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

Knot surgery formulae for instanton Floer homology I: the main theorem

Zhenkun Li Department of Mathematics and Statistics, University of South Florida [email protected]  and  Fan Ye Department of Mathematics, Harvard University [email protected]
Abstract.

We prove an integral surgery formula for framed instanton homology I(Ym(K))I^{\sharp}(Y_{m}(K)) for any knot KK in a 33-manifold YY with [K]=0H1(Y;)[K]=0\in H_{1}(Y;\mathbb{Q}) and m0m\neq 0. Although the statement is similar to Ozsváth-Szabó’s integral surgery formula for Heegaard Floer homology, the proof is new and based on sutured instanton homology SHISHI and the octahedral lemma in the derived category. As byproducts, we obtain a formula computing instanton knot homology of the dual knot analogous to Eftekhary’s and Hedden-Levine’s work, and also an exact triangle between I(Ym(K))I^{\sharp}(Y_{m}(K)), I(Ym+k(K))I^{\sharp}(Y_{m+k}(K)) and kk copies of I(Y)I^{\sharp}(Y) for any m0m\neq 0 and large kk. In the proof of the formula, we discover many new exact triangles for sutured instanton homology and relate some surgery cobordism map to the sum of bypass maps, which are of independent interest. In a companion paper, we derive many applications and computations based on the integral surgery formula.

1. Introduction

The framed instanton homology I(Y)I^{\sharp}(Y) for a closed 3-manifold YY was introduced by Kronheimer and Mrowka in [17] and has been conjectured to be isomorphic to the hat version of Heegaard Floer homology HF^(Y)\widehat{HF}(Y). This conjecture is still widely open and, due to the computational difficulty of instanton Floer homology, not many examples has been known. In recent years, many people have done computations of the framed instanton homology of special families of 33-manifolds, see for example [21, 9, 5]. Yet, most results have focused on computing the dimension of framed instanton Floer homology and many techniques only work for S3S^{3} or rational homology spheres, however, a general structural theorem that relates the framed instanton homology of Dehn surgeries to the information from the knot complement still remains elusive.

In [22], the authors of the current paper proved a large surgery formula for framed instanton homology which led to a series of applications in computing the framed instanton homology and studying the representations of the fundamental groups of Dehn surgeries of some families of knots. However, in that work, the Dehn surgery slope must be large (at least 2g+12g+1 where gg is the Seifert genus of the knot), and thus still not much is known about the framed instanton homology of small Dehn surgery slopes. In this paper, we further prove an integral surgery formula for rationally null-homologous knots, inspired by Ozsváth-Szabó’s surgery formula for Heegaard Floer homology [27, 28]. For simplicity, in the introduction, we present only the discussions and results for (integral) null-homologous knots (e.g. knots in S3S^{3}) and leave the general setup to Section 3.3.

First, let us recall the results from [22]. Suppose KYK\subset Y is a null-homologous knot. Let Y\N(K)Y\backslash N(K) be the knot complement, and let Γμ\Gamma_{\mu} be the union of two oppositely oriented meridians of the knot on (Y\N(K))\partial(Y\backslash N(K)). Let SHI(Y\N(K),Γμ)SHI(-Y\backslash N(K),-\Gamma_{\mu}) be the corresponding sutured instanton homology introduced by Kronheimer-Mrowka [16], where the minus sign denotes orientation reversal for technical needs (note that SHI(M,γ)SHI(M,γ)SHI(-M,-\gamma)\cong SHI(M,\gamma) and in particular I(Ym(K))I(Ym(K))I^{\sharp}(-Y_{-m}(K))\cong I^{\sharp}(Y_{-m}(K))). A Seifert surface of KK induces a \mathbb{Z}-grading on SHI(Y\N(K),Γμ)SHI(-Y\backslash N(K),-\Gamma_{\mu}). In [22], we constructed a set of differentials on SHI(Y\N(K),Γμ)SHI(-Y\backslash N(K),-\Gamma_{\mu})

dji:SHI(Y\N(K),Γμ,i)SHI(Y\N(K),Γμ,j)d^{i}_{j}:SHI(-Y\backslash N(K),-\Gamma_{\mu},i)\rightarrow SHI(-Y\backslash N(K),-\Gamma_{\mu},j)

for any gradings iji\neq j\in\mathbb{Z}. We then constructed bent complexes

As=(SHI(Y\N(K),Γμ),si<jdji+si>jdji),A_{s}=\bigg{(}SHI(-Y\backslash N(K),-\Gamma_{\mu}),\sum_{s\leq i<j}d^{i}_{j}+\sum_{s\geq i>j}d^{i}_{j}\bigg{)},
B+=(SHI(Y\N(K),Γμ),i<jdji),andB=(SHI(Y\N(K),Γμ),i>jdji).B^{+}=\bigg{(}SHI(-Y\backslash N(K),-\Gamma_{\mu}),\sum_{i<j}d^{i}_{j}\bigg{)},~{}{\rm and}~{}B^{-}=\bigg{(}SHI(-Y\backslash N(K),-\Gamma_{\mu}),\sum_{i>j}d^{i}_{j}\bigg{)}.

From [22], the homologies of these complexes are related to the Dehn surgeries of KK as follows:

(1.1) H(B+)H(B)I(Y),H(B^{+})\cong H(B^{-})\cong I^{\sharp}(-Y),
(1.2) I(Ym(K))s=1m2m12H(As) for any integer m2g(K)+1.I^{\sharp}(Y_{-m}(K))\cong\bigoplus_{s=\lfloor\frac{1-m}{2}\rfloor}^{\lfloor\frac{m-1}{2}\rfloor}H(A_{s})\text{ for any integer }m\geq 2g(K)+1.

To state the integral surgery formula, we introduce more notations. For ss\in\mathbb{Z}, let Bs±B^{\pm}_{s} be identical copies of B±B^{\pm}. Define chain maps

π±,s:AsBs±\pi^{\pm,s}:A_{s}\rightarrow B^{\pm}_{s}

as follows. For x(SHI(Y\N(K),Γμ),i)x\in(SHI(-Y\backslash N(K),-\Gamma_{\mu}),i),

π+,s(x)={xis,0i<s,andπ,s(x)={xis,0i>s.\pi^{+,s}(x)=\begin{cases}x&i\geq s,\\ 0&i<s,\end{cases}~{}{\rm and}~{}\pi^{-,s}(x)=\begin{cases}x&i\leq s,\\ 0&i>s.\end{cases}

Let π±\pi^{\pm} denote the direct sum of all π±,s\pi^{\pm,s}. While this slightly abuses the notation, we also use them to denote the induced maps on the homologies. The main result of the paper is the following.

Theorem 1.1 (Integral surgery formula).

Suppose KYK\subset Y is a null-homologous knot. Let AsA_{s}, Bs±B_{s}^{\pm}, and π±\pi^{\pm} be defined as above. Then for any m\{0}m\in\mathbb{Z}\backslash\{0\}, there exists an isomorphism

Ξm:sH(Bs+)sH(Bs+m)\Xi_{m}:\bigoplus_{s\in\mathbb{Z}}H(B^{+}_{s})\xrightarrow{\cong}\bigoplus_{s\in\mathbb{Z}}H(B^{-}_{s+m})

as the direct sum of isomorphisms

Ξm,s:H(Bs+)H(Bs+m)\Xi_{m,s}:H(B^{+}_{s})\xrightarrow{\cong}H(B^{-}_{s+m})

such that

I(Ym(K))H(Cone(π+Ξmπ+:sH(As)sH(Bs))).I^{\sharp}(-Y_{-m}(K))\cong H\bigg{(}{\rm Cone}(\pi^{-}+\Xi_{m}\circ\pi^{+}:\bigoplus_{s\in\mathbb{Z}}H(A_{s})\to\bigoplus_{s\in\mathbb{Z}}H(B^{-}_{s}))\bigg{)}.
Remark 1.2.

As an analog of the surgery formula in Heegaard Floer homology, the map π\pi^{-} is related to the vertical projection map vv, and the map Ξmπ+\Xi_{m}\circ\pi^{+} is related to the map hh, which is defined by the composition of the horizontal projection and a chain homotopy equivalence between C{j0}C\{j\geq 0\} to C{i0}C\{i\geq 0\}. Here we bend the horizontal part of the hook complex to become vertical, so the differentials go upwards. The homotopy equivalence in Heegaard Floer homology depends on many auxiliary choices (c.f. Construction before [13, Lemma 2.16]). The same situation applies to Ξm\Xi_{m}. Hence we only state the existence of the isomorphism.

Remark 1.3.

The hypothesis of Theorem 1.1 excludes the case where m=0m=0. This is due to the sign ambiguity in the definition of sutured instanton homology. The original version of sutured instanton homology defined by Kronheimer-Mrowka [16] was only well-defined up to isomorphisms, and then Baldwin-Sivek [2] proved that they are well-defined up to a scalar in \mathbb{C}. As a result, all related maps are only well-defined up to scalars. When m0m\neq 0, the maps π+,s\pi^{+,s} and Ξmπ,s\Xi_{m}\circ\pi^{-,s} have distinct target spaces, namely BsB_{s} and Bs+mB_{s+m}. As a result, the scalar ambiguity for individual maps does not influence the dimension of the homology of the mapping cone. However, when m=0m=0, different scalars would indeed make differences. For an example of this subtlety, see the end of Section 4.

Remark 1.4.

We also obtain a formula computing instanton knot homology KHI(Ym(K),Km)KHI(-Y_{-m}(K),K_{-m}) of the dual knot KmK_{-m} inside the resulting manifold in Theorem 3.22, which is analogous to the results by Eftekhary for knot Floer homology HFK^\widehat{HFK} [10, Proposition 1.5] and by Hedden-Levine [13]. In this formula, we may assume m=0m=0 because the scalar issue in Remark 1.3 does not appear.

With the isomorphisms in (1.2) and (1.1), we can truncate the above formula for I(Ym(K))I^{\sharp}(Y_{-m}(K)) to obtain the following exact triangle.

Corollary 1.5 (Generalized surgery exact triangle).

Suppose KYK\subset Y is a null-homologous knot, and mm is a fixed non-zero integer. Then for any sufficiently large integer kk, there exists an exact triangle

(1.3) I(Ymk(K))\textstyle{I^{\sharp}(-Y_{-m-k}(K))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}i=1kI(Y)\textstyle{\mathop{\bigoplus}_{i=1}^{k}I^{\sharp}(-Y)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}I(Ym(K))\textstyle{I^{\sharp}(-Y_{-m}(K))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
Remark 1.6.

The analogous result of the exact triangle (1.3) in Heegaard Floer theory was proved by Ozsváth-Szabó [27] using twisted coefficients, which is a crucial step towards proving the integral surgery formula in their setup. The proof cannot be applied to instanton theory directly. Thus, in this paper, we adopt a reversed approach: we use sutured instanton theory to prove Theorem 1.1 and derive Corollary 1.5 as a direct application. The strategy to prove Theorem 1.1 can be found in Section 3.1 and Section 3.2.

The analogs of π±,s\pi^{\pm,s} in Heegaard Floer theory can be interpreted as cobordism maps associated to some particular spinc structures. In instanton theory, there is a decomposition of cobordism maps along basic classes. However, currently such a decomposition is only known to exist for cobordisms whose first Betti number is zero. So for the moment let us assume the ambient 33-manifold YY is a rational homology sphere. For any integer mm, there is a natural cobordism WmW_{m} from Ym3(K)-Y^{3}_{-m}(K) to Y3-Y^{3}. From [9, Section 1.2], there exists a decomposition of the cobordism map I(Wm)I^{\sharp}(W_{m}) along basic classes

I(Wm)=sI(Wm,[s]),I^{\sharp}(W_{m})=\sum_{s\in\mathbb{Z}}I^{\sharp}(W_{m},[s]),

where [s]H2(W)[s]\in H^{2}(W) denotes the class that satisfies the equality

[s]([S¯])=2sm.[s]([\widebar{S}])=2s-m.

We make the following conjecture.

Conjecture 1.7.

Suppose KYK\subset Y is a null-homologous knot. Suppose b1(Y)=0b_{1}(Y)=0 and mm\in\mathbb{Z} with m2g(K)+1m\geq 2g(K)+1. Let AsA_{s}, BsB^{-}_{s}, π±,s\pi^{\pm,s}, WmW_{m}, I(Wm,[s])I^{\sharp}(W_{m},[s]) be defined as above. Then for any s[m12,1m2]s\in[\lfloor\frac{m-1}{2}\rfloor,\lfloor\frac{1-m}{2}\rfloor]\cap\mathbb{Z}, there are commutative diagrams

H(As)\textstyle{H(A_{s})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π,s\scriptstyle{\pi^{-,s}}I(Ym(K))\textstyle{I^{\sharp}(-Y_{-m}(K))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}I(Wm,[s])\scriptstyle{I^{\sharp}(W_{m},[s])}H(Bs)\textstyle{H(B_{s})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}I(Y)\textstyle{I^{\sharp}(-Y)}H(As)\textstyle{H(A_{s})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π+,s\scriptstyle{\pi^{+,s}}I(Ym(K))\textstyle{I^{\sharp}(-Y_{-m}(K))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}I(Wm,[s+m])\scriptstyle{I^{\sharp}(W_{m},[s+m])}H(Bs)\textstyle{H(B_{s})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}I(Y)\textstyle{I^{\sharp}(-Y)}
Remark 1.8.

In Heegaard Floer theory, the large surgery formula in [26, Theorem 4.1] states that the homology of the bent complex AsA_{s} is isomorphic to the Heegaard Floer homology of Ym(K)Y_{-m}(K) together with a spinc structure specified by ss. In instanton theory, we do not have the spinc structures in the construction of instanton Floer homology but a similar decomposition was introduced in [23, 25]. However, involving the spinc-type decomposition in the statement of Conjecture 1.7 would make the statement more complicated. So here we only write the top horizontal map in each commutative diagram as an inclusion.

The obstacle to obtaining a decomposition of the instanton cobordism map, in general, is one of the difficulties in exporting the original proof of the integral surgery formula in Heegaard Floer theory to the instanton setup. To overcome this problem, we need to work with a suitable setup for which some kind of decompositions do exist. A good candidate is sutured instanton theory. In sutured instanton theory, properly embedded surfaces induce \mathbb{Z}-gradings on the homology, and bypass maps relating different sutures are homogeneous with respect to such gradings. We have already used this setup to construct spinc-like decompositions for the framed instanton homology of Dehn surgeries of knots, construct bent complexes in instanton theory, and establish a large surgery formula in our previous work [23, 25, 22].

In this paper, to prove the integral surgery formula, we further study the relations between different sutures on the knot complement and establish some new exact triangles and commutative diagrams that may be of independent interest. Then these new and old algebraic structures relating to different sutures enable us to apply the octahedral lemma to prove the desired integral surgery formula. It is worth mentioning that ultimately the whole proof in the current paper depends only on some most fundamental properties of Floer theory: the surgery exact triangle, the functoriality of the cobordism maps, and the adjunction inequality. This implies that the existence of the surgery formula is an inherent property of Floer theory.

The surgery formula developed in the current paper is a powerful tool to study the Dehn surgeries along knots. It enables us to do explicit computations in many cases, even when the ambient 33-manifold has a positive first Betti number. In a companion paper [24], we will use the surgery formula and the techniques developed in this paper to derive many new applications and computations. We sketch the results as follows.

  1. (1)

    We study the behavior of the integral surgery formula under the connected sum with a core knot in a lens space (whose complement is a solid torus) and then derive a rational surgery formula for framed instanton homology.

  2. (2)

    We study the 0-surgery on a knot KK inside S3S^{3}. We derive a formula computing the non-zero grading part of I(S0(K))I^{\sharp}(S_{0}(K)) with respect to the grading induced by the Seifert surface.

  3. (3)

    We study non-zero integral surgeries on Boromean knots inside Y=#2gS1×S2Y=\#^{2g}S^{1}\times S^{2}, which gives nontrivial circle bundles over surfaces. In this case the bent complexes AsA_{s} and Bs±B^{\pm}_{s} can be computed directly and the maps π±\pi^{\pm} between them can be fixed with the help of the ΛH1(Y;)\Lambda^{*}H_{1}(Y;\mathbb{C})-action. Moreover, we show dimI(Y)=dim𝔽2HF^(Y)\dim_{\mathbb{C}}I^{\sharp}(Y)=\dim_{{\mathbb{F}_{2}}}\widehat{HF}(Y) for most Seifert fibered manifolds YY with non-zero orbifold degrees.

  4. (4)

    We study a family of alternating knots. Using an inductive argument by the oriented skein relation, we can describe their bent complexes explicitly and then the surgery formula applies routinely.

  5. (5)

    Using the same technique as above, we also study the non-zero integral surgery of twisted Whitehead doubles. The results for Whitehead doubles can also tell us the framed instanton Floer homology of the splicing of two knot complements in S3S^{3}, where one knot is a twist knot, i.e. the Whitehead double of the unknot.

  6. (6)

    We study almost L-space knots, i.e., a non-L-space knot KK such that there exists n+n\in\mathbb{N}_{+} with dimI(Sn3(K))=n+2\dim I^{\sharp}(S^{3}_{n}(K))=n+2 (see [6] for the results in Heegaard Floer theory). We prove a genus one almost LL-space knot is either the figure-eight or the mirror of the knot 525_{2}. We also show that almost LL-space knots of genus at least 22 are fibered and strongly quasi-positive.

Organization.

The paper is organized as follows. In Section 2, we introduce basic setup, the notations in sutured instanton homology, and deal with the scalar ambiguity mentioned in Remark 2.4. We also present some algebraic lemmas including the octahedral lemma in the derived category that are used in latter sections. In Section 3, we present the strategy to prove the integral surgery formula. We first restate the integral surgery formula using sutured instanton homology, and explain how to apply the octahedral lemma to prove it. Then we explain how to translate the integral surgery formula from the language of sutured instanton theory to the language of bent complexes, which coincides with the discussions in the introduction. All the rest of the sections are devoted to prove the three exact triangles and three commutative diagrams that are involved in the octahedral lemma, i.e., Equation (3.2) to Equation (3.7). In Section 4, we study the relation between the (1)(-1)-Dehn surgery map associated to a curve intersecting the suture twice and the two natural bypass maps associated to that curve. This helps us to prove Equation (3.2) and Equation (3.5). In Section 5, Equation (3.3), Equation (3.6) and part of Equation (3.4) are proved. The last two sections are devoted to prove Equation (3.4) and Equation (3.7), which is the most technical part of the paper. In Section 6 we prove some technical lemmas that are finally used in Section 7 to finish the proof.

Acknowledgement.

The authors thank John A. Baldwin and Steven Sivek for the discussion on the proof of Proposition 4.1, and thank Zekun Chen and Linsheng Wang for the discussion on homological algebra. The authors would like to thank Ciprian Manolescu and Jacob Rasmussen, and the anonymous referee for helpful comments. The authors also thank Sudipta Ghosh, Jianfeng Lin, Yi Xie and Ian Zemke for valuable discussions. The second author is also grateful to Yi Liu for inviting him to BICMR, Peking University when he was writing the early version of this paper.

2. Basic setup

2.1. Conventions

If it is not mentioned, all manifolds are smooth, oriented, and connected. Homology groups and cohomology groups are with \mathbb{Z} coefficients. We write n\mathbb{Z}_{n} for /n\mathbb{Z}/n\mathbb{Z} and 𝔽2\mathbb{F}_{2} for the field with two elements.

A knot KYK\subset Y is called null-homologous if it represents the trivial homology class in H1(Y;)H_{1}(Y;\mathbb{Z}), while it is called rationally null-homologous if it represents the trivial homology class in H1(Y;)H_{1}(Y;\mathbb{Q}).

For any oriented 3-manifold MM, we write M-M for the manifold obtained from MM by reversing the orientation. For any surface SS in MM and any suture γM\gamma\subset\partial M, we write SS and γ\gamma for the same surface and suture in M-M, without reversing their orientations. For a knot KK in a 3-manifold YY, we write (Y,K)(-Y,K) for the induced knot in Y-Y with induced orientation, called the mirror knot of KK. The corresponding balanced sutured manifold is (Y\N(K),γK)(-Y\backslash N(K),-\gamma_{K}).

2.2. Sutured instanton homology

For any balanced sutured manifold (M,γ)(M,\gamma) [15, Definition 2.2], Kronheimer-Mrowka [16, Section 7] constructed an isomorphism class of \mathbb{C}-vector spaces SHI(M,γ)SHI(M,\gamma). Later, Baldwin-Sivek [2, Section 9] dealt with the naturality issue and constructed (untwisted and twisted vesions of) projectively transitive systems related to SHI(M,γ)SHI(M,\gamma). We will use the twisted version, which we write as SHI¯(M,γ)\underline{\rm SHI}(M,\gamma) and call sutured instanton homology.

In this paper, when considering maps between sutured instanton homology, we can regard them as linear maps between actual vector spaces, at the cost that equations (or commutative diagrams) between maps only hold up to a non-zero scalar due to the projectivity. A more detailed discussion on the projectivity can be found in the next subsection.

Moreover, there is a relative 2\mathbb{Z}_{2}-grading on SHI¯(M,γ)\underline{\rm SHI}(M,\gamma) obtained from the construction of sutured instanton homology, which we consider as a homological grading and use to take Euler characteristic.

Definition 2.1.

Suppose KK is a knot in a closed 3-manifold YY. Let Y(1)\colonequalsY\B3Y(1)\colonequals Y\backslash B^{3} and let δ\delta be a simple closed curve on Y(1)S2\partial Y(1)\cong S^{2}. Let Y\N(K)Y\backslash N(K) be the knot complement and let Γμ\Gamma_{\mu} be two oppositely oriented meridians of KK on (Y\N(K))T2\partial(Y\backslash N(K))\cong T^{2}. Define

I(Y)\colonequalsSHI¯(Y(1),δ)andKHI¯(Y,K)\colonequalsSHI¯(Y\N(K),Γμ).I^{\sharp}(Y)\colonequals\underline{\rm SHI}(Y(1),\delta)~{}{\rm and}~{}\underline{\rm KHI}(Y,K)\colonequals\underline{\rm SHI}(Y\backslash N(K),\Gamma_{\mu}).
Remark 2.2.

By the naturality results, we should specify the places of the removing ball, the neighborhood of the knot, and the sutures to define I(Y)I^{\sharp}(Y) and KHI¯(Y,K)\underline{\rm KHI}(Y,K). These data can be fixed by choosing a basepoint in YY or KK. For simplicity, we omit those choices in the notations.

From now on, we will suppose KYK\subset Y is a rationally null-homologous knot and fix some notations. Let μ\mu be the meridian of KK and pick a longitude λ\lambda (such that λμ=1\lambda\cdot\mu=1) to fix a framing of KK. We will always assume Y\N(K)Y\backslash N(K) is irreducible, but many results still hold due to the following connected sum formula of sutured instanton homology [18, Remark 1.6]:

SHI¯(Y#Y\N(K),γ)I(Y)SHI¯(Y\N(K),γ).\underline{\rm SHI}(Y^{\prime}\#Y\backslash N(K),\gamma)\cong I^{\sharp}(Y^{\prime})\otimes\underline{\rm SHI}(Y\backslash N(K),\gamma).

Given coprime integers rr and ss, let Γr/s\Gamma_{r/s} be the suture on (Y\N(K))\partial(Y\backslash N(K)) consists of two oppositely oriented, simple closed curves of slope r/s-r/s, with respect to the chosen framing (i.e. the homology of the curves are ±(rμ+sλ)H1(N(K))\pm(-r\mu+s\lambda)\in H_{1}(\partial N(K))). Pick SS to be a minimal genus Seifert surface of KK, with genus g=g(S)g=g(S). Note that S\partial S may have multiple components.

Convention.

For a fixed pair (λ,μ)(\lambda,\mu) as above, we write p=Sμp=\partial S\cdot\mu and q=Sλq=\partial S\cdot\lambda. Note that when an orientation of the knot KK is chosen, the orientation of SS is induced by the knot. The orientation of μ\mu is chosen such that p>0p>0 and the orientation of λ\lambda is chosen such that λμ=1\lambda\cdot\mu=1. Note that pp is the order of [K]H1(Y)[K]\in H_{1}(Y), i.e., pp is the minimal positive integer satisfying p[K]=0H1(Y)p[K]=0\in H_{1}(Y). When KK is null-homologous, we always choose the Seifert framing λ=S\lambda=\partial S. In such case, we have (p,q)=(1,0)(p,q)=(1,0).

Remark 2.3.

The meanings of pp and qq above are different from our previous papers [23, 22]. Before, we used μ^\hat{\mu} and λ^\hat{\lambda} to denote the meridian of the knot KK and the preferred framing. In particular, the framing is fixed by [23, Definition 4.2]. Note that in that case, we assume that S\partial S is connected, and hence it is the same as the homological longitude (with the notation λ\lambda in previous papers, while we use μ\mu to denote the homological meridian). Also, the numbers pp and qq in this paper should be qq and q0q_{0} in previous papers.

For simplicity, we use the bold symbol of the suture to represent the sutured instanton homology of the balanced sutured manifold with the reversed orientation:

𝚪r/s\colonequalsSHI¯((Y\N(K)),Γr/s).\mathbf{\Gamma}_{r/s}\colonequals\underline{\rm SHI}(-(Y\backslash N(K)),-\Gamma_{r/s}).

When (r,s)=(±1,0)(r,s)=(\pm 1,0), we write Γr/s=Γμ\Gamma_{r/s}=\Gamma_{\mu}. When s=±1s=\pm 1, we write Γn=Γn/1=Γ(n)/(1)\Gamma_{n}=\Gamma_{n/1}=\Gamma_{(-n)/(-1)}. We also write 𝚪μ\mathbf{\Gamma}_{\mu} and 𝚪n\mathbf{\Gamma}_{n} for the corresponding sutured instanton homologies.

Remark 2.4.

Strictly speaking, the sutures corresponding to (r,s)=(1,0)(r,s)=(1,0) and (1,0)(-1,0) are not identical because the orientations are opposite. Since both sutures are on (Y\N(K))\partial(Y\backslash N(K)) of the same slope, they are isotopic. Moreover, we can choose a canonical isotopy by rotating the suture along the direction specified by the orientation of the knot. Due to discussion in Heegaard Floer theory [29, 32] and the conjectural relation between Heegaard Floer theory and instanton theory [16], it is expected that rotating the suture back to the original position induces a nontrivial isomorphism of the sutured instanton homology. So we pick the canonical isotopy to be the minimal rotation of the suture. Hence we can abuse notations and write Γμ\Gamma_{\mu} for both sutures. The same discussion also applies to the relation between Γn/1\Gamma_{n/1} and Γ(n)/(1)\Gamma_{(-n)/(-1)}.

We always assume that S\partial S has minimal intersections with Γr/s\Gamma_{r/s}, i.e. |Sγ|=2|rpsq||\partial S\cap\gamma|=2|rp-sq|. When the intersection number rpsqrp-sq is odd, then SS induces a \mathbb{Z}-grading on 𝚪r/s\mathbf{\Gamma}_{r/s}. When rpsqrp-sq is even, we need to perform either a positive stabilization or negative stabilization on SS to induce a \mathbb{Z}-grading, and the two gradings are related by an overall grading shift of 11. To get rid of stabilizations, we can equivalently regard that, in this case, the surface SS induces a (+12)(\mathbb{Z}+\frac{1}{2})-grading. We write the graded part of 𝚪r/s\mathbf{\Gamma}_{r/s} as

(𝚪r/s,i)\colonequalsSHI¯((Y\N(K)),Γr/s,S,i)(\mathbf{\Gamma}_{r/s},i)\colonequals\underline{\rm SHI}(-(Y\backslash N(K)),-\Gamma_{r/s},S,i)

with ii\in\mathbb{Z} or i+12i\in\mathbb{Z}+\frac{1}{2}, depending on the parity of the intersection number. From the construction of the grading in [20], we have the following vanishing theorem due to the adjunction inequality.

Lemma 2.5.

We have (𝚪r/s,i)=0(\mathbf{\Gamma}_{r/s},i)=0 when

|i|>|rpsq|χ(S)2.|i|>\frac{|rp-sq|-\chi(S)}{2}.
Proof.

This follows from [23, Theorem 2.21 (1)] (which is ultimately based on [16, Proposition 7.5]) and a direct computation in the new notations. ∎

The bypass exact triangle for sutured instanton homology was introduced by Baldwin-Sivek in [8, Section 4]. In [23, Section 4.2], we applied the triangle to sutures on knot complements and computed the grading shifts. We restate the results in the notation introduced before.

Lemma 2.6.

For any nn\in\mathbb{Z}, there are two graded bypass exact triangles

(𝚪n,i+p2)\textstyle{(\mathbf{\Gamma}_{n},i+\frac{p}{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n+1n\scriptstyle{\psi^{n}_{+,n+1}}(𝚪n+1,i)\textstyle{(\mathbf{\Gamma}_{n+1},i)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,μn+1\scriptstyle{\psi^{n+1}_{+,\mu}}(𝚪μ,inpq2)\textstyle{(\mathbf{\Gamma}_{\mu},i-\frac{np-q}{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,nμ\scriptstyle{\psi^{\mu}_{+,n}}
(𝚪n,ip2)\textstyle{(\mathbf{\Gamma}_{n},i-\frac{p}{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n+1n\scriptstyle{\psi^{n}_{-,n+1}}(𝚪n+1,i)\textstyle{(\mathbf{\Gamma}_{n+1},i)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,μn+1\scriptstyle{\psi^{n+1}_{-,\mu}}(𝚪μ,i+npq2)\textstyle{(\mathbf{\Gamma}_{\mu},i+\frac{np-q}{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,nμ\scriptstyle{\psi^{\mu}_{-,n}}

where the maps are homogeneous with respect to the homological 2\mathbb{Z}_{2}-gradings.

Proof.

This is [23, Proposition 4.14] in the new notations. The idea of the proof can be found in [23, Lemma 3.18] (see also [23, Remark 3.19]). Roughly, we perturb the surface SS by stabilizations so that its boundary is disjoint from the bypass arc. Then the grading shifts are obtained by counting the number of positive or negative stabilizations.

Unlike the setup in [23, Section 4], here KK is not necessarily a dual knot of the Dehn surgery on a null-homologous knot, so we adopt the remarks in the beginning of [23, Section 5]. For example, when nn is large enough so that npq0np-q\geq 0 and [23, Proposition 4.14 (1)] applies, we have

i^maxn=npqχ(S)2,i^minn=npqχ(S)2,i^maxμ=pχ(S)2,i^minμ=pχ(S)2,\hat{i}_{\max}^{n}=\frac{np-q-\chi(S)}{2},~{}\hat{i}_{\min}^{n}=-\frac{np-q-\chi(S)}{2},~{}\hat{i}_{\max}^{\mu}=\frac{p-\chi(S)}{2},~{}\hat{i}_{\min}^{\mu}=-\frac{p-\chi(S)}{2},

where we omit \lceil\cdot\rceil since we think about +12\mathbb{Z}+\frac{1}{2} if necessary. Then

i^minn+1i^minn=p2andi^maxn+1i^maxμ=npq2\hat{i}_{\min}^{n+1}-\hat{i}_{\min}^{n}=-\frac{p}{2}~{}{\rm and}~{}\hat{i}_{\max}^{n+1}-\hat{i}_{\max}^{\mu}=\frac{np-q}{2}

An easy way to understand the grading shift was described in [23, Remark 4.15]. Note that the grading shift of a map between two spaces equals half of the intersection number between S\partial S and the curve corresponding to the third space up to the sign, while the sign depends on the choice of the sign in the bypass map. For example, we have Sμ=p\partial S\cdot\mu=p, so the grading shifts of ψ±,n+1n\psi_{\pm,n+1}^{n} are p/2\mp p/2.

Remark 2.7.

The reason to use balanced sutured manifolds with reversed orientation is because of the above bypass exact triangles.

Remark 2.8.

If we do not mention gradings, the above triangles and the results in the rest of this subsection (except Corollary 2.9 and Lemma 2.19 since the statements involve gradings) also hold without the assumption that KK is rationally null homologous since the proofs only involve the neighborhood of (Y\N(K))\partial(-Y\backslash N(K)).

Corollary 2.9.

For any sufficiently large integer nn, we have the following properties for restrictions of maps.

  1. (1)

    The map ψ+,n+1n|(𝚪n,i)\psi^{n}_{+,n+1}|_{(\mathbf{\Gamma}_{n},i)} is an isomorphism when

    i<12(npq+χ(S)).i<\frac{1}{2}\bigg{(}np-q+\chi(S)\bigg{)}.
  2. (2)

    The map ψ,n+1n|(𝚪n,i)\psi^{n}_{-,n+1}|_{(\mathbf{\Gamma}_{n},i)} is an isomorphism when

    i>12(npq+χ(S))i>-\frac{1}{2}\bigg{(}np-q+\chi(S)\bigg{)}
  3. (3)

    For any grading ii such that

    12(npq+χ(S))<i<12((n2)pq+χ(S)),-\frac{1}{2}\bigg{(}np-q+\chi(S)\bigg{)}<i<\frac{1}{2}\bigg{(}(n-2)p-q+\chi(S)\bigg{)},

    there is an isomorphism

    (ψ+,n+1n)1ψ,n+1n:(𝚪n,i)(𝚪n,i+p).(\psi^{n}_{+,n+1})^{-1}\circ\psi^{n}_{-,n+1}:(\mathbf{\Gamma}_{n},i)\xrightarrow{\cong}(\mathbf{\Gamma}_{n},i+p).
  4. (4)

    The map ψ,1nn|(𝚪n,i)\psi^{-n}_{-,1-n}|_{(\mathbf{\Gamma}_{-n},i)} is an isomorphism when

    i<12((n2)p+q+χ(S))i<\frac{1}{2}\bigg{(}(n-2)p+q+\chi(S)\bigg{)}
  5. (5)

    The map ψ+,1nn|(𝚪n,i)\psi^{-n}_{+,1-n}|_{(\mathbf{\Gamma}_{-n},i)} is an isomorphism when

    i>12((n2)p+q+χ(S))i>-\frac{1}{2}\bigg{(}(n-2)p+q+\chi(S)\bigg{)}
  6. (6)

    For any grading ii such that

    12((n2)p+q+χ(S))<i<12((n4)p+q+χ(S))-\frac{1}{2}\bigg{(}(n-2)p+q+\chi(S)\bigg{)}<i<\frac{1}{2}\bigg{(}(n-4)p+q+\chi(S)\bigg{)}

    there is an isomorphism

    (ψ+,1nn)1ψ,1nn:(𝚪n,i)(𝚪n,i+p).(\psi^{-n}_{+,1-n})^{-1}\circ\psi^{-n}_{-,1-n}:(\mathbf{\Gamma}_{-n},i)\xrightarrow{\cong}(\mathbf{\Gamma}_{-n},i+p).
Proof.

It is a combination of Lemma 2.5 and Lemma 2.6. ∎

Definition 2.10.

The maps in Lemma 2.6 are called bypass maps. The ones with subscripts ++ and - are called positive and negative bypass maps, respectively. We will use ±\pm to denote one of the bypass maps. For any integer nn and any positive integer kk, define

Ψ±,n+kn\colonequalsψ±,n+kn+k1ψ±,n+1n:𝚪n𝚪n+k.\Psi_{\pm,n+k}^{n}\colonequals\psi_{\pm,n+k}^{n+k-1}\circ\cdots\circ\psi_{\pm,n+1}^{n}:\mathbf{\Gamma}_{n}\rightarrow\mathbf{\Gamma}_{n+k}.

In [23, Section 4.4], we proved many commutative diagrams for bypass maps, which we restate as follows by notations introduced before.

Lemma 2.11.

For any nn\in\mathbb{Z}, we have the following commutative diagrams up to scalars.

𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n+1n\scriptstyle{\psi^{n}_{+,n+1}}ψ,n+1n\scriptstyle{\psi^{n}_{-,n+1}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n+2n+1\scriptstyle{\psi^{n+1}_{-,n+2}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n+2n+1\scriptstyle{\psi^{n+1}_{+,n+2}}𝚪n+2\textstyle{\mathbf{\Gamma}_{n+2}}𝚪n+2\textstyle{\mathbf{\Gamma}_{n+2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,μn+2\scriptstyle{\psi^{n+2}_{+,\mu}}ψ,μn+2\scriptstyle{\psi^{n+2}_{-,\mu}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,nμ\scriptstyle{\psi^{\mu}_{+,n}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,nμ\scriptstyle{\psi^{\mu}_{-,n}}𝚪n\textstyle{\mathbf{\Gamma}_{n}}
Proof.

The first diagram follows from [23, Lemma 4.33]. Note that the proof only used the functionality of the contact gluing map and did not depend on the assumption that KK is rationally null-homologous. The second diagram is obtained from the first diagram by changing the choice of the framed knot. Explicitly, let KK^{\prime} be the dual knot corresponding to 𝚪n+1\mathbf{\Gamma}_{n+1}. Let μ=(n+1)μ+λ\mu^{\prime}=-(n+1)\mu+\lambda denote its meridian. Then λ=μ\lambda^{\prime}=-\mu is a framing of KK^{\prime}. Applying the first diagram to KK^{\prime}, we will obtain the second diagram for the original KK. Note that the sign of the bypass map may switch when we regard it as the bypass map for the original knot. That is the reason for the signs in the second diagram. This can be double-checked by keeping track of the grading shifts. ∎

Lemma 2.12.

For any nn\in\mathbb{Z}, we have the following commutative diagrams up to scalars

𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n+1n\scriptstyle{\psi^{n}_{+,n+1}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,nμ\scriptstyle{\psi^{\mu}_{-,n}}ψ,n+1μ\scriptstyle{\psi^{\mu}_{-,n+1}}
𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n+1n\scriptstyle{\psi^{n}_{-,n+1}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,nμ\scriptstyle{\psi^{\mu}_{+,n}}ψ+,n+1μ\scriptstyle{\psi^{\mu}_{+,n+1}}
𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n+1n\scriptstyle{\psi^{n}_{+,n+1}}ψ,μn\scriptstyle{\psi^{n}_{-,\mu}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,μn+1\scriptstyle{\psi^{n+1}_{-,\mu}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}}
𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n+1n\scriptstyle{\psi^{n}_{-,n+1}}ψ+,μn\scriptstyle{\psi^{n}_{+,\mu}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,μn+1\scriptstyle{\psi^{n+1}_{+,\mu}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}}

There are more bypass triangles involving more complicated sutures, which are obtained from changing the choice of the framed knot as in the proof of Lemma 2.11.

Lemma 2.13.

