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

Unknottedness of real Lagrangian tori in S2×S2S^{2}\times S^{2}

Joontae Kim School of Mathematics, Korea Institute for Advanced Study, 85 Hoegiro, Dongdaemun-gu, Seoul 02455, Republic of Korea, joontae@kias.re.kr
(Recent modification; )
Abstract

We prove the Hamiltonian unknottedness of real Lagrangian tori in the monotone S2×S2S^{2}\times S^{2}, namely any real Lagrangian torus in S2×S2S^{2}\times S^{2} is Hamiltonian isotopic to the Clifford torus. The proof is based on a neck-stretching argument, Gromov’s foliation theorem, and the Cieliebak–Schwingenheuer criterion.

Mathematics Subject Classification (2000) 53D12, 53D35, 54H25

1 Introduction

An even dimensional smooth manifold MM equipped with a closed non-degenerate 2-form ω\omega is a symplectic manifold. By Darboux’s theorem [34, Theorem 3.2.2], symplectic manifolds are locally standard, and hence only global properties in symplectic topology are interesting; in particular, the study of middle dimensional submanifolds along which the symplectic form vanishes, namely Lagrangian submanifolds.

In 1986, as one of the first steps in symplectic topology [1, Section 6], Arnold proposed the Lagrangian knot problem asking whether two given Lagrangians are isotopic. As formulated in a survey of Eliashberg–Polterovich [22], there are different flavors of isotopy, namely smooth, Lagrangian and Hamiltonian. Hamiltonian isotopies are Lagrangian, and Lagrangian isotopies are smooth. Two Lagrangians are said to be unknotted if they are isotopic to each other in one of these three ways.

A remarkable result of Gromov [25] says that there are no closed exact Lagrangians in (2n,i=1ndxidyi)({\mathbb{R}}^{2n},\sum_{i=1}^{n}dx_{i}\wedge dy_{i}), and hence the extensive study of Lagrangian tori in 2n{\mathbb{R}}^{2n} has been made for a long time. Chekanov [9] first constructed a monotone Lagrangian torus in 2n{\mathbb{R}}^{2n} for n2n\geq 2, which is Lagrangian isotopic while not Hamiltonian isotopic to the Clifford torus 𝕋Clifn=×nS1\mathbb{T}^{n}_{\operatorname{Clif}}=\times_{n}S^{1}, i.e., products of circles in 2{\mathbb{R}}^{2} of equal radius. This exhibits that a Lagrangian isotopy cannot be in general deformed into a Hamiltonian isotopy. His result is even more interesting since the classical invariants cannot detect this phenomenon. Indeed, the Audin conjecture [2, Section 6.4], which is proved by Polterovich [37] and Viterbo [40] in dimension 4 and Cieliebak–Mohnke [11] in any dimension, says that the minimal Maslov number (one of the classical invariants) of any Lagrangian torus in 2n{\mathbb{R}}^{2n} is always two, so exotic monotone Lagrangian tori are hard to discover. Auroux [3] constructed infinitely many monotone Lagrangian tori in 6{\mathbb{R}}^{6} up to Hamiltonian isotopy, while all of them are Lagrangian isotopic. We refer to the work of Dimitroglou Rizell–Evans [17, Corollary C] about the smooth unknottedness of monotone Lagrangian tori inside 2n{\mathbb{R}}^{2n} for n5n\geq 5 odd.

Since Lagrangian 2-planes in 4{\mathbb{R}}^{4} that are asymptotically linear are trivial by Eliashberg–Polterovich [21] (see also [20]), one may expect reasonable unknottedness results in symplectic 4-manifolds. Symplectic Field Theory developed by Eliashberg–Givental–Hofer [19] has become a core technique to address the Lagrangian unknottedness problems. By a neck-stretching argument [5], Hind [26] showed the Hamiltonian unknottedness of spheres in S2×S2S^{2}\times S^{2}.

In contrast to spheres, the Hamiltonian unknottedness of monotone Lagrangian tori in S2×S2S^{2}\times S^{2} fails. Chekanov–Schlenk [8] found a monotone Lagrangian torus in S2×S2S^{2}\times S^{2} which is not Hamiltonian isotopic to the Clifford torus 𝕋Clif=S1×S1\mathbb{T}_{\operatorname{Clif}}=S^{1}\times S^{1}, defined as the product of the equators. This torus, in another description, was also found by Entov–Polterovich [23, Example 1.22]. Vianna [39] showed that there are infinitely many Hamiltonian isotopy classes of monotone Lagrangian tori in S2×S2S^{2}\times S^{2}. Dimitroglou Rizell–Goodman–Ivrii [18, Theorem A] established the Lagrangian unknottedness of tori in S2×S2S^{2}\times S^{2}, namely Lagrangian tori in S2×S2S^{2}\times S^{2} are unique up to Lagrangian isotopy. As a result, infinitely many monotone Lagrangian tori in S2×S2S^{2}\times S^{2} are Hamiltonianly knotted.

In this paper, we are interested in a class more rigid than monotone Lagrangian tori, namely real Lagrangian tori. By a real Lagrangian submanifold LL in a symplectic manifold (M,ω)(M,\omega) we mean a Lagrangian submanifold that is the fixed point set of an antisymplectic involution RR of MM, i.e., R2=idMR^{2}=\operatorname{id}_{M} and Rω=ωR^{*}\omega=-\omega. The fixed point set Fix(R)={xMR(x)=x}\operatorname{Fix}(R)=\{x\in M\mid R(x)=x\} of an antisymplectic involution RR of a symplectic manifold (M,ω)(M,\omega) is Lagrangian if it is nonempty. If (M,ω)(M,\omega) is monotone, then every real Lagrangian of MM is monotone, see Lemma 2.1. Recall that the Clifford torus 𝕋Clif=S1×S1\mathbb{T}_{\operatorname{Clif}}=S^{1}\times S^{1} is a real Lagrangian torus in S2×S2S^{2}\times S^{2} whose antisymplectic involution RClifR_{\operatorname{Clif}} is given by the product of the reflection of S2S^{2} fixing the equator.

The main result of this paper is to prove the Hamiltonian unknottedness of real Lagrangian tori in S2×S2S^{2}\times S^{2} in contrast to the monotone case.

Main Theorem.

Any real Lagrangian torus in S2×S2S^{2}\times S^{2} is Hamiltonian isotopic to the Clifford torus 𝕋Clif\mathbb{T}_{\operatorname{Clif}}.

As we discussed, being real plays a key role, and the result shows a non-trivial phenomenon of Hamiltonian unknottedness. An immediate consequence of the main theorem together with a known result [30, Proposition B] is the complete classification of real Lagrangian submanifolds in S2×S2S^{2}\times S^{2}.

Theorem A.

Any real Lagrangian submanifold in S2×S2S^{2}\times S^{2} is Hamiltonian isotopic to either the antidiagonal sphere Δ¯\overline{\Delta} or the Clifford torus 𝕋Clif\mathbb{T}_{\operatorname{Clif}}.

From a real symplectic perspective, the study of real Lagrangian tori in S2×S2S^{2}\times S^{2} is the simplest non-trivial case in dimension 4. For topological reasons, there is no closed real Lagrangian in 2n{\mathbb{R}}^{2n} at all. Known monotone symplectic 4-manifolds containing real Lagrangian tori are S2×S2S^{2}\times S^{2} and the three-fold monotone blow-up of P2{\mathbb{C}}P^{2}, see [7, Theorem D]. In [29], it was proved that the Chekanov–Schlenk torus in S2×S2S^{2}\times S^{2} is not real, from which we believed that our main result holds. We refer to a recent work of Brendel [6] that extends the result in [29].

