. This is because [e]P1[e]\in{\mathbb{C}}P^{1} corresponds to (1,0,0)S2(1,0,0)\in S^{2} under (2.5), and so for any AA, P(A)[e]P(A)\cdot[e] is the first column of the 3×33\times 3 matrix P(A)P(A) appearing in (2.4). That is,

P(A)[e]=(|α|2|β|2,2Im(α¯β),2Re(α¯β)),P(A)\cdot[e]=(|\alpha|^{2}-|\beta|^{2},-2\text{Im}(\bar{\alpha}\beta),-2\text{Re}(\bar{\alpha}\beta)), (2.6)

(where (α,β)S3(\alpha,\beta)\in S^{3} is the unique element corresponding to ASU(2)A\in\text{SU}(2), using (2.1)). By (2.5), the point (2.6) equals [(α,β)]=[(α,β)(1,0)]=[Ae][(\alpha,\beta)]=[(\alpha,\beta)\cdot(1,0)]=[A\cdot e], and so the result holds when z=ez=e.

For the general case, note that any zS3z\in S^{3} equals BeB\cdot e for some BSU(2)B\in\text{SU}(2), and use the fact that PP is a group homomorphism. ∎

The Reeb flow of λ\lambda is given by the S1S^{1}\subset{\mathbb{C}}^{*} (Hopf) action, peitpp\mapsto e^{it}\cdot p. Thus, all γ𝒫(λ)\gamma\in\mathcal{P}(\lambda) have period 2kπ2k\pi, with linearized return maps equal to Id:ξγ(0)ξγ(0)\text{Id}:\xi_{\gamma(0)}\to\xi_{\gamma(0)}, and are degenerate.

Notation 2.2.

Following the general recipe of perturbing the degenerate contact form on a prequantization bundle, outlined in [Ne20, §1.5], we establish the following notation:

  • 𝔓:S3S2\mathfrak{P}:S^{3}\to S^{2} is the Hopf fibration.

  • ff is a Morse-Smale function on (S2,ωFS(,j))(S^{2},\omega_{\text{FS}}(\cdot,j\cdot)) and Crit(f)\text{Crit}(f) is its set of critical points,

  • For ε>0\varepsilon>0,

    • fε:=1+εf:S2f_{\varepsilon}:=1+\varepsilon f:S^{2}\to{\mathbb{R}},

    • Fε:=fε𝔓:S3F_{\varepsilon}:=f_{\varepsilon}\circ\mathfrak{P}:S^{3}\to{\mathbb{R}},

    • λε:=FελΩ1(S3)\lambda_{\varepsilon}:=F_{\varepsilon}\lambda\in\Omega^{1}(S^{3}).

  • Xf~ξ\widetilde{X_{f}}\in\xi is the horizontal lift of XfTS2X_{f}\in TS^{2} using the fiberwise linear symplectomorphism d𝔓|ξ:(ξ,dλ|ξ)(TS2,ωFS)d\mathfrak{P}|_{\xi}:(\xi,d\lambda|_{\xi})\to(TS^{2},\omega_{\text{FS}}), where XfX_{f} denotes the Hamiltonian vector field of ff on S2S^{2} with respect to ωFS\omega_{\text{FS}}.

Remark 2.3.

Our convention is that for a smooth, real valued function ff on symplectic manifold (M,ω)(M,\omega), the Hamiltonian vector field XfX_{f} uniquely satisfies ιXfω=df\iota_{X_{f}}\omega=-df.

For small ε\varepsilon, ker λε=kerλ\lambda_{\varepsilon}=\mbox{ker}\lambda. We refer to λε\lambda_{\varepsilon} as the perturbed contact form on S3S^{3}. Although λ\lambda and λε\lambda_{\varepsilon} define the same contact structure, their Reeb dynamics differ.

Lemma 2.4.

The following relationship between vector fields on S3S^{3} holds:

Rλε=RλFεεXf~Fε2.R_{\lambda_{\varepsilon}}=\frac{R_{\lambda}}{F_{\varepsilon}}-\varepsilon\frac{\widetilde{X_{f}}}{F_{\varepsilon}^{2}}.
Proof.

This is [Ne20, Prop. 4.10]. Note that the sign discrepancy is a result of our convention regarding Hamiltonian vector fields, see Remark 2.3. ∎

We now explore how the relationship between vector fields from Lemma 2.4 provides a relationship between Reeb and Hamiltonian flows.

Notation 2.5.

(Reeb and Hamiltonian flows). For any tt\in{\mathbb{R}},

  • ϕ0t:S3S3\phi_{0}^{t}:S^{3}\to S^{3} denotes the time tt flow of the unperturbed Reeb vector field RλR_{\lambda},

  • ϕt:S3S3\phi^{t}:S^{3}\to S^{3} denotes the time tt flow of the perturbed Reeb vector field RλεR_{\lambda_{\varepsilon}},

  • φt:S2S2\varphi^{t}:S^{2}\to S^{2} denotes the time tt flow of the vector field V:=εXffε2.V:=-\frac{\varepsilon X_{f}}{f_{\varepsilon}^{2}}.

Lemma 2.6.

For all tt values, we have 𝔓ϕt=φt𝔓\mathfrak{P}\circ\phi^{t}=\varphi^{t}\circ\mathfrak{P} as smooth maps S3S2S^{3}\to S^{2}.

Proof.

Pick zS3z\in S^{3} and let γ~:S3\widetilde{\gamma}:{\mathbb{R}}\to S^{3} denote the unique integral curve for RλεR_{\lambda_{\varepsilon}} which passes through zz at time t=0t=0, i.e., γ~(t)=ϕt(z)\widetilde{\gamma}(t)=\phi^{t}(z). By Lemma 2.4, d𝔓d\mathfrak{P} carries the derivative of γ~\widetilde{\gamma} precisely to the vector VTS2V\in TS^{2}. Thus, 𝔓γ~:S2\mathfrak{P}\circ\widetilde{\gamma}:{\mathbb{R}}\to S^{2} is the unique integral curve, γ\gamma, of VV passing through p:=𝔓(z)p:=\mathfrak{P}(z) at time t=0t=0, i.e., γ(t)=φt(p)\gamma(t)=\varphi^{t}(p). Combining these facts provides

𝔓(γ~(t))=γ(t)𝔓(ϕt(z))=φt(p)𝔓(ϕt(z))=φt(𝔓(z)).\mathfrak{P}(\widetilde{\gamma}(t))=\gamma(t)\,\,\implies\,\,\mathfrak{P}(\phi^{t}(z))=\varphi^{t}(p)\,\,\implies\mathfrak{P}(\phi^{t}(z))=\varphi^{t}(\mathfrak{P}(z)).

Lemma 2.7 describes the orbits γ𝒫(λε)\gamma\in\mathcal{P}(\lambda_{\varepsilon}) projecting to critical points of ff under 𝔓\mathfrak{P}.

Lemma 2.7.

Let pCrit(f)p\in\mbox{\em Crit}(f) and take z𝔓1(p)z\in\mathfrak{P}^{-1}(p). Then the map

γp:[0,2πfε(p)]S3,teitfε(p)z\gamma_{p}:[0,2\pi f_{\varepsilon}(p)]\to S^{3},\,\,\,\,t\mapsto e^{\frac{it}{f_{\varepsilon}(p)}}\cdot z

descends to a closed, embedded Reeb orbit γp:/2πfε(p)S3\gamma_{p}:{\mathbb{R}}/2\pi f_{\varepsilon}(p){\mathbb{Z}}\to S^{3} of λε\lambda_{\varepsilon}, passing through point zz in S3S^{3}, whose image under 𝔓\mathfrak{P} is {p}S2\{p\}\subset S^{2}, where \cdot denotes the S1S^{1}\subset{\mathbb{C}}^{*} action on S3S^{3}.

Proof.

The map /2πS3{\mathbb{R}}/2\pi{\mathbb{Z}}\to S^{3}, teitzt\mapsto e^{it}\cdot z is a closed, embedded integral curve for the degenerate Reeb field RλR_{\lambda}, and so by the chain rule we have that γp˙(t)=Rλ(γp(t))/fε(p)\dot{\gamma_{p}}(t)=R_{\lambda}(\gamma_{p}(t))/f_{\varepsilon}(p). Note that 𝔓(γ(t))=𝔓(eitfε(p)z)=𝔓(z)=p\mathfrak{P}(\gamma(t))=\mathfrak{P}(e^{\frac{it}{f_{\varepsilon}(p)}}\cdot z)=\mathfrak{P}(z)=p and, because Xf~(γ(t))\widetilde{X_{f}}(\gamma(t)) is a lift of Xf(p)=0X_{f}(p)=0, we have Xf~(γ(t))=0\widetilde{X_{f}}(\gamma(t))=0. By the description of RλεR_{\lambda_{\varepsilon}} in Lemma 2.4, we have γp˙(t)=Rλε(γp(t))\dot{\gamma_{p}}(t)=R_{\lambda_{\varepsilon}}(\gamma_{p}(t)). ∎

Next we set notation to be used in describing the local models for the linearized Reeb flow along the orbits γp\gamma_{p} from Lemma 2.7. For ss\in{\mathbb{R}}, (s)\mathcal{R}(s) denotes the 2×22\times 2 rotation matrix:

(s):=(cos((s))sin((s))sin((s))cos((s)))SO(2).\mathcal{R}(s):=\begin{pmatrix}\cos{(s)}&-\sin{(s)}\\ \sin{(s)}&\cos{(s)}\end{pmatrix}\in\text{SO}(2).

Note that J0=(π/2)J_{0}=\mathcal{R}(\pi/2). For pCrit(f)S2p\in\text{Crit}(f)\subset S^{2}, pick coordinates ψ:2S2\psi:{\mathbb{R}}^{2}\to S^{2}, so that ψ(0,0)=p\psi(0,0)=p. Then we let H(f,ψ)H(f,\psi) denote the Hessian of ff in these coordinates at pp.

Notation 2.8.

The term stereographic coordinates at pS2p\in S^{2} describes a smooth ψ:2S2\psi:{\mathbb{R}}^{2}\to S^{2} with ψ(0,0)=p\psi(0,0)=p, which has a factorization ψ=ψ1ψ0\psi=\psi_{1}\circ\psi_{0}, where ψ0:2S2\psi_{0}:{\mathbb{R}}^{2}\to S^{2} is the map

(x,y)11+x2+y2(2x,2y,1x2y2),(x,y)\mapsto\frac{1}{1+x^{2}+y^{2}}(2x,2y,1-x^{2}-y^{2}),

taking (0,0)(0,0) to (0,0,1)(0,0,1), and ψ1:S2S2\psi_{1}:S^{2}\to S^{2} is given by the action of some element of SO(3)\text{SO}(3) taking (0,0,1)(0,0,1) to pp. If ψ\psi and ψ\psi^{\prime} are both stereographic coordinates at pp, then they differ by a precomposition with some (s)\mathcal{R}(s) in SO(2)\text{SO}(2). Note that ψωFS=dxdy(1+x2+y2)2\psi^{*}\omega_{\text{FS}}=\frac{dx\wedge dy}{(1+x^{2}+y^{2})^{2}} ([MS15, Ex. 4.3.4]).

Lemma 2.9 describes the linearized Reeb flow of the unperturbed λ\lambda with respect to τ\tau.

Lemma 2.9.

For any zS3z\in S^{3}, the linearization dϕ0t|ξz:ξzξeitzd\phi_{0}^{t}|_{\xi_{z}}:\xi_{z}\to\xi_{e^{it}\cdot z} is represented by (2t)\mathcal{R}(2t), with respect to ordered bases (V1(z),V2(z))(V_{1}(z),V_{2}(z)) of ξz\xi_{z} and (V1(eitz),V2(eitz))(V_{1}(e^{it}\cdot z),V_{2}(e^{it}\cdot z)) of ξeitz\xi_{e^{it}\cdot z}.

Proof.

Since the ViV_{i} are SU(2)\text{SU}(2)-invariant and the SU(2)\text{SU}(2)-action commutes with ϕ0t\phi_{0}^{t}, we may reduce to the case z=e=(1,0)S3z=e=(1,0)\in S^{3}. That is, we must show for all tt values that

(dϕ0t)(1,0)V1(1,0)\displaystyle(d\phi_{0}^{t})_{(1,0)}V_{1}(1,0) =cos((2t))V1(eit,0)+sin((2t))V2(eit,0)\displaystyle=\cos{(2t)}V_{1}(e^{it},0)+\sin{(2t)}V_{2}(e^{it},0) (2.7)
(dϕ0t)(1,0)V2(1,0)\displaystyle(d\phi_{0}^{t})_{(1,0)}V_{2}(1,0) =sin((2t))V1(eit,0)+cos((2t))V2(eit,0).\displaystyle=-\sin{(2t)}V_{1}(e^{it},0)+\cos{(2t)}V_{2}(e^{it},0). (2.8)

Note that (2.8) follows from (2.7) by applying the endomorphism J2J_{{\mathbb{C}}^{2}} to both sides of (2.7), and noting both that dϕ0td\phi_{0}^{t} commutes with J2J_{{\mathbb{C}}^{2}}, and that J2(V1)=V2J_{{\mathbb{C}}^{2}}(V_{1})=V_{2}. We now prove (2.7).

The coordinate descriptions (2.3) tell us that V1(eit,0)V_{1}(e^{it},0) and V2(eit,0)V_{2}(e^{it},0) can be respectively written as 0,0,cos(t),sin(t)\langle 0,0,\cos{t},-\sin{t}\rangle and 0,0,sin(t),cos(t)\langle 0,0,\sin{t},\cos{t}\rangle. Angle sum formulas now imply

cos((2t))V1(eit,0)+sin((2t))V2(eit,0)=0,0,cos(t),sin(t).\cos{(2t)}V_{1}(e^{it},0)+\sin{(2t)}V_{2}(e^{it},0)=\langle 0,0,\cos{t},\sin{t}\rangle.

The vector on the right is precisely (dϕ0t)(1,0)V1(1,0)(d\phi_{0}^{t})_{(1,0)}V_{1}(1,0), and so we have proven (2.7). ∎

Proposition 2.10 and Corollary 2.11 conclude our discussion of dynamics on S3S^{3}.

Proposition 2.10.

Fix a critical point pp of ff in S2S^{2} and stereographic coordinates ψ:2S2\psi:{\mathbb{R}}^{2}\to S^{2} at pp, and suppose γ𝒫(λε)\gamma\in\mathcal{P}(\lambda_{\varepsilon}) projects to pp under 𝔓\mathfrak{P}. Let MtSp(2)M_{t}\in\mbox{\em Sp}(2) denote dϕt|ξγ(0):ξγ(0)ξγ(t)d\phi^{t}|_{\xi_{\gamma(0)}}:\xi_{\gamma(0)}\to\xi_{\gamma(t)} with respect to the trivialization τ\tau. Then MtM_{t} is a conjugate of the matrix

(2tfε(p))exp(tεfε(p)2J0H(f,ψ)missing)\mathcal{R}\bigg{(}\frac{2t}{f_{\varepsilon}(p)}\bigg{)}\cdot\exp\bigg(\frac{-t\varepsilon}{f_{\varepsilon}(p)^{2}}J_{0}\cdot H(f,\psi)\bigg{missing})

by some element of SO(2)\mbox{\em SO}(2), which is independent of tt.

Proof.

Let z:=γ(0)z:=\gamma(0). We linearize the identity 𝔓ϕt=φt𝔓\mathfrak{P}\circ\phi^{t}=\varphi^{t}\circ\mathfrak{P} from Lemma 2.6, restrict to ξz\xi_{z}, and rearrange to recover the equality dϕt|ξz=abc:ξzξϕt(z)d\phi^{t}|_{\xi_{z}}=a\circ b\circ c:\xi_{z}\to\xi_{\phi^{t}(z)}, where

a=(d𝔓|ξϕt(z))1:TpS2ξϕt(z),b=dφpt:TpS2TpS2,andc=d𝔓|ξz:ξzTpS2.a=\left(d\mathfrak{P}|_{\xi_{\phi^{t}(z)}}\right)^{-1}:T_{p}S^{2}\to\xi_{\phi^{t}(z)},\,\,b=d\varphi^{t}_{p}:T_{p}S^{2}\to T_{p}S^{2},\,\,\text{and}\,\,c=d\mathfrak{P}|_{\xi_{z}}:\xi_{z}\to T_{p}S^{2}.

Let vi:=d𝔓(Vi(z))TpS2v_{i}:=d\mathfrak{P}(V_{i}(z))\in T_{p}S^{2}, then (v1,v2)(v_{1},v_{2}) and (V1,V2)(V_{1},V_{2}) provide an oriented basis of each of the three vector spaces appearing in the above composition of linear maps. Let AA, BB, and CC denote the matrix representations of aa, bb, and cc with respect to these ordered bases. We have Mt=ABCM_{t}=A\cdot B\cdot C. Note that C=IdC=\text{Id}. We compute AA and BB:

To compute AA, recall that ϕ0t:S3S3\phi^{t}_{0}:S^{3}\to S^{3} denotes the time tt flow of the unperturbed Reeb field (alternatively, the Hopf action). Linearize the equality 𝔓ϕ0t=𝔓\mathfrak{P}\circ\phi^{t}_{0}=\mathfrak{P}, then use ϕ0t(z)=ϕt/fε(p)(z)\phi_{0}^{t}(z)=\phi^{t/f_{\varepsilon}(p)}(z) from Lemma 2.7 and Lemma 2.9 to find that

A=(2t/fε(p)).A=\mathcal{R}(2t/f_{\varepsilon}(p)). (2.9)

To compute BB, note that φt\varphi^{t} is the Hamiltonian flow of the function 1/fε1/f_{\varepsilon} with respect to ωFS\omega_{\text{FS}}. It is advantageous to study this flow using our stereographic coordinates: recall that ψ\psi defines a symplectomorphism (2,dxdy(1+x2+y2)2)(S2{p},ωFS)\left({\mathbb{R}}^{2},\frac{dx\wedge dy}{(1+x^{2}+y^{2})^{2}}\right)\to(S^{2}\setminus\{p^{\prime}\},\omega_{\text{FS}}), where pS2p^{\prime}\in S^{2} is antipodal to pp in S2S^{2} (see Notation 2.8). Because symplectomorphisms preserve Hamiltonian data, we have that ψ1φtψ:22\psi^{-1}\circ\varphi^{t}\circ\psi:{\mathbb{R}}^{2}\to{\mathbb{R}}^{2} is given near the origin as the time tt flow of the Hamiltonian vector field for ψ(1/fε)\psi^{*}(1/f_{\varepsilon}) with respect to dxdy(1+x2+y2)2\frac{dx\wedge dy}{(1+x^{2}+y^{2})^{2}}. That is,

ψ1φtψis the time t flow ofε(1+x2+y2)2fyfε2xε(1+x2+y2)2fxfε2y.\psi^{-1}\circ\varphi^{t}\circ\psi\,\,\,\,\text{is the time $t$ flow of}\,\,\,\,\frac{\varepsilon(1+x^{2}+y^{2})^{2}f_{y}}{f_{\varepsilon}^{2}}\partial_{x}-\frac{\varepsilon(1+x^{2}+y^{2})^{2}f_{x}}{f_{\varepsilon}^{2}}\partial_{y}.

Recall that if Px+QyP\partial_{x}+Q\partial_{y} is a smooth vector field on 2{\mathbb{R}}^{2}, vanishing at (0,0)(0,0), then the linearization of its time tt flow evaluated at the origin is represented by the 2×22\times 2 matrix exp(tX)\exp(tX), with respect to the standard ordered basis (x,y)(\partial_{x},\partial_{y}) of T(0,0)2T_{(0,0)}{\mathbb{R}}^{2}, where

X=(PxPyQxQy).X=\begin{pmatrix}P_{x}&P_{y}\\ Q_{x}&Q_{y}\end{pmatrix}.

Here, the partial derivatives of PP and QQ are implicitly assumed to be evaluated at the origin. In this spirit, set P=ε(1+x2+y2)2fyfε2P=\frac{\varepsilon(1+x^{2}+y^{2})^{2}f_{y}}{f_{\varepsilon}^{2}} and Q=ε(1+x2+y2)2fxfε2Q=\frac{-\varepsilon(1+x^{2}+y^{2})^{2}f_{x}}{f_{\varepsilon}^{2}}, and we compute

X=(PxPyQxQy)=εfϵ(p)2(fyx(0,0)fyy(0,0)fxx(0,0)fxy(0,0))=εfε(p)2J0H(f,ψ).X=\begin{pmatrix}P_{x}&P_{y}\\ Q_{x}&Q_{y}\end{pmatrix}=\frac{\varepsilon}{f_{\epsilon}(p)^{2}}\begin{pmatrix}f_{yx}(0,0)&f_{yy}(0,0)\\ -f_{xx}(0,0)&-f_{xy}(0,0)\end{pmatrix}=\frac{-\varepsilon}{f_{\varepsilon}(p)^{2}}J_{0}\cdot H(f,\psi).

This implies that B=D1exp(tεfε(p)2J0H(f,ψ)missing)DB=D^{-1}\exp\big(\frac{-t\varepsilon}{f_{\varepsilon}(p)^{2}}J_{0}H(f,\psi)\big{missing})D, where DD is a change of basis matrix relating (v1,v2)(v_{1},v_{2}) and the pushforward of (x,y)(\partial_{x},\partial_{y}) by ψ\psi in TpS2T_{p}S^{2}. Because ψ\psi is holomorphic, DD must equal r(s)r\cdot\mathcal{R}(s) for some r>0r>0 and some ss\in{\mathbb{R}}. This provides that

B=(s)exp(tεfε(p)2J0H(f,ψ)missing)(s).B=\mathcal{R}(-s)\exp\bigg(\frac{-t\varepsilon}{f_{\varepsilon}(p)^{2}}J_{0}H(f,\psi)\bigg{missing})\mathcal{R}(s). (2.10)

Finally, we combine (2.9) and (2.10) to conclude

Mt=AB\displaystyle M_{t}=A\cdot B =(2tfε(p))(s)exp(tεfε(p)2J0H(f,ψ)missing)(s)\displaystyle=\mathcal{R}\bigg{(}\frac{2t}{f_{\varepsilon}(p)}\bigg{)}\cdot\mathcal{R}(-s)\exp\bigg(\frac{-t\varepsilon}{f_{\varepsilon}(p)^{2}}J_{0}H(f,\psi)\bigg{missing})\mathcal{R}(s)
=(s)(2tfε(p))exp(tεfε(p)2J0H(f,ψ)missing)(s).\displaystyle=\mathcal{R}(-s)\mathcal{R}\bigg{(}\frac{2t}{f_{\varepsilon}(p)}\bigg{)}\exp\bigg(\frac{-t\varepsilon}{f_{\varepsilon}(p)^{2}}J_{0}H(f,\psi)\bigg{missing})\mathcal{R}(s).

Corollary 2.11.

Fix L>0L>0. Then there exists some ε0>0\varepsilon_{0}>0 such that for ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}], all Reeb orbits γ𝒫L(λε)\gamma\in\mathcal{P}^{L}(\lambda_{\varepsilon}) are nondegenerate, take the form γpk\gamma_{p}^{k}, a kk-fold cover of an embedded Reeb orbit γp\gamma_{p} as in in Lemma 2.7, and μCZ(γ)=4k+indf(p)1\mu_{\operatorname{CZ}}(\gamma)=4k+\mbox{\em ind}_{f}(p)-1 for pCrit(f)p\in\mbox{\em Crit}(f).

Proof.

That γ\gamma is nondegenerate and projects to a critical point pp of ff is proven in [Ne20, Lemma 4.11]. To compute μCZ(γ)\mu_{\operatorname{CZ}}(\gamma), we apply the naturality, loop, and signature properties of the Conley Zehnder index (see [Sal99, §2.4]) to our path {Mt}Sp(2)\{M_{t}\}\subset\text{Sp}(2). By Proposition 2.10, this family of matrices has a factorization, up to SO(2)\text{SO}(2)-conjugation, Mt=ΦtΨtM_{t}=\Phi_{t}\Psi_{t}, where Φ\Phi is the loop of symplectic matrices /2πkfε(p)Sp(2){\mathbb{R}}/2\pi kf_{\varepsilon}(p){\mathbb{Z}}\to\text{Sp}(2), t(2t/fε(p))t\mapsto\mathcal{R}(2t/f_{\varepsilon}(p)), and Ψ\Psi is the path of matrix exponentials texp(tεfε(p)2J0H(f,ψ))t\mapsto\exp(\frac{-t\varepsilon}{f_{\varepsilon}(p)^{2}}J_{0}H(f,\psi)), where ψ\psi denotes a choice of stereographic coordinates at pp. In total,

μCZ(γpk)\displaystyle\mu_{\operatorname{CZ}}(\gamma_{p}^{k}) =2μ(Φ)+μCZ(Ψ)\displaystyle=2\mu(\Phi)+\mu_{\operatorname{CZ}}(\Psi)
=22k+12sign(H(f,ψ))\displaystyle=2\cdot 2k+\frac{1}{2}\text{sign}(-H(f,\psi))
=4k+indf(p)1.\displaystyle=4k+\text{ind}_{f}(p)-1.

Here, μ\mu denotes the Maslov index of a loop of symplectic matrices (see [MS15, §2]). ∎

2.2 Geometry of S3/GS^{3}/G and associated Reeb dynamics

The previous process of perturbing a degenerate contact form on prequantization bundles, is often used to compute Floer theories, for example, their cylindrical contact homology [Ne20] and embedded contact homology [NW23]. Although the quotients S3/GS^{3}/G are not prequantization bundles, they do admit an S1S^{1}-action (with fixed points), and are examples of Seifert fiber spaces which are realizable as principal S1S^{1}-orbibundles over integral symplectic orbifolds.

Let GSU(2)G\subset\text{SU}(2) be a finite nontrivial group. Since GG acts on S3S^{3} without fixed points, S3/GS^{3}/G inherits smooth structure. The quotient πG:S3S3/G\pi_{G}:S^{3}\to S^{3}/G is a universal cover, thus π1(S3/G)G\pi_{1}(S^{3}/G)\cong G is completely torsion, and rank H1(S3/G)=0\mbox{rank }H_{1}(S^{3}/G)=0. Because the GG-action preserves λΩ1(S3)\lambda\in\Omega^{1}(S^{3}), we have a descent of λ\lambda to a contact form on S3/GS^{3}/G, denoted λGΩ1(S3/G)\lambda_{G}\in\Omega^{1}(S^{3}/G), with ξG:=ker(λG)\xi_{G}:=\text{ker}(\lambda_{G}). As the actions of S1S^{1} and GG on S3S^{3} commute, we obtain an S1S^{1}-action on S3/GS^{3}/G, which realizes the Reeb flow of λG\lambda_{G}. Hence, λG\lambda_{G} is degenerate.

Let HSO(3)H\subset\text{SO}(3) denote P(G)P(G), the image of GG under P:SU(2)SO(3)P:\text{SU}(2)\to\text{SO}(3). The HH-action on S2S^{2} has fixed points, and so the quotient S2/HS^{2}/H inherits orbifold structure. Lemma 2.1 provides a unique map 𝔭:S3/GS2/H\mathfrak{p}:S^{3}/G\to S^{2}/H, making the following diagram commute

S3{S^{3}}S3/G{S^{3}/G}S2{S^{2}}S2/H{S^{2}/H}πG\scriptstyle{\pi_{G}}𝔓\scriptstyle{\mathfrak{P}}𝔭\scriptstyle{\mathfrak{p}}πH\scriptstyle{\pi_{H}} (2.11)

where πG\pi_{G} is a finite cover, πH\pi_{H} is an orbifold cover, 𝔓\mathfrak{P} is a projection of a prequantization bundle, and 𝔭\mathfrak{p} is identified with the Seifert fibration.

Remark 2.12.

(Global trivialization of ξG\xi_{G}). Recall the SU(2)\text{SU}(2)-invariant vector fields ViV_{i} spanning ξ\xi on S3S^{3} (2.2). Because these ViV_{i} are GG-invariant, they descend to smooth sections of ξG\xi_{G}, providing a global unitary trivialization, τG\tau_{G}, of ξG\xi_{G}, hence c1(ξG)=0c_{1}(\xi_{G})=0. Given a Reeb orbit γ\gamma of some contact form on S3/GS^{3}/G, we denote by μCZ(γ)\mu_{\operatorname{CZ}}(\gamma) the Conley Zehnder index of γ\gamma with respect to this global trivialization.

We assume that the Morse function f:S2f:S^{2}\to{\mathbb{R}} is HH-invariant and descends to an orbifold Morse function, fH:S2/Hf_{H}:S^{2}/H\to{\mathbb{R}}, in the language of [CH14]. The HH-invariance of ff provides that the smooth F=f𝔓F=f\circ\mathfrak{P} is GG-invariant, and descends to a smooth function, FG:S3/GF_{G}:S^{3}/G\to{\mathbb{R}}. We define, analogously to Notation 2.2,

fH,ε:=1+εfH,FG,ε:=1+εFG,λG,ε:=FG,ελG.f_{H,\varepsilon}:=1+\varepsilon f_{H},\hskip 42.67912ptF_{G,\varepsilon}:=1+\varepsilon F_{G},\hskip 42.67912pt\lambda_{G,\varepsilon}:=F_{G,\varepsilon}\lambda_{G}.

For sufficiently small ε\varepsilon, λG,ε\lambda_{G,\varepsilon} is a contact form on S3/GS^{3}/G with kernel ξG\xi_{G}. The condition πGλG,ε=λε\pi_{G}^{*}\lambda_{G,\varepsilon}=\lambda_{\varepsilon} implies that γ:[0,T]S3\gamma:[0,T]\to S^{3} is an integral curve of RλεR_{\lambda_{\varepsilon}} if and only if πGγ:[0,T]S3/G\pi_{G}\circ\gamma:[0,T]\to S^{3}/G is an integral curve of RλG,εR_{\lambda_{G,\varepsilon}}.

Remark 2.13.

(Local models on S3S^{3} and S3/GS^{3}/G agree). Suppose γ:[0,T]S3\gamma:[0,T]\to S^{3} is a Reeb trajectory of λε\lambda_{\varepsilon}, so that πGγ:[0,T]S3/G\pi_{G}\circ\gamma:[0,T]\to S^{3}/G is a Reeb trajectory of λG,ε\lambda_{G,\varepsilon}. For t[0,T]t\in[0,T], let MtSp(2)M_{t}\in\text{Sp}(2) denote the time tt linearized Reeb flow of λε\lambda_{\varepsilon} along γ\gamma with respect to τ\tau, and let NtSp(2)N_{t}\in\text{Sp}(2) denote that of λG,ε\lambda_{G,\varepsilon} along πGγ\pi_{G}\circ\gamma with respect to τG\tau_{G}. Then Mt=NtM_{t}=N_{t}, because the local contactomorphism πG\pi_{G} preserves the trivializations in addition to the contact forms.

Let 𝒪(p):={hp|hH}\mathcal{O}(p):=\{h\cdot p\ |\,h\in H\} to be the orbit of pp, and denote the isotropy subgroup of pp by

Hp:={hH|hp=p}H.H_{p}:=\{h\in H\ |\,h\cdot p=p\}\subset H.

Recall that |𝒪(p)||Hp|=|H||\mathcal{O}(p)||H_{p}|=|H| for any pS2p\in S^{2}. A point pS2p\in S^{2} is a fixed point if |Hp|>1|H_{p}|>1. The set of fixed points of HH is Fix(H)\text{Fix}(H). The point qS2/Hq\in S^{2}/H is an orbifold point if q=πH(p)q=\pi_{H}(p) for some pFix(H)p\in\text{Fix}(H). We now additionally assume that ff satisfies Crit(f)=Fix(H)\text{Crit}(f)=\text{Fix}(H); this will be the case in Section 3. The Reeb orbit γp𝒫(λε)\gamma_{p}\in\mathcal{P}(\lambda_{\varepsilon}) from Lemma 2.7 projects to pCrit(f)p\in\text{Crit}(f) under 𝔓\mathfrak{P}, and thus πGγp𝒫(λG,ε)\pi_{G}\circ\gamma_{p}\in\mathcal{P}(\lambda_{G,\varepsilon}) projects to the orbifold point πH(p)\pi_{H}(p) under 𝔭\mathfrak{p}. Lemma 2.14 computes the Reeb orbit multiplicity d(πGγp)d(\pi_{G}\circ\gamma_{p}).

Lemma 2.14.

Let γp𝒫(λε)\gamma_{p}\in\mathcal{P}(\lambda_{\varepsilon}) be the embedded Reeb orbit in S3S^{3} from Lemma 2.7. Then the multiplicity of πGγp𝒫(λG,ε)\pi_{G}\circ\gamma_{p}\in\mathcal{P}(\lambda_{G,\varepsilon}) is 2|Hp|2|H_{p}| if |G||G| is even, and is |Hp||H_{p}| if |G||G| is odd.

Proof.

Recall that |G||G| is even if and only if |G|=2|H||G|=2|H|, and that |G||G| is odd if and only if |G|=|H||G|=|H| (the only element of SU(2)\text{SU}(2) of order 2 is Id-\text{Id}, the generator of ker(P)\text{ker}(P)). By the classification of finite subgroups of SU(2)\text{SU}(2), if |G||G| is odd, then GG is cyclic.