For a knot KYK\subset Y and nn\in\mathbb{Z}, there are two graded bypass exact triangles

(𝚪n1,i+npq2)\textstyle{(\mathbf{\Gamma}_{n-1},i+\frac{np-q}{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,2n12n1\scriptstyle{\psi^{n-1}_{+,\frac{2n-1}{2}}}(𝚪2n12,i)\textstyle{(\mathbf{\Gamma}_{\frac{2n-1}{2}},i)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n2n12\scriptstyle{\psi^{\frac{2n-1}{2}}_{+,n}}(𝚪n,i(n1)pq2)\textstyle{(\mathbf{\Gamma}_{n},i-\frac{(n-1)p-q}{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n1n\scriptstyle{\psi^{n}_{+,n-1}}
(𝚪n1,inpq2)\textstyle{(\mathbf{\Gamma}_{n-1},i-\frac{np-q}{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,2n12n1\scriptstyle{\psi^{n-1}_{-,\frac{2n-1}{2}}}(𝚪2n12,i)\textstyle{(\mathbf{\Gamma}_{\frac{2n-1}{2}},i)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n2n12\scriptstyle{\psi^{\frac{2n-1}{2}}_{-,n}}(𝚪n,i+(n1)pq2)\textstyle{(\mathbf{\Gamma}_{n},i+\frac{(n-1)p-q}{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n1n\scriptstyle{\psi^{n}_{-,n-1}}
Lemma 2.14.

For a knot KYK\subset Y and nn\in\mathbb{Z}, there are commutative diagrams up to scalars

𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n1μ\scriptstyle{\psi^{\mu}_{+,n-1}}ψ,n1μ\scriptstyle{\psi^{\mu}_{-,n-1}}𝚪n1\textstyle{\mathbf{\Gamma}_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,2n12n1\scriptstyle{\psi^{n-1}_{+,\frac{2n-1}{2}}}𝚪n1\textstyle{\mathbf{\Gamma}_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,2n12n1\scriptstyle{\psi^{n-1}_{-,\frac{2n-1}{2}}}𝚪2n12\textstyle{\mathbf{\Gamma}_{\frac{2n-1}{2}}}
𝚪2n12\textstyle{\mathbf{\Gamma}_{\frac{2n-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n2n12\scriptstyle{\psi^{\frac{2n-1}{2}}_{+,n}}ψ,n2n12\scriptstyle{\psi^{\frac{2n-1}{2}}_{-,n}}𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,μn\scriptstyle{\psi^{n}_{-,\mu}}𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,μn\scriptstyle{\psi^{n}_{+,\mu}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}}
𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n1μ\scriptstyle{\psi^{\mu}_{+,n-1}}𝚪n1\textstyle{\mathbf{\Gamma}_{n-1}}𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,μn\scriptstyle{\psi^{n}_{-,\mu}}ψ+,n1n\scriptstyle{\psi^{n}_{+,n-1}}
𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n1μ\scriptstyle{\psi^{\mu}_{-,n-1}}𝚪n1\textstyle{\mathbf{\Gamma}_{n-1}}𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,μn\scriptstyle{\psi^{n}_{+,\mu}}ψ,n1n\scriptstyle{\psi^{n}_{-,n-1}}
𝚪n1\textstyle{\mathbf{\Gamma}_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n12n12\scriptstyle{\psi^{\frac{2n-1}{2}}_{+,n-1}}ψ+,nn1\scriptstyle{\psi^{n-1}_{+,n}}𝚪2n12\textstyle{\mathbf{\Gamma}_{\frac{2n-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n2n12\scriptstyle{\psi^{\frac{2n-1}{2}}_{-,n}}𝚪n\textstyle{\mathbf{\Gamma}_{n}}
𝚪n1\textstyle{\mathbf{\Gamma}_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n12n12\scriptstyle{\psi^{\frac{2n-1}{2}}_{-,n-1}}ψ,nn1\scriptstyle{\psi^{n-1}_{-,n}}𝚪2n12\textstyle{\mathbf{\Gamma}_{\frac{2n-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n2n12\scriptstyle{\psi^{\frac{2n-1}{2}}_{+,n}}𝚪n\textstyle{\mathbf{\Gamma}_{n}}
Remark 2.15.

The choices of positive and negative bypass maps in Lemma 2.14 seem to be different from Lemma 2.11 and Lemma 2.12. But indeed they are the same up to changing the framed knot. In particular, the grading shifts match. Note that similar to the second diagram in Lemma 2.11, we always use the notations of the bypass maps for the original knot, while the signs may change if the maps are regarded as the bypass maps of the dual knot.

Suppose α\alpha is a connected non-separating simple closed curve on (Y\N(K))\partial(Y\backslash N(K)). We can push α\alpha into the interior of Y\N(K)Y\backslash N(K). For any fixed suture on (Y\N(K))\partial(Y\backslash N(K)) and a closure of the sutured manifold, the push-off of α\alpha is inside the closure, which is a closed 33-manifold. We can then take the framing on α\alpha induced by the surface (Y\N(K))\partial(Y\backslash N(K)) and there is an exact triangle associated to the instanton Floer homology of the (1)(-1)- 0- and \infty-surgeries along the push-off of α\alpha. Since the push-off of α\alpha is disjoint from (Y\N(K))\partial(Y\backslash N(K)), the exact triangle descends to one between corresponding sutured instanton Floer homologies.

According to [3, Section 4], when α\alpha intersects the suture at two points, the 0-surgery along the push-off of α\alpha (with framing induced by (Y\N(K))\partial(Y\backslash N(K))) corresponds to a 22-handle attachment along α\alpha. Note that attaching a 22-handle along α(Y\N(K))\alpha\subset\partial(Y\backslash N(K)) will change the 33-manifold from Y\N(K)Y\backslash N(K) to Yα(K)\B3Y_{\alpha}(K)\backslash B^{3}, where Yα(K)Y_{\alpha}(K) is the Dehn surgery along KK with slope specified by α\alpha. We write

𝐘𝐫/𝐬\colonequalsI(Yr/s(K)),\mathbf{Y_{r/s}}\colonequals I^{\sharp}(-Y_{-r/s}(K)),

and in particular

𝐘n\colonequalsI(Yn(K))and𝐘\colonequalsI(Y).\mathbf{Y}_{n}\colonequals I^{\sharp}(-Y_{-n}(K))~{}{\rm and}~{}\mathbf{Y}\colonequals I^{\sharp}(-Y).
Lemma 2.16 ([23, Lemma 3.21]).

For any nn\in\mathbb{Z}, we have the following exact triangles.

𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hn\scriptstyle{H_{n}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fn+1\scriptstyle{F_{n+1}}𝐘\textstyle{\mathbf{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gn\scriptstyle{G_{n}}
𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}An1\scriptstyle{A_{n-1}}𝚪n1\textstyle{\mathbf{\Gamma}_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Bn1\scriptstyle{B_{n-1}}𝐘n\textstyle{\mathbf{Y}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Cn\scriptstyle{C_{n}}
Proof.

To obtain the first exact triangle, we can take the sutured manifold ((Y\N(K)),Γn)(-(Y\backslash N(K)),-\Gamma_{n}), and take a meridian α(Y\N(K))\alpha\subset\partial(Y\backslash N(K)). As explained before the lemma, there is a surgery exact triangle associated to the sutured instanton Floer homology of the three sutured manifolds obtained by taking (1)(-1)-, 0-, and \infty-surgeries [30, Theorem 2.1]; see also [11] for the original construction and [8, Proof of Theorem 1.21, especially (16)-(19)] for the resolution of the subtlety of the bundle data.

The \infty-surgery will keep the manifold ((Y\N(K)),Γn)(-(Y\backslash N(K)),-\Gamma_{n}). The (1)(-1)-surgery changes the framing and hence we obtain ((Y\N(K)),Γn+1)(-(Y\backslash N(K)),-\Gamma_{n+1}). The 0-surgery, as discussed above, gives rise to the manifold Yα(K)\B3Y_{\alpha}(K)\backslash B^{3} which is Y\B3Y\backslash B^{3} since α\alpha is the meridian. Hence we obtain the expected triangle. The second exact triangle in the statement of the lemma is obtained similarly by taking α\alpha to be a curve on (Y\N(K))\partial(Y\backslash N(K)) having slope n-n instead of a meridian. ∎

Remark 2.17.

From [3, Section 4], we know the 0-surgery corresponds to a 22-handle attachment and a 11-handle attachment. Hence 𝐘\mathbf{Y} in the above lemma is indeed KHI¯(Y,U)\underline{\rm KHI}(-Y,U), where UU is the unknot in Y-Y bounding an embedded disk. By [3, Section 4], a 11-handle attachment does not change the closure of the balanced sutured manifold, and then there is a canonical identification between KHI¯(Y,U)\underline{\rm KHI}(-Y,U) and I(Y)I^{\sharp}(-Y). Hence we can abuse the notations. The same discussion also applies to 𝐘n\mathbf{Y}_{n}.

Furthermore, we proved the following properties in [23]. Note that the assumption that KK is the dual knot of a null-homologous knot in that paper is inessential by remarks in the beginning of [23, Section 5]. The inequalities of the gradings are from Corollary 2.9.

Lemma 2.18 ([23, Lemma 3.21 and Lemma 4.9]).

For any nn\in\mathbb{Z}, we have the following commutative diagrams up to scalars

𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n+1n\scriptstyle{\psi^{n}_{+,n+1}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}}𝐘\textstyle{\mathbf{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gn\scriptstyle{G_{n}}Gn+1\scriptstyle{G_{n+1}}
𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n+1n\scriptstyle{\psi^{n}_{-,n+1}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}}𝐘\textstyle{\mathbf{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gn\scriptstyle{G_{n}}Gn+1\scriptstyle{G_{n+1}}
𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n+1n\scriptstyle{\psi^{n}_{+,n+1}}Fn\scriptstyle{F_{n}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fn+1\scriptstyle{F_{n+1}}𝐘\textstyle{\mathbf{Y}}
𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n+1n\scriptstyle{\psi^{n}_{-,n+1}}Fn\scriptstyle{F_{n}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fn+1\scriptstyle{F_{n+1}}𝐘\textstyle{\mathbf{Y}}
Lemma 2.19 ([23, Lemma 4.17, Proposition 4.26, Lemma 4.29 and Proposition 4.32]).

Let FnF_{n} and GnG_{n} be defined as in Lemma 2.16. Then for any sufficiently large integer nn, we have the following properties.

  1. (1)

    The map Gn1G_{n-1} is zero and FnF_{n} is surjective. Moreover, for any grading

    (npq+χ(S))/2<i0<(npq+χ(S))/2p+1,-(np-q+\chi(S))/2<i_{0}<(np-q+\chi(S))/2-p+1,

    the restriction of the map

    Fn:i=0p1(𝚪n,i0+i)𝐘F_{n}:\bigoplus_{i=0}^{p-1}(\mathbf{\Gamma}_{n},i_{0}+i)\to\mathbf{Y}

    is an isomorphism.

  2. (2)

    The map Fn+1F_{-n+1} is zero and GnG_{-n} is injective. Moreover, for any grading

    ((n2)p+q+χ(S))/2<i0<((n2)p+q+χ(S))/2p+1,-((n-2)p+q+\chi(S))/2<i_{0}<((n-2)p+q+\chi(S))/2-p+1,

    the map

    ProjGn:𝐘i=0p1(𝚪n,i0+i),{\rm Proj}\circ G_{-n}:\mathbf{Y}\to\bigoplus_{i=0}^{p-1}(\mathbf{\Gamma}_{-n},i_{0}+i),

    is an isomorphism, where

    Proj:𝚪ni=0p1(𝚪n,i0+i){\rm Proj}:\mathbf{\Gamma}_{-n}\rightarrow\bigoplus_{i=0}^{p-1}(\mathbf{\Gamma}_{-n},i_{0}+i)

    is the projection.

The following lemma is a special case of Proposition 4.1, which we will prove later.

Lemma 2.20.

For any nn\in\mathbb{Z}, let the maps HnH_{n} and ψ±,n+1n\psi^{n}_{\pm,n+1} be defined as in Lemma 2.16 and Lemma 2.6 respectively. Then there exist c1,c2\{0}c_{1},c_{2}\in\mathbb{C}\backslash\{0\} such that

Hn=c1ψ+,n+1n+c2ψ,n+1nH_{n}=c_{1}\psi^{n}_{+,n+1}+c_{2}\psi^{n}_{-,n+1}

2.3. Fixing the scalars

By construction, sutured instanton homology forms a projectively transitive system, which means all the spaces and maps between spaces are well-defined only up to non-zero scalars. When the balanced sutured manifold is obtained from a framed knot as in the last subsection, we can make some canonical choices to reduce the projective ambiguities.

Suppose KYK\subset Y is a framed knot with the meridian μ\mu and the framing λ\lambda. Fix a knot complement Y\N(K)Y\backslash N(K) and the suture Γμ\Gamma_{\mu}. We fix a special choice of a (marked odd) closure of (Y\N(K),Γμ)(Y\backslash N(K),\Gamma_{\mu}) following the construction in [16, Formula (18)].

Let FF be a closed, oriented, connected surface of genus at least 22. Suppose c0c_{0} is a non-separating curve in FF. Let c=pt×c0S1×Fc=\operatorname{pt}\times c_{0}\subset S^{1}\times F and let (μc,λc)(\mu_{c},\lambda_{c}) be the meridian and the longitude of cc (the latter comes from the surface framing). Let

(2.1) (Y¯μ,R)=(S1×F\N(c)(μc,λc)(λ,μ)Y\N(K),pt×F),(\widebar{Y}_{\mu},R)=(S^{1}\times F\backslash N(c)\cup_{(\mu_{c},\lambda_{c})\sim(\lambda,\mu)}Y\backslash N(K),\operatorname{pt}^{\prime}\times F),

where pt\operatorname{pt}^{\prime} is a point different from pt\operatorname{pt}. We can pick α=S1×pt′′\alpha=S^{1}\times\operatorname{pt}^{\prime\prime} and ηR\eta\subset R be a curve intersecting pt×c0\operatorname{pt}^{\prime}\times c_{0} once. Since αR=1\alpha\cdot R=1 and ηR=0\eta\cdot R=0, the pair (Y¯μ,αη)(\widebar{Y}_{\mu},\alpha\cup\eta) defines an instanton Floer homology in the setting of [16, Section 7.1]. Moreover, 𝒟μ=(Y,R,η,α)\mathcal{D}_{\mu}=(Y,R,\eta,\alpha) forms a marked odd closure in [2, Definition 9.2], which was used in the naturality result [2, Theorem 9.17]. The reason for g(F)2g(F)\geq 2 is to apply the naturality result (c.f. [2, Remark 9.4]).

Similarly, for (Y\N(K),Γn)(Y\backslash N(K),\Gamma_{n}) and (Y\N(K),Γ2n12)(Y\backslash N(K),\Gamma_{\frac{2n-1}{2}}), we fix closures 𝒟n\mathcal{D}_{n} and 𝒟2n12\mathcal{D}_{\frac{2n-1}{2}} as in (2.1), except replacing the gluing map (μc,λc)(λ,μ)(\mu_{c},\lambda_{c})\sim(\lambda,\mu) by (μc,λc)(μ,nμ+λ)(\mu_{c},\lambda_{c})\sim(-\mu,-n\mu+\lambda) and (μc,λc)(nμ+λ,(12n)μ+2λ)(\mu_{c},\lambda_{c})\sim(-n\mu+\lambda,(1-2n)\mu+2\lambda), respectively.

For the sutured manifold (Y\B3,δ)(Y\backslash B^{3},\delta), we regard it as (Y\N(U),Γμ,U)(Y\backslash N(U),\Gamma_{\mu,U}) by Remark 2.17, where UU is the unknot and Γμ,U\Gamma_{\mu,U} is meridian suture on the unknot complement. Then we apply the above construction to obtain a special closure of (Y\B3,δ)(Y\backslash B^{3},\delta). We reverse the orientations of the chosen closures when the orientations of the sutured manifolds are reversed. Note that we do not choose canonical closures for (Yn(K)\B3,δ)(Y_{n}(K)\backslash B^{3},\delta) since we only care about the dimension of its framed instanton homology.

After fixing the choices of closures, we can view 𝚪n\mathbf{\Gamma}_{n} and 𝐘\mathbf{Y} as actual vector spaces, and then the elements in them are well-defined. Strictly speaking, we also need to choose extra data such as the metric and the perturbation on the closure to define the instanton Floer homology of the closure, but different choices of metrics and perturbations now induce a transitive system of vector spaces, from which we can construct an actual vector space. So, we omit the discussion on those extra data.

The construction of bypass maps and surgery maps may not be realized as cobordism maps between the chosen closures, but the construction of the projectively transitive system (c.f. [2, Definition 9.18]) guarantees the existence of such maps up to scalars. Now we make (non-canonical) choices of the maps to get rid of the scalar ambiguities in the commutative diagrams mentioned in the last subsection.

We first assume that I(Y)0I^{\sharp}(Y)\neq 0. When I(Y)=0I^{\sharp}(Y)=0, the first exact triangle in Lemma 2.16 is trivial for any nn and hence the maps FnF_{n} and GnG_{n} that play an important role in later sections are both zero. This makes fixing the scalars a somewhat straightforward job: we just need to fix the scalars for the bypass maps. Also, it is worth mentioning that, the Euler characteristic result in [30, Corollary 1.4] implies that I(Y)0I^{\sharp}(Y)\neq 0 for any rational homology sphere YY and, up to author’s knowledge, there is no known closed oriented 33-manifold YY with I(Y)=0I^{\sharp}(Y)=0.

To help us fixing the scalars, suppose the maps FnF_{n} and GnG_{n} are defined as in the proof of Lemma 2.16, and we define

(2.2) nG=min{n|Gn=0} and nF=max{n|Fn=0}.n_{G}=\min\{n\in\mathbb{Z}~{}|~{}G_{n}=0\}\text{ and }n_{F}=\max\{n\in\mathbb{Z}~{}|~{}F_{n}=0\}.

We have the following basic properties for these indices.

Lemma 2.21.

Assuming I(Y)0I^{\sharp}(Y)\neq 0. Suppose nGn_{G} and nFn_{F} are defined as in Equation (2.2). Then we have

<nFnG<.-\infty<n_{F}\leq n_{G}<\infty.

Moreover, we have Gn=0G_{n}=0 if and only if nnGn\geq n_{G} and Fn=0F_{n}=0 if and only if nnFn\leq n_{F}.

Proof.

The fact that <nF<-\infty<n_{F}<\infty and <nG<-\infty<n_{G}<\infty follow from Lemma 2.19 and the fact that they fits into an exact triangle as in Lemma 2.16. Next, the commutative diagrams in Lemma 2.18 implies that Gn+1=0G_{n+1}=0 whenever Gn=0G_{n}=0 and thus we know Gn=0G_{n}=0 if and only if nnGn\geq n_{G}. The argument for FnF_{n} is similar. Finally, by definition we know GnG=0G_{n_{G}}=0 and hence from the exact triangle we know that

ImFn+1=kerGnG=I(Y)0.\operatorname{Im}F_{n+1}=\ker G_{n_{G}}=I^{\sharp}(Y)\neq 0.

Hence we conclude that nFnGn_{F}\leq n_{G}. ∎

By Lemma 2.19, we can pick a sufficiently large integer n0n_{0} such that n0<nFnG-n_{0}<n_{F}\leq n_{G}. Pick arbitrary representatives of the maps

Gn0,ψ+,μn0,ψ,μn0,ψ+,n0μ,ψ,n0μG_{-n_{0}},\psi^{-n_{0}}_{+,\mu},\psi^{-n_{0}}_{-,\mu},\psi^{\mu}_{+,-n_{0}},\psi^{\mu}_{-,-n_{0}}

and we also pick arbitrary representatives of the maps

ψ+,n+1n,ψ+,2n12n1\psi^{n}_{+,n+1},\psi^{n-1}_{+,\frac{2n-1}{2}}

for all nn\in\mathbb{Z}.

Now we explain how to fix the scalars for maps with nn0n\geq n_{0}. Note that we have already chosen a representative for ψ+,n0+1n0\psi^{-n_{0}}_{+,-n_{0}+1} and Gn0G_{-n_{0}}. From Lemma 2.18, we have

Gn0+1ψ+,n0+1n0Gn0G_{-n_{0}+1}\doteq\psi^{-n_{0}}_{+,-n_{0}+1}\circ G_{-n_{0}}

where \doteq means commutative up to a non-zero scalar. We can choose a representative of Gn0+1G_{-n_{0}+1} to obtain an equality

Gn0+1=ψ+,n0+1n0Gn0G_{-n_{0}+1}=\psi^{-n_{0}}_{+,-n_{0}+1}\circ G_{-n_{0}}

We then choose a representative of ψ,n0+1n0\psi^{-n_{0}}_{-,-n_{0}+1} so that

Gn0+1=ψ,n0+1n0Gn0.G_{-n_{0}+1}=\psi^{-n_{0}}_{-,-n_{0}+1}\circ G_{-n_{0}}.

Next, we pick representatives of the maps ψ,n+1n\psi^{n}_{-,n+1} inductively, with the base case n=n0n=-n_{0} constructed above, so that

ψ,n+1nψ+,nn1=ψ+,n+1nψ,nn1\psi^{n}_{-,n+1}\circ\psi^{n-1}_{+,n}=\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n}

hold for all nn0+1n\geq-n_{0}+1. If the compositions happen to be zero, we could pick an arbitrary representative since the diagram will be trivially satisfied. We will discuss the ambiguity arising from the possibility that ψ+,n+1nψ,nn1=0\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n}=0 more carefully later.

Similarly, we pick the maps ψ+,μn\psi^{n}_{+,\mu} inductively to satisfy the commutative diagram

ψ+,μnψ,nn1=ψ+,μn1.\psi^{n}_{+,\mu}\circ\psi^{n-1}_{-,n}=\psi^{n-1}_{+,\mu}.

We can choose representatives of ψ,μn\psi^{n}_{-,\mu}, ψ+,nμ\psi^{\mu}_{+,n}, and ψ,nμ\psi^{\mu}_{-,n} in a similar manner.

Furthermore, the representatives of

ψ,n2n12,ψ+,n1n,ψ,n1n,ψ,2n12n1,ψ+,n2n12\psi^{\frac{2n-1}{2}}_{-,n},\psi^{n}_{+,n-1},\psi^{n}_{-,n-1},\psi^{n-1}_{-,\frac{2n-1}{2}},\psi^{\frac{2n-1}{2}}_{+,n}

can be chosen according to Lemma 2.14. As mentioned in Remark 2.15, we always use the notations of the bypass maps for the original knot even though we consider some dual knots in the proofs. Hence here we first fix the knot KYK\subset Y and then fix the representatives, while we do not fix the representatives by any commutative diagrams for the dual knot in the proofs.

Next, we deal with maps GnG_{n} and FnF_{n} in Lemma 2.16. We choose representatives of the maps GnG_{n} inductively so that

Gn+1=ψ+,n+1nGnG_{n+1}=\psi^{n}_{+,n+1}\circ G_{n}

is satisfied for all nn0n\geq-n_{0}. We pick arbitrary representatives of the map FnF+1F_{n_{F}+1} and then pick FnF_{n} inductively so that

Fn+1ψ+,n+1n=FnF_{n+1}\circ\psi^{n}_{+,n+1}=F_{n}

is satisfied for all nnF+1n\geq n_{F}+1. We can then use induction to prove the following two equalities.

  • Gn+1=ψ,n+1nGnG_{n+1}=\psi^{n}_{-,n+1}\circ G_{n} for all nn0n\geq-n_{0}, and

  • Fn+1ψ,n+1n=cFnF_{n+1}\circ\psi^{n}_{-,n+1}=c\cdot F_{n} for a non-zero scalar cc that is independent of nn with nnF+1n\geq n_{F}+1.

We verify the equality for GG first. The base case n=n0n=n_{0} is by construction. Assuming we have already established the equality for nn, from Lemma 2.11 and Lemma 2.18, we have

Gn+2\displaystyle G_{n+2} =ψ+,n+2n+1Gn+1\displaystyle=\psi^{n+1}_{+,n+2}\circ G_{n+1}
(Inductive hypothesis)\displaystyle(\text{Inductive hypothesis}) =ψ+,n+2n+1ψ,n+1nGn\displaystyle=\psi^{n+1}_{+,n+2}\circ\psi^{n}_{-,n+1}\circ G_{n}
(Lemma 2.11)\displaystyle(\text{Lemma }\ref{lem: com diag for n,n+1,n+2}) =ψ,n+2n+1ψ+,n+1nGn\displaystyle=\psi^{n+1}_{-,n+2}\circ\psi^{n}_{+,n+1}\circ G_{n}
=ψ,n+2n+1Gn+1.\displaystyle=\psi^{n+1}_{-,n+2}\circ G_{n+1}.

The argument for FnF_{n} is similar, once we take c0c\neq 0 to be the complex number such that

(2.3) FnF+2ψ,nF+2nF+1=cFnF+1.F_{n_{F}+2}\circ\psi^{n_{F}+1}_{-,n_{F}+2}=c\cdot F_{n_{F}+1}.

The remaining issues are summarized as follows:

  • (i)

    When choosing representatives of ψ,n+1n\psi^{n}_{-,n+1}, we use the commutative diagram

    ψ,n+1nψ+,nn1ψ+,n+1nψ,nn1\psi^{n}_{-,n+1}\circ\psi^{n-1}_{+,n}\doteq\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n}

    However, when ψ+,n+1nψ,nn1=0\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n}=0, there is no unique choice of ψ,n+1n\psi^{n}_{-,n+1} and one might worry that different choices of ψ,n+1n\psi^{n}_{-,n+1} may affect the commutative diagrams in Lemma 2.18.

  • (ii)

    By Proposition 4.1, we can assume that the map HnH_{n} in the exact triangle in Lemma 2.16 to have the form

    Hn=ψ+,n+1ncnψ,n+1n.H_{n}=\psi^{n}_{+,n+1}-c_{n}\cdot\psi^{n}_{-,n+1}.

    We want to pin down the values of cnc_{n}.

  • (iii)

    We want to get rid of the scalar cc in Equation 2.3.

We treat these issues in several different cases.

Case 1. nF+3nGn_{F}+3\leq n_{G}. In this case, we know that

ψ+,n+1nψ,nn10\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n}\neq 0

for any integer nn. Indeed, if n+1<nGn+1<n_{G}, we know from Lemma 2.18 that

ψ+,n+1nψ,nn1Gn1=Gn+10.\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n}\circ G_{n-1}=G_{n+1}\neq 0.

If n+1nGnF+3n+1\geq n_{G}\geq n_{F}+3, instead of the above equation involving GG, we have

Fn+1ψ+,n+1nψ,nn1=cFn10.F_{n+1}\circ\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n}=c\cdot F_{n-1}\neq 0.

This implies that inductively we can fix a unique representative of ψ,n+1n\psi^{n}_{-,n+1} for all nn0+1n\geq-n_{0}+1.

For any integer nn with n0n<nG1-n_{0}\leq n<n_{G}-1, we have Gn+10G_{n+1}\neq 0. Then we take an element α𝐘\alpha\in\mathbf{Y} so that Gn+1(α)0G_{n+1}(\alpha)\neq 0. Then we can solve the scalar cnc_{n} as follows.

0\displaystyle 0 =HnGn(α)\displaystyle=H_{n}\circ G_{n}(\alpha)
=(ψ+,n+1ncnψ,n+1n)Gn(α)\displaystyle=(\psi^{n}_{+,n+1}-c_{n}\cdot\psi^{n}_{-,n+1})\circ G_{n}(\alpha)
=ψ+,n+1nGn(α)cnψ,n+1nGn(α)\displaystyle=\psi^{n}_{+,n+1}\circ G_{n}(\alpha)-c_{n}\cdot\psi^{n}_{-,n+1}\circ G_{n}(\alpha)
=(1cn)Gn+1(α)\displaystyle=(1-c_{n})\cdot G_{n+1}(\alpha)

Hence we conclude that cn=1c_{n}=1. In particular, we can take n=nF+1<nG1n=n_{F}+1<n_{G}-1. Note Fn0F_{n}\neq 0 so we can take x𝚪nx\in\mathbf{\Gamma}_{n} so that Fn(x)0F_{n}(x)\neq 0. Then we have

0\displaystyle 0 =Fn+1Hn(x)\displaystyle=F_{n+1}\circ H_{n}(x)
=Fn+1(ψ+,n+1nψ,n+1n)(x)\displaystyle=F_{n+1}\circ(\psi^{n}_{+,n+1}-\psi^{n}_{-,n+1})(x)
=Fn+1ψ+,n+1n(x)Fn+1ψ,n+1n(x)\displaystyle=F_{n+1}\circ\psi^{n}_{+,n+1}(x)-F_{n+1}\circ\psi^{n}_{-,n+1}(x)
=(1c)Fn(x)\displaystyle=(1-c)\cdot F_{n}(x)

This implies that c=1c=1 as well. Now for any nnG1nF+2n\geq n_{G}-1\geq n_{F}+2, we can take x𝚪nx\in\mathbf{\Gamma}_{n} such that Fn(x)0F_{n}(x)\neq 0. we have

0\displaystyle 0 =Fn+1Hn(x)\displaystyle=F_{n+1}\circ H_{n}(x)
=Fn+1(ψ+,n+1ncnψ,n+1n)(x)\displaystyle=F_{n+1}\circ(\psi^{n}_{+,n+1}-c_{n}\cdot\psi^{n}_{-,n+1})(x)
=(1cn)Fn(x)\displaystyle=(1-c_{n})\cdot F_{n}(x)

Hence we conclude that cn=1c_{n}=1. In summary, in Case 1, we have the following.

  • We can fix a unique representative of ψ,n+1n\psi^{n}_{-,n+1} for any nn0n\geq-n_{0}.

  • We have cn=1c_{n}=1 for any nn0n\geq-n_{0}.

  • We have c=1c=1.

Case 2. nG=nF+2n_{G}=n_{F}+2. Note that some arguments in Case 1 still apply. We summarize as follows.

  • For n<nG1n<n_{G}-1, we have Gn+10G_{n+1}\neq 0 so there is a unique choice of ψ,n+1n\psi^{n}_{-,n+1}.

  • For nnG=nF+2n\geq n_{G}=n_{F}+2, we have Fn10F_{n-1}\neq 0 so again there is a unique choice of ψ,n+1n\psi^{n}_{-,n+1}.

  • For n<nG1n<n_{G}-1, we have Gn+10G_{n+1}\neq 0 so cn=1c_{n}=1 in the expression of HnH_{n}.

  • For nnG1=nF+1n\geq n_{G}-1=n_{F}+1, we have Fn0F_{n}\neq 0 so cn=c1c_{n}=c^{-1}.

To resolve the issue (i), the only nonfixed index is n=nF+1=nG1n=n_{F}+1=n_{G}-1. In case

ψ+,n+1nψ,nn10,\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n}\neq 0,

there is a unique choice of ψ,n+1n\psi^{n}_{-,n+1} so that

ψ,n+1nψ+,nn1=ψ+,n+1nψ,nn1.\psi^{n}_{-,n+1}\circ\psi^{n-1}_{+,n}=\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n}.

Otherwise, we just fix any representative of ψ,n+1n\psi^{n}_{-,n+1}.

To resolve issues (ii) and (iii), we rescale the maps FnF_{n} according to the grading on 𝚪n\mathbf{\Gamma}_{n}. To do this, for an integer nn and a grading ii, define

j(n,i)=ipn2.j(n,i)=\lfloor\frac{i}{p}-\frac{n}{2}\rfloor.

It is straightforward that

j(n+1,i+p2)=j(n,i) and j(n+1,ip2)=j(n,i)1.j(n+1,i+\frac{p}{2})=j(n,i)\text{ and }j(n+1,i-\frac{p}{2})=j(n,i)-1.

Hence we define

(2.4) F~n=icj(n,i)FnProjni,\widetilde{F}_{n}=\sum_{i}c^{j(n,i)}\cdot F_{n}\circ\text{Proj}^{i}_{n},

where the map

Projni:𝚪n(𝚪n,i)\text{Proj}^{i}_{n}:\mathbf{\Gamma}_{n}\to(\mathbf{\Gamma}_{n},i)

is the projection. Equivalently, if we have an element x(𝚪n,i)x\in(\mathbf{\Gamma}_{n},i), we take

F~n(x)=cj(n,i)Fn(x).\widetilde{F}_{n}(x)=c^{j(n,i)}\cdot F_{n}(x).

We then need to verify the following two equalities for F~\widetilde{F}:

  • F~n+1ψ+,n+1n=F~n\widetilde{F}_{n+1}\circ\psi^{n}_{+,n+1}=\widetilde{F}_{n}, and

  • F~n+1ψ,n+1n=F~n\widetilde{F}_{n+1}\circ\psi^{n}_{-,n+1}=\widetilde{F}_{n}.

For the first one, note that ψ+,n+1n\psi^{n}_{+,n+1} increases the grading by p/2p/2 from Lemma 2.6. So assume x(𝚪n,i)x\in(\mathbf{\Gamma}_{n},i), we have

F~n+1ψ+,n+1n(x)\displaystyle\widetilde{F}_{n+1}\circ\psi^{n}_{+,n+1}(x) =cj(n+1,i+p2)Fn+1(x)ψ+,n+1n\displaystyle=c^{j(n+1,i+\frac{p}{2})}\cdot F_{n+1}(x)\circ\psi^{n}_{+,n+1}
=cj(n,i)Fn(x)\displaystyle=c^{j(n,i)}\cdot F_{n}(x)
=F~n(x)\displaystyle=\widetilde{F}_{n}(x)

On the other hand, the map ψ,n+1n\psi^{n}_{-,n+1} decreases the grading by p/2p/2 from Lemma 2.6, and so for x(𝚪n,i)x\in(\mathbf{\Gamma}_{n},i),

F~n+1ψ,n+1n(x)\displaystyle\widetilde{F}_{n+1}\circ\psi^{n}_{-,n+1}(x) =cj(n+1,ip2)Fn+1(x)ψ,n+1n\displaystyle=c^{j(n+1,i-\frac{p}{2})}\cdot F_{n+1}(x)\circ\psi^{n}_{-,n+1}
=cj(n,i)1cFn(x)\displaystyle=c^{j(n,i)-1}\cdot c\cdot F_{n}(x)
=F~n(x)\displaystyle=\widetilde{F}_{n}(x)

Hence we can use F~n\widetilde{F}_{n} instead of FnF_{n} to get rid of the scalar cc. However, changing FnF_{n} to F~n\widetilde{F}_{n}, we might break the exact triangle in Lemma 2.16. Hence we need to alter the definition of HnH_{n} as well. We define

(2.5) H~n=ψ+,n+1nψ,n+1n\widetilde{H}_{n}=\psi^{n}_{+,n+1}-\psi^{n}_{-,n+1}

and it remains to establish the following exact triangle:

𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H~n\scriptstyle{\widetilde{H}_{n}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F~n+1\scriptstyle{\widetilde{F}_{n+1}}𝐘\textstyle{\mathbf{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gn\scriptstyle{G_{n}}

When n<nG2=nFn<n_{G}-2=n_{F}, the triangle automatically holds. Indeed, for such nn, we have cn=1c_{n}=1 so H~n=Hn\widetilde{H}_{n}=H_{n}; the fact Fn+1=0F_{n+1}=0 implies that F~n+1=0\widetilde{F}_{n+1}=0, so the new exact triangle is exactly the original one in Lemma 2.16.

When n=nFn=n_{F}, we still have cn=1c_{n}=1 and hence H~n=Hn\widetilde{H}_{n}=H_{n}. So the exactness at 𝚪n\mathbf{\Gamma}_{n} is from Lemma 2.16. By construction we have ImF~n+1=ImFn+1\operatorname{Im}\widetilde{F}_{n+1}=\operatorname{Im}F_{n+1} for any nn so we conclude the exactness at 𝐘\mathbf{Y}. This also implies that

dimkerF~n+1=dimkerFn+1=dimImHn.\dim\ker\widetilde{F}_{n+1}=\dim\ker F_{n+1}=\dim\operatorname{Im}{H}_{n}.

Hence to show the exactness at 𝚪n+1\mathbf{\Gamma}_{n+1}, it remains to show that

ImHnkerF~n+1.\operatorname{Im}H_{n}\subset\ker\widetilde{F}_{n+1}.

For any x𝚪nx\in\mathbf{\Gamma}_{n}, we have

F~n+1Hn(x)\displaystyle\widetilde{F}_{n+1}\circ H_{n}(x) =\displaystyle=
=F~n+1(ψ+,n+1nψ,n+1n)(x)\displaystyle=\widetilde{F}_{n+1}\circ(\psi^{n}_{+,n+1}-\psi^{n}_{-,n+1})(x)
=F~n+1ψ+,n+1n(x)F~n+1ψ,n+1n(x)\displaystyle=\widetilde{F}_{n+1}\circ\psi^{n}_{+,n+1}(x)-\widetilde{F}_{n+1}\circ\psi^{n}_{-,n+1}(x)
=F~n(x)F~n(x)\displaystyle=\widetilde{F}_{n}(x)-\widetilde{F}_{n}(x)
=0.\displaystyle=0.

Hence we are done.

Finally, we verify the exact triangle for nnF+1n\geq n_{F}+1. Define a homomorphism

ιn:𝚪n𝚪n\iota_{n}:\mathbf{\Gamma}_{n}\to\mathbf{\Gamma}_{n}

as

ιn(x)=icj(n,i)Projni(x).\iota_{n}(x)=\sum_{i}c^{-j(n,i)}\text{Proj}^{i}_{n}(x).

It is clear that ι\iota is an isomorphism as its inverse is

ιn1(x)=icj(n,i)Projni(x),\iota_{n}^{-1}(x)=\sum_{i}c^{j(n,i)}\text{Proj}^{i}_{n}(x),

and from the construction of F~\widetilde{F} in Equation (2.4), we have

ιn+1(kerF~n+1)=kerFn+1.\iota_{n+1}(\ker\widetilde{F}_{n+1})=\ker F_{n+1}.

From the fact that Hn=ψ+,n+1nc1ψ,n+1nH_{n}=\psi^{n}_{+,n+1}-{c}^{-1}\cdot\psi^{n}_{-,n+1} (we have cn=c1c_{n}=c^{-1}), H~n=ψ+,n+1nψ,n+1n\widetilde{H}_{n}=\psi^{n}_{+,n+1}-\psi^{n}_{-,n+1}, and that ψ±,n+1n\psi^{n}_{\pm,n+1} shift the grading by p/2\mp p/2, we conclude that

ιn+1(ImH~n)=ImHn.\iota_{n+1}(\operatorname{Im}\widetilde{H}_{n})=\operatorname{Im}H_{n}.

As a result, we conclude the exactness at Γn+1\Gamma_{n+1}. The exactness at 𝐘\mathbf{Y} holds as above, since we still have

ImF~n+1=ImFn+1=kerGn.\operatorname{Im}\widetilde{F}_{n+1}=\operatorname{Im}F_{n+1}=\ker G_{n}.