In order to prove the main theorem, we employ Gromov’s foliation theorem (Section 2.2) together with the Cieliebak–Schwingenheuer criterion (Section 2.4). Gromov’s strategy was to foliate the monotone symplectic manifold S2×S2S^{2}\times S^{2} by a family of JJ-holomorphic spheres of minimal symplectic area, and hence to reduce symplectic questions to two dimensional fibered versions. In particular, when one deforms the split complex structure iii\oplus i on S2×S2S^{2}\times S^{2} to any tame almost complex structure JJ, two transversal foliations by JJ-holomorphic spheres still exist (even smoothly depending on JJ), and yield a symplectic S2S^{2}-fibration of S2×S2S^{2}\times S^{2} over S2S^{2}. In the case of studying a monotone Lagrangian torus, the nicest situation is when a fiber symplectic sphere intersects the torus along a circle or does not intersect at all. Cieliebak–Schwingenheuer provided a criterion for the Hamiltonian unknottedness of tori in S2×S2S^{2}\times S^{2} by means of this fibered structure. In particular, the existence of two suitable symplectic sections of the fibration is the precise condition for a given monotone Lagrangian torus being Hamiltonianly deformable into the Clifford torus 𝕋Clif\mathbb{T}_{\operatorname{Clif}}. In general, one symplectic section always exists (essentially by an SFT argument, see Proposition 2.6), but the second symplectic section may be missing. For real Lagrangian tori, an antisymplectic involution will provide the second symplectic section as desired. To realize this heuristic argument precisely, we should deal with an almost complex structure that is compatible with an antisymplectic involution.

It might be interesting to mention interactions between real symplectic topology and real algebraic varieties. We refer to [14, 16, 28] for expositions about the topology of real algebraic varieties. Inspired by the notion of quasi-simplicity for real algebraic varieties [28], we formulate a counterpart in symplectic topology. A closed monotone symplectic manifold (M,ω)(M,\omega) is called symplectically quasi-simple if the diffeomorphism type of connected real Lagrangian submanifolds of MM uniquely determines its Hamiltonian isotopy class.

Problem.

Show that symplectic del Pezzo surfaces are symplectically quasi-simple.

Notice that one has to impose a condition on the homology classes of real Lagrangians if necessary. It is known that the problem is true for P2{\mathbb{C}}P^{2} [30, Proposition A] and S2×S2S^{2}\times S^{2} (Theorem A). Note that quasi-simplicity in real algebraic varieties is known for all real del Pezzo surfaces by Degtyarev–Itenberg–Kharlamov [14, 15]. Together with Brendel and Moon [7, Theorem D], we obtain the complete list of diffeomorphism types of connected real Lagrangians in toric symplectic del Pezzo surfaces. For symplectic results, we refer to the works of Evans [24] and Seidel [38], which deal with the (un)knotting problems of Lagrangian spheres in del Pezzo surfaces. See also the work of Borman–Li–Wu [4]. Finally, we mention a work of Welschinger [41] which might be related to the problem. In that paper, he defines an invariant under deformation of real symplectic 4-manifolds, called Welschinger invariant.

This paper is organized as follows. In Section 2, we give results relevant for the proof of the main theorem. In Section 3, we prove the main theorem.

2 Structural results for real Lagrangian tori in S2×S2S^{2}\times S^{2}

Recall that a symplectic manifold (M,ω)(M,\omega) is monotone if there exists C>0C>0 such that c1(M)=C[ω]c_{1}(M)=C\cdot[\omega] and that a Lagrangian submanifold LL in MM is monotone if there exists C>0C^{\prime}>0 such that μL(β)=Cω(β)\mu_{L}(\beta)=C^{\prime}\cdot\omega(\beta) for all βπ2(M,L)\beta\in\pi_{2}(M,L), where μL:π2(M,L)\mu_{L}\colon\pi_{2}(M,L)\to{\mathbb{Z}} denotes the Maslov class of LL, see [34, Definition 3.4.4].

Lemma 2.1.

Every real Lagrangian in a monotone symplectic manifold (M,ω)(M,\omega) is monotone.

Proof.

Let L=Fix(R)L=\operatorname{Fix}(R) be a real Lagrangian in MM for an antisymplectic involution RR. Note that S2S^{2} is the union of two closed discs D1D_{1} and D2D_{2} such that D2=ρ(D1)D_{2}=\rho(D_{1}), where ρ\rho is an orientation reversing involution of S2S^{2} whose fixed point set is D1=D2\partial D_{1}=\partial D_{2}. For a smooth map β:(D2,D2)(M,L)\beta\colon(D^{2},\partial D^{2})\to(M,L), its double is defined by

β:S2M,β(z)={β(z),zD1D2,R(β(ρ(z))),zD2.\beta^{\sharp}\colon S^{2}\to M,\quad\beta^{\sharp}(z)=\begin{cases}\beta(z),&z\in D_{1}\cong D^{2},\\ R(\beta(\rho(z))),&z\in D_{2}.\end{cases}

Then one can check that c1(β)=μL(β)c_{1}(\beta^{\sharp})=\mu_{L}(\beta) and ω(β)=2ω(β)\omega(\beta^{\sharp})=2\cdot\omega(\beta). Since MM is monotone, LL is monotone as well. ∎

Throughout this paper, the symplectic manifold S2×S2S^{2}\times S^{2} is equipped with the split symplectic form ωω\omega\oplus\omega, where ω\omega is a Euclidean area form on S2S^{2}. This symplectic form is monotone, and hence every real Lagrangian in S2×S2S^{2}\times S^{2} is monotone.

2.1 Topology of antisymplectic involutions of S2×S2S^{2}\times S^{2}

We fix generators of H2(S2×S2)2H_{2}(S^{2}\times S^{2})\cong{\mathbb{Z}}^{2},

A1=[S2×{pt}]andA2=[{pt}×S2].A_{1}=[S^{2}\times\{\operatorname{pt}\}]\quad\text{and}\quad A_{2}=[\{\operatorname{pt}\}\times S^{2}].

The following simple topological result plays a crucial role. This result says that any antisymplectic involution RR of S2×S2S^{2}\times S^{2} with fixed point set Fix(R)T2\operatorname{Fix}(R)\cong T^{2} is homologically the standard antisymplectic involution RClifR_{\operatorname{Clif}} for the Clifford torus 𝕋Clif\mathbb{T}_{\operatorname{Clif}}.

Lemma 2.2.

Let RR be an antisymplectic involution on S2×S2S^{2}\times S^{2} whose fixed point set Fix(R)\operatorname{Fix}(R) is diffeomorphic to T2T^{2}. Then the map RR_{*} induced in homology H2(S2×S2)H_{2}(S^{2}\times S^{2}) is given by RAi=AiR_{*}A_{i}=-A_{i} for i=1,2i=1,2.

Proof.

Write R=(a1a2b1b2)R_{*}=\begin{pmatrix}a_{1}&a_{2}\\ b_{1}&b_{2}\end{pmatrix} for the induced map of RR on H2(S2×S2)2H_{2}(S^{2}\times S^{2})\cong{\mathbb{Z}}^{2}. It satisfies R2=idR_{*}^{2}=\operatorname{id} and

R(A1+A2)=(A1+A2)(a1+a2b1+b2)=(11).R_{*}(A_{1}+A_{2})=-(A_{1}+A_{2})\iff\begin{pmatrix}a_{1}+a_{2}\\ b_{1}+b_{2}\end{pmatrix}=\begin{pmatrix}-1\\ -1\end{pmatrix}.

The second condition holds since RR is antisymplectic. Since RR is orientation-preserving, RR_{*} preserves the intersection form of S2×S2S^{2}\times S^{2}. In particular, for i=1,2i=1,2 we obtain

0\displaystyle 0 =AiAi=RAiRAi=(aibi)(aibi)=2aibi.\displaystyle=A_{i}\bullet A_{i}=R_{*}A_{i}\bullet R_{*}A_{i}=\begin{pmatrix}a_{i}\\ b_{i}\end{pmatrix}\bullet\begin{pmatrix}a_{i}\\ b_{i}\end{pmatrix}=2a_{i}b_{i}.

Hence, RR_{*} must be either (1001)\begin{pmatrix}-1&0\\ 0&-1\end{pmatrix} or (0110)\begin{pmatrix}0&-1\\ -1&0\end{pmatrix}, and the same result holds for H2(S2×S2;)H_{2}(S^{2}\times S^{2};{\mathbb{Q}}). The Lefschetz fixed point theorem [14, Section 1.3] implies that

χ(Fix(R))\displaystyle\chi(\operatorname{Fix}(R)) =i=04(1)itrace[R:Hi(S2×S2;)Hi(S2×S2;)]\displaystyle=\sum_{i=0}^{4}(-1)^{i}\operatorname{trace}\Big{[}R_{*}\colon H_{i}(S^{2}\times S^{2};{\mathbb{Q}})\to H_{i}(S^{2}\times S^{2};{\mathbb{Q}})\Big{]}
=2+trace[R:H2(S2×S2;)H2(S2×S2;)].\displaystyle=2+\operatorname{trace}\Big{[}R_{*}\colon H_{2}(S^{2}\times S^{2};{\mathbb{Q}})\to H_{2}(S^{2}\times S^{2};{\mathbb{Q}})\Big{]}.