Let q:=πH(p)S2/Hq:=\pi_{H}(p)\in S^{2}/H, let d:=|𝒪(p)|d:=|\mathcal{O}(p)| so that d|Hp|=|H|d|H_{p}|=|H| and |G|=rd|Hp||G|=rd|H_{p}|, where r=2r=2 when |G||G| is even and r=1r=1 when odd. Label the points of 𝒪(p)\mathcal{O}(p) by p1=p,p2,,pdp_{1}=p,p_{2},\dots,p_{d}. Now 𝔓1(𝒪(p))\mathfrak{P}^{-1}(\mathcal{O}(p)) is a disjoint union of dd Hopf fibers, CiC_{i}, where Ci=𝔓1(pi)C_{i}=\mathfrak{P}^{-1}(p_{i}). Let CC denote the embedded circle 𝔭1(q)S3/G\mathfrak{p}^{-1}(q)\subset S^{3}/G. By commutativity of (2.11), we have that πG1(C)=iCi=𝔓1(𝒪(p))\pi_{G}^{-1}(C)=\sqcup_{i}C_{i}=\mathfrak{P}^{-1}(\mathcal{O}(p)). We have the following commutative diagram of circles and points:

i=1dCi=πG1(C){\sqcup_{i=1}^{d}C_{i}=\pi_{G}^{-1}(C)}C=𝔭1(q){C=\mathfrak{p}^{-1}(q)}{p1,,pd}=𝒪(p){\{p_{1},\dots,p_{d}\}=\mathcal{O}(p)}{q}{\{q\}}πG\scriptstyle{\pi_{G}}𝔓\scriptstyle{\mathfrak{P}}𝔭\scriptstyle{\mathfrak{p}}πH\scriptstyle{\pi_{H}}

We must have that πG:i=1dCiC\pi_{G}:\sqcup_{i=1}^{d}C_{i}\to C is a |G|=rd|Hp||G|=rd|H_{p}|-fold cover from the disjoint union of dd Hopf fibers to one embedded circle. The restriction of πG\pi_{G} to any one of these circles CiC_{i} provides a smooth covering map, CiCC_{i}\to C; let nin_{i} denote the degree of this cover. Because GG acts transitively on these circles, all of the degrees nin_{i} are equal to some nn. Thus, idni=dn=|G|=rd|Hp|\sum_{i}^{d}n_{i}=dn=|G|=rd|H_{p}| which implies that n=r|Hp|n=r|H_{p}| is the covering multiplicity of πGγp\pi_{G}\circ\gamma_{p}. ∎

We conclude this section with an analogue of Corollary 2.11 for the Reeb orbits of λG,ε\lambda_{G,\varepsilon}.

Lemma 2.15.

Fix L>0L>0. Then there exists some ε0>0\varepsilon_{0}>0 such that, for ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}], all Reeb orbits γ𝒫L(λG,ε)\gamma\in\mathcal{P}^{L}(\lambda_{G,\varepsilon}) are nondegenerate, project to an orbifold critical point of fHf_{H} under 𝔭:S3/GS2/H\mathfrak{p}:S^{3}/G\to S^{2}/H, and μCZ(γ)=4k+indf(p)1\mu_{\operatorname{CZ}}(\gamma)=4k+\mbox{\em ind}_{f}(p)-1 whenever γ\gamma is contractible with a lift to some orbit γpk\gamma_{p}^{k} in S3S^{3} as in Lemma 2.7, where pCrit(f)p\in\mbox{\em Crit}(f).

Proof.

Let L:=|G|LL^{\prime}:=|G|L and take the corresponding ε0\varepsilon_{0} as appearing in Corollary 2.11, applied to LL^{\prime}. Now, for ε(0,ε0]\varepsilon\in(0,\varepsilon_{0}], elements of 𝒫L(λε)\mathcal{P}^{L^{\prime}}(\lambda_{\varepsilon}) are nondegenerate and project to critical points of ff. Let γ𝒫L(λG,ε)\gamma\in\mathcal{P}^{L}(\lambda_{G,\varepsilon}) and let nn\in{\mathbb{N}} denote the order of [γ][\gamma] in π1(S3/G)G\pi_{1}(S^{3}/G)\cong G. Now we see that γn\gamma^{n} is contractible and lifts to an orbit γ~𝒫L(λε)\widetilde{\gamma}\in\mathcal{P}^{L^{\prime}}(\lambda_{\varepsilon}), which must be nondegenerate and must project to some critical point pp of ff under 𝔓\mathfrak{P}.

If the orbit γ\gamma is degenerate, then γn\gamma^{n} is degenerate and, by the discussion in Remark 2.13, γ~\widetilde{\gamma} would be degenerate. Commutativity of (2.11) implies that γ\gamma projects to the orbifold critical point πH(p)\pi_{H}(p) of fHf_{H} under 𝔭\mathfrak{p}. Finally, if n=1n=1 then again by Remark 2.13, the local model {Nt}t[0,T]\{N_{t}\}_{t\in[0,T]} of the Reeb flow along γ\gamma matches that of γ~\widetilde{\gamma}, {Mt}t[0,T]\{M_{t}\}_{t\in[0,T]}, and thus μCZ(γ)=μCZ(γ~)\mu_{\operatorname{CZ}}(\gamma)=\mu_{\operatorname{CZ}}(\widetilde{\gamma}). The latter index is computed in Corollary 2.11.

2.3 Construction of HH-invariant Morse-Smale functions

We now produce the HH-invariant, Morse-Smale functions on S2S^{2} for the dihedral H=𝔻2nH={\mathbb{D}}_{2n} and polyhedral H=H={\mathbb{P}} subgroups of SO(3)\text{SO}(3). Table 3 describes three finite subsets, X0X_{0}, X1X_{1}, and X2X_{2}, of S2S^{2} which depend on HSO(3)H\subset\text{SO}(3). We construct an HH-invariant, Morse-Smale function ff on (S2,ωFS(,j))(S^{2},\omega_{\text{FS}}(\cdot,j\cdot)), whose set of critical points of index ii is XiX_{i}, so that Crit(f)=X:=X0X1X2\text{Crit}(f)=X:=X_{0}\cup X_{1}\cup X_{2}. Additionally, X=Fix(H)X=\text{Fix}(H), the fixed point set of the HH-action on S2S^{2}. This constructed ff is perfect in the sense that it features the minimal number of required critical points, because Fix(H)Crit(f)\text{Fix}(H)\subset\text{Crit}(f) must always hold. In the case that HH is a polyhedral group, X0X_{0} is the set of vertex points, X1X_{1} is the set of edge midpoints, and X2X_{2} is the set of face barycenters.

We illustrate the constructions of our perfect HH-invariant, Morse-Smale functions on S2S^{2} in Figures 2.2, 2.2, 2.4, and 2.4, wherein the blue, violet, and red critical points are of Morse index 2, 1, and 0 respectively. Further details are also given in the following proof of Lemma 2.16

HH X0X_{0} X1X_{1} X2X_{2}
𝔻2n{\mathbb{D}}_{2n} {p,k| 1kn}\{p_{-,k}\,|\,1\leq k\leq n\} {ph,k| 1kn}\{p_{h,k}\,|\,1\leq k\leq n\} {p+,k| 1k2}\{p_{+,k}\,|\,1\leq k\leq 2\}
{\mathbb{P}} {𝔳k| 1kA}\{\mathfrak{v}_{k}\,|\,1\leq k\leq A\} {𝔢k| 1kB}\{\mathfrak{e}_{k}\,|\,1\leq k\leq B\} {𝔣k| 1kC}\{\mathfrak{f}_{k}\,|\,1\leq k\leq C\}
Table 3: Fixed points of HH sorted by Morse index
Lemma 2.16.

Let HSO(3)H\subset\mbox{\em SO}(3) be either 𝔻2n{\mathbb{D}}_{2n} or {\mathbb{P}}. Then there exists an HH-invariant, Morse function ff on S2S^{2}, with Crit(f)=X\mbox{\em Crit}(f)=X, such that indf(p)=i\mbox{\em ind}_{f}(p)=i if pXip\in X_{i}. Furthermore, there are stereographic coordinates at pXp\in X in which ff takes the form

  1. (i)

    q0(x,y):=(x2+y2)/21q_{0}(x,y):=(x^{2}+y^{2})/2-1,    if pX0p\in X_{0}

  2. (ii)

    q1(x,y):=(y2x2)/2q_{1}(x,y):=(y^{2}-x^{2})/2,              if pX1p\in X_{1}

  3. (iii)

    q2(x,y):=1(x2+y2)/2q_{2}(x,y):=1-(x^{2}+y^{2})/2,     if pX2p\in X_{2}

Proof.

We first produce an auxiliary Morse function f~\widetilde{f}, which might not be HH-invariant, then ff is taken to be the HH-average of f~\widetilde{f}. Fix δ>0\delta>0; for pXp\in X, let DpS2D_{p}\subset S^{2} be the open geodesic disc centered at pp with radius δ\delta with respect to the metric ωFS(,j)\omega_{\text{FS}}(\cdot,j\cdot). Define f~\widetilde{f} on DpD_{p} to be the pullback of q0q_{0}, q1q_{1}, or q2q_{2}, for pX0p\in X_{0}, X1X_{1}, or X2X_{2} respectively, by stereographic coordinates at pp. Set D:=pXDpD:=\cup_{p\in X}D_{p}. For δ\delta small, DD is a disjoint union and f~:D\widetilde{f}:D\to{\mathbb{R}} is Morse. We can arrange for our selection of stereographic coordinates to satisfy

()for allpX1,andhH,f~|Dp=f~ϕh|Dp,(*)\,\,\text{for all}\,\,p\in X_{1},\,\,\text{and}\,\,h\in H,\,\,\widetilde{f}|_{D_{p}}=\widetilde{f}\circ\phi_{h}|_{D_{p}},

where ϕh:S2S2\phi_{h}:S^{2}\to S^{2} denotes xhxx\mapsto h\cdot x. This ensures that the “saddles are rotated in the same HH-direction”. Notice that ()(*) automatically holds for pX0X2p\in X_{0}\cup X_{2}, for any choice of coordinates, because of the rotational symmetry of the quadratics q0q_{0} and q2q_{2}. Note that:

  1. (a)

    The δ\delta-neighborhood DD of XX in S2S^{2}, is an HH-invariant set, and for all pXp\in X, the HpH_{p}-action restricts to an action on DpD_{p}, where HpHH_{p}\subset H denotes the stabilizer subgroup.

  2. (b)

    For all pXp\in X, f~|Dp:Dp\widetilde{f}|_{D_{p}}:D_{p}\to{\mathbb{R}} is HpH_{p}-invariant.

  3. (c)

    The function f~:D\widetilde{f}:D\to{\mathbb{R}} is an HH-invariant Morse function, with Crit(f~)=X\text{Crit}(\widetilde{f})=X.

The HH-invariance and HpH_{p}-invariance in (a) hold, because HSO(3)H\subset\text{SO}(3) acts on S2S^{2} by ωFS(,j)\omega_{\text{FS}}(\cdot,j\cdot)-isometries (rotations about axes through pXp\in X), and because XX is an HH-invariant set. The HpH_{p}-invariance of (b) holds because, in stereographic coordinates, the HpH_{p}-action pulled back to 2{\mathbb{R}}^{2} is always generated by some linear rotation about the origin, (θ)\mathcal{R}(\theta). Both q0q_{0} and q2q_{2} are invariant with respect to any (θ)\mathcal{R}(\theta), whereas q1q_{1} is invariant with respect to the action generated by (π)\mathcal{R}(\pi), which is precisely the action by HpH_{p} when pX1p\in X_{1}, so (b) holds. Finally, the HH-invariance in (c) holds directly by the HpH_{p} invariance from (b), and by ()(*). Now, extend the domain of f~\widetilde{f} from DD to all of S2S^{2} so that f~\widetilde{f} is smooth and Morse, with Crit(f~)=X\text{Crit}(\widetilde{f})=X. Figures 2.2 and 2.2 depict possible extensions f~\widetilde{f} in the H=𝔻2nH={\mathbb{D}}_{2n} and H=𝕋H={\mathbb{T}} cases, for example.

For hHh\in H, let ϕh:S2S2\phi_{h}:S^{2}\to S^{2} denote the group action, phpp\mapsto h\cdot p. Define

f:=1|H|hHϕhf~,f:=\frac{1}{|H|}\sum_{h\in H}\phi_{h}^{*}\widetilde{f},

where |H||H|\in{\mathbb{N}} is the group order of HH. This HH-invariant ff is smooth and agrees with f~\widetilde{f} on DD. If no critical points are created in the averaging process of f~\widetilde{f}, then we have that Crit(f)=X\text{Crit}(f)=X, implying that ff is Morse, and we are done.

We say that the extension f~\widetilde{f} to S2S^{2} from DD is roughly HH-invariant, if for any pS2Xp\in S^{2}\setminus X and hHh\in H, the angle between the nonzero gradient vectors

grad(f~)andgrad(ϕhf~)\text{grad}(\widetilde{f})\,\,\,\text{and}\,\,\,\text{grad}(\phi_{h}^{*}\widetilde{f})

in TpS2T_{p}S^{2} is less than π/2\pi/2. If f~\widetilde{f} if roughly HH-invariant, then for pXp\notin X, grad(f)(p)\text{grad}(f)(p) is an average of a collection of nonzero vectors in the same convex half space of TpS2T_{p}S^{2} and must be nonzero, implying pCrit(f)p\notin\text{Crit}(f). That is, if f~\widetilde{f} is roughly HH-invariant, then Crit(f)=X\text{Crit}(f)=X, as desired. The extensions f~\widetilde{f} in Figures 2.2, 2.2, 2.4, and 2.4 are all roughly HH-invariant by inspection, and the proof is complete. ∎

Refer to caption
Figure 2.1: A dihedral f~\widetilde{f}
Refer to caption
Figure 2.2: A tetrahedral f~\widetilde{f}
Refer to caption
Figure 2.3: An octahedral f~\widetilde{f}
Refer to caption
Figure 2.4: An icosahedral f~\widetilde{f}
Lemma 2.17.

If ff is a Morse function on a 2-dimensional manifold SS such that f(p1)=f(p2)f(p_{1})=f(p_{2}) for all p1,p2Crit(f)p_{1},p_{2}\in\mbox{\em Crit}(f) with Morse index 1, then ff is Smale, given any metric on SS.

Proof.

Given metric gg on SS, ff fails to be Smale with respect to gg if and only if there are two distinct critical points of ff of Morse index 1 that are connected by a gradient flow line of ff. Because all such critical points have the same ff value, no such flow line exists. ∎

Remark 2.18.

By Lemma 2.17, the Morse function ff provided in Lemma 2.16 is Smale for ωFS(,j)\omega_{\text{FS}}(\cdot,j\cdot).

2.4 Cylinders over orbifold Morse trajectories

In Section 3 we will compute the action filtered cylindrical contact homology groups using the preceding set up. In particular we will show that the grading of any generator of the filtered chain complex is even444Recall that the degree of a generator γ\gamma is given by |γ|=μCZ(γ)1|\gamma|=\mu_{\operatorname{CZ}}(\gamma)-1., implying that the action filtered differential vanishes.

It is interesting to note however that not all moduli spaces of holomorphic cylinders are empty. In this section, we elucidate the correspondence between moduli spaces of certain JJ-holomorphic cylinders and the moduli spaces of orbifold Morse trajectories in the base; the latter is often nonempty. We establish an orbifold version of the correspondence between cylinders and flow lines, in particular constructing a holomorphic cylinder from an orbifold Morse trajectory. We do not provide the full details as to why all holomorphic cylinders arise this way; this direction of the correspondence follows by way of the arguments as collected in [Ne20, §5] in the context of prequantization bundles, which may be invoked in the setting at hand as a result of [HN16, §2, 4], [Wen-SFT, §10], [SZ92]. While nothing presented in this section is necessary to the proof of Theorem 1.2, the correspondence may be of value for computing other contact homology theories.

As discussed in Section 1.4, the Seifert projection 𝔭:S3/GS2/H\mathfrak{p}:S^{3}/G\to S^{2}/H highlights many of the interplays between orbifold Morse theory and cylindrical contact homology. In particular, the projection geometrically relates holomorphic cylinders in ×S3/G{\mathbb{R}}\times S^{3}/G to the orbifold Morse trajectories in S2/HS^{2}/H. This necessitates a discussion about the complex structure on ξG\xi_{G} that we will use.

Remark 2.19.

(The canonical complex structure JJ on ξG\xi_{G})
Recall that the GG-action on S3S^{3} preserves the standard complex structure J2J_{{\mathbb{C}}^{2}} on ξTS3\xi\subset TS^{3}. Thus, J2J_{{\mathbb{C}}^{2}} descends to a complex structure on ξG\xi_{G}, which we will denote simply by JJ in this section. Note that for any sufficiently small ε>0\varepsilon>0, and for any ff on S2S^{2}, J2J_{{\mathbb{C}}^{2}} is λε\lambda_{\varepsilon}-compatible, thus JJ is λG,ε\lambda_{G,\varepsilon}-compatible as well.555This JJ might not be one of the generic JNJ_{N} used to compute the filtered homology groups in the later Sections 3.1, 3.2, or 3.3. A generic choice of JNJ_{N} is necessary to ensure transversality of the cylinders in symplectic cobordisms, which are used to define the chain maps later in Section 4.

Remark 2.20.

(J2J_{{\mathbb{C}}^{2}}-holomorphic cylinders in ×S3{\mathbb{R}}\times S^{3} over Morse flow lines in S2S^{2})
Fix critical points pp and qq of a Morse-Smale function ff on (S2,ωFS(,j))(S^{2},\omega_{\text{FS}}(\cdot,j\cdot)). Fix ε>0\varepsilon>0 sufficiently small. By [Ne20, Propositions 5.4, 5.5], we have a bijective correspondence between (p,q)\mathcal{M}(p,q) and J2(γpk,γqk)/\mathcal{M}^{J_{{\mathbb{C}}^{2}}}(\gamma_{p}^{k},\gamma_{q}^{k})/{\mathbb{R}}, for any kk\in{\mathbb{N}}, where γp\gamma_{p} and γq\gamma_{q} are the embedded Reeb orbits in S3S^{3} projecting to pp and qq under 𝔓\mathfrak{P}. Given a Morse trajectory x(p,q)x\in\mathcal{M}(p,q), the components of the corresponding cylinder ux:×S1×S3u_{x}:{\mathbb{R}}\times S^{1}\to{\mathbb{R}}\times S^{3} are explicitly written down in [Ne20, §5] in terms of a parametrization of xx, the Hopf action on S3S^{3}, the Morse function ff, and the horizontal lift of its gradient to ξ\xi. The resulting uxJ2(γp,γq)/u_{x}\in\mathcal{M}^{J_{{\mathbb{C}}^{2}}}(\gamma_{p},\gamma_{q})/{\mathbb{R}} is J2J_{{\mathbb{C}}^{2}}-holomorphic.666We are abusing notation by conflating the parametrized cylindrical map uxu_{x} with the equivalence class [ux]J2(γp,γq)/[u_{x}]\in\mathcal{M}^{J_{{\mathbb{C}}^{2}}}(\gamma_{p},\gamma_{q})/{\mathbb{R}}; we will continue to abuse notation in this way. Furthermore, the Fredholm index of uxu_{x} agrees with that of xx. The image of the composition

×S1ux×S3πS3S3𝔓S2{\mathbb{R}}\times S^{1}\xrightarrow{u_{x}}{\mathbb{R}}\times S^{3}\xrightarrow{\pi_{S^{3}}}S^{3}\xrightarrow{\mathfrak{P}}S^{2}

equals the image of xx in S2S^{2}. We call uxu_{x} the cylinder over xx.

The following procedure uses Remark 2.20 and Diagram 2.11 to establish a similar correspondence between moduli spaces of orbifold flow lines of S2/HS^{2}/H and moduli spaces of JJ-holomorphic cylinders in ×S3/G{\mathbb{R}}\times S^{3}/G, where JJ is taken to be the J2J_{{\mathbb{C}}^{2}}-descended complex structure on ξG\xi_{G} from Remark 2.19.

  1. 1.

    Take x(p,q)x\in\mathcal{M}(p,q), for orbifold Morse critical points p,qS2/Hp,q\in S^{2}/H of fHf_{H}.

  2. 2.

    Take a πH\pi_{H}-lift, x~:S2\widetilde{x}:{\mathbb{R}}\to S^{2} of xx, from p~\widetilde{p} to q~\widetilde{q} in S2S^{2}, for some preimages p~\widetilde{p} and q~\widetilde{q} of pp and qq. We have x~(p~,q~)\widetilde{x}\in\mathcal{M}(\widetilde{p},\widetilde{q}).

  3. 3.

    Let ux~u_{\widetilde{x}} be the J2J_{{\mathbb{C}}^{2}}-holomorphic cylinder in ×S3{\mathbb{R}}\times S^{3} over x~\widetilde{x} (see Remark 2.20). We now have ux~J2(γp~,γq~)/u_{\widetilde{x}}\in\mathcal{M}^{J_{{\mathbb{C}}^{2}}}(\gamma_{\widetilde{p}},\gamma_{\widetilde{q}})/{\mathbb{R}}.

  4. 4.

    Let ux:×S1×S3/Gu_{x}:{\mathbb{R}}\times S^{1}\to{\mathbb{R}}\times S^{3}/G denote the composition

    ×S1ux~×S3Id×πG×S3/G.{\mathbb{R}}\times S^{1}\xrightarrow{u_{\widetilde{x}}}{\mathbb{R}}\times S^{3}\xrightarrow{\text{Id}\times\pi_{G}}{\mathbb{R}}\times S^{3}/G.

    Because JJ is the πG\pi_{G}-descent of J2J_{{\mathbb{C}}^{2}}, we have that

    Id×πG:(×S3,J2)(×S3/G,J)\text{Id}\times\pi_{G}:({\mathbb{R}}\times S^{3},J_{{\mathbb{C}}^{2}})\to({\mathbb{R}}\times S^{3}/G,J)

    is a holomorphic map. This implies that uxu_{x} is JJ-holomorphic;

    uxJ(πGγp~,πGγq~)/.u_{x}\in\mathcal{M}^{J}(\pi_{G}\circ\gamma_{\widetilde{p}},\pi_{G}\circ\gamma_{\widetilde{q}})/{\mathbb{R}}. (2.12)

Note that πGγp~\pi_{G}\circ\gamma_{\widetilde{p}} and πGγq~\pi_{G}\circ\gamma_{\widetilde{q}} are contractible Reeb orbits of λG,ε\lambda_{G,\varepsilon} projecting to pp and qq, respectively. Thus, if γp\gamma_{p} and γq\gamma_{q} are the embedded (potentially non-contractible) Reeb orbits of λG,ε\lambda_{G,\varepsilon} in S3/GS^{3}/G over pp and qq, then we have that

πGγp~=γpmp,andπGγq~=γqmq,\pi_{G}\circ\gamma_{\widetilde{p}}=\gamma_{p}^{m_{p}},\,\,\,\text{and}\,\,\,\pi_{G}\circ\gamma_{\widetilde{q}}=\gamma_{q}^{m_{q}},

where mp,mqm_{p},m_{q}\in{\mathbb{N}} are the orders of [γp][\gamma_{p}] and [γq][\gamma_{q}] in π1(S3/G)\pi_{1}(S^{3}/G). In particular, we can simplify equation (2.12):

uxJ(γpmp,γqmq)/.u_{x}\in\mathcal{M}^{J}(\gamma_{p}^{m_{p}},\gamma_{q}^{m_{q}})/{\mathbb{R}}.

This allows us to establish an orbifold version of the correspondence in [Ne20, §5]:

(p,q)\displaystyle\mathcal{M}(p,q) J(γpmp,γqmq)/\displaystyle\cong\mathcal{M}^{J}(\gamma_{p}^{m_{p}},\gamma_{q}^{m_{q}})/{\mathbb{R}} (2.13)
x\displaystyle x ux.\displaystyle\sim u_{x}.
Remark 2.21.

In order to conclude that all holomorphic cylinders arise as lifts of orbifold Morse trajectories, one must make a straightforward modification of the arguments as explained in [Ne20, §5], which adapts [Wen-SFT, Thm. 10.30, 10.32], which in turn is a modification of the original arguments by Salamon and Zehnder [SZ92]. The proof of [Ne20, Thm. 5.5] holds in the present setting as a result of the compactness results established in [HN16, Prop. 2.8] and automatic transversality results [HN16, Prop. 4.2(b)], [Wen10].

2.5 Orbifold and contact interplays: an example

Before giving the proof of the main theorem, we continue with our digression establishing connections between the contact data of S3/GS^{3}/G and the orbifold Morse data of S2/HS^{2}/H, as previously alluded to in Section 1.4. As before, for each pCrit(f)p\in\text{Crit}(f), select an orientation of the embedded disc W(p)W^{-}(p). The action of the stabilizer (equivalently, isotropy) subgroup of pp,

Hp:={hH|hp=p}H,H_{p}:=\{h\in H\,|\,h\cdot p=p\}\subset H,

on S2S^{2} restricts to an action on W(p)W^{-}(p) by diffeomorphisms. We say that the critical point pp is orientable if this action is by orientation preserving diffeomorphisms. Let Crit+(f)Crit(f)\text{Crit}^{+}(f)\subset\text{Crit}(f) denote the set of orientable critical points.

Note that the HH-action on S2S^{2} permutes Crit(f)\text{Crit}(f), and the action restricts to a permutation of Crit+(f)\text{Crit}^{+}(f). Furthermore, the index of a critical point is preserved by the action. Let Crit(fH)\text{Crit}(f_{H}), Crit+(fH)\text{Crit}^{+}(f_{H}), and Critk+(fH)\text{Crit}^{+}_{k}(f_{H}) denote the quotients Crit(f)/H\text{Crit}(f)/H, Crit+(f)/H\text{Crit}^{+}(f)/H, and Critk+(f)/H\text{Crit}^{+}_{k}(f)/H, respectively. As in the smooth case, we define the kthk^{\text{th}}-orbifold Morse chain group, denoted CMkorbCM_{k}^{\text{orb}}, to be the free abelian group generated by Critk+(fH)\text{Crit}_{k}^{+}(f_{H}). The differential will be defined by a signed and weighted count of negative gradient trajectories in S2/HS^{2}/H. The homology of this chain complex is, as in the smooth case, isomorphic to the singular homology of S2/HS^{2}/H ([CH14, Theorem 2.9]).

First, we demonstrate why it is necessary to discard the non-orientable critical points.

Remark 2.22.

(Discarding non-orientable critical points to recover singular homology)
Every index 1 critical point of f:S2f:S^{2}\to{\mathbb{R}} depicted in Figures 2.2, 2.2, 2.4, and 2.4 is non-orientable. This is because the unstable submanifolds associated to each of these critical points is an open interval, and the action of the stabilizer of each such critical point is a 180-degree rotation of S2S^{2} about an axis through the critical point. Thus, this action reverses the orientation of the embedded open intervals. If we were to include these index 1 critical points in the chain complex, then CMorbCM^{\text{orb}}_{*} would have rank three, with

CM0orbCM1orbCM2orb.CM^{\text{orb}}_{0}\cong CM^{\text{orb}}_{1}\cong CM^{\text{orb}}_{2}\cong{\mathbb{Z}}.

Note that it is not possible to define a differential on this purported chain complex with homology isomorphic to H(S2/H;)H(S2;)H_{*}(S^{2}/H;{\mathbb{Z}})\cong H_{*}(S^{2};{\mathbb{Z}}). Indeed, the correct chain complex, obtained by discarding the non-orientable index 1 critical points, has rank two:

CM0orbCM2orb,CM1orb=0,CM_{0}^{\text{orb}}\cong CM_{2}^{\text{orb}}\cong{\mathbb{Z}},\,\,\,CM_{1}^{\text{orb}}=0,

and has vanishing differential, producing isomorphic homology H(S2/H;)H(S2;)H_{*}(S^{2}/H;{\mathbb{Z}})\cong H_{*}(S^{2};{\mathbb{Z}}).

Next, we explain why it is necessary to discard the non-orientable critical points to orient the gradient trajectories.

Remark 2.23.

(Discarding non-orientable critical points to orient the gradient trajectories)
Let pp and qq be orientable critical points in S2S^{2} with Morse index difference equal to 1. Let x:S2/Hx:{\mathbb{R}}\to S^{2}/H be a negative gradient trajectory of fHf_{H} from [p][p] to [q][q]. Because pp and qq are orientable, the value of ϵ(x~){±1}\epsilon(\widetilde{x})\in\{\pm 1\} is independent of any choice of lift of xx to a negative gradient trajectory x~:M\widetilde{x}:{\mathbb{R}}\to M of ff in S2S^{2}. We define ϵ(x)\epsilon(x) to be this value. Conversely, if one of the points pp or qq is non-orientable, then there exist two lifts of xx with opposite signs, making the choice ϵ(x)\epsilon(x) dependent on choice of lift.

Recall that orb:CMkorbCMk1orb\partial^{\text{orb}}:CM_{k}^{\text{orb}}\to CM_{k-1}^{\text{orb}} is defined as follows. Let pCritk+(fH)p\in\text{Crit}_{k}^{+}(f_{H}) and qCritk1+(fH)q\in\text{Crit}_{k-1}^{+}(f_{H}) be orientable critical points then

orbp,q:=x(p,q)ϵ(x)|Hp||Hx|.\langle\partial^{\text{orb}}p,q\rangle:=\sum_{x\in\mathcal{M}(p,q)}\epsilon(x)\frac{|H_{p}|}{|H_{x}|}\in{\mathbb{Z}}.

As previously mentioned, for any finite HSO(3)H\subset\text{SO}(3), the quotient S2/HS^{2}/H is an orbifold 2-sphere. When HH is cyclic, the orbifold 2-sphere S2/HS^{2}/H resembles a lemon shape, featuring two orbifold points, and is immediately homeomorphic to S2S^{2}. If not cyclic, HH is dihedral, or polyhedral. A fundamental domain for the HH-action on S2S^{2} in these two latter cases can be taken to be an isosceles, closed, geodesic triangle, denoted ΔHS2\Delta_{H}\subset S^{2}. These geodesic triangles ΔH\Delta_{H} are identified by the shaded regions of S2S^{2} in Figure 2.6 for H=𝔻2nH={\mathbb{D}}_{2n} and Figure 2.6 for H=𝕋H={\mathbb{T}}.

Refer to caption
Figure 2.5: Fundamental domain Δ𝔻2n\Delta_{{\mathbb{D}}_{2n}}
Refer to caption
Figure 2.6: Fundamental domain Δ𝕋\Delta_{{\mathbb{T}}}
  • In Figure 2.6, the three vertices of Δ𝔻2n\Delta_{{\mathbb{D}}_{2n}}are p+,1p_{+,1}, p,1p_{-,1}, and p,2p_{-,2}. We have that S2S^{2} is tessellated by |𝔻2n|=2n|{\mathbb{D}}_{2n}|=2n copies of Δ𝔻2n\Delta_{{\mathbb{D}}_{2n}} and the internal angles of Δ𝔻2n\Delta_{{\mathbb{D}}_{2n}} are π/2\pi/2, π/2\pi/2, and 2π/n2\pi/n; Δ𝔻2n\Delta_{{\mathbb{D}}_{2n}} is isosceles.

  • In Figure 2.6, the three vertices of Δ𝕋\Delta_{{\mathbb{T}}} are 𝔣1\mathfrak{f}_{1}, 𝔳1\mathfrak{v}_{1}, and 𝔳2\mathfrak{v}_{2}. We have that S2S^{2} is tessellated by |𝕋|=12|{\mathbb{T}}|=12 copies of Δ𝕋\Delta_{{\mathbb{T}}}, and the internal angles of Δ𝕋\Delta_{{\mathbb{T}}} are π/3\pi/3, π/3\pi/3, and 2π/32\pi/3; Δ𝕋\Delta_{{\mathbb{T}}} is isosceles.

The triangular fundamental domains Δ𝕆\Delta_{{\mathbb{O}}} and Δ𝕀\Delta_{{\mathbb{I}}} for the 𝕆{\mathbb{O}} and 𝕀{\mathbb{I}} actions on S2S^{2} are constructed analogously to Δ𝕋\Delta_{{\mathbb{T}}}. Ultimately, in every (non-cyclic) case, we have a closed, isosceles, geodesic triangle serving as a fundamental domain for the HH-action on S2S^{2}. Applying the HH-identifications on the boundary of ΔH\Delta_{H} produces S2/HS^{2}/H, a quotient that is homeomorphic to S2S^{2} with three orbifold points. Specifically, under the surjective quotient map restricted to the closed ΔH\Delta_{H}, depicted in Figure 2.7,

πH|ΔH:ΔHS2S2/H.\pi_{H}|_{\Delta_{H}}:\Delta_{H}\subset S^{2}\to S^{2}/H.

In terms of the critical points we obtain:

  • (blue maximum) one orbifold point of S2/HS^{2}/H has a preimage consisting of a single vertex of ΔH\Delta_{H}, this is an index 2 critical point in S2S^{2};

  • (violet saddle) one orbifold point of S2/HS^{2}/H has a preimage consisting of a single midpoint of an edge of ΔH\Delta_{H}, this is an index 1 critical point in S2S^{2};

  • (red minimum) one orbifold point of S2/HS^{2}/H has a preimage consisting of two vertices of ΔH\Delta_{H}; both are index 0 critical points in S2S^{2}.

Figure 2.7 depicts the attaching map for ΔH\Delta_{H} along the boundary and these points.

Refer to caption
Figure 2.7: A triangular fundamental domain ΔH\Delta_{H} produces S2/HS^{2}/H, a topological S2S^{2} with three orbifold points when HH is non-cyclic. The directional markers on the boundary of ΔH\Delta_{H} indicating the identifications under the HH-action simultaneously realize the orbifold Morse trajectories.