Dimension counting similar to the above argument then implies that

dimkerH~n=dimImGn.\dim\ker\widetilde{H}_{n}=\dim\operatorname{Im}G_{n}.

We then verify that

ImGnkerH~n.\operatorname{Im}G_{n}\subset\ker\widetilde{H}_{n}.

For any α𝐘\alpha\in\mathbf{Y}, we have

H~nGn(α)\displaystyle\widetilde{H}_{n}\circ G_{n}(\alpha) =(ψ+,n+1nψ,n+1n)Gn(α)\displaystyle=(\psi^{n}_{+,n+1}-\psi^{n}_{-,n+1})\circ G_{n}(\alpha)
=ψ+,n+1nGn(α)ψ,n+1nGn(α)\displaystyle=\psi^{n}_{+,n+1}\circ G_{n}(\alpha)-\psi^{n}_{-,n+1}\circ G_{n}(\alpha)
=Gn+1(α)Gn+1(α)\displaystyle=G_{n+1}(\alpha)-G_{n+1}(\alpha)
=0.\displaystyle=0.

In summary, in Case 2, we did the following (extra things):

  • We choose representatives of ψ,n+1n\psi^{n}_{-,n+1} for all nn0+1n\geq-n_{0}+1 so that

    ψ,n+1nψ+,nn1=ψ+,n+1nψ,nn1.\psi^{n}_{-,n+1}\circ\psi^{n-1}_{+,n}=\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n}.
  • We define new maps F~n\widetilde{F}_{n} for all nn so that

    (2.6) F~n=F~n+1ψ±,n+1n.\widetilde{F}_{n}=\widetilde{F}_{n+1}\circ\psi^{n}_{\pm,n+1}.
  • We define new maps H~n\widetilde{H}_{n} for all nn so that we have the following exact triangle for all nn.

    (2.7) 𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H~n\scriptstyle{\widetilde{H}_{n}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F~n+1\scriptstyle{\widetilde{F}_{n+1}}𝐘\textstyle{\mathbf{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gn\scriptstyle{G_{n}}

Case 3. nG=nF+1n_{G}=n_{F}+1 or nG=nFn_{G}=n_{F}. The situation and argument are similar to those in Case 2. We summarize the differences here:

  • In Case 3, the composition ψ+,n+1nψ,nn1\psi^{n}_{+,n+1}\circ\psi^{n-1}_{-,n} could only be zero when nG1nnF+1n_{G}-1\leq n\leq n_{F}+1. In case the composition is indeed 0, we choose an arbitrary representative of the map ψ,n+1n\psi^{n}_{-,n+1}.

  • We still define the maps F~n\widetilde{F}_{n} as in Equation (2.4) and the maps H~n\widetilde{H}_{n} as in Equation (2.5), and can verify the Equation (2.6) and the exact triangle (2.7) as in Case 2.

Note from Lemma 2.21, we must have nGnFn_{G}\geq n_{F}, so the above three cases cover all situations.

Finally, we could extend the choice of representatives for all relevant maps for the indices n<n0n<-n_{0}. Note that when n<n0n<-n_{0}, we have that GnG_{n} is injective and Fn=0F_{n}=0. Hence we do not need to worry about the issues (i), (ii), and (iii).

Convention.

From now on, we write the maps H~n\widetilde{H}_{n} and F~n\widetilde{F}_{n} simply as HnH_{n} and FnF_{n}, respectively. From the above discussion, when KYK\subset Y is a fixed rationally null-homologous knot, we can assume the first commutative diagram in Lemma 2.11 and all commutative diagrams in Lemma 2.12, Lemma 2.14 and Lemma 2.18 hold without introducing scalars.

2.4. Algebraic lemmas

In this subsection, we introduce some lemmas in homological algebra. All graded vector spaces in this subsection are finite dimensional and over \mathbb{C} and all maps are complex linear maps. For convenience, we will switch freely between long exact sequences and exact triangles.

From Section 2.2, we know the sutured instanton homology is usually 2\mathbb{Z}\oplus\mathbb{Z}_{2}-graded, where we regard the 2\mathbb{Z}_{2}-grading as a homological grading. Many results in this subsection come from properties of the derived category of vector spaces over \mathbb{C}, for which people usually consider cochain complexes. However, for a 2\mathbb{Z}_{2}-graded space there is no difference between the chain complex and the cochain complex. Hence by saying a complex we mean a 2\mathbb{Z}_{2}-graded (co)chain complex, though all results apply to \mathbb{Z}-graded cochain complexes verbatim.

For a complex CC and an integer nn, we write CnC^{n} for its grading nn part (under the natural map 2\mathbb{Z}\to\mathbb{Z}_{2}). With this notation, we suppose the differential dCd_{C} on CC sends CnC^{n} to Cn+1C^{n+1}. For any integer kk, we write C{k}C\{k\} for the complex obtained from CC by the grading shift C{k}n=Cn+kC\{k\}^{n}=C^{n+k}. We write H(C,dC)H(C,d_{C}) or H(C)H(C) for the homology of a complex CC with differential dCd_{C}. A 2\mathbb{Z}_{2}-graded vector space is regarded as a complex with the trivial differential.

For a chain map f:CDf:C\to D, we write Cone(f)\operatorname{Cone}(f) for the mapping cone of ff, i.e., the complex consisting of the space DC{1}D\oplus C\{1\} and the differential

dCone(f)\colonequals[dDf0dC].d_{\operatorname{Cone}(f)}\colonequals\begin{bmatrix}d_{D}&-f\\ 0&-d_{C}\end{bmatrix}.

Then there is a long exact sequence

H(C)𝑓H(D)𝑖H(Cone(f))𝑝H(C){1}\cdots\to H(C)\xrightarrow{f}H(D)\xrightarrow{i}H(\operatorname{Cone}(f))\xrightarrow{p}H(C)\{1\}\to\cdots

where ii sends xDx\in D to (x,0)(x,0) and pp sends (x,y)DC{1}(x,y)\in D\oplus C\{1\} to y-y. If the differentials of CC and DD are trivial, then we know

(2.8) H(Cone(f))ker(f)coker(f).H(\operatorname{Cone}(f))\cong\ker(f)\oplus\operatorname{coker}(f).
Remark 2.22.

Our definitions about mapping cones follow from [31], which are different from those in [27, 28].

Note that the derived category is a triangulated category, so it satisfies the octahedral lemma (for example, see [31, Proposition 10.2.4]).

Lemma 2.23 (octahedral lemma).

Suppose X,Y,Z,X,Y,ZX,Y,Z,X^{\prime},Y^{\prime},Z^{\prime} are 2\mathbb{Z}_{2}-graded vector spaces satisfying the following long exact sequences

\displaystyle\cdots X𝑓YZX{1}\displaystyle\to X\xrightarrow{f}Y\xrightarrow{h}Z^{\prime}\to X\{1\}\to\cdots
\displaystyle\cdots XgfZhYlX{1}\displaystyle\to X\xrightarrow{g\circ f}Z\xrightarrow{h^{\prime}}Y^{\prime}\xrightarrow{l^{\prime}}X\{1\}\to\cdots
\displaystyle\cdots Y𝑔ZX𝑙Y{1}\displaystyle\to Y\xrightarrow{g}Z\to X^{\prime}\xrightarrow{l}Y\{1\}\to\cdots

Then we have the fourth long exact sequence

Z𝜓YϕXh{1}lZ{1}\cdots\to Z^{\prime}\xrightarrow{\psi}Y^{\prime}\xrightarrow{\phi}X^{\prime}\xrightarrow{h\{1\}\circ l}Z^{\prime}\{1\}\to\cdots

such that the following diagram commutes

(2.9) Y\textstyle{Y^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕ\scriptstyle{\phi}Z\textstyle{Z^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\scriptstyle{\psi}X\textstyle{X^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h{1}l\scriptstyle{h\{1\}\circ l}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}gf\scriptstyle{g\circ f}Z\textstyle{Z\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Y\textstyle{Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}

where we omit the grading shifts and the notation for the maps h,l,jh,l,j. We can also write (2.9) in another form so that there is enough room to write the maps

(2.10) Z\textstyle{Z^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\scriptstyle{\psi}X{1}\textstyle{X\{1\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f{1}\scriptstyle{f\{1\}}Y\textstyle{Y\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}g\scriptstyle{g}h\scriptstyle{h}Y\textstyle{Y^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕ\scriptstyle{\phi}l\scriptstyle{l^{\prime}}Z\textstyle{Z\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h^{\prime}}Y{1}\textstyle{Y\{1\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h{1}\scriptstyle{h\{1\}}X\textstyle{X\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f\scriptstyle{f}gf\scriptstyle{g\circ f}X\textstyle{X^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h{1}l\scriptstyle{h\{1\}\circ l}l\scriptstyle{l}Z{1}\textstyle{Z^{\prime}\{1\}}

The maps ψ\psi and ϕ\phi in (2.10) can be written explicitly as follows. By the long exact sequences in the assumption of Lemma 2.23, we know that Z,X,YZ^{\prime},X^{\prime},Y^{\prime} are chain homotopic to the mapping cones Cone(f),Cone(g),Cone(gf)\operatorname{Cone}(f),\operatorname{Cone}(g),\operatorname{Cone}(g\circ f), respectively. Under such homotopies, we can write

ψ:YX{1}\displaystyle\psi:Y\oplus X\{1\} ZX{1}\displaystyle\to Z\oplus X\{1\}
ψ(y,x)\displaystyle\psi(y,x) (g(y),x)\displaystyle\mapsto(g(y),x)

and

ϕ:ZX{1}\displaystyle\phi:Z\oplus X\{1\} ZY{1}\displaystyle\to Z\oplus Y\{1\}
ϕ(z,x)\displaystyle\phi(z,x) (z,f{1}(x))\displaystyle\mapsto(z,f\{1\}(x))

However, the chain homotopies are not canonical, and hence the maps ψ\psi and ϕ\phi are also not canonical. Thus, usually we cannot identify them with other given maps. Fortunately, with an extra \mathbb{Z}-grading, we may identify H(Cone(ϕ))H(\operatorname{Cone}(\phi)) with H(Cone(ϕ))H(\operatorname{Cone}(\phi^{\prime})) for another map ϕ:YX\phi^{\prime}:Y^{\prime}\to X^{\prime}.

First, we introduce the following lemma to deal with the projectivity of maps (i.e. maps well-defined only up to scalars). Note that the \mathbb{Z}-grading in the following lemma is not the homological grading used before.

Lemma 2.24.

Suppose XX and YY are \mathbb{Z}-graded vector spaces and suppose f,g:XYf,g:X\to Y are homogeneous maps with different grading shifts k1k_{1} and k2k_{2}. Then Cone(f+g)\operatorname{Cone}(f+g) is isomorphic to Cone(c1f+c2g)\operatorname{Cone}(c_{1}f+c_{2}g) for any c1,c2\{0}c_{1},c_{2}\in\mathbb{C}\backslash\{0\}.

Proof.

For simplicity, we can suppose k1=0k_{1}=0 and k2=1k_{2}=1. The proof for the general case is similar. For i,ji,j\in\mathbb{Z}, we write (X,i)(X,i) and (Y,j)(Y,j) for grading summands of XX and YY, respectively. Suppose TT is an automorphism of XYX\oplus Y that acts by

c2ic1i+1Id on (X,i) and c2jc1jId on (Y,j).\frac{c_{2}^{i}}{c_{1}^{i+1}}{\rm Id}\text{ on }(X,i)\text{ and }\frac{c_{2}^{j}}{c_{1}^{j}}{\rm Id}\text{ on }(Y,j).

Then TT is an isomorphism between Cone(f+g)\operatorname{Cone}(f+g) and Cone(c1f+c2g)\operatorname{Cone}(c_{1}f+c_{2}g). ∎

Then we state the lemma that relates the map ϕ\phi in Lemma 2.23 to another map ϕ\phi^{\prime} constructed explicitly.

Lemma 2.25.

Suppose X,Y,Z,X,YX,Y,Z,X^{\prime},Y^{\prime} are 2\mathbb{Z}\oplus\mathbb{Z}_{2}-graded vector spaces satisfying the following horizontal exact sequences.

Z\textstyle{Z\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}h\scriptstyle{h^{\prime}}=\scriptstyle{=}Y\textstyle{Y^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l^{\prime}}ϕ\scriptstyle{\phi}ϕ=a+b\scriptstyle{\phi^{\prime}=a^{\prime}+b^{\prime}}X{1}\textstyle{X\{1\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}f{1}=a+b\scriptstyle{f\{1\}=a+b}Z\textstyle{Z\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕh=ϕh\scriptstyle{\phi\circ h^{\prime}=\phi^{\prime}\circ h^{\prime}}X\textstyle{X^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l}Y{1}\textstyle{Y\{1\}}

where the shift {1}\{1\} is for the 2\mathbb{Z}_{2}-grading. Suppose ϕ:YX\phi:Y^{\prime}\to X^{\prime} satisfies the two commutative diagrams and suppose ϕ:YX\phi^{\prime}:Y^{\prime}\to X^{\prime} satisfies the left commutative diagram. Suppose ll and ll^{\prime} are homogeneous with respect to the \mathbb{Z}-grading. Suppose f{1}=a+bf\{1\}=a+b and ϕ=a+b\phi^{\prime}=a^{\prime}+b^{\prime} are sums of homogeneous maps with different grading shifts with respect to the \mathbb{Z}-grading. Moreover, suppose the following diagrams hold up to scalars.

Y\textstyle{Y^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l^{\prime}}a\scriptstyle{a^{\prime}}X{1}\textstyle{X\{1\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}a\scriptstyle{a}X\textstyle{X^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l}Y{1}\textstyle{Y\{1\}}
Y\textstyle{Y^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l^{\prime}}b\scriptstyle{b^{\prime}}X{1}\textstyle{X\{1\}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}b\scriptstyle{b}X\textstyle{X^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l}Y{1}\textstyle{Y\{1\}}

Then there is an isomorphism between H(Cone(ϕ))H(\operatorname{Cone}(\phi)) and H(Cone(ϕ))H(\operatorname{Cone}(\phi^{\prime})).

Proof.

Since ϕ\phi and ϕ\phi^{\prime} share the same domain and codomain, it suffices to show that they have the same rank. Fix inner products on YY^{\prime} and XX^{\prime} such that we have orthogonal decompositions

Y=Im(h)Y′′andX=Im(ϕh)X′′.Y^{\prime}=\operatorname{Im}(h^{\prime})\oplus Y^{\prime\prime}~{}{\rm and}~{}X^{\prime}=\operatorname{Im}(\phi\circ h^{\prime})\oplus X^{\prime\prime}.

By commutativity, we know both ϕ\phi and ϕ\phi^{\prime} send Im(h)\operatorname{Im}(h^{\prime}) onto Im(ϕh)\operatorname{Im}(\phi\circ h^{\prime}). Hence if we choose bases with respect to the decompositions so that linear maps are represented by matrices (we use row vectors), then we have

ϕ=[AB0C]andϕ=[AB0C],\phi=\begin{bmatrix}A&B\\ 0&C\end{bmatrix}~{}{\rm and}~{}\phi^{\prime}=\begin{bmatrix}A^{\prime}&B^{\prime}\\ 0&C^{\prime}\end{bmatrix},

where A=A:Im(h)Im(ϕh)A=A^{\prime}:\operatorname{Im}(h^{\prime})\to\operatorname{Im}(\phi\circ h^{\prime}) has full row rank. Then it suffices to show CC and CC^{\prime} have the same row rank.

By the exactness at YY^{\prime} and XX^{\prime}, we know the restriction of ll^{\prime} on Y′′Y^{\prime\prime} is an isomorphism between Y′′Y^{\prime\prime} and Im(l)\operatorname{Im}(l^{\prime}) and the restriction of ll on X′′X^{\prime\prime} is an isomorphism between X′′X^{\prime\prime} and Im(l)\operatorname{Im}(l). By commutativity, we know that both aa and bb send Im(l)\operatorname{Im}(l^{\prime}) to Im(l)\operatorname{Im}(l) and

rowrank(C)=rank(f{1}|Im(l))androwrank(C)=rank((c1a+c2b)|Im(l))\operatorname{rowrank}(C)=\operatorname{rank}(f\{1\}|{\operatorname{Im}(l^{\prime})})~{}{\rm and}~{}\operatorname{rowrank}(C^{\prime})=\operatorname{rank}((c_{1}a+c_{2}b)|{\operatorname{Im}(l^{\prime})})

for some c1,c2\{0}c_{1},c_{2}\in\mathbb{C}\backslash\{0\}. Since ll and ll^{\prime} are homogeneous, there exist induced \mathbb{Z}-gradings on Im(l)\operatorname{Im}(l) and Im(l)\operatorname{Im}(l^{\prime}). The maps aa and bb are still homogeneous with different grading shifts with respect to these induced gradings. Then we can apply Lemma 2.24 to obtain that the ranks of the restrictions of f{1}=a+bf\{1\}=a+b and c1a+c2bc_{1}a+c_{2}b on Im(l)\operatorname{Im}(l^{\prime}) are the same. ∎

3. Integral surgery formulae

3.1. A formula for framed instanton homology

In this subsection, we propose an integral surgery formula based on sutured instanton homology and package it into the language of bent complexes in a later subsection.

Suppose KYK\subset Y is a framed rationally null-homologous knot, and we adopt the notations introduced in Section 2.2. Define

πm,k±\colonequalsΨ±,m1+2km+kψ,m+k2m+2k12:𝚪2m+2k12𝚪m+2k1\pi^{\pm}_{m,k}\colonequals\Psi_{\pm,m-1+2k}^{m+k}\circ\psi^{\frac{2m+2k-1}{2}}_{\mp,m+k}:\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\rightarrow\mathbf{\Gamma}_{m+2k-1}

and write πm,k±,i\pi^{\pm,i}_{m,k} as the restriction of πm,k±\pi^{\pm}_{m,k} on (𝚪2m+2k12,i)(\mathbf{\Gamma}_{\frac{2m+2k-1}{2}},i). From Lemma 2.13 and Lemma 2.6, we can verify that πm,k±\pi^{\pm}_{m,k} shifts grading by ±(mpq)/2\pm(mp-q)/2, and then the integral surgery formula can be stated as follow.

Theorem 3.1.

Suppose mm is a fixed integer such that mpq0mp-q\neq 0. Then for any sufficiently large integer kk, there exists an exact triangle

𝚪2m+2k12\textstyle{\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πm,k++πm,k\scriptstyle{\pi_{m,k}^{+}+\pi_{m,k}^{-}}𝚪m+2k1\textstyle{\mathbf{\Gamma}_{m+2k-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐘m\textstyle{\mathbf{Y}_{m}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

Hence we have an isomorphism

𝐘mH(Cone(πm,k++πm,k)).\mathbf{Y}_{m}\cong H(\operatorname{Cone}(\pi_{m,k}^{+}+\pi_{m,k}^{-})).
Remark 3.2.

Let μ\mu and λ\lambda represent the meridian and the longitude of the knot KK, respectively. Then, mpq0mp-q\neq 0 is equivalent to the fact that mμ+λ-m\mu+\lambda is not isotopic to a connected component of the boundary of the Seifert surface. Specifically, if KK is null-homologous, we must have m0m\neq 0.

In the rest of this subsection and in the next subsection, we state the strategy to prove Theorem 3.1, and defer the proofs of some propositions to the remaining sections. An important step is to apply the octahedral axiom mentioned in Section 2.4 to the following diagram:

(3.1) 𝚪2m+2k12\textstyle{\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝐘m\textstyle{\mathbf{Y}_{m}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪m1+2k\textstyle{\mathbf{\Gamma}_{m-1+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪m1+k𝚪m1+k\textstyle{\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪m1\textstyle{\mathbf{\Gamma}_{m-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

To obtain the dotted exact triangle, we need to establish the following three exact triangles:

(3.2) I(Sm3(K))\textstyle{I^{\sharp}(-S^{3}_{-m}(K))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪m1\textstyle{\mathbf{\Gamma}_{m-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
(3.3) 𝚪2m+2k12\textstyle{\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪m1+k𝚪m1+k\textstyle{\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
(3.4) 𝚪m1+2k\textstyle{\mathbf{\Gamma}_{m-1+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪m1+k𝚪m1+k\textstyle{\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪m1\textstyle{\mathbf{\Gamma}_{m-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

and establish the following commutative diagram:

(3.5) 𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪m1+k𝚪m1+k\textstyle{\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}}𝚪m1\textstyle{\mathbf{\Gamma}_{m-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

The octahedral lemma then implies the existence of the dotted triangle and ensure that all diagrams in (3.1) other than exact triangles commute.

We will then use Lemma 2.25 to identify the map coming from the octahedral lemma with πm,k++πm,k\pi_{m,k}^{+}+\pi_{m,k}^{-}. We also require the following two extra diagrams to commute, where the maps other than πm,k++πm,k\pi_{m,k}^{+}+\pi_{m,k}^{-} come from (3.1).

(3.6) 𝚪2m+2k12\textstyle{\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πm,k++πm,k\scriptstyle{\pi^{+}_{m,k}+\pi^{-}_{m,k}}𝚪m1+2k\textstyle{\mathbf{\Gamma}_{m-1+2k}}𝚪m1+k𝚪m1+k\textstyle{\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
(3.7) 𝚪2m+2k12\textstyle{\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πm,k++πm,k\scriptstyle{\pi^{+}_{m,k}+\pi^{-}_{m,k}}𝚪m1+2k\textstyle{\mathbf{\Gamma}_{m-1+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪μ\textstyle{\quad\quad\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\quad\quad}𝚪m1\textstyle{\mathbf{\Gamma}_{m-1}}

Indeed, by applying Lemma 2.25, it suffices to prove some weaker commutative diagrams involving πm,k±\pi^{\pm}_{m,k} separately.

3.2. A strategy of the proof

In this subsection, we provide more details of the strategy mentioned in Section 3.1. For simplicity, we fix the scalar ambiguities of commutative diagrams as in Section 2.3. To write down the maps, we redraw the octahedral diagram (3.1) as follows:

(3.8) 𝐘m\textstyle{\mathbf{Y}_{m}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ\scriptstyle{\psi}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,m1μ+ψ,m1μ\scriptstyle{\begin{aligned} \psi_{+,m-1}^{\mu}\\ +\psi_{-,m-1}^{\mu}\end{aligned}}𝚪m1\textstyle{\mathbf{\Gamma}_{m-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}(Ψ+,m1+km1,Ψ,m1+km1)\scriptstyle{\begin{aligned} (\Psi_{+,m-1+k}^{m-1},\\ \Psi_{-,m-1+k}^{m-1})\end{aligned}}𝚪2m+2k12\textstyle{\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ϕ\scriptstyle{\phi}l\scriptstyle{l^{\prime}}𝚪m1+k𝚪m1+k\textstyle{\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ,m1+2km1+kΨ+,m1+2km1+k\scriptstyle{\begin{aligned} \Psi_{-,m-1+2k}^{m-1+k}\\ -\Psi_{+,m-1+2k}^{m-1+k}\end{aligned}}h\scriptstyle{h^{\prime}}𝚪m1\textstyle{\mathbf{\Gamma}_{m-1}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,m1μ+ψ,m1μ\scriptstyle{\begin{aligned} \psi_{+,m-1}^{\mu}\\ +\psi_{-,m-1}^{\mu}\end{aligned}}(ψ,m1+kμ,ψ+,m1+kμ)\scriptstyle{\begin{aligned} (\psi_{-,m-1+k}^{\mu},\\ \psi_{+,m-1+k}^{\mu})\end{aligned}}𝚪m1+2k\textstyle{\mathbf{\Gamma}_{m-1+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l}

where

h=ψ,2m+2k12m+k1ψ+,2m+2k12m+k1.h^{\prime}=\psi_{-,\frac{2m+2k-1}{2}}^{m+k-1}-\psi_{+,\frac{2m+2k-1}{2}}^{m+k-1}.

The reader can compare (3.8) with (2.9) and (2.10). We omit the term corresponding to Z{1}Z^{\prime}\{1\} because there is not enough room, and the maps involving it are not important in our proof.

The first exact sequence of (3.8)

(3.9) 𝚪μψ+,m1μ+ψ,m1μ𝚪m1𝐘m𝚪μ\mathbf{\Gamma}_{\mu}\xrightarrow{\psi_{+,m-1}^{\mu}+\psi_{-,m-1}^{\mu}}\mathbf{\Gamma}_{m-1}\xrightarrow{}\mathbf{Y}_{m}\xrightarrow{}\mathbf{\Gamma}_{\mu}

follows from the second exact triangle in Lemma 2.16. Though the map Am1A_{m-1} may not be the same as the sum ψ+,m1μ+ψ,m1n\psi_{+,m-1}^{\mu}+\psi_{-,m-1}^{n}, we can use the following proposition and Lemma 2.24 (another special case of Proposition 4.1) to identify Cone(Am1)\operatorname{Cone}(A_{m-1}) with Cone(ψ+,m1μ+ψ,m1n)\operatorname{Cone}(\psi_{+,m-1}^{\mu}+\psi_{-,m-1}^{n}). Here we use the assumption that mpq0mp-q\neq 0.

Proposition 3.3.

Suppose An1A_{n-1} is the map in Lemma 2.16. For any integer nn, there exist scalars c1,c2\{0}c_{1},c_{2}\in\mathbb{C}\backslash\{0\} such that

An1=c1ψ+,n1μ+c2ψ,n1μ.A_{n-1}=c_{1}\psi_{+,n-1}^{\mu}+c_{2}\psi_{-,n-1}^{\mu}.

The exactness at

𝚪m1+k𝚪m1+k\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}

in the second and the third exact sequences are both special cases of the following proposition, which will be proved in Section 5.1 by diagram chasing.

Proposition 3.4.

Fixing the scalars as in Section 2.3, and given nn\in\mathbb{Z} and k0+k_{0}\in\mathbb{N}_{+}. Then, for any c1,c2,c3,c4c_{1},c_{2},c_{3},c_{4} satisfying the equation

c1c3=c2c4,c_{1}c_{3}=-c_{2}c_{4},

the following sequence is exact

𝚪n(c1Ψ+,n+k0n,c2Ψ,n+k0n)𝚪n+k0𝚪n+k0c3Ψ,n+2k0n+k0+c4Ψ+,n+2k0n+k0𝚪n+2k0\mathbf{\Gamma}_{n}\xrightarrow{(c_{1}\Psi_{+,n+k_{0}}^{n},c_{2}\Psi_{-,n+k_{0}}^{n})}\mathbf{\Gamma}_{n+k_{0}}\oplus\mathbf{\Gamma}_{n+k_{0}}\xrightarrow{c_{3}\Psi_{-,n+2k_{0}}^{n+k_{0}}+c_{4}\Psi_{+,n+2k_{0}}^{n+k_{0}}}\mathbf{\Gamma}_{n+2k_{0}}
Remark 3.5.

The exactness at the direct summand for the second exact sequence (the one involving 𝚪μ\mathbf{\Gamma}_{\mu}) might not be as clear from Proposition 3.4. Explicitly, we apply the proposition to the dual knot KK^{\prime} corresponding to Γm+k\Gamma_{m+k} with framing λ=μ\lambda^{\prime}=-\mu and n=0,k0=1n=0,k_{0}=1.

The exactness at 𝚪μ\mathbf{\Gamma}_{\mu} and 𝚪2m+2k12\mathbf{\Gamma}_{\frac{2m+2k-1}{2}} in the second exact sequence of (3.8)

(3.10) 𝚪μ(ψ,m1+kμ,ψ+,m1+kμ)𝚪m1+k𝚪m1+kψ,2m+2k12m+k1ψ+,2m+2k12m+k1𝚪2m+2k12l𝚪μ\mathbf{\Gamma}_{\mu}\xrightarrow{(\psi_{-,m-1+k}^{\mu},\psi_{+,m-1+k}^{\mu})}\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}\xrightarrow{\psi_{-,\frac{2m+2k-1}{2}}^{m+k-1}-\psi_{+,\frac{2m+2k-1}{2}}^{m+k-1}}\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\xrightarrow{l^{\prime}}\mathbf{\Gamma}_{\mu}

will also be proved by diagram chasing. We can explicitly construct the map ll^{\prime} by the composition of bypass maps

l\colonequalsψ,μm+kψ+,m+k2m+2k12=ψ+,μm+kψ,m+k2m+2k12,l^{\prime}\colonequals\psi_{-,\mu}^{m+k}\circ\psi_{+,m+k}^{\frac{2m+2k-1}{2}}=\psi_{+,\mu}^{m+k}\circ\psi_{-,m+k}^{\frac{2m+2k-1}{2}},

where the last equation follows from Lemma 2.14 and the conventions in Section 2.3. The following proposition will be proved in Section 5.2 by diagram chasing.

Proposition 3.6.

Suppose ll^{\prime} is constructed as above. For any c1,c2,c3,c4\{0}c_{1},c_{2},c_{3},c_{4}\in\mathbb{C}\backslash\{0\}, the following sequence is exact

𝚪m1+k𝚪m1+kc3ψ,2m+2k12m+k1+c4ψ+,2m+2k12m+k1𝚪2m+2k12l𝚪μ\displaystyle\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}\xrightarrow{c_{3}\psi_{-,\frac{2m+2k-1}{2}}^{m+k-1}+c_{4}\psi_{+,\frac{2m+2k-1}{2}}^{m+k-1}}\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\xrightarrow{l^{\prime}}\mathbf{\Gamma}_{\mu}
(c1ψ,m1+kμ,c2ψ+,m1+kμ)𝚪m1+k𝚪m1+k\displaystyle\xrightarrow{(c_{1}\psi_{-,m-1+k}^{\mu},c_{2}\psi_{+,m-1+k}^{\mu})}\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}
Remark 3.7.

In the proof of [22, Theorem 3.23], we obtained a long exact sequence

𝚪μ(ψ+,n1μ,ψ,n1μ)𝚪n1𝚪n1𝚪2n12𝚪μ\mathbf{\Gamma}_{\mu}\xrightarrow{(\psi_{+,n-1}^{\mu},\psi_{-,n-1}^{\mu})}\mathbf{\Gamma}_{n-1}\oplus\mathbf{\Gamma}_{n-1}\xrightarrow{}\mathbf{\Gamma}_{\frac{2n-1}{2}}\xrightarrow{}\mathbf{\Gamma}_{\mu}

by the octahedral lemma. However, we did not know the two maps involving 𝚪2n12\mathbf{\Gamma}_{\frac{2n-1}{2}} explicitly. Thus, the second exact sequence here is stronger than the one from octahedral lemma.

Remark 3.8.

The reason that Proposition 3.6 holds for any choices of c1,c2,c3,c4c_{1},c_{2},c_{3},c_{4} is because

ker((c1ψ,m1+kμ,c2ψ+,m1+kμ))=ker(c1ψ,m1+kμ)ker(c2ψ+,m1+kμ)\ker((c_{1}\psi_{-,m-1+k}^{\mu},c_{2}\psi_{+,m-1+k}^{\mu}))=\ker(c_{1}\psi_{-,m-1+k}^{\mu})\cap\ker(c_{2}\psi_{+,m-1+k}^{\mu})

and

Im(c3ψ,2m+2k12m+k1+c4ψ+,2m+2k12m+k1)=Im(c3ψ,2m+2k12m+k1)+Im(c4ψ+,2m+2k12m+k1),\operatorname{Im}(c_{3}\psi_{-,\frac{2m+2k-1}{2}}^{m+k-1}+c_{4}\psi_{+,\frac{2m+2k-1}{2}}^{m+k-1})=\operatorname{Im}(c_{3}\psi_{-,\frac{2m+2k-1}{2}}^{m+k-1})+\operatorname{Im}(c_{4}\psi_{+,\frac{2m+2k-1}{2}}^{m+k-1}),

where the right hand sides of the equations are independent of scalars.

The exactness at 𝚪m1\mathbf{\Gamma}_{m-1} and 𝚪m1+2k\mathbf{\Gamma}_{m-1+2k} in the third exact sequence of (3.8)

(3.11) 𝚪m1(Ψ+,m1+km1,Ψ,m1+km1)𝚪m1+k𝚪m1+kΨ,m1+2km1+kΨ+,m1+2km1+k𝚪m1+2k𝑙𝚪m1\mathbf{\Gamma}_{m-1}\xrightarrow{(\Psi_{+,m-1+k}^{m-1},\Psi_{-,m-1+k}^{m-1})}\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}\xrightarrow{\Psi_{-,m-1+2k}^{m-1+k}-\Psi_{+,m-1+2k}^{m-1+k}}\mathbf{\Gamma}_{m-1+2k}\xrightarrow{l}\mathbf{\Gamma}_{m-1}

is harder to prove since the map ll cannot be constructed by bypass maps. We expect that there are many equivalent constructions of ll and we will use the one for which the exactness is easiest to prove. Even so, we only prove the exactness with the assumption that kk is large. See Section 7.2 for more details.

Proposition 3.9.

Suppose c1,c2,c3,c4\{0}c_{1},c_{2},c_{3},c_{4}\in\mathbb{C}\backslash\{0\} and suppose k0k_{0} is a large integer. For any nn\in\mathbb{Z}, there exists a map ll such that the following sequence is exact

𝚪n+k0𝚪n+k0c3Ψ,n+2k0n+k0+c4Ψ+,n+2k0n+k0𝚪n+2k0𝑙𝚪n(c1Ψ+,n+k0n,c2Ψ,n+k0n)𝚪n+k0𝚪n+k0\mathbf{\Gamma}_{n+k_{0}}\oplus\mathbf{\Gamma}_{n+k_{0}}\xrightarrow{c_{3}\Psi_{-,n+2k_{0}}^{n+k_{0}}+c_{4}\Psi_{+,n+2k_{0}}^{n+k_{0}}}\mathbf{\Gamma}_{n+2k_{0}}\xrightarrow{l}\mathbf{\Gamma}_{n}\xrightarrow{(c_{1}\Psi_{+,n+k_{0}}^{n},c_{2}\Psi_{-,n+k_{0}}^{n})}\mathbf{\Gamma}_{n+k_{0}}\oplus\mathbf{\Gamma}_{n+k_{0}}
Remark 3.10.

In the first arXiv version of this paper, we only proved Proposition 3.9 for knots in S3S^{3} because we had to use the fact that dimI(S3)=1\dim_{\mathbb{C}}I^{\sharp}(-S^{3})=1 and S3S^{3} has an orientation-reversing involution. The construction of ll for knots in general 33-manifolds is inspired by the original proof for S3S^{3} and the proof in Section 7 is a generalization of the previous proof.

Remark 3.11.

For the same reason as in Remark 3.8, the coefficients in Proposition 3.9 are not important.

Then we consider the commutative diagrams mentioned in Section 3.1. By Lemma 2.6 and Lemma 2.12, we have

(Ψ+,m1+km1,Ψ,m1+km1)(ψ+,m1μ+ψ,m1μ)=(ψ,m1+kμ,ψ+,m1+kμ),(\Psi_{+,m-1+k}^{m-1},\Psi_{-,m-1+k}^{m-1})\circ(\psi_{+,m-1}^{\mu}+\psi_{-,m-1}^{\mu})=(\psi_{-,m-1+k}^{\mu},\psi_{+,m-1+k}^{\mu}),

which verifies the commutative diagram in the assumption of the octahedral axiom.

Define

ϕ\colonequalsπm,k++πm,k=Ψ+,m1+2km+kψ,m+k2m+2k12+Ψ,m1+2km+kψ+,m+k2m+2k12\phi^{\prime}\colonequals\pi_{m,k}^{+}+\pi_{m,k}^{-}=\Psi_{+,m-1+2k}^{m+k}\circ\psi^{\frac{2m+2k-1}{2}}_{-,m+k}+\Psi_{-,m-1+2k}^{m+k}\circ\psi^{\frac{2m+2k-1}{2}}_{+,m+k}

By Lemma 2.13 and Lemma 2.14 with n=m+kn=m+k, we have

ϕh=\displaystyle\phi^{\prime}\circ h^{\prime}= (Ψ+,m1+2km+kψ,m+k2m+2k12+Ψ,m1+2km+kψ+,m+k2m+2k12)(ψ,2m+2k12m+k1ψ+,2m+2k12m+k1)\displaystyle(\Psi_{+,m-1+2k}^{m+k}\circ\psi^{\frac{2m+2k-1}{2}}_{-,m+k}+\Psi_{-,m-1+2k}^{m+k}\circ\psi^{\frac{2m+2k-1}{2}}_{+,m+k})\circ(\psi_{-,\frac{2m+2k-1}{2}}^{m+k-1}-\psi_{+,\frac{2m+2k-1}{2}}^{m+k-1})
=\displaystyle= Ψ,m1+2km+kψ+,m+k2m+2k12ψ,2m+2k12m+k1Ψ+,m1+2km+kψ,m+k2m+2k12ψ+,2m+2k12m+k1\displaystyle\Psi_{-,m-1+2k}^{m+k}\circ\psi^{\frac{2m+2k-1}{2}}_{+,m+k}\circ\psi_{-,\frac{2m+2k-1}{2}}^{m+k-1}-\Psi_{+,m-1+2k}^{m+k}\circ\psi^{\frac{2m+2k-1}{2}}_{-,m+k}\circ\psi_{+,\frac{2m+2k-1}{2}}^{m+k-1}
=\displaystyle= Ψ,m1+2km+kψ,m+km+k1Ψ+,m1+2km+kψ+,m+km+k1\displaystyle\Psi_{-,m-1+2k}^{m+k}\circ\psi^{m+k-1}_{-,m+k}-\Psi_{+,m-1+2k}^{m+k}\circ\psi^{m+k-1}_{+,m+k}
=\displaystyle= Ψ,m1+2km1+kΨ+,m1+2km1+k\displaystyle\Psi_{-,m-1+2k}^{m-1+k}-\Psi_{+,m-1+2k}^{m-1+k}

This verifies the second commutative diagram mentioned in Section 3.1.

Finally, we state a weaker version of the third commutative diagram mentioned in Section 3.1, which is enough to apply Lemma 2.25. The following proposition will be proved in Section 7.4.

Proposition 3.12.

Suppose ll^{\prime} and ll are the maps in Proposition 3.6 and Proposition 3.9. Then, there are two commutative diagrams up to scalars.

𝚪2m+2k12\textstyle{\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l^{\prime}}πm,k+\scriptstyle{\pi_{m,k}^{+}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,m1μ\scriptstyle{\psi_{+,m-1}^{\mu}}𝚪m1+2k\textstyle{\mathbf{\Gamma}_{m-1+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l}𝚪m1\textstyle{\mathbf{\Gamma}_{m-1}}
𝚪2m+2k12\textstyle{\mathbf{\Gamma}_{\frac{2m+2k-1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l^{\prime}}πm,k\scriptstyle{\pi_{m,k}^{-}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,m1μ\scriptstyle{\psi_{-,m-1}^{\mu}}𝚪m1+2k\textstyle{\mathbf{\Gamma}_{m-1+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}l\scriptstyle{l}𝚪m1\textstyle{\mathbf{\Gamma}_{m-1}}
Proof of Theorem 3.1.