Since χ(T2)=0\chi(T^{2})=0, the lemma follows. ∎

Remark 2.3.

By the above proof, for given antisymplectic involution RR of S2×S2S^{2}\times S^{2} with possibly empty fixed point set, the induced map RR_{*} on H2(S2×S2)H_{2}(S^{2}\times S^{2}) is either

I1=(1001)orI2=(0110).I_{1}=\begin{pmatrix}-1&0\\ 0&-1\end{pmatrix}\quad\text{or}\quad I_{2}=\begin{pmatrix}0&-1\\ -1&0\end{pmatrix}.

It is known that any real Lagrangian in S2×S2S^{2}\times S^{2} is diffeomorphic to either T2T^{2} or S2S^{2}, see [30, Proposition B]. Hence, if Fix(R)\operatorname{Fix}(R) is nonempty we obtain

  • Fix(R)\operatorname{Fix}(R) is diffeomorphic to T2T^{2} if and only if R=I1R_{*}=I_{1}.

  • Fix(R)\operatorname{Fix}(R) is diffeomorphic to S2S^{2} if and only if R=I2R_{*}=I_{2}.

2.2 Gromov’s foliation theorem

We recall the celebrated Gromov foliation theorem in S2×S2S^{2}\times S^{2}, see [25, Theorem 2.4.A1A_{1}]. Let 𝒥\mathcal{J} denote the space of compatible almost complex structures on S2×S2S^{2}\times S^{2}. We emphasize that for the following result it is crucial that the symplectic form ωω\omega\oplus\omega is monotone.

Theorem 2.4 (Gromov).

For every J𝒥J\in\mathcal{J} there exist two transversal foliations 1\mathcal{F}_{1} and 2\mathcal{F}_{2} of S2×S2S^{2}\times S^{2} whose leaves are (unparametrized) embedded JJ-holomorphic spheres in the homology class A1A_{1} and A2A_{2}, respectively.

Here, by transversal foliations 1\mathcal{F}_{1} and 2\mathcal{F}_{2}, we mean that any two leaves of 1\mathcal{F}_{1} and 2\mathcal{F}_{2} are transverse. By positivity of intersections [33, Section 2.6], the two leaves intersects transversely at a single point.

Remark 2.5.
  1. (i)

    Actually, Theorem 2.4 holds for tame almost complex structures.

  2. (ii)

    By topological argument as in [12, Remark 2.6 (a)], any smooth FF-fibration of S2×S2S^{2}\times S^{2} over a closed surface BB must satisfy BFS2B\cong F\cong S^{2}. Gromov’s foliations yield plenty of symplectic S2S^{2}-fibrations of S2×S2S^{2}\times S^{2} over S2S^{2} [32, Proposition 4.1]. See also Example 2.10 for an easy illustration.

  3. (iii)

    The embeddeness of the leaves of a Gromov’s foliation follows from the adjunction inequality, see [33, Theorem 2.6.4]. More precisely, any JJ-holomorphic sphere of S2×S2S^{2}\times S^{2} in the homology class AiA_{i} is embedded.

2.3 The neck-stretching for a real Lagrangian torus

Let RR be an antisymplectic involution of S2×S2S^{2}\times S^{2}. A compatible almost complex structure JJ on S2×S2S^{2}\times S^{2} is called RR-anti-invariant if J=RJJ=-R^{*}J. We abbreviate by 𝒥R\mathcal{J}_{R} the space of RR-anti-invariant compatible almost complex structures on S2×S2S^{2}\times S^{2}, which is nonempty and contractible, see [41, Proposition 1.1].

The following is an application of a neck-stretching argument combined with Gromov’s theorem, which is a version of Dimitroglou Rizell–Goodman–Ivrii [18, Theorem C]. See also a related result of Welschinger [42, Theorem 1.3].

Proposition 2.6.

Let L=Fix(R)L=\operatorname{Fix}(R) be a real Lagrangian torus in S2×S2S^{2}\times S^{2} for an antisymplectic involution RR. Then there exists J𝒥RJ\in\mathcal{J}_{R} and a JJ-holomorphic sphere in S2×S2S^{2}\times S^{2} which represents the homology class A1A_{1} and is disjoint from LL.

The rest of this section is devoted to the proof of this statement. Below, we mainly follow the descriptions given in [18, Sections 2, 3, and 4].

The real splitting construction.

We explain the splitting construction for a real Lagrangian torus L=Fix(R)L=\operatorname{Fix}(R) in S2×S2S^{2}\times S^{2}, which matches well with the antisymplectic involution RR. We refer to [18, Section 2], [11, Example 2.5] or [19, Example 1.3.1] for the split construction for a Lagrangian.

Fix a flat metric on the torus LL and write (𝜽,𝐩)=(θ1,θ2,p1,p2)TLT2×2(\boldsymbol{\theta},{\bf p})=(\theta_{1},\theta_{2},p_{1},p_{2})\in T^{*}L\cong T^{2}\times{\mathbb{R}}^{2} for the coordinates. Let λcan=p1dθ1+p2dθ2\lambda_{\operatorname{can}}=p_{1}d\theta_{1}+p_{2}d\theta_{2} be the Liouville form on TLT^{*}L. The cotangent bundle (TL,dλcan)(T^{*}L,d\lambda_{\operatorname{can}}) carries the canonical antisymplectic involution Rcan(𝜽,𝐩)=(𝜽,𝐩)R_{\operatorname{can}}(\boldsymbol{\theta},{\bf p})=(\boldsymbol{\theta},-{\bf p}), which is exact, i.e., Rcanλcan=λcanR_{\operatorname{can}}^{*}\lambda_{\operatorname{can}}=-\lambda_{\operatorname{can}}, and is an isometry. This map restricts to the (strict) anticontact involution Rcan|SLR_{\operatorname{can}}|_{S^{*}L} on the unit cotangent bundle SLS^{*}L equipped with the contact form α=λcan|SL{\alpha}=\lambda_{\operatorname{can}}|_{S^{*}L}. This involution extends to the exact antisymplectic involution on the symplectization (×SL,d(etα))({\mathbb{R}}\times S^{*}L,d(e^{t}\alpha)) by

Rcancyl(t,𝜽,𝐩)=(t,𝜽,𝐩).R_{\operatorname{can}}^{\operatorname{cyl}}(t,\boldsymbol{\theta},{\bf p})=(t,\boldsymbol{\theta},-{\bf p}).

For r>0r>0 we let TrL={(𝜽,𝐩)TL𝐩r}T^{*}_{r}L=\{(\boldsymbol{\theta},{\bf p})\in T^{*}L\mid\|{\bf p}\|\leq r\}. By the equivariant Weinstein neighborhood theorem [35, Theorem 2], there exists a symplectic embedding

(2.1) Ψ:(T4εL,dλcan)⸦-→(S2×S2,ωω)\Psi\colon(T^{*}_{4\varepsilon}L,d\lambda_{\operatorname{can}})\lhook\joinrel\relbar\joinrel\rightarrow(S^{2}\times S^{2},\omega\oplus\omega)

such that Ψ(𝟎L)=L\Psi({\bf 0}_{L})=L and RΨ=ΨRcanR\circ\Psi=\Psi\circ R_{\operatorname{can}}. Together with the exact symplectomorphism

(2.2) (×SL,d(etα))(TL𝟎L,dλcan),(t,𝜽,𝐩)(𝜽,et𝐩),({\mathbb{R}}\times S^{*}L,d(e^{t}\alpha))\overset{\cong}{\longrightarrow}(T^{*}L\setminus{\bf 0}_{L},d\lambda_{\operatorname{can}}),\qquad(t,\boldsymbol{\theta},{\bf p})\longmapsto(\boldsymbol{\theta},e^{t}{\bf p}),

which identifies the antisymplectic involutions RcancylR_{\operatorname{can}}^{\operatorname{cyl}} and RcanR_{\operatorname{can}}, we see that S2×S2LS^{2}\times S^{2}\setminus L and TLT^{*}L have a cylindrical end over the real contact manifold (SL,α,Rcan|SL)(S^{*}L,\alpha,R_{\operatorname{can}}|_{S^{*}L}). Therefore, we obtain the real split symplectic manifold consisting of

