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

Ellipsoidal superpotentials and singular curve counts

Dusa McDuff and Kyler Siegel
Abstract.

Given a closed symplectic manifold, we construct invariants which count (a) closed rational pseudoholomorphic curves with prescribed cusp singularities and (b) punctured rational pseudoholomorphic curves with ellipsoidal negative ends. We prove an explicit equivalence between these two frameworks, which in particular gives a new geometric interpretation of various counts in symplectic field theory. We show that these invariants encode important information about singular symplectic curves and stable symplectic embedding obstructions. We also prove a correspondence theorem between rigid unicuspidal curves and perfect exceptional classes, which we illustrate by classifying rigid unicuspidal (symplectic or algebraic) curves in the first Hirzebruch surface.

K.S. is partially supported by NSF grant DMS-2105578

1. Introduction

1.1. Motivation

In algebraic and symplectic geometry one often considers curves which satisfy various types of geometric constraints. For instance, Gromov–Witten invariants count curves passing through one or more cycles representing chosen homology classes, and relative Gromov–Witten invariants further specify tangency conditions with a chosen divisor. For certain applications, a distinguished role is played by constraints which are local near points, as these make no assumption on the ambient topology. For example, curves with several generic point constraints are the subjects of the celebrated formulas of Kontsevich [KM, KV] and Caporaso–Harris [CH], and they form the basis of the UU-map in embedded contact homology used to define the ECH capacities [Hut1].

More recently, curves with local tangency constraints have been used to put various obstructions on Lagrangain submanifolds and symplectic embeddings – see e.g. [CM1, CM2, Ton, MS3, MS5]. The local tangency constraint can be formulated in terms of a generic local divisor and it essentially amounts to specifying the mm-jet of a curve at a marked point, for some m0m\in\mathbb{Z}_{\geq 0}. In particular, for m=0m=0 this reduces to an ordinary point constraint.

In this paper we introduce a family of local geometric constraints which naturally generalize local tangency constraints and are related to singularities of algebraic curves. The resulting curve counts can be interpreted in several ways:

  1. (a)

    as closed curves satisfying local multidirectional tangency constraints

  2. (b)

    as closed curves with prescribed cusp singularities

  3. (c)

    as punctured curves which are negatively asymptotic (in the sense of symplectic field theory) to Reeb orbits in ellipsoids

  4. (d)

    as relative Gromov–Witten invariants with respect to certain nongeneric chains of divisors.

Interpretation (a) leads to a natural definition of these counts in the spirit of symplectic Gromov–Witten theory, while (b) will relate these to classical existence problems about singular algebraic curves, particularly those with one (p,q)(p,q) cusp singularity (this is locally modeled on {xp+yq=0}2\{x^{p}+y^{q}=0\}\subset\mathbb{C}^{2}). Interpretation (c) embeds these counts into the framework of SFT and will allow us to use them to obstruct (stabilized) symplectic embeddings of ellipsoids. Finally, (d) is related most directly to (b) via embedded resolution of curve singularities.

The main goals of this paper are:

  • to give rigorous self-contained definitions of these counts (under suitable assumptions)

  • to formalize the above interpretations (a)-(d) into precise equivalences between invariants

  • to discuss applications to symplectic embeddings and classifying singular (algebraic or symplectic) curves.

We will restrict to genus zero and primarily focus on constraints of an essentially (real) four-dimensional nature, although we allow target spaces of any dimension. For the remainder of this introduction we outline our main results in more detail.

1.2. Ellipsoidal superpotentials

We begin by discussing punctured curves with ellipsoidal negative ends. Symplectic field theory packages punctured curve counts in symplectizations and completed symplectic cobordisms into algebraic invariants which are known to provide powerful symplectic embedding obstructions. Of particular importance for this paper are cobordisms of the form MaM_{\vec{a}}, given by excising from a closed symplectic manifold MM a (rescaling of) the ellipsoid E(a)E({\vec{a}}) with area factors a=(a1,,an)>0n{\vec{a}}=(a_{1},\dots,a_{n})\in\mathbb{R}_{>0}^{n} (see §2). Given a homology class AH2(M)A\in H_{2}(M), we denote by 𝐓M,Aa\mathbf{T}_{M,A}^{\vec{a}} the SFT count of index zero planes in the symplectic completion M^a\widehat{M}_{\vec{a}} of MaM_{\vec{a}} with one negative end asymptotic to a Reeb orbit in E(a)\partial E({\vec{a}}). In [MS6] this count is called the ellipsoidal superpotential.

In general 𝐓M,Aa\mathbf{T}_{M,A}^{\vec{a}} is a virtual count of pseudoholomorphic buildings in a compactified moduli space and takes values in \mathbb{Q}. However, we will show that in favorable situations the count 𝐓M,Aa\mathbf{T}_{M,A}^{\vec{a}} can be defined by classical pseudoholomorphic curve techniques and takes values in \mathbb{Z}. In order to give a precise formulation let us first introduce some relevant notation. For a=(a1,,an)>0n{\vec{a}}=(a_{1},\dots,a_{n})\in\mathbb{R}_{>0}^{n} rationally independent111That is, a1,,ana_{1},\dots,a_{n} are linearly independent over \mathbb{Q}. We will frequently make this assumption out of convenience in order to ensure nondegenerate Reeb dynamics in E(a)\partial E({\vec{a}})., let 𝔬1a,𝔬2a,𝔬3a,\mathfrak{o}^{\vec{a}}_{1},\mathfrak{o}^{\vec{a}}_{2},\mathfrak{o}^{\vec{a}}_{3},\dots denote the Reeb orbits in E(a)\partial E({\vec{a}}) in order of increasing action. In particular, the action of 𝔬ka\mathfrak{o}^{\vec{a}}_{k} is given by 𝒜(𝔬ka)=𝕄ka\mathcal{A}(\mathfrak{o}^{\vec{a}}_{k})={\mathbb{M}}_{k}^{\vec{a}}, where 𝕄ka{\mathbb{M}}^{\vec{a}}_{k} denotes the kkth smallest positive integer multiple of one of a1,,ana_{1},\dots,a_{n}.222In general 𝒜(γ)\mathcal{A}(\gamma) will denote the action of a Reeb orbit γ\gamma, and we use the notation 𝕄ka{\mathbb{M}}_{k}^{\vec{a}} when we wish to emphasize its combinatorial nature in the case of ellipsoids. Recall that each Reeb orbit in E(a)\partial E({\vec{a}}) is an iterate of one of the simple orbits ν1,,νn\nu_{1},\dots,\nu_{n}, where νi\nu_{i} denotes the intersection of E(a)\partial E({\vec{a}}) with the iith complex axis. We will denote the covering multiplicity of any Reeb orbit γ\gamma by mult(γ){{\operatorname{mult}}}(\gamma).

Let 𝒥(Ma)\mathcal{J}(M_{\vec{a}}) denote the set of SFT admissible almost complex structures on M^a\widehat{M}_{\vec{a}} (see §2.2). Given J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}), let Ma,AJ(𝔬c1(A)1a)\mathcal{M}_{M_{\vec{a}},A}^{J}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) denote the moduli space of JJ-holomorphic planes333These need not be embedded. u:M^au:\mathbb{C}\rightarrow\widehat{M}_{\vec{a}} which lie in the homology class AA and are negatively asymptotic to 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}. Here c1(A)c_{1}(A)\in\mathbb{Z} denotes the first Chern number of AA, and 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1} is precisely the negative asymptotic Reeb orbit needed for this moduli space to have index zero.

The following theorem gives precise conditions under which the ellipsoidal superpotential is classically defined and integer-valued, in which case we get obstructions for stabilized symplectic embeddings of ellipsoids.

Theorem A (specialization of Theorem 2.3.5 and Corollary 2.7.2).

Let M2nM^{2n} be a semipositive closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, and a=(a1,,an){\vec{a}}=(a_{1},\dots,a_{n}) a rationally independent tuple satisfying Assumption (*), and such that AA and 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1} have no common divisibility.444That is, there is no κ2\kappa\in\mathbb{Z}_{\geq 2} such that 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1} is a κ\kappa-fold cover of another Reeb orbit and we have A=κBA=\kappa B for some BH2(M)B\in H_{2}(M). This holds for example if c1(A)c_{1}(A) and mult(𝔬c1(A)1a){{\operatorname{mult}}}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) are relatively prime. Then:

  1. (a)

    For generic J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}), the moduli space Ma,AJ(𝔬c1(A)1a)\mathcal{M}_{M_{\vec{a}},A}^{J}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is finite and regular, and its signed count is independent of the choice of generic JJ.

  2. (b)

    Suppose that this count is nonzero, and further that M×NM\times\mathbb{C}^{N} is semipositive for some N0N\in\mathbb{Z}_{\geq 0}. Then given any symplectic embedding E(ca)×N𝑠M×NE(c{\vec{a}})\times\mathbb{C}^{N}\overset{s}{\hookrightarrow}M\times\mathbb{C}^{N} we must have c[ωM]A𝕄c1(A)1ac\leq\frac{[\omega_{M}]\cdot A}{{\mathbb{M}}^{\vec{a}}_{c_{1}(A)-1}}.

Here [ωM]A[\omega_{M}]\cdot A denotes the area of AA with respect to the symplectic form ωM\omega_{M} on MM. Assumption (*) is our key numerical condition on a{\vec{a}} and c1(A)c_{1}(A), formulated in §2.3:

Assumption (*) (Assumptions A and B).

Assume that for any k2k\in\mathbb{Z}_{\geq 2} and i1,,ik1i_{1},\dots,i_{k}\in\mathbb{Z}_{\geq 1} satisfying s=1kis+k1c1(A)1\sum\limits_{s=1}^{k}i_{s}+k-1\leq c_{1}(A)-1 we have

𝕄i1a++𝕄ika𝕄i1++ik+k1a,\displaystyle{\mathbb{M}}^{\vec{a}}_{i_{1}}+\cdots+{\mathbb{M}}^{\vec{a}}_{i_{k}}\leq{\mathbb{M}}^{\vec{a}}_{i_{1}+\cdots+i_{k}+k-1},

with strict inequality if s=1kis+k1=c1(A)1\sum\limits_{s=1}^{k}i_{s}+k-1=c_{1}(A)-1.

Notation 1.2.1.

For M,A,aM,A,{\vec{a}} as in Theorem A, the ellipsoidal superpotential is

𝐓M,Aa:=#Ma,AJ(𝔬c1(A)1a)\displaystyle\mathbf{T}_{M,A}^{\vec{a}}:=\#\mathcal{M}_{M_{\vec{a}},A}^{J}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})\in\mathbb{Z}

for generic J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}).

Special cases of Theorem A in the case M=2M=\mathbb{CP}^{2} have appeared in e.g. [HK, CGH, McD2, CGHM]. Theorem A is applied in [MS4] to the monotone toric surfaces from [CGHMP], and in [MMW] to one point blowups of 2\mathbb{CP}^{2}. For the stabilized embedding obstructions in (b) it is essential that we count planes, as curves with higher genus or several negative ends do not typically behave well under stabilization (see §2.6). Note that M×NM\times\mathbb{C}^{N} is semipositive e.g. whenever MM is monotone or dimM=4\dim_{\mathbb{R}}M=4 and N=1N=1.

While Assumption (*) is somewhat mysterious in general, we have the following important specialization for which Assumption (*) always holds (see §2.3). Fix p,q1p,q\in\mathbb{Z}_{\geq 1} with p>qp>q relatively prime satisfying p+q=c1(A)p+q=c_{1}(A), and put a=(q,p±δ,a3,,an){\vec{a}}=(q,p\pm\delta,a_{3},\dots,a_{n}) with a3,,an>pqa_{3},\dots,a_{n}>pq and δ>0\delta>0 sufficiently small. In this case the Reeb orbit 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1} is either ν1p\nu_{1}^{p} or ν2q\nu_{2}^{q}, and in either case its action 𝕄c1(A)1a{\mathbb{M}}^{\vec{a}}_{c_{1}(A)-1} is approximately pqpq. In dimension four Assumption (*) is closely related to the ECH partition conditions – see Remark 2.3.4.

Remark 1.2.2.

The obstruction in Theorem A(b) actually follows from a stronger result (see §2) which obstructs embeddings of the form E(ca,cb)𝑠M×Q2NE(c{\vec{a}},c{\vec{b}})\overset{s}{\hookrightarrow}M\times Q^{2N} for any b=(b1,,bN){\vec{b}}=(b_{1},\dots,b_{N}) with b1,,bN>𝕄c1(A)1ab_{1},\dots,b_{N}>{\mathbb{M}}^{\vec{a}}_{c_{1}(A)-1} and Q2NQ^{2N} any closed symplectic manifold such that M×QM\times Q is semipositive. \Diamond

Remark 1.2.3 (on the assumptions in Theorem A).

Semipositivity is a standard assumption in symplectic Gromov–Witten theory which is needed to rule out multiple covers of negative index – see [MS1] and §2.3 below. The non-divisibility assumption on AA and 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1} is also needed is rule out multiple covers in the moduli space Ma,AJ(𝔬c1(A)1a)\mathcal{M}_{M_{\vec{a}},A}^{J}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}).

It is tempting to relax the Assumption (*) in Theorem A, but this likely requires a more sophisticated transversality setup (and results in values in \mathbb{Q} rather than \mathbb{Z}), due to the possibility of branched covers of trivial cylinders in ×E(a)\mathbb{R}\times\partial E({\vec{a}}) with nonpositive index (see Remark 2.4.3). In fact, [MS6, §7] gives examples where 𝐓M,Aa\mathbf{T}_{M,A}^{\vec{a}} cannot be an integer, e.g. 𝐓2,5[L](1,8+)=113/13\mathbf{T}_{\mathbb{CP}^{2},5[L]}^{(1,8^{+})}=113/13. Similar considerations will apply to Theorem B below. \Diamond

1.3. Multidirectional tangency constraints

We next discuss closed curve counts with local multidirectional tangency constraints, denoted by NM,A<𝒞mpt>N_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle. Roughly speaking, for m=(m1,,mn){\vec{m}}=(m_{1},\dots,m_{n}) this constraint requires curves to pass through a specified point in M2nM^{2n} with contact order mim_{i} to the iith complex direction. The most relevant case for us will be when m3==mn=1m_{3}=\cdots=m_{n}=1, i.e. we only impose constraints in the first two complex directions. For simplicity we restrict to curves carrying a single multidirectional tangency constraint, although this could be readily generalized to multiple constraints.

Given a symplectic manifold M2nM^{2n}, we will say that a collection of smooth local symplectic divisors 𝐃=(𝐃1,,𝐃n){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\dots,\mathbf{D}_{n}) through a point x0Mx_{0}\in M is spanning if their tangent spaces at x0x_{0} span Tx0MT_{x_{0}}M. For example, the complex hyperplanes {zi=0}\{z_{i}=0\} give a set of spanning local divisors through 0n\vec{0}\in\mathbb{C}^{n}. We denote by 𝒥(M,𝐃)\mathcal{J}(M,{\vec{\mathbf{D}}}) the space of all tame almost complex structures on MM which are integrable near x0x_{0} and preserve each of 𝐃1,,𝐃n\mathbf{D}_{1},\dots,\mathbf{D}_{n}.

Given J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}), we define M,AJ<𝒞𝐃mx0>\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle to be the moduli space of JJ-holomorphic maps u:1Mu:\mathbb{CP}^{1}\rightarrow M such that [u]=A[u]=A, u([0:0:1])=x0u([0:0:1])=x_{0}, and uu has contact order at least mim_{i} with 𝐃i\mathbf{D}_{i} at [0:0:1][0:0:1] for i=1,,ni=1,\dots,n.

Theorem B (specialization of Theorem 3.3.2).

Let M2nM^{2n} be a semipositive closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, 𝐃=(𝐃1,𝐃n){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\dots\mathbf{D}_{n}) a collection of spanning local divisors at a point x0Mx_{0}\in M, and m=(p,q,1,,1)1{\vec{m}}=(p,q,1,\dots,1)\in\mathbb{Z}_{\geq 1} a tuple such that p+q=c1(A)p+q=c_{1}(A) and gcd(p,q)=1\gcd(p,q)=1. Then for generic J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}), the moduli space M,AJ<𝒞𝐃mx0>\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}^{{\vec{m}}}_{{\vec{\mathbf{D}}}}x_{0}\Rangle is finite and regular, and its signed count is independent of the choices of x0,𝐃x_{0},{\vec{\mathbf{D}}}, and generic JJ.

Notation 1.3.1.

For M,A,mM,A,{\vec{m}} as in Theorem B, we put

NM,A<𝒞mpt>:=#M,AJ<𝒞𝐃mx0>\displaystyle N_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle:=\#\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}^{{\vec{m}}}_{{\vec{\mathbf{D}}}}x_{0}\Rangle\in\mathbb{Z}

for generic J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}).

Remark 1.3.2.

In the special case m2==mn=1m_{2}=\cdots=m_{n}=1, the constraint <𝒞𝐃mx0>\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle reduces to the local tangency constraint denoted by <𝒯𝐃1(m1)x0>\Langle\mathcal{T}_{\mathbf{D}_{1}}^{(m_{1})}x_{0}\Rangle in [MS3]. \Diamond

Remark 1.3.3.

Suppose dimM=4\dim_{\mathbb{R}}M=4, and we have p+q=c1(A)p+q=c_{1}(A) and gcd(p,q)=1\gcd(p,q)=1. As we make precise in §3.5, the local multidirectional tangency constraint <𝒞(p,q)pt>\Langle\mathcal{C}^{(p,q)}{\operatorname{pt}}\Rangle is akin to prescribing a (p,q)(p,q) cusp at x0x_{0} along with its maximal jet. In particular, for generic J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}), every curve CM,AJ<𝒞𝐃(p,q)x0>C\in\mathcal{M}^{J}_{M,A}\Langle\mathcal{C}^{(p,q)}_{{\vec{\mathbf{D}}}}x_{0}\Rangle has a (p,q)(p,q) cusp singularity. Conversely, if CC is a JJ-holomorphic curve with a (p,q)(p,q) cusp at x0x_{0}, then we can find spanning local divisors 𝐃=(𝐃1,𝐃2){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\mathbf{D}_{2}) at x0x_{0} such that the constraint <𝒞𝐃(p,q)x0>\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{(p,q)}x_{0}\Rangle is satisfied. Incidentally, the choice of 𝐃2\mathbf{D}_{2} is irrelevant if we assume p>qp>q. \Diamond

Example 1.3.4.

The constraint <𝒞𝐃(3,2)x0>\Langle\mathcal{C}^{(3,2)}_{{\vec{\mathbf{D}}}}x_{0}\Rangle corresponds to having an ordinary cusp at a specified point x0x_{0} and with specified tangent line. Recall that there is indeed a well-defined tangent line at an ordinary cusp even though it is a singular point. \Diamond

Let us briefly comment on the proof of Theorem B. Generalizing [MS3, Prop. 2.2.2] for local tangency constraints, the basic strategy is to show that any bad degenerations (i.e. multiple covers or stable maps with more than one component) only appear in real codimension at least two. In the case of local tangency constraints, the main difficulty comes from configurations involving ghost bubbles (i.e. the constraint is carried by a constant component), and these are handled by observing (via [CM1, Lem. 7.2]) that the nearby nonconstant components satisfy tangency constraints which “remember” the main constraint. In the case of local multidirectional tangency constraints this approach apparently fails to produce enough codimension, but we show in §3.2 that such ghost degenerations carry additional “hidden constraints”:

Proposition C (specialization of Proposition 3.2.4).

Suppose that a curve carrying the constraint <𝒞𝐃(p,q)x0>\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{(p,q)}x_{0}\Rangle degenerates into curves carrying constraints <𝒞𝐃(p1,q1)x0>,,<𝒞𝐃(pk,qk)x0>\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{(p_{1},q_{1})}x_{0}\Rangle,\dots,\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{(p_{k},q_{k})}x_{0}\Rangle respectively. Then we must have

i=1kmin(api,bqi)min(ap,bq)\displaystyle\sum_{i=1}^{k}\min(ap_{i},bq_{i})\geq\min(ap,bq) (1.3.1)

for any choice of a,b>0a,b\in\mathbb{R}_{>0}.

A more precise statement which applies in all dimensions is given in §3.2. Assuming gcd(p,q)=1\gcd(p,q)=1, (1.3.1) readily implies the inequality i=1k(pi+qi)p+q+1\sum_{i=1}^{k}(p_{i}+q_{i})\geq p+q+1, which is the main ingredient needed to show that bad degenerations have codimension at least two.

1.4. Correspondence theorem

We now formulate a precise relationship between negative ellipsoidal ends and multidirectional tangency constraints. Recall that M^a\widehat{M}_{\vec{a}} has a negative end modeled on 0×E(εa)\mathbb{R}_{\leq 0}\times\partial E(\varepsilon{\vec{a}}), and that a curve CMa,AJ(𝔬c1(A)1a)C\in\mathcal{M}_{M_{\vec{a}},A}^{J}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) has a negative end asymptotic to the Reeb orbit 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1} in (a rescaling of) E(a)\partial E({\vec{a}}). We can view the image of a small loop around the puncture as an embedded loop in E(a)\partial E({\vec{a}}) which has a well-defined linking number mim_{i}\in\mathbb{Z} with the (2n3)(2n-3)-sphere E(a){zi=0}\partial E({\vec{a}})\cap\{z_{i}=0\} for i=1,,ni=1,\dots,n. Using results from [MS6, §4], for a careful choice of JJ we can directly transform CC into a JJ^{\prime}-holomorphic sphere in MM satisfying the constraint <𝒞𝐃mx0>\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{\vec{m}}x_{0}\Rangle for suitable J,𝐃J^{\prime},{\vec{\mathbf{D}}}.

Furthermore, the linking numbers m=(m1,,mn){\vec{m}}=(m_{1},\dots,m_{n}) associated to the puncture of CC are generically given by Δc1(A)1a\Delta^{\vec{a}}_{c_{1}(A)-1}, which is defined combinatorially as follows.

Definition 1.4.1.

For rationally independent a>0n{\vec{a}}\in\mathbb{R}_{>0}^{n} and k1k\in\mathbb{Z}_{\geq 1}, let Δka\Delta_{k}^{\vec{a}} denote the tuple (i1,,in)1(i_{1},\dots,i_{n})\in\mathbb{Z}_{\geq 1} which maximizes min1snasis\min\limits_{1\leq s\leq n}a_{s}i_{s} subject to s=1nis=n+k1\sum\limits_{s=1}^{n}i_{s}=n+k-1.

Example 1.4.2.

For a=(2,3+){\vec{a}}=(2,3^{+}), we have

kk 1 2 3 4 5 6 7 8
Δka\Delta^{\vec{a}}_{k} (1,1) (2,1) (2,2) (3,2) (4,2) (4,3) (5,3) (5,4)

.

\Diamond

The lattice path Δa=(Δ1a,Δ2a,Δ3a,)\Delta^{\vec{a}}=(\Delta^{\vec{a}}_{1},\Delta^{\vec{a}}_{2},\Delta^{\vec{a}}_{3},\dots) in 1n\mathbb{Z}_{\geq 1}^{n} will play a central role in relating negative ellipsoidal ends with multidirectional tangencies. It is the natural analogue of the lattice path Γa\Gamma^{\vec{a}} from [MS6, §1.1] which instead relates to positive ends (in fact we have Δka=Γk1a+(1,,1)\Delta^{\vec{a}}_{k}=\Gamma^{\vec{a}}_{k-1}+(1,\dots,1)).

Note that for a3,,ana1,a2a_{3},\dots,a_{n}\gg a_{1},a_{2} we have Δka=(p,q,1,,1)\Delta^{\vec{a}}_{k}=(p,q,1,\dots,1) for some p,q1p,q\in\mathbb{Z}_{\geq 1}. In this case it is easy to check that 𝔬ka\mathfrak{o}^{\vec{a}}_{k} is either ν1p\nu_{1}^{p} or ν2q\nu_{2}^{q}, depending on whether a1pa_{1}p or a2qa_{2}q is smaller.

Theorem D (specialization of Theorem 3.4.1).

Let M2nM^{2n} be a semipositive symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, and a=(a1,,an)>0n{\vec{a}}=(a_{1},\dots,a_{n})\in\mathbb{R}_{>0}^{n} a rationally independent tuple with a3,,an>pqa_{3},\dots,a_{n}>pq. Put (p,q,1,1):=Δc1(A)1a=:m(p,q,1\dots,1):=\Delta^{\vec{a}}_{c_{1}(A)-1}=:{\vec{m}}, and assume that gcd(p,q)=1\gcd(p,q)=1. Then we have 𝐓M,Aa=NM,A<𝒞mpt>\mathbf{T}_{M,A}^{{\vec{a}}}=N_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle.

Remark 1.4.3.

Let Esk4=E(a1,a2)E_{\operatorname{sk}}^{4}=E(a_{1},a_{2}) with a2a1a_{2}\gg a_{1} denote the four-dimensional “skinny ellipsoid” (up to rescaling). It is shown in [MS3] that the local tangency constraint <𝒯(m)pt>\Langle\mathcal{T}^{(m)}{\operatorname{pt}}\Rangle is interchangeable with a negative end asymptotic to the Reeb orbit 𝔬k=ν1k\mathfrak{o}_{k}=\nu_{1}^{k} in Esk4\partial E_{\operatorname{sk}}^{4}. Theorem D generalizes this to multidirectional tangency constraints and arbitrary ellipsoids. \Diamond

1.5. Singular symplectic and algebraic curves

It is important to understand when 𝐓M,Aa\mathbf{T}_{M,A}^{\vec{a}} is nonzero, as by Theorem A this obstructs symplectic embeddings of the form E(ca)×N𝑠M×NE(c{\vec{a}})\times\mathbb{C}^{N}\overset{s}{\hookrightarrow}M\times\mathbb{C}^{N}. In fact, for N1N\geq 1, all such obstructions known to us are of this form, and at least in the case M=2M=\mathbb{CP}^{2} it is expected that these provide a complete set of obstructions. The recent paper [MS6] provides tools to compute 𝐓M,Aa\mathbf{T}_{M,A}^{\vec{a}} (and hence also NM,A<𝒞mpt>N_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle) by purely combinatorial methods, but many mysteries remain.

Using Theorem D, we recast this problem in §3.5 in much more geometric terms, opening new avenues to prove nonvanishing results. A (𝐩,𝐪)\mathbf{(p,q)}-sesquicuspidal symplectic curve in a symplectic four-manifold M4M^{4} is a subset which has one singularity modeled on the (p,q)(p,q) cusp, and which is otherwise a positively immersed symplectic submanifold (see Definition 3.5.1). One can show using the adjunction formula that the number of double points must be 12(2+AAc1(A)(p1)(q1))\tfrac{1}{2}(2+A\cdot A-c_{1}(A)-(p-1)(q-1)). In §3.5 we prove:

Theorem E.

Fix M4M^{4} a four-dimensional closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, and p,q1p,q\in\mathbb{Z}_{\geq 1} with gcd(p,q)=1\gcd(p,q)=1 and p+q=c1(A)p+q=c_{1}(A). Then we have NM,A<𝒞(p,q)pt>0N_{M,A}\Langle\mathcal{C}^{(p,q)}{\operatorname{pt}}\Rangle\in\mathbb{Z}_{\geq 0}, with NM,A<𝒞(p,q)pt>>0N_{M,A}\Langle\mathcal{C}^{(p,q)}{\operatorname{pt}}\Rangle>0 if and only if there exists a rational (p,q)(p,q)-sesquicuspidal symplectic curve in MM lying in homology class AA.

Since any singularity of an algebraic curve can be symplectically perturbed into a finite set of positive double points, we have:

Corollary F.

If M4M^{4} is a smooth projective surface containing an index zero irreducible rational algebraic curve in homology class AA with a (p,q)(p,q) cusp (and possibly other singularities), then we have 𝐓M,A(p,q)>0\mathbf{T}_{M,A}^{(p,q)}>0.

This allows us to study 𝐓M,Aa\mathbf{T}_{M,A}^{\vec{a}} (or equivalently NM,A<𝒞mpt>N_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle) by constructing algebraic curves and potentially importing techniques from algebraic geometry. As an illustration, Orevkov [Ore] constructed an infinite sequence of index zero unicuspidal rational algebraic curves in 2\mathbb{CP}^{2}, so by Corollary F there is a corresponding infinite sequence of nonvanishing values for 𝐓2,d[L](p,q)\mathbf{T}_{\mathbb{CP}^{2},d[L]}^{(p,q)}, and these turn out to give precisely the symplectic embedding obstructions at the outer corners of the Fibonacci staircase [MS2] via Theorem A(b). In the forthcoming work [MS4] we generalize Orevkov’s construction in two independent directions, by constructing new families of index zero rational sesquicuspidal algebraic plane curves and by replacing 2\mathbb{CP}^{2} with various other toric surfaces.

Remark 1.5.1.

Under the assumptions of Theorem E, one can similarly show that we have 𝐓M,A(p,q)>0\mathbf{T}_{M,A}^{(p,q)}>0 if and only if the transverse (p,q)(p,q) torus knot 𝕋p,qS3\mathbb{T}_{p,q}\subset S^{3} with maximal self-linking number has a genus zero positively immersed symplectic hat (in the sense of [EG]) in MM in homology class AA. \Diamond

1.6. Unicuspidal curves and perfect exceptional classes

Lastly, we discuss applications to existence questions for singular curves. While the previous subsection highlights the relevance of index zero sesquicuspidal curves in symplectic four-manifolds, we focus here on the special case of curves which are unicuspidal curves, i.e. having a (p,q)(p,q) cusp and no other singularities.

According to [FdBLMHN], the aforementioned curves constructed by Orevkov are the only index zero unicuspidal rational algebraic plane curves. Before generalizing this result we must recall a bit more terminology. Firstly, given p,q1p,q\in\mathbb{Z}_{\geq 1} with gcd(p,q)=1\gcd(p,q)=1 and p>qp>q, let 𝒲(p,q)=(m1,,mL)1L{\mathcal{W}}(p,q)=(m_{1},\dots,m_{L})\in\mathbb{Z}_{\geq 1}^{L} denote the corresponding weight sequence. As we recall in §4.1, this is related to the continued fraction expansion of p/qp/q and controls (among other things) the resolution of the (p,q)(p,q) cusp singularity.

