<<

Symplectic and -valued measures

Adi Dickstein1,3, Yaniv Ganor2, Leonid Polterovich3, and Frol Zapolsky

Abstract We adapt Gromov’s notion of ideal-valued measures to symplectic topology, and use it for proving new results on symplectic rigidity and symplectic inter- sections. Furthermore, it enables us to discuss three “big fiber theorems”, the Centerpoint Theorem in combinatorial , the Maximal Fiber Inequality in topology, and the Non-displaceable Fiber Theorem in symplectic topology, from a unified viewpoint. Our main technical tool is an enhancement of the symplectic cohomology theory recently developed by Varolgunes.

Contents

1 Introduction and main results 2 1.1 Three big fiber theorems ...... 2 1.2 Gromov’s ideal-valued measures: a review ...... 4 1.3 Symplectic ideal-valued quasi-measures ...... 10 1.4 Quantitative non-displaceable fiber theorem ...... 14 1.5 Central fibers of involutive maps and symplectic rigidity ...... 16 1.6 SH-heavy sets ...... 17 1.7 Quantum IVQM versus cohomology IVM ...... 20 1.8 Discussion ...... 20

2 Relative symplectic cohomology of pairs 23 arXiv:2107.10012v1 [math.SG] 21 Jul 2021 2.1 Axioms for relative symplectic cohomology of pairs ...... 24 2.2 Discussion ...... 26

3 Proof of Main Theorem 1.34 28 1)Partially supported by the Milner Foundation 2)Partially supported by the Israel Science Foundation grant 1715/18 3)Partially supported by the Israel Science Foundation grant 1102/20

1 4 Proofs of properties of symplectic cohomology of pairs 33 4.1 Completion for Λ≥0-modules ...... 34 4.2 Cubes ...... 34 4.3 Direct sums and tensor products ...... 36 4.4 Cones, cocones ...... 38 4.5 Compositions of maps of cubes ...... 41 4.6 Folding and pullbacks of V-shaped diagrams ...... 43 4.7 Rays and telescopes ...... 44 4.8 Floer data, rays, and symplectic cohomology of pairs ...... 47 4.9 Constructing the product ...... 51 4.10 Well-definedness and properties of SH(K,K0) ...... 58

5 Symplectic rigidity and computations 65 5.1 Proof of Theorem 1.43 ...... 65 5.2 Proof of Theorem 1.48 ...... 68 5.3 Proof of Theorem 1.55 ...... 68 5.4 Proof of Theorem 2.4 ...... 76 5.5 Proof of Theorem 1.54 ...... 77 5.6 Proof of Theorem 1.52 ...... 79

1 Introduction and main results

1.1 Three big fiber theorems In various fields of there exist “big fiber” theorems, which are of the following type:

−1 For any f: X → Y in a suitable class there is y0 ∈ Y such that the fiber f (y0) is big.

The “suitable class” and “big” have different interpretations in different fields. Here we will present three results which exemplify this principle.

Theorem 1.1 (Maximal fiber inequality, Gromov [10, p.758], [11, p.425]). Let Y be a of covering d, and let p, n be positive such that n n ≥ p(d + 1). Then for any continuous map f: T → Y there is y0 ∈ Y such that ˇ ∗ n ˇ ∗ −1  p rk H (T ) → H (f (y0)) ≥ 2 .

2 Here Hˇ ∗ stands for Cechˇ cohomology with field coefficients. The suitable class consists of continuous maps into metric spaces of a given covering dimension and the result is that there is a “big” fiber, namely the restriction map from the cohomology of the ambient space Tn to that of the fiber has large rank. Theorem 1.2 (Topological centerpoint theorem, Karasev, [12, Theorem 5.1]). Let Y be a metric space of covering dimension d, let p be a positive , let n = p(d + 1) n n and let ∆ be the n-simplex. Then for any continuous map f: ∆ → Y there is y0 ∈ Y −1 n such that f (y0) intersects all the pd-dimensional faces of ∆ . Here we have continuous maps from simplexes into metric spaces of a given covering dimension, while a fiber is “big” if it intersects all the high-dimensional faces of the simplex. The affine version of this theorem can be found in [5], a classical result proved in a slightly different language by Rado, 1946 [19]. Our final sample result belongs to the field of symplectic topology. Recall that a symplectic is a pair (M, ω) where M is a manifold and ω is a closed ∧ 1 dim M 2-form on M which is also nondegenerate, meaning that ω 2 is a volume form. ∞ Given f ∈ C (M) its Hamiltonian vector field Xf is uniquely determined by the ∞ equation ιXf ω = −df. The Poisson bracket of f, g ∈ C (M) is the {f, g} = ∞ −ω(Xf ,Xg) = df(Xg); f, g Poisson commute if {f, g} = 0. If f ∈ C (M × [0, 1]) is a t time-dependent function, then integrating the Hamiltonian vector field Xf of ft ≡ f(·, t) t 1 yields a Hamiltonian isotopy φf . The set of φf for all such f is the Hamiltonian Ham(M, ω) of (M, ω). A set S ⊂ M is displaceable if there is φ ∈ Ham(M, ω) such that φ(S) ∩ S = ∅, and non-displaceable otherwise. Let us describe the suitable class of maps defined on symplectic . Definition 1.3. Let (M, ω) be a symplectic manifold and let B be a smooth mani- fold. We call a smooth map π: M → B involutive if for all f, g ∈ C∞(B) we have {π∗f, π∗g} ≡ 0. Theorem 1.4 (Non-displaceable fiber theorem, Entov–Polterovich, [6]). Let (M, ω) be a closed symplectic manifold. Then for any involutive map π: M → B −1 there is b0 ∈ B such that π (b0) is non-displaceable. Non-displaceable sets in a symplectic manifold are “big,” so this can be interpreted as a “big fiber” theorem in symplectic topology.

Concepts used in proofs of the above theorems. Both Gromov’s maximal fiber inequality and Karasev’s topological centerpoint theorem can be proved using Gromov’s notion of ideal-valued measures coming from Cechˇ cohomology. Ideal-valued measures are the subject of Definition 1.6, and they are used to prove the aforementioned results

3 in 1.2. By contrast, Entov–Polterovich’s result is proved using partial symplectic quasi-states defined using Floer , which are seemingly unrelated concepts. Our idea is to unify the two approaches, using a generalization of ideal-valued mea- sures to what we call symplectic ideal-valued quasi-measures (Definition 1.24), defined on symplectic manifolds using U. Varolgunes’s relative symplectic cohomology [23]. Armed with these, we

• Refine the non-displaceable fiber theorem, see Theorem 1.37 and Corollary 1.41;

• Prove a symplectic version of the centerpoint theorem, see Theorem 1.42, and use it to produce a new example of symplectic rigidity, Theorem 1.43;

• Define SH-heavy subsets of a symplectic manifold, which are a variant of Entov– Polterovich’s notion of heavy subsets [7], and provide a simple algebraic criterion which guarantees that two SH-heavy sets intersect, see Definition 1.45 and Propo- sition 1.47. In Section 1.6 we adress a question about the connection between SH-heavy and heavy sets and prove that under certain assumptions heavy sets are necessarily SH-heavy.

Remark 1.5. In [7] it is shown that a heavy set is non-displaceable, but beyond that it was unclear when two heavy sets should intersect. For instance, if L, L0 ⊂ T2 are meridians, then both of them are heavy, but they may or may not be displaceable from one another. The two situations are when L, L0 lie in distinct homology classes versus when they represent the same homology class and have zero geometric intersection number. Using our notion of SH-heavy sets and the intersection criterion, we are able to distinguish between the two situations, see Example 1.49.

1.2 Gromov’s ideal-valued measures: a review In this section we review Gromov’s notion of ideal-valued measures, provide exam- ples, and indicate how to use them to prove Theorems 1.1 and 1.2. L i An algebra (A, ∗) is called graded if it decomposes as a A = i A of i j i+j graded components, such that A ∗A ⊂ A , where we take the grading in Z or Zk, and in the latter case assume that k is even. We say that A is skew-commutative if for homogeneous a, b ∈ A we have ab = (−1)|a||b|ba, where | · | stands for the graded degree. If the ground field has characteristic 2, skew-commutativity is equivalent to commu- tativity. In the rest of the paper by an algebra we mean a graded skew-commutative associative unital algebra. A typical example is the cohomology of a space.

4 For future use note that if A is a Z-graded algebra, then we can produce a Zk-graded algebra B as follows: [i] M j B = A for [i] ∈ Zk . j≡i mod k We denote this algebra by A∗ mod k and call it the mod k regrading of A. A graded ideal in A is an ideal I ⊂ A which decomposes as the direct sum of its L i graded components, that is I = i(I ∩ A ). Equivalently, I is the kernel of a graded algebra morphism A → B. Note that in skew-commutative algebras left, right, and two-sided ideals are equivalent notions. We let I(A) be the collection of graded ideals of A. We say that A is graded Noetherian if every ascending sequence of graded ideals stabilizes. This is the case, for instance, if A is Noetherian, in particular if it is finite-dimenisional.

Definition 1.6. (Gromov, [11, Section 4.1]) Let X be a and let A be an algebra. An A-ideal-valued measure (an A-IVM) is an assignment U 7→ µ(U) ∈ I(A), where U ⊂ X runs over open sets, such that the following properties hold:

(i) (normalization): µ(∅) = 0 and µ(X) = A; (ii) (monotonicity): µ(U) ⊂ µ(U 0) if U ⊂ U 0; S S (iii) (continuity): if U1 ⊂ U2 ⊂ ... and U = i Ui, then µ(U) = i µ(Ui); (iv) (additivity): µ(U ∪ U 0) = µ(U) + µ(U 0) for disjoint U, U 0;

(v) (multiplicativity): µ(U) ∗ µ(U 0) ⊂ µ(U ∩ U 0);

(vi) (intersection): if U, U 0 cover X, then µ(U ∩ U 0) = µ(U) ∩ µ(U 0).1)

Remark 1.7. The condition µ(X) = A is called fullness in [11], however we opted to include it as part of the normalization condition because we will only use IVMs which satisfy it. Also, in Gromov’s definition additivity and intersection are stated for countably many sets.

Remark 1.8. Given an A-IVM µ on a compact Hausdorff space we will use its natural extension to compact sets defined by \ µ(K) = µ(U) . U open⊃K

1)This only holds for covers of X by two subsets. The generalization to multiple sets is as follows: Q T  if U1,...,Uk ⊂ X and for each i 6= j, Ui ∪ Uj = X, then i µ(Ui) = µ i Ui .

5 This extended function then satisfies the analogs of the monotonicity, additivity, mul- tiplicativity, and intersection properties. Note that the values of µ on open sets are S recoverable from those on compact sets via µ(U) = K compact⊂U µ(K). Example 1.9. Given any algebra A and a compact connected Hausdorff space X, the trivial A-IVM on X is given by µ(U) = 0 for all U ( X and µ(X) = A. Example 1.10. Here we describe two fundamental examples of IVMs: the Cechˇ coho- mology IVM and the singular cohomology IVM. • Let Hˇ ∗ denote Cechˇ cohomology with coefficients in a field. Letting X be a compact Hausdorff space and A = Hˇ ∗(X) and putting µ(U) = ker Hˇ ∗(X) → Hˇ ∗(X \ U) for open U ⊂ X yields an IVM, called the Cechˇ cohomology IVM (Gromov, [11, Section 4.1]). Note that A is Z-graded. • In this example H∗ stands for singular cohomology with fixed field coefficients. We will describe the singular cohomology IVM on a compact Hausdorff space X. Let A = H∗(X) and note that A is likewise Z-graded. The idea is to take the construction of the previous example and regularize it: for compact K ⊂ X we put \ µ(K) = ker H∗(X) → H∗(X \ U) , U open⊃K and for open U ⊂ X we let [ µ(U) = µ(K) . K compact⊂U

Remark 1.11. • Regularization refers to the standard approximation of compact sets by open sets and of open sets by compacts.

• The reason we first define the singular cohomology IVM for compact sets is to make it similar to our definition of ideal-valued quasi-measures based on Varol- gunes’s relative symplectic cohomology, where we must use restriction maps to compact sets, see Definition 1.32.

• If X is a closed manifold, then the singular cohomology and the Cechˇ cohomology IVM coincide, because singular cohomology and Cechˇ cohomology coincide on zero compact submanifolds of X with boundary, and any compact set K can be approximated by such submanifolds containing K in their interior. Henceforth we will refer to either IVM on X as the cohomology IVM, and denote it by µcoh.

6 IVMs provide a conceptual framework in which to prove Karasev’s topological cen- terpoint theorem 1.2. We start with the following result which will also be used for the proof of Gromov’s version of the centerpoint theorem 1.21. Theorem 1.12. Let Y be a compact metric space of covering dimension d, let A be a graded Noetherian algebra, and let I ∈ I(A) be a graded ideal with Id+1 6= 0. For every A-IVM ν on Y , there exists a point y0 ∈ Y such that I 6⊂ ν(Y \{y0}). Proof. Assume on the contrary that I ⊂ ν(Y \{y}) for all y ∈ Y . By continuity we S 1 have for each y ∈ Y : ν(Y \{y}) = n≥1 ν(Y \ By( n )), which by the graded Noetherian property of A implies I ⊂ ν(Y \ B ( 1 )) for some n ∈ . Since Y has covering y ny y N dimension d, and by the Palais lemma (see [12]), the open cover (B ( 1 )) admits a y ny y∈Y 0 S finite refinement {Vij}i,j where i = 0, . . . , d and Vij ∩ Vij0 = ∅ if j 6= j . Let Yi = j Vij. c c c 0 Note that I ⊂ ν(Vij) by monotonicity. Since Y = Vij ∪ Vij0 for any j 6= j , by the intersection property we have

c  T c T c ν(Yi ) = ν j Vij = j ν(Vij) ⊃ I, and by multiplicativity, normalization and monotonicity we obtain

d+1 Qd c Td c 0 6= I ⊂ i=0 ν(Yi ) ⊂ ν i=0 Yi = ν(∅) = 0 , which is a contradiction.

We call such y0 a centerpoint of the IVM ν with respect to the ideal I. This terminology is motivated by the following result.

Corollary 1.13. The point y0 belongs to every closed subset Z ⊂ Y with I ⊂ ν(Z).

Proof. Otherwise Z ⊂ Y \{y0}, and hence I ⊂ ν(Y \{y0}), a contradiction. We will derive the following from Corollary 1.13: Corollary 1.14. Let X be a compact Hausdorff space, let A be an algebra, and let µ be an A-IVM. Let I ∈ I(A) be a graded ideal such that Id+1 6= 0 for some d ≥ 1. Consider the collection XI = {Z ⊂ X | Z compact with I ⊂ µ(Z)} . Let Y be a metric space of covering dimension d. Then any continuous map f: X → Y has a fiber which intersects all the members of XI . We call such a fiber a central fiber of f with respect to I. To deduce Corollary 1.14 from Corollary 1.13, we need the notion of pushforward for IVMs.

7 Definition 1.15. Let A be an algebra, let f: X → Y be a continuous map, and let µ be an A-IVM on X. The pushforward of µ by f is the set function f∗µ defined on −1 the open sets of Y by f∗µ(V ) = µ(f (V )).

Remark 1.16. • The pushforward construction makes sense for any I(A)-valued function defined on open or closed subsets of X, not necessarily an A-IVM. We will use this below.

• It is easy to see that f∗µ is an A-IVM on Y . If X,Y are Hausdorff and X is in addition compact, then the extension of f∗µ to the compact subsets of Y , as described in Remark 1.8, coincides with the pushforward of the extension of µ to the compact subsets of X.

Proof of Corollary 1.14. Without loss of generality assume that Y is compact and that f is onto. Applying Corollary 1.13 to the A-IVM ν = f∗µ, we obtain a point y0 contained 0 0 −1 in any compact Y ⊂ Y with I ⊂ ν(Y ). If Z ∈ XI , then, since Z ⊂ f (f(Z)), we have

I ⊂ µ(Z) ⊂ µ(f −1(f(Z)) = ν(f(Z)) , therefore y0 ∈ f(Z), as claimed. Remark 1.17. It is not hard to prove that Corollaries 1.13, 1.14 hold for general A.

The topological centerpoint theorem readily follows from this corollary combined with a trick from toric geometry appearing in Karasev’s original proof [12].

Proof of Theorem 1.2. Consider a continuous map f: ∆n → Y , where Y is a metric space of covering dimension d and n = p(d + 1). Consider the standard toric moment map Φ: CP n → ∆n. Let µ be the cohomology IVM on CP n and consider its pushforward 2 n p ν = Φ∗µ. Let h ∈ H (CP ) be a generator and consider the graded ideal I = hh i. We claim that the collection of pd-dimensional faces of ∆n is contained in the col- n lection XI of the subsets Z ⊂ ∆ with I ⊂ ν(Z). Indeed, if X is such a face, then Φ−1(X) is a projective subspace of CP n of complex dimension pd, and as such it is Poincar´edual to hp, which implies that for every open neighborhood U of Φ−1(X) we p p −1 have h |CP n\U = 0. It follows that h ∈ µ(Φ (X)) = ν(X), that is I ⊂ ν(X). Corollary −1 1.14 now yields a point y0 ∈ Y such that f (y0) intersects all the members of XI , in particular all the pd-dimensional faces of ∆n.

Similarly, Gromov’s Theorem 1.1 can be proved using a general result about IVMs, Theorem 1.21, also due to Gromov [11, Section 4.2]. To state it, we need to define multiplicative ranks of algebras, defined ibid.

8 Definition 1.18. Let A be a finite-dimensional algebra. For r ≥ 1 put \ A/r = {I ∈ I(A) | dim A/I < r} .

For d ≥ 1, the d-rank of A, denoted rkd A, is defined as the maximal r for which the restriction of the d-fold product map A⊗d → A to (A/r)⊗d does not vanish.

/1 Example 1.19. Note that A = A. If A 6= 0, this implies that rkd A ≥ 1 for all d. / dim A+1 Also note that A = 0, therefore rkd A ≤ dim A for all d. Example 1.20. ([11, Section 4.1]) If X is a closed oriented manifold, then Poincar´e implies that any nonzero graded ideal in H∗(X) must contain the orientation ∗ class [X]. If X1,...,Xd are closed oriented manifolds such that dim H (Xi) ≥ m for all Qd i, and X = i=1 Xi, K¨unneth’sformula implies that

d ∗ O ∗ H (X) = H (Xi) i=1 as graded skew-commutative algebras. The natural inclusion map

∗ ∗ ⊗(i−1) ⊗(d−i) ιi: H (Xi) → H (X) , a 7→ 1 ⊗ a ⊗ 1 is a graded algebra morphism. If K ⊂ H∗(X) is a graded ideal of codimension < m, −1 ∗ then ιi (K) is a nonzero graded ideal in H (Xi), which then contains [Xi] by the ⊗(i−1) ⊗(d−i) above. It follows that K contains ιi([Xi]) = 1 ⊗ [Xi] ⊗ 1 , and therefore so does H∗(X)/m. Since

d Y ιi([Xi]) = [X1] ⊗ · · · ⊗ [Xd] = [X] 6= 0 , i=1 it follows that ∗ ∗ rkd H (X) ≥ min dim H (Xi) . i If n ≥ p(d + 1), then for the Tn = (Tp)d × Tn−pd, we obtain n ∗ p p rkd+1 T ≥ dim H (T ) = 2 . Gromov’s centerpoint theorem now reads as follows. Theorem 1.21 (Gromov, [11]). Let Y be a compact metric space of covering dimension d, let A be a finite-dimensional algebra, and let µ be an A-IVM on X. Then there is y0 ∈ Y such that  dim A/µ Y \{y0} ≥ rkd+1 A.

9 /r d+1 Proof. Put r = rkd+1(A). By definition, (A ) 6= 0, therefore by Theorem 1.12, for /r /r some y0, A 6⊂ µ(Y \{y0}). Thus dim A/µ(Y \{y0}) ≥ r by the definition of A . This result implies Gromov’s Theorem 1.1.

Proof of Theorem 1.1. Recall that we have a map f: Tn → Y , where Y is a metric space of covering dimension d and n ≥ p(d + 1). Without loss of generality assume that n f is onto and that Y is compact. Let µ be the cohomology IVM on T and let ν = f∗µ. Theorem 1.21 implies that there is y0 ∈ Y such that

ˇ ∗ n ˇ ∗ n dim H (T )/ν(Y \{y0}) ≥ rkd+1 H (T ) .

ˇ ∗ n p Example 1.20 implies that rkd+1 H (T ) ≥ 2 , therefore ˇ ∗ n n −1 p dim H (T )/µ(T \ f (y0)) ≥ 2 . By the definition in Example 1.10 we have

n −1 ˇ ∗ n ˇ ∗ −1  µ(T \ f (y0)) = ker H (T ) → H (f (y0)) , therefore

ˇ ∗ n ˇ ∗ −1  ˇ ∗ n ˇ ∗ n ˇ ∗ −1  p rk H (T ) → H (f (y0)) = dim H (T ) − dim ker H (T ) → H (f (y0)) ≥ 2 .

1.3 Symplectic ideal-valued quasi-measures Here we present the main contribution of the present paper, symplectic ideal-valued quasi-measures, which are a suitable generalization of IVMs for symplectic manifolds.

Definition 1.22. Let (M, ω) be a closed symplectic manifold. We say that two compact subsets K,K0 ⊂ M commute if there are Poisson-commuting f, f 0 ∈ C∞(M, [0, 1]) such that K = f −1(0),K0 = f 0−1(0). Two open sets commute if their complements commute.

Example 1.23. If M is closed and π: M → B is an involutive map, then preimages of closed sets commute, as do preimages of open sets. In particular if f, f 0 are as in the definition, then for any s, s0 ∈ [0, 1] the sets f −1([0, s]), f 0−1([0, s0]) commute, and so do the sets f −1([0, s)), f 0−1([0, s0)).

In what follows, Symp0(M, ω) stands for the identity component of the group of symplectomorphisms of (M, ω).

10 Definition 1.24. Fix an algebra A.A symplectic A-ideal-valued quasi-measure (symplectic A-IVQM) on a closed symplectic manifold (M, ω) is an assignment U 7→ τ(U) ∈ I(A), where U ranges over open subsets of M, such that τ satisfies all of the properties of an A-IVM, with the exception of the multiplicativity axiom, which is replaced by the weaker

• (quasi-multiplicativity): τ(U) ∗ τ(U 0) ⊂ τ(U ∩ U 0) whenever U, U 0 commute. Definition 1.25. For a symplectic A-IVQM τ we define two further properties:

• (invariance): if φ ∈ Symp0(M, ω), then τ(φ(U)) = τ(U) for all U. • (vanishing): if K ⊂ M is a displaceable compact set, then τ(M − K) = A, and there is an open U such that K ⊂ U and τ(U) = 0.