(2.3) {(S2×S2L,ωω,R,J)(Top level)(×SL,d(etα),Rcancyl,Jcyl)(Middle level)(TL,dλcan,Rcan,Jstd)(Bottom level)\left\{\begin{aligned} &(S^{2}\times S^{2}\setminus L,\;&&\omega\oplus\omega,\;&&R,\quad\;&&J_{\infty}&&)\qquad\qquad&&\text{(Top level)}\\ &({\mathbb{R}}\times S^{*}L,&&d(e^{t}\alpha),&&R_{\operatorname{can}}^{\operatorname{cyl}},&&J_{\operatorname{cyl}}&&)&&\text{(Middle level)}\\ &(T^{*}L,&&d\lambda_{\operatorname{can}},&&R_{\operatorname{can}},&&J_{\operatorname{std}}&&)&&\text{(Bottom level)}\end{aligned}\right.

endowed with the RR-anti-invariant almost complex structures JJ_{\infty}, JcylJ_{\operatorname{cyl}}, and JstdJ_{\operatorname{std}}, which are explained below.

A neck-stretching family of RR-anti-invariant almost complex structures.

A compatible almost complex structure JJ on the symplectization (×SL,d(etα))({\mathbb{R}}\times S^{*}L,d(e^{t}\alpha)) is called cylindrical if JJ is {\mathbb{R}}-translation invariant, J(t)=αJ(\partial_{t})=\mathcal{R}_{\alpha}, and J(kerα)=kerαJ(\ker\alpha)=\ker\alpha. Here, α\mathcal{R}_{\alpha} is the Reeb vector field on (SL,α)(S^{*}L,\alpha) uniquely determined by the equations dα(α,)=0d\alpha(\mathcal{R}_{\alpha},\cdot)=0 and α(α)=1\alpha(\mathcal{R}_{\alpha})=1. We recall the construction in [18, Section 4] of the specific almost complex structures on the split symplectic manifold (2.3). We refer to [18, Lemma 4.1] for details.

Using the identification (2.2), we define the cylindrical compatible almost complex structure JcylJ_{\operatorname{cyl}} on (×SL,d(etα))(TL𝟎L,dλcan)({\mathbb{R}}\times S^{*}L,d(e^{t}\alpha))\cong(T^{*}L\setminus{\bf 0}_{L},d\lambda_{\operatorname{can}}) by

Jcylθi\displaystyle J_{\operatorname{cyl}}\partial_{\theta_{i}} =𝐩pi,for (𝜽,𝐩)TL𝟎L×SL and i=1,2.\displaystyle=-\|{\bf p}\|\partial_{p_{i}},\,\qquad\text{for $(\boldsymbol{\theta},{\bf p})\in T^{*}L\setminus{\bf 0}_{L}\cong{\mathbb{R}}\times S^{*}L$\; and\; $i=1,2$}.
We consider the tame almost complex structure JstdJ_{\operatorname{std}} on (TL,dλcan)(T^{*}L,d\lambda_{\operatorname{can}}) given by
Jstdθi\displaystyle J_{\operatorname{std}}\partial_{\theta_{i}} =f(𝐩)pi,for (𝜽,𝐩)TL and i=1,2,\displaystyle=-f(\|{\bf p}\|)\partial_{p_{i}},\quad\text{for $(\boldsymbol{\theta},{\bf p})\in T^{*}L$\; and\; $i=1,2$},

where f:[0,)[ε,)f\colon[0,\infty)\to[\varepsilon,\infty) is a smooth function such that

  • f(t)0f^{\prime}(t)\geq 0 for t0t\geq 0,

  • f(t)=εf(t)=\varepsilon for 0tε0\leq t\leq\varepsilon, and

  • f(t)=tf(t)=t for t2εt\geq 2\varepsilon.

Here, ε>0\varepsilon>0 is chosen such that (2.1) exists. On TLT2εLT^{*}L\setminus T^{*}_{2\varepsilon}L, the almost complex structure JstdJ_{\operatorname{std}} agrees with the cylindrical almost complex structure JcylJ_{\operatorname{cyl}}. We can readily check that JstdJ_{\operatorname{std}} and JcylJ_{\operatorname{cyl}} are RR-anti-invariant, i.e.,

Jstd=RcanJstd,Jcyl=(Rcancyl)Jcyl.J_{\operatorname{std}}=-R_{\operatorname{can}}^{*}J_{\operatorname{std}},\qquad J_{\operatorname{cyl}}=-(R_{\operatorname{can}}^{\operatorname{cyl}})^{*}J_{\operatorname{cyl}}.

Take the neighborhood U:=Ψ(T4εL)U:=\Psi(T^{*}_{4\varepsilon}L) of LL in S2×S2S^{2}\times S^{2}, where Ψ:T4εLS2×S2\Psi\colon T^{*}_{4\varepsilon}L\hookrightarrow S^{2}\times S^{2} is the symplectic embedding from (2.1). Let 𝒥R(S2×S2L)\mathcal{J}_{R}(S^{2}\times S^{2}\setminus L) be the space of RR-anti-invariant compatible almost complex structures on S2×S2LS^{2}\times S^{2}\setminus L. The set

𝒥Rcyl(S2×S2L)={J𝒥R(S2×S2L)ΨJ=Jcyl on UL},\mathcal{J}^{\operatorname{cyl}}_{R}(S^{2}\times S^{2}\setminus L)=\{J_{\infty}\in\mathcal{J}_{R}(S^{2}\times S^{2}\setminus L)\mid\text{$\Psi^{*}J_{\infty}=J_{\operatorname{cyl}}$ on $U\setminus L$}\},

is nonempty and contractible.

Suppose that we are given J𝒥Rcyl(S2×S2L)J_{\infty}\in\mathcal{J}_{R}^{\operatorname{cyl}}(S^{2}\times S^{2}\setminus L). A neck-stretching family {Jτ}τ0\{J_{\tau}\}_{\tau\geq 0} in [18, Section 2.5] of RR-anti-invariant tame almost complex structures JτJ_{\tau} on S2×S2S^{2}\times S^{2} is defined by

Jτ={Jon S2×S2Ψ(T4εL),ΦτJcylon Ψ(T4εLT2εL),Jstdon Ψ(T2εL),J_{\tau}=\begin{cases}J_{\infty}&\text{on $S^{2}\times S^{2}\setminus\Psi(T^{*}_{4\varepsilon}L)$},\\ \Phi_{\tau}^{*}J_{\operatorname{cyl}}&\text{on $\Psi(T_{4\varepsilon}^{*}L\setminus T_{2\varepsilon}^{*}L)$},\\ J_{\operatorname{std}}&\text{on $\Psi(T^{*}_{2\varepsilon}L)$},\end{cases}

where

  • The map Φτ\Phi_{\tau} is a diffeomorphism

    Φτ:[log2ε,log4ε)×SL\displaystyle\Phi_{\tau}\colon[\log 2\varepsilon,\log 4\varepsilon)\times S^{*}L [log2ε,log4ε+τ)×SL\displaystyle\longrightarrow[\log 2\varepsilon,\log 4\varepsilon+\tau)\times S^{*}L
    (t,𝜽,𝐩)\displaystyle(t,\boldsymbol{\theta},{\bf p}) (ϕτ(t),𝜽,𝐩),\displaystyle\longmapsto(\phi_{\tau}(t),\boldsymbol{\theta},{\bf p}),

    induced by an orientation-preserving diffeomorphism

    ϕτ:[log2ε,log4ε)[log2ε,log4ε+τ).\phi_{\tau}\colon[\log 2\varepsilon,\log 4\varepsilon)\longrightarrow[\log 2\varepsilon,\log 4\varepsilon+\tau).
  • The identifications (2.1) and (2.2) are used in JstdJ_{\operatorname{std}} and ΦτJcyl\Phi_{\tau}^{*}J_{\operatorname{cyl}}.

Remark 2.7.