Given a closed symplectic four-manifold M4M^{4}, a homology class BH2(M)B\in H_{2}(M) is exceptional if we have BB=1B\cdot B=-1, c1(B)=1c_{1}(B)=1, and BB is represented by a symplectically embedded two-sphere. We will say that AH2(M)A\in H_{2}(M) is (𝐩,𝐪)\mathbf{(p,q)}-perfect exceptional if A~:=Ai=1Lm1e1mLeLH2(M~)\widetilde{A}:=A-\sum_{i=1}^{L}m_{1}e_{1}-\cdots-m_{L}e_{L}\in H_{2}(\widetilde{M}) is an exceptional class, where 𝒲(p,q)=(m1,,mL){\mathcal{W}}(p,q)=(m_{1},\dots,m_{L}). Here M~\widetilde{M} is the LL-point blowup of MM, and we have the identification H2(M~)H2(M)e1,,eLH_{2}(\widetilde{M})\cong H_{2}(M)\oplus\langle e_{1},\dots,e_{L}\rangle, where e1,,eLe_{1},\dots,e_{L} are exceptional classes.

In §4.4 we prove:

Theorem G.

Fix a M4M^{4} a symplectic four-manifold and AH2(M)A\in H_{2}(M) a homology class. There is an index zero rational (p,q)(p,q)-unicuspidal symplectic curve in MM in homology class AA if and only if AA is (p,q)(p,q)-perfect exceptional. Moreover, in this case we have NM,A<𝒞(p,q)pt>=1N_{M,A}\Langle\mathcal{C}^{(p,q)}{\operatorname{pt}}\Rangle=1.

As an illustration of Theorem G, we note that the perfect exceptional homology classes in the first Hirzebruch surface F1F_{1} were studied comprehensively in [MMW] as part of their study of infinite staircases in the symplectic ellipsoid embedding functions of one-point blowups of 2\mathbb{CP}^{2}. Put H2(F1)=,eH_{2}(F_{1})=\langle\ell,e\rangle, where \ell is the line class and ee is the exceptional divisor class. Let a1,a2,a3,a_{1},a_{2},a_{3},\dots be the sequence defined by the recursion aj+6=6aj+3aja_{j+6}=6a_{j+3}-a_{j} with initial values 1,1,1,1,2,41,1,1,1,2,4, and put

  • tj:=aj+326aj+3aj+aj2+8t_{j}:=\sqrt{a_{j+3}^{2}-6a_{j+3}a_{j}+a_{j}^{2}+8}

  • dj:=18(3aj+3+3aj+(1)j+1tj)d_{j}:=\tfrac{1}{8}(3a_{j+3}+3a_{j}+(-1)^{j+1}t_{j})

  • mj:=18(aj+3+aj+(1)j+13tj)m_{j}:=\tfrac{1}{8}(a_{j+3}+a_{j}+(-1)^{j+1}3t_{j})

for j1j\in\mathbb{Z}_{\geq 1}.

Corollary H.

There is an index zero (p,q)(p,q)-unicuspidal rational symplectic curve in F1F_{1} in homology class A=dmeA=d\ell-me and satisfying p/q<3+22p/q<3+2\sqrt{2} if and only if (p,q,d,m)=(aj+3,aj,dj,mj)(p,q,d,m)=(a_{j+3},a_{j},d_{j},m_{j}) for some j1j\in\mathbb{Z}_{\geq 1}.

In fact, the constructions in [MS4] show that each of these is realized by an algebraic curve. This could be viewed as a version of symplectic isotopy problem for singular symplectic curves in the spirit of [GS].

Corollary H should be compared with (the index zero part of) [FdBLMHN, Thm 1.1], but with 2\mathbb{CP}^{2} replaced by F1F_{1}. The case p/q>3+22p/q>3+2\sqrt{2} (sans algebraicity) can also be deduced from the results in [MMW], but it is considerably more complicated; we discuss this briefly in §4.5.

Acknowledgements

K.S. benefited from many helpful discussions with Grisha Mikhalkin.

2. Ellipsoidal superpotentials

The main technical result in this section is Theorem 2.3.5, which establishes conditions under which the ellipsoidal superpotential is robust (see Definition 2.3.1), which roughly means well-defined by classical transversality techniques and independent of any auxiliary choices. Together with Corollary 2.7.2, this implies Theorem A.

After some setting up some geometric preliminaries in §2.1 and §2.2, we state our main result and some of its consequences in §2.3. In §2.4 we give the proof of Theorem 2.3.5 assuming several lemmas about curves and their branched covers, which are proved in §2.5. In §2.6 we give criteria under which moduli spaces are stabilization invariant. Finally, in §2.7, we illustrate how (stable) symplectic embedding obstructions are naturally seen from this framework.

2.1. Symplectic embeddings of small ellipsoids

Let M2nM^{2n} be a closed symplectic manifold, and let E(a)E({\vec{a}}) be the symplectic ellipsoid with area factors a=(a1,,an)>0n{\vec{a}}=(a_{1},\dots,a_{n})\in\mathbb{R}_{>0}^{n}. For ε>0\varepsilon>0 sufficiently small, there exists a symplectic embedding ι\iota of the scaled ellipsoid E(εa)=E(εa1,,εan)E(\varepsilon{\vec{a}})=E(\varepsilon a_{1},\dots,\varepsilon a_{n}) into MM. In fact, up to further shrinking ε\varepsilon, this embedding is unique up to Hamiltonian isotopy:

Lemma 2.1.1.

Let ι0,ι1:E(εa)𝑠M\iota_{0},\iota_{1}:E(\varepsilon{\vec{a}})\overset{s}{\hookrightarrow}M be symplectic embeddings. Then for any ε>0\varepsilon^{\prime}>0 sufficiently small there is a smooth family of symplectic embeddings {ιt:E(εa)𝑠M|t[0,1]}\{\iota^{\prime}_{t}:E(\varepsilon^{\prime}{\vec{a}})\overset{s}{\hookrightarrow}M\;|\;t\in[0,1]\} such that ιi=ιi|E(εa)\iota^{\prime}_{i}=\iota_{i}|_{E(\varepsilon^{\prime}{\vec{a}})} for i=0,1i=0,1.

Proof.

After post-composing with a Hamiltonian isotopy of MM which sends ι1(0)\iota_{1}(0) to ι0(0)\iota_{0}(0), we can assume ι0(0)=ι1(0)=:p\iota_{0}(0)=\iota_{1}(0)=:p. We can then find ε>0\varepsilon^{\prime}>0 sufficiently small so that the images of both ι0:=ι0|E(εa)\iota_{0}^{\prime}:=\iota_{0}|_{E(\varepsilon^{\prime}{\vec{a}})} and ι1:=ι1|E(εa)\iota_{1}^{\prime}:=\iota_{1}|_{E(\varepsilon^{\prime}{\vec{a}})} lie in a Darboux ball in MM centered at pp. We thereby view ι0,ι1\iota_{0}^{\prime},\iota_{1}^{\prime} as symplectic embeddings E(εa)𝑠B2nE(\varepsilon^{\prime}{\vec{a}})\overset{s}{\hookrightarrow}B\subset\mathbb{R}^{2n}, where BB is a 2n2n-dimensional ball of some radius centered at the origin. After further shrinking ε\varepsilon^{\prime}, we can assume that BB also contains the standard E(εa)2nE(\varepsilon^{\prime}{\vec{a}})\subset\mathbb{R}^{2n}. It suffices to find a family of symplectic embeddings E(εa)𝑠BE(\varepsilon^{\prime}{\vec{a}})\overset{s}{\hookrightarrow}B interpolating between ι0\iota_{0}^{\prime} and ι1\iota_{1}^{\prime}.

By the “extension after restriction” principle (see e.g. [Sch, §4.4]), ι0:E(εa)𝑠B2n\iota_{0}^{\prime}:E(\varepsilon^{\prime}{\vec{a}})\overset{s}{\hookrightarrow}B^{2n} extends to a Hamiltonian diffeomorphism of 2n\mathbb{R}^{2n}. In particular, there is a Hamiltonian isotopy {ϕtHam(2n)|t[0,1]}\{\phi_{t}\in{\operatorname{Ham}}(\mathbb{R}^{2n})\;|\;t\in[0,1]\} with ϕ0=𝟙\phi_{0}=\mathbb{1} such that ϕ1|E(εa)=ι0\phi_{1}|_{E(\varepsilon^{\prime}{\vec{a}})}=\iota_{0}^{\prime}. Note that ϕt|E(εa)\phi_{t}|_{E(\varepsilon^{\prime}{\vec{a}})} has image in BB for t=0,1t=0,1, but not necessarily for all t(0,1)t\in(0,1). However, an inspection of the construction of the Hamiltonian isotopy via the Alexander trick as (c.f. the explanation in [Sch, §4.4]) shows that we can further assume ϕt(0)=0\phi_{t}(0)=0 for all t[0,1]t\in[0,1]. Therefore, after further shrinking ε\varepsilon^{\prime}, we can arrange that ϕt|E(εa)\phi_{t}|_{E(\varepsilon^{\prime}{\vec{a}})} has image in BB for all t[0,1]t\in[0,1]. By concatening this family with (the time reversal of) the analogous one given by applying the same considerations to ι1\iota^{\prime}_{1}, this concludes the proof. ∎

Remark 2.1.2.

The proof above actually applies equally if we replace the ellipsoid E(a)E({\vec{a}}) with any star-shaped domain in 2n\mathbb{R}^{2n}. \Diamond

Remark 2.1.3.

Since H1(E(a);)=0H^{1}(E({\vec{a}});\mathbb{R})=0, given any family of symplectic embeddings {ιt:E(εa)𝑠M}\{\iota_{t}^{\prime}:E(\varepsilon^{\prime}{\vec{a}})\overset{s}{\hookrightarrow}M\} there is a Hamiltonian isotopy {ϕtHam(M)}\{\phi_{t}\in{\operatorname{Ham}}(M)\} such that ϕ0=𝟙\phi_{0}=\mathbb{1} and ιt=ϕtι0\iota_{t}^{\prime}=\phi_{t}\circ\iota^{\prime}_{0}. In particular, there is a symplectomorphism between Mι0(E̊(εa))M\setminus\iota_{0}^{\prime}({\mathring{E}}(\varepsilon^{\prime}{\vec{a}})) and Mι1(E̊(εa))M\setminus\iota_{1}^{\prime}({\mathring{E}}(\varepsilon^{\prime}{\vec{a}})). Here E̊(a){\mathring{E}}({\vec{a}}) denotes the open ellipsoid {πi=1n|zi|2/ai<1}\{\pi\sum_{i=1}^{n}|z_{i}|^{2}/a_{i}<1\}. \Diamond

2.2. Moduli spaces of punctured curves

Here we briefly discuss some geometric preliminaries and relevant notions about pseudoholomorphic curves, taking the occasion to set notation and terminology.

2.2.1. Spaces with negative ellipsoidal ends

Fix a closed symplectic manifold M2nM^{2n}, a homology class AH2(M)A\in H_{2}(M), and a vector a=(a1,,an)>0n{\vec{a}}=(a_{1},\dots,a_{n})\in\mathbb{R}_{>0}^{n} whose components are rationally independent.

Notation 2.2.1.

We put Ma:=Mι(E̊(εa))M_{\vec{a}}:=M\,\setminus\,\iota({\mathring{E}}(\varepsilon{\vec{a}})), where ι:E(εa)𝑠M\iota:E(\varepsilon{\vec{a}})\overset{s}{\hookrightarrow}M is a symplectic embedding for some ε>0\varepsilon>0.

Such an embedding ι\iota always exists for ε>0\varepsilon>0 sufficiently small, e.g. with image contained in a Darboux chart. We denote the symplectic completion of MaM_{\vec{a}} by

M^a:=Ma(0×Ma).\displaystyle\widehat{M}_{\vec{a}}:=M_{\vec{a}}\cup(\mathbb{R}_{\leq 0}\times\partial M_{\vec{a}}).

2.2.2. Almost complex structures

Given a contact manifold YY (typically Y=E(a)Y=\partial E({\vec{a}})), we denote by 𝒥(Y)\mathcal{J}(Y) the space of admissible almost complex structures on the symplectization ×Y\mathbb{R}\times Y (see e.g. [MS5, §2.1.2]). In particular, any J𝒥(Y)J\in\mathcal{J}(Y) is invariant under translations in the first factor of ×Y\mathbb{R}\times Y.

Similarly, given a compact symplectic cobordism XX with positive contact boundary +X\partial^{+}X and negative contact boundary X\partial^{-}X, we denote by 𝒥(X)\mathcal{J}(X) the space of admissible almost complex structures on the symplectic completion X^\widehat{X} of XX. It will be more convenient to require J𝒥(X)J\in\mathcal{J}(X) to be only tame rather than compatible on the compact part XX^X\subset\widehat{X}, which suffices for all of our purposes.

Our main example is X=MaX=M_{\vec{a}}, which has empty positive boundary and negative boundary MaE(εa)\partial M_{\vec{a}}\cong\partial E(\varepsilon{\vec{a}}) for some small ε>0\varepsilon>0. By definition J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}) is an almost complex structure on M^a\widehat{M}_{\vec{a}} which is tame on the compact piece MaM_{\vec{a}} and whose restriction to the negative end 0×Ma\mathbb{R}_{\leq 0}\times\partial M_{\vec{a}} agrees with the restriction of some admissible almost complex structure J|Ma𝒥(Ma)J|_{M_{\vec{a}}}\in\mathcal{J}(\partial M_{\vec{a}}) on ×Ma\mathbb{R}\times\partial M_{\vec{a}}.

2.2.3. Moduli spaces

Given a collection of Reeb orbits Γ=(γ1,,γk)\Gamma=(\gamma_{1},\dots,\gamma_{k}) in E(a)\partial E({\vec{a}}) and an almost complex structure J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}), we denote by MaJ(Γ)\mathcal{M}^{J}_{M_{\vec{a}}}(\Gamma) the moduli space of JJ-holomorphic maps uu from a kk-punctured Riemann sphere to M^a\widehat{M}_{\vec{a}}, such that uu is asymptotic at the iith puncture to γi\gamma_{i} for i=1,,ki=1,\dots,k. Here the conformal structure on the domain (or equivalently the locations of the punctures) is arbitrary, and elements of MaJ(Γ)\mathcal{M}^{J}_{M_{\vec{a}}}(\Gamma) (usually referred to as simply “curves”) are taken modulo biholomorphic reparametrization.

Since H1(E(a);)=H2(E(a);)=0H^{1}(E({\vec{a}});\mathbb{R})=H^{2}(E({\vec{a}});\mathbb{R})=0, any curve CMaJ(Γ)C\in\mathcal{M}^{J}_{M_{\vec{a}}}(\Gamma) can be uniquely filled so as to give a well-defined homology class [C]H2(M)[C]\in H_{2}(M). Given AH2(M)A\in H_{2}(M), we put Ma,AJ(Γ):={CMaJ(Γ)|[C]=A}.\mathcal{M}^{J}_{M_{\vec{a}},A}(\Gamma):=\{C\in\mathcal{M}^{J}_{M_{\vec{a}}}(\Gamma)\;|\;[C]=A\}. We will sometimes suppress the almost complex structure from the notation and write simply Ma,A(Γ)\mathcal{M}_{M_{\vec{a}},A}(\Gamma) when the choice of JJ is implicit or immaterial. We write Ma,A(γ)\mathcal{M}_{M_{\vec{a}},A}(\gamma) when Γ\Gamma consists of a single Reeb orbit γ\gamma.

Similarly, given collections of Reeb orbits Γ+=(γ1+,,γk++)\Gamma^{+}=(\gamma_{1}^{+},\dots,\gamma^{+}_{k^{+}}) and Γ=(γ1,,γk)\Gamma^{-}=(\gamma^{-}_{1},\dots,\gamma^{-}_{k^{-}}) in E(a)\partial E({\vec{a}}) and an almost complex structure J𝒥(E(a))J\in\mathcal{J}(\partial E({\vec{a}})), we denote by E(a)J(Γ+;Γ)\mathcal{M}^{J}_{\partial E({\vec{a}})}(\Gamma^{+};\Gamma^{-}) the moduli space of maps uu from a Riemann sphere having k+k^{+} positive punctures and kk^{-} negative punctures to the symplectization ×E(a)\mathbb{R}\times\partial E({\vec{a}}), such that uu is asymptotic to the respective orbits Γ+\Gamma^{+} at the positive punctures and to Γ\Gamma^{-} at the negative punctures. Since JE(a)J\in\partial E({\vec{a}}) is translation invariant, this moduli space inherits an \mathbb{R} action which translates in the first factor of the target space, and we denote the quotient space by E(a)J(Γ+;Γ)/\mathcal{M}^{J}_{\partial E({\vec{a}})}(\Gamma^{+};\Gamma^{-})/\mathbb{R}.

2.2.4. Index formulas

Given a homology class AH2(M)A\in H_{2}(M) and Reeb orbits Γ=(γ1,,γk)\Gamma=(\gamma_{1},\dots,\gamma_{k}) in E(a)\partial E({\vec{a}}), recall that the expected dimension of the moduli space Ma,A(Γ)\mathcal{M}_{M_{\vec{a}},A}(\Gamma) is given by

indMa,A(Γ)=(n3)(2k)+2c1(A)i=1kCZ(γi).\displaystyle{\operatorname{ind}}\,\mathcal{M}_{M_{\vec{a}},A}(\Gamma)=(n-3)(2-k)+2c_{1}(A)-\sum_{i=1}^{k}{\operatorname{CZ}}(\gamma_{i}).

Here c1(A)c_{1}(A) is the first Chern number of AA, and CZ(γi){\operatorname{CZ}}(\gamma_{i}) is the Conley–Zehnder index of γi\gamma_{i} with respect to the (unique up to homotopy) global trivialization of the contact hyperplane distribution over E(a)\partial E({\vec{a}}). In particular, for the Reeb orbit 𝔬ja\mathfrak{o}_{j}^{\vec{a}} in E(a)\partial E({\vec{a}}) we have CZ(𝔬ja)=n1+2j{\operatorname{CZ}}(\mathfrak{o}_{j}^{\vec{a}})=n-1+2j (see e.g. equation (2.3.1) below and [GH, §2.1]). As usual we put dimM=2n\dim M=2n.

By definition the ellipsoidal superpotential 𝐓M,Aa\mathbf{T}_{M,A}^{\vec{a}} counts planes, i.e. k=1k=1, and we have indMa,A(γ)=0{\operatorname{ind}}\,\mathcal{M}_{M_{\vec{a}},A}(\gamma)=0 if and only if CZ(γ)=n3+2c1(A){\operatorname{CZ}}(\gamma)=n-3+2c_{1}(A), i.e. γ=𝔬c1(A)1a\gamma=\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}. Standard transversality results (see e.g. [Wen2, §8]) show that for generic J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}) the moduli space Ma,AJ(γ)\mathcal{M}^{J}_{M_{\vec{a}},A}(\gamma) is smooth and of the expected dimension near any simple curve CC.

Similarly, given Reeb orbits Γ+=(γ1+,,γk++)\Gamma^{+}=(\gamma_{1}^{+},\dots,\gamma^{+}_{k^{+}}) and Γ=(γ1,,γk)\Gamma^{-}=(\gamma_{1}^{-},\dots,\gamma^{-}_{k^{-}}), the index of the moduli space E(a)(Γ+;Γ)/\mathcal{M}_{\partial E({\vec{a}})}(\Gamma^{+};\Gamma^{-})/\mathbb{R} is given by

indE(a)(Γ+;Γ)1=(n3)(2k+k)+i=1k+CZ(γi+)j=1kCZ(γj)1.\displaystyle{\operatorname{ind}}\,\mathcal{M}_{\partial E({\vec{a}})}(\Gamma^{+};\Gamma^{-})-1=(n-3)(2-k^{+}-k^{-})+\sum_{i=1}^{k^{+}}{\operatorname{CZ}}(\gamma^{+}_{i})-\sum_{j=1}^{k^{-}}{\operatorname{CZ}}(\gamma^{-}_{j})-1.

Standard transversality results show that, for generic J𝒥(E(a))J\in\mathcal{J}(\partial E({\vec{a}})), the moduli space E(a)(Γ+;Γ)/\mathcal{M}_{\partial E({\vec{a}})}(\Gamma^{+};\Gamma^{-})/\mathbb{R} is smooth and of the expected dimension near any simple curve CC which is not a trivial cylinder.

We will also need to consider transversality in generic one-parameter families. Given a one-parameter family of almost complex structures {Jt𝒥(Ma)|t[0,1]}\{J_{t}\in\mathcal{J}(M_{\vec{a}})\;|\;t\in[0,1]\}, we consider the associated the parametrized moduli space

Ma,A{Jt}(Γ)={(u,t)|t[0,1],uMa,AJt(Γ)}.\displaystyle\mathcal{M}_{M_{\vec{a}},A}^{\{J_{t}\}}(\Gamma)=\{(u,t)\;|\;t\in[0,1],\;u\in\mathcal{M}^{J_{t}}_{M_{\vec{a}},A}(\Gamma)\}.

If the family {Jt}\{J_{t}\} is generic, this is a smooth manifold of dimension 2n5+2c1(A)i=1kCZ(γi)2n-5+2c_{1}(A)-\sum_{i=1}^{k}{\operatorname{CZ}}(\gamma_{i}) near any simple curve (i.e. a pair (u,t)(u,t) with uu simple). Given a one-parameter family {Jt𝒥(E(a))|t[0,1]}\{J_{t}\in\mathcal{J}(\partial E({\vec{a}}))\;|\;t\in[0,1]\}, we define the parametrized moduli space E(a){Jt}(Γ+,Γ)\mathcal{M}_{\partial E({\vec{a}})}^{\{J_{t}\}}(\Gamma^{+},\Gamma^{-}) in a similar manner. We denote the quotient by target translations by E(a){Jt}(Γ+,Γ)/\mathcal{M}_{\partial E({\vec{a}})}^{\{J_{t}\}}(\Gamma^{+},\Gamma^{-})/\mathbb{R}. For a generic one-parameter family {Jt}\{J_{t}\}, this is a smooth manifold of dimension 2n6+i=1k+CZ(γi+)k=1kCZ(γj)2n-6+\sum_{i=1}^{k^{+}}{\operatorname{CZ}}(\gamma_{i}^{+})-\sum_{k=1}^{k^{-}}{\operatorname{CZ}}(\gamma_{j}^{-}) near any simple curve.

2.2.5. Formal curves

Following the usage in [MS5], the language of formal curves provides a convenient bookkeeping tool for making index arguments. The basic idea is that, for certain combinatorial purposes, we can formally glue together several curve components in a pseudoholomorphic building in order to treat them at a single formal curve. In such contexts, it is usually irrelevant whether or not an actual analytic gluing exists.

Here we give definitions specialized to the situations most relevant for us.

Definition 2.2.2.

A genus zero formal curve component CC in MaM_{\vec{a}} is a pair (Γ,A)(\Gamma,A), where

  • Γ=(γ1,,γk)\Gamma=(\gamma_{1},\dots,\gamma_{k}) is a tuple of Reeb orbits in E(a)\partial E({\vec{a}})

  • AH2(M)A\in H_{2}(M) is a homology class

  • we require the energy

    (C):=Aωi=1k𝒜(γi)\displaystyle\mathcal{E}(C):=\int_{A}\omega-\sum_{i=1}^{k}\mathcal{A}(\gamma_{i})

    to be nonnegative.

Recall that 𝒜(γ)\mathcal{A}(\gamma) denotes the action of γ\gamma, and for the Reeb orbit γ=𝔬ka\gamma=\mathfrak{o}^{\vec{a}}_{k} in E(a)\partial E({\vec{a}}) we have 𝒜(𝔬ka)=𝕄ka\mathcal{A}(\mathfrak{o}^{\vec{a}}_{k})={\mathbb{M}}^{\vec{a}}_{k}, where 𝕄ka{\mathbb{M}}^{\vec{a}}_{k} is the kkth smallest element of the multiset {iaj|i1,1jn}\{ia_{j}\;|\;i\in\mathbb{Z}_{\geq 1},1\leq j\leq n\}, or equivalently

𝕄ka=mini1,,in0,i1++in=kmax1snasis{\mathbb{M}}^{\vec{a}}_{k}=\min\limits_{\begin{subarray}{c}i_{1},\dots,i_{n}\in\mathbb{Z}_{\geq 0},\\ i_{1}+\cdots+i_{n}=k\end{subarray}}\;\max\limits_{1\leq s\leq n}a_{s}i_{s}

(see [GH, §1.2]). The index of CC is defined to be

ind(C)=(n3)(2k)+2c1(A)i=1kCZ(γi),\displaystyle{\operatorname{ind}}(C)=(n-3)(2-k)+2c_{1}(A)-\sum_{i=1}^{k}{\operatorname{CZ}}(\gamma_{i}), (2.2.1)

with CZ(γi){\operatorname{CZ}}(\gamma_{i}) calculated as in (2.3.1).

Definition 2.2.3.

Similarly, a genus zero formal curve component CC in E(a)\partial E({\vec{a}}) is a pair (Γ+,Γ)(\Gamma^{+},\Gamma^{-}), where

  • Γ+=(γ1+,,γk++)\Gamma^{+}=(\gamma_{1}^{+},\dots,\gamma^{+}_{k^{+}}) and Γ=(γ1,,γk)\Gamma^{-}=(\gamma^{-}_{1},\dots,\gamma^{-}_{k^{-}}) are tuples of Reeb orbits in E(a)\partial E({\vec{a}})

  • we require the energy

    (C):=i=1k+𝒜(γi+)j=1k𝒜(γj)\displaystyle\mathcal{E}(C):=\sum_{i=1}^{k^{+}}\mathcal{A}(\gamma_{i}^{+})-\sum_{j=1}^{k^{-}}\mathcal{A}(\gamma_{j}^{-}) (2.2.2)

    to be nonnegative.

The index of CC is given by

ind(C)=(n3)(2k+k)+i=1k+CZ(γi+)j=1kCZ(γj).\displaystyle{\operatorname{ind}}(C)=(n-3)(2-k^{+}-k^{-})+\sum_{i=1}^{k^{+}}{\operatorname{CZ}}(\gamma_{i}^{+})-\sum_{j=1}^{k^{-}}{\operatorname{CZ}}(\gamma_{j}^{-}). (2.2.3)

A formal cylinder in E(a)\partial E({\vec{a}}) is a formal curve with one positive end and one negative end (i.e. k+=k=1k^{+}=k^{-}=1), and it is trivial if Γ+=Γ\Gamma^{+}=\Gamma^{-}.

2.3. Main robustness result

We now discuss notions of robustness for moduli spaces. In the following, M2nM^{2n} is a closed symplectic manifold, AH2(M)A\in H_{2}(M) is a homology class, a=(a1,,an)>0n{\vec{a}}=(a_{1},\dots,a_{n})\in\mathbb{R}_{>0}^{n} is a rationally independent tuple, and we put Ma=Mι(E̊(εa))M_{\vec{a}}=M\setminus\iota({\mathring{E}}(\varepsilon{\vec{a}})), which implicitly depends on a choice of ε>0\varepsilon>0 and ι:E(εa)𝑠M\iota:E(\varepsilon{\vec{a}})\overset{s}{\hookrightarrow}M.

Definition 2.3.1.

We will say that Ma,A(γ)\mathcal{M}_{M_{\vec{a}},A}(\gamma) is robust if Ma,AJ(γ)\mathcal{M}_{M_{\vec{a}},A}^{J}(\gamma) is finite and regular for any choice of ε,ι\varepsilon,\iota, and generic J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}), and moreover the (signed) count #Ma,AJ(γ)\#\mathcal{M}_{M_{\vec{a}},A}^{J}(\gamma) is independent of these choices.

We will say that Ma,A(γ)\mathcal{M}_{M_{\vec{a}},A}(\gamma) is strongly robust if it is robust and moreover we have ¯Ma,AJ(γ)=Ma,AJ(γ)\overline{\mathcal{M}}_{M_{\vec{a}},A}^{J}(\gamma)=\mathcal{M}_{M_{\vec{a}},A}^{J}(\gamma) for any choice of ε,ι\varepsilon,\iota, and generic J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}).

Here ¯Ma,AJ(γ)\overline{\mathcal{M}}_{M_{\vec{a}},A}^{J}(\gamma) denotes the SFT compactification of Ma,AJ(γ)\mathcal{M}_{M_{\vec{a}},A}^{J}(\gamma) via pseudoholomorphic buildings à la [BEH+]. Note that ¯Ma,AJ(γ)=Ma,AJ(γ)\overline{\mathcal{M}}_{M_{\vec{a}},A}^{J}(\gamma)=\mathcal{M}_{M_{\vec{a}},A}^{J}(\gamma) implies that Ma,AJ(γ)\mathcal{M}_{M_{\vec{a}},A}^{J}(\gamma) is already compact, and hence finite if it is regular with index zero. If Ma,A(γ)\mathcal{M}_{M_{\vec{a}},A}(\gamma) is robust, we will write #Ma,A(γ)\#\mathcal{M}_{M_{\vec{a}},A}(\gamma) to refer to the (signed) count #Ma,AJ(γ)\#\mathcal{M}_{M_{\vec{a}},A}^{J}(\gamma) for any choice of generic J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}). We will say that a robust moduli space Ma,A(γ)\mathcal{M}_{M_{\vec{a}},A}(\gamma) is deformation invariant if it remains robust under deformations of the symplectic form on MM, and moreover the count #Ma,A(γ)\#\mathcal{M}_{M_{\vec{a}},A}(\gamma) is unchanged under such deformations.

Remark 2.3.2.