We verified all assumptions of the octahedral lemma (Lemma 2.23) for the diagram (3.8). Hence, there exists a map ϕ\phi such that

𝐘mH(Cone(ϕ)).\mathbf{Y}_{m}\cong H(\operatorname{Cone}(\phi)).

We also verified all assumptions of Lemma 2.25 for ϕ=πm,k++πm,k\phi^{\prime}=\pi_{m,k}^{+}+\pi_{m,k}^{-}. Thus, we have

H(Cone(ϕ))H(Cone(ϕ))𝐘m.H(\operatorname{Cone}(\phi^{\prime}))\cong H(\operatorname{Cone}(\phi))\cong\mathbf{Y}_{m}.

Then the desired triangle in the theorem holds. ∎

3.3. Reformulation by bent complexes

In this subsection, we restate Theorem 3.1 using the language of bent complexes introduced in [22]. Suppose KK is a rationally null-homologous knot in a closed 3-manifold YY. We continue to adopt the notations and conventions from Section 2.2 and Section 2.3.

Putting bypass triangles in Lemma 2.6 for different nn together, we obtain the following diagram:

(3.12) \textstyle{\cdots}𝚪n+1\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,μn+1\scriptstyle{\psi_{+,\mu}^{n+1}}𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n+1n\scriptstyle{\psi_{+,n+1}^{n}}ψ+,μn\scriptstyle{\psi_{+,\mu}^{n}}𝚪n1\textstyle{\mathbf{\Gamma}_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,nn1\scriptstyle{\psi_{+,n}^{n-1}}ψ+,μn1\scriptstyle{\psi_{+,\mu}^{n-1}}𝚪n2\textstyle{\mathbf{\Gamma}_{n-2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n1n2\scriptstyle{\psi_{+,n-1}^{n-2}}\textstyle{\ignorespaces\ignorespaces\ignorespaces\ignorespaces\cdots}\textstyle{\cdots}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,nμ\scriptstyle{\psi_{+,n}^{\mu}}ψ,n2μ\scriptstyle{\psi_{-,n-2}^{\mu}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n1μ\scriptstyle{\psi_{+,n-1}^{\mu}}ψ,n1μ\scriptstyle{\psi_{-,n-1}^{\mu}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,n2μ\scriptstyle{\psi_{+,n-2}^{\mu}}ψ,nμ\scriptstyle{\psi_{-,n}^{\mu}}\textstyle{\cdots}\textstyle{\cdots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}𝚪n2\textstyle{\mathbf{\Gamma}_{n-2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n1n2\scriptstyle{\psi_{-,n-1}^{n-2}}𝚪n1\textstyle{\mathbf{\Gamma}_{n-1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,nn1\scriptstyle{\psi_{-,n}^{n-1}}ψ,μn1\scriptstyle{\psi_{-,\mu}^{n-1}}𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,n+1n\scriptstyle{\psi_{-,n+1}^{n}}ψ,μn\scriptstyle{\psi_{-,\mu}^{n}}𝚪n+1\textstyle{\mathbf{\Gamma}_{n+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,μn+1\scriptstyle{\psi_{-,\mu}^{n+1}}\textstyle{\cdots}

where the \mathbb{Z}-grading shift of ψ±,kμψ±,μk\psi_{\pm,k}^{\mu}\circ\psi_{\pm,\mu}^{k} is ±p\pm p for any kk\in\mathbb{Z}. From (3.12), we constructed in [22, Section 3.4] two spectral sequences {Er,+,dr,+}r1\{E_{r,+},d_{r,+}\}_{r\geq 1} and {Er,,dr,}r1\{E_{r,-},d_{r,-}\}_{r\geq 1} from 𝚪μ\mathbf{\Gamma}_{\mu} to 𝐘\mathbf{Y}, where dr,±d_{r,\pm} is roughly

(3.13) ψ±,μk(Ψ±,k+rk)1ψ±,k+rμ for any k.\psi_{\pm,\mu}^{k}\circ(\Psi_{\pm,k+r}^{k})^{-1}\circ\psi_{\pm,k+r}^{\mu}\text{ for any }k\in\mathbb{Z}.

The composition with the inverse map is well-defined on the rr-th page, and the independence of kk (and hence nn in (3.13)) follows from Lemma 2.12. The \mathbb{Z}-grading shift of dr,±d_{r,\pm} is ±rp\pm rp. By fixing an inner product on 𝚪μ\mathbf{\Gamma}_{\mu}, we then lifted those spectral sequences to two differentials d+d_{+} and dd_{-} on 𝚪μ\mathbf{\Gamma}_{\mu} such that

H(𝚪μ,d+)H(𝚪μ,d)𝐘.H(\mathbf{\Gamma}_{\mu},d_{+})\cong H(\mathbf{\Gamma}_{\mu},d_{-})\cong\mathbf{Y}.

In such way, the inverses of Ψ±,k+rk\Psi_{\pm,k+r}^{k} are also well-defined, which we will use freely later.

Then we propose an integral surgery formula for 𝐘m\mathbf{Y}_{m} using differentials d+d_{+} and dd_{-} on 𝚪μ\mathbf{\Gamma}_{\mu}. To state the formula, we introduce the following notations.

Definition 3.13 ([22, Construction 3.27 and Definition 5.12]).

For any integer ss, define the complexes

B±(s)\colonequals(k(𝚪μ,s+kp),d±),B+(s)\colonequals(k0(𝚪μ,s+kp),d+),B^{\pm}(s)\colonequals(\bigoplus_{k\in\mathbb{Z}}(\mathbf{\Gamma}_{\mu},s+kp),d_{\pm}),\quad B^{+}({\geq s})\colonequals(\bigoplus_{k\geq 0}(\mathbf{\Gamma}_{\mu},s+kp),d_{+}),
andB(s)\colonequals(k0(𝚪μ,s+kp),d).~{}{\rm and}~{}B^{-}({\leq s})\colonequals(\bigoplus_{k\leq 0}(\mathbf{\Gamma}_{\mu},s+kp),d_{-}).

Furthermore, define

I+(s):B+(s)B+(s)andI(s):B(s)B(s)I^{+}(s):B^{+}(\geq s)\to B^{+}(s)~{}{\rm and}~{}I^{-}(s):B^{-}(\leq s)\to B^{-}(s)

to be the inclusion maps. We also write the same notation for the induced map on homology.

Remark 3.14.

By Lemma 2.5, we know that the nontrivial gradings of 𝚪μ\mathbf{\Gamma}_{\mu} are finite. Then, for any sufficiently large integer s0s_{0} satisfying

ss0ppχ(S)2ands+s0ppχ(S)2,s-s_{0}p\leq-\frac{p-\chi(S)}{2}~{}{\rm and}~{}s+s_{0}p\geq\frac{p-\chi(S)}{2},

we have

B+(s)=B+(ss0p)andB(s)=B(s+s0p).B^{+}(s)=B^{+}(\geq s-s_{0}p)~{}{\rm and}~{}B^{-}(s)=B^{-}({\leq s+s_{0}p}).

In such case, I+(ss0p)I^{+}(s-s_{0}p) and I(s+s0p)I^{-}(s+s_{0}p) are identities.

By splitting the diagram (3.12) into \mathbb{Z}-gradings, we can calculate homologies of the complexes defined in Definition 3.13.

Proposition 3.15.

Suppose n+n\in\mathbb{N}_{+} and ii is a grading. Fix an inner product on 𝚪n\mathbf{\Gamma}_{n}. If i>(pχ(S))/2npi>(p-\chi(S))/2-np, then there exists a canonical isomorphism

H(B+(i))(𝚪n,i+(n1)pq2).H(B^{+}({\geq i}))\cong(\mathbf{\Gamma}_{n},i+\frac{(n-1)p-q}{2}).

If i<(pχ(S))/2+npi<-(p-\chi(S))/2+np, then there exists a canonical isomorphism

H(B(i))(𝚪n,i(n1)pq2).H(B^{-}({\leq i}))\cong(\mathbf{\Gamma}_{n},i-\frac{(n-1)p-q}{2}).
Proof.

The proof mirrors that of [22, Lemma 5.13]. Following the notation in [22, (3.9) and (3.10)], if

i>i^maxμnq=(pχ(S))/2np,i>\hat{i}_{max}^{\mu}-nq=(p-\chi(S))/2-np,

then 𝚪0i,+=0\mathbf{\Gamma}_{0}^{i,+}=0 (the corresponding grading summand of 𝚪0\mathbf{\Gamma}_{0}) and the isomorphism follows from the convergence theorem of the unrolled spectral sequence [22, Theorem 2.4] (see also [1, Theorem 6.1]). Note that the unrolled spectral sequence induces a filtration on 𝚪n\mathbf{\Gamma}_{n}, and the homology is canonically isomorphic to the direct sum of all associated graded objects of the filtration. Then we use the inner product to identify the direct sum with the total space 𝚪n\mathbf{\Gamma}_{n}. The other statement holds for the same reason. ∎

Definition 3.16 ([22, Construction 3.27 and Definition 5.12]).

For any integer ss, define the bent complex

A(s)\colonequals(k(𝚪μ,s+kp),ds),A(s)\colonequals(\bigoplus_{k\in\mathbb{Z}}(\mathbf{\Gamma}_{\mu},s+kp),d_{s}),

where for any element x(𝚪μ,s+kp)x\in(\mathbf{\Gamma}_{\mu},s+kp),

ds(x)={d+(x)k>0,d+(x)+d(x)k=0,d(x)k<0.d_{s}(x)=\begin{cases}d_{+}(x)&k>0,\\ d_{+}(x)+d_{-}(x)&k=0,\\ d_{-}(x)&k<0.\end{cases}

Define

π+(s):A(s)B+(s)andπ(s):A(s)B(s)\pi^{+}({s}):A(s)\to B^{+}(s)~{}{\rm and}~{}\pi^{-}({s}):A(s)\to B^{-}(s)

by

π+(s)(x)={xk0,0k<0,andπ(s)(x)={xk0,0k>0,\pi^{+}(s)(x)=\begin{cases}x&k\geq 0,\\ 0&k<0,\end{cases}~{}{\rm and}~{}\pi^{-}(s)(x)=\begin{cases}x&k\leq 0,\\ 0&k>0,\end{cases}

where x(𝚪μ,s+kp)x\in(\mathbf{\Gamma}_{\mu},s+kp). Define

π±:sA(s)sB±(s)\pi^{\pm}:\bigoplus_{s\in\mathbb{Z}}A(s)\to\bigoplus_{s\in\mathbb{Z}}B^{\pm}(s)

by putting π±(s)\pi^{\pm}(s) together for all ss . We also use the same notation for the induced map on homology.

Remark 3.17.

Similar to Remark 3.14, according to Lemma 2.5, there are only finitely many the nontrivial gradings of 𝚪μ\mathbf{\Gamma}_{\mu}. Then, for any sufficiently large integer s0s_{0} such that s0(pχ(S))/2s_{0}\geq(p-\chi(S))/2, we have

A(s0)=B(s0)andA(s0)=B+(s0).A(s_{0})=B^{-}(s_{0})~{}{\rm and}~{}A(-s_{0})=B^{+}(-s_{0}).

In such case, π(s0)\pi^{-}(s_{0}) and π+(s0)\pi^{+}(-s_{0}) are identities.

Now, we state the integral surgery formula in the above setup.

Theorem 3.18.

Suppose mm is a fixed integer such that mpq0mp-q\neq 0. Then there exists an isomorphism

Ξm:sH(B+(s))sH(B(s+mpq))\Xi_{m}:\bigoplus_{s\in\mathbb{Z}}H(B^{+}(s))\xrightarrow{\cong}\bigoplus_{s\in\mathbb{Z}}H(B^{-}(s+mp-q))

as the direct sum of isomorphisms

Ξm,s:H(Bs+)H(Bs+mpq)\Xi_{m,s}:H(B^{+}_{s})\xrightarrow{\cong}H(B^{-}_{s+mp-q})

so that

𝐘mH(Cone(π+Ξmπ+:sH(A(s))sH(B(s)))).\mathbf{Y}_{m}\cong H\bigg{(}\operatorname{Cone}(\pi^{-}+\Xi_{m}\circ\pi^{+}:\bigoplus_{s\in\mathbb{Z}}H(A(s))\to\bigoplus_{s\in\mathbb{Z}}H(B^{-}(s)))\bigg{)}.
Proof.

According to Remark 3.17, we only need to consider the maps π±(s)\pi^{\pm}(s) for |s||s| less than a fixed integer. For such values of ss, we can apply the following proposition.

Proposition 3.19 ([22, Proposition 3.28]).

Fix m,sm,s\in\mathbb{Z} such that |s|(pχ(S))/2|s|\leq(p-\chi(S))/2. For any large integer kk, fix inner products on 𝚪2m+2k12\mathbf{\Gamma}_{\frac{2m+2k-1}{2}} and 𝚪m1+2k\mathbf{\Gamma}_{m-1+2k}. Then there exist s1,s2+,s2,s3+,s3s_{1},s_{2}^{+},s_{2}^{-},s_{3}^{+},s_{3}^{-}\in\mathbb{Z} such that the following diagram commutes

H(A(s))\textstyle{H(A(s))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π±(s)\scriptstyle{\pi^{\pm}(s)}\scriptstyle{\cong}H(B±(s))\textstyle{H(B^{\pm}(s))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}(𝚪2m+2k12,s1)\textstyle{(\mathbf{\Gamma}_{\frac{2m+2k-1}{2}},s_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}πm,k±,s1\scriptstyle{\pi^{\pm,s_{1}}_{m,k}}(𝚪m1+2k,s3±)\textstyle{(\mathbf{\Gamma}_{m-1+2k},s_{3}^{\pm})}

where the maps πm,k±,s1\pi^{\pm,s_{1}}_{m,k}, defined in Section 3.1, factor through (𝚪m+k,s2±)(\mathbf{\Gamma}_{m+k},s_{2}^{\pm}).

Remark 3.20.

The maps π±(s)\pi^{\pm}(s) factor through I±(s)I^{\pm}(s) constructed in Definition 3.13. We denote

π±(s)=I+(s)π±,(s).\pi^{\pm}(s)=I^{+}(s)\circ\pi^{\pm,\prime}(s).

This corresponds to the factorization about (𝚪m+k,s2±)(\mathbf{\Gamma}_{m+k},s_{2}^{\pm}) in Proposition 3.19 (we fix an inner product on 𝚪m+k\mathbf{\Gamma}_{m+k} to apply Proposition 3.15), i.e., the following diagrams commute

H(A(s))\textstyle{H(A(s))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π+,(s)\scriptstyle{\pi^{+,\prime}(s)}\scriptstyle{\cong}H(B+(s))\textstyle{H(B^{+}(\geq s))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}I+(s)\scriptstyle{I^{+}(s)}H(B+(s))\textstyle{H(B^{+}(s))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}(𝚪2m+2k12,s1)\textstyle{(\mathbf{\Gamma}_{\frac{2m+2k-1}{2}},s_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,m+k2m+2k12\scriptstyle{\psi_{-,m+k}^{\frac{2m+2k-1}{2}}}(𝚪m+k,s2+)\textstyle{(\mathbf{\Gamma}_{m+k},s_{2}^{+})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ+,m1+2km+k\scriptstyle{\Psi_{+,m-1+2k}^{m+k}}(𝚪m1+2k,s3+)\textstyle{(\mathbf{\Gamma}_{m-1+2k},s_{3}^{+})}
H(A(s))\textstyle{H(A(s))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}π,(s)\scriptstyle{\pi^{-,\prime}(s)}\scriptstyle{\cong}H(B(s))\textstyle{H(B^{-}(\leq s))\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}I(s)\scriptstyle{I^{-}(s)}H(B(s))\textstyle{H(B^{-}(s))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{\cong}(𝚪2m+2k12,s1)\textstyle{(\mathbf{\Gamma}_{\frac{2m+2k-1}{2}},s_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,m+k2m+2k12\scriptstyle{\psi_{+,m+k}^{\frac{2m+2k-1}{2}}}(𝚪m+k,s2)\textstyle{(\mathbf{\Gamma}_{m+k},s_{2}^{-})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ,m1+2km+k\scriptstyle{\Psi_{-,m-1+2k}^{m+k}}(𝚪m1+2k,s3)\textstyle{(\mathbf{\Gamma}_{m-1+2k},s_{3}^{-})}

From the calculation in [22, Remark 3.29] (we replace nn and ll there by m+km+k and k1k-1, and note that there is a typo about sign in the first arXiv version of [22]), the difference of the grading shifts is

s3+s3=(m+k(k1)1)pq=mpq.s_{3}^{+}-s_{3}^{-}=(m+k-(k-1)-1)p-q=mp-q.

Note that the notations in this paper and [22] are different (c.f. Remark 2.3).

Then we can construct the isomorphism

Ξm,s:H(B+(s))H(B(s+mpq))\Xi_{m,s}:H(B^{+}(s))\xrightarrow{\cong}H(B^{-}(s+mp-q))

for |s|(pχ(S))/2|s|\leq(p-\chi(S))/2 by identifying both H(B+(s))H(B^{+}(s)) and H(B(s+mpq))H(B^{-}(s+mp-q)) with (𝚪m1+2k,s3+)(\mathbf{\Gamma}_{m-1+2k},s_{3}^{+}) for a sufficiently large kk. A priori, this isomorphism depends on inner products on

𝚪μ,𝚪2m+2k12,𝚪m1+2kand𝚪m+k.\mathbf{\Gamma}_{\mu},\mathbf{\Gamma}_{\frac{2m+2k-1}{2}},\mathbf{\Gamma}_{m-1+2k}~{}{\rm and}~{}\mathbf{\Gamma}_{m+k}.

For other ss, we can take any isomorphism Ξm,s\Xi_{m,s} since the choice does not affect the computation of the mapping cone.

Consequently, we obtain

H(Cone(π+Ξmπ+))H(Cone(πm,k+πm,k+))𝐘m,H(\operatorname{Cone}(\pi^{-}+\Xi_{m}\circ\pi^{+}))\cong H(\operatorname{Cone}(\pi_{m,k}^{-}+\pi_{m,k}^{+}))\cong\mathbf{Y}_{m},

where the last isomorphism comes from Theorem 3.1. ∎

Remark 3.21.

Theorem 3.18 is slightly weaker than Theorem 3.1. Indeed, when we use the integral surgery formula to calculate surgeries on the Boromean knot in the companion paper [24], we have to study the H1(Y)H_{1}(Y) action on sutured instanton homology, where Y=#2nS1×S2Y=\#^{2n}S^{1}\times S^{2} is the ambient manifold of the knot. This action vanishes on 𝚪μ\mathbf{\Gamma}_{\mu} so vanishes on the bent complex. But it is nonvanishing on 𝚪m+k\mathbf{\Gamma}_{m+k} and 𝚪m1+2k\mathbf{\Gamma}_{m-1+2k} and we use this information to realize the computation. This issue for the bent complex might be resolved by introducing some E0E_{0}-pages for differentials d+d_{+} and dd_{-} such that the action is nontrivial on E0E_{0}-pages.

3.4. A formula for instanton knot homology

The third exact sequence (3.11) implies

𝚪m1H(Cone(Ψ,m1+2km1+kΨ+,m1+2km1+k:𝚪m1+k𝚪m1+k𝚪m1+2k))\mathbf{\Gamma}_{m-1}\cong H(\operatorname{Cone}(\Psi_{-,m-1+2k}^{m-1+k}-\Psi_{+,m-1+2k}^{m-1+k}:\mathbf{\Gamma}_{m-1+k}\oplus\mathbf{\Gamma}_{m-1+k}\to\mathbf{\Gamma}_{m-1+2k}))

for any sufficiently large integer kk. Since there are two copies 𝚪m1+k\mathbf{\Gamma}_{m-1+k}, we can always regard the grading shifts of the maps Ψ,m1+2km1+k\Psi_{-,m-1+2k}^{m-1+k} as different ones by rescaling the grading of the first summand from ii to 2i12i-1 and the second summand from ii to 2i2i. Hence we do not need the assumption mpq0mp-q\neq 0 as in the previous mapping cone formula in Theorem 3.1. By Lemma 2.24, we can replace the minus sign with any coefficient.

In this subsection, we restate this result in the language of bent complexes. The formula is inspired by Eftekhary’s formula for knot Floer homology HFK^\widehat{HFK} [10, Proposition 1.5] (see also Hedden-Levine’s work [13]). Since mm can be any integer, we replace m1m-1 by mm.

Theorem 3.22.

Suppose m,jm,j\in\mathbb{Z}. Define

j+=j(m1)pq2andj=j+(m1)pq2.j^{+}=j-\frac{(m-1)p-q}{2}~{}{\rm and}~{}j^{-}=j+\frac{(m-1)p-q}{2}.

Then there exists an isomorphism

Ξm,j:H(B+(j+))H(B(j))\Xi_{m,j}^{\prime}:H(B^{+}(j^{+}))\xrightarrow{\cong}H(B^{-}(j^{-}))

such that

(𝚪m,j)H(Cone(I(j)+Ξm,jI+(j+):H(B(j))H(B+(j+))H(B(j)))).(\mathbf{\Gamma}_{m},j)\cong H\bigg{(}\operatorname{Cone}(I^{-}(j^{-})+\Xi_{m,j}^{\prime}\circ I^{+}(j^{+}):H(B^{-}(\leq j^{-}))\oplus H(B^{+}(\geq j^{+}))\to H(B^{-}(j^{-})))\bigg{)}.
Proof.

As mentioned before, we have

𝚪mH(Cone(Ψ,m+2km+kΨ+,m+2km+k))H(Cone(Ψ,m+2km+k+Ψ+,m+2km+k))\mathbf{\Gamma}_{m}\cong H(\operatorname{Cone}(\Psi_{-,m+2k}^{m+k}-\Psi_{+,m+2k}^{m+k}))\cong H(\operatorname{Cone}(\Psi_{-,m+2k}^{m+k}+\Psi_{+,m+2k}^{m+k}))

for any sufficiently large integer kk.

Since bypass maps are homogeneous, the above mapping cone splits into \mathbb{Z}-gradings (or (+12)(\mathbb{Z}+\frac{1}{2})-gradings). Hence we can use it to calculate (𝚪m,j)(\mathbf{\Gamma}_{m},j). By Lemma 2.6, the corresponding spaces are

(𝚪m+k,jkp2)(𝚪m+k,j+kp2)and(𝚪m+2k,j).(\mathbf{\Gamma}_{m+k},j-\frac{kp}{2})\oplus(\mathbf{\Gamma}_{m+k},j+\frac{kp}{2})~{}{\rm and}~{}(\mathbf{\Gamma}_{m+2k},j).

From Proposition 3.15 with i=j±kp/2i=j\pm kp/2, by fixing an inner product on 𝚪m+k\mathbf{\Gamma}_{m+k}, we know that

(𝚪m+k,jkp2)H(B(j)) for jkp2<pχ(S)2+(m+k)p(\mathbf{\Gamma}_{m+k},j-\frac{kp}{2})\cong H(B^{-}(\leq j^{-}))\text{ for }j-\frac{kp}{2}<-\frac{p-\chi(S)}{2}+(m+k)p

and

(𝚪m+k,j+kp2)H(B+(j+)) for j+kp2>pχ(S)2(m+k)p.(\mathbf{\Gamma}_{m+k},j+\frac{kp}{2})\cong H(B^{+}(\geq j^{+}))\text{ for }j+\frac{kp}{2}>\frac{p-\chi(S)}{2}-(m+k)p.

Since mm is fixed, when kk is sufficiently large, we know that any jj with (𝚪m,j)(\mathbf{\Gamma}_{m},j) nontrivial (i.e. |j|(|mpq|χ(S))/2|j|\leq(|mp-q|-\chi(S))/2 by Lemma 2.5) satisfies the above inequalities. By Proposition 3.15 again (fixing an inner product on 𝚪m+2k\mathbf{\Gamma}_{m+2k}) and Remark 3.14, for kk sufficiently large, we know that

(𝚪m+2k,j)H(B(j))H(B+(j+))(\mathbf{\Gamma}_{m+2k},j)\cong H(B^{-}(j^{-}))\cong H(B^{+}(j^{+}))

for such jj. By unpackaging the construction of differentials d+d_{+} and dd_{-} in [22, Section 3.4], we know that the restrictions of maps Ψ,m+2km\Psi_{-,m+2k}^{m} and Ψ+,m+2km\Psi_{+,m+2k}^{m} on the corresponding gradings coincide with the maps induced by the inclusions I(j)I^{-}(j^{-}) and I+(j+)I^{+}(j^{+}) under the canonical isomorphisms, respectively.

For |j|(|mpq|χ(S))/2|j|\leq(|mp-q|-\chi(S))/2, let

Ξm,j:H(B+(j+))H(B(j))\Xi^{\prime}_{m,j}:H(B^{+}(j^{+}))\xrightarrow{\cong}H(B^{-}(j^{-}))

be the isomorphism obtained from identifying both spaces to the corresponding grading summand of 𝚪m+2k\mathbf{\Gamma}_{m+2k}. Note that it depends on inner products on 𝚪μ,𝚪m+k\mathbf{\Gamma}_{\mu},\mathbf{\Gamma}_{m+k} and 𝚪m+2k\mathbf{\Gamma}_{m+2k}. For other jj, we can take any isomorphism Ξm,j\Xi_{m,j}^{\prime} since the choice does not affect the computation of the mapping cone. Then we know that

(𝚪m,j)\displaystyle(\mathbf{\Gamma}_{m},j)\cong H(Cone(Ψ,m+2km+k+Ψ+,m+2km+k|(𝚪m+k,j+kp2)(𝚪m+k,jkp2)))\displaystyle H(\operatorname{Cone}(\Psi_{-,m+2k}^{m+k}+\Psi_{+,m+2k}^{m+k}|(\mathbf{\Gamma}_{m+k},j+\frac{kp}{2})\oplus(\mathbf{\Gamma}_{m+k},j-\frac{kp}{2})))
\displaystyle\cong H(Cone(I(j)+Ξm,jI+(j+))).\displaystyle H(\operatorname{Cone}(I^{-}(j^{-})+\Xi_{m,j}^{\prime}\circ I^{+}(j^{+}))).

4. Dehn surgery and bypass maps

In this section, we prove a generalization of Lemma 2.20 and Proposition 3.3.

Suppose (M,γ)(M,\gamma) is a balanced sutured manifold and αM\alpha\subset\partial M is a connected simple closed curve that intersects the suture γ\gamma twice. There are two natural bypass arcs associated to α\alpha, each of which intersects the suture at three points and induces a bypass triangle (c.f. [8, Section 4])

SHI¯(M,γ)\textstyle{\underline{\rm SHI}(-M,-\gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ±\scriptstyle{\psi_{\pm}}SHI¯(M,γ2)\textstyle{\underline{\rm SHI}(-M,-\gamma_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SHI¯(M,γ3)\textstyle{\underline{\rm SHI}(-M,-\gamma_{3})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

where γ2\gamma_{2} and γ3\gamma_{3} are the sutures coming from bypass attachments. Note that the two bypass exact triangles involve the same set of balanced sutured manifolds but have different maps between them. Let (M0,γ0)(M_{0},\gamma_{0}) be obtained from (M,γ)(M,\gamma) by attaching a contact 22-handle along α\alpha. From [4, Section 3.3], it has been shown that a closure of (M0,γ0)(-M_{0},-\gamma_{0}) coincides with a closure of the sutured manifold obtained from (M,γ)(-M,-\gamma) by 0-surgery along α\alpha with respect to the surface framing. Hence there is also a surgery exact triangle (c.f. [23, Lemma 3.21])

SHI¯(M,γ)\textstyle{\underline{\rm SHI}(-M,-\gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hα\scriptstyle{H_{\alpha}}SHI¯(M,γ2)\textstyle{\underline{\rm SHI}(-M,-\gamma_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SHI¯(M0,γ0)\textstyle{\underline{\rm SHI}(-M_{0},-\gamma_{0})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

The map HαH_{\alpha} is related to the bypass maps ψ±\psi_{\pm} as follows:

Proposition 4.1.

There exist c1,c2\{0}c_{1},c_{2}\in\mathbb{C}\backslash\{0\}, such that

Hα=c1ψ++c2ψ.H_{\alpha}=c_{1}\psi_{+}+c_{2}\psi_{-}.
Remark 4.2.

The proof of Proposition 4.1 was developed through the discussions with John A. Baldwin and Steven Sivek.

Proof of Proposition 4.1.

Let AMA\subset\partial M be a tubular neighborhood of αM\alpha\subset\partial M. Push the interior of AA into the interior of MM to make it a properly embedded surface. By a standard argument in [14], we can assume that a collar of M\partial M is equipped with a product contact structure such that γ\gamma is (isotopic to) the dividing set, α\alpha is a Legendrian curve, AA is in the contact collar, and AA is a convex surface with Legendrian boundary that separates a standard contact neighborhood of α\alpha off MM. The convex decomposition of MM along AA yields two pieces

M=MAV,M=M^{\prime}\mathop{\cup}_{A}V,

where MM^{\prime} is diffeomorphic to MM and VV is the contact neighborhood of α\alpha. It is straightforward to check that, after rounding the corners, the contact structure near the boundary of MM^{\prime} is still a product contact structure with M\partial M^{\prime} being a convex boundary. Let γ\gamma^{\prime} be the dividing set on M\partial M^{\prime}. Also, after rounding the corners, with the contact structure on VS1×D2V\cong S^{1}\times D^{2}, we can suppose V\partial V is a convex surface with dividing set being the union of two connected simple closed curves on V\partial V of slope 1-1. When viewing VV as the complement of an unknot in S3S^{3}, the dividing set coincides with the suture Γ1V\Gamma_{1}\subset V, so from now on we call it Γ1\Gamma_{1}. By the construction of the gluing map in [19], there exists a map

G1:SHI¯(M,γ)SHI¯(V,Γ1)SHI¯(M,γ).G_{1}:\underline{\rm SHI}(-M^{\prime},-\gamma^{\prime})\otimes\underline{\rm SHI}(-V,-\Gamma_{1})\rightarrow\underline{\rm SHI}(-M,-\gamma).

As in [19], the map G1G_{1} comes from attaching contact handles to (M,γ)(V,Γ1)(M^{\prime},\gamma^{\prime})\sqcup(V,\Gamma_{1}) to recover the gluing along AA. From [20, Proposition 1.4], we know that

SHI¯(V,Γ1).\underline{\rm SHI}(-V,-\Gamma_{1})\cong\mathbb{C}.

Note that MM^{\prime} and MM are equipped with the product contact structure near their boundaries. From the functoriality of the contact gluing map in [19], we know that G1G_{1} is an isomorphism. Now both the (1)(-1)-surgery along a push off of α\alpha and the bypass attachments can be thought of as happening in the piece VV. Note that the result of both (1)(-1)-surgery and the bypass attachments for Γ1\Gamma_{1} is Γ2\Gamma_{2}. Hence we have the following commutative diagram.

(4.1) SHI¯(M,γ)SHI¯(V,Γ1)\textstyle{\underline{\rm SHI}(-M^{\prime},-\gamma^{\prime})\otimes\underline{\rm SHI}(-V,-\Gamma_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G1\scriptstyle{\quad\quad\quad\quad\quad G_{1}}IdH^α\scriptstyle{\operatorname{Id}\otimes\widehat{H}_{\alpha}}SHI¯(M,γ)\textstyle{\underline{\rm SHI}(-M,-\gamma)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hα\scriptstyle{H_{\alpha}}SHI¯(M,γ)SHI¯(V,Γ2)\textstyle{\underline{\rm SHI}(-M^{\prime},-\gamma^{\prime})\otimes\underline{\rm SHI}(-V,-\Gamma_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G2\scriptstyle{\quad\quad\quad\quad\quad G_{2}}SHI¯(M,γ2)\textstyle{\underline{\rm SHI}(-M,-\gamma_{2})}

where H^α\widehat{H}_{\alpha} denotes the surgery map for the manifold VV and G2G_{2} is the gluing map obtained by attaching the same set of contact handles as G1G_{1}. A similar commutative diagram holds when replacing HαH_{\alpha} and H^α\widehat{H}_{\alpha} by ψ±\psi_{\pm} and

ψ^±:SHI¯(V,Γ1)SHI¯(V,Γ2)\hat{\psi}_{\pm}:\underline{\rm SHI}(-V,-\Gamma_{1})\rightarrow\underline{\rm SHI}(-V,-\Gamma_{2})

in (4.1), respectively.

Since G1G_{1} is an isomorphism, to obtain a relation between HαH_{\alpha} and ψ±\psi_{\pm}, it suffices to understand the relation between H^α\widehat{H}_{\alpha} and ψ^±\hat{\psi}_{\pm}. From [20, Proposition 1.4], we know that

SHI¯(V,Γ2)2.\underline{\rm SHI}(-V,-\Gamma_{2})\cong\mathbb{C}^{2}.

Moreover, the meridian disk of VV induces a (+12)(\mathbb{Z}+\frac{1}{2}) grading on SHI¯(V,Γ2)\underline{\rm SHI}(-V,-\Gamma_{2}) and we have

SHI¯(V,Γ2)SHI¯(V,Γ2,12)SHI¯(V,Γ2,12),\underline{\rm SHI}(-V,-\Gamma_{2})\cong\underline{\rm SHI}(-V,-\Gamma_{2},\frac{1}{2})\oplus\underline{\rm SHI}(-V,-\Gamma_{2},-\frac{1}{2}),

with

SHI¯(V,Γ2,12)SHI¯(V,Γ2,12).\underline{\rm SHI}(-V,-\Gamma_{2},\frac{1}{2})\cong\underline{\rm SHI}(-V,-\Gamma_{2},-\frac{1}{2})\cong\mathbb{C}.

Let

𝟏SHI¯(V,Γ1)\mathbf{1}\in\underline{\rm SHI}(-V,-\Gamma_{1})\cong\mathbb{C}

be a generator. In [20, Section 4.3] it is shown that

ψ^(𝟏)SHI¯(V,Γ2,12)andψ^+(𝟏)SHI¯(V,Γ2,12)\hat{\psi}_{-}(\mathbf{1})\in\underline{\rm SHI}(-V,-\Gamma_{2},\frac{1}{2}){\rm~{}and~{}}\hat{\psi}_{+}(\mathbf{1})\in\underline{\rm SHI}(-V,-\Gamma_{2},-\frac{1}{2})

are non-zero. Also, when viewing VV as the complement of the unknot UU, there is an exact triangle

(4.2) SHI¯(V,Γ1)\textstyle{\underline{\rm SHI}(-V,-\Gamma_{1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}H^α\scriptstyle{\widehat{H}_{\alpha}}SHI¯(V,Γ2)\textstyle{\underline{\rm SHI}(-V,-\Gamma_{2})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F2\scriptstyle{F_{2}}I(S3)\textstyle{I^{\sharp}(-S^{3})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}G1\scriptstyle{G_{1}}

as in Lemma 2.16. Comparing the dimensions of the spaces in (4.2), we have G1=0G_{1}=0 and H^α\widehat{H}_{\alpha} is injective. From the fact that τI(U)=0\tau_{I}(U)=0, we know from [12, Corollary 3.5] that

F2|SHI¯(V,Γ2,12)0andF2|SHI¯(V,Γ2,12)0,F_{2}\big{|}{\underline{\rm SHI}(-V,-\Gamma_{2},\frac{1}{2})}\neq 0{\rm~{}and~{}}F_{2}\big{|}{\underline{\rm SHI}(-V,-\Gamma_{2},-\frac{1}{2})}\neq 0,

By the exactness in (4.2), we have ker(F2)=Im(H^α)\ker(F_{2})=\operatorname{Im}(\widehat{H}_{\alpha}) and then H^α(𝟏)\widehat{H}_{\alpha}(\mathbf{1}) is not in SHI¯(V,Γ2,±12)\underline{\rm SHI}(-V,-\Gamma_{2},\pm\frac{1}{2}), i.e., it is a linear combination of generators of SHI¯(V,Γ2,±12)\underline{\rm SHI}(-V,-\Gamma_{2},\pm\frac{1}{2}). Hence we know that there are c1,c2\{0}c_{1},c_{2}\in\mathbb{C}\backslash\{0\} such that

H^α(𝟏)=c1ψ^+(𝟏)+c2ψ^(𝟏).\widehat{H}_{\alpha}(\mathbf{1})=c_{1}\hat{\psi}_{+}(\mathbf{1})+c_{2}\hat{\psi}_{-}(\mathbf{1}).

Then the proposition follows from the commutative diagram (4.1). ∎

In Remark 1.3, we discussed the ambiguity arising from scalars. It is worth mentioning that such ambiguity already exists in instanton theory. For example, if MM is the complement of a knot KS3K\subset S^{3} and γ\gamma consists of two meridians of the knot, which we denote by Γμ\Gamma_{\mu}, we can choose α\alpha to be a curve on (S3\N(K))\partial(S^{3}\backslash N(K)) of slope n-n. Then we have a surgery triangle:

SHI¯(M,Γμ)\textstyle{\underline{\rm SHI}(-M,-\Gamma_{\mu})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hn\scriptstyle{H_{n}}SHI¯(M,Γn1)\textstyle{\underline{\rm SHI}(-M,-\Gamma_{n-1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}I(Sn3(K))\textstyle{I^{\sharp}(-S^{3}_{-n}(K))\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

Note that this triangle is not the one from Floer’s original exact triangle, but rather one with a slight modification on the choice of 1-cycles inside the 33-manifold that represents the second Stiefel-Whitney class of the relevant SO(3)SO(3)-bundle; see [5, Section 2.2] for more details. Floer’s original exact triangle, on the other hand, yields a different triangle

SHI¯(S3\N(K),Γμ)\textstyle{\underline{\rm SHI}(-S^{3}\backslash N(K),-\Gamma_{\mu})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Hn\scriptstyle{H^{\prime}_{n}}SHI¯(S3\N(K),Γn1)\textstyle{\underline{\rm SHI}(-S^{3}\backslash N(K),-\Gamma_{n-1})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}I(Sn3(K),μ)\textstyle{I^{\sharp}(-S^{3}_{-n}(K),\mu)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

where μSn3(K)\mu\subset-S^{3}_{-n}(K) denotes a meridian of the knot. Note the difference between HαH_{\alpha} and HαH^{\prime}_{\alpha} is that they come from the same cobordism but the SO(3)SO(3)-bundles over the cobordism are different. The local argument to prove Proposition 4.1 works for both HαH_{\alpha} and HαH^{\prime}_{\alpha}. Hence there exists non-zero complex numbers c1,c2,c1,c2c_{1},c_{2},c_{1}^{\prime},c_{2}^{\prime} such that

Hα=c1ψ+,nμ+c2ψ,n1μandHα=c1ψ+,nμ+c2ψ,n1μH_{\alpha}=c_{1}\psi^{\mu}_{+,n}+c_{2}\psi^{\mu}_{-,n-1}~{}{\rm and~{}}H^{\prime}_{\alpha}=c^{\prime}_{1}\psi^{\mu}_{+,n}+c^{\prime}_{2}\psi^{\mu}_{-,n-1}

where the maps

ψ±,n1μ:SHI¯(S3\N(K),Γμ)SHI¯(S3\N(K),Γn1)\psi^{\mu}_{\pm,n-1}:\underline{\rm SHI}(-S^{3}\backslash N(K),-\Gamma_{\mu})\rightarrow\underline{\rm SHI}(-S^{3}\backslash N(K),-\Gamma_{n-1})

are the two related bypass maps. When n0n\neq 0, these two bypass maps have different grading shifting behavior, so by Lemma 2.24, different choice of non-zero coefficients does not change the dimensions of kernel and cokernel of the map. Hence we conclude that for n0n\neq 0,

I(Sn3(K),μ)I(Sn3(K)).I^{\sharp}(-S^{3}_{-n}(K),\mu)\cong I^{\sharp}(-S^{3}_{-n}(K)).