In the limit τ\tau\to\infty of the quadruple (S2×S2,ωω,R,Jτ)(S^{2}\times S^{2},\omega\oplus\omega,R,J_{\tau}) we obtain the real split symplectic manifold (2.3), see for instance [10, Section 2.7] or [5, Section 3.4].

We are now in position to prove Proposition 2.6. The SFT analysis that we will use is carried out in [18, Sections 2 and 3] and we closely follow their arguments adapted to our purpose.

Proof of Proposition 2.6.

Pick a regular J𝒥Rcyl,reg(S2×S2L)J_{\infty}\in\mathcal{J}_{R}^{\operatorname{cyl},\,\operatorname{reg}}(S^{2}\times S^{2}\setminus L) as in Lemma 2.8 and consider the associated neck-stretching family {Jτ}τ0\{J_{\tau}\}_{\tau\geq 0} of RR-anti-invariant tame almost complex structures JτJ_{\tau} on S2×S2S^{2}\times S^{2}. It follows from Gromov’s result together with positivity of intersections that for any τ0\tau\geq 0 there exists a unique embedded JτJ_{\tau}-holomorphic sphere in the homology class A1A_{1} passing through any given point in S2×S2S^{2}\times S^{2}. We can therefore apply the SFT compactness theorem [18, Theorem 2.2] or [5, 10] to these spheres to obtain a limit holomorphic building in the real split symplectic manifold (S2×S2L,ωω,R)(TL,dλcan,Rcan)(S^{2}\times S^{2}\setminus L,\omega\oplus\omega,R)\sqcup(T^{*}L,d\lambda_{\operatorname{can}},R_{\operatorname{can}}) given in (2.3). Note that the limit holomorphic building possibly consists of components in the middle levels (×SL)(×SL)({\mathbb{R}}\times S^{*}L)\sqcup\cdots\sqcup({\mathbb{R}}\times S^{*}L). By convexity and exactness of (TL,dλcan)(T^{*}L,d\lambda_{\operatorname{can}}) and (×SL,d(etα))({\mathbb{R}}\times S^{*}L,d(e^{t}\alpha)), the limit holomorphic building must contain at least one component in the top level S2×S2LS^{2}\times S^{2}\setminus L.
Claim. If a limit holomorphic building is broken, then its top level consists of two simple JJ_{\infty}-holomorphic planes in S2×S2LS^{2}\times S^{2}\setminus L of index 1.

In [18, Proposition 3.5], a similar version of the claim is obtained under the choice of a regular J𝒥cyl(S2×S2L)J_{\infty}\in\mathcal{J}^{\operatorname{cyl}}(S^{2}\times S^{2}\setminus L) for all simple punctured JJ_{\infty}-holomorphic spheres in S2×S2LS^{2}\times S^{2}\setminus L, where LL is not necessarily monotone. The generic choice of JJ_{\infty} was crucially required to guarantee the index non-negativity result for punctured spheres in S2×S2LS^{2}\times S^{2}\setminus L. In our case, the monotonicity of LL controls the index of punctured spheres.
Step 1. Any punctured JJ_{\infty}-holomorphic sphere in S2×S2LS^{2}\times S^{2}\setminus L satisfies ind(u)1\operatorname{ind}(u)\geq 1.

Fixing a symplectic trivialization Φ\Phi of the contact structure kerα\ker\alpha on SLS^{*}L, we denote by CZΦ(Γ)\operatorname{CZ}^{\Phi}(\Gamma) the Conley–Zehnder index of a Morse–Bott manifold Γ\Gamma of periodic Reeb orbits in SLS^{*}L and by c1,relΦ()c_{1,\operatorname{rel}}^{\Phi}(\cdot) the relative first Chern number, see [18, Section 3.1]. By [18, Equations (1) and (2) in Section 3.1], the index of a punctured JJ_{\infty}-holomorphic sphere uu in S2×S2LS^{2}\times S^{2}\setminus L having 1\ell\geq 1 negative punctures asymptotic to periodic orbits in Morse–Bott manifolds Γ1,,Γ\Gamma_{1},\dots,\Gamma_{\ell} is given by

ind(u)\displaystyle\operatorname{ind}(u) =2+i=1(CZΦ(Γi)1)+2c1,relΦ(u)\displaystyle=-2+\ell-\sum_{i=1}^{\ell}(\operatorname{CZ}^{\Phi}(\Gamma_{i}^{-})-1)+2c_{1,\operatorname{rel}}^{\Phi}(u)
=2++μL(u¯),\displaystyle=-2+\ell+\mu_{L}(\overline{u}),

where u¯\overline{u} is a surface in S2×S2S^{2}\times S^{2} with boundary on LL which is the boundary compactification of uu obtained by adding (geodesic) circles in LL corresponding to the asymptotic orbits. Since LL is orientable, u¯\overline{u} has Maslov number μL(u¯)2\mu_{L}(\overline{u})\in 2{\mathbb{Z}}. By the monotonicity of LL together with the fact that uu is non-constant, we have μL(u¯)2\mu_{L}(\overline{u})\geq 2. Hence, we deduce that

ind(u)=2++μL(u¯)1,\operatorname{ind}(u)=-2+\ell+\mu_{L}(\overline{u})\geq 1,

which completes Step 1.
Step 2. Proof of the claim.

This essentially follows from the proof of [18, Proposition 3.5]. Indeed, that proof yields the identity

i=1N1ind(vi)i=1N2χ(wi)=2,\sum_{i=1}^{N_{1}}\operatorname{ind}(v_{i})-\sum_{i=1}^{N_{2}}\chi(w_{i})=2,

where v1,,vN1v_{1},\dots,v_{N_{1}} and w1,,wN2w_{1},\dots,w_{N_{2}} denote the components of the limit holomorphic building in the top level S2×S2LS^{2}\times S^{2}\setminus L and the remaining levels, respectively. Here, χ(wi)\chi(w_{i}) is the Euler characteristic of the domain of wiw_{i}. Since there are no contractible periodic orbits in SLS^{*}L, a pseudoholomorphic plane in TLT^{*}L or ×SL{\mathbb{R}}\times S^{*}L asymptotic to a periodic orbit cannot exist, which shows χ(wi)0\chi(w_{i})\leq 0. By Step 1, we know that ind(vi)1\operatorname{ind}(v_{i})\geq 1. Since the components of the limit holomorphic building in the split symplectic manifold (S2×S2L)TL(S^{2}\times S^{2}\setminus L)\sqcup T^{*}L glue together to form a sphere, we conclude that the top level of the limit holomorphic building consists of two planes of index 1. The simplicity of the plane follows from [18, Lemmata 3.3 and 3.4]. Hence, the claim follows.

We now crucially use that our J𝒥Rcyl,reg(S2×S2L)J_{\infty}\in\mathcal{J}_{R}^{\operatorname{cyl},\,\operatorname{reg}}(S^{2}\times S^{2}\setminus L) is regular. By the claim, if the limit building is broken, then its top level consists of two simple JJ_{\infty}-holomorphic planes in S2×S2LS^{2}\times S^{2}\setminus L of index 1. Such planes are contained in the moduli space of index 1 simple JJ_{\infty}-holomorphic planes in S2×S2LS^{2}\times S^{2}\setminus L, which is a smooth manifold of dimension 1. Since the total collection of components of broken limit buildings in the top level S2×S2LS^{2}\times S^{2}\setminus L forms a subset of dimension at most 3 in S2×S2LS^{2}\times S^{2}\setminus L, we conclude that there must be a non-broken limit building, that is an embedded JJ_{\infty}-holomorphic sphere uu in S2×S2LS^{2}\times S^{2}\setminus L whose homology class represents A1A_{1}. It remains to show that uu can be seen as a JJ-holomorphic sphere in S2×S2S^{2}\times S^{2} for some J𝒥RJ\in\mathcal{J}_{R}. Since the image of uu is a compact set in S2×S2LS^{2}\times S^{2}\setminus L, we can choose an RR-invariant neighborhood of LL in S2×S2S^{2}\times S^{2} which is disjoint from uu. We then modify JJ_{\infty} on this neighborhood so that the resulting one can be extended to an RR-anti-invariant compatible almost complex structure JJ defined on S2×S2S^{2}\times S^{2}, namely J𝒥RJ\in\mathcal{J}_{R}. This completes the proof. ∎

Following ideas of [29, Section 3], we show the equivariant transversality for simple pseudoholomorphic planes in S2×S2LS^{2}\times S^{2}\setminus L asymptotic to a periodic orbit.