Although robust moduli spaces are already well-suited for enumerative purposes, the added benefit of strong robustness is that it guarantees that the counts agree with their SFT counterparts, which a priori depend on the full SFT compactification (c.f. [MS5, §3.4]). Strong robustness is also closely related to the notion of “formal perturbation invariance” utilized in [MS5], but due to some technical differences in the setup we use a different term here to avoid confusion. \Diamond

Recall that a closed symplectic manifold M2nM^{2n} is semipositive if any Aπ2(M)A\in\pi_{2}(M) with positive symplectic area and c1(A)3nc_{1}(A)\geq 3-n satisfies c1(A)0c_{1}(A)\geq 0 (this is automatic if dimM6\dim M\leq 6). The upshot is that, for generic JJ, all JJ-holomorphic spheres CC have ind(C)0{\operatorname{ind}}(C)\geq 0. Indeed, by standard transversality the underlying simple curve C¯\underline{C} must have ind(C¯)0{\operatorname{ind}}(\underline{C})\geq 0, and hence c1(C¯)0c_{1}(\underline{C})\geq 0 by semipositivity, whence ind(C)ind(C¯)0{\operatorname{ind}}(C)\geq{\operatorname{ind}}(\underline{C})\geq 0.

Before stating our main result on robust moduli spaces, we will need to make some numerical assumptions on the data a{\vec{a}} and c1(A)c_{1}(A).

Assumption A.

We have

𝕄i1a++𝕄ika𝕄i1++ik+k1a\displaystyle{\mathbb{M}}^{\vec{a}}_{i_{1}}+\cdots+{\mathbb{M}}^{\vec{a}}_{i_{k}}\leq{\mathbb{M}}^{\vec{a}}_{i_{1}+\cdots+i_{k}+k-1}

for any i1,,ik1i_{1},\dots,i_{k}\in\mathbb{Z}_{\geq 1} with k2k\geq 2 satisfying s=1kis+k1c1(A)1\sum\limits_{s=1}^{k}i_{s}+k-1\leq c_{1}(A)-1.

Assumption B.

We have

𝕄i1a++𝕄ika<𝕄c1(A)1a\displaystyle{\mathbb{M}}^{\vec{a}}_{i_{1}}+\cdots+{\mathbb{M}}^{\vec{a}}_{i_{k}}<{\mathbb{M}}^{\vec{a}}_{c_{1}(A)-1}

for any i1,,ik1i_{1},\dots,i_{k}\in\mathbb{Z}_{\geq 1} with k2k\geq 2 satisfying s=1kis+k1=c1(A)1\sum\limits_{s=1}^{k}i_{s}+k-1=c_{1}(A)-1.

Assumption C.

The homology class AA and Reeb orbit 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1} have no common divisibility.

Remark 2.3.3.

Assumption A equivalently states that any index zero (rational) formal curve in E(a)\partial E({\vec{a}}) with one negative end 𝔬ma\mathfrak{o}^{\vec{a}}_{m} must have nonpositive energy, provided that mc1(A)1m\leq c_{1}(A)-1. Incidentally, a formal curve with nonpositive energy necessarily has zero energy since by definition formal curves cannot have negative energy. Assumption B equivalently states that there are no nontrivial index zero (rational) formal curves in E(a)\partial E({\vec{a}}) with one negative end 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}. Note that for a{\vec{a}} rationally independent a formal curve with one negative end can only have zero energy if all Reeb orbits involved are covers of the same underlying simple orbit. \Diamond

Remark 2.3.4.

Note that for a{\vec{a}} rationally independent, When n=2n=2, Assumption B is equivalent to the statement that 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1} is maximal with respect to the Hutchings–Taubes partial order [HT, Def. 1.8], i.e. a negative end asymptotic to 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1} satisfies the negative ECH partition condition (c.f. [Hut2, Ex. 3.14]). \Diamond

We can now state our main robustness result:

Theorem 2.3.5.

Let M2nM^{2n} be a semipositive closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, and a=(a1,,an)>0n{\vec{a}}=(a_{1},\dots,a_{n})\in\mathbb{R}_{>0}^{n} a rationally independent tuple such that Assumptions A, B, and C hold. Then the moduli space Ma,A(𝔬c1(A)1a)\mathcal{M}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is strongly robust and deformation invariant.

Recall that we put 𝐓M,Aa:=#Ma,A(𝔬c1(A)1a)\mathbf{T}_{M,A}^{\vec{a}}:=\#\mathcal{M}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}).

Although we do not a priori have a uniform way to verify Assumptions A and B in general, the next two lemmas give useful sufficient conditions.

Lemma 2.3.6.

If a3,,an>𝕄c1(A)1(a1,a2)a_{3},\dots,a_{n}>{\mathbb{M}}^{(a_{1},a_{2})}_{c_{1}(A)-1}, then Assumption A holds.

Lemma 2.3.7.

Fix p,q1p,q\in\mathbb{Z}_{\geq 1} relatively prime such that p+q=c1(A)p+q=c_{1}(A), and suppose we have a=(q,p±δ,a3,,an){\vec{a}}=(q,p\pm\delta,a_{3},\dots,a_{n}) with a3,,an>pqa_{3},\dots,a_{n}>pq and δ>0\delta\in\mathbb{R}_{>0} sufficiently small. Then Assumptions A and B hold.

Note that Assumptions A,B,C depend only on a{\vec{a}} up to scalar multiplication, i.e. if they hold for a{\vec{a}} then they also hold for λa\lambda{\vec{a}} for any λ>0\lambda\in\mathbb{R}_{>0}.

Deferring the proofs of these lemmas for the moment, let us consider the four-dimensional case n=2n=2, or more generally a3,,ana1,a2a_{3},\dots,a_{n}\gg a_{1},a_{2}. In this situation it is customary to take without loss of generality (a1,a2)=(1,a)(a_{1},a_{2})=(1,a) with a>1a>1, and we refer to ν1\nu_{1} as the “short orbit” 𝔰\mathfrak{s} and ν2\nu_{2} as the “long orbit” 𝔩\mathfrak{l}. In the following, for p/qp/q\in\mathbb{Q} with p>qp>q and gcd(p,q)=1\gcd(p,q)=1 we put (p/q)+:=p/q+δ(p/q)^{+}:=p/q+\delta and (p/q):=p/qδ(p/q)^{-}:=p/q-\delta for δ>0\delta>0 arbitrarily small. We will sometimes write MaM_{a} in place of M(1,a)M_{(1,a)}, and 𝔬ka\mathfrak{o}^{a}_{k} in place of 𝔬k(1,a)\mathfrak{o}^{(1,a)}_{k}, and so on. In particular, assuming p+q=c1(A)p+q=c_{1}(A), note that if a=(p/q)+a=(p/q)^{+} then we have 𝔬c1(A)1a=𝔰p\mathfrak{o}^{a}_{c_{1}(A)-1}=\mathfrak{s}^{p}, while if a=(p/q)a=(p/q)^{-} then we have 𝔬c1(A)1a=𝔩q\mathfrak{o}^{a}_{c_{1}(A)-1}=\mathfrak{l}^{q}. We thus have:

Corollary 2.3.8.

Let M4M^{4} be a closed symplectic four-manifold, and let AH2(M)A\in H_{2}(M) be a homology class such that c1(A)=p+qc_{1}(A)=p+q for some p>q1p>q\in\mathbb{Z}_{\geq 1} relatively prime.

  1. (1)

    For a=(p/q)+a=(p/q)^{+}, the moduli space Ma,A(𝔰p)\mathcal{M}_{M_{a},A}(\mathfrak{s}^{p}) is strongly robust.

  2. (2)

    For a=(p/q)a=(p/q)^{-}, the moduli space Ma,A(𝔩q)\mathcal{M}_{M_{a},A}(\mathfrak{l}^{q}) is strongly robust.

In either case, the signed count is a nonnegative integer.

More generally, the same is true if M2nM^{2n} is any semipositive closed symplectic manifold and we replace MaM_{a} with MaM_{\vec{a}} for a=(1,a,a3,,an){\vec{a}}=(1,a,a_{3},\dots,a_{n}) with a3,,an>pa_{3},\dots,a_{n}>p.

The last sentence of the first paragraph is a standard consequence of automatic transversality as in [Wen1] – see [MS5, §5.2].

Remark 2.3.9.

If we put a=p/qa=p/q, E(1,a)\partial E(1,a) has a two-parameter family of Reeb orbits which becomes the two orbits 𝔰p,𝔩q\mathfrak{s}^{p},\mathfrak{l}^{q} after slightly perturbing aa, and the moduli spaces described in cases (1) and (2) of Corollary 2.3.8 can be viewed as two different ways of Morsifying this family. \Diamond

We note that the above discussion gives sufficient but not necessary conditions for Assumptions A and B to hold, as the following simple example illustrates.

Example 2.3.10.

Consider the case with a=(8,13,22){\vec{a}}=(8,13,22) (or a small perturbation thereof) and c1(A)=5c_{1}(A)=5. Then we have 𝔬c1(A)1a=ν3\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}=\nu_{3} and 𝒜(𝔬c1(A)1a)=22\mathcal{A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})=22, and one readily checks that Assumptions A and B hold, even though the hypotheses of Lemma 2.3.6 and Lemma 2.3.7 do not hold (even after permuting a{\vec{a}}). \Diamond

We end this subsection by proving Lemmas 2.3.6 and 2.3.7.

Proof of Lemma 2.3.6.

Fix i1,,ik1i_{1},\dots,i_{k}\in\mathbb{Z}_{\geq 1} satisfying m:=s=1kis+k1c1(A)1m:=\sum\limits_{s=1}^{k}i_{s}+k-1\leq c_{1}(A)-1. Considering the contrapositive of Assumption A, it suffices to show that any formal curve CC in E(a)\partial E({\vec{a}}) with strictly positive energy and with positive ends 𝔬i1a,,𝔬ika\mathfrak{o}_{i_{1}}^{\vec{a}},\dots,\mathfrak{o}^{\vec{a}}_{i_{k}} and negative end 𝔬ma\mathfrak{o}^{\vec{a}}_{m} must satisfy ind(C)>0{\operatorname{ind}}(C)>0.

Note that for any ic1(A)1i\leq c_{1}(A)-1, the Reeb orbit 𝔬ia\mathfrak{o}_{i}^{\vec{a}} is a multiple of either ν1\nu_{1} or ν2\nu_{2}. Suppose that we have 𝔬ia=ν1j\mathfrak{o}_{i}^{\vec{a}}=\nu_{1}^{j} for some j1j\in\mathbb{Z}_{\geq 1}. Then we have

ja1=𝒜(ν1j)=𝒜(𝔬ia)𝒜(𝔬c1(A)1a)<a3,,an\displaystyle ja_{1}=\mathcal{A}(\nu_{1}^{j})=\mathcal{A}(\mathfrak{o}^{\vec{a}}_{i})\leq\mathcal{A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})<a_{3},\dots,a_{n}

and hence

CZ(𝔬ia)=CZ(ν1j)\displaystyle{\operatorname{CZ}}(\mathfrak{o}_{i}^{\vec{a}})={\operatorname{CZ}}(\nu_{1}^{j}) =(n1)+2j+2a1j/a2++a1j/an\displaystyle=(n-1)+2j+2\lfloor a_{1}j/a_{2}\rfloor+\cdots+\lfloor a_{1}j/a_{n}\rfloor (2.3.1)
=(n1)+2j+2a1j/a2,\displaystyle=(n-1)+2j+2\lfloor a_{1}j/a_{2}\rfloor,

i.e. a3,,ana_{3},\dots,a_{n} are too large to contribute to the Conley–Zehnder index of 𝔬ia\mathfrak{o}_{i}^{\vec{a}}. Similarly, if 𝔬ia=ν2j\mathfrak{o}_{i}^{\vec{a}}=\nu_{2}^{j} for some j1j\in\mathbb{Z}_{\geq 1} then we have

CZ(𝔬ia)=CZ(ν2j)=(n1)+2j+2a2j/a1.\displaystyle{\operatorname{CZ}}(\mathfrak{o}_{i}^{\vec{a}})={\operatorname{CZ}}(\nu_{2}^{j})=(n-1)+2j+2\lfloor a_{2}j/a_{1}\rfloor.

It follows that for index purposes we can assume n=2n=2, and since we are assuming (C)>0\mathcal{E}(C)>0 we then have ind(C)2{\operatorname{ind}}(C)\geq 2 by Lemma 2.3.11 below. ∎

Lemma 2.3.11.

Fix a1,a2>0a_{1},a_{2}\in\mathbb{R}_{>0} rationally independent, and let CC be a (rational) formal curve in E(a1,a2)\partial E(a_{1},a_{2}). Then we have ind(C)0{\operatorname{ind}}(C)\geq 0, and in fact ind(C)2{\operatorname{ind}}(C)\geq 2 unless (C)=0\mathcal{E}(C)=0.

Proof.

Up to permutation we can write the positive asymptotics of CC as ν1i1+,,ν1ik++,ν2j1+,,ν2jl++\nu_{1}^{i_{1}^{+}},\dots,\nu_{1}^{i_{k^{+}}^{+}},\nu_{2}^{j_{1}^{+}},\dots,\nu_{2}^{j_{l^{+}}^{+}} and the negative asymptotics as ν1i1,,ν1ik,ν2j1,,ν2jl\nu_{1}^{i_{1}^{-}},\dots,\nu_{1}^{i_{k^{-}}^{-}},\nu_{2}^{j_{1}^{-}},\dots,\nu_{2}^{j_{l^{-}}^{-}}. We then have

(C)=s=1k+is+a1+s=1l+js+a2s=1kisa1s=1ljsa20.\displaystyle\mathcal{E}(C)=\sum_{s=1}^{k^{+}}i_{s}^{+}a_{1}+\sum_{s=1}^{l^{+}}j_{s}^{+}a_{2}-\sum_{s=1}^{k^{-}}i_{s}^{-}a_{1}-\sum_{s=1}^{l^{-}}j_{s}^{-}a_{2}\geq 0. (2.3.2)

Slightly rearranging the terms in the index formula, we have

ind(C)=2+2(s=1k+is++s=1l+js+a2/a1s=1kiss=1ljsa2/a1)\displaystyle{\operatorname{ind}}(C)=-2+2\left(\sum_{s=1}^{k^{+}}i_{s}^{+}+\sum_{s=1}^{l^{+}}\lceil j_{s}^{+}a_{2}/a_{1}\rceil-\sum_{s=1}^{k^{-}}i_{s}^{-}-\sum_{s=1}^{l^{-}}\lfloor j_{s}^{-}a_{2}/a_{1}\rfloor\right)
+2(s=1k+is+a1/a2+s=1l+js+s=1kisa1/a2s=1ljs).\displaystyle+2\left(\sum_{s=1}^{k^{+}}\lceil i_{s}^{+}a_{1}/a_{2}\rceil+\sum_{s=1}^{l^{+}}j_{s}^{+}-\sum_{s=1}^{k^{-}}\lfloor i_{s}^{-}a_{1}/a_{2}\rfloor-\sum_{s=1}^{l^{-}}j_{s}^{-}\right).

By  2.3.2, each of the terms in parentheses must be nonnegative, and at least one of them must be strictly positive (and hence at least 11 by integrality) since a2/a1a_{2}/a_{1} is irrational. If (C)\mathcal{E}(C) is strictly positive then both of the terms in parentheses must be strictly positive.

Proof of Lemma 2.3.7.

To verify Assumption A, by Lemma 2.3.6 it suffices to establish 𝕄p+q1(q,p±δ)<a3,,an{\mathbb{M}}_{p+q-1}^{(q,p\pm\delta)}<a_{3},\dots,a_{n}. Note that for δ>0\delta>0 sufficiently small we have 𝕄p+q1(q,p+δ)=pq{\mathbb{M}}_{p+q-1}^{(q,p+\delta)}=pq and 𝕄p+q1(q,pδ)=q(pδ){\mathbb{M}}_{p+q-1}^{(q,p-\delta)}=q(p-\delta).

We will verify Assumption B in the case a=(q,p+δ,a3,,an){\vec{a}}=(q,p+\delta,a_{3},\dots,a_{n}), the case a=(q,pδ,a3,,an){\vec{a}}=(q,p-\delta,a_{3},\dots,a_{n}) being nearly identical. It suffices to show that any nontrivial formal (rational) curve CC in E(a)\partial E({\vec{a}}) with negative end 𝔬p+q1a\mathfrak{o}^{\vec{a}}_{p+q-1} and (C)=0\mathcal{E}(C)=0 satisfies ind(C)2{\operatorname{ind}}(C)\geq 2.

As in the proof of Lemma 2.3.6, it suffices to consider the case n=2n=2. Since 𝔬p+q1a=ν1p\mathfrak{o}^{\vec{a}}_{p+q-1}=\nu_{1}^{p} and (C)=0\mathcal{E}(C)=0, we can assume that the positive ends of CC are ν1p1,,ν1pm\nu_{1}^{p_{1}},\dots,\nu_{1}^{p_{m}} for some p1,,pm1p_{1},\dots,p_{m}\in\mathbb{Z}_{\geq 1} with m2m\geq 2 and p1++pm=pp_{1}+\cdots+p_{m}=p. Then we have

ind(C)=2i=1mpia1a22pa1a2=2i=1mpiqp+δ2q.\displaystyle{\operatorname{ind}}(C)=2\sum_{i=1}^{m}\lceil\tfrac{p_{i}a_{1}}{a_{2}}\rceil-2\lceil\tfrac{pa_{1}}{a_{2}}\rceil=2\sum_{i=1}^{m}\lceil\tfrac{p_{i}q}{p+\delta}\rceil-2q.

Note that piq/pp_{i}q/p is not an integer for i=1,,mi=1,\dots,m, since we have gcd(p,q)=1\gcd(p,q)=1 and 1pi<p1\leq p_{i}<p. Therefore, for δ>0\delta>0 sufficiently small we have

ind(C)=2i=1mpiqp2q>2i=1mpiqp2q=0.\displaystyle{\operatorname{ind}}(C)=2\sum_{i=1}^{m}\lceil\tfrac{p_{i}q}{p}\rceil-2q>2\sum_{i=1}^{m}\tfrac{p_{i}q}{p}-2q=0.

This completes the proof. ∎

2.4. Proof modulo lemmas

In this subsection we give the proof of Theorem 2.3.5, modulo several lemmas whose proofs are deferred to §2.5. Our goal is to establish the following:

  1. (I)

    for any choice of ε>0\varepsilon>0, ι:E(εa)𝑠M\iota:E(\varepsilon{\vec{a}})\overset{s}{\hookrightarrow}M, and generic J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}), the moduli space Ma,AJ(𝔬c1(A)1a)\mathcal{M}^{J}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is regular and equal to its SFT compactification

  2. (II)

    the count #Ma,AJ(𝔬c1(A)1a)\#\mathcal{M}^{J}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is independent of ι,ε\iota,\varepsilon and generic JJ, and is invariant under deformations of the symplectic form of MM.

By the following claim, we can work with fixed ε,ι\varepsilon,\iota throughout the argument.

Claim 2.4.1.

Independence of #Ma,AJ(γ0)\#\mathcal{M}^{J}_{M_{\vec{a}},A}(\gamma_{0}) of the choice of ε\varepsilon and ι\iota follows from independence of the choice of generic J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}).

Proof.

For i=0,1i=0,1, consider symplectic embeddings ιi:E(εia)𝑠M\iota_{i}:E(\varepsilon_{i}{\vec{a}})\overset{s}{\hookrightarrow}M for some εi>0\varepsilon_{i}>0, along with generic Ji𝒥(Mιi(E̊(εia)))J_{i}\in\mathcal{J}(M\setminus\iota_{i}({\mathring{E}}(\varepsilon_{i}{\vec{a}}))). By Lemma 2.1.1, we can find ε>0\varepsilon^{\prime}>0 sufficiently small such that the restrictions ι0:=ι0|E(εa)\iota_{0}^{\prime}:=\iota_{0}|_{E(\varepsilon^{\prime}{\vec{a}})} and ι1:=ι1|E(εa)\iota_{1}^{\prime}:=\iota_{1}|_{E(\varepsilon^{\prime}{\vec{a}})} are isotopic through symplectic embeddings. Note that, for i=0,1i=0,1, the symplectic completions of Mιi(E̊(εia))M\setminus\iota_{i}({\mathring{E}}(\varepsilon_{i}{\vec{a}})) and Mιi(E̊(εa))M\setminus\iota_{i}({\mathring{E}}(\varepsilon^{\prime}{\vec{a}})) are naturally identified such that we have an inclusion 𝒥(Mιi(E̊(εia)))𝒥(Mιi(E̊(εa)))\mathcal{J}(M\setminus\iota_{i}({\mathring{E}}(\varepsilon_{i}{\vec{a}})))\subset\mathcal{J}(M\setminus\iota_{i}({\mathring{E}}(\varepsilon^{\prime}{\vec{a}}))). Moreover, as in Remark 2.1.3, there is a symplectomorphism Mι0(E̊(εa))Mι1(E̊(εa))M\setminus\iota_{0}({\mathring{E}}(\varepsilon^{\prime}{\vec{a}}))\cong M\setminus\iota_{1}({\mathring{E}}(\varepsilon^{\prime}{\vec{a}})), and this sets up a bijection between 𝒥(Mι0(E̊(εa)))\mathcal{J}(M\setminus\iota_{0}({\mathring{E}}(\varepsilon^{\prime}{\vec{a}}))) and 𝒥(Mι1(E̊(εa)))\mathcal{J}(M\setminus\iota_{1}({\mathring{E}}(\varepsilon^{\prime}{\vec{a}}))). Under these identifications, we can thus view J0,J1J_{0},J_{1} as two generic almost complex structures on the symplectic completion of a fixed symplectic manifold MaM_{\vec{a}}. ∎

Now fix a generic one-parameter family of almost complex structures

{Jt𝒥(Ma)|t[0,1]},\{J_{t}\in\mathcal{J}(M_{\vec{a}})\;|\;t\in[0,1]\},

and let

{Jt𝒥(E(a))|t[0,1]}\{J_{t}^{-}\in\mathcal{J}(E({\vec{a}}))\;|\;t\in[0,1]\}

be the induced family given by restricting each JtJ_{t} to the negative end. Here we could also allow a one-parameter family of symplectic forms on MM, but to keep the exposition simpler we will assume the symplectic form on MM is fixed.

Proof of Theorem 2.3.5.

By Assumption C, every curve in the parametrized moduli space Ma,A{Jt}(𝔬c1(A)1a)\mathcal{M}^{\{J_{t}\}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is necessarily simple. Therefore by genericity of the family {Jt}\{J_{t}\} we can assume that the moduli space Ma,A{Jt}(𝔬c1(A)1a)\mathcal{M}^{\{J_{t}\}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is regular and hence a smooth 11-dimensional manifold. Similarly, Ma,AJi(𝔬c1(A)1a)\mathcal{M}^{J_{i}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is regular for i=0,1i=0,1.

Now consider the SFT compactification

¯Ma,A{Jt}(𝔬c1(A)1a)={(u,t)|t[0,1],u¯Ma,AJt(𝔬c1(A)1a)}.\displaystyle\overline{\mathcal{M}}^{\{J_{t}\}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})=\{(u,t)\;|\;t\in[0,1],u\in\overline{\mathcal{M}}^{J_{t}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})\}.

To prove Theorem 2.3.5, it will suffice to show that ¯Ma,A{Jt}(𝔬c1(A)1a)=Ma,A{Jt}(𝔬c1(A)1a)\overline{\mathcal{M}}^{\{J_{t}\}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})=\mathcal{M}^{\{J_{t}\}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}), which implies that Ma,A{Jt}(𝔬c1(A)1a)\mathcal{M}^{\{J_{t}\}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is already compact. In particular, this implies that ¯Ma,AJi(𝔬c1(A)1a)\overline{\mathcal{M}}^{J_{i}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is equal to its SFT compactification for i=0,1i=0,1, which confirms (I). Then Ma,A{Jt}(𝔬c1(A)1a)\mathcal{M}^{\{J_{t}\}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is a smooth 11-dimensional compact cobordism between finite zero-dimensional manifolds Ma,AJ0(𝔬c1(A)1a)\mathcal{M}^{J_{0}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) and Ma,AJ1(𝔬c1(A)1a)\mathcal{M}^{J_{1}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}), which confirms (II).

A priori, an element of ¯Ma,A{Jt}(𝔬c1(A)1a)\overline{\mathcal{M}}^{\{J_{t}\}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is a stable pseudoholomorphic building \mathcal{B} consisting of, for some te[0,1]t_{e}\in[0,1],

  • a level in M^a\widehat{M}_{\vec{a}} consisting of one or more JteJ_{t_{e}}-holomorphic curve components whose total homology class is AA

  • some number (possibly zero) of symplectization levels ×E(a)\mathbb{R}\times\partial E({\vec{a}}), each consisting of one or more JteJ^{-}_{t_{e}}-holomorphic curve components, not all of which are trivial cylinders.

Our goal is to show that in fact there cannot be any symplectization levels, and that the level in M^a\widehat{M}_{{\vec{a}}} has a single component. See Figure 1 for a cartoon.

Refer to caption
Figure 1. A potential building \mathcal{B} in ¯Ma,A{Jt}(𝔬c1(A)1a)\overline{\mathcal{M}}^{\{J_{t}\}}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}). After formally gluing as described above, C0C_{0} has 44 positive ends, and there are 44 planes C1,C2,C3,C4C_{1},C_{2},C_{3},C_{4} and 22 spheres Q1,Q2Q_{1},Q_{2}. Here one plane and one sphere have components in all three levels, while the others are entirely contained in MaM_{{\vec{a}}}.

After formally gluing various pairs of curve components in adjacent levels along shared Reeb orbits, we arrive at the following picture:

  • a formal curve component C0C_{0} in E(a)\partial E({\vec{a}}) with kk positive ends and a single negative end 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}, for some k1k\in\mathbb{Z}_{\geq 1}

  • formal planes C1,,CkC_{1},\dots,C_{k} in MaM_{\vec{a}}

  • formal spheres Q1,,QmQ_{1},\dots,Q_{m} in MaM_{\vec{a}}, for some m0m\in\mathbb{Z}_{\geq 0}.

Here the positive ends of C0C_{0} match up with the respective negative ends of C1,,CkC_{1},\dots,C_{k}. Note that each of C0,,Ck,Q1,,QmC_{0},\dots,C_{k},Q_{1},\dots,Q_{m} is composed of some number of curve components of \mathcal{B}.

In the next subsection we will prove that Assumptions A and B imply the following conditions:

Condition (a).

Let J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}) be part of a generic one-parameter family. Let CC be a genus zero JJ-holomorphic curve component in M^a\widehat{M}_{\vec{a}} with mm negative ends, all but one of which bounds a formal plane in E(a)E({\vec{a}}). Assume also that we have c1([C])c1(A)c_{1}([C])\leq c_{1}(A). Then, denoting the formal planes by P1,,Pm1P_{1},\dots,P_{m-1}, we have

ind(C)+i=1m1ind(Pi)0,\displaystyle{\operatorname{ind}}(C)+\sum_{i=1}^{m-1}{\operatorname{ind}}(P_{i})\geq 0,

with equality only if m=1m=1.

Condition (b).

Let CC be a genus zero formal curve in E(a)\partial E({\vec{a}}) with one negative end 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}. Then we have ind(C)0{\operatorname{ind}}(C)\geq 0, with equality only if (C)=0\mathcal{E}(C)=0.

Condition (c).

Let CC be a genus zero formal curve in E(a)\partial E({\vec{a}}) with one negative end 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}, and suppose that ind(C)=0{\operatorname{ind}}(C)=0 and (C)=0\mathcal{E}(C)=0. Then CC is a trivial cylinder.

We also have the following elementary lemma, whose proof is deferred to the next subsection:

Lemma 2.4.2.

Let J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}) be part of a generic one-parameter family, and let CC be a genus zero JJ-holomorphic curve component in M^a\widehat{M}_{\vec{a}}. Then we have c1(C)0c_{1}(C)\geq 0, and in fact c1(C)2c_{1}(C)\geq 2 if CC has any negative ends.

Taking these for granted for the moment, we complete the proof of Theorem 2.3.5 as follows.

For i=1,,ki=1,\dots,k, we claim that ind(Ci)0{\operatorname{ind}}(C_{i})\geq 0, with ind(Ci)2{\operatorname{ind}}(C_{i})\geq 2 unless CiC_{i} is composed of a single JteJ_{t_{e}}-holomorphic plane in M^a\widehat{M}_{\vec{a}}. To see this, note that a priori CiC_{i} is composed of some number mim_{i} of JteJ_{t_{e}}-holomorphic curve components in M^a\widehat{M}_{\vec{a}} and some number of JteJ_{t_{e}}^{-}-holomorphic curve components in ×E(a)\mathbb{R}\times\partial E({\vec{a}}). If mi=1m_{i}=1, then we can view CiC_{i} as a JteJ_{t_{e}}-holomorphic curve component in M^a\widehat{M}_{\vec{a}} such that all but one of its ends bounds a formal plane in E(a)\partial E({\vec{a}}), whence the claim follows directly by Condition (a). A similar argument holds in the case mi2m_{i}\geq 2, noting that by Lemma 2.4.2 any extra components in M^a\widehat{M}_{\vec{a}} will only push up the total index of CiC_{i}.

Now let 0\mathcal{B}_{0} denote the sub-building of \mathcal{B} given by throwing away all components corresponding to Q1,,QmQ_{1},\dots,Q_{m}. Since c1(Q1),,c1(Qm)0c_{1}(Q_{1}),\dots,c_{1}(Q_{m})\geq 0 by Lemma 2.4.2, the total index of all curve components satisfies

ind(0)ind()=0.\displaystyle{\operatorname{ind}}(\mathcal{B}_{0})\leq{\operatorname{ind}}(\mathcal{B})=0.

Since ind(C0)0{\operatorname{ind}}(C_{0})\geq 0 by (b) and ind(C1),,ind(Ck)0{\operatorname{ind}}(C_{1}),\dots,{\operatorname{ind}}(C_{k})\geq 0 by the above discussion, we must have ind(C0)==ind(Ck){\operatorname{ind}}(C_{0})=\cdots={\operatorname{ind}}(C_{k}), and hence

  • (C0)=0\mathcal{E}(C_{0})=0

  • CiC_{i} is a JteJ_{t_{e}}-holomorphic plane in M^a\widehat{M}_{\vec{a}} for i=1,,ki=1,\dots,k.

Moreover, by Condition (c) C0C_{0} must be a trivial formal cylinder, and in particular k=1k=1. It also follows that the total index of 0\mathcal{B}_{0} is zero, and hence c1(Q1)==c1(Qm)=0c_{1}(Q_{1})=\cdots=c_{1}(Q_{m})=0, so by Lemma 2.4.2 each QiQ_{i} is actually a JteJ_{t_{e}}-holomorphic sphere in M^a\widehat{M}_{\vec{a}}.