Remark 1.26. • Any A-IVM on M is also a symplectic A-IVQM.

• For brevity, from this point on we will refer to symplectic IVQMs simply as IVQMs.

• We only construct and use A-IVQMs satisfying invariance and vanishing, therefore henceforth all IVQMs are assumed to have these two properties.

• Analogously to IVMs, we extend an A-IVQM to compact sets by \ τ(K) = τ(U) . U open⊃K This satisfies the analogs of the monotonicity, additivity, quasi-multiplicativity, intersection, and invariance properties. Thanks to the second half of the vanishing property, if K is a displaceable compact set, then the extension satsifes τ(K) = 0. If A is in addition graded Noetherian, then the second half of the vanishing property can be restated as follows: if K is displaceable, then τ(K) = 0.

• In the concrete examples of IVQMs we will construct the quasi-multiplicativity actually holds for more general pairs of subsets, see Sections 2.1, 4.8.3, but the stated property suffices for applications. Remark 1.27. The trivial A-IVM (see Example 1.9 above) satisfies the invariance property. If A 6= 0 and M has positive dimension, then the trivial A-IVM does not satisfy the vanishing property. Theorem 1.28 (Main theorem). Every closed symplectic manifold carries an IVQM with respect to some nonzero finite-dimensional algebra over a suitable field.

11 Example 1.29. Let A be an algebra. Take M = S2 and let ω be the standard area form on M normalized to have area 1. If U ⊂ S2 is an open subset, put τ(U) = 0 if each 1 connected component of U is contained in an open disk of area ≤ 2 , and let τ(U) = A otherwise. Then τ is an A-IVQM. Our results are based on a specific IVQM coming from Varolgunes’s relative symplec- tic cohomology [23], which we now briefly review. Fix a base field F. The corresponding Novikov field is ( ∞ )

X λi Λ = ciT ci ∈ , 3 λi −−−→ ∞ , F R i→∞ i=0 and the Novikov ring is the subring

( ∞ ) X λi Λ≥0 = ciT ∈ Λ ∀i : λi ≥ 0 . i=0 The of M with coefficients in Λ is additively the singular cohomology QH∗(M) = H∗ mod 2NM (M; Λ) regraded 2NM , where NM is the minimal Chern number of M. The product operation on QH∗(M) is given by the quantum product [14], where the matrix coefficient E(u) R ∗ corresponding to a holomorphic sphere u is multiplied by T , where E(u) = S2 u ω is its symplectic area. Given a compact set K ⊂ M the relative symplectic cohomology of K inside M is   SH∗(K; Λ) = H∗ lim CF ∗(H ) ⊗ Λ , (1) −→di i Λ≥0 where Hi is a pointwise increasing sequence of non-degenerate time-periodic Hamilto- nians on M such that Hi|K < 0, and such that  0 , x ∈ K lim Hi(x) = ; i→∞ ∞ , x∈ / K

∗ CF (Hi) stands for the Floer complex

∗ M CF (Hi) = Λ≥0 · x , (2) ◦ x∈P (Hi)

◦ ∗ where P (Hi) is the set of contractible 1-periodic orbits of Hi. The complexes CF (Hi) are connected by Floer continuation maps. Finally, the hat stands for the completion of Λ≥0-modules. See Section 4 and [23] for more details.

12 ∗ Remark 1.30. Varolgunes uses the notation SHM to explicitly point out the symplectic manifold. We drop the subscript M throughout, trusting that the context will resolve the ambiguity.

This invariant has the following properties, among others:

∗ • Each SH (K; Λ) is a Z2NM -graded unital associative skew-commutative algebra, and SH∗(M; Λ) = QH∗(M) as an algebra [21];

0 K0 ∗ 0 ∗ • For K ⊂ K there is a restriction map resK : SH (K ; Λ) → SH (K; Λ), which 0 00 K0 K00 is a graded unital algebra morphism, and if K ⊂ K ⊂ K , then resK ◦ resK0 = K00 resK ; • The Mayer–Vietoris property: if K,K0 commute, the restriction maps fit into a long

K∪K0 K∪K0  res ,res 0 · · · → SH∗(K ∪ K0; Λ) −−−−−−−−−−−−→K K SH∗(K) ⊕ SH∗(K0) 0 resK − resK −−−−−−−−−−−→K∩K0 K∩K0 SH∗(K ∩ K0) −→+1 ...

• If K is displaceable, then SH∗(K; Λ) = 0. Remark 1.31. Unlike the original definition by Varolgunes, we only consider con- tractible periodic orbits. The reason for this is that we are only interested in kernels of restriction maps defined on SH∗(M; Λ), which on the homology level is generated by contractible orbits, and chain level restriction maps preserve the free class of periodic orbits and thus preserve the contractible component. Also, since we only consider contractible orbits, the grading can be taken modulo 2NM rather than merely modulo 2.

Definition 1.32. Let (M, ω) be a closed symplectic manifold and let A = QH∗(M) be its quantum cohomology algebra. The quantum cohomology IVQM τ is defined as follows: \ M τ(K) = ker resM\U U open⊃K for compact K ⊂ M while for open U ⊂ M we put [ τ(U) = τ(K) . K compact⊂U

13 T Remark 1.33. Note that if K is compact, then τ(K) = U open⊃K τ(U). In other words, had we used the values of τ on open sets and extended it to compact sets as in Remark 1.8, we would have recovered the values in Definition 1.32. The next result is a more explicit version of our main result, Theorem 1.28 above; it is proved in Section 3. Theorem 1.34. The quantum cohomology IVQM is a symplectic QH∗(M)-IVQM. Proof of Theorem 1.28. The quantum cohomology algebra QH∗(M) is a nonzero finite- dimensional algebra over the Novikov field Λ. The quantum cohomology IVQM τ is a symplectic QH∗(M)-IVQM, thanks to Theorem 1.34. Example 1.35 (Compare with Example 1.29). Let M = S2 with an area form ω which is normalized to have area 1. The quantum cohomology algebra of M is QH∗(M) = Λh1, hi, where 1 is the unit class while h = [ω] is the area class. The grading is modulo 2 2NS2 = 4, |1| = 0, |h| = 2. The multiplication is completely determined by h = T · 1. If K ⊂ M is a compact two-dimensional submanifold with boundary, and the com- plement of K decomposes into connected components, each of which is contained in 1 ∗ ∗ an open disk of area ≤ 2 , then SH (K; Λ) = QH (M), and the restriction map SH∗(M; Λ) → SH∗(K; Λ) is the identity. Otherwise SH∗(K; Λ) = 0. From this we can deduce that the quantum cohomology IVQM is given as follows: if U ⊂ M is open 1 and each connected component of U is contained in an open disk of area ≤ 2 , then τ(U) = 0, and τ(U) = QH∗(M) otherwise. A similar description applies to compact sets. All of this can be obtained directly from the above properties.

Let us pass to applications of IVQMs to symplectic topology.

1.4 Quantitative non-displaceable fiber theorem In this section we formulate and prove a refinement of Theorem 1.4. Definition 1.36. Let (M, ω) be a symplectic manifold and let Y be a topological space. We call a continuous map f: M → Y involutive if there is a (smooth) involutive map π: M → B and a continuous map f: B → Y such that f = f ◦ π. Note that smooth involutive maps are involutive in this generalized sense. Theorem 1.37. Let A be a finite-dimensional algebra. Let (M, ω) be a closed symplectic manifold equipped with an A-IVQM τ. Let Y be a metric space of covering dimension d. Then for any involutive map f: M → Y there is y0 ∈ Y such that

−1  dim A/τ M \ f (y0) ≥ rkd+1 A.

14 The proof appears below; it relies on Corollary 1.13 and the relation of the pushfor- ward construction, see Remark 1.16, to involutive maps. Proposition 1.38. Let (M, ω) be a closed symplectic manifold and let π: M → B be a smooth involutive map. If A is an algebra and τ is a A-IVQM on M, then π∗τ is an A-IVM on B. Proof. All the axioms of an A-IVM are satisfied automatically, except multiplicativity. If U, U 0 ⊂ B are open, then by Example 1.23 their preimages Poisson commute, therefore quasi-multiplicativity implies

0 −1 −1 0 −1 −1 0 0 π∗τ(U) ∗ π∗τ(U ) = τ(π (U)) ∗ τ(π (U )) ⊂ τ(π (U) ∩ π (U )) = π∗τ(U ∩ U ) , as required. Remark 1.39. Remark 1.16 applies here as well, that is the pushforward of the ex- tension of τ to compact subsets of M coincides with the extension to compact subsets of B of the pushforward of τ, because M and B are Hausdorff and M is in addition compact. Corollary 1.40. Let (M, ω) be a closed symplectic manifold. Then for any continuous involutive map f: M → Y the pushforward f∗τ of an A-IVQM is an A-IVM.

π f Proof. Factor f as M −→ B −→ Y , where π is a smooth involutive map. Then π∗τ is an A-IVM on B according to Proposition 1.38. Therefore f∗τ = f ∗(π∗τ) is an A-IVM on Y .

Proof of Theorem 1.37. Apply Theorem 1.21 to the A-IVM f∗τ. In the next corollary τ stands for the quantum cohomology IVQM we described above. Corollary 1.41. Let (M, ω) be a closed symplectic manifold, let Y be a metric space of covering dimension d, and let f: M → Y be a continuous involutive map. Then there is y0 ∈ Y such that

∗ −1  ∗ dimΛ QH (M)/τ M \ f (y0) ≥ rkd+1 QH (M) .

−1 In particular f (y0) is non-displaceable. Proof. This is a special case of Theorem 1.37. Since QH∗(M) is a nonzero unital finite- ∗ dimensional algebra, it follows that rkd+1 QH (M) ≥ 1 for all d, and in particular −1 ∗ −1 τ(M \ f (y0)) 6= QH (M), which by the vanishing property of τ implies that f (y0) is non-displaceable.

15 1.5 Central fibers of involutive maps and symplectic rigidity Here we extend the existence of central fibers (as in Corollary 1.14 above) to the context of involutive maps of symplectic manifolds. Let A be an algebra.

Theorem 1.42. Let (M, ω) be a closed symplectic manifold equipped with an A-IVQM τ. Let I ∈ I(A) be a graded ideal such that Id+1 6= 0 for some d ≥ 1. Consider the collection of subsets

XI = {Z ⊂ M | Z compact with I ⊂ τ(Z)} .

If Y is a metric space of covering dimension d, then any continuous involutive map f: M → Y has a fiber which intersects all the members of XI . Proof. Without loss of generality assume that Y is compact and that f is onto. Let µ = f∗τ and note that this is an A-IVM on Y thanks to Corollary 1.40. Corollary 1.13 0 0 implies that there is y0 ∈ Y contained in every compact Y ⊂ Y such that I ⊂ µ(Y ). If Z ∈ XI , then I ⊂ τ(Z) ⊂ τ(f −1(f(Z))) = µ(f(Z)) , therefore y0 ∈ f(Z), as claimed. We will now deduce a new example of symplectic rigidity from this result. Consider 6 the standard symplectic 6-torus T with coordinates pi, qi, i = 1, 2, 3 and symplectic form ω = dp ∧ dq. For a, b, c ∈ T2 consider the following coisotropic subtori:

6 T1(a) = {(p, q) ∈ T | (q1, q2) = a} , 6 T2(b) = {(p, q) ∈ T | (p1, p3) = b} , 6 T3(c) = {(p, q) ∈ T | (p2, q3) = c} . Let T (a, b, c) = T1(a) ∪ T2(b) ∪ T3(c) . In the next theorem an equator is any smoothly embedded circle in S2 dividing it into two disks of equal area.

Theorem 1.43. Let B be a surface. Then any involutive map T6 × S2 → B has a fiber which intersects every set of the form

T (a, b, c) × equator .

Remark 1.44. Let us comment on the sharpness of the various assumptions in this theorem:

16 • The dimension of B cannot be increased. Indeed, consider the involutive map 6 2 3 T × S → T ,(p, q; z) 7→ (q1, p2, p3). It maps T (a, b, c) × equator to the union of the coordinate planes {q1 = a1} ∪ {p2 = c1} ∪ {p3 = b2}, and in particular any fiber over a point not in this union is disjoint from T (a, b, c) × equator.

• The involutivity assumption is essential. Consider the non-involutive projection 6 2 2 2 2 π: T × S → S . If y0 ∈ S and L ⊂ S is an equator not containing y0, −1 2 then π (y0) is disjoint from T (a, b, c) × L, meaning that for any y0 ∈ S the corresponding fiber doesn’t meet all the above sets.

• The union of just two coisotropic tori does not work: consider the involutive map 6 2 2 2 T × S → T ,(p, q; z) 7→ (q1, p3). Then (T1(a) ∪ T2(b)) × S maps to the union of coordinate planes {q1 = a1} ∪ {p3 = b2} and thus the fibers over points not in 2 this union are disjoint from (T1(a) ∪ T2(b)) × S .

1.6 SH-heavy sets In [7] Entov–Polterovich defined a special class of compact subsets of a closed sym- plectic manifold (M, ω), the so-called heavy sets. They proved that a heavy subset is non-displaceable, however it remained unclear in general when two heavy sets should intersect. Here we address this problem to a degree. First, let us introduce the suitable class of subsets. Definition 1.45. Let (M, ω) be a closed symplectic manifold and let τ be the quantum cohomology IVQM on M. We call a compact set K ⊂ M SH-heavy if τ(K) 6= 0. The following is an immediate consequence of the vanishing property of τ: Proposition 1.46. SH-heavy sets are non-displaceable. Next we formulate an algebraic criterion which guarantees that two SH-heavy sets cannot be displaced from one another. Proposition 1.47. Let K,K0 ⊂ M be compact sets. If τ(K) ∗ τ(K0) 6= 0, then K,K0 are SH-heavy and cannot be displaced from one another by a symplectic isotopy. Proof. The assumption clearly implies that τ(K), τ(K0) 6= 0, whence the first assertion. 0 If φ ∈ Symp0(M, ω) displaces K from K, then by the invariance and multiplicativity properties we have:

0 0 0 τ(K) ∗ τ(K ) = τ(K) ∗ τ(φ(K )) ⊂ τ(K ∩ φ(K )) = τ(∅) = 0 , contradicting the assumption.

17 The following theorem provides examples of SH-heavy subsets in standard symplectic n 1 tori. Let us call a subset S ⊂ T a polyinterval if there are intervals J1,...,Jn ⊂ S such that S = J1 × · · · × Jn, where an interval can be a point or the whole circle.

Theorem 1.48. Let M = T2n = Tn(p) × Tn(q) be endowed with the symplectic form ω = dp ∧ dq, let S ⊂ Tn be a closed polyinterval, let K = S × Tn ⊂ Tn × Tn = M, and let τ be the quantum cohomology IVQM on M. Then

τ(K) = ker H∗(M; Λ) → H∗(M \ K; Λ) = ∗ n ∗ n  ∗ n ∗ n ∗ n ∗ ker H (T ; Λ) → H (T \S; Λ) ⊗H (T ; Λ) ⊂ H (T ; Λ)⊗H (T ; Λ) = H (M; Λ) .

It follows that if S 6= ∅, then in particular K is SH-heavy. The same results hold if S is an open polyinterval.

Here we used the K¨unnethformula H∗(Tn; Λ) ⊗ H∗(Tn; Λ) = H∗(Tn × Tn; Λ). This theorem is proved in Section 5.2 as a consequence of Theorem 1.54, see Section 1.7, where we also present additional computations of symplectic cohomology.

Example 1.49. Let L = {pt} × Tn(q) ⊂ T2n be a linear Lagrangian torus. Since ∗ n ∗ n  n n ker H (T ; Λ) → H (T \ pt; Λ) is spanned by [dp1 ∧ · · · ∧ dpn] ∈ H (T ; Λ), we have, according to the theorem:

∗ n ∗ 2n τ(L) = Λ · [dp1 ∧ · · · ∧ dpn] ⊗ H (T ; Λ) = H (T ; Λ) · h[dp1 ∧ · · · ∧ dpn]i .

If we take L0 = Tn(p) × {pt} ⊂ M, then using the same reasoning we obtain

0 ∗ 2n τ(L ) = H (T ; Λ) · h[dq1 ∧ · · · ∧ dqn]i . In particular we see that τ(L) ∗ τ(L0) is spanned by [ωn], therefore nonzero. By Propo- sition 1.47, L, L0 cannot be displaced from one another by a symplectic isotopy. Both L, L0 are also heavy, but the technology of [7] cannot guarantee a rigid intersection for L, L0.

It would be interesting to understand the connection between heavy and SH-heavy sets. In fact, we propose the following

Conjecture 1.50. A compact subset of a closed symplectic manifold is heavy if and only if it is SH-heavy.

Here we would like to prove that under certain assumptions, heavy sets are SH-heavy. First, recall that a symplectic manifold (M, ω) is symplectically aspherical if ω and c1(M) vanish on π2(M). Next, a cooriented hypersurface Σ ⊂ M is of contact type if

18 there exists a vector field Y defined on a neighborhood of Σ satisfying LY ω = ω, and such that along Σ, Y points everywhere in the positive direction. Note that in this case α := ιY ω|Σ is a contact form on Σ. Next we say that Σ is incompressible if the map π1(Σ) → π1(M), induced by the inclusion, is injective. If Σ is an incompressible cooriented hypersurface of contact type in a symplectically aspherical manifold (M, ω) and all the Reeb orbits of α on Σ, which are contractible in M, are nondegenerate, then we can unambiguously assign a Conley–Zehnder index to each such orbit γ, as follows. Since γ is contractible in M and Σ is incompressible, γ is contractible in Σ. Let u be a contracting disk in Σ and let ξ = ker α be the contact structure on Σ; then u∗ξ is trivializable and a trivialization allows us to assign an index to γ. Since c1(ξ) = c1(M)|Σ and c1(M)|π2(M) = 0 by the symplectic asphericity, the index is independent of the choice of the contracting disk. Definition 1.51 (See [21]). Let (M, ω) be symplectically aspherical and let Σ ⊂ M be an incompressible cooriented hypersurface of contact type, such that all the Reeb orbits, which are contractible in M, are nondegenerate. We say that Σ is index-bounded if for each k ∈ Z there is a bound on the set of periods of the Reeb orbits which are contractible in M and have Conley–Zehnder index k. Convention. For the rest of the paper by a domain we mean a compact codimension zero submanifold with boundary. If W ⊂ M is a domain, we say that W has contact-type boundary or that it is a contact-type domain if ∂W is of contact type relative to the outward coorientation. We then have: Theorem 1.52. Let (M, ω) be symplectically aspherical. Let K ⊂ M be a compact set such that there is a sequence Wi of contact-type domains with incompressible index- T bounded boundary such that K ⊂ Int Wi for each i and such that K = i Wi. If K is heavy, then for each i we have M [Vol] ∈ ker res c , Wi T M which implies that [Vol] ∈ τ(K) = i ker res c , and in particular that K is SH-heavy. Wi Here [Vol] ∈ H2n(M; Λ) is the volume class. Theorem 1.52 is proved in Section 5.6. We have the following immediate consequence. Corollary 1.53. If (M, ω) is symplectically aspherical and K is a heavy contact-type domain with incompressible index-bounded boundary, then K is SH-heavy. Proof. Let Σ = ∂K. Then there is a neighborhood of Σ which is diffeomorphic to Σ × (1 − , 1 + )r, such that the vector field ∂r points outward along Σ = Σ × {0}, and such that L∂r ω = ω. In particular Wi = K ∪ [0, /2i] are as in the theorem and the assertion follows.

19 1.7 Quantum IVQM versus cohomology IVM In this section (M, ω) stands for a closed symplectically aspherical symplectic man- ifold. Theorem 1.48 is a special case of the following result, proved in Section 5.5: Theorem 1.54. Let K ⊂ M be a domain with Σ = ∂K, such that Σ ,→ M extends to a smooth embedding (−, ) × Σ ,→ M such that no {ρ} × Σ carries closed characteristics which are contractible in M. Then SH∗(K; Λ) = H∗(K; Λ) and the restriction map SH∗(M; Λ) → SH∗(K; Λ) coincides with H∗(M; Λ) → H∗(K; Λ). We also include results regarding domains with contact-type incompressible index- bounded boundary. In Section 5.3 we will prove the following result: Theorem 1.55. Let K ⊂ M be a contact-type domain with incompressible index- bounded boundary. Then SH∗(K; Λ) is canonically isomorphic to the classical sym- plectic cohomology SH∗(Kb; Λ). Moreover ∗ ∗  M ∗ ∗ ker H (M; Λ) → H (K; Λ) ⊂ ker resK : SH (M; Λ) → SH (K; Λ) .

Here Kb stands for the completion of K, obtained by attaching to it the positive end of the symplectization of ∂K. We specify in Section 5.3.2 what exactly we mean by SH∗(Kb; Λ). Write now µ for the cohomology IVM on M (see Example 1.10 above) and τ for the quantum cohomology IVQM. As an immediate consequence we get that under assumptions of Theorems 1.54 and 1.55, τ(M \ K) ⊃ µ(M \ K).

1.8 Discussion 1.8.1 Persistence modules and symplectic cohomology Varolgunes’s symplectic cohomology is constructed as a over the Novikov ring Λ≥0, and here we have only used its torsion-free part, obtained by tensoring with the Novikov field Λ. The passage to the torsion-free part is unnecessary for many con- structions in this paper. It would be interesting to understand and apply the additional information carried by the torsion part. Novikov modules, that is modules over Λ≥0, can carry an additional structure, namely a valuation, which in turn gives rise to persistence modules (see for instance [18] for preliminaries on persistence modules in the symplectic context). All the sym- plectic cohomology modules come with a natural valuation. Some elementary examples show that the cokernel of the restriction map SH∗(M) → SH∗(M \ U) is a Novikov module which can have non-trivial torsion. The latter can be interpreted as a per- sistence module whose structure is encoded by combinatorial invariants, the barcodes

20 (see ibid.). Thus, to every subset of M there correspond a persistence module and its barcodes. Understanding the algebraic structure behind this correspondence should undoubtedly lead to new applications of symplectic cohomology.