We now specialize to the case H=𝕋H={\mathbb{T}} and study the geometry of S3/𝕋S^{3}/{\mathbb{T}}^{*} and S2/𝕋S^{2}/{\mathbb{T}}. This choice H=𝕋H={\mathbb{T}} makes the examples and diagrams concrete; note that a choice of H=𝔻H={\mathbb{D}}, 𝕆{\mathbb{O}}, or 𝕀{\mathbb{I}} produces similar geometric scenarios. More generally it is expected that for prequantization orbibundles that the orbifold Morse flow lines are in correspondence with S1S^{1}-invariant holomorphic cylinders, but this has only been established in certain cases, cf. Haney-Mark [HM22] and Nelson-Weiler [NW2].

Using the notation to be introduced in Section 3.3, the three orbifold points of S2/𝕋S^{2}/{\mathbb{T}} are denoted 𝔳\mathfrak{v}, 𝔢\mathfrak{e}, and 𝔣\mathfrak{f}, which are critical points of the orbifold Morse function f𝕋f_{{\mathbb{T}}} of index 0, 1, and 2, respectively. Furthermore, for small ε>0\varepsilon>0, we have three embedded nondegenerate Reeb orbits, 𝒱\mathcal{V}, \mathcal{E}, and \mathcal{F} in S3/𝕋S^{3}/{\mathbb{T}}^{*} of λ𝕋,ε\lambda_{{\mathbb{T}}^{*},\varepsilon}, projecting to the respective orbifold critical points under 𝔭:S3/𝕋S2/𝕋\mathfrak{p}:S^{3}/{\mathbb{T}}^{*}\to S^{2}/{\mathbb{T}}. Figure 2.8 illustrates this data.

Refer to caption
Figure 2.8: The Seifert projection 𝔭\mathfrak{p} takes Reeb orbits and cylinders of S3/𝕋S^{3}/{\mathbb{T}}^{*} to orbifold critical points and Morse trajectories of S2/𝕋S^{2}/{\mathbb{T}}. The depicted cylinders in S3/𝕋S^{3}/{\mathbb{T}}^{*} should be understood as the images of infinite cylinders in the symplectization under the projection S3/𝕋×S3/𝕋S^{3}/{\mathbb{T}}^{*}\times{\mathbb{R}}\to S^{3}/{\mathbb{T}}^{*}.

In Section 1.4 we explained how bad Reeb orbits in cylindrical contact homology are analogous to non-orientable critical points in orbifold Morse theory. We explicitly realize this analogy geometrically with 𝔭:S3/𝕋S2/𝕋\mathfrak{p}:S^{3}/{\mathbb{T}}^{*}\to S^{2}/{\mathbb{T}}. In Section 3.3, we will show that the even iterates 2k\mathcal{E}^{2k} are examples of bad Reeb orbits, and by Remark 2.22, 𝔢\mathfrak{e} is a non-orientable critical point of f𝕋f_{{\mathbb{T}}}. The projection 𝔭\mathfrak{p} maps the bad Reeb orbits 2k\mathcal{E}^{2k} to the non-orientable critical point 𝔢\mathfrak{e}.

Next we consider the relationships between the moduli spaces of JJ-holomorphic cylinders and gradient flow lines. The orders of 𝒱\mathcal{V}, \mathcal{E}, and \mathcal{F} in π1(S3/𝕋)\pi_{1}(S^{3}/{\mathbb{T}}^{*}) are 6, 4, and 6, respectively (see Section 4.3.3). Thus, by the correspondence (2.13) in Section 2.4, we have the following identifications between moduli spaces of orbifold Morse flow lines of S2/𝕋S^{2}/{\mathbb{T}} and JJ-holomorphic cylinders in ×S3/𝕋{\mathbb{R}}\times S^{3}/{\mathbb{T}}^{*} (with respect to the complex structure JJ described in Remark 2.19):

(𝔣,𝔢)\displaystyle\mathcal{M}(\mathfrak{f},\mathfrak{e}) J(6,4)/,\displaystyle\cong\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{E}^{4})/{\mathbb{R}}, (2.14)
(𝔢,𝔳)\displaystyle\mathcal{M}(\mathfrak{e},\mathfrak{v}) J(4,𝒱6)/,\displaystyle\cong\mathcal{M}^{J}(\mathcal{E}^{4},\mathcal{V}^{6})/{\mathbb{R}}, (2.15)
(𝔣,𝔳)\displaystyle\mathcal{M}(\mathfrak{f},\mathfrak{v}) J(6,𝒱6)/.\displaystyle\cong\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{V}^{6})/{\mathbb{R}}. (2.16)

Correspondences (2.14) and (2.15) are between singleton sets. Indeed, let x+x_{+} be the unique orbifold Morse flow line from 𝔣\mathfrak{f} to 𝔢\mathfrak{e} in S2/𝕋S^{2}/{\mathbb{T}}, and let xx_{-} be the unique orbifold Morse flow line from 𝔢\mathfrak{e} to 𝔳\mathfrak{v}, depicted in Figure 2.8. Then we have corresponding cylinders u+u_{+} and uu_{-} from 6\mathcal{F}^{6} to 4\mathcal{E}^{4}, and from 4\mathcal{E}^{4} to 𝒱6\mathcal{V}^{6}, respectively:

{x+}=(𝔣,𝔢)\displaystyle\{x_{+}\}=\mathcal{M}(\mathfrak{f},\mathfrak{e}) J(6,4)/={u+},\displaystyle\cong\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{E}^{4})/{\mathbb{R}}=\{u_{+}\}, (2.17)
{x}=(𝔢,𝔳)\displaystyle\{x_{-}\}=\mathcal{M}(\mathfrak{e},\mathfrak{v}) J(4,𝒱6)/={u}.\displaystyle\cong\mathcal{M}^{J}(\mathcal{E}^{4},\mathcal{V}^{6})/{\mathbb{R}}=\{u_{-}\}. (2.18)

These cylinders u±u_{\pm} are depicted in Figure 2.8. Additionally, note that the indices of the corresponding objects agree:

ind(x+)\displaystyle\text{ind}(x_{+}) =indf𝕋(𝔣)indf𝕋(𝔢)=21=1,\displaystyle=\text{ind}_{f_{{\mathbb{T}}}}(\mathfrak{f})-\text{ind}_{f_{{\mathbb{T}}}}(\mathfrak{e})=2-1=1,
ind(u+)\displaystyle\text{ind}(u_{+}) =μCZ(6)μCZ(4)=54=1,\displaystyle=\mu_{\operatorname{CZ}}(\mathcal{F}^{6})-\mu_{\operatorname{CZ}}(\mathcal{E}^{4})=5-4=1,

and

ind(x)\displaystyle\text{ind}(x_{-}) =indf𝕋(𝔢)indf𝕋(𝔳)=10=1,\displaystyle=\text{ind}_{f_{{\mathbb{T}}}}(\mathfrak{e})-\text{ind}_{f_{{\mathbb{T}}}}(\mathfrak{v})=1-0=1,
ind(u)\displaystyle\text{ind}(u_{-}) =μCZ(4)μCZ(𝒱6)=43=1.\displaystyle=\mu_{\operatorname{CZ}}(\mathcal{E}^{4})-\mu_{\operatorname{CZ}}(\mathcal{V}^{6})=4-3=1.

Next, we consider the third correspondence of moduli spaces in (2.16). As in Figure 2.8, we have that (𝔣,𝔳)\mathcal{M}(\mathfrak{f},\mathfrak{v}) is diffeomorphic to a 1-dimensional open interval. For any x(𝔣,𝔳)x\in\mathcal{M}(\mathfrak{f},\mathfrak{v}), we have that

ind(x)=indf𝕋(𝔣)indf𝕋(𝔳)=2,\text{ind}(x)=\text{ind}_{f_{{\mathbb{T}}}}(\mathfrak{f})-\text{ind}_{f_{{\mathbb{T}}}}(\mathfrak{v})=2,

algebraically verifying that the moduli space of orbifold flow lines must be 1-dimensional. On the other hand, take any cylinder uJ(6,𝒱6)/u\in\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{V}^{6})/{\mathbb{R}}. Now,

ind(u)=μCZ(6)μCZ(𝒱6)=53=2,\text{ind}(u)=\mu_{\operatorname{CZ}}(\mathcal{F}^{6})-\mu_{\operatorname{CZ}}(\mathcal{V}^{6})=5-3=2,

verifying that this moduli space of cylinders is 11-dimensional, as expected by (2.16).

Note that both of the open 1-dimensional moduli spaces in (2.16) admit a compactification by broken objects. We can see explicitly from Figure 2.8 that both ends of the 1-dimensional moduli space (𝔣,𝔳)\mathcal{M}(\mathfrak{f},\mathfrak{v}) converge to the same once-broken orbifold Morse trajectory, (x+,x)(𝔣,𝔢)×(𝔢,𝔳)(x_{+},x_{-})\in\mathcal{M}(\mathfrak{f},\mathfrak{e})\times\mathcal{M}(\mathfrak{e},\mathfrak{v}). In particular, the compactification (𝔣,𝔳)¯\overline{\mathcal{M}(\mathfrak{f},\mathfrak{v})} is a topological S1S^{1}, obtained by adding the single point (x+,x)(x_{+},x_{-}) to an open interval, and we write

(𝔣,𝔳)¯=(𝔣,𝔳)((𝔣,𝔢)×(𝔢,𝔳)).\overline{\mathcal{M}(\mathfrak{f},\mathfrak{v})}=\mathcal{M}(\mathfrak{f},\mathfrak{v})\,\bigsqcup\,\bigg{(}\mathcal{M}(\mathfrak{f},\mathfrak{e})\times\mathcal{M}(\mathfrak{e},\mathfrak{v})\bigg{)}.

An identical phenomenon occurs for the compactification of the cylinders. That is, both ends of the 1-dimensional interval J(6,𝒱6)/\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{V}^{6})/{\mathbb{R}} converge to the same once broken building of cylinders (u+,u)(u_{+},u_{-}). The compactification J(6,𝒱6)/¯\overline{\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{V}^{6})/{\mathbb{R}}} is a topological S1S^{1}, obtained by adding a single point (u+,u)(u_{+},u_{-}) to an open interval, and we write

J(6,𝒱6)/¯=J(6,𝒱6)/(J(6,4)/×J(4,𝒱6/)).\overline{\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{V}^{6})/{\mathbb{R}}}=\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{V}^{6})/{\mathbb{R}}\,\bigsqcup\,\bigg{(}\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{E}^{4})/{\mathbb{R}}\times\mathcal{M}^{J}(\mathcal{E}^{4},\mathcal{V}^{6}/{\mathbb{R}})\bigg{)}.

In Section 1.4, we argued that the differentials of cylindrical contact homology and orbifold Morse homology are structurally identical due to the similarities in the compactifications of the 1-dimensional moduli spaces. This is due to the fact that in both theories, a once broken building can serve as a limit of multiple ends of a 1-dimensional moduli space. Our examples depict this phenomenon:

  • The broken building (x+,x)(x_{+},x_{-}) of orbifold Morse flow lines serves as the limit of both ends of the open interval (𝔣,𝔳)\mathcal{M}(\mathfrak{f},\mathfrak{v}).

  • The broken building (u+,u)(u_{+},u_{-}) of JJ-holomorphic cylinders serves as the limit of both ends of the open interval J(6,𝒱6)/\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{V}^{6})/{\mathbb{R}}.

Another analogy is highlighted in this example. In both homology theories, it is possible for a sequence of flow lines or cylinders between orientable objects to break along an intermediate non-orientable object (see [CH14, Example 2.10]). For example:

  • There is a sequence xn(𝔣,𝔳)x_{n}\in\mathcal{M}(\mathfrak{f},\mathfrak{v}) converging to the broken building (x+,x)(x_{+},x_{-}), which breaks at 𝔢\mathfrak{e}. The critical points 𝔣\mathfrak{f} and 𝔳\mathfrak{v} are orientable, whereas 𝔢\mathfrak{e} is non-orientable.

  • There is a sequence unJ(6,𝒱6)/u_{n}\in\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{V}^{6})/{\mathbb{R}} converging to the broken building (u+,u)(u_{+},u_{-}), which breaks along the orbit 4\mathcal{E}^{4}. The Reeb orbits 6\mathcal{F}^{6} and 𝒱6\mathcal{V}^{6} are good, whereas 4\mathcal{E}^{4} is bad.

As explained in Remark 2.23, we cannot assign a value ϵ(x±){±1}\epsilon(x_{\pm})\in\{\pm 1\} nor a value ϵ(u±){±1}\epsilon(u_{\pm})\in\{\pm 1\} to the objects x±x_{\pm} and u±u_{\pm}, because they have a non-orientable limiting object. This difficulty complicates the proof that 2=0\partial^{2}=0, as we cannot write the signed count as a sum involving terms of the form ϵ(x+)ϵ(x)\epsilon(x_{+})\epsilon(x_{-}) or ϵ(u+)ϵ(u)\epsilon(u_{+})\epsilon(u_{-}). One can nevertheless show that a once broken building breaking along a non-orientable object is utilized by an even number of ends of the 1-dimensional moduli space, and that a cyclic group action on this set of even number of ends interchanges the orientations. Using this fact, one shows 2=0\partial^{2}=0 (see [CH14, Remark 5.3] and [HN16, §4.4]). We see this explicitly in our examples:

  • The once broken building (x+,x)(x_{+},x_{-}) is the limit of two ends of the 1-dimensional moduli space of orbifold trajectories; one positive and one negative end.

  • The once broken building (u+,u)(u_{+},u_{-}) is the limit of two ends of the 1-dimensional moduli space of orbifold trajectories; one positive and one negative end.

Remark 2.24.

(Including non-orientable objects complicates 2=0\partial^{2}=0) In Section 1.4 we saw that one discards bad Reeb orbits and non-orientable orbifold critical points as generators in cylindrical contact homology and orbifold Morse homology respectively, in order to achieve 2=0\partial^{2}=0. Using our understanding of moduli spaces from Figure 2.8, we show why 2=0\partial^{2}=0 could not reasonably hold if we were to include the non-orientable critical point 𝔢\mathfrak{e} and bad Reeb orbit 4\mathcal{E}^{4} in the corresponding chain complexes. Suppose that we have some coherent way of assigning ±1\pm 1 to the trajectories x±x_{\pm} and cylinders u±u_{\pm}. Now, due to equations (2.17) and (2.18), we would have in the orbifold case

orb𝔣\displaystyle\partial^{\text{orb}}\mathfrak{f} =ϵ(x+)|𝕋𝔣||𝕋x+|𝔢=3ϵ(x+)𝔢\displaystyle=\dfrac{\epsilon(x_{+})|{\mathbb{T}}_{\mathfrak{f}}|}{|{\mathbb{T}}_{x_{+}}|}\mathfrak{e}=3\epsilon(x_{+})\mathfrak{e}
orb𝔢\displaystyle\partial^{\text{orb}}\mathfrak{e} =ϵ(x)|𝕋𝔢||𝕋x|𝔳=2ϵ(x)𝔳\displaystyle=\dfrac{\epsilon(x_{-})|{\mathbb{T}}_{\mathfrak{e}}|}{|{\mathbb{T}}_{x_{-}}|}\mathfrak{v}=2\epsilon(x_{-})\mathfrak{v}
\displaystyle\implies (orb)2𝔣,𝔳=6ϵ(x+)ϵ(x)0,\displaystyle\langle\left(\partial^{\text{orb}}\right)^{2}\mathfrak{f},\mathfrak{v}\rangle=6\epsilon(x_{+})\epsilon(x_{-})\neq 0,

where we have used that |𝕋𝔣|=3|{\mathbb{T}}_{\mathfrak{f}}|=3, |𝕋𝔢|=2|{\mathbb{T}}_{\mathfrak{e}}|=2, and |𝕋x±|=1|{\mathbb{T}}_{x_{\pm}}|=1. Similarly, again due to equations (2.17) and (2.18), we would have

6\displaystyle\partial\mathcal{F}^{6} =ϵ(u+)d(6)d(u+)4=3ϵ(u+)4\displaystyle=\dfrac{\epsilon(u_{+})d(\mathcal{F}^{6})}{d(u_{+})}\mathcal{E}^{4}=3\epsilon(u_{+})\mathcal{E}^{4}
4\displaystyle\partial\mathcal{E}^{4} =ϵ(u)d(4)d(u)𝒱6=2ϵ(u)𝒱6\displaystyle=\dfrac{\epsilon(u_{-})d(\mathcal{E}^{4})}{d(u_{-})}\mathcal{V}^{6}=2\epsilon(u_{-})\mathcal{V}^{6}
\displaystyle\implies 26,𝒱6=6ϵ(u+)ϵ(u)0,\displaystyle\langle\partial^{2}\mathcal{F}^{6},\mathcal{V}^{6}\rangle=6\epsilon(u_{+})\epsilon(u_{-})\neq 0,

where we have used that d(6)=6d(\mathcal{F}^{6})=6, d(4)=4d(\mathcal{E}^{4})=4, and d(u±)=2d(u_{\pm})=2.

Recall that the multiplicity of a JJ-holomorphic cylinder uu divides the multiplicity of the limiting Reeb orbits γ±\gamma_{\pm}. The following remark uses the example of this section to demonstrate it need not be the case that d(u)=GCD(d(γ+),d(γ))d(u)=\text{GCD}(d(\gamma_{+}),d(\gamma_{-})).

Remark 2.25.

By (2.16), J(6,𝒱6)/\mathcal{M}^{J}(\mathcal{F}^{6},\mathcal{V}^{6})/{\mathbb{R}} is nonempty, and must contain some uu. We see that m(6)=m(𝒱6)=6m(\mathcal{F}^{6})=m(\mathcal{V}^{6})=6, so that

GCD(d(6),d(𝒱6))=6.\text{GCD}(d(\mathcal{F}^{6}),d(\mathcal{V}^{6}))=6.

Suppose for contradiction’s sake that that d(u)=6d(u)=6, and consider the underlying somewhere injective JJ-holomorphic cylinder vv. It must be the case that u=v6u=v^{6} and that vJ(,𝒱)/v\in\mathcal{M}^{J}(\mathcal{F},\mathcal{V})/{\mathbb{R}}. The existence of such a vv implies that \mathcal{F} and 𝒱\mathcal{V} represent the same free homotopy class of loops in S3/𝕋S^{3}/{\mathbb{T}}^{*}. However, we will determine in Section 4.3.3 that these Reeb orbits represent distinct homotopy classes (see Table 14). Thus, d(u)d(u) is not equal to the GCD.

3 Filtered cylindrical contact homology

A finite subgroup of SU(2)\text{SU}(2) is either cyclic, conjugate to the binary dihedral group 𝔻2n{\mathbb{D}}^{*}_{2n}, or conjugate to a binary polyhedral group 𝕋{\mathbb{T}}^{*}, 𝕆{\mathbb{O}}^{*}, or 𝕀{\mathbb{I}}^{*}. If subgroups G1G_{1} and G2G_{2} satisfy A1G2A=G1A^{-1}G_{2}A=G_{1}, for ASU(2)A\in\text{SU}(2), then the map S3S3S^{3}\to S^{3}, pApp\mapsto A\cdot p, descends to a strict conactomorphism (S3/G1,λG1)(S3/G2,λG2)(S^{3}/G_{1},\lambda_{G_{1}})\to(S^{3}/G_{2},\lambda_{G_{2}}), preserving the Reeb dynamics. Thus, we compute the contact homology of S3/GS^{3}/G for a particular choice of GG. The action threshold used to compute the filtered homology depends on GG; for NN\in{\mathbb{N}}, LNL_{N}\in{\mathbb{R}} is given by:

  1. (i)

    2πNπn2\pi N-\frac{\pi}{n} when GG is cyclic of order nn;

  2. (ii)

    2πNπ2n2\pi N-\frac{\pi}{2n} when GG is conjugate to 𝔻2n{\mathbb{D}}_{2n}^{*};

  3. (iii)

    2πNπ102\pi N-\frac{\pi}{10} when GG is conjugate to 𝕋{\mathbb{T}}^{*}, 𝕆{\mathbb{O}}^{*}, or 𝕀{\mathbb{I}}^{*}.

3.1 Cyclic subgroups

From [AHNS17, Theorem 1.5], the positive S1S^{1}-equivariant symplectic homology of the link of the AnA_{n} singularity, LAn3L_{A_{n}}\subset{\mathbb{C}}^{3}, with contact structure ξ0=TLAnJ3(TLAn)\xi_{0}=TL_{A_{n}}\cap J_{{\mathbb{C}}^{3}}(TL_{A_{n}}), satisfies:

SH+,S1(LAn,ξ0)={n=1,n+13and odd0else.SH_{*}^{+,S^{1}}(L_{A_{n}},\xi_{0})=\begin{cases}{\mathbb{Q}}^{n}&*=1,\\ {\mathbb{Q}}^{n+1}&*\geq 3\,\,\mbox{and odd}\\ 0&\mbox{else.}\end{cases}

Furthermore, [BO17] prove that there are restricted classes of contact manifolds whose cylindrical contact homology (with a degree shift) is isomorphic to its positive S1S^{1}-equivariant symplectic homology, when both are defined over {\mathbb{Q}} coefficients. Indeed, we note an isomorphism by inspection when we compare this symplectic homology with the cylindrical contact homology of (LAn,ξ0)(S3/G,ξG)(L_{A_{n}},\xi_{0})\cong(S^{3}/G,\xi_{G}) for Gn+1G\cong{\mathbb{Z}}_{n+1} from Theorem 1.2. Although this cylindrical contact homology is computed in [Ne20, Theorem 1.36], we recompute these groups using a direct limit of filtered contact homology to present the general structure of the computations to come in the dihedral and polyhedral cases.

Let GnG\cong{\mathbb{Z}}_{n} be a finite cyclic subgroup of GG of order nn. If |G|=n|G|=n is even, with n=2mn=2m, then P:GH:=P(G)P:G\to H:=P(G) has nontrivial two element kernel, and HH is cyclic of order mm. Otherwise, nn is odd, P:GH=P(G)P:G\to H=P(G) has trivial kernel, and HH is cyclic of order nn.

By conjugating GG if necessary, we can assume that HH acts on S2S^{2} by rotations around the vertical axis through S2S^{2}. The height function f:S2[1,1]f:S^{2}\to[-1,1] is Morse, HH-invariant, and provides precisely two fixed points; the north pole, 𝔫S2\mathfrak{n}\in S^{2} featuring f(𝔫)=1f(\mathfrak{n})=1, and the south pole, 𝔰S2\mathfrak{s}\in S^{2} where f(𝔰)=1f(\mathfrak{s})=-1. For small ε\varepsilon, we can expect to see iterates of two embedded Reeb orbits, denoted γ𝔫\gamma_{\mathfrak{n}} and γ𝔰\gamma_{\mathfrak{s}}, of λG,ε:=(1+ε𝔭fH)λG\lambda_{G,\varepsilon}:=(1+\varepsilon\mathfrak{p}^{*}f_{H})\lambda_{G} in S3/GS^{3}/G as the only generators of the filtered chain groups. Both γ𝔫\gamma_{\mathfrak{n}} and γ𝔰\gamma_{\mathfrak{s}} are elliptic and parametrize the exceptional fibers in S3/GS^{3}/G over the two orbifold points of S2/HS^{2}/H.

Select NN\in{\mathbb{N}}. Lemma 2.15 produces an εN>0\varepsilon_{N}>0 for which if ε(0,εN]\varepsilon\in(0,\varepsilon_{N}], then all orbits in 𝒫LN(λG,ε)\mathcal{P}^{L_{N}}(\lambda_{G,\varepsilon}) are nondegenerate and are iterates of γ𝔰\gamma_{\mathfrak{s}} or γ𝔫\gamma_{\mathfrak{n}}, whose actions satisfy

𝒜(γ𝔰k)=2πk(1ε)n,𝒜(γ𝔫k)=2πk(1+ε)n.\mathcal{A}(\gamma_{\mathfrak{s}}^{k})=\frac{2\pi k(1-\varepsilon)}{n},\,\,\,\,\,\mathcal{A}(\gamma_{\mathfrak{n}}^{k})=\frac{2\pi k(1+\varepsilon)}{n}. (3.1)

Thus the LNL_{N}-filtered chain complex is {\mathbb{Q}}-generated by the Reeb orbits γ𝔰k\gamma_{\mathfrak{s}}^{k} and γ𝔫k\gamma_{\mathfrak{n}}^{k}, for 1k<nN1\leq k<nN. With respect to the trivialization τG\tau_{G}, the rotation numbers of γ𝔰\gamma_{\mathfrak{s}} and γ𝔫\gamma_{\mathfrak{n}} satisfy

θ𝔰=2nε1n(1ε),θ𝔫=2n+ε1n(1+ε).\theta_{\mathfrak{s}}=\frac{2}{n}-\varepsilon\frac{1}{n(1-\varepsilon)},\,\,\,\,\,\,\theta_{\mathfrak{n}}=\frac{2}{n}+\varepsilon\frac{1}{n(1+\varepsilon)}.

(See [Ne20, §2.2] for a definition of rotation numbers.) If εN\varepsilon_{N} is sufficiently small then

μCZ(γ𝔰k)=22kn1,μCZ(γ𝔫k)=22kn+1.\mu_{\operatorname{CZ}}(\gamma_{\mathfrak{s}}^{k})=2\Bigl{\lceil}\frac{2k}{n}\Bigr{\rceil}-1,\,\,\,\,\,\mu_{\operatorname{CZ}}(\gamma_{\mathfrak{n}}^{k})=2\Bigl{\lfloor}\frac{2k}{n}\Bigr{\rfloor}+1. (3.2)

For ii\in{\mathbb{Z}}, let cic_{i} denote the number of γ𝒫LN(λG,εN)\gamma\in\mathcal{P}^{L_{N}}(\lambda_{G,\varepsilon_{N}}) with |γ|=μCZ(γ)1=i|\gamma|=\mu_{\operatorname{CZ}}(\gamma)-1=i. Then:

  • c0=n1c_{0}=n-1,

  • c2i=nc_{2i}=n for 0<i<2N10<i<2N-1,

  • c4N2=n1c_{4N-2}=n-1,

  • ci=0c_{i}=0 for all other ii values.

By (3.2), if μCZ(γ𝔫k)=1\mu_{\operatorname{CZ}}(\gamma^{k}_{\mathfrak{n}})=1, then k<n/2k<n/2, so the orbit γ𝔫k\gamma^{k}_{\mathfrak{n}} is not contractible. If μCZ(γ𝔰k)=1\mu_{\operatorname{CZ}}(\gamma^{k}_{\mathfrak{s}})=1, then by (3.2), kn/2k\leq n/2 and γ𝔰k\gamma^{k}_{\mathfrak{s}} is not contractible. Thus, λN:=λG,εN\lambda_{N}:=\lambda_{G,\varepsilon_{N}} is LNL_{N}-dynamically convex and so by [HN16, Thm. 1.3],777One hypothesis of [HN16, Thm. 1.3] requires that all contractible Reeb orbits γ\gamma satisfying μCZ(γ)=3\mu_{\operatorname{CZ}}(\gamma)=3 must be embedded. This fails in our case by considering the contractible γ𝔰n\gamma_{\mathfrak{s}}^{n}, which is not embedded yet satisfies μCZ(γ𝔰n)=3\mu_{\operatorname{CZ}}(\gamma_{\mathfrak{s}}^{n})=3. In the arXiv v2 of this paper, we proved in §4.3 why we do not need this additional hypothesis. a generic choice JN𝒥(λN)J_{N}\in\mathcal{J}(\lambda_{N}) provides a well defined filtered chain complex, yielding the isomorphism

CHLN(S3/G,λN,JN)\displaystyle CH_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N}) i=02N1n2[2i]i=02N2H(S2;)[2i]\displaystyle\cong\bigoplus_{i=0}^{2N-1}{\mathbb{Q}}^{n-2}[2i]\oplus\bigoplus_{i=0}^{2N-2}H_{*}(S^{2};{\mathbb{Q}})[2i]
={n1=0, 4N2n=2i,0<i<2N10else.\displaystyle=\begin{cases}{\mathbb{Q}}^{n-1}&*=0,\,4N-2\\ {\mathbb{Q}}^{n}&*=2i,0<i<2N-1\\ 0&\mbox{else.}\end{cases}

This follows from the good contributions to cic_{i}, which is 0 for odd ii, implying LN=0\partial^{L_{N}}=0. This proves Theorem 1.2 in the cyclic case, because GG is abelian and |Conj(G)|=n|\text{Conj}(G)|=n, after appealing to Theorem 4.4, which permits taking a direct limit over inclusions of these groups.

3.2 Binary dihedral groups 𝔻2n{\mathbb{D}}^{*}_{2n}

The binary dihedral group 𝔻2nSU(2){\mathbb{D}}^{*}_{2n}\subset\text{SU}(2) has order 4n4n and projects to the dihedral group 𝔻2nSO(3){\mathbb{D}}_{2n}\subset\text{SO}(3), which has order 2n2n, under the cover P:SU(2)SO(3)P:\text{SU}(2)\to\text{SO}(3). With the quantity nn understood, these groups will respectively be denoted 𝔻{\mathbb{D}}^{*} and 𝔻{\mathbb{D}}. The group 𝔻{\mathbb{D}}^{*} is generated by the two matrices

A=(ζn00ζn¯)B=(0110),A=\begin{pmatrix}\zeta_{n}&0\\ 0&\overline{\zeta_{n}}\end{pmatrix}\hskip 28.45274ptB=\begin{pmatrix}0&-1\\ 1&0\end{pmatrix},

where ζn:=exp(iπ/n)\zeta_{n}:=\exp(i\pi/n) is a primitive 2nth2n^{\text{th}} root of unity. These matrices satisfy the relations B2=An=IdB^{2}=A^{n}=-\text{Id} and BAB1=A1BAB^{-1}=A^{-1}. The group elements may be enumerated as follows:

𝔻={AkBl:0k<2n, 0l1}.{\mathbb{D}}^{*}=\{A^{k}B^{l}:0\leq k<2n,\,0\leq l\leq 1\}.

By applying (2.4), the following matrices generate 𝔻SO(3){\mathbb{D}}\subset\text{SO}(3):

a:=P(A)=(1000cos((2π/n))sin((2π/n))0sin((2π/n))cos((2π/n)))b:=P(B)=(100010001).a:=P(A)=\begin{pmatrix}1&0&0\\ 0&\cos{(2\pi/n)}&-\sin{(2\pi/n)}\\ 0&\sin{(2\pi/n)}&\cos{(2\pi/n)}\end{pmatrix}\hskip 28.45274ptb:=P(B)=\begin{pmatrix}-1&0&0\\ 0&1&0\\ 0&0&-1\end{pmatrix}.

There are three types of fixed points in S2S^{2} of the 𝔻{\mathbb{D}}-action, categorized as follows:

Morse index 0: p,k:=(0,cos(((π+2kπ)/n)),sin(((π+2kπ)/n)))Fix(𝔻)p_{-,k}:=(0,\cos{((\pi+2k\pi)/n)},\sin{((\pi+2k\pi)/n)})\in\text{Fix}({\mathbb{D}}), for k{1,,n}k\in\{1,\dots,n\}. We have that p,np_{-,n} is a fixed point of abab and p,k=akp,np_{-,k}=a^{k}\cdot p_{-,n}. Thus, p,kp_{-,k} is a fixed point of ak(ab)ak=a2k+1ba^{k}(ab)a^{-k}=a^{2k+1}b. These nn points enumerate a 𝔻{\mathbb{D}}-orbit in S2S^{2}, and so the isotropy subgroup of 𝔻{\mathbb{D}} associated to any of the p,kp_{-,k} is of order 2 and is generated by a2k+1b𝔻a^{2k+1}b\in{\mathbb{D}}. The point pS2/𝔻p_{-}\in S^{2}/{\mathbb{D}} denotes the image of any p,kp_{-,k} under π𝔻\pi_{{\mathbb{D}}}.

Morse index 1: ph,k:=(0,cos((2kπ/n)),sin((2kπ/n)))Fix(𝔻)p_{h,k}:=(0,\cos{(2k\pi/n)},\sin{(2k\pi/n)})\in\text{Fix}({\mathbb{D}}), for k{1,,n}k\in\{1,\dots,n\}. We have that ph,np_{h,n} is a fixed point of of bb and ph,k=akph,np_{h,k}=a^{k}\cdot p_{h,n}. Thus, ph,kp_{h,k} is a fixed point of ak(b)ak=a2kba^{k}(b)a^{-k}=a^{2k}b. These nn points enumerate a 𝔻{\mathbb{D}}-orbit in S2S^{2}, and so the isotropy subgroup of 𝔻{\mathbb{D}} associated to any of the ph,kp_{h,k} is of order 2 and is generated by a2kb𝔻a^{2k}b\in{\mathbb{D}}. The point phS2/𝔻p_{h}\in S^{2}/{\mathbb{D}} denotes the image of any ph,kp_{h,k} under π𝔻\pi_{{\mathbb{D}}}.

Morse index 2: p+,1=(1,0,0)p_{+,1}=(1,0,0) and p+,2=(1,0,0)p_{+,2}=(-1,0,0). These are the fixed points of aka^{k}, for 0<k<n0<k<n, and together enumerate a single two element 𝔻{\mathbb{D}}-orbit. The isotropy subgroup associated to either of the points is cyclic of order nn in 𝔻{\mathbb{D}}, generated by aa. The point p+S2/𝔻p_{+}\in S^{2}/{\mathbb{D}} denotes the image of any one of these two points under π𝔻\pi_{{\mathbb{D}}}.