In principle C0C_{0} could be composed of several JteJ_{t_{e}}^{-}-holomorphic curve components in ×E(a)\mathbb{R}\times\partial E({\vec{a}}) which formally glue together to give a cylinder. However, since the total energy must be zero, C0C_{0} can only consist of trivial cylinders over 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}. This contradicts stability of the pseudoholomorphic building \mathcal{B}.

It remains that \mathcal{B} does not have any symplectization levels, and consists of a JteJ_{t_{e}}-holomorphic plane C0C_{0} and JteJ_{t_{e}}-holomorphic spheres Q1,,QmQ_{1},\dots,Q_{m}. By a standard argument using semipositivity, we must have m=0m=0, which completes the proof. Indeed, letting Q¯i\underline{Q}_{i} denote the underlying simple curve of QiQ_{i} for i=1,,mi=1,\dots,m, we have

(ind(Q¯i)+2)+(ind(C0)+2)=2n2,\displaystyle({\operatorname{ind}}(\underline{Q}_{i})+2)+({\operatorname{ind}}(C_{0})+2)=2n-2,

so by genericity of {Jt}\{J_{t}\} and standard tranversality techniques Q¯i\underline{Q}_{i} and C0C_{0} cannot intersect for codimension reasons. ∎

Remark 2.4.3.

If it easy to find counterexamples to Condition (b) if we relax the Assumption A. For instance, put a=(1,1+δ1,1+δ2){\vec{a}}=(1,1+\delta_{1},1+\delta_{2}) with 0<δ1<δ210<\delta_{1}<\delta_{2}\ll 1, and let CC be the branched cover of the trivial cylinder over ν1\nu_{1} in ×E(a)\mathbb{R}\times\partial E({\vec{a}}) having negative end ν1k\nu_{1}^{k} and kk positive ends on ν1\nu_{1}, for some k2k\in\mathbb{Z}_{\geq 2}. We have CZ(ν1k)=6k2{\operatorname{CZ}}(\nu_{1}^{k})=6k-2 and hence ind(C)=22k<0{\operatorname{ind}}(C)=2-2k<0. \Diamond

2.5. Proof of lemmas

In this subsection we now prove that Assumptions  A and B imply Conditions (a),(b),(c), and we also give the proof of Lemma 2.4.2.

Lemma 2.5.1.

Condition (a) holds under Assumption A.

Proof.

Let γ\gamma be the negative end of CC which does not bound a formal plane, and put B:=[C]H2(M)B:=[C]\in H_{2}(M). Our goal is to establish

n3+2c1(B)CZ(γ)0,\displaystyle n-3+2c_{1}(B)-{\operatorname{CZ}}(\gamma)\geq 0, (2.5.1)

with equality only if CC has precisely one negative end. Note that the left hand side of (2.5.1) can be viewed as the index of CC after capping off all of its other negative ends by formal planes.

Let κ\kappa be the covering index of CC over its underlying simple curve C¯\underline{C}, with B¯:=[C¯]H2(M)\underline{B}:=[\underline{C}]\in H_{2}(M), and let γ¯\underline{\gamma} denote the negative end of C¯\underline{C} which is covered by γ\gamma. Since C¯\underline{C} is simple, by the genericity assumption on JJ we have ind(C¯)1{\operatorname{ind}}(\underline{C})\geq-1 and hence ind(C¯)0{\operatorname{ind}}(\underline{C})\geq 0 by parity considerations. Since all formal planes in E(a)E({\vec{a}}) have positive index, we have

n3+2c1(B¯)CZ(γ¯)ind(C¯)0,\displaystyle n-3+2c_{1}(\underline{B})-{\operatorname{CZ}}(\underline{\gamma})\geq{\operatorname{ind}}(\underline{C})\geq 0, (2.5.2)

where the first inequality is strict unless C¯\underline{C} has precisely one negative end (i.e. m=1m=1). We will assume κ2\kappa\geq 2, since otherwise C=C¯C=\underline{C} and we are done.

Put γ=𝔬ia\gamma=\mathfrak{o}^{\vec{a}}_{i} and γ¯=𝔬i¯a\underline{\gamma}=\mathfrak{o}^{\vec{a}}_{\underline{i}} for some i,i¯1i,\underline{i}\in\mathbb{Z}_{\geq 1}. Using (2.5.2) we have

n1+2i¯=CZ(γ¯)n3+2c1(B¯),\displaystyle n-1+2\underline{i}={\operatorname{CZ}}(\underline{\gamma})\leq n-3+2c_{1}(\underline{B}),

i.e. i¯c1(B¯)1\underline{i}\leq c_{1}(\underline{B})-1, and hence κi¯+κ1κc1(B¯)1\kappa\underline{i}+\kappa-1\leq\kappa c_{1}(\underline{B})-1, with strict inequality unless m=1m=1.

We can therefore apply Assumption A in the case k=κk=\kappa and i1==ik=i¯i_{1}=\cdots=i_{k}=\underline{i} to get

κ𝕄i¯a𝕄κi¯+κ1a𝕄κc1(B¯)1a.\displaystyle\kappa{\mathbb{M}}_{\underline{i}}^{\vec{a}}\leq{\mathbb{M}}^{\vec{a}}_{\kappa\underline{i}+\kappa-1}\leq{\mathbb{M}}^{\vec{a}}_{\kappa c_{1}(\underline{B})-1}.

Since γ\gamma is at most a κ\kappa-fold cover of γ¯\underline{\gamma}, we also have

𝒜(γ)=𝒜(𝔬ia)κ𝒜(𝔬i¯a)𝒜(𝔬κc1(B¯)1a),\displaystyle\mathcal{A}(\gamma)=\mathcal{A}(\mathfrak{o}^{\vec{a}}_{i})\leq\kappa\mathcal{A}(\mathfrak{o}^{\vec{a}}_{\underline{i}})\leq\mathcal{A}(\mathfrak{o}^{\vec{a}}_{\kappa c_{1}(\underline{B})-1}), (2.5.3)

with the first inequality strict unless γ\gamma is precisely a κ\kappa-fold cover of γ¯\underline{\gamma} and the second inequality strict unless m=1m=1. Note that 𝒜(γ)𝒜(𝔬κc1(B¯)1a)\mathcal{A}(\gamma)\leq\mathcal{A}(\mathfrak{o}^{\vec{a}}_{\kappa c_{1}(\underline{B})-1}) is equivalent to iκc1(B¯)1i\leq\kappa c_{1}(\underline{B})-1.

Finally, note that (2.5.1) is equivalent to ic1(B)1i\leq c_{1}(B)-1, i.e. iκc1(B¯)1i\leq\kappa c_{1}(\underline{B})-1. This holds by the above, with strictly inequality only if m=1m=1 and γ\gamma is a κ\kappa-fold cover of γ¯\underline{\gamma}, in which case CC also has precisely one negative end. ∎

Lemma 2.5.2.

Condition (b) holds under Assumption A.

Proof.

Since CC has positive ends 𝔬i1a,,𝔬ika\mathfrak{o}^{\vec{a}}_{i_{1}},\dots,\mathfrak{o}^{\vec{a}}_{i_{k}} and negative end 𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}, we have

12ind(C)=s=1kis+k1(c1(A)1).\displaystyle\tfrac{1}{2}{\operatorname{ind}}(C)=\sum_{s=1}^{k}i_{s}+k-1-(c_{1}(A)-1). (2.5.4)

If ind(C)=0{\operatorname{ind}}(C)=0, then by Assumption A we have (C)0\mathcal{E}(C)\leq 0, and hence (C)=0\mathcal{E}(C)=0 since the energy must be nonnegative.

Otherwise, suppose by contradiction that we have ind(C)<0{\operatorname{ind}}(C)<0. Then s=1kis+k1<c1(A)1\sum\limits_{s=1}^{k}i_{s}+k-1<c_{1}(A)-1, so by Assumption A we have

𝕄i1a++𝕄ika𝕄i1++ik+k1a<𝕄c1(A)1a,\displaystyle{\mathbb{M}}_{i_{1}}^{\vec{a}}+\cdots+{\mathbb{M}}_{i_{k}}^{\vec{a}}\leq{\mathbb{M}}^{\vec{a}}_{i_{1}+\cdots+i_{k}+k-1}<{\mathbb{M}}^{\vec{a}}_{c_{1}(A)-1},

and hence (C)<0\mathcal{E}(C)<0, which is impossible. ∎

Lemma 2.5.3.

Condition (c) holds under Assumption B.

Proof.

If CC has index zero is not a trivial cylinder, then (2.5.4) together with Assumption B implies that (C)<0\mathcal{E}(C)<0, which is a contradiction since formal curves by definition have nonnegative energy. ∎

Proof of Lemma 2.4.2.

Since the first Chern number is multiplicative under taking covers, we can assume that CC is simple. Let k0k\in\mathbb{Z}_{\geq 0} denote the number of negative ends of CC, and let γ1,,γk\gamma_{1},\dots,\gamma_{k} denote its asymptotic Reeb orbits. By genericity we have ind(C)1{\operatorname{ind}}(C)\geq-1, and hence ind(C)0{\operatorname{ind}}(C)\geq 0 since the index

ind(C)=(n3)(2k)+2c1(C)i=1kCZ(γi)\displaystyle{\operatorname{ind}}(C)=(n-3)(2-k)+2c_{1}(C)-\sum_{i=1}^{k}{\operatorname{CZ}}(\gamma_{i})

is necessarily even. If CC is a closed curve, i.e. k=0k=0, then we have c1(A)3nc_{1}(A)\geq 3-n, and hence c1(A)0c_{1}(A)\geq 0 by semipositivity. Otherwise, we have k1k\geq 1 and

2c1(A)\displaystyle 2c_{1}(A) i=1kCZ(γi)+(k2)(n3)\displaystyle\geq\sum_{i=1}^{k}{\operatorname{CZ}}(\gamma_{i})+(k-2)(n-3)
k(n+1)+(k2)(n3)\displaystyle\geq k(n+1)+(k-2)(n-3)
=2n(k1)2k+6\displaystyle=2n(k-1)-2k+6
2(k1)2k+6=4.\displaystyle\geq 2(k-1)-2k+6=4.

2.6. Stabilization invariance I

We now discuss the effect on the ellipsoidal superpotential of stabilizing the target space, i.e. multiplying it with a given closed symplectic manifold Q2NQ^{2N}. This will be used in the next subsection to obstruct stabilized symplectic embeddings as in Theorem A(b).

Theorem 2.6.1.

Let M2nM^{2n} be a closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, and a=(a1,,an){\vec{a}}=(a_{1},\dots,a_{n}) a rationally independent tuple. Let Q2NQ^{2N} be another closed symplectic manifold and b=(b1,,bN){\vec{b}}=(b_{1},\dots,b_{N}) another tuple satisfying b1,,bN>𝕄c1(A)1ab_{1},\dots,b_{N}>{\mathbb{M}}^{\vec{a}}_{c_{1}(A)-1}. Then we have

𝐓M,Aa=𝐓M×Q,A×[pt](a,b),\displaystyle\mathbf{T}_{M,A}^{\vec{a}}=\mathbf{T}_{M\times Q,A\times[{\operatorname{pt}}]}^{({\vec{a}},{\vec{b}})}, (2.6.1)

provided that the moduli spaces underlying both sides of (2.6.1) are robust and deformation invariant.

Here (a,b)({\vec{a}},{\vec{b}}) denotes the concatenated tuple (a1,,an,b1,,bN)(a_{1},\dots,a_{n},b_{1},\dots,b_{N}).

In the special case of Corollary 2.3.8 we have:

Corollary 2.6.2.

Let M4M^{4} be a closed symplectic four-manifold, and let AH2(M)A\in H_{2}(M) be a homology class such that c1(A)=p+qc_{1}(A)=p+q for some p,q1p,q\in\mathbb{Z}_{\geq 1} relatively prime. Let Q2NQ^{2N} be another closed symplectic manifold such that M×QM\times Q is semipositive. Then for any b1,,bN>pqb_{1},\dots,b_{N}>pq and δ>0\delta>0 sufficiently small we have

𝐓M×Q,A×[pt](q,p±δ,b1,,bN)=𝐓M,A(q,p±δ).\displaystyle\mathbf{T}_{M\times Q,A\times[{\operatorname{pt}}]}^{(q,p\pm\delta,b_{1},\dots,b_{N})}=\mathbf{T}_{M,A}^{(q,p\pm\delta)}.
Remark 2.6.3.

Note that M×QM\times Q is semipositive if e.g. dim(M)=4\dim(M)=4 and dim(Q)=2\dim(Q)=2, or if MM is monotone and π2(Q)=0\pi_{2}(Q)=0. \Diamond

We naturally view any Reeb orbit γ\gamma in E(a)\partial E({\vec{a}}) as lying also in E(a,b)\partial E({\vec{a}},{\vec{b}}) via the inclusion E(a)E(a,b)\partial E({\vec{a}})\subset\partial E({\vec{a}},{\vec{b}}). For b1,,bNa1,,anb_{1},\dots,b_{N}\gg a_{1},\dots,a_{n}, the Conley–Zehnder index of γ\gamma as an orbit of E(a,b)\partial E({\vec{a}},{\vec{b}}) is NN greater than its Conley–Zehnder index as an orbit of E(a)\partial E({\vec{a}}). Using the shorthand Mastab:=(M×Q)(a,b)M^{\operatorname{stab}}_{\vec{a}}:=(M\times Q)_{({\vec{a}},{\vec{b}})}, we have by (2.2.1) and (2.3.1) that

indMastab,A×[pt](γ)=indMa,A(γ).\displaystyle{\operatorname{ind}}\,\mathcal{M}_{M_{\vec{a}}^{\operatorname{stab}},A\times[{\operatorname{pt}}]}(\gamma)={\operatorname{ind}}\,\mathcal{M}_{M_{\vec{a}},A}(\gamma). (2.6.2)

This simple observation goes back to [HK], and is the key starting point for obstructing stabilized symplectic embeddings. Here is it crucial that we count planes. Indeed, by constrast note that for Γ=(γ1,,γk)\Gamma=(\gamma_{1},\dots,\gamma_{k}) with k2k\geq 2 we have

indMastab,A×[pt](Γ)=indMa,A(Γ)+2N(1k)<indMa,A(Γ).\displaystyle{\operatorname{ind}}\,\mathcal{M}_{M_{\vec{a}}^{\operatorname{stab}},A\times[{\operatorname{pt}}]}(\Gamma)={\operatorname{ind}}\,\mathcal{M}_{M_{\vec{a}},A}(\Gamma)+2N(1-k)<{\operatorname{ind}}\,\mathcal{M}_{M_{\vec{a}},A}(\Gamma).

If QQ is a symplectic surface, Theorem 2.6.1 can be proved by a straightforward adaptation of the technique in [MS5, §3.6]. In brief, we can arrange that M^a\widehat{M}_{\vec{a}} sits inside M^astab\widehat{M}^{\operatorname{stab}}_{\vec{a}} as a symplectic divisor, and we work with Jstab𝒥(Mastab)J^{\operatorname{stab}}\in\mathcal{J}(M_{\vec{a}}^{\operatorname{stab}}) which preserves this divisor and restricts to some J𝒥(Ma)J\in\mathcal{J}(M_{\vec{a}}). Then we have a natural inclusion

Ma,AJ(𝔬c1(A)1a)Mastab,A×[pt]Jstab.\displaystyle\mathcal{M}^{J}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})\subset\mathcal{M}^{J^{\operatorname{stab}}}_{M^{\operatorname{stab}}_{\vec{a}},A\times[{\operatorname{pt}}]}. (2.6.3)

In fact, any curve CMastab,A×[pt]JstabC\in\mathcal{M}^{J^{\operatorname{stab}}}_{M^{\operatorname{stab}}_{\vec{a}},A\times[{\operatorname{pt}}]} must be entirely contained in M^a\widehat{M}_{\vec{a}}, since otherwise its intersection number with the divisor M^a\widehat{M}_{\vec{a}} must be zero for homological reasons, but also positive due to positivity of intersections and winding number estimates. Furthermore, by [MS5, §A] (see also [Per]), the inclusion (2.6.3) preserves regularity, i.e. any regular curve CMa,AJ(𝔬c1(A)1a)C\in\mathcal{M}^{J}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is also regular in Mastab,A×[pt]Jstab\mathcal{M}^{J^{\operatorname{stab}}}_{M^{\operatorname{stab}}_{\vec{a}},A\times[{\operatorname{pt}}]}. Therefore we have

𝐓M×Q,A×[pt](a,b)=#Mastab,A×[pt]Jstab(𝔬c1(A)1a)=#Ma,AJ(𝔬c1(A)1a)=𝐓M,Aa,\displaystyle\mathbf{T}_{M\times Q,A\times[{\operatorname{pt}}]}^{({\vec{a}},{\vec{b}})}=\#\mathcal{M}_{M_{\vec{a}}^{\operatorname{stab}},A\times[{\operatorname{pt}}]}^{J^{\operatorname{stab}}}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})=\#\mathcal{M}_{M_{\vec{a}},A}^{J}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})=\mathbf{T}_{M,A}^{\vec{a}},

as desired.

This argument extends to the case of general Q2NQ^{2N} by first deforming the symplectic form to make it integral, and then applying Donaldson’s theorem to find a full flag of smooth symplectic divisors

{y0}=Q0Q1QN=Q,\displaystyle\{y_{0}\}=Q_{0}\subset Q_{1}\subset\cdots\subset Q_{N}=Q,

with dimQi=2i\dim Q_{i}=2i for i=0,,Ni=0,\dots,N. For suitable Jstab𝒥(Mastab)J^{\operatorname{stab}}\in\mathcal{J}(M_{\vec{a}}^{\operatorname{stab}}), the above argument can be applied iteratively to show that that inclusion map (2.6.3) is again a regularity-preserving bijection, whence 𝐓M×Q,A×[pt](a,b)=𝐓M,Aa\mathbf{T}_{M\times Q,A\times[{\operatorname{pt}}]}^{({\vec{a}},{\vec{b}})}=\mathbf{T}_{M,A}^{\vec{a}}.

We will explain the closely analogous argument for closed curves in detail in §3.6. The case of closed curves is slightly cleaner since it does not require intersection theory for punctured curves, and it also implies Theorem 2.6.1 via the equivalence discussed in §3.4.

2.7. Symplectic embedding obstructions

We end this section by discussing the relationship between robust moduli spaces and (stable) symplectic embedding obstructions.

Proposition 2.7.1.

Let M2nM^{2n} be a closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, and a=(a1,,an)>0n{\vec{a}}=(a_{1},\dots,a_{n})\in\mathbb{R}_{>0}^{n} a rationally independent tuple. Assume that the moduli space Ma,A(𝔬c1(A)1a)\mathcal{M}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) is robust and we have 𝐓M,Aa0\mathbf{T}_{M,A}^{\vec{a}}\neq 0. Then given any symplectic embedding E(ca)𝑠ME(c{\vec{a}})\overset{s}{\hookrightarrow}M we must have c[ωM]A𝒜(𝔬c1(A)1a)c\leq\frac{[\omega_{M}]\cdot A}{\mathcal{A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})}.

Proof.

Let ι:E(ca)𝑠M\iota:E(c{\vec{a}})\overset{s}{\hookrightarrow}M be a symplectic embedding. By robustness of Ma,A(𝔬c1(A)1a)\mathcal{M}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}), for any generic J𝒥(Mι(E̊(ca))J\in\mathcal{J}(M\setminus\iota({\mathring{E}}(c{\vec{a}})) we have

#Mι(E̊(ca)),AJ(𝔬c1(A)1a)=#Ma,A(𝔬c1(A)1a)0.\displaystyle\#\mathcal{M}^{J}_{M\,\setminus\,\iota({\mathring{E}}(c{\vec{a}})),A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})=\#\mathcal{M}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})\neq 0.

Any curve CMι(E̊(ca)),AJ(Γ)C\in\mathcal{M}^{J}_{M\,\setminus\,\iota({\mathring{E}}(c{\vec{a}})),A}(\Gamma) must have nonnegative energy, i.e. we have

0(C)=[ωM]Ac𝒜(𝔬c1(A)1a).\displaystyle 0\leq\mathcal{E}(C)=[\omega_{M}]\cdot A-c\mathcal{A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}).

As for stable obstructions, we have:

Corollary 2.7.2.

Let M2nM^{2n} be a closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, and a=(a1,,an)>0n{\vec{a}}=(a_{1},\dots,a_{n})\in\mathbb{R}_{>0}^{n} a rationally independent tuple. Let Q2NQ^{2N} be another closed symplectic manifold and b=(b1,,bN){\vec{b}}=(b_{1},\dots,b_{N}) another tuple satisfying b1,,bN>𝒜(𝔬c1(A)1a)b_{1},\dots,b_{N}>\mathcal{A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}). Assume that the moduli spaces underlying 𝐓M,Aa\mathbf{T}_{M,A}^{\vec{a}} and 𝐓M×Q,A×[pt](a,b)\mathbf{T}_{M\times Q,A\times[{\operatorname{pt}}]}^{({\vec{a}},{\vec{b}})} are both robustly defined and deformation invariant, and we have 𝐓M,Aa0\mathbf{T}_{M,A}^{\vec{a}}\neq 0. Then any symplectic embedding E(ca,cb)𝑠M×QE(c{\vec{a}},c{\vec{b}})\overset{s}{\hookrightarrow}M\times Q must satisfy c[ωM]A𝒜(𝔬c1(A)1a)c\leq\frac{[\omega_{M}]\cdot A}{\mathcal{A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})}.

Proof.

By Theorem 2.6.1 we have 𝐓M×Q,A×[pt](a,b)=𝐓M,Aa0\mathbf{T}_{M\times Q,A\times[{\operatorname{pt}}]}^{({\vec{a}},{\vec{b}})}=\mathbf{T}_{M,A}^{\vec{a}}\neq 0. The result then follows by Proposition 2.7.1. ∎

Note that under the hypotheses of Corollary 2.7.2, any symplectic embedding of the form ι:E(ca)×N𝑠M×N\iota:E(c{\vec{a}})\times\mathbb{C}^{N}\overset{s}{\hookrightarrow}M\times\mathbb{C}^{N} must also satisfy c[ωM]A𝒜(𝔬c1(A)1a)c\leq\frac{[\omega_{M}]\cdot A}{\mathcal{A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})}. Indeed, we can restrict ι\iota to an embedding of the form E(ca,cb)𝑠M×NE(c{\vec{a}},c{\vec{b}})\overset{s}{\hookrightarrow}M\times\mathbb{C}^{N}, and by compactifying the target space we get a symplectic embedding E(ca,cb)𝑠M×QE(c{\vec{a}},c{\vec{b}})\overset{s}{\hookrightarrow}M\times Q, after suitably scaling up the symplectic form on QQ. The claim then follows by applying Corollary 2.7.2 to this embedding.

3. Multidirectional tangency constraints

In this section we construct Gromov–Witten type invariants, denoted by NM,A<𝒞mpt>N_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle, which count closed curves with local multidirectional tangency constraints. Here M2nM^{2n} is a closed symplectic manifold, AH2(M)A\in H_{2}(M) is a homology class, and m=(m1,,mn)1{\vec{m}}=(m_{1},\dots,m_{n})\in\mathbb{Z}_{\geq 1} is a tuple satisfying

s=1nms=c1(A)+n2\displaystyle\sum_{s=1}^{n}m_{s}=c_{1}(A)+n-2 (3.0.1)

(so that we expect a finite count). Theorem 3.3.2 establishes robustness of these counts under suitable assumptions, and Theorem 3.4.1 proves equivalence with the counts of the previous section. In §3.1 we import a technical tool from [MS6, §4] which will be used to relate multidirectional tangencies and ellipsoidal ends. In §3.2 we discuss “hidden constraints”, essentially showing that multidirectional tangency constraints degenerate in the same way as ellipsoidal negative ends. Using this, the proof of Theorem 3.3.2 is nearly identical mutatis mutandis to the proof of Theorem 2.3.5, after which Theorem 3.4.1 follows directly by making a special choice of almost complex structure. Finally, in §3.5 we discuss relations between multidirectional tangencies and singular curves, and in §3.6 we establish invariance under stabilization.

3.1. Multidirectional tangencies and ellipsoidal ends: local equivalence

We first discuss the local relationship between multidirectional tangency constraints and ellipsoidal negative ends. This will be used in the next subsection to compare degenerations of multidirectional tangency constraints with degenerations of ellipsoidal ends. In the following, we consider n\mathbb{C}^{n} with its standard integrable almost complex structure JstdJ_{\operatorname{std}}. We take x0=0nx_{0}=\vec{0}\in\mathbb{C}^{n}, and the standard set of spanning local divisors at 0\vec{0} given by 𝐃std=(𝐃1std,,𝐃nstd){\vec{\mathbf{D}}}^{\operatorname{std}}=(\mathbf{D}_{1}^{\operatorname{std}},\dots,\mathbf{D}_{n}^{\operatorname{std}}), with 𝐃istd={zi=0}n\mathbf{D}_{i}^{\operatorname{std}}=\{z_{i}=0\}\subset\mathbb{C}^{n} for i=1,,ni=1,\dots,n.

Notation 3.1.1 ([MS6]).

For v=(v1,,vn)1n{\vec{v}}=(v_{1},\dots,v_{n})\in\mathbb{Z}_{\geq 1}^{n}, let 𝔬va\mathfrak{o}^{\vec{a}}_{-{\vec{v}}} denote the Reeb orbit in E(a)\partial E({\vec{a}}) given by νimvim\nu_{i_{m}}^{v_{i_{m}}}, where imi_{m} is the index 1in1\leq i\leq n for which aivia_{i}v_{i} is minimal.

The following proposition shows that JstdJ_{\operatorname{std}} becomes SFT admissible under a suitable diffeomorphism, and for a JstdJ_{\operatorname{std}}-holomorphic curve we can explicitly describe the resulting Reeb orbit asymptotics. Let 𝔻:={|z|<1}\mathbb{D}:=\{|z|<1\} denote the open unit disk in \mathbb{C}, and let 𝔻:=𝔻{0}\mathbb{D}_{\circ}:=\mathbb{D}\setminus\{0\} denote the result after puncturing the origin.

Proposition 3.1.2 ([MS6, §4]).

For each a>0n{\vec{a}}\in\mathbb{R}_{>0}^{n} rationally independent, there is an explicit diffeomorphism

Φa:×E(a)n{0}\displaystyle\Phi_{\vec{a}}:\mathbb{R}\times\partial E({\vec{a}})\rightarrow\mathbb{C}^{n}\setminus\{\vec{0}\}

such that JE(a):=(Φa)JstdJ_{\partial E({\vec{a}})}:=(\Phi_{\vec{a}})^{*}J_{\operatorname{std}} lies in 𝒥(E(a))\mathcal{J}(\partial E({\vec{a}})).

Moreover, suppose that u:𝔻nu:\mathbb{D}\rightarrow\mathbb{C}^{n} is a JstdJ_{\operatorname{std}}-holomorphic map which strictly satisfies <𝒞𝐃stdm0>\Langle\mathcal{C}_{{\vec{\mathbf{D}}}^{\operatorname{std}}}^{\vec{m}}\vec{0}\Rangle at 0𝔻0\in\mathbb{D} for some m1n{\vec{m}}\in\mathbb{Z}_{\geq 1}^{n}, and is otherwise disjoint from 0\vec{0}. Then the JE(a)J_{\partial E({\vec{a}})}-holomorphic map

Φa1u|𝔻:𝔻×E(a)\displaystyle\Phi_{\vec{a}}^{-1}\circ u|_{\mathbb{D}_{\circ}}:\mathbb{D}_{\circ}\rightarrow\mathbb{R}\times\partial E({\vec{a}})

is negatively asymptotic to the Reeb orbit 𝔬ma\mathfrak{o}^{\vec{a}}_{-{\vec{m}}} in E(a)\partial E({\vec{a}}) at the puncture 0 .

Explicitly, we have

Φa(r,z1,,zn)=(e2πr/a1a1z1,,e2πr/ananzn)πi=1n|zi|2,\displaystyle\Phi_{\vec{a}}(r,z_{1},\dots,z_{n})=\frac{(e^{2\pi r/a_{1}}\sqrt{a_{1}}z_{1},\dots,e^{2\pi r/a_{n}}\sqrt{a_{n}}z_{n})}{\sqrt{\pi\sum_{i=1}^{n}|z_{i}|^{2}}}, (3.1.1)

where rr is the coordinate on the first factor of ×E(a)\mathbb{R}\times\partial E({\vec{a}}) and (z1,,zn)(z_{1},\dots,z_{n}) are coordinates on n\mathbb{C}^{n}. The second statement in Proposition 3.1.2 can be seen by analyzing the asymptotics as |z|0|z|\rightarrow 0 of (Φa1u)(z)(\Phi_{\vec{a}}^{-1}\circ u)(z), with uu of the form u(z)=(C1zm1,,Cnzmn)u(z)=(C_{1}z^{m_{1}},\dots,C_{n}z^{m_{n}}).