1.8.2 Rigidity of open covers and multi-intersections In [6], Entov and Polterovich showed that no closed symplectic manifold M admits a finite cover by displaceable Poisson commuting subsets. Using symplectic IVQMs one can generalize this result as follows. Fix an algebra A and write τ for a symplectic A-IVQM on M.

Theorem 1.56. Let {Ui} be a finite open cover of M such that Y τ(M \ Ui) 6= 0 . (3) i

Then there are i 6= j such that Ui and Uj do not commute.

Note that if each Ui is displaceable, τ(M \Ui) = A by the vanishing property and hence (3) holds. For the proof of Theorem 1.56, we need the following auxiliary result.

Lemma 1.57. Let P, Q, R ⊂ M be three closed subsets, such that the pairs P,R and Q, R commute. Then P ∩ Q, R commute.

Proof. There are smooth functions p, q, r: M → R such that P = p−1(0), Q = q−1(0), R = r−1(0) and {p, r} = {q, r} = 0. Thus P ∩ Q = (p2 + q2)−1(0) and {p2 + q2, r} = 0, as required.

Proof of Theorem 1.56. Assume on the contrary that the sets Di := M \ Ui pairwise commute. Applying Lemma 1.57 along with the quasi-multiplicativity property of τ we get that  \  Y 0 = τ Di ⊃ τ(Di) 6= 0 , i and we get a contradiction.

Example 1.58. Let M be the two-dimensional torus R2/Z2 equipped with the form dp ∧ dq and let τ be the quantum cohomology IVQM on M. Consider annuli Q = {0 < q < a} and P = {0 < p < b} with a, b ∈ (0, 1). Consider the rectangle M \ (P ∪ Q), and denote by R a slightly bigger open rectangle. Clearly, P, Q, R is an open cover of M by non-commuting subsets. By Theorem 1.48 we have τ(P c) = A·[dq], τ(Qc) = A·[dq] and

21 τ(Rc) = A, where A = QH∗(M). It follows that (3) holds for P, Q, R. Therefore, by the invariance property of τ and Theorem 1.56, for every triple of symplectomorphisms f, g, h ∈ Symp0(M) the images f(P ), g(Q), h(R) cannot form a Poisson commuting cover of M.

We say that two sets P,Q in a symplectic manifold M virtually commute if there exists a symplectomorphism f ∈ Symp0(M) such that f(P ) and Q Poisson commute. Example 1.58 naturally leads to the following question.

Question 1.59. Do the annuli P,Q ⊂ T2 virtually commute? What about their stabilizations P × equator and Q × equator in T2 × S2? It is likely that using methods of two-dimensional topology one can answer the first question in the negative. The second one apparently requires new ideas. The following multi-intersection problem is a counterpart of the covering problem considered above.

Question 1.60. Given closed subsets D1,...,DN of a closed symplectic manifold (M, ω), do there exist Hamiltonian diffeomorphisms ϕ1, . . . , ϕN such that \ ϕi(Di) = ∅ ? (4) i

For N = 2, non-trivial constraints come from “hard” symplectic topology. In contrast to this, for N ≥ 3 it is quite likely that the multi-intersection problem is “soft.” Conjecture 1.61. For N ≥ 3, Question 1.60 has an affirmative answer provided there exist elements ϕi lying in the identity component of the group of volume-preserving diffeomorphisms and satisfying (4). There are two obvious necessary conditions for the existence of such volume-preserving diffeomorphisms: Q (i) (no topological obstructions) The i µcoh(Di) vanishes; P (ii) (no volume obstructions) i Vol(Di) < Vol(M). It would be interesting to understand whether these conditions are also sufficient. The techniques of [20] should be useful in proving Conjecture 1.61. However, if one additionally requires that in (4) the sets ϕi(Di) pairwise commute, Q equation (3) gives rise to a “hard” constraint i τ(Di) 6= 0. Here the product is the quantum product in QH∗(M). Note that this constraint is the symplectic counterpart of the above topological assumption (i).

22 1.8.3 Lagrangian IVQMs Fix a closed Lagrangian submanifold L ⊂ M. In [21] Tonkonog and Varolgunes pre- sented an extension of Varolgunes’s relative symplectic cohomology to the Lagrangian setting, where Hamiltonian Floer cohomology HF ∗(H) of a Hamiltonian H is replaced by Lagrangian Floer cohomology HF ∗(L, H). It is natural to expect that, arguing along the lines of the present paper, one arrives at an IVQM on M based on the Lagrangian quantum ring QH∗(L) (see a paper [1] by Biran and Cornea for a detailed introduction into this ring for monotone Lagrangian submanifolds). Note that the latter ring is in general non-commutative (even as a ), so one has to work out an extension the notion of an IVQM to non-commutative algebras. Organization of the paper: In Section 2 we introduce axioms for a relative symplectic cohomology of a pair. In Section 3 we use this notion in order to prove Main Theorem 1.34 stating that the quantum cohomology IVQM is indeed an IVQM. In Section 4 we present a construction of a symplectic cohomology of pairs by combining methods of Floer theory with homotopical algebra. Finally, in Section 5 we present the proofs of the results formulated in Section 1.7, and also proving our results on symplectic intersections from Section 1.5 and on SH-heavy sets stated in Section 1.6.

2 Relative symplectic cohomology of pairs

Fix a closed symplectic manifold (M, ω). Here we formulate the axioms for rela- tive symplectic cohomology of pairs of compact subsets of M. It is an extension of Varolgunes’s relative symplectic cohomology. The axioms are reminiscent of those of Eilenberg–Steenrod for cohomology, see a paper [3] by Cieliebak and Oancea. This is not the most extensive list, but it is reasonably complete, and it contains all the prop- erties we need to prove our results. Section 3 contains the proof of our Main Theorem 1.34 based on the axioms. Let C denote the of compact subsets of M, where the morphisms are in- clusions, and let CP denote the category of compact pairs of subsets of M, that is the objects are pairs (K,K0) of compact subsets of M such that K0 ⊂ K, and there is exactly one morphism (K,K0) → (L, L0) if K ⊂ L and K0 ⊂ L0. Recall that if P,Q are graded modules with graded components P i,Qi, respectively, then a module map f: P → Q is graded of degree d if f(P i) ⊂ Qi+d for all i. We write |f| for the degree of f.

23 2.1 Axioms for relative symplectic cohomology of pairs

Fix a commutative Λ≥0-algebra R; in applications R is either Λ≥0 or Λ. Let M be the category of Z2NM -graded R-modules and graded module maps of arbitrary de- gree and let Ae ⊂ M be the subcategory consisting of associative skew-commutative non-unital R-algebras and degree zero algebra morphisms. Varolgunes’s symplectic co- ∗ homology is a contravariant SH (·): C → Ae where the coefficients are in Λ≥0; if R is a general commutative Λ≥0-algebra, we extend the definition in (1) by setting ∗ ∗ SH (·; R) := SH (·) ⊗Λ≥0 R. We refer the reader to the discussion in Section 2.2 regarding the product structures on SH∗(·) and SH∗(·; Λ). For the definition we need the so-called descent property for pairs of compact sets. It is defined in [23], and we recall the definition in Section 4.8.3. As Varolgunes proves in [23], if two compact sets commute, they are in particular in descent; therefore descent can be interpreted as a kind of “algebraic commutation” condition. We define a relative symplectic cohomology of pairs with coefficients in R as any contravariant functor SH∗(·, ·; R): CP → Ae satisfying the properties appearing below. Note that the functor property means that for each compact pair (K,K0) we have the corresponding graded algebra SH∗(K,K0; R), for each inclusion of pairs (K,K0) ⊂ (L, L0) there is a restriction map

(L,L0) ∗ 0 ∗ 0 res(K,K0) : SH (L, L ; R) → SH (K,K ; R) , which is a degree zero algebra morphism, such that if (P,P 0) is another pair containing (L, L0), then (L,L0) (P,P 0) (P,P 0) res(K,K0) ◦ res(L,L0) = res(K,K0) , (K,K0) and such that res(K,K0) = idSH∗(K,K0;R). The properties are as follows:

• (normalization): Let ι: C → CP be the inclusion functor ι: K 7→ (K, ∅). There is a natural isomorphism between the composition SH∗(·, ·; R) ◦ ι and SH∗(·; R), meaning that for each K ∈ C there is an isomorphism SH∗(K, ∅; R) = SH∗(K; R) in Ae commuting with restrictions.

∗ ∗ Notation 2.1. We let qK : SH (M,K; R) → SH (M; R) be the composition of res(M,K) and the isomorphism SH∗(M, ; R) = SH∗(M; R). (M,∅) ∅

• (triangle): Let Π: CP → C be the functor projecting onto the second factor, that is (K,K0) 7→ K0. Then there is a δ∗ from the composition of the forgetful functor Ae ,→ M with SH∗(·; R) ◦ Π to the composition of the forgetful functor Ae ,→ M with SH∗(·, ·; R)[1], that is for each compact pair

24 (K,K0) we have a degree zero module map δ∗: SH∗(K0; R) → SH∗(K,K0; R)[1], compatible with restrictions. Moreover, this morphism enters an exact triangle

SH∗(K,K0; R) / SH∗(K; R) h

δ∗ resK w K0 SH∗(K0; R)

0 where the horizontal arrow is the restriction res(K,K ) composed with the natural (K,∅) isomorphism SH∗(K, ∅; R) = SH∗(K; R). • (product): Let CTD = {(K,K0,K00) ∈ C3 | K0,K00 ⊂ K and K0,K00 are in descent} be the category of triples where the last two sets are subsets of the first one and 0 00 are in descent. Let Π1,2: CT D → CP be the projection Π1(K,K ,K ) = 0 0 00 00 (K,K ), Π2(K,K ,K ) = (K,K ), and let U: CT D → CP be the union functor, that is (K,K0,K00) 7→ (K,K0 ∪K00). There is a natural transformation of functors CT D → Ae, called product:

∗: SH(·, ·; R) ◦ Π1 ⊗ SH(·, ·; R) ◦ Π2 → SH(·, ·; R) ◦ U, that is for (K,K0,K00) ∈ CT D there is a degree zero algebra morphism ∗: SH(K,K0; R) ⊗ SH(K,K00; R) → SH(K,K0 ∪ K00; R) , compatible with restriction morphisms.

• (Mayer–Vietoris): Let I: CT D → CP be the intersection functor (K,K0,K00) 7→ (K,K0 ∩ K00). There is then a natural transformation from the composition of the forgetful functor Ae ,→ M with SH(·, ·; R) ◦ I to the composition of the for- getful functor Ae ,→ M with SH(·, ·; R)[1] ◦ U, as functors CT D → M, meaning whenever (K,K0,K00) ∈ CT D, then we have a degree zero linear map SH∗(K,K0 ∩ K00; R) → SH∗(K,K0 ∪ K00; R)[1] compatible with restrictions. It moreover enters the exact Mayer–Vietoris tri- angle

SH∗(K,K0 ∪ K00; R) / SH∗(K,K0; R) ⊕ SH∗(K,K00; R) j

+1 s SH∗(K,K0 ∩ K00; R)

25 where the top arrow is the direct sum of restrictions and the right slanted arrow is the difference of restrictions. In Section 4, we prove Theorem 2.2. There exists a relative symplectic cohomology of pairs with coefficients in Λ≥0. We prove the normalization and the triangle property in full detail in Section 4.10. The Mayer–Vietoris property is elaborated upon, with the exception of the compatibility of the natural transformation with restrictions, but this compatibility can be proved using techniques appearing in Section 4.10. Finally, we only elaborate on the existence of the product, but not its compatibility with restrictions, which likewise can be proved using the techniques of Section 4.10. It is plausible that the triangle property can be generalized to a long exact sequence of a general triple, just like in ordinary cohomology. It is also conceivable that other axioms can be formulated and proved for the relative symplectic cohomology of a pair. If R is any flat Λ≥0-algebra, for instance R = Λ, then we have the following result. Corollary 2.3. Let SH∗(·, ·) be a relative symplectic cohomology of pairs with coeffi- ∗ ∗ cients in Λ≥0. Then SH (·, ·; R) := SH (·, ·)⊗Λ≥0 R is a relative symplectic cohomology of pairs with coefficients in R.

2.2 Discussion The product structure on SH∗(·). In [21] Tonkonog and Varolgunes construct an associative skew-commutative product on SH∗(K; Λ), and show that it possesses a unit, and that restriction maps are unital algebra morphisms. To this end they define so- called raised symplectic cohomology, and construct the product and the unit on the level of cohomology. They note in Remark 5.16, that it is possible to construct the product directly on SH∗(K) without tensoring with Λ. It is in fact possible to construct such a product already on the chain level, without the need for raised cohomology, and we use this in order to construct the product on SH∗(·, ·; R). The main technical issue seems to be the ability to choose a Hamiltonian 1-form with nonnegative curvature, which extends the given Hamiltonians on the ends of the pair-of-pants surface, as in [21]. An elementary construction shows that it is indeed possible. Although this product cannot carry a unit, as explained in Remark 5.16 in [21], it does have a similar structure, which upon tensoring with Λ furnishes a λ 0 unit. This structure consists of a family of elements TK ∈ SH (K) for λ > 0, with λ λ ∗ K λ λ the property that TK ∗ α = T α for α ∈ SH (K), and such that resK0 (TK ) = TK0 for K0 ⊂ K. These elements are defined as follows: under the canonical isomorphism

26 ∗ ∗ λ 0 λ SH (M) = H (M) ⊗ Λ>0, let TM ∈ SH (M) be the element corresponding to 1 ⊗ T , λ M λ and put TK := resK (TM ). It is possible to show, using standard Floer-theoretic techniques, that the product

SH∗(M) ⊗ SH∗(K) → SH∗(K)

λ λ maps TM ⊗ α 7→ T α. It then follows from the compatibility of the product with restrictions that λ M λ λ λ TK ∗ α = resK (TM ) ∗ α = TM ∗ α = T α , as claimed.

The product structure on SH∗(·, ·) and the Mayer–Vietoris property. Let us provide intuition for the fact that the descent property (which holds, for instance, when the sets Poisson commute) appears in the seemingly unrelated narrative about the product on symplectic cohomology of pairs. We employ an informal analogy between symplectic topology and basic via the following glossary: the relative symplectic cohomology corresponds to the singular cohomology, and the relative sym- plectic cohomology of a pair corresponds to the singular cohomology of a pair. Since the relative symplectic cohomologies of subsets in descent satisfy the Mayer-Vietoris property (see [23]), the starting point of our discussion is a pair of subspaces A, B of a topological space X satisfying the Mayer-Vietoris short exact sequence

0 → C∗(A ∩ B) → C∗(A) ⊕ C∗(B) → C∗(A ∪ B) → 0 , where C∗ stands for the singular complex. Compare it with the obvious short exact sequence 0 → C∗(A ∩ B) → C∗(A) ⊕ C∗(B) → C∗(A) + C∗(B) → 0 , where C∗(A) + C∗(B) ⊂ C∗(X) stands for the sum of the subspaces C∗(A),C∗(B) ⊂ C∗(X). From long exact homology sequences and the 5-lemma, we see that the natural chain map C∗(A) + C∗(B) → C∗(A ∪ B) induces an isomorphism in homology. Thus A, B is an excisive pair in the terminology of [4, Definition 3.2]. As it is explained ibid., for such a pair one has a well defined cup product on relative cohomology

Hp(X,A) ⊗ Hq(X,B) → Hp+q(X,A ∪ B) .

Roughly speaking, in Section 4.9 below we elaborate on the implication “Varolgunes’s Mayer-Vietoris ⇒ product on SH∗(·, ·)” for subsets in descent in the context of sym- plectic cohomology, which is the subject of the product property stated above. The

27 main technical difficulty is that various diagrams, including those containing the Mayer- Vietoris short exact sequence, commute only up to homotopy, forcing us to use tools of homotopical algebra.

An exact triangle. For domains appearing in Theorem 1.55 we will prove the fol- lowing result, relating SH∗(M,K; Λ) to the SH+-invariant of K (see [3] and Section 5.4 for the definition) and to H∗(M,K; Λ). Theorem 2.4. Under the assumptions of Theorem 1.55 there exists an exact triangle

H∗(M,K; Λ) / SH∗(M,K) , (5) h

+1 v SH+,∗(Kb)[−1] The proof is given is Section 5.4 below. It would be interesting to explore possible extensions of this triangle to more general symplectic manifolds and domains.

3 Proof of Main Theorem 1.34

For open U ⊂ M put

M ∗ ∗  ∗ ∗ θ(U) := ker resM−U : SH (M; Λ) → SH (M \ U; Λ) ⊂ SH (M; Λ) = QH (M) . The idea is to express τ via θ, prove that θ behaves similarly to an IVQM, and deduce that τ is an actual IVQM based on that. First, let us relate θ and τ. Using Definition 1.32 it can easily be shown that [ τ(U) = {θ(U 0) | U 0 open with U 0 ⊂ U} . (6)

Since SH∗(·; Λ) is a functor with respect to inclusion, it follows that θ is monotone: θ(U) ⊂ θ(V ) if U ⊂ V . Given an open U ⊂ M, an approximating chain for U is a sequence {Uj}j of open subsets of U with Uj ⊂ Uj+1 for all j and such that U = S j Uj. Note that every has an approximating chain. We will use the following elementary fact multiple times: if K ⊂ U is compact and {Uj}j is an approximating chain for U, then there is j such that K ⊂ Uj. In particular, any two approximating 0 0 chains {Uj}j, {Uj}j dominate each other, meaning for each j there is j such that Uj ⊂ 0 ∗ Uj0 and vice versa. Since QH (M) is finite-dimensional, the increasing sequence of ∗ ideals {θ(Uj)}j stabilizes, that is there is I ∈ I(QH (M)) such that θ(Uj) = I for all j large enough. It follows from (6) that τ(U) = I. Next, we have the following result.

28 Proposition 3.1. The function θ satisfies all the properties of an IVQM, except con- tinuity and additivity, instead of which we have (weak additivity): θ(U ∪V ) = θ(U)+θ(V ) if U, V ⊂ M are open sets with disjoint closures.

Let us use this to show that τ is an IVQM. (Normalization): Trivial. (Monotonicity): Let U, V ⊂ M be open and assume U ⊂ V . Let {Uj}j, {Vj}j be approximating chains for U, V , respectively. We can pick j such that τ(U) = θ(Uj), τ(V ) = θ(Vj), and Uj ⊂ Vj. It follows that τ(U) ⊂ τ(V ) by the monotonicity of θ. (Continuity): Let W ⊂ M be open and let

W (1) ⊂ W (2) ⊂ · · · ⊂ W

(k) be open subsets whose union equals W . Denote by {Wi } an approximating chain (k) (k)  (k) for W . For every k, one can choose i(k) large enough so that θ Wi(k) = τ(W ) (k) and so that the sequence {Wi(k)} is an approximating chain for W . It follows that τ(W ) = τ(W (k)) for k large enough, and the desired continuity follows. (Additivity): Let U, V ⊂ M be disjoint open sets and let {Uj}j, {Vj}j be ap- proximating chains for U, V , respectively. It is easy to show that {Uj ∪ Vj}j is an approximating chain for U ∪ V . There is j such that τ(U) = θ(Uj), τ(V ) = θ(Vj), τ(U ∪ V ) = θ(Uj ∪ Vj). Since Uj,Vj clearly have disjoint closures, it follows from the weak additivity of θ that θ(Uj ∪ Vj) = θ(Uj) + θ(Vj) and the claim follows. (Quasi-multiplicativity): First, we have the following lemma, proved below.

Lemma 3.2. If U, V ⊂ M are open commuting sets, then there are approximating chains {Uj}j, {Vj}j such that Uj,Vj commute for each j. Assuming the lemma for the moment, let U, V ⊂ M be open commuting sets and pick approximating chains as in the lemma. Since [ [ [ Uj ∩ Vj ⊂ Uj ∩ Vj ⊂ Uj+1 ∩ Vj+1 and (Uj ∩ Vj) = Uj ∩ Vj = U ∩ V, j j j it follows that {Uj ∩ Vj}j is an approximating chain for U ∩ V , therefore there is j such that τ(U) = θ(Uj), τ(V ) = θ(Vj), τ(U ∩ V ) = θ(Uj ∩ Vj), and thus

τ(U) ∗ τ(V ) = θ(Uj) ∗ θ(Vj) ⊂ θ(Uj ∩ Vj) = τ(U ∩ V ) , where the containment is thanks to the quasi-multiplicativity of θ.

29 (Intersection): Let U, V ⊂ M be open sets which cover M and let {Uj}j, {Vj}j be approximating chains for them. We claim that there is j such that Uj,Vj cover M. Assuming this, we can increase j so that in addition we have τ(U) = θ(Uj), τ(V ) = θ(Vj). Since, as we have seen, {Uj ∩ Vj}j is an approximating chain for U ∩ V , we can further increase j so that we also have τ(U ∩ V ) = θ(Uj ∩ Vj). Since Uj,Vj cover M, by the intersection property of θ we have

τ(U ∩ V ) = θ(Uj ∩ Vj) = θ(Uj) ∩ θ(Vj) = τ(U) ∩ τ(V ) .

c c To prove the claim, note that U ⊂ V is compact, therefore there is j so that U ⊂ Vj, c c whence Vj ⊂ U, therefore there is k so that Vj ⊂ Uk, which means that Vj,Uk cover M, therefore a fortiori Ul,Vl cover M for l = max{j, k}. (Invariance): If U ⊂ M is open, {Uj}j is an approximating chain for U, and φ ∈ Symp0(M), then {φ(Uj)}j is an approximation chain for φ(U). We can pick j so that τ(U) = θ(Uj), τ(φ(U)) = θ(φ(Uj)), and the invariance of τ follows from that of θ. (Vanishing): Let K ⊂ M be displaceable and let {Uj}j be an approximating chain for M \ K. It follows that for j large enough, M \ Uj is displaceable, whence  ∗ θ(Uj) = θ M\(M\Uj) = QH (M), and taking j large enough so that τ(M\K) = θ(Uj) we obtain τ(M \ K) = QH∗(M). For the second part of the vanishing property, let U ⊃ K be open such that θ(U) = 0. If V is an open set with V ⊂ U, then by the monotonicity of θ we have θ(V ) = 0, whence [ τ(U) = {θ(V ) | V open with V ⊂ U} = 0 . This completes the proof of Theorem 1.34 modulo Lemma 3.2 and Proposition 3.1, which we will now prove. Proof of Lemma 3.2. By definition there are Poisson-commuting f, g ∈ C∞(M, [0, 1]) such that M \ U = f −1(0),M \ V = g−1(0). For every α ∈ (0, 1] consider the compact −1 −1 sets Uα = f ((α, 1]) and Vα = g ((α, 1]). Note that for every 0 < β < α ≤ 1 we have Uα ⊂ Uβ and Vα ⊂ Vβ, moreover, [ [ U = Uα, and V = Vα α∈(0,1] α∈(0,1]

We claim that Uα and Vα commute. Indeed, let ψα : [0, 1] → [0, 1] be a smooth function −1 such that ψα (0) = [0, α]. The functions ψα ◦ f, ψα ◦ g Poisson commute and satisfy −1 −1 M \ Uα = (ψα ◦ f) (0),M \ Vα = (ψα ◦ g) (0).