There exists a 𝔻{\mathbb{D}}-invariant, Morse-Smale function ff on (S2,ωFS(,j))(S^{2},\omega_{\text{FS}}(\cdot,j\cdot)), with Crit(f)=Fix(𝔻)\text{Crit}(f)=\text{Fix}({\mathbb{D}}), which descends to an orbifold Morse function f𝔻:S2/𝔻f_{{\mathbb{D}}}:S^{2}/{\mathbb{D}}\to{\mathbb{R}}, constructed in Section 2.3. Furthermore, there are stereographic coordinates at:

  1. (i)

    the points p,kp_{-,k}, in which ff takes the form (x2+y2)/21(x^{2}+y^{2})/2-1 near (0,0)(0,0);

  2. (ii)

    the points ph,kp_{h,k}, in which ff takes the form (x2y2)/2(x^{2}-y^{2})/2 near (0,0)(0,0);

  3. (iii)

    the points p+,kp_{+,k}, in which ff takes the form 1(x2+y2)/21-(x^{2}+y^{2})/2 near (0,0)(0,0).

The orbifold surface S2/𝔻S^{2}/{\mathbb{D}} is homeomorphic to S2S^{2} and has three orbifold points. Lemma 3.1 identifies the Reeb orbits of λ𝔻,ε=(1+ε𝔭f𝔻)λ𝔻\lambda_{{\mathbb{D}}^{*},\varepsilon}=(1+\varepsilon\mathfrak{p}^{*}f_{{\mathbb{D}}})\lambda_{{\mathbb{D}}^{*}} that appear in the filtered chain complex and computes their Conley Zehnder indices.

Lemma 3.1.

Fix NN\in{\mathbb{N}}. Then there exists an εN>0\varepsilon_{N}>0 such that for all ε(0,εN]\varepsilon\in(0,\varepsilon_{N}], every γ𝒫LN(λ𝔻,ε)\gamma\in\mathcal{P}^{L_{N}}(\lambda_{{\mathbb{D}}^{*},\varepsilon}) is nondegenerate and projects to an orbifold critical point of f𝔻f_{{\mathbb{D}}} under 𝔭\mathfrak{p}, where LN=2πNπ/2nL_{N}=2\pi N-\pi/2n. If cic_{i} denotes the number of γ𝒫LN(λ𝔻,ε)\gamma\in\mathcal{P}^{L_{N}}(\lambda_{{\mathbb{D}}^{*},\varepsilon}) with |γ|=i|\gamma|=i, then

  • ci=0c_{i}=0 if i<0i<0 or i>4N2i>4N-2;

  • ci=n+2c_{i}=n+2 for i=0i=0 and i=4N2i=4N-2, with all n+2n+2 contributions by good Reeb orbits;

  • ci=n+3c_{i}=n+3 for even ii, 0<i<4N20<i<4N-2, with all n+3n+3 contributions by good Reeb orbits;

  • ci=1c_{i}=1 for odd ii, 0<i<4N20<i<4N-2, and this contribution is by a bad Reeb orbit.

Proof.

Apply Lemma 2.15 to LN=2πNπ2nL_{N}=2\pi N-\frac{\pi}{2n} to obtain εN\varepsilon_{N}. Now, if ε(0,εN]\varepsilon\in(0,\varepsilon_{N}], we have that every γ𝒫LN(λ,ε)\gamma\in\mathcal{P}^{L_{N}}(\lambda_{{\mathbb{P}}^{*},\varepsilon}) is nondegenerate and projects to an orbifold critical point of f𝔻f_{{\mathbb{D}}}. We now study the actions and indices of these orbits.

Orbits over pp_{-}: Let ee_{-} denote the embedded Reeb orbit of λ𝔻,ε\lambda_{{\mathbb{D}}^{*},\varepsilon} which projects to pS2/𝔻p_{-}\in S^{2}/{\mathbb{D}}. By Lemmas 2.14 and 2.7, e4e_{-}^{4} lifts to an embedded Reeb orbit of λε\lambda_{\varepsilon} in S3S^{3} with action 2π(1ε)2\pi(1-\varepsilon), projecting to some p,jp_{-,j}. Thus, 𝒜(e)=π(1ε)/2\mathcal{A}(e_{-})=\pi(1-\varepsilon)/2, so 𝒜(ek)=kπ(1ε)/2\mathcal{A}(e^{k}_{-})=k\pi(1-\varepsilon)/2. Hence, our chain complex will be generated by only the k=1,2,,4N1k=1,2,\dots,4N-1 iterates. Using Remark 2.13 and Proposition 2.10, we see that the linearized Reeb flow of λ𝔻,ε\lambda_{{\mathbb{D}}^{*},\varepsilon} along eke^{k}_{-} with respect to trivialization τ𝔻\tau_{{\mathbb{D}}^{*}} is given by the family of matrices Mt=(2ttε)M_{t}=\mathcal{R}(2t-t\varepsilon) for t[0,kπ(1ε)/2]t\in[0,k\pi(1-\varepsilon)/2], where we have used that f(p,j)=1f(p_{-,j})=-1 and that we have stereographic coordinates ψ\psi at the point p,jp_{-,j} such that H(f,ψ)=IdH(f,\psi)=\text{Id}. We see that eke_{-}^{k} is elliptic with rotation number θk=k/2kε/4(1ε)\theta_{-}^{k}=k/2-k\varepsilon/4(1-\varepsilon), thus

μCZ(ek)=2k2kε4(1ε)1=2k21,\mu_{\operatorname{CZ}}(e_{-}^{k})=2\Bigl{\lceil}\frac{k}{2}-\frac{k\varepsilon}{4(1-\varepsilon)}\Bigr{\rceil}-1=2\Bigl{\lceil}\frac{k}{2}\Bigr{\rceil}-1,

where the last step is valid by reducing εN\varepsilon_{N} if necessary.

Orbits over php_{h}: Let hh denote the embedded Reeb orbit of λ𝔻,ε\lambda_{{\mathbb{D}}^{*},\varepsilon} which projects to phS2/𝔻p_{h}\in S^{2}/{\mathbb{D}}. By Lemmas 2.14 and 2.7, h4h^{4} lifts to an embedded Reeb orbit of λε\lambda_{\varepsilon} in S3S^{3} with action 2π2\pi, projecting to some ph,jp_{h,j}. Thus, 𝒜(h)=π/2\mathcal{A}(h)=\pi/2, so 𝒜(hk)=kπ/2\mathcal{A}(h^{k})=k\pi/2. Hence, our chain complex will be generated only the k=1,2,,4N1k=1,2,\dots,4N-1 iterates.

To see that hh is a hyperbolic Reeb orbit, we consider its 4-fold cover h4h^{4}. By again using Remark 2.13 and Proposition 2.10, one may compute the linearized Reeb flow, noting that the lifted Reeb orbit projects to ph,jp_{h,j} where f(ph,j)=0f(p_{h,j})=0, and that we have stereographic coordinates ψ\psi at ph,jp_{h,j} such that H(f,ψ)=Diag(1,1)H(f,\psi)=\text{Diag}(1,-1). We evaluate the matrix at t=2πt=2\pi to see that the linearized return map associated to h4h^{4} is

exp(02πε2πε0)=(cosh((2πε))sinh(2πε)sinh(2πε)cosh((2πε))).\exp\begin{pmatrix}0&2\pi\varepsilon\\ 2\pi\varepsilon&0\end{pmatrix}=\begin{pmatrix}\cosh{(2\pi\varepsilon)}&\sinh{(2\pi\varepsilon)}\\ \sinh{(2\pi\varepsilon)}&\cosh{(2\pi\varepsilon)}\end{pmatrix}.

The eigenvalues of this matrix are cosh((2πε))±sinh(2πε)\cosh{(2\pi\varepsilon)}\pm\sinh{(2\pi\varepsilon)}. So long as ε\varepsilon is small, these eigenvalues are real and positive, so that h4h^{4} is positive hyperbolic, implying that hh is also hyperbolic. If μCZ(h)=I\mu_{\operatorname{CZ}}(h)=I, then by Corollary 2.11, 4I=μCZ(h4)=44I=\mu_{\operatorname{CZ}}(h^{4})=4. Hence I=1I=1 and hh is negative hyperbolic.

Orbits over p+p_{+}: Let e+e_{+} denote the embedded Reeb orbit of λ𝔻,ε\lambda_{{\mathbb{D}}^{*},\varepsilon} which projects to p+S2/𝔻p_{+}\in S^{2}/{\mathbb{D}}. By Lemmas 2.14 and 2.7, the 2n2n-fold cover e2ne_{-}^{2n} lifts to some embedded Reeb orbit of λε\lambda_{\varepsilon} in S3S^{3} with action 2π(1+ε)2\pi(1+\varepsilon), projecting to some p+,jp_{+,j}. Thus, 𝒜(e+)=π(1+ε)/n\mathcal{A}(e_{+})=\pi(1+\varepsilon)/n, so 𝒜(e+k)=kπ(1+ε)/n\mathcal{A}(e^{k}_{+})=k\pi(1+\varepsilon)/n and so our chain complex will be generated by only the k=1,2,,2nN1k=1,2,\dots,2nN-1 iterates. Using Remark 2.13 and Proposition 2.10, we see that the linearized Reeb flow of λ𝔻,ε\lambda_{{\mathbb{D}}^{*},\varepsilon} along e+ke^{k}_{+} with respect to trivialization τ𝔻\tau_{{\mathbb{D}}^{*}} is given by the family of matrices

Mt=(2t1+ε+tε(1+ε)2)fort[0,kπ(1+ε)/n],M_{t}=\mathcal{R}\bigg{(}\frac{2t}{1+\varepsilon}+\frac{t\varepsilon}{(1+\varepsilon)^{2}}\bigg{)}\,\,\,\text{for}\,\,\,t\in[0,k\pi(1+\varepsilon)/n],

where we have used that f(p+,j)=1f(p_{+,j})=1 and that we have stereographic coordinates ψ\psi at p+,jp_{+,j} such that H(f,ψ)=IdH(f,\psi)=-\text{Id}. Thus, e+ke_{+}^{k} is elliptic with

θ+k=kn+εk2n(1+ε),μCZ(e+k)=1+2kn+εk2n(1+ε)=1+2kn,\theta_{+}^{k}=\frac{k}{n}+\frac{\varepsilon k}{2n(1+\varepsilon)},\,\,\mu_{\operatorname{CZ}}(e_{+}^{k})=1+2\Bigl{\lfloor}\frac{k}{n}+\frac{\varepsilon k}{2n(1+\varepsilon)}\Bigr{\rfloor}=1+2\Bigl{\lfloor}\frac{k}{n}\Bigr{\rfloor},

where the last step is valid for sufficiently small εN\varepsilon_{N}. ∎

Lemma 3.1 produces the sequence (εN)N=1(\varepsilon_{N})_{N=1}^{\infty}, which we can assume decreases monotonically to 0 in {\mathbb{R}}. Define the sequence of 1-forms (λN)N=1(\lambda_{N})_{N=1}^{\infty} on S3/𝔻S^{3}/{\mathbb{D}}^{*} by λN:=λ𝔻,εN\lambda_{N}:=\lambda_{{\mathbb{D}}^{*},\varepsilon_{N}}.

Summary.

(Dihedral data). We have

μCZ(ek)=2k21,μCZ(hk)=k,μCZ(e+k)=2kn+1,\mu_{\operatorname{CZ}}(e_{-}^{k})=2\Bigl{\lceil}\frac{k}{2}\Bigr{\rceil}-1,\,\,\,\mu_{\operatorname{CZ}}(h^{k})=k,\,\,\,\mu_{\operatorname{CZ}}(e_{+}^{k})=2\Bigl{\lfloor}\frac{k}{n}\Bigr{\rfloor}+1, (3.3)
𝒜(ek)=kπ(1ε)2,𝒜(hk)=kπ2,𝒜(e+k)=kπ(1+ε)n.\mathcal{A}(e_{-}^{k})=\frac{k\pi(1-\varepsilon)}{2},\,\,\,\mathcal{A}(h^{k})=\frac{k\pi}{2},\,\,\,\mathcal{A}(e_{+}^{k})=\frac{k\pi(1+\varepsilon)}{n}. (3.4)
Grading Index Orbits cic_{i}
0 1 e,e2,h,e+,e+2,,e+n1e_{-},e_{-}^{2},h,e_{+},e_{+}^{2},\dots,e_{+}^{n-1} n+2n+2
1 2 h2h^{2} 1
2 3 e3,e4,h3,e+n,,e+2n1e_{-}^{3},e_{-}^{4},h^{3},e_{+}^{n},\dots,e_{+}^{2n-1} n+3n+3
4N44N-4 4N34N-3 e4N3,e4N2,h4N3,e+(2N2)n,e+(2N1)n1e_{-}^{4N-3},e_{-}^{4N-2},h^{4N-3},e_{+}^{(2N-2)n},\dots e_{+}^{(2N-1)n-1} n+3n+3
4N34N-3 4N24N-2 h4N2h^{4N-2} 1
4N24N-2 4N14N-1 e4N1,h4N1,e+(2N1)n,,e+2Nn1e_{-}^{4N-1},h^{4N-1},e_{+}^{(2N-1)n},\dots,e_{+}^{2Nn-1} n+2n+2
Table 4: Reeb orbits of 𝒫LN(λ𝔻2n,εN)\mathcal{P}^{L_{N}}(\lambda_{{\mathbb{D}}_{2n}^{*},\varepsilon_{N}}) and their Conley Zehnder indices

None of the orbits in the first two rows of Table 4 are contractible, thus λN=λ𝔻,εN\lambda_{N}=\lambda_{{\mathbb{D}}^{*},\varepsilon_{N}} is LNL_{N}-dynamically convex and by [HN16, Thm. 1.3],888One hypothesis of [HN16, Thm. 1.3] requires that all contractible Reeb orbits γ\gamma satisfying μCZ(γ)=3\mu_{\operatorname{CZ}}(\gamma)=3 must be embedded. This fails in our case by considering the contractible e4e_{-}^{4}, which is not embedded yet satisfies μCZ(e4)=3\mu_{\operatorname{CZ}}(e_{-}^{4})=3. In the arXiv v2 of this paper, we proved in §4.3 why we do not need this additional hypothesis. a generic choice JN𝒥(λN)J_{N}\in\mathcal{J}(\lambda_{N}) provides a well-defined filtered chain complex, yielding the isomorphism of {\mathbb{Z}}-graded vector spaces

CHLN(S3/𝔻,λN,JN)\displaystyle CH_{*}^{L_{N}}(S^{3}/{\mathbb{D}}^{*},\lambda_{N},J_{N}) i=02N1n+1[2i]i=02N2H(S2;)[2i]\displaystyle\cong\bigoplus_{i=0}^{2N-1}{\mathbb{Q}}^{n+1}[2i]\oplus\bigoplus_{i=0}^{2N-2}H_{*}(S^{2};{\mathbb{Q}})[2i]
={n+2=0, 4N2n+3=2i,0<i<2N10else..\displaystyle=\begin{cases}{\mathbb{Q}}^{n+2}&*=0,\,4N-2\\ {\mathbb{Q}}^{n+3}&*=2i,0<i<2N-1\\ 0&\mbox{else.}\end{cases}.

This follows from investigating the good contributions to cic_{i}, which is 0 for odd ii, implying LN=0\partial^{L_{N}}=0. This proves Theorem 1.2 in the dihedral case because |Conj(𝔻2n)|=n+3|\text{Conj}({\mathbb{D}}_{2n}^{*})|=n+3, after appealing to Theorem 4.4, which permits taking a direct limit over inclusions of these groups.

3.3 Binary polyhedral groups 𝕋{\mathbb{T}}^{*}, 𝕆{\mathbb{O}}^{*}, and 𝕀{\mathbb{I}}^{*}

In contrast to the dihedral case, we opt not work with explicit matrix generators of the polyhedral groups, because the computations of the fixed points are too involved. Instead, we will take a more geometric approach. Let SU(2){\mathbb{P}}^{*}\subset\text{SU}(2) be some binary polyhedral group so that it is congruent to either 𝕋{\mathbb{T}}^{*}, 𝕆{\mathbb{O}}^{*} or 𝕀{\mathbb{I}}^{*}, with ||=24|{\mathbb{P}}^{*}|=24, 48, or 120, respectively. Let SO(3){\mathbb{P}}\subset\text{SO}(3) denote the image of {\mathbb{P}}^{*} under the group homomorphism PP. This group {\mathbb{P}} is conjugate to one of 𝕋{\mathbb{T}}, 𝕆{\mathbb{O}}, or 𝕀{\mathbb{I}} in SO(3)\text{SO}(3), and its order satisfies ||=2|||{\mathbb{P}}^{*}|=2|{\mathbb{P}}|. It is known that the {\mathbb{P}} action on S2S^{2} is given by the symmetries of a regular polyhedron inscribed in S2S^{2}. The fixed point set Fix()\text{Fix}({\mathbb{P}}) is partitioned into three {\mathbb{P}}-orbits. Let the number of vertices, edges, and faces of the polyhedron in question be 𝒱\mathscr{V}, \mathscr{E}, and \mathscr{F} respectively (see Table 5).

Vertex type fixed points: The set {𝔳1,𝔳2,,𝔳𝒱}Fix()\{\mathfrak{v}_{1},\mathfrak{v}_{2},\dots,\mathfrak{v}_{\mathscr{V}}\}\subset\text{Fix}({\mathbb{P}}) constitutes a single {\mathbb{P}}-orbit, where each 𝔳i\mathfrak{v}_{i} is an inscribed vertex of the polyhedron in S2S^{2}. Let 𝒱\mathscr{I}_{\mathscr{V}}\in{\mathbb{N}} denote ||/𝒱|{\mathbb{P}}|/\mathscr{V}, so that the isotropy subgroup associated to any of the 𝔳i\mathfrak{v}_{i} is cyclic of order 𝒱\mathscr{I}_{\mathscr{V}}. Let 𝔳S2/\mathfrak{v}\in S^{2}/{\mathbb{P}} denote the image of any of the 𝔳i\mathfrak{v}_{i} under the orbifold covering map π:S2S2/\pi_{{\mathbb{P}}}:S^{2}\to S^{2}/{\mathbb{P}}.

Edge type fixed points: The set {𝔢1,𝔢2,,𝔢}Fix()\{\mathfrak{e}_{1},\mathfrak{e}_{2},\dots,\mathfrak{e}_{\mathscr{E}}\}\subset\text{Fix}({\mathbb{P}}) constitutes a single {\mathbb{P}}-orbit, where each 𝔢i\mathfrak{e}_{i} is the image of a midpoint of one of the edges of the polyhedron under the radial projection 3{0}S2{\mathbb{R}}^{3}\setminus\{0\}\to S^{2}. Let \mathscr{I}_{\mathscr{E}}\in{\mathbb{N}} denote ||/|{\mathbb{P}}|/\mathscr{E}, so that the isotropy subgroup associated to any of the 𝔢i\mathfrak{e}_{i} is cyclic of order \mathscr{I}_{\mathscr{E}}. One can see that =2\mathscr{I}_{\mathscr{E}}=2 for any choice of {\mathbb{P}}. Let 𝔢S2/\mathfrak{e}\in S^{2}/{\mathbb{P}} denote the image of any of the 𝔢i\mathfrak{e}_{i} under the orbifold covering map π:S2S2/\pi_{{\mathbb{P}}}:S^{2}\to S^{2}/{\mathbb{P}}.

Face type fixed points: The set {𝔣1,𝔣2,,𝔣}Fix()\{\mathfrak{f}_{1},\mathfrak{f}_{2},\dots,\mathfrak{f}_{\mathscr{F}}\}\subset\text{Fix}({\mathbb{P}}) constitutes a single {\mathbb{P}}-orbit, where each 𝔣i\mathfrak{f}_{i} is the image of a barycenter of one of the faces of the polyhedron under the radial projection 3{0}S2{\mathbb{R}}^{3}\setminus\{0\}\to S^{2}. Let \mathscr{I}_{\mathscr{F}} denote ||/|{\mathbb{P}}|/\mathscr{F}, so that the isotropy subgroup associated to any of the 𝔣i\mathfrak{f}_{i} is cyclic of order \mathscr{I}_{\mathscr{F}}. One can see that =3\mathscr{I}_{\mathscr{F}}=3 for any choice of {\mathbb{P}}. Let 𝔣S2/\mathfrak{f}\in S^{2}/{\mathbb{P}} denote the image of any of the 𝔣i\mathfrak{f}_{i} under the orbifold covering map π:S2S2/\pi_{{\mathbb{P}}}:S^{2}\to S^{2}/{\mathbb{P}}.

Group Group order 𝒱\mathscr{V} \mathscr{E} \mathscr{F} 𝒱\mathscr{I}_{\mathscr{V}} \mathscr{I}_{\mathscr{E}} \mathscr{I}_{\mathscr{F}} |Conj()||\text{Conj}({\mathbb{P}}^{*})|
𝕋{\mathbb{T}} 12 4 6 4 3 2 3 7
𝕆{\mathbb{O}} 24 6 12 8 4 2 3 8
𝕀{\mathbb{I}} 60 12 30 20 5 2 3 9
Table 5: Polyhedral quantities. Note |Cong()|=𝒱++1.|\text{Cong}({\mathbb{P}}^{*})|=\mathscr{I}_{\mathscr{V}}+\mathscr{I}_{\mathscr{E}}+\mathscr{I}_{\mathscr{F}}-1.
Remark 3.2.

(Dependence on choice of {\mathbb{P}}^{*}). The coordinates of the fixed point set of {\mathbb{P}} are determined by the initial selection of SU(2){\mathbb{P}}^{*}\subset\text{SU}(2). More precisely, if A12A=1A^{-1}{\mathbb{P}}_{2}^{*}A={\mathbb{P}}_{1}^{*} for ASU(2)A\in\text{SU}(2), then the rigid motion of 3{\mathbb{R}}^{3} given by P(A)SO(3)P(A)\in\text{SO}(3) takes the fixed point set of 1{\mathbb{P}}_{1} to that of 2{\mathbb{P}}_{2}.

There exists a {\mathbb{P}}-invariant, Morse-Smale function ff on (S2,ωFS(,j))(S^{2},\omega_{\text{FS}}(\cdot,j\cdot)), with Crit(f)=Fix()\text{Crit}(f)=\text{Fix}({\mathbb{P}}), which descends to an orbifold Morse function f:S2/f_{{\mathbb{P}}}:S^{2}/{\mathbb{P}}\to{\mathbb{R}}, constructed in Section 2.3. Furthermore, there are stereographic coordinates at

  1. (i)

    the points 𝔳i\mathfrak{v}_{i}, in which ff takes the form (x2+y2)/21(x^{2}+y^{2})/2-1 near (0,0)(0,0);

  2. (ii)

    the points 𝔢i\mathfrak{e}_{i}, in which ff takes the form (x2y2)/2(x^{2}-y^{2})/2 near (0,0)(0,0);

  3. (iii)

    the points 𝔣i\mathfrak{f}_{i}, in which ff takes the form 1(x2+y2)/21-(x^{2}+y^{2})/2 near (0,0)(0,0).

The orbifold surface S2/S^{2}/{\mathbb{P}} is homeomorphic to S2S^{2} and has three orbifold points. Lemma 3.3 identifies the Reeb orbits of λ,ε=(1+ε𝔭f)λ\lambda_{{\mathbb{P}}^{*},\varepsilon}=(1+\varepsilon\mathfrak{p}^{*}f_{{\mathbb{P}}})\lambda_{{\mathbb{P}}^{*}} that appear in the filtered chain complex and computes their Conley Zehnder indices. Let mm\in{\mathbb{N}} denote the integer 𝒱++1\mathscr{I}_{\mathscr{V}}+\mathscr{I}_{\mathscr{E}}+\mathscr{I}_{\mathscr{F}}-1 (equivalently, m=|Conj()|m=|\text{Conj}({\mathbb{P}}^{*})|, see Table 5).

Lemma 3.3.

Fix NN\in{\mathbb{N}}. Then there exists an εN>0\varepsilon_{N}>0 such that, for all ε(0,εN]\varepsilon\in(0,\varepsilon_{N}], every γ𝒫LN(λ,ε)\gamma\in\mathcal{P}^{L_{N}}(\lambda_{{\mathbb{P}}^{*},\varepsilon}) is nondegenerate and projects to an orbifold critical point of ff_{{\mathbb{P}}} under 𝔭\mathfrak{p}, where LN:=2πNπ/10L_{N}:=2\pi N-\pi/10. If cic_{i} denotes the number of γ𝒫LN(λ,ε)\gamma\in\mathcal{P}^{L_{N}}(\lambda_{{\mathbb{P}}^{*},\varepsilon}) with |γ|=i|\gamma|=i, then

  1. 1.

    ci=0c_{i}=0 if i<0i<0 or i>4N2i>4N-2,

  2. 2.

    ci=m1c_{i}=m-1 for i=0i=0 and i=4N2i=4N-2, with all contributions by good Reeb orbits;

  3. 3.

    ci=mc_{i}=m for even ii, 1<i<4N11<i<4N-1, with all contributions by good Reeb orbits;

  4. 4.

    ci=1c_{i}=1 for odd ii, 0<i<4N20<i<4N-2, and this contribution is by a bad Reeb orbit.

Proof.

Apply Lemma 2.15 to LN=2πNπ10L_{N}=2\pi N-\frac{\pi}{10} to obtain εN\varepsilon_{N}. If ε(0,εN]\varepsilon\in(0,\varepsilon_{N}], then every γ𝒫LN(λ,ε)\gamma\in\mathcal{P}^{L_{N}}(\lambda_{{\mathbb{P}}^{*},\varepsilon}) is nondegenerate and projects to an orbifold critical point of ff_{{\mathbb{P}}}. We investigate the actions and Conley Zehnder indices of these three types of orbits. Our reasoning will largely follow that used in the proof of Lemma 3.1, and so some details will be omitted.

Orbits over 𝔳\mathfrak{v}: Let 𝒱\mathcal{V} denote the embedded Reeb orbit of λ,ε\lambda_{{\mathbb{P}}^{*},\varepsilon} in S3/S^{3}/{\mathbb{P}}^{*} which projects to 𝔳S2/\mathfrak{v}\in S^{2}/{\mathbb{P}}. One computes that 𝒜(𝒱k)=kπ(1ε)/𝒱\mathcal{A}(\mathcal{V}^{k})=k\pi(1-\varepsilon)/\mathscr{I}_{\mathscr{V}}, and so the iterates 𝒱k\mathcal{V}^{k} are included for all k<2N𝒱k<2N\mathscr{I}_{\mathscr{V}}. The orbit 𝒱k\mathcal{V}^{k} is elliptic with:

θ𝒱k=k𝒱εk2𝒱(1ε),μCZ(𝒱k)=2k𝒱1.\theta_{\mathcal{V}}^{k}=\frac{k}{\mathscr{I}_{\mathscr{V}}}-\frac{\varepsilon k}{2\mathscr{I}_{\mathscr{V}}(1-\varepsilon)},\,\,\,\,\,\mu_{\operatorname{CZ}}(\mathcal{V}^{k})=2\Bigl{\lceil}\frac{k}{\mathscr{I}_{\mathscr{V}}}\Bigr{\rceil}-1.

Orbits over 𝔢\mathfrak{e}: Let \mathcal{E} denote the embedded Reeb orbit of λ,ε\lambda_{{\mathbb{P}}^{*},\varepsilon} in S3/S^{3}/{\mathbb{P}}^{*} which projects to 𝔢S2/\mathfrak{e}\in S^{2}/{\mathbb{P}}. By a similar study of the orbit hh of Lemma 3.1, one sees that 𝒜(k)=kπ/2\mathcal{A}(\mathcal{E}^{k})=k\pi/2, so the iterates k\mathcal{E}^{k} are included for all k<4Nk<4N. Like the dihedral Reeb orbit hh, \mathcal{E} is negative hyperbolic with μCZ()=1\mu_{\operatorname{CZ}}(\mathcal{E})=1, thus μCZ(k)=k\mu_{\operatorname{CZ}}(\mathcal{E}^{k})=k. The even iterates of \mathcal{E} are bad Reeb orbits.

Orbits over 𝔣\mathfrak{f}: Let \mathcal{F} denote the embedded Reeb orbit of λ,ε\lambda_{{\mathbb{P}}^{*},\varepsilon} in S3/S^{3}/{\mathbb{P}}^{*} which projects to 𝔣S2/\mathfrak{f}\in S^{2}/{\mathbb{P}}. One computes that 𝒜(k)=kπ(1+ε)/Z\mathcal{A}(\mathcal{F}^{k})=k\pi(1+\varepsilon)/Z, and so the iterates k\mathcal{F}^{k} are included for all k<6Nk<6N. The orbit k\mathcal{F}^{k} is elliptic with:

θk=k3+εk6(1+ε),μCZ(k)=2k3+1=2k3+1.\theta_{\mathcal{F}}^{k}=\frac{k}{3}+\frac{\varepsilon k}{6(1+\varepsilon)},\,\,\,\,\,\mu_{\operatorname{CZ}}(\mathcal{F}^{k})=2\Bigl{\lfloor}\frac{k}{3}\Bigr{\rfloor}+1=2\Bigl{\lfloor}\frac{k}{3}\Bigr{\rfloor}+1.

Lemma 3.3 produces the sequence (εN)N=1(\varepsilon_{N})_{N=1}^{\infty}. Define the sequence of 1-forms (λN)N=1(\lambda_{N})_{N=1}^{\infty} on S3/S^{3}/{\mathbb{P}}^{*} by λN:=λ,εN\lambda_{N}:=\lambda_{{\mathbb{P}}^{*},\varepsilon_{N}}.

Summary.

(Polyhedral data). We have

μCZ(𝒱k)=2k𝒱1,μCZ(k)=k,μCZ(k)=2k3+1,\mu_{\operatorname{CZ}}(\mathcal{V}^{k})=2\Bigl{\lceil}\frac{k}{\mathscr{I}_{\mathscr{V}}}\Bigr{\rceil}-1,\,\,\,\mu_{\operatorname{CZ}}(\mathcal{E}^{k})=k,\,\,\,\mu_{\operatorname{CZ}}(\mathcal{F}^{k})=2\Bigl{\lfloor}\frac{k}{3}\Bigr{\rfloor}+1, (3.5)
𝒜(𝒱k)=kπ(1ε)𝒱,𝒜(k)=kπ2,𝒜(k)=kπ(1+ε)3.\mathcal{A}(\mathcal{V}^{k})=\frac{k\pi(1-\varepsilon)}{\mathscr{I}_{\mathscr{V}}},\,\,\,\mathcal{A}(\mathcal{E}^{k})=\frac{k\pi}{2},\,\,\,\mathcal{A}(\mathcal{F}^{k})=\frac{k\pi(1+\varepsilon)}{3}. (3.6)
Grading Index Orbits cic_{i}
0 1 𝒱,,𝒱𝒱,,,2\mathcal{V},\dots,\mathcal{V}^{\mathscr{I}_{\mathscr{V}}},\mathcal{E},\mathcal{F},\mathcal{F}^{2} m1m-1
1 2 2\mathcal{E}^{2} 1
2 3 𝒱𝒱+1,,𝒱2𝒱,3,3,4,5\mathcal{V}^{\mathscr{I}_{\mathscr{V}}+1},\dots,\mathcal{V}^{2\mathscr{I}_{\mathscr{V}}},\mathcal{E}^{3},\mathcal{F}^{3},\mathcal{F}^{4},\mathcal{F}^{5} mm
4N44N-4 4N34N-3 𝒱(2N2)𝒱+1,,𝒱(2N1)𝒱,4N3,6N6,6N5,6N4\mathcal{V}^{(2N-2)\mathscr{I}_{\mathscr{V}}+1},\dots,\mathcal{V}^{(2N-1)\mathscr{I}_{\mathscr{V}}},\mathcal{E}^{4N-3},\mathcal{F}^{6N-6},\mathcal{F}^{6N-5},\mathcal{F}^{6N-4} mm
4N34N-3 4N24N-2 4N2\mathcal{E}^{4N-2} 1
4N24N-2 4N14N-1 𝒱(2N1)𝒱+1,,𝒱2N𝒱1,4N1,6N3,6N2,6N1\mathcal{V}^{(2N-1)\mathscr{I}_{\mathscr{V}}+1},\dots,\mathcal{V}^{2N\mathscr{I}_{\mathscr{V}}-1},\mathcal{E}^{4N-1},\mathcal{F}^{6N-3},\mathcal{F}^{6N-2},\mathcal{F}^{6N-1} m1m-1
Table 6: Reeb orbits of 𝒫LN(λ,εN)\mathcal{P}^{L_{N}}(\lambda_{{\mathbb{P}}^{*},\varepsilon_{N}}) and their Conley Zehnder indices

None of the orbits in the first two rows of Table 6 are contractible, so λN=λ,εN\lambda_{N}=\lambda_{{\mathbb{P}}^{*},\varepsilon_{N}} is LNL_{N}-dynamically convex and so by [HN16, Theorem 1.3],999One hypothesis of [HN16, Th. 1.3] requires that all contractible Reeb orbits γ\gamma satisfying μCZ(γ)=3\mu_{\operatorname{CZ}}(\gamma)=3 must be embedded. This fails in our case by considering the contractible 𝒱2𝒱\mathcal{V}^{2\mathscr{I}_{\mathscr{V}}}, which is not embedded yet satisfies μCZ(𝒱2𝒱)=3\mu_{\operatorname{CZ}}(\mathcal{V}^{2\mathscr{I}_{\mathscr{V}}})=3. In the arXiv v2 of this paper, we proved in §4.3 why we do not need this additional hypothesis. a generic choice JN𝒥(λN)J_{N}\in\mathcal{J}(\lambda_{N}) provides a well defined filtered chain complex, yielding the isomorphism of {\mathbb{Z}}-graded vector spaces

CHLN(S3/,λN,JN)\displaystyle CH_{*}^{L_{N}}(S^{3}/{\mathbb{P}}^{*},\lambda_{N},J_{N}) i=02N1m2[2i]i=02N2H(S2;)[2i]\displaystyle\cong\bigoplus_{i=0}^{2N-1}{\mathbb{Q}}^{m-2}[2i]\oplus\bigoplus_{i=0}^{2N-2}H_{*}(S^{2};{\mathbb{Q}})[2i]
={m1=0, 4N2m=2i,0<i<2N10else.\displaystyle=\begin{cases}{\mathbb{Q}}^{m-1}&*=0,\,4N-2\\ {\mathbb{Q}}^{m}&*=2i,0<i<2N-1\\ 0&\mbox{else.}\end{cases}

This follows from investigating the good contributions to cic_{i}, which is 0 for odd ii, implying LN=0\partial^{L_{N}}=0. This proves Theorem 1.2 in the polyhedral case because |Conj()|=m|\text{Conj}({\mathbb{P}}^{*})|=m, after appealing to Theorem 4.4, which permits taking a direct limit over inclusions of these groups.