For completeness let us elaborate on how the Reeb orbit 𝔬ma\mathfrak{o}^{\vec{a}}_{-{\vec{m}}} arises in Proposition 3.1.2, refering the reader to [MS6, §4] for full details.555Here the seemingly superfluous minus sign in the notation 𝔬ma\mathfrak{o}^{\vec{a}}_{-{\vec{m}}} is inherited from [MS6], where the discussion applies uniformly to both positive and negative ellipsoidal ends, whereas in this paper we only consider negative ellipsoidal ends. If π:×E(a)E(a)\pi:\mathbb{R}\times\partial E({\vec{a}})\to\partial E({\vec{a}}) is the projection, one can show that the map

𝔻E(a),z(πΦa1u)(z)\mathbb{D}_{\circ}\rightarrow\partial E({\vec{a}}),\quad z\mapsto(\pi\circ\Phi_{\vec{a}}^{-1}\circ u)(z)

is a composite zGflρa(u(z))(u(z))z\mapsto G\circ{\rm fl}^{\rho_{{\vec{a}}}(u(z))}(u(z)), where G:E(a)E(a)G:\partial E({\vec{a}})\to\partial E({\vec{a}}) is a diffeomorphism, flt:nn,t,{\rm fl}^{t}:\mathbb{C}^{n}\to\mathbb{C}^{n},t\in\mathbb{R}, is the flow z(e2πt/a1z1,,e2πt/anzn)\vec{z}\mapsto(e^{2\pi t/a_{1}}z_{1},\dots,e^{2\pi t/a_{n}}z_{n}), and ρa(w)\rho_{{\vec{a}}}(w) is the unique real number such that flρa(w)(w)E(a){\rm fl}^{\rho_{{\vec{a}}}(w)}(w)\in\partial E({\vec{a}}). It turns out that if we are only interested in the asymptotics of (Φa1u)(z)(\Phi_{\vec{a}}^{-1}\circ u)(z) as |z|0|z|\rightarrow 0 then we can ignore GG and assume that u(z)=(zm1,,zmn)u(z)=(z^{m_{1}},\dots,z^{m_{n}}), so that ρa(u(seiθ))\rho_{{\vec{a}}}(u(se^{i\theta})) depends only on the absolute values sm1,,smns^{m_{1}},\dots,s^{m_{n}} as s0s\to 0. When n=1n=1 we have ρa(u(seiθ))=r1\rho_{{\vec{a}}}(u(se^{i\theta}))=r_{1} where e2πr1/a1sm1=a1/πe^{2\pi r_{1}/a_{1}}s^{m_{1}}=\sqrt{a_{1}/\pi}; that is, 2πr1/a1=m1|log(s)|2\pi r_{1}/a_{1}=m_{1}|\log(s)| as s0s\rightarrow 0, i.e. r1=a1m12π|log(s)|r_{1}=\tfrac{a_{1}m_{1}}{2\pi}|\log(s)|. When n>1n>1 the flow is a product of the flows in each factor i\mathbb{C}_{i}, and the time taken to flow the circle (imu)i({{\operatorname{im}}}\,u)\cap\mathbb{C}_{i} back to E(a)i\partial E({\vec{a}})\cap\mathbb{C}_{i} is approximately aimi2π|log(s)|\tfrac{a_{i}m_{i}}{2\pi}|\log(s)| for s0s\approx 0. Thus ρa(u(z))min1inaimi2π|log(s)|\rho_{{\vec{a}}}(u(z))\approx\min\limits_{1\leq i\leq n}\tfrac{a_{i}m_{i}}{2\pi}|\log(s)|, and, if this minimum is attained for i=i0i=i_{0}, the limiting orbit is an mi0m_{i_{0}}-fold cover of the circle E(a)i0\partial E({\vec{a}})\cap\mathbb{C}_{i_{0}}.

Proposition 3.1.2 also has a more global analogue which applies to any closed symplectic manifold MM and which will form the basis of our proof of Theorem 3.4.1. However, we must first establish Theorem 3.3.2, namely that the counts NM,A<𝒞𝐃mx0>N_{M,A}\Langle\mathcal{C}^{\vec{m}}_{\vec{\mathbf{D}}}x_{0}\Rangle are robustly defined.

3.2. Hidden constraints for cuspidal degenerations

The basic strategy for establishing robustness of moduli spaces M,A<𝒞mx0>\mathcal{M}_{M,A}\Langle\mathcal{C}^{\vec{m}}x_{0}\Rangle will closely parallel the argument we used for Ma,A(𝔬c1(A)1a)\mathcal{M}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}) in §2, i.e. we seek to show that no bad degenerations can occur for a generic one-parameter family of almost complex structures. For punctured curves in MaM_{\vec{a}} we considered the SFT compactification by pseudoholomorphic buildings, and we ultimately showed that this agrees with the uncompactified moduli space. Similarly, for closed curves in MM with multidirectional tangency constraints we will consider a compatification by stable maps, and seek to show that this in fact agrees with the uncompactified moduli space.

As a warmup, recall that the multidirectional tangency constraint <𝒞𝐃mx0>\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle reduces to a unidirectional tangency constraint <𝒯D1(m)x0>\Langle\mathcal{T}^{(m)}_{D_{1}}x_{0}\Rangle in the case m=(m,1,,1){\vec{m}}=(m,1,\dots,1), and robustness of the moduli spaces M,A<𝒯D1(m)x0>\mathcal{M}_{M,A}\Langle\mathcal{T}_{D_{1}}^{(m)}x_{0}\Rangle was established in [MS3, §2] for M2nM^{2n} semipositive (here m=c1(A)1m=c_{1}(A)-1). The main subtlety in establishing robustness comes from the possibility of ghost degenerations, since strictly speaking a marked point on a constant curve component is tangent to any local divisor through its image to infinite order. This issue is resolved by observing that the nearby nonconstant curve components satisfy tangency conditions which collectively “remember” the initial constraint <𝒯D1(m)x0>\Langle\mathcal{T}_{D_{1}}^{(m)}x_{0}\Rangle.

The naive extension of this argument is actually insufficient to rule out ghost degenerations for multidirectional tangency constraints, as we now explain. Fix a closed symplectic manifold M2nM^{2n}, a homology class AH2(M)A\in H_{2}(M), and m=(m1,,mn)1n{\vec{m}}=(m_{1},\dots,m_{n})\in\mathbb{Z}_{\geq 1}^{n} with s=1mms=c1(A)+n2\sum_{s=1}^{m}m_{s}=c_{1}(A)+n-2. Fix also a point x0Mx_{0}\in M, a set 𝐃=(𝐃1,,𝐃n){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\dots,\mathbf{D}_{n}) of spanning local divisors at x0x_{0}, and an almost complex structure J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}).666More generally we could take a sequence J1,J2,J3,𝒥(M,𝐃)J_{1},J_{2},J_{3},\dots\in\mathcal{J}(M,{\vec{\mathbf{D}}}) converging to some J𝒥(M,𝐃)J_{\infty}\in\mathcal{J}(M,{\vec{\mathbf{D}}}), but we will suppress this to keep the notation simpler. Let C1,C2,C3,M,AJ<𝒞𝐃m>C_{1},C_{2},C_{3},\dots\in\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{\vec{m}}\Rangle be a sequence of curves which converges to some stable map C¯M,AJ<x0>C_{\infty}\in\overline{\mathcal{M}}_{M,A}^{J}\Langle x_{0}\Rangle. Let Q0Q_{0} be the component of CC_{\infty} which carries the constraint <𝒞𝐃m>\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{\vec{m}}\Rangle, and suppose that Q0Q_{0} is a ghost component. Let GG be the maximal ghost tree containing Q0Q_{0}, i.e. the set of all ghost components in CC_{\infty} which are connected to Q0Q_{0} through ghost components. Let Q1,,QkQ_{1},\dots,Q_{k} (for some k2k\in\mathbb{Z}_{\geq 2}) denote the nonconstant curve components of CC_{\infty} which are nodally adjacent to a curve component of GG, and let z1,,zkz_{1},\dots,z_{k} denote the corresponding special points of Q1,,QkQ_{1},\dots,Q_{k} respectively. For i=1,,ki=1,\dots,k, ziz_{i} strictly carries a constraint <𝒞𝐃mix0>\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{{\vec{m}}_{i}}x_{0}\Rangle for some mi=(m1i,,mni)1n{\vec{m}}_{i}=(m^{i}_{1},\dots,m^{i}_{n})\in\mathbb{Z}_{\geq 1}^{n}.

Lemma 3.2.1 ([CM1, Lem. 7.2]).

In the above situation we have

i=1kmsims\displaystyle\sum_{i=1}^{k}m_{s}^{i}\geq m_{s} (3.2.1)

for s=1,,ns=1,\dots,n.

Example 3.2.2.

Suppose that CC_{\infty} consists precisely of the curve components Q0,,QkQ_{0},\dots,Q_{k}, with k2k\in\mathbb{Z}_{\geq 2}. Using (3.2.1)\eqref{eq:CM_ineq} and the index zero assumption s=1nms=c1(A)+n2\sum_{s=1}^{n}m_{s}=c_{1}(A)+n-2, we have

i=1kind(Qi)\displaystyle\sum_{i=1}^{k}{\operatorname{ind}}(Q_{i}) =i=1k(2c1([Qi]+2n42s=1nmsi))\displaystyle=\sum_{i=1}^{k}\left(2c_{1}([Q_{i}]+2n-4-2\sum_{s=1}^{n}m_{s}^{i})\right)
=2c1(A)+k(2n4)2i=1ks=1nmsi\displaystyle=2c_{1}(A)+k(2n-4)-2\sum_{i=1}^{k}\sum_{s=1}^{n}m_{s}^{i}
2c1(A)+k(2n4)2s=1nms\displaystyle\leq 2c_{1}(A)+k(2n-4)-2\sum_{s=1}^{n}m_{s}
=(2n4)(k1).\displaystyle=(2n-4)(k-1).

Even if we assume ind(Qi)0{\operatorname{ind}}(Q_{i})\geq 0 for i=1,,ki=1,\dots,k (e.g. if each QiQ_{i} is simple and J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}) is generic), this does not immediately give any contradiction. \Diamond

Example 3.2.3.

Let us further specialize the previous example by taking m=(m,1,,1){\vec{m}}=(m,1,\dots,1) (i.e. the case of local tangency constraints). Since each msim_{s}^{i} is at least 11, for s=2,,ns=2,\dots,n we have i=1kmsik2>1=ms\sum_{i=1}^{k}m_{s}^{i}\geq k\geq 2>1=m_{s}, and hence

i=1ks=1nmsim+k(n1).\displaystyle\sum_{i=1}^{k}\sum_{s=1}^{n}m_{s}^{i}\geq m+k(n-1).

Using m=c1(A)1m=c_{1}(A)-1, this gives

i=1kind(Qi)22k2\displaystyle\sum_{i=1}^{k}{\operatorname{ind}}(Q_{i})\leq 2-2k\leq-2

which is a contradiction at least if ind(Q1),,ind(Qk)0{\operatorname{ind}}(Q_{1}),\dots,{\operatorname{ind}}(Q_{k})\geq 0. \Diamond

Although Lemma 3.2.1 is generally insufficient for ruling out bad degenerations, the following proposition puts stronger restrictions on degenerations of multidirectional tangency constraints.

Proposition 3.2.4.

In the same situation as Lemma 3.2.1, we have

i=1kmin1snasmsimin1snasms\displaystyle\sum_{i=1}^{k}\min\limits_{1\leq s\leq n}a_{s}m^{i}_{s}\geq\min\limits_{1\leq s\leq n}a_{s}m_{s} (3.2.2)

for any choice of a=(a1,,an)>0n{\vec{a}}=(a_{1},\dots,a_{n})\in\mathbb{R}_{>0}^{n}.

Remark 3.2.5.

Observe that (3.2.2) imples (3.2.1). Indeed by choosing a=(a1,,an){\vec{a}}=(a_{1},\dots,a_{n}) so that aj=1a_{j}=1 and as1a_{s}\gg 1 for sjs\neq j, we get min1snasmsi=mji\min\limits_{1\leq s\leq n}a_{s}m_{s}^{i}=m_{j}^{i} and min1snasms=mj\min\limits_{1\leq s\leq n}a_{s}m_{s}=m_{j}. \Diamond

Proof of Proposition 3.2.4.

Recall that JJ is assumed to be integrable near x0x_{0}. Pick local complex coordinates z1,,znz_{1},\dots,z_{n} for (M,J)(M,J) centered at x0x_{0} and defined in some open neighborhood UMU\subset M such that 𝐃s={zs=0}\mathbf{D}_{s}=\{z_{s}=0\} for s=1,,ns=1,\dots,n. Let ζ:Un\zeta:U\rightarrow\mathbb{C}^{n} denote the corresponding holomorphic chart.

For i=1,,ki=1,\dots,k, the restriction of QiQ_{i} to a small neighborhood of ziz_{i} has image contained in UU, with only ziz_{i} mapping to x0x_{0}. By choosing a complex coordinate near ziz_{i} we view this as a holomorphic map Q̊i:𝔻U{x0}\mathring{Q}_{i}:\mathbb{D}_{\circ}\rightarrow U\setminus\{x_{0}\}. Recalling the diffeomorphism Φa:×E(a)n{0}\Phi_{\vec{a}}:\mathbb{R}\times\partial E({\vec{a}})\rightarrow\mathbb{C}^{n}\;\setminus\;\{\vec{0}\} from Proposition 3.1.2, we consider the JE(a)J_{\partial E({\vec{a}})}-holomorphic composition

ηi:=Φa1ζQ̊i:𝔻×E(a),\displaystyle\eta_{i}:=\Phi_{\vec{a}}^{-1}\circ\zeta\circ\mathring{Q}_{i}:\mathbb{D}_{\circ}\rightarrow\mathbb{R}\times\partial E({\vec{a}}),

By Proposition 3.1.2, ηi\eta_{i} is negatively asymptotic at the puncture 0𝔻0\in\mathbb{D}_{\circ} to the Reeb orbit 𝔬mia\mathfrak{o}^{\vec{a}}_{-{\vec{m}}_{i}} in E(a)\partial E({\vec{a}}), which has action 𝒜(𝔬mia)=min1snasmsi\mathcal{A}(\mathfrak{o}^{\vec{a}}_{-{\vec{m}}_{i}})=\min\limits_{1\leq s\leq n}a_{s}m^{i}_{s}. In particular, for any ε>0\varepsilon>0 we can find R0R\ll 0 so that the loop γi\gamma_{i} given by restricting ηi\eta_{i} to the preimage of {r=R}\{r=R\} satisfies

|(γi)αamin1snasmsi|<ε.\displaystyle\left|\int(\gamma_{i})^{*}\alpha_{\vec{a}}-\min\limits_{1\leq s\leq n}a_{s}m_{s}^{i}\right|<\varepsilon.

Here αa\alpha_{\vec{a}} denotes the standard contact one-form on E(a)\partial E({\vec{a}}), i.e. the restriction of the Liouville one-form λstd=12s=1n(xidyiyidxi)\lambda_{\operatorname{std}}=\tfrac{1}{2}\sum_{s=1}^{n}(x_{i}dy_{i}-y_{i}dx_{i}) on n\mathbb{C}^{n}.

For j1j\in\mathbb{Z}_{\geq 1} sufficiently large, we can restrict the curve CjC_{j} to the preimage of U{x0}U\setminus\{x_{0}\}, then postcompose with Φa1ζ\Phi_{\vec{a}}^{-1}\circ\zeta, and finally restrict to the preimage of {rR}\{r\leq R\} to obtain a JE(a)J_{\partial E({\vec{a}})}-holomorphic map μj:Σj×E(a)\mu_{j}:\Sigma_{j}\rightarrow\mathbb{R}\times\partial E({\vec{a}}). Here Σj\Sigma_{j} is a Riemann surface with kk boundary circles c1j,,ckjc^{j}_{1},\dots,c^{j}_{k} and one interior puncture, such that for i=1,,ki=1,\dots,k we have

|(γij)αamin1snasmsi|<2ε,\displaystyle\left|\int(\gamma^{j}_{i})^{*}\alpha_{\vec{a}}-\min\limits_{1\leq s\leq n}a_{s}m_{s}^{i}\right|<2\varepsilon,

where γij\gamma^{j}_{i} denotes the restriction of μj\mu_{j} to cijc^{j}_{i}. Since by Proposition 3.1.2 μj\mu_{j} is negatively asymptotic at its interior puncture to the Reeb orbit 𝔬ma\mathfrak{o}^{\vec{a}}_{-{\vec{m}}} in E(a)\partial E({\vec{a}}) of action min1snasms\min\limits_{1\leq s\leq n}a_{s}m_{s}, we can find RRR^{\prime}\ll R so that the loop ρj\rho_{j} given by restricting μj\mu_{j} to {r=R}\{r=R^{\prime}\} satisfies

|ρjαamin1snasms|<ε.\displaystyle\left|\int\rho_{j}^{*}\alpha_{\vec{a}}-\min\limits_{1\leq s\leq n}a_{s}m_{s}\right|<\varepsilon.

Now put (Σj)RR:=μj1({RrR})(\Sigma_{j})^{R}_{R^{\prime}}:=\mu_{j}^{-1}(\{R^{\prime}\leq r\leq R\}). This is a Riemann surface with boundary circles c1j,,ckjc_{1}^{j},\dots,c_{k}^{j} mapping to {r=R}\{r=R\} and an additional boundary circle c0jc^{j}_{0} which maps to {r=R}\{r=R^{\prime}\}. By Stokes’ theorem and nonnegativity of energy, we have

0(Σj)RRμj𝑑αa\displaystyle 0\leq\int_{(\Sigma_{j})^{R}_{R^{\prime}}}\mu_{j}^{*}d\alpha_{\vec{a}} =i=1k(γij)αaρjαa\displaystyle=\sum_{i=1}^{k}\int(\gamma^{j}_{i})^{*}\alpha_{\vec{a}}-\int\rho_{j}^{*}\alpha_{\vec{a}}
i=1kmin1snasmsimin1snasms+(k+1)ε.\displaystyle\leq\sum_{i=1}^{k}\min\limits_{1\leq s\leq n}a_{s}m_{s}^{i}-\min\limits_{1\leq s\leq n}a_{s}m_{s}+(k+1)\varepsilon.

By ε>0\varepsilon>0 was arbitrarily small, this gives (3.2.2). ∎

Remark 3.2.6.

We can view Proposition 3.2.4 as saying that whenever a constraint <𝒞mpt>\Langle\mathcal{C}^{{\vec{m}}}{\operatorname{pt}}\Rangle degenerates into constraints <𝒞m1pt>,,<𝒞mkpt>\Langle\mathcal{C}^{{\vec{m}}_{1}}{\operatorname{pt}}\Rangle,\dots,\Langle\mathcal{C}^{{\vec{m}}_{k}}{\operatorname{pt}}\Rangle there must exist a formal rational curve in E(a)\partial E({\vec{a}}) with positive ends 𝔬m1a,,𝔬mka\mathfrak{o}^{\vec{a}}_{-{\vec{m}}_{1}},\dots,\mathfrak{o}^{\vec{a}}_{-{\vec{m}}_{k}} and negative end 𝔬ma\mathfrak{o}^{\vec{a}}_{-{\vec{m}}}, for any choice of (rationally independent) a>0n{\vec{a}}\in\mathbb{R}_{>0}^{n}. Recall that Conditions (b) and (c) put restrictions on formal curves in E(a)\partial E({\vec{a}}), and we have seen that these conditions hold e.g. under the assumptions of Lemma 2.3.7. \Diamond

Another way to infer hidden constraints is via iterated blowups. Since the numerics become rather complicated we just illustrate this idea with a simple example.

Example 3.2.7.

We will show that the constraint <𝒞(3,2)pt>\Langle\mathcal{C}^{(3,2)}{\operatorname{pt}}\Rangle cannot degenerate into <𝒞(2,1)pt>\Langle\mathcal{C}^{(2,1)}{\operatorname{pt}}\Rangle and <𝒞(1,1)pt>\Langle\mathcal{C}^{(1,1)}{\operatorname{pt}}\Rangle. More precisely, consider a sequence of curves C1,C2,C3,M,BJ<𝒞(𝐃1,𝐃2)(3,2)x0>C_{1},C_{2},C_{3},\dots\in\mathcal{M}_{M,B}^{J}\Langle\mathcal{C}^{(3,2)}_{(\mathbf{D}_{1},\mathbf{D}_{2})}x_{0}\Rangle, each of which strictly satisfies the constraint, and suppose by contradiction that these converge to a curve CC_{\infty} consisting of two irreducible components C1,C2C_{\infty}^{1},C_{\infty}^{2} which strictly satisfy the constraints <𝒞(𝐃1,𝐃2)(3,2)x0>\Langle\mathcal{C}^{(3,2)}_{(\mathbf{D}_{1},\mathbf{D}_{2})}x_{0}\Rangle and <𝒞(𝐃1,𝐃1)(3,2)x0>\Langle\mathcal{C}^{(3,2)}_{(\mathbf{D}_{1},\mathbf{D}_{1})}x_{0}\Rangle respectively. We assume also that all of these curves are disjoint from x0x_{0} away from the main constraints.

After blowing up at x0x_{0}, the proper transforms C~1,C~2,C~3,\widetilde{C}_{1},\widetilde{C}_{2},\widetilde{C}_{3},\dots all lie in homology class B2eB-2e, where e=[𝔼]H2(Bl1M)e=[\mathbb{E}]\in H_{2}({{\operatorname{Bl}}}^{1}M) is the homology class of the exceptional divisor 𝔼\mathbb{E}. After possibly passing to a subsequence, these converge to a curve C~\widetilde{C}_{\infty} which projects to CC_{\infty}, and hence consists of the proper transforms C~1,C~2\widetilde{C}_{\infty}^{1},\widetilde{C}_{\infty}^{2} of C1,C2C_{\infty}^{1},C_{\infty}^{2} respectively. A priori C~\widetilde{C}_{\infty} could also have some additional components which are covers of 𝔼\mathbb{E}, but this is ruled out since [C~i]=[Ci]e[\widetilde{C}_{\infty}^{i}]=[C_{\infty}^{i}]-e for i=1,2i=1,2 and hence [C~1]+[C~2]=B2e[\widetilde{C}_{\infty}^{1}]+[\widetilde{C}_{\infty}^{2}]=B-2e. But this is a contradiction, since C~1\widetilde{C}_{\infty}^{1} and C~2\widetilde{C}_{\infty}^{2} are disjoint near 𝔼\mathbb{E}.

Suppose by contradiction that CC is a curve in homology class BB which strictly satisfies <𝒞(𝐃1,𝐃2)(3,2)x0>\Langle\mathcal{C}^{(3,2)}_{(\mathbf{D}_{1},\mathbf{D}_{2})}x_{0}\Rangle, and which degenerates into curves C1C_{1} and C2C_{2} which strictly satisfy <𝒞(𝐃1,𝐃2)(2,1)x0>\Langle\mathcal{C}^{(2,1)}_{(\mathbf{D}_{1},\mathbf{D}_{2})}x_{0}\Rangle and <𝒞(𝐃1,𝐃2)(1,1)x0>\Langle\mathcal{C}^{(1,1)}_{(\mathbf{D}_{1},\mathbf{D}_{2})}x_{0}\Rangle. Blowing up at x0x_{0}, the proper transform C~\widetilde{C} of CC lies in homology class B2eB-2e, while the proper transforms of C~1\widetilde{C}_{1} and C~2\widetilde{C}_{2} and C1C_{1} and C2C_{2} lie in homology classes B1eB_{1}-e and B2eB_{2}-e respectively. Then C~\widetilde{C} degenerates into C~1C~2\widetilde{C}_{1}\cup\widetilde{C}_{2} along with some number of copies of (covers of) 𝔼\mathbb{E}. However this is not possible since C~1\widetilde{C}_{1} and C~2\widetilde{C}_{2} are locally disjoint. \Diamond

3.3. Counting curves with multidirectional tangency constraints

As before, let M2nM^{2n} be a closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, and m=(m1,,mn)1n{\vec{m}}=(m_{1},\dots,m_{n})\in\mathbb{Z}_{\geq 1}^{n} a tuple satisfying s=1nms=c1(A)+n2\sum_{s=1}^{n}m_{s}=c_{1}(A)+n-2. As usual, 𝐃=(𝐃1,,𝐃n){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\dots,\mathbf{D}_{n}) denotes a collection of smooth local symplectic divisors which span at a point x0Mx_{0}\in M.

Definition 3.3.1.

We will say that M,A<𝒞mpt>\mathcal{M}_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle is robust if M,AJ<𝒞𝐃mx0>\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}^{\vec{m}}_{\vec{\mathbf{D}}}x_{0}\Rangle is finite and regular for any choice of x0,𝐃x_{0},{\vec{\mathbf{D}}}, and generic J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}), and moreover the (signed) count #M,AJ<𝒞𝐃mx0>\#\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}^{\vec{m}}_{\vec{\mathbf{D}}}x_{0}\Rangle is independent of these choices.

We will say a robust moduli space M,A<𝒞mpt>\mathcal{M}_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle is deformation invariant if it remains robust under deformations of the symplectic form on MM, and moreover the count #M,A<𝒞mpt>\#\mathcal{M}_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle is unchanged under such deformations.

Given a tuple a>0n{\vec{a}}\in\mathbb{R}_{>0}^{n}, recall that we have the lattice path Δ1a,Δ2a,Δ3a,1n\Delta^{\vec{a}}_{1},\Delta^{\vec{a}}_{2},\Delta^{\vec{a}}_{3},\dots\in\mathbb{Z}_{\geq 1}^{n} from Definition 1.4.1.

Theorem 3.3.2.

Let M2nM^{2n} be a semipositive closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, 𝐃=(𝐃1,𝐃n){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\dots\mathbf{D}_{n}) a collection of spanning local divisors at a point x0Mx_{0}\in M, and m=(m1,,mn)1n{\vec{m}}=(m_{1},\dots,m_{n})\in\mathbb{Z}_{\geq 1}^{n} a tuple satisfying s=1nms=c1(A)+n2\sum_{s=1}^{n}m_{s}=c_{1}(A)+n-2. Assume that there exists a tuple a>0n{\vec{a}}\in\mathbb{R}_{>0}^{n} satisfying Assumptions A, B, and C, and such that Δc1(A)1a=m\Delta^{\vec{a}}_{c_{1}(A)-1}={\vec{m}}. Then the moduli space M,A<𝒞mpt>\mathcal{M}_{M,A}\Langle\mathcal{C}^{{\vec{m}}}{\operatorname{pt}}\Rangle is robust and deformation invariant.

Remark 3.3.3.

Theorem 3.3.2 implies Theorem B as follows. Taking a=(p,q+,a3,,an){\vec{a}}=(p,q^{+},a_{3},\dots,a_{n}) with a3,,an>pqa_{3},\dots,a_{n}>pq, we have Δp+q1a=(p,q,1,,1)\Delta^{\vec{a}}_{p+q-1}=(p,q,1,\dots,1) whence Assumptions A and B hold by Lemma 2.3.7. \Diamond

Proof sketch of Theorem 3.3.2.

Given J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}), we denote by ¯M,AJ<𝒞𝐃mx0>\overline{\mathcal{M}}^{J}_{M,A}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle the closure of M,AJ<𝒞𝐃mx0>\mathcal{M}^{J}_{M,A}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle in the compactified moduli space ¯M,AJ<x0>\overline{\mathcal{M}}_{M,A}^{J}\Langle x_{0}\Rangle (i.e. the space of maps passing through x0x_{0}). Similarly, given a one-parameter family {Jt𝒥(M,𝐃)|t[0,1]}\{J_{t}\in\mathcal{J}(M,{\vec{\mathbf{D}}})\;|\;t\in[0,1]\}, we define ¯M,A{Jt}<𝒞𝐃mx0>\overline{\mathcal{M}}_{M,A}^{\{J_{t}\}}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle to be the closure of M,A{Jt}<𝒞𝐃mx0>\mathcal{M}_{M,A}^{\{J_{t}\}}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle in ¯M,A{Jt}<x0>\overline{\mathcal{M}}_{M,A}^{\{J_{t}\}}\Langle x_{0}\Rangle. The main task is to establish ¯M,A{Jt}<𝒞𝐃mx0>=M,A{Jt}<𝒞𝐃mx0>\overline{\mathcal{M}}_{M,A}^{\{J_{t}\}}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle=\mathcal{M}_{M,A}^{\{J_{t}\}}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle for generic {Jt}\{J_{t}\}.

The most interesting point is to rule out degenerations as in Example 3.2.2, i.e. with C¯M,A{Jt}<𝒞𝐃mx0>C_{\infty}\in\overline{\mathcal{M}}_{M,A}^{\{J_{t}\}}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle composed of a ghost component Q0Q_{0} and nonconstant components Q1,,QnQ_{1},\dots,Q_{n} which carry respective constraints m1,,mk{\vec{m}}_{1},\dots,{\vec{m}}_{k}. Given such a degeneration, by Proposition 3.2.4 there exists a formal curve in E(a)\partial E({\vec{a}}) with positive ends 𝔬m1a,,𝔬mka\mathfrak{o}^{\vec{a}}_{-{\vec{m}}_{1}},\dots,\mathfrak{o}^{\vec{a}}_{-{\vec{m}}_{k}} and negative end 𝔬ma\mathfrak{o}^{\vec{a}}_{-{\vec{m}}}, where by assumption a>0n{\vec{a}}\in\mathbb{R}_{>0}^{n} satisfies Assumptions A and B and we have Δc1(A)1a=m\Delta^{\vec{a}}_{c_{1}(A)-1}={\vec{m}} (c.f. Remark 3.2.6).

For the purposes of index calculations, we could view Q1,,QkQ_{1},\dots,Q_{k} as curve components Q~1,,Q~k\widetilde{Q}_{1},\dots,\widetilde{Q}_{k} in M^a\widehat{M}_{\vec{a}}, where Q~i\widetilde{Q}_{i} has negative end 𝔬mia\mathfrak{o}^{\vec{a}}_{-{\vec{m}}_{i}} for i=1,,ki=1,\dots,k. On a purely formal level, note that the index of Ma,[Qi](𝔬mia)\mathcal{M}_{M_{\vec{a}},[Q_{i}]}(\mathfrak{o}^{\vec{a}}_{-{\vec{m}}_{i}}) is at least that of M,[Qi]<𝒞mipt>\mathcal{M}_{M,[Q_{i}]}\Langle\mathcal{C}^{{\vec{m}}_{i}}{\operatorname{pt}}\Rangle for i=1,,ki=1,\dots,k. The upshot is that any index lower bound for the closed curve QiQ_{i} is a fortiori true for its punctured counterpart Q~i\widetilde{Q}_{i}. Also, similar to the arguments in §2.5, we can assume by genericity of {Jt}\{J_{t}\} and standard transversality techniques that the underlying simple curves of Q1,,QkQ_{1},\dots,Q_{k} have nonnegative indices. But now the existence of the configuration Q~0,,Q~k\widetilde{Q}_{0},\dots,\widetilde{Q}_{k} is ruled out essentially by the argument in §2.4.

The rest of the proof also closely follows that of Theorem 2.3.5, after formally trading closed curves with multidirectional tangency constraints in MM for punctured curves in M^a\widehat{M}_{\vec{a}}, and ghost trees in MM for formal curves in E(a)\partial E({\vec{a}}). We leave the details to the reader. ∎

3.4. Multidirectional tangencies and ellipsoidal ends: global equivalence

The following is our main result equating closed curves with multidirectional tangency constraints and punctured curves with ellipsoidal ends.