Therefore Uα and Vα commute.

Finally, the sequences {Uj := U1/j}j∈N and {Vj := V1/j}j∈N are approximating chains as claimed.

30 Proof of Proposition 3.1. We have already remarked that θ is monotone. Thus it re- mains to show that θ satisfies normalization, weak additivity, quasi-multiplicativity, intersection, invariance, and vanishing. M ∗ ∗ (Normalization): Since resM : SH (M; Λ) → SH (M; Λ) is the identity, we con- clude that θ(∅) = 0; θ(M) = ker(SH∗(M; Λ) → SH∗(∅; Λ) = 0) = SH∗(M; Λ). (Weak additivity): Let U, V be open subsets with disjoint closures. Denote A = M \ U and B = M \ V . Note that {A, B} is a cover of M by compact sets with disjoint boundaries. By definition we have

M M M θ(U) = ker resA , θ(V ) = ker resB , θ(U ∪ V ) = ker resA∩B .

Thus we need to show that

M M M ker resA∩B = ker resA + ker resB .

Note that the inclusion

M M M ker resA∩B ⊃ ker resA + ker resB is trivial. On the other hand, since A, B cover M we get that SH∗(M,A ∪ B; Λ) = SH∗(M,M; Λ) = 0. Since ∂A ∩ ∂B = ∅, the subsets A, B commute. Therefore, from the Mayer-Vietoris property for a relative symplectic cohomology of pairs and the Mayer-Vietoris sequence for relative symplectic cohomology [23] we get the commutative diagram

0 m / SH∗(M,A; Λ) ⊕ SH∗(M,B; Λ) =∼

+ ∗ qA⊕qB SH (M,A ∩ B; Λ)

  ∗ ∗ ∗ q SH (M; Λ) m / SH (M; Λ) ⊕ SH (M; Λ) A∩B proj1 − proj2 +  M M ∗ resA ⊕ resB SH (M; Λ)

 ∗ ∗ ∗ resM SH (M; Λ) m / SH (A; Λ) ⊕ SH (B; Λ) A∩B

+  SH∗(A ∩ B; Λ)

31 consisting of exact triangles in the top, middle and bottom slices, while the vertical sequences are exact at the middle term. Using the exactness of the top triangle we get that

SH∗(M,A; Λ) ⊕ SH∗(M,B; Λ) ∼= SH∗(M,A ∩ B; Λ).

Denote this isomorphism by f: SH∗(M,A; Λ) ⊕ SH∗(M,B; Λ) → SH∗(M,A ∩ B; Λ). M Now, let x ∈ ker resA∩B. From the exactness of the right vertical column, there ∗ exists y ∈ SH (M,A ∩ B; Λ) such that qA∩B(y) = x. Since f is an isomorphism, we ∗ ∗ can find a unique pair (y1, y1) ∈ SH (M,A; Λ) ⊕ SH (M,B; Λ) such that f(y1, y2) = y. M Denote x1 = qA(y1), x2 = qA(y2). Using the exactness, we get that x1 ∈ ker resA M and x2 ∈ ker resB . Finally, by the commutativity of the diagram, we get that x = M M (proj1 − proj2)(x1, x2) = x1 − x2 ∈ ker resA + ker resB . M M M This proves that ker resA∩B ⊂ ker resA + ker resB , and therefore that

M M M ker resA∩B = ker resA + ker resB .

(Quasi-multiplicativity): Let U, V ⊂ M be open commuting subsets. Then their complements M \ U, M \ V are compact commuting subsets. Let a ∈ θ(U) and b ∈ θ(V ). By the triangle property of a relative symplectic cohomology of pairs there ∗ ∗ are x ∈ SH (M,M \ U; Λ) and y ∈ SH (M,M \ V ; Λ) such that qM\U (x) = a and ∗ qM\V (y) = b. From the product property of SH (·, ·; Λ) we get that

a ∗ b = qM\U (x) ∗ qM\V (y) = q(M\U)∪(M\V )(x ∗ y) = qM\(U∩V )(x ∗ y).

Using the triangle property again, we obtain

M im qM\(U∩V ) = ker resM\(U∩V ) = θ(U ∩ V ).

This implies that a ∗ b ∈ θ(U ∩ V ), hence θ(U) ∗ θ(V ) ⊂ θ(U ∩ V ). (Intersection): Let U, V be two open subsets that cover of M. Note that M \ U, M \ V are disjoint compact subsets, therefore M \ U, M \ V commute. By Theorem 4.8.1 and Lemma 2.4.3 from [23] we can use Mayer-Vietoris, that is

SH∗(M \(U ∩V ); Λ) = SH∗((M \U)∪(M \V ); Λ) ∼= SH∗(M \U; Λ)⊕SH∗(M \V ; Λ), since (M \ U) ∩ (M \ V ) = M \ (U ∪ V ) = ∅ and SH∗(∅; Λ) = 0. The isomorphism is given by

 M\(U∩V ) M\(U∩V )  ∗ ∗ ∗ resM\U , resM\V : SH (M \ (U ∩ V ); Λ) → SH (M \ U; Λ) ⊕ SH (M \ V ; Λ).

32 Since

M M\(U∩V ) M M M\(U∩V ) M resM\U = resM\U ◦ resM\(U∩V ) and resM\V = resM\V ◦ resM\(U∩V ) , we conclude that

M M M θ(U ∩ V ) = ker resM\(U∩V ) = ker resM\U ∩ ker resM\V = θ(U) ∩ θ(V ) , as required. (Invariance): Varolgunes proved that the symplectic cohomology and the restric- tion maps are invariant under Ham(M, ω)-action, see [22, Theorem 1.3.1]. It is not hard to show, using essentially Morse-theoretic arguments, that the natural action ∗ of Symp0(M, ω) on SH (M) is trivial, which implies that φ ∈ Symp0(M, ω), then θ(φ(K)) = θ(K), and this yields the invariance property of θ. (Vanishing): Let K ⊂ M be a displaceable compact subset. Varolgunes proved that SH∗(K; Λ) = 0, see [22, Theorem 1.3.1]. This shows that

M M ∗ θ(M \ K) = ker resM\(M−K) = ker resK = SH (M; Λ). Additionally, since K is a displaceable compact subset, there exists a displaceable do- main W which contains K in its interior. Let Z be the closure of the complement of W and denote U = Int(W ). The compact sets W, Z commute and since their intersection is a displaceable compact, we get that SH∗(W ∩ Z) = {0}. Using Mayer-Vietoris we get that SH∗(M; Λ) = SH∗(Z ∪ W ; Λ) ∼= SH∗(Z; Λ) ⊕ SH∗(W ; Λ) = SH∗(Z; Λ).

M ∗ ∗ Thus resZ : SH (M; Λ) → SH (Z; Λ) is an isomorphism, therefore M M θ(U) = ker resM\U = ker resZ = {0}.

4 Proofs of properties of symplectic cohomology of pairs

In this section we construct a relative symplectic cohomology of pairs with coef- ficients in Λ≥0 and prove its properties as announced in Section 2, thereby proving Theorem 2.2. The construction and the proofs occupy Sections 4.9, 4.10. This relies on the algebraic language of cubes, which we outline in Sections 4.2–4.7. The language of cubes was introduced by Varolgunes in [23]. This theory is an explicit realization of some ∞-categorical aspects of cochain complexes. The highlight of this section is the construction of the product on relative symplectic cohomology of pairs in Section 4.9.

33 4.1 Completion for Λ≥0-modules

Completion of Λ≥0-modules plays an essential role in Varolgunes’s definition of sym- plectic cohomology, therefore we describe it here. First, the Novikov field Λ carries a valuation ν:Λ → R ∪ {+∞} given by ν(0) = ∞ and   P∞ λi ν i=1 ciT = λ1 provided the λi form a strictly increasing sequence and c1 6= 0. Using ν, we can define the interval modules

−1 −1 Λ≥r = ν ([r, ∞]) , Λ>r = ν ((r, ∞]) , Λ[a,b) = Λ≥a/Λ≥b for a < b .

Given a Λ≥0-module A, its completion is the Λ≥0-module

A := lim A ⊗ Λ . b ←− [0,r) r→∞

Completion is an endofunctor on the category of Λ≥0-modules. We extend it to the category of graded Λ≥0-modules by completing degree-wise. From the universal property of inverse limits we obtain a morphism

A → Ab called completion. We say that A is complete if the completion morphism is an isomorphism. The interval modules are typical examples. On the other hand, Λb = 0. Since completion commutes with finite direct sums, a finite direct sum of interval modules is likewise complete.

4.2 Cubes The language of cubes is very convenient in order to work with the algebraic data arising when defining relative symplectic cohomology. Throughout we fix a unital com- mutative ring R and we work in the category of graded R-modules and graded maps, where the grading is over Z or over Zk with k even. Recall that given graded modules C,D with graded components Ci,Di, a module map f: C → D is graded of degree d if f(Ci) ⊂ Di+d. We denote the degree of f by |f|. Fix a nonnegative integer n. We call a subset of [0, 1]n a face if it is given by setting some of the coordinates to either 0 or 1; the rest of the coordinates are referred to as the free coordinates of F . Given a face F its dimension, denoted |F |, is the number of its free coordinates, while its initial vertex ini F and terminal vertex ter F , are

34 the points of F closest to or farthest from the origin, respectively; in this case we write F : v → v0, where v = ini F , v0 = ter F . Note that F is determined by its initial and final vertices. If F 0,F 00 are faces, we write F = F 0 · F 00 to denote the situation in which

ini F = ini F 0 , ter F 0 = ini F 00 , ter F 00 = ter F.

v C v An (algebraic) n-cube C is a pair ({C }v∈{0,1}n , {fF }F ⊂[0,1]n a face), where each C is a graded module, and for every face F ⊂ [0, 1]n we have a graded module morphism C ini F ter F fF : C → C of degree 1 − |F |, subject to the condition that for each face F we have X |F 0| 0 00 C C (−1) sgn(F ,F )fF 00 fF 0 = 0 , (7) F =F 0·F 00 where sgn(F 0,F 00) is a sign defined as follows. Any face of [0, 1]n comes equipped with a natural orientation coming from the ordering of its free coordinates. Then sgn(F 0,F 00) is the intersection index of F 0,F 00 inside F . For a more explicit description, assume first that we have a linearly ordered finite set S = {s1 < ··· < sk+l}. Recall that a (k, l)-shuffle on S is a permutation σ ∈ SS such that σ(s1) < ··· < σ(sk) and 0 00 0 00 σ(sk+1) < ··· < σ(sk+l). If S = S ] S , then there is a unique (|S |, |S |)-shuffle σS0,S00 0 00 such that σ({s1, . . . , sk}) = S , σ({sk+1, . . . , sk+l}) = S . Now let S ⊂ {1, . . . , n} be the set of free coordinates of F , S0 the set of free coordinates of F 0 and S00 the set of free coordinates of F 00, so that S = S0 ] S00. Endow S with the order induced from 0 00 {1, . . . , n}. Then sgn(F ,F ) := sgn σS0,S00 . v C The above relations mean in particular that for each vertex v,(C , f{v}) is a cochain complex, each 1-dimensional edge yields a cochain map between its vertex modules, each 2-dimensional face yields a homotopy between the two compositions of the maps running along its perimeter, and so on. Note that the above sign only depends on the internal ordering of the free coordinates 0 00 n v C of F ,F , which means that for any face G ⊂ [0, 1] the pair ({C }v, {fF }F ) where v ranges over the vertices of G while F over the subfaces of G, is itself a cube, provided we renumber the coordinates in their natural order in {1, . . . , n}. We will refer to such a cube as a subcube of C, or, when the face G is given, the subcube obtained by restricting C to G, and we denote it by C|G. If A0, A1 are n-cubes, a map from A0 to A1 is an (n + 1)-cube C such that C Ai = C|[0,1]n×{i}, i = 0, 1. In this case we write A0 −→A1. We define the corresponding −C negated map A0 −→A1 by negating all the maps going from the vertices of A0 to those n of A1. A trivial check shows that this is indeed a map. Note that if F ⊂ [0, 1] is a 0 0 0 0 0 face and C = C|F ×[0,1], Ai = Ai|F = C|F ×{i}, i = 0, 1, then C is a map from A0 to A1: 0 0 C 0 A0 −→A1.

35 In Section 4.6 below we’ll define pullbacks for V-shaped diagrams, which are a particular case of a so-called partial cube, which is given by vertex modules assigned to a subset of the vertices, and face maps between these, such that they satisfy the cube relation (7) whenever all the maps in it are defined.

4.3 Direct sums and tensor products In various constructions related to the definition and proof of properties of the symplectic cohomology of a pair we will need direct sums and tensor products of cubes. If A, B are n-cubes, their direct sum A ⊕ B is defined as the cube with vertices v v v A⊕B A B (A ⊕ B) = A ⊕ B and face maps fF = fF ⊕ fF . A trivial verification shows that this is indeed a cube. Note that in particular taking direct sums commutes with C⊕C0 passing to subcubes, thus restricting A ⊕ A0 −−−→B ⊕ B0 to a face Fb = F × [0, 1] yields C| ⊕C| 0 Fb Fb 0 A|F ⊕ C |F −−−−−→B|F ⊕ B |F . Before we define tensor products of cubes, let us recall the notion of graded tensor product of graded maps: if V,V 0, W, V 0 are graded modules and if f: V → W , f 0: V 0 → W 0 are graded maps, their graded tensor product f ⊗ f 0: V ⊗ V 0 → W ⊗ W 0 is defined by (f ⊗ f 0)(v ⊗ v0) = (−1)|f 0||v|f(v) ⊗ f 0(v0). All the tensor products of graded maps below are taken in the graded sense. If (C, d), (C0, d0) are modules endowed with differentials, that is degree 1 maps 0 0 squaring to zero, their tensor product is the graded module (C ⊗ C , d ⊗ idC0 + idC ⊗d ), where the tensor product is taken in the graded sense. This amounts to the usual Leibnitz rule for differentials on the tensor product of graded modules. Let now A, B be a k-cube and an l-cube, respectively. We are going to define their tensor product A ⊗ B, which is a (k + l)-cube, as follows. The vertex modules are

(A ⊗ B)i1...ik+l = Ai1...ik ⊗ Bik+1...ik+l .

A⊗B The face maps fF are defined as follows. If F = {(v1, . . . , vk, w1, . . . , wl)} has di- mension zero and we let v = (v1, . . . , vk), w = (w1, . . . , wl), then the corresponding face map, that is the differential on the module Av ⊗ Bw, is simply the differential on the tensor product (in the graded sense!):

A⊗B A B fF = f{v} ⊗ idBw + idAv ⊗f{w} .

If F has positive dimension, let F 0 ⊂ [0, 1]k, F 00 ⊂ [0, 1]l be faces such that F = F 0 × F 00 ⊂ [0, 1]k × [0, 1]l = [0, 1]k+l. We have three cases:

0 00 A⊗B (i) |F |, |F | > 0: in this case fF = 0.

36 (ii) |F 00| = 0: let F 00 = {w}; then

A⊗B A v w v0 w fF = fF 0 ⊗ idBw : A ⊗ B → A ⊗ B ,

where v = ini F 0, v0 = ter F 0;

(iii) |F 0| = 0: let F 0 = {v}; then

A⊗B B v w v w0 fF = idAv ⊗fF 00 : A ⊗ B → A ⊗ B ,

where w = ini F 00, w0 = ter F 00.

The following is obtained by unraveling the definitions:

Proposition 4.1. The tensor product of cubes is a cube.

We will also use a more general tensor product corresponding to a (k, l)-shuffle σ ∈ Sk+l, denoted A ⊗σ B. The above case corresponds to the identity shuffle. The vertex modules of the general tensor product are given by

i1...ik+l iσ(1)...iσ(k) iσ(k+1)...iσ(k+l) (A ⊗σ B) = A ⊗ B , and the face maps are defined analogously to the case σ = id, except we need to renumber coordinates according to σ. The formulas are exactly the same, except the notation becomes more complicated. We will need special cases of the tensor product construction: tensoring with a , the identity map, the diagonal map and the sum map.

Tensoring with a chain complex Let C be a cube and let (A, d) be a chain complex, that is a 0-cube. We can then form their tensor product C ⊗ A. Its vertices are (C ⊗ A)v = Cv ⊗ A, the differential on such a vertex is the above tensor of the respective n C⊗A C differentials, and if F is a positive-dimensional face of [0, 1] , then fF = fF ⊗ idA. We can likewise form the tensor product A ⊗ C, which differs from C ⊗ A by signs of the differentials. The simplest example is as follows. Let R be viewed as a graded module over itself, concentrated in degree zero, and given the zero differential. Then, using the canonical isomorphism C ⊗R R = C = R⊗R C for an R-module C, we obtain canonically C ⊗ R = C = R ⊗ C. Similarly we can consider the module R ⊕ R sitting in degree zero, also with the zero differential. In this case, given a cube C, we have a canonical isomorphism C ⊕ C ≡ C ⊗ (R ⊕ R).

37 The identity map Given an n-cube C and a direction i ∈ {1, . . . , n+1}, the identity map in the i-th direction is the tensor product C ⊗σi idR, where σi is the unique (n, 1)-shuffle mapping n + 1 7→ i, and idR : R → R is the identity map, viewed as a 1-cube. The most common case we will need is when i = n + 1, in which case σ = id. id In this case we will denote C ⊗ idR ≡ C −→C. Note that this is a map from C to itself in the above sense. Note that if F ⊂ [0, 1]n is a face, then the restriction of the identity map C −→Cid to id the face F × [0, 1] yields the identity map C|F −→C|F .

The diagonal map Consider the diagonal map ∆R: R → R ⊕ R as a 1-cube. If C is an n-cube, the corresponding diagonal map is the (n + 1)-cube C ⊗ ∆R. Computing, we see that it is a map C −→∆C C ⊕ C, where the only nonzero maps in the (n + 1)-st direction are the diagonal maps Cv → Cv ⊕ Cv.

The sum map Consider now the sum map Σ: R ⊕ R → R, again considered as a 1-cube. Tensoring a cube C with it, we obtain, after suitable identifications, the sum map C ⊕ C −→CΣC , whose only nonzero horizontal maps are sums Cv ⊕ Cv → Cv.

4.4 Cones, cocones In , cones and cocones afford explicit models for homotopy cokernels and homotopy kernels, respectively. Here we review the corresponding notions for cubes, as defined in [23]. n−1 n n Fix i ∈ {1, . . . , n}. Identify R with the {xi = 0} ⊂ R , let π: R → n−1 n−1 n R be the map turning the i-th coordinate to 0, and let ιj: R → R be the n−1 inclusions (x1,..., 0, . . . , xn) 7→ (x1,...,j,...,xn) for j = 0, 1. For F ⊂ [0, 1] ⊂ n−1 −1 R let F = (π|[0,1]n ) (F ), Fj = ιj(F ). Let C be an n-cube. Its cone in the i-th n−1 direction is the (n − 1)-cube conei C, defined as follows. For a vertex v ∈ {0, 1} we have v ι0(v) ι1(v) (conei C) = C [1] ⊕ C . For a face F ⊂ [0, 1]n−1 the corresponding face map f conei C is given by the triangular F matrix  |F0| C  −(−1) fF0 0 ](i,F ) C , −(−1) fF fF1 where ](i, F ) denotes the number of the i-th coordinate among the free coordinates of F . Note in particular that the differentials are negated in the shifted modules, while

38 those in unshifted ones retain their sign, and that face maps corresponding to edges always come with the positive sign. The following is obtained from the definitions:

Proposition 4.2. conei C is a cube. It also follows that the cone operation behaves well with respect to restrictions: if F ⊂ [0, 1]n is a face containing the i-the direction as one of its free coordinates, then for any n-cube C we have

(conei C)|π(F ) = cone](i,F )(C|F ) .

◦k We will need the k-th iterated cone in the first direction, cone := cone1 ◦ · · ·◦cone1, which maps n-cubes into (n − k)-cubes. We also need to discuss a particular property of cones, also described in [23]. Given a module A, we call it k-coniform if it comes with a decomposition into 2k submodules:

1 M A = Ai1...ik .

i1,...,ik=0 Given two such modules A, B, we say that a map f: A → B is k-coniform if the composition i ...i f i0 ...,i0 A 1 k → A −→ B → B 1 k 0 vanishes unless ij ≤ ij for j = 1, . . . , k; here the first map is the inclusion while the last one is the projection. Finally, we say that a cube C is k-coniform if so is every vertex module and every face map of C. The very definition of cones implies the following Proposition 4.3. The iterated cone operation cone◦k establishes a bijective correspon- dence between n-cubes and k-coniform (n−k)-cubes. In particular, given an n-coniform cochain complex (A, d), there is a well-defined n-cube (cone◦n)−1A. If C is a cube, we let C[k] be the cube whose vertex modules have been shifted by k in degree, and whose structure maps have all been multiplied by (−1)k. It is trivially a cube. We define the cocone of C in the i-th direction by

coi C := conei C[−1] . Combining the fact that shifts commute with passage to subcubes, we see that cocones are also compatible with it, namely if F ⊂ [0, 1]n is a face containing the i-th direction as one of its free coordinates, then

(coi C)|π(F ) = co](i,F )(C|F ) .