4 Direct limits of filtered cylindrical contact homology

In the previous section we computed the action filtered cylindrical contact homology of the links of the simple singularities. For any finite, nontrivial subgroup GSU(2)G\subset\text{SU}(2), we have a sequence (LN,λN,JN)N=1(L_{N},\lambda_{N},J_{N})_{N=1}^{\infty}, where LNL_{N}\to\infty monotonically in {\mathbb{R}}, λN\lambda_{N} is an LNL_{N}-dynamically convex contact form on S3/GS^{3}/G with kernel ξG\xi_{G}, and JN𝒥(λN)J_{N}\in\mathcal{J}(\lambda_{N}) is generically chosen so that

CHLN(S3/G,λN,JN)i=02N1m2[2i]i=02N2H(S2;)[2i],CH_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N})\cong\bigoplus_{i=0}^{2N-1}{\mathbb{Q}}^{m-2}[2i]\oplus\bigoplus_{i=0}^{2N-2}H_{*}(S^{2};{\mathbb{Q}})[2i], (4.1)

where m=|Cong(G)|m=|\text{Cong}(G)|. For NMN\leq M, there is a natural inclusion of the vector spaces on the right hand side of (4.1).

This section establishes Theorem 4.4, yielding that exact symplectic cobordisms obtained from decreasing the Morse-Bott perturbation induce well-defined maps on filtered homology, which can be identified with these inclusions, which completes the proof of Theorem 1.2 as

limNCHLN(S3/G,λN,JN)i0m2[2i]i0H(S2;)[2i].\varinjlim_{N}CH_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N})\cong\bigoplus_{i\geq 0}{\mathbb{Q}}^{m-2}[2i]\oplus\bigoplus_{i\geq 0}H_{*}(S^{2};{\mathbb{Q}})[2i].

We first explain the cobordisms and the maps they induce.

Definition 4.1.

An exact symplectic cobordism is from (Y+,λ+)(Y_{+},\lambda_{+}) to (Y,λ)(Y_{-},\lambda_{-}) is a pair (X¯,λ)(\overline{X},\lambda) where X¯\overline{X} is a compact symplectic manifold with boundary X¯=Y+Y\partial\overline{X}=Y_{+}-Y_{-} and dλd\lambda is a symplectic form on X¯\overline{X} with λ|Y±=λ±\lambda|_{Y_{\pm}}=\lambda_{\pm}. Given an exact symplectic cobordism (X¯,λ)(\overline{X},\lambda), we form its completion

X=((,0]×Y)YX¯Y+([0,)×Y+){X}=((-\infty,0]\times Y_{-})\sqcup_{Y_{-}}\overline{X}\sqcup_{Y_{+}}([0,\infty)\times Y_{+})

using the gluing under the following identifications. A neighborhood of Y+Y_{+} in (X¯,λ)(\overline{X},\lambda) can be canonically identified with (ϵ,0]s×Y+(-\epsilon,0]_{s}\times Y_{+} for some ϵ>0\epsilon>0 so that λ\lambda is identified with esλ+e^{s}\lambda_{+}. Moreover, this identification is defined so that s\partial_{s} corresponds to the unique vector field VV such that ιVdλ=λ\iota_{V}d\lambda=\lambda. Similarly, a neighborhood of YY_{-} can be canonically identified with [0,ϵ)×Y[0,\epsilon)\times Y_{-} so that λ\lambda is identified with esλe^{s}\lambda_{-}.

An almost complex structure JJ on the completion X{X} is said to be cobordism compatible if JJ is dλd\lambda-compatible on XX (meaning that dλ(,J)d\lambda(\cdot,J\cdot) is a Riemannian metric on XX), and there are λ±\lambda_{\pm}-compatible almost complex structures J±J_{\pm} on ×Y±{\mathbb{R}}\times Y_{\pm} such that JJ agrees with J+J_{+} on [0,)×Y+[0,\infty)\times Y_{+} and with JJ_{-} on (,0]×Y(-\infty,0]\times Y_{-}.

We will make use of some further notation.

Notation 4.2.

For γ+𝒫LN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}(\lambda_{N}) and γ𝒫LM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}(\lambda_{M}), we write γ+γ\gamma_{+}\sim\gamma_{-} whenever m(γ+)=m(γ)m(\gamma_{+})=m(\gamma_{-}) and both orbits project to the same orbifold point under 𝔭\mathfrak{p}. If either condition does not hold, we write γ+γ\gamma_{+}\nsim\gamma_{-}. Note that \sim defines an equivalence relation on the disjoint union N𝒫LN(λN)\bigsqcup_{N\in{\mathbb{N}}}\mathcal{P}^{L_{N}}(\lambda_{N}). If γ\gamma is contractible in S3/GS^{3}/G, then we write [γ]=0[\gamma]=0 in [S1,S3/G][S^{1},S^{3}/G].

In the following, we consider cobordism maps ΦNM\Phi^{M}_{N} induced by decreasing the Morse-Bott perturbation. We more precisely define the exact completed cobordism and space of cobordism compatible almost complex structures that we will use below.

Definition 4.3.

We want to ensure that there are ‘obvious’ index zero cylinders, which we henceforth refer to as cobordism trivial cylinders connecting the nondegenerate Reeb orbits γ+𝒫LN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}(\lambda_{N}) and γ𝒫LM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}(\lambda_{M}) satisfying γ+γ\gamma_{+}\sim\gamma_{-}. To do so, we will restrict to the following contractible space of JJ satisfying the following conditions.

Since kerλN=kerλM\ker\lambda_{N}=\ker\lambda_{M}, we know λN=egNλM\lambda_{N}=e^{g_{N}}\lambda_{M} for some gN(Y,)g_{N}\in{\mathbb{C}}^{\infty}(Y,{\mathbb{R}}). We may assume gN>0g_{N}>0 everywhere. Let g(×Y,)g\in{\mathbb{C}}^{\infty}({\mathbb{R}}\times Y,{\mathbb{R}}) such that

  • g(s,y)=sg(s,y)=s for s(,ς)s\in(-\infty,\varsigma) for some ς>0\varsigma>0;

  • g(s,y)=gN(y)+s1g(s,y)=g_{N}(y)+s-1 for s(1ς,)s\in(1-\varsigma,\infty) for some ς>0\varsigma>0;

  • sg>0\partial_{s}g>0.

Since d(eg(s,)λM)d(e^{g(s,\cdot)}\lambda_{M}) is symplectic, ([0,1]×S3/G,eg(s,)λM)([0,1]\times S^{3}/G,e^{g(s,\cdot)}\lambda_{M}) is an exact symplectic cobordism from (S3/G,λN)(S^{3}/G,\lambda_{N}) to (S3/G,λM)(S^{3}/G,\lambda_{M}).

Let JJ be a cobordism compatible almost complex structure on ×S3/G{\mathbb{R}}\times S^{3}/G, which agrees with JNJ_{N}, a generic λN\lambda_{N}-compatible almost complex structure on [1,)×S3/G[1,\infty)\times S^{3}/G, and with JMJ_{M}, a generic λM\lambda_{M}-compatible almost complex structure on (,0]×S3/G(-\infty,0]\times S^{3}/G. We can choose JJ so that the cobordism trivial cylinder connecting γ+\gamma_{+} to γ\gamma_{-}, where γ+γ\gamma_{+}\sim\gamma_{-} is a JJ-holomorphic curve because [1,)×γ+[1,\infty)\times{\gamma_{+}} is a JNJ_{N}-holomorphic submanifold, (,0]×γ(-\infty,0]\times\gamma_{-} is a JMJ_{M}-holomorphic submanifold, and the compact portion connecting them, defined by the union of the Reeb orbits of eg(s,)λMe^{g(s,\cdot)}\lambda_{M} in the same equivalence class, is a symplectic submanifold of ([0,1]×S3/G,eg(s,)λM)([0,1]\times S^{3}/G,e^{g(s,\cdot)}\lambda_{M}). Note that the space of JJ satisfying these conditions is contractible.

Next we consider the space of Fredholm index zero JJ-holomorphic cylinders in XX, denoted by 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}), where JJ is subject to the conditions in Definition 4.3, γ+𝒫good(λN)\gamma_{+}\in\mathcal{P}_{\text{good}}(\lambda_{N}), and γ𝒫good(λM)\gamma_{-}\in\mathcal{P}_{\text{good}}(\lambda_{M}). We will show that 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) is a compact 0-manifold, which is nonempty only when γ+γ\gamma_{+}\sim\gamma_{-} and that #0J(γ+,γ)=1\#\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-})=1. From this, the map ΦNM\Phi_{N}^{M}, defined by

ΦNM:CC(S3/G,λN,JN)CC(S3/G,λM,JM),ΦNM(γ+),γ:=u0J(γ+,γ)ϵ(u)m(γ+)m(u),\Phi_{N}^{M}:CC_{*}(S^{3}/G,\lambda_{N},J_{N})\to CC_{*}(S^{3}/G,\lambda_{M},J_{M}),\,\,\,\,\,\,\,\langle\Phi_{N}^{M}(\gamma_{+}),\gamma_{-}\rangle:=\sum_{u\in\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-})}\epsilon(u)\frac{m(\gamma_{+})}{m(u)},

is well defined. That ΦNM\Phi_{N}^{M} is a chain map follows from a careful analysis of the moduli spaces of index 1 cylinders that appear in these completed cobordisms. These chain maps induce continuation homomorphisms on the cylindrical contact homology groups. Our main result in Section 4 is the following:

Theorem 4.4.

An exact completed symplectic cobordism (X,λ,J)(X,\lambda,J) from (S3/G,λN,JN)(S^{3}/G,\lambda_{N},J_{N}) to (S3/G,λM,JM)(S^{3}/G,\lambda_{M},J_{M}) as in Definition 4.3, for NMN\leq M, induces a well defined chain map between filtered chain complexes. The induced maps on homology ΨNM:CHLN(S3/G,λN,JN)CHLM(S3/G,λM,JM)\Psi_{N}^{M}:CH_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N})\to CH_{*}^{L_{M}}(S^{3}/G,\lambda_{M},J_{M}) may be identified with the standard inclusions

i=02N1m2[2i]i=02N2H(S2;)[2i]i=02M1m2[2i]i=02M2H(S2;)[2i]\bigoplus_{i=0}^{2N-1}{\mathbb{Q}}^{m-2}[2i]\oplus\bigoplus_{i=0}^{2N-2}H_{*}(S^{2};{\mathbb{Q}})[2i]\hookrightarrow\bigoplus_{i=0}^{2M-1}{\mathbb{Q}}^{m-2}[2i]\oplus\bigoplus_{i=0}^{2M-2}H_{*}(S^{2};{\mathbb{Q}})[2i]

and form a directed system of graded {\mathbb{Q}}-vector spaces over {\mathbb{N}}.

There are two main steps in the proof of Theorem 4.4. The first is to establish compactness of the 0-dimensional moduli spaces. This is shown in Section 4.1 via the study of free homotopy classes of Reeb orbits in Section 4.3. The second is to establish that the so-called cobordism trivial cylinders are unique, namely that #0J(γ+,γ)=1\#\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-})=1. This is carried out in Section 4.2 and relies on intersection theoretic arguments of Hutchings [Hu02b, Hu09] and Siefring [Si11], as elucidated by Wendl [We20], in conjunction with establishing that the theoretical bounds on the winding of asymptotic eigenfunctions are achieved via a strengthening of results due to Hofer, Wysocki, and Zehnder [HWZ95].

The proof of Theorem 4.4 is as follows. Automatic transversality will be used in Corollary 4.8 to prove that 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) is a 0-dimensional manifold. By Proposition 4.7, 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) will be shown to be compact. From this, one concludes that 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) is a finite set for γ+𝒫goodLN(λN)\gamma_{+}\in\mathcal{P}_{\text{good}}^{L_{N}}(\lambda_{N}) and γ𝒫goodLM(λM)\gamma_{-}\in\mathcal{P}_{\text{good}}^{L_{M}}(\lambda_{M}). Thus for NMN\leq M, we have the well-defined chain map:

ΦNM:(CCLN(S3/G,λN,JN),LN)(CCLN(S3/G,λM,JM),LN).\Phi_{N}^{M}:\big{(}CC_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N}),\partial^{L_{N}}\big{)}\to\big{(}CC_{*}^{L_{N}}(S^{3}/G,\lambda_{M},J_{M}),\partial^{L_{N}}\big{)}.

The uniqueness result of Section 4.2 allows us to conclude that the finite set of cylinders in XX, equals that of JN(γ+,γ+)\mathcal{M}^{J_{N}}(\gamma_{+},\gamma_{+}) and that of JM(γ,γ)\mathcal{M}^{J_{M}}(\gamma_{-},\gamma_{-}), the moduli spaces of cylinders in the symplectizations of λN\lambda_{N} and λM\lambda_{M}, which is known to be 1, given by the contribution of a single trivial cylinder. If γ+γ\gamma_{+}\nsim\gamma_{-}, then Corollary 4.8 will imply that 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) is empty, and so ΦNM(γ+),γ=0\langle\Phi_{N}^{M}(\gamma_{+}),\gamma_{-}\rangle=0. Ultimately we conclude that, given γ+𝒫goodLN(λN)\gamma_{+}\in\mathcal{P}_{\text{good}}^{L_{N}}(\lambda_{N}), our chain map ΦNM\Phi_{N}^{M} takes the form ΦNM(γ+)=γ\Phi_{N}^{M}(\gamma_{+})=\gamma_{-}, where γ𝒫goodLN(λM)\gamma_{-}\in\mathcal{P}_{\text{good}}^{L_{N}}(\lambda_{M}) is the unique Reeb orbit satisfying γ+γ\gamma_{+}\sim\gamma_{-}.

We let ιNM\iota_{N}^{M} denote the chain map given by the inclusion of subcomplexes,

ιNM:(CCLN(S3/G,λM,JM),LN)(CCLM(S3/G,λM,JM),LM).\iota_{N}^{M}:\big{(}CC_{*}^{L_{N}}(S^{3}/G,\lambda_{M},J_{M}),\partial^{L_{N}}\big{)}\hookrightarrow\big{(}CC_{*}^{L_{M}}(S^{3}/G,\lambda_{M},J_{M}),\partial^{L_{M}}\big{)}.

The composition ιNMΦNM\iota_{N}^{M}\circ\Phi_{N}^{M} a chain map. Let ΨNM\Psi_{N}^{M} denote the map on homology induced by this composition, that is, ΨNM=(ιNMΦNM)\Psi_{N}^{M}=(\iota_{N}^{M}\circ\Phi_{N}^{M})_{*}. We see that

ΨNM:CHLN(S3/G,λN,JN)CHLM(S3/G,λM,JM)\Psi_{N}^{M}:CH_{*}^{L_{N}}(S^{3}/G,\lambda_{N},J_{N})\to CH_{*}^{L_{M}}(S^{3}/G,\lambda_{M},J_{M})

satisfies ΨNM([γ+])=[γ]\Psi_{N}^{M}([\gamma_{+}])=[\gamma_{-}] whenever γ+γ\gamma_{+}\sim\gamma_{-}, and thus ΨNM\Psi_{N}^{M} takes the form of the standard inclusions after making the identifications (4.1).

4.1 Compactness

A holomorphic building BB is a tuple (u1,u2,,un)(u_{1},u_{2},\dots,u_{n}), where each uiu_{i} is a potentially disconnected holomorphic curve in a completed (exact) symplectic cobordism equipped with a cobordism compatible almost complex structure. The curve uiu_{i} is the ithi^{\text{th}} level of BB. Each uiu_{i} has a set of positive (resp. negative) ends which are positively (resp. negatively) asymptotic to a set of Reeb orbits. For i{1,,n1}i\in\{1,\dots,n-1\}, there is a bijection between the negative ends of uiu_{i} and the positive ends of ui+1u_{i+1}, such that paired ends are asymptotic to the same Reeb orbit. The height of BB is nn. The positive ends of BB are given by the positive ends of u1u_{1} and the negative ends of BB are given by the negative ends of unu_{n}. The genus of BB is the genus of the Riemann surface SS obtained by attaching ends of the domains of the uiu_{i} to those of ui+1u_{i+1} according to the bijections; BB is connected if SS is connected. The index of BB is ind(B):=i=1nind(ui)\text{ind}(B):=\sum_{i=1}^{n}\text{ind}(u_{i}).

All buildings that we consider will have at most one level in a nontrivial exact symplectic cobordism, with the rest in a symplectization. Unless otherwise stated, we require that no level of BB be solely given in terms of a union of trivial cylinders in a symplectization, e.g. BB is without trivial levels.

Remark 4.5.

The index of a connected genus 0 building BB with one positive end at α\alpha, and with kk negative ends at β1,βk\beta_{1},...\beta_{k}, is given by:

ind(B)=k1+μCZ(α)i=1kμCZ(βi).\text{ind}(B)=k-1+\mu_{\operatorname{CZ}}(\alpha)-\sum_{i=1}^{k}\mu_{\operatorname{CZ}}(\beta_{i}). (4.2)

This fact follows from an inductive argument applied to the height of BB.

The following proposition considers the relationships between the Conley Zehnder indices and actions of a pair of Reeb orbits representing the same free homotopy class in [S1,S3/G][S^{1},S^{3}/G].

Proposition 4.6.

Suppose [γ+]=[γ][S1,S3/G][\gamma_{+}]=[\gamma_{-}]\in[S^{1},S^{3}/G] for γ+𝒫LN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}(\lambda_{N}) and γ𝒫LM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}(\lambda_{M}).

  1. (a)

    If μCZ(γ+)=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=\mu_{\operatorname{CZ}}(\gamma_{-}) then γ+γ\gamma_{+}\sim\gamma_{-}.

  2. (b)

    If μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) then 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}).

The proof of Proposition 4.6 is postponed to Section 4.3, where we will show that it holds for cyclic, dihedral, and polyhedral groups GG (Lemmas 4.23 and 4.24, and Proposition 4.29). For any exact symplectic cobordism (X,λ,J)(X,\lambda,J), Proposition 4.6 (a) implies that 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) is empty whenever γ+γ\gamma_{+}\nsim\gamma_{-}, and (b) crucially implies that there do not exist cylinders of negative Fredholm index in XX. Using Proposition 4.6, we now prove a compactness argument:

Proposition 4.7.

Fix N<MN<M, γ+𝒫LN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}(\lambda_{N}), and γ𝒫LM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}(\lambda_{M}). For n±0n_{\pm}\in{\mathbb{Z}}_{\geq 0}, consider a connected, genus zero building B=(un+,,u0,,un)B=(u_{n_{+}},\dots,u_{0},\dots,u_{-n_{-}}), where uiu_{i} is in the symplectization of λN\lambda_{N} for i>0i>0, uiu_{i} is in the symplectization of λM\lambda_{M} for i<0i<0, and u0u_{0} is in a generic, completed, exact symplectic cobordism (X,λ,J)(X,\lambda,J) from (λN,JN)(\lambda_{N},J_{N}) to (λM,JM)(\lambda_{M},J_{M}). If ind(B)=0\mbox{\em ind}(B)=0, with single positive puncture at γ+\gamma_{+} and single negative puncture at γ\gamma_{-}, then n+=n=0n_{+}=n_{-}=0 and u00J(γ+,γ)u_{0}\in\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}).

Proof.

This building BB provides the following sub-buildings, some of which may be empty:

[Uncaptioned image]
  • B+B_{+}, a building in the symplectization of λN\lambda_{N}, with no level consisting entirely of trivial cylinders, with one positive puncture at γ+\gamma_{+}, k+1k+1 negative punctures at αi𝒫LN(λN)\alpha_{i}\in\mathcal{P}^{L_{N}}(\lambda_{N}), for i=0,,ki=0,\dots,k, with [αi]=0[\alpha_{i}]=0 for i>0i>0, and [γ+]=[α0][\gamma_{+}]=[\alpha_{0}].

  • B0B_{0}, a height 1 building in the cobordism with one positive puncture at α0\alpha_{0}, l+1l+1 negative punctures at βi𝒫LN(λM)\beta_{i}\in\mathcal{P}^{L_{N}}(\lambda_{M}), for i=0,,li=0,\dots,l, with [βi]=0[\beta_{i}]=0 for i>0i>0, and [α0]=[β0][\alpha_{0}]=[\beta_{0}].

  • BiB_{i}, a building with one positive end at αi\alpha_{i}, without negative ends, for i=1,,ki=1,\dots,k.

  • CiC_{i}, with one positive end at βi\beta_{i}, without negative ends, for i=1,2,,li=1,2,\dots,l.

  • BB_{-}, a building in the symplectization of λM\lambda_{M}, with one positive puncture at β0\beta_{0} and one negative puncture at γ\gamma_{-}.

Contractible Reeb orbits are shown in red, while Reeb orbits representing the free homotopy class [γ±][\gamma_{\pm}] are shown in blue. Each sub-building is connected and has genus zero. Although no level of BB consists entirely of trivial cylinders, some of these sub-buildings may have entirely trivial levels in a symplectization. Note that 0=ind(B)0=\text{ind}(B) equals the sum of indices of the above buildings. Write 0=ind(B)=U+V+W0=\text{ind}(B)=U+V+W, where

U:=ind(B+)+i=1kind(Bi),V:=ind(B0)+i=1lind(Ci),andW:=ind(B).U:=\text{ind}(B_{+})+\sum_{i=1}^{k}\text{ind}(B_{i}),\,\,\,\,V:=\text{ind}(B_{0})+\sum_{i=1}^{l}\text{ind}(C_{i}),\,\,\text{and}\,\,W:=\text{ind}(B_{-}).

We will first argue that UU, VV, and W0W\geq 0.

To see U0U\geq 0, apply the index formula (4.2) to each summand to compute U=μCZ(γ+)μCZ(α0)U=\mu_{\operatorname{CZ}}(\gamma_{+})-\mu_{\operatorname{CZ}}(\alpha_{0}). If U<0U<0, then Proposition 4.6 (b) implies 𝒜(γ+)<𝒜(α0)\mathcal{A}(\gamma_{+})<\mathcal{A}(\alpha_{0}), which would violate the fact that action decreases along holomorphic buildings. We must have that U0U\geq 0.

To see V0V\geq 0, again apply the index formula (4.2) to find V=μCZ(α0)μCZ(β0)V=\mu_{\operatorname{CZ}}(\alpha_{0})-\mu_{\operatorname{CZ}}(\beta_{0}). Suppose V<0V<0. Now Proposition 4.6 (b) implies 𝒜(α0)<𝒜(β0)\mathcal{A}(\alpha_{0})<\mathcal{A}(\beta_{0}), contradicting the decrease of action.

To see W0W\geq 0, consider that either BB_{-} consists entirely of trivial cylinders, or it doesn’t. In the former case W=0W=0. In the latter case, [HN16, Prop. 2.8] implies that W>0W>0.

Because 0 is written as the sum of three non-negative integers, we conclude that U=V=W=0U=V=W=0. We will combine this fact with Proposition 4.6 to conclude that B±B_{\pm} are empty buildings and that B0B_{0} is a cylinder, concluding the proof.

Note that U=0U=0 implies μCZ(γ+)=μCZ(α0)\mu_{\operatorname{CZ}}(\gamma_{+})=\mu_{\operatorname{CZ}}(\alpha_{0}). Because [γ+]=[α0][\gamma_{+}]=[\alpha_{0}], Proposition 4.6 (a) implies that γ+α0\gamma_{+}\sim\alpha_{0}. This is enough to conclude γ+=α0𝒫(λN)\gamma_{+}=\alpha_{0}\in\mathcal{P}(\lambda_{N}), and importantly, 𝒜(γ+)=𝒜(α0)\mathcal{A}(\gamma_{+})=\mathcal{A}(\alpha_{0}). Noting that 𝒜(γ+)i=0k𝒜(αi)\mathcal{A}(\gamma_{+})\geq\sum_{i=0}^{k}\mathcal{A}(\alpha_{i}) (again, by decrease of action), we must have that k=0k=0 and that this inequality is an equality. Thus, the buildings BiB_{i} are empty for i0i\neq 0, and the building B+B_{+} has index 0 with only one negative end, α0\alpha_{0}. If B+B_{+} has some nontrivial levels then [HN16, Prop. 2.8] implies 0=ind(B+)>00=\text{ind}(B_{+})>0. Thus, B+B_{+} consists only of trivial levels. Because B+B_{+} has no trivial levels, it is empty, and n+=0n_{+}=0.

Similarly, V=0V=0 implies μCZ(α0)=μCZ(β0)\mu_{\operatorname{CZ}}(\alpha_{0})=\mu_{\operatorname{CZ}}(\beta_{0}). Again, because [α0]=[β0][\alpha_{0}]=[\beta_{0}], Proposition 4.6 (a) implies that α0β0\alpha_{0}\sim\beta_{0}. Although we cannot write α0=β0\alpha_{0}=\beta_{0}, we can conclude that the difference 𝒜(α0)𝒜(β0)\mathcal{A}(\alpha_{0})-\mathcal{A}(\beta_{0}) may be made arbitrarily small, by rescaling {εK}K=1\{\varepsilon_{K}\}_{K=1}^{\infty} by some c(0,1)c\in(0,1). Thus, the inequality 𝒜(α0)i=0l𝒜(βi)\mathcal{A}(\alpha_{0})\geq\sum_{i=0}^{l}\mathcal{A}(\beta_{i}) forces l=0l=0, implying that each CiC_{i} is empty, and that B0B_{0} has a single negative puncture at β0\beta_{0}, i.e. B0=(u0)B_{0}=(u_{0}) for u00J(α0,β0)u_{0}\in\mathcal{M}_{0}^{J}(\alpha_{0},\beta_{0}).

Finally, we consider our index 0 building BB_{-} in the symplectization of λM\lambda_{M}, with one positive end at β0\beta_{0} and one negative end at γ\gamma_{-}. Again, [HN16, Prop. 2.8] tells us that if BB_{-} has nontrivial levels, then ind(B)>0\text{ind}(B_{-})>0. Thus, all levels of BB_{-} must be trivial (this implies β0=γ\beta_{0}=\gamma_{-}). However, because the BiB_{i} and CjC_{j} are empty, a trivial level of BB_{-} is a trivial level of BB itself, contradicting our hypothesis on BB. Thus, BB_{-} is empty and n=0n_{-}=0. ∎

Corollary 4.8.

The moduli space 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) is a compact, 0-dimensional manifold for generic JJ, where γ+𝒫LN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}(\lambda_{N}) and γ𝒫LM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}(\lambda_{M}) are good, and N<MN<M. Furthermore, 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) is empty if γ+γ\gamma_{+}\nsim\gamma_{-}.

Proof.

To prove regularity of the cylinders u0J(γ+,γ)u\in\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}), we invoke automatic transversality ([Wen10, Thm. 1]), providing regularity of all such uu, because the inequality

0=ind(u)>2b2+2g(u)+h+(u)=20=\text{ind}(u)>2b-2+2g(u)+h_{+}(u)=-2

holds. Here, b=0b=0 is the number of branched points of uu over its underlying somewhere injective cylinder, g(u)=0g(u)=0 is the genus of the curve, and h+(u)h_{+}(u) is the number of ends of uu asymptotic to positive hyperbolic Reeb orbits. Because good Reeb orbits in 𝒫LN(λN)\mathcal{P}^{L_{N}}(\lambda_{N}) and 𝒫LM(λM)\mathcal{P}^{L_{M}}(\lambda_{M}) are either elliptic or negative hyperbolic, we have that h+(u)=0h_{+}(u)=0. To prove compactness of the moduli spaces, note that a sequence in 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) has a subsequence converging in the sense of [BEHWZ03] to a building with the properties detailed in Proposition 4.7. Proposition 4.7 proves that such an object is single cylinder in 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}), proving compactness of 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}). Finally, by Proposition 4.6 (a), the existence of u0J(γ+,γ)u\in\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) implies that γ+γ\gamma_{+}\sim\gamma_{-}. Thus, γ+γ\gamma_{+}\nsim\gamma_{-} implies that 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) is empty. ∎

4.2 Uniqueness of cobordism trivial cylinders

The purpose of this section is to prove uniqueness of cobordism trivial cylinders. The main technical obstruction to carrying out a proof of this is that in the N=N=\infty Morse-Bott limit, these cylinders are no longer isolated but come in families, so one cannot naively invoke [Wen-SFT, Thm. 10.32]. However, since we are in dimension four, we can appeal to some intersection theory due to Hutchings and Siefring, as summarized by Wendl to establish the desired result, namely that #0J(γ+,γ)=1\#\mathcal{M}^{J}_{0}(\gamma_{+},\gamma_{-})=1. Our arguments rely on the special situation at hand, because the theoretical bounds on the winding of asymptotic eigenfunctions controlling the approach of the cylinders at infinity is achieved, by way of a strengthening of results originating in the work of Hofer, Wysocki, and Zehnder. We first give a brief, mostly contained summary of the results utilized, and then combine them give the proof of our main result below.

Proposition 4.9.

Let (X,λ,J)(X,\lambda,J) be an exact completed symplectic cobordism from (S3/G,λN,JN)(S^{3}/G,\lambda_{N},J_{N}) to (S3/G,λM,JM)(S^{3}/G,\lambda_{M},J_{M}) as in Definition 4.3, for NMN\leq M. Then #0J(γ+,γ)=1\#\mathcal{M}^{J}_{0}(\gamma_{+},\gamma_{-})=1.

First, we recall some results about the asymptotics of holomorphic curves and the winding of the associated eigenfunctions. Next we give a mostly self contained review of intersection theory. Finally, we put everything together to give a proof of uniqueness of cobordism trivial cylinders.

4.2.1 Winding of asymptotic eigenfunctions

Let γ\gamma be a nondegenerate embedded Reeb orbit. W define the asymptotic operator Lγ:C(γξ)C(γξ)L_{\gamma}:C^{\infty}(\gamma^{*}\xi)\longrightarrow C^{\infty}(\gamma^{*}\xi) from the space of smooth sections of ξ|γd\xi|_{\gamma^{d}} to itself by

Lγ=Jγ(t)tR,L_{\gamma}=-J_{\gamma(t)}\nabla_{t}^{R},

where R\nabla^{R} denotes the connection on ξ|γd\xi|_{\gamma^{d}} defined by the linearized Reeb flow along γ\gamma. The operator LL is symmetric and so its eigenvalues are real. Nondegeneracy of the Reeb orbit γ\gamma implies that γξ\gamma^{*}\xi does not have any nonzero section which is parallel with respect to R\nabla^{R}, and so zero is not an eigenvalue of LL.

We first recall some facts pertaining to the spectral properties of the asymptotic operator and the winding numbers of its eigenfunctions, going back to [HWZ95, §3].

Lemma 4.10.
  1. (a)

    Every (nonzero) eigenfunction φ\varphi of LγL_{\gamma} is a nonvanishing section of γξ\gamma^{*}\xi, and hence has a well defined winding number (with respect to a chosen unitary trivialization of γξ\gamma^{*}\xi), which we denote by wind(φ)\operatorname{wind}(\varphi). Any two nontrivial eigenfunctions in the same eigenspace have the same winding number.

  2. (b)

    Let φ1\varphi_{1} and φ2\varphi_{2} be linearly independent eigenfunctions of LL with the same winding number. Then any nontrivial linear combination of φ1\varphi_{1} and φ2\varphi_{2} is a nonvanishing section of γξ\gamma^{*}\xi.

  3. (c)

    If φ1\varphi_{1} and φ2\varphi_{2} are eigenfunctions of LL with eigenvalues λ1λ2\lambda_{1}\leq\lambda_{2} then wind(φ1)wind(φ2)\operatorname{wind}(\varphi_{1}){\leq}\operatorname{wind}(\varphi_{2}).

  4. (d)

    For every ww\in{\mathbb{Z}}, LγL_{\gamma} has exactly two eigenvalues (counting multiplicity) for which the corresponding eigenfunctions have winding number equal to ww.