Lemma 2.8.

There exists a Baire subset 𝒥Rcyl,reg(S2×S2L)𝒥Rcyl(S2×S2L)\mathcal{J}_{R}^{\operatorname{cyl},\,\operatorname{reg}}(S^{2}\times S^{2}\setminus L)\subset\mathcal{J}_{R}^{\operatorname{cyl}}(S^{2}\times S^{2}\setminus L) which has the property that every J𝒥Rcyl,reg(S2×S2L)J_{\infty}\in\mathcal{J}_{R}^{\operatorname{cyl},\,\operatorname{reg}}(S^{2}\times S^{2}\setminus L) is regular for simple JJ_{\infty}-holomorphic planes in S2×S2LS^{2}\times S^{2}\setminus L asymptotic to a periodic orbit in SLS^{*}L.

Proof.

We first observe that every simple JJ_{\infty}-holomorphic plane u:S2{}S2×S2Lu\colon{\mathbb{C}}\cong S^{2}\setminus\{\infty\}\to S^{2}\times S^{2}\setminus L is not RR-invariant for topological reasons, namely ImuR(Imu)\operatorname{Im}u\neq R(\operatorname{Im}u). Assume to the contrary that Imu=R(Imu)\operatorname{Im}u=R(\operatorname{Im}u). Let ρ0\rho_{0} be an antiholomorphic involution on S2S^{2} leaving S2\infty\in S^{2} invariant. Since uu and Ruρ0|S2{}R\circ u\circ\rho_{0}|_{S^{2}\setminus\{\infty\}} are simple JJ_{\infty}-holomorphic planes whose images coincide, there exists σAut(S2)\sigma\in\operatorname{Aut}(S^{2}) such that σ()=\sigma(\infty)=\infty and

(2.4) u=Ruρ0σ|S2{},u=R\circ u\circ\rho_{0}\circ\sigma|_{S^{2}\setminus\{\infty\}},

see [36, Theorem 3.7] and [33, Corollary 2.5.4]. By applying (2.4) twice, we obtain that u(ρ0σ|S2{})2=uu\circ(\rho_{0}\circ\sigma|_{S^{2}\setminus\{\infty\}})^{2}=u. Since uu is simple, ρ0σ|S2{}\rho_{0}\circ\sigma|_{S^{2}\setminus\{\infty\}} is an involution of {\mathbb{C}} and hence has a fixed point, say z0z_{0}\in{\mathbb{C}}. We then see that u(z0)Fix(R)=Lu(z_{0})\in\operatorname{Fix}(R)=L, which is a contradiction. Since JJ_{\infty}-holomorphic planes (asymptotic to periodic orbits) are not RR-invariant and are proper, the lemma follows almost verbatim from the proof of [29, Theorem 3.9]. We emphasize that our J𝒥Rcyl,reg(S2×S2L)J_{\infty}\in\mathcal{J}^{\operatorname{cyl},\,\operatorname{reg}}_{R}(S^{2}\times S^{2}\setminus L) is required to satisfy ΨJ=Jcyl\Psi^{*}J_{\infty}=J_{\operatorname{cyl}} on ULU\setminus L, but this is not problematic. By convexity, uu cannot be entirely contained in a cylindrical end ULU\setminus L since otherwise it violates the maximum principle. Therefore, the transversality argument works by perturbing JJ_{\infty} outside of ULU\setminus L. ∎

Remark 2.9.

Since 𝒥R\mathcal{J}_{R} is closed in 𝒥\mathcal{J} (in the CC^{\infty}-topology), the Baire category theorem and Lemma 2.8 imply that 𝒥Rreg\mathcal{J}_{R}^{\operatorname{reg}} is dense in 𝒥R\mathcal{J}_{R}, and hence not empty.

2.4 Cieliebak–Schwingenheuer’s criterion

We give a review on the Cieliebak–Schwingenheuer criterion [12, Theorem 1.1] when a monotone Lagrangian torus in S2×S2S^{2}\times S^{2} is Hamiltonian isotopic to the Clifford torus 𝕋Clif\mathbb{T}_{\operatorname{Clif}}. A monotone Lagrangian torus LL in S2×S2S^{2}\times S^{2} is called fibered if there exist a foliation \mathcal{F} of S2×S2S^{2}\times S^{2} by symplectic 2-spheres in the homology class A2A_{2} and a symplectic 2-sphere Σ\Sigma in the homology class A1A_{1} such that

  • Σ\Sigma is transverse to the leaves of \mathcal{F}.

  • Σ\Sigma is disjoint from LL.

  • The leaves of \mathcal{F} intersect LL in a circle or not at all.

In this case, we say that LL is fibered by \mathcal{F} and Σ\Sigma. It is a highly non-trivial result that any monotone Lagrangian torus in S2×S2S^{2}\times S^{2} is fibered. This result, which originally goes back to Ivrii’s thesis [27], is proved by Dimitroglou Rizell–Goodman–Ivrii [18, Theorem D] based on the neck-stretching argument [5].

Consider a monotone Lagrangian torus LL in S2×S2S^{2}\times S^{2} which is fibered by \mathcal{F} and Σ\Sigma. Each leaf of the foliation \mathcal{F} intersecting LL is written as a union of two closed discs glued along the embedded loop given by the intersection of LL and the leaf. Hence, the discs intersecting Σ\Sigma form a solid torus TT with boundary T=L\partial T=L. Note that the discs which do not intersect LL also define a solid torus TT^{\prime} with T=L\partial T^{\prime}=L.

Example 2.10 (Clifford torus).

Let iii\oplus i be the split complex structure on S2×S2S^{2}\times S^{2}. Gromov’s foliations are given by 1\mathcal{F}_{1} and 2\mathcal{F}_{2} whose leaves are the holomorphic spheres F1,y:=S2×{y}F_{1,y}:=S^{2}\times\{y\} for yS2y\in S^{2} and F2,x:={x}×S2F_{2,x}:=\{x\}\times S^{2} for xS2x\in S^{2}, respectively. The associated symplectic S2S^{2}-fibration is defined as follows. Fix one leaf F1,y0F_{1,y_{0}} of 1\mathcal{F}_{1}. Since each leaf F2,xF_{2,x} of 2\mathcal{F}_{2} intersects F1,y0F_{1,y_{0}} transversely at a unique point (x,y0)(x,y_{0}), we can define a symplectic S2S^{2}-fibration of S2×S2S^{2}\times S^{2} by sending F2,xF_{2,x} to (x,y0)F1,y0S2(x,y_{0})\in F_{1,y_{0}}\cong S^{2}. In this case, this is the projection π(x,y)=x\pi(x,y)=x of S2×S2S^{2}\times S^{2} onto the first S2S^{2}-factor. One checks that the Clifford torus 𝕋Clif=S1×S1\mathbb{T}_{\operatorname{Clif}}=S^{1}\times S^{1} is fibered by 2\mathcal{F}_{2} and F1,y0F_{1,y_{0}}. Each fiber F2,xS2F_{2,x}\cong S^{2} intersecting 𝕋Clif\mathbb{T}_{\operatorname{Clif}} is a union of two holomorphic discs of Maslov index 2 with boundary on 𝕋Clif\mathbb{T}_{\operatorname{Clif}}. The S1S^{1}-family of the discs intersecting F1,y0F_{1,y_{0}} forms a solid torus TT with T=𝕋Clif\partial T=\mathbb{T}_{\operatorname{Clif}}. It is worth noting that we can find another symplectic section F1,y1F_{1,y_{1}} for some y1S2y_{1}\in S^{2} (for example, the antipodal point of y0S2y_{0}\in S^{2}) which does not intersect the solid torus TT.

The criterion of Cieliebak–Schwingenheuer says that the existence of the second nice symplectic 2-sphere in the homology class A1A_{1} guarantees that a given monotone Lagrangian torus in S2×S2S^{2}\times S^{2} is Hamiltonian isotopic to the Clifford torus.

Theorem 2.11 (Cieliebak–Schwingenheuer).