39 Remark 4.4. This is a rather lengthy remark, relevant to the constructions described in Sections 4.9, 4.10. Although a cocone of a cube is again a cube, it has some peculiar properties. For instance if C is a cube which is the identity map in a direction i, its cocone in any other direction is rather minus the identity map. This is because any cone in a direction other than i is the identity, and in the cocone all the maps are negated. The main consequence for us is that when trying to define the symplectic cohomology of a pair of subsets, the comparison maps between the various homologies, coming from cocones of certain maps, need to be negated in order to form a direct system whose direct limit we need to take. In practice, this takes the following form. Varolgunes defines special kinds of cubes: triangles and slits. An n-triangle is an n-cube of the form

A / B

id   A / C where the diagram directions are the last two ones, so that A, B, C are the corresponding subcubes of dimension n − 2. If we take its cocone in any direction other than the last two ones, we obtain a cube of the form

coi A / coi B

− id  #  coi A / coi C which is not a triangle because of the minus sign in − id. We will only need to use the case when this last cube is actually a square:

f A / B h − id g   A / C k

This cube simply means that g ◦ f and −k are homotopic. Another way of saying this is that (−g) ◦ (−f) and −k are homotopic, that is we have the following square:

−f A / B h id −g   A / C −k

40 which is in fact a 2-triangle. Thus we see that if we have a 3-triangle and we take its cocone in the first direction, we obtain a square which can be modified as above to obtain a triangle, which is equivalent to a homotopy commutative triangle involving maps which are negated. A similar remark applies to slits. An n-slit is an n-cube of the form

A / B

id id   A / B

It can be thought of as two maps A → B and a homotopy between them. If we have a 3-slit like this, taking its cocone in the first direction results in a square of the form

f A / B h − id − id   A / B k or, equivalently, a square of the form

−f A / B h id id   A / B −k which is a 2-slit. These squares express the fact that f, k: A → B are homotopic maps, which is all we need.

4.5 Compositions of maps of cubes In the construction of the product on the relative symplectic cohomology of pairs in Section 4.9 we’ll need to use compositions of maps of cubes. Recall that a map between two n-cubes is simply an (n + 1)-cube whose restrictions to the corresponding n-faces are the given n-cubes. Let A −→BF −→CG be two maps of n-cubes. We will define their composition A −−→CG◦F as follows. Taking the n-th iterated cone in the first direction, we arrive at a sequence of chain complexes and chain maps

◦n ◦n cone◦n A −−−−−→cone F cone◦n B −−−−−→cone G cone◦n C .

41 The chain maps are n-coniform, as is their composition cone◦n G ◦ cone◦n F, therefore we can apply (cone◦n)−1 to the chain map

◦n ◦n cone◦n A −−−−−−−−−−→cone G◦cone F cone◦n C , and the result is the desired composition

A −−→CG◦F .

The following material will be relevant in Section 4.10, where we prove that the relative symplectic cohomology of a pair is well-defined, as well as in Section 4.9, where we construct the product. If F ⊂ [0, 1]n is a face, we let Fb := F × [0, 1] ⊂ [0, 1]n+1. The main property of com- positions we’ll need is that the 1-dimensional edges in the direction of the composition simply compose and they acquire no signs. That is if v is a vertex, then

f G◦F = f G ◦ f F , {dv} {dv} {dv}

Let us call a map of n-cubes A −→BF straight if, whenever v, v0 ∈ {0, 1}n are vertices and F : v → v0 is the corresponding face, the map f F vanishes unless v = v0. Examples Fb of straight maps include the identity, diagonal, and the sum map on a given cube, and more generally the tensor product of any cube with a 1-cube. Another property of straight maps we’ll use in the sequel is as follows. Assume that A −→BF is a straight map between two n-cubes. It follows that F is completely determined by the structure maps of A, B and the maps fbF := f F : Av → Bv for the v {dv} n v v vertices v of [0, 1] . If we are given two cubes A, B and a collection of maps fbv: A → B , they are the structure maps of a straight map A → B if and only if for every face of [0, 1]n+1 of the form Fb, for F ⊂ [0, 1]n, we have

B A fF ◦ fbini F = fbter F ◦ fF .

We will apply this fact in the following form. Let A −→BF be a map of n-cubes and let Fb = conen+1 F be the corresponding cone, which is an n-cube. We claim that there are natural straight maps ι π B −→ Fb −→A[1] . These are defined as follows. Let v ∈ {0, 1}n be a vertex. Then the defining morphisms of the straight maps ι, π are as follows:

v v v v ιv: B → Fb = A [1] ⊕ B is the inclusion map ,

42 v v v v kvk1 πv: Fb = A [1] ⊕ B → A [1] is (−1) · the projection map , where kvk1 is the `1-norm of v, that is the number of coordinates of v equal to 1. Shifting everything by −1, we see that we also have natural straight maps

ι[−1] π[−1] B[−1] −−→ Fb[−1] = con+1 F −−−→A , which we’ll use in the definition of relative symplectic cohomology of pairs. We’ll also need another feature of this sequence, namely its exactness. Let’s call a sequence

0 → A −→BF −→CG → 0 of maps between n-cubes exact if the following sequence is:

◦n ◦n 0 → cone◦n A −−−−−→cone F cone◦n B −−−−−→cone G cone◦n C → 0 .

We claim that ι[−1] π[−1] 0 → B[−1] −−→ con+1 F −−−→A → 0 (8) is exact. In fact, since the maps here are straight, applying cone◦n to the sequence results in M M M 0 → Bv[−1] → (Av ⊕ Bv) → Av → 0 , v v v where the horizontal maps are the direct sums of the components of ι[−1] and π, and this sequence is clearly exact, whence the claim. In particular if both A, B are acyclic, then so is con+1 F.

4.6 Folding and pullbacks of V-shaped diagrams When defining the product on the relative symplectic cohomology of a pair in Section 4.9, we’ll need to use the technique of folding a cube and then taking its cocone. Here we describe this technique. Consider an (n + 2)-cube, n ≥ 0, of the form

F A / B (9) H G I   C / D K

Here A,..., D are n-cubes, and F, G, I, K are maps of n-cubes, that is (n+1)-cubes where the last direction is the one marked with the corresponding letter; H stands for

43 all the maps running from vertices of A to those of D. The horizontal and the vertical directions have numbers n + 1, n + 2, respectively. We define the corresponding folded cube to be as follows: (F,−G) A / B ⊕ C H I+K  #  0 / D (F,−G) Here A −−−−→B ⊕ C is the composition of the diagonal map A → A ⊕ A and the direct sum F ⊕ (−G): A ⊕ A → B ⊕ C, while I + K is the composition of the direct sum I ⊕ K: B ⊕ C → D ⊕ D with the sum map D ⊕ D → D. It is a matter of routine verification that this is indeed a cube. Note that folding commutes with passage to subcubes in the obvious sense. A V-shaped diagram is a partial cube of the form B

I  C / D K

I+K We define its pullback to be the n-cube Q = con+1(B ⊕ C −−→D). If we have a cube of the form (9), we can fold it and then apply cocone in the last ((n + 2)-nd) direction. We then obtain an (n + 1)-cube of the form

L I+K con+1(A → 0) −→Q = con+1(B ⊕ C −−→D) , which is the point of the current section.2)

4.7 Rays and telescopes For n ≥ 0, an (n + 1)-ray is a diagram of the form

F1 F2 R = A1 −→A2 −→A3 → ...

n consisting of n-cubes A1, A2,... and maps between them F1, F2,... If F ⊂ [0, 1] is a face, such a ray defines a restriction to F , which is the (|F | + 1)-ray

F | 1 Fb R|F = A1|F −−→A2|F → ...

2) Note that even though the vertex modules of con+1(A → 0) are all canonically isomorphic to those of A, the structure maps gain signs which depend on the dimension of the corresponding face (in fact n con+1(A→0) |F | A if F ⊂ [0, 1] is a face, fF = (−1) fF ).

44 Here we will define the telescope tel R of such a diagram, which is an n-cube. It will have the property that for any face F ⊂ [0, 1]n we have

tel(R|F ) = (tel R)|F , (10) which is crucial in the applications of telescopes below.

Remark 4.5. To motivate the definition of telescopes, recall that given modules Ai, i ∈ , and module maps f : A → A , the corresponding direct limit lim A can be N i i i+1 −→i i taken as the cokernel of the map M M id −f: Ai → Ai , where f(a1, a2,..., ) = (0, f1(a1),... ) . i i In this paper we are working with homotopical constructions, therefore we need the analog of the direct limit in homotopy theory, also known as the emphhomotopy colimit. The telescope of a 1-ray is a model for it. Since the above direct limit is the cokernel of a map, it is expected that the corresponding homotopy colimit is given by the homotopy cokernel of a map, or in other words, by its cone. Consider the map of n-cubes ∞ ∞ M F M Ai −→ Ai , i=1 i=1 L symbolically defined as F = id + i Fi, and explicitly given as follows. First, both the domain and the target cubes have their own structure maps given by the direct sums n of the structure maps of the Ai. Now, given a vertex v ∈ {0, 1} , both the domain and target have the corresponding vertex module

∞ !v ∞ M M v Ai = Ai . i=1 i=1 Given vertices v, v0 ∈ {0, 1}n spanning a face F with v = ini F , v0 = ter F , the corre- sponding face map ∞ ∞ M M 0 f F : Av → Av Fb i i i=1 i=1 has matrix representation  1  F 0 0 0 ... f F1 2 0 0 ...  Fb F   F  ,  0 f 2 3 0 ...  Fb F ...

45 i 0 i where  = 0 unless v = v in which case  = id v . F {v} Ai We can now define the telescope of R:

tel R := conen+1 F .

It can be shown that the telescope of a subray is the corresponding subcube of the telescope, as claimed in equation (10). Another feature is that if we have an (n + 2)-ray consisting of the identity maps between n-cubes as follows:

F1 F2 A1 / A2 / A3 / ...

id id id

 F1  F2  A1 / A2 / A3 / ... then its telescope is tel R −→id tel R , where R = A1 → A2 → ... Another crucial property of telescopes is their behavior relative to tensor products. 0 f1 0 0 f1 0 If R = A1 −→ A2 → ... and R = A1 −→ A2 → ... are 1-rays, their tensor product is 0 0 0 f1⊗f1 0 defined to be the 1-ray R⊗R = A1⊗A1 −−−→ A2⊗A2 → ... There is a canonical quasi- isomorphism tel(R ⊗ R0) → tel R ⊗ tel R0, see [23] and [9, Lemma 0.1]. Moreover, the induced map on completions, tel(\R ⊗ R0) → tel R\ ⊗ tel R0, is also a quasi-isomorphism, provided all the modules in sight are flat. Assume now that there are two 2-rays T , T 0:

0 0 A1 / A2 / ... A1 / A2 / ...

   0  0 B1 / B2 / ... B1 / B2 / ...

0 0 0 0 and let A, B, A B be the 1-rays comprised of the Ai, Bi, Ai, Bi, respectively. We obtain a natural 3-ray whose constituent 1-rays are the tensor products A⊗A0, A⊗B0, B ⊗A0, and B ⊗ B0. Taking the telescope, we obtain the square

tel(A ⊗ A0) / tel(A ⊗ B0) (11)

 '  tel(B ⊗ A0) / tel(B ⊗ B0)

46 On the other hand, we have the tensor product of the cochain maps tel T = tel A → tel B, tel T 0 = tel A0 → tel B0, that is the square

tel A ⊗ tel A0 / tel A ⊗ tel B0 (12)

 (  tel B ⊗ tel A0 / tel B ⊗ tel B0

The point is that the above quasi-isomorphism extends to a map of cubes from the square (11) to the square (12), in the sense that each edge map going in the direction of the cube map is a quasi-isomorphism.

4.8 Floer data, rays, and symplectic cohomology of pairs Here we discuss the notions of Floer data, acceleration data, the resulting rays of Floer complexes, and show how these are used to define the symplectic cohomology of a pair. Also we recall Varolgunes’s notion of descent for a pair.

4.8.1 Hamiltonian rays In [23], Varolgunes defined monotone cubes, triangles, and slits of Hamiltonians. Let us recall what this means. Consider a strictly increasing Morse function ρ1: [0, 1] → R n with exactly two critical points at 0, 1. It gives rise to the Morse function ρn: [0, 1] → R by ρn(x) = ρ1(x1) + ··· + ρ1(xn). We also endow the cube with the standard Rieman- nian metric. A monotone cube of Hamiltonians is a suitably smooth map [0, 1]n → ∞ C (M), which is monotone nondecreasing along the gradient lines of ρn; Varolgunes’s definition in particular assumes that the Hamiltonians at the vertices of such a shape are nondegenerate, and that they are also constant on a neighborhood of each vertex. We will impose these assumptions throughout. Similarly monotone triangles and slits are defined as smooth maps into C∞(M) defined on subsets of [0, 1]n called triangles and slits, see ibid.

Definition 4.6. An n-ray of Hamiltonians or a Hamiltonian n-ray is a sequence Hi, i ∈ N, of monotone n-cubes of Hamiltonians, such that the face of Hi corresponding to {xn = 1} coincides with the face of Hi+1 corresponding to {xn = 0}. Similarly we define triangular n-rays and slit-like n-rays of Hamiltonians: a triangular n-ray of Hamiltonians, defined for n ≥ 3, is a sequence of monotone n-triangles of Hamiltonians which agree along faces as above. A slit-like n-ray of Hamiltonians, defined for n ≥ 3, is likewise a sequence of monotone n-slits of Hamiltonians which agree along the appropriate faces.

47 Remark 4.7. Such configurations will give rise to rays, triangular and slit-like rays in the algebraic sense, as we will describe below. To associate such an algebraic object to a configuration of Hamiltonians, additional structure is needed, such as choices of almost complex structures and Pardon data, depending on the level of generality. A suitable choice of such structures always exists given a configuration of Hamiltonians, therefore we will suppress such choices both from notation and from the discussion below.

Varolgunes proves the following result, itself a consequence of Pardon’s constructions [15, 16]:

Theorem 4.8. Given a monotone n-cube H of Hamiltonians and a suitable choice of additional data such as almost complex structures or Pardon data, there is an n- cube whose vertices are the Floer complexes of the Hamiltonians at the vertices of H, and whose higher maps are obtained by counting elements of suitable moduli spaces of parametrized Floer equations. Similarly, a monotone n-triangle gives rise to an n- triangle of Floer complexes, and a monotone n-slit yields an n-slit.

Since an n-ray of Hamiltonians consists of monotone n-cubes of Hamiltonians glued along the n-th direction, we have the following

Corollary 4.9. An n-ray of Hamiltonians H gives rise to an n-ray whose vertices are the Floer complexes of the Hamiltonians at the vertices of H. Similarly, a triangular or a slit-like n-ray of Hamiltonians gives rise to a triangular or a slit-like n-ray of Floer complexes, respectively.

This is applied as follows: we take the telescope of such an n-ray to obtain an (n − 1)- cube, to which we then apply the completion functor.

4.8.2 Weighted Floer complexes Here we specify the Floer complexes we will be using. If H ∈ C∞(M × S1) is a non-degenerate Hamiltonian, its Floer complex, generated by the set P◦(H) of its ∗ L contractible 1-periodic orbits, was defined as CF (H) = x∈P◦(H) Λ≥0 · x in equation

(2). This complex is graded over Z2NM by the Conley–Zehnder index. The differential is determined by its matrix elements, given by X hdx, yi = #M(H; x, y; A) T E(A) ,

A∈π2(M,x,y)

1 where π2(M, x, y) is the set of homotopy classes of smooth maps of the cylinder R × S to M which are asymptotic at ±∞ to x, y, M(H; x, y; A) stands for the moduli space of

48 unparametrized Floer trajectories corresponding to H, running from x to y, and repre- senting the class A, such that every solution has index 1; #M(H; x, y; A) is a suitable virtual count as in Pardon [15, 16] or just a signed count if M is assumed to be semi- R  positive. Finally E(A) = S1 Ht(y(t)) − Ht(x(t)) dt − hω, Ai is the topological energy of solutions in class A, or equivalently the increase in the action along such solutions. Continuation maps between such complexes corresponding to monotone nondecreasing of Hamiltonians are defined in a similar manner, with weighting by suitable powers of T .

4.8.3 Acceleration data and descent We need the following definitions. Definition 4.10. • An acceleration datum is a 1-ray of Hamiltonians. Given ∞ an acceleration datum H, we will write H = (Hi)i=1, meaning that Hi is the nondegenerate Hamiltonian at the i-th vertex of the ray, with the monotone 1- cubes of Hamiltonians between them being implicit.

0 0 0 • Given two acceleration data H = (Hi)i and H = (Hi)i, we write H  H if 0 Hi ≤ Hi for all i. Note that we do not require anything regarding the monotone 1-cubes between the Hamiltonians at the vertices.

• If H, H0 are two acceleration data and H  H0, a filling F: H → H0 is a 2-ray of Hamiltonians whose top and bottom 1-rays are H, H0. Remark 4.11. For the rest of Section 4 we drop the superscript indicating the grading from all Floer complexes and complexes derived therefrom.

Notation 4.12. Given an acceleration datum H = (Hi)i we denote the corresponding Floer 1-ray by CF (H) = CF (H1) → CF (H2) → ... Definition 4.13. Let K ⊂ M be compact. We say that an acceleration datum H = ∞ (Hi)i is an acceleration datum for K if (Hi)i is a cofinal sequence in CK⊂M = {H ∈ ∞ 1 3) C (M × S ) | H|K×S1 < 0} relative to the usual order on functions.

Definition 4.14. (Varolgunes [23]) Let K ⊂ M be compact and let H = (Hi)i be an acceleration datum for K. The corresponding complex is defined to be

SC(H) := telc CF (H) . The relative symplectic cohomology of K inside M is SH(K) := HSC(H) . 3)This is a variant of the notation used in [23].

49 Remark 4.15. Varolgunes proves in [23] that this is well-defined, in the sense that given another acceleration datum H0 for K, the two cohomology modules are canonically isomorphic. We will provide details of this proof in Section 4.10 below.

Let us now recall what it means for two sets to be in descent; see [23]. Let K,K0 ⊂ M be compact subsets. Choose an acceleration datum H• for • = K,K0,K ∪ K0,K ∩ K0, such that H•  H•0 whenever • ⊃ •0. These acceleration data can be fitted into a Hamiltonian 3-ray by choosing suitable fillings for the corresponding Hamiltonian cubes, as proved in [23]. Passing to the corresponding 3-ray of Floer complexes, taking its telescope and completing yields the following square:

SC(K ∪ K0) / SC(K0)

  SC(K) / SC(K0 ∩ K0) where SC(•) stands for SC(H•). Denoting this square by C, we say that K,K0 are in descent if C acyclic, meaning that its repeated cone cone◦2 C is an acyclic complex.

4.8.4 The relative symplectic cohomology of a pair Next we define one of the main characters in our story, the relative symplectic coho- mology of a pair.

Definition 4.16. Let (K,K0) be a compact pair in M. Let H, H0 be acceleration data for K,K0, respectively, assume that H  H0, and fix a filling F: H → H0. There is a 2-ray of Floer complexes whose top and bottom 1-rays are CF (H) and CF (H0), respectively. Taking its completed telescope, we arrive at a cochain map ΦF : SC(H) → SC(H0). We define the relative complex corresponding to F to be

SC(F) := co ΦF , and the relative symplectic cohomology of the pair (K,K0) as its cohomology:

SH(K,K0) := HSC(F) .

Theorem 2.2 is an immediate consequence of the following result:

Theorem 4.17. The relative symplectic cohomology of a pair is well-defined indepen- dently of the chosen data. Moreover, it is a relative symplectic cohomology of pairs with coefficients in Λ≥0.

50 The assertion that the symplectic cohomology of a pair is well-defined is proved in Section 4.10, which also includes all the other properties with the exception of the product, which is constructed in the next subsection. Remark 4.18. Defining relative cohomology on the chain level by taking a cocone is a standard construction in homotopy theory when dealing with relative (co)homology, see for instance [2, p.78] for such a definition for relative .

4.9 Constructing the product The goal here is to prove the existence of a diagram

∗ SH(K,K0) ⊗ SH(K,K00) / SH(K,K0 ∪ K00) (13)

 ∗  SH(K) ⊗ SH(K) / SH(K) provided K0,K00 are subsets of K, which are in descent. Here the vertical arrows are the canonical restrictions while the bottom arrow is the Tonkonog–Varolgunes product [21]. We will construct the top arrow and prove that the resulting diagram commutes. In the following description SC(K) and so on stand for SC of a suitably chosen acceleration datum for K. The above diagram is constructed as follows: Step 1: Using the construction of Section 4.6, we will construct a module Q, which is the pullback of the V-shaped diagram consisting of SC(K0),SC(K00),SC(K0 ∩ K00) and restrictions, and a natural morphism p: SC(K) → Q, whose cocone co(p) is then shown to be naturally quasi-isomorphic to SC(K,K0 ∪ K00); it is here that Varolgunes’s Mayer–Vietoris sequence enters, as alluded to in Section 2.2; Step 2: We will construct a “zigzag” diagram, which on passing to cohomology yields a commutative diagram

HSC(K,K0) ⊗ SC(K,K00) / H co(p) (14)

  HSC(K) ⊗ SC(K) / SH(K)

Step 3: Finally, using the isomorphism H co(p) ' SH(K,K0 ∪ K00) and the natural transformation H(·) ⊗ H(·) → H(· ⊗ ·), we arrive at the diagram

SH(K,K0) ⊗ SH(K,K00) / HSC(K,K0) ⊗ SC(K,K00) / H co(p) / SH(K,K0 ∪ K00)

   w SH(K) ⊗ SH(K) / HSC(K) ⊗ SC(K) / SH(K)

51 Suitably composing arrows, we arrive at the desired diagram (13). In the next two subsections we will describe Steps 1, 2 in more detail. Remark 4.19. It is possible to show, using the techniques appearing in Section 4.10, that the product we construct is independent of the choices of acceleration data, almost complex structures, and so on, and that it indeed commutes with restriction morphisms.