Definition 4.11.

Given a nondegenerate closed Reeb orbit γ\gamma and a trivialization τ:γξS1×2\tau:\gamma^{*}\xi\to S^{1}\times{\mathbb{R}}^{2}, we define

windτ:σ(Lγ)\operatorname{wind}^{\tau}:\sigma(L_{\gamma})\to{\mathbb{Z}}

by windτ(λ):=windτ(φ)\operatorname{wind}^{\tau}(\lambda):=\operatorname{wind}^{\tau}(\varphi) where φ:S12\varphi:S^{1}\to{\mathbb{R}}^{2} is the expression via τ\tau of any nontrivial eigenfunction φΓ(γξ)\varphi\in\Gamma(\gamma^{*}\xi) with Lγφ=λφL_{\gamma}\varphi=\lambda\varphi. By way of Lemma 4.10, we can sensibly define the positive / negative extremal winding numbers to be:

α+τ(γ)\displaystyle\alpha_{+}^{\tau}(\gamma) :=min{windτ(η)|ησ(Lγ)(0,)},\displaystyle:=\min\{\operatorname{wind}_{\tau}(\eta)\ |\ \eta\in\sigma(L_{\gamma})\cap(0,\infty)\},
ατ(γ)\displaystyle\alpha_{-}^{\tau}(\gamma) :=max{windτ(η)|ησ(Lγ)(,0)}.\displaystyle:=\max\{\operatorname{wind}_{\tau}(\eta)\ |\ \eta\in\sigma(L_{\gamma})\cap(-\infty,0)\}.

The parity of a Reeb orbit is given by:

p(γ):=α+τ(γ)ατ(γ),p(\gamma):=\alpha_{+}^{\tau}(\gamma)-\alpha_{-}^{\tau}(\gamma),

which is equal to 0 when γ\gamma is positive hyperbolic or an even iterate of a negative hyperbolic orbit and 1 when γ\gamma is elliptic or an odd iterate of a negative hyperbolic orbit, cf. [Wen-SFT, §3].

From [HWZ95, §3], we also have the following definition for the Conley-Zehnder index in terms of the winding invariants. For any nondegenerate Reeb orbit γ\gamma and any trivialization τ\tau of γξ\gamma^{*}\xi we have

μCZτ(γ)=2ατ(γ)+p(γ)=2α+τ(γ)p(γ)\mu_{\operatorname{CZ}}^{\tau}(\gamma)=2\alpha_{-}^{\tau}(\gamma)+p(\gamma)=2\alpha_{+}^{\tau}(\gamma)-p(\gamma) (4.3)

Next we review how the asymptotic behavior of the end of an asymptotically cylindrical JJ-holomorphic curve at a Reeb orbit is determined by an asymptotic eigenfunction of the corresponding asymptotic operator. First, to warm up and set up some notation, we recall what it means for a JJ-holomorphic half cylinders to be asymptotically cylindrical to a Reeb orbit γ\gamma. Denote the positive and negative half-cylinders by

Z+:=[0,)×S1,Z:=(,0]×S1.Z_{+}:=[0,\infty)\times S^{1},\ \ \ Z_{-}:=(-\infty,0]\times S^{1}.

A JJ-holomorphic half-cylinder u:Z±×Mu:Z_{\pm}\to{\mathbb{R}}\times M is said to be (positively / negatively) asymptotic to a nondegenerate TT-periodic Reeb orbit γ\gamma if, after a positive reparametrization near infinity,

u(s,t)=exp(Ts,γ(t))η(s,t) for |s| large,u(s,t)=\exp_{(Ts,\gamma(t))}\eta(s,t)\ \ \ \mbox{ for $|s|$ large}, (4.4)

where the exponential map is assumed to be translation invariant and η(s,t)\eta(s,t) is a vector field along the trivial cylinder101010The trivial cylinder is the JJ-holomorphic “orbit cylinder” ×γ{\mathbb{R}}\times\gamma where JJ is {\mathbb{R}}-invariant, which is expressed as u:×S1×Y:(s,t)(Ts,γ(Tt))u:{\mathbb{R}}\times S^{1}\to{\mathbb{R}}\times Y:(s,t)\mapsto(Ts,\gamma(Tt)), where TT is the period of the orbit γ\gamma. with η(s,)0\eta(s,\cdot)\to 0 in C(S1)C^{\infty}(S^{1}) as s±s\to\pm\infty. Thus as |s||s| approaches infinity, u(s,t)u(s,t) becomes CC^{\infty} close to the trivial cylinder, which is an immersion with normal bundle equivalent to γξ.\gamma^{*}\xi.

After a further reparametrization of Z±Z_{\pm}, we can arrange for (4.4) to hold for a unique section ηu(s,t)ξγ(t)\eta_{u}(s,t)\in\xi_{\gamma(t)}; the section ηu(s,t)\eta_{u}(s,t) is called the asymptotic representative of uu. The uniqueness of η\eta depends on the choice of parametrization of the Reeb orbit γ:S1Y\gamma:S^{1}\to Y; different choices change ηu\eta_{u} by a shift in the tt-coordinate.

Proposition 4.12.

[HWZ96], [HT09, Prop. 2.4] Suppose u:Z±×Yu:Z_{\pm}\to{\mathbb{R}}\times Y is a JJ-holomorphic half cylinder which is positively/negatively asymptotic to the nondegenerate Reeb orbit γ:/M\gamma:{\mathbb{R}}/{\mathbb{Z}}\to M and let η(s,t)ξγ(t)\eta(s,t)\in\xi_{\gamma(t)} denote its asymptotic representative. Then if η\eta is not identically zero, there exists a unique nontrivial eigenfunction φ\varphi of LγL_{\gamma} with

Lγφ=λφ,±λ<0L_{\gamma}\varphi=\lambda\varphi,\ \ \ \pm\lambda<0

and a section r(s,t)ξγ(t)r(s,t)\in\xi_{\gamma(t)} satisfying r(s,)0r(s,\cdot)\to 0 uniformly as ss\to\infty such that for sufficiently large |s||s|,

η(s,t)=eλs(φ+r(s,t)).\eta(s,t)=e^{\lambda s}\left(\varphi+r(s,t)\right). (4.5)

In the context of Proposition 4.12, we say that uu approaches the Reeb orbit γ\gamma along the asymptotic eigenfunction φλ\varphi_{\lambda} with decay rate |λ||\lambda|. This proposition establishes the asymptotic approach between the half cylinder given by the end of uu and the trivial cylinder ×γ{\mathbb{R}}\times\gamma. A similar result holds for any two curves approaching the same orbit, and from this one can establish a lower bound on the resulting “relative” decay rate and a generalization of this asymptotic formula to “higher orders” was established in [Si08, §2].

Remark 4.13.

An important consequence of Proposition 4.12, and its relative decay rate improvement for any two curves asymptotic to the same Reeb orbit, is that any two asymptotically cylindrical JJ-holomorphic having nonidentical images must have at most finitely many intersections, cf. [We20, Cor. 3.13]

In particular, the asymptotic representative encodes the isotopy class of the braid ζ\zeta, which is given in terms of the intersection of the end of the JJ-holomorphic curve asymptotic to γd\gamma^{d} and a tubular neighborhood NN of γ\gamma for |s|0|s|\gg 0. Let N(/)×D2N\simeq({\mathbb{R}}/{\mathbb{Z}})\times D^{2} be a tubular neighborhood of γ\gamma which has been identified with a disk bundle in the normal bundle to γ\gamma as well as in γξ\gamma^{*}\xi so that γ\gamma corresponds to (/)×{0}({\mathbb{R}}/{\mathbb{Z}})\times\{0\} and the derivative of the identification along γ\gamma sends γξ\gamma^{*}\xi to {0}\{0\}\oplus{\mathbb{C}} in agreement with the unitary trivialization τ\tau. Let uu be an asymptotically cylindrical JJ-holomorphic curve with a positive (or negative) end at γd\gamma^{d}, which is not part of a multiply covered component. Then results of Siefring [Si08, Cor. 2.5, 2.6] allow us to conclude that if |s||s| is sufficiently large then the intersection of this end of uu with {s}×N{s}×Y\{s\}\times N\subset\{s\}\times Y is a braid ζ\zeta, whose isotopy class is independent of ss.

We conclude by providing the following bounds on the winding number in terms of the Conley-Zehnder index. We deduce that the ends of the JJ-holomorphic curves in 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}) meet the extremal bounds, by way of a mild generalization of arguments previously appearing in [HWZ95, §3], [HN16, Lem. 3.2], [HT09, §3].

Lemma 4.14.

[HWZ95, §3], [HN16, Lem. 3.2] Let (X,λ,J)(X,\lambda,J) be an exact completed symplectic cobordism from (S3/G,λN,JN)(S^{3}/G,\lambda_{N},J_{N}) to (S3/G,λM,JM)(S^{3}/G,\lambda_{M},J_{M}) as in Definition 4.3, for NMN\leq M. Let γ\gamma be an embedded Reeb orbit, let uu be a JJ-holomorphic curve in XX with a positive end at γd\gamma^{d} and a single negative end. Let ζ\zeta denote the intersection of an end with {s}×Y\{s\}\times Y. If ss is sufficiently large, then

  1. (a)

    ζ\zeta is the graph in NN of a nonvanishing section of ξ|γd\xi|_{\gamma^{d}} and thus has a well-defined winding number windτ(ζ)\operatorname{wind}_{\tau}(\zeta).

  2. (b)

    windτ(ζ)μCZτ(γd)/2.\operatorname{wind}_{\tau}(\zeta)\leq\lfloor\mu_{\operatorname{CZ}}^{\tau}(\gamma^{d})/2\rfloor.

  3. (c)

    If μCZτ(γd)\mu_{CZ}^{\tau}(\gamma^{d}) is odd and ind(u)=0\operatorname{ind}(u)=0, then equality holds in (b).

Proof.

Choose an identification N(/)×D2N\simeq({\mathbb{R}}/{\mathbb{Z}})\times D^{2} compatible with the trivialization τ\tau. The asymptotic behavior of pseudoholomorphic curves as described in conjunction with Lemma 4.10 above implies (a).

Hofer, Wysocki, and Zehnder demonstrated in [HWZ95, §3] that for each integer nn, there are exactly two eigenvalues of LγL_{\gamma} whose eigenfunctions have winding number nn. (Here and below we count eigenvalues with multiplicity.) Moreover, larger winding numbers correspond to larger eigenvalues, and the largest possible winding number for a negative eigenvalue is μCZτ(γd)/2\lfloor\mu_{\operatorname{CZ}}^{\tau}(\gamma^{d})/2\rfloor, thus (b) follows.

To establish (c), note that the same argument in [HWZ95, §3] also shows that the smallest possible winding number of an eigenfunction of LγL_{\gamma} with positive eigenvalue is μCZτ(γd)/2\lceil\mu_{\operatorname{CZ}}^{\tau}(\gamma^{d})/2\rceil. Since μCZτ(γd)\mu_{\operatorname{CZ}}^{\tau}(\gamma^{d}) is assumed odd, we have a strict inequality μCZτ(γd)/2<μCZτ(γd)/2\lfloor\mu_{\operatorname{CZ}}^{\tau}(\gamma^{d})/2\rfloor<\lceil\mu_{\operatorname{CZ}}^{\tau}(\gamma^{d})/2\rceil. Consequently, the two (possibly equal) eigenvalues of LγL_{\gamma} whose eigenfunctions have winding number μCZτ(γd)/2\lfloor\mu_{\operatorname{CZ}}^{\tau}(\gamma^{d})/2\rfloor are both negative. Thus, if equality does not hold in (b), then the associated eigenvalue λ\lambda is not one of the two largest negative eigenvalues of LγL_{\gamma} (counted with multiplicity as usual).

Wendl observed that one can use exponentially weighted Sobolev spaces to set up the moduli space of irreducible holomorphic curves in which the eigenvalue λ\lambda is not one of the two largest negative eigenvalues, see [HT09, Rmk. 3.3], but beware of their sign convention for the asymptotic operator.

Automatic transversality holds under the assumptions in (c), e.g. [Wen10, Thm. 1], because

0=ind(u)>2+h+(u),0=\operatorname{ind}(u)>-2+h_{+}(u),

since h+(u)=0h_{+}(u)=0, the number of ends of uu asymptotic to positive hyperbolic Reeb orbits. Thus any curves in this moduli space are cut out transversely, but the dimension of the moduli space is 22 less than usual. Consequently there are no JJ-holomorphic curves uu in this moduli space with ind(u)=0\text{ind}(u)=0.

Symmetrically to Lemma 4.14, we also have the following:

Lemma 4.15.

Let γ\gamma be an embedded Reeb orbit, let uu be a JJ-holomorphic curve in XX with a negative end at γd\gamma^{d} which is not part of a trivial cylinder in ×Y{\mathbb{R}}\times Y with JJ chosen to be {\mathbb{R}}-invariant, and a single positive end let ζ\zeta denote the intersection of this end with {s}×Y\{s\}\times Y. If s<<0s<<0, then the following hold:

  1. (a)

    ζ\zeta is the graph of a nonvanishing section of ξ|γd\xi|_{\gamma^{d}}, and thus has a well-defined winding number windτ(ζ)\operatorname{wind}_{\tau}(\zeta).

  2. (b)

    windτ(ζ)μCZτ(γd)/2\operatorname{wind}_{\tau}(\zeta)\geq\lceil\mu_{CZ}^{\tau}(\gamma^{d})/2\rceil.

  3. (c)

    If μCZτ(γd)\mu_{CZ}^{\tau}(\gamma^{d}) is odd and ind(u)=0\operatorname{ind}(u)=0, then equality holds in (b).

4.2.2 Recap of intersection theory

We now briefly review some intersection theory due to Hutchings [Hu02b, Hu09] and Siefring [Si08, Si11], which has been nicely summarized by Wendl [We20]. The Siefring intersection pairing gives a relation between homotopy invariant quantities that have geometric meanings independent of any choice of trivializations and surface representatives of the holomorphic curves. This improves upon Hutchings’ work, which defined a relative intersection pairing and provided a relative adjunction formula for somewhere injective asymptotically cylindrical curves.

Theorem 4.16 ([Si11], [We20, Thm. 4.1]).

For any two asymptotically cylindrical maps in a 4-dimensional completed symplectic cobordism with uu and uu^{\prime} with nondegenerate asymptotic Reeb orbits, there exists the so-called Siefring intersection pairing

uvu*v\in\mathbb{Z}

which satisfies the following properties.

  1. (a)

    The pairing uvu*v depends only on the homotopy classes of uu and vv as asymptotically cylindrical maps.

  2. (b)

    If uu and vv are asymptotically cylindrical JJ-holomorphic curves that have nonidentical images then

    uv:=uv+ι(u,v),u*v:=u\cdot v+\iota_{\infty}(u,v), (4.6)

    where uvu\cdot v is the usual algebraic count of intersections which is nonnegative by intersection positivity and ι(u,v)\iota_{\infty}(u,v) is a nonnegative integer realizing “hidden intersections at infinity.”

Remark 4.17.

There is the following Folk Theorem, which we do not make use of, but helps elucidate the Siefring intersection pairing, whose statement is as follows. In Theorem 4.16(b), there exists a perturbation JϵJ_{\epsilon} which is CC^{\infty} close to JJ and a pair of JϵJ_{\epsilon}-holomorphic asymptotically cylindrical curves uϵu_{\epsilon} and vϵv_{\epsilon} having the same domains as uu and vv and close to uu and vv in their respective modui spaces, such that

uϵvϵ=uv.u_{\epsilon}\cdot v_{\epsilon}=u*v.

Thus we should interpret uvu*v as the count of intersections between generic curves homotopic to uu and vv.

An important corollary of Theorem 4.16 that we will make use of is as follows.

Corollary 4.18.

If uu and vv are JJ-holomorphic curves satisfying uv=0u*v=0 then any two JJ-holomorphic curves that have nonidentical images and are homotopic to uu and vv respectively are disjoint.

The final foundational result that we need concerns nonnegativity of asymptotic intersections.

Theorem 4.19 ([We20, Thm. 4.13]).

If uu and vv are asymptotically cylindrical JJ-holomorphic maps with nonidentical images, then ι(u,v)0\iota_{\infty}(u,v)\geq 0 with equality if and only if for all pairs of ends of uu and vv respectively asymptotic to covers of the same Reeb orbit, all of the resulting relative asymptotic eigenfunctions have extremal winding.

To actually compute uvu*v, we need a few more notions, including the relative intersection number, which is due to Hutchings [Hu02b, Hu09]. The relative intersection number is denoted in the work of Siefring and Wendl by uτvu\bullet_{\tau}v. The relative intersection pairing previously appeared in Hutchings’ work as one of the a key ingredients to the definition of embedded contact homology, where it is denoted by Qτ([u],[v])Q_{\tau}([u],[v]) and was established to depend only on the homotopy class of the trivialization τ\tau and relative homology class of the curves. These intersection numbers coincide:

uτv=Qτ([u],[v]).u\bullet_{\tau}v=Q_{\tau}([u],[v]).

(We go into further details in regards to how to more precisely define and compute the relative intersection number for JJ-holomorphic cylinders à la Hutchings in the proof of Lemma 4.22.)

In the interim, a colloquial description is that the relative intersection number is an algebraic count of intersections between uu and a generic perturbation of vv so that they have at most finitely many intersections and the perturbation of vv is shifted by an arbitrarily small positive distance in directions determined by the chosen trivialization τ\tau at infinity. (One could symmetrically perturb uu and leave vv alone.)

If uu and vv have finitely many intersections then the algebraic intersection number uvu\cdot v is well-defined. In this situation,

uτv=uv+ιτ(u,v),u\bullet_{\tau}v=u\cdot v+\iota_{\infty}^{\tau}(u,v), (4.7)

where ιτ(u,v)\iota_{\infty}^{\tau}(u,v) is a count of additional intersections near infinity that appear when vv is perturbed, because the perturbation of vv can be assumed to be nontrivial only in some neighborhood of infinity where uu and vv are disjoint. (Again, the content of (4.7) previously appeared in the work of Hutchings with different notation, which we elucidate in the proof of Lemma 4.22, including the definition of ιτ(u,v)\iota_{\infty}^{\tau}(u,v).)

Modulo having deferred more precise definitions of uτvu\bullet_{\tau}v and ιτ(u,v)\iota_{\infty}^{\tau}(u,v) to later, we are nearly ready to define the trivialization independent asymptotic intersection count, ι(u,v).\iota_{\infty}(u,v). The last ingredient we need are the terms Ω±τ(γk,γm)\Omega_{\pm}^{\tau}(\gamma^{k},\gamma^{m}), which are the theoretical bounds on the possible asymptotic winding of ends of uu around the ends of vv, as discussed in Section 4.2.1, and defined below.

Definition 4.20.

Assume uu and vv are asymptotically cylindrical JJ-holomorphic curves, so that u(s,t)u(s,t) and v(s,t)v(s,t) approach their respective covers of γ\gamma (both at either a negative end or both at a positive end) along asymptotic eigenfunctions φu\varphi_{u} and φv\varphi_{v} with decay rates |λu||\lambda_{u}| and |λv||\lambda_{v}|, respectively. We define the theoretical extremal bounds at a positive / negative puncture by

Ω±τ(γk,γm):=min{kατ(γm),mατ(γk)}\Omega_{\pm}^{\tau}(\gamma^{k},\gamma^{m}):=\operatorname{min}\left\{\mp k\alpha_{\mp}^{\tau}(\gamma^{m}),\mp m\alpha_{\mp}^{\tau}(\gamma^{k})\right\}

We extend the definition of Ω±τ\Omega_{\pm}^{\tau} by setting

Ω±τ(γ1k,γ2m):=0whenever γ1γ2.\Omega_{\pm}^{\tau}(\gamma^{k}_{1},\gamma^{m}_{2}):=0\quad\mbox{whenever }\gamma_{1}\neq\gamma_{2}.

As explained in [We20, §4.2], these theoretical extremal bounds provides a universal lower bound on ιτ(u,v)\iota_{\infty}^{\tau}(u,v), namely

ιτ(u,v)(z,z)Γu±×Γv±Ω±τ(γzkz,γzkz),\iota_{\infty}^{\tau}(u,v)\geq\sum_{(z,z^{\prime})\in\Gamma_{u}^{\pm}\times\Gamma_{v}^{\pm}}\Omega_{\pm}^{\tau}(\gamma^{k_{z}}_{z},\gamma^{k_{z^{\prime}}}_{z^{\prime}}),

where Γu±\Gamma_{u}^{\pm} denote the positive  negative punctures of uu and γzkz\gamma^{k_{z}}_{z} denotes the associated Reeb orbit to which the puncture zz asymptotes, and likewise for vv.

We can now define the Siefring intersection pairing in terms of these more concrete and computable terms as follows.

Definition 4.21.

For asymptotically cylindrical maps uu and vv with finitely many intersections, we define

ι(u,v):=ιτ(u,v)(z,z)Γu±×Γv±Ω±τ(γzkz,γzkz).\iota_{\infty}(u,v):=\iota_{\infty}^{\tau}(u,v)-\sum_{(z,z^{\prime})\in\Gamma_{u}^{\pm}\times\Gamma_{v}^{\pm}}\Omega_{\pm}^{\tau}(\gamma^{k_{z}}_{z},\gamma^{k_{z^{\prime}}}_{z^{\prime}}).

and similarly for any asymptotically cylindrical maps uu and vv (with)not necessarily with finitely many intersections), we can define

uv:=uτv(z,z)Γu±×Γv±Ω±τ(γzkz,γzkz)u*v:=u\bullet_{\tau}v-\sum_{(z,z^{\prime})\in\Gamma_{u}^{\pm}\times\Gamma_{v}^{\pm}}\Omega_{\pm}^{\tau}(\gamma^{k_{z}}_{z},\gamma^{k_{z^{\prime}}}_{z^{\prime}})

Crucially, the results in Section 4.2.1 will allow us to conclude that in the case of interest, the winding numbers are extremal in the sense that they achieve the corresponding theoretical bound.

We have one last computational ingredient that we will rely on, cf. [We20, Ex. 4.19(b)], detailing some of the aforementioned notions along the way as we proceed with the proof.

Lemma 4.22.

Let (X,λ,J)(X,\lambda,J) be an exact completed symplectic cobordism from (S3/G,λN,JN)(S^{3}/G,\lambda_{N},J_{N}) to (S3/G,λM,JM)(S^{3}/G,\lambda_{M},J_{M}) as in Definition 4.3, for NMN\leq M. Let u[γ]u_{[\gamma]} be the cobordism trivial cylinder between orbits in the same equivalence class, which are further assumed to be either both elliptic or both odd iterates of a negative hyperbolic Reeb orbit. Then the Siefring self-intersection pairing is given by the multiplicity of the Reeb orbit equivalence class:

u[γ]u[γ]=m([γ]).u_{[\gamma]}*u_{[\gamma]}=-m([\gamma]). (4.8)
Proof.

By Definition 4.21

u[γ]u[γ]=u[γ]τu[γ]+Ω+τΩτ.u_{[\gamma]}*u_{[\gamma]}=u_{[\gamma]}\bullet_{\tau}u_{[\gamma]}+{\Omega_{+}^{\tau}-\Omega_{-}^{\tau}}. (4.9)

Here u[γ]τu[γ]u_{[\gamma]}\bullet_{\tau}u_{[\gamma]} is the relative intersection number, which appears in the ECH literature as Qτ(u[γ],u[γ]),Q_{\tau}(u_{[\gamma]},u_{[\gamma]}), which we detail further and compute to be 0. We compute Ω±τ\Omega_{\pm}^{\tau} momentarily in (4.10).

Using the definition of the relative intersection pairing by Hutchings [Hu09, §2.7], it follows that the relative intersection number of any cobordism trivial cylinder is 0, e.g. that

Qτ(u[γ])=u[γ]τu[γ]=0{Q_{\tau}(u_{[\gamma]})=u_{[\gamma]}\bullet_{\tau}u_{[\gamma]}=0}

To elucidate this point, recall that the relative intersection number Qτ(u)Q_{\tau}(u) of any JJ-holomorphic curve uu is a well-defined algebraic count of interior intersections of two compact oriented surfaces whose interiors are transverse and do not intersect at the boundary S,S[1,1]×YS,S^{\prime}\in[-1,1]\times Y representing [u]H2(Y,β+,β)[u]\in H_{2}(Y,\beta_{+},\beta_{-}), where [β+]=[β]H1(Y)[\beta_{+}]=[\beta_{-}]\in H_{1}(Y), whose definition in the case when uu is a cylinder we more precisely explain as follows.

If SS is one of these so called admissible representatives of a JJ-holomorphic cylinder uu (with one positive end at β+m+\beta_{+}^{m_{+}} and one negative end at βn\beta_{-}^{n_{-}}), then for ϵ\epsilon sufficiently small, S({1ϵ}×Y)S\cap(\{1-\epsilon\}\times Y) consists of a braid ζ+\zeta_{+} with m+m_{+} strands in a tubular neighborhood of β+\beta_{+}, which is well-defined up to isotopy. Likewise S({1+ϵ}×Y)S\cap(\{-1+\epsilon\}\times Y) consists of a braid ζ\zeta_{-} with mm_{-} strands in a tubular neighborhood of β\beta_{-}, which is well-defined up to isotopy. (Here β±\beta_{\pm} are embedded Reeb orbits.)

The linking number of two disjoint braids ζ1\zeta_{1} and ζ2\zeta_{2} around an embedded Reeb orbit β\beta is denoted by τ(ζ1,ζ2)\ell_{\tau}(\zeta_{1},\zeta_{2})\in\mathbb{Z} and is defined to be the linking number of the oriented links ϕτ(ζ1)\phi_{\tau}(\zeta_{1}) and ϕτ(ζ2)\phi_{\tau}(\zeta_{2}) in 3{\mathbb{R}}^{3}, where ϕτ\phi_{\tau} is an embedding of the tubular neighborhood NN of β\beta which has been identified via the trivialization τ\tau with S1×D2S^{1}\times D^{2}, so that the projection of ζi\zeta_{i} to to S1S^{1} is a submersion, which has further been identified with a solid torus in 3{\mathbb{R}}^{3}, cf. [Hu09, §2.6] for more details. The linking number is by definition, one half of the signed count of crossings of a strand of ϕτ(ζ1)\phi_{\tau}(\zeta_{1}) with a strand of ϕτ(ζ2)\phi_{\tau}(\zeta_{2}) in the projection to 2×{0}{\mathbb{R}}^{2}\times\{0\}. The sign convention is that counterclockwise twists count positively, which differs from knot theory literature [Hu02b, §3.1].

In the case of the aforementioned cylinder uu, we define the linking number

τ(S,S):=τ(ζ+,ζ+)τ(ζ,ζ);\ell_{\tau}(S,S^{\prime}):=\ell_{\tau}(\zeta_{+},\zeta_{+}^{\prime})-\ell_{\tau}(\zeta_{-},\zeta_{-}^{\prime});

for more general JJ-holomorphic curves, one obtains a collection of disjoint braids, whose linking numbers one sums over. The linking number τ\ell_{\tau} as defined above is Siefring’s relative asymptotic intersection number ιτ\iota_{\infty}^{\tau} with an opposite sign convention for the two punctures.

The relative intersection pairing is defined by

Qτ([u],[u]):=#(S˙S˙)τ(S,S),Q_{\tau}([u],[u]):=\#(\dot{S}\cap\dot{S}^{\prime})-\ell_{\tau}(S,S^{\prime}),

and one further denotes Qτ([u]):=Qτ([u],[u]).Q_{\tau}([u]):=Q_{\tau}([u],[u]). With the proceeding understood, we can now see by the definition detailed above that

Qτ(u[γ])=u[γ]τu[γ]=0{Q_{\tau}(u_{[\gamma]})=u_{[\gamma]}\bullet_{\tau}u_{[\gamma]}=0}

as claimed.

The remaining terms in (4.9), Ω±\Omega_{\pm} are the theoretical bounds on the possible relative asymptotic winding of ends of uu around itself at the appropriate puncture. By definition, for u[γ]u_{[\gamma]} we have that

Ω±τ(γz±m,γz±m):=mατ(γm).\Omega_{\pm}^{\tau}(\gamma_{z_{\pm}}^{m},\gamma_{z_{\pm}}^{m}):=\mp m\alpha_{\mp}^{\tau}(\gamma^{m}). (4.10)

Here Ω+τ\Omega_{+}^{\tau} corresponds to the positive asymptotic orbit of u[γ]u_{[\gamma]} at the puncture z+z_{+} and Ωτ\Omega_{-}^{\tau} corresponds to the negative asymptotic orbit of u[γ]u_{[\gamma]} at the puncture zz_{-}.

In the below, let

m=m([γ]).m=m([\gamma]).

Thus (4.9) can be computed to be

u[γ]u[γ]=u[γ]τu[γ]Ω+τ+Ωτ=0+mατ([γ])mα+τ([γ])=m.\begin{split}u_{[\gamma]}*u_{[\gamma]}&=u_{[\gamma]}\bullet_{\tau}u_{[\gamma]}-{\Omega_{+}^{\tau}+\Omega_{-}^{\tau}}\\ &=0+m\ \alpha_{-}^{\tau}([\gamma])-m\ \alpha_{+}^{\tau}([\gamma])\\ &=m.\end{split}

Here the last line follows from a basic manipulation of the Conley-Zehnder index formula (4.3) in terms of the positive and negative extremal winding numbers and parity of the orbit, yielding

0=2α(γ)2α+(γ)+2p(γ).0=2\alpha_{-}(\gamma)-2\alpha_{+}(\gamma)+2p(\gamma).

Here p(γ)p(\gamma) is the parity of the Reeb orbit γ\gamma, which is by definition is 1 whenever γ\gamma is elliptic or an odd multiple of a negative hyperbolic orbit. ∎

4.2.3 Proof of Proposition 4.9

We now combine Sections 4.2.1 and 4.2.2 to establish the desired uniqueness of cobordism trivial cylinders as follows.

Proof of Proposition 4.9.

In the below, denote TT to be the cobordism trivial cylinder between γ+\gamma_{+} and γ\gamma_{-}, where γ+γ\gamma_{+}\sim\gamma_{-}, γ+𝒫goodLN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}_{\text{good}}(\lambda_{N}), and γ𝒫goodLM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}_{\text{good}}(\lambda_{M}). Let vv be a curve with the same asymptotics, which is not a cobordism trivial cylinder. In particular, TT and vv have nonidentical images.

Recall that by Theorem 4.19 the count of “hidden intersections at infinity” satisfies

ι(T,v)0\iota_{\infty}(T,v)\geq 0

with equality if and only if for all pairs of ends of TT and vv respectively asymptotic to covers of the same Reeb orbit, all of the resulting relative asymptotic eigenfunctions have extremal winding. We established that the resulting relative asymptotic eigenfunctions have extremal winding in §4.2.1, cf. Lemmas 4.14 and 4.15.

Thus the Siefring intersection ‘*’ number coincides with the algebraic intersection ‘\cdot’ number:

Tv=Tv,T*v=T\cdot v,

which must be non-negative by the positivity of intersections (cf. [We20, App. B]), with equality if and only if TT and vv are disjoint by Corollary 4.18. However, the Siefring intersection number TvT*v depends only on the homotopy classes of TT and vv as asymptotically cylindrical maps by Theorem 4.16(a). We computed that all asymptotically cylindrical maps with the same asymptotics and relative homology class have negative Siefring intersection number in Lemma 4.22, a contradiction. Thus TT and vv must have identical images, hence #0J(γ+,γ)=1\#\mathcal{M}^{J}_{0}(\gamma_{+},\gamma_{-})=1, as desired. ∎

4.3 Homotopy classes of Reeb orbits and proof of Proposition 4.6

Proposition 4.6 was key in arguing compactness of 0J(γ+,γ)\mathcal{M}_{0}^{J}(\gamma_{+},\gamma_{-}). To prove Proposition 4.6, we will make use of a bijection [S1,S3/G]Conj(G)[S^{1},S^{3}/G]\cong\text{Conj}(G) to identify the free homotopy classes represented by orbits in S3/GS^{3}/G in terms of GG. A loop in S3/GS^{3}/G is a map γ:[0,T]S3/G\gamma:[0,T]\to S^{3}/G, satisfying γ(0)=γ(T)\gamma(0)=\gamma(T). Selecting a lift γ~:[0,T]S3\widetilde{\gamma}:[0,T]\to S^{3} of γ\gamma to S3S^{3} determines a unique gGg\in G, for which gγ~(0)=γ~(T)g\cdot\widetilde{\gamma}(0)=\widetilde{\gamma}(T). The map [γ][S1,S3/G][g]Conj(G)[\gamma]\in[S^{1},S^{3}/G]\mapsto[g]\in\text{Conj}(G) is well defined, bijective, and respects iterations; that is, if [γ][g][\gamma]\cong[g], then [γm][gm][\gamma^{m}]\cong[g^{m}] for mm\in{\mathbb{N}}.