However, when n=0n=0, the two bypass maps ψ±,n1μ\psi^{\mu}_{\pm,n-1} both preserves gradings, making the coefficients significant, i.e., I(S03(K),μ)I^{\sharp}(-S^{3}_{0}(K),\mu) and I(S03(K))I^{\sharp}(-S^{3}_{0}(K)) might have different dimensions for different choices of coefficients. Indeed, it is observed by Baldwin-Sivek [5] that for what they called as W-shaped knots (which is clearly a non-empty class, e.g. the figure-8 knot [7, Proposition 10.4]), these two framed instanton homologies have dimensions differing by 22.

5. Some exactness by diagram chasing

5.1. At the direct summand

In this subsection, we prove Proposition 3.4 by diagram chasing. We restate the result in Proposition 5.1. We also adopt the conventions for scalars from Section 2.3, and this together with Lemma 2.11 implies that

Ψ+,n+2k0n+k0Ψ,n+k0n=Ψ,n+2k0n+k0Ψ+,n+k0n.\Psi_{+,n+2k_{0}}^{n+k_{0}}\circ\Psi_{-,n+k_{0}}^{n}=\Psi_{-,n+2k_{0}}^{n+k_{0}}\circ\Psi_{+,n+k_{0}}^{n}.

for any nn and k0k_{0}.

Proposition 5.1.

Given nn\in\mathbb{Z} and k0+k_{0}\in\mathbb{N}_{+}, for any c1,c2,c3,c4c_{1},c_{2},c_{3},c_{4} satisfying the equation

c1c3=c2c4,c_{1}c_{3}=-c_{2}c_{4},

the following sequence is exact

𝚪n(c1Ψ+,n+k0n,c2Ψ,n+k0n)𝚪n+k0𝚪n+k0c3Ψ,n+2k0n+k0+c4Ψ+,n+2k0n+k0𝚪n+2k0\mathbf{\Gamma}_{n}\xrightarrow{(c_{1}\Psi_{+,n+k_{0}}^{n},c_{2}\Psi_{-,n+k_{0}}^{n})}\mathbf{\Gamma}_{n+k_{0}}\oplus\mathbf{\Gamma}_{n+k_{0}}\xrightarrow{c_{3}\Psi_{-,n+2k_{0}}^{n+k_{0}}+c_{4}\Psi_{+,n+2k_{0}}^{n+k_{0}}}\mathbf{\Gamma}_{n+2k_{0}}
Proof.

For simplicity, we only prove the proposition for n=0n=0. The proof for any general nn is similar (replacing all 𝚪m\mathbf{\Gamma}_{m} below by 𝚪n+m\mathbf{\Gamma}_{n+m} and modifying the notations for bypass maps). Also, we only prove the case when

c1=c2=c3=1,c4=1.c_{1}=c_{2}=c_{3}=1,c_{4}=-1.

The proof for general scalars can be obtained similarly.

We prove the proposition by induction on k0k_{0}. We will use the exactness in Lemma 2.6 and the commutative diagrams in Lemma 2.12 and Lemma 2.11 for many times. For simplicity, we will use them without mentioning the lemmas.

First, we assume k0=1k_{0}=1. The proposition reduces to

ker(ψ,21ψ+,21)=Im((ψ+,10,ψ,10)).\ker(\psi_{-,2}^{1}-\psi_{+,2}^{1})=\operatorname{Im}((\psi_{+,1}^{0},\psi_{-,1}^{0})).

The commutative diagram in Lemma 2.11 implies

ker(ψ,21ψ+,21)Im((ψ+,10,ψ,10)).\ker(\psi_{-,2}^{1}-\psi_{+,2}^{1})\supset\operatorname{Im}((\psi_{+,1}^{0},\psi_{-,1}^{0})).

We then prove

ker(ψ,21ψ+,21)Im((ψ+,10,ψ,10)).\ker(\psi_{-,2}^{1}-\psi_{+,2}^{1})\subset\operatorname{Im}((\psi_{+,1}^{0},\psi_{-,1}^{0})).

Suppose

(x1,x2)ker(ψ,21ψ+,21),i.e.,ψ,21(x1)ψ+,21(x2)=0.(x_{1},x_{2})\in\ker(\psi_{-,2}^{1}-\psi_{+,2}^{1}),\text{{i.e.}},\psi_{-,2}^{1}(x_{1})-\psi_{+,2}^{1}(x_{2})=0.

Then we have

ψ+,μ1(x1)=ψ+,μ2ψ,21(x1)=ψ+,μ2ψ+,21(x2)=0.\psi_{+,\mu}^{1}(x_{1})=\psi_{+,\mu}^{2}\circ\psi_{-,2}^{1}(x_{1})=\psi_{+,\mu}^{2}\circ\psi_{+,2}^{1}(x_{2})=0.

By exactness, there exists y𝚪0y\in\mathbf{\Gamma}_{0} such that ψ+,10(y)=x1\psi_{+,1}^{0}(y)=x_{1}. Then

ψ+,21ψ,10(y)=ψ,21ψ+,10(y)=ψ,21(x1)andψ+,21(x2ψ,10(y))=0.\psi_{+,2}^{1}\circ\psi_{-,1}^{0}(y)=\psi_{-,2}^{1}\circ\psi_{+,1}^{0}(y)=\psi_{-,2}^{1}(x_{1})~{}{\rm and}~{}\psi_{+,2}^{1}(x_{2}-\psi_{-,1}^{0}(y))=0.

By exactness, there exists z𝚪μz\in\mathbf{\Gamma}_{\mu} such that

ψ+,1μ(z)=x2ψ,10(y).\psi_{+,1}^{\mu}(z)=x_{2}-\psi_{-,1}^{0}(y).

Let y=y+ψ+,0μ(z).y^{\prime}=y+\psi_{+,0}^{\mu}(z). Then

ψ+,10(y)=ψ+,10(y)=x1\psi_{+,1}^{0}(y^{\prime})=\psi_{+,1}^{0}(y)=x_{1}

and

ψ,10(y)=ψ,10(y)+ψ,10ψ+,0μ(z)=ψ,10(y)+ψ+,1μ(z)=x2,\psi_{-,1}^{0}(y^{\prime})=\psi_{-,1}^{0}(y)+\psi_{-,1}^{0}\circ\psi_{+,0}^{\mu}(z)=\psi_{-,1}^{0}(y)+\psi_{+,1}^{\mu}(z)=x_{2},

which concludes the proof for k0=1k_{0}=1.

Suppose the proposition holds for k0=kk_{0}=k. We prove it also holds for k0=k+1k_{0}=k+1. The proof is similar to the case for k0=1k_{0}=1. Again by Lemma 2.11, we have

ker(Ψ,2k+2k+1Ψ+,2k+2k+1)Im((Ψ+,k+10,Ψ,k+10)).\ker(\Psi_{-,2k+2}^{k+1}-\Psi_{+,2k+2}^{k+1})\supset\operatorname{Im}((\Psi_{+,k+1}^{0},\Psi_{-,k+1}^{0})).

Then we prove

ker(Ψ,2k+2k+1Ψ+,2k+2k+1)Im((Ψ+,k+10,Ψ,k+10)).\ker(\Psi_{-,2k+2}^{k+1}-\Psi_{+,2k+2}^{k+1})\subset\operatorname{Im}((\Psi_{+,k+1}^{0},\Psi_{-,k+1}^{0})).

Suppose

(x1,x2)ker(Ψ,2k+2k+1Ψ+,2k+2k+1),i.e.,Ψ,2k+2k+1(x1)Ψ+,2k+2k+1(x2)=0.(x_{1},x_{2})\in\ker(\Psi_{-,2k+2}^{k+1}-\Psi_{+,2k+2}^{k+1}),\text{{i.e.}},\Psi_{-,2k+2}^{k+1}(x_{1})-\Psi_{+,2k+2}^{k+1}(x_{2})=0.

Then we have

ψ+,μk+1(x1)=ψ+,μ2k+2Ψ,2k+2k+1(x1)=ψ+,μ2k+2Ψ+,2k+2k+1(x2)=0.\psi_{+,\mu}^{k+1}(x_{1})=\psi_{+,\mu}^{2k+2}\circ\Psi_{-,2k+2}^{k+1}(x_{1})=\psi_{+,\mu}^{2k+2}\circ\Psi_{+,2k+2}^{k+1}(x_{2})=0.

By exactness, there exists y1𝚪ky_{1}\in\mathbf{\Gamma}_{k} such that ψ+,k+1k(y1)=x1\psi_{+,k+1}^{k}(y_{1})=x_{1}. By a similar reason, there exists y2𝚪ky_{2}\in\mathbf{\Gamma}_{k} such that ψ,k+1k(y2)=x2\psi_{-,k+1}^{k}(y_{2})=x_{2}. The goal is to prove

Ψ,2kk(y1)=Ψ+,2kk(y2)\Psi_{-,2k}^{k}(y_{1}^{\prime})=\Psi_{+,2k}^{k}(y_{2}^{\prime})

for some modifications y1y^{\prime}_{1} and y2y^{\prime}_{2} of y1y_{1} and y2y_{2} as for yy^{\prime} in the case of k0=1k_{0}=1. Then the induction hypothesis will imply that there exists w𝚪0w\in\mathbf{\Gamma}_{0} such that

Ψ+,k0(w)=y1andΨ,k0(w)=y2.\Psi_{+,k}^{0}(w)=y_{1}^{\prime}~{}{\rm and}~{}\Psi_{-,k}^{0}(w)=y_{2}^{\prime}.

Hence we will have

Ψ+,k+10(w)=ψ+,k+1k(y1)=x1andΨ,k+10(w)=ψ,k+1k(y2)=x2.\Psi_{+,k+1}^{0}(w)=\psi_{+,k+1}^{k}(y_{1}^{\prime})=x_{1}~{}{\rm and}~{}\Psi_{-,k+1}^{0}(w)=\psi_{-,k+1}^{k}(y_{2}^{\prime})=x_{2}.

This will conclude the proof for k0=k+1k_{0}=k+1.

Now we start to construct y1y_{1}^{\prime}. We have

ψ+,2k+22k+1(Ψ+,2k+1k+1(x2)Ψ,2k+1k(y1))\displaystyle\psi_{+,2k+2}^{2k+1}(\Psi_{+,2k+1}^{k+1}(x_{2})-\Psi_{-,2k+1}^{k}(y_{1})) =ψ+,2k+22k+1Ψ+,2k+1k+1(x2)ψ+,2k+22k+1Ψ,2k+1k(y1)\displaystyle=\psi_{+,2k+2}^{2k+1}\circ\Psi_{+,2k+1}^{k+1}(x_{2})-\psi_{+,2k+2}^{2k+1}\circ\Psi_{-,2k+1}^{k}(y_{1})
=Ψ+,2k+2k+1(x2)ψ+,2k+22k+1Ψ,2k+1k(y1)\displaystyle=\Psi_{+,2k+2}^{k+1}(x_{2})-\psi_{+,2k+2}^{2k+1}\circ\Psi_{-,2k+1}^{k}(y_{1})
=Ψ,2k+2k+1(x2)Ψ+,2k+2k+1(x1)\displaystyle=\Psi_{-,2k+2}^{k+1}(x_{2})-\Psi_{+,2k+2}^{k+1}(x_{1})
=0.\displaystyle=0.

By exactness, there exists z1𝚪μz_{1}\in\mathbf{\Gamma}_{\mu} such that

ψ+,2k+1μ(z1)=Ψ+,2k+1k+1(x2)Ψ,2k+1k(y1).\psi_{+,2k+1}^{\mu}(z_{1})=\Psi_{+,2k+1}^{k+1}(x_{2})-\Psi_{-,2k+1}^{k}(y_{1}).

Let y1=y1+ψ+,kμ(z1).y_{1}^{\prime}=y_{1}+\psi_{+,k}^{\mu}(z_{1}). Then

ψ+,k+1k(y1)=ψ+,k+1k(y1)=x1\psi_{+,k+1}^{k}(y_{1}^{\prime})=\psi_{+,k+1}^{k}(y_{1})=x_{1}

and

Ψ,2k+1k(y1)\displaystyle\Psi_{-,2k+1}^{k}(y_{1}^{\prime}) =Ψ,2k+1k(y1)+Ψ,2k+1kψ+,kμ(z1)\displaystyle=\Psi_{-,2k+1}^{k}(y_{1})+\Psi_{-,2k+1}^{k}\circ\psi_{+,k}^{\mu}(z_{1})
=Ψ,2k+1k(y1)+ψ+,2k+1μ(z1)\displaystyle=\Psi_{-,2k+1}^{k}(y_{1})+\psi_{+,2k+1}^{\mu}(z_{1})
=Ψ+,2k+1k+1(x2),\displaystyle=\Psi_{+,2k+1}^{k+1}(x_{2}),

Then we start to construct y2y_{2}^{\prime}. We have

ψ,2k+12k(Ψ,2kk(y1)Ψ+,2kk(y2))\displaystyle\psi_{-,2k+1}^{2k}(\Psi_{-,2k}^{k}(y_{1}^{\prime})-\Psi_{+,2k}^{k}(y_{2})) =Ψ,2k+1k(y1)ψ,2k+12kΨ+,2kk(y2)\displaystyle=\Psi_{-,2k+1}^{k}(y_{1}^{\prime})-\psi_{-,2k+1}^{2k}\circ\Psi_{+,2k}^{k}(y_{2})
=Ψ,2k+1k(y1)Ψ,2k+1k+1(x2)\displaystyle=\Psi_{-,2k+1}^{k}(y_{1}^{\prime})-\Psi_{-,2k+1}^{k+1}(x_{2})
=0.\displaystyle=0.

By exactness, there exists z2𝚪μz_{2}\in\mathbf{\Gamma}_{\mu} such that

ψ,2kμ(z2)=Ψ,2kk(y1)Ψ+,2kk(y2).\psi_{-,2k}^{\mu}(z_{2})=\Psi_{-,2k}^{k}(y_{1}^{\prime})-\Psi_{+,2k}^{k}(y_{2}).

Let y2=y2+ψ,kμ(z2).y_{2}^{\prime}=y_{2}+\psi_{-,k}^{\mu}(z_{2}). Then

ψ,k+1k(y2)=ψ,k+1k(y2)=x2\psi_{-,k+1}^{k}(y_{2}^{\prime})=\psi_{-,k+1}^{k}(y_{2})=x_{2}

and

Ψ+,2kk(y2)\displaystyle\Psi_{+,2k}^{k}(y_{2}^{\prime}) =Ψ+,2kk(y2)+Ψ+,2kkψ,kμ(z2)\displaystyle=\Psi_{+,2k}^{k}(y_{2})+\Psi_{+,2k}^{k}\circ\psi_{-,k}^{\mu}(z_{2})
=Ψ+,2kk(y2)+ψ,2kμ(z2)\displaystyle=\Psi_{+,2k}^{k}(y_{2})+\psi_{-,2k}^{\mu}(z_{2})
=Ψ,2kk(y1),\displaystyle=\Psi_{-,2k}^{k}(y_{1}^{\prime}),

Then we have the following commutative diagrams

x1𝚪k+1\textstyle{x_{1}\in\mathbf{\Gamma}_{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{-}y1𝚪k\textstyle{y_{1}^{\prime}\in\mathbf{\Gamma}_{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}+\scriptstyle{+}\scriptstyle{-}𝚪0\textstyle{\quad\mathbf{\Gamma}_{0}\quad\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}+\scriptstyle{+}\scriptstyle{-}𝚪2k\textstyle{\mathbf{\Gamma}_{2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{-}𝚪2k+2\textstyle{\mathbf{\Gamma}_{2k+2}}𝚪2k+1\textstyle{\mathbf{\Gamma}_{2k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}+\scriptstyle{+}y2𝚪k\textstyle{y_{2}^{\prime}\in\mathbf{\Gamma}_{k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\scriptstyle{-}+\scriptstyle{+}x2𝚪k+1\textstyle{x_{2}\in\mathbf{\Gamma}_{k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}+\scriptstyle{+}

By the induction hypothesis, there exists w𝚪0w\in\mathbf{\Gamma}_{0} such that

Ψ+,k0(w)=y1andΨ,k0(w)=y2,\Psi_{+,k}^{0}(w)=y_{1}^{\prime}~{}{\rm and}~{}\Psi_{-,k}^{0}(w)=y_{2}^{\prime},

which concludes the proof for k0=k+1k_{0}=k+1. ∎

Remark 5.2.

By similar arguments, we can prove that the following sequence is exact for any k1,k2+k_{1},k_{2}\in\mathbb{N}_{+}

𝚪n(c1Ψ+,n+k1n,c2Ψ,n+k2n)𝚪n+k1𝚪n+k2c3Ψ,n+k1+k2n+k1+c4Ψ+,n+k1+k2n+k2𝚪n+k1+k2,\mathbf{\Gamma}_{n}\xrightarrow{(c_{1}\Psi_{+,n+k_{1}}^{n},c_{2}\Psi_{-,n+k_{2}}^{n})}\mathbf{\Gamma}_{n+k_{1}}\oplus\mathbf{\Gamma}_{n+k_{2}}\xrightarrow{c_{3}\Psi_{-,n+k_{1}+k_{2}}^{n+k_{1}}+c_{4}\Psi_{+,n+k_{1}+k_{2}}^{n+k_{2}}}\mathbf{\Gamma}_{n+k_{1}+k_{2}},

where the scalars satisfies the equality c1c3=c2c4c_{1}c_{3}=-c_{2}c_{4}.

5.2. The second exact triangle

In this subsection, we prove Proposition 3.6 by diagram chasing. For convenience, we restate it as follows, which is a little stronger than the previous version. Replacing the original knot in the proposition by the dual knot in the Dehn filling of slope (m+k)μ+λ-(m+k)\mu+\lambda with framing μ-\mu and setting n=1n=-1 will recover Proposition 3.6.

Proposition 5.3.

Suppose

l=ψ+,n1μψ+,μn+1=ψ,n1μψ,μn+1.l^{\prime}=\psi_{+,n-1}^{\mu}\circ\psi_{+,\mu}^{n+1}=\psi_{-,n-1}^{\mu}\circ\psi_{-,\mu}^{n+1}.

Then for any c1,c2,c3,c4\{0}c_{1},c_{2},c_{3},c_{4}\in\mathbb{C}\backslash\{0\}, the following sequence is exact

𝚪n𝚪nc3ψ,n+1n+c4ψ+,n+1n𝚪n+1l𝚪n1(c1ψ,nn1,c2ψ+,nn1)𝚪n𝚪n.\mathbf{\Gamma}_{n}\oplus\mathbf{\Gamma}_{n}\xrightarrow{c_{3}\psi_{-,n+1}^{n}+c_{4}\psi_{+,n+1}^{n}}\mathbf{\Gamma}_{n+1}\xrightarrow{l^{\prime}}\mathbf{\Gamma}_{n-1}\xrightarrow{(c_{1}\psi_{-,n}^{n-1},c_{2}\psi_{+,n}^{n-1})}\mathbf{\Gamma}_{n}\oplus\mathbf{\Gamma}_{n}.
Proof.

We adopt the conventions from Section 2.3. We will use Lemma 2.6, Lemma 2.11 and Lemma 2.12 without mentioning them. We prove the exactness at 𝚪n1\mathbf{\Gamma}_{n-1} first. We have

ψ±,nn1l=ψ±,nn1ψ±,n1μψ±,μn+1=0.\psi_{\pm,n}^{n-1}\circ l^{\prime}=\psi_{\pm,n}^{n-1}\circ\psi_{\pm,n-1}^{\mu}\circ\psi_{\pm,\mu}^{n+1}=0.

Hence

ker((c1ψ,nn1,c2ψ+,nn1))Im(l).\ker((c_{1}\psi_{-,n}^{n-1},c_{2}\psi_{+,n}^{n-1}))\supset\operatorname{Im}(l^{\prime}).

Then we prove

ker((c1ψ,nn1,c2ψ+,nn1))Im(l).\ker((c_{1}\psi_{-,n}^{n-1},c_{2}\psi_{+,n}^{n-1}))\subset\operatorname{Im}(l^{\prime}).

Suppose

xker((c1ψ,nn1,c2ψ+,nn1))=ker(ψ,nn1)ker(ψ+,nn1).x\in\ker((c_{1}\psi_{-,n}^{n-1},c_{2}\psi_{+,n}^{n-1}))=\ker(\psi_{-,n}^{n-1})\cap\ker(\psi_{+,n}^{n-1}).

By exactness, there exists y𝚪μy\in\mathbf{\Gamma}_{\mu} such that ψ+,n1μ(y)=x\psi_{+,n-1}^{\mu}(y)=x. Then we have

ψ+,nμ(y)=ψ,nn1ψ+,n1μ(y)=ψ,nn1(x)=0.\psi_{+,n}^{\mu}(y)=\psi_{-,n}^{n-1}\circ\psi_{+,n-1}^{\mu}(y)=\psi_{-,n}^{n-1}(x)=0.

By exactness, there exists z𝚪n+1z\in\mathbf{\Gamma}_{n+1} such that ψ+,μn+1(z)=y\psi_{+,\mu}^{n+1}(z)=y. Thus, we have l(z)=xl^{\prime}(z)=x, which concludes the proof for the exactness at 𝚪n1\mathbf{\Gamma}_{n-1}.

Then we prove the exactness at 𝚪n+1\mathbf{\Gamma}_{n+1}. Similarly by exactness, we have

ker(l)Im(c3ψ,n+1n+c4ψ+,n+1n)=Im(ψ,n+1n)+Im(ψ+,n+1n).\ker(l^{\prime})\supset\operatorname{Im}(c_{3}\psi_{-,n+1}^{n}+c_{4}\psi_{+,n+1}^{n})=\operatorname{Im}(\psi_{-,n+1}^{n})+\operatorname{Im}(\psi_{+,n+1}^{n}).

Suppose xker(l)x\in\ker(l^{\prime}). If ψ+,μn+1(x)=0\psi_{+,\mu}^{n+1}(x)=0, then by the exactness, we know xIm(ψ+,n+1n)x\in\operatorname{Im}(\psi_{+,n+1}^{n}). If ψ+,μn+1(x)0\psi_{+,\mu}^{n+1}(x)\neq 0, then by the exactness, there exists y𝚪ny\in\mathbf{\Gamma}_{n} such that

ψ+,μn(y)=ψ+,μn+1(x).\psi_{+,\mu}^{n}(y)=\psi_{+,\mu}^{n+1}(x).

Then we know

xψ,n+1n(y)ker(ψ+,μn+1)=Im(ψ+,n+1n).x-\psi_{-,n+1}^{n}(y)\in\ker(\psi_{+,\mu}^{n+1})=\operatorname{Im}(\psi_{+,n+1}^{n}).

Thus, we have

xIm(ψ,n+1n)+Im(ψ+,n+1n),x\in\operatorname{Im}(\psi_{-,n+1}^{n})+\operatorname{Im}(\psi_{+,n+1}^{n}),

which concludes the proof for the exactness at 𝚪n+1\mathbf{\Gamma}_{n+1}. ∎

6. Some technical constructions

6.1. Filtrations

In this subsection, we study some filtrations on 𝐘\mathbf{Y} and 𝚪μ\mathbf{\Gamma}_{\mu} that will be important in later sections. We continue to adopt conventions from Section 2.3. In particular, KYK\subset Y is a rationally null-homologous knot and SS is a rational Seifert surface of KK.

Lemma 6.1.

The maps GnG_{n} in Lemma 2.16 lead to a filtration on 𝐘\mathbf{Y}: for a sufficiently large integer n0n_{0},

0=kerGn0kerGnkerGn+1kerGn0=𝐘.0=\ker G_{-n_{0}}\subset\dots\subset\ker G_{n}\subset\ker G_{n+1}\subset\dots\subset\ker G_{n_{0}}=\mathbf{Y}.
Proof.

It follows from Lemma 2.19 that when n0n_{0} is sufficiently large we have

0=kerGn0andkerGn0=𝐘.0=\ker G_{-n_{0}}~{}{\rm and~{}}\ker G_{n_{0}}=\mathbf{Y}.

It follows from Lemma 2.18 that for any nn\in\mathbb{Z},

kerGnkerGn+1.\ker G_{n}\subset\ker G_{n+1}.

Lemma 6.2.

For any nn\in\mathbb{Z}, the map GnG_{n} induces an isomorphism

Gn:(kerGn+1/kerGn)kerψ+,n+1nkerψ,n+1n.G_{n}:\bigg{(}\ker G_{n+1}/\ker G_{n}\bigg{)}\xrightarrow{\cong}\ker\psi^{n}_{+,n+1}\cap\ker\psi^{n}_{-,n+1}.
Proof.

Suppose xkerGn+1x\in\ker G_{n+1}. Then from Lemma 2.18 we know that

ψ±,n+1nGn(x)=Gn+1(x)=0.\psi_{\pm,n+1}^{n}\circ G_{n}(x)=G_{n+1}(x)=0.

Hence we have

Gn(kerGn+1))kerψn+,n+1kerψn,n+1.G_{n}(\ker G_{n+1}))\subset\ker\psi^{n}_{+,n+1}\cap\ker\psi^{n}_{-,n+1}.

Clearly GnG_{n} is injective on kerGn+1/kerGn\ker G_{n+1}/\ker G_{n} so it suffices to show that the image is kerψ+,n+1nkerψ,n+1n\ker\psi^{n}_{+,n+1}\cap\ker\psi^{n}_{-,n+1}. To achieve this, for any element xkerψ+,n+1nkerψ,n+1nx\in\ker\psi^{n}_{+,n+1}\cap\ker\psi^{n}_{-,n+1}, Lemma 2.20 implies that

xkerHn=ImGn.x\in\ker H_{n}=\operatorname{Im}G_{n}.

As a result, there exists α𝐘\alpha\in\mathbf{Y} such that

x=Gn(α).x=G_{n}(\alpha).

Again from Lemma 2.18 we know that

Gn+1(α)=ψ+,nn+1Gn(α)=ψ+,nn+1(x)=0.G_{n+1}(\alpha)=\psi^{n+1}_{+,n}\circ G_{n}(\alpha)=\psi^{n+1}_{+,n}(x)=0.

This implies that αkerGn+1.\alpha\in\ker G_{n+1}.

Lemma 6.3.

For any nn\in\mathbb{Z}, the maps ψ±,nμ\psi_{\pm,n}^{\mu} induce isomorphisms

ψ+,nμ:(Imψ+,μn+2/Imψ+,μn+1)kerψ+,n+1nkerψ,n+1n\psi^{\mu}_{+,n}:\bigg{(}\operatorname{Im}\psi^{n+2}_{+,\mu}/\operatorname{Im}\psi^{n+1}_{+,\mu}\bigg{)}\xrightarrow{\cong}\ker\psi^{n}_{+,n+1}\cap\ker\psi^{n}_{-,n+1}
ψ,nμ:(Imψ,μn+2/Imψ,μn+1)kerψ+,n+1nkerψ,n+1n\psi^{\mu}_{-,n}:\bigg{(}\operatorname{Im}\psi^{n+2}_{-,\mu}/\operatorname{Im}\psi^{n+1}_{-,\mu}\bigg{)}\xrightarrow{\cong}\ker\psi^{n}_{+,n+1}\cap\ker\psi^{n}_{-,n+1}
Proof.

We only prove the lemma for positive bypasses. The proof for the negative bypasses is similar. Let uImψ+,μn+2u\in\operatorname{Im}\psi^{n+2}_{+,\mu}. By Lemma 2.6 and Lemma 2.12, we have

ψ+,n+1nψ+,nμ(u)=0andψ,n+1nψ+,nμ(u)=ψ+,n+1μ(u)=0.\psi^{n}_{+,n+1}\circ\psi^{\mu}_{+,n}(u)=0~{}{\rm and}~{}\psi^{n}_{-,n+1}\circ\psi^{\mu}_{+,n}(u)=\psi^{\mu}_{+,n+1}(u)=0.

Hence we know

ψ+,nμ(Imψ+,μn+2)kerψ+,n+1nkerψ,n+1n.\psi^{\mu}_{+,n}(\operatorname{Im}\psi^{n+2}_{+,\mu})\subset\ker\psi^{n}_{+,n+1}\cap\ker\psi^{n}_{-,n+1}.

Since kerψ+,nμ=Imψ+,μn+1\ker\psi^{\mu}_{+,n}=\operatorname{Im}\psi^{n+1}_{+,\mu}, the map ψ+,nμ\psi^{\mu}_{+,n} is injective on Imψ+,μn+2/Imψ+,μn+1\operatorname{Im}\psi^{n+2}_{+,\mu}/\operatorname{Im}\psi^{n+1}_{+,\mu}. To show it is surjective as well, pick xkerψ+,n+1nkerψ,n+1nx\in\ker\psi^{n}_{+,n+1}\cap\ker\psi^{n}_{-,n+1}. Note xkerψ+,n+1n=Imψ+,nμx\in\ker\psi^{n}_{+,n+1}=\operatorname{Im}\psi^{\mu}_{+,n} implies that there exists u𝚪μu\in\mathbf{\Gamma}_{\mu} such that ψ+,nμ(u)=x\psi^{\mu}_{+,n}(u)=x. Lemma 2.12 then implies that

ψ+,n+1μ(u)=ψ,n+1nψ+,nμ(u)=ψ,n+1n(x)=0.\psi^{\mu}_{+,n+1}(u)=\psi^{n}_{-,n+1}\circ\psi^{\mu}_{+,n}(u)=\psi^{n}_{-,n+1}(x)=0.

As a result, ukerψ+,n+1μ=Imψ+,μn+2u\in\ker\psi^{\mu}_{+,n+1}=\operatorname{Im}\psi^{n+2}_{+,\mu}. ∎

Corollary 6.4.
  1. (1)

    For any nn\in\mathbb{Z}, there is a canonical isomorphism

    (kerGn+1/kerGn)(Imψ+,μn+2/Imψ+,μn+1)(Imψ,μn+2/Imψ,μn+1).\bigg{(}\ker G_{n+1}/\ker G_{n}\bigg{)}\cong\bigg{(}\operatorname{Im}\psi^{n+2}_{+,\mu}/\operatorname{Im}\psi^{n+1}_{+,\mu}\bigg{)}\cong\bigg{(}\operatorname{Im}\psi^{n+2}_{-,\mu}/\operatorname{Im}\psi^{n+1}_{-,\mu}\bigg{)}.
  2. (2)

    For sufficiently large n0n_{0}, there exists a (noncanonical) isomorphism

    𝐘(Imψ+,μn0/Imψ+,μn0)(Imψ,μn0/Imψ,μn0)\mathbf{Y}\cong\bigg{(}\operatorname{Im}\psi^{n_{0}}_{+,\mu}/\operatorname{Im}\psi^{-n_{0}}_{+,\mu}\bigg{)}\cong\bigg{(}\operatorname{Im}\psi^{n_{0}}_{-,\mu}/\operatorname{Im}\psi^{-n_{0}}_{-,\mu}\bigg{)}
Definition 6.5.

For any integer nn\in\mathbb{Z} and any grading ii, define the map FniF_{n}^{i} as the restriction

Fni=Fn|(𝚪n,i).F_{n}^{i}=F_{n}|{(\mathbf{\Gamma}_{n},i)}.

where FnF_{n} is the map from Lemma 2.16.

Lemma 6.6.

Suppose n0n_{0}\in\mathbb{Z} is small enough such that Fn0=0F_{n_{0}}=0 (c.f. Lemma 2.19). Then for any integer nn0n\geq n_{0} and any grading ii, we have

ψ±,μn(kerFni)=Im(Projμi(n1)pq2ψ±,μn0),\psi^{n}_{\pm,\mu}(\ker F_{n}^{i})=\operatorname{Im}\bigg{(}{\rm Proj}^{i\mp\frac{(n-1)p-q}{2}}_{\mu}\circ\psi^{n_{0}}_{\pm,\mu}\bigg{)},

where

Projμi(n1)pq2:𝚪μ(𝚪μ,i(n1)pq2){\rm Proj}^{i\mp\frac{(n-1)p-q}{2}}_{\mu}:\mathbf{\Gamma}_{\mu}\to(\mathbf{\Gamma}_{\mu},i\mp\frac{(n-1)p-q}{2})

is the projection.

Proof.

We only prove the lemma for positive bypasses and the proof for negative bypasses is similar. First, suppose

uIm(Projμi(n1)pq2ψ+,μn0)=Imψ+,μn0(𝚪μ,i(n1)pq2).u\in\operatorname{Im}\bigg{(}{\rm Proj}^{i-\frac{(n-1)p-q}{2}}_{\mu}\circ\psi^{n_{0}}_{+,\mu}\bigg{)}=\operatorname{Im}\psi^{n_{0}}_{+,\mu}\cap(\mathbf{\Gamma}_{\mu},i-\frac{(n-1)p-q}{2}).

Pick x(𝚪n0,i(nn0)p2)x\in(\mathbf{\Gamma}_{n_{0}},i-\frac{(n-n_{0})p}{2}) such that

ψ+,μn0(x)=u.\psi^{n_{0}}_{+,\mu}(x)=u.

Taking y=Ψ,nn0(x)y=\Psi^{n_{0}}_{-,n}(x), we know from Lemma 2.6 that y(𝚪n,i),y\in(\mathbf{\Gamma}_{n},i), from Lemma 2.18 that Fn(y)=Fn0(x)=0,F_{n}(y)=F_{n_{0}}(x)=0, and from Lemma 2.12 that ψ+,μn(y)=u.\psi^{n}_{+,\mu}(y)=u. As a result, we conclude uψ+,μn(kerFni)u\in\psi^{n}_{+,\mu}(\ker F_{n}^{i}).

Second, suppose uψ+,μn(kerFni)u\in\psi^{n}_{+,\mu}(\ker F_{n}^{i}) is non-zero. Pick x1kerFnix_{1}\in\ker F_{n}^{i} such that

ψ+,μn(x1)=u.\psi^{n}_{+,\mu}(x_{1})=u.

By Lemma 2.5 and Lemma 2.6, the fact that ψ+,μn(x1)=u0\psi^{n}_{+,\mu}(x_{1})=u\neq 0 implies that

(6.1) pχ(S)2i(n1)pq2pχ(S)2.-\frac{p-\chi(S)}{2}\leq i-\frac{(n-1)p-q}{2}\leq\frac{p-\chi(S)}{2}.

Pick a sufficiently large integer kk and then take

x2=Ψ,n+kn(x1)andx3=Ψ+,2n+2kn0n+k(x2).x_{2}=\Psi^{n}_{-,n+k}(x_{1})~{}{\rm and}~{}x_{3}=\Psi^{n+k}_{+,2n+2k-n_{0}}(x_{2}).

By Lemma 2.18 we have

F2n+2kn0(x3)=Fn+k(x2)=Fn(x1)=0.F_{2n+2k-n_{0}}(x_{3})=F_{n+k}(x_{2})=F_{n}(x_{1})=0.

Note that the grading jj of x3x_{3} equals to

(6.2) j=i+kp2(n+kn0)p2=i(nn0)p2.j=i+\frac{kp}{2}-\frac{(n+k-n_{0})p}{2}=i-\frac{(n-n_{0})p}{2}.

Combining 6.1 and 6.2, we obtain

(n02)pqχ(S)2jn0pqχ(S)2.\frac{(n_{0}-2)p-q-\chi(S)}{2}\leq j\leq\frac{n_{0}p-q-\chi(S)}{2}.

Note that we pick kk to be a sufficiently large integer. In particular, we can assume

(2n+2kn0)p+q+χ(S)2+1(n02)pqχ(S)2\frac{-(2n+2k-n_{0})p+q+\chi(S)}{2}+1\leq\frac{(n_{0}-2)p-q-\chi(S)}{2}

and

n0pqχ(S)2(2n+2kn0)pq+χ(S)2.\frac{n_{0}p-q-\chi(S)}{2}\leq\frac{(2n+2k-n_{0})p-q+\chi(S)}{2}.

Thus jj is in this range as well and then Lemma 2.19 implies that F2n+2kn0F_{2n+2k-n_{0}} is injective on the grading jj. Hence x3=0x_{3}=0. Then the following Lemma 6.7 applies to (x,y)=(x2,0)(x,y)=(x_{2},0) and there exists x4𝚪n0x_{4}\in\mathbf{\Gamma}_{n_{0}} such that

Ψ,n+kn0(x4)=x2.\Psi^{n_{0}}_{-,n+k}(x_{4})=x_{2}.

Thus by Lemma 2.12,

u=ψ+,μn(x1)=ψ+,μn+k(x2)=ψ+,μn0(x4)Im(Projμi(n1)pq2ψ+,μn0).u=\psi^{n}_{+,\mu}(x_{1})=\psi^{n+k}_{+,\mu}(x_{2})=\psi^{n_{0}}_{+,\mu}(x_{4})\in\operatorname{Im}\bigg{(}{\rm Proj}^{i-\frac{(n-1)p-q}{2}}_{\mu}\circ\psi^{n_{0}}_{+,\mu}\bigg{)}.

Lemma 6.7.