Let LL be a monotone Lagrangian torus in S2×S2S^{2}\times S^{2} which is fibered by \mathcal{F} and Σ\Sigma. Suppose that there exists a second symplectic 2-sphere Σ\Sigma^{\prime} of S2×S2S^{2}\times S^{2} in the homology class A1A_{1} such that

  • Σ\Sigma^{\prime} is transverse to the leaves of \mathcal{F},

  • Σ\Sigma^{\prime} is disjoint from Σ\Sigma and the solid torus TT.

Then LL is Hamiltonian isotopic to the Clifford torus 𝕋Clif\mathbb{T}_{\operatorname{Clif}}.

The proof is based on a sophisticated version of the Lalonde–Mcduff inflation procedure [31].

Remark 2.12.

The converse of Theorem 2.11 obviously holds, namely if a monotone Lagrangian torus LL in S2×S2S^{2}\times S^{2} is Hamiltonian isotopic to the Clifford torus 𝕋Clif\mathbb{T}_{\operatorname{Clif}}, then LL is fibered by some \mathcal{F}, Σ\Sigma, and the second symplectic 2-sphere Σ\Sigma^{\prime} as well.

3 Proof of the Main Theorem

Throughout this section, we let L=Fix(R)L=\operatorname{Fix}(R) be a real Lagrangian torus in S2×S2S^{2}\times S^{2} for some antisymplectic involution RR of S2×S2S^{2}\times S^{2}. Recall that every JJ-holomorphic sphere of S2×S2S^{2}\times S^{2} in the homology class AiA_{i} for i=1,2i=1,2 is embedded, see Remark 2.5. We start with the following simple observation.

Lemma 3.1.

Let J𝒥RJ\in\mathcal{J}_{R} and i=1,2i=1,2. Suppose that u:S2S2×S2u\colon S^{2}\to S^{2}\times S^{2} is a JJ-holomorphic sphere in the homology class AiA_{i}. If uu intersects LL, then the embedded 2-sphere Imu\operatorname{Im}u is RR-invariant and ImuL\operatorname{Im}u\cap L is diffeomorphic to S1S^{1}.

Proof.

Consider the JJ-holomorphic sphere in S2×S2S^{2}\times S^{2} given by

u:=Ruρ,u^{\prime}:=R\circ u\circ\rho,

where ρ(z)=z¯1\rho(z)=\bar{z}^{-1} denotes the antiholomorphic involution of S2{}S^{2}\cong{\mathbb{C}}\cup\{\infty\} with Fix(ρ)S1\operatorname{Fix}(\rho)\cong S^{1}. Here, uu^{\prime} is JJ-holomorphic since JJ is RR-anti-invariant. By Lemma 2.2, uu^{\prime} represents the homology class AiA_{i}. By positivity of intersections, uu^{\prime} must be one leaf of Gromov’s foliation i\mathcal{F}_{i} associated to JJ as in Theorem 2.4. Pick any xImuLx\in\operatorname{Im}u\cap L. Since u(z)=xu(z)=x for some zS2z\in S^{2}, we see that

u(ρ(z))=Ruρ(ρ(z))=R(u(z))=R(x)=x,u^{\prime}(\rho(z))=R\circ u\circ\rho(\rho(z))=R(u(z))=R(x)=x,

which shows that uu^{\prime} passes through xx. This implies that Imu=Imu\operatorname{Im}u=\operatorname{Im}u^{\prime}, and hence Imu\operatorname{Im}u is RR-invariant as desired. Recalling that uu is embedded, the antisymplectic involution RR of S2×S2S^{2}\times S^{2} restricts to a smooth involution τ:=R|Imu\tau:=R|_{\operatorname{Im}u} of the sphere S2ImuS^{2}\cong\operatorname{Im}u. Since RAi=AiR_{*}A_{i}=-A_{i} by Lemma 2.2, the involution τ\tau is orientation-reversing. Recall that every smooth involution of S2S^{2} is conjugated to a map in O(3)O(3), see [13, Theorem 4.1]. Hence τ\tau is conjugated to either the reflection or the antipodal map on S2S^{2}. Since xFix(τ)x\in\operatorname{Fix}(\tau) and, in particular, Fix(τ)\operatorname{Fix}(\tau) is nonempty, τ\tau is conjugated to the reflection, and hence Fix(τ)=ImuLS1\operatorname{Fix}(\tau)=\operatorname{Im}u\cap L\cong S^{1}. ∎

Remark 3.2.

We notice that any smooth involution σ\sigma of S2S^{2} with fixed point set Fix(σ)S1\operatorname{Fix}(\sigma)\cong S^{1} must interchange the two discs obtained by cutting S2S^{2} along the embedded loop Fix(σ)\operatorname{Fix}(\sigma). This again follows from the fact that σ\sigma is conjugated to a map τO(3)\tau\in O(3). Since Fix(σ)Fix(τ)S1\operatorname{Fix}(\sigma)\cong\operatorname{Fix}(\tau)\cong S^{1}, we deduce that τ\tau is a reflection. Hence, σ\sigma must interchange the two discs as desired.

An immediate consequence of Lemma 3.1 together with Proposition 2.6 is that the real Lagrangian torus L=Fix(R)L=\operatorname{Fix}(R) is fibered by Gromov’s foliations 1\mathcal{F}_{1} and 2\mathcal{F}_{2} associated to some J𝒥RJ\in\mathcal{J}_{R}. More precisely, we have the following.

Corollary 3.3.

Let L=Fix(R)L=\operatorname{Fix}(R) be a real Lagrangian torus in S2×S2S^{2}\times S^{2} for an antisymplectic involution RR. Then there exist J𝒥RJ\in\mathcal{J}_{R} and a leaf Σ1\Sigma\in\mathcal{F}_{1} such that LL is fibered by 2\mathcal{F}_{2} and Σ\Sigma, that is,

  • Σ\Sigma is transverse to the leaves of 2\mathcal{F}_{2}.

  • Σ\Sigma is disjoint from LL.

  • The leaves of 2\mathcal{F}_{2} intersect LL in a circle or not at all.

Here, 1\mathcal{F}_{1} and 2\mathcal{F}_{2} denote Gromov’s foliations associated to JJ.

Proof.

By Proposition 2.6, we can choose an embedded JJ-holomorphic sphere Σ\Sigma which represents the homology class A1A_{1} and is disjoint from LL. It follows from positivity of intersections that Σ\Sigma is transverse to the leaves of 2\mathcal{F}_{2}. By Lemma 3.1, we see that the leaves of \mathcal{F} intersect LL in a circle or not at all. Since JJ-holomorphic spheres are symplectic, the corollary follows. ∎

We are now in position to prove the main theorem.

Proof of the Main Theorem.

Applying Corollary 3.3, choose J𝒥RJ\in\mathcal{J}_{R} and a leaf Σ1\Sigma\in\mathcal{F}_{1} such that the real Lagrangian torus L=Fix(R)L=\operatorname{Fix}(R) is fibered by 2\mathcal{F}_{2} and Σ\Sigma. We parametrize Σ\Sigma by an embedded JJ-holomorphic sphere u:S2S2×S2u\colon S^{2}\to S^{2}\times S^{2} in the homology class A1A_{1}. Consider another embedded JJ-holomorphic sphere u:=Ruρu^{\prime}:=R\circ u\circ\rho which is disjoint from LL as well. By Lemma 2.2, uu^{\prime} also represents the homology class A1A_{1}. We write Σ=Imu\Sigma^{\prime}=\operatorname{Im}u^{\prime} for its image. By positivity of intersections, we know that

  • Σ\Sigma^{\prime} is transverse to the leaves of 2\mathcal{F}_{2}.

  • We have Σ=Σ\Sigma=\Sigma^{\prime} or ΣΣ=\Sigma\cap\Sigma^{\prime}=\emptyset.