4.3.1 Cyclic subgroups

We may assume that G=gnG=\langle g\rangle\cong{\mathbb{Z}}_{n}, where gg is the diagonal matrix g=Diag(ϵ,ϵ¯)g=\text{Diag}(\epsilon,\overline{\epsilon}), and ϵ:=exp(2πi/n)\epsilon:=\text{exp}{(2\pi i/n)}\in{\mathbb{C}}. We have Conj(G)={[gm]:0m<n}\text{Conj}(G)=\{[g^{m}]:0\leq m<n\}, each class is a singleton because GG is abelian. We select explicit lifts γ~𝔰\widetilde{\gamma}_{\mathfrak{s}} and γ~𝔫\widetilde{\gamma}_{\mathfrak{n}} of γ𝔰\gamma_{\mathfrak{s}} and γ𝔫\gamma_{\mathfrak{n}} to S3S^{3} given by:

γ~𝔫(t)=(eit,0),andγ~𝔰(t)=(0,eit),fort[0,2π/n].\widetilde{\gamma}_{\mathfrak{n}}(t)=(e^{it},0),\,\,\,\text{and}\,\,\,\widetilde{\gamma}_{\mathfrak{s}}(t)=(0,e^{it}),\,\,\text{for}\,\,t\in[0,2\pi/n].

Because gγ~𝔫(0)=γ~𝔫(2π/n)g\cdot\widetilde{\gamma}_{\mathfrak{n}}(0)=\widetilde{\gamma}_{\mathfrak{n}}(2\pi/n), [γ𝔫][g][\gamma_{\mathfrak{n}}]\cong[g], and thus [γ𝔫m][gm][\gamma_{\mathfrak{n}}^{m}]\cong[g^{m}] for mm\in{\mathbb{N}}. Similarly, because gn1γ~𝔰(0)=γ~𝔰(2π/n)g^{n-1}\cdot\widetilde{\gamma}_{\mathfrak{s}}(0)=\widetilde{\gamma}_{\mathfrak{s}}(2\pi/n), [γ𝔰][gn1]=[g1][\gamma_{\mathfrak{s}}]\cong[g^{n-1}]=[g^{-1}], and thus [γ𝔰m][gm][\gamma_{\mathfrak{s}}^{m}]\cong[g^{-m}] for mm\in{\mathbb{N}}.

Class Represented orbits
[gm][g^{m}], for 0m<n0\leq m<n γ𝔫m+kn,γ𝔰nm+kn\gamma_{\mathfrak{n}}^{m+kn},\,\,\gamma_{\mathfrak{s}}^{n-m+kn}
Lemma 4.23.

Suppose [γ+]=[γ][S1,S3/n][\gamma_{+}]=[\gamma_{-}]\in[S^{1},S^{3}/{\mathbb{Z}}_{n}] for γ+𝒫LN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}(\lambda_{N}) and γ𝒫LM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}(\lambda_{M}).

  1. (a)

    If μCZ(γ+)=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=\mu_{\operatorname{CZ}}(\gamma_{-}) then γ+γ\gamma_{+}\sim\gamma_{-}.

  2. (b)

    If μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) then 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}).

Proof.

Write [γ±][gm][\gamma_{\pm}]\cong[g^{m}], for some 0m<n0\leq m<n. To prove (a), assume μCZ(γ+)=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=\mu_{\operatorname{CZ}}(\gamma_{-}). Recall the Conley Zehnder index formulas (3.2) from Section 3.1:

μCZ(γ𝔰k)=22kn1,μCZ(γ𝔫k)=22kn+1.\mu_{\operatorname{CZ}}(\gamma_{\mathfrak{s}}^{k})=2\Bigl{\lceil}\frac{2k}{n}\Bigr{\rceil}-1,\,\,\,\,\,\mu_{\operatorname{CZ}}(\gamma_{\mathfrak{n}}^{k})=2\Bigl{\lfloor}\frac{2k}{n}\Bigr{\rfloor}+1. (4.11)

We first argue that γ±\gamma_{\pm} both project to the same orbifold point. To see why, note that an iterate of γ𝔫\gamma_{\mathfrak{n}} representing the same homotopy class as an iterate of γ𝔰\gamma_{\mathfrak{s}} cannot share the same Conley Zehnder indices, as the equality μCZ(γ𝔫m+k1n)=μCZ(γ𝔰nm+k2n)\mu_{\operatorname{CZ}}(\gamma_{\mathfrak{n}}^{m+k_{1}n})=\mu_{\operatorname{CZ}}(\gamma_{\mathfrak{s}}^{n-m+k_{2}n}) implies, by (4.11), that 22mn+1=2+2k22k12\lfloor\frac{2m}{n}\rfloor+1=2+2k_{2}-2k_{1}. This equality cannot hold because the left hand side is odd while the right hand side is even. So, without loss of generality, suppose both γ±\gamma_{\pm} are iterates of γ𝔫\gamma_{\mathfrak{n}}. Then we must have that γ±=γ𝔫m+k±n\gamma_{\pm}=\gamma_{\mathfrak{n}}^{m+k_{\pm}n}. Again by (4.11), the equality μCZ(γ+)=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=\mu_{\operatorname{CZ}}(\gamma_{-}) implies k+=kk_{+}=k_{-} and thus γ+γ\gamma_{+}\sim\gamma_{-}.

To prove (b), we suppose μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}). Recall the action formulas (3.1) from Section 3.1:

𝒜(γ𝔰k)=2πk(1ε)n,𝒜(γ𝔫k)=2πk(1+ε)n.\mathcal{A}(\gamma_{\mathfrak{s}}^{k})=\frac{2\pi k(1-\varepsilon)}{n},\,\,\,\,\,\mathcal{A}(\gamma_{\mathfrak{n}}^{k})=\frac{2\pi k(1+\varepsilon)}{n}. (4.12)

If both γ±\gamma_{\pm} project to the same point, then m(γ+)<m(γ)m(\gamma_{+})<m(\gamma_{-}), because the Conley Zehnder index is a non-decreasing function of the multiplicity, and thus, 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}). In the case that γ±\gamma_{\pm} project to different orbifold points there are two possibilities for the pair (γ+,γ)(\gamma_{+},\gamma_{-}):

Case 1: (γ+,γ)=(γ𝔫m+nk+,γ𝔰nm+nk)(\gamma_{+},\gamma_{-})=(\gamma_{\mathfrak{n}}^{m+nk_{+}},\gamma_{\mathfrak{s}}^{n-m+nk_{-}}). By (4.12),

𝒜(γ)𝒜(γ+)=x+δ,wherex=2π(1+kk+2mn)\mathcal{A}(\gamma_{-})-\mathcal{A}(\gamma_{+})=x+\delta,\,\,\,\,\text{where}\,\,\,\,x=2\pi\bigg{(}1+k_{-}-k_{+}-\frac{2m}{n}\bigg{)}

and δ\delta can be made arbitrarily small, independent of mm, nn, and k±k_{\pm}. Thus, 0<x0<x would imply 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}), after reducing εN\varepsilon_{N} and εM\varepsilon_{M} if necessary. By (4.11), the inequality μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) yields

k++2mnk.k_{+}+\Bigl{\lfloor}\frac{2m}{n}\Bigr{\rfloor}\leq k_{-}. (4.13)

If 2m<n2m<n, then 2mn=0\lfloor\frac{2m}{n}\rfloor=0 and (4.13) implies k+kk_{+}\leq k_{-}. Now we see that

x=2π(1+kk+2mn)2π(12mn)>0,x=2\pi\bigg{(}1+k_{-}-k_{+}-\frac{2m}{n}\bigg{)}\geq 2\pi\bigg{(}1-\frac{2m}{n}\bigg{)}>0,

thus 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}). If 2mn2m\geq n, then 2mn=1\lfloor\frac{2m}{n}\rfloor=1 and (4.13) implies k++1kk_{+}+1\leq k_{-}. Now we see that

x=2π(1+kk+2mn)2π(22mn)>0,x=2\pi\bigg{(}1+k_{-}-k_{+}-\frac{2m}{n}\bigg{)}\geq 2\pi\bigg{(}2-\frac{2m}{n}\bigg{)}>0,

hence 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}).

Case 2: (γ+,γ)=(γ𝔰nm+nk+,γ𝔫m+nk)(\gamma_{+},\gamma_{-})=(\gamma_{\mathfrak{s}}^{n-m+nk_{+}},\gamma_{\mathfrak{n}}^{m+nk_{-}}). By (4.12),

𝒜(γ)𝒜(γ+)=x+δ,wherex=2π(2mn+kk+1)\mathcal{A}(\gamma_{-})-\mathcal{A}(\gamma_{+})=x+\delta,\,\,\,\,\text{where}\,\,\,\,x=2\pi\bigg{(}\frac{2m}{n}+k_{-}-k_{+}-1\bigg{)}

and δ\delta is a small positive number. Thus, 0x0\leq x would imply 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}). Applying (4.11), μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) yields

k+2mn<k.k_{+}-\Bigl{\lfloor}\frac{2m}{n}\Bigr{\rfloor}<k_{-}. (4.14)

If 2m<n2m<n, then 2mn=0\lfloor\frac{2m}{n}\rfloor=0 and (4.14) implies k++1kk_{+}+1\leq k_{-}. Now we see that

x=2π(2mn+kk+1)2π(2mn)0,x=2\pi\bigg{(}\frac{2m}{n}+k_{-}-k_{+}-1\bigg{)}\geq 2\pi\bigg{(}\frac{2m}{n}\bigg{)}\geq 0,

thus 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}). If 2mn2m\geq n, then 2mn=1\lfloor\frac{2m}{n}\rfloor=1 and (4.14) implies k+kk_{+}\leq k_{-}. Now we see that

x=2π(2mn+kk+1)2π(2mn1)0,x=2\pi\bigg{(}\frac{2m}{n}+k_{-}-k_{+}-1\bigg{)}\geq 2\pi\bigg{(}\frac{2m}{n}-1\bigg{)}\geq 0,

thus 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}). ∎

4.3.2 Binary dihedral groups 𝔻2n{\mathbb{D}}_{2n}^{*}

The nonabelian group 𝔻2n{\mathbb{D}}_{2n}^{*} has 4n4n elements and has n+3n+3 conjugacy classes. For 0<m<n0<m<n, the conjugacy class of AmA^{m} is [Am]={Am,A2nm}[A^{m}]=\{A^{m},A^{2n-m}\} (see Section 3.2 for notation). Because Id-\text{Id} generates the center, we have [Id]={Id}[-\text{Id}]=\{-\text{Id}\}. We also have the conjugacy class [Id]={Id}[\text{Id}]=\{\text{Id}\}. The final two conjugacy classes are

[B]={B,A2B,,A2n2B},and[AB]={AB,A3B,,A2n1B}.[B]=\{B,A^{2}B,\dots,A^{2n-2}B\},\,\,\,\text{and}\,\,\,[AB]=\{AB,A^{3}B,\dots,A^{2n-1}B\}.

Note that B1=B3=B=AnBB^{-1}=B^{3}=-B=A^{n}B is conjugate to BB if and only if nn is even. Table 7 records lifts of our three embedded Reeb orbits, e,h,e_{-},h, and e+e_{+}, to paths γ~\widetilde{\gamma} in S3S^{3}, along with the group element g𝔻2ng\in{\mathbb{D}}^{*}_{2n} satisfying gγ~(0)=γ~(T)g\cdot\widetilde{\gamma}(0)=\widetilde{\gamma}(T).

γ~\widetilde{\gamma} S3S^{3} expression interval g𝔻2ng\in{\mathbb{D}}^{*}_{2n}
e~(t)\widetilde{e_{-}}(t) teit2(1,iϵ)t\mapsto\frac{e^{it}}{\sqrt{2}}\cdot(1,-i\epsilon) t[0,π2]t\in[0,\frac{\pi}{2}] ABAB
h~(t)\widetilde{h}(t) teit2(eiπ/4,eiπ/4)t\mapsto\frac{e^{it}}{\sqrt{2}}\cdot(e^{i\pi/4},-e^{i\pi/4}) t[0,π2]t\in[0,\frac{\pi}{2}] BB
e+~(t)\widetilde{e_{+}}(t) t(eit,0)t\mapsto(e^{it},0) t[0,πn]t\in[0,\frac{\pi}{n}] AA
Table 7: Lifts of Dihedral Reeb orbits to S3S^{3}

The homotopy classes of ee_{-}, hh, and e+e_{+} and their iterates are determined by this data and given in Table 8. We now prove Lemma 4.24, the dihedral case of Proposition 4.6.

Free homotopy/conjugacy class Represented orbits (nn even) Represented orbits (nn odd)
[Id][\text{Id}] e4k,h4k,e+2nke_{-}^{4k},\,h^{4k},\,e_{+}^{2nk} e4k,h4k,e+2nke_{-}^{4k},\,h^{4k},\,e_{+}^{2nk}
[Id][-\text{Id}] e2+4k,h2+4k,e+n+2nke_{-}^{2+4k},\,h^{2+4k},\,e_{+}^{n+2nk} e2+4k,h2+4k,e+n+2nke_{-}^{2+4k},\,h^{2+4k},\,e_{+}^{n+2nk}
[Am][A^{m}], for 0<m<n0<m<n e+m+2nk,e+2nm+2nke_{+}^{m+2nk},\,e_{+}^{2n-m+2nk} e+m+2nk,e+2nm+2nke_{+}^{m+2nk},\,e_{+}^{2n-m+2nk}
[B][B] h1+4k,h3+4kh^{1+4k},\,h^{3+4k} h1+4k,e3+4kh^{1+4k},\,e_{-}^{3+4k}
[AB][AB] e1+4k,e3+4ke_{-}^{1+4k},\,e_{-}^{3+4k} e1+4k,h3+4ke_{-}^{1+4k},\,h^{3+4k}
Table 8: Dihedral homotopy classes of Reeb orbits
Lemma 4.24.

Suppose [γ+]=[γ][S1,S3/𝔻2n][\gamma_{+}]=[\gamma_{-}]\in[S^{1},S^{3}/{\mathbb{D}}^{*}_{2n}] for γ+𝒫LN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}(\lambda_{N}) and γ𝒫LM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}(\lambda_{M}).

  1. (a)

    If μCZ(γ+)=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=\mu_{\operatorname{CZ}}(\gamma_{-}) then γ+γ\gamma_{+}\sim\gamma_{-}.

  2. (b)

    If μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) then 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}).

Proof.

We first prove (a). Recall the Conley Zehnder index formulas (3.3) from Section 3.2:

μCZ(ek)=2k21,μCZ(hk)=k,μCZ(e+k)=2kn+1.\mu_{\operatorname{CZ}}(e_{-}^{k})=2\Bigl{\lceil}\frac{k}{2}\Bigr{\rceil}-1,\,\,\,\mu_{\operatorname{CZ}}(h^{k})=k,\,\,\,\mu_{\operatorname{CZ}}(e_{+}^{k})=2\Bigl{\lfloor}\frac{k}{n}\Bigr{\rfloor}+1. (4.15)

By Table 8, there are five possible values of the class [γ±][\gamma_{\pm}]:

Case 1: [γ±][Id][\gamma_{\pm}]\cong[\text{Id}]. Then {γ±}{e4k1,h4k2,e+2nk3|ki}\{\gamma_{\pm}\}\subset\{e_{-}^{4k_{1}},h^{4k_{2}},e_{+}^{2nk_{3}}\,|\,k_{i}\in{\mathbb{N}}\}. The Conley Zehnder index modulo 4 of e4k1e_{-}^{4k_{1}}, h4k2h^{4k_{2}}, or e+2nk3e_{+}^{2nk_{3}} is 1-1, 0, or 11, respectively. Thus, γ±\gamma_{\pm} must project to the same orbifold point. Now, because (1) γ±\gamma_{\pm} are both type ee_{-}, type hh, or type e+e_{+}, and (2) share a homotopy class, μCZ(γ+)=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=\mu_{\operatorname{CZ}}(\gamma_{-}) and (4.15) imply m(γ+)=m(γ)m(\gamma_{+})=m(\gamma_{-}), thus γ+γ\gamma_{+}\sim\gamma_{-}.

Case 2: [γ±][Id][\gamma_{\pm}]\cong[-\text{Id}]. Then {γ±}{e2+4k1,h2+4k2,e+n+2nk3|ki0}\{\gamma_{\pm}\}\subset\{e_{-}^{2+4k_{1}},h^{2+4k_{2}},e_{+}^{n+2nk_{3}}\,|\,k_{i}\in{\mathbb{Z}}_{\geq 0}\}. The Conley Zehnder index modulo 4 of e4k1e_{-}^{4k_{1}}, h4k2h^{4k_{2}}, or e+2nk3e_{+}^{2nk_{3}} is 11, 2, or 33, respectively. By the reasoning as in the previous case, we obtain γ+γ\gamma_{+}\sim\gamma_{-}.

Case 3: [γ±][Am][\gamma_{\pm}]\cong[A^{m}] for some 0<m<n0<m<n. If γ+γ\gamma_{+}\nsim\gamma_{-}, then by Table 8 and (4.15), we must have k1,k2k_{1},k_{2}\in{\mathbb{Z}} such that μCZ(e+m+2nk1)=μCZ(e+2nm+2nk2)\mu_{\operatorname{CZ}}(e_{+}^{m+2nk_{1}})=\mu_{\operatorname{CZ}}(e_{+}^{2n-m+2nk_{2}}). This equation becomes mn+mn=2(1+k2k1)\lfloor\frac{m}{n}\rfloor+\lceil\frac{m}{n}\rceil=2(1+k_{2}-k_{1}). By the bounds on mm, the ratio mn\frac{m}{n} is not an integer, and so the quantity on the left hand side is odd, which is impossible. Thus, we must have γ+γ\gamma_{+}\sim\gamma_{-}.

Case 4: [γ±][B][\gamma_{\pm}]\cong[B]. If nn is even, then both γ±\gamma_{\pm} are iterates of hh, and by (4.15), these multiplicities agree, so γ+γ\gamma_{+}\sim\gamma_{-}. If nn is odd and γ±\gamma_{\pm} project to the same orbifold point, then because their homotopy classes and Conley Zehnder indices agree, Table 8 and (4.15) imply their multiplicities must agree, so γ+γ\gamma_{+}\sim\gamma_{-}. If they project to different orbifold points then 1+4k+=μCZ(h1+4k+)=μCZ(e3+4k)=3+4k1+4k_{+}=\mu_{\operatorname{CZ}}(h^{1+4k_{+}})=\mu_{\operatorname{CZ}}(e_{-}^{3+4k_{-}})=3+4k_{-}, which would imply 1=31=3 mod 4, impossible.

Case 5: [γ±][AB][\gamma_{\pm}]\cong[AB]. If nn is even, then both γ±\gamma_{\pm} are iterates of ee_{-}, and by (4.15), their multiplicities agree, thus γ+γ\gamma_{+}\sim\gamma_{-}. If nn is odd and γ±\gamma_{\pm} project to the same point, then γ+γ\gamma_{+}\sim\gamma_{-} for the same reasons as in Case 4. Thus γ+γ\gamma_{+}\nsim\gamma_{-} implies 3+4k+=μCZ(h3+4k+)=μCZ(e1+4k)=1+4k3+4k_{+}=\mu_{\operatorname{CZ}}(h^{3+4k_{+}})=\mu_{\operatorname{CZ}}(e_{-}^{1+4k_{-}})=1+4k_{-}, impossible mod 4.

To prove (b), let μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}). Recall the action formulas (3.4) from Section 3.2:

𝒜(ek)=kπ(1ε)2,𝒜(hk)=kπ2,𝒜(e+k)=kπ(1+ε)n.\mathcal{A}(e_{-}^{k})=\frac{k\pi(1-\varepsilon)}{2},\,\,\,\mathcal{A}(h^{k})=\frac{k\pi}{2},\,\,\,\mathcal{A}(e_{+}^{k})=\frac{k\pi(1+\varepsilon)}{n}. (4.16)

First, note that if γ±\gamma_{\pm} project to the same orbifold critical point, then m(γ+)<m(γ)m(\gamma_{+})<m(\gamma_{-}), because the Conley Zehnder index as a function of multiplicity of the iterate is non-decreasing. This implies 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}), because the action strictly increases as a function of the iterate. We now prove (b) for pairs of orbits γ±\gamma_{\pm} projecting to different orbifold critical points.

Case 1: [γ±][Id][\gamma_{\pm}]\cong[\text{Id}]. Then {γ±}{e4k1,h4k2,e+2nk3|ki}\{\gamma_{\pm}\}\subset\{e_{-}^{4k_{1}},h^{4k_{2}},e_{+}^{2nk_{3}}\,|\,k_{i}\in{\mathbb{N}}\}. There are six combinations of the value of (γ+,γ)(\gamma_{+},\gamma_{-}), whose index and action are compared using (4.15) and (4.16):

  1. (i)

    γ+=e4k+\gamma_{+}=e_{-}^{4k_{+}} and γ=h4k\gamma_{-}=h^{4k_{-}}. The inequality 4k+1=μCZ(γ+)<μCZ(γ)=4k4k_{+}-1=\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-})=4k_{-} implies k+kk_{+}\leq k_{-}. This verifies that 𝒜(γ+)=2πk+(1εN)<2πk=𝒜(γ)\mathcal{A}(\gamma_{+})=2\pi k_{+}(1-\varepsilon_{N})<2\pi k_{-}=\mathcal{A}(\gamma_{-}).

  2. (ii)

    γ+=h4k+\gamma_{+}=h^{4k_{+}} and γ=e4k\gamma_{-}=e_{-}^{4k_{-}}. The inequality 4k+=μCZ(γ+)<μCZ(γ)=4k14k_{+}=\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-})=4k_{-}-1 implies k+<kk_{+}<k_{-}. This verifies that 𝒜(γ+)=2πk+<2πk(1εM)=𝒜(γ)\mathcal{A}(\gamma_{+})=2\pi k_{+}<2\pi k_{-}(1-\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

  3. (iii)

    γ+=e4k+\gamma_{+}=e_{-}^{4k_{+}} and γ=e+2nk\gamma_{-}=e_{+}^{2nk_{-}}. The inequality 4k+1=μCZ(γ+)<μCZ(γ)=4k+14k_{+}-1=\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-})=4k_{-}+1 implies k+kk_{+}\leq k_{-}. This verifies that 𝒜(γ+)=2πk+(1εN)<2πk(1+εM)=𝒜(γ)\mathcal{A}(\gamma_{+})=2\pi k_{+}(1-\varepsilon_{N})<2\pi k_{-}(1+\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

  4. (iv)

    γ+=e+2nk+\gamma_{+}=e_{+}^{2nk_{+}} and γ=e4k\gamma_{-}=e_{-}^{4k_{-}}. The inequality 4k++1=μCZ(γ+)<μCZ(γ)=4k14k_{+}+1=\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-})=4k_{-}-1 implies k+<kk_{+}<k_{-}. This verifies that 𝒜(γ+)=2πk+(1+εN)<2πk(1εM)=𝒜(γ)\mathcal{A}(\gamma_{+})=2\pi k_{+}(1+\varepsilon_{N})<2\pi k_{-}(1-\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

  5. (v)

    γ+=h4k+\gamma_{+}=h^{4k_{+}} and γ=e+2nk\gamma_{-}=e_{+}^{2nk_{-}}. The inequality 4k+=μCZ(γ+)<μCZ(γ)=4k+14k_{+}=\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-})=4k_{-}+1 implies k+kk_{+}\leq k_{-}. This verifies that 𝒜(γ+)=2πk+<2πk(1+εM)=𝒜(γ)\mathcal{A}(\gamma_{+})=2\pi k_{+}<2\pi k_{-}(1+\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

  6. (vi)

    γ+=e+2nk+\gamma_{+}=e_{+}^{2nk_{+}} and γ=h4k\gamma_{-}=h^{4k_{-}}. The inequality 4k++1=μCZ(γ+)<μCZ(γ)=4k4k_{+}+1=\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-})=4k_{-} implies k+<kk_{+}<k_{-}. This verifies that 𝒜(γ+)=2πk+(1+εN)<2πk=𝒜(γ)\mathcal{A}(\gamma_{+})=2\pi k_{+}(1+\varepsilon_{N})<2\pi k_{-}=\mathcal{A}(\gamma_{-}).

Case 2: [γ±][Id][\gamma_{\pm}]\cong[-\text{Id}]. Then {γ±}{e2+4k1,h2+4k2,e+n+2nk3|ki0}\{\gamma_{\pm}\}\subset\{e_{-}^{2+4k_{1}},h^{2+4k_{2}},e_{+}^{n+2nk_{3}}\,|\,k_{i}\in{\mathbb{Z}}_{\geq 0}\}. The possible values of (γ+,γ)(\gamma_{+},\gamma_{-}) and the arguments for 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}) are identical to the above case.

Case 3: [γ±][Am][\gamma_{\pm}]\cong[A^{m}] for 0<m<n0<m<n and both γ±\gamma_{\pm} are iterates of e+e_{+}, so (b) holds.

Case 4: [γ±][B][\gamma_{\pm}]\cong[B]. If nn is even, then both γ±\gamma_{\pm} are iterates of hh, and so (b) holds. Otherwise, nn is odd and the pair (γ+,γ)(\gamma_{+},\gamma_{-}) is either (e3+4k+,h1+4k)(e_{-}^{3+4k_{+}},h^{1+4k_{-}}) or (h1+4k+,e3+4k)(h^{1+4k_{+}},e_{-}^{3+4k_{-}}). In the former case μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) and (4.15) imply k+<kk_{+}<k_{-}, so, by (4.16),
𝒜(γ+)=(3+4k+)π2(1εN)<(1+4k)π2=𝒜(γ)\mathcal{A}(\gamma_{+})=\frac{(3+4k_{+})\pi}{2}(1-\varepsilon_{N})<\frac{(1+4k_{-})\pi}{2}=\mathcal{A}(\gamma_{-}). In the latter case, μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) and (4.15) imply k+kk_{+}\leq k_{-}, thus 𝒜(γ+)=(1+4k+)π2<(3+4k)π2(1εM)=𝒜(γ)\mathcal{A}(\gamma_{+})=\frac{(1+4k_{+})\pi}{2}<\frac{(3+4k_{-})\pi}{2}(1-\varepsilon_{M})=\mathcal{A}(\gamma_{-}) by (4.16).

Case 5: [γ±][AB][\gamma_{\pm}]\cong[AB]. If nn is even, then both γ±\gamma_{\pm} are iterates of ee_{-}, and so 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}) holds. Otherwise, nn is odd and the pair (γ+,γ)(\gamma_{+},\gamma_{-}) is either (e1+4k+,h3+4k)(e_{-}^{1+4k_{+}},h^{3+4k_{-}}) or (h3+4k+,e1+4k)(h^{3+4k_{+}},e_{-}^{1+4k_{-}}). If the former holds, then μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) and (4.15) imply k+kk_{+}\leq k_{-}, and so, by (4.16),
𝒜(γ+)=(1+4k+)π2(1εN)<(3+4k)π2=𝒜(γ)\mathcal{A}(\gamma_{+})=\frac{(1+4k_{+})\pi}{2}(1-\varepsilon_{N})<\frac{(3+4k_{-})\pi}{2}=\mathcal{A}(\gamma_{-}). If the latter holds, then μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) and (4.15) imply k+<kk_{+}<k_{-}, and so 𝒜(γ+)=(3+4k+)π2<(1+4k)π2(1εM)=𝒜(γ)\mathcal{A}(\gamma_{+})=\frac{(3+4k_{+})\pi}{2}<\frac{(1+4k_{-})\pi}{2}(1-\varepsilon_{M})=\mathcal{A}(\gamma_{-}) by (4.16). ∎

4.3.3 Binary polyhedral groups 𝕋{\mathbb{T}}^{*}, 𝕆{\mathbb{O}}^{*}, and 𝕀{\mathbb{I}}^{*}

We will describe the homotopy classes of the Reeb orbits in S3/S^{3}/{\mathbb{P}}^{*} using a more geometric point of view than in the dihedral case. If a loop γ\gamma in S3/S^{3}/{\mathbb{P}}^{*} and cConj(G)c\in\text{Conj}(G) satisfy [γ]c[\gamma]\cong c, then the order of γ\gamma in π1(S3/)\pi_{1}(S^{3}/{\mathbb{P}}^{*}) equals the group order of cc, defined to be the order of any gg\in{\mathbb{P}}^{*} representing cc. If γ\gamma has order kk in the fundamental group and cConj()c\in\text{Conj}({\mathbb{P}}^{*}) is the only class with group order kk, then we can immediately conclude that [γ]c[\gamma]\cong c. Determining the free homotopy class of γ\gamma via the group order is more difficult when there are multiple conjugacy classes of {\mathbb{P}}^{*} with the same group order.

Tables 9, 10, and 11 provide notation for the distinct conjugacy classes of 𝕋{\mathbb{T}}^{*}, 𝕆{\mathbb{O}}^{*}, and 𝕀{\mathbb{I}}^{*}, along with their group orders. Our notation indicates when there exist multiple conjugacy classes featuring the same group order - the notation Pi,AP_{i,A} and Pi,BP_{i,B} provides labels for the two distinct conjugacy classes of {\mathbb{P}}^{*} (for P{T,O,I}P\in\{T,O,I\}) of group order ii. For P{T,O,I}P\in\{T,O,I\}, PIdP_{\text{Id}} and PIdP_{-\text{Id}} denote the singleton conjugacy classes {Id}\{\text{Id}\} and {Id}\{-\text{Id}\}, respectively, and PiP_{i} denotes the unique conjugacy class of group order ii.

Conjugacy class TIdT_{\text{Id}} TIdT_{-\text{Id}} T4T_{4} T6,AT_{6,A} T6,BT_{6,B} T3,AT_{3,A} T3,BT_{3,B}
Group order 1 2 4 6 6 3 3
Table 9: The 7 conjugacy classes of 𝕋{\mathbb{T}}^{*}
Conjugacy class OIdO_{\text{Id}} OIdO_{-\text{Id}} O8,AO_{8,A} O8,BO_{8,B} O4,AO_{4,A} O4,BO_{4,B} O6O_{6} O3O_{3}
Group order 1 2 8 8 4 4 6 3
Table 10: The 8 conjugacy classes of 𝕆{\mathbb{O}}^{*}
Conjugacy class IIdI_{\text{Id}} IIdI_{-\text{Id}} I10,AI_{10,A} I10,BI_{10,B} I5,AI_{5,A} I5,BI_{5,B} I4I_{4} I6I_{6} I3I_{3}
Group order 1 2 10 10 5 5 4 6 3
Table 11: The 9 conjugacy classes of 𝕀{\mathbb{I}}^{*}

Lemmas 4.25, 4.26, and 4.27 allow us to record the homotopy classes represented by any iterate of 𝒱\mathcal{V}, \mathcal{E}, or \mathcal{F} in Tables 14, 14, and 14. We explain how the table is set up in the tetrahedral case: it must be true that [𝒱][][\mathcal{V}]\neq[\mathcal{F}], otherwise taking the 2-fold iterate would violate Lemma 4.26(i) so without loss of generality write [𝒱]T6,A[\mathcal{V}]\cong T_{6,A} and []T6,B[\mathcal{F}]\cong T_{6,B}, and similarly [𝒱2]T3,A[\mathcal{V}^{2}]\cong T_{3,A} and [2]T3,B[\mathcal{F}^{2}]\cong T_{3,B}. By Lemma 4.25(i), we must have that [5]T6,A[\mathcal{F}^{5}]\cong T_{6,A}, [4]T3,A[\mathcal{F}^{4}]\cong T_{3,A}. Taking the 4-fold iterate of [𝒱]=[5][\mathcal{V}]=[\mathcal{F}^{5}] provides [𝒱4]=[2]T3,B[\mathcal{V}^{4}]=[\mathcal{F}^{2}]\cong T_{3,B}, and the 5-fold iterate provides [𝒱5]=[1]T6,B[\mathcal{V}^{5}]=[\mathcal{F}^{1}]\cong T_{6,B}, and we have resolved all ambiguity regarding the tetrahedral classes of group orders 6 and 3. Analogous arguments apply in the octahedral and icosahedral cases.