Suppose nn\in\mathbb{Z} and k1,k2+k_{1},k_{2}\in\mathbb{N}_{+}. Suppose x𝚪n+k1,y𝚪n+k2x\in\mathbf{\Gamma}_{n+k_{1}},y\in\mathbf{\Gamma}_{n+k_{2}} such that

Ψ+,n+k1+k2n+k1(x)=Ψ,n+k1+k2n+k2(y)\Psi^{n+k_{1}}_{+,n+k_{1}+k_{2}}(x)=\Psi^{n+k_{2}}_{-,n+k_{1}+k_{2}}(y)

Then there exists z𝚪nz\in\mathbf{\Gamma}_{n} such that

Ψ,n+k1n(z)=xandΨ+,n+k2n(z)=y.\Psi^{n}_{-,n+k_{1}}(z)=x~{}{\rm and~{}}\Psi^{n}_{+,n+k_{2}}(z)=y.
Proof.

This is a restatement of Remark 5.2. The proof is similar to that of Proposition 5.1. ∎

6.2. Tau invariants in a general 33-manifold

Definition 6.8.

An element α𝐘\alpha\in\mathbf{Y} is called homogeneous if there exists an nn\in\mathbb{Z} and a grading ii such that αImFni\alpha\in\operatorname{Im}F_{n}^{i}. Note that from 2.18 and Corollary 2.9, we know that

αImFniαImFn+1i±p2.\alpha\in\operatorname{Im}F_{n}^{i}\Rightarrow\alpha\in\operatorname{Im}F_{n+1}^{i\pm\frac{p}{2}}.

For a homogeneous element α𝐘\alpha\in\mathbf{Y}, we pick a sufficiently large n0n_{0} and define

τ+(α)\colonequalsmaxi{i|x(𝚪n0,i),Fn0(x)=α}(n01)pq2\tau^{+}(\alpha)\colonequals\max_{i}\{i~{}|~{}\exists x\in(\mathbf{\Gamma}_{n_{0}},i),~{}F_{n_{0}}(x)=\alpha\}-\frac{(n_{0}-1)p-q}{2}
τ(α)\colonequalsmini{i|x(𝚪n0,i),Fn0(x)=α}+(n01)pq2\tau^{-}(\alpha)\colonequals\min_{i}\{i~{}|~{}\exists x\in(\mathbf{\Gamma}_{n_{0}},i),~{}F_{n_{0}}(x)=\alpha\}+\frac{(n_{0}-1)p-q}{2}
τ(α)\colonequals1+τ(α)τ+(α)+qp=minmaxp+n0.\tau(\alpha)\colonequals 1+\frac{\tau^{-}(\alpha)-\tau^{+}(\alpha)+q}{p}=\frac{\min-\max}{p}+n_{0}.

We will prove the independence of these τ\tau invariants about n0n_{0} later in Lemma 6.12.

Remark 6.9.

Here we fix the knot KYK\subset Y and define the tau invariants for a homogeneous element αI(Y)\alpha\in I^{\sharp}(Y). The reason why we go in this order is because (1) currently the definition of homogeneous elements depends on the choice of the knot and (2) in this paper we only focus on the Dehn surgeries of a fixed knot.

Remark 6.10.

The normalization (n01)pq2\mp\frac{(n_{0}-1)p-q}{2} comes from the grading shifts of ψ±,μn0\psi_{\pm,\mu}^{n_{0}} in Lemma 2.6. When KK is a knot inside Y=S3Y=S^{3}, we have that τ±(α)\tau^{\pm}(\alpha) is equal to the tau invariant τI(K)\tau_{I}(K) defined in [12], where α\alpha is the unique generator of I(S3)I^{\sharp}(-S^{3})\cong\mathbb{C} up to a scalar. Then τ(α)=12τI(K)\tau(\alpha)=1-2\tau_{I}(K).

Lemma 6.11.

We have the following properties.

  1. (1)

    Suppose n1,n2n_{1},n_{2} are two integers and i1,i2i_{1},i_{2} are two gradings such that there exist x1(𝚪n1,i1)x_{1}\in(\mathbf{\Gamma}_{n_{1}},i_{1}) and x2(𝚪n2,i2)x_{2}\in(\mathbf{\Gamma}_{n_{2}},i_{2}) with

    Fn1(x1)=Fn2(x2)0.F_{n_{1}}(x_{1})=F_{n_{2}}(x_{2})\neq 0.

    Then there exists an integer NN such that

    i2=i1(n2n1)p2+Npi_{2}=i_{1}-\frac{(n_{2}-n_{1})p}{2}+Np

    i.e. when we send x1x_{1} and x2x_{2} into the same 𝚪n3\mathbf{\Gamma}_{n_{3}} with n3>n1,n2n_{3}>n_{1},n_{2} by bypass maps, then the difference of the expected gradings of the images is divisible by pp (the grading shifts of the bypass maps ψ±,n+1n\psi_{\pm,n+1}^{n} are p/2\mp p/2).

  2. (2)

    Suppose we have an integer n1n_{1}, a grading i1i_{1}, and an element x1(𝚪n1,i1)x_{1}\in(\mathbf{\Gamma}_{n_{1}},i_{1}). Then for any integer n2n1n_{2}\geq n_{1} and grading i2i_{2} such that there exists an integer N[0,n2n1]N\in[0,n_{2}-n_{1}] with

    i2=i1(n2n1)p2+Np,i_{2}=i_{1}-\frac{(n_{2}-n_{1})p}{2}+Np,

    there exists an element x2(𝚪n2,i2)x_{2}\in(\mathbf{\Gamma}_{n_{2}},i_{2}) such that

    Fn1(x1)=Fn2(x2).F_{n_{1}}(x_{1})=F_{n_{2}}(x_{2}).
  3. (3)

    Suppose nn\in\mathbb{Z} and for 1jl1\leq j\leq l we have a grading iji_{j} and an element xj(𝚪n,ij)x_{j}\in(\mathbf{\Gamma}_{n},i_{j}) such that Fn(x1)F_{n}(x_{1}),…, Fn(xl)F_{n}(x_{l}) are linearly independent. Then the element

    α=j=1lFn(xj)\alpha=\sum_{j=1}^{l}F_{n}(x_{j})

    is homogeneous if and only if for any 1jl1\leq j\leq l, we have

    iji1(modp)i_{j}\equiv i_{1}~{}({\rm mod~{}}p)
Proof.

(1). Take n0n_{0} a sufficiently large integer. For j=1,2j=1,2, take ij(p2,p2]i^{\prime}_{j}\in(-\frac{p}{2},\frac{p}{2}] to be the unique grading such that there exists an integer NjN_{j} with

ij=ij(n0nj)p2+Njp.i_{j}^{\prime}=i_{j}-\frac{(n_{0}-n_{j})p}{2}+N_{j}p.

Take

xj=Ψ+,n0nj+NjΨ,nj+Njnj(xj).x_{j}^{\prime}=\Psi^{n_{j}+N_{j}}_{+,n_{0}}\circ\Psi^{n_{j}}_{-,n_{j}+N_{j}}(x_{j}).

From Lemma 2.18 we know that

xj(𝚪n0,ij)andFn0(x1)=Fn1(x1)=Fn2(x2)=Fn0(x2).x^{\prime}_{j}\in(\mathbf{\Gamma}_{n_{0}},i_{j}^{\prime})~{}{\rm and~{}}F_{n_{0}}(x_{1}^{\prime})=F_{n_{1}}(x_{1})=F_{n_{2}}(x_{2})=F_{n_{0}}(x_{2}^{\prime}).

By Lemma 2.19, we know that x1=x2x_{1}^{\prime}=x_{2}^{\prime} and in particular, i1=i2i_{1}^{\prime}=i_{2}^{\prime}. As a result, we can take N=N1N2N=N_{1}-N_{2} then it is straightforward to verify that

i2=i1(n2n1)p2+Np.i_{2}=i_{1}-\frac{(n_{2}-n_{1})p}{2}+Np.

(2). We can take

x2=Ψ,n2n1+NΨ+,n1+Nn1(x1)x_{2}=\Psi^{n_{1}+N}_{-,n_{2}}\circ\Psi^{n_{1}}_{+,n_{1}+N}(x_{1})

Then it follows from Lemma 2.6 that x2(𝚪n2,i2)x_{2}\in(\mathbf{\Gamma}_{n_{2}},i_{2}) and follows from Lemma 2.18 that

Fn2(x2)=Fn1(x1).F_{n_{2}}(x_{2})=F_{n_{1}}(x_{1}).

(3). The proof is similar to that of (1). ∎

Lemma 6.12.

For a homogeneous element α\alpha, we have the following.

  1. (1)

    τ±(α)\tau^{\pm}(\alpha) and hence τ(α)\tau(\alpha) are well-defined. (i.e. they are independent of the choice of the large integer n0n_{0}.)

  2. (2)

    We have τ(α)\tau(\alpha)\in\mathbb{Z}.

  3. (3)

    For any integer nn and grading ii, the following two statements are equivalent.

    1. (a)

      There exists x(𝚪n,i)x\in(\mathbf{\Gamma}_{n},i) such that Fn(x)=αF_{n}(x)=\alpha.

    2. (b)

      We have nτ(α)n\geq\tau(\alpha) and there exists NN\in\mathbb{Z} such that N[0,nτ(α)]N\in[0,n-\tau(\alpha)] and

      i=τ+(α)+τ(α)(nτ(α))p2+Np.i=\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)-(n-\tau(\alpha))p}{2}+Np.
  4. (4)

    We have

    τ+(α)pχ(S)2andτ(α)pχ(S)2.\tau^{+}(\alpha)\geq-\frac{p-\chi(S)}{2}~{}{\rm and}~{}\tau^{-}(\alpha)\leq\frac{p-\chi(S)}{2}.
Proof.

(1). Suppose α\alpha is a homogeneous element. Then by definition there exists x(𝚪n,i)x\in(\mathbf{\Gamma}_{n},i) for some integer nn and grading ii such that

Fn(x)=α.F_{n}(x)=\alpha.

Then for sufficiently large n0n_{0}, we can take

y=ψ+,n0n(x)y=\psi^{n}_{+,n_{0}}(x)

and from Lemma 2.18 implies that

Fn0(y)=αF_{n_{0}}(y)=\alpha

and hence τ±(α)\tau^{\pm}(\alpha) exists.

To show the value of τ±(α)\tau^{\pm}(\alpha) is independent of n0n_{0} as long as it is sufficiently large, a combination of Lemma 2.5 and Lemma 2.6 implies that the map

ψ,n0+1n0:(𝚪n0,i)(𝚪n0+1,i+p2)\psi^{n_{0}}_{-,n_{0}+1}:(\mathbf{\Gamma}_{n_{0}},i)\to(\mathbf{\Gamma}_{n_{0}+1},i+\frac{p}{2})

is an isomorphism for any i>gn0pq12i>g-\frac{n_{0}p-q-1}{2}. Then Lemma 2.18 implies that τ+\tau^{+} is well-defined. The argument for τ\tau^{-} is similar.

(2). It follows directly from Lemma 6.11 part (1).

(3). We first establish the following claim.

Claim. There exists an element

z(𝚪τ(α),τ+(α)+τ(α)2)z\in(\mathbf{\Gamma}_{\tau(\alpha)},\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2})

such that

Fτ(α)(z)=α.F_{\tau(\alpha)}(z)=\alpha.
Proof of the claim.

Suppose n0n_{0}\in\mathbb{Z} is sufficiently large and

x±(𝚪n0,τ±(α)±(n01)pq2)x_{\pm}\in(\mathbf{\Gamma}_{n_{0}},\tau^{\pm}(\alpha)\pm\frac{(n_{0}-1)p-q}{2})

such that Fn0(x±)=αF_{n_{0}}(x_{\pm})=\alpha. Note that the existence of x±x_{\pm} follows from the definition of τ±(α)\tau^{\pm}(\alpha). Let

x±=Ψ±,2n0τ(α)n(x±).x_{\pm}^{\prime}=\Psi^{n}_{\pm,2n_{0}-\tau(\alpha)}(x_{\pm}).

It follows from Lemma 2.6 that

x±(𝚪2n0τ(α),τ+(α)+τ(α)2).x_{\pm}^{\prime}\in(\mathbf{\Gamma}_{2n_{0}-\tau(\alpha)},\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2}).

From Lemma 2.18 we know that

F2n0τ(α)(x+)=α=F2n0τ(α)(x).F_{2n_{0}-\tau(\alpha)}(x_{+}^{\prime})=\alpha=F_{2n_{0}-\tau(\alpha)}(x_{-}^{\prime}).

By Lemma 2.19 this implies that

x+=x.x_{+}^{\prime}=x_{-}^{\prime}.

Hence Lemma 6.7 applies and there exists z𝚪τ(α)z\in\mathbf{\Gamma}_{\tau(\alpha)} such that

Ψ,nτ(α)(z)=x±.\Psi^{\tau(\alpha)}_{\mp,n}(z)=x_{\pm}.

Again Lemma 2.6 implies that zz is in the grading

z(𝚪τ(α),τ+(α)+τ(α)2)z\in(\mathbf{\Gamma}_{\tau(\alpha)},\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2})

and Lemma 2.18 implies

Fτ(α)(z)=α.F_{\tau(\alpha)}(z)=\alpha.

Now if an integer nn and a grading ii satisfy statement (b), then (a) is a direct consequence of the above claim and Lemma 6.11 part (2).

It remains to show that (a) implies (b). Suppose there exists x(𝚪n,i)x\in(\mathbf{\Gamma}_{n},i) such that Fn(x)=αF_{n}(x)=\alpha. From the above claim, we already know that there exists

z(𝚪τ(α),τ+(α)+τ(α)2)z\in(\mathbf{\Gamma}_{\tau(\alpha)},\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2})

such that

Fτ(α)(z)=αF_{\tau(\alpha)}(z)=\alpha

Hence Lemma 6.11 part (1) implies that there exists NN\in\mathbb{Z} such that

i=τ+(α)+τ(α)(nτ(α))p2+Np.i=\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)-(n-\tau(\alpha))p}{2}+Np.

If N>nτ(α)N>n-\tau(\alpha), we can take a sufficiently large n0n_{0} and

x=Ψ,n0n(x).x^{\prime}=\Psi^{n}_{-,n_{0}}(x).

It follows from Lemma 2.6 that

x(𝚪n0,i)withi>τ+(α)+(n01)pq2.x^{\prime}\in(\mathbf{\Gamma}_{n_{0}},i^{\prime})~{}{\rm with~{}}i^{\prime}>\tau^{+}(\alpha)+\frac{(n_{0}-1)p-q}{2}.

Then Lemma 2.18 implies that

Fn0(x)=αF_{n_{0}}(x^{\prime})=\alpha

which contradicts the definition of τ+\tau^{+} in Definition 6.8. Similarly if N<0N<0 we can take

x=Ψ+,n0n(x)x^{\prime}=\Psi^{n}_{+,n_{0}}(x)

which would be an element contradicting the definition of τ\tau^{-}. When n<τ(α)n<\tau(\alpha) we have nτ(α)<0n-\tau(\alpha)<0 so there is always a contradiction by the above argument. This concludes (b).

(4). It follows from the definition of τ±\tau^{\pm} and Lemma 2.19 that Fn0F_{n_{0}} is an isomorphism when restricted to the direct sum of pp consecutive middle gradings of 𝚪n0\mathbf{\Gamma}_{n_{0}} when n0{n_{0}} is large. ∎

Lemma 6.13.

For any nn\in\mathbb{Z} we have that

ImFn=Span{α𝐘|αhomogeneousandτ(α)n}\operatorname{Im}F_{n}={\rm Span}\{\alpha\in\mathbf{Y}~{}|~{}\alpha~{}{\rm homogeneous~{}and}~{}\tau(\alpha)\leq n\}
Proof.

Suppose αImFn\alpha\in\operatorname{Im}F_{n}. Let

α=iαiwhereαiImFniishomogeneous.\alpha=\sum_{i}\alpha_{i}~{}{\rm where~{}}\alpha_{i}\in\operatorname{Im}F_{n}^{i}~{}{\rm is~{}homogeneous}.

From Lemma 6.12 we know that τ(αi)n\tau(\alpha_{i})\leq n for all ii. On the other hand, suppose

α=iαiwhereτ(αi)nforalli.\alpha=\sum_{i}\alpha_{i}~{}{\rm where}~{}\tau(\alpha_{i})\leq n~{}{\rm for~{}all~{}}i.

By Lemma 6.12 part (3) we can pick zi𝚪τ(αi)z_{i}\in\mathbf{\Gamma}_{\tau(\alpha_{i})} such that

Fτ(αi)(zi)=αi.F_{\tau(\alpha_{i})}(z_{i})=\alpha_{i}.

Then from Lemma 2.18 we know

α=Fn(iΨ+,nτ(αi)(zi)).\alpha=F_{n}(\sum_{i}\Psi^{\tau(\alpha_{i})}_{+,n}(z_{i})).

6.3. A basis for framed instanton homology

We pick a basis 𝔅\mathfrak{B} for 𝐘\mathbf{Y} as follows. First

𝔅=n𝔅n.\mathfrak{B}=\mathop{\bigcup}_{n\in\mathbb{Z}}\mathfrak{B}_{n}.

To construct the set 𝔅n\mathfrak{B}_{n}, first, let 𝔅n=\mathfrak{B}_{n}=\emptyset if Fn=0F_{n}=0. By Lemma 2.19 this means 𝔅n=\mathfrak{B}_{n}=\emptyset for all small enough nn. Write

𝔅n=kn𝔅k.\mathfrak{B}_{\leq n}=\mathop{\bigcup}_{k\leq n}\mathfrak{B}_{k}.

We pick the set 𝔅n\mathfrak{B}_{n} inductively. Note that we have taken 𝔅n=\mathfrak{B}_{n}=\emptyset for nn with Fn=0F_{n}=0. Suppose we have already constructed the set 𝔅n1\mathfrak{B}_{\leq n-1} that consists of homogeneous elements and is a basis of ImFn1\operatorname{Im}F_{n-1}, we pick the set 𝔅n\mathfrak{B}_{n} such that 𝔅n\mathfrak{B}_{n} consists of homogeneous elements with τ=n\tau=n, and the set

𝔅n=𝔅n1𝔅n\mathfrak{B}_{\leq n}=\mathfrak{B}_{\leq n-1}\cup\mathfrak{B}_{n}

forms a basis of ImFn\operatorname{Im}F_{n}. Note that Lemma 6.13 implies that 𝔅n\mathfrak{B}_{n} exists and

|𝔅n|=dim(ImFn/ImFn1).|\mathfrak{B}_{n}|=\dim_{\mathbb{C}}\bigg{(}\operatorname{Im}F_{n}/\operatorname{Im}F_{n-1}\bigg{)}.

For any n,kn,k\in\mathbb{Z} such that kn2k\leq n-2, define maps

η±,kn:𝔅n𝚪k\eta^{n}_{\pm,k}:\mathfrak{B}_{n}\to\mathbf{\Gamma}_{k}

as follows: for any α𝔅nImFn\alpha\in\mathfrak{B}_{n}\subset\operatorname{Im}F_{n}, since α\alpha is homogeneous and τ(α)=n\tau(\alpha)=n, we can pick

z(𝚪n,τ+(α)+τ(α)2)z\in(\mathbf{\Gamma}_{n},\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2})

by Lemma 6.12 part (3) such that Fn(z)=αF_{n}(z)=\alpha. Then define

η±,kn(α)=ψ±,kμψ±,μn(z)\eta^{n}_{\pm,k}(\alpha)=\psi^{\mu}_{\pm,k}\circ\psi^{n}_{\pm,\mu}(z)

.

Lemma 6.14.

Suppose n,kn,k\in\mathbb{Z} such that kn2k\leq n-2.

  1. (1)

    The maps η±,kn\eta^{n}_{\pm,k} are all well-defined.

  2. (2)

    We have η+,n2n=cnη,n2n.\eta^{n}_{+,n-2}=c_{n}\cdot\eta^{n}_{-,n-2}. for some scalar cn\{0}c_{n}\in\mathbb{C}\backslash\{0\}.

  3. (3)

    Elements in Imη±,kn𝚪k\operatorname{Im}\eta^{n}_{\pm,k}\subset\mathbf{\Gamma}_{k} are linearly independent.

  4. (4)

    Imη±,n2n\operatorname{Im}\eta^{n}_{\pm,n-2} forms a basis for kerψ+,n1n2kerψ,n1n2\ker\psi^{n-2}_{+,n-1}\cap\ker\psi^{n-2}_{-,n-1}.

  5. (5)

    For any α𝔅n\alpha\in\mathfrak{B}_{n} we have

    η±,kn(α)(𝚪k,τ+(α)+τ(α)2(n2k)p2).\eta^{n}_{\pm,k}(\alpha)\in(\mathbf{\Gamma}_{k},\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2}\mp\frac{(n-2-k)p}{2}).
  6. (6)

    We have

    ψ,kk1η±,k1n=η±,kn,andψ±,kk1η±,k1n=0.\psi^{k-1}_{\mp,k}\circ\eta^{n}_{\pm,k-1}=\eta^{n}_{\pm,k},~{}{\rm and~{}}\psi^{k-1}_{\pm,k}\circ\eta^{n}_{\pm,k-1}=0.
Proof.

(1). We only work with η+,kn\eta^{n}_{+,k} and the arguments for η,kn\eta^{n}_{-,k} are similar. Suppose there are z1,z2(𝚪n,i)z_{1},z_{2}\in(\mathbf{\Gamma}_{n},i) such that Fn(z1)=Fn(z2)=αF_{n}(z_{1})=F_{n}(z_{2})=\alpha, where i=τ+(α)+τ(α)2i=\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2}. Then

z1z2kerFniz_{1}-z_{2}\in\ker F_{n}^{i}

and by Lemma 6.6 we have

ψ+,μn(z1z2)ψ+,μn(kerFni)Imψ+,μn0Imψ+,μk+1.\psi^{n}_{+,\mu}(z_{1}-z_{2})\in\psi^{n}_{+,\mu}(\ker F_{n}^{i})\subset\operatorname{Im}\psi^{n_{0}}_{+,\mu}\subset\operatorname{Im}\psi^{k+1}_{+,\mu}.

Here n0n_{0}\in\mathbb{Z} is a small enough integer. As a result,

η+,kn(α)=ψ+,kμψ+,μn(z1)=ψ+,kμψ+,μn(z2)\eta^{n}_{+,k}(\alpha)=\psi^{\mu}_{+,k}\circ\psi^{n}_{+,\mu}(z_{1})=\psi^{\mu}_{+,k}\circ\psi^{n}_{+,\mu}(z_{2})

is well-defined.

(2). This follows directly from Lemma 2.11. Note that in Section 2.3 we do not fix the scalars of the second commutative diagram of Lemma 2.11, and hence a non-zero coefficient cnc_{n} would possibly arise.

(3). We only work with η+,kn\eta^{n}_{+,k} and the arguments for η,kn\eta^{n}_{-,k} are similar. Suppose

𝔅n={α1,,αl},wherel=|𝔅n|=dim(ImFn/ImFn1).\mathfrak{B}_{n}=\{\alpha_{1},\dots,\alpha_{l}\},~{}{\rm where~{}}l=|\mathfrak{B}_{n}|=\dim_{\mathbb{C}}\bigg{(}\operatorname{Im}F_{n}/\operatorname{Im}F_{n-1}\bigg{)}.

Suppose there exists λ1,,λl\lambda_{1},\dots,\lambda_{l} such that

j=1lλjη+,kn(αj)=0.\sum_{j=1}^{l}\lambda_{j}\cdot\eta^{n}_{+,k}(\alpha_{j})=0.

Pick zj(𝚪n,τ+(αj)+τ(αj)2)z_{j}\in(\mathbf{\Gamma}_{n},\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})}{2}) such that Fn(zj)=αjF_{n}(z_{j})=\alpha_{j}. Then we have

ψ+,kμψ+,μn(j=1lλjzj)=0.\psi^{\mu}_{+,k}\circ\psi^{n}_{+,\mu}\bigg{(}\sum_{j=1}^{l}\lambda_{j}z_{j}\bigg{)}=0.

As a result, there exists x𝚪k+1x\in\mathbf{\Gamma}_{k+1} such that

ψ+,μk+1(x)=ψ+,μn(j=1lλjzj).\psi^{k+1}_{+,\mu}(x)=\psi^{n}_{+,\mu}\bigg{(}\sum_{j=1}^{l}\lambda_{j}z_{j}\bigg{)}.

Note that, from Lemma 2.12, we know

ψ+,μnΨ,nk+1(x)=ψ+,μk+1(x)=ψ+,μn(j=1lλjzj)\psi^{n}_{+,\mu}\circ\Psi^{k+1}_{-,n}(x)=\psi^{k+1}_{+,\mu}(x)=\psi^{n}_{+,\mu}\bigg{(}\sum_{j=1}^{l}\lambda_{j}z_{j}\bigg{)}

so as a result there exists y𝚪n1y\in\mathbf{\Gamma}_{n-1} such that

j=1lλjzj=Ψ,nk+1(x)+ψ+,nn1(y).\sum_{j=1}^{l}\lambda_{j}z_{j}=\Psi^{k+1}_{-,n}(x)+\psi^{n-1}_{+,n}(y).

Hence by Lemma 2.18 we have

j=1lλjαj\displaystyle\sum_{j=1}^{l}\lambda_{j}\alpha_{j} =Fn(j=1lλjzj)\displaystyle=F_{n}(\sum_{j=1}^{l}\lambda_{j}z_{j})
=FnΨ,nk+1(x)+Fnψ+,nn1(y)\displaystyle=F_{n}\circ\Psi^{k+1}_{-,n}(x)+F_{n}\circ\psi^{n-1}_{+,n}(y)
=Fk+1(x)+Fn1(y)\displaystyle=F_{k+1}(x)+F_{n-1}(y)
ImFn1.\displaystyle\subset\operatorname{Im}F_{n-1}.

Since αj\alpha_{j} form a basis of 𝔅n\mathfrak{B}_{n}, the sum cannot be in ImFn1\operatorname{Im}F_{n-1} except λi=0\lambda_{i}=0 for all ii.

(4). For α𝔅n\alpha\in\mathfrak{B}_{n}, pick z(𝚪n,τ+(α)+τ(α)2)z\in(\mathbf{\Gamma}_{n},\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2}) such that Fn(z)=αF_{n}(z)=\alpha. Then by definition

η+,n2n(α)=ψ+,n2μψ+,μn(z).\eta^{n}_{+,n-2}(\alpha)=\psi^{\mu}_{+,n-2}\circ\psi^{n}_{+,\mu}(z).

Now we can compute

ψ+,n1n2η+,n2n(α)=ψ+,n1n2ψ+,n2μψ+,μn(z)=0,\psi^{n-2}_{+,n-1}\circ\eta^{n}_{+,n-2}(\alpha)=\psi^{n-2}_{+,n-1}\circ\psi^{\mu}_{+,n-2}\circ\psi^{n}_{+,\mu}(z)=0,

and by Lemma 2.12

ψ,n1n2η+,n2n(α)=ψ,n1n2ψ+,n2μψ+,μn(z)=ψ+,n1μψ+,μn(z)=0.\psi^{n-2}_{-,n-1}\circ\eta^{n}_{+,n-2}(\alpha)=\psi^{n-2}_{-,n-1}\circ\psi^{\mu}_{+,n-2}\circ\psi^{n}_{+,\mu}(z)=\psi^{\mu}_{+,n-1}\circ\psi^{n}_{+,\mu}(z)=0.

Hence

η+,n2n(α)kerψ+,n1n2kerψ,n1n2.\eta^{n}_{+,n-2}(\alpha)\in\ker\psi^{n-2}_{+,n-1}\cap\ker\psi^{n-2}_{-,n-1}.

Then (4) follows from (3), Lemma 6.2, and ImFn=kerGn1\operatorname{Im}F_{n}=\operatorname{ker}G_{n-1}.

(5). It follows directly from the construction of η±,kn\eta^{n}_{\pm,k} and Lemma 2.6.

(6). It follows from the construction of η±,kn\eta^{n}_{\pm,k}, the commutativity in Lemma 2.12 and the exactness in Lemma 2.6. ∎

Convention.

We can define

η~+,kn=η+,knandη~,kn=cnη,kn\tilde{\eta}_{+,k}^{n}=\eta_{+,k}^{n}~{}{\rm and}~{}\tilde{\eta}_{-,k}^{n}=c_{n}\cdot\eta_{-,k}^{n}

such that

η~+,n2n=η~,n2n\tilde{\eta}_{+,n-2}^{n}=\tilde{\eta}_{-,n-2}^{n}

and the new maps satisfy all properties in Lemma 6.14 except (2). We will use η+,kn\eta_{+,k}^{n} to denote η~+,kn\tilde{\eta}_{+,k}^{n} in latter sections.

7. The map in the third exact triangle

In this section, we construct the map ll in Proposition 3.9 and Proposition 3.12 and show it satisfies the exactness and the commutative diagram. We continue to adopt conventions from Section 2.3. We restate the propositions as follows and no longer use the notations l,ll,l^{\prime} for maps.

Proposition 7.1.

Suppose nn\in\mathbb{Z} is fixed and kk\in\mathbb{Z} is sufficiently large. Then there is an exact triangle

𝚪n\textstyle{\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φn+kn\scriptstyle{\Phi^{n}_{n+k}}𝚪n+k𝚪n+k\textstyle{\mathbf{\Gamma}_{n+k}\oplus\mathbf{\Gamma}_{n+k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φn+2kn+k\scriptstyle{\Phi^{n+k}_{n+2k}}𝚪n+2k.\textstyle{\mathbf{\Gamma}_{n+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces.}Φnn+2k\scriptstyle{\Phi^{n+2k}_{n}}

where two of the maps are already constructed

Φn+kn\colonequals(Ψ+,n+kn,Ψ,n+kn):𝚪n𝚪n+k𝚪n+k\Phi^{n}_{n+k}\colonequals(\Psi^{n}_{+,n+k},\Psi^{n}_{-,n+k}):\mathbf{\Gamma}_{n}\to\mathbf{\Gamma}_{n+k}\oplus\mathbf{\Gamma}_{n+k}
Φn+2kn+k\colonequalsΨ,n+2kn+kΨ+,n+2kn+k:𝚪n+k𝚪n+k𝚪n+2k.\Phi^{n+k}_{n+2k}\colonequals\Psi^{n+k}_{-,n+2k}-\Psi^{n+k}_{+,n+2k}:\mathbf{\Gamma}_{n+k}\oplus\mathbf{\Gamma}_{n+k}\to\mathbf{\Gamma}_{n+2k}.
Proposition 7.2.

Suppose nn\in\mathbb{Z} is fixed and kk\in\mathbb{Z} is sufficiently large. Suppose Φn1n+2k1\Phi_{n-1}^{n+2k-1} is constructed in Proposition 7.1. Then, there are two commutative diagrams up to scalars.

𝚪2n+2k+12\textstyle{\mathbf{\Gamma}_{\frac{2n+2k+1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ+,n+2kn+k+1ψ,n+k+12n+2k+12\scriptstyle{\Psi^{n+k+1}_{+,n+2k}\circ\psi^{\frac{2n+2k+1}{2}}_{-,n+k+1}}ψ+,μn+k+1ψ,n+k+12n+2k+12\scriptstyle{\psi^{n+k+1}_{+,\mu}\circ\psi^{\frac{2n+2k+1}{2}}_{-,n+k+1}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,nμ\scriptstyle{\psi^{\mu}_{+,n}}𝚪n+2k\textstyle{\mathbf{\Gamma}_{n+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φnn+2k\scriptstyle{\Phi^{n+2k}_{n}}𝚪n\textstyle{\mathbf{\Gamma}_{n}}
𝚪2n+2k+12\textstyle{\mathbf{\Gamma}_{\frac{2n+2k+1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ,n+2k+1n+k+1ψ+,n+k+12n+2k+12\scriptstyle{\Psi^{n+k+1}_{-,n+2k+1}\circ\psi^{\frac{2n+2k+1}{2}}_{+,n+k+1}}ψ,μn+k+1ψ+,n+k+12n+2k+12\scriptstyle{\psi^{n+k+1}_{-,\mu}\circ\psi^{\frac{2n+2k+1}{2}}_{+,n+k+1}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ,nμ\scriptstyle{\psi^{\mu}_{-,n}}𝚪n+2k\textstyle{\mathbf{\Gamma}_{n+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φnn+2k\scriptstyle{\Phi^{n+2k}_{n}}𝚪n\textstyle{\mathbf{\Gamma}_{n}}

7.1. Characterizations of the kernel and the image

Before constructing Φnn+2k\Phi^{n+2k}_{n}, we characterize the spaces kerΦn+kn\ker\Phi^{n}_{n+k} and ImΦn+2kn+k\operatorname{Im}\Phi^{n+k}_{n+2k}. These results will motivate the construction of Φnn+2k\Phi^{n+2k}_{n} to ensure that

ImΦnn+2k=kerΦn+knandkerΦnn+2k=ImΦn+2kn+k.\operatorname{Im}\Phi^{n+2k}_{n}=\ker\Phi^{n}_{n+k}~{}{\rm and}~{}\ker\Phi^{n+2k}_{n}=\operatorname{Im}\Phi^{n+k}_{n+2k}.

Since Φn+kn\Phi^{n}_{n+k} and Φn+2kn+k\Phi^{n+k}_{n+2k} are constructed using bypass maps, it suffices to consider their restrictions on each grading.

Lemma 7.3.

Suppose nn\in\mathbb{Z} is fixed and kk\in\mathbb{Z} is sufficiently large. Let

Projni:𝚪n(𝚪n,i){\rm Proj}^{i}_{n}:\mathbf{\Gamma}_{n}\rightarrow(\mathbf{\Gamma}_{n},i)

be the projection. Then we have

kerΦn+kn(𝚪n,i)=Im(ProjniGn).\ker\Phi^{n}_{n+k}\cap(\mathbf{\Gamma}_{n},i)=\operatorname{Im}\bigg{(}{\rm Proj}^{i}_{n}\circ G_{n}\bigg{)}.
Proof.

We need to apply Lemma 2.20. Following conventions in Section 2.3, we have

(7.1) Hn=ψ+,n+1nψ,n+1n.H_{n}=\psi^{n}_{+,n+1}-\psi^{n}_{-,n+1}.

Suppose xIm(ProjniGn)x\in\operatorname{Im}\bigg{(}{\rm Proj}^{i}_{n}\circ G_{n}\bigg{)}. Pick α𝐘\alpha\in\mathbf{Y} and y𝚪ny\in\mathbf{\Gamma}_{n} such that

Gn(α)=x+yG_{n}(\alpha)=x+y

where Projni(y)=0{\rm Proj}^{i}_{n}(y)=0. When kk is sufficiently large, we know from Lemma 2.19 that

Gn+k0.G_{n+k}\equiv 0.

In particular, from Lemma 2.18

Ψ±,n+kn(x)+Ψ±,n+kn(y)=Gn+k(α)=0.\Psi^{n}_{\pm,n+k}(x)+\Psi^{n}_{\pm,n+k}(y)=G_{n+k}(\alpha)=0.

Since the maps Ψ±,n+kn\Psi^{n}_{\pm,n+k} are homogeneous, we know that

Ψ±,n+kn(x)=0,\Psi^{n}_{\pm,n+k}(x)=0,

which implies that xkerΦn+kn(𝚪n,i)x\in\ker\Phi^{n}_{n+k}\cap(\mathbf{\Gamma}_{n},i).

Next, suppose xkerΦn+kn(𝚪n,i)x\in\ker\Phi^{n}_{n+k}\cap(\mathbf{\Gamma}_{n},i). We take xni=xx^{i}_{n}=x and we will pick xnj(𝚪n,j)x^{j}_{n}\in(\mathbf{\Gamma}_{n},j) for all jij\neq i such that

jxnjkerHn=ImGn.\sum_{j}x^{j}_{n}\in\ker H_{n}=\operatorname{Im}G_{n}.

We will use the notation xabx_{a}^{b} to denote an element in (𝚪a,b)(\mathbf{\Gamma}_{a},b). Recall that from Lemma 2.6, the grading shifts of ψ±,n+1n\psi_{\pm,n+1}^{n} are p2\mp\frac{p}{2}. Take

xn+k1i+(k1)p2=Ψ,n+k1n(x)andxn+k1i+(k+1)p2=0.x^{i+\frac{(k-1)p}{2}}_{n+k-1}=\Psi^{n}_{-,n+k-1}(x)~{}{\rm and~{}}x^{i+\frac{(k+1)p}{2}}_{n+k-1}=0.

Since xkerΦn+kn(𝚪n,i)x\in\ker\Phi^{n}_{n+k}\cap(\mathbf{\Gamma}_{n},i) we know that

(7.2) ψ,n+kn+k1(xn+k1i+(k1)p2)=Ψ,n+kn(x)=0=ψ+,n+kn+k1(xn+k1i+(k+1)p2).\psi^{n+k-1}_{-,n+k}(x^{i+\frac{(k-1)p}{2}}_{n+k-1})=\Psi^{n}_{-,n+k}(x)=0=\psi^{n+k-1}_{+,n+k}(x^{i+\frac{(k+1)p}{2}}_{n+k-1}).

Hence from Lemma 6.7, there exists

xn+k2i+kp2(𝚪n+k2,i+kp2)x_{n+k-2}^{i+\frac{kp}{2}}\in(\mathbf{\Gamma}_{n+k-2},i+\frac{kp}{2})

such that

ψ,n+k1n+k2(xn+k2i+kp2)=xn+k1i+(k+1)p2=0andψ+,n+k1n+k2(xn+k2i+kp2)=xn+k1i+(k1)p2.\psi^{n+k-2}_{-,n+k-1}(x_{n+k-2}^{i+\frac{kp}{2}})=x^{i+\frac{(k+1)p}{2}}_{n+k-1}=0~{}{\rm and~{}}\psi^{n+k-2}_{+,n+k-1}(x_{n+k-2}^{i+\frac{kp}{2}})=x^{i+\frac{(k-1)p}{2}}_{n+k-1}.

Then we can take

xn+k2i+(k+2)p2=0andxn+k2i+(k2)p2=Ψ,n+k2n(x).x_{n+k-2}^{i+\frac{(k+2)p}{2}}=0~{}{\rm and}~{}x_{n+k-2}^{i+\frac{(k-2)p}{2}}=\Psi^{n}_{-,n+k-2}(x).