4.9.1 Step 1 In what follows we implicitly choose acceleration data for all the sets appearing in the diagrams, such that if two sets A, B satisfy B ⊃ A, then the corresponding acceleration data are related by . Moreover, we choose suitable fillings between those acceleration data, as well as higher-dimensional Hamiltonian rays as necessary. The discussion in Section 4.8 shows that this is indeed possible. Consider the cube

SC(K) / SC(K00)

' SC(K0 ∪ K00) / SC(K00)

  SC(K0) / SC(K0 ∩ K00)

  SC(K0) / SC(K0 ∩ K00) Here the order of coordinates is as follows: the first one is toward the reader, the second one is to the right, and the third one is down. This cube is obtained by choosing a 4-ray of Hamiltonians as indicated in the previous paragraph, passing to the corresponding 4- ray of Floer complexes, and taking its completed telescope. Note that the non-identity arrows are chain level restriction maps. Let us denote 00 00 resK + resK Q = co SC(K00) ⊕ SC(K0) −−−−−−−−−−−−→K0∩K00 K0∩K00 SC(K0 ∩ K00) ; this is the pullback of the suitable V-shaped diagram, which can be seen as part of the front and the back faces of the above cube. Now let us apply folding to the above cube and then take cocone in the vertical (that is third) direction. We obtain a square p SC(K) = co(SC(K) → 0) / Q

− resK K0∪K00 − id   SC(K0 ∪ K00) = co(SC(K0 ∪ K00) → 0) / Q

52 where the minus signs come out of the definition of cocones, see Remark 4.4. Negating the vertical and diagonal maps, we obtain a homotopy commutative triangle, from which we can build the following square:

res SC(K) / SC(K0 ∪ K00)

 SC(K) p / Q

Applying cocone in the horizontal direction, we arrive at the following morphism of short exact sequences, see equation (8):

0 / SC(K0 ∪ K00)[−1] / SC(K,K0 ∪ K00) / SC(K) / 0

  0 / Q[−1] / co(p) / SC(K) / 0

Passing to homology and noting that the first and the third arrow are quasi-isomorphisms (the first one by [23]), we arrive at the conclusion that so is SC(K,K0 ∪ K00) → co(p). This completes Step 1.

4.9.2 Step 2 Here we will prove the existence of the following commutative “zigzag” diagram:

SC(K,K0) ⊗ SC(K,K00) SC(K,K0) ⊗ SC(K,K00) (15)

 · / SC(K) ⊗ SC(K) O O qis qis

· / ·

  co(p) / SC(K) where · stands for unspecified modules, which we will describe below, and “qis” stands for “quasi-isomorphism.” Passing to cohomology, we will obtain the diagram (14) above as claimed. This will complete the construction of the diagram (13). This diagram is obtained as follows. Below we describe seven 3-cubes numbered I–VII. We will compose cubes I through V, juxtapose the result with cubes VI, VII, and the result is three 3-cubes written side-by-side as a “zigzag” diagram of 2-cubes

53 and maps between them of the form · → · ← · → ·. To this diagram we apply folding and cocone as in Section 4.6, and the above diagram (15) is obtained as a result. Let us now describe this in detail. In the cubes, the order of coordinates is as follows: the first is down, the second is right, and the third is perpendicular to the page. Cubes I–IV are tensor products of lower-dimensional ones, cubes V, VI come from the functoriality of completed telescopes (see Section 4.7), while the last cube comes from Floer theory. Cube I:

SC(K,K0) ⊗ SC(K,K00) / 0

' SC(K,K0) ⊗ SC(K,K00) / SC(K,K0) ⊗ SC(K00)

 0 0

  0 0 This cube is obtained as follows. Consider the square

SC(K,K00) / 0

00  00 SC(K,K ) res ◦ pr/ SC(K ) and tensor it with SC(K,K0) to obtain

SC(K,K0) ⊗ SC(K,K00) / 0

 SC(K,K0) ⊗ SC(K,K00) / SC(K,K0) ⊗ SC(K00)

This is the top face of our cube and it remains to append the bottom face comprised of zeros.

54 Cube II:

SC(K, K0) ⊗ SC(K, K00) / SC(K, K0) ⊗ SC(K00)

SC(K, K0) ⊗ SC(K, K00) / SC(K, K0) ⊗ SC(K00)

0 0

*  )  SC(K0) ⊗ SC(K, K00) / SC(K0) ⊗ SC(K00)

This is obtained by tensoring the square

SC(K,K0) SC(K,K0)

res ◦ pr   0 / SC(K0) by the map SC(K,K00) → SC(K00). Cube III:

SC(K, K0) ⊗ SC(K, K00) / SC(K, K0) ⊗ SC(K00)

* ) SC(K) ⊗ SC(K, K00) / SC(K) ⊗ SC(K00)

  SC(K0) ⊗ SC(K, K00) / SC(K0) ⊗ SC(K00)

  SC(K0) ⊗ SC(K, K00) / SC(K0) ⊗ SC(K00)

This is obtained by tensoring the square

pr SC(K,K0) / SC(K)

res ◦ pr res   SC(K0) SC(K0)

res ◦ pr by the map SC(K,K00) −−−−→ SC(K00).

55 Cube IV:

SC(K) ⊗ SC(K,K00) / SC(K) ⊗ SC(K00)

) SC(K) ⊗ SC(K) / SC(K) ⊗ SC(K00)

  SC(K0) ⊗ SC(K,K00) / SC(K0) ⊗ SC(K00)

)   SC(K0) ⊗ SC(K) / SC(K0) ⊗ SC(K00)

This is obtained by tensoring the square

pr SC(K,K00) / SC(K)

res ◦ pr res   SC(K00) SC(K00) by the map SC(K) −→res SC(K0). Cube V:

SC(K) ⊗ SC(K) / SC(K) ⊗ SC(K00)

) ) tel CF (H\) ⊗ tel CF (H) / tel CF (H)\⊗ tel CF (H00)

  SC(K0) ⊗ SC(K) / SC(K0) ⊗ SC(K00)

)  )  tel CF (H0\) ⊗ tel CF (H) / tel CF (H0)\⊗ tel CF (H00)

This cube is obtained by applying the natural transformation b· ⊗b· → ·d ⊗ · termwise. Note that the cube is straight in the direction pointing into the page.

56 Cube VI: tel CF (H\) ⊗ tel CF (H) / tel CF (H)\⊗ tel CF (H00) i j

tel(c CF (H) ⊗ CF (H)) / tel(c CF (H) ⊗ CF (H00))

  tel CF (H0\) ⊗ tel CF (H) / tel CF (H0)\⊗ tel CF (H00) i j

  tel(c CF (H0) ⊗ CF (H)) / tel(c CF (H0) ⊗ CF (H00))

The construction of this cube before completion is outlined at the end of Section 4.7. Note that even after completion, the diagonal arrows remain quasi-isomorphisms. Cube VII:

tel(c CF (H) ⊗ CF (H)) / tel(c CF (H) ⊗ CF (H00))

) ) SC(K) / SC(K00)

  tel(c CF (H0) ⊗ CF (H)) / tel(c CF (H0) ⊗ CF (H00))

)  )  SC(K0) / SC(K0 ∩ K00)

This is the Floer theoretic cube corresponding to products and restriction maps. We note here that the back face of this cube coincides with the front face of cube VI. We now compose cubes I-V, take the resulting cube and juxtapose it with cubes VI, VII. There results a diagram of four 2-cubes and three maps between them going as follows: · → · ← · → ·, that is we have a “zigzag” in the middle. We can now apply folding to this diagram and then take cocone in the vertical direction, which results in the diagram on the left. The diagram on the right is obtained from it by taking its cocone in the horizontal direction and appealing to the natural map from the cocone to the domain of a map, see equation (8):

57 SC(K,K0) ⊗ SC(K,K00) / 0 SC(K,K0) ⊗ SC(K,K00) SC(K,K0) ⊗ SC(K,K00)

  SC(K) ⊗ SC(K) / · · / SC(K) ⊗ SC(K) O O O O qis qis qis qis

tel(c CF (H) ⊗ CF (H)) / · · / tel(c CF (H) ⊗ CF (H))

 p    SC(K) / Q co(p) / SC(K) This completes Step 2 and therefore the construction of the product on relative symplectic cohomology.

4.10 Well-definedness and properties of SH(K,K0) Here we prove that SH∗(·, ·) is well-defined, that is it is independent of the various choices of acceleration data, fillings, and so on, and then we prove that it satisfies the properties announced in Section 2.1, thereby proving Theorem 2.2. We start with a proof that SH(K) is well-defined. This serves as a framework for the analogous proof for SH(K,K0).

Definition 4.20. • If H = (Hi)i is an acceleration datum, we say that an accelera- tion datum He = (Hej)j is a subdatum of H if there is a strictly increasing sequence of natural numbers i(j) such that Hej = Hi(j); note that we do not require anything of the 1-cubes between them. We write He ⊂ H. Note that H  He. • If this is the case, we call H, H0 equivalent if there is a subdatum He ⊂ H such that H0  He. We let SK = {H = (Hi)i | H is an acceleration datum for K} .

Note that the collection of all acceleration data is directed relative to , and that SK is a directed subset. Moreover, any two acceleration data for K are equivalent. Below is a reformulation of Varolgunes’s construction of the relative symplectic cohomology. (i) Given an acceleration datum H, in Section 4.8 we have defined the corresponding 1-ray of Floer complexes CF (H), and the corresponding complex SC(H) and homology SH(H) = H(SC(H)).

58 (ii) Given another acceleration datum H0 such that H  H0, and a filling F: H → H0, there is a 2-ray of Floer complexes such that CF (H) and CF (H0) are its top and bottom 1-rays. Taking its telescope and completing, we arrive at a 1-cube of 0 the form ΦF : SC(H) → SC(H ), which is simply a chain map between the com- plexes of H, H0. We use the same notation for the induced map on cohomology: 0 ΦF : SH(H) → SH(H ).

(iii) If Fe: H → H0 is another filling, F and Fe fit into a slit-like 3-ray of Floer data like this: F H G H0 , Fe where the third dimension is perpendicular to the page. This yields a slit-like 3-ray of Floer complexes, whose completed telescope is a 2-slit of the form

ΦF SC(H) - SC(H0) , 1 Φ Fe

in other words, we obtain a homotopy between Φ , Φ , and therefore the two F Fe induce the same morphism on homology. It follows that the maps induced on homology by various fillings all coincide, and we denote the resulting map by H0 0 ΦH : SH(H) → SH(H ).

0 00 0 0 00 (iv) If H  H  H are acceleration data and F0: H → H , F1: H → H , F2: H → H00 are fillings, they fit into a triangular 3-ray of Hamiltonians. To it there cor- responds a triangular 3-ray of Floer complexes, and its completed telescope is a 2-triangle

ΦF0 SC(H) / SC(H0)

ΦF Φ 1 F2 %  SC(H00)

H00 H0 H00 which yields the relation ΦH0 ◦ ΦH = ΦH on homology.

He (v) If He ⊂ H is a subdatum, Varolgunes proves in [23] that ΦH is an isomorphism. In H H H H particular, since ΦH ◦ ΦH = ΦH, we see that ΦH is the identity. All of the above H0 means that (SH(H), ΦH ) is a direct system.

59 (vi) If H  H0 are equivalent, then there are subdata He ⊂ H, He0 ⊂ H0 such that H  H0  He  He0. Considering the resulting commutative diagram

He ΦH ΦHe 0 , He 0 SH(H) 0 / SH(H ) / SH(He) / SH(He ) ΦH ΦHe H H0 4 0 ΦHe H0

He He0 H0 and the fact that ΦH and ΦH0 are isomorphisms, we see that ΦH is likewise an isomorphism. (vii) If K ⊂ M is a compact subset, we define SH(K) := lim SH(H) , −→ H∈SK where the connecting maps are the above morphisms. Since all the connecting maps are isomorphisms, every natural morphism SH(H) → SH(K) is in fact an isomorphism. (viii) If L ⊂ K is another compact set, the restriction morphism K resL : SH(K) → SH(L) 0 0 is defined as follows. Let H ∈ SK . There is H ∈ SL with H  H . Consider 0 ΦH the composition SH(H) −−→H SH(H0) → SH(L), where the second map is the natural morphism into the direct limit. Thus we get a map SH(H) → SH(L). Using the cocycle identities for the comparison maps Φ as above, as well as prop- erties of direct limits, we can show that this map is independent of the choice of 0 H . Moreover if H1 ∈ SK is such that H  H1, then the map SH(H) → SH(L) H1 ΦH equals the composition SH(H) −−→ SH(H1) → SH(L). It follows from the uni- K versal property of limits that we have described a map resL : SH(K) → SH(L). Moreover, since natural maps SH(H) → SH(K) and SH(H0) → SH(L) for 0 H ∈ SK , H ∈ SL are isomorphisms, the diagram

SH(H) / SH(K)

  SH(H0) / SH(L) commutes, a fact which is convenient when we actually have to compute the K L K restriction. It is also clear that resK = id and that restrictions satisfy resP ◦ resL = K resP if P ⊂ L ⊂ K.

60 We will now prove that SH(K,K0) is well-defined, define the restriction morphisms, and prove the properties formulated in Section 2.1. Items (i-viii) contain the proof of well-definedness, restrictions are the subject of item (ix), while items (x-xii) contains the proofs of the normalization, triangle, and the Mayer–Vietoris properties. The main obstacle to overcome is to prove that this cohomology is independent of the choice of acceleration data and the filling. This is done using the same scheme as in ∗ the above proof that SHM (K) is well-defined independently of the acceleration datum used, except that we have to add a dimension throughout, and that now we also have to invoke properties of cocones.

0 (i) If F: H → H is a filling between acceleration data, we let SC(F) := co ΦF . Note that we have an exact sequence

0 → SC(H0)[−1] → SC(F) → SC(H) → 0 ,

see equation (8). We let SH(F) = H(SC(F)).

0 0 (ii) If Hi, Hi, i = 0, 1 are acceleration data such that Hi  Hi for i = 0, 1, H0  H1 0 0 0 and H0  H1, consider fillings Fi: Hi → Hi. These fit into a 3-ray G of Hamilto- nians, which then produces a 3-ray of Floer complexes. Taking its telescope and completing, we arrive at a 2-cube, where the vertical direction is the first one and the horizontal one is the second one:

Φ F0 0 SC(H0) / SC(H0)

Φ  F1  0 SC(H1) / SC(H1)

Taking the cocone in the horizontal direction, we obtain a chain map SC(F0) → SC(F1). We let BG be the negative of this map. We use the same notation for the induced map on cohomology. Note that we have the following commutative diagram with rows being exact sequences:

0 0 / SC(H0)[−1] / SC(F0) / SC(H0) / 0 (16)

−BG

0   0 / SC(H1)[−1] / SC(F1) / SC(H1) / 0

which is a particular case of equation (8).

61 (iii) If in the previous situation we have another filling G0 of the same square, the two fit into a slit-like 3-ray of Hamiltonians, and taking its telescope and completion, we obtain a 3-slit as follows:

Φ F0 0 SC(H0) / SC(H0)

Φ  F1  0 SC(H1) / SC(H1)

Taking cocone in the horizontal direction yields a 2-slit giving a homotopy between BG,BG0 (see Remark 4.4). We denote the resulting well-defined map on homology F 1 by BF0 : SH(F0) → SH(F1).

(iv) If we have three pairs of acceleration data Hij, i = 0, 1, j = 0, 1, 2, such that H0j  H1j and Hi0  Hi1  Hi2, and we have fillings Fj: H0j → H1j, then all of this fits into a triangular 4-ray of Hamiltonians, whose completed telescope is the following 3-triangle:

ΦF0 SC(H00) / SC(H10)

& ΦF1 & SC(H01) / SC(H11)

  ΦF2   SC(H02) / SC(H12)

Taking cocone in the horizontal direction yields the homotopy-commutative tri- angle (see Remark 4.4): BF1 F0 SC(F0) / SC(F1)

BF2 F1 BF2 F0 %  SC(F2)

It follows that on homology we have BF2 ◦ BF1 = BF2 . F1 F0 F0

(v) If F: H → H0 is a filling and Fe: He → He0 is a filling between subdata, then using any 3-ray G of Hamiltonians extending F, Fe, we arrive at the following diagram,

62 which is a particular case of (16):

0 / SC(H0)[−1] / SC(F) / SC(H) / 0

−BG    0 / SC(He0)[−1] / SC(Fe) / SC(He) / 0

Since the right and the left vertical arrows are quasi-isomorphisms, from the long exact sequence of homology it follows that so is the middle one. There- Fe F fore BF : SH(F) → SH(Fe) is an isomorphism. In particular BF is the identity map.

(vi) Given a pair of acceleration data H, H0 with H  H0, consider the set of fillings 0 F 0 {F: H → H }. It parametrizes the system (SH(F),BF ) of modules and isomor- phisms satisfying the cocycle identity. It follows that we can take its (co)limit:

SH(H, H0) := lim SH(F) , F: H→H0 and that for any F the natural map SH(F) → SH(H, H0) is an isomorphism.

(vii) If Hij are acceleration data with H0j  H1j and Hi0  Hi1, then the above discussion yields a natural map SH(H00, H01) → SH(H10, H11). If we have a third such pair, then the corresponding maps obey the composition rule. Moreover the natural map SH(H, H0) → SH(H, H0) is the identity.

(viii) If K0 ⊂ K ⊂ M are compact sets, put

0 0 SK,K0 = {(H, H ) ∈ SK × SK0 | H  H } .

This then is a directed set of pairs of acceleration data with respect to the product order induced by , and we put

SH(K,K0) := lim SH(H, H0) . −→ 0 (H,H )∈SK,K0

0 Using reasoning as above, we see that for any (H, H ) ∈ SK,K0 the natural mor- phism SH(H, H0) → SH(K,K0) is an isomorphism.

(ix) If (L, L0) ⊂ (K,K0) is a compact subpair, we can define the corresponding restric- tion morphism (K,K0) 0 0 res(L,L0) : SH(K,K ) → SH(L, L )

63 similarly to the restriction map for the absolute case. Namely, we pick (H, H0) ∈ 0 0 0 SK,K0 and (G, G ) ∈ SL,L0 with (H, H )  (G, G ). We have the composition SH(H, H0) → SH(G, G0) → SH(L, L0), the latter being the natural map into the direct limit. It is easy to show that this composition is independent of the choice of (G, G0), and that moreover these maps SH(H, H0) → SH(L, L0) form a 0 0 morphism from the direct system (SH(H, H )) 0 to SH(L, L ). In par- (H,H )∈SK,K0 (K,K0) (K,K0) ticular it yields a morphism res(L,L0) as claimed. It also follows that res(K,K0) is the identity and that these restriction morphisms satisfy the cocycle identity.

0 0 2 (x) Let us prove normalization: if (H, H ) ∈ SK,∅, and H consists of a given C -small Morse function plus i, then SC(H0) = 0 as Varolgunes shows in [23]. Therefore the canonical map SH(H, H0) → SH(H) is an isomorphism. It is also easy to see the compatibility with restrictions.

0 0 (xi) Let us prove the triangle property. Let K ⊂ K and pick H ∈ SK , H ∈ SK0 such that H  H0. Pick a filling F: H → H0. The morphism SH(K0) → SH(K,K0)[1] is defined as follows. We have a natural morphism SC(H0)[−1] → SC(F), that is SC(H0) → SC(F)[1] after shifting. Passing to homology, we obtain SH(H0) → SH(F)[1]. Using the above methods, it is easy to show that this morphism is compatible with the various comparison morphisms, and therefore induces a well- defined map SH(K0) → SH(K,K0)[1]. We have the short exact sequence

0 → SC(H0)[−1] → SC(F) → SC(H) → 0 .

Its long exact homology sequence reads

· · · → SH(H0)[−1] → SH(F) → SH(H) → ...

It is also compatible with the various comparison morphisms, therefore we have the long exact sequence

...SH(K0)[−1] → SH(K,K0) → SH(K) → ...

as claimed.

(xii) Let us prove the Mayer–Vietoris property. Fix acceleration data for K,K0,K00,K0∪ K00,K0 ∩ K00, and consider the following cube, where SC(·) means SC for the cor-

64 responding acceleration datum:

SC(K) / SC(K0 ∪ K00)

( SC(K) / SC(K00)

 SC(K) / SC(K0)

(  SC(K) / SC(K0 ∩ K00)

Let us call the square coming from the left face of this cube by A, the right square by B, and the resulting map by A −→BF . There results a short exact sequence of squares 0 → B[−1] → co3 F → A → 0 .

Since A is clearly acyclic, and since B is acyclic by [23], it follows that co3 F is acyclic, which, thanks to [23] results in a long exact homology sequence, which is precisely the Mayer–Vietoris one.

5 Symplectic rigidity and computations

5.1 Proof of Theorem 1.43 Here we prove Theorem 1.43, based on Theorem 1.48, which is proved in Section 5.2. The other ingredient we need is the following lemma. Lemma 5.1. Let M,N be closed symplectic manifolds and let K ⊂ M, L ⊂ N be compact sets. If N \ L decomposes into a finite number of pairwise disjoint displaceable subsets, then τ(L) = SH∗(N; Λ), and moreover the K¨unnethisomorphism ψ : SH∗(M; Λ) ⊗ SH∗(N; Λ) → SH∗(M × N; Λ) maps τ(K) ⊗ τ(L) = τ(K) ⊗ SH∗(N; Λ) into τ(K × L). We refer the reader to [22] for the defintion of the K¨unnethmorphism SH∗(A) ⊗ SH∗(B) → SH∗(A × B) for compact sets A ⊂ M,B ⊂ N. In general this is nei- ther injective nor surjective, however when A = M and B = N, it can be shown to be an isomorphism. The above isomorphism is then obtained by tensoring with Λ. Assuming this, we have:

65 Proof of Theorem 1.43. Let τ be the quantum cohomology IVQM on T6 × S2. The K¨unnethformula yields

∗ 6 2 ∗ 6 ∗ 2 ∗ mod 4 6 QH (T × S ) = QH (T ) ⊗ QH (S ) = H (T ; Λ) ⊗ Λh1, hi .

∗ 6 Let α = [dq1 ∧ dq2], β = [dp1 ∧ dp3], γ = [dp2 ∧ dq3] ∈ QH (T ). Consider the graded ideal I ⊂ QH∗(T6 × S2) generated by α ⊗ h, β ⊗ h, γ ⊗ h. Since

α ∗ β ∗ γ = [dq1 ∧ dq2 ∧ dq3 ∧ dp1 ∧ dp2 ∧ dp3] 6= 0 and h3 = T h 6= 0, we have

(α ⊗ h) ∗ (β ⊗ h) ∗ (γ ⊗ h) = (α ∗ β ∗ γ) ⊗ h3 6= 0 , which implies that I3 6= 0. Thanks to Theorem 1.48 we have

α ∈ τ(T1(a)) , β ∈ τ(T2(b)) , γ ∈ τ(T3(c)) , which by monotonicity implies that α, β, γ ∈ τ(T (a, b, c)). Also h ∈ τ(L) for any equator L ⊂ S2, thanks to Example 1.35. It follows from Lemma 5.1 that α ⊗ h, β ⊗ h, γ ⊗ h ∈ τ(T (a, b, c) × L), and consequently that I ⊂ τ(T (a, b, c) × L). Thus sets of the form T (a, b, c) × equator are all members of the collection XI of all the compact subsets K with I ⊂ τ(K), and the result now follows from Theorem 1.42. It remains to prove the lemma.