Theorem 3.4.1.

Let M2nM^{2n} be a semipositive closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, and a>0n{\vec{a}}\in\mathbb{R}_{>0}^{n} a rationally independent tuple satisfying Assumptions A, B, and C. Put m:=Δc1(A)1a1n{\vec{m}}:=\Delta^{\vec{a}}_{c_{1}(A)-1}\in\mathbb{Z}_{\geq 1}^{n}. Then we have

𝐓M,Aa=NM,A<𝒞mpt>.\displaystyle\mathbf{T}_{M,A}^{\vec{a}}=N_{M,A}\Langle\mathcal{C}^{{\vec{m}}}{\operatorname{pt}}\Rangle.
Remark 3.4.2.

Similar to Remark 3.3.3, Theorem 3.4.1 implies Theorem D via Lemma 2.3.7. \Diamond

One can imagine proving Theorem D by surrounding the constraint <𝒞mpt>\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle with an ellipsoid of shape a{\vec{a}} and stretching the neck. This would a detailed understanding of curves with local multidirectional tangency constraints in E^(a)\widehat{E}({\vec{a}}) and their gluings. Instead, we will give a more direct correspondence by establishing an explicit bijection

Ma,AJE(𝔬c1(A)1a)M,AJC<𝒞𝐃mx0>\displaystyle\mathcal{M}_{M_{\vec{a}},A}^{J_{E}}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})\cong\mathcal{M}_{M,A}^{J_{C}}\Langle\mathcal{C}^{{\vec{m}}}_{{\vec{\mathbf{D}}}}x_{0}\Rangle

for certain special choices of JE𝒥(Ma)J_{E}\in\mathcal{J}(M_{\vec{a}}) and JC𝒥(M,𝐃)J_{C}\in\mathcal{J}(M,{\vec{\mathbf{D}}}).

In the following, by complex Darboux coordinates we mean local coordinates which simultaneously trivialize a symplectic form and an almost complex structure. For a Kähler manifold, such coordinates exist if and only if the Kähler metric is flat, e.g. 𝕋2n\mathbb{T}^{2n} works but not n\mathbb{CP}^{n}.

Proposition 3.4.3 ([MS6, §4]).

Let M2nM^{2n} be a closed symplectic manifold and J𝒥(M)J\in\mathcal{J}(M) a tame almost complex structure such that there exist complex Darboux coordinates z1,,znz_{1},\dots,z_{n} centered at x0x_{0}. Then, for any given a>0n{\vec{a}}\in\mathbb{R}_{>0}^{n} rationally independent, there exists an explicit diffeomorphism

Q:M^aM{x0}\displaystyle Q:\widehat{M}_{\vec{a}}\rightarrow M\setminus\{x_{0}\}

such that QJQ^{*}J lies in 𝒥(Ma)\mathcal{J}(M_{\vec{a}}).

Moreover, put 𝐃=(𝐃1,,𝐃n){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\dots,\mathbf{D}_{n}), with 𝐃i={zi=0}\mathbf{D}_{i}=\{z_{i}=0\} for i=1,,ni=1,\dots,n, suppose that u:1Mu:\mathbb{CP}^{1}\rightarrow M is a JJ-holomorphic map which strictly satisfies <𝒞𝐃mx0>\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{\vec{m}}x_{0}\Rangle at a point z01z_{0}\in\mathbb{CP}^{1} for some m1n{\vec{m}}\in\mathbb{Z}_{\geq 1}^{n}, and is otherwise disjoint from x0x_{0}. Then the QJQ^{*}J-holomorphic map

Φa1u|1{z0}:1{z0}M^a\displaystyle\Phi_{\vec{a}}^{-1}\circ u|_{\mathbb{CP}^{1}\setminus\{z_{0}\}}:\mathbb{CP}^{1}\setminus\{z_{0}\}\rightarrow\widehat{M}_{\vec{a}}

is negatively asymptotic to the Reeb orbit 𝔬ma\mathfrak{o}^{\vec{a}}_{-{\vec{m}}} in E(a)\partial E({\vec{a}}) at the puncture z0z_{0}.

Here the map QQ is given by the formula in (3.1.1) near the negative end of M^a\widehat{M}_{\vec{a}} and is slowly deformed to the identity away from this end.

We now complete the proof of Theorem 3.4.1.

Proof of Theorem 3.4.1.

Pick J𝒥(M)J\in\mathcal{J}(M) as in Proposition 3.4.3, which we can assume is generic away from x0Mx_{0}\in M. Then Proposition 3.4.3 sets up a bijection

mM,AJ<𝒞𝐃mx0>Ma,AQJ(𝔬c1(A)1a),\displaystyle\bigcup\limits_{{\vec{m}}^{\prime}}\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{{\vec{m}}^{\prime}}x_{0}\Rangle\cong\mathcal{M}^{Q^{*}J}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}), (3.4.1)

where the union is over all m1n{\vec{m}}^{\prime}\in\mathbb{Z}_{\geq 1}^{n} satisfying 𝔬ma=𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{-{\vec{m}}^{\prime}}=\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}. By definition, m=Δc1(A)1a{\vec{m}}=\Delta^{\vec{a}}_{c_{1}(A)-1} is the unique tuple m=(m1,,mn){\vec{m}}^{\prime}=(m_{1}^{\prime},\dots,m_{n}^{\prime}) which minimizes s=1nms\sum_{s=1}^{n}m_{s}^{\prime} subject to 𝔬ma=𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{-{\vec{m}}^{\prime}}=\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}. It follows from (3.0.1) that for any other m{\vec{m}}^{\prime} satisfying 𝔬ma=𝔬c1(A)1a\mathfrak{o}^{\vec{a}}_{-{\vec{m}}^{\prime}}=\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1} we have

indM,AJ<𝒞𝐃mx0><indM,AJ<𝒞𝐃mx0>=0,\displaystyle{\operatorname{ind}}\,\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{{\vec{m}}^{\prime}}x_{0}\Rangle\;\,<\;\,{\operatorname{ind}}\,\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{\vec{m}}x_{0}\Rangle=0,

and hence M,A<𝒞𝐃mx0>=\mathcal{M}_{M,A}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{{\vec{m}}^{\prime}}x_{0}\Rangle=\varnothing by our genericity assumption on JJ. It follows that (3.4.1) is in fact a bijection

M,AJ<𝒞𝐃mx0>Ma,AQJ(𝔬c1(A)1a),\displaystyle\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{{\vec{m}}}x_{0}\Rangle\cong\mathcal{M}^{Q^{*}J}_{M_{\vec{a}},A}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1}),

and hence we have

𝐓M,Aa=#Ma,AQJ(𝔬c1(A)1a)=#M,AJ<𝒞𝐃mx0>=NM,A<𝒞mpt>.\displaystyle\mathbf{T}_{M,A}^{\vec{a}}=\#\mathcal{M}_{M_{\vec{a}},A}^{Q^{*}J}(\mathfrak{o}^{\vec{a}}_{c_{1}(A)-1})=\#\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{{\vec{m}}}x_{0}\Rangle=N_{M,A}\Langle\mathcal{C}^{{\vec{m}}}{\operatorname{pt}}\Rangle.

3.5. Cusps, multidirectional tangencies, and singular symplectic curves

We now elaborate on the relationship between cusp singularities and multidirectional tangency constraints. Let C2C\subset\mathbb{C}^{2} be an algebraic curve. A singular point x0Cx_{0}\in C is a (𝐩,𝐪)\mathbf{(p,q)} cusp if its link is the (p,q)(p,q) torus knot, i.e. if for ε>0\varepsilon>0 sufficiently small there is a diffeomorphism

(Sε3,CSε3)(S3,{xp+yq=0}S3),\displaystyle(S^{3}_{\varepsilon},C\cap S^{3}_{\varepsilon})\cong(S^{3},\{x^{p}+y^{q}=0\}\cap S^{3}),

where Sε32S^{3}_{\varepsilon}\subset\mathbb{C}^{2} denotes the sphere of radius ε\varepsilon centered at x0x_{0}. More generally, any singularity of an algebraic curve in 2\mathbb{C}^{2} has a link which is an iterated torus link (see [EN]).

The singular curves most relevant to the ellipsoidal superpotential are those having one main singularity which is a (p,q)(p,q) cusp, and possibly some additional singularities which are ordinary double points (i.e. modeled on {x2+y2=0}\{x^{2}+y^{2}=0\}).

Definition 3.5.1.

Given a symplectic four-manifold M4M^{4} and p,q1p,q\in\mathbb{Z}_{\geq 1} with gcd(p,q)=1\gcd(p,q)=1, a (𝐩,𝐪)\mathbf{(p,q)}-sesquicuspidal symplectic curve is a subset CMC\subset M such that

  • there is a point x0Cx_{0}\in C with an open neighborhood UMU\subset M such that (U,CU)(U,C\cap U) is symplectomorphic to (U,CU)(U^{\prime},C^{\prime}\cap U^{\prime}), where CC^{\prime} is an algebraic curve in 2\mathbb{C}^{2} having a (p,q)(p,q) cusp at x0Cx_{0}^{\prime}\in C^{\prime} and U2U^{\prime}\subset\mathbb{C}^{2} is an open neighborhood of x0x_{0}^{\prime}

  • C{x0}C\setminus\{x_{0}\} is a positively immersed symplectic submanifold of MM.

Remark 3.5.2.

Every sesquicuspidal curve is in particular a singular symplectic curve in the sense of [GS, Def. 2.5]. \Diamond

Lemma 3.5.3.

Let C2C\subset\mathbb{C}^{2} be an irreducible local algebraic curve near a point x02x_{0}\in\mathbb{C}^{2}, and fix p,q1p,q\in\mathbb{Z}_{\geq 1} relatively prime.

  1. (a)

    Suppose that CC strictly satisfies the constraint <𝒞(𝐃1,𝐃2)(p,q)x0>\Langle\mathcal{C}^{(p,q)}_{(\mathbf{D}_{1},\mathbf{D}_{2})}x_{0}\Rangle for some smooth local holomorphic divisors 𝐃1,𝐃2\mathbf{D}_{1},\mathbf{D}_{2} which span at x0x_{0}. Then CC has a (p,q)(p,q) cusp at x0x_{0}.

  2. (b)

    Conversely, suppose that CC has a (p,q)(p,q) cusp at x0x_{0}. Then we can find a smooth local divisors (𝐃1,𝐃2)(\mathbf{D}_{1},\mathbf{D}_{2}) intersecting transversely at x0x_{0} so that CC strictly satisfies the constraint <𝒞(𝐃1,𝐃2)(p,q)x0>\Langle\mathcal{C}^{(p,q)}_{(\mathbf{D}_{1},\mathbf{D}_{2})}x_{0}\Rangle.

Remark 3.5.4.

If CC has a (p,q)(p,q) cusp at x0x_{0}, then the multiplicity of CC at x0x_{0} is min{p,q}\min\{p,q\}, i.e. we have (𝐃C)x0min{p,q}(\mathbf{D}\cdot C)_{x_{0}}\geq\min\{p,q\} for any divisor 𝐃\mathbf{D} passing through x0x_{0}. If say p>qp>q and we have (𝐃1C)x0=p(\mathbf{D}_{1}\cdot C)_{x_{0}}=p, then we have (𝐃2C)x0=q(\mathbf{D}_{2}\cdot C)_{x_{0}}=q for any choice of smooth local divisor 𝐃2\mathbf{D}_{2} which intersects 𝐃1\mathbf{D}_{1} transversely at x0x_{0}. In other words, the choice of 𝐃2\mathbf{D}_{2} is essentially irrelevant. \Diamond

Proof.

To prove (a), let x,yx,y be complex coordinates for 2\mathbb{C}^{2} centered at the origin such that 𝐃1={x=0}\mathbf{D}_{1}=\{x=0\} and 𝐃2={y=0}\mathbf{D}_{2}=\{y=0\}. Recall that CC has a Newton–Puiseux parametrization of the form

y=c1xr1p1+c2xr2p1p2+c3xr3p1p2p3+,\displaystyle y=c_{1}x^{\frac{r_{1}}{p_{1}}}+c_{2}x^{\frac{r_{2}}{p_{1}p_{2}}}+c_{3}x^{\frac{r_{3}}{p_{1}p_{2}p_{3}}}+\cdots,

where

  • cic_{i}\in\mathbb{C}^{*} for i1i\in\mathbb{Z}_{\geq 1}

  • the Puiseux pairs (pi,ri)12(p_{i},r_{i})\in\mathbb{Z}_{\geq 1}^{2} are relatively prime for all i1i\in\mathbb{Z}_{\geq 1}

  • there exists L1L\in\mathbb{Z}_{\geq 1} such that pi=1p_{i}=1 for all i>Li>L

  • we have increasing exponents r1p1<r2p1p2<r3p1p2p3<\tfrac{r_{1}}{p_{1}}<\tfrac{r_{2}}{p_{1}p_{2}}<\tfrac{r_{3}}{p_{1}p_{2}p_{3}}<\cdots

(here we follow the conventions of [Neu]). The link of CC at x0x_{0} is then an iterated torus knot, with cabling parameters (pi,si)(p_{i},s_{i}) determined by s1=r1s_{1}=r_{1} and si=riri1pi+pi1pisi1s_{i}=r_{i}-r_{i-1}p_{i}+p_{i-1}p_{i}s_{i-1} for i2i\geq 2. This gives a parametrization of CC of the form

t(tp1pL,c1tr1p2pL+c2tr2p3pL+).\displaystyle t\mapsto(t^{p_{1}\cdots p_{L}},c_{1}t^{r_{1}p_{2}\cdots p_{L}}+c_{2}t^{r_{2}p_{3}\cdots p_{L}}+\cdots). (3.5.1)

and in particular we have intersection multiplicities p=C𝐃1=p1pLp=C\cdot\mathbf{D}_{1}=p_{1}\cdots p_{L} and q=C𝐃2=r1p2pLq=C\cdot\mathbf{D}_{2}=r_{1}p_{2}\cdots p_{L}. Since pp and qq are relatively prime, we must have p2==pL=1p_{2}=\cdots=p_{L}=1, i.e. there is just a single cabling parameter (p1,s1)=(p1,r1)=(p,q)(p_{1},s_{1})=(p_{1},r_{1})=(p,q).

To prove (b), let x,yx,y be complex coordinates for 2\mathbb{C}^{2} and consider the parameterization of CC as in (3.5.1). Observe that if CC has a (p,q)(p,q) cusp at x0x_{0} then there is at most one topologically nontrivial cabling parameter (pi,si)(p_{i},s_{i}), i.e for some k1k\in\mathbb{Z}_{\geq 1} we have pi=1p_{i}=1 for all iki\neq k. Then the parameterization (3.5.1) takes the form

t(tp1,c1tr1p1++ck1trk1p1+cktrk+ck+1trk+1+).\displaystyle t\mapsto(t^{p_{1}},c_{1}t^{r_{1}p_{1}}+\cdots+c_{k-1}t^{r_{k-1}p_{1}}+c_{k}t^{r_{k}}+c_{k+1}t^{r_{k+1}}+\cdots). (3.5.2)

Assuming p>qp>q without loss of generality, then the multiplicity of CC at x0x_{0} is qq, so we must have p1=qp_{1}=q. Note that si=ris_{i}=r_{i} for i=1,,ki=1,\dots,k, so have rk=sk=pr_{k}=s_{k}=p, and (3.5.2) becomes

t(tq,c1tr1q++ck1trk1q+cktp+).\displaystyle t\mapsto(t^{q},c_{1}t^{r_{1}q}+\cdots+c_{k-1}t^{r_{k-1}q}+c_{k}t^{p}+\cdots). (3.5.3)

Then the polynomial

F(x,y):=yc1xr1ck1xrk1,\displaystyle F(x,y):=y-c_{1}x^{r_{1}}-\cdots-c_{k-1}x^{r_{k-1}},

has lowest degree term tpt^{p} when composed with the parameterization (3.5.3), which means that 𝐃1:={F(x,y)=0}\mathbf{D}_{1}:=\{F(x,y)=0\} satisfies (𝐃1C)x0=p(\mathbf{D}_{1}\cdot C)_{x_{0}}=p. ∎

We are now ready to complete the proof of Theorem E.

Proof of Theorem E.

Suppose first that NM,A<𝒞(p,q)pt>N_{M,A}\Langle\mathcal{C}^{(p,q)}{\operatorname{pt}}\Rangle is nonzero. which means that for any choice of 𝐃=(𝐃1,𝐃2){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\mathbf{D}_{2}) and generic J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}) the moduli space M,AJ<𝒞𝐃(p,q)x0>\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{(p,q)}x_{0}\Rangle is nonempty. By genericity any curve CM,AJ<𝒞𝐃(p,q)x0>C\in\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{(p,q)}x_{0}\Rangle satisfies the constraint <𝒞𝐃(p,q)x0>\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{(p,q)}x_{0}\Rangle strictly, and hence has a (p,q)(p,q) cusp by Lemma 3.5.3(a). After a small perturbation we can further assume that all singularities of CC away from the main cusp are positive ordinary double points.

Conversely, suppose that there is a (p,q)(p,q)-sesquicuspidal symplectic curve CC in MM lying in homology class AA which is positively immersed away from the cusp point x0x_{0}. Let JJ be a tame almost complex structure on MM which is integrable near x0x_{0} and preserves CC. By Lemma 3.5.3(b) we can find 𝐃=(𝐃1,𝐃2){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\mathbf{D}_{2}) such that CC strictly satisfies <𝒞𝐃(p,q)x0>\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{(p,q)}x_{0}\Rangle. Finally, by Proposition 3.5.6 below we have NM,A<𝒞(p,q)x0>1N_{M,A}\Langle\mathcal{C}^{(p,q)}x_{0}\Rangle\geq 1. ∎

Note that by combining Theorem E and the symplectic deformation invariance part of Theorem 3.3.2, we have:

Corollary 3.5.5.

Let M4M^{4} be a closed symplectic four-manifold, and suppose that CC is an index zero (p,q)(p,q)-sesquicuspidal rational symplectic curve in MM. Then for any symplectic deformation MM^{\prime} of MM there exists an index zero (p,q)(p,q)-unicuspidal rational symplectic curve CC^{\prime} in MM^{\prime} satisfying [C]=[C][C^{\prime}]=[C].

Lastly, the following Proposition 3.5.6 allows us to give lower bounds for the counts NM,A<𝒞mpt>N_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle when MM is a smooth complex projective surface by constructing sesquicuspidal algebraic curves, even if the natural Kähler form on MM is not flat (e.g. M=2M=\mathbb{CP}^{2}).

Proposition 3.5.6.

Let M4M^{4} be a closed four-dimensional symplectic manifold and AH2(M)A\in H_{2}(M) a homology class such that p+q=c1(A)p+q=c_{1}(A) for some p,q1p,q\in\mathbb{Z}_{\geq 1} relatively prime. Let 𝐃=(𝐃1,𝐃2){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\mathbf{D}_{2}) be spanning local divisors at x0Mx_{0}\in M, and let J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}) be any almost complex structure. Suppose that there exist, for some k1k\in\mathbb{Z}_{\geq 1}, curves C1,,CkM,AJ<𝒞𝐃(p,q)x0>C_{1},\dots,C_{k}\in\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{(p,q)}x_{0}\Rangle which are immersed away from the constraint. Then we have NM,A<𝒞mpt>kN_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle\geq k.

Proof.

We do not assume that JJ is generic, but by a version of automatic transversality (see [Bar]) the curves C1,,CkC_{1},\dots,C_{k} are regular. Then for J(M,𝐃)J^{\prime}\in\mathcal{M}(M,{\vec{\mathbf{D}}}) any sufficiently small generic perturbation of JJ, the curves C1,,CkC_{1},\dots,C_{k} deform to curves C1,,CkM,AJ<𝒞𝐃mx0>C_{1}^{\prime},\dots,C_{k}^{\prime}\in\mathcal{M}^{J^{\prime}}_{M,A}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle. Again by automatic transversality, all curves in M,AJ<𝒞𝐃mx0>\mathcal{M}^{J^{\prime}}_{M,A}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle are regular and count positively (c.f. [MS5, §5.2]). Since JJ^{\prime} is generic, we have by definition NM,A<𝒞mpt>=#M,AJ<𝒞𝐃mx0>kN_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle=\#\mathcal{M}^{J^{\prime}}_{M,A}\Langle\mathcal{C}_{\vec{\mathbf{D}}}^{\vec{m}}x_{0}\Rangle\geq k. ∎

3.6. Stabilization invariance II

We end this section by discussing the closed curve counterpart of Theorem 2.6.1:

Theorem 3.6.1.

Let M2nM^{2n} be a closed symplectic manifold, AH2(M)A\in H_{2}(M) a homology class, and m=(m1,,mn)1{\vec{m}}=(m_{1},\dots,m_{n})\in\mathbb{Z}_{\geq 1} a tuple such that s=1nms=c1(A)+n2\sum_{s=1}^{n}m_{s}=c_{1}(A)+n-2. For Q2NQ^{2N} another closed symplectic manifold, we have

NM×Q,A×[pt]<𝒞(m,𝟙)pt>=NM,A<𝒞mpt>,\displaystyle N_{M\times Q,A\times[{\operatorname{pt}}]}\Langle\mathcal{C}^{({\vec{m}},\mathbb{1})}{\operatorname{pt}}\Rangle=N_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle, (3.6.1)

provided that the moduli spaces underlying both sides of (3.6.1) are robust and deformation invariant.

Here (m,𝟙)({\vec{m}},\mathbb{1}) denotes the tuple (m1,,mn,1,,1N)(m_{1},\dots,m_{n},\underbrace{1,\dots,1}_{N}) given by padding m{\vec{m}} with ones. Together with the equivalence from §3.4, this imples Theorem 2.6.1.

To prove Theorem 3.6.1, observe that, by the deformation invariance assumption, we can assume without loss of generality that the symplectic form on QQ is integral. Then by Donaldson’s theorem we can find a full flag of smooth symplectic divisors:

{y0}=Q0Q1QN=Q,\displaystyle\{y_{0}\}=Q_{0}\subset Q_{1}\subset\cdots\subset Q_{N}=Q,

where dimQi=2i\dim Q_{i}=2i for i=0,,Ni=0,\dots,N. Here y0y_{0} is simply a point in QQ, and we pick also a preferred basepoint x0Mx_{0}\in M.

Let 𝐃=(𝐃1,,𝐃n+N){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\dots,\mathbf{D}_{n+N}) be a collection of spanning smooth local symplectic divisors at the point (x0,y0)M×Q(x_{0},y_{0})\in M\times Q, such that 𝐃1,,𝐃n\mathbf{D}_{1},\dots,\mathbf{D}_{n} are tangent to M×{y0}M\times\{y_{0}\}, and 𝐃n+i\mathbf{D}_{n+i} is tangent to M×QiM\times Q_{i} for i=1,,Ni=1,\dots,N. As a shorthand, we put 𝐃i=(𝐃1,,𝐃n+i){\vec{\mathbf{D}}}_{i}=(\mathbf{D}_{1},\dots,\mathbf{D}_{n+i}) for i=0,,Ni=0,\dots,N.

Choose an admissible almost complex structure JN𝒥(M×Q,𝐃)J_{N}\in\mathcal{J}(M\times Q,{\vec{\mathbf{D}}}) which preserves M×QiM\times Q_{i} for i=0,,Ni=0,\dots,N and is otherwise generic. Then for i=0,,N1i=0,\dots,N-1 we have a natural inclusion map

M×Qi,A×[pt]Ji<𝒞𝐃i(m,𝟙)(x0,y0)>M×Qi+1,A×[pt]Ji+1<𝒞𝐃i+1(m,𝟙)(x0,y0)>.\displaystyle\mathcal{M}_{M\times Q_{i},A\times[{\operatorname{pt}}]}^{J_{i}}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}_{i}}^{({\vec{m}},\mathbb{1})}(x_{0},y_{0})\Rangle\subset\mathcal{M}_{M\times Q_{i+1},A\times[{\operatorname{pt}}]}^{J_{i+1}}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}_{i+1}}^{({\vec{m}},\mathbb{1})}(x_{0},y_{0})\Rangle. (3.6.2)

Here (m,𝟙)({\vec{m}},\mathbb{1}) is shorthand for (m1,,mn,1,,1i)(m_{1},\dots,m_{n},\underbrace{1,\dots,1}_{i}) when appearing in a constraint on a space of dimension 2n+2i2n+2i.

Lemma 3.6.2.

For i=0,,N1i=0,\dots,N-1, the inclusion map (3.6.2) is a regularity-preserving bijection.

Proof.

Suppose by contradiction that there is some curve

CM×Qi+1,A×[pt]Ji+1<𝒞𝐃i+1(m,𝟙)(x0,y0)>M×Qi,A×[pt]Ji<𝒞𝐃i(m,𝟙)(x0,y0)>.\displaystyle C\in\mathcal{M}_{M\times Q_{i+1},A\times[{\operatorname{pt}}]}^{J_{i+1}}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}_{i+1}}^{({\vec{m}},\mathbb{1})}(x_{0},y_{0})\Rangle\setminus\mathcal{M}_{M\times Q_{i},A\times[{\operatorname{pt}}]}^{J_{i}}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}_{i}}^{({\vec{m}},\mathbb{1})}(x_{0},y_{0})\Rangle.

Viewing CC as a curve in M×Qi+1M\times Q_{i+1}, note that by positivity of intersections CC must have nonnegative homological intersection number with the Ji+1J_{i+1}-holomorphic divisor M×QiM×Qi+1M\times Q_{i}\subset M\times Q_{i+1}, and in fact this is positive since CC passes through the point (x0,y0)M×Qi(x_{0},y_{0})\in M\times Q_{i}. On the other hand, the classes [C]=A×[pt]H2(M×Qi+1)[C]=A\times[{\operatorname{pt}}]\in H_{2}(M\times Q_{i+1}) and [M×Qi]H2n+2i(M×Qi+1)[M\times Q_{i}]\in H_{2n+2i}(M\times Q_{i+1}) evidently have trivial homological intersection number, so this is a contradiction.

It follows that any curve CM×Qi+1,A×[pt]Ji+1<𝒞𝐃i+1(m,𝟙)(x0,y0)>C\in\mathcal{M}_{M\times Q_{i+1},A\times[{\operatorname{pt}}]}^{J_{i+1}}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}_{i+1}}^{({\vec{m}},\mathbb{1})}(x_{0},y_{0})\Rangle must be entirely contained in M×QiM\times Q_{i}, which shows that the inclusion (3.6.2) is a bijection. The fact that a regular curve in M×Qi,A×[pt]Ji<𝒞𝐃i(m,𝟙)(x0,y0)>\mathcal{M}_{M\times Q_{i},A\times[{\operatorname{pt}}]}^{J_{i}}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}_{i}}^{({\vec{m}},\mathbb{1})}(x_{0},y_{0})\Rangle is also regular in M×Qi+1,A×[pt]Ji+1<𝒞𝐃i+1(m,𝟙)(x0,y0)>\mathcal{M}_{M\times Q_{i+1},A\times[{\operatorname{pt}}]}^{J_{i+1}}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}_{i+1}}^{({\vec{m}},\mathbb{1})}(x_{0},y_{0})\Rangle follows by a straightforward adaptation of the argument in [MS5, §A]. ∎

Proof of Theorem 3.6.1.

By applying Lemma 3.6.2 NN times, we find that the natural inclusion map

M,AJ0<𝒞𝐃0mx0>M×Q,A×[pt]JN<𝒞𝐃N(m,𝟙)(x0,y0)>\displaystyle\mathcal{M}_{M,A}^{J_{0}}\Langle\mathcal{C}^{{\vec{m}}}_{{\vec{\mathbf{D}}}_{0}}x_{0}\Rangle\subset\mathcal{M}_{M\times Q,A\times[{\operatorname{pt}}]}^{J_{N}}\Langle\mathcal{C}^{({\vec{m}},\mathbb{1})}_{{\vec{\mathbf{D}}}_{N}}(x_{0},y_{0})\Rangle

is a regularity-preserving bijection, whence we have

NM×Q,A×[pt]<𝒞(m,𝟙)pt>\displaystyle N_{M\times Q,A\times[{\operatorname{pt}}]}\Langle\mathcal{C}^{({\vec{m}},\mathbb{1})}{\operatorname{pt}}\Rangle =#M×Q,A×[pt]JN<𝒞𝐃N(m,𝟙)(x0,y0)>\displaystyle=\#\mathcal{M}^{J_{N}}_{M\times Q,A\times[{\operatorname{pt}}]}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}_{N}}^{({\vec{m}},\mathbb{1})}(x_{0},y_{0})\Rangle
=#M,AJ0<𝒞𝐃0mx0>=NM,A<𝒞mpt>.\displaystyle=\#\mathcal{M}^{J_{0}}_{M,A}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}_{0}}^{{\vec{m}}}x_{0}\Rangle=N_{M,A}\Langle\mathcal{C}^{\vec{m}}{\operatorname{pt}}\Rangle.

This completes the proof. ∎

4. From perfect exceptional classes to unicuspidal curves

We first discuss resolution of singularities in §4.1 from the point of view of the box diagram. In §4.2 we elaborate on the role of the local divisor in guiding the resolution, with a view towards enumerative problems. In §4.3 we prove a correspondence theorem relating multidirectional tangency counts with relative Gromov–Witten invariants. In §4.4 to prove Theorem G by a degeneration argument involving generic and nongeneric almost complex structures on blowups. Finally, in §4.5 we apply these techniques to classify rigid unicuspidal curves in the first Hirzebruch surface.

4.1. Resolution of singularities and the box diagram

We begin by discussing some aspects of embedded resolution of singularities for curve singularities. In brief, blowing up a (p,q)(p,q) cusp results in a (pq,q)(p-q,q) cusp, and by repeatedly blowing up we arrive at a smooth resolution C~\widetilde{C}. The numerics of the resulting chain of divisors can be neatly encoded in a device called the “box diagram”, which also manifests connections with ellipsoidal ends.

In more detail, let M4M^{4} be a closed symplectic four-manifold and J𝒥(M)J\in\mathcal{J}(M) a tame almost complex structure which is integrable near x0Mx_{0}\in M. Let CMC\subset M be a JJ-holomorphic curve with a (p,q)(p,q) cusp singularity at x0x_{0} for some p>q2p>q\geq 2 relatively prime. We denote by Bl1M{{\operatorname{Bl}}}^{1}M the (complex analytic) blowup of MM at x0x_{0}, and by C1C^{1} the proper transform of CC. Let 𝐃1,𝐃2\mathbf{D}_{1},\mathbf{D}_{2} be smooth local JJ-holomorphic divisors in MM which intersect transversely at x0x_{0} such that CC strictly satisfies the constraint <𝒞(𝐃1,𝐃2)(p,q)x0>\Langle\mathcal{C}_{(\mathbf{D}_{1},\mathbf{D}_{2})}^{(p,q)}x_{0}\Rangle (these exist by Lemma 3.5.3(b)). Since CC has multiplicity qq at x0x_{0}, C1C^{1} has contact order pqp-q with the proper transform 𝐃11Bl1M\mathbf{D}_{1}^{1}\subset{{\operatorname{Bl}}}^{1}M of 𝐃1\mathbf{D}_{1}, while C1C^{1} is disjoint from the proper transform of 𝐃2\mathbf{D}_{2}. We take 𝐃21:=𝔼Bl1M\mathbf{D}_{2}^{1}:=\mathbb{E}\subset{{\operatorname{Bl}}}^{1}M to be the exceptional divisor resulting from the blowup Bl1MM{{\operatorname{Bl}}}^{1}M\rightarrow M. Then C1C^{1} strictly satisfies the constraint <𝒞(𝐃11,𝐃21)(p1,q1)x1>\Langle\mathcal{C}_{(\mathbf{D}_{1}^{1},\mathbf{D}_{2}^{1})}^{(p^{1},q^{1})}x_{1}\Rangle, where we put x1:=𝐃11𝐃21x_{1}:=\mathbf{D}_{1}^{1}\cap\mathbf{D}_{2}^{1} and (p1,q1):=(pq,q)(p^{1},q^{1}):=(p-q,q). After possibly swapping 𝐃11\mathbf{D}_{1}^{1} and 𝐃21\mathbf{D}_{2}^{1}, we can assume that p1q1p^{1}\geq q^{1}.

Remark 4.1.1.

Note that {xq=yp}2\{x^{q}=y^{p}\}\subset\mathbb{C}^{2} has the parametrization t(tp,tq)t\mapsto(t^{p},t^{q}), and its blowup at the origin has the parametrization

t([tp,tq],tp,tq)=([tpq:1],tp,tq)=([x(t):1],x(t)y(t),y(t)),t\mapsto\bigl{(}[t^{p},t^{q}],t^{p},t^{q}\bigr{)}=\bigl{(}[t^{p-q}:1],t^{p},t^{q}\bigr{)}=\bigl{(}[x(t):1],x(t)y(t),y(t)\bigr{)},

where (x(t),y(t))=(tpq,tq)(x(t),y(t))=(t^{p-q},t^{q}), i.e. the blowup has a (pq,q)(p-q,q) cusp at ([0:1],0,0)([0:1],0,0). \Diamond

We proceed by blowing up Bl1M{{\operatorname{Bl}}}^{1}M at x1x_{1}, giving a proper transform C2Bl2MC^{2}\subset{{\operatorname{Bl}}}^{2}M which strictly satisfies <𝒞𝐃12,𝐃22(p2,q2)x2>\Langle\mathcal{C}_{\mathbf{D}_{1}^{2},\mathbf{D}_{2}^{2}}^{(p^{2},q^{2})}x_{2}\Rangle, with x2:=𝐃12𝐃22x_{2}:=\mathbf{D}_{1}^{2}\cap\mathbf{D}_{2}^{2} and (p2,q2):=(p1q1,q1)(p^{2},q^{2}):=(p^{1}-q^{1},q^{1}). Continuing in this manner, we eventually arrive, after say K1K\in\mathbb{Z}_{\geq 1} blowups, at the minimal resolution CKBlKMC^{K}\subset{{\operatorname{Bl}}}^{K}M, which is smooth. After some additional blowups, say LL overall, we arrive at the normal crossings resolution CLBlLMC^{L}\subset{{\operatorname{Bl}}}^{L}M, which further satisfies that the total transform of CC in BlLM{{\operatorname{Bl}}}^{L}M (i.e. the preimage of CC under the blowup map BlLMM{{\operatorname{Bl}}}^{L}M\rightarrow M) is a normal crossings divisor.

For j=1,,Lj=1,\dots,L, let 𝔽jjBljM\mathbb{F}_{j}^{j}\subset{{\operatorname{Bl}}}^{j}M denote the exceptional divisor which is the preimage of xj1x_{j-1} under the blowup map BljMBlj1M{{\operatorname{Bl}}}^{j}M\rightarrow{{\operatorname{Bl}}}^{j-1}M, and let 𝔽jk\mathbb{F}_{j}^{k} denote its proper transform in BlkM{{\operatorname{Bl}}}^{k}M for k=j+1,,Lk=j+1,\dots,L. We also put 𝔽i:=𝔽iL\mathbb{F}_{i}:=\mathbb{F}_{i}^{L} for i=1,,Li=1,\dots,L. Note that each blowup adds a \mathbb{Z} summand to the second homology, and we have a natural identification

H2(BlLM)H2(M)e1,,eL,\displaystyle H_{2}({{\operatorname{Bl}}}^{L}M)\cong H_{2}(M)\oplus\mathbb{Z}\langle e_{1},\dots,e_{L}\rangle,

with eiei=1e_{i}\cdot e_{i}=-1 and Aei=ejei=0A\cdot e_{i}=e_{j}\cdot e_{i}=0 for all 1i<jL1\leq i<j\leq L and AH2(M)A\in H_{2}(M).

In the study of symplectic embeddings of four-dimensional ellipsoids E(q,p)E(q,p), a key role is played by the weight sequence 𝒲(p,q){\mathcal{W}}(p,q) [MS2, §1], which is a sequence of positive integers associated to (p,q)(p,q). The following pictorial perspective will be useful:

Definition 4.1.2.

Given p,q1p,q\in\mathbb{Z}_{\geq 1} relatively prime, the box diagram Box(p,q){\operatorname{Box}}(p,q) is the unique decomposition into squares of the rectangle Rect(p,q){{\operatorname{Rect}}}(p,q) with horizontonal length pp and vertical length qq, subject to the following rules:

  1. (1)

    for p=q=1p=q=1, we have Box(1,1)=Rect(1,1){\operatorname{Box}}(1,1)={{\operatorname{Rect}}}(1,1) (i.e. we take the trivial decomposition)

  2. (2)

    if p>qp>q, then Box(p,q){\operatorname{Box}}(p,q) consists of the square Rect(q,q){{\operatorname{Rect}}}(q,q) flush with the left side of Rect(p,q){{\operatorname{Rect}}}(p,q), along with the decomposition of the remainder according to Box(pq,q){\operatorname{Box}}(p-q,q)

  3. (3)

    if q>pq>p, then Box(p,q){\operatorname{Box}}(p,q) consists of the square Rect(p,p){{\operatorname{Rect}}}(p,p) flush with the bottom side of Rect(p,q){{\operatorname{Rect}}}(p,q), along with the decomposition of the remainder according to Box(p,qp){\operatorname{Box}}(p,q-p).

Refer to caption
Figure 2. Below: the box diagram Box(51,23){\operatorname{Box}}(51,23). Above: the spheres 𝔽1,,𝔽L\mathbb{F}_{1},\dots,\mathbb{F}_{L} (labeled by self-intersection numbers) arising in the resolution of the (51,23)(51,23) cusp singularity.

The squares in Box(p,q){\operatorname{Box}}(p,q) are totally ordered by the rule that a square comes before any other square which lies to its right or above it, and we view the iith square as representing 𝔽i\mathbb{F}_{i} for i=1,,Li=1,\dots,L. The weight sequence of (p,q)(p,q) is then 𝒲(p,q):=(m1,,mL){\mathcal{W}}(p,q):=(m_{1},\dots,m_{L}), where mim_{i} is the side length of the iith square in Box(p,q){\operatorname{Box}}(p,q) for i=1,,Li=1,\dots,L. Note that each square in Box(p,q){\operatorname{Box}}(p,q) except for the last one touches either the top side or the right side of Rect(p,q){{\operatorname{Rect}}}(p,q), and we denote the corresponding spheres by 𝔽1hor,,𝔽Lhorhor\mathbb{F}_{1}^{\operatorname{hor}},\dots,\mathbb{F}_{L_{\operatorname{hor}}}^{\operatorname{hor}} and 𝔽1ver,,𝔽Lverver\mathbb{F}_{1}^{\operatorname{ver}},\dots,\mathbb{F}_{L_{\operatorname{ver}}}^{\operatorname{ver}} respectively. Thus the preimage of x0x_{0} under the blowup map BlLMM{{\operatorname{Bl}}}^{L}M\rightarrow M is

𝔽1𝔽L=(𝔽1hor𝔽Lhorhor)(𝔽1ver𝔽Lverver)𝔽L.\mathbb{F}_{1}\cup\cdots\cup\mathbb{F}_{L}=(\mathbb{F}_{1}^{\operatorname{hor}}\cup\cdots\cup\mathbb{F}_{L_{\operatorname{hor}}}^{\operatorname{hor}})\cup\left(\mathbb{F}_{1}^{\operatorname{ver}}\cup\cdots\cup\mathbb{F}_{L_{\operatorname{ver}}}^{\operatorname{ver}}\right)\cup\mathbb{F}_{L}.

The following properties of the box diagram Box(p,q){\operatorname{Box}}(p,q) are elementary to verify and show that it encodes the numerics of the resolution of a (p,q)(p,q) cusp.

Lemma 4.1.3.

(see also [MS2, §A] and [McD1, Fig. 3.4])

  1. (a)

    For i=1,,li=1,\dots,l we have [𝔽i]=eici+1iei+1cLieL[\mathbb{F}_{i}]=e_{i}-c_{i+1}^{i}e_{i+1}-\cdots-c_{L}^{i}e_{L}, where cji=1c^{i}_{j}=1 if the jjth square in Box(p,q){\operatorname{Box}}(p,q) lies either immediately to the right of or immediately above the iith square, and cji=0c^{i}_{j}=0 otherwise.

  2. (b)

    For 1i<jLhor1\leq i<j\leq L_{\operatorname{hor}}, 𝔽ihor\mathbb{F}_{i}^{\operatorname{hor}} and 𝔽jhor\mathbb{F}_{j}^{\operatorname{hor}} meet in a single point if |ji|=1|j-i|=1, and 𝔽ihor𝔽jhor=\mathbb{F}_{i}^{\operatorname{hor}}\cap\mathbb{F}_{j}^{\operatorname{hor}}=\varnothing otherwise. Similarly, for 1i<jLver1\leq i<j\leq L_{\operatorname{ver}}, 𝔽iver\mathbb{F}_{i}^{\operatorname{ver}} and 𝔽jver\mathbb{F}_{j}^{\operatorname{ver}} meet in a single point if |ji|=1|j-i|=1, and 𝔽iver𝔽jver=\mathbb{F}_{i}^{\operatorname{ver}}\cap\mathbb{F}_{j}^{\operatorname{ver}}=\varnothing otherwise.

  3. (c)

    𝔽L\mathbb{F}_{L} meets each of 𝔽Lhorhor,𝔽Lverver\mathbb{F}^{\operatorname{hor}}_{L_{\operatorname{hor}}},\mathbb{F}^{\operatorname{ver}}_{L_{\operatorname{ver}}} in a single point and is disjoint from 𝔽1hor,,𝔽Lhor1hor\mathbb{F}_{1}^{\operatorname{hor}},\dots,\mathbb{F}^{\operatorname{hor}}_{L_{\operatorname{hor}}-1} and 𝔽1ver,,𝔽Lver1ver\mathbb{F}_{1}^{\operatorname{ver}},\dots,\mathbb{F}^{\operatorname{ver}}_{L_{\operatorname{ver}}-1}

  4. (d)

    The normal crossing resolution CLC^{L} intersects 𝔽L\mathbb{F}_{L} in one point and is disjoint from 𝔽1,,𝔽L1\mathbb{F}_{1},\dots,\mathbb{F}_{L-1}.

In particular, the self-intersection number [𝔽i][𝔽i][\mathbb{F}_{i}]\cdot[\mathbb{F}_{i}] is 1k-1-k, where kk is the number of squares in Box(p,q){\operatorname{Box}}(p,q) which lie immediately to the right of or immediately above the iith square, and we have [𝔽L][𝔽L]=1[\mathbb{F}_{L}]\cdot[\mathbb{F}_{L}]=-1.

Example 4.1.4.

Figure 2 shows the box diagram for (51,23)(51,23), which has weight sequence is 23,23,5,5,5,5,3,2,1,123,23,5,5,5,5,3,2,1,1. The curves 𝔽1hor,,𝔽4hor\mathbb{F}_{1}^{\operatorname{hor}},\dots,\mathbb{F}_{4}^{\operatorname{hor}} appear at the top, with respective self-intersection numbers 2,6,3,2-2,-6,-3,-2, and the curves 𝔽1ver,,𝔽5ver\mathbb{F}_{1}^{\operatorname{ver}},\dots,\mathbb{F}_{5}^{\operatorname{ver}} appear at the right, with respective self-intersection numbers 2,2,2,3,3-2,-2,-2,-3,-3. \Diamond

Remark 4.1.5.

We can also write the weight sequence of (p,q)(p,q) without repetitions in the form (w1×r1,,w×r)(w_{1}^{\times r_{1}},\dots,w_{\ell}^{\times r_{\ell}}), with w1>>ww_{1}>\cdots>w_{\ell}. Then (r1,,r)(r_{1},\dots,r_{\ell}) is precisely the continued fraction expansion coefficients of p/qp/q, i.e. we have

p/q=[r1,,r]:=r1+1r2+1+1r.\displaystyle p/q=[r_{1},\dots,r_{\ell}]:=r_{1}+\cfrac{1}{r_{2}+\cfrac{1}{\ddots+\cfrac{1}{r_{\ell}}}}.

\Diamond

4.2. The role of the local divisor

In the above description of BlLM{{\operatorname{Bl}}}^{L}M, the blowup points x0,x1,,xL1x_{0},x_{1},\dots,x_{L-1} all depend on the initial curve CC, so a priori another curve CC^{\prime} with a (p,q)(p,q) cusp at x0x_{0} would have a resolution living in a different blowup on MM, thereby complicating our counting efforts. The next lemma shows that in fact BlLM{{\operatorname{Bl}}}^{L}M depends only on a certain jet of the divisor 𝐃1\mathbf{D}_{1}. As before we put 𝒲(p,q)=(m1,,mL)=(w1×r1,,w×r){\mathcal{W}}(p,q)=(m_{1},\dots,m_{L})=(w_{1}^{\times r_{1}},\dots,w_{\ell}^{\times r_{\ell}}) and assume pqp\geq q.

Lemma 4.2.1.

Let 𝐃,𝐃M\mathbf{D},\mathbf{D}^{\prime}\subset M be smooth local JJ-holomorphic divisors passing through x0x_{0} and having the same r1r_{1}-jet at x0x_{0}. Let C,CMC,C^{\prime}\subset M be JJ-holomorphic curves, each having a (p,q)(p,q) cusp at x0x_{0}, and such that (C𝐃)x0=(C𝐃)x0=p(C\cdot\mathbf{D})_{x_{0}}=(C^{\prime}\cdot\mathbf{D}^{\prime})_{x_{0}}=p. Then the sequence of blowups of MM which achieves the normal crossing resolution of CC is the same as that for CC^{\prime}.

Proof.

We proceed by induction on p+qp+q. Here we allow the case q=1q=1, which corresponds to C,CC,C^{\prime} being smooth but having contact order pp with 𝐃,𝐃\mathbf{D},\mathbf{D}^{\prime} respectively. For the base case we suppose p=q=1p=q=1, so that CC and CC^{\prime} are both smooth and pass through x0x_{0} with multiplicity 11. In this case by convention the normal crossing resolution for both CC and CC^{\prime} is obtained by blowing up MM once at x0x_{0}.

For the inductive step, note that the first blowup for both CC and CC^{\prime} occurs at x0x_{0}. Let 𝔼\mathbb{E} denote the resulting exceptional divisor, and let C~,C~,𝐃~,𝐃~\widetilde{C},\widetilde{C}^{\prime},\widetilde{\mathbf{D}},\widetilde{\mathbf{D}}^{\prime} denote the respective proper transforms of C,C,𝐃,𝐃C,C^{\prime},\mathbf{D},\mathbf{D}^{\prime}. Since CC and CC^{\prime} are tangent at x0x_{0}, C~\widetilde{C} and C~\widetilde{C}^{\prime} intersect 𝔼\mathbb{E} at the same point, say x1x_{1}.

Suppose first that p<2qp<2q, or equivalently r1=1r_{1}=1, so that C~\widetilde{C} and C~\widetilde{C}^{\prime} both have a (q,pq)(q,p-q) cusp at x1x_{1}, and we have C~𝔼=C~𝔼=q\widetilde{C}\cdot\mathbb{E}=\widetilde{C}^{\prime}\cdot\mathbb{E}=q. Then by the inductive hypothesis, with C~1,C~2\widetilde{C}_{1},\widetilde{C}_{2} playing the respective roles of C,CC,C^{\prime} and 𝔼\mathbb{E} playing the roles of both 𝐃\mathbf{D} and 𝐃\mathbf{D}^{\prime}, the remaining blowups for C~\widetilde{C} and C~\widetilde{C}^{\prime} coincide.

Now suppose that p>2qp>2q, so that C~\widetilde{C} and C~\widetilde{C}^{\prime} each has a (pq,q)(p-q,q) cusp at x1x_{1}, and the weight sequence of (pq,q)(p-q,q) is (w1×(r11),,w×r)(w_{1}^{\times(r_{1}-1)},\dots,w_{\ell}^{\times r_{\ell}}). Since CC has multiplicity qq at x0x_{0}, we have

(C~𝐃~)x1=(C𝐃)x0q=pq,\displaystyle(\widetilde{C}\cdot\widetilde{\mathbf{D}})_{x_{1}}=(C\cdot\mathbf{D})_{x_{0}}-q=p-q,

and similarly C~𝐃~=pq\widetilde{C}^{\prime}\cdot\widetilde{\mathbf{D}}^{\prime}=p-q. Note also that 𝐃~\widetilde{\mathbf{D}} and 𝐃~\widetilde{\mathbf{D}}^{\prime} have the same (r11)(r_{1}-1)-jet at x1x_{1}. Therefore we may again apply the inductive hypothesis to conclude that the remaining blowups for C~\widetilde{C} and C~\widetilde{C}^{\prime} coincide. ∎

In light of the above lemma, the following notation is well-defined, i.e. independent of the choice of CC. As above we put 𝒲(p,q)=(m1,,mL){\mathcal{W}}(p,q)=(m_{1},\dots,m_{L}), with p>qp>q relatively prime positive integers, and let M4M^{4} be a closed symplectic manifold and J𝒥(M)J\in\mathcal{J}(M) a tame almost complex structure which is integrable near a point x0Mx_{0}\in M.

Notation 4.2.2.

Put Res(p,q)𝐃,x0(M,J):=BlLM{\operatorname{Res}}_{(p,q)}^{\mathbf{D},x_{0}}(M,J):={{\operatorname{Bl}}}^{L}M, where BlLM{{\operatorname{Bl}}}^{L}M is the iterated (complex analytic) blowup of MM which achieves the normal crossing resolution for a local JJ-holomorphic curve CC which has a (p,q)(p,q) cusp at x0x_{0} and satisfies (C𝐃)x0=p(C\cdot\mathbf{D})_{x_{0}}=p.

We will sometimes use the shorthand Res(p,q)(M):=Res(p,q)𝐃,x0(M,J){\operatorname{Res}}_{(p,q)}(M):={\operatorname{Res}}_{(p,q)}^{\mathbf{D},x_{0}}(M,J) when the data 𝐃,x0,J\mathbf{D},x_{0},J is implicit or immaterial.

Note that by construction Res(p,q)𝐃,x0(M,J){\operatorname{Res}}_{(p,q)}^{\mathbf{D},x_{0}}(M,J) inherits an almost complex structure J~\widetilde{J} which is integrable near the spheres 𝔽1,,𝔽L\mathbb{F}_{1},\dots,\mathbb{F}_{L} and preserves each of them. In fact, we can also identify Res(p,q)𝐃,x0(M,J){\operatorname{Res}}_{(p,q)}^{\mathbf{D},x_{0}}(M,J) diffeomorphically with the corresponding symplectic blowup using small Darboux balls, and we thereby equip Res(p,q)𝐃,x0(M,J){\operatorname{Res}}_{(p,q)}^{\mathbf{D},x_{0}}(M,J) with a symplectic form which tames J~\widetilde{J} (and is uniquely defined up to symplectic deformation).

4.3. Relationship with relative Gromov–Witten theory

Following Notation 4.2.2, put M~:=Res(p,q)𝐃,x0(M,J)\widetilde{M}:={\operatorname{Res}}_{(p,q)}^{\mathbf{D},x_{0}}(M,J), with almost complex structure J~\widetilde{J}. Put also A~=Am1e1mLeLH2(M~)\widetilde{A}=A-m_{1}e_{1}-\cdots-m_{L}e_{L}\in H_{2}(\widetilde{M}), where 𝒲(p,q)=(m1,,mL){\mathcal{W}}(p,q)=(m_{1},\dots,m_{L}). Note that we have A~[𝔽L]=1\widetilde{A}\cdot[\mathbb{F}_{L}]=1 and A~[𝔽L]=0\widetilde{A}\cdot[\mathbb{F}_{L}]=0 for i=1,,L1i=1,\dots,L-1, so by positivity of intersections any J~\widetilde{J}-holomorphic curve in M~\widetilde{M} in homology class A~\widetilde{A} intersects 𝔽L\mathbb{F}_{L} in one point and is disjoint from 𝔽1,,𝔽L1\mathbb{F}_{1},\dots,\mathbb{F}_{L-1}.

Put 𝐃1=𝐃\mathbf{D}_{1}=\mathbf{D}, and let 𝐃2\mathbf{D}_{2} be any smooth local JJ-holomorphic divisor in MM which passes through x0x_{0} and intersects 𝐃\mathbf{D} transversely. By the above discussion, since any JJ-holomorphic curve in MM has a proper transform in M~\widetilde{M} and conversely any J~\widetilde{J}-holomorphic curve in M~\widetilde{M} can be projected to a curve in MM, we have:

Proposition 4.3.1.

There is a natural bijective correspondence

M,AJ<𝒞(𝐃1,𝐃2)(p,q)x0>M~,A~J~.\displaystyle\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}^{(p,q)}_{(\mathbf{D}_{1},\mathbf{D}_{2})}x_{0}\Rangle\cong\mathcal{M}_{\widetilde{M},\widetilde{A}}^{\widetilde{J}}.
Corollary 4.3.2.

Let M4M^{4} be a closed symplectic four-manifold and AH2(M)A\in H_{2}(M) a homology class such that c1(A)=p+qc_{1}(A)=p+q for some p,q1p,q\in\mathbb{Z}_{\geq 1} relatively prime. Then the multidirectional tangency count NM,A<𝒞(p,q)pt>N_{M,A}\Langle\mathcal{C}^{(p,q)}{\operatorname{pt}}\Rangle agrees with the genus zero symplectic Gromov–Witten invariant of M~\widetilde{M} in homology class A~\widetilde{A} relative to the norming crossing divisor 𝔽1𝔽L\mathbb{F}_{1}\cup\cdots\cup\mathbb{F}_{L}, with specified intersection pattern as above.

Example 4.3.3.

It is important to note that the relative Gromov–Witten invariant in Corollary 4.3.2 is not generally equal to the absolute Gromov–Witten invariant NM~,A~N_{\widetilde{M},\widetilde{A}} in the same homology class (although we see in §4.4 that they do agree when AA is a perfect exceptional class, in which case both counts are equal to 11). As a simple example, N2,3[L]<𝒞(8,1)pt>N_{\mathbb{CP}^{2},3[L]}\Langle\mathcal{C}^{(8,1)}{\operatorname{pt}}\Rangle coincides with the local tangency invariant N2,3[L]<𝒯(8)pt>=4N_{\mathbb{CP}^{2},3[L]}\Langle\mathcal{T}^{(8)}{\operatorname{pt}}\Rangle=4 from [MS3], whereas the corresponding absolute Gromov–Witten invariant is NBl82,3[L]e1e8=12N_{{{\operatorname{Bl}}}^{8}\mathbb{CP}^{2},3[L]-e_{1}-\cdots-e_{8}}=12. Incidentally, the “combining constraints” formula in [MS3, §4.2] describes precisely how the 1212 curves in Bl82,3[L]e1e8J~\mathcal{M}_{{{\operatorname{Bl}}}^{8}\mathbb{CP}^{2},3[L]-e_{1}-\dots-e_{8}}^{\widetilde{J}^{\prime}} degenerate as we deform a generic J~𝒥(Bl82)\widetilde{J}^{\prime}\in\mathcal{J}({{\operatorname{Bl}}}^{8}\mathbb{CP}^{2}) to the nongeneric blowup almost complex structure J~𝒥(Bl82)\widetilde{J}\in\mathcal{J}({{\operatorname{Bl}}}^{8}\mathbb{CP}^{2}), with 44 of them landing in Bl82,3[L]e1e8J~\mathcal{M}_{{{\operatorname{Bl}}}^{8}\mathbb{CP}^{2},3[L]-e_{1}-\dots-e_{8}}^{\widetilde{J}} and the remaining 88 limiting to reducible configurations in ¯Bl82,3[L]e1e8J~\overline{\mathcal{M}}_{{{\operatorname{Bl}}}^{8}\mathbb{CP}^{2},3[L]-e_{1}-\dots-e_{8}}^{\widetilde{J}}. For a different approach to this question via curves with negative ellipsoidal ends see [CGHM, Prop.3.5.1]. \Diamond

Remark 4.3.4.

Let m1hor,,mLhorhorm_{1}^{\operatorname{hor}},\dots,m_{L_{\operatorname{hor}}}^{\operatorname{hor}} and m1ver,,mLververm_{1}^{\operatorname{ver}},\dots,m_{L_{\operatorname{ver}}}^{\operatorname{ver}} denote the self-intersection numbers of 𝔽1hor,,𝔽Lhorhor\mathbb{F}_{1}^{\operatorname{hor}},\dots,\mathbb{F}_{L_{\operatorname{hor}}}^{\operatorname{hor}} and 𝔽1ver,,𝔽Lverver\mathbb{F}_{1}^{\operatorname{ver}},\dots,\mathbb{F}_{L_{\operatorname{ver}}}^{\operatorname{ver}} respectively. As explained in [GS, §2.1], these give the negative (a.k.a. Hirzebruch–Jung) continued fraction expansions of ppq\tfrac{p}{p-q} and qqr(p,q)\tfrac{q}{q-r(p,q)} respectively, where r(p,q)r(p,q) is the remainder of the division of pp by qq. Namely, we have ppq=[m1hor,,mLhorhor]\tfrac{p}{p-q}=[m_{1}^{\operatorname{hor}},\dots,m_{L_{\operatorname{hor}}}^{\operatorname{hor}}]^{-} and qqr(p,q)=[m1ver,,mLverver]\tfrac{q}{q-r(p,q)}=[m_{1}^{\operatorname{ver}},\dots,m_{L_{\operatorname{ver}}}^{\operatorname{ver}}]^{-}, where in general we put

[c1,,ck]:=c11c211ck.\displaystyle[c_{1},\dots,c_{k}]^{-}:=c_{1}-\cfrac{1}{c_{2}-\cfrac{1}{\ddots-\cfrac{1}{c_{k}}}}.

If p/q=[c1,,ck]p/q=[c_{1},\dots,c_{k}]^{-}, then the cyclic quotient surface singularity777This is modeled on the quotient of 2\mathbb{C}^{2} by the action (z1,z2)(e2πi/pz1,e2πiq/pz2)(z_{1},z_{2})\mapsto(e^{2\pi i/p}z_{1},e^{2\pi iq/p}z_{2}) of the group of ppth roots of unity. 1p(1,q)\tfrac{1}{p}(1,q) has a resolution which introduces a chain of spheres with self-intersection numbers c1,,ck-c_{1},\dots,-c_{k}. \Diamond

Example 4.3.5.