We can apply the same argument and use Lemma 6.7 to find

xn+k3i+(k+3)p2,xn+k3i+(k+1)p2,xn+k3i+(k1)p2,xn+k3i+(k3)p2𝚪n+k3x_{n+k-3}^{i+\frac{(k+3)p}{2}},~{}x_{n+k-3}^{i+\frac{(k+1)p}{2}},~{}x_{n+k-3}^{i+\frac{(k-1)p}{2}},~{}x_{n+k-3}^{i+\frac{(k-3)p}{2}}\in\mathbf{\Gamma}_{n+k-3}

such that ψ±,n+k2n+k3\psi^{n+k-3}_{\pm,n+k-2} send them to corresponding elements in 𝚪n+k2\mathbf{\Gamma}_{n+k-2}. Repeating this argument, we can obtain elements

xni+pj(𝚪n,i+pj)forj[1,k]x_{n}^{i+pj}\in(\mathbf{\Gamma}_{n},i+pj)~{}{\rm for~{}}j\in[1,k]\cap\mathbb{Z}

such that xni=xx_{n}^{i}=x, ψ,n+1n(xni+pk)=0\psi^{n}_{-,n+1}(x_{n}^{i+pk})=0, ψ+,n+1n(xnipk)=0\psi^{n}_{+,n+1}(x_{n}^{i-pk})=0, and for any j[1,k1]j\in[1,k-1]\cap\mathbb{Z} we have

ψ,n+1n(xni+pj)=ψ+,n+1n(xni+p(j+1)).\psi^{n}_{-,n+1}(x_{n}^{i+pj})=\psi^{n}_{+,n+1}(x_{n}^{i+p(j+1)}).

Note that we obtain the above xni+pjx_{n}^{i+pj} for j[1,k]j\in[1,k]\cap\mathbb{Z} essentially from the fact that Ψ,n+kn(x)=0\Psi^{n}_{-,n+k}(x)=0 as in Equation (7.2). However, xkerΦn+knx\in{\rm ker}\Phi^{n}_{n+k} so we have Ψ+,n+kn(x)=0\Psi^{n}_{+,n+k}(x)=0 as well. A similar argument as above then yields

xni+pj(𝚪n,i+pj)forj[k,1].x_{n}^{i+pj}\in(\mathbf{\Gamma}_{n},i+pj)~{}{\rm for~{}}j\in[-k,-1]\cap\mathbb{Z}.

Together with xni=xx_{n}^{i}=x, we obtain xni+pjx_{n}^{i+pj} for all j[k,k]j\in[-k,k]\cap\mathbb{Z}.

It is then straightforward to check that

y=j=kkxni+pjker(ψ+,n+1nψ,n+1n)=kerHn=ImGn.y=\sum_{j=-k}^{k}x_{n}^{i+pj}\in\ker(\psi^{n}_{+,n+1}-\psi^{n}_{-,n+1})=\ker H_{n}=\operatorname{Im}G_{n}.

Lemma 7.4.

Suppose α𝐘\alpha\in\mathbf{Y} is a homogeneous element and

α=j=1lλjαj\alpha=\sum_{j=1}^{l}\lambda_{j}\cdot\alpha_{j}

where λj0\lambda_{j}\neq 0 and αj𝔅\alpha_{j}\in\mathfrak{B} for 1jl1\leq j\leq l. Let nn be an integer, ii be a grading and kk be a sufficiently large integer. For an element x(𝚪n+2k,i)x\in(\mathbf{\Gamma}_{n+2k},i) such that Fn+2k(x)=αF_{n+2k}(x)=\alpha, the following is true.

  1. (1)

    We have

    τ+(α)=min1jl{τ+(αj)}andτ(α)=max1jl{τ(αj)}.\tau^{+}(\alpha)=\min_{1\leq j\leq l}\{\tau^{+}(\alpha_{j})\}~{}{\rm and~{}}\tau^{-}(\alpha)=\max_{1\leq j\leq l}\{\tau^{-}(\alpha_{j})\}.
  2. (2)

    We have xImΦn+2kn+kx\in\operatorname{Im}\Phi^{n+k}_{n+2k} if and only if for any 1jl1\leq j\leq l, at least one of the following inequalities holds

    iτ(αj)(n1)pq2andiτ+(αj)+(n1)pq2.i\geq\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2}~{}{\rm and}~{}i\leq\tau^{+}(\alpha_{j})+\frac{(n-1)p-q}{2}.
  3. (3)

    If xImΦn+2kn+kx\notin\operatorname{Im}\Phi^{n+k}_{n+2k} then there exists j,Nj,N\in\mathbb{Z} such that 1jl1\leq j\leq l, 0Nτ(αj)n20\leq N\leq\tau(\alpha_{j})-n-2, and

    i=τ+(αj)+(n1)pq2+(N+1)p.i=\tau^{+}(\alpha_{j})+\frac{(n-1)p-q}{2}+(N+1)p.
Proof.

(1). We only demonstrate the proof of the result for τ+\tau^{+} and the proof for τ\tau^{-} is similar. First, We make the following two claims.

Claim 1. For any homogeneous elements (not necessarily elements in 𝔅\mathfrak{B}) α1\alpha_{1} and α2\alpha_{2} such that α1+α2\alpha_{1}+\alpha_{2} is also homogeneous, if τ+(α1)>τ+(α2)\tau^{+}(\alpha_{1})>\tau^{+}(\alpha_{2}) then τ+(α1+α2)=τ+(α2).\tau^{+}(\alpha_{1}+\alpha_{2})=\tau^{+}(\alpha_{2}).

To prove Claim 1, let n0n_{0} be sufficiently large. From Lemma 6.11 part (3) we know that

(7.3) τ+(α1)τ+(α2)τ+(α1+α2)(modp).\tau^{+}(\alpha_{1})\equiv\tau^{+}(\alpha_{2})\equiv\tau^{+}(\alpha_{1}+\alpha_{2})~{}({\rm mod~{}}p).

Assume τ+(α1+α2)>τ+(α2)\tau^{+}(\alpha_{1}+\alpha_{2})>\tau^{+}(\alpha_{2}). Let

τ+=min{τ+(α1),τ+(α1+α2)}>τ+(α2).\tau^{+}=\min\{\tau^{+}(\alpha_{1}),\tau^{+}(\alpha_{1}+\alpha_{2})\}>\tau^{+}(\alpha_{2}).

We claim that there exist

x1,x3(𝚪n0,τ++(n01)pq2)x_{1},x_{3}\in(\mathbf{\Gamma}_{n_{0}},\tau^{+}+\frac{(n_{0}-1)p-q}{2})

such that

Fn0(x1)=α1andFn0(x3)=α1+α2.F_{n_{0}}(x_{1})=\alpha_{1}~{}{\rm and}~{}F_{n_{0}}(x_{3})=\alpha_{1}+\alpha_{2}.

We prove only the existence of x1x_{1}, and the argument for the existence of x3x_{3} is similar. By Definition 6.8, we know that

τ+(α1)+(n01)pq2=τ+(α1)+τ(α1)(n0τ(α1))p2+(n0τ(α1))p.\tau^{+}(\alpha_{1})+\frac{(n_{0}-1)p-q}{2}=\frac{\tau^{+}(\alpha_{1})+\tau^{-}(\alpha_{1})-(n_{0}-\tau(\alpha_{1}))p}{2}+(n_{0}-\tau(\alpha_{1}))p.

Taking

N=(n0τ(α1))1p(τ+(α1)τ+),N=(n_{0}-\tau(\alpha_{1}))-\frac{1}{p}(\tau^{+}(\alpha_{1})-\tau^{+}),

we know that

(7.4) τ++(n01)pq2=τ+(α1)+τ(α1)(n0τ(α1))p2+Np.\tau^{+}+\frac{(n_{0}-1)p-q}{2}=\frac{\tau^{+}(\alpha_{1})+\tau^{-}(\alpha_{1})-(n_{0}-\tau(\alpha_{1}))p}{2}+Np.

Equation (7.3) implies that NN\in\mathbb{Z}. The definition of τ+\tau^{+} makes sure that Nn0τ(α1)N\leq n_{0}-\tau(\alpha_{1}). The fact that n0n_{0} is sufficiently large and Lemma 6.12 part (4) implies that N0N\geq 0. Hence Lemma 6.12 part (3) implies the existence of x1x_{1} such that

x1(𝚪n0,τ++(n01)pq2) and Fn0(x1)=α1.x_{1}\in(\mathbf{\Gamma}_{n_{0}},\tau^{+}+\frac{(n_{0}-1)p-q}{2})\text{ and }F_{n_{0}}(x_{1})=\alpha_{1}.

Now the existence of x1x_{1} and x3x_{3} implies that

Fn(x3x1)=α2F_{n}(x_{3}-x_{1})=\alpha_{2}

which contradicts the definition of τ+(α2)\tau^{+}(\alpha_{2}).

Claim 2. Suppose α1,,αu𝔅\alpha_{1},\dots,\alpha_{u}\in\mathfrak{B} are pairwise distinct elements in 𝔅\mathfrak{B} such that

τ+(α1)=τ+(α2)==τ+(αu)=τ+.\tau^{+}(\alpha_{1})=\tau^{+}(\alpha_{2})=\dots=\tau^{+}(\alpha_{u})=\tau^{+}.

Suppose

α=i=1uλiαi\alpha^{\prime}=\sum_{i=1}^{u}\lambda_{i}\cdot\alpha_{i}

and suppose it is homogeneous. Then τ+(α)=τ+\tau^{+}(\alpha^{\prime})=\tau^{+}.

To prove Claim 2, assume that τ+(α)>τ+\tau^{+}(\alpha^{\prime})>\tau^{+}. Without loss of generality, assume that λ10\lambda_{1}\neq 0 and

τ(α1)=min1ju{τ(αj)}.\tau^{-}(\alpha_{1})=\min_{1\leq j\leq u}\{\tau^{-}(\alpha_{j})\}.

Then a similar argument as in the proof of Claim 1 implies that

τ(α)τ(α1).\tau^{-}(\alpha^{\prime})\leq\tau^{-}(\alpha_{1}).

Note that we have assumed τ+(α)>τ+=τ+(α1)\tau^{+}(\alpha^{\prime})>\tau^{+}=\tau^{+}(\alpha_{1}). Hence by Definition 6.8, τ(α)<τ(α1)\tau(\alpha^{\prime})<\tau(\alpha_{1}), which contradicts the construction of the set 𝔅\mathfrak{B}.

Now we prove part (1). Suppose α1,,αl𝔅\alpha_{1},\dots,\alpha_{l}\in\mathfrak{B} are pairwise distinct elements in 𝔅\mathfrak{B}. Let

α=j=1lλjαj.\alpha=\sum_{j=1}^{l}\lambda_{j}\cdot\alpha_{j}.

We want to show that

τ+(α)=min1jl{τ+(αj)}.\tau^{+}(\alpha)=\min_{1\leq j\leq l}\{\tau^{+}(\alpha_{j})\}.

To do this, relabel the elements αj\alpha_{j} if necessary such that

τ+(α1)=τ+(α2)==τ+(αu)<τ+(αu+1)τ+(αu+2)τ+(αl).\tau^{+}(\alpha_{1})=\tau^{+}(\alpha_{2})=\dots=\tau^{+}(\alpha_{u})<\tau^{+}(\alpha_{u+1})\leq\tau^{+}(\alpha_{u+2})\leq\dots\leq\tau^{+}(\alpha_{l}).

Since α\alpha is homogeneous, from Lemma 6.11 part (3), we know that the sum

j=1vλjαj\sum_{j=1}^{v}\lambda_{j}\cdot\alpha_{j}

is also homogeneous for any v=1,,lv=1,\dots,l. Applying Claim 2, we conclude that

τ+(j=1uλjαj)=τ+(α1).\tau^{+}\bigg{(}\sum_{j=1}^{u}\lambda_{j}\cdot\alpha_{j}\bigg{)}=\tau^{+}(\alpha_{1}).

Hence we can apply Claim 1 repeatedly to conclude that

τ+(j=1lλjαj)=τ+(α1)=min1jl{τ+(αj)}.\tau^{+}\bigg{(}\sum_{j=1}^{l}\lambda_{j}\cdot\alpha_{j}\bigg{)}=\tau^{+}(\alpha_{1})=\min_{1\leq j\leq l}\{\tau^{+}(\alpha_{j})\}.

(2). If xImΦn+2kn+kx\in\operatorname{Im}\Phi^{n+k}_{n+2k}, then there exists y(𝚪n+k,ikp2)y\in(\mathbf{\Gamma}_{n+k},i-\frac{kp}{2}) and z(𝚪n+k,i+kp2)z\in(\mathbf{\Gamma}_{n+k},i+\frac{kp}{2}) such that

x=Ψ,n+2kn+k(y)+Ψ+,n+2kn+k(z).x=\Psi^{n+k}_{-,n+2k}(y)+\Psi^{n+k}_{+,n+2k}(z).

By assumption

Fn+2k(x)=α=j=1lλjαjF_{n+2k}(x)=\alpha=\sum_{j=1}^{l}\lambda_{j}\cdot\alpha_{j}

with λj0\lambda_{j}\neq 0 and α\alpha homogeneous. By Lemma 2.18 we have

α=Fn+k(y+z).\alpha=F_{n+k}(y+z).

Since 𝔅\mathfrak{B} forms a basis for 𝐘\mathbf{Y}, we can write

Fn+k(y)=j=1lλjαjandFn+k(z)=j=1lλj′′αj,F_{n+k}(y)=\sum_{j=1}^{l}\lambda_{j}^{\prime}\cdot\alpha_{j}~{}{\rm and~{}}F_{n+k}(z)=\sum_{j=1}^{l}\lambda_{j}^{\prime\prime}\cdot\alpha_{j},

where l=|𝔅|l=|\mathfrak{B}|. Then for any 1jl1\leq j\leq l, at least one of λj\lambda_{j}^{\prime} and λj′′\lambda_{j}^{\prime\prime} is non-zero. Since both Fn+k(y)F_{n+k}(y) and Fn+k(z)F_{n+k}(z) are homogeneous, from part (1) we know

ikp2τ(αj)(n+k1)pq2 when λj0i-\frac{kp}{2}\geq\tau^{-}(\alpha_{j})-\frac{(n+k-1)p-q}{2}\text{ when }\lambda_{j}^{\prime}\neq 0
and i+kp2τ+(αj)+(n+k1)pq2 when λj′′0.\text{and }i+\frac{kp}{2}\leq\tau^{+}(\alpha_{j})+\frac{(n+k-1)p-q}{2}\text{ when }\lambda_{j}^{\prime\prime}\neq 0.

Conversely, suppose for any 1jl1\leq j\leq l at least one of the following inequalities holds

iτ(αj)(n1)pq2andiτ+(αj)+(n1)pq2.i\geq\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2}~{}{\rm and}~{}i\leq\tau^{+}(\alpha_{j})+\frac{(n-1)p-q}{2}.

We need to show xImΦn+2kn+kx\in\operatorname{Im}\Phi^{n+k}_{n+2k}. We deal with three cases.

Case 1. The grading ii satisfies

i(n+2k2)pq+χ(S)2.i\geq\frac{(n+2k-2)p-q+\chi(S)}{2}.

We want to argue that

Ψ,n+2kn+k:(𝚪n+k,ikp2)(𝚪n+2k,i)\Psi^{n+k}_{-,n+2k}:(\mathbf{\Gamma}_{n+k},i-\frac{kp}{2})\to(\mathbf{\Gamma}_{n+2k},i)

is surjective and hence conclude that xImΦn+2kn+kx\in\operatorname{Im}\Phi^{n+k}_{n+2k}. To do this, note that Ψ,n+2kn+k|(𝚪n+k,ikp2)\Psi^{n+k}_{-,n+2k}|_{(\mathbf{\Gamma}_{n+k},i-\frac{kp}{2})} is the composition of maps ψ,n+k+j+1n+k+j|(𝚪n+k+j,ikp2+jp2)\psi^{n+k+j}_{-,n+k+j+1}|_{(\mathbf{\Gamma}_{n+k+j},i-\frac{kp}{2}+\frac{jp}{2})} for j=0,1,,k1j=0,1,...,k-1. With the assumption of Case 1, we have

ikp2+jp2(n+k+j2)pq+χ(S)2>(n+k+j)pq+χ(S)2.i-\frac{kp}{2}+\frac{jp}{2}\geq\frac{(n+k+j-2)p-q+\chi(S)}{2}>-\frac{(n+k+j)p-q+\chi(S)}{2}.

(Note that since kk is sufficiently large this is a very loose inequality.) Then Corollary 2.9 part (2) applies and we conclude that ψ,n+k+j+1n+k+j|(𝚪n+k+j,ikp2+jp2)\psi^{n+k+j}_{-,n+k+j+1}|_{(\mathbf{\Gamma}_{n+k+j},i-\frac{kp}{2}+\frac{jp}{2})} is an isomorphism for all j=0,1,,k1j=0,1,...,k-1. Hence we conclude Case 1.

Case 2. The grading ii satisfies

i(n+2k2)pq+χ(S)2.i\leq-\frac{(n+2k-2)p-q+\chi(S)}{2}.

The argument is similar to that for Case 1, except for using Ψ+,n+2kn+k\Psi^{n+k}_{+,n+2k} instead of Ψ,n+2kn+k\Psi^{n+k}_{-,n+2k}.

Case 3. If the grading ii satisfies

|i|<(n+2k2)pq+χ(S)2|i|<\frac{(n+2k-2)p-q+\chi(S)}{2}

Under the assumption of Case 3, Lemma 2.19 part (1) implies that Fn+2kF_{n+2k} is injective when restricted to (𝚪n+2k,i)(\mathbf{\Gamma}_{n+2k},i).

Now, for each j=1,2,,lj=1,2,...,l, if we have

iτ(αj)(n1)pq2,i\geq\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2},

we claim that there exists yj(𝚪n+k,ikp2)y_{j}\in(\mathbf{\Gamma}_{n+k},i-\frac{kp}{2}) such that

Fn+k(yj)=λjαj.F_{n+k}(y_{j})=\lambda_{j}\cdot\alpha_{j}.

If instead

iτ+(αj)+(n1)pq2,i\leq\tau^{+}(\alpha_{j})+\frac{(n-1)p-q}{2},

we claim that there exists zj(𝚪n+k,i+kp2)z_{j}\in(\mathbf{\Gamma}_{n+k},i+\frac{kp}{2}) such that

Fn+k(zj)=λjαjF_{n+k}(z_{j})=\lambda_{j}\cdot\alpha_{j}

We will verify the existence of yjy_{j} or zjz_{j} in a moment, but for now let yy be the sum of all yjy_{j}’s and zz be the sum of all zjz_{j}’s. Then from Lemma 2.18 it is straightforward to check that

Fn+2k(Ψ,n+2kn+k(y)+Ψ+,n+2kn+k(z))=α=Fn+2k(x).F_{n+2k}(\Psi^{n+k}_{-,n+2k}(y)+\Psi^{n+k}_{+,n+2k}(z))=\alpha=F_{n+2k}(x).

Since in Case 3 the restriction of Fn+2kF_{n+2k} on (𝚪n+2k,i)(\mathbf{\Gamma}_{n+2k},i) is injective, we conclude that

x=Ψ,n+2kn+k(y)+Ψ+,n+2kn+k(z)ImΦn+2kn+k.x=\Psi^{n+k}_{-,n+2k}(y)+\Psi^{n+k}_{+,n+2k}(z)\in\operatorname{Im}\Phi^{n+k}_{n+2k}.

It remains to show that the desired yjy_{j} or zjz_{j} exists. We only prove the existence of yjy_{j} and the argument for zjz_{j} is similar. Now assume that

iτ(αj)(n1)pq2.i\geq\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2}.

This implies that

ikp2τ(αj)(n+k1)pq2.i-\frac{kp}{2}\geq\tau^{-}(\alpha_{j})-\frac{(n+k-1)p-q}{2}.

The hypothesis of the Lemma and the definition of τ(αj)\tau^{-}(\alpha_{j}) in Definition 6.8, together with Lemma 6.11 part (3) imply that

iτ(αj)(n1)pq2(mod p).i\equiv\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2}~{}(\text{mod }p).

As a result, there exists an integer N0N\geq 0 such that

ikp2=τ(αj)(n+k1)pq2+Np.i-\frac{kp}{2}=\tau^{-}(\alpha_{j})-\frac{(n+k-1)p-q}{2}+Np.

Note that, from Definition 6.8, we know

τ(αj)(n+k1)pq2=τ+(αj)+τ(αj)(n+kτ(αj))p2.\tau^{-}(\alpha_{j})-\frac{(n+k-1)p-q}{2}=\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(n+k-\tau(\alpha_{j}))p}{2}.

As a result, we have

ikp2=τ+(αj)+τ(αj)(n+kτ(αj))p2+Np.i-\frac{kp}{2}=\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(n+k-\tau(\alpha_{j}))p}{2}+Np.

The assumption in Case 3 and Lemma 6.12 part (4) then implies that N(n+k)τ(αj)N\leq(n+k)-\tau(\alpha_{j}). Hence Lemma 6.12 part (3) implies the existence of yjy_{j}.

(3). If xImΦn+2kn+kx\notin\operatorname{Im}\Phi^{n+k}_{n+2k}, then part (2) means that there exists some jj such that

τ+(αj)+(n1)pq2<i<τ(αj)(n1)pq2.\tau^{+}(\alpha_{j})+\frac{(n-1)p-q}{2}<i<\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2}.

Note that, by Lemma 6.11 part (3), we must have

iτ(αj)(n1)pq2τ+(αj)+(n1)pq2(modp).i\equiv\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2}\equiv\tau^{+}(\alpha_{j})+\frac{(n-1)p-q}{2}~{}({\rm mod}~{}p).

By direct calculation, we have

(τ(αj)(n1)pq2)(τ+(αj)+(n1)pq2)=(τ(αj)n)p(\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2})-(\tau^{+}(\alpha_{j})+\frac{(n-1)p-q}{2})=(\tau(\alpha_{j})-n)p

Then we can choose NN with 0Nτ(αj)n20\leq N\leq\tau(\alpha_{j})-n-2 as desired. ∎

7.2. The construction of the map

Since Φn+kn\Phi^{n}_{n+k} and Φn+2kn+k\Phi^{n+k}_{n+2k} are homogeneous, we can construct Φnn+2k\Phi^{n+2k}_{n} for each grading to achieve both the exactness and the commutativity. Given the grading shifts in Lemma 2.6 and Lemma 2.13, the map Φnn+2k\Phi^{n+2k}_{n} preserves the gradings. From Lemma 2.5, for any grading ii with

|i|>|npq|χ(S)2,|i|>\frac{|np-q|-\chi(S)}{2},

we have (𝚪n,i)=0(\mathbf{\Gamma}_{n},i)=0. From Corollary 2.9, we know either Ψ+,n+2kn+k\Psi_{+,n+2k}^{n+k} or Ψ,n+2kn+k\Psi_{-,n+2k}^{n+k} is surjective onto (𝚪n+2k,i)(\mathbf{\Gamma}_{n+2k},i) for such grading ii. Thus, on such grading ii, the zero map satisfies the exactness for Φnn+2k\Phi^{n+2k}_{n} (though we still have to verify the commutativity in Proposition 7.2).

On the other hand, from Lemma 2.19, the restriction of Fn+2kF_{n+2k} on the consecutive pp middle gradings is an isomorphism. In particular, when p=1p=1, it is an isomorphism when restricted to each middle grading. Also from Lemma 7.3, it seems that the definition of Φn+kn+2k\Phi^{n+2k}_{n+k} on (𝚪n+2k,i)(\mathbf{\Gamma}_{n+2k},i) should involve ProjniGn{\rm Proj}^{i}_{n}\circ G_{n}. However, if we simply take

ProjniGnFn+2k{\rm Proj}^{i}_{n}\circ G_{n}\circ F_{n+2k}

as the definition, the current techniques fall short of demonstrating exactness and commutativity.

We resolve this issue by introducing an isomorphism

I:𝐘𝐘I:\mathbf{Y}\xrightarrow{\cong}\mathbf{Y}

and define

(7.5) Φnn+2k(x)=ProjniGnIFn+2k(x) for x(𝚪n+2k,i).\Phi_{n}^{n+2k}(x)={\rm Proj}^{i}_{n}\circ G_{n}\circ I\circ F_{n+2k}(x)\text{ for }x\in(\mathbf{\Gamma}_{n+2k},i).

The construction of II is noncanonical but it helps us to prove the exactness and commutativity.

Remark 7.5.

In the first arXiv version of this paper, we deal with the special case Y=S3Y=S^{3}. In this case 𝐘\mathbf{Y}\cong\mathbb{C} so up to a scalar we have I=IdI=\operatorname{Id}. In this special case indeed we could prove the exactness and commutativity without explicitly writing down the isomorphism II as follows.

We first define the map II on the basis

𝔅=n𝔅n\mathfrak{B}=\mathop{\cup}_{n\in\mathbb{Z}}\mathfrak{B}_{n}

of 𝐘\mathbf{Y} chosen in Section 6.3 that consists of homogeneous elements and then extend the map on the whole space linearly. We will show it is an isomorphism.

Fix n0n_{0}\in\mathbb{Z} small enough such that Corollary 2.9 and Lemma 2.19 apply. For any α𝔅n\alpha\in\mathfrak{B}_{n}, there exists a grading i(α)(p2,p2]i(\alpha)\in(-\frac{p}{2},\frac{p}{2}] such that there exists N(α)N(\alpha)\in\mathbb{Z} with

i(α)=τ+(α)+τ(α)2(τ(α)2n0)p2+N(α)p.i(\alpha)=\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2}-\frac{(\tau(\alpha)-2-n_{0})p}{2}+N(\alpha)p.

Note that, from the above equality, we know that

N(α)=τ(α)2n02+2i(α)τ+(α)τ(α)2p.N(\alpha)=\frac{\tau(\alpha)-2-n_{0}}{2}+\frac{2i(\alpha)-\tau^{+}(\alpha)-\tau^{-}(\alpha)}{2p}.

Note that, except for n0n_{0}, the rest of the terms are bounded, so N(α)0N(\alpha)\geq 0 when n0n_{0} is small enough. Similarly,

N(α)+n0=τ(α)2+n02+2i(α)τ+(α)τ(α)2pN(\alpha)+n_{0}=\frac{\tau(\alpha)-2+n_{0}}{2}+\frac{2i(\alpha)-\tau^{+}(\alpha)-\tau^{-}(\alpha)}{2p}

so we have N(α)+n0τ(α)2N(\alpha)+n_{0}\leq\tau(\alpha)-2 when n0n_{0} is small enough. As a result, by Lemma 6.14 part (2) and (6) (and the convention after the lemma), we know that for any α𝔅n\alpha\in\mathfrak{B}_{n}

Ψ,τ(α)2n0+N(α)(η+,n0+N(α)τ(α)(α))=η+,τ(α)2τ(α)(α)=η,τ(α)2τ(α)(α)=Ψ+,τ(α)2τ(α)2N(α)(η,τ(α)2N(α)τ(α)(α)).\Psi^{n_{0}+N(\alpha)}_{-,\tau(\alpha)-2}\bigg{(}\eta^{\tau(\alpha)}_{+,n_{0}+N(\alpha)}(\alpha)\bigg{)}=\eta^{\tau(\alpha)}_{+,\tau(\alpha)-2}(\alpha)=\eta^{\tau(\alpha)}_{-,\tau(\alpha)-2}(\alpha)=\Psi^{\tau(\alpha)-2-N(\alpha)}_{+,\tau(\alpha)-2}\bigg{(}\eta^{\tau(\alpha)}_{-,\tau(\alpha)-2-N(\alpha)}(\alpha)\bigg{)}.

Then by Lemma 6.7, there exists w(𝚪n0,i(α))w\in(\mathbf{\Gamma}_{n_{0}},i(\alpha)) such that

(7.6) Ψ+,n0+N(α)n0(w)=η+,n0+N(α)τ(α)(α)andΨ,τ(α)2N(α)n0(w)=η,τ(α)2N(α)τ(α)(α).\Psi^{n_{0}}_{+,n_{0}+N(\alpha)}(w)=\eta^{\tau(\alpha)}_{+,n_{0}+N(\alpha)}(\alpha)~{}{\rm and}~{}\Psi^{n_{0}}_{-,\tau(\alpha)-2-N(\alpha)}(w)=\eta^{\tau(\alpha)}_{-,\tau(\alpha)-2-N(\alpha)}(\alpha).

Let

Proj:𝚪n0i(p2,p2](𝚪n0,i).{\rm Proj}:\mathbf{\Gamma}_{n_{0}}\to\bigoplus_{i\in(-\frac{p}{2},\frac{p}{2}]}(\mathbf{\Gamma}_{n_{0}},i).

From Lemma 2.19, we know

ProjGn0:𝐘i(p2,p2](𝚪n0,i){\rm Proj}\circ G_{n_{0}}:\mathbf{Y}\to\bigoplus_{i\in(-\frac{p}{2},\frac{p}{2}]}(\mathbf{\Gamma}_{n_{0}},i)

is an isomorphism. Hence we define

I(α)=(ProjGn0)1(w).I(\alpha)=({\rm Proj}\circ G_{n_{0}})^{-1}(w).

The following diagram might be helpful for understanding the construction of II. (We write n=τ(α)n=\tau(\alpha), n1=τ(α)2N(α)n_{1}=\tau(\alpha)-2-N(\alpha), and n2=n0+N(α)n_{2}=n_{0}+N(\alpha).)

η,n1n(α)𝚪n1\textstyle{\eta^{n}_{-,n_{1}}(\alpha)\in\mathbf{\Gamma}_{n_{1}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ+,n2n1\scriptstyle{\Psi^{n_{1}}_{+,n-2}}w𝚪n0\textstyle{w\in\mathbf{\Gamma}_{n_{0}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ,n1n0\scriptstyle{\Psi^{n_{0}}_{-,n_{1}}}Ψ+,n2n0\scriptstyle{\Psi^{n_{0}}_{+,n_{2}}}η,n2n(α)=η+,n2n(α)𝚪n2\textstyle{\eta^{n}_{-,n-2}(\alpha)=\eta^{n}_{+,n-2}(\alpha)\in\mathbf{\Gamma}_{n-2}}z𝚪n\textstyle{z\in\mathbf{\Gamma}_{n}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fn\scriptstyle{F_{n}}η+,n2n(α)𝚪n2\textstyle{\eta^{n}_{+,n_{2}}(\alpha)\in\mathbf{\Gamma}_{n_{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ,n2n2\scriptstyle{\Psi^{n_{2}}_{-,n-2}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}I(α)𝐘\textstyle{I(\alpha)\in\mathbf{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Gn0\scriptstyle{G_{n_{0}}}α𝐘\textstyle{\alpha\in\mathbf{Y}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}I\scriptstyle{I}
Remark 7.6.

For a general 33-manifold YY, our construction of II is noncanonical since there are many choices such as the basis 𝔅\mathfrak{B} and the element ww for each α𝔅\alpha\in\mathfrak{B}. However, one could still ask whether we could simply pick I=IdI=\operatorname{Id} or not. If we take I=IdI=\operatorname{Id}, then Proposition 7.2 can finally be reduced to Conjecture 7.7 which we state below. We believe that the following conjecture is true, though currently, we do not find a proof for it. Hence in order to fulfill the main purpose of the paper, we introduce the isomorphism II to bypass this conjecture.

Conjecture 7.7.

For any α𝔅\alpha\in\mathfrak{B}, and any integer nτ(α)2n\leq\tau(\alpha)-2, we have

η±,nτ(α)(α)=Projnj±Gn(α),\eta^{\tau(\alpha)}_{\pm,n}(\alpha)={\rm Proj}^{j_{\pm}}_{n}\circ G_{n}(\alpha),

where

j±=τ+(α)+τ(α)2(τ(α)2n)p2j_{\pm}=\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2}\mp\frac{(\tau(\alpha)-2-n)p}{2}

and

Projnj±:𝚪n(𝚪n,j±){\rm Proj}^{j_{\pm}}_{n}:\mathbf{\Gamma}_{n}\to(\mathbf{\Gamma}_{n},j_{\pm})

is the projection.

Lemma 7.8.

We have the following.

  1. (1)

    Suppose α𝔅\alpha\in\mathfrak{B} and n0n_{0}, ww, N(α)N(\alpha) are chosen as above. Suppose n,kn,k are two integers such that n0knn_{0}\leq k\leq n. Then (a) Ψ,nkΨ+,kn0(w)0\Psi^{k}_{-,n}\circ\Psi^{n_{0}}_{+,k}(w)\neq 0 if and only if (b) kn0+N(α)k\leq n_{0}+N(\alpha) and nkτ(α)2n0N(α)n-k\leq\tau(\alpha)-2-n_{0}-N(\alpha) (in particular, we have nτ(α)2n\leq\tau(\alpha)-2).

  2. (2)

    The map I:𝐘𝐘I:\mathbf{Y}\to\mathbf{Y} is an isomorphism.

  3. (3)

    For an element α𝔅\alpha\in\mathfrak{B}, an integer nn and a grading ii, the following two statements are equivalent.

    1. (a)

      We have ProjniGnI(α)0.{\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha)\neq 0.

    2. (b)

      We have nτ(α)2n\leq\tau(\alpha)-2 and there exists NN\in\mathbb{Z} such that N[0,τ(α)2n]N\in[0,\tau(\alpha)-2-n] and

      i=τ+(α)+τ(α)(τ(α)2n)p2+Np.i=\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)-(\tau(\alpha)-2-n)p}{2}+Np.
  4. (4)

    Suppose for an integer nn and a grading ii we have α1,,αL𝔅\alpha_{1},\dots,\alpha_{L}\in\mathfrak{B} such that ProjniGnI(αj)0{\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha_{j})\neq 0 for all 1jL1\leq j\leq L, then ProjniGnI(α1){\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha_{1}),…, ProjniGnI(αL){\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha_{L}) are linearly independent.

  5. (5)

    Suppose α𝔅\alpha\in\mathfrak{B}. For any nn\in\mathbb{Z} such that nτ(α)2n\leq\tau(\alpha)-2, we have

    Projni±GnI(α)=η±,nτ(α)(α)wherei±=τ+(α)+τ(α)(τ(α)2n)2{\rm Proj}^{i_{\pm}}_{n}\circ G_{n}\circ I(\alpha)=\eta^{\tau(\alpha)}_{\pm,n}(\alpha)~{}{\rm where~{}}i_{\pm}=\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)\mp(\tau(\alpha)-2-n)}{2}
Proof.

(1). First, when k>n0+N(α)k>n_{0}+N(\alpha), from the construction of ww, we know that

Ψ+,kn0(w)=Ψ+,kn0+N(α)Ψ+,n0+N(α)n0(w)=Ψ+,kn0+N(α)η+,n0+N(α)τ(α)(α)=0.\Psi^{n_{0}}_{+,k}(w)=\Psi^{n_{0}+N(\alpha)}_{+,k}\circ\Psi^{n_{0}}_{+,n_{0}+N(\alpha)}(w)=\Psi^{n_{0}+N(\alpha)}_{+,k}\circ\eta^{\tau(\alpha)}_{+,n_{0}+N(\alpha)}(\alpha)=0.

The last equality is from Lemma 6.14 part (6). Similarly, if nk>τ(α)2n0N(α)n-k>\tau(\alpha)-2-n_{0}-N(\alpha), we know from Lemma 2.11 that

Ψ,nkΨ+,kn0(w)\displaystyle\Psi^{k}_{-,n}\circ\Psi^{n_{0}}_{+,k}(w) =Ψ+,nn+n0kΨ,n+n0kn0(w)\displaystyle=\Psi^{n+n_{0}-k}_{+,n}\circ\Psi^{n_{0}}_{-,n+n_{0}-k}(w)
=Ψ+,nn+n0kΨ,n+n0kτ(α)2N(α)Ψ,τ(α)2N(α)n0(w)\displaystyle=\Psi^{n+n_{0}-k}_{+,n}\circ\Psi^{\tau(\alpha)-2-N(\alpha)}_{-,n+n_{0}-k}\circ\Psi^{n_{0}}_{-,\tau(\alpha)-2-N(\alpha)}(w)
(Definitionofw)\displaystyle({\rm Definition~{}of~{}}w) =Ψ+,nn+n0kΨ,n+n0kτ(α)2N(α)η,τ(α)2N(α)τ(α)(α)\displaystyle=\Psi^{n+n_{0}-k}_{+,n}\circ\Psi^{\tau(\alpha)-2-N(\alpha)}_{-,n+n_{0}-k}\circ\eta^{\tau(\alpha)}_{-,\tau(\alpha)-2-N(\alpha)}(\alpha)
(Lemma6.14part(6))\displaystyle({\rm Lemma}~{}\ref{lem: the map eta^n_pm,k}~{}{\rm part~{}(6)}) =0.\displaystyle=0.

Next, we need to show that Ψ,nkΨ+,kn0(w)0\Psi^{k}_{-,n}\circ\Psi^{n_{0}}_{+,k}(w)\neq 0 when kn0+N(α)k\leq n_{0}+N(\alpha) and nkτ(α)2n0N(α)n-k\leq\tau(\alpha)-2-n_{0}-N(\alpha). Again, from Lemma 2.11 we have

Ψ+,n+N(α)+n0knΨ,nkΨ+,kn0(w)\displaystyle\Psi^{n}_{+,n+N(\alpha)+n_{0}-k}\circ\Psi^{k}_{-,n}\circ\Psi^{n_{0}}_{+,k}(w) =Ψ,n0+N(α)+nkn0+N(α)Ψ+,n0+N(α)n0(w)\displaystyle=\Psi^{n_{0}+N(\alpha)}_{-,n_{0}+N(\alpha)+n-k}\circ\Psi^{n_{0}}_{+,n_{0}+N(\alpha)}(w)
(Definitionofw)\displaystyle({\rm Definition~{}of~{}}w) =Ψ,n0+N(α)+nkn0+N(α)η+,n0+N(α)τ(α)(α)\displaystyle=\Psi^{n_{0}+N(\alpha)}_{-,n_{0}+N(\alpha)+n-k}\circ\eta^{\tau(\alpha)}_{+,n_{0}+N(\alpha)}(\alpha)
(Lemma6.14part(6))\displaystyle({\rm Lemma}~{}\ref{lem: the map eta^n_pm,k}~{}{\rm part~{}(6)}) =η+,n0+N(α)+nkτ(α)(α)\displaystyle=\eta^{\tau(\alpha)}_{+,n_{0}+N(\alpha)+n-k}(\alpha)
(Lemma6.14part(3))\displaystyle({\rm Lemma}~{}\ref{lem: the map eta^n_pm,k}~{}{\rm part~{}(3)}) 0.\displaystyle\neq 0.

(2). Suppose 𝔅={α1,,αL}\mathfrak{B}=\{\alpha_{1},\dots,\alpha_{L}\} where L=dim𝐘L=\dim_{\mathbb{C}}\mathbf{Y}. We order the elements αi\alpha_{i} such that

τ(αi)τ(αi+1).\tau(\alpha_{i})\geq\tau(\alpha_{i+1}).