M Proof of Lemma 5.1. In this proof we abbreviate resK ≡ resK and similarly for N,M × N. For the first assertion it suffices to show that for any neighborhood V of L we ∗ c have SH (V ; Λ) = 0. Let N \ L = W1 ∪ · · · ∪ Wk be a decomposition into pairwise c Sk c disjoint displaceable open sets. It follows that V = i=1(V ∩ Wi), and moreover that c each V ∩ Wi is displaceable, being contained in Wi, and compact, because it is the c S c c ∗ c complement in V of j6=i(V ∩ Wj), which is open in V . Thus SH (V ∩ Wi; Λ) = 0, and by the Mayer–Vietoris property

∗ c M ∗ c SH (V ; Λ) = SH (V ∩ Wi; Λ) = 0 . i For the second assertion it is enough to prove that if α ∈ τ(K) and β ∈ τ(L) = SH∗(N; Λ), then for every pair of neighborhoods U ⊃ K, V ⊃ L we have

res(U×V )c (ψ(α ⊗ β)) = 0 .

To prove this, we need the following

66 (U×V )c Claim: The restriction resU c×N is an isomorphism. Assuming this for a moment, and using the naturality of the K¨unnethmorphism with respect to restrictions, we obtain the commutative diagram

ψ res(U×V )c SH∗(M; Λ) ⊗ SH∗(N; Λ) / SH∗(M × N; Λ) / SH∗((U × V )c; Λ)

res c ⊗ id resUc×N U c res(U×V )  ψ  t Uc×N SH∗(U c; Λ) ⊗ SH∗(M; Λ) / SH∗(U c × N; Λ)

It follows that

(U×V )c −1  res(U×V )c (ψ(α ⊗ β)) = resU c×N ψ((resU c ⊗ id)(α ⊗ β)) (U×V )c −1  = resU c×N ψ(resU c (α) ⊗β) = 0 , | {z } =0 as claimed. Here we used that resU c (α) = 0, which follows from α ∈ τ(K). It remains to prove the above claim. First, we claim that the sets U c × N,M × V c commute. Indeed, let f: M → [0, 1] and g: N → [0, 1] be smooth functions with f −1(0) = U c and g−1(0) = V c; then the map f × g: M × N → [0, 1]2 is involutive and

U c × N = (f × g)−1({0} × [0, 1]) ,M × V c = (f × g)−1([0, 1] × {0}) , and therefore these sets commute thanks to Example 1.23. Noting that (U c × N) ∪ (M × V c) = (U × V )c and (U c × N) ∩ (M × V c) = U c × V c, the corresponding exact Mayer–Vietoris triangle reads

(U×V )c (U×V )c  resUc×N ,resM×V c SH∗(U × V )c;Λ / SH∗(U c × N; Λ) ⊕ SH∗(M × V c; Λ) i

s SH∗(U c × V c; Λ)

Since V c is a finite union of pairwise disjoint displaceable compact sets, so are M × V c and U c × V c, therefore SH∗(M × V c; Λ) = SH∗(U c × V c; Λ) = 0 by the Mayer–Vietoris property, whence the top arrow in the triangle is the desired isomorphism

(U×V )c res c SH∗(U × V )c;Λ −−−−−→U ×N SH∗(U c × N; Λ) .

67 5.2 Proof of Theorem 1.48 It suffices to prove the case when S is closed, because the open case follows from the closed one by regularization: if S ⊂ Tn is an open polyinterval, then there is an S increasing sequence of closed polyintervals Sk such that S = k Sk, and such that n τ(Sk × T ) stabilizes, therefore τ(S) = τ(Sk) for k large enough. By the results of Section 3, it suffices to prove that for any neighborhood U 0 of K 0 M ∗ there is another one U with U ⊂ U and such that θ(U) = ker resU c = ker H (M; Λ) → ∗  H (M−K; Λ) . By the definition of polyintervals there are compact intervals J1,...,Jn ⊂ 1 Qn n n n 2n S such that K = i=1 Ji ×T (q) ⊂ T (p)×T (q) = T = M. It follows that there ex- ists a domain Wf ⊂ Tn such that W := Wf×Tn ⊂ T2n satisfies the following: K∩W = ∅, W c ⊂ U 0, W is a deformation retract of Kc. We claim that U = W c is the desired neighborhood. Indeed, by linear algebra considerations, any closed characteristic of ∂W has the form {x} × γ, where x ∈ Tn and γ ⊂ Tn is a linear circle. In particular, any such M closed characteristic is noncontractible in M. Theorem 1.54 implies that ker resW = ker H∗(M; Λ) → H∗(W ; Λ) = ker H∗(M; Λ) → H∗(Kc; Λ), the latter equality being due to the fact that W is a deformation retract of Kc.

5.3 Proof of Theorem 1.55 First, in Section 5.3.1 we describe the relation between the Floer complexes we use in this paper (described in Section 4.8), where the differentials, continuation maps, and so on, carry weights which are suitable powers of the Novikov parameter T , and “classical” Floer complexes, wherein maps carry no such weights. We then recall the definition of the classical symplectic cohomology in Section 5.3.2, then in Section 5.3.3 we prove the first assertion of Theorem 1.55, which states that SH∗(K; Λ) = SH∗(Kb; Λ) for a contact-type domain with incompressible index-bounded boundary, and finally in M Section 5.3.4 we prove the containment assertion for ker resK . Throughout this section (M, ω) is symplectically aspherical. We also omit the almost complex structures from the notation.

5.3.1 Classical and weighted Floer cohomology Here we establish an isomorphism between the classical and weighted Floer theories. Given a nondegenerate Hamiltonian H, its classical Floer cochain complex over Λ is

∗ M CFcl(H; Λ) = Λ · x x∈P◦(H)

68 carrying the differential X dclx = #M(H; x, y) y , y∈P◦(H) where M(H; x, y) is the moduli space of Floer trajectories of H from x to y of index difference 1. As described in Section 4.8, the complex we use in this paper is the Λ≥0-module ∗ M CF (H) = Λ≥0 · x x∈P◦(H) with the weighted differential

X AH (y)−AH (x) dclx = #M(H; x, y)T y , y∈P◦(H) where AH is the action functional. The relation between the two complexes is given by the chain isomorphism

∗ ∗ −AH (x) CF (H) ⊗Λ≥0 Λ → CFcl(H; Λ) , x 7→ T x .

∗ ∗ Letting HFcl(H; Λ) be the cohomology of (CFcl(H; Λ), dcl), by the flatness of Λ we obtain the isomorphism

∗ ∗ ∗ ∼ ∗ ∗ ∗ HF (H) ⊗ Λ = H (CF (H) ⊗ Λ) = H (CFcl(H; Λ)) = HFcl(H; Λ) . (17)

The above isomorphism also intertwines continuations maps, and thus yields a complete identification between classical Floer theory with coefficients in Λ and its weighted counterpart. Moreover, given a negative C2-small Hamiltonian H, consider a positive decreasing ∗ sequence si → 0 and construct SC (M; Λ) using the acceleration datum (siH)i. Then from the above equation, together with a computation of the injective limit it follows that ∗ ∼ ∗ SC (M; Λ) = CFcl(H; Λ) , (18) see [22].

5.3.2 Symplectic cohomology of domains with contact-type boundary Here we take K ⊂ M to be a domain with contact-type incompressible boundary. We denote by Y the Liouville vector field defined on a neighborhood of Σ = ∂K. We consider the symplectomorphism between a neighborhood of Σ and the subset Σ × (1 − ρ , 1 + ε)r of the symplectization of (Σ, ιY ω|Σ) given by the flow of Y , so that φY (z)

69 corresponds to (z, eρ); in particular ρ = ln r. Recall the index boundedness condition, Definition 1.51. Throughout we pick a generic almost complex structure which is cylindrical near Σ, that is in Liouville coordinates in the neighborhood Σ × (1 − , 1 + ) it is r-invariant, preserves the contact structure on Σ, and maps Y to the Reeb vector field on Σ × {1}. We omit it from notation. Let λ = ιY ω be the Liouville form on Σ×(1−, 1+). Let Kb denote the completion of K, obtained by attaching the positive end ([1, ∞) × Σ, d(rλ|Σ)) of the symplectization of Σ to K along Σ. Let us first precisely define the cohomology SH(Kb; Λ) in our setting. µ µ We consider for µ > 0 Hamiltonians Hf on Kb such that Hf |K = −1/µ and µ Hf |Σ×[1,∞) = µr − µ − 1/µ. We then pick smooth non-degenerate time-dependent perturbations Hµ of these Hamiltonians, such that • over K they are obtained from Hfµ by multiplying by a positive C2-small Morse perturbation, so that in particular Hµ < 0 and Hµ → 0 as µ → ∞; µ→∞ • they have all their nontrivial 1-perioidic orbits in the region [1, εµ] with εµ −−−→ 0;

• and they are all linear in r for r ≥ 1 + εµ.

We choose a strictly increasing sequence µn which tends to +∞ as n → ∞, and choose the perturbations such that Hµn is an increasing sequence. We consider the Floer complexes CF ∗(Hµ), defined in Section 5.3.1, and for each n we define a continuation ∗ µn ∗ µn+1 s map Φn: CF (H ) → CF (H ) by choosing a generic monotone homotopy (Hn)s∈R and counting the corresponding continuation solutions, weighted by their topological energy, namely,

X µ 0 µ s 0 AH n+1 (γ )−AH n (γ) 0 Φn(γ) = #M((Hn)s; γ, γ ) · T · γ . γ0∈P◦(Hµn+1 ) These yield a 1-ray:

Φ Φ CF ∗(Hµ1 ) −→1 CF ∗(Hµ2 ) −→2 CF ∗(Hµ3 ) → ... We put   SC∗(Kb) = lim CF ∗(Hµn ) and SH∗(Kb; Λ) := H∗ SC∗(Kb) ⊗Λ = H(SC(Kb)⊗Λ) . µn→∞

∗ The “classical” symplectic cohomology SHcl(Kb; Λ) is defined with unweighted dif- ferentials and continuation maps, namely:   ∗ ∗ µn SHcl(Kb; Λ) := H lim CFcl(H ; Λ) . µn→∞

70 ∗ ∗ Since by equation (17), HFcl(H; Λ) = HF (H) ⊗Λ≥0 Λ, the two versions agree:

∗ ∗ SHcl(Kb; Λ) = SH (Kb; Λ) .

In the next section we will show that SH∗(K; Λ) coincides with SH∗(Kb; Λ).

5.3.3 SH∗(K; Λ) = SH∗(Kb; Λ) In this and in the following subsections, we will use the following fact, whose proof is an immediate verification from the definitions:

Lemma 5.2. Consider a 1-ray of modules Ci over Λ≥0,

ψ1 ψ2 ψ3 C1 −→ C2 −→ C3 −→· · · .

Let x1 ∈ C1 and denote its images in the modules Ci by xi, namely, xi = ψi−1(xi−1). c Assume that there exists c > 0, so that for all i, ψi(xi) ∈ T · Ci+1. Then, the image of x in lim\C is zero, and moreover, if for every i, we have 1 −→ i ψ (C ) ⊂ T c · C , then lim\C = 0 i i i+1 −→ i To compute SH∗(K; Λ), we construct a suitable acceleration datum where we can separate the generators lying inside and outside of K by action. We keep the notations from the beginning of the previous subsection. Choose a positive decreasing sequence εn → 0. We define a sequence of smooth autonomous Hamiltonians as follows, referring to the r coordinate of the point x where applicable:  −1/n x ∈ K  Monotone increasing r ∈ (1, 1 + εn]  Of the form ar + b Hen(x) = r ∈ (1 + εn, 1 + 3εn) (19) for a noncharacteristic a   Monotone increasing r ∈ [1 + 3εn, 1 + 4εn)  n x∈ / K ∪ Σ×(1, 4εn]

2 To achieve non-degeneracy we then perturb Hen by multiplying Hen by a C -small 2 c Morse function in K, by adding a C -small Morse function in (K ∪ Σ×(1, 1 + 4εn)) and perform a small time-dependent perturbation in the nonlinear parts, leaving the linear part with slope a untouched. We denote the perturbed Hamiltonian by Hn. Moreover, one can assume that the critical points of the Morse perturbations do not depend on n.

71 Note that the 1-periodic orbits of Hn fall into four groups. The first two are Morse critical points that lie in K, which we denote by PL (for “Low”) and Morse critical points which lie in the complement of K ∪ Σ×(1, 1 + 4εn] which we denote by PH (for “High”). The rest are 1-periodic orbits corresponding to the Reeb orbits on Σ, which

↑ fall into two groups, P∂K↓ , the “low” orbits, lying in Σ×(1, 1+εn) and P∂K , the “high” orbits, lying in Σ × (1 + 3εn, 1 + 4εn). They are all depicted in Figure 1.

P∂K↑ PE

K   1 1+2εn 1+4εn

PI P ∂K↓

Figure 1: The Hamiltonian Hn and the four distinguished sets of 1-periodic orbits

For what follows we fix an index j ∈ Z and restrict our attention to orbits of indices j − 1, j, j + 1.

Notation 5.3. Given an index j and a Z-graded cochain complex A∗, we let A∼j be the truncation of A∗ to degrees j − 1, j, j + 1. Note that Hj(A∗) = Hj(A∼j).

It follows that in order to compute SHj(K; Λ), it is enough to consider the complexes ∼j CF (Hn). Note that for n large enough the actions of the orbits in PH ∪ P∂K↑ (having indices j − 1, j, j + 1) are strictly larger than those of the orbits in P∂K↓ ∪ PL, thanks to the index boundedness; this is due to the fact that the action of a Hamiltonian orbit corresponding to a Reeb orbit is given by the y-intercept of the tangent to the graph of Hamiltonian in the r coordinate. We may pick acceleration data with n large enough so that the action separation guarantees two essential properties:

72 • There are no Floer trajectories going from elements in P∂K↑ ∪ PH to elements

in P∂K↓ ∪ PL, since the Floer differential is action increasing; it follows that the ∼j submodule of CF (Hn) generated P∂K↑ ∪ PH is in fact a subcomplex, which ∼j ∼j ∼j ∼j we denote by CF+ (Hn). We let CF− (Hn) = CF (Hn)/CF+ (Hn) be the corresponding quotient complex, which is the free Λ≥0-module whose basis is

formed by the elements of PL ∪ PK↓ of indices j − 1, j, j + 1.

• Since the homotopy from Hn to Hn+1 is increasing, continuation trajectories in- ∼j ∼j crease action, therefore no continuation trajectory from CF (Hn) to CF (Hn+1)

sends an element of P∂K↑ ∪ PH to P∂K↓ ∪ PL, that is the continuation morphisms ∼j ∼j map CF+ (Hn) into CF+ (Hn+1). Consider the following commutative diagram of cochain complexes and cochain maps, where the rows are short exact sequences while the middle vertical arrows are the continuation maps, while the left and right ones are induced from them:

∼j i ∼j p ∼j 0 / CF+ (Hn) / CF (Hn) / CF− (Hn) / 0

   ∼j i ∼j p ∼j 0 / CF+ (Hn+1) / CF (Hn+1) / CF− (Hn+1) / 0

   ......

One can verify from the definitions that taking telescope of the 1-rays involved preserves the exactness, namely we obtain the following short exact sequence:

∼j i ∼j p ∼j 0 / tel CF+ (Hn) / tel CF (Hn) / tel CF− (Hn) / 0 n n n

Note that each element in this short exact sequence is a free Λ≥0 module, in par- ticular the rightmost module, being free, is flat, which implies that for each r > 0, 1 ∼j  Tor teln CF− (Hn), Λ[0,r) = 0, therefore using the long exact sequence associated to the Tor functors, we obtain an exact sequence:

∼j i ∼j p ∼j 0 → tel CF+ (Hn) ⊗ Λ[0,r) −→ tel CF (Hn) ⊗ Λ[0,r) −−→ tel CF− (Hn) ⊗ Λ[0,r) → 0 . n n n Now we show that taking a projective limit over r still yields a short exact sequence. 0 Note that for all r > r, the projection Λ[0,r0) → Λ[0,r) is surjective, therefore by the right exactness of tensor products, so is the induced map N ⊗ Λ[0,r0) → N ⊗ Λ[0,r) for

73 each module N. This is a special case of the Mittag–Leffler condition, by which lim ←− preserves exactness, thus the following sequence is exact:

∼j i ∼j p ∼j 0 / telc CF+ (Hn) / telc CF (Hn) / telc CF− (Hn) / 0 . n n n We now analyze the homologies of each chain complex in this sequence. First, we can assume without loss of generality that the Hamiltonians Hn are such that the ∼j c ∼j continuation morphisms map CF+ (Hn) into T CF+ (Hn+1) for some fixed c > 0, which implies by Lemma 5.2 that the complex lim CF ∼j(H ), is acyclic, and therefore −→c n n ∼j so is the complex teldn CF+ (Hn), since the two are quasi-isomorphic. It follows that p is a quasi-isomorphism. ∼j On the other hand, the differential of CF− (Hn) counts Floer trajectories connect- ing orbits in P∂K↓ ∪ PL. Since our almost complex structures are cylindrical in the “neck” region Σ×(1, 1 + 4εn), by the “no escape” lemma, [3, 8], all these Floer so- lutions are contained in K, and the same applies to the continuation solutions go- ∼j ing between those generators. It follows that teldn CF− (Hn) is quasi-isomorphic to SC\∼j(K) = lim CF ∼j(H ). It can be checked that in our case SC∼j(K) is complete, b −→c n − n b namely SC∼j(Kb) ∼= SC\∼j(Kb). Therefore j j ∼j  SH (K; Λ) = H telcnCF (Hn) ⊗ Λ j ∼j  = H telcnCF− (Hn) ⊗ Λ p is a quasi-isomorphism = Hjlim CF ∼j(H ) ⊗ Λ −→c n − n = Hj lim CF ∼j(H ) ⊗ Λ lim CF ∼j(H ) is complete −→ − n −→ − n n n = Hj lim(CF ∼j(H ) ⊗ Λ) −→ − n n = Hj lim CF ∼j(H ; Λ) −→ cl n n j j = SHcl(Kb; Λ) = SH (Kb; Λ) .

M 5.3.4 The kernel of resK Here we prove the last assertion of Theorem 1.55, that is ∗ ∗ M ker (H (M; Λ) → H (K; Λ)) ⊂ ker resK .

We retain the acceleration datum (Hn)n specified in the previous subsection. We com- M ∗ ∗ pute the restriction map resK : SC (M) → SC (K) using a suitable sequence of Hamil- ∗ tonians Ln to compute SC(M), and monotone homotopies from Ln to Hn. The Hamil- tonians Ln are assumed to satisfy the following:

74 2 (i) Fix a C -small everywhere negative Morse function F , and put Ln = snF , where sn → 0 a decreasing positive sequence; (ii) The values of F on K are strictly smaller than those on Kc;

(iii) Along ∂K the gradient of F with respect to a fixed Riemannian metric is outward pointing;

2 (iv) By the C -smallness assumption, the only 1-periodic orbits of Ln are the critical points of F ;

(v) Their levels are in relation with those of Hn as seen in Figure 2.

P∂K↑ PE

K   1 1+4εn 1+δn

P∂K PI ↓ PH

PL

Figure 2: The Hamiltonians Ln and Hn with the distinguished sets of 1-periodic orbits

We denote the critical points of Ln that lie in K by PI (for “interior”) and those c that lie in (K ∪ Σ×(1, 1 + δn)) by PE (for “Exterior”). Note that the actions of the ∗ ∗ generators of CF (Ln) in PE are greater than those of the generators of CF (Hn) in ∼j P∂K↓ ∪ PL. Define CF+ (Ln) to be the subcomplex generated by the generators in PE ∼j ∼j ∼j and let CF− (Ln) := CF (Ln)/CF+ (Ln) be the quotient complex. Again, by action considerations, and arguing similarly to the end of Section 5.3.3 we obtain the following

75 homotopy-commutative diagram

∼j i ∼j p ∼j 0 / telc CF+ (Ln) / telc CF (Ln) / telc CF− (Ln) / 0 n n n (20)    ∼j i ∼j p ∼j 0 / telc CF+ (Hn) / telc CF (Hn) / telc CF− (Hn) / 0 n n n

∼j Let us identify the relevant complexes. By a calculation as in [23], teldn CF (Ln) j j ∼j computes SH (M) = H (M) ⊗ Λ>0. The complex telc CF− (Ln) is quasi-isomorphic n to lim CF ∼j(L ) = Λ ⊗ {x ∈ P | x has index j − 1, j, j + 1}, where the differential −→dn − n >0 I counts Morse flow lines of F between the critical points of F in PI . It follows that this j complex computes H (K) ⊗ Λ>0. Similarly to the considerations of Section 5.3.3, the ∼j complex telcn CF+ (Hn) is acyclic. Therefore, taking the homology of the right square of (20) we obtain the commutative diagram

j j i∗ j SH (M) = H (M) ⊗ Λ>0 / H (K) ⊗ Λ>0

M resK  =∼  SHj(K) / · where i∗ is the restriction on singular cohomology. It follows from this diagram that ∗ M ker i ⊂ ker resK . This completes the proof of Theorem 1.55.

5.4 Proof of Theorem 2.4 Here we prove Theorem 2.4, which asserts the existence of an exact triangle relating SH∗(M,K; Λ), H∗(M,K; Λ), and SH+(Kb; Λ), which is an invariant defined, for in- stance in [3]. The complex computing this invariant is obtained by quotienting SC∗(Kb) µn by the subcomplex generated by the critical points of the Morse functions (H |K )n, see Section 5.3.2. We repeat the arguments of the previous section, only this time we impose an extra condition on Ln and Hn: they are chosen so that they coincide in K. Picking up from the commutative diagram (20), labeling arrows relevant to us and identifying the chain complexes with known ones, we arrive at:

L ∼j i ∼j p ∼j 0 / C (M,K) ⊗ Λ>0 / C (M) ⊗ Λ>0 / C (K) ⊗ Λ>0 / 0

Φ+ Φ ΦQ   H  ∼j i ∼j p ∼j 0 / telc CF+ (Hn) / SC (K) / SC (Kb) / 0 . n qis

76 Note that pH is a quasi-isomorphism. Passing to cones of the vertical morphisms, we obtain the following 3x3 square of exact sequences (both horizontal and vertical)

0 0 0

  H  ∼j i ∼j p ∼j 0 / telc CF+ (Hn) / SC (K) / SC (Kb) / 0 n M qis

   . 0 / cone(Φ+) / cone(Φ) / cone(ΦQ) / 0

 i  pL  0 / C∼j(M,K)[1] / C∼j(M)[1] / C∼j(K)[1] / 0

   0 0 0 Let us identify the cones with known complexes. First, cone(Φ) = SC∼j(M,K)[1] +,∼j by definition. Second, cone(ΦQ) is quasi-isomorphic to SC (Kb) which can be seen as follows: The complex C∼j(K) can be identified with the subcomplex of SC∼j(Kb) generated by the orbits which are the critical points of the Hamiltonians on K, and ΦQ with the corresponding inclusion map. Since the cone of the inclusion of a subcom- plex, which is a direct summand, is quasi-isomorphic to the quotient complex (see e.g. +,∼j Exercise 1.59 on p.38 in [24]), cone(ΦQ) is quasi-isomorphic to SC (Kb). The direct summand condition holds in our case since the suitable exact sequence consisting of telescopes splits due to the fact that the telescopes are all free, and a splitting persists to the completion. ∼j Finally, note that telcn CF+ (Hn) is acyclic by the arguments of Section 5.3.2, hence ∼j cone(Φ+) → C (M,K)[1] is a quasi-isomorphism. Therefore the long exact sequence obtained by combining those associated to the middle row for all j, we obtain the desired triangle ∗ H (M,K; Λ) / SHM (M,K) . (21) h

[+1] v SH+(Kb)[−1]

5.5 Proof of Theorem 1.54 Recall that the theorem states the following: assume that K is a domain with boundary Σ whose neighborhood is the image of a smooth embedding (−, ) × Σ ,→ M

77 extending id : Σ = {0} × Σ → M, such that no {ρ} × Σ carries closed characteristics ∗ ∗ M ∗ ∗ contractible in M; then SH (K; Λ) = H (K; Λ) and resK : SH (M; Λ) → SH (K; Λ) coincides with the restriction on singular cohomology H∗(M; Λ) → H∗(K; Λ).

Remark 5.4. Note that K satisfying these assumptions implies that M \ K satisfies them as well.

We pick acceleration data for K and for M similarly to the previous section, only this time r is the first coordinate of the aforementioned embedding (−, ) × Σ ,→ M. Similarly to the previous subsection, we require that Ln and Hn coincide in K. The conditions on the closed characteristics near ∂K imply that the only 1-periodic orbits of Hn are the critical points in PL and PH . We use the same notation PL for the critical points on Ln in K, as well, since the functions coincide in K. See Figure 3.

PH

No contractible periodic orbits

K   1 1+4εn

PE PL

Figure 3: The Hamiltonians Ln and Hn with the distinguished sets of 1-periodic orbits

As in subsection 5.3.4, we obtain two short exact sequences with morphisms between them ∗ i ∗ p ∗ 0 / telc CF+(Ln) / telc CF (Ln) / telc CF−(Ln) / 0 n n n

H ΦL    ∗ i ∗ p ∗ 0 / telc CF+(Hn) / telc CF (Hn) / telc CF−(Hn) / 0 . n n n

78 ∗ ∗ ∗ ∗ As before, teldn CF+(Hn) is acyclic, teldn CF (Ln) computes SH (M) = H (M)⊗Λ>0, ∗ ∗ H while teldn CF−(Ln) computes H (K) ⊗ Λ>0. What is different is the map ΦL , which we − ∗ will now show to be a quasi-isomorphism. Fix n and consider the map Φn : CF−(Ln) → ∗ CF−(Hn). Since Ln ≡ Hn on K, the constant Floer solution sitting at a genera- tor x ∈ PL contributes to the continuation map and is the only energy zero con- − necting trajectory, therefore Φn (x) = x+ terms with positive powers of T . This − means that the reduction of Φn to the residue field is just the identity, in particu- − lar invertible, which in turn implies that Φn is invertible. Therefore the 2-ray con- ∗ ∗ sisting of the complexes CF−(Ln),CF−(Hn) has the property that the maps in the finite direction are all isomorphisms. It is not hard to show that the induced map ∗ ∗ tel CF−(Ln) → tel CF−(Hn) is then an isomorphism, therefore so is the map on com- H ∗ ∗ pletions ΦL : telcnCF−(Ln) → telcnCF−(Hn). Taking the homology of the right square, we arrive at the following commutative diagram: ∗ ∗ i∗ ∗ SH (M) = H (M) ⊗ Λ>0 / H (K) ⊗ Λ>0

M ∼ resK =  =∼  SH∗(K) / · Composing the right arrow and the inverse of the bottom arrow, we conclude that ∗ ∼ ∗ SH (K) = H (K) ⊗ Λ>0 and that under this isomorphism the restriction maps on M ∗ singular and symplectic cohomologies coincide. In particular ker resK = ker i .

5.6 Proof of Theorem 1.52 Recall that the theorem states that if (M, ω) is symplectically aspherical, K ⊂ M is a compact heavy set, and there is a sequence of contact-type domains Wi with incompressible index-bounded boundary, all containing K in the interior, such that K = T M T M i Wi, then for all i, [Vol] ∈ ker res c , and in particular [Vol] ∈ τ(K) = i ker res c , Wi Wi therefore K is SH-heavy. For the rest of the proof we let W be one of the Wi. Varolgunes proves in [23] that ∗ ∗ SH (M) = H (M) ⊗ Λ>0. We will prove the following Claim 5.5. There exists µ > 0 such that

T µ[Vol] ∈ ker SH∗(M) → SH∗(W c) .

That [Vol] ∈ ker resM follows from this upon tensoring with Λ. W c Let us first recall what it means for a set to be heavy. Like in Section 5.3.1, given a ground field F, the classical Floer cochain complex of a nondegenerate Hamiltonian

79 ∗ ◦ H is the F- CFcl(H; F) = FhP (H)i. Since the differential increases the ∗ ◦ action, for each a ∈ R, the subspace CFcl,≥a(H; F) generated by x ∈ P (H) with ∗ ∗ AH (x) ≥ a is a subcomplex. Let HFcl,≥a(H; F) be the cohomology of CFcl,≥a(H; F) ∗ ∗ ∗ and let ja: HFcl,≥a(H; F) → HFcl(H; F) be the morphism induced by the inclusion. ∗ ∗ ∗ Letting PSS : HFcl(H; F) → QH (M; F) = H (M; F) be the PSS isomorphism [17], we have the cohomological spectral invariants

∨ ∗ ∗ c (A, H) = sup{a ∈ R | A ∈ im(PSS ◦ja)} for A ∈ QH (M; F) . The corresponding partial symplectic quasi-state ζ: C∞(M) → R is given by c∨([Vol], kH) ζ(H) = lim , k→∞ k where [Vol] ∈ H2n(M; F) is the volume class. The original definition in [7] used homo- logical spectral invariants c([M], ·) relative to the [M] ∈ H2n(M; F), but c([M], ·) = c∨([Vol], ·) thanks to the duality formula, see for instance [13, Section 4.2]. For K to be heavy means that for each F ∈ C∞(M) we have ζ(F ) ≥ min F. K For us, the crucial consequence of the assumption that K is heavy is the following

∞ Lemma 5.6. For any L > 0 there exists F ∈ C (M) such that F |W c < 0 and such that c∨([Vol],F ) > L.

∞ Proof. Let F0 ∈ C (M) satisfy F0|W c < 0 and F0|K > 0. Since K is heavy, we have ∨ c ([Vol], kF0) 0 < min F0 ≤ ζ(F0) = lim . K k→∞ k ∨ In particular there is k0 such that c ([Vol], k0F0) > L. Now put F = k0F0. Let us identify a neighborhood of the boundary Σ = ∂W with the piece Σ × (1 − , 1 + )r of the symplectization of Σ, so that ∂r is the Liouville vector field. Let L be the maximum of the actions of all the Reeb orbits on Σ, contractible in M, of Conley– Zehnder indices 2n − 1, 2n, 2n + 1, and fix F as in the lemma for this L. Then there c exists an acceleration datum (Hi)i≥0 for W such that for all i we have:

• Hi ≥ F ;

2 • On W \ Σ × (1 − , 1], Hi is a C -small Morse function plus a constant depending c 2 on i, and on W , Hi equals a C -small Morse function plus a constant depending on i; in particular the only 1-periodic orbits of Hi on M \ Σ × (1 − , 1) are the critical points of these Morse functions;

80 • In the region Σ × (1 − , 1), Hi is a sufficiently small perturbation of fi ◦ r, where fi: (1 − , 1) → R is a strictly decreasing smooth function;

• The 1-periodic orbits of Hi fall into two groups, the “upper” and the “lower” orbits; the upper orbits are situated in W \Σ×(1−2/3, 1], and these in particular include the critical points of the Morse function Hi|W \Σ×(1−,1]; the “lower” orbits are in W c ∪ Σ × [1 − /3, 1], and these include the critical points of the Morse

function Hi|W c ;

• The lower orbits of Hi of index 2n have actions < L, while all the upper orbits of index 2n have actions > L;

• There exists c > 0 such that the minimal action of an upper orbit of Hi+1 of index 2n is at least c+ the maximal action of an upper orbit of Hi of index 2n. See Figure 4 for an illustration.

L upper orbits

lower orbits

1

Figure 4: The action separation between the lower and upper orbits is demonstrated on the on the Hamiltonians Hn. The value L is marked. The tangents whose y intercepts are the actions are depicted in gray, as well as dots representing the actions of the Morse critical points.

This acceleration datum yields a 1-ray of Floer complexes and continuation maps

ΦH1 ∗ H0 ∗ CF (H0) −−→ CF (H1) → ...

81 2n Lemma 5.7. Let y ∈ CF (H0) be a linear combination of the upper orbits of H0. Then it is killed by the composition

Φ CF ∗(H ) −−→H0 lim CF ∗(H ) −→b· lim CF ∗(H ) , 0 −→i i −→c i i where ΦH0 is the natural map into the direct limit.

Proof. The map ΦH0 factors as follows:

H j H1 ΦH ◦···◦ΦH ΦH CF ∗(H ) −−−−−−−−→j−1 0 CF ∗(H ) −−→j lim CF ∗(H ) , 0 j −→ i i where ΦHi is again the natural map into the direct limit. By the action separation property of the acceleration datum (H ) ,ΦHi+1 : CF 2n(H ) → CF 2n(H ) maps any i i Hi i i+1 linear combination z of upper orbits of Hi to a linear combination of upper orbits of H , since it is action-increasing. Thus ΦHi+1 (z) ∈ T c · CF 2n(H ), and by induction i+1 Hi i+1 we have (ΦHj ◦ · · · ◦ ΦH1 )(y) ∈ T jc · CF ∗(H ) , Hj−1 H0 j whence Hj H1  jc ∗ ΦH (y) = ΦH (Φ ◦ · · · ◦ Φ )(y) ∈ T lim CF (Hi) 0 j Hj−1 H0 −→ i for all j, that is \ Φ (y) ∈ T λ lim CF ∗(H ) ⊂ ker · . H0 −→ i b λ≥0 i

2 We now fix a C -small negative Morse function G: M → R, let (si)i≥0 be a strictly decreasing sequence of positive numbers < 1, and put Gi = siG. Then (Gi)i≥0 is ∗ an acceleration datum for M. The Floer complexes CF (Gi) are all isomorphic to ∗ the Morse complex CM (G) ⊗ Λ≥0 with the differential weighted by suitable powers ∗ ∗ of T , while the continuation map CF (Gi) → CF (Gi+1) is given by Crit G 3 p 7→ T Gi+1(p)−Gi(p)p. It follows that lim CF ∗(G ) is canonically isomorphic to CM ∗(G)⊗Λ −→i i >0 with the usual Morse differential; in particular it is a complete Λ≥0-module. The natural map CM ∗(G)⊗Λ = CF ∗(G ) → lim CF ∗(G ) = CM ∗(G)⊗Λ is given by Crit G 3 ≥0 j −→i i >0 p 7→ T −Gj (p)p. ∗ ∗  ∗ In particular SH (M) = H CM (G)⊗Λ>0 = HM (G)⊗Λ>0 and if q is a maximum −G0(q) ∗ −G0(q) ∗ of G, then [Vol]⊗T ∈ SH (M) is represented by T q ∈ CM (G)⊗Λ>0 and is the image of q ∈ CF ∗(G ) under the natural map CF ∗(G ) → lim CF ∗(G ). Note that 0 0 −→i i we can assume that Gi ≤ Hi for all i, therefore as in Section 4.8 there exists a filling between the acceleration data (Gi)i, (Hi)i and in particular we have the corresponding continuation maps ΦHi : CF ∗(G ) → CF ∗(H ). Gi i i

82 2n Lemma 5.8. Let q ∈ CF (G0) be a maximum. Then there is µ ≥ −G0(q) such that T µ+G0(q)ΦH0 (q) is cohomologous in CF ∗(H ) to a linear combination of upper orbits. G0 0 Proof of Claim 5.5. We have the following commutative diagram:

∗ ∗ ∗ ∗ CF (G0) / telciCF (Gi) / limdiCF (Gi) = limi CF (Gi)

ΦH0 G0  ∗ ∗ ∗ CF (H0) / telciCF (Hi) / limdiCF (Hi)

The middle vertical arrow comes from the 2-ray of Floer complexes coming from the ∗ aforementioned filling. The top left arrow is the inclusion of CF (G0) into the second ∗ L ∗ L ∗ direct sum in teli CF (Gi) = i CF (Gi)[1] ⊕ i CF (Gi), followed by completion, and similarly for the bottom left arrow. It follows that the square commutes. The top ∗ right arrow is the completion of the composition of the projection of teli CF (Gi) onto L ∗ i CF (Gi) followed by the natural map into the direct limit. Varolgunes proves in [23] that this arrow is a quasi-isomorphism. The same considerations hold for the bottom right arrow. Note that the composition of the top arrows is the natural map into the direct limit, and the same is true for the composition of the bottom arrows, except the completion is then tagged on. Fix µ as in Lemma 5.8. We will show that T µ[Vol] is killed by SH∗(M) → SH∗(W c). µ 2n 2n 2n ∗  The element T [Vol] ∈ SH (M) = H (M) ⊗ Λ>0 = H telciCF (Gi) is represented µ+G0(q) 2n by the image of T q ∈ CF (G0) under the top left arrow. It follows that the im- age of T µ[Vol] by SH∗(M) → SH∗(W c) is represented by the image of ΦH0 (T µ+G0(q)q) ∈ G0 CF 2n(H ) in lim CF ∗(H ). By Lemma 5.8, ΦH0 (T µ+G0(q)q) = T µ+G0(q)ΦH0 (q) is coho- 0 di i G0 G0 mologous to a linear combination of upper orbits of H0, and thus by Lemma 5.7, this ∗ ∗ linear combination is killed by the map CF (H0) → limdiCF (Hi). It follows that the image of T µ+G0(q)ΦH0 (q) in tel CF ∗(H ) is cohomologous to zero, and thus we have our G0 ci i claim.

83 Proof of Lemma 5.8. Consider the following commutative diagram:

∗ CFcl(G0; F)

∗ ∗  ( ∗ CFcl,≥0(G0; Λ) / CFcl(G0; Λ) CFcl(H0; F)

( ∗ ( ∗  CFcl,≥0(H0; Λ) / CFcl(H0; Λ)

∗ ∗  CF (G0) / CF (G0) ⊗ Λ

) ∗ ( ∗  CF (H0) / CF (H0) ⊗ Λ

∗ Here for a nondegenerate Hamiltonian F we have a valuation on CFcl(F ; Λ) given λ ∗ by T x 7→ λ + AF (x) and CFcl,≥0(F ; Λ) stands for the Λ≥0-submodule of elements with nonnegative valuation. The oblique arrows are continuation maps in respective Floer theories (classical or T -weighted), the upper vertical arrows are the embeddings induced by the field extension F ⊂ Λ, the horizontal arrows are inclusions, while the rest of the vertial arrows are the isomorphism ψ from Section 5.3.1. For instance ∗ ∗ AG (x) ψ: CFcl(G0; Λ) → CF (G0) ⊗ Λ is defined by x 7→ T 0 x, and similarly for H0. It is ∗ ∗ easy to check that ψ maps CFcl,≥0(G0; Λ) isomorphically onto CF (G0) and same for H0. 2n Let now q ∈ CFcl (G0; F) be a maximum, that is a representative of the volume class. Since continuation maps commute on cohomology with the PSS isomorphisms, it follows 2n ∗ that ye := Φ(q) ∈ CFcl (H0; F) also represents the volume class, where Φ: CFcl(G0; F) → ∗ CFcl(H0; F) is the continuation map on classical Floer cohomology. On the other hand, ∨ Lemma 5.6 and the assumption H0 ≥ F imply that c ([Vol],H0) > L, therefore there is 2n a representative of the volume class y ∈ CFcl (H0; F) consisting of orbits of action > L, and by our choice of acceleration datum this forces y to be a linear combination of the 2n−1 upper orbits of H0. It follows that there is b ∈ CFcl (H0; F) such that ye − y = dclb. ∗ We can now look at this equation in CFcl(H0; Λ). The elements y,e b do not necessarily ∗ µ µ ∗ lie in CFcl,≥0(H0; Λ). Let µ ≥ −G0(q) > 0 be such that T y,e T b ∈ CFcl,≥0(H0; Λ). Applying ψ and noting that it is Λ≥0-linear and commutes with continuation maps, we ∗ obtain the following equation in CF (H0): d(T µψ(b)) = ψ(T µΦ(q))−ψ(T µy) = T µΦH0 (ψ(q))−ψ(T µy) = T µ+G0(q)ΦH0 (q)−ψ(T µy) . G0 G0

84 This means that T µ+G0(q)ΦH0 (q) is cohomologous in CF ∗(H ) to ψ(T µy), which is a G0 0 linear combination of the upper orbits of H0, as claimed.

Acknowledgments: We thank Shachar Carmeli, Liran Shaul and Moshe White for useful discussions about higher category theory, completions of modules, and combina- torial geometry, respectively.

References

[1] P. Biran and O. Cornea. Quantum structures for Lagrangian submanifolds. arXiv preprint arXiv:0708.4221, 2007.

[2] R. Bott, L. W. Tu, et al. Differential forms in algebraic topology, volume 82. Springer, 1982.

[3] K. Cieliebak and A. Oancea. Symplectic homology and the Eilenberg–Steenrod axioms. Algebraic & , 18(4):1953–2130, 2018.

[4] J. F. Davis and P. Kirk. Lecture notes in algebraic topology, volume 35. American Mathematical Soc., 2001.

[5] H. Edelsbrunner. Algorithms in combinatorial geometry, volume 10. Springer Science & Business Media, 2012.

[6] M. Entov and L. Polterovich. Quasi-states and symplectic intersections. Commen- tarii Mathematici Helvetici, 81(1):75–99, 2006.

[7] M. Entov and L. Polterovich. Rigid subsets of symplectic manifolds. Compositio Mathematica, 145(3):773–826, 2009.

[8] Y. Ganor and S. Tanny. Floer theory of disjointly supported Hamiltonians on symplectically aspherical manifolds. arXiv preprint arXiv:2005.11096, 2020.

[9] J. P. Greenlees and J. P. May. Derived functors of I-adic completion and local homology. Journal of Algebra, 149(2):438–453, 1992.

[10] M. Gromov. Singularities, expanders and topology of maps. part 1: Homology ver- sus volume in the spaces of cycles. Geometric and Functional Analysis, 19(3):743– 841, 2009.

85 [11] M. Gromov. Singularities, expanders and topology of maps. part 2: From combina- torics to topology via algebraic isoperimetry. Geometric and Functional Analysis, 20(2):416–526, 2010. [12] R. N. Karasev. Covering dimension using toric varieties. Topology and its Appli- cations, 177:59–65, 2014. [13] R. Leclercq and F. Zapolsky. Spectral invariants for monotone Lagrangians. Journal of Topology and Analysis, 10(03):627–700, 2018. [14] D. McDuff and D. Salamon. J-holomorphic curves and symplectic topology, vol- ume 52. American Mathematical Soc., 2012. [15] J. Pardon. An algebraic approach to virtual fundamental cycles on moduli spaces of pseudo-holomorphic curves. Geometry & Topology, 20(2):779–1034, 2016. [16] J. Pardon. Contact homology and virtual fundamental cycles. Journal of the American Mathematical Society, 32(3):825–919, 2019. [17] S. Piunikhin, D. Salamon, and M. Schwarz. Symplectic Floer-Donaldson theory and quantum cohomology. In Contact and symplectic geometry, pages 171–200. Cambridge Univ. Press, Publ. Newton Instit. 8, Cambridge, 1996. [18] L. Polterovich, D. Rosen, K. Samvelyan, and J. Zhang. Topological persistence in geometry and analysis, volume 74. American Mathematical Soc., 2020. [19] R. Rado. A theorem on general measure. Journal of the London Mathematical Society, 1(4):291–300, 1946. [20] Y. B. Rudyak and F. Schlenk. Minimal atlases of closed symplectic manifolds. Communications in Contemporary Mathematics, 9(06):811–855, 2007. [21] D. Tonkonog and U. Varolgunes. Super-rigidity of certain skeleta using relative symplectic cohomology. arXiv preprint arXiv:2003.07486, to appear in Journal of Topology and Analysis, 2020. [22] U. Varolgunes. Mayer - Vietoris property for relative symplectic cohomology. Mas- sachusetts Institute of Technology, 2018. [23] U. Varolgunes. Mayer–Vietoris property for relative symplectic cohomology. Ge- ometry & Topology, 25(2):547–642, 2021. [24] C. A. Weibel. An introduction to . Number 38. Cambridge university press, 1994.

86 Adi Dickstein Tel Aviv University School of Mathematical Sciences 69978, Tel Aviv, Israel E-mail: [email protected] Yaniv Ganor Technion Department of Mathematics 32000, Haifa, Israel E-mail: [email protected] Leonid Polterovich Tel Aviv University School of Mathematical Sciences 69978, Tel Aviv, Israel E-mail: [email protected] Frol Zapolsky University of Haifa Department of Mathematics Faculty of Natural Sciences 3498838, Haifa, Israel E-mail: [email protected]

87