Let 2\ell\subset\mathbb{R}^{2} denote the line passing through (q,0)(q,0) and (0,p)(0,p), and let Ω\Omega denote the set of points in the first quadrant 02\mathbb{R}_{\geq 0}^{2} which lie on or above \ell. Consider the noncompact toric surface X(q,p)X_{(q,p)} with moment map polytope Ω\Omega. Then X(q,p)X_{(q,p)} is a weighted blow up of 2\mathbb{C}^{2} at the origin, and it has two cyclic quotient singularities 1p(1,pq)\tfrac{1}{p}(1,p-q) and 1q(1,qr(p,q))\tfrac{1}{q}(1,q-r(p,q)). We can resolve both of these by toric blowups, giving a smooth surface X~(q,p)\widetilde{X}_{(q,p)} having two chains of spheres with self-intersection numbers (m1hor,,mLhorhor)(-m_{1}^{\operatorname{hor}},\dots,-m_{L_{\operatorname{hor}}}^{\operatorname{hor}}) and (m1ver,,mLverver)(-m_{1}^{\operatorname{ver}},\dots,-m_{L_{\operatorname{ver}}}^{\operatorname{ver}}) respectively. These chains are joined by the (1)(-1)-sphere in X~(q,p)\widetilde{X}_{(q,p)} which is the proper transform of the toric divisor in X(q,p)X_{(q,p)} lying over the slant edge in Ω\Omega. \Diamond

Observe that in Example 4.3.5 the resolved surface X~(q,p)\widetilde{X}_{(q,p)} bears striking resemblance to the result of embedded resolution of singularities for a curve with a (p,q)(p,q) cusp at the origin. This offers the following perspective on Theorem D. Put 𝐃1std={z1=0}\mathbf{D}_{1}^{\operatorname{std}}=\{z_{1}=0\} and 𝐃2std={z2=0}\mathbf{D}_{2}^{\operatorname{std}}=\{z_{2}=0\}. Given a local curve C2C\subset\mathbb{C}^{2} which strictly satisfies <𝒞(𝐃1std,𝐃2std)(p,q)(0,0)>\Langle\mathcal{C}_{(\mathbf{D}_{1}^{\operatorname{std}},\mathbf{D}_{2}^{\operatorname{std}})}^{(p,q)}(0,0)\Rangle, by iteratively blowing up we arrive at the normal crossing resolution C~BlL2X~(q,p)\widetilde{C}\subset{{\operatorname{Bl}}}^{L}\mathbb{C}^{2}\cong\widetilde{X}_{(q,p)}, which intersects the (1)(-1)-sphere 𝔽L\mathbb{F}_{L} in one point and is disjoint from the other spheres 𝔽1,,𝔽L1\mathbb{F}_{1},\dots,\mathbb{F}_{L-1}. Then after blowing down 𝔽1,,𝔽L1\mathbb{F}_{1},\dots,\mathbb{F}_{L-1} we get a smooth curve CC^{\prime} in X(q,p)X_{(q,p)} which is disjoint from the two orbifold points. At least heuristically, this is akin to a curve in the negative symplectic completion of 2E̊(q,p)\mathbb{C}^{2}\setminus{\mathring{E}}(q,p) with negative end asymptotic to one of the Reeb orbits in the family from Remark 2.3.9. This picture easily extends to any closed symplectic manifold M4M^{4}, noting that all relevant blowups can taken in a small neighborhood of the image of an ellipsoid embedding E(εq,εp)𝑠ME(\varepsilon q,\varepsilon p)\overset{s}{\hookrightarrow}M for ε>0\varepsilon>0 sufficiently small.

4.4. Degenerating to the nongeneric blowup

The goal of this subsection is to prove Theorem G. As before, let M4M^{4} be a closed symplectic manifold, and let J𝒥(M)J\in\mathcal{J}(M) be a tame almost complex structure which is integrable near a point x0Mx_{0}\in M. Let p>qp>q be relatively prime positive integers, and put 𝒲(p,q)=(m1,,mL){\mathcal{W}}(p,q)=(m_{1},\dots,m_{L}). Let 𝐃M\mathbf{D}\subset M be a smooth local JJ-holomorphic divisor passing through x0x_{0}, and put Res(p,q)(M):=Res(p,q)𝐃,x0(M,J){\operatorname{Res}}_{(p,q)}(M):={\operatorname{Res}}_{(p,q)}^{\mathbf{D},x_{0}}(M,J).

Recall that we have the identification

H2(Res(p,q)(M))H2(M)e1,,eL.\displaystyle H_{2}({\operatorname{Res}}_{(p,q)}(M))\cong H_{2}(M)\oplus\mathbb{Z}\langle e_{1},\dots,e_{L}\rangle.

In particular, note that H2(Res(p,q)(M))H_{2}({\operatorname{Res}}_{(p,q)}(M)) depends only on H2(M)H_{2}(M) and (p,q)(p,q) and not on the data J,𝐃,x0J,\mathbf{D},x_{0}. A general element in H2(Res(p,q)(M))H_{2}({\operatorname{Res}}_{(p,q)}(M)) takes the form A=Bk1e1kLeLA=B-k_{1}e_{1}-\cdots-k_{L}e_{L} for some k1,,kLk_{1},\dots,k_{L}\in\mathbb{Z}, with first Chern number c1(A)=c1(B)k1kLc_{1}(A)=c_{1}(B)-k_{1}-\cdots-k_{L} and self-intersection number AA=BBk12kL2A\cdot A=B\cdot B-k_{1}^{2}-\cdots-k_{L}^{2}.

Definition 4.4.1.

A homology class BH2(M)B\in H_{2}(M) is exceptional if it satisfies c1(B)=1c_{1}(B)=1 and BB=1B\cdot B=-1, and BB is represented by a symplectically embedded two-sphere in MM.

A priori the last condition in Definition 4.4.1 is nontrivial to check, but for blowups of 2\mathbb{CP}^{2} there is a purely combinatorial characterization in terms of elementary Cremona transformations (see [MS2, Prop. 1.2.12]).

Definition 4.4.2.

A homology class AH2(M)A\in H_{2}(M) is (𝐩,𝐪)\mathbf{(p,q)}-perfect exceptional888This usage differs slightly from other sources. if A~:=Am1e1mLeLH2(Res(p,q)(M))\widetilde{A}:=A-m_{1}e_{1}-\cdots-m_{L}e_{L}\in H_{2}({\operatorname{Res}}_{(p,q)}(M)) is an exceptional class.

Note that a (p,q)(p,q)-perfect exceptional class AA satisfies c1(A)=p+qc_{1}(A)=p+q.

A standard fact about exceptional homology classes BH2(M)B\in H_{2}(M) is that the corresponding Gromov–Witten invariant NM,BN_{M,B} is equal to 11.

Lemma 4.4.3.

Let M4M^{4} be a closed symplectic four-manifold, and suppose that BH2(M)B\in H_{2}(M) is an exceptional class. Then we have NM,B=1N_{M,B}=1, and in particular ¯M,BJ\overline{\mathcal{M}}_{M,B}^{J}\neq\varnothing for any J𝒥(M)J\in\mathcal{J}(M).

The basic idea is that using BB=1B\cdot B=-1 and positivity of intersections there cannot be two irreducible JJ-holomorphic spheres in the same exceptional class. Meanwhile, using the fact that BB has a symplectically embedded representative we can engineer a choice of almost complex structure JJ which preserves it. Also, any irreducible pseudoholomorphic sphere in an exceptional class BB is necessarily nonsingular by the adjunction formula. Note that we have ¯M,BJ=M,BJ\overline{\mathcal{M}}_{M,B}^{J}=\mathcal{M}_{M,B}^{J} for generic J𝒥(M)J\in\mathcal{J}(M) (but not necessarily for arbitrary JJ).

Observe that the almost complex structure J~\widetilde{J} on Res(p,q)(M){\operatorname{Res}}_{(p,q)}(M) is not generic, since it preserves the spheres 𝔽1,,𝔽L1\mathbb{F}_{1},\dots,\mathbb{F}_{L-1} which all have negative index. Therefore a priori an exceptional class A~H2(Res(p,q)(M))\widetilde{A}\in H_{2}({\operatorname{Res}}_{(p,q)}(M)) might be represented by a degenerate configuration, in which case it does not necessarily blow down to a nice curve in MM. The next lemma, which is the final ingredient to prove Theorem G, shows that this cannot occur provided that JJ is generic away from x0x_{0}.

Lemma 4.4.4.

Let M4M^{4} be a closed symplectic four-manifold, and let AH2(M)A\in H_{2}(M) be an (p,q)(p,q)-perfect exceptional class for some p>qp>q relatively prime. Let 𝐃\mathbf{D} be smooth local symplectic divisor passing through x0Mx_{0}\in M, and suppose that J𝒥(M)J\in\mathcal{J}(M) is integrable near x0x_{0}, preserves 𝐃\mathbf{D}, and is otherwise generic. Put M~:=Res(p,q)𝐃,x0(M,J)\widetilde{M}:={\operatorname{Res}}_{(p,q)}^{\mathbf{D},x_{0}}(M,J) with its induced almost complex structure J~𝒥(M~)\widetilde{J}\in\mathcal{J}(\widetilde{M}), and put A~:=Am1e1mLeL\widetilde{A}:=A-m_{1}e_{1}-\cdots-m_{L}e_{L}, where 𝒲(p,q)=(m1,,mL){\mathcal{W}}(p,q)=(m_{1},\dots,m_{L}). Then we have ¯M~,A~J~=M~,A~J~\overline{\mathcal{M}}_{\widetilde{M},\widetilde{A}}^{\widetilde{J}}=\mathcal{M}_{\widetilde{M},\widetilde{A}}^{\widetilde{J}}.

Deferring the proof for the moment, we complete the proof of Theorem G.

Proof of Theorem G.

Pick any smooth local divisor divisors 𝐃=(𝐃1,𝐃2){\vec{\mathbf{D}}}=(\mathbf{D}_{1},\mathbf{D}_{2}) which span at a point x0x_{0}, and generic J𝒥(M,𝐃)J\in\mathcal{J}(M,{\vec{\mathbf{D}}}). We assume pqp\geq q without loss of generality, and put M~:=Res(p,q)𝐃1,x0(M,J)\widetilde{M}:={\operatorname{Res}}_{(p,q)}^{\mathbf{D}_{1},x_{0}}(M,J) with its induced almost complex structure J~\widetilde{J}.

Suppose first that there is an index zero rational (p,q)(p,q)-unicuspidal symplectic curve CC in MM in homology class AA. By Proposition 4.3.1 and Theorem E we have

#M~,A~J~=#M,AJ<𝒞𝐃(p,q)x0>=NM,A<𝒞(p,q)pt>>0.\#\mathcal{M}^{\widetilde{J}}_{\widetilde{M},\widetilde{A}}=\#\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}^{(p,q)}_{{\vec{\mathbf{D}}}}x_{0}\Rangle=N_{M,A}\Langle\mathcal{C}^{(p,q)}{\operatorname{pt}}\Rangle>0.

In particular, A~\widetilde{A} is represented by a symplectically embedded sphere in M~\widetilde{M}. The fact that CC has index zero translates into c1(A~)=1c_{1}(\widetilde{A})=1, while CC being (p,q)(p,q)-unicuspidal translates into A~A~=1\widetilde{A}\cdot\widetilde{A}=-1.

Conversely, suppose that AH2(M)A\in H_{2}(M) is a (p,q)(p,q)-perfect exceptional, i.e. A~=Am1e1mLeLH2(M~)\widetilde{A}=A-m_{1}e_{1}-\cdots-m_{L}e_{L}\in H_{2}(\widetilde{M}) is an exceptional class. By Lemma 4.4.4 we have ¯M~,A~J~=M~,A~J~\overline{\mathcal{M}}_{\widetilde{M},\widetilde{A}}^{\widetilde{J}}=\mathcal{M}_{\widetilde{M},\widetilde{A}}^{\widetilde{J}}, so by Lemma 4.4.3 we have

#M~,A~J~=NM~,A~=1.\displaystyle\#\mathcal{M}_{\widetilde{M},\widetilde{A}}^{\widetilde{J}}=N_{\widetilde{M},\widetilde{A}}=1.

By Proposition 4.3.1 and genericity of JJ we thus have

NM,A<𝒞(p,q)pt>=#M,AJ<𝒞𝐃(p,q)x0>=#M~,A~J~=1.\displaystyle N_{M,A}\Langle\mathcal{C}^{(p,q)}{\operatorname{pt}}\Rangle=\#\mathcal{M}_{M,A}^{J}\Langle\mathcal{C}_{{\vec{\mathbf{D}}}}^{(p,q)}x_{0}\Rangle=\#\mathcal{M}_{\widetilde{M},\widetilde{A}}^{\widetilde{J}}=1.

Finally, by Theorem E there exists a rational (p,q)(p,q)-sesquicuspidal symplectic curve in MM lying in homology class AA, and this is necessarily unicuspidal by the adjunction formula.

Remark 4.4.5.

If M4M^{4} is a smooth complex projective surface, it is not a priori clear whether the analogue of Lemma 4.4.4 holds for its prefered integrable almost complex structure, which is not necessarily generic. However, if it does, then the argument in the proof of Theorem G shows that any perfect exceptional class corresponds to an algebraic unicuspidal curve. \Diamond

Proof of Lemma 4.4.4.

As in §4.1, let 𝔽1,,𝔽L\mathbb{F}_{1},\dots,\mathbb{F}_{L} denote the J~\widetilde{J}-holomorphic spheres in M~\widetilde{M} which project to x0x_{0} under the blowup map M~M\widetilde{M}\rightarrow M, with [𝔽i]=eici+1iei+1cLieL[\mathbb{F}_{i}]=e_{i}-c^{i}_{i+1}e_{i+1}-\cdots-c_{L}^{i}e_{L} with cji{0,1}c_{j}^{i}\in\{0,1\} for 1i<jL1\leq i<j\leq L.

Given C¯M~,AJ~C\in\overline{\mathcal{M}}_{\widetilde{M},A}^{\widetilde{J}}, it suffices to show that CC is irreducible. Suppose by contradiction that CC has more than one component. Since J~\widetilde{J} is generic outside of a small neighborhood of 𝔽1𝔽L\mathbb{F}_{1}\cup\cdots\cup\mathbb{F}_{L}, by index considerations we can assume that at least one component of CC covers one of 𝔽1,,𝔽L\mathbb{F}_{1},\dots,\mathbb{F}_{L}. Put C=C1C2C=C_{1}\cup C_{2}, where no component of C1C_{1} is contained in 𝔽1𝔽L\mathbb{F}_{1}\cup\cdots\cup\mathbb{F}_{L} and each component of C2C_{2} is contained in 𝔽1𝔽L\mathbb{F}_{1}\cup\cdots\cup\mathbb{F}_{L}, with C1,C2C_{1},C_{2} both nonempty.

Claim 4.4.6.

We have [C2]A~0[C_{2}]\cdot\widetilde{A}\geq 0.

Proof.

It suffices to show that [𝔽i]A~0[\mathbb{F}_{i}]\cdot\widetilde{A}\geq 0 for i=1,,Li=1,\dots,L. Recall that we have [𝔽i]A~=0[\mathbb{F}_{i}]\cdot\widetilde{A}=0 for i=1,,L1i=1,\dots,L-1 and [𝔽L]A~=1[\mathbb{F}_{L}]\cdot\widetilde{A}=1. ∎

Let J~𝒥(M~)\widetilde{J}^{\prime}\in\mathcal{J}(\widetilde{M}) be a sufficiently small generic perturbation of J~\widetilde{J} (note that 𝔽1,,𝔽L\mathbb{F}_{1},\dots,\mathbb{F}_{L} are not J~\widetilde{J}^{\prime}-holomorphic). Then we can deform C1C_{1} to a J~\widetilde{J}^{\prime}-holomorphic curve C1C_{1}^{\prime} with [C1]=[C1]H2(M~)[C_{1}^{\prime}]=[C_{1}]\in H_{2}(\widetilde{M}). Indeed, since J~\widetilde{J} is generic outside of a small neighborhood of 𝔽1𝔽m\mathbb{F}_{1}\cup\cdots\cup\mathbb{F}_{m}, we can assume that the underlying simple curve of each component C1C_{1} is regular, and hence unobstructed under small perturbations of J~\widetilde{J}.

Moreover, since A~H2(M~)\widetilde{A}\in H_{2}(\widetilde{M}) is an exceptional class, by Lemma 4.4.3 and genericity of J~\widetilde{J}^{\prime} there exists a J~\widetilde{J}^{\prime}-holomorphic sphere CM~,A~J~C^{\prime}\in\mathcal{M}_{\widetilde{M},\widetilde{A}}^{\widetilde{J}^{\prime}}.

Claim 4.4.7.

C1C_{1}^{\prime} must intersect CC^{\prime} in isolated points, i.e. C1C_{1}^{\prime} does not have any component with the same image as CC^{\prime}.

Proof.

We will argue using Lemma 4.4.8 below that no component of C1C_{1}^{\prime} can lie in a homology class which is a nonzero multiple of A~\widetilde{A}. Firstly, note that no component of C1C_{1}^{\prime} can lie in a homology class of the form ceice_{i} for some c1c\in\mathbb{Z}_{\geq 1} and i{1,,L}i\in\{1,\dots,L\}. Indeed, by the way C1C_{1}^{\prime} was constructed we would then have a component QQ of C1C_{1} of the same form, but then we have

[Q][𝔽i]=cei(eici+1iei+1cLiei)=c,\displaystyle[Q]\cdot[\mathbb{F}_{i}]=ce_{i}\cdot(e_{i}-c_{i+1}^{i}e_{i+1}-\cdots-c_{L}^{i}e_{i})=-c,

which violates positivity of intersections.

Now suppose by contradiction that C1C_{1}^{\prime} has a component QQ^{\prime} such that [Q]=kA~[Q^{\prime}]=k\widetilde{A} for some nonzero kk\in\mathbb{Z}, and let BB denote the total homology class of the remaining components of C1C_{1}^{\prime}, so that we have [C1]=kA~+B[C_{1}^{\prime}]=k\widetilde{A}+B. Note that the natural projection map π:H2(M~)H2(M)\pi:H_{2}(\widetilde{M})\rightarrow H_{2}(M) satisfies π(A~)=A\pi(\widetilde{A})=A and π([C2])=0\pi([C_{2}])=0 and hence π([C1])=π([C1])=A\pi([C_{1}^{\prime}])=\pi([C_{1}])=A, so we have

c1(A)=c1(π([C1]))=kc1(A)+c1(π(B)).\displaystyle c_{1}(A)=c_{1}(\pi([C_{1}^{\prime}]))=kc_{1}(A)+c_{1}(\pi(B)).

By Lemma 4.4.8 we have c1(A),c1(π(B)),c1(kA)1c_{1}(A),c_{1}(\pi(B)),c_{1}(kA)\geq 1, and hence k1k\geq 1, which is a contradiction. ∎

Finally, by positivity of intersections we have [C1][C2]0[C_{1}]\cdot[C_{2}]\geq 0 and [C1][C]0[C_{1}^{\prime}]\cdot[C^{\prime}]\geq 0, and hence

0[C1][C]=(A~[C2])A~=1[C2]A~1,\displaystyle 0\leq[C_{1}^{\prime}]\cdot[C^{\prime}]=(\widetilde{A}-[C_{2}])\cdot\widetilde{A}=-1-[C_{2}]\cdot\widetilde{A}\leq-1,

which is a contradiction. ∎

Lemma 4.4.8.

Any J~\widetilde{J}^{\prime}-holomorphic curve in M~\widetilde{M} represents a homology class which is either of the form (i) ceice_{i} for some c1c\in\mathbb{Z}_{\geq 1} and i{1,,L}i\in\{1,\dots,L\}, or else it is of the form (ii) Ak1e1kLeLA^{\prime}-k_{1}e_{1}-\cdots-k_{L}e_{L}, where k1,,kL0k_{1},\dots,k_{L}\in\mathbb{Z}_{\geq 0} and AH2(M)A^{\prime}\in H_{2}(M) satisfies c1(A)1c_{1}(A^{\prime})\geq 1.

Proof.

Let C0C_{0} be a J~\widetilde{J}^{\prime}-holomorphic curve in M~\widetilde{M}, and put [C0]=Ak1e1kLeL[C_{0}]=A^{\prime}-k_{1}e_{1}-\cdots-k_{L}e_{L} for some AH2(M)A^{\prime}\in H_{2}(M) and k1,,kLk_{1},\dots,k_{L}\in\mathbb{Z}. Evidently it suffices to prove the lemma in the case that C0C_{0} is simple, whence ind(C0)0{\operatorname{ind}}(C_{0})\geq 0. Since J~\widetilde{J}^{\prime} is generic and e1,,enH2(M~)e_{1},\dots,e_{n}\in H_{2}(\widetilde{M}) are exceptional classes, there is a curve component 𝔼iM~,eiJ~\mathbb{E}_{i}\in\mathcal{M}_{\widetilde{M},e_{i}}^{\widetilde{J}^{\prime}} for i=1,,Li=1,\dots,L. If C0C_{0} does not cover any of 𝔼1,,𝔼L\mathbb{E}_{1},\dots,\mathbb{E}_{L} then by positivity of intersections we have 0[C0][𝔼i]=ki0\leq[C_{0}]\cdot[\mathbb{E}_{i}]=k_{i} for i=1,,Li=1,\dots,L, so we have

0ind(C0)=2+2c1(A)2k12kL,\displaystyle 0\leq{\operatorname{ind}}(C_{0})=-2+2c_{1}(A^{\prime})-2k_{1}-\cdots-2k_{L},

and hence c1(A)1+k1++kL1c_{1}(A^{\prime})\geq 1+k_{1}+\cdots+k_{L}\geq 1. ∎

4.5. The case of the first Hirzebruch surface

Let F1F_{1} denote the first Hirzebruch surface, i.e. the one-point blowup of 2\mathbb{CP}^{2}. We can also view this as the toric Kähler manifold whose moment polygon has vertices (0,0),(3,0),(1,2),(0,2)(0,0),(3,0),(1,2),(0,2). This choice of normalization makes the symplectic form monotone, although in light of Corollary 3.5.5 we will only be concerned with the symplectic form up to symplectic deformation.

Let =[L]H2(F1)\ell=[L]\in H_{2}(F_{1}) denote the line class and e=[𝔼]H2(F1)e=[\mathbb{E}]\in H_{2}(F_{1}) the exceptional class, so that any AH2(F1)A\in H_{2}(F_{1}) takes the form dmed\ell-me for some d,md,m\in\mathbb{Z}. By Theorem G, index zero rational unicuspidal symplectic curves in F1F_{1} are in bijective correspondence with perfect exceptional classes in H2(F1)H_{2}(F_{1}). Putting

Perf(F1):={(p,q,d,m)|A=dmeis a(p,q)-perfect exceptional class},\displaystyle{\operatorname{Perf}}(F_{1}):=\{(p,q,d,m)\;|\;A=d\ell-me\;\text{is a}\;\text{$(p,q)$-perfect exceptional class}\},

we can reformulate this as:

Corollary 4.5.1.

There is an index zero (p,q)(p,q)-unicuspidal symplectic curve in F1F_{1} in homology class A=dmeA=d\ell-me if and only if we have (p,q,d,m)Perf(F1)(p,q,d,m)\in{\operatorname{Perf}}(F_{1}).

Put also

Perf¯(F1):={(p,q)|(p,q,d,m)Perf(F1) for some d,m}.\displaystyle\underline{{\operatorname{Perf}}}(F_{1}):=\{(p,q)\;|\;(p,q,d,m)\in{\operatorname{Perf}}(F_{1})\text{ for some }d,m\}.

A comprehensive description of Perf(F1){\operatorname{Perf}}(F_{1}) is given in [MMW], so together with Corollary 4.5.1 this describes all index zero (p,q)(p,q)-unicuspidal rational symplectic curves in F1F_{1}. Here we make just a few remarks about Perf(F1){\operatorname{Perf}}(F_{1}), referring the reader to loc. cit. for more details.

Theorem 4.5.2 ([MM, MMW]).
  • The forgetful map Perf(F1)Perf¯(F1){\operatorname{Perf}}(F_{1})\rightarrow\underline{{\operatorname{Perf}}}(F_{1}) is injective, i.e. for each (p,q)(p,q) there is at most one (p,q,d,m)Perf(F)(p,q,d,m)\in{\operatorname{Perf}}(F). More precisely, for any (p,q,d,m)Perf(F1)(p,q,d,m)\in{\operatorname{Perf}}(F_{1}) we must have d=18(3p+3q+εt)d=\tfrac{1}{8}(3p+3q+\varepsilon t) and m=18(p+q+3εt)m=\tfrac{1}{8}(p+q+3\varepsilon t), where t=p26pq+q2+8t=\sqrt{p^{2}-6pq+q^{2}+8} and ε{1,1}\varepsilon\in\{-1,1\} (and one can check that d,md,m are integers for at most one value of ε\varepsilon).

  • If (p,q)Perf¯(F1)(p,q)\in\underline{{\operatorname{Perf}}}(F_{1}), then we also have S(p,q):=(6pq,p)Perf¯(F1)S(p,q):=(6p-q,p)\in\underline{{\operatorname{Perf}}}(F_{1}).

  • If (p,q)Perf¯(F1)(p,q)\in\underline{{\operatorname{Perf}}}(F_{1}) with p/q>6p/q>6, then we also have R(p,q):=(6p35q,p6q)Perf¯(F1)R(p,q):=(6p-35q,p-6q)\in\underline{{\operatorname{Perf}}}(F_{1}).

Remark 4.5.3.

As we show in [MS4], the symmetry S(p,q)S(p,q) has a natural geometric interpretation in terms of the generalized Orevkov twist, which is a birational transformation F1F1F_{1}\dashrightarrow F_{1}. We currently do not of any geometric interpretation of the symmetry R(p,q)R(p,q). It is not currently clear whether every index zero rational (p,q)(p,q)-cuspidal symplectic curve in F1F_{1} is realized by an algebraic curve. \Diamond

References

  • [Bar] Jean-François Barraud. Courbes pseudo-holomorphes équisingulieres en dimension 4. Bulletin de la Société Mathématique de France 128(2000), 179–206.
  • [BEH+] Frédéric Bourgeois, Yakov Eliashberg, Helmut Hofer, Krzysztof Wysocki, and Eduard Zehnder. Compactness results in symplectic field theory. Geom. Topol. 7(2003), 799–888.
  • [CH] Lucia Caporaso and Joe Harris. Counting plane curves of any genus. Invent. Math. 131(1998), 345–392.
  • [CM1] Kai Cieliebak and Klaus Mohnke. Symplectic hypersurfaces and transversality in Gromov-Witten theory. J. Symplectic Geom. 5(2007), 281–356.
  • [CM2] Kai Cieliebak and Klaus Mohnke. Punctured holomorphic curves and Lagrangian embeddings. Invent. Math. 212(2018), 213–295.
  • [CGHMP] Dan Cristofaro-Gardiner, Tara S Holm, Alessia Mandini, and Ana Rita Pires. On infinite staircases in toric symplectic four-manifolds. arXiv:2004.13062 (2020).
  • [CGH] Daniel Cristofaro-Gardiner and Richard Hind. Symplectic embeddings of products. Comment. Math. Helv. 93(2018), 1–32.
  • [CGHM] Daniel Cristofaro-Gardiner, Richard Hind, and Dusa McDuff. The ghost stairs stabilize to sharp symplectic embedding obstructions. J. Topol. 11(2018), 309–378.
  • [EN] David Eisenbud and Walter D Neumann. Three-dimensional link theory and invariants of plane curve singularities. Princeton University Press, 1985.
  • [EG] John B Etnyre and Marco Golla. Symplectic hats. arXiv:2001.08978 (2020).
  • [FdBLMHN] Javier Fernández de Bobadilla, Ignacio Luengo, Alejandro Melle Hernández, and Andras Némethi. Classification of rational unicuspidal projective curves whose singularities have one Puiseux pair. pages 31–45. Springer, 2006.
  • [GS] Marco Golla and Laura Starkston. The symplectic isotopy problem for rational cuspidal curves. Compositio Mathematica 158(2022), 1595–1682.
  • [GH] Jean Gutt and Michael Hutchings. Symplectic capacities from positive S1S^{1}-equivariant symplectic homology. Algebr. Geom. Topol. 18(2018), 3537–3600.
  • [HK] Richard Hind and Ely Kerman. New obstructions to symplectic embeddings. Invent. Math. 196(2014), 383–452.
  • [Hut1] Michael Hutchings. Quantitative embedded contact homology. J. Differential Geom. 88(2011), 231–266.
  • [Hut2] Michael Hutchings. Lecture notes on embedded contact homology. pages 389–484. Springer, 2014.
  • [HT] Michael Hutchings and Clifford Henry Taubes. Gluing pseudoholomorphic curves along branched covered cylinders. I. J. Symplectic Geom. 5(2007), 43–137.
  • [KV] Joachim Kock and Israel Vainsencher. An invitation to quantum cohomology: Kontsevich’s formula for rational plane curves. Springer Science & Business Media, 2007.
  • [KM] Maxim Kontsevich and Yu Manin. Gromov-Witten classes, quantum cohomology, and enumerative geometry. Communications in Mathematical Physics 164(1994), 525–562.
  • [MM] Nicki Magill and Dusa McDuff. Staircase symmetries in Hirzebruch surfaces. Algebr. Geom. Topol. (To appear).
  • [MMW] Nicki Magill, Dusa McDuff, and Morgan Weiler. Staircase patterns in Hirzebruch surfaces. arXiv:2203.06453 (2022).
  • [McD1] Dusa McDuff. Symplectic embeddings of 4-dimensional ellipsoids. Journal of Topology 2(2009), 1–22.
  • [McD2] Dusa McDuff. A remark on the stabilized symplectic embedding problem for ellipsoids. Eur. J. Math. 4(2018), 356–371.
  • [MS1] Dusa McDuff and Dietmar Salamon. J-holomorphic curves and symplectic topology. American Mathematical Soc., 2012.
  • [MS2] Dusa McDuff and Felix Schlenk. The embedding capacity of 4-dimensional symplectic ellipsoids. Ann. of Math. (2) 175(2012), 1191–1282.
  • [MS3] Dusa McDuff and Kyler Siegel. Counting curves with local tangency constraints. Journal of Topology 14(2021), 1176–1242.
  • [MS4] Dusa McDuff and Kyler Siegel. Rational cuspidal curves and symplectic staircases. (In preparation).
  • [MS5] Dusa McDuff and Kyler Siegel. Symplectic capacities, unperturbed curves, and convex toric domains. Geometry and topology (To appear).
  • [MS6] Grigory Mikhalkin and Kyler Siegel. Ellipsoidal superpotentials and stationary descendants. arXiv:2307.13252 (2023).
  • [Neu] Walter D Neumann. Topology of hypersurface singularities. arXiv:1706.04386 (2017).
  • [Ore] S Yu Orevkov. On rational cuspidal curves: I. Sharp estimate for degree via multiplicities. Mathematische Annalen 324(2002), 657–673.
  • [Per] Miguel Pereira. Equivariant symplectic homology, linearized contact homology and the Lagrangian capacity. arXiv:2205.13381 (2022).
  • [Sch] Felix Schlenk. Symplectic embedding problems, old and new. Bull. Amer. Math. Soc. (N.S.) 55(2018), 139–182.
  • [Ton] Dmitry Tonkonog. String topology with gravitational descendants, and periods of Landau-Ginzburg potentials. arXiv:1801.06921 (2018).
  • [Wen1] Chris Wendl. Automatic transversality and orbifolds of punctured holomorphic curves in dimension four. Comment. Math. Helv. 85(2010), 347–407.
  • [Wen2] Chris Wendl. Lectures on Symplectic Field Theory. https://www.mathematik.hu-berlin.de/~wendl/Sommer2020/SFT/lecturenotes.pdf, 2020.