For each leaf FF of 2\mathcal{F}_{2} intersecting LL, the antisymplectic involution RR of S2×S2S^{2}\times S^{2} restricts to the orientation-reversing involution τ\tau on the leaf FS2F\cong S^{2} with fixed point set Fix(τ)=LFS1\operatorname{Fix}(\tau)=L\cap F\cong S^{1}. The embedded loop Fix(τ)\operatorname{Fix}(\tau) cuts FF into two closed discs glued along Fix(τ)\operatorname{Fix}(\tau). Since τ\tau must interchange the two discs by Remark 3.2, we deduce that ΣF\Sigma\cap F and ΣF\Sigma^{\prime}\cap F are disjoint. Hence, Σ\Sigma and Σ\Sigma^{\prime} must be disjoint. Moreover, the second symplectic sphere Σ\Sigma^{\prime} is disjoint from the solid torus TT which is the union of one of the two discs of each leaf intersecting Σ\Sigma. Now we can apply the Cieliebak–Schwingenheuer criterion (Theorem 2.11) to complete the proof. ∎

Acknowledgement

This paper came out of ideas I learned from Kai Cieliebak. The author cordially thanks Kai Cieliebak for invaluable comments, and Felix Schlenk and Urs Frauenfelder for fruitful discussions. The author also deeply thanks the anonymous referee who pointed out a mistake in the original proof of the main theorem. This work is supported by Samsung Science and Technology Foundation under Project Number SSTF-BA1901-01. Finally, the author always thanks KIAS for providing a great place to conduct research.

References

  • [1] V. I. Arnol’d. The first steps of symplectic topology. Uspekhi Mat. Nauk, 41(6(252)):3–18, 229, 1986.
  • [2] M. Audin. Fibrés normaux d’immersions en dimension double, points doubles d’immersions lagragiennes et plongements totalement réels. Comment. Math. Helv., 63(4):593–623, 1988.
  • [3] D. Auroux. Infinitely many monotone Lagrangian tori in 6\mathbb{R}^{6}. Invent. Math., 201(3):909–924, 2015.
  • [4] M. S. Borman, T.-J. Li, and W. Wu. Spherical Lagrangians via ball packings and symplectic cutting. Selecta Math. (N.S.), 20(1):261–283, 2014.
  • [5] F. Bourgeois, Y. Eliashberg, H. Hofer, K. Wysocki, and E. Zehnder. Compactness results in symplectic field theory. Geom. Topol., 7:799–888, 2003.
  • [6] J. Brendel. Real Lagrangian Tori and Versal Deformations. arXiv:2002.03696, 2020.
  • [7] J. Brendel, J. Kim, and J. Moon. On the topology of real Lagrangians in toric symplectic manifolds. arXiv:1912.10470, 2019.
  • [8] Y. Chekanov and F. Schlenk. Notes on monotone Lagrangian twist tori. Electron. Res. Announc. Math. Sci., 17:104–121, 2010.
  • [9] Y. V. Chekanov. Lagrangian tori in a symplectic vector space and global symplectomorphisms. Math. Z., 223(4):547–559, 1996.
  • [10] K. Cieliebak and K. Mohnke. Compactness for punctured holomorphic curves. J. Symplectic Geom., 3(4):589–654, 2005. Conference on Symplectic Topology.
  • [11] K. Cieliebak and K. Mohnke. Punctured holomorphic curves and Lagrangian embeddings. Invent. Math., 212(1):213–295, 2018.
  • [12] K. Cieliebak and M. Schwingenheuer. Hamiltonian unknottedness of certain monotone Lagrangian tori in S2×S2S^{2}\times S^{2}. Pacific J. Math., 299(2):427–468, 2019.
  • [13] A. Constantin and B. Kolev. The theorem of Kerékjártó on periodic homeomorphisms of the disc and the sphere. Enseign. Math. (2), 40(3-4):193–204, 1994.
  • [14] A. Degtyarev, I. Itenberg, and V. Kharlamov. Real Enriques surfaces, volume 1746 of Lecture Notes in Mathematics. Springer-Verlag, Berlin, 2000.
  • [15] A. Degtyarev and V. Kharlamov. Real rational surfaces are quasi-simple. J. Reine Angew. Math., 551:87–99, 2002.
  • [16] A. I. Degtyarev and V. M. Kharlamov. Topological properties of real algebraic varieties: Rokhlin’s way. Uspekhi Mat. Nauk, 55(4(334)):129–212, 2000.
  • [17] G. Dimitroglou Rizell and J. D. Evans. Unlinking and unknottedness of monotone Lagrangian submanifolds. Geom. Topol., 18(2):997–1034, 2014.
  • [18] G. Dimitroglou Rizell, E. Goodman, and A. Ivrii. Lagrangian isotopy of tori in S2×S2S^{2}\times S^{2} and P2\mathbb{C}P^{2}. Geom. Funct. Anal., 26(5):1297–1358, 2016.
  • [19] Y. Eliashberg, A. Givental, and H. Hofer. Introduction to symplectic field theory. Number Special Volume, Part II, pages 560–673. 2000. GAFA 2000 (Tel Aviv, 1999).
  • [20] Y. Eliashberg and L. Polterovich. Unknottedness of Lagrangian surfaces in symplectic 44-manifolds. Internat. Math. Res. Notices, (11):295–301, 1993.
  • [21] Y. Eliashberg and L. Polterovich. Local Lagrangian 22-knots are trivial. Ann. of Math. (2), 144(1):61–76, 1996.
  • [22] Y. Eliashberg and L. Polterovich. The problem of Lagrangian knots in four-manifolds. In Geometric topology (Athens, GA, 1993), volume 2 of AMS/IP Stud. Adv. Math., pages 313–327. Amer. Math. Soc., Providence, RI, 1997.
  • [23] M. Entov and L. Polterovich. Rigid subsets of symplectic manifolds. Compos. Math., 145(3):773–826, 2009.
  • [24] J. D. Evans. Lagrangian spheres in del Pezzo surfaces. J. Topol., 3(1):181–227, 2010.
  • [25] M. Gromov. Pseudo holomorphic curves in symplectic manifolds. Invent. Math., 82(2):307–347, 1985.
  • [26] R. Hind. Lagrangian spheres in S2×S2S^{2}\times S^{2}. Geom. Funct. Anal., 14(2):303–318, 2004.
  • [27] A. Ivrii. Lagrangian unknottedness of tori in certain symplectic 4-manifolds. 2003. Thesis (Ph.D.)–Stanford University.
  • [28] V. Kharlamov. Topology, moduli and automorphisms of real algebraic surfaces. Milan J. Math., 70:25–37, 2002.
  • [29] J. Kim. The Chekanov torus in S2×S2{S}^{2}\times{S}^{2} is not real. to appear in J. Symplectic Geom.
  • [30] J. Kim. Uniqueness of Real Lagrangians up to Cobordism. Int. Math. Res. Not. IMRN, online ready. rnz345.
  • [31] F. Lalonde and D. McDuff. The classification of ruled symplectic 44-manifolds. Math. Res. Lett., 3(6):769–778, 1996.
  • [32] D. McDuff. The structure of rational and ruled symplectic 44-manifolds. J. Amer. Math. Soc., 3(3):679–712, 1990.
  • [33] D. McDuff and D. Salamon. JJ-holomorphic curves and symplectic topology, volume 52 of American Mathematical Society Colloquium Publications. American Mathematical Society, Providence, RI, second edition, 2012.
  • [34] D. McDuff and D. Salamon. Introduction to symplectic topology. Oxford Graduate Texts in Mathematics. Oxford University Press, Oxford, third edition, 2017.
  • [35] K. R. Meyer. Hamiltonian systems with a discrete symmetry. J. Differential Equations, 41(2):228–238, 1981.
  • [36] J. Nelson. Automatic transversality in contact homology I: regularity. Abh. Math. Semin. Univ. Hambg., 85(2):125–179, 2015.
  • [37] L. V. Polterovich. The Maslov class of the Lagrange surfaces and Gromov’s pseudo-holomorphic curves. Trans. Amer. Math. Soc., 325(1):241–248, 1991.
  • [38] P. Seidel. Lectures on four-dimensional Dehn twists. In Symplectic 4-manifolds and algebraic surfaces, volume 1938 of Lecture Notes in Math., pages 231–267. Springer, Berlin, 2008.
  • [39] R. Vianna. Infinitely many monotone Lagrangian tori in del Pezzo surfaces. Selecta Math. (N.S.), 23(3):1955–1996, 2017.
  • [40] C. Viterbo. A new obstruction to embedding Lagrangian tori. Invent. Math., 100(2):301–320, 1990.
  • [41] J.-Y. Welschinger. Invariants of real symplectic 4-manifolds and lower bounds in real enumerative geometry. Invent. Math., 162(1):195–234, 2005.
  • [42] J.-Y. Welschinger. Effective classes and Lagrangian tori in symplectic four-manifolds. J. Symplectic Geom., 5(1):9–18, 2007.