Let wiw_{i}, Ni=N(αi)N_{i}=N(\alpha_{i}) be the data associated to αi\alpha_{i} as above. Since

wij(p2,p2](𝚪n0,j)w_{i}\in\bigoplus_{j\in(-\frac{p}{2},\frac{p}{2}]}(\mathbf{\Gamma}_{n_{0}},j)

for any ii, and by Lemma 2.19, the map

ProjGn0:𝐘j(p2,p2](𝚪n0,j){\rm Proj}\circ G_{n_{0}}:\mathbf{Y}\to\bigoplus_{j\in(-\frac{p}{2},\frac{p}{2}]}(\mathbf{\Gamma}_{n_{0}},j)

is an isomorphism, in order to show that II is an isomorphism, it suffices to show that w1w_{1},…, wLw_{L} are linearly independent.

Now suppose there are complex numbers λ1\lambda_{1},…, λL\lambda_{L} such that

i=1Lλiwi=0.\sum_{i=1}^{L}\lambda_{i}w_{i}=0.

Our goal is to show that all λi\lambda_{i} are zero. The idea is to apply various maps Ψ+,τ(α1)2n0+N1Ψ,n0+N1n0\Psi^{n_{0}+N_{1}}_{+,\tau(\alpha_{1})-2}\circ\Psi^{n_{0}}_{-,n_{0}+N_{1}} to filter out different indices by part (1) of the lemma.

Applying the map Ψ+,τ(α1)2n0+N1Ψ,n0+N1n0\Psi^{n_{0}+N_{1}}_{+,\tau(\alpha_{1})-2}\circ\Psi^{n_{0}}_{-,n_{0}+N_{1}}, from the construction of wiw_{i}, the order of αi\alpha_{i} and part (1) of the lemma, we know

0\displaystyle 0 =Ψ+,τ(α1)2n0+N1Ψ,n0+N1n0(i=1Lλiwi)\displaystyle=\Psi^{n_{0}+N_{1}}_{+,\tau(\alpha_{1})-2}\circ\Psi^{n_{0}}_{-,n_{0}+N_{1}}(\sum_{i=1}^{L}\lambda_{i}w_{i})
=αiλiη±τ(α1)2τ(α1)(αi)\displaystyle=\sum_{\alpha_{i}}\lambda_{i}\eta^{\tau(\alpha_{1})}_{\pm\tau(\alpha_{1})-2}(\alpha_{i})

where the summation in the second line is over all αi\alpha_{i} with

τ(αi)=τ(α1),andNi=N1.\tau(\alpha_{i})=\tau(\alpha_{1}),~{}{\rm and}~{}N_{i}=N_{1}.

Note that, from Lemma 6.14 part (2) and the convention after the lemma, we know η+τ(α1)2τ(α1)=ητ(α1)2τ(α1)\eta^{\tau(\alpha_{1})}_{+\tau(\alpha_{1})-2}=\eta^{\tau(\alpha_{1})}_{-\tau(\alpha_{1})-2}. From Lemma 6.14 again, we know that η±τ(α1)2τ(α1)(αi)\eta^{\tau(\alpha_{1})}_{\pm\tau(\alpha_{1})-2}(\alpha_{i}) are linearly independent, and as a result all relevant λi\lambda_{i} must be zero. Suppose i0i_{0} is the smallest index in the rest. By our choice of αi\alpha_{i}, the element αi0\alpha_{i_{0}} has the largest τ\tau among the rest of the αi\alpha_{i}. Hence we can apply the map Ψ+,τ(αi0)2n0+Ni0Ψ,n0+Ni0n0\Psi^{n_{0}+N_{i_{0}}}_{+,\tau(\alpha_{i_{0}})-2}\circ\Psi^{n_{0}}_{-,n_{0}+N_{i_{0}}} to filter out αi\alpha_{i} with smaller τ\tau. Repeating this argument, we could prove that all λi\lambda_{i} must be zero.

(3). Let n0n_{0}, ww, i(α)i(\alpha), and N(α)N(\alpha) be constructed as above. We first prove that (b) \Rightarrow (a). Note that, when constructing the isomorphism II, from Corollary 2.9 and Lemma 2.18 we can take n0=n02n_{0}^{\prime}=n_{0}-2 and w=(ψ+,n01n02)1(ψ+,n0n01)1(w)w^{\prime}=(\psi^{n_{0}-2}_{+,n_{0}-1})^{-1}\circ(\psi^{n_{0}-1}_{+,n_{0}})^{-1}(w) that will lead to the same II as n0n_{0}, ww. (Note that, by construction, passing from n0n_{0} to n02n_{0}-2 will increase N(α)N(\alpha) by 11.)

Remark 7.9.

Note that the main goal of the current paper is to derive an integral surgery formula. For nn that is sufficiently large, we already know a large surgery as in [22]. When nn is small enough, we can pass to n-n for the mirror of the knot. As a result, instead of changing n0n_{0} for particular nn, we could assume a universal bound for all the integers nn that we care about and make n0n_{0} universally small.

As a result, we can always assume that n0n_{0} is small enough compared with any given nn. Now recall by construction

i(α)=τ+(α)+τ(α)2(τ(α)2n0)p2+N(α)pi(\alpha)=\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)}{2}-\frac{(\tau(\alpha)-2-n_{0})p}{2}+N(\alpha)p

and by the assumption in (b) we have

i=τ+(α)+τ(α)(τ(α)2n)p2+Np.i=\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)-(\tau(\alpha)-2-n)p}{2}+Np.

We can assume that n0n_{0} is small enough such that N(α)>NN(\alpha)>N. Take k=N(α)N+n0k=N(\alpha)-N+n_{0}. Note that, by construction we have w(𝚪n0,i(α))w\in(\mathbf{\Gamma}_{n_{0}},i(\alpha)), so from Lemma 2.6 we know that

Ψ,nkΨ+,kn0(w)(𝚪n,i(α)(kn0)p2+(nk)p2)=(𝚪n,i).\Psi^{k}_{-,n}\circ\Psi^{n_{0}}_{+,k}(w)\in(\mathbf{\Gamma}_{n},i(\alpha)-\frac{(k-n_{0})p}{2}+\frac{(n-k)p}{2})=(\mathbf{\Gamma}_{n},i).

As a result, we conclude from the definition of ww and Lemma 2.18 that

Ψ,nkΨ+,kn0(w)\displaystyle\Psi^{k}_{-,n}\circ\Psi^{n_{0}}_{+,k}(w) =Ψ,nkΨ+,kn0Projnoi(α)Gn0I(α)\displaystyle=\Psi^{k}_{-,n}\circ\Psi^{n_{0}}_{+,k}\circ{\rm Proj}_{n_{o}}^{i(\alpha)}\circ G_{n_{0}}\circ I(\alpha)
=ProjniΨ,nkΨ+,kn0Gn0I(α)\displaystyle={\rm Proj}^{i}_{n}\circ\Psi^{k}_{-,n}\circ\Psi^{n_{0}}_{+,k}\circ G_{n_{0}}\circ I(\alpha)
=ProjniGnI(α)\displaystyle={\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha)

Then it is straightforward to verify that kn0N(α)k-n_{0}\leq N(\alpha) and nkτ(α)2n0N(α)n-k\leq\tau(\alpha)-2-n_{0}-N(\alpha). As a result, we conclude from part (1) that

ProjniGnI(α)0.{\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha)\neq 0.

Next we show that (a) \Rightarrow (b). Again assume that n0n_{0} is small enough compared with the given nn. Then there exists i(p2,p2]i^{\prime}\in(-\frac{p}{2},\frac{p}{2}] such that there exists NN^{\prime}\in\mathbb{Z} with

i=i(nn0)p2+Np.i^{\prime}=i-\frac{(n-n_{0})p}{2}+N^{\prime}p.

By Lemma 2.18, we know that

ProjniGnI(α)=Ψ,nn0+NΨ+,n0+Nn0Projn0iGn0I(α).{\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha)=\Psi^{n_{0}+N^{\prime}}_{-,n}\circ\Psi^{n_{0}}_{+,n_{0}+N^{\prime}}\circ{\rm Proj}^{i^{\prime}}_{n_{0}}\circ G_{n_{0}}\circ I(\alpha).

From the construction of I(α)I(\alpha) and Lemma 2.19 we know ProjniGnI(α)0{\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha)\neq 0 only if i=i(α)i^{\prime}=i(\alpha), in which case

ProjniGnI(α)=Ψ,nn0+NΨ+,n0+Nn0Projn0iGn0I(α)=Ψ,nn0+NΨ+,n0+Nn0(w).{\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha)=\Psi^{n_{0}+N^{\prime}}_{-,n}\circ\Psi^{n_{0}}_{+,n_{0}+N^{\prime}}\circ{\rm Proj}^{i^{\prime}}_{n_{0}}\circ G_{n_{0}}\circ I(\alpha)=\Psi^{n_{0}+N^{\prime}}_{-,n}\circ\Psi^{n_{0}}_{+,n_{0}+N^{\prime}}(w).

Hence ProjniGnI(α)0{\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha)\neq 0 implies that

NN(α)andnNτ(α)2N(α)N^{\prime}\leq N(\alpha)~{}{\rm and~{}}n-N^{\prime}\leq\tau(\alpha)-2-N(\alpha)

by part (1). Taking N=N(α)NN=N(\alpha)-N^{\prime}, it is then straightforward to check that

N[0,τ(α)2n]andi=τ+(α)+τ(α)(τ(α)2n)p2+Np.N\in[0,\tau(\alpha)-2-n]~{}{\rm and~{}}i=\frac{\tau^{+}(\alpha)+\tau^{-}(\alpha)-(\tau(\alpha)-2-n)p}{2}+Np.

(4). The proof is similar to that of (2).

(5). It follows from the proofs of parts (1) and (3). ∎

7.3. The exact triangle

In this subsection, we prove the exact triangle. Note that we choose the basis 𝔅\mathfrak{B} of 𝐘\mathbf{Y} as in Section 6.3.

Proof of Proposition 7.1.

We will verify the exactness at each space of the triangle.

The exactness at 𝚪n+k𝚪n+k\mathbf{\Gamma}_{n+k}\oplus\mathbf{\Gamma}_{n+k}. This follows from Proposition 5.1.

The exactness at 𝚪n\mathbf{\Gamma}_{n}. From Lemma 7.3 and the construction of Φnn+2k\Phi^{n+2k}_{n} in (7.5), we know that ImΦnn+2kkerΦn+kn\operatorname{Im}\Phi^{n+2k}_{n}\subset\ker\Phi^{n}_{n+k}. Now pick an arbitrary

x(𝚪n,i)kerΦn+kn=Im(ProjniGn)x\in(\mathbf{\Gamma}_{n},i)\cap\ker\Phi^{n}_{n+k}=\operatorname{Im}\bigg{(}{\rm Proj}^{i}_{n}\circ G_{n}\bigg{)}

Since II is an isomorphism, we can assume that

x=j=1lProjniGn(λjI(αj))x=\sum_{j=1}^{l}{\rm Proj}^{i}_{n}\circ G_{n}(\lambda_{j}\cdot I(\alpha_{j}))

where αj𝔅\alpha_{j}\in\mathfrak{B} and ProjniGnI(αj)0{\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha_{j})\neq 0. From Lemma 7.8 part (3), we know that this implies that for any j[1,l]j\in[1,l]\cap\mathbb{Z}, we have nτ(αj)2n\leq\tau(\alpha_{j})-2 and there exists NjN_{j}\in\mathbb{Z} such that Nj[0,τ(αj)2n]N_{j}\in[0,\tau(\alpha_{j})-2-n]

i=τ+(αj)+τ(αj)(τ(αj)2n)p2+Njp.i=\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(\tau(\alpha_{j})-2-n)p}{2}+N_{j}p.

Now, for kk sufficiently large, we have n+2k>τ(αj)n+2k>\tau(\alpha_{j}). Taking

Nj=n+k+1τ(αj)+Nj,N_{j}^{\prime}=n+k+1-\tau(\alpha_{j})+N_{j}\in\mathbb{Z},

it is straightforward to verify that when kk is sufficiently large, we have

Nj[0,n+2kτ(αj)]andi=τ+(αj)+τ(αj)(n+2kτ(αj))p2+Njp.N_{j}^{\prime}\in[0,n+2k-\tau(\alpha_{j})]~{}{\rm and~{}}~{}i=\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(n+2k-\tau(\alpha_{j}))p}{2}+N_{j}^{\prime}p.

Hence by Lemma 6.12 part (3), there exists yj(𝚪n+2k,i)y_{j}\in(\mathbf{\Gamma}_{n+2k},i) such that Fn+2k(yj)=αjF_{n+2k}(y_{j})=\alpha_{j}. As a result, it is straightforward to check that

x=Φnn+2k(j=1lλjyj)ImΦnn+2k.x=\Phi^{n+2k}_{n}\bigg{(}\sum_{j=1}^{l}\lambda_{j}\cdot y_{j}\bigg{)}\in\operatorname{Im}\Phi^{n+2k}_{n}.

The exactness at 𝚪n+2k\mathbf{\Gamma}_{n+2k}. Suppose x(𝚪n+2k,i)x\in(\mathbf{\Gamma}_{n+2k},i) and

Fn+2k(x)=j=1lλjαjF_{n+2k}(x)=\sum_{j=1}^{l}\lambda_{j}\cdot\alpha_{j}

with λj0\lambda_{j}\neq 0 and αj𝔅\alpha_{j}\in\mathfrak{B}.

First, if xImΦn+2kn+kx\in\operatorname{Im}\Phi^{n+k}_{n+2k}, then from Lemma 7.4 part (2), we know that for any 1jl1\leq j\leq l, we have

eitheriτ(αj)(n1)pq2oriτ+(αj)+(n1)pq2.{\rm either}~{}i\geq\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2}~{}{\rm or~{}}i\leq\tau^{+}(\alpha_{j})+\frac{(n-1)p-q}{2}.

If we write

i=τ+(αj)+τ(αj)(τ(αj)2n)p2+Njp,i=\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(\tau(\alpha_{j})-2-n)p}{2}+N_{j}p,

for some NjN_{j} then the inequality

iτ(αj)(n1)pq2i\geq\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2}

implies that

Nj\displaystyle N_{j} 1p(τ(αj)(n1)pq2τ+(αj)+τ(αj)(τ(αj)2n)p2)\displaystyle\geq\frac{1}{p}\bigg{(}\tau^{-}(\alpha_{j})-\frac{(n-1)p-q}{2}-\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(\tau(\alpha_{j})-2-n)p}{2}\bigg{)}
=12(1+τ(αj)τ+(αj)+qp)+τ(αj)21n\displaystyle=\frac{1}{2}\bigg{(}1+\frac{\tau^{-}(\alpha_{j})-\tau^{+}(\alpha_{j})+q}{p}\bigg{)}+\frac{\tau(\alpha_{j})}{2}-1-n
=τ(αj)1n.\displaystyle=\tau(\alpha_{j})-1-n.

Note that the last equality uses the definition of τ(α)\tau(\alpha) in Definition 6.8. Similarly, we can compute that

iτ+(αj)+(n1)pq2i\leq\tau^{+}(\alpha_{j})+\frac{(n-1)p-q}{2}

implies that

Nj\displaystyle N_{j} 1p(τ+(αj)+(n1)pq2τ+(αj)+τ(αj)(τ(αj)2n)p2)\displaystyle\leq\frac{1}{p}\bigg{(}\tau^{+}(\alpha_{j})+\frac{(n-1)p-q}{2}-\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(\tau(\alpha_{j})-2-n)p}{2}\bigg{)}
=12(1+τ+(αj)τ(αj)qp)+τ(αj)21\displaystyle=\frac{1}{2}\bigg{(}-1+\frac{\tau^{+}(\alpha_{j})-\tau^{-}(\alpha_{j})-q}{p}\bigg{)}+\frac{\tau(\alpha_{j})}{2}-1
=1.\displaystyle=-1.

In summary, xImΦn+2kn+kx\in\operatorname{Im}\Phi^{n+k}_{n+2k} implies that for all 1jl1\leq j\leq l, either Njτ(αj)1nN_{j}\geq\tau(\alpha_{j})-1-n or Nj1N_{j}\leq-1. Hence from Lemma 7.8 part (3), we know that

ProjniGnI(αj)=0{\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha_{j})=0

for all 1jl1\leq j\leq l and as a result, Φnn+2k(x)=0\Phi^{n+2k}_{n}(x)=0.

Second, suppose xImΦn+2kn+kx\notin\operatorname{Im}\Phi^{n+k}_{n+2k}. For any 1jl1\leq j\leq l, we can write

i=τ+(αj)+τ(αj)(τ(αj)2n)p2+Njpi=\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(\tau(\alpha_{j})-2-n)p}{2}+N_{j}p

for some NjN_{j}. Then from Lemma 7.4 part (3) we know that there exists jj such that 1jl1\leq j\leq l, and

Nj[0,τ(αj)2n].N_{j}\in[0,\tau(\alpha_{j})-2-n]\cap\mathbb{Z}.

Hence by Lemma 7.8 part (3) and (4) we know that

ProjniGnI(αj)0Φnn+2k(x)0.{\rm Proj}^{i}_{n}\circ G_{n}\circ I(\alpha_{j})\neq 0\Rightarrow\Phi^{n+2k}_{n}(x)\neq 0.

Hence we conclude that

ImΦn+2kn+k=kerΦnn+2k.\operatorname{Im}\Phi^{n+k}_{n+2k}=\ker\Phi^{n+2k}_{n}.

7.4. The commutative diagram

In this subsection, we will prove the commutative diagram presented at the beginning of the section. Note that we choose the basis 𝔅\mathfrak{B} of 𝐘\mathbf{Y} as in Section 6.3.

Lemma 7.10.

Suppose nn\in\mathbb{Z} and ii is a grading. Suppose x(𝚪n,i)x\in(\mathbf{\Gamma}_{n},i) such that

Fn(x)=jlλjαj,F_{n}(x)=\sum_{j}^{l}\lambda_{j}\alpha_{j},

with λj0\lambda_{j}\neq 0 and αj𝔅\alpha_{j}\in\mathfrak{B} for all 1jl1\leq j\leq l. Then for any 1jl1\leq j\leq l, there exists Nj[0,n+1τ(αj)]N_{j}\in[0,n+1-\tau(\alpha_{j})] such that

i=τ+(αj)+τ(αj)(n+1τ(αj))p2+Njp.i=\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(n+1-\tau(\alpha_{j}))p}{2}+N_{j}p.
Proof.

This is a combination of Lemma 6.11 part (3), Lemma 6.12 part (3), and Lemma 7.4 part (1). The proof is similar to that of Lemma 7.4 part (2). ∎

Proof of Proposition 7.2.

We only prove the first commutative diagram

𝚪2n+2k+12\textstyle{\mathbf{\Gamma}_{\frac{2n+2k+1}{2}}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ+,n+2kn+k+1ψ,n+k+12n+2k+12\scriptstyle{\Psi^{n+k+1}_{+,n+2k}\circ\psi^{\frac{2n+2k+1}{2}}_{-,n+k+1}}ψ+,μn+k+1ψ,n+k+12n+2k+12\scriptstyle{\psi^{n+k+1}_{+,\mu}\circ\psi^{\frac{2n+2k+1}{2}}_{-,n+k+1}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,nμ\scriptstyle{\psi^{\mu}_{+,n}}𝚪n+2k\textstyle{\mathbf{\Gamma}_{n+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φnn+2k\scriptstyle{\Phi^{n+2k}_{n}}𝚪n\textstyle{\mathbf{\Gamma}_{n}}

The other is similar. Note that at the end of Section 6.3, we introduce new notations of η±,n2n\eta_{\pm,n-2}^{n} to remove the scalars. Then the second commutative diagram only holds up to a scalar.

First, note that the maps from 𝚪2n+2k+12\mathbf{\Gamma}_{\frac{2n+2k+1}{2}} to 𝚪μ\mathbf{\Gamma}_{\mu} and 𝚪n+2k\mathbf{\Gamma}_{n+2k} both factor through 𝚪n+k+1\mathbf{\Gamma}_{n+k+1}. As a result, we only need to prove the following commutative diagram for sufficiently large kk.

𝚪n+k+1\textstyle{\mathbf{\Gamma}_{n+k+1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ψ+,n+2kn+k+1\scriptstyle{\Psi^{n+k+1}_{+,n+2k}}ψ+,μn+k+1\scriptstyle{\psi^{n+k+1}_{+,\mu}}𝚪μ\textstyle{\mathbf{\Gamma}_{\mu}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ψ+,nμ\scriptstyle{\psi^{\mu}_{+,n}}𝚪n+2k\textstyle{\mathbf{\Gamma}_{n+2k}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Φnn+2k\scriptstyle{\Phi^{n+2k}_{n}}𝚪n\textstyle{\mathbf{\Gamma}_{n}}

Now suppose x(𝚪n+k+1,i)x\in(\mathbf{\Gamma}_{n+k+1},i). Write

Fn+k+1(x)=j=1lλiαjF_{n+k+1}(x)=\sum_{j=1}^{l}\lambda_{i}\cdot\alpha_{j}

with λj0\lambda_{j}\neq 0 and αj𝔅\alpha_{j}\in\mathfrak{B} for 1jl1\leq j\leq l. We want to first establish an identity

(7.7) ψ+,nμψ+,μn+k+1(x)=1jlnτ(αj)2Nj=n+k+1τ(αj)λjη+,nτ(αj)(αj)\psi^{\mu}_{+,n}\circ\psi^{n+k+1}_{+,\mu}(x)=\sum_{\begin{subarray}{c}1\leq j\leq l\\ n\leq\tau(\alpha_{j})-2\\ N_{j}=n+k+1-\tau(\alpha_{j})\end{subarray}}\lambda_{j}\cdot\eta^{\tau(\alpha_{j})}_{+,n}(\alpha_{j})

and then show that the other composition has exactly the same expression.

From Lemma 7.10, we know for any 1jl1\leq j\leq l, there exists Nj[0,n+k+1τ(αj)]N_{j}\in[0,n+k+1-\tau(\alpha_{j})] such that

(7.8) i=τ+(αj)+τ(αj)(n+k+1τ(αj))p2+Njp.i=\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(n+k+1-\tau(\alpha_{j}))p}{2}+N_{j}p.

Taking nj=τ(αj)n^{\prime}_{j}=\tau(\alpha_{j}) and Nj=0N_{j}^{\prime}=0, we can apply Lemma 6.12 part (3) to find an element

xj(𝚪τ(αj),τ+(αj)+τ(αj)2)x_{j}\in(\mathbf{\Gamma}_{\tau(\alpha_{j})},\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})}{2})

such that

Fτ(αj)(xj)=αj.F_{\tau(\alpha_{j})}(x_{j})=\alpha_{j}.

It is then straightforward to check that

(7.9) yj=Ψ+,n+k+1τ(αj)+NjΨ,τ(αj)+Njτ(αj)(xj)(𝚪n+k+1,i).y_{j}=\Psi^{\tau(\alpha_{j})+N_{j}}_{+,n+k+1}\circ\Psi^{\tau(\alpha_{j})}_{-,\tau(\alpha_{j})+N_{j}}(x_{j})\in(\mathbf{\Gamma}_{n+k+1},i).

Write

y=xj=1lλjyj(𝚪n+k+1,i).y=x-\sum_{j=1}^{l}\lambda_{j}\cdot y_{j}\in(\mathbf{\Gamma}_{n+k+1},i).

From Lemma 2.18 we know that

Fn+k+1(y)=0.F_{n+k+1}(y)=0.

As a result, by Lemma 6.6,

ψ+,nμψ+,μn+k+1(x)\displaystyle\psi^{\mu}_{+,n}\circ\psi^{n+k+1}_{+,\mu}(x) =j=1lλjψ+,nμψ+,μn+k+1(yj)\displaystyle=\sum_{j=1}^{l}\lambda_{j}\cdot\psi^{\mu}_{+,n}\circ\psi^{n+k+1}_{+,\mu}(y_{j})

Note that, unless Nj=n+k+1τ(αj)N_{j}=n+k+1-\tau(\alpha_{j}), we have

ψ+,μn+k+1Ψ+,n+k+1τ(αj)+Nj=0\psi^{n+k+1}_{+,\mu}\circ\Psi^{\tau(\alpha_{j})+N_{j}}_{+,n+k+1}=0

by the exactness. As a result,

ψ+,nμψ+,μn+k+1(x)\displaystyle\psi^{\mu}_{+,n}\circ\psi^{n+k+1}_{+,\mu}(x) =1jlNj=n+k+1τ(αj)λjψ+,nμψ+,μn+k+1Ψ,τ(αj)+Njτ(αj)(xj)\displaystyle=\sum_{\begin{subarray}{c}1\leq j\leq l\\ N_{j}=n+k+1-\tau(\alpha_{j})\end{subarray}}\lambda_{j}\cdot\psi^{\mu}_{+,n}\circ\psi^{n+k+1}_{+,\mu}\circ\Psi^{\tau(\alpha_{j})}_{-,\tau(\alpha_{j})+N_{j}}(x_{j})
(Lemma2.12)\displaystyle({\rm Lemma~{}\ref{lem: comm diag for n,n+1,mu}}) =1jlnτ(αj)2Nj=n+k+1τ(αj)λjψ+,nμψ+,μτ(αj)(xj)+1jlnτ(αj)1Nj=n+k+1τ(αj)λjψ+,nμψ+,μτ(αj)(xj)\displaystyle=\sum_{\begin{subarray}{c}1\leq j\leq l\\ n\leq\tau(\alpha_{j})-2\\ N_{j}=n+k+1-\tau(\alpha_{j})\end{subarray}}\lambda_{j}\cdot\psi^{\mu}_{+,n}\circ\psi^{\tau(\alpha_{j})}_{+,\mu}(x_{j})+\sum_{\begin{subarray}{c}1\leq j\leq l\\ n\geq\tau(\alpha_{j})-1\\ N_{j}=n+k+1-\tau(\alpha_{j})\end{subarray}}\lambda_{j}\cdot\psi^{\mu}_{+,n}\circ\psi^{\tau(\alpha_{j})}_{+,\mu}(x_{j})
Equation (7.10)\displaystyle\text{Equation }(\ref{eq: prop 7.2, 3}) =1jlnτ(αj)2Nj=n+k+1τ(αj)λjψ+,nμψ+,μτ(αj)(xj)\displaystyle=\sum_{\begin{subarray}{c}1\leq j\leq l\\ n\leq\tau(\alpha_{j})-2\\ N_{j}=n+k+1-\tau(\alpha_{j})\end{subarray}}\lambda_{j}\cdot\psi^{\mu}_{+,n}\circ\psi^{\tau(\alpha_{j})}_{+,\mu}(x_{j})
(Definitionofη+,nτ(αj))\displaystyle({\rm Definition~{}of~{}}\eta^{\tau(\alpha_{j})}_{+,n}) =1jlnτ(αj)2Nj=n+k+1τ(αj)λjη+,nτ(αj)(αj)\displaystyle=\sum_{\begin{subarray}{c}1\leq j\leq l\\ n\leq\tau(\alpha_{j})-2\\ N_{j}=n+k+1-\tau(\alpha_{j})\end{subarray}}\lambda_{j}\cdot\eta^{\tau(\alpha_{j})}_{+,n}(\alpha_{j})

This verifies Equation (7.7) if we show that

1jlnτ(αj)1Nj=n+k+1τ(αj)λjψ+,nμψ+,μτ(αj)(xj)=0.\sum_{\begin{subarray}{c}1\leq j\leq l\\ n\geq\tau(\alpha_{j})-1\\ N_{j}=n+k+1-\tau(\alpha_{j})\end{subarray}}\lambda_{j}\cdot\psi^{\mu}_{+,n}\circ\psi^{\tau(\alpha_{j})}_{+,\mu}(x_{j})=0.

To verify this last equality, assume that nτ(αj)1n\geq\tau(\alpha_{j})-1. Then from Lemma 2.12 and the exactness of the bypass maps we have

(7.10) ψ+,nμψ+,μτ(αj)(xj)=ψ+,nμψ+,μn+1Ψ,n+1τ(αj)(x)=0.\psi^{\mu}_{+,n}\circ\psi^{\tau(\alpha_{j})}_{+,\mu}(x_{j})=\psi^{\mu}_{+,n}\circ\psi^{n+1}_{+,\mu}\circ\Psi^{\tau(\alpha_{j})}_{-,n+1}(x)=0.

Now we deal with Φnn+2kΨ+,n+2kn+k+1(x)\Phi^{n+2k}_{n}\circ\Psi^{n+k+1}_{+,n+2k}(x). Since Fn+k+1(y)=0F_{n+k+1}(y)=0, Lemma 2.19 implies that

Ψ+,n+2kn+k+1(y)=0.\Psi^{n+k+1}_{+,n+2k}(y)=0.

Hence

Ψ+,n+2kn+k+1(x)=j=1lλjΨ+,n+2kn+k+1(yj)\Psi^{n+k+1}_{+,n+2k}(x)=\sum_{j=1}^{l}\lambda_{j}\cdot\Psi^{n+k+1}_{+,n+2k}(y_{j})

where yjy_{j} is defined as in (7.9). Note that, by definition we have yj(𝚪n+1+k,i)y_{j}\in(\mathbf{\Gamma}_{n+1+k},i), so from Lemma 2.6, we know

Ψ+,n+2kn+k+1(yj)(𝚪n+2k,i(k1)p2).\Psi^{n+k+1}_{+,n+2k}(y_{j})\in(\mathbf{\Gamma}_{n+2k},i-\frac{(k-1)p}{2}).

Note that, by (7.9) and Lemma 2.18, we know that

Fn+2kΨ+,n+2kn+k+1(yj)=Fτ(αj)(xj)=αj.F_{n+2k}\circ\Psi^{n+k+1}_{+,n+2k}(y_{j})=F_{\tau(\alpha_{j})}(x_{j})=\alpha_{j}.

Hence

Φnn+2kΨ+,n+2kn+k+1(x)=j=1lλjProjni(k1)p2GnI(αj).\Phi^{n+2k}_{n}\circ\Psi^{n+k+1}_{+,n+2k}(x)=\sum_{j=1}^{l}\lambda_{j}\cdot{\rm Proj}^{i-\frac{(k-1)p}{2}}_{n}\circ G_{n}\circ I(\alpha_{j}).

We write

i(k1)p2=τ+(αj)+τ(αj)(τ(αj)2n)p2+Njpi-\frac{(k-1)p}{2}=\frac{\tau^{+}(\alpha_{j})+\tau^{-}(\alpha_{j})-(\tau(\alpha_{j})-2-n)p}{2}+N_{j}^{\prime}p

Comparing the above formula with (7.8), we know

Nj=Nj+τ(αj)nk1.N_{j}^{\prime}=N_{j}+\tau(\alpha_{j})-n-k-1.

Note that, by construction, Njn+k+1τ(αj)N_{j}\leq n+k+1-\tau(\alpha_{j}), which means Nj0N_{j}^{\prime}\leq 0. Hence from Lemma 7.8 we know

Projni(k1)p2GnI(αj)0{\rm Proj}^{i-\frac{(k-1)p}{2}}_{n}\circ G_{n}\circ I(\alpha_{j})\neq 0

if and only if Nj=0N_{j}^{\prime}=0, i.e., Nj=n+k+1τ(α)N_{j}=n+k+1-\tau(\alpha). Also when Nj=0N_{j}^{\prime}=0 from Lemma 7.8 part (5) we know

Projni(k1)p2GnI(αj)=η+,nτ(αj)(αj).{\rm Proj}^{i-\frac{(k-1)p}{2}}_{n}\circ G_{n}\circ I(\alpha_{j})=\eta^{\tau(\alpha_{j})}_{+,n}(\alpha_{j}).

Note that we could focus on indices jj such that Projni(k1)p2GnI(αj)0.{\rm Proj}^{i-\frac{(k-1)p}{2}}_{n}\circ G_{n}\circ I(\alpha_{j})\neq 0. This is because if an index jj makes Projni(k1)p2GnI(αj)=0{\rm Proj}^{i-\frac{(k-1)p}{2}}_{n}\circ G_{n}\circ I(\alpha_{j})=0, then on one hand it does not contribute to Φnn+2kΨ+,n+2kn+k+1(x)\Phi^{n+2k}_{n}\circ\Psi^{n+k+1}_{+,n+2k}(x) since the corresponding summand is 0, on the other hand, we have Njn+k+1τ(α)N_{j}\neq n+k+1-\tau(\alpha) hence per Equation (7.7) it does not contribute to ψ+,nμψ+,μn+k+1(x)\psi^{\mu}_{+,n}\circ\psi^{n+k+1}_{+,\mu}(x), either. Also, we know from Lemma 7.8 part (3) that when Projni(k1)p2GnI(αj)0{\rm Proj}^{i-\frac{(k-1)p}{2}}_{n}\circ G_{n}\circ I(\alpha_{j})\neq 0 we must have nτ(αj)2n\leq\tau(\alpha_{j})-2. As a result, we know

Φnn+2kΨ+,n+2kn+k+1(x)\displaystyle\Phi^{n+2k}_{n}\circ\Psi^{n+k+1}_{+,n+2k}(x) =j=1lλjProjnj(k1)p2GnI(αj)\displaystyle=\sum_{j=1}^{l}\lambda_{j}\cdot{\rm Proj}^{j-\frac{(k-1)p}{2}}_{n}\circ G_{n}\circ I(\alpha_{j})
=1jlnτ(αj)2Nj=n+k+1τ(αj)λjη+,nτ(αj)(αj)\displaystyle=\sum_{\begin{subarray}{c}1\leq j\leq l\\ n\leq\tau(\alpha_{j})-2\\ N_{j}=n+k+1-\tau(\alpha_{j})\end{subarray}}\lambda_{j}\cdot\eta^{\tau(\alpha_{j})}_{+,n}(\alpha_{j})
Equation (7.7)\displaystyle\text{Equation }(\ref{eq: lem 7.9, 1}) =ψ+,nμψ+,μn+k+1(x)\displaystyle=\psi^{\mu}_{+,n}\circ\psi^{n+k+1}_{+,\mu}(x)

References

  • Boa [99] J. Michael Boardman. Conditionally convergent spectral sequences. In Homotopy invariant algebraic structures, volume 239 of Contemporary Mathematics, page 49–84. American Mathematical Society, Providence, RI, 1999.
  • BS [15] John A. Baldwin and Steven Sivek. Naturality in sutured monopole and instanton homology. J. Differ. Geom., 100(3):395–480, 2015.
  • [3] John A. Baldwin and Steven Sivek. A contact invariant in sutured monopole homology. Forum Math. Sigma, 4:e12, 82, 2016.
  • [4] John A. Baldwin and Steven Sivek. Instanton Floer homology and contact structures. Selecta Math. (N.S.), 22(2):939–978, 2016.
  • BS [21] John A. Baldwin and Steven Sivek. Framed instanton homology and concordance. J. Topol., 14(4):1113–1175, 2021.
  • [6] John A. Baldwin and Steven Sivek. Characterizing slopes for 525_{2}. ArXiv:2209.09805, v1, 2022.
  • [7] John A. Baldwin and Steven Sivek. Framed instanton Floer homology and concordance, II. ArXiv:2206.11531, v1, 2022.
  • [8] John A. Baldwin and Steven Sivek. Khovanov homology detects the trefoils. Duke Math. J., 174(4):885–956, 2022.
  • BS [23] John A. Baldwin and Steven Sivek. Instantons and L-space surgeries. J. Eur. Math. Soc. (JEMS), 25(10):4033–4122, 2023.
  • Eft [18] Eaman Eftekhary. Bordered Floer homology and existence of incompressible tori in homology spheres. Compositio Math., 154(6):1222–1268, 2018.
  • Flo [90] Andreas Floer. Instanton homology, surgery, and knots. In Geometry of low-dimensional manifolds, 1 (Durham, 1989), volume 150 of London Math. Soc. Lecture Note Ser., pages 97–114. Cambridge Univ. Press, Cambridge, 1990.
  • GLW [19] Sudipta Ghosh, Zhenkun Li, and C.-M. Michael Wong. Tau invariants in monopole and instanton theories. ArXiv:1910.01758, v3, 2019.
  • HL [21] Mathew Hedden and Adam Simon Levine. A surgery formula for knot Floer homology. arXiv:1901.02488, v2, 2021.
  • Hon [00] Ko Honda. On the classification of tight contact structures I. Geom. Topol., 4:309–368, 2000.
  • Juh [06] András Juhász. Holomorphic discs and sutured manifolds. Algebr. Geom. Topol., 6:1429–1457, 2006.
  • KM [10] Peter B. Kronheimer and Tomasz S. Mrowka. Knots, sutures, and excision. J. Differ. Geom., 84(2):301–364, 2010.
  • KM [11] Peter B. Kronheimer and Tomasz S. Mrowka. Knot homology groups from instantons. J. Topol., 4(4):835–918, 2011.
  • Li [20] Zhenkun Li. Contact structures, excisions and sutured monopole Floer homology. Algebr. Geom. Topol., 20(5):2553–2588, 2020.
  • [19] Zhenkun Li. Gluing maps and cobordism maps in sutured monopole and instanton Floer theories. Algebr. Geom. Topol., 21(6):3019–3071, 2021.
  • [20] Zhenkun Li. Knot homologies in monopole and instanton theories via sutures. J. Symplectic Geom., 19(6):1339–1420, 2021.
  • LPCS [22] Tye Lidman, Juanita Pinzón-Caicedo, and Christopher Scaduto. Framed instanton homology of surgeries on L-space knots. Indiana Univ. Math. J., 71(3):1317–1347, 2022.
  • LY [21] Zhenkun Li and Fan Ye. SU(2) representations and a large surgery formula. ArXiv:2107.11005, v1, 2021.
  • [23] Zhenkun Li and Fan Ye. Instanton Floer homology, sutures, and Heegaard diagrams. J. Topol., 15(1):39–107, 2022.
  • [24] Zhenkun Li and Fan Ye. Knot surgery formulae for instanton Floer homology II: applications. ArXiv:2209.11018, v1, 2022.
  • LY [23] Zhenkun Li and Fan Ye. An enhanced Euler characteristic of sutured instanton homology. Int. Math. Res. Not. IMRN, page rnad066, 04 2023.
  • OS [04] Peter S. Ozsváth and Zoltán Szabó. Holomorphic disks and knot invariants. Adv. Math., 186(1):58–116, 2004.
  • OS [08] Peter S. Ozsváth and Zoltán Szabó. Knot Floer homology and integer surgeries. Algebr. Geom. Topol., 8(1):101–153, 2008.
  • OS [11] Peter S. Ozsváth and Zoltán Szabó. Knot Floer homology and rational surgeries. Algebr. Geom. Topol., 11(1):1–68, 2011.
  • Sar [15] Sucharit Sarkar. Moving basepoints and the induced automorphisms of link Floer homology. Algebr. Geom. Topol., 15(5):2479–2515, 2015.
  • Sca [15] Christopher Scaduto. Instantons and odd Khovanov homology. J. Topol., 8(3):744–810, 2015.
  • Wei [94] Charles A. