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

On the Hofer-Zehnder conjecture for semipositive symplectic manifolds

Marcelo S. Atallah and Han Lou
Abstract.

We show that, on a closed semipositive symplectic manifold with semisimple quantum homology, any Hamiltonian diffeomorphism possessing more contractible fixed points, counted homologically, than the total Betti number of the manifold, must have infinitely many periodic points. This generalizes to the semipositive setting the beautiful result of Shelukhin on the Hofer-Zehnder conjecture.

1. Introduction

1.1. Introduction and Main Results

The Hofer-Zehnder conjecture, which concerns the existence of infinitely many periodic points of Hamiltonian systems, originated in the study of celestial mechanics. In one of Poincaré’s remarkable contributions to the three-body problem, he proved the existence of infinitely many periodic solutions provided the planetary masses are sufficiently small. In attempting to generalize such a result for larger planetary masses, he was led to conjecture the existence of at least two fixed points of any time-one map of an area-preserving isotopy of the annulus satisfying a twist condition on the boundary [Poi12]. He proved some particular cases of the conjecture, while Birkhoff established it in general; this is the content of the celebrated Poincaré-Birkhoff theorem [Bir13, Bir26]. It was later shown [BN77] that a similar proof yields the existence of infinitely many periodic points111For ϕ:MM\phi:M\rightarrow M, a point xMx\in M is periodic if there is an integer k1k\geq 1 such that ϕk(x)=x\phi^{k}(x)=x.. In a similar vein, Franks proved that, without the twisting condition, the existence of an interior fixed point is sufficient to guarantee the existence of infinitely many periodic points [Fra92, Fra96]. Furthermore, in what is one of the first results in the direction of the Hofer-Zehnder conjecture, and probably central to its motivation, Franks showed that any time-one map of an area-preserving isotopy of the sphere with at least three fixed points must have infinitely many periodic points. Note that having at least three fixed points is essential since any rotation of the sphere by an irrational fraction of π\pi has only two fixed points, which are also the only periodic points, namely, the North and South poles.

A conjecture by Arnol’d [Arn14, Arn65], relating Hamiltonian dynamics to the topology of the ambient symplectic manifold, has been a driving force of symplectic topology. A homological version of the conjecture states that any non-degenerate222Non-degeneracy means that for every fixed point xx, the linearization dϕxd\phi_{x} of ϕ\phi at xx does not have 11 as an eigenvalue. Equivalently, the graph of ϕ\phi intersects the diagonal Δ\Delta transversely in M×MM\times M. Hamiltonian diffeomorphism ϕ\phi of a closed symplectic manifold must have at least dimH(M;𝕂)\dim\operatorname{\mathrm{H}}_{*}(M;{\mathbb{K}}) fixed points, where 𝕂{\mathbb{K}} is a choice of coefficient field. After some particular cases of the conjecture were established [Eli79, FW85, For85, CZ83, Gro85], Floer proved it for all symplectically aspherical and spherically monotone symplectic manifolds [Flo86, Flo87, Flo89]. To this end, Floer developed a groundbreaking homology theory generated by the contractible orbits of a Hamiltonian diffeomorphism, inspired by the seminal work of Gromov, who introduced the study of pseudo-holomorphic curves to symplectic topology [Gro85]. With the advent of Floer theory the homological lower bound is achieved for the number of contractible fixed points of a Hamiltonian diffeomorphism. Proving the conjecture in full generality, not only in terms of weakening the topological hypothesis on the symplectic manifold but also allowing for more general coefficient rings, is still the subject of ongoing research [HS95, LT98, FO99, Rua99, PSS96, AB21, BX22, Rez22]; all of the proofs use the machinery invented by Floer. From a viewpoint, Franks’s result on S2\mathrm{S}^{2} indicates that having more fixed points than what is required by the Arnol’d conjecture is sufficient to guarantee infinitely many periodic points. Indeed, any time-one map of an area-preserving isotopy on S2\mathrm{S}^{2} is Hamiltonian and, thus, must have at least two fixed points. Hofer and Zehnder conjectured the following for closed symplectic manifolds.

Conjecture 1.1 (Hofer-Zehnder conjecture).

“One is tempted to conjecture that every Hamiltonian map on a compact symplectic manifold (M,ω)(M,\omega) possessing more fixed points than necessarily required by the V. Arnold conjecture possesses always infinitely many periodic orbits…”

In the non-degenerate setting, a natural interpretation of Conjecture 1.1 leads to the statement that any Hamiltonian diffeomorphism ϕ\phi of a closed symplectic manifold (M,ω)(M,\omega) possessing more contractible fixed points Fix(ϕ)\mathrm{Fix}(\phi) than the total Betti number of MM must have infinitely many periodic points. Shelukhin suggested the following interpretation of Conjecture 1.1.

Conjecture 1.2.

Let (M,ω)(M,\omega) be a closed symplectic manifold and 𝕂{\mathbb{K}} a choice of ground field. If ϕ\phi is a (possibly degenerate) Hamiltonian diffeomorphism satisfying

N(ϕ,𝕂)=xFix(ϕ)dim𝕂HFloc(ϕ,x)>dim𝕂H(M;𝕂),N(\phi,{\mathbb{K}})=\displaystyle\sum_{x\in\mathrm{Fix}(\phi)}\dim_{{\mathbb{K}}}\operatorname{\mathrm{HF}}^{\operatorname{\mathrm{loc}}}(\phi,x)>\dim_{{\mathbb{K}}}\operatorname{\mathrm{H}}_{*}(M;{\mathbb{K}}),

then, it must have infinitely many periodic points.

Here, HFloc(ϕ,x)\operatorname{\mathrm{HF}}^{\operatorname{\mathrm{loc}}}(\phi,x) is a local version of Floer homology for a contractible fixed point xx of ϕ\phi. When xx is a non-degenerate fixed point, we have HFloc(ϕ,x)𝕂\operatorname{\mathrm{HF}}^{\operatorname{\mathrm{loc}}}(\phi,x)\cong{\mathbb{K}}, recovering the non-degenerate interpretation of Conjecture 1.1. In essence, one can think of N(ϕ,𝕂)N(\phi,{\mathbb{K}}) as a weighted sum of the number of contractible fixed points of ϕ\phi. In a striking result [She22, Theorem A], Shelukhin proved Conjecture 1.2 for spherically monotone symplectic manifolds with semisimple even quantum homology; a class of manifolds that includes complex projective spaces, complex Grassmannians, and their products. Furthermore, if 𝕂{\mathbb{K}} has characteristic 0, then there is a simple pp-periodic orbit for each sufficiently large prime pp. This article aims to generalize Shelukhin’s theorem to the semipositive setting. We obtain the following result:

Theorem 1.3.

Let (M,ω)(M,\omega) be a closed semipositive symplectic manifold with semisimple even quantum homology QHev(M;Λ𝕂,univ)\operatorname{\mathrm{QH}}_{ev}(M;\Lambda_{{\mathbb{K}},univ}) for a ground field 𝕂{\mathbb{K}}. Then, any Hamiltonian diffeomorphism ϕ\phi with finitely many contractible fixed points such that

N(ϕ,𝕂)=xFix(ϕ)dim𝕂HFloc(ϕ,x)>dim𝕂H(M;𝕂)N(\phi,{\mathbb{K}})=\displaystyle\sum_{x\in\mathrm{Fix}(\phi)}\dim_{{\mathbb{K}}}\operatorname{\mathrm{HF}}^{\operatorname{\mathrm{loc}}}(\phi,x)>\dim_{{\mathbb{K}}}\operatorname{\mathrm{H}}_{*}(M;{\mathbb{K}})

must have infinitely many periodic points. If 𝕂{\mathbb{K}} has characteristic zero, then ϕ\phi has a simple333A fixed point of ϕk\phi^{k} is said to be simple if it is not a fixed point of ϕl\phi^{l} for any proper divisor ll of kk. contractible pp-periodic point for each sufficiently large prime pp.

Remark 1.4.

There are known examples of closed semipositive symplectic manifolds that are not monotone and that have semisimple even quantum homology. Indeed, following [Ost06], (S2×S2,ωλω)(\mathrm{S}^{2}\times\mathrm{S}^{2},\omega\oplus\lambda\omega), for λ>1\lambda>1, and the symplectic one point blow-up (P2#P2¯,ωμ)({\mathbb{C}}P^{2}\#\overline{{\mathbb{C}}P^{2}},\omega_{\mu}) of P2{\mathbb{C}}P^{2}, where ωμ\omega_{\mu} integrates to μ(0,1){1/3}\mu\in(0,1)\setminus\{1/3\} on the exceptional divisor and to 11 on [P1][{\mathbb{C}}P^{1}], are non-monotone, semipositve, and have semisimple even quantum homology. Furthermore, any toric Fano manifold has semisimple even quantum homology for a generic toric symplectic form [Iri07, FOOO10, OT09], and if the even quantum homology of a toric Fano manifold XX is semisimple for the distinguished monotone symplectic form, then it is also semisimple for any toric symplectic form on XX. If in addition XX is at most six-dimensional, it is also semipositive and, therefore, within the scope of Theorem 1.3. Recent work of Bai and Xu [BX23] confirms Conjecture 1.2 for all compact toric symplectic manifolds. There are non-toric, non-monotone symplectic manifolds that satisfy the conditions of Theorem 1.3. For example, (Q3×P2,ωFSλωFS)(Q_{3}\times{\mathbb{C}}P^{2},\omega_{FS}\oplus\lambda\cdot\omega_{FS}), where λ>0\lambda>0 and Q3Q_{3} is the quadric in P4{\mathbb{C}}P^{4}. Indeed, it follows from the arguments in the proof of [McD11, Proposition 1.8], that if Q3×P2Q_{3}\times{\mathbb{C}}P^{2} were symplectomorphic to a toric manifold, then it would have the toric structure of P3×P2{\mathbb{C}}P^{3}\times{\mathbb{C}}P^{2}, which contradicts the fact that their Chern numbers are distinct. Moreover a kk-point blow-up of size ϵ\epsilon of P2{\mathbb{C}}P^{2}, is non-toric if 1/ϵ1/\epsilon is an integer and k4k\geq 4; see [Bay04, Ush11b, KK07].

The Hofer-Zehnder conjecture is related to a conjecture by Conley [CZ83, CZ84], which postulates that for a broad class of symplectic manifolds, any Hamiltonian diffeomorphism must have infinitely many periodic points. Therefore, the Hofer-Zehnder conjecture is automatically satisfied when the Conley conjecture holds. Following plenty of previous works that confirmed the conjecture in several cases [SZ92, CZ84, CZ86, FH03, LC06, Hin09, Gin10, GG09, Hei12, GG12], Ginzburg and Gürel [GG19] proved the most general statement known to hold. They showed that the existence of a Hamiltonian diffeomorphism with finitely many periodic points implies there is a spherical homology class AA with [ω],A>0,c1(M),A>0{\left<[\omega],A\right>}>0,{\left<c_{1}(M),A\right>}>0, in particular, the Conley conjecture holds for closed symplectically aspherical, negative-monotone and Calabi-Yau symplectic manifolds. However, the simple example of irrational rotation of the two-sphere shows that the Conley conjecture does not hold in general, and those are the cases where the Hofer-Zehnder conjecture is interesting. Furthermore, the Conley conjecture also fails for Pn{\mathbb{C}}\mathrm{P}^{n} and complex Grassmannians, all of which have semisimple quantum homology [EP03].

Lastly, we point to other possible interpretations of the meaning of a fixed point not being “necessarily required by the V. Arnold conjecture” in Conjecture 1.1. More precisely, the presence of a hyperbolic fixed point [GG14, GG18], and the existence of a non-contractible fixed point [Gür13, GG16, Ori17, Ori19], are considered “unnecessary” from this viewpoint and were shown to imply the existence of infinitely many periodic points for a large class of symplectic manifolds.

1.2. Setup

We recall some notions about Hamiltonian diffeomorphisms and symplectic topology required to follow an overview of the main result.

Definition 1.5.

A closed symplectic manifold (M,ω)(M,\omega) is called semipositive if for every sphere class AA in the image H2S(M;)\operatorname{\mathrm{H}}^{S}_{2}(M;{\mathbb{Z}}) of the Hurewicz map π2(M)H2(M;)\pi_{2}(M)\rightarrow\operatorname{\mathrm{H}}_{2}(M;{\mathbb{Z}})

c1(M),A3n,[ω],A>0c1(M),A0.{\left<c_{1}(M),A\right>}\geq 3-n,\quad{\left<[\omega],A\right>}>0\quad\Rightarrow\quad{\left<c_{1}(M),A\right>}\geq 0.

Where c1(M)H2(M;)c_{1}(M)\in\operatorname{\mathrm{H}}^{2}(M;{\mathbb{Z}}) denotes the first Chern class444An almost complex structure JJ on MM is an automorphism of the tangent bundle TMTM satisfying J2=idTMJ^{2}=-\operatorname{\mathrm{id}}_{TM}. It is called ω\omega-compatible if ω(,J)\omega(-,J-) is a Riemannian metric. The space 𝒥(M,ω)\mathcal{J}(M,\omega) of ω\omega-compatible almost complex structures is contractible, therefore, the first Chern class c1(M)=c1(M,ω)c_{1}(M)=c_{1}(M,\omega) of (TM,J)(TM,J) is independent of J𝒥(M,ω)J\in\mathcal{J}(M,\omega). associated with the symplectic manifold.

From this point forward, (M,ω)(M,\omega) denotes a closed semipositive symplectic manifold of dimension 2n2n. Let ϕ=ϕH\phi=\phi_{H} be a Hamiltonian diffeomorphism generated by a Hamiltonian function555We usually consider normalized Hamiltonians HH, i.e. H(t,)H(t,-) has zero ωn\omega^{n} mean for all t[0,1]t\in[0,1] HC(/×M,)H\in C^{\infty}({\mathbb{R}}/{\mathbb{Z}}\times M,{\mathbb{R}}). We say that a fixed point xx of ϕ\phi is contractible when the loop {ϕHt(x)}t[0,1]\{\phi_{H}^{t}(x)\}_{t\in[0,1]} is contractible in MM. It is a deep fact of symplectic topology that the class this loop represents in π1(M)\pi_{1}(M) is independent of the choice of HH generating ϕ\phi. We denote by Fix(ϕ)\mathrm{Fix}(\phi) the collection of contractible fixed points. Observe that Fix(ϕ)\mathrm{Fix}(\phi) includes naturally in Fix(ϕk)\mathrm{Fix}(\phi^{k}) for all k1k\geq 1. For a fixed point xx, we denote by x(k)x^{(k)} its image under this inclusion. We call a fixed point xx of ϕk\phi^{k} simple if it is not a fixed point of ϕl\phi^{l} for any proper divisor ll of kk.

Definition 1.6.

A fixed point xx of ϕ\phi is called non-degenerate if 11 is not an eigenvalue of the linearized time-one map dϕxd\phi_{x}. A Hamiltonian function HH is called non-degenerate if all contractible fixed points of ϕH\phi_{H} are non-degenerate.

Whenever a Hamiltonian diffeomorphism ϕHam(M,ω)\phi\in\operatorname{\mathrm{Ham}}(M,\omega) has finitely many fixed points, one can associate to it a filtered Floer homology theory [Flo86, Flo87, Flo89] with coefficients over a base field 𝕂{\mathbb{K}}, whose information is partially captured by a finite set of positive real numbers

β1(ϕ,𝕂)βK(ϕ,𝕂)(ϕ,𝕂),\beta_{1}(\phi,{\mathbb{K}})\leq\dots\leq\beta_{K(\phi,{\mathbb{K}})}(\phi,{\mathbb{K}}),

depending only on ϕ\phi and 𝕂{\mathbb{K}}, called the bar-length spectrum of ϕ\phi. We describe these notions, which were first introduced in symplectic topology by Polterovich and Shelukhin [PS16] (see also [PSS17, UZ16]) in Section 2.6. There are K(ϕ,𝕂)K(\phi,{\mathbb{K}}) elements, counted with multiplicity, in the bar-length spectrum of ϕ\phi. The length β(ϕ,𝕂)=βK(ϕ,𝕂)(ϕ,𝕂)\beta(\phi,{\mathbb{K}})=\beta_{K(\phi,{\mathbb{K}})}(\phi,{\mathbb{K}}) of the largest bar is called the boundary depth and was introduced by Usher [Ush11a, Ush13]. By definition, the boundary depth β(ϕ,𝕂)\beta(\phi,{\mathbb{K}}) is zero when K(ϕ,𝕂)K(\phi,{\mathbb{K}}) is zero. We denote by

βtot(ϕ,𝕂)=β1(ϕ,𝕂)++βK(ϕ,𝕂)(ϕ,𝕂)\beta_{tot}(\phi,{\mathbb{K}})=\beta_{1}(\phi,{\mathbb{K}})+\dots+\beta_{K(\phi,{\mathbb{K}})}(\phi,{\mathbb{K}})

the total bar-length. Finally, Shelukhin [She22] showed that

(1.1) N(ϕ,𝕂)=dim𝕂H(M;𝕂)+2K(ϕ,𝕂),N(\phi,{\mathbb{K}})=\dim_{{\mathbb{K}}}\operatorname{\mathrm{H}}_{*}(M;{\mathbb{K}})+2K(\phi,{\mathbb{K}}),

where N(ϕ,𝕂)N(\phi,{\mathbb{K}}) is as in the statement of Theorem 1.3. This equality also holds in the semipositive setting, we summarize the details of the underlying theorem in Section 2.7. In particular, the condition N(ϕ,𝕂)>dim𝕂H(M;𝕂)N(\phi,{\mathbb{K}})>\dim_{{\mathbb{K}}}\operatorname{\mathrm{H}}_{*}(M;{\mathbb{K}}) in the statement of the theorem implies that β(ϕ,𝕂)\beta(\phi,{\mathbb{K}}) is positive.

Remark 1.7.

We note that for sufficiently large primes pp we have by [She22, Lemma 16] the following equalities N(ϕ,)=N(ϕ,𝔽p),dimH(M;)=dim𝔽pH(M;𝔽p)N(\phi,{\mathbb{Q}})=N(\phi,{\mathbb{F}}_{p}),\dim_{{\mathbb{Q}}}\operatorname{\mathrm{H}}_{*}(M;{\mathbb{Q}})=\dim_{{\mathbb{F}}_{p}}\operatorname{\mathrm{H}}(M;{\mathbb{F}}_{p}), and β(ϕ,)=β(ϕ,𝔽p)\beta(\phi,{\mathbb{Q}})=\beta(\phi,{\mathbb{F}}_{p}) for any Hamiltonian diffeomorphism ϕ\phi. If 𝕂{\mathbb{K}} has characteristic 0, it is a field extension of {\mathbb{Q}}, which by [She22, Section 4.4.4] implies that N(ϕ,)=N(ϕ,𝕂),dimH(M;)=dim𝕂(M;𝕂)N(\phi,{\mathbb{Q}})=N(\phi,{\mathbb{K}}),\dim_{{\mathbb{Q}}}\operatorname{\mathrm{H}}_{*}(M;{\mathbb{Q}})=\dim_{{\mathbb{K}}}(M;{\mathbb{K}}), and β(ϕ,)=β(ϕ,𝕂)\beta(\phi,{\mathbb{Q}})=\beta(\phi,{\mathbb{K}}).

1.3. Overview

We summarize the proof of Theorem 1.3 while pointing to the technical developments required to obtain the result in the semipositive setting.

There are two main components to the proof. One is a generalization to the semipositive setting, stated as Theorem 5.1, of the Smith type inequality

(1.2) pβtot(ϕ,𝔽p)βtot(ϕp,𝔽p),p\cdot\beta_{tot}(\phi,{\mathbb{F}}_{p})\leq\beta_{tot}(\phi^{p},{\mathbb{F}}_{p}),

shown to hold by Shelukhin for spherically monotone symplectic manifolds under the assumption that ϕp\phi^{p} has finitely many fixed points. The main technical difficulty can be overcome using the recent work of Sugimoto [Sug21], generalizing the /(p){\mathbb{Z}}/(p)-equivariant product-isomorphism to the semipositive setting. The Smith inequality is then obtained by following Shelukhin’s proof [She22] in the spherically monotone setting. The equivariant Floer theory required to obtain Inequality (1.2) is detailed in Section 5.

The other component of the proof, proven as Theorem 4.11, is the main technical achievement of this paper. We uniformly bound the boundary-depth

(1.3) β(ϕk,𝔽p)C\beta(\phi^{k},{\mathbb{F}}_{p})\leq C

for sufficiently large iterations of ϕ\phi. In the monotone case, Shelukhin [She22] uses the relationship between indices and actions of contractible fixed points to obtain an upper bound for the boundary-depth that only depends on the dimension of the manifold. In the semipositive case, there is no such uniform relation between the actions and indices, therefore we carefully choose the coefficients of the relevant filtered Floer homology groups to have a good relation, before and after reducing coefficients, between the idempotents generating the even quantum homology of MM. We are then able to bound the boundary-depth by a constant, independent of pp, depending on the idempotents of the even quantum homology whose coefficient has characteristic 0.

With the Smith-type inequality and the uniform bound on the boundary-depth established, we now summarize Shelukhin’s argument to prove Theorem 1.3. First, we consider the case where 𝕂{\mathbb{K}} has characteristic 0 and the Hamiltonian diffeomorphism ϕ\phi and all of its iterates are non-degenerate. By Inequality (1.2) and the simple observation that βtot(ψ,𝕂)K(ψ,𝕂)β(ψ,𝕂)\beta_{tot}(\psi,{\mathbb{K}})\leq K(\psi,{\mathbb{K}})\cdot\beta(\psi,{\mathbb{K}}) for any Hamiltonian diffeomorphism ψ\psi and base field 𝕂{\mathbb{K}}, we have

pβtot(ϕ,𝔽p)βtot(ϕp,𝔽p)K(ϕp,𝔽p)β(ϕp,𝔽p).p\cdot\beta_{\mathrm{tot}}(\phi,{\mathbb{F}}_{p})\leq\beta_{\mathrm{tot}}(\phi^{p},{\mathbb{F}}_{p})\leq K(\phi^{p},{\mathbb{F}}_{p})\cdot\beta(\phi^{p},{\mathbb{F}}_{p}).

The assumption that N(ϕ,𝕂)>dim𝕂H(M;𝕂)N(\phi,{\mathbb{K}})>\dim_{{\mathbb{K}}}H_{*}(M;{\mathbb{K}}) and Remark 1.7 imply that the total bar-length βtot(ϕ,𝔽p)\beta_{\mathrm{tot}}(\phi,{\mathbb{F}}_{p}) is positive for a sufficiently large prime pp. Furthermore, Inequality (1.3) yields

pβtot(ϕ,𝔽p)CK(ϕp,𝔽p),p\cdot\beta_{\mathrm{tot}}(\phi,{\mathbb{F}}_{p})\leq C\cdot K(\phi^{p},{\mathbb{F}}_{p}),

which means that K(ϕp,𝕂)K(\phi^{p},{\mathbb{K}}) grows at least linearly with respect to pp. We now observe that in the non-degenerate setting Equation (1.1) yields

Fix(ϕp)=dim𝕂H(M;𝕂)+2K(ϕp,𝕂),\mathrm{Fix}(\phi^{p})=\dim_{{\mathbb{K}}}\operatorname{\mathrm{H}}_{*}(M;{\mathbb{K}})+2K(\phi^{p},{\mathbb{K}}),

which implies that ϕ\phi must have infinitely many contractible periodic points. In general, when the Hamiltonian diffeomorphism ϕ\phi is possibly degenerate, we can again achieve Inequality (1.3) by the local equivariant Floer homology argument [She22, Section 7.4]. Furthermore, by the canonical complex whose properties are listed in Theorem 2.20, the upper bound for the boundary-depth, which is also independent of pp, continues to hold. Therefore, we can use the same argument as in the non-degenerate case to obtain the linear growth of K(ϕp,𝕂)K(\phi^{p},{\mathbb{K}}) and, thus, of N(ϕp,𝕂)N(\phi^{p},{\mathbb{K}}). To conclude the argument, we assume that pp is large enough to guarantee ϕp\phi^{p} is an admissible iteration in the sense of Definition 2.17, it then follows by Theorem 2.18, [GG10, Theorem 1.1],[She22, Theorem C], that HFloc(ϕp1,x)HFloc(ϕp2,x)\operatorname{\mathrm{HF}}^{\operatorname{\mathrm{loc}}}(\phi^{p_{1}},x)\cong\operatorname{\mathrm{HF}}^{\operatorname{\mathrm{loc}}}(\phi^{p_{2}},x) for all xFix(ϕp1)x\in\mathrm{Fix}(\phi^{p_{1}}) for any two primes p2p1pp_{2}\geq p_{1}\geq p. In particular, there must be a new simple pp^{\prime}-periodic point for each prime p>pp^{\prime}>p. In fact, if Fix(ϕp1)=Fix(ϕp2)\mathrm{Fix}(\phi^{p_{1}})=\mathrm{Fix}(\phi^{p_{2}}) for p2p1pp_{2}\geq p_{1}\geq p, then N(ϕp1,𝕂)=N(ϕp2,𝕂)N(\phi^{p_{1}},{\mathbb{K}})=N(\phi^{p_{2}},{\mathbb{K}}) contradicting the linear growth of N(ϕp,𝕂)N(\phi^{p^{\prime}},{\mathbb{K}}) for ppp^{\prime}\geq p. A similar argument works when 𝕂{\mathbb{K}} has characteristic pp the details of which can be found in [She22, Section 8].

Acknowledgements

We thank Egor Shelukhin and Michael Usher for bringing us together to work on this project, for their support, and for the numerous helpful discussions. H.L. thanks Shengzhen Ning for the example of the four-point blowup of P2{\mathbb{C}}P^{2} and for pointing out the work [LMN22] which was helpful to understand toric structures under blowups. This work is part of both authors’ Ph.D. theses. H.L.’s Ph.D is taking place at the University of Georgia under the supervision of Michael Usher and M.S.A’s Ph.D is being carried out at the Université de Montréal under the supervision of Egor Shelukhin. M.S.A was partially supported by Fondation Courtois, by the ISM’s excellence scholarship, and by the J.Armand Bombardier’s excellence scholarhip. This material is based upon work supported by the National Science Foundation under Grant No. DMS-1928930, while M.S.A was in residence at the Simons Laufer Mathematical Sciences Institute (previously known as MSRI) Berkeley, California during the Fall 2022 semester.

2. Preliminaries

2.1. Basic setup

2.1.1. Fixed points of Hamiltonian diffeomorphisms

Let (M,ω)(M,\omega) be a closed symplectic manifold of dimension 2n2n. Denote by \mathcal{H} the space of 11-periodic Hamiltonian functions HC(/×M,)H\in C^{\infty}({\mathbb{R}}/{\mathbb{Z}}\times M,{\mathbb{R}}) normalized so that H(t,)H(t,-) has zero ωn\omega^{n} mean for all t[0,1]t\in[0,1]. We denote by XHX_{H} the time-dependent vector-field induced by Hamilton’s equation

ιXHtω=dHt,\iota_{X_{H}^{t}}\omega=-dH_{t},

and by {ϕHt}t[0,1]\{\phi_{H}^{t}\}_{t\in[0,1]} the induced symplectic isotopy i.e. ϕHt\phi_{H}^{t} satisfies

dϕHtdt=XHtϕHt\frac{d\phi^{t}_{H}}{dt}=X_{H}^{t}\circ\phi_{H}^{t}

with initial condition ϕH0=id\phi_{H}^{0}=\operatorname{\mathrm{id}}. We denote by ϕH=ϕH1\phi_{H}=\phi_{H}^{1} its time-one map. Diffeomorphisms obtained in this manner are called Hamiltonian diffeomorphisms. There is a bijective correspondence between the 11-periodic orbits of XHtX_{H}^{t} and the fixed points Fix(ϕH)\mathrm{Fix}(\phi_{H}) of ϕH\phi_{H}, therefore, periodic points of ϕH\phi_{H} correspond to

Per(ϕH)=k1Fix(ϕHk).{\mathrm{Per}}(\phi_{H})=\bigcup_{k\geq 1}\mathrm{Fix}(\phi_{H}^{k}).

For F,GF,G\in\mathcal{H}, we denote by F#GF\#G the Hamiltonian function

F#G(t,x)=F(t,x)+G(t,(ϕFt)1(x))F\#G(t,x)=F(t,x)+G(t,(\phi_{F}^{t})^{-1}(x))

that induces the isotopy {ϕFtϕGt}\{\phi_{F}^{t}\phi_{G}^{t}\}, in particular H##HH\#\cdots\#H (kk-times) generates {(ϕHt)k}\{(\phi_{H}^{t})^{k}\}. Note that H(k)(t,x)=kH(kt,x)H^{(k)}(t,x)=kH(kt,x) generates a homotopic path (rel. ends), therefore ϕH(k)=ϕHk\phi_{H^{(k)}}=\phi_{H}^{k}. We denote by H¯\overline{H} the Hamiltonian function H¯(t,x)=H(t,ϕHt(x))\overline{H}(t,x)=-H(t,\phi_{H}^{t}(x)), which generates {(ϕHt)1}\{(\phi_{H}^{t})^{-1}\}.

2.1.2. The Hamiltonian action funcitonal

Let M\mathcal{L}M denote the component of contractible loops of the free loop-space of MM. For a pair (x,x¯)(x,\overline{x}) consisting of a loop xMx\in\mathcal{L}M and a smooth capping x¯:𝔻M\overline{x}:{\mathbb{D}}\rightarrow M, with x¯|𝔻=x\overline{x}|_{\partial{\mathbb{D}}}=x, consider the following equivalence relation: (x,x¯)(y,y¯)(x,\overline{x})\sim(y,\overline{y}) if and only if,

x=y,and[x¯#(y¯)]ker[ω]kerc1.\displaystyle x=y,\quad\text{and}\quad[\overline{x}\#(-\overline{y})]\in\ker[\omega]\cap\ker c_{1}.

Here, x¯#(y¯)\overline{x}\#(-\overline{y}) stands for gluing the disks along their boundaries with the orientation of y¯\overline{y} reversed, and [x¯#(y¯)][\overline{x}\#(-\overline{y})] its class in H2S(M;)\operatorname{\mathrm{H}}^{S}_{2}(M;{\mathbb{Z}}). Denote by ~M\widetilde{\mathcal{L}}M the covering space of M\mathcal{L}M given by the collection of such pairs modulo the equivalence relation \sim. For the sake of brevity, we shall write x¯\overline{x} instead of (x,x¯)(x,\overline{x}). Note that, the group of deck transformations of ~M\widetilde{\mathcal{L}}M is isomorphic to

Gω=π2(M)/ker[ω]kerc1,G_{\omega}=\pi_{2}(M)/\ker[\omega]\cap\ker c_{1},

where the transformation associated with Aπ2(M)A\in\pi_{2}(M) is given by sending x¯\overline{x} to x¯#A\overline{x}\#A. To a Hamiltonian function HH, we associate an action functional 𝒜H:~M:\mathcal{A}_{H}:\widetilde{\mathcal{L}}M:\rightarrow{\mathbb{R}} defined by

𝒜H(x¯)=01H(t,x(t))𝑑tx¯ω.\mathcal{A}_{H}(\overline{x})=\int^{1}_{0}H(t,x(t))dt-\int_{\overline{x}}\omega.

The critical points of 𝒜H\mathcal{A}_{H} are the lifts 𝒫~(H)\widetilde{\mathcal{P}}(H) of the contractible 11-periodic orbits 𝒫(H)\mathcal{P}(H) satisfying the equation x(t)=XHt(x(t))x^{\prime}(t)=X_{H}^{t}(x(t)). The action spectrum of HH is defined as the subset of {\mathbb{R}} given by the critical values Spec(H)=𝒜H(𝒫~(H))\operatorname{\mathrm{Spec}}(H)=\mathcal{A}_{H}(\widetilde{\mathcal{P}}(H)) of the action functional. If AGωA\in G_{\omega} then,

𝒜H(x¯#A)=𝒜H(x¯)[ω],A.\mathcal{A}_{H}(\overline{x}\#A)=\mathcal{A}_{H}(\overline{x})-{\left<[\omega],A\right>}.

Furthermore, the action functional behaves well with iterations in the sense that

𝒜H(k)(x¯(k))=k𝒜H(x¯),\mathcal{A}_{H^{(k)}}(\overline{x}^{(k)})=k\mathcal{A}_{H}(\overline{x}),

where x¯(k)\overline{x}^{(k)} inherits the natural capping induced by x¯\overline{x}.

2.2. Filtered Hamiltonian Floer homology

Floer theory was first developed by A. Floer in [Flo86, Flo87, Flo89] as a generalization of Morse-Novikov homology, to prove the non-degenerate Arnold conjecture. We refer to [HS95, MS17, Oh15] and references therein for the details of the construction and to [Abo15, Sei02, Zap] for in-depth discussions of canonical orientations. We shall consider the construction of Hamiltonian Floer homology in the semipositive setting.

Let (M,ω)(M,\omega) be a closed symplectic manifold and 𝕂{\mathbb{K}} a choice of base field. Let HH be a non-degenerate Hamiltonian function on MM and J={Jt}J=\{J_{t}\} be a 11-periodic family of ω\omega-compatible almost complex structures. For aSpec(H)a\in{\mathbb{R}}\setminus\operatorname{\mathrm{Spec}}(H) the Floer chain complex at filtration level aa is defined by

CF(H;J)<a={aix¯i|ai𝕂,x¯i𝒫~(H),𝒜H(x¯i)<a},\operatorname{\mathrm{CF}}_{*}(H;J)^{<a}=\Big{\{}\sum a_{i}\overline{x}_{i}\,\Big{|}\,a_{i}\in{\mathbb{K}},\,\overline{x}_{i}\in\widetilde{\mathcal{P}}(H),\,\mathcal{A}_{H}(\overline{x}_{i})<a\Big{\}},

where every summation satisfies the condition that the set {i|ai0,𝒜H(x¯i)>c}\{i\,|\,a_{i}\neq 0,\mathcal{A}_{H}(\overline{x}_{i})>c\} is finite for all cc\in{\mathbb{R}}. The complex is graded by the Conley-Zehnder index, which assigns an integer CZ(x¯)\operatorname{\mathrm{CZ}}(\overline{x}) to each x¯𝒫~(H)\overline{x}\in\widetilde{\mathcal{P}}(H) of 𝒜H\mathcal{A}_{H} and satisfies CZ(x¯#A)=CZ(x¯)c1(M),A\operatorname{\mathrm{CZ}}(\overline{x}\#A)=\operatorname{\mathrm{CZ}}(\overline{x})-{\left<c_{1}(M),A\right>}. Note that CF(H;J)=CF(H;J)+\operatorname{\mathrm{CF}}_{*}(H;J)=\operatorname{\mathrm{CF}}_{*}(H;J)^{+\infty} is naturally a finitely generated module over the Novikov ring

Λω,𝕂={AGωaATω(A)|aA𝕂,#{A|aA0,ω(A)<c}<,c}.\Lambda_{\omega,{\mathbb{K}}}=\Bigg{\{}\sum_{A\in G_{\omega}}a_{A}T^{\omega(A)}\,\Big{|}\,a_{A}\in{\mathbb{K}},\,\#\{A\,|\,a_{A}\neq 0,\omega(A)<c\}<\infty,\,\forall c\in{\mathbb{R}}\Bigg{\}}.

A Floer trajectory between capped orbits x¯,x¯+\overline{x}_{-},\overline{x}_{+}, is a smooth map u:×S1Mu:{\mathbb{R}}\times\mathrm{S}^{1}\rightarrow M satisfying the Floer equation

us+Jt(u)(utXHt(u))=0,\frac{\partial u}{\partial s}+J_{t}(u)\Bigg{(}\frac{\partial u}{\partial t}-X_{H}^{t}(u)\Bigg{)}=0,

with asymptotics,

lims±u(s,t)=x±(t),\lim_{s\rightarrow\pm\infty}u(s,t)=x_{\pm}(t),

such that the capping u#x¯+u\#\overline{x}_{+} is equivalent to x¯\overline{x}_{-}, and CZ(x¯)CZ(x¯+)=1\operatorname{\mathrm{CZ}}(\overline{x}_{-})-\operatorname{\mathrm{CZ}}(\overline{x}_{+})=1. In the semipositive setting, the compactified moduli space (x¯,y¯;J)\mathcal{M}(\overline{x},\overline{y};J) of Floer-trajectories from x¯\overline{x} to y¯\overline{y} (modulo the natural {\mathbb{R}}-action) is a manifold of dimension CZ(x¯)CZ(y¯)1\operatorname{\mathrm{CZ}}(\overline{x})-\operatorname{\mathrm{CZ}}(\overline{y})-1. The Floer differential

dF:CFk(H;J)<aCFk1(H;J)<ad_{F}:\operatorname{\mathrm{CF}}_{k}(H;J)^{<a}\rightarrow\operatorname{\mathrm{CF}}_{k-1}(H;J)^{<a}

is defined by

dF(x¯)=CZ(x¯)CZ(y¯)=1(#(x¯,y¯;J))y¯.d_{F}(\overline{x})=\sum_{\operatorname{\mathrm{CZ}}(\overline{x})-\operatorname{\mathrm{CZ}}(\overline{y})=1}\big{(}\#\mathcal{M}(\overline{x},\overline{y};J)\big{)}\cdot\overline{y}.

It squares to zero and preserves the filtration induced by 𝒜H\mathcal{A}_{H}. For an interval I=(a,b)I=(a,b), a<ba<b, where a,bSpec(H)a,b\in{\mathbb{R}}\setminus\operatorname{\mathrm{Spec}}(H) we define the Floer complex in the action window II as the quotient complex

CF(H;J)I=CF(H;J)<b/CF(H;J)<a.\operatorname{\mathrm{CF}}_{*}(H;J)^{I}=\operatorname{\mathrm{CF}}_{*}(H;J)^{<b}/\operatorname{\mathrm{CF}}_{*}(H;J)^{<a}.

We denote by HF(H)I\operatorname{\mathrm{HF}}_{*}(H)^{I} the resulting homology of this complex with the differential induced by dFd_{F} and call it the Floer homology of HH in the action window II. Note that it is independent of the generic choice of almost complex structure JJ. The (total) Floer homology HF(H)\operatorname{\mathrm{HF}}_{*}(H) of HH is obtained by setting a=a=-\infty and b=+b=+\infty and does not depend on the choice of Hamiltonian (by a standard continuation argument). Furthermore, for all HH\in\mathcal{H}, HF(H)I\operatorname{\mathrm{HF}}_{*}(H)^{I} depends only on the homotopy class of {ϕHt}t[0,1]\{\phi_{H}^{t}\}_{t\in[0,1]} in the universal cover Ham~(M,ω)\widetilde{\operatorname{\mathrm{Ham}}}(M,\omega) of the group of Hamiltonian diffeomorphisms Ham(M,ω)\operatorname{\mathrm{Ham}}(M,\omega).

When HH is degenerate, we have to consider perturbation data 𝒟=(KH,JH)\mathcal{D}=(K^{H},J^{H}), where KK\in\mathcal{H} is such that H𝒟=H#KHH^{\mathcal{D}}=H\#K^{H} is a non-degenerate Hamiltonian and JHJ^{H} is a choice of generic almost complex structure with respect to H𝒟H^{\mathcal{D}}. We take the action functional to be 𝒜H;𝒟=𝒜H𝒟\mathcal{A}_{H;\mathcal{D}}=\mathcal{A}_{H^{\mathcal{D}}}. When (M,ω)(M,\omega) is rational, an admissible action window I=(a,b)I=(a,b) for HH, i.e. a,bSpec(H)a,b\in{\mathbb{R}}\setminus\operatorname{\mathrm{Spec}}(H), will remain so for H𝒟H^{\mathcal{D}} for sufficiently C2C^{2}-small KHK^{H}, furthermore, the groups HF(H;𝒟)I\operatorname{\mathrm{HF}}(H;\mathcal{D})^{I} are canonically isomorphic. Therefore, HF(H)I\operatorname{\mathrm{HF}}(H)^{I} is defined as the colimit of the induced directed system. The general case is dealt with by taking the colimit over partially ordered non-degenerate perturbations whose action spectrums do not include aa or bb; refer to [Hei12] or [AS23, Section 2.2.2] for a detailed exposition.

2.3. Novikov ring and non-Archimedean valuation

2.3.1. Novikov ring and extending coefficients

Let RR be a commutative unital ring. The universal Novikov ring over RR is defined as

(2.1) ΛR,univ={i=KaiTλi|aiR,λi+}.\Lambda_{R,univ}=\bigg{\{}\sum_{i=-K}^{\infty}a_{i}T^{\lambda_{i}}\,|\,a_{i}\in R,\,\lambda_{i}\nearrow+\infty\bigg{\}}.

It follows from [HS95, Theorem 4.2] that ΛR,univ\Lambda_{{R},univ} is a principal ideal domain (resp. a field) whenever RR is a principal ideal domain (resp. a field). In particular, when RR is a principal ideal domain, every nonzero prime ideal in ΛR,univ\Lambda_{{R},univ} is maximal. We are particularly interested in the case R=R={\mathbb{Z}}, when RR is not a field, and the case R=𝔽p,R={\mathbb{F}}_{p},{\mathbb{Q}}, for pp prime, when RR is a field.

For a base field 𝕂{\mathbb{K}}, denote by CF(H;Λ𝕂,ω)\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},\omega}) the Floer chain complex defined in Section 2.2, where we omit the choice of generic almost complex structure JJ. It will often be convenient to extend the coefficients to the universal Novikov field Λ𝕂,univ\Lambda_{{\mathbb{K}},univ} and to its algebraic closure Λ¯𝕂,univ\overline{\Lambda}_{{\mathbb{K}},univ}. We define

CF(H;Λ𝕂,univ)=CF(H;Λ𝕂,ω)Λ𝕂,ωΛ𝕂,univ\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},univ})=\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},\omega})\otimes_{\Lambda_{{\mathbb{K}},\omega}}\Lambda_{{\mathbb{K}},univ}

and

CF(H;Λ¯𝕂,univ)=CF(H;Λ𝕂,univ)Λ𝕂,univΛ¯𝕂,univ,\operatorname{\mathrm{CF}}(H;\overline{\Lambda}_{{\mathbb{K}},univ})=\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},univ})\otimes_{\Lambda_{{\mathbb{K}},univ}}\overline{\Lambda}_{{\mathbb{K}},univ},

the differentials are extended by linearity. In particular, if {x1,,xB}\{x_{1},\cdots,x_{B}\} is a Λ𝕂,ω\Lambda_{{\mathbb{K}},\omega}-basis of CF(H;Λ𝕂,ω)\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},\omega}), then the elements of CF(H;Λ𝕂,univ)\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},univ}) (resp. CF(H;Λ¯𝕂,univ)\operatorname{\mathrm{CF}}(H;\overline{\Lambda}_{{\mathbb{K}},univ})) are of the form λixi\sum\lambda_{i}x_{i} where λiΛ𝕂,univ\lambda_{i}\in\Lambda_{{\mathbb{K}},univ} (resp. Λ¯𝕂,univ\overline{\Lambda}_{{\mathbb{K}},univ}).

2.3.2. Non-Archimedean valuation

Definition 2.1.

A non-Archimedean valuation on a field Λ\Lambda is a function

ν:Λ{+}\nu:\Lambda\rightarrow{\mathbb{R}}\cup\{+\infty\}

satisfying the following properties:

  1. (1)

    ν(x)=+\nu(x)=+\infty if and only if x=0x=0,

  2. (2)

    ν(xy)=ν(x)+ν(y)\nu(xy)=\nu(x)+\nu(y) for all x,yΛx,y\in\Lambda,

  3. (3)

    ν(x+y)min{ν(x),ν(y)}\nu(x+y)\geq\min\{\nu(x),\nu(y)\} for all x,yΛx,y\in\Lambda.

Furthermore, we set Λ0=ν1([0,+))\Lambda^{0}=\nu^{-1}([0,+\infty)) to be the subring of elements of non-negative valuation.

It shall often be the case that Λ=Λ𝕂,univ\Lambda=\Lambda_{{\mathbb{K}},univ}, where Λ𝕂,univ\Lambda_{{\mathbb{K}},univ} is the universal Novikov field over a ground field 𝕂{\mathbb{K}},

(2.2) Λ𝕂,univ={i=KaiTλi|ai𝕂,λi+}.\Lambda_{{\mathbb{K}},univ}=\bigg{\{}\sum_{i=-K}^{\infty}a_{i}T^{\lambda_{i}}\,|\,a_{i}\in{\mathbb{K}},\,\lambda_{i}\nearrow+\infty\bigg{\}}.

In this case, Λ\Lambda can be endowed with a non-Archimedean valuation given by setting ν(0)=+\nu(0)=+\infty and

(2.3) ν(i=KaiTλi)=λK\nu\bigg{(}\sum_{i=-K}^{\infty}a_{i}T^{\lambda_{i}}\bigg{)}=\lambda_{-K}

on Λ{0}\Lambda\setminus\{0\}. The universal Novikov ring over {\mathbb{Z}}, denoted Λ,univ\Lambda_{{\mathbb{Z}},univ}, is defined just as in Equation (2.2), however, it is not a field. Its field of fractions Q(Λ,univ)Q(\Lambda_{{\mathbb{Z}},univ}) can be identified with a subfield of Λ,univ\Lambda_{{\mathbb{Q}},univ}, therefore, its elements can be as expressed as

i=KciTλi,ci.\sum_{i=-K}^{\infty}c_{i}T^{\lambda_{i}},\quad c_{i}\in{\mathbb{Q}}.

We can hence define a valuation ν:Q(Λ,univ){+}\nu:Q(\Lambda_{{\mathbb{Z}},univ})\rightarrow{\mathbb{R}}\cup\{+\infty\} as in Equation (2.3).

2.3.3. Field norms and extension of valuations

The notion of valuation on a field Λ\Lambda is closely related to that of a field norm defined below.

Definition 2.2.

A non-Archimedean norm on a field Λ\Lambda is a map ||:Λ0|\cdot|:\Lambda\rightarrow{\mathbb{R}}_{\geq 0} satisfying the following properties:

  1. (1)

    |x|=0|x|=0 if and only if x=0x=0,

  2. (2)

    |xy|=|x||y||xy|=|x||y| for all x,yΛx,y\in\Lambda,

  3. (3)

    |x+y|max{|x|,|y|}|x+y|\leq\max\{|x|,|y|\} for all x,yΛx,y\in\Lambda.

Furthermore, Λ\Lambda is said to be complete with respect to |||\cdot| if it is a complete metric space with respect to the induced topology.

Note that given a non-Archimedean valuation ν\nu, one can define a non-Archimedean norm by setting |x|=eν(x)|x|=e^{-\nu(x)}. Conversely, if |||\cdot| is a non-Archimedean norm, a non-Archimedean valuation can be obtained by setting ν(x)=ln(|x|)\nu(x)=-\ln(|x|). The following results in [Cas86] allows one to extend a given valuation to certain field extensions.

Proposition 2.3 (Chapter 7, Theorem 1.1 in [Cas86]).

Let 𝕂{\mathbb{K}} be a field that is complete with respect to a norm |||\cdot| and let 𝕃{\mathbb{L}} be a finite extension of degree nn. Then, there is precisely one extension \parallel\cdot\parallel of |||\cdot| to 𝕃{\mathbb{L}}. It is given by

A=|N𝕃/𝕂(A)|1/n\parallel A\parallel=|N_{{\mathbb{L}}/{\mathbb{K}}}(A)|^{1/n}

where A𝕃A\in{\mathbb{L}} and N𝕃/𝕂(A)N_{{\mathbb{L}}/{\mathbb{K}}}(A) is the determinant of the map BABB\mapsto AB for B𝕃B\in{\mathbb{L}}. Furthermore, 𝕃{\mathbb{L}} is complete with respect to \parallel\cdot\parallel.

Proposition 2.4 (Chapter 9, Lemma 2.1 in [Cas86]).

Let 𝕃=𝕂(A){\mathbb{L}}={\mathbb{K}}(A) be a separable extension and let F(x)𝕂[x]F(x)\in{\mathbb{K}}[x] be the minimal polynomial for AA. Let 𝔎\mathfrak{K} be the completion of 𝕂{\mathbb{K}} with respect to a norm |||\cdot|. Let F(x)=ϕ1(x)ϕJ(x)F(x)=\phi_{1}(x)\cdots\phi_{J}(x) be the decomposition of F(x)F(x) into irreducibles in 𝔎[x]\mathfrak{K}[x]. Then the ϕj\phi_{j} are distinct. Let 𝕃j=𝔎(Bj){\mathbb{L}}_{j}=\mathfrak{K}(B_{j}) where BjB_{j} is a root of ϕj(x)\phi_{j}(x). Then there is an injection

(2.4) 𝕃𝕃j{\mathbb{L}}\hookrightarrow{\mathbb{L}}_{j}

extending 𝕂𝔎{\mathbb{K}}\hookrightarrow\mathfrak{K} under which ABjA\mapsto B_{j}. Denote by ||j|\cdot|_{j} the norm on 𝕃{\mathbb{L}} induced by equation 2.4 and the unique norm on 𝕃j{\mathbb{L}}_{j} extending |||\cdot|. Then the ||j|\cdot|_{j} (1jJ)(1\leq j\leq J) are precisely all the extensions of |||\cdot| from 𝕂{\mathbb{K}} to 𝕃{\mathbb{L}}. Furthermore, 𝕃j{\mathbb{L}}_{j} is the completion of 𝕃{\mathbb{L}} with respect to ||j|\cdot|_{j}.

Remark 2.5.

We apply these extension propositions in the following two situations. First, we describe how to extend the valuation ν\nu on Λ𝔽p,univ\Lambda_{{\mathbb{F}}_{p},univ} to a finite field extension Λ𝔽p,univ(γ)\Lambda_{{\mathbb{F}}_{p},univ}(\gamma) for an algebraic element γ\gamma over Λ𝔽p,univ\Lambda_{{\mathbb{F}}_{p},univ}. In this case, we note that Λ𝔽p,univ\Lambda_{{\mathbb{F}}_{p},univ} is complete with respect to the norm ||=eν()|\cdot|=e^{-\nu(\cdot)}, therefore, Proposition 2.3 gives us an extension of the norm and, hence, the valuation, to Λ𝔽p,univ(γ)\Lambda_{{\mathbb{F}}_{p},univ}(\gamma). The same argument applies to the extension of ν\nu to the algebraic closure Λ¯𝔽p,univ\overline{\Lambda}_{{\mathbb{F}}_{p},univ}. Similarly, we now explain how to extend the valuation ν\nu on Q(Λ,univ)Q(\Lambda_{{\mathbb{Z}},univ}) to a valuation on a field extension Q(Λ,univ)(α)Q(\Lambda_{{\mathbb{Z}},univ})(\alpha), where α\alpha is algebraic over Q(Λ,univ)Q(\Lambda_{{\mathbb{Z}},univ}). We note that Q(Λ,univ)Q(\Lambda_{\mathbb{Z},univ}) has characteristic zero making it a perfect field, which implies that Q(Λ,univ)(α)Q(\Lambda_{\mathbb{Z},univ})(\alpha) is a separable extension. Therefore, by Proposition 2.4 one obtains a norm |||\cdot| and, thus, a valuation ν\nu on Q(Λ,univ)(α)Q(\Lambda_{\mathbb{Z},univ})(\alpha).

Proposition 2.6 (Chapter 7, Corollary 1 in [Cas86]).

There is a unique extension of |||\cdot| to the algebraic closure 𝕂¯\overline{{\mathbb{K}}} of 𝕂{\mathbb{K}}

Remark 2.7.

Proposition 2.6 implies that the extensions of the valuations on Λ𝔽p,univ\Lambda_{{\mathbb{F}}_{p},univ} and Λ𝔽p,univ(γ)\Lambda_{{\mathbb{F}}_{p},univ}(\gamma) to a valuation on Λ¯𝔽p,univ\overline{\Lambda}_{{\mathbb{F}}_{p},univ} coincide.

2.3.4. Non-Archimedean filtrations

Let Λ\Lambda be a field with a non-Archimedean valuation ν\nu. Suppose CC is a finite dimensional module over Λ\Lambda.

Definition 2.8.

A non-Archimedean filtration is a function l:C{}l:C\rightarrow{\mathbb{R}}\cup\{-\infty\} satisfying the following properties:

  1. (1)

    l(x)=l(x)=-\infty if and only if x=0x=0,

  2. (2)

    l(λx)=l(x)ν(λ)l(\lambda x)=l(x)-\nu(\lambda) for all λΛ,xC\lambda\in\Lambda,x\in C,

  3. (3)

    l(x+y)max{l(x),l(y)}l(x+y)\leq\max\{l(x),l(y)\}

It is not hard to check that the maximum property (3) implies that when l(x)l(y)l(x)\neq l(y), we have that l(x+y)=max{l(x),l(y)}l(x+y)=\max\{l(x),l(y)\}, see [EP03, UZ16]. We call a Λ\Lambda-basis (x1,,xB)(x_{1},\cdots,x_{B}) of (C,l)(C,l) orthogonal if

l(λixi)=max{l(xi)ν(λi)}l\Big{(}\sum\lambda_{i}x_{i}\Big{)}=\max\{l(x_{i})-\nu(\lambda_{i})\}

for all λiΛ\lambda_{i}\in\Lambda. It is called orthonormal if it also satisfies l(xi)=0l(x_{i})=0 for all ii.

In order to define a non-Archimedean filtration 𝒜:CF(H;Λ𝕂,univ){}\mathcal{A}:\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},univ})\rightarrow{\mathbb{R}}\cup\{-\infty\}, we choose a Λ𝕂,ω\Lambda_{{\mathbb{K}},\omega}-basis {x1,,xN}\{x_{1},\cdots,x_{N}\} of CF(H;Λ𝕂,ω)\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},\omega}) and set

(2.5) 𝒜(λixi)=max{𝒜H(xi)ν(λi)}.\mathcal{A}\Big{(}\sum\lambda_{i}x_{i}\Big{)}=\max\{\mathcal{A}_{H}(x_{i})-\nu(\lambda_{i})\}.

Equivalently, we declare {x1,,xN}\{x_{1},\dots,x_{N}\} to be an orthogonal basis of (CF(H;Λ𝕂,univ),𝒜)(\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},univ}),\mathcal{A}). We note that for any non-trivial xCF(H;Λ𝕂,univ)x\in\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},univ}) we have 𝒜(dF(x))<𝒜(x)\mathcal{A}(d_{F}(x))<\mathcal{A}(x). In this case, the filtered complex is said to be strict. The basis given by {x~1,,x~N}={T𝒜(x1)x1,,T𝒜(xN)xN}\{\widetilde{x}_{1},\cdots,\widetilde{x}_{N}\}=\{T^{\mathcal{A}(x_{1})}x_{1},\cdots,T^{\mathcal{A}(x_{N})}x_{N}\} is orthonormal.

We now consider the case where we extend the coefficients of CF(H;Λ𝕂,univ)\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},univ}) to Λ¯𝕂,univ\overline{\Lambda}_{{\mathbb{K}},univ}. For an orthogonal basis {x1,,xN}\{x_{1},\dots,x_{N}\} of (CF(H;Λ𝕂,univ),𝒜)(\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},univ}),\mathcal{A}), {x11,,xN1}\{x_{1}\otimes 1,\dots,x_{N}\otimes 1\} is an orthogonal basis of CF(H;Λ¯𝕂,univ)\operatorname{\mathrm{CF}}(H;\overline{\Lambda}_{{\mathbb{K}},univ}). We define a non-Archimedean filtration 𝒜¯\overline{\mathcal{A}} on CF(H;Λ¯𝕂,univ)\operatorname{\mathrm{CF}}(H;\overline{\Lambda}_{{\mathbb{K}},univ}) by setting

𝒜¯(λ¯ixi1)=max{𝒜H(xi)ν¯(λ¯i)},\overline{\mathcal{A}}\Big{(}\sum\overline{\lambda}_{i}x_{i}\otimes 1\Big{)}=\max\{\mathcal{A}_{H}(x_{i})-\overline{\nu}(\overline{\lambda}_{i})\},

where ν¯\overline{\nu} is the non-Archimedean valuation on Λ¯𝕂,univ\overline{\Lambda}_{{\mathbb{K}},univ} described in Remark 2.5. One verifies that {T𝒜(x1)x11,,T𝒜(xN)xN1}\{T^{\mathcal{A}(x_{1})}x_{1}\otimes 1,\dots,T^{\mathcal{A}(x_{N})}x_{N}\otimes 1\} is an orthonormal basis of (CF(H;Λ¯𝕂,univ),𝒜¯)(\operatorname{\mathrm{CF}}(H;\overline{\Lambda}_{{\mathbb{K}},univ}),\overline{\mathcal{A}}). Any orthonormal basis {y1,,yB}\{y_{1},\dots,y_{B}\} of (CF(H;Λ𝕂,univ),𝒜)(\operatorname{\mathrm{CF}}(H;\Lambda_{{\mathbb{K}},univ}),\mathcal{A}) is related to the orthonormal basis {x¯1,,x¯B}\{\overline{x}_{1},\dots,\overline{x}_{B}\} by an invertible matrix AGL(B,Λ𝕂,univ0)A\in{\mathrm{GL}}(B,\Lambda^{0}_{{\mathbb{K}},univ}) in the sense that A(T𝒜(xj)xj)=yjA(T^{\mathcal{A}(x_{j})}x_{j})=y_{j}, furthermore,

{y11,,yN1}=A{x11,,xN1}\{y_{1}\otimes 1,\dots,y_{N}\otimes 1\}=A\{x^{\prime}_{1}\otimes 1,\dots,x^{\prime}_{N}\otimes 1\}

is an orthonormal basis of (CF(H;Λ¯𝕂,univ),𝒜¯)(\operatorname{\mathrm{CF}}(H;\overline{\Lambda}_{{\mathbb{K}},univ}),\overline{\mathcal{A}}).

2.4. Quantum homology

2.4.1. Definition of quantum homology

We follow [MS12] for the definition of quantum homology. The quantum homology of MM is defined by QH(M)=QH(M,Λ,univ)=H(M)Λ,univ\operatorname{\mathrm{QH}}_{*}(M)=\operatorname{\mathrm{QH}}_{*}(M,\Lambda_{{\mathbb{Z}},univ})=\operatorname{\mathrm{H}}_{*}(M)\otimes\Lambda_{\mathbb{Z},univ}. There is a product on

QHev(M)=iQH2i(M)\operatorname{\mathrm{QH}}_{ev}(M)=\bigoplus_{i}\operatorname{\mathrm{QH}}_{2i}(M)

defined as follows. Choose an integer basis e0,,eNe_{0},\dots,e_{N} of the free part of H(M;)\operatorname{\mathrm{H}}_{*}(M;\mathbb{Z}) such that e0=[M]H2n(M)e_{0}=[M]\in\operatorname{\mathrm{H}}_{2n}(M) and each basis element eνe_{\nu} has pure degree. Define the integer matrix gνμg_{\nu\mu} by

gνμ:=MPD(eν)PD(eμ),g_{\nu\mu}:=\displaystyle\int_{M}PD(e_{\nu})\smile PD(e_{\mu}),

and let gνμg^{\nu\mu} denote the inverse matrix. Then the product of a,bHev(M)a,b\in\operatorname{\mathrm{H}}_{ev}(M) is defined by

ab:=Aν,μGWA,3M(a,b,eν)gνμeμTω(A)a*b:=\displaystyle\sum_{A}\sum_{\nu,\mu}GW^{M}_{A,3}(a,b,e_{\nu})g^{\nu\mu}e_{\mu}T^{\omega(A)}.

The product of aa and bb can also be expressed as

ab=A(ab)ATω(A)a*b=\displaystyle\sum_{A}(a*b)_{A}T^{\omega(A)}

where

(ab)A:=ν,μGWA,3M(a,b,eν)gνμeμHdeg(a)+deg(b)+2c1(A)2n(M)(a*b)_{A}:=\displaystyle\sum_{\nu,\mu}GW^{M}_{A,3}(a,b,e_{\nu})g^{\nu\mu}e_{\mu}\in\operatorname{\mathrm{H}}_{deg(a)+deg(b)+2c_{1}(A)-2n}(M)

which is characterized by the condition

MPD((ab)A)c:=GWA,3M(a,b,PD(c))\displaystyle\int_{M}PD((a*b)_{A})\smile c:=GW^{M}_{A,3}(a,b,PD(c))

for cH(M)c\in\operatorname{\mathrm{H}}^{*}(M). For a base field, 𝕂{\mathbb{K}} we define QH(M,Λ𝕂,univ)=H(M)Λ𝕂,univ\operatorname{\mathrm{QH}}_{*}(M,\Lambda_{{\mathbb{K}},univ})=\operatorname{\mathrm{H}}_{*}(M)\otimes\Lambda_{\mathbb{{\mathbb{K}}},univ}, and the product is defined in the same way. Also, if 𝕃{\mathbb{L}} is a the field containing Λ,univ\Lambda_{{\mathbb{Z}},univ} or if it is a field extension of Λ𝕂,univ\Lambda_{{\mathbb{K}},univ}, we set QH(H,𝕃)=QH(M)𝕃\operatorname{\mathrm{QH}}(H,{\mathbb{L}})=\operatorname{\mathrm{QH}}(M)\otimes{\mathbb{L}} and extend the quantum product linearly.

2.4.2. Semisimplicity of quantum homology

We note that the quantum product is graded commutative, however, since we are considering only the even degrees, QHev(M,Λ𝕂,univ)\operatorname{\mathrm{QH}}_{ev}(M,\Lambda_{{\mathbb{K}},univ}) has the structure of a commutative algebra over Λ𝕂,univ\Lambda_{{\mathbb{K}},univ}. Hence, it is semiseimple if it splits as an algebra into a direct sum of fields F1FmF_{1}\oplus\cdots\oplus F_{m}. In particular FjF_{j} is a finite dimension vector space over Λ𝕂,univ\Lambda_{{\mathbb{K}},univ} for each jj.

Remark 2.9.

Other notions of semisimplicity have been considered in the non-monotone setting, for instance we can ask that QH2n(M,Λω,𝕂)\operatorname{\mathrm{QH}}_{2n}(M,\Lambda_{\omega,{\mathbb{K}}}) is a semisimple algebra over the field Λω,𝕂0\Lambda_{\omega,{\mathbb{K}}}^{0}, which is the degree 0 component of Λω,𝕂\Lambda_{\omega,{\mathbb{K}}}. We now show that this condition implies that QHev(M,Λ𝕂,univ)\operatorname{\mathrm{QH}}_{ev}(M,\Lambda_{{\mathbb{K}},univ}) is semisimple. Indeed, by [EP08, Proposition 2.1(A)] it follows that if QH2n(M,Λω,𝕂)\operatorname{\mathrm{QH}}_{2n}(M,\Lambda_{\omega,{\mathbb{K}}}) is semisimple over Λω,𝕂0\Lambda_{\omega,{\mathbb{K}}}^{0}, then QH2n(M,Λω,𝕂)Λ𝕂,univ\operatorname{\mathrm{QH}}_{2n}(M,\Lambda_{\omega,{\mathbb{K}}})\otimes\Lambda_{{\mathbb{K}},univ} is semisimple over Λ𝕂,univ\Lambda_{{\mathbb{K}},univ}. Denote by Λω,𝕂j\Lambda_{\omega,{\mathbb{K}}}^{j} the degree 2j2j component of Λω,𝕂\Lambda_{\omega,{\mathbb{K}}}. Note that,

QH2n(M,Λω,𝕂)=H2i(H;𝕂)Λω,𝕂ni.\operatorname{\mathrm{QH}}_{2n}(M,\Lambda_{\omega,{\mathbb{K}}})=\bigoplus\operatorname{\mathrm{H}}_{2i}(H;{\mathbb{K}})\otimes\Lambda_{\omega,{\mathbb{K}}}^{n-i}.

Therefore, QHev(M,Λ𝕂,univ)=QH2n(M,Λω,𝕂)Λ𝕂,univ\operatorname{\mathrm{QH}}_{ev}(M,\Lambda_{{\mathbb{K}},univ})=\operatorname{\mathrm{QH}}_{2n}(M,\Lambda_{\omega,{\mathbb{K}}})\otimes\Lambda_{{\mathbb{K}},univ} is semisimple.

2.4.3. PSS isomorphism

Piunikhin, Salamon and Schwarz [PSS96] defined the PSSPSS isomorphism between Hamiltonian Floer homology and quantum homology

PSSH:QH(M)HF(H)\operatorname{\mathrm{PSS}}_{H}:\operatorname{\mathrm{QH}}(M)\to\operatorname{\mathrm{HF}}(H).

On the chain level and for generic auxiliary data, the map is defined by counting certain isolated configurations consisting of negative gradient trajectories γ:(,0]M\gamma:(-\infty,0]\rightarrow M of a generic Morse-Smale pair666A Morse function ff and Riemannian metric gg on M,M, satisfying the Morse-Smale condition. incident at γ(0)\gamma(0) with the asymptotic limsu(s,)\lim_{s\to-\infty}u(s,\cdot) of a map u:×S1Mu:{\mathbb{R}}\times\mathrm{S}^{1}\rightarrow M of finite energy, satisfying the Floer equation

us+Jt(u)(utXKt(u))=0,\frac{\partial u}{\partial s}+J_{t}(u)\left(\frac{\partial u}{\partial t}-X_{K}^{t}(u)\right)=0,

for (s,t)×S1(s,t)\in{\mathbb{R}}\times\mathrm{S}^{1} and K(s,t)C(M,)K(s,t)\in C^{\infty}(M,{\mathbb{R}}) a small perturbation of β(s)Ht\beta(s)H_{t} such that K(s,t)=β(s)HtK(s,t)=\beta(s)H_{t} for s1s\ll-1 and for s+1s\gg+1. Here β:[0,1]\beta:{\mathbb{R}}\rightarrow[0,1] is a smooth function satisfying β(s)=0\beta(s)=0 for s1s\ll-1 and β(s)=1\beta(s)=1 for s+1s\gg+1. This map produces an isomorphism of Λ𝕂,ω\Lambda_{{\mathbb{K}},\omega}-modules, which intertwines the quantum product on QH(M)\operatorname{\mathrm{QH}}(M) with the pair-of-pants product on Hamiltonian Floer homology. It is extended by linearity when we work over the universal Novikov ring Λ𝕂,univ\Lambda_{{\mathbb{K}},univ} and its algebraic closure Λ¯𝕂,univ\overline{\Lambda}_{{\mathbb{K}},univ}.

2.4.4. Filtration on quantum homology

Consider the non-Archimedean valuation ν\nu on Λ,univ\Lambda_{\mathbb{Z},univ} defined in Section 2.3.2. For each element fiαi\sum f_{i}\alpha_{i} where fiΛ,univf_{i}\in\Lambda_{\mathbb{Z},univ} and αiH(M)\alpha_{i}\in\operatorname{\mathrm{H}}_{*}(M) define the filtration l:QH(M){}l:\operatorname{\mathrm{QH}}(M)\rightarrow{\mathbb{R}}\cup\{-\infty\} to be l(fiαi)=max{ν(fi)}l(\sum f_{i}\alpha_{i})=\max\{-\nu(f_{i})\}. Now, as in [PSS17, She22], for each elementor each αQH(M)\alpha\in\operatorname{\mathrm{QH}}(M), we have a map

α:HF(H)<a\displaystyle\alpha*:\operatorname{\mathrm{HF}}(H)^{<a} HF(H)<a+l(α)\displaystyle\longrightarrow\operatorname{\mathrm{HF}}(H)^{<a+l(\alpha)}

defined by counting negative gg-gradient trajectories v:(,0]Mv:(-\infty,0]\rightarrow M of a Morse function ff on MM, for a Morse-Smale pair (f,g)(f,g), asymptotic to critical points of ff as ss\rightarrow-\infty, and with γ(0)\gamma(0) incident to Floer cylinders u:×S1Mu:{\mathbb{R}}\times\mathrm{S}^{1}\rightarrow M at u(0,0)u(0,0). This construction is reminiscent of the quantum cap product as in [PSS96, Sch00, Sei02, Flo89].

2.5. Spectral invariants

2.5.1. Definitions and basic properties

Consider the filtered complex (CF(H;Λ),𝒜)(\operatorname{\mathrm{CF}}(H;\Lambda),\mathcal{A}), where Λ\Lambda is one of the following Λ𝕂,ω,Λ𝕂,univ,Λ¯𝕂,univ\Lambda_{{\mathbb{K}},\omega},\Lambda_{{\mathbb{K}},univ},\overline{\Lambda}_{{\mathbb{K}},univ} and 𝒜\mathcal{A} is as in Section 2.3.4. Denote CF(H;Λ)<c=𝒜1(,c)\operatorname{\mathrm{CF}}(H;\Lambda)^{<c}=\mathcal{A}^{-1}(-\infty,c) and HF(H;Λ)<c\operatorname{\mathrm{HF}}(H;\Lambda)^{<c} the filtered homology groups. The spectral invariant associated to a non-trivial αQH(M)\alpha\in\operatorname{\mathrm{QH}}(M) is defined as

c(α,H)=inf{a|PSSH(α)im(HF(H;Λ)<aHF(H;Λ))}.c(\alpha,H)=\inf\{a\in{\mathbb{R}}\,|\,\operatorname{\mathrm{PSS}}_{H}(\alpha)\in\operatorname{\mathrm{im}}(\operatorname{\mathrm{HF}}(H;\Lambda)^{<a}\rightarrow\operatorname{\mathrm{HF}}(H;\Lambda))\}.

By [BC09], spectral invariants do not change under extension of coefficients, in particular, we do not need to specify the Λ\Lambda in the notation. Spectral invariants enjoy a wealth of useful properties established by Schwarz [Sch00], Viterbo [Vit92], Oh [Oh05, Oh06, MS12] and generalized by Usher [Ush08, Ush10]. We summarize some of their properties.

Proposition 2.10.

The spectral invariants satisfy the following:

  1. (1)

    Stability: for all H,GH,G\in\mathcal{H} and αQH(M){0}\alpha\in\operatorname{\mathrm{QH}}(M)\setminus\{0\},

    01min(HtGt)𝑑tc(α,H)c(α,G)01max(HtGt)𝑑t.\int^{1}_{0}\min(H_{t}-G_{t})dt\leq c(\alpha,H)-c(\alpha,G)\leq\int^{1}_{0}\max(H_{t}-G_{t})dt.
  2. (2)

    Triangle inequality: for all H,GH,G\in\mathcal{H} and α,βQH(M){0}\alpha,\beta\in\operatorname{\mathrm{QH}}(M)\setminus\{0\},

    c(αβ,H#G)c(α,H)+c(β,G).c(\alpha*\beta,H\#G)\leq c(\alpha,H)+c(\beta,G).
  3. (3)

    Novikov action: for all H,αQH(M){0}H\in\mathcal{H},\alpha\in\operatorname{\mathrm{QH}}(M)\setminus\{0\}, and λΛ\lambda\in\Lambda,

    c(λα,H)=c(α,H)ν(λ).c(\lambda\alpha,H)=c(\alpha,H)-\nu(\lambda).
  4. (4)

    Non-Archimedean property: for all HH\in\mathcal{H} and α,βQH(M){0}\alpha,\beta\in\operatorname{\mathrm{QH}}(M)\setminus\{0\},

    c(α+β,H)max{c(α,H),c(β,H)}.c(\alpha+\beta,H)\leq\max\{c(\alpha,H),c(\beta,H)\}.

We remark that by the stability property, the spectral invariants are defined for all HH\in\mathcal{H} and all the above listed properties apply in this generality.

2.5.2. Poincare Duality for Spectral Invariants

Let ϕ\phi be a non-degenerate Hamiltonian diffeomorphism and fix a choice of capping x¯k\overline{x}_{k} for each 11-periodic orbit xkx_{k}. Define a bilinear pairing

Δ:CF(H,Λ¯𝔽p,univ)×CF(H¯,Λ¯𝔽p,univ)\displaystyle\Delta:\operatorname{\mathrm{CF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\times\operatorname{\mathrm{CF}}_{*}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) Λ¯𝔽p,univ\displaystyle\longrightarrow\overline{\Lambda}_{{\mathbb{F}}_{p},univ}
(aix¯i,bjx¯j)\displaystyle\Big{(}\sum a_{i}\overline{x}_{i},\sum b_{j}\overline{x}^{\prime}_{j}\Big{)} aibi\displaystyle\longmapsto\sum a_{i}b_{i}

where the sums are finite and for all kk, the capped orbit x¯k\overline{x}^{\prime}_{k} is equal to x¯k\overline{x}_{k} with reversed orientation, in particular, xk(t)=xk(1t)x^{\prime}_{k}(t)=x_{k}(1-t).

Lemma 2.11.

The bilinear pairing Δ\Delta is non-degenerate.

Proof.

Suppose Δ(aix¯i,)=0\Delta(\sum a_{i}\overline{x}_{i},\cdot)=0. Then, for every x¯i\overline{x}^{\prime}_{i},

Δ(aix¯i,x¯i)=0.\Delta\Big{(}\sum a_{i}\overline{x}_{i},\overline{x}^{\prime}_{i}\Big{)}=0.

On the other hand, Δ(aix¯i,x¯i)=ai\Delta(\sum a_{i}\overline{x}_{i},\overline{x}^{\prime}_{i})=a_{i}. Thus ai=0a_{i}=0 for each ii, i.e. aix¯i=0\sum a_{i}\overline{x}_{i}=0. Similarly, if Δ(,bjx¯j)=0\Delta(\cdot,\sum b_{j}\overline{x}^{\prime}_{j})=0, then bjx¯j=0\sum b_{j}\overline{x}^{\prime}_{j}=0. ∎

Lemma 2.12.

For any real number α\alpha the composition νΔ\nu\circ\Delta is positive on

CF(H,Λ¯𝔽p,univ)<α×CF(H¯,Λ¯𝔽p,univ)<α.\operatorname{\mathrm{CF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{<\alpha}\times\operatorname{\mathrm{CF}}_{*}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{<-\alpha}.
Proof.

Suppose that

(aix¯i,bix¯i)CF(H,Λ¯𝔽p,univ)<α×CF(H¯,Λ¯𝔽p,univ)<α.\Big{(}\sum a_{i}\overline{x}_{i},\sum b_{i}\overline{x}^{\prime}_{i}\Big{)}\in\operatorname{\mathrm{CF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{<\alpha}\times\operatorname{\mathrm{CF}}_{*}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{<-\alpha}.

Then,

𝒜(aix¯i)=max{𝒜H(x¯i)ν(ai)}<α.\mathcal{A}\big{(}\sum a_{i}\overline{x}_{i}\big{)}=\max\{\mathcal{A}_{H}(\overline{x}_{i})-\nu(a_{i})\}<\alpha.

In particular, we have that ν(ai)>𝒜H(x¯i)α\nu(a_{i})>\mathcal{A}_{H}(\overline{x}_{i})-\alpha for all ii. Following the same logic, ν(bi)>𝒜H(x¯i)+α\nu(b_{i})>-\mathcal{A}_{H}(\overline{x}_{i})+\alpha for all ii. Thus,

ν(Δ(aix¯i,bix¯i))\displaystyle\nu\Big{(}\Delta\Big{(}\sum a_{i}\overline{x}_{i},\sum b_{i}\overline{x}^{\prime}_{i}\Big{)}\Big{)} =ν(aibi)\displaystyle=\nu\Big{(}\sum a_{i}b_{i}\Big{)}
min{ν(aibi)}\displaystyle\geq\min\{\nu(a_{i}b_{i})\}
=min{ν(ai)+ν(bi)}\displaystyle=\min\{\nu(a_{i})+\nu(b_{i})\}
>𝒜H(x¯i)α𝒜H(x¯i)+α=0\displaystyle>\mathcal{A}_{H}(\overline{x}_{i})-\alpha-\mathcal{A}_{H}(\overline{x}_{i})+\alpha=0

Lemma 2.13.

Let aCF(H,Λ¯𝔽p,univ)a\in\operatorname{\mathrm{CF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) and bCF(H¯,Λ¯𝔽p,univ)b\in\operatorname{\mathrm{CF}}_{*}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ}). Then, Δ((a),b)=±Δ(a,(b))\Delta(\partial(a),b)=\pm\Delta(a,\partial(b)). In particular, there is an induced pairing on homology

Δ:HF(H,Λ¯𝔽p,univ)×HF(H¯,Λ¯𝔽p,univ)Λ¯𝔽p,univ\Delta:\operatorname{\mathrm{HF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\times\operatorname{\mathrm{HF}}_{*}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\longrightarrow\overline{\Lambda}_{{\mathbb{F}}_{p},univ}
Proof.

Choose a basis {x¯i}\{\overline{x}_{i}\} of CFk+1(H,Λ¯𝔽p,univ)\operatorname{\mathrm{CF}}_{k+1}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) and a basis {y¯j}\{\overline{y}_{j}\} of CFk(H,Λ¯𝔽p,univ)\operatorname{\mathrm{CF}}_{k}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}). Then {x¯i}\{\overline{x}^{\prime}_{i}\} is a basis of CFk1(H¯,Λ¯𝔽p,univ)\operatorname{\mathrm{CF}}_{-k-1}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) and {y¯j}\{\overline{y}^{\prime}_{j}\} is basis of CFk(H¯,Λ¯𝔽p,univ)\operatorname{\mathrm{CF}}_{-k}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ}). Suppose x¯i=ajy¯j\partial\overline{x}_{i}=\sum a_{j}\overline{y}_{j} and y¯j=bix¯i\partial\overline{y}^{\prime}_{j}=\sum b_{i}\overline{x}^{\prime}_{i}. Then,

Δ(x¯i,y¯j)=Δ(ajy¯j,y¯j)=aj.\Delta(\partial\overline{x}_{i},\overline{y}^{\prime}_{j})=\Delta\big{(}\sum a_{j}\overline{y}_{j},\overline{y}^{\prime}_{j}\big{)}=a_{j}.

Similarly, Δ(x¯i,y¯j)=bi\Delta(\overline{x}_{i},\partial\overline{y}_{j})=b_{i}. By definition, aja_{j} is the number of Floer trajectories connecting x¯i\overline{x}_{i} and y¯j\overline{y}_{j} and bib_{i} is the number of the Floer trajectory connecting y¯j\overline{y}^{\prime}_{j} and x¯i\overline{x}^{\prime}_{i}. Thus aj=bia_{j}=b_{i}, i.e.

Δ(x¯i,y¯j)=Δ(x¯i,y¯j).\Delta(\partial\overline{x}_{i},\overline{y}_{j})=\Delta(\overline{x}_{i},\partial\overline{y}_{j}).

Since Δ\Delta is bilinear, we have that Δ(a,b)=Δ(a,b)\Delta(\partial a,b)=\Delta(a,\partial b) for any aCF(H,Λ¯𝔽p,univ)a\in\operatorname{\mathrm{CF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) and bCF(H¯,Λ¯𝔽p,univ)b\in\operatorname{\mathrm{CF}}_{*}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ}). ∎

The following proposition is interpreted as Poincaré duality in Floer theory. The equality follows from standard arguments in [Ush10, Corollary 1.4] and [EP03, Lemma 2.2].

Proposition 2.14.

Let aQH(M,Λ¯𝔽p,univ)a\in\operatorname{\mathrm{QH}}_{*}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) be non-trivial, then,

c(a,H)=inf{c(b,H¯)|bQH(M,Λ¯𝔽p,univ),ν(Δ(PSSH(a),PSSH¯(b)))0}.c(a,H)=-\inf\{c(b,\overline{H})\,|\,b\in\operatorname{\mathrm{QH}}_{*}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}),\,\nu(\Delta(\operatorname{\mathrm{PSS}}_{H}(a),\operatorname{\mathrm{PSS}}_{\overline{H}}(b)))\leq 0\}.
Proof.

We first show that

c(a,H)inf{c(b,H¯)|ν(Δ(PSSH(a),PSSH¯(b))0}.c(a,H)\geq-\inf\{c(b,\overline{H})\,|\,\nu(\Delta(\operatorname{\mathrm{PSS}}_{H}(a),\operatorname{\mathrm{PSS}}_{\overline{H}}(b))\geq 0\}.

Suppose that α<c(a,H)\alpha<c(a,H). We have a short exact sequence of chain complexes

0CF(H,Λ¯𝔽p,univ)<αCF(H,Λ¯𝔽p,univ)CF(H,Λ¯𝔽p,univ)(α,)0,0\rightarrow\operatorname{\mathrm{CF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{<\alpha}\rightarrow\operatorname{\mathrm{CF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\rightarrow\operatorname{\mathrm{CF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{(\alpha,\infty)}\rightarrow 0,

inducing an exact sequence on homology

HF(H,Λ¯𝔽p,univ)<αiαHF(H,Λ¯𝔽p,univ)παHF(H,Λ¯𝔽p,univ)(α,).\operatorname{\mathrm{HF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{<\alpha}\xrightarrow{i_{\alpha}}\operatorname{\mathrm{HF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\xrightarrow{\pi_{\alpha}}\operatorname{\mathrm{HF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{(\alpha,\infty)}.

The fact that α<c(a,H)\alpha<c(a,H) means that PSSH(a)\operatorname{\mathrm{PSS}}_{H}(a) is not represented by any chains of filtration level at most α\alpha, so that PSSH(a)im(iα)\operatorname{\mathrm{PSS}}_{H}(a)\notin\operatorname{\mathrm{im}}(i_{\alpha}), thus πα(PSSH(a))0\pi_{\alpha}(\operatorname{\mathrm{PSS}}_{H}(a))\neq 0. Fix a representative a~\widetilde{a} of PSSH(a)\operatorname{\mathrm{PSS}}_{H}(a). There is bQH(M,Λ¯𝔽p,univ)b\in\operatorname{\mathrm{QH}}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) such that ν(Δ(PSSH(a),PSSH¯(b)))=0\nu(\Delta(\operatorname{\mathrm{PSS}}_{H}(a),\operatorname{\mathrm{PSS}}_{\overline{H}}(b)))=0 and 𝒜(b~)𝒜(a~)<α\mathcal{A}(\widetilde{b})\leq-\mathcal{A}(\widetilde{a})<-\alpha, b~\widetilde{b} represents PSSH¯(b)\operatorname{\mathrm{PSS}}_{\overline{H}}(b). Thus,

inf{c(b,H¯)|bQH(M,Λ¯𝔽p,univ),ν(Δ(PSSH(a),PSSH¯(b))0}c(a,H)-\inf\big{\{}c(b,\overline{H})\,|\,b\in\operatorname{\mathrm{QH}}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}),\,\nu(\Delta(\operatorname{\mathrm{PSS}}_{H}(a),\operatorname{\mathrm{PSS}}_{\overline{H}}(b))\leq 0\big{\}}\leq c(a,H)

because α\alpha is an arbitrary number smaller than c(a,H)c(a,H). We now detail how to find such a class bb. Let {x¯1,,x¯N}\{\overline{x}^{*}_{1},\dots,\overline{x}^{*}_{N}\} be a Λ¯𝔽p,univ\overline{\Lambda}_{{\mathbb{F}}_{p},univ}-basis of the dual vector space CF(H,Λ¯𝔽p,univ)\operatorname{\mathrm{CF}}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{*}. It not hard to check that x¯i=Δ(,x¯i)\overline{x}^{*}_{i}=\Delta(-,\overline{x}^{\prime}_{i}) for all ii, which yields an identification

CF(H,Λ¯𝔽p,univ)CF(H¯,Λ¯𝔽p,univ).\operatorname{\mathrm{CF}}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{*}\cong\operatorname{\mathrm{CF}}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ}).

Let {ξ1,,ξB,η1,,ηK,ζ1,,ζK}\{\xi_{1},\dots,\xi_{B},\eta_{1},\dots,\eta_{K},\zeta_{1},\dots,\zeta_{K}\} be a singular value decomposition for the complex CF(H,Λ¯𝔽p,univ)\operatorname{\mathrm{CF}}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) as detailed in Section 2.6.3. We recall from [UZ16, Proposition 2.20] that there is an 𝒜\mathcal{A}^{*}-orthogonal dual basis {ξ1,,ξB,η1,,ηK,ζ1,,ζK}\{\xi^{*}_{1},\dots,\xi^{*}_{B},\eta^{*}_{1},\dots,\eta^{*}_{K},\zeta^{*}_{1},\dots,\zeta^{*}_{K}\} of CF(H,Λ¯𝔽p,univ)\operatorname{\mathrm{CF}}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{*} such that 𝒜(ξi)=𝒜(ξi)\mathcal{A}^{*}(\xi_{i})=-\mathcal{A}(\xi_{i}), 𝒜(ηi)=𝒜(ηi)\mathcal{A}^{*}(\eta_{i})=-\mathcal{A}(\eta_{i}), and 𝒜(ζi)=𝒜(ζi)\mathcal{A}^{*}(\zeta_{i})=-\mathcal{A}(\zeta_{i}), where

𝒜(f)=sup{𝒜(θ)ν(f(θ))|θCF(H,Λ¯𝔽p,univ),θ0}.\mathcal{A}^{*}(f)=\sup\{-\mathcal{A}(\theta)-\nu(f(\theta))\,|\,\theta\in\operatorname{\mathrm{CF}}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}),\theta\neq 0\}.

Note that if f=fix¯if=\sum f_{i}\overline{x}_{i}^{*}, then 𝒜(fix¯i)𝒜(f)\mathcal{A}(\sum f_{i}\overline{x}_{i}^{\prime})\leq\mathcal{A}^{*}(f). Let, ξi,ηi\xi_{i}^{\prime},\eta_{i}^{\prime} and ζi\zeta_{i}^{\prime} be defined by the property that Δ(,ξi)=ξi\Delta(-,\xi_{i}^{\prime})=\xi_{i}^{*}, Δ(,ηi)=ηi\Delta(-,\eta_{i}^{\prime})=\eta_{i}^{*}, and Δ(,ζi)=ζi\Delta(-,\zeta_{i}^{\prime})=\zeta_{i}^{*}. We verify that dξi=0d\xi_{i}^{\prime}=0. Suppose,

dξi=aijξj+bijηj+cijζj,d\xi_{i}^{\prime}=\sum a_{ij}\xi_{j}^{\prime}+\sum b_{ij}\eta_{j}^{\prime}+c_{ij}\zeta_{j}^{\prime},

then, aij=Δ(ξj,dξi)=Δ(dξj,ξi)=0a_{ij}=\Delta(\xi_{j},d\xi_{i}^{\prime})=\Delta(d\xi_{j},\xi_{i}^{\prime})=0 and similarly, bij=cij=0b_{ij}=c_{ij}=0. Thus, dξi=0d\xi_{i}^{\prime}=0. Let bb be the class PSSH¯1([b~])\operatorname{\mathrm{PSS}}_{\overline{H}}^{-1}([\widetilde{b}]), where b~=a11ξ1\widetilde{b}=a_{1}^{-1}\xi_{1}^{\prime} and a~=aiξi\widetilde{a}=\sum a_{i}\xi_{i}. We now show that

c(a,H)inf{c(b,H¯)|ν(Δ(PSSH(a),PSSH¯(b))0}.c(a,H)\leq-\inf\{c(b,\overline{H})\,|\,\nu(\Delta(\operatorname{\mathrm{PSS}}_{H}(a),\operatorname{\mathrm{PSS}}_{\overline{H}}(b))\leq 0\}.

Suppose that α>c(a,H)\alpha>c(a,H). Thus, there must be some cycle cCF(H,Λ¯𝔽p,univ)<αc\in\operatorname{\mathrm{CF}}_{*}(H,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{<\alpha} representing the class PSSH(a)\operatorname{\mathrm{PSS}}_{H}(a). If bQH(M,Λ¯𝔽p,univ)b\in\operatorname{\mathrm{QH}}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) is an arbitrary class satisfying ν(Δ(PSSH(a),PSSH¯(b)))0\nu(\Delta(\operatorname{\mathrm{PSS}}_{H}(a),\operatorname{\mathrm{PSS}}_{\overline{H}}(b)))\geq 0, then by the definition of Δ\Delta it must hold that every representative dCF(H¯,Λ¯𝔽p,univ)d\in\operatorname{\mathrm{CF}}_{*}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) of the class PSSH¯(b)\operatorname{\mathrm{PSS}}_{\overline{H}}(b) satisfies ν(Δ(c,d))0\nu(\Delta(c,d))\geq 0. By Lemma 2.12, this can only be true if no representative dd belongs to CF(H¯,Λ¯𝔽p,univ)(,α)\operatorname{\mathrm{CF}}_{*}(\overline{H},\overline{\Lambda}_{{\mathbb{F}}_{p},univ})^{(-\infty,-\alpha)}, which amounts to the statement that c(b,H¯)αc(b,\overline{H})\geq-\alpha. Note that bb is an arbitrary class with ν(Δ(PSSH(a),PSSH¯(b)))0\nu(\Delta(\operatorname{\mathrm{PSS}}_{H}(a),\operatorname{\mathrm{PSS}}_{\overline{H}}(b)))\leq 0, while α\alpha is an arbitrary number exceeding c(a,H)c(a,H), and so we obtain the desired inequality. ∎

2.6. Floer persistence

2.6.1. Overview of persistence modules

The theory of persistence modules has origins in the field of topological data analysis. It was introduced by Carlsson and Zomordian [ZC04] as an algebraic tool whose purpose was to deal with persistence homology invented by Edelsbrunner, Letscher and Zomordian [ELZ00] to study topological aspects of large data sets. Persistence modules have since then proven useful in many disciplines of pure mathematics such as metric geometry and calculus of variations. Polterovich and Shelukhin [PS16] (see also [PSS17, UZ16]) were the first to view filtered Floer homology as a persistence module in order to prove interesting results about autonomous Hamiltonian diffeomorphisms. Since then, this viewpoint has led to several applications such as Shelukhin’s proof [She22] of the Hofer-Zehnder conjecture in the monotone setting (under the semisimplicity assumption) and to obstructions to the existence of non-trivial finite subgroups of Ham(M,ω)\operatorname{\mathrm{Ham}}(M,\omega) [AS23]. However, in order to view filtered Floer homology as a persistence module a certain finiteness assumption is required. In practice this means that going beyond the monotone setting requires some work. In [UZ16], Usher and Zhang generalized the notion of a barcode (the main invariant obtained from a persistence module) to the semipositive setting.

2.6.2. Persistence modules

In this section we follow [She22, Section 4.4.1] in order define persistence modules and their associated barcodes and discuss the relation between them.

Let 𝕂{\mathbb{K}} be a field. Denote by Vect𝕂{\mathrm{Vect}}_{{\mathbb{K}}} the category of finite dimensional 𝕂{\mathbb{K}}-vector spaces and by (,)({\mathbb{R}},\leq) the poset category of {\mathbb{R}}. A persistence module over 𝕂{\mathbb{K}} is a functor

V:(,)Vect𝕂.V:({\mathbb{R}},\leq)\rightarrow{\mathrm{Vect}}_{{\mathbb{K}}}.

The collection of such functors together with their natural transformations form an abelian category Fun((,),Vect𝕂){\mathrm{Fun}}(({\mathbb{R}},\leq),{\mathrm{Vect}}_{{\mathbb{K}}}). We consider a full abelian subcategory

𝐩𝐦𝐨𝐝Fun((,),Vect𝕂),{\mathbf{pmod}}\subset{\mathrm{Fun}}(({\mathbb{R}},\leq),{\mathrm{Vect}}_{{\mathbb{K}}}),

which is defined by requiring that certain technical assumptions are satisfied. The following definition summarizes the data of such a persistence module.

Definition 2.15.

A persistence module VV in 𝐩𝐦𝐨𝐝\mathbf{pmod} consists of a family

{VaVect𝕂}a\{V^{a}\in{\mathrm{Vect}}_{{\mathbb{K}}}\}_{a\in{\mathbb{R}}}

of vector spaces and 𝕂{\mathbb{K}}-linear maps πVa,b:VaVb\pi_{V}^{a,b}:V^{a}\rightarrow V^{b} for each aba\leq b such that πVa,a=idVa\pi_{V}^{a,a}=\operatorname{\mathrm{id}}_{V^{a}}, and πVb,cπVa,b=πVa,c\pi_{V}^{b,c}\circ\pi_{V}^{a,b}=\pi_{V}^{a,c} for all abca\leq b\leq c. Furthermore, we require them to satisfy the following:

  1. (1)

    Support: Va=0V^{a}=0 for all a0a\ll 0.

  2. (2)

    Finiteness: there exists a finite subset SS\subset{\mathbb{R}} such that for all a,ba,b in the same connected components of S{\mathbb{R}}\setminus S, the map πVa,b\pi_{V}^{a,b} is an isomorphism.

  3. (3)

    Continuity: for every two consecutive elements s<ss<s^{\prime} of SS, and any a(s,s)a\in(s,s^{\prime}), the map πVa,s\pi^{a,s^{\prime}}_{V} is an isomorphism.

We define V=limaVaV^{\infty}=\lim_{a\rightarrow\infty}V^{a}.

The normal form theorem [ZC04, CB15] implies that the isomorphism classes of a persistence module V𝐩𝐦𝐨𝐝V\in\mathbf{pmod} is determined by its barcode, that is, a multiset (V)={(Ik,mk)}1kN\mathcal{B}(V)=\{(I_{k},m_{k})\}_{1\leq k\leq N} of intervals IkI_{k}\subset{\mathbb{R}} with multiplicities mk>0m_{k}\in{\mathbb{Z}}_{>0}. The intervals are of two types, K=K(V)K=K(V) of them are finite, Ik=(ak,bk]I_{k}=(a_{k},b_{k}], and B=B(V)=NKB=B(V)=N-K are infinite, Ik=(ak,)I_{k}=(a_{k},\infty). The intervals are called bars and the bar-lengths are defined as |Ik|=bkak|I_{k}|=b_{k}-a_{k} in the finite case, and |Ik|=+|I_{k}|=+\infty otherwise.

The isometry theorem [CDSGO16, BL15, CCSG+09, CSEH05] shows that the barcode assignment map

:(𝐩𝐦𝐨𝐝,dinter)\displaystyle\mathcal{B}:(\mathbf{pmod},d_{inter}) (𝐛𝐚𝐫𝐜𝐨𝐝𝐞𝐬,dbottle)\displaystyle\rightarrow(\mathbf{barcodes},d_{bottle})
V\displaystyle V (V)\displaystyle\mapsto\mathcal{B}(V)

is an isometry. The interleaving distance is defined by setting

dinter(V,W)=inf{δ0|δ-interleaving,fhom(V,W[δ]),ghom(W,V[δ])},d_{inter}(V,W)=\inf\{\delta\geq 0\,|\,\exists\,\delta\text{-interleaving},\,f\in{\mathrm{hom}}(V,W[\delta]),g\in{\mathrm{hom}}(W,V[\delta])\},

where for V𝐩𝐦𝐨𝐝V\in\mathbf{pmod} and cc\in{\mathbb{R}}, V[c]𝐩𝐦𝐨𝐝V[c]\in\mathbf{pmod} is given by pre-composing with the functor Tc:(,)(,)T_{c}:({\mathbb{R}},\leq)\rightarrow({\mathbb{R}},\leq), tt+ct\mapsto t+c. We say that a pair fhom(V,W[c]),ghom(W,V[c])f\in{\mathrm{hom}}(V,W[c]),g\in{\mathrm{hom}}(W,V[c]) is a cc-interleaving if

g[c]f=sh2δ,V,f[c]g=sh2δ,W,g[c]\circ f={\mathrm{sh}}_{2\delta,V},\quad f[c]\circ g={\mathrm{sh}}_{2\delta,W},

where for c0c\geq 0, shc,Vhom(V,V[c]){\mathrm{sh}}_{c,V}\in{\mathrm{hom}}(V,V[c]) is the natural transformation id(,)Tc\operatorname{\mathrm{id}}_{({\mathbb{R}},\leq)}\rightarrow T_{c}. Note that, dinter(V,W)0{}d_{inter}(V,W)\in{\mathbb{R}}_{\geq 0}\cup\{\infty\}, and it is finite if and only if VWV^{\infty}\cong W^{\infty}.

The bottleneck distance is defined as

dbottle(,𝒞)=inf{δ>0|δ-matching between ,𝒞},d_{bottle}(\mathcal{B},\mathcal{C})=\inf\{\delta>0\,|\,\exists\,\delta\text{-matching between }\mathcal{B},\mathcal{C}\},

where a δ\delta-matching between ,𝒞\mathcal{B},\mathcal{C} is defined as bijection σ:2δ𝒞2δ\sigma:\mathcal{B}^{2\delta}\rightarrow\mathcal{C}^{2\delta} between the sub-multisets 2δ\mathcal{B}^{2\delta}\subset\mathcal{B}, 𝒞2δ𝒞\mathcal{C}^{2\delta}\subset\mathcal{C}, which contain the bars of ,𝒞\mathcal{B},\mathcal{C}, respectively, with bar-length greater than 2δ2\delta, such that if σ((a,b])=(c,d]\sigma((a,b])=(c,d], then max{|ac|,|bd|}δ\max\{|a-c|,|b-d|\}\leq\delta. We have that dbottle(,𝒞)0{}d_{bottle}(\mathcal{B},\mathcal{C})\in{\mathbb{R}}_{\geq 0}\cup\{\infty\}, with it being finite if and only if B()=B(𝒞)B(\mathcal{B})=B(\mathcal{C}).

Note that for each cc\in{\mathbb{R}} there is an isometry given by sending a barcode ={(Ik,mk)}\mathcal{B}=\{(I_{k},m_{k})\} to [c]={(Ikc,mk)}\mathcal{B}[c]=\{(I_{k}-c,m_{k})\}. We can therefore consider the quotient space (𝐛𝐚𝐫𝐜𝐨𝐝𝐞𝐬,dbottle)(\mathbf{barcodes^{\prime}},{d}_{bottle}^{\prime}) by this isometric {\mathbb{R}}-action, where

dbottle([],[𝒞])=infcdbottle(,𝒞[c]).d_{bottle}^{\prime}([\mathcal{B}],[\mathcal{C}])=\inf_{c\in{\mathbb{R}}}d_{bottle}(\mathcal{B},\mathcal{C}[c]).

We observe that bar-lengths are well-defined in the quotient.

2.6.3. Barcodes of Hamiltonian diffeomorphisms

In the symplectically aspherical and monotone settings, filtered Hamiltonian Floer homology together with the maps

HFk(H)<aHFk(H)<b,\operatorname{\mathrm{HF}}_{k}(H)^{<a}\rightarrow\operatorname{\mathrm{HF}}_{k}(H)^{<b},

for a<ba<b, induced by the natural inclusions, were studied in [PS16, PSS17, She22] from the viewpoint of persistence modules. In particular, one can associate a barcode to each Hamiltonian diffeomorphism with finitely many fixed points, and, hence, a bar-length spectrum. We refer to [PS16, PSS17, UZ16] for details of the construction and for first properties. Following the discussion in [She22, Section 4.4], we describe two alternative descriptions of the bar-length spectrum, which coincide in the semipositive setting. All three descriptions coincide in the monotone setting [She22, Lemma 9].

Consider the filtered Floer chain complex (CF(H;Λ),d,𝒜)(\operatorname{\mathrm{CF}}(H;\Lambda),d,\mathcal{A}), where Λ\Lambda is one of the following Λ𝕂,univ,Λ¯𝕂,univ\Lambda_{{\mathbb{K}},univ},\overline{\Lambda}_{{\mathbb{K}},univ}, the non-Archimidean filtration 𝒜\mathcal{A} is as in Section 2.3.4, and dd is the Floer differential. Then, by [UZ16], the complex (CF(H;Λ),d)(\operatorname{\mathrm{CF}}(H;\Lambda),d) admits an orthogonal basis

E=(ξ1,,ξB,η1,,ηK,ζ1,,ζK)E=(\xi_{1},\dots,\xi_{B},\eta_{1},\dots,\eta_{K},\zeta_{1},\dots,\zeta_{K})

such that dξj=0d\xi_{j}=0 for all j{1,,B}j\in\{1,\dots,B\}, and dζj=ηjd\zeta_{j}=\eta_{j} for all j{1,,K}j\in\{1,\dots,K\}. The lengths of the finite bars are given by

βj=βj(ϕH,𝕂)=𝒜(ζj)𝒜(ηj),\beta_{j}=\beta_{j}(\phi_{H},{\mathbb{K}})=\mathcal{A}(\zeta_{j})-\mathcal{A}(\eta_{j}),

which we can assume to satisfy β1βK\beta_{1}\leq\dots\leq\beta_{K}. The length of the largest finite bar, is the boundary-depth introduced by Usher [Ush11a, Ush13], and denoted by β(ϕH,𝕂)\beta(\phi_{H},{\mathbb{K}}). There are BB infinite bar-lengths corresponding to each ξj\xi_{j}. This description yields the identity N=B+2KN=B+2K, where N,BN,B, and KK can be computed by N=dimΛCF(H,Λ),B=dimΛHF(H,Λ)N=\dim_{\Lambda}\operatorname{\mathrm{CF}}(H,\Lambda),B=\dim_{\Lambda}\operatorname{\mathrm{HF}}(H,\Lambda), and K=dimΛim(d)K=\dim_{\Lambda}\operatorname{\mathrm{im}}(d). We denote by

βtot(ϕ,𝕂)=β1(ϕ,𝕂)++βK(ϕ,𝕂)(ϕ,𝕂)\beta_{tot}(\phi,{\mathbb{K}})=\beta_{1}(\phi,{\mathbb{K}})+\cdots+\beta_{K(\phi,{\mathbb{K}})}(\phi,{\mathbb{K}})

the total bar-length associated to the barcode.

Following [FOOO13], the Floer differential dd in the orthonormal basis described in Section 2.3.4 has coefficients in Λ0\Lambda^{0}. Therefore, one defines a Floer complex (CF(H,Λ0),d)(\operatorname{\mathrm{CF}}(H,\Lambda^{0}),d) whose homology is a finitely generated Λ0\Lambda^{0}-module, and is therefore of the form 𝒯\mathcal{F}\oplus\mathcal{T}, where \mathcal{F} is a free Λ0\Lambda^{0}-module and 𝒯\mathcal{T} is a torsion Λ0\Lambda^{0}-module. The bar-lengths are given by the isomorphism

𝒯1jKΛ0/(Tβj).\mathcal{T}\cong\bigoplus_{1\leq j\leq K}\Lambda^{0}/(T^{\beta_{j}}).

Combining the ideas in the proof of [She22, Lemma 16] and combining with Proposition 3.1 we show that the bar-length spectrum over 𝔽p{\mathbb{F}}_{p} coincides with that over {\mathbb{Q}} for a sufficiently large prime pp.

Lemma 2.16.

Let ϕ\phi be a Hamiltonian diffeormorphism. Then, the bar-length spectrum

0<β1(ϕ,𝔽p)βK(ϕ,𝔽p)(ϕ,𝔽p)0<\beta_{1}(\phi,{\mathbb{F}}_{p})\leq\cdots\leq\beta_{K(\phi,{\mathbb{F}}_{p})}(\phi,{\mathbb{F}}_{p})

over 𝔽p{\mathbb{F}}_{p} coincides with the bar-length spectrum

0<β1(ϕ,)βK(ϕ,)(ϕ,)0<\beta_{1}(\phi,{\mathbb{Q}})\leq\cdots\leq\beta_{K(\phi,{\mathbb{Q}})}(\phi,{\mathbb{Q}})

over {\mathbb{Q}}. In particular β(ϕ,𝔽p)=β(ϕ,)\beta(\phi,{\mathbb{F}}_{p})=\beta(\phi,{\mathbb{Q}}) for sufficiently large pp.

Proof.

Let {ξ1,,ξB,η1,,ηK,ζ1,,ζK}\{\xi_{1},\dots,\xi_{B},\eta_{1},\dots,\eta_{K},\zeta_{1},\dots,\zeta_{K}\} be an orthonormal singular value decomposition of CF(H,Q(Λ,univ))\operatorname{\mathrm{CF}}(H,Q(\Lambda_{{\mathbb{Z}},univ})) satisfying dξi=0d\xi_{i}=0 for all i{0,,B}i\in\{0,\dots,B\} and dζj=Tβjηjd\zeta_{j}=T^{\beta_{j}}\eta_{j} for all j{1,,K}j\in\{1,\dots,K\}, where βj\beta_{j} is the jj-th bar-length in the spectrum. Note that there is a canonical orthonormal basis {x¯1,,x¯N}\{\overline{x}_{1},\dots,\overline{x}_{N}\} where, x¯i=T𝒜(xi)\overline{x}_{i}=T^{\mathcal{A}}(x_{i}) for all ii, and {xi}i=1N\{x_{i}\}_{i=1}^{N} is the collection of contractible fixed points of ϕ\phi. Recall from the discussion in Section 2.3.4 that these two orthonormal basis will be related by an matrix QGL(N,Q(Λ,univ))Q\in{\mathrm{GL}}(N,Q(\Lambda_{{\mathbb{Z}},univ})) whose coefficients have non-negative valuation, in particular, its filtration-preserving. Since QQ has only finitely many coefficients, Proposition 3.1 implies that for a sufficiently large prime pp, it is possible to reduce QQ to a matrix [Q]pGL(N,Λ𝔽p,univ)[Q]_{p}\in{\mathrm{GL}}(N,\Lambda_{{\mathbb{F}}_{p},univ}), i.e. [detQ]p0[\det Q]_{p}\neq 0. We can then obtain a singular value decomposition {[ξ1]p,,[ξB]p,[η1]p,,[ηK]p,[ζ1]p,,[ζK]p}\{[\xi_{1}]_{p},\dots,[\xi_{B}]_{p},[\eta_{1}]_{p},\dots,[\eta_{K}]_{p},[\zeta_{1}]_{p},\dots,[\zeta_{K}]_{p}\} of CF(H,Λ𝔽p,univ)\operatorname{\mathrm{CF}}(H,\Lambda_{{\mathbb{F}}_{p},univ}), satisfying the same relations as before, by applying [Q]p[Q]_{p} to the canonical orthonormal basis given by the contractible fixed points of ϕ\phi. In particular, it will have the same bar-length spectrum. On the other hand, {ξ11,,ξB1,η11,,ηK1,ζ11,,ζK1}\{\xi_{1}\otimes 1,\dots,\xi_{B}\otimes 1,\eta_{1}\otimes 1,\dots,\eta_{K}\otimes 1,\zeta_{1}\otimes 1,\dots,\zeta_{K}\otimes 1\} is an orthogonal singular value decomposition of CF(H,Λ,univ)\operatorname{\mathrm{CF}}(H,\Lambda_{{\mathbb{Q}},univ}) with the same bar-length spectrum. ∎

2.7. Local Floer homology

The definition of local Floer homology and its properties can be found in [GG10]. We include them in this section for the convenience of the readers.

Let xx be an isolated fixed point of a Hamiltonian diffeomorphism ϕ\phi and {ϕt}\{\phi_{t}\} be a Hamiltonian isotopy with ϕ1=ϕ\phi_{1}=\phi. Then x(t)=ϕt(x)x(t)=\phi_{t}(x) is an 11-periodic orbit. Denote by x~:S1S1×M\tilde{x}:S^{1}\to S^{1}\times M the graph of xx. Take U~\tilde{U} to be a small enough neighborhood of x~\tilde{x} and put U=pM(U~)U=p_{M}(\tilde{U}), where pM:S1×MMp_{M}:S^{1}\times M\to M is the projection. When xx is a degenerate fixed point, we can take a sufficiently small non-degenerate perturbation ϕ\phi^{\prime} of ϕ\phi with support in UU such that the Floer trajectories with sufficiently small energy connecting the 11-periodic orbits of ϕ\phi^{\prime} in UU are contained in UU. Thus, every broken trajectory is also contained in UU. Denote by CF(ϕ,x)\operatorname{\mathrm{CF}}(\phi^{\prime},x) the 𝕂{\mathbb{K}}-vector space generated by the 11-periodic orbits of ϕ\phi^{\prime} in UU. Then, we can define the Floer homology in UU, which is independent of the choice of the perturbation and of the almost complex structure. We call this Floer homology in UU the local Floer homology at xx and denote it by HFloc(ϕ,x)\operatorname{\mathrm{HF}}^{loc}(\phi,x). By the definition, one can easily see that HFloc(ϕ,x)𝕂\operatorname{\mathrm{HF}}^{loc}(\phi,x)\cong\mathbb{K} whenever xx is a non-degenerate fixed point. Furthermore, if (x,x¯1)(x,\overline{x}_{1}) and (x,x¯2)(x,\overline{x}_{2}) are two distinct cappings of a 11-periodic orbit xx. Then, CZ((x,x¯1))=CZ((x,x¯2))mod2\operatorname{\mathrm{CZ}}((x,\overline{x}_{1}))=\operatorname{\mathrm{CZ}}((x,\overline{x}_{2}))\mod 2. Thus, there is a /2\mathbb{Z}/2-grading on HFloc(ϕ,x)\operatorname{\mathrm{HF}}^{loc}(\phi,x). When xx is non-degenerate, HFloc(ϕ,x)𝕂Λ𝕂,univ\operatorname{\mathrm{HF}}^{loc}(\phi,x)\otimes_{\mathbb{K}}\Lambda_{\mathbb{K},univ} contributes a copy of Λ𝕂,univ\Lambda_{\mathbb{K},univ} to the Floer complex CFk(ϕ,Λ𝕂,univ)\operatorname{\mathrm{CF}}_{k}(\phi,\Lambda_{\mathbb{K},univ}). Following the arguments in [GG19, GG14, McL12, SZ21], for any two distinct capped periodic orbits x¯\overline{x} and y¯\overline{y} of ϕ\phi, there exists a crossing energy 2ϵ0>02\epsilon_{0}>0 such that all Floer trajectories, or product structures with x¯\overline{x} and y¯\overline{y} among their asymptotes carry energy at least 2ϵ02\epsilon_{0}.

Definition 2.17.

An iteration ϕk\phi^{k} of ϕ\phi is admissible at a fixed point xx of ϕ\phi if λk1\lambda^{k}\neq 1 for all eigenvalues λ1\lambda\neq 1 of dϕxd\phi_{x}.

For example, when none of λ1\lambda\neq 1 are roots of unity, ϕk\phi^{k} is admissible for k>0k>0. Otherwise, ϕpn\phi^{p^{n}} is admissible for sufficiently large pp and n>0n>0. By [GG10, Theorem 1.1, Remark 1.2 ], we have the following theorem.

Theorem 2.18.

Let ϕk\phi^{k} be an admissible iteration of ϕ\phi at an isolated fixed point xx. Then, the kk-iteration x(k)x^{(k)} of xx is an isolated fixed point of ϕk\phi^{k} and

HFloc(ϕk,xk)HFloc(ϕ,x).\operatorname{\mathrm{HF}}^{loc}(\phi^{k},x^{k})\cong\operatorname{\mathrm{HF}}^{loc}(\phi,x).
Remark 2.19.

Let ϕk1\phi^{k_{1}} and ϕk2\phi^{k_{2}} be admissible iterations of ϕ\phi at an isolated fixed point xx of ϕ\phi, then HFloc(ϕk1,xk1)HFloc(ϕk2,xk2)\operatorname{\mathrm{HF}}^{loc}(\phi^{k_{1}},x^{k_{1}})\cong\operatorname{\mathrm{HF}}^{loc}(\phi^{k_{2}},x^{k_{2}}) by Theorem 2.18.

The following theorem is a summary of the canonical Λ0\Lambda^{0}-complex constructed in [She22, Section 4.4.7] and upgraded to our setting in [Sug21] (see also [SW]), we only change the coefficients to fit our setting. Let ϕ\phi be a Hamiltonian diffeomorphism. For each isolated 11-periodic orbit xx of ϕ\phi, there is an isolating neighborhood UxU_{x} of xx. Let ϕ1\phi_{1} be a sufficiently small non-degenerate perturbation of ϕ\phi. Because there are finitely many isolated 11-periodic orbits, one can choose ϕ1=ϕ\phi_{1}=\phi outside of xFix(ϕ)Ux\bigcup_{x\in\mathrm{Fix}(\phi)}U_{x}.

Theorem 2.20.

Suppose ϕ\phi has finitely many fixed points. Then, there is a homotopically canonical Λ𝕂,univ0\Lambda_{\mathbb{K},univ}^{0}-complex denoted by CF(ϕ,Λ𝕂,univ0)\operatorname{\mathrm{CF}}(\phi,\Lambda_{\mathbb{K},univ}^{0}) that satisfies the following properties:

  1. (1)

    As a Λ𝕂,univ0\Lambda_{\mathbb{K},univ}^{0}-module,

    CF(ϕ,Λ𝕂,univ0)xFix(ϕ)HFloc(ϕ,x)𝕂Λ𝕂,univ0.\operatorname{\mathrm{CF}}(\phi,\Lambda_{\mathbb{K},univ}^{0})\cong\displaystyle\bigoplus_{x\in\mathrm{Fix}(\phi)}\operatorname{\mathrm{HF}}^{loc}(\phi,x)\otimes_{\mathbb{K}}\Lambda_{\mathbb{K},univ}^{0}.
  2. (2)

    Its differential is defined over Λ𝕂,univ0\Lambda_{\mathbb{K},univ}^{0}.

  3. (3)

    The homology of

    CF(ϕ,Λ𝕂,univ)=CF(ϕ,Λ𝕂,univ0)Λ𝕂,univ0Λ𝕂,univ\operatorname{\mathrm{CF}}(\phi,\Lambda_{\mathbb{K},univ})=\operatorname{\mathrm{CF}}(\phi,\Lambda_{\mathbb{K},univ}^{0})\otimes_{\Lambda_{\mathbb{K},univ}^{0}}\Lambda_{\mathbb{K},univ}

    is isomorphic to HF(ϕ1,Λ𝕂,univ)\operatorname{\mathrm{HF}}(\phi_{1},\Lambda_{\mathbb{K},univ}).

  4. (4)

    The bar-length spectrum associated to CF(ϕ,Λ𝕂,univ)\operatorname{\mathrm{CF}}(\phi,\Lambda_{\mathbb{K},univ}), denoted by

    β1(ϕ,Λ𝕂,univ)βK(ϕ,Λ𝕂,univ)(ϕ,Λ𝕂,univ),\beta^{\prime}_{1}(\phi,\Lambda_{\mathbb{K},univ})\leq\dots\leq\beta^{\prime}_{K(\phi,\Lambda_{\mathbb{K},univ})}(\phi,\Lambda_{\mathbb{K},univ}),

    satisfies β1(ϕ,Λ𝕂,univ)>ϵ0\beta^{\prime}_{1}(\phi,\Lambda_{\mathbb{K},univ})>\epsilon_{0} and is 2δ02\delta_{0}-close to the part

    βK+1(ϕ1,Λ𝕂,univ)βK+K(ϕ,Λ𝕂,univ)(ϕ1,Λ𝕂,univ)\beta_{K^{\prime}+1}(\phi_{1},\Lambda_{\mathbb{K},univ})\leq\dots\leq\beta_{K^{\prime}+K(\phi,\Lambda_{\mathbb{K},univ})}(\phi_{1},\Lambda_{\mathbb{K},univ})

    of the bar-length spectrum of ϕ1\phi_{1} above ϵ0\epsilon_{0} where βK(ϕ1,Λ𝕂,univ)<2δ0ϵ0\beta_{K^{\prime}}(\phi_{1},\Lambda_{\mathbb{K},univ})<2\delta_{0}\ll\epsilon_{0} and δ0ϵ0\delta_{0}\ll\epsilon_{0} is a small parameter converging to 0 as ϕ1\phi_{1} converges to ϕ\phi in the C2C^{2}-topology.

  5. (5)

    The bar-lengths βj(ϕ,Λ𝕂,univ)\beta^{\prime}_{j}(\phi,\Lambda_{\mathbb{K},univ}) for 1jK(ϕ,Λ𝕂,univ)1\leq j\leq K(\phi,\Lambda_{\mathbb{K},univ}) have a limit βj(ϕ,Λ𝕂,univ)\beta_{j}(\phi,\Lambda_{\mathbb{K},univ}) as the Hamiltonian perturbation tends to zero in the C2C^{2}-topology.

3. Semisimplicity of Quantum homology with different coefficients

3.1. Reduction modulo pp

The following proposition implies that for a sufficiently large prime pp, elements in the field of fractions Q(Λ,univ)Q(\Lambda_{{\mathbb{Z}},univ}) of Λ,univ\Lambda_{{\mathbb{Z}},univ} can be reduced modulo pp.

Proposition 3.1.

If fQ(Λ,univ)f\in Q(\Lambda_{\mathbb{Z},univ}), then it can be written as cjTμj\sum c_{j}T^{\mu_{j}}, where cjc_{j}\in\mathbb{Q} and only finitely many primes appear in the denominators of the coefficients of ff.

Proof.

Note that if h=i=NhiTνih=\sum_{i=-N}^{\infty}h_{i}T^{\nu_{i}} is an element in Λ,univ\Lambda_{\mathbb{Z},univ}, then TνNhT^{-\nu_{-N}}h only has non-negative exponents. Suppose fQ(Λ,univ)f\in Q(\Lambda_{\mathbb{Z},univ}), We may assume without loss of generality that

f=a0+i=1aiTλib0+j=1bjTθj,f=\frac{a_{0}+\displaystyle\sum_{i=1}^{\infty}a_{i}T^{\lambda_{i}}}{b_{0}+\displaystyle\sum_{j=1}^{\infty}b_{j}T^{\theta_{j}}},

where 0<λ1<λ2<0<\lambda_{1}<\lambda_{2}<\cdots and 0<θ1<θ2<0<\theta_{1}<\theta_{2}<\cdots. Set

A:=a0+i=1aiTλiandB:=b0+j=1bjTθj.A:=a_{0}+\displaystyle\sum_{i=1}^{\infty}a_{i}T^{\lambda_{i}}\quad\text{and}\quad B:=b_{0}+\displaystyle\sum_{j=1}^{\infty}b_{j}T^{\theta_{j}}.

and let c0=a0b0c_{0}=\frac{a_{0}}{b_{0}} and g0=a0b0g_{0}=\frac{a_{0}}{b_{0}}. Then

Ag0B=i=1aiTλij=1a0bjb0Tθj.A-g_{0}B=\displaystyle\sum_{i=1}^{\infty}a_{i}T^{\lambda_{i}}-\displaystyle\sum_{j=1}^{\infty}\frac{a_{0}b_{j}}{b_{0}}T^{\theta_{j}}.

The leading term has a non-negative exponent. Consider

c1Tμ1B=c1b0Tμ1+j=1c1bjTθj+μ1c_{1}T^{\mu_{1}}B=c_{1}b_{0}T^{\mu_{1}}+\displaystyle\sum_{j=1}^{\infty}c_{1}b_{j}T^{\theta_{j}+\mu_{1}}

where μ1>0\mu_{1}>0. Then, leading term is c1b0Tμ1c_{1}b_{0}T^{\mu_{1}}. There are a few cases to be considered:

Case 1. λ1<θ1\lambda_{1}<\theta_{1} In this case, the leading term in Ag0BA-g_{0}B is a1Tλ1a_{1}T^{\lambda_{1}}. Let c1Tμ1=a1b0Tλ1c_{1}T^{\mu_{1}}=\frac{a_{1}}{b_{0}}T^{\lambda_{1}}. Then, a1Tλ1=c1b0Tμ1a_{1}T^{\lambda_{1}}=c_{1}b_{0}T^{\mu_{1}}. Define g1=a0b0+a1b0Tλ1g_{1}=\frac{a_{0}}{b_{0}}+\frac{a_{1}}{b_{0}}T^{\lambda_{1}}. Then

Ag1B=i=2aiTλij=1a0bjb0Tθjj=1a1bjb0Tθj+λ1,A-g_{1}B=\displaystyle\sum_{i=2}^{\infty}a_{i}T^{\lambda_{i}}-\sum_{j=1}^{\infty}\frac{a_{0}b_{j}}{b_{0}}T^{\theta_{j}}-\sum_{j=1}^{\infty}\frac{a_{1}b_{j}}{b_{0}}T^{\theta_{j}+\lambda_{1}},

which has the exponent of the leading term greater than λ1\lambda_{1}.

Case 2. λ1=θ1\lambda_{1}=\theta_{1} The leading term in Ag0BA-g_{0}B is given by (a1a0b1b0)Tλ1(a_{1}-\frac{a_{0}b_{1}}{b_{0}})T^{\lambda_{1}}. Let c1Tμ1=(a1b0a0b1b02)Tλ1c_{1}T^{\mu_{1}}=(\frac{a_{1}}{b_{0}}-\frac{a_{0}b_{1}}{b_{0}^{2}})T^{\lambda_{1}}. Then, (a1a0b1b0)Tλ1=c1b0Tμ1(a_{1}-\frac{a_{0}b_{1}}{b_{0}})T^{\lambda_{1}}=c_{1}b_{0}T^{\mu_{1}}. Define g1=a0b0+(a1b0a0b1b02)Tλ1g_{1}=\frac{a_{0}}{b_{0}}+(\frac{a_{1}}{b_{0}}-\frac{a_{0}b_{1}}{b_{0}^{2}})T^{\lambda_{1}}. Then

Ag1B=i=2aiTλij=2a0bjb0Tθjj=1(a1b0a0b1b02)bjTλ1+θj,A-g_{1}B=\displaystyle\sum_{i=2}^{\infty}a_{i}T^{\lambda_{i}}-\sum_{j=2}^{\infty}\frac{a_{0}b_{j}}{b_{0}}T^{\theta_{j}}-\sum_{j=1}^{\infty}(\frac{a_{1}}{b_{0}}-\frac{a_{0}b_{1}}{b_{0}^{2}})b_{j}T^{\lambda_{1}+\theta_{j}},

which has the exponent of the leading term greater than λ1\lambda_{1}.

Case 3. λ1>θ1\lambda_{1}>\theta_{1} The leading term in Ag0BA-g_{0}B is a0b1b0Tθ1-\frac{a_{0}b_{1}}{b_{0}}T^{\theta_{1}}. Let c1Tμ1=a0b1b02Tθ1c_{1}T^{\mu_{1}}=-\frac{a_{0}b_{1}}{b_{0}^{2}}T^{\theta_{1}}. Then, a0b1b0Tθ1=c1b0Tμ1-\frac{a_{0}b_{1}}{b_{0}}T^{\theta_{1}}=c_{1}b_{0}T^{\mu_{1}}. Define g1=a0b0a0b1b02Tθ1g_{1}=\frac{a_{0}}{b_{0}}-\frac{a_{0}b_{1}}{b_{0}^{2}}T^{\theta_{1}}. Then,

Ag1B=i=1aiTλij=2a0bjb0Tθj+j=1a0b1bjb02Tθ1+θj,A-g_{1}B=\displaystyle\sum_{i=1}^{\infty}a_{i}T^{\lambda_{i}}-\sum_{j=2}^{\infty}\frac{a_{0}b_{j}}{b_{0}}T^{\theta_{j}}+\sum_{j=1}^{\infty}\frac{a_{0}b_{1}b_{j}}{b_{0}^{2}}T^{\theta_{1}+\theta_{j}},

which has the exponent of the leading term greater than θ1\theta_{1}.

One can repeat the process to get gng_{n} for nn\in\mathbb{N}. Then f=limgnf=\lim g_{n}. Hence, the primes appearing in the denominators of the coefficients of ff are the primes dividing b0b_{0}, of which there are finitely many. ∎

Remark 3.2.

Note that reduction of coefficients is not possible for arbitrary finite sets {f1,,fk}Λ,univ\{f_{1},\cdots,f_{k}\}\subset\Lambda_{{\mathbb{Q}},univ}. Indeed, you can have elements of the form

l=K1l!Tl,\sum_{l=-K}^{\infty}\frac{1}{l!}T^{l},

which has infinitely many primes in the denominators. In particular, one cannot reduce the coefficients. The previous proposition, shows that this does not happen in Q(Λ,univ)Q(\Lambda_{{\mathbb{Z}},univ}).

3.2. Semisimplicity and idempotents

The following corollary is a particular case of [EP08, Proposition 2.1(A)].

Corollary 3.3.

Let 𝕂\mathbb{K} be a field of characteristic 0. If QHev(M,Λ𝕂,univ)\operatorname{\mathrm{QH}}_{ev}(M,\Lambda_{\mathbb{K},univ}) is semisimple, then QHev(M,Q(Λ,univ))\operatorname{\mathrm{QH}}_{ev}(M,Q(\Lambda_{\mathbb{Z},univ})) and QHev(M,Q(Λ,univ)¯)\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})}) are semisimple.

Under the assumption that QHev(M,Q(Λ,univ)¯)\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})}) is semisimple, let {e¯i}i=1m\{\overline{e}_{i}\}_{i=1}^{m} be a collection of idempotents such that

QHev(M,Q(Λ,univ)¯)=i=1me¯iQHev(M,Q(Λ,univ)¯).\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})})=\displaystyle\bigoplus_{i=1}^{m}\overline{e}_{i}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})}).

Then, each e¯i\overline{e}_{i} is of the form

e¯i=j=0nk¯ijhj\overline{e}_{i}=\displaystyle\sum_{j=0}^{n}\overline{k}_{ij}h_{j}

where k¯ijQ(Λ,univ)¯\overline{k}_{ij}\in\overline{Q(\Lambda_{\mathbb{Z},univ})} and hjH2j(M)h_{j}\in\operatorname{\mathrm{H}}_{2j}(M). Let pijQ(Λ,univ)[x]p_{ij}\in Q(\Lambda_{\mathbb{Z},univ})[x] be the minimial polynomial of k¯ij\overline{k}_{ij} and let {αijl}\{\alpha_{ij}^{l}\} be all of its roots. Since

Q(Λ,univ)({αijl}i,j,l)Q(\Lambda_{\mathbb{Z},univ})(\{\alpha_{ij}^{l}\}_{i,j,l})

is a finite separable extension, the primitive element theorem implies there is an element αQ(Λ,univ)¯\alpha\in\overline{Q(\Lambda_{\mathbb{Z},univ})} such that

Q(Λ,univ)({αijl}i,j,l)=Q(Λ,univ)(α).Q(\Lambda_{\mathbb{Z},univ})(\{\alpha_{ij}^{l}\}_{i,j,l})=Q(\Lambda_{\mathbb{Z},univ})(\alpha).

In particular, k¯ijQ(Λ,univ)(α)\overline{k}_{ij}\in Q(\Lambda_{\mathbb{Z},univ})(\alpha) for all i=1,,mi=1,\cdots,m and j=1,,nj=1,\cdots,n. Let ff be the minimal polynomial of α\alpha. Denote

f=arxr+ar1xr1+a1x+a0f=a_{r}x^{r}+a_{r-1}x^{r-1}+\cdots a_{1}x+a_{0}

where aiQ(Λ,univ)a_{i}\in Q(\Lambda_{\mathbb{Z},univ}). Assume ai=aiai′′a_{i}=\frac{a^{\prime}_{i}}{a^{\prime\prime}_{i}} where ai,ai′′Λ,univa^{\prime}_{i},a^{\prime\prime}_{i}\in\Lambda_{\mathbb{Z},univ}. Then

i=1rai′′fΛ,univ[x],\displaystyle\prod_{i=1}^{r}a^{\prime\prime}_{i}\cdot f\in\Lambda_{\mathbb{Z},univ}[x],

which we still denote by ff. Finally we have

Q(Λ,univ)(α)Q(Λ,univ)[x](f).Q(\Lambda_{\mathbb{Z},univ})(\alpha)\cong\frac{Q(\Lambda_{\mathbb{Z},univ})[x]}{(f)}.

By reducing the coefficients of ff to Λ𝔽p,univ\Lambda_{{\mathbb{F}}_{p},univ}, we get

[f]p=g1m1g2m2gsms[f]_{p}=g_{1}^{m_{1}}g_{2}^{m_{2}}\cdots g_{s}^{m_{s}}

in Λ𝔽p,univ[x]\Lambda_{{\mathbb{F}}_{p},univ}[x], where gig_{i} are irreducible and distinct from each other.

Claim 3.4.

m1=m2==ms=1m_{1}=m_{2}=\cdots=m_{s}=1 for sufficiently large pp.

Proof.

Since ff is an irreducible polynomial over Q(Λ,univ)[x]Q(\Lambda_{\mathbb{Z},univ})[x] we have gcd(f,f)=1{\mathrm{gcd}}(f,f^{\prime})=1. So,

rf+qf=1rf+qf^{\prime}=1

for some r,qQ(Λ,univ)[x]r,q\in Q(\Lambda_{\mathbb{Z},univ})[x]. Let ΘΛ,univ\Theta\in\Lambda_{\mathbb{Z},univ} be the product of the denominators of the coefficient of rr and qq. Then

(3.1) r~f+q~f=Θ\widetilde{r}f+\widetilde{q}f^{\prime}=\Theta

where r~=rΘ\widetilde{r}=r\cdot\Theta and q~=qΘ\widetilde{q}=q\cdot\Theta. Write

Θ=θsTλs+θs+1Tλs+1+.\Theta=\theta_{-s}T^{\lambda_{-s}}+\theta_{-s+1}T^{\lambda_{-s+1}}+\cdots.

Then, for all primes p>|θs|p>|\theta_{-s}|, Θ\Theta reduces to a nonzero element [Θ]p[\Theta]_{p} in Λ𝔽p,univ\Lambda_{{\mathbb{F}}_{p},univ}. Now, by reducing Equation (3.1) for sufficiently large primes pp, we have

[r~]p[f]p+[q~]p[f]p=[Θ]p[\widetilde{r}]_{p}[f]_{p}+[\widetilde{q}]_{p}[f^{\prime}]_{p}=[\Theta]_{p}

Thus, gcd([f]p,[f]p)=1{\mathrm{gcd}}([f]_{p},[f^{\prime}]_{p})=1 in Λ𝔽p,univ[x]\Lambda_{{\mathbb{F}}_{p},univ}[x], hence, m1=m2==ms=1m_{1}=m_{2}=\cdots=m_{s}=1. ∎

Thus, for a sufficiently large prime pp, [f]p=g1g2gs[f]_{p}=g_{1}g_{2}\cdots g_{s} where gig_{i} are irreducible and distinct from each other.

3.3. Reduction of idempotents

We now detail the process of reducing the idempotents e¯i\overline{e}_{i} to elements [e¯i]p[\overline{e}_{i}]_{p} in QHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}), where Λ¯𝔽p\overline{\Lambda}_{{\mathbb{F}}_{p}} is the algebraic closure of Λ𝔽p\Lambda_{{\mathbb{F}}_{p}}. Recall that we can write e¯i=k¯ijhj\overline{e}_{i}=\sum\overline{k}_{ij}h_{j} where k¯ijQ(Λ,univ)(α)\overline{k}_{ij}\in Q(\Lambda_{\mathbb{Z},univ})(\alpha) is non-zero and that

Q(Λ,univ)(α)Q(Λ,univ)[x](f).Q(\Lambda_{\mathbb{Z},univ})(\alpha)\cong\frac{Q(\Lambda_{\mathbb{Z},univ})[x]}{(f)}.

Therefore, we can write k¯ij\overline{k}_{ij} as Kij+(f)K_{ij}+(f) for KijQ(Λ,univ)[x]K_{ij}\in Q(\Lambda_{{\mathbb{Z}},univ})[x]. Since k¯ij\overline{k}_{ij} is invertible, there is an element LijQ(Λ,univ)[x]L_{ij}\in Q(\Lambda_{\mathbb{Z},univ})[x] such that KijLij+(f)=1+(f)K_{ij}L_{ij}+(f)=1+(f), in particular, KijLij=1+MijfK_{ij}L_{ij}=1+M_{ij}f for MijQ(Λ,univ)[x]M_{ij}\in Q(\Lambda_{{\mathbb{Z}},univ})[x]. Let ΥK\Upsilon_{K}, ΥL\Upsilon_{L} and ΥM\Upsilon_{M} be the product of the denominators of the coefficients of KijK_{ij}, LijL_{ij} and MijM_{ij}, respectively. Then

(3.2) K~ijL~ijΥM=ΥKΥLΥM+M~ijfΥKΥL\widetilde{K}_{ij}\widetilde{L}_{ij}\Upsilon_{M}=\Upsilon_{K}\Upsilon_{L}\Upsilon_{M}+\widetilde{M}_{ij}f\Upsilon_{K}\Upsilon_{L}

where

K~ij=KijΥK,L~ij=LijΥL,M~ij=MijΥM\widetilde{K}_{ij}=K_{ij}\cdot\Upsilon_{K},\quad\widetilde{L}_{ij}=L_{ij}\cdot\Upsilon_{L},\quad\widetilde{M}_{ij}=M_{ij}\cdot\Upsilon_{M}

all belong to Λ,univ[x]\Lambda_{{\mathbb{Z}},univ}[x]. For sufficiently large prime pp, [ΥK]p0[\Upsilon_{K}]_{p}\neq 0, [ΥL]p0[\Upsilon_{L}]_{p}\neq 0, [ΥM]p0[\Upsilon_{M}]_{p}\neq 0 and [f]p0[f]_{p}\neq 0. By reducing the Equation (3.2), we have

(3.3) [K~ij]p[L~ij]p[ΥM]p=[ΥK]p[ΥL]p[ΥM]p+[M~ij]p[f]p[ΥK]p[ΥL]p[\widetilde{K}_{ij}]_{p}[\widetilde{L}_{ij}]_{p}[\Upsilon_{M}]_{p}=[\Upsilon_{K}]_{p}[\Upsilon_{L}]_{p}[\Upsilon_{M}]_{p}+[\widetilde{M}_{ij}]_{p}[f]_{p}[\Upsilon_{K}]_{p}[\Upsilon_{L}]_{p}

Thus,

[K~ij]p[L~ij]p0inΛ𝔽p,univ[x]([f]p).[\widetilde{K}_{ij}]_{p}[\widetilde{L}_{ij}]_{p}\neq 0\quad\text{in}\quad\frac{\Lambda_{{\mathbb{F}}_{p},univ}[x]}{([f]_{p})}.

In particular, [K~ij]p0[\widetilde{K}_{ij}]_{p}\neq 0. Thus, one can reduce Kij+(f)K_{ij}+(f) to

[K~ij]p[ΥK]p1+([f]p)Λ𝔽p,univ[x]([f]p).[\widetilde{K}_{ij}]_{p}[\Upsilon_{K}]_{p}^{-1}+([f]_{p})\in\frac{\Lambda_{{\mathbb{F}}_{p},univ}[x]}{([f]_{p})}.

Note that

Λ𝔽p,univ[x]([f]p)=Λ𝔽p,univ[x]g1g2gsl=1sΛ𝔽p,univ[x](gl)\frac{\Lambda_{{\mathbb{F}}_{p},univ}[x]}{([f]_{p})}=\frac{\Lambda_{{\mathbb{F}}_{p},univ}[x]}{{\left<g_{1}g_{2}\cdots g_{s}\right>}}\cong\displaystyle\prod_{l=1}^{s}\frac{\Lambda_{{\mathbb{F}}_{p},univ}[x]}{(g_{l})}

and each term in the product is a field because gig_{i} is irreducible. Let

Pl:Λ𝔽p,univ[x]([f]p)Λ𝔽p,univ[x](gl)P_{l}:\frac{\Lambda_{{\mathbb{F}}_{p},univ}[x]}{([f]_{p})}\longrightarrow\frac{\Lambda_{{\mathbb{F}}_{p},univ}[x]}{(g_{l})}

be the projection. Then, there exists at least one PlP_{l} such that

Pl([K~ij]p[ΥK]p1+([f]p))0.P_{l}([\widetilde{K}_{ij}]_{p}[\Upsilon_{K}]_{p}^{-1}+([f]_{p}))\neq 0.

Furthermore, let

ιl:Λ𝔽p,univ[x](gl)Λ¯𝔽p,univ\iota_{l}:\frac{\Lambda_{{\mathbb{F}}_{p},univ}[x]}{(g_{l})}\longrightarrow\overline{\Lambda}_{{\mathbb{F}}_{p},univ}

be the inclusion, then ιl(Pl([K~ij]p[ΥK]p1+([f]p)))0.\iota_{l}(P_{l}([\widetilde{K}_{ij}]_{p}[\Upsilon_{K}]_{p}^{-1}+([f]_{p})))\neq 0. Thus, we have reduced k¯ij\overline{k}_{ij} to an nonzero element in Λ¯𝔽p,univ\overline{\Lambda}_{{\mathbb{F}}_{p},univ}. In particular, e¯i\overline{e}_{i} can be reduced to a nonzero element in QHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}). Finally, note that each e¯i\overline{e}_{i} is represented by a finite linear combination of elements in H(M)\operatorname{\mathrm{H}}_{*}(M) and there are finitely many e¯i\overline{e}_{i}, thus, one can choose a sufficiently large pp to assure that k¯ij\overline{k}_{ij} can be reduced to a nonzero element in Λ¯𝔽p,univ\overline{\Lambda}_{{\mathbb{F}}_{p},univ} for any i,ji,j.

Remark 3.5.

In fact, for any l=1,,sl=1,\cdots,s and sufficiently large pp,

Pl([K~ij]p[ΥK]p1+([f]p))0.P_{l}([\widetilde{K}_{ij}]_{p}[\Upsilon_{K}]_{p}^{-1}+([f]_{p}))\neq 0.

Indeed, suppose otherwise then, there exists δΛ,univ[x]\delta\in\Lambda_{\mathbb{Z},univ}[x] such that

[K~ij]p[ΥK]p1=glδ.[\widetilde{K}_{ij}]_{p}[\Upsilon_{K}]_{p}^{-1}=g_{l}\delta.

Now, we can rewrite the Equation (3.3) as:

glδ[ΥK]p[L~ij]p[ΥM]p=[ΥK]p[ΥL][ΥM]p+[M~ij]p[f]p[ΥK]p[ΥL]pg_{l}\delta[\Upsilon_{K}]_{p}[\widetilde{L}_{ij}]_{p}[\Upsilon_{M}]_{p}=[\Upsilon_{K}]_{p}[\Upsilon_{L}][\Upsilon_{M}]_{p}+[\widetilde{M}_{ij}]_{p}[f]_{p}[\Upsilon_{K}]_{p}[\Upsilon_{L}]_{p}

Let η\eta be a root of glg_{l}. By plugging η\eta into the equation, we get that the product [ΥK]p[ΥL]p[ΥM]p=0[\Upsilon_{K}]_{p}[\Upsilon_{L}]_{p}[\Upsilon_{M}]_{p}=0, which contradicts our choice of pp.

Proposition 3.6.

The reductions [e¯i]p[\overline{e}_{i}]_{p} are idempotents in QHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) for all i=1,,mi=1,\cdots,m. In addition, [e¯i]p[e¯j]p=0[\overline{e}_{i}]_{p}*[\overline{e}_{j}]_{p}=0 for iji\neq j, and i=1m[e¯i]p=1\sum_{i=1}^{m}[\overline{e}_{i}]_{p}=1 where 11 is the multiplicative identity in QHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}).

Proof.

Since QHev(M,Q(Λ,univ)¯)\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})}) is semisimple and {e¯i}i=1m\{\overline{e}_{i}\}_{i=1}^{m} are idempotents such that

QHev(M,Q(Λ,univ)¯)=i=1me¯iQHev(M,Q(Λ,univ)¯),\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})})=\displaystyle\bigoplus_{i=1}^{m}\overline{e}_{i}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})}),

we have the following equations:

e¯ie¯i=e¯i\overline{e}_{i}*\overline{e}_{i}=\overline{e}_{i}

for i=1,,mi=1,\cdots,m,

e¯ie¯j=0,ij\overline{e}_{i}*\overline{e}_{j}=0,i\neq j

for i,j=1,,mi,j=1,\cdots,m,

i=1me¯i=1\displaystyle\sum_{i=1}^{m}\overline{e}_{i}=1

One can reduce the three equations to QHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}). Thus,

QHev(M,Λ¯𝔽p,univ)=i=1m[e¯i]pQHev(M,Λ¯𝔽p,univ).\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})=\displaystyle\bigoplus_{i=1}^{m}[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}).

Remark 3.7.

Note that each e¯iQHev(M,Q(Λ,univ)¯)\overline{e}_{i}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})}) is an algebraic field extension of Q(Λ,univ)¯\overline{Q(\Lambda_{\mathbb{Z},univ})}, which in turn, is the algebraic closure of Q(Λ,univ)Q(\Lambda_{\mathbb{Z},univ}). Thus,

e¯iQHev(M,Q(Λ,univ)¯)Q(Λ,univ)¯.\overline{e}_{i}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})})\cong\overline{Q(\Lambda_{\mathbb{Z},univ})}.

Also,

QHev(M,Q(Λ,univ)¯)=i=1me¯iQHev(M,Q(Λ,univ)¯)i=1mQ(Λ,univ)¯.\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})})=\displaystyle\bigoplus_{i=1}^{m}\overline{e}_{i}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})})\cong\bigoplus_{i=1}^{m}\overline{Q(\Lambda_{\mathbb{Z},univ})}.

Thus m=rank(Hev(M))m=\operatorname{\mathrm{rank}}(\operatorname{\mathrm{H}}_{ev}(M)). Since QHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) is a free Λ¯𝔽p,univ\overline{\Lambda}_{{\mathbb{F}}_{p},univ}-module and [e¯i]pQHev(M,Λ¯𝔽p,univ)[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) is a submodule, [e¯i]pQHev(M,Λ¯𝔽p,univ)[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) can be written as a direct sum of copies of Λ¯𝔽p,univ\overline{\Lambda}_{{\mathbb{F}}_{p},univ}. The facts that m=rank(Hev(M,))=rank(Hev(M,𝔽p))m=\operatorname{\mathrm{rank}}(\operatorname{\mathrm{H}}_{ev}(M,\mathbb{Q}))=\operatorname{\mathrm{rank}}(\operatorname{\mathrm{H}}_{ev}(M,{\mathbb{F}}_{p})) for sufficiently large pp, [e¯i]p[\overline{e}_{i}]_{p} is nonzero and

[e¯i]pQHev(M,Λ¯𝔽p,univ)[e¯j]pQHev(M,Λ¯𝔽p,univ)=0[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\cap[\overline{e}_{j}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})=0

for iji\neq j, imply that

[e¯i]pQHev(M,Λ¯𝔽p,univ)Λ¯𝔽p,univ.[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\cong\overline{\Lambda}_{{\mathbb{F}}_{p},univ}.

Thus QHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) is semisimple and is generated by the idempotents {[e¯i]p}i=1m\{[\overline{e}_{i}]_{p}\}_{i=1}^{m}.

4. Upper bound on the boundary depth

Recall from Section 3.2 that ff is the minimal polynomial of α\alpha and, reducing to Λ𝔽p,univ[x]\Lambda_{{\mathbb{F}}_{p},univ}[x], we obtain [f]p=g1g2gs[f]_{p}=g_{1}g_{2}\cdots g_{s} where gig_{i} are irreducible and distinct from each other for sufficiently large pp. Therefore,

|[f(0)]p|=i=1s|gi(0)|.|[f(0)]_{p}|=\displaystyle\prod_{i=1}^{s}|g_{i}(0)|.

Thus, for some ii, we have |gi(0)||[f(0)]p|1/s|g_{i}(0)|\leq|[f(0)]_{p}|^{1/s}. Without loss of generality, assume that |g1(0)||[f(0)]p|1/s|g_{1}(0)|\leq|[f(0)]_{p}|^{1/s}. There is an algebraic element γ\gamma over Λ𝔽p,univ\Lambda_{{\mathbb{F}}_{p},univ} such that

Λ𝔽p,univ[x](g1)=Λ𝔽p,univ(γ).\frac{\Lambda_{{\mathbb{F}}_{p},univ}[x]}{(g_{1})}=\Lambda_{{\mathbb{F}}_{p},univ}(\gamma).

Since Λ𝔽p,univ\Lambda_{{\mathbb{F}}_{p},univ} is complete, by Remark 2.5, one can get a norm |||\cdot| and, hence, a valuation ln(||)-\ln(|\cdot|) on Λ𝔽p,univ(γ)\Lambda_{{\mathbb{F}}_{p},univ}(\gamma).

Lemma 4.1.

Let bQ(Λ,univ)b\in Q(\Lambda_{\mathbb{Z},univ}). Then |b||[b]p||b|\geq|[b]_{p}| for sufficiently large pp.

Proof.

Suppose b=i=KbiTλib=\sum_{i=-K}^{\infty}b_{i}T^{\lambda_{i}} with λK<λK+1<\lambda_{-K}<\lambda_{-K+1}<\cdots. Then ν(b)=λK\nu(b)=\lambda_{-K}. Write bK=bK,0bK,1b_{-K}=\frac{b_{-K,0}}{b_{-K,1}}. For a sufficiently large prime pp, we have that [bK]p[b_{-K}]_{p} is invertible, thus, [bK]p=[bK,0]p[bK,1]p1[b_{-K}]_{p}=[b_{-K,0}]_{p}[b_{-K,1}]_{p}^{-1}. If [bK]p=0[b_{-K}]_{p}=0, then ν(b)<ν([b]p)\nu(b)<\nu([b]_{p}). If [bK]p0[b_{-K}]_{p}\neq 0, then ν(b)=ν([b]p)\nu(b)=\nu([b]_{p}). Since ||=eν()|\cdot|=e^{-\nu(\cdot)}, then |b||[b]p||b|\geq|[b]_{p}|. ∎

Remark 4.2.

The prime pp in the above lemma depends on bb and there is not a uniform pp for all elements in Q(Λ,univ)Q(\Lambda_{\mathbb{Z},univ}).

Proposition 4.3.

|γ|max{1,|f(0)|}|\gamma|\leq\max\{1,|f(0)|\}.

Proof.

One can see that g1g_{1} is both the characteristic polynomial and the minimal polynomial of γ\gamma. Thus NΛ𝔽p,univ(γ)/Λ𝔽p,univ(γ)=(1)Mg1(0)N_{\Lambda_{{\mathbb{F}}_{p},univ}(\gamma)/\Lambda_{{\mathbb{F}}_{p},univ}}(\gamma)=(-1)^{M}g_{1}(0) where MM is the degree of g1g_{1}. Then |γ|=|NΛ𝔽p,univ(γ)/Λ𝔽p,univ(γ)|1/M=|g1(0)|1/M|\gamma|=|N_{\Lambda_{{\mathbb{F}}_{p},univ}(\gamma)/\Lambda_{{\mathbb{F}}_{p},univ}}(\gamma)|^{1/M}=|g_{1}(0)|^{1/M}. Furthermore |γ||f(0)|1sM|\gamma|\leq|f(0)|^{\frac{1}{sM}}. Let NN be the degree of ff. If |f(0)|1|f(0)|\leq 1, then |f(0)|1sM1|f(0)|^{\frac{1}{sM}}\leq 1. If |f(0)|>1|f(0)|>1, |f(0)|1sM|f(0)||f(0)|^{\frac{1}{sM}}\leq|f(0)|. Hence, |γ|max{1,|f(0)|}|\gamma|\leq\max\{1,|f(0)|\} as required. ∎

Recall that {e¯i=j=0nk¯ijhj}i=1m\{\overline{e}_{i}=\sum_{j=0}^{n}\overline{k}_{ij}h_{j}\}_{i=1}^{m} and {[e¯i]p=j=0n[k¯ij]phj}i=1m\{[\overline{e}_{i}]_{p}=\sum_{j=0}^{n}[\overline{k}_{ij}]_{p}h_{j}\}_{i=1}^{m} are the idempotents of QHev(M,Q(Λ,univ)¯)\operatorname{\mathrm{QH}}_{ev}(M,\overline{Q(\Lambda_{\mathbb{Z},univ})}) and QHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}), respectively, where k¯ijQ(Λ,univ)(α)\overline{k}_{ij}\in Q(\Lambda_{\mathbb{Z},univ})(\alpha) and [k¯ij]pΛ𝔽p,univ(γ)[\overline{k}_{ij}]_{p}\in\Lambda_{{\mathbb{F}}_{p},univ}(\gamma). Suppose k¯ij=s=0Nbijsαs\overline{k}_{ij}=\sum_{s=0}^{N}b_{ijs}\alpha^{s}. Then, [k¯ij]p=s=0N[bijs]pγs[\overline{k}_{ij}]_{p}=\sum_{s=0}^{N}[b_{ijs}]_{p}\gamma^{s}. Denote by Ξ=maxi,j,s{|bijs|}\Xi=\displaystyle\max_{i,j,s}\{|b_{ijs}|\}. Thus,

|[k¯ij]p|max0sN{|[bijs]pγs|}=max0sN{|[bijs]p||γs|}Ξ(max{1,|f(0)|})N.|[\overline{k}_{ij}]_{p}|\leq\displaystyle\max_{0\leq s\leq N}\{|[b_{ijs}]_{p}\gamma^{s}|\}=\displaystyle\max_{0\leq s\leq N}\{|[b_{ijs}]_{p}||\gamma^{s}|\}\leq\Xi(\max\{1,|f(0)|\})^{N}.

Thus,

l([e¯i]p)=max0jn{ν([k¯ij]p)}ln(Ξ(max{1,f(0)})N).l([\overline{e}_{i}]_{p})=\displaystyle\max_{0\leq j\leq n}\{-\nu([\overline{k}_{ij}]_{p})\}\leq\ln(\Xi(\max\{1,f(0)\})^{N}).

In particular, we have the following proposition.

Proposition 4.4.

There is a number δ\delta, independent of pp, such that l([e¯i]p)δl([\overline{e}_{i}]_{p})\leq\delta for all i=1,,mi=1,\cdots,m and all sufficiently large pp.

Definition 4.5.

Suppose QHev(M,𝕂)\operatorname{\mathrm{QH}}_{ev}(M,\mathbb{K}) is semisimple and E={e1,,em}E=\{e_{1},\cdots,e_{m}\} are idempotents. Then, define

γej(H,𝕂)\displaystyle\gamma_{e_{j}}(H,\mathbb{K}) =c(ej,H,𝕂)+c(ej,H¯,𝕂)\displaystyle=c(e_{j},H,\mathbb{K})+c(e_{j},\overline{H},\mathbb{K})
γej(ϕ,𝕂)\displaystyle\gamma_{e_{j}}(\phi,\mathbb{K}) =infϕH1=ϕγej(H,𝕂)\displaystyle=\displaystyle\inf_{\phi_{H}^{1}=\phi}\gamma_{e_{j}}(H,\mathbb{K})

and

γE(ϕ,𝕂)=max1imγej(ϕ,𝕂).\displaystyle\gamma_{E}(\phi,\mathbb{K})=\displaystyle\max_{1\leq i\leq m}\gamma_{e_{j}}(\phi,\mathbb{K}).
Lemma 4.6.

Let θ\theta be a class in QH(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{*}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}). Then, we have c(θ,id,Λ¯𝔽p,univ)0c(\theta,id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\geq 0 if ν(Δ(PSSH(θ),PSSH¯([M]))0\nu(\Delta(\operatorname{\mathrm{PSS}}_{H}(\theta),\operatorname{\mathrm{PSS}}_{\overline{H}}([M]))\leq 0.

Proof.

Suppose that θ=θihi\theta=\sum\theta_{i}h_{i} with θiΛ¯𝔽p,univ\theta_{i}\in\overline{\Lambda}_{{\mathbb{F}}_{p},univ} and hiH(M,𝔽p)h_{i}\in\operatorname{\mathrm{H}}_{*}(M,{\mathbb{F}}_{p}). Then

l(θ)=max{ν(θi)}=min{ν(θi)}.l(\theta)=\max\{-\nu(\theta_{i})\}=-\min\{\nu(\theta_{i})\}.

Observe that Δ(PSSH(θ),PSSH¯([M]))=θj\Delta(\operatorname{\mathrm{PSS}}_{H}(\theta),\operatorname{\mathrm{PSS}}_{\overline{H}}([M]))=\theta_{j} with hj=[pt]h_{j}=[pt]. Thus,

ν(Δ(PSSH(θ),PSSH¯([M])))=ν(θj)l(θ)=c(θ,id,Λ¯𝔽p,univ).\nu(\Delta(\operatorname{\mathrm{PSS}}_{H}(\theta),\operatorname{\mathrm{PSS}}_{\overline{H}}([M])))=\nu(\theta_{j})\geq-l(\theta)=-c(\theta,id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}).

Since ν(Δ(PSSH(θ),PSSH¯([M])))0\nu(\Delta(\operatorname{\mathrm{PSS}}_{H}(\theta),\operatorname{\mathrm{PSS}}_{\overline{H}}([M])))\leq 0, we have that c(θ,id,Λ¯𝔽p,univ)0c(\theta,id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\geq 0. ∎

Lemma 4.7.

Let θQHev(M,Λ¯𝔽p,univ)\theta\in\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) and [e¯i]pθ0[\overline{e}_{i}]_{p}*\theta\neq 0. Then we have that

c(([e¯i]pθ)1,id,Λ¯𝔽p,univ)+c([e¯i]pθ,id,Λ¯𝔽p,univ)=2c([e¯i]p,id,Λ¯𝔽p,univ),c(([\overline{e}_{i}]_{p}*\theta)^{-1},id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})+c([\overline{e}_{i}]_{p}*\theta,id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})=2c([\overline{e}_{i}]_{p},id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}),

where the inversion is taken in the field [e¯i]pQHev(M,Λ¯𝔽p,univ)[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}).

Proof.

Note that [e¯i]pQHev(M,Λ¯𝔽p,univ)Λ¯𝔽p,univ[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\cong\overline{\Lambda}_{{\mathbb{F}}_{p},univ}. Thus, there is a δΛ¯𝔽p,univ\delta\in\overline{\Lambda}_{{\mathbb{F}}_{p},univ} such that [e¯i]pθ=δ[e¯i]p[\overline{e}_{i}]_{p}*\theta=\delta[\overline{e}_{i}]_{p}. Hence, ([e¯i]pθ)1=δ1[e¯i]p([\overline{e}_{i}]_{p}*\theta)^{-1}=\delta^{-1}[\overline{e}_{i}]_{p}. We then obtain

c(([e¯i]pθ)1,id,Λ¯𝔽p,univ)+c([e¯i]pθ,id,Λ¯𝔽p,univ)\displaystyle c(([\overline{e}_{i}]_{p}*\theta)^{-1},id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})+c([\overline{e}_{i}]_{p}*\theta,id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})
=c(δ1[e¯i]p,id,Λ¯𝔽p,univ)+c(δ[e¯i]p,id,Λ¯𝔽p,univ)\displaystyle=c(\delta^{-1}[\overline{e}_{i}]_{p},id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})+c(\delta[\overline{e}_{i}]_{p},id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})
=c([e¯i]p,id,Λ¯𝔽p,univ)ν(δ1)+c([e¯i]p,id,Λ¯𝔽p,univ)ν(δ)\displaystyle=c([\overline{e}_{i}]_{p},id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})-\nu(\delta^{-1})+c([\overline{e}_{i}]_{p},id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})-\nu(\delta)
=2c([e¯i]p,id,Λ¯𝔽p,univ).\displaystyle=2c([\overline{e}_{i}]_{p},id,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}).

The proof of the following proposition is similar to that of Theorem 3.1 in [EP03].

Proposition 4.8.

For sufficiently large pp, let QHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) be semisimple, and let Ep={[e¯1]p,,[e¯m]p}E_{p}=\{[\overline{e}_{1}]_{p},\cdots,[\overline{e}_{m}]_{p}\} be the idempotents such that

QHev(M,Λ¯𝔽p,univ)=i=1m[e¯i]pQHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})=\displaystyle\bigoplus_{i=1}^{m}[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})

and [e¯i]pQHev(M,Λ¯𝔽p,univ)Λ¯𝔽p,univ[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\cong\overline{\Lambda}_{{\mathbb{F}}_{p},univ}. Then, there is a constant DD, independent of pp, such that γEp(ϕ)D\gamma_{E_{p}}(\phi)\leq D for each ϕHam(M,ω)\phi\in\operatorname{\mathrm{Ham}}(M,\omega).

Proof.

By the triangle inequality of spectral invariants we have

c([e¯i]p,ϕ)\displaystyle c([\overline{e}_{i}]_{p},\phi) c([e¯i]pb,ϕ)+c(([e¯i]pb)1,id)\displaystyle\leq c([\overline{e}_{i}]_{p}*b,\phi)+c(([\overline{e}_{i}]_{p}*b)^{-1},\operatorname{\mathrm{id}})
c([e¯i]p,id)+c(b,ϕ)+c(([e¯i]pb)1,id)\displaystyle\leq c([\overline{e}_{i}]_{p},\operatorname{\mathrm{id}})+c(b,\phi)+c(([\overline{e}_{i}]_{p}*b)^{-1},\operatorname{\mathrm{id}})
c(b,ϕ)+c(([e¯i]pb)1,id)+δ,\displaystyle\leq c(b,\phi)+c(([\overline{e}_{i}]_{p}*b)^{-1},\operatorname{\mathrm{id}})+\delta,

where the final inequality follows from Proposition 4.4. Furthermore,

c([e¯i]p,ϕ)+c([e¯i]p,ϕ1)\displaystyle c([\overline{e}_{i}]_{p},\phi)+c([\overline{e}_{i}]_{p},\phi^{-1})
=c([e¯i]p,ϕ)inf{c(b,ϕ)|bQH(M),ν(Δ(PSS([e¯i]p),PSS(b))0}\displaystyle=c([\overline{e}_{i}]_{p},\phi)-\inf\{c(b,\phi)|b\in\operatorname{\mathrm{QH}}_{*}(M),\nu(\Delta(\operatorname{\mathrm{PSS}}([\overline{e}_{i}]_{p}),\operatorname{\mathrm{PSS}}(b))\leq 0\}
=sup{c([e¯i]p,ϕ)c(b,ϕ)|bQH(M),ν(Δ(PSS([e¯i]pb),PSS([M])))0}\displaystyle=\displaystyle\sup\{c([\overline{e}_{i}]_{p},\phi)-c(b,\phi)|b\in\operatorname{\mathrm{QH}}_{*}(M),\nu(\Delta(\operatorname{\mathrm{PSS}}([\overline{e}_{i}]_{p}*b),\operatorname{\mathrm{PSS}}([M])))\leq 0\}
sup{c(([e¯i]pb)1,id)|bQH(M),ν(Δ(PSS([e¯i]pb),PSS([M])))0}+δ\displaystyle\leq\displaystyle\sup\{c(([\overline{e}_{i}]_{p}*b)^{-1},id)|b\in\operatorname{\mathrm{QH}}_{*}(M),\nu(\Delta(\operatorname{\mathrm{PSS}}([\overline{e}_{i}]_{p}*b),\operatorname{\mathrm{PSS}}([M])))\leq 0\}+\delta
supb{c(([e¯i]pb)1,id)+c([e¯i]pb,id)|ν(Δ(PSS([e¯i]pb),PSS([M])))0}+δ\displaystyle\leq\displaystyle\sup_{b}\{c(([\overline{e}_{i}]_{p}*b)^{-1},id)+c([\overline{e}_{i}]_{p}*b,id)|\nu(\Delta(\operatorname{\mathrm{PSS}}([\overline{e}_{i}]_{p}*b),\operatorname{\mathrm{PSS}}([M])))\leq 0\}+\delta
=2c([e¯i]p,id)+δ\displaystyle=2c([\overline{e}_{i}]_{p},id)+\delta
3δ\displaystyle\leq 3\delta

The first equality follows from Proposition 2.14 and the second inequality follows from Lemma 4.6. Hence, γ[e¯i]p(ϕ)3δ\gamma_{[\overline{e}_{i}]_{p}}(\phi)\leq 3\delta for all ii, which implies γEp(ϕ)3δ\gamma_{E_{p}}(\phi)\leq 3\delta. ∎

The proof of the following proposition is the same as that of Proposition 12 in [She22].

Proposition 4.9.

For sufficiently large pp, let QHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ}) be semisimple, and Ep={[e¯1]p,,[e¯m]p}E_{p}=\{[\overline{e}_{1}]_{p},\cdots,[\overline{e}_{m}]_{p}\} be the idempotents such that

QHev(M,Λ¯𝔽p,univ)=i=1m[e¯i]pQHev(M,Λ¯𝔽p,univ)\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})=\displaystyle\bigoplus_{i=1}^{m}[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})

and [e¯i]pQHev(M,Λ¯𝔽p,univ)Λ¯𝔽p,univ[\overline{e}_{i}]_{p}*\operatorname{\mathrm{QH}}_{ev}(M,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\cong\overline{\Lambda}_{{\mathbb{F}}_{p},univ}. Then

|βj(ϕ,Λ¯𝔽p,univ)βj(ψ,Λ¯𝔽p,univ)|γEp(ϕψ1,Λ¯𝔽p,univ)+δ|\beta_{j}(\phi,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})-\beta_{j}(\psi,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})|\leq\gamma_{E_{p}}(\phi\psi^{-1},\overline{\Lambda}_{{\mathbb{F}}_{p},univ})+\delta

By Proposition 4.8 and Proposition 4.9, we have the following theorem.

Theorem 4.10.

Suppose that QHev(M,Λ𝕂,univ)\operatorname{\mathrm{QH}}_{ev}(M,\Lambda_{\mathbb{K},univ}) is semisimple. Then the boundary depth of each ψHam(M,ω)\psi\in\operatorname{\mathrm{Ham}}(M,\omega) satisfies β(ψ,Λ¯𝔽p,univ)C\beta(\psi,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})\leq C, where CC is independent of pp.

Since β(ψ,Λ¯𝔽p,univ)=β(ψ,Λ𝔽p,univ)\beta(\psi,\overline{\Lambda}_{{\mathbb{F}}_{p},univ})=\beta(\psi,\Lambda_{{\mathbb{F}}_{p},univ}), we have the following corollary.

Theorem 4.11.

Suppose that QHev(M,Λ𝕂,univ)\operatorname{\mathrm{QH}}_{ev}(M,\Lambda_{\mathbb{K},univ}) is semisimple. Then the boundary depth of each ψHam(M,ω)\psi\in\operatorname{\mathrm{Ham}}(M,\omega) satisfies β(ψ,Λ𝔽p,univ)C\beta(\psi,\Lambda_{{\mathbb{F}}_{p},univ})\leq C, where CC is independent of pp.

5. The p{\mathbb{Z}}_{p}-equivariant Floer homology

This section is a combination of [She22, SZ21, Sei15, Sug21]. We mainly follow the ideas of Sugimoto in [Sug21] to extend the definition of the p{\mathbb{Z}}_{p}-equivariant pair-of-paints product introduced in [She22] to the semipositive setting.

5.1. The p{\mathbb{Z}}_{p}-equivariant Floer homology of CF(ϕ,Λ𝔽p0)p\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p}

Let HkH_{k} be a sequence of Hamiltonian functions CC^{\infty}-converging to HH with (Hk,J)(H_{k},J) a regular pair for all kk. We proceed to construct an XX_{\infty}-module that yield H(p,CF(ϕ,Λ𝔽p0)p)\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p}). Let {x1,,xl}\{x_{1},\dots,x_{l}\} be the 11-periodic orbits of HH. For each ii, xix_{i} splits into 11-periodic orbits {xi1,,xili}\{x_{i}^{1},\dots,x_{i}^{l_{i}}\} of HkH_{k}. For a small cylinder vijv_{i}^{j} connecting xix_{i} to xijx_{i}^{j}, define the action gap

c(xi,xij)=vijω+01H(t,xi(t))Hk(t,xij(t))dtc(x_{i},x_{i}^{j})=\int_{v_{i}^{j}}\omega+\int^{1}_{0}H(t,x_{i}(t))-H_{k}(t,x_{i}^{j}(t))dt

and let τ\tau be the map on CF(Hk,Λ𝔽p,univ)\operatorname{\mathrm{CF}}(H_{k},\Lambda_{{\mathbb{F}}_{p},univ}) defined by xijTc(xi,xij)xijx_{i}^{j}\mapsto T^{c(x_{i},x_{i}^{j})}x_{i}^{j}. The modified Floer differential

dp:CF(Hk,Λ𝔽p0)pΛ0[[u]]𝔽pθCF(Hk,Λ𝔽p0)pΛ𝔽p0[[u]]θd_{{\mathbb{Z}}_{p}}:\operatorname{\mathrm{CF}}(H_{k},\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p}\otimes\Lambda^{0}[[u]]_{{\mathbb{F}}_{p}}{\left<\theta\right>}\rightarrow\operatorname{\mathrm{CF}}(H_{k},\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p}\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}[[u]]{\left<\theta\right>}

is defined by

dp(x1)\displaystyle d_{{\mathbb{Z}}_{p}}(x\otimes 1) =dF(x)1+(1τ)(x)θ\displaystyle=d_{F}^{\prime}(x)\otimes 1+(1-\tau)(x)\otimes\theta
dp(xθ)\displaystyle d_{{\mathbb{Z}}_{p}}(x\otimes\theta) =dF(x)θ+(1+τ++τp1)(x)uθ\displaystyle=d_{F}^{\prime}(x)\otimes\theta+(1+\tau+\cdots+\tau^{p-1})(x)\otimes u\theta

where dF=τ1dFτd_{F}^{\prime}=\tau^{-1}d_{F}\tau. Then, dpd_{{\mathbb{Z}}_{p}} determines an XX_{\infty}-module structure on

CF(Hk,Λ𝔽p0)pΛ𝔽p0θ.\operatorname{\mathrm{CF}}(H_{k},\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p}\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}{\left<\theta\right>}.

Observe that for each ii, (CF(Hi,Λ𝔽p0)pΛ𝔽p0θ,dp)(\operatorname{\mathrm{CF}}(H_{i},\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p}\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}{\left<\theta\right>},d_{{\mathbb{Z}}_{p}}) is an ϵ\epsilon-gapped XKX_{K}-modules, with an associated directed system.

5.2. The p{\mathbb{Z}}_{p}-equivariant Floer homology of CF(ϕp,Λ𝔽p0)\operatorname{\mathrm{CF}}(\phi^{p},\Lambda^{0}_{{\mathbb{F}}_{p}})

Let ff be a p{\mathbb{Z}}_{p}-invariant Morse function on SS^{\infty}, where the p{\mathbb{Z}}_{p}-action on SS^{\infty} is given by scalar multiplication by the pp-th root of unity. For each degree kk\in\mathbb{N}, there are pp critical points denoted by ZkmZ_{k}^{m}, m{0,1,,p1}m\in\{0,1,\cdots,p-1\}. The critical points contained in S2k1S^{2k-1} are {Zjm}\{Z_{j}^{m}\} with j{0,1,,2k+1}j\in\{0,1,\cdots,2k+1\} and m{0,1,,p1}m\in\{0,1,\cdots,p-1\}. Fix an almost complex structure JJ and let H(p)H^{(p)} be a generator of ϕp\phi^{p}. We now proceed as in [Sug21] to construct the relevant ϵ\epsilon-gapped XKX_{K}-modules. Choose ϵ>0\epsilon>0 to be less than the minimal symplectic area of a JJ-holomorphic sphere and the minimal energy of a non-constant Floer cylinder with respect to H(p)H^{(p)}. Let {G(K,i)}\{G_{(K,i)}\} be a family of Hamiltonian functions such that G(K,i)𝑖H(p)G_{(K,i)}\xrightarrow{i}H^{(p)} in the CC^{\infty}-topology and (G(K,i),J)(G_{(K,i)},J) is a regular pair. Furthermore, let 𝒢(K,i)\mathcal{G}^{(K,i)} be a family of Hamiltonian functions parametrized by (w,t)S2K+1×S1(w,t)\in S^{2K+1}\times S^{1} satisfying:

  1. (i)

    For all ww in a small neighbourhood of {Zjm}\{Z_{j}^{m}\}, 𝒢w,t(K,i)(x)=G(K,i)(tm/p,x)\mathcal{G}^{(K,i)}_{w,t}(x)=G_{(K,i)}(t-m/p,x),

  2. (ii)

    For all mZpm\in Z_{p} and wSw\in S^{\infty}, 𝒢mw,t(K,i)=𝒢w,tm/p(K,i)\mathcal{G}^{(K,i)}_{m\cdot w,t}=\mathcal{G}^{(K,i)}_{w,t-m/p},

  3. (iii)

    𝒦\mathcal{K} is invariant under shift by τ\tau, i.e. 𝒢τ(w),t(K,i)=𝒢w,t(K,i)\mathcal{G}^{(K,i)}_{\tau(w),t}=\mathcal{G}^{(K,i)}_{w,t}.

Let x,yx,y be 11-periodic orbits of G(K,i)G_{(K,i)}, mpm\in{\mathbb{Z}}_{p}, λ0\lambda\geq 0, α{0,1}\alpha\in\{0,1\} and 0l2K+10\leq l\leq 2K+1. Suppose that the 11-periodic orbits x~,y~\widetilde{x},\widetilde{y} of H(p)H^{(p)} split into {x,}\{x,\dots\} and {y,}\{y,\dots\}, respectively. Consider the following perturbed Cauchy-Riemann equation

{¯Ju+X𝒢t,w(s)(K,i)(u)0,1=0sw+f(w)=0\begin{cases}\overline{\partial}_{J}u+X_{\mathcal{G}^{(K,i)}_{t,w(s)}}(u)^{0,1}=0\\ \partial_{s}w+\nabla f(w)=0\end{cases}

subject to

{lims(u(s,t),w(s))=(x(t),Zα0)lims+(u(s,t),w(s))=(y(tm/p),Zlm)uω+01H(p)(t,x~(t))H(p)(t,y~(t))dt=λ.\begin{cases}\lim_{s\to-\infty}(u(s,t),w(s))=(x(t),Z_{\alpha}^{0})\\ \lim_{s\to+\infty}(u(s,t),w(s))=(y(t-m/p),Z_{l}^{m})\\ \int_{u}\omega+\int^{1}_{0}H^{(p)}(t,\widetilde{x}(t))-H^{(p)}(t,\widetilde{y}(t))dt=\lambda.\end{cases}

Where, G(K,i)G_{(K,i)} is sufficiently close to H(p)H^{(p)} that either λ=0\lambda=0 or λϵ\lambda\geq\epsilon. Counting solutions to the above perturbed Cauchy-Riemann equation allows one to define a map

dα,l(K,i):CF(G(K,i),Λ𝔽p0)CF(G(K,i),Λ𝔽p0),d^{(K,i)}_{\alpha,l}:\operatorname{\mathrm{CF}}(G_{(K,i)},\Lambda^{0}_{{\mathbb{F}}_{p}})\rightarrow\operatorname{\mathrm{CF}}(G_{(K,i)},\Lambda^{0}_{{\mathbb{F}}_{p}}),

which induces an ϵ\epsilon-gapped XKX_{K}-module (CF(G(K,i),Λ𝔽p0)Λ𝔽p0θ,{δl(K,i)}l=0K)(\operatorname{\mathrm{CF}}(G_{(K,i)},\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}{\left<\theta\right>},\{\delta^{(K,i)}_{l}\}_{l=0}^{K}), where

δl(K,i)(x1)\displaystyle\delta^{(K,i)}_{l}(x\otimes 1) =d0,2l(K,i)(x)1+d0,2l+1(K,i)(x)θ\displaystyle=d^{(K,i)}_{0,2l}(x)\otimes 1+d^{(K,i)}_{0,2l+1}(x)\otimes\theta
δl(K,i)(xθ)\displaystyle\delta^{(K,i)}_{l}(x\otimes\theta) =d1,2l(K,i)(x)1+d1,2l+1(K,i)(x)θ.\displaystyle=d^{(K,i)}_{1,2l}(x)\otimes 1+d^{(K,i)}_{1,2l+1}(x)\otimes\theta.

In particular, (δ(K,i))(K)(δ(K,i))(K)=0mod(uK+1)(\delta^{(K,i)})^{(K)}\circ(\delta^{(K,i)})^{(K)}=0\mod(u^{K+1}), where

(δ(K,i))(K)=δ0(K,i)+δ1(K,i)u++δK(K,i)uK.(\delta^{(K,i)})^{(K)}=\delta^{(K,i)}_{0}+\delta^{(K,i)}_{1}\otimes u+\cdots+\delta^{(K,i)}_{K}\otimes u^{K}.

We next define ϵ\epsilon-gapped XKX_{K}-morphisms

ι(K,ij):CF(G(K,i),Λ𝔽p0)Λ𝔽p0θCF(G(K,j),Λ𝔽p0)Λ𝔽p0θ\iota_{(K,i\rightarrow j)}:\operatorname{\mathrm{CF}}(G_{(K,i)},\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}{\left<\theta\right>}\rightarrow\operatorname{\mathrm{CF}}(G_{(K,j)},\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}{\left<\theta\right>}

as follows. Consider a family 𝒢s,w,t(K,ij)\mathcal{G}^{(K,i\rightarrow j)}_{s,w,t} of Hamiltonian functions connecting 𝒢w,t(K,i)\mathcal{G}^{(K,i)}_{w,t} to 𝒢w,t(K,j)\mathcal{G}^{(K,j)}_{w,t}, which is parametrized by ×S2K+1×S1{\mathbb{R}}\times S^{2K+1}\times S^{1} and satisfies:

  1. (i)

    For s0s\ll 0, 𝒢s,w,t(K,ij)(x)=𝒢w,t(K,i)(x)\mathcal{G}^{(K,i\rightarrow j)}_{s,w,t}(x)=\mathcal{G}^{(K,i)}_{w,t}(x), and for s0s\gg 0, 𝒢s,w,t(K,j)=𝒢w,t(K,j)\mathcal{G}^{(K,j)}_{s,w,t}=\mathcal{G}^{(K,j)}_{w,t}.

  2. (ii)

    For all mZpm\in Z_{p} and wSw\in S^{\infty}, 𝒢s,mw,t(K,ij)=𝒢s,w,tm/p(K,ij)\mathcal{G}^{(K,i\rightarrow j)}_{s,m\cdot w,t}=\mathcal{G}^{(K,i\rightarrow j)}_{s,w,t-m/p},

  3. (iii)

    𝒢\mathcal{G} is invariant under shift by τ\tau, i.e. 𝒢s,τ(w),t(K,ij)=𝒢s,w,t(K,ij)\mathcal{G}^{(K,i\rightarrow j)}_{s,\tau(w),t}=\mathcal{G}^{(K,i\rightarrow j)}_{s,w,t}.

Let x,yx,y be 11-periodic orbits of G(K,i)G_{(K,i)} and G(K,j)G_{(K,j)}, respectively, mpm\in{\mathbb{Z}}_{p}, λ0\lambda\geq 0, α{0,1}\alpha\in\{0,1\} and 0l2K+10\leq l\leq 2K+1. Suppose that the 11-periodic orbits x~,y~\widetilde{x},\widetilde{y} of H(p)H^{(p)} split into {x,}\{x,\dots\} and {y,}\{y,\dots\}, respectively. Define ια,l,(K,ij)\iota_{\alpha,l,(K,i\rightarrow j)} by counting solutions to the following perturbed Cauchy-Riemann equation

{¯Ju+X𝒢s,w(s),t(K,ij)(u)0,1=0sw+f(w)=0\begin{cases}\overline{\partial}_{J}u+X_{\mathcal{G}^{(K,i\rightarrow j)}_{s,w(s),t}}(u)^{0,1}=0\\ \partial_{s}w+\nabla f(w)=0\end{cases}

subject to

{lims(u(s,t),w(s))=(x(t),Zα0)lims+(u(s,t),w(s))=(y(tm/p),Zlm)uω+01H(p)(t,x~(t))H(p)(t,y~(t))dt=λ.\begin{cases}\lim_{s\to-\infty}(u(s,t),w(s))=(x(t),Z_{\alpha}^{0})\\ \lim_{s\to+\infty}(u(s,t),w(s))=(y(t-m/p),Z_{l}^{m})\\ \int_{u}\omega+\int^{1}_{0}H^{(p)}(t,\widetilde{x}(t))-H^{(p)}(t,\widetilde{y}(t))dt=\lambda.\end{cases}

Set,

ι(K,ij),l(x1)\displaystyle\iota_{(K,i\rightarrow j),l}(x\otimes 1) =ι0,2l,(K,ij)(x)1+ι0,2l+1,(K,ij)(x)θ\displaystyle=\iota_{0,2l,(K,i\rightarrow j)}(x)\otimes 1+\iota_{0,2l+1,(K,i\rightarrow j)}(x)\otimes\theta
ι(K,ij),l(xθ)\displaystyle\iota_{(K,i\rightarrow j),l}(x\otimes\theta) =ι1,2l,(K,ij)(x)1+ι1,2l+1,(K,ij)(x)θ.\displaystyle=\iota_{1,2l,(K,i\rightarrow j)}(x)\otimes 1+\iota_{1,2l+1,(K,i\rightarrow j)}(x)\otimes\theta.

It follows from [Sug21, Section 6] that ι(K,jk)ι(K,ij)\iota_{(K,j\rightarrow k)}\circ\iota_{(K,i\rightarrow j)} is ϵ\epsilon-gapped homotopic to ι(K,ik)\iota_{(K,i\rightarrow k)}. In a similar fashion, it is possible to define ϵ\epsilon-gapped XKX_{K}-morphisms

τ(KK+1,i):CF(G(K,i),Λ𝔽p0)Λ𝔽p0θCF(G(K+1,i),Λ𝔽p0)Λ𝔽p0θ.\tau_{(K\rightarrow K+1,i)}:\operatorname{\mathrm{CF}}(G_{(K,i)},\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}{\left<\theta\right>}\rightarrow\operatorname{\mathrm{CF}}(G_{(K+1,i)},\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}{\left<\theta\right>}.

We note that ii has to be sufficiently large so that τ(KK+1,i)\tau_{(K\rightarrow K+1,i)} is defined over Λ𝔽p0\Lambda^{0}_{{\mathbb{F}}_{p}}. It turns out that

{CF(G(K,i),Λ𝔽p0)Λ𝔽p0θ,ι(K,ij),τ(KK+1,i)}\bigg{\{}\operatorname{\mathrm{CF}}(G_{(K,i)},\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}{\left<\theta\right>},\,\,\iota_{(K,i\rightarrow j)},\,\,\tau_{(K\rightarrow K+1,i)}\bigg{\}}

is a directed family of XKX_{K}-modules and, thus, we have a unique up to homotopy ϵ\epsilon-gapped XX_{\infty}-module.

5.3. The p{\mathbb{Z}}_{p}-equivariant pair-of-pants product

The p{\mathbb{Z}}_{p}-equivariant pair-of-pants product is an ϵ\epsilon-gapped XX_{\infty}-morphism induced by the ϵ\epsilon-gapped XKX_{K}-morphism

𝔣(K,i):CF(Hi,Λ𝔽p0)pΛ0θCF(G(K,i),Λ𝔽p0)Λ𝔽p0θ\mathfrak{f}^{(K,i)}:\operatorname{\mathrm{CF}}(H_{i},\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p}\otimes\Lambda^{0}{\left<\theta\right>}\rightarrow\operatorname{\mathrm{CF}}(G_{(K,i)},\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}{\left<\theta\right>}

between the two directed families of ϵ\epsilon-gapped morphisms that were discussed above. Therefore, we obtain a map

H(p,CF(ϕ,Λ𝔽p0)p)H(p,CF(ϕp,Λ𝔽p0))\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p})\rightarrow\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi^{p},\Lambda^{0}_{{\mathbb{F}}_{p}}))

as desired. We proceed to recall the construction of 𝔣(K,i)\mathfrak{f}^{(K,i)}. Set

Σp=(0kp1×[k,k+1])/,\Sigma_{p}=\Bigg{(}\bigsqcup_{0\leq k\leq p-1}{\mathbb{R}}\times[k,k+1]\Bigg{)}/\sim,

where the equivalence relation is given by

  1. (i)

    For all 0kp10\leq k\leq p-1, (s,k)[0,)×[k1,k](s,k)[0,)×[k,k+1](s,k)\in[0,\infty)\times[k-1,k]\sim(s,k)\in[0,\infty)\times[k,k+1],

  2. (ii)

    for all 0kp10\leq k\leq p-1, (s,k)(,0]×[k,k+1](s,k+1)(,0]×[k,k+1](s,k)\in(-\infty,0]\times[k,k+1]\sim(s,k+1)\in(-\infty,0]\times[k,k+1],

  3. (iii)

    and (s,0)[0,)×[0,1](s,p)[0,)×[p1,p](s,0)\in[0,\infty)\times[0,1]\sim(s,p)\in[0,\infty)\times[p-1,p].

Determine a complex structure near the point [(0,0)]Σp[(0,0)]\in\Sigma_{p}, which is in the image of the following coordinate chart w:B(1/2)Σpw:B(1/2)\subset{\mathbb{C}}\rightarrow\Sigma_{p} defined as

w(z)={[zp],0arg(z)π/p,[zp+ki],(2k1)π/parg(z)(2k+1)π/p,[zp+pi],(2p1)π/parg(z)2π.w(z)=\begin{cases}[z^{p}],\qquad&0\leq\arg(z)\leq\pi/p,\\ [z^{p}+ki],&(2k-1)\pi/p\leq\arg(z)\leq(2k+1)\pi/p,\\ [z^{p}+pi],&(2p-1)\pi/p\leq\arg(z)\leq 2\pi.\end{cases}

Assume that H(t,x)=0H(t,x)=0 near t=0t=0 and let 𝒦(K,i)\mathcal{K}^{(K,i)} be a family of Hamiltonians parametrized by (w,z)S2K+1×Σp(w,z)\in S^{2K+1}\times\Sigma_{p} satisfying the following:

  1. (i)

    For all z=[s,t]Σpz=[s,t]\in\Sigma_{p}, 𝒦w,z(K,i)(x)=HK([t],x)\mathcal{K}^{(K,i)}_{w,z}(x)=H_{K}([t],x) when s0s\ll 0, and 𝒦w,z(K,i)(x)=(1/p)G(K,i)(x,t/p)\mathcal{K}^{(K,i)}_{w,z}(x)=(1/p)G_{(K,i)}(x,t/p), when s0s\gg 0,

  2. (ii)

    For all z=[s,t]Σpz=[s,t]\in\Sigma_{p}, 𝒦w,z(K,i)iH(t,x)\mathcal{K}^{(K,i)}_{w,z}\xrightarrow{i\rightarrow\infty}H(t,x),

  3. (iii)

    For all mZpm\in Z_{p} and wS2K+1w\in S^{2K+1}, 𝒦mw,t(K,i)=𝒦w,z+mi(K,i)\mathcal{K}^{(K,i)}_{m\cdot w,t}=\mathcal{K}^{(K,i)}_{w,z+mi},

  4. (iv)

    𝒦\mathcal{K} is invariant under shift by τ\tau, i.e. 𝒦τ(w),t(K,i)=𝒦w,z(K,i)\mathcal{K}^{(K,i)}_{\tau(w),t}=\mathcal{K}^{(K,i)}_{w,z}.

For 11-periodic orbits {xk}k=1p\{x_{k}\}_{k=1}^{p} and yy of G(K,i)G_{(K,i)} and HKH_{K}, respectively, mpm\in{\mathbb{Z}}_{p}, λ0\lambda\geq 0, α{0,1}\alpha\in\{0,1\}, and 0l2K+10\leq l\leq 2K+1, consider the following equation:

{¯Ju+X𝒦w(s),[s,t](K,i)(u)0,1=0sw+f(w)=0\begin{cases}\overline{\partial}_{J}u+X_{\mathcal{K}^{(K,i)}_{w(s),[s,t]}}(u)^{0,1}=0\\ \partial_{s}w+\nabla f(w)=0\end{cases}

subject to

{lims(u([s,t]),w(s))=(xi(t),Zα0)lims+(u([s,t]),w(s))=(y(tm/p),Zlm)u~ω+01H(p)(t,y~(t))𝑑tiH(p)(t,xi~(t))dt=λ.\begin{cases}\lim_{s\to-\infty}(u([s,t]),w(s))=(x_{i}(t),Z_{\alpha}^{0})\\ \lim_{s\to+\infty}(u([s,t]),w(s))=(y(t-m/p),Z_{l}^{m})\\ \int_{\widetilde{u}}\omega+\int^{1}_{0}H^{(p)}(t,\widetilde{y}(t))dt-\sum_{i}H^{(p)}(t,\widetilde{x_{i}}(t))dt=\lambda.\end{cases}

Here, {x~i}\{\widetilde{x}_{i}\} and y~\widetilde{y} are the corresponding periodic orbits of the unperturbed system and u~\widetilde{u} is a map connecting them. Counting solutions to this equation, yields the map

fα,l(K,i):CF(Hi,Λ𝔽p0)pΛ0θCF(G(K,i),Λ𝔽p0)Λ𝔽p0θ.f^{(K,i)}_{\alpha,l}:\operatorname{\mathrm{CF}}(H_{i},\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p}\otimes\Lambda^{0}{\left<\theta\right>}\rightarrow\operatorname{\mathrm{CF}}(G_{(K,i)},\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes\Lambda^{0}_{{\mathbb{F}}_{p}}{\left<\theta\right>}.

Finally, one defines 𝔣(K,i)={fl(K,i)}l=0K\mathfrak{f}^{(K,i)}=\{f^{(K,i)}_{l}\}^{K}_{l=0} where

fl(K,i)((x1xp)1)\displaystyle f^{(K,i)}_{l}((x_{1}\otimes\cdots\otimes x_{p})\otimes 1) =f0,2l(K,i)(x1xp)1+f0,2l+1(K,i)(x1xp)θ\displaystyle=f^{(K,i)}_{0,2l}(x_{1}\otimes\cdots\otimes x_{p})\otimes 1+f^{(K,i)}_{0,2l+1}(x_{1}\otimes\cdots\otimes x_{p})\otimes\theta
fl(K,i)((x1xp)θ)\displaystyle f^{(K,i)}_{l}((x_{1}\otimes\cdots\otimes x_{p})\otimes\theta) =f1,2l(K,i)(x1xp)1+f1,2l+1(K,i)(x1xp)θ.\displaystyle=f^{(K,i)}_{1,2l}(x_{1}\otimes\cdots\otimes x_{p})\otimes 1+f^{(K,i)}_{1,2l+1}(x_{1}\otimes\cdots\otimes x_{p})\otimes\theta.

It follows from [Sug21] that 𝔣(K,i)\mathfrak{f}^{(K,i)} is a morphism between the desired directed family of ϵ\epsilon-gapped XKX_{K}-modules.

5.4. The Smith type inequality for total bar-lengths

With the equivariant pair-of-pants product defined, we can prove the following theorem in the same way as in [She22].

Theorem 5.1.

Let ϕHam(M,ω)\phi\in\operatorname{\mathrm{Ham}}(M,\omega) be a Hamiltonian diffeomorphism of a closed semipositive symplectic manifold (M,ω)(M,\omega). Suppose that Fix(ϕp)\mathrm{Fix}(\phi^{p}) is finite. Then

pβtot(ϕ,𝔽p)βtot(ϕp,𝔽p)p\cdot\beta_{\mathrm{tot}}(\phi,{\mathbb{F}}_{p})\leq\beta_{\mathrm{tot}}(\phi^{p},{\mathbb{F}}_{p}).

Proof.

First, note that HF(ϕ,Λ𝔽p0)Λ𝔽p0Λ𝒦0θ\operatorname{\mathrm{HF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes_{\Lambda^{0}_{{\mathbb{F}}_{p}}}\Lambda^{0}_{\mathcal{K}}{\left<\theta\right>} is isomorphic to H(p,CF(ϕ,Λ𝔽p0)){\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})) since the p{\mathbb{Z}}_{p}-action on CF(ϕ,Λ𝔽p0)\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}}) is trivial. Denote by

β1,,βs\beta_{1},\dots,\beta_{s}

the bar-lengths of H(p,CF(ϕ,Λ𝔽p0))\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})). Then, s=2K(ϕ,𝔽p)s=2K(\phi,{\mathbb{F}}_{p}) and β2i1=β2i=βi(ϕ,𝔽p)\beta_{2i-1}=\beta_{2i}=\beta_{i}(\phi,{\mathbb{F}}_{p}) since

HF(ϕ,Λ𝔽p0)Λ𝔽p0Λ𝒦0θ(HF(ϕ,Λ𝔽p0)Λ𝔽p0Λ𝒦0)(HF(ϕ,Λ𝔽p0)Λ𝔽p0Λ𝒦0).\operatorname{\mathrm{HF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes_{\Lambda^{0}_{{\mathbb{F}}_{p}}}\Lambda^{0}_{\mathcal{K}}{\left<\theta\right>}\cong(\operatorname{\mathrm{HF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes_{\Lambda^{0}_{{\mathbb{F}}_{p}}}\Lambda^{0}_{\mathcal{K}})\oplus(\operatorname{\mathrm{HF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})\otimes_{\Lambda^{0}_{{\mathbb{F}}_{p}}}\Lambda^{0}_{\mathcal{K}}).

Secondly, as in [She22, Section 7.2], the map

xCF(ϕ,Λ𝔽p0)xxCF(ϕ,Λ𝔽p0)px\in\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})\mapsto x\otimes\cdots\otimes x\in\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p}

induces a quasi-Frobenius isomorphism

rpH(p,CF(ϕ,Λ𝔽p0))H(p,CF(ϕ,Λ𝔽p0)p),r_{p}^{*}\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}}))\rightarrow\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p}),

where rp:Λ𝒦0Λ𝒦0r_{p}:\Lambda^{0}_{\mathcal{K}}\rightarrow\Lambda^{0}_{\mathcal{K}} is the homomorphism defined by TT1pT\mapsto T^{\frac{1}{p}}. By adjusting by rpr_{p}, the isomorphism is defined over Λ𝒦0\Lambda^{0}_{\mathcal{K}}. Denote the bar-lengths of H(p,CF(ϕ,Λ𝔽p0))\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})) by βp,1,,βp,s\beta_{p,1},\cdots,\beta_{p,s}, which are also the bar-lengths of rpH(p,CF(ϕ,Λ𝔽p0))r_{p}^{*}\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})). Then, βp,i=pβi\beta_{p,i}=p\cdot\beta_{i} by the map rpr_{p}^{*} and s=2K(ϕ,𝔽p)s=2K(\phi,{\mathbb{F}}_{p}). Thus, βp,2i1=βp,2i=pβ(ϕ,𝔽p)\beta_{p,2i-1}=\beta_{p,2i}=p\cdot\beta(\phi,{\mathbb{F}}_{p}). Next, by the equivariant pair-of-pants product, there is an isomoprhims

H(p,CF(ϕ,Λ𝔽p0)p)H(p,CF(ϕp,Λ𝔽p0))\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi,\Lambda^{0}_{{\mathbb{F}}_{p}})^{\otimes p})\rightarrow\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi^{p},\Lambda^{0}_{{\mathbb{F}}_{p}}))

over Λ𝒦0\Lambda^{0}_{\mathcal{K}}. Denote by β^1,,β^s\widehat{\beta}_{1},\cdots,\widehat{\beta}_{s} the bar-lengths of H(p,CF(ϕp,Λ𝔽p0))\operatorname{\mathrm{H}}({\mathbb{Z}}_{p},\operatorname{\mathrm{CF}}(\phi^{p},\Lambda^{0}_{{\mathbb{F}}_{p}})). Then, β^2i1=β^2i=pβ^(ϕ,𝔽p)\widehat{\beta}_{2i-1}=\widehat{\beta}_{2i}=p\cdot\widehat{\beta}(\phi,{\mathbb{F}}_{p}) as in [She22, Corollary 23]. Finally, by [She22, Proposition 1 and Lemma 18] we have that β^tot=β^i2βtot(ϕp,𝔽p)\widehat{\beta}_{tot}=\sum\widehat{\beta}_{i}\leq 2\cdot\beta_{tot}(\phi^{p},{\mathbb{F}}_{p}), in particular, 2pβ(ϕ,𝔽p)2βtot(ϕp,𝔽p)2p\cdot\beta(\phi,{\mathbb{F}}_{p})\leq 2\cdot\beta_{tot}(\phi^{p},{\mathbb{F}}_{p}). When the fixed points are degenerate, one can argues using local equivariant Floer homology as in [She22, Section 7.4]. ∎

Combining Theorem 4.11 and Theorem 5.1, one obtains the main result, Theorem 1.3, as in Section 1.3 in the introduction.

References

  • [AB21] M. Abouzaid and A.J. Blumberg. Arnold conjecture and Morava K-theory. arXiv preprint arXiv:2103.01507, 2021.
  • [Abo15] M. Abouzaid. Symplectic cohomology and Viterbo’s theorem. In Free loop spaces in geometry and topology, volume 24 of IRMA Lect. Math. Theor. Phys., pages 271–485. Eur. Math. Soc., Zürich, 2015.
  • [Arn65] V.I. Arnold. Sur une proprietes topologique des applications globalment canonique de la mechanique classique. CR. Acad. Sci. Paris, 261:3719–3722, 1965.
  • [Arn14] V.I. Arnold. The stability problem and ergodic properties for classical dynamical systems. Vladimir I. Arnold-Collected Works: Hydrodynamics, Bifurcation Theory, and Algebraic Geometry 1965-1972, pages 107–113, 2014.
  • [AS23] M.S. Atallah and E. Shelukhin. Hamiltonian no-torsion. Geometry & Topology, 27(7):2833–2897, 2023.
  • [Bay04] A. Bayer. Semisimple quantum cohomology and blowups. International Mathematics Research Notices, 2004(40):2069–2083, 2004.
  • [BC09] P. Biran and O. Cornea. Rigidity and uniruling for Lagrangian submanifolds. Geometry & Topology, 13(5):2881–2989, 2009.
  • [Bir13] G.D. Birkhoff. Proof of Poincaré’s geometric theorem. Transactions of the American Mathematical Society, 14(1):14–22, 1913.
  • [Bir26] G.D. Birkhoff. An extension of Poincaré’s last geometric theorem. Acta mathematica, 47(4):297–311, 1926.
  • [BL15] U. Bauer and M. Lesnick. Induced matchings and the algebraic stability of persistence barcodes. Journal of Computational Geometry, 6(2):162–191, 2015.
  • [BN77] M. Brown and W.D. Neumann. Proof of the Poincaré-Birkhoff fixed point theorem. Michigan Math. J, 24(1):21–31, 1977.
  • [BX22] S. Bai and G. Xu. Arnold conjecture over integers. arXiv preprint arXiv:2209.08599, 2022.
  • [BX23] S. Bai and G. Xu. Hofer-Zehnder conjecture for toric manifolds. arXiv preprint arXiv:2309.07991, 2023.
  • [Cas86] J.W.S. Cassels. Local fields, volume 3. Cambridge University Press Cambridge, 1986.
  • [CB15] W. Crawley-Boevey. Decomposition of pointwise finite-dimensional persistence modules. Journal of Algebra and its Applications, 14(05):1550066, 2015.
  • [CCSG+09] F. Chazal, D. Cohen-Steiner, M. Glisse, L.J. Guibas, and S.Y. Oudot. Proximity of persistence modules and their diagrams. In Proceedings of the twenty-fifth annual symposium on Computational geometry, pages 237–246, 2009.
  • [CDSGO16] F. Chazal, V. De Silva, M. Glisse, and S. Oudot. The structure and stability of persistence modules, volume 10. Springer, 2016.
  • [CSEH05] D. Cohen-Steiner, H. Edelsbrunner, and J. Harer. Stability of persistence diagrams. In Proceedings of the twenty-first annual symposium on Computational geometry, pages 263–271, 2005.
  • [CZ83] C. Conley and E. Zehnder. The Birkhoff-Lewis fixed point theorem and a conjecture of VI Arnold. Inventiones Mathematicae, 73:33–50, 1983.
  • [CZ84] C. Conley and E. Zehnder. Subharmonic solutions and Morse theory. Physica A: Statistical Mechanics and its Applications, 124(1-3):649–657, 1984.
  • [CZ86] C. Conley and E. Zehnder. for symplectic maps and subharmonic solutions of Hamiltonian equations on tori. Nonlinear Functional Analysis and Its Applications, Part 1: Proceedings of the Summer Research Institute: the Result of the Thirty-first Summer Research Institute of the American Mathematical Society; Berkeley-Calif., July 11-29, 1983, 45(Part 1):283, 1986.
  • [Eli79] Y. Eliashberg. Estimates on the number of fixed points of area preserving transformations. Syktyvkar University, 1979.
  • [ELZ00] H. Edelsbrunner, D. Letscher, and A. Zomorodian. Topological persistence and simplification. In Proceedings 41st annual symposium on foundations of computer science, pages 454–463. IEEE, 2000.
  • [EP03] M. Entov and L. Polterovich. Calabi quasimorphism and quantum homology. International Mathematics Research Notices, 2003(30):1635–1676, 2003.
  • [EP08] M. Entov and L. Polterovich. Symplectic quasi-states and semi-simplicity of quantum homology. Toric Topology, pages 47–70, 2008.
  • [FH03] J. Franks and M. Handel. Periodic points of Hamiltonian surface diffeomorphisms. Geometry & Topology, 7(2):713–756, 2003.
  • [Flo86] A. Floer. Proof of the Arnol’d conjecture for surfaces and generalizations to certain Kähler manifolds. Duke Math. J., 53(1):1–32, 1986.
  • [Flo87] A. Floer. Morse theory for fixed points of symplectic diffeomorphisms. Bull. Amer. Math. Soc. (N.S.), 16(2):279–281, 1987.
  • [Flo89] A. Floer. Symplectic fixed points and holomorphic spheres. Comm. Math. Phys., 120(4):575–611, 1989.
  • [FO99] K. Fukaya and K. Ono. Arnold conjecture and Gromov–Witten invariant. Topology, 38(5):933–1048, 1999.
  • [FOOO10] K. Fukaya, Y.-G. Oh, H. Ohta, and K. Ono. Lagrangian intersection Floer theory: anomaly and obstruction, Part II, volume 2. American Mathematical Soc., 2010.
  • [FOOO13] K. Fukaya, Y.-G. Oh, H. Ohta, and K. Ono. Displacement of polydisks and Lagrangian Floer theory. Journal of Symplectic Geometry, 11(2):231–268, 2013.
  • [For85] B. Fortune. A symplectic fixed point theorem for Pn\mathbb{C}\mathrm{P}^{n}. Inventiones Mathematicae, 81:29–46, 1985.
  • [Fra92] J. Franks. Geodesics on S2{S}^{2} and periodic points of annulus homeomorphisms. Inventiones Mathematicae, 108(1):403–418, 1992.
  • [Fra96] J. Franks. Area preserving homeomorphisms of open surfaces of genus zero. New York J. Math, 2(1):19, 1996.
  • [FW85] B. Fortune and A. Weinstein. A symplectic fixed point theorem for complex projective spaces. Bull. Amer. Math. Soc.(NS), 12(11):128–130, 1985.
  • [GG09] V.L. Ginzburg and B.Z. Gürel. Action and index spectra and periodic orbits in Hamiltonian dynamics. Geometry & Topology, 13(5):2745–2805, 2009.
  • [GG10] V.L. Ginzburg and B.Z. Gürel. Local Floer homology and the action gap. Journal of Symplectic Geometry, 8(3):323–357, 2010.
  • [GG12] V.L. Ginzburg and B.Z. Gürel. Conley conjecture for negative monotone symplectic manifolds. IMRN: International Mathematics Research Notices, 2012(8), 2012.
  • [GG14] V.L. Ginzburg and B.Z. Gürel. Hyperbolic fixed points and periodic orbits of Hamiltonian diffeomorphisms. Duke Mathematical Journal, 163(3), 2014.
  • [GG16] V.L. Ginzburg and B.Z. Gürel. Non-contractible periodic orbits in Hamiltonian dynamics on closed symplectic manifolds. Compositio Mathematica, 152(9):1777–1799, 2016.
  • [GG18] V.L. Ginzburg and B.Z. Gürel. Hamiltonian pseudo-rotations of projective spaces. Inventiones Mathematicae, 214(3):1081–1130, 2018.
  • [GG19] V.L. Ginzburg and B.Z. Gürel. Conley conjecture revisited. International Mathematics Research Notices, 2019(3):761–798, 2019.
  • [Gin10] V.L. Ginzburg. The Conley conjecture. Annals of Mathematics, pages 1127–1180, 2010.
  • [Gro85] M. Gromov. Pseudo holomorphic curves in symplectic manifolds. Inventiones Mathematicae, 82(2):307–347, 1985.
  • [Gür13] B.Z. Gürel. On non-contractible periodic orbits of Hamiltonian diffeomorphisms. Bulletin of the London Mathematical Society, 45(6):1227–1234, 2013.
  • [Hei12] D. Hein. The Conley conjecture for irrational symplectic manifolds. Journal of Symplectic Geometry, 10(2):183–202, 2012.
  • [Hin09] N. Hingston. Subharmonic solutions of Hamiltonian equations on tori. Annals of Mathematics, pages 529–560, 2009.
  • [HS95] H. Hofer and D. Salamon. Floer homology and Novikov rings. In The Floer memorial volume, pages 483–524. Springer, 1995.
  • [Iri07] H. Iritani. Convergence of quantum cohomology by quantum Lefschetz. 2007.
  • [KK07] Y. Karshon and L. Kessler. Circle and torus actions on equal symplectic blow-ups of P2\mathbb{C}\mathrm{P}^{2}. Mathematical research letters, 14(5):807–823, 2007.
  • [LC06] P. Le Calvez. Periodic orbits of Hamiltonian homeomorphisms of surfaces. Duke Mathematical Journal, 133(1), 2006.
  • [LMN22] T.-J. Li, J. Min, and S. Ning. Enumerative aspect of symplectic log Calabi-Yau divisors and almost toric fibrations. arXiv preprint arXiv:2203.08544, 2022.
  • [LT98] G. Liu and G. Tian. Floer homology and Arnold conjecture. Journal of Differential Geometry, 49(1):1–74, 1998.
  • [McD11] D. McDuff. The topology of toric symplectic manifolds. Geometry & Topology, 15(1):145–190, 2011.
  • [McL12] M. McLean. Local floer homology and infinitely many simple Reeb orbits. Algebraic & Geometric Topology, 12(4):1901–1923, 2012.
  • [MS12] D. McDuff and D. Salamon. J-holomorphic curves and symplectic topology, volume 52. American Mathematical Soc., 2012.
  • [MS17] D. McDuff and D. Salamon. Introduction to symplectic topology. Oxford Graduate Texts in Mathematics. Oxford University Press, Oxford, third edition, 2017.
  • [Oh05] Y.-G. Oh. Construction of spectral invariants of Hamiltonian paths on closed symplectic manifolds. The Breadth of Symplectic and Poisson Geometry: Festschrift in Honor of Alan Weinstein, pages 525–570, 2005.
  • [Oh06] Y.-G. Oh. Lectures on Floer theory and spectral invariants of Hamiltonian flows. In Morse theoretic methods in nonlinear analysis and in symplectic topology, pages 321–416. Springer, 2006.
  • [Oh15] Y.-G. Oh. Symplectic topology and Floer homology. Vols. 1 and 2, volume 27 and 28 of New Mathematical Monographs. Cambridge University Press, Cambridge, 2015.
  • [Ori17] R. Orita. Non-contractible periodic orbits in Hamiltonian dynamics on tori. Bulletin of the London Mathematical Society, 49(4):571–580, 2017.
  • [Ori19] R. Orita. On the existence of infinitely many non-contractible periodic orbits of Hamiltonian diffeomorphisms of closed symplectic manifolds. Journal of Symplectic Geometry, 17(6):1893–1927, 2019.
  • [Ost06] Y. Ostrover. Calabi quasi-morphisms for some non-monotone symplectic manifolds. Algebraic & Geometric Topology, 6(1):405–434, 2006.
  • [OT09] Y. Ostrover and I. Tyomkin. On the quantum homology algebra of toric Fano manifolds. Selecta Mathematica, 15:121–149, 2009.
  • [Poi12] H. Poincaré. Sur un théoreme de géométrie. Rendiconti del Circolo Matematico di Palermo (1884-1940), 33(1):375–407, 1912.
  • [PS16] L. Polterovich and E. Shelukhin. Autonomous Hamiltonian flows, Hofer’s geometry and persistence modules. Selecta Mathematica, 22(1):227–296, 2016.
  • [PSS96] S. Piunikhin, D. Salamon, and M. Schwarz. Symplectic Floer-Donaldson theory and quantum cohomology. Contact and symplectic geometry (Cambridge, 1994), 8:171–200, 1996.
  • [PSS17] L. Polterovich, E. Shelukhin, and V. Stojisavljević. Persistence modules with operators in Morse and Floer theory. Moscow Mathematical Journal, 17(4):757–786, 2017.
  • [Rez22] S. Rezchikov. Integral Arnol’d conjecture. arXiv preprint arXiv:2209.11165, 2022.
  • [Rua99] Y. Ruan. Virtual neighborhoods and pseudo-holomorphic curves. Turkish Journal of Mathematics, 23(1):161–232, 1999.
  • [Sch00] M. Schwarz. On the action spectrum for closed symplectically aspherical manifolds. Pacific Journal of Mathematics, 193(2):419–461, 2000.
  • [Sei02] P. Seidel. Symplectic Floer homology and the mapping class group. Pacific J. Math., 206(1):219–229, 2002.
  • [Sei15] P. Seidel. The equivariant pair-of-pants product in fixed point Floer cohomology. Geometric and Functional Analysis, 25(3):942–1007, 2015.
  • [She22] E. Shelukhin. On the Hofer-Zehnder conjecture. Annals of Mathematics(2), 195(3):775–839, 2022.
  • [Sug21] Y. Sugimoto. On the Hofer-Zehnder conjecture for non-contractible periodic orbits in Hamiltonian dynamics. arXiv preprint arXiv:2102.05273, 2021.
  • [SW] E. Shelukhin and N. Wilkins. Quantum Steenrod powers and Hamiltonian maps. Work in progress.
  • [SZ92] D. Salamon and E. Zehnder. Morse theory for periodic solutions of Hamiltonian systems and the Maslov index. Communications on pure and applied mathematics, 45(10):1303–1360, 1992.
  • [SZ21] E. Shelukhin and J. Zhao. The /p\mathbb{Z}/p\mathbb{Z}-equivariant product-isomorphism in fixed point Floer cohomology. Journal of Symplectic Geometry, 19(5):1101–1188, 2021.
  • [Ush08] M. Usher. Spectral numbers in Floer theories. Compositio Mathematica, 144(6):1581–1592, 2008.
  • [Ush10] M. Usher. Duality in filtered Floer–Novikov complexes. Journal of Topology and Analysis, 2(02):233–258, 2010.
  • [Ush11a] M. Usher. Boundary depth in Floer theory and its applications to Hamiltonian dynamics and coisotropic submanifolds. Israel Journal of Mathematics, 184:1–57, 2011.
  • [Ush11b] M. Usher. Deformed Hamiltonian Floer theory, capacity estimates and Calabi quasimorphisms. Geometry & Topology, 15(3):1313–1417, 2011.
  • [Ush13] M. Usher. Hofer’s metrics and boundary depth. In Annales scientifiques de l’École Normale Supérieure, volume 46, pages 57–129, 2013.
  • [UZ16] M. Usher and J. Zhang. Persistent homology and Floer–Novikov theory. Geometry & Topology, 20(6):3333–3430, 2016.
  • [Vit92] C. Viterbo. Symplectic topology as the geometry of generating functions. Mathematische Annalen, 292(1):685–710, 1992.
  • [Zap] F. Zapolsky. The Lagrangian Floer-quantum-PSS package and canonical orientations in Floer theory. Preprint, arXiv:1507.02253, 2015.
  • [ZC04] A. Zomorodian and G. Carlsson. Computing persistent homology. In Proceedings of the twentieth annual symposium on Computational geometry, pages 347–356, 2004.