Class Represented orbits TIdT_{\text{Id}} 𝒱6k,4k,6k\mathcal{V}^{6k},\,\mathcal{E}^{4k},\,\mathcal{F}^{6k} TIdT_{-\text{Id}} 𝒱3+6k,2+4k,3+6k\mathcal{V}^{3+6k},\,\mathcal{E}^{2+4k},\,\mathcal{F}^{3+6k} T4T_{4} 1+4k,3+4k\mathcal{E}^{1+4k},\,\mathcal{E}^{3+4k} T6,AT_{6,A} 𝒱1+6k,5+6k\mathcal{V}^{1+6k},\,\mathcal{F}^{5+6k} T6,BT_{6,B} 1+6k,𝒱5+6k\mathcal{F}^{1+6k},\,\mathcal{V}^{5+6k} T3,AT_{3,A} 𝒱2+6k,4+6k\mathcal{V}^{2+6k},\,\mathcal{F}^{4+6k} T3,BT_{3,B} 2+6k,𝒱4+6k\mathcal{F}^{2+6k},\,\mathcal{V}^{4+6k} Table 12: S3/𝕋S^{3}/{\mathbb{T}}^{*} Reeb Orbits Class Represented orbits OIdO_{\text{Id}} 𝒱8k,4k,6k\mathcal{V}^{8k},\,\mathcal{E}^{4k},\,\mathcal{F}^{6k} OIdO_{-\text{Id}} 𝒱4+8k,2+4k,3+6k\mathcal{V}^{4+8k},\,\mathcal{E}^{2+4k},\,\mathcal{F}^{3+6k} O8,AO_{8,A} 𝒱1+8k,𝒱7+8k\mathcal{V}^{1+8k},\,\mathcal{V}^{7+8k} O8,BO_{8,B} 𝒱3+8k,𝒱5+8k\mathcal{V}^{3+8k},\,\mathcal{V}^{5+8k} O4,AO_{4,A} 𝒱2+8k,𝒱6+8k\mathcal{V}^{2+8k},\,\mathcal{V}^{6+8k} O4,BO_{4,B} 1+4k,3+4k\mathcal{E}^{1+4k},\,\mathcal{E}^{3+4k} O6O_{6} 1+6k,5+6k\mathcal{F}^{1+6k},\,\mathcal{F}^{5+6k} O3O_{3} 2+6k,4+6k\mathcal{F}^{2+6k},\,\mathcal{F}^{4+6k} Table 13: S3/𝕆S^{3}/{\mathbb{O}}^{*} Reeb Orbits Class Represented orbits IIdI_{\text{Id}} 𝒱10k,4k,6k\mathcal{V}^{10k},\,\mathcal{E}^{4k},\,\mathcal{F}^{6k} IIdI_{-\text{Id}} 𝒱5+10k,2+4k,3+6k\mathcal{V}^{5+10k},\,\mathcal{E}^{2+4k},\,\mathcal{F}^{3+6k} I10,AI_{10,A} 𝒱1+10k,𝒱9+10k\mathcal{V}^{1+10k},\,\mathcal{V}^{9+10k} I10,BI_{10,B} 𝒱3+10k,𝒱7+10k\mathcal{V}^{3+10k},\,\mathcal{V}^{7+10k} I5,AI_{5,A} 𝒱2+10k,𝒱8+10k\mathcal{V}^{2+10k},\,\mathcal{V}^{8+10k} I5,BI_{5,B} 𝒱4+10k,𝒱6+10k\mathcal{V}^{4+10k},\,\mathcal{V}^{6+10k} I4I_{4} 1+4k,3+4k\mathcal{E}^{1+4k},\,\mathcal{E}^{3+4k} I6I_{6} 1+6k,5+6k\mathcal{F}^{1+6k},\,\mathcal{F}^{5+6k} I3I_{3} 2+6k,4+6k\mathcal{F}^{2+6k},\,\mathcal{F}^{4+6k} Table 14: S3/𝕀S^{3}/{\mathbb{I}}^{*} Reeb Orbits

Lemma 4.25.

We have the following identifications between free homotopy classes represented by orbits:

  1. (i)

    [𝒱]=[5][\mathcal{V}]=[\mathcal{F}^{5}], for =𝕋{\mathbb{P}}^{*}={\mathbb{T}}^{*},

  2. (ii)

    [𝒱]=[𝒱7][\mathcal{V}]=[\mathcal{V}^{7}], for =𝕆{\mathbb{P}}^{*}={\mathbb{O}}^{*},

  3. (iii)

    [𝒱]=[𝒱9][\mathcal{V}]=[\mathcal{V}^{9}], for =𝕀{\mathbb{P}}^{*}={\mathbb{I}}^{*}.

Proof.

Suppose p1p_{1}, p2Fix()S2p_{2}\in\text{Fix}({\mathbb{P}})\subset S^{2} are antipodal, i.e. p1=p2p_{1}=-p_{2}. Select zi𝔓1(pi)S3z_{i}\in\mathfrak{P}^{-1}(p_{i})\subset S^{3}. The Hopf fibration 𝔓\mathfrak{P} has the property that 𝔓(v1)=𝔓(v2)\mathfrak{P}(v_{1})=-\mathfrak{P}(v_{2}) in S2S^{2} if and only if v1v_{1} and v2v_{2} are orthogonal vectors in 2{\mathbb{C}}^{2}, so z1z_{1} and z2z_{2} must be orthogonal. Now, pip_{i} is either of vertex, edge, or face type, so let γi\gamma_{i} denote the orbit 𝒱\mathcal{V}, \mathcal{E}, or \mathcal{F}, depending on this type of pip_{i}. Because p1p_{1} and p2p_{2} are antipodal, the order of γ1\gamma_{1} equals that of γ2\gamma_{2} in π1(S3/)\pi_{1}(S^{3}/{\mathbb{P}}^{*}), call this order dd, and let T:=2πdT:=\frac{2\pi}{d}. Now, consider that the map

Γ1:[0,T]S3,teitz1\Gamma_{1}:[0,T]\to S^{3},\,\,\,t\mapsto e^{it}\cdot z_{1}

is a lift of γ1\gamma_{1} to S3S^{3}. Thus, z1z_{1} is an eigenvector with eigenvalue eiTe^{iT} for some gg\in{\mathbb{P}}^{*}, and [γ1][g][\gamma_{1}]\cong[g] because gΓ1(0)=Γ1(T)g\cdot\Gamma_{1}(0)=\Gamma_{1}(T). Because gg is special unitary, we must also have that z2z_{2} is an eigenvector of gg with eigenvalue ei(2πT)=ei(d1)Te^{i(2\pi-T)}=e^{i(d-1)T}. Now the map

Γ2d1:[0,(d1)T]S3,teitz2\Gamma_{2}^{d-1}:[0,(d-1)T]\to S^{3},\,\,\,t\mapsto e^{it}\cdot z_{2}

is a lift of γ2d1\gamma_{2}^{d-1} to S3S^{3}. This provides gΓ2d1(0)=Γ2d1((d1)T)g\cdot\Gamma_{2}^{d-1}(0)=\Gamma_{2}^{d-1}((d-1)T) which implies [γ1]=[γ2d1][\gamma_{1}]=[\gamma_{2}^{d-1}], as both are identified with [g][g], and thereby establishing the above identifications. ∎

Lemma 4.26.

The following identifications fail to hold between free homotopy classes represented by orbits:

  1. (i)

    [𝒱2][2][\mathcal{V}^{2}]\neq[\mathcal{F}^{2}], for =𝕋{\mathbb{P}}^{*}={\mathbb{T}}^{*},

  2. (ii)

    [][𝒱2][\mathcal{E}]\neq[\mathcal{V}^{2}], for =𝕆{\mathbb{P}}^{*}={\mathbb{O}}^{*}

Proof.

Suppose, for i=1,2i=1,2, γi\gamma_{i} is one of 𝒱\mathcal{V}, \mathcal{E}, or \mathcal{F}, and suppose γ1γ2\gamma_{1}\neq\gamma_{2}. Let did_{i} be the order of γi\gamma_{i} in π1(S3/)\pi_{1}(S^{3}/{\mathbb{P}}^{*}), and select kik_{i}\in{\mathbb{N}} for i=1,2i=1,2. If 2πk1d12πk2d2\frac{2\pi k_{1}}{d_{1}}\equiv\frac{2\pi k_{2}}{d_{2}} modulo 2π2\pi{\mathbb{Z}} and if π2πk1d1\pi\nmid\frac{2\pi k_{1}}{d_{1}}, then [γ1k1][γ2k2][\gamma_{1}^{k_{1}}]\neq[\gamma_{2}^{k_{2}}]. To prove this, consider that we have gig_{i}\in{\mathbb{P}}^{*} with eigenvector ziz_{i} in S3S^{3}, with eigenvalue λ:=e2πk1/d1=e2πk2i/d2\lambda:=e^{2\pi k_{1}/d_{1}}=e^{2\pi k_{2}i/d_{2}} so that [γiki][gi][\gamma_{i}^{k_{i}}]\cong[g_{i}]. Note that γ1γ2\gamma_{1}\neq\gamma_{2} implies 𝔓(z1)\mathfrak{P}(z_{1}) is not in the same {\mathbb{P}}-orbit as 𝔓(z2)\mathfrak{P}(z_{2}) in S2S^{2}, i.e. π(𝔓(z1))π(𝔓(z2))\pi_{{\mathbb{P}}}(\mathfrak{P}(z_{1}))\neq\pi_{{\mathbb{P}}}(\mathfrak{P}(z_{2})). Now, [γ1k1]=[γ2k2][\gamma_{1}^{k_{1}}]=[\gamma_{2}^{k_{2}}] would imply g1=x1g2xg_{1}=x^{-1}g_{2}x, for some xx\in{\mathbb{P}}^{*}, ensuring that xz1x\cdot z_{1} is a λ\lambda eigenvector of g2g_{2}. Because λ±1\lambda\neq\pm 1, we know that the λ\lambda-eigenspace of g2g_{2} is complex 1-dimensional, so we must have that xz1x\cdot z_{1} and z2z_{2} are co-linear. That is, xz1=αz2x\cdot z_{1}=\alpha z_{2} for some αS1\alpha\in S^{1}, which implies that

π(𝔓(z1))=π(𝔓(xz1))=π(𝔓(αz2))=π(𝔓(z2)),\pi_{{\mathbb{P}}}(\mathfrak{P}(z_{1}))=\pi_{{\mathbb{P}}}(\mathfrak{P}(x\cdot z_{1}))=\pi_{{\mathbb{P}}}(\mathfrak{P}(\alpha\cdot z_{2}))=\pi_{{\mathbb{P}}}(\mathfrak{P}(z_{2})),

a contradiction. This yields the desired inequivalences.

Lemma 4.27.

The following identifications fail to hold between free homotopy classes represented by iterates of the same orbits:

  1. (i)

    [𝒱1][𝒱3][\mathcal{V}^{1}]\neq[\mathcal{V}^{3}], for =𝕆{\mathbb{P}}^{*}={\mathbb{O}}^{*},

  2. (ii)

    [𝒱2][𝒱4][\mathcal{V}^{2}]\neq[\mathcal{V}^{4}], for =𝕀{\mathbb{P}}^{*}={\mathbb{I}}^{*}.

Proof.

For i=1,2i=1,2, select kik_{i}\in{\mathbb{N}} and let γi\gamma_{i} denote one of 𝒱\mathcal{V}, \mathcal{E}, or \mathcal{F}. Let did_{i} denote the order of γi\gamma_{i} in π1(S3/)\pi_{1}(S^{3}/{\mathbb{P}}^{*}). Suppose that 2πkidi\frac{2\pi k_{i}}{d_{i}} is not a multiple of 2π2\pi, and 2πk1d12πk2d2\frac{2\pi k_{1}}{d_{1}}\ncong\frac{2\pi k_{2}}{d_{2}} mod 2π2\pi{\mathbb{Z}}. If 2πk1d1+2πk2d2\frac{2\pi k_{1}}{d_{1}}+\frac{2\pi k_{2}}{d_{2}} is not a multiple of 2π2\pi, then [γ1k1][γ2k2][\gamma_{1}^{k_{1}}]\neq[\gamma_{2}^{k_{2}}]. To prove this, consider that we have gig_{i}\in{\mathbb{P}}^{*} with [γiki][gi][\gamma_{i}^{k_{i}}]\cong[g_{i}]. This tells us that e2πkji/dje^{2\pi k_{j}i/d_{j}} is an eigenvalue of gjg_{j}. If it were the case that [γ1k1]=[γ2k2][\gamma_{1}^{k_{1}}]=[\gamma_{2}^{k_{2}}] in [S1,S3/][S^{1},S^{3}/{\mathbb{P}}^{*}], then we would have [g1]=[g2][g_{1}]=[g_{2}] in Conj(G)\text{Conj}(G). Because conjugate elements share eigenvalues, we would have

Spec(g1)=Spec(g2)={e2πk1i/d1,e2πk2i/d2}.\text{Spec}(g_{1})=\text{Spec}(g_{2})=\{e^{2\pi k_{1}i/d_{1}},e^{2\pi k_{2}i/d_{2}}\}.

However, the product of these eigenvalues is not 1, contradicting that giSU(2)g_{i}\in\text{SU}(2). ∎

We are ready to prove Proposition 4.29, which is the polyhedral case of Proposition 4.6. Lemma 4.28 will streamline some casework.

Lemma 4.28.

If γ+𝒫LN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}(\lambda_{N}) and γ𝒫LM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}(\lambda_{M}) project to the same orbifold critical point and are in the same free homotopy class, then

  1. (i)

    μCZ(γ+)=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=\mu_{\operatorname{CZ}}(\gamma_{-}) implies m(γ+)=m(γ)m(\gamma_{+})=m(\gamma_{-}), i.e., γ+γ\gamma_{+}\sim\gamma_{-};

  2. (ii)

    μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) implies m(γ+)<m(γ)m(\gamma_{+})<m(\gamma_{-}), and so 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}).

Proof.

For =𝕋{\mathbb{P}}^{*}={\mathbb{T}}^{*}, 𝕆{\mathbb{O}}^{*}, or 𝕀{\mathbb{I}}^{*}, fix cConj()c\in\text{Conj}({\mathbb{P}}^{*}) and let γ\gamma denote one of 𝒱\mathcal{V}, \mathcal{E}, or \mathcal{F}. Define Sγ,c:={γk|[γk]c}S_{\gamma,c}:=\{\gamma^{k}\,|\,[\gamma^{k}]\cong c\}; note that this set may potentially be empty. Observe that the map Sγ,cS_{\gamma,c}\to{\mathbb{Z}}, γkμCZ(γk)\gamma^{k}\mapsto\mu_{\operatorname{CZ}}(\gamma^{k}), is injective. Thus the result holds. ∎

Proposition 4.29.

Suppose [γ+]=[γ][S1,S3/][\gamma_{+}]=[\gamma_{-}]\in[S^{1},S^{3}/{\mathbb{P}}^{*}] for γ+𝒫LN(λN)\gamma_{+}\in\mathcal{P}^{L_{N}}(\lambda_{N}) and γ𝒫LM(λM)\gamma_{-}\in\mathcal{P}^{L_{M}}(\lambda_{M}).

  1. (a)

    If μCZ(γ+)=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=\mu_{\operatorname{CZ}}(\gamma_{-}), then γ+γ\gamma_{+}\sim\gamma_{-}.

  2. (b)

    If μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}), then 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}).

Proof.

We first prove (a). Recall the Conley Zehnder index formulas (3.5) from Section 3.3:

μCZ(𝒱k)=2k𝒱1,μCZ(k)=k,μCZ(k)=2k3+1,\mu_{\operatorname{CZ}}(\mathcal{V}^{k})=2\Bigl{\lceil}\frac{k}{\mathscr{I}_{\mathscr{V}}}\Bigr{\rceil}-1,\,\,\,\mu_{\operatorname{CZ}}(\mathcal{E}^{k})=k,\,\,\,\mu_{\operatorname{CZ}}(\mathcal{F}^{k})=2\Bigl{\lfloor}\frac{k}{3}\Bigr{\rfloor}+1, (4.17)

Consider the following possible values of the homotopy class [γ±][\gamma_{\pm}]:

Case 1: [γ±]TId[\gamma_{\pm}]\cong T_{\text{Id}}, OIdO_{\text{Id}}, or IIdI_{\text{Id}}, so that {γ±}{𝒱2𝒱k1,4k2,6k3|ki}\{\gamma_{\pm}\}\subset\{\mathcal{V}^{2\mathscr{I}_{\mathscr{V}}k_{1}},\mathcal{E}^{4k_{2}},\mathcal{F}^{6k_{3}}\,|\,k_{i}\in{\mathbb{N}}\}. By reasoning identical to the analogous case of 𝔻2n{\mathbb{D}}_{2n}^{*} (Lemma 4.24(a), Case 1), γ+γ\gamma_{+}\sim\gamma_{-}.

Case 2: [γ±]TId[\gamma_{\pm}]\cong T_{-\text{Id}}, OIdO_{-\text{Id}}, or IIdI_{-\text{Id}}, so that {γ±}{𝒱𝒱+2𝒱k1,2+4k2,3+6k3|ki0}\{\gamma_{\pm}\}\subset\{\mathcal{V}^{\mathscr{I}_{\mathscr{V}}+2\mathscr{I}_{\mathscr{V}}k_{1}},\mathcal{E}^{2+4k_{2}},\mathcal{F}^{3+6k_{3}}\,|\,k_{i}\in{\mathbb{Z}}_{\geq 0}\}. Again, as in the the analogous case of 𝔻2n{\mathbb{D}}_{2n}^{*} (Lemma 4.24(a), Case 2), γ+γ\gamma_{+}\sim\gamma_{-}.

Case 3: [γ±]T6,A,T6,B,T3,A[\gamma_{\pm}]\cong T_{6,A},T_{6,B},T_{3,A}, or T3,BT_{3,B}. If both γ±\gamma_{\pm} are iterates of 𝒱\mathcal{V}, then by Lemma 4.28(i), they must share the same multiplicity, i.e. γ+γ\gamma_{+}\sim\gamma_{-}. If both γ±\gamma_{\pm} are iterates of \mathcal{F} then again by Lemma 4.28(i), they must share the same multiplicity, i.e. γ+γ\gamma_{+}\sim\gamma_{-}. So we must argue, using (4.17), that in each of these four free homotopy classes that it is impossible that γ±\gamma_{\pm} project to different orbifold points whenever μCZ(γ+)=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=\mu_{\operatorname{CZ}}(\gamma_{-}).

  • If [γ±]T6,A[\gamma_{\pm}]\cong T_{6,A} and γ±\gamma_{\pm} project to different orbifold points then, up to relabeling,
    γ+=𝒱1+6k+\gamma_{+}=\mathcal{V}^{1+6k_{+}} and γ=5+6k\gamma_{-}=\mathcal{F}^{5+6k_{-}}. Now, μCZ(γ+)=4k++14k+3=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=4k_{+}+1\neq 4k_{-}+3=\mu_{\operatorname{CZ}}(\gamma_{-}).

  • If [γ±]T6,B[\gamma_{\pm}]\cong T_{6,B} and γ±\gamma_{\pm} project to different orbifold points, write γ+=𝒱5+6k+\gamma_{+}=\mathcal{V}^{5+6k_{+}}
    and γ=1+6k\gamma_{-}=\mathcal{F}^{1+6k_{-}}. Now, μCZ(γ+)=4k++34k+1=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=4k_{+}+3\neq 4k_{-}+1=\mu_{\operatorname{CZ}}(\gamma_{-}).

  • If [γ±]T3,A[\gamma_{\pm}]\cong T_{3,A} and γ±\gamma_{\pm} project to different orbifold points, write γ+=𝒱2+6k+\gamma_{+}=\mathcal{V}^{2+6k_{+}}
    and γ=4+6k\gamma_{-}=\mathcal{F}^{4+6k_{-}}. Now, μCZ(γ+)=4k++14k+3=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=4k_{+}+1\neq 4k_{-}+3=\mu_{\operatorname{CZ}}(\gamma_{-}).

  • If [γ±]T3,B[\gamma_{\pm}]\cong T_{3,B} and γ±\gamma_{\pm} project to different orbifold points, write γ+=𝒱4+6k+\gamma_{+}=\mathcal{V}^{4+6k_{+}}
    and γ=2+6k\gamma_{-}=\mathcal{F}^{2+6k_{-}}. Now, μCZ(γ+)=4k++34k+1=μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})=4k_{+}+3\neq 4k_{-}+1=\mu_{\operatorname{CZ}}(\gamma_{-}).

Case 4: [γ±][\gamma_{\pm}] is a homotopy class not covered in Cases 1 - 3. Because every such homotopy class is represented by Reeb orbits either of type 𝒱\mathcal{V}, of type \mathcal{E}, or of type \mathcal{F}, then we see that γ±\gamma_{\pm} project to the same orbifold point of S2/S^{2}/{\mathbb{P}}. By Lemma 4.28(i), we have that γ+γ\gamma_{+}\sim\gamma_{-}.

We now prove (b). Recall the action formulas (3.6) from Section 3.3:

𝒜(𝒱k)=kπ(1ε)𝒱,𝒜(k)=kπ2,𝒜(k)=kπ(1+ε)3.\mathcal{A}(\mathcal{V}^{k})=\frac{k\pi(1-\varepsilon)}{\mathscr{I}_{\mathscr{V}}},\,\,\,\mathcal{A}(\mathcal{E}^{k})=\frac{k\pi}{2},\,\,\,\mathcal{A}(\mathcal{F}^{k})=\frac{k\pi(1+\varepsilon)}{3}. (4.18)

Consider the following possible values of the homotopy class [γ±][\gamma_{\pm}]:

Case 1: [γ±]TId[\gamma_{\pm}]\cong T_{\text{Id}}, OIdO_{\text{Id}}, or IIdI_{\text{Id}}, so that {γ±}{𝒱2𝒱k1,4k2,6k3|ki}\{\gamma_{\pm}\}\subset\{\mathcal{V}^{2\mathscr{I}_{\mathscr{V}}k_{1}},\mathcal{E}^{4k_{2}},\mathcal{F}^{6k_{3}}\,|\,k_{i}\in{\mathbb{N}}\}. By reasoning identical to the analogous case of 𝔻2n{\mathbb{D}}_{2n}^{*} (Lemma 4.24(b), Case 1), 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}).

Case 2: [γ±]TId[\gamma_{\pm}]\cong T_{-\text{Id}}, OIdO_{-\text{Id}}, or IIdI_{-\text{Id}}, so that {γ±}{𝒱𝒱+2𝒱k1,2+4k2,3+6k3|ki0}\{\gamma_{\pm}\}\subset\{\mathcal{V}^{\mathscr{I}_{\mathscr{V}}+2\mathscr{I}_{\mathscr{V}}k_{1}},\mathcal{E}^{2+4k_{2}},\mathcal{F}^{3+6k_{3}}\,|\,k_{i}\in{\mathbb{Z}}_{\geq 0}\}. Again, as in the the analogous case of 𝔻2n{\mathbb{D}}_{2n}^{*} (Lemma 4.24(b), Case 2), 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}).

Case 3: [γ±]T6,A,T6,B,T3,A[\gamma_{\pm}]\cong T_{6,A},T_{6,B},T_{3,A}, or T3,BT_{3,B}. If both γ±\gamma_{\pm} are of type 𝒱\mathcal{V}, then by Lemma 4.28(ii), 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}). If both γ±\gamma_{\pm} are of type \mathcal{F}, then again by Lemma 4.28(ii), 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}). So we must argue, using (4.17) and (4.18), that for each of these four free homotopy classes that if γ+\gamma_{+} and γ\gamma_{-} are not of the same type, and if μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}), then 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}).

3.A If [γ±]T6,A[\gamma_{\pm}]\cong T_{6,A}, then the pair (γ+,γ)(\gamma_{+},\gamma_{-}) is either (𝒱1+6k+,5+6k)(\mathcal{V}^{1+6k_{+}},\mathcal{F}^{5+6k_{-}}) or (5+6k+,𝒱1+6k)(\mathcal{F}^{5+6k_{+}},\mathcal{V}^{1+6k_{-}}). If the former holds, then μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) implies k+kk_{+}\leq k_{-}, and so

𝒜(γ+)=(1+6k+)π3(1εN)<(5+6k)π3(1+εM)=𝒜(γ).\mathcal{A}(\gamma_{+})=\frac{(1+6k_{+})\pi}{3}(1-\varepsilon_{N})<\frac{(5+6k_{-})\pi}{3}(1+\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

If the latter holds, then μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) implies k+<kk_{+}<k_{-}, and again the action satisfies

𝒜(γ+)=(5+6k+)π3(1+εN)<(1+6k)π3(1εM)=𝒜(γ).\mathcal{A}(\gamma_{+})=\frac{(5+6k_{+})\pi}{3}(1+\varepsilon_{N})<\frac{(1+6k_{-})\pi}{3}(1-\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

3.B If [γ±]T6,B[\gamma_{\pm}]\cong T_{6,B}, then the pair (γ+,γ)(\gamma_{+},\gamma_{-}) is either (𝒱5+6k+,1+6k)(\mathcal{V}^{5+6k_{+}},\mathcal{F}^{1+6k_{-}}) or (1+6k+,𝒱5+6k)(\mathcal{F}^{1+6k_{+}},\mathcal{V}^{5+6k_{-}}). If the former holds, then μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) implies k+<kk_{+}<k_{-}, and so

𝒜(γ+)=(5+6k+)π3(1εN)<(1+6k)π3(1+εM)=𝒜(γ).\mathcal{A}(\gamma_{+})=\frac{(5+6k_{+})\pi}{3}(1-\varepsilon_{N})<\frac{(1+6k_{-})\pi}{3}(1+\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

If the latter holds, then μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) implies k+kk_{+}\leq k_{-}, and so

𝒜(γ+)=(1+6k+)π3(1+εN)<(5+6k)π3(1εM)=𝒜(γ).\mathcal{A}(\gamma_{+})=\frac{(1+6k_{+})\pi}{3}(1+\varepsilon_{N})<\frac{(5+6k_{-})\pi}{3}(1-\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

3.C If [γ±]T3,A[\gamma_{\pm}]\cong T_{3,A}, then the pair (γ+,γ)(\gamma_{+},\gamma_{-}) is either (𝒱2+6k+,4+6k)(\mathcal{V}^{2+6k_{+}},\mathcal{F}^{4+6k_{-}}) or (4+6k+,𝒱2+6k)(\mathcal{F}^{4+6k_{+}},\mathcal{V}^{2+6k_{-}}). If the former holds, then μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) implies k+kk_{+}\leq k_{-}, and so

𝒜(γ+)=(2+6k+)π3(1εN)<(4+6k)π3(1+εM)=𝒜(γ).\mathcal{A}(\gamma_{+})=\frac{(2+6k_{+})\pi}{3}(1-\varepsilon_{N})<\frac{(4+6k_{-})\pi}{3}(1+\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

If the latter holds, then μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) implies k+<kk_{+}<k_{-}, and so

𝒜(γ+)=(4+6k+)π3(1+εN)<(2+6k)π3(1εM)=𝒜(γ).\mathcal{A}(\gamma_{+})=\frac{(4+6k_{+})\pi}{3}(1+\varepsilon_{N})<\frac{(2+6k_{-})\pi}{3}(1-\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

3.D If [γ±]T3,B[\gamma_{\pm}]\cong T_{3,B}, then the pair (γ+,γ)(\gamma_{+},\gamma_{-}) is either (𝒱4+6k+,2+6k)(\mathcal{V}^{4+6k_{+}},\mathcal{F}^{2+6k_{-}}) or (2+6k+,𝒱4+6k)(\mathcal{F}^{2+6k_{+}},\mathcal{V}^{4+6k_{-}}). If the former holds, then μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) implies k+<kk_{+}<k_{-}, and so

𝒜(γ+)=(4+6k+)π3(1εN)<(2+6k)π3(1+εM)=𝒜(γ).\mathcal{A}(\gamma_{+})=\frac{(4+6k_{+})\pi}{3}(1-\varepsilon_{N})<\frac{(2+6k_{-})\pi}{3}(1+\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

If the latter holds, then μCZ(γ+)<μCZ(γ)\mu_{\operatorname{CZ}}(\gamma_{+})<\mu_{\operatorname{CZ}}(\gamma_{-}) implies k+kk_{+}\leq k_{-}, and so

𝒜(γ+)=(2+6k+)π3(1+εN)<(4+6k)π3(1εM)=𝒜(γ).\mathcal{A}(\gamma_{+})=\frac{(2+6k_{+})\pi}{3}(1+\varepsilon_{N})<\frac{(4+6k_{-})\pi}{3}(1-\varepsilon_{M})=\mathcal{A}(\gamma_{-}).

Case 4: [γ±][\gamma_{\pm}] is a homotopy class not covered in Cases 1 - 3. Because every such homotopy class is represented by Reeb orbits either of type 𝒱\mathcal{V}, of type \mathcal{E}, or of type \mathcal{F}, then we see that γ±\gamma_{\pm} project to the same orbifold point of S2/S^{2}/{\mathbb{P}}. By Lemma 4.28(ii), 𝒜(γ+)<𝒜(γ)\mathcal{A}(\gamma_{+})<\mathcal{A}(\gamma_{-}).

References

  • [AHNS17] L. Abbrescia, I. Huq-Kuruvilla, J. Nelson, and N. Sultani, Reeb dynamics of the link of the AnA_{n} singularity, Involve, Vol. 10, no. 3, Mathematical Sciences Publishers, 2017.
  • [BEHWZ03] F. Bourgeois, Y. Eliashberg, H. Hofer, K. Wysocki, and E. Zehnder, Compactness results in symplectic field theory, Geom. Topol. 7 (2003), 799-888.
  • [BO17] F. Bourgeois and A. Oancea, S1S^{1}-equivariant symplectic homology and linearized contact homology, Int. Math. Res. Not. Volume 2017, Issue 13, July 2017, Pages 3849–3937.
  • [CH14] C. Cho and H. Hong, Orbifold Morse–Smale–Witten complexes, Internat. J. Math. 25 (2014), no. 5, 1450040, 35 pp.
  • [EGH00] Y. Eliashberg, A. Givental and H. Hofer, Introduction to symplectic field theory, Geom. Funct. Anal. 2000, Special Volume, Part II, 560–673.
  • [HM22] S. Haney and T. Mark, Cylindrical contact homology of 3-dimensional Brieskorn manifolds, Algebr. Geom. Topol. 22 (2022), no. 1, 153–187.
  • [HWZ96] H. Hofer, K. Wysocki and E. Zehnder, Properties of pseudoholomorphic curves in symplectisations. I. Asymptotics, Ann. Inst. H. Poincare Anal. Non Lineaire 13 (1996), 337–379.
  • [HWZ95] H. Hofer, K. Wysocki and E. Zehnder, Properties of pseudo-holomorphic curves in symplectisations. II. Embedding controls and algebraic invariants, Geom. Funct. Anal. 5 (1995), 270–328.
  • [Hu02b] M. Hutchings, An index inequality for embedded pseudoholomorphic curves in symplectizations, J. Eur. Math. Soc. 4 (2002), 313–361.
  • [Hu09] M. Hutchings, The embedded contact homology index revisited, New perspectives and challenges in symplectic field theory, 263–297, CRM Proc. Lecture Notes 49, Amer. Math. Soc., 2009.
  • [HN16] M. Hutchings and J. Nelson, Cylindrical contact homology for dynamically convex contact forms in three dimensions, J. Symplectic Geom. Volume 14 (2016), no. 4, 983-1012.
  • [HN22] M. Hutchings and J. Nelson, S1S^{1}-equivariant contact homology for hypertight contact forms, J. Topology 2022, (15) 1455-1539.
  • [HT09] M. Hutchings and C. H. Taubes, Gluing pseudoholomorphic curves along branched covered cylinders II, J. Symplectic Geom. 7 (2009), 29–133.
  • [McK80] J. McKay, Graphs, singularities, and finite groups, The Santa Cruz Conference on Finite Groups, pp. 183-186, Proc. Symp. Pure Math. Vol. 37. No. 183. 1980, AMS.
  • [MR] M. McLean and A. Ritter The McKay correspondence via Floer theory, J. Differential Geom. 124 (2023), no. 1, 113–168.
  • [MS15] D. McDuff and D. Salamon, Introduction to symplectic topology, 3rd ed., Oxford University Press, 2015.
  • [Ne15] J. Nelson, Automatic transversality in contact homology I: regularity, Abh. Math. Semin. Univ. Hambg. 85 (2015), no. 2, 125-179.
  • [Ne20] J. Nelson, Automatic transversality in contact homology II: filtrations and computations, Proc. Lond. Math. Soc. (3) 120 (2020), 853-917.
  • [Ne1] J. Nelson, A symplectic perspective of the simple singularities, https://math.rice.edu/~jkn3/nelson_simple.pdf, Accessed 7/8/21.
  • [NW23] J. Nelson and M. Weiler, Embedded contact homology of prequantization bundles, J. Symp. Geom. vol 21, issue 6, 2023, 1077-1189.
  • [NW24] J. Nelson and M. Weiler, Torus knot filtered embedded contact homology of the tight contact 3-sphere, J. Topology 2024, (7) Issue 2, 74 pages.
  • [NW2] J. Nelson and M. Weiler, Torus knotted Reeb dynamics and the Calabi invariant, arXiv:2310.18307
  • [Sal99] D. Salamon, Lectures on Floer homology, Symplectic geometry and topology (Park City, UT, 1997), volume 7 of IAS/Park City Math. Ser., 143–229. Amer. Math. Soc., Providence, RI, 1999.
  • [SZ92] D. Salamon and E. Zehnder, Morse theory for periodic solutions of Hamiltonian systems and the Maslov index, Comm. Pure Appl. Math. 45 (1992), no 10, 1303-1360.
  • [Si08] R. Siefring, Relative asymptotic behavior of pseudoholomorphic half-cylinders, Pure Appl. Math. 61 (2008).
  • [Si11] R. Siefring, Intersection theory of punctured pseudoholomorphic curves, Geom. Topol. 15 (2011), 2351–2457.
  • [Sl80] P. Slodowy, Simple Singularities and Simple Algebraic Groups, Lecture Notes in Math. 815, Springer-Verlag, New York (1980).
  • [St85] R. Steinberg, Finite subgroups of SU2SU_{2}, dynkin diagrams, and affine coxeter elements, Pacific Journal of Math. Vol. 118, No. 2, 1985.
  • [Wen10] C. Wendl, Automatic transversality and orbifolds of punctured holomorphic curves in dimension four, Comment. Math. Helv. 85 (2010), no. 2, 347-407.
  • [We20] C. Wendl, Lectures on contact 3-manifolds, holomorphic curves and intersection theory. Cambridge Tracts in Mathematics, 220, 2020.
  • [Wen-SFT] C. Wendl, Lectures on Symplectic Field Theory, to appear in EMS Lectures in Mathematics series, link to version 3, 2021; arXiv:1612.01009v2
  • [Za58] H. Zassenhaus, The Theory of Groups, 2nd ed., Chelsea Publishing Company, New York, 1958.

Leo Digiosia
Rice University, PhD 2022
Quantitative Model Development Analyst at US Bank
email:
[email protected]

Jo Nelson
Rice University
email:
[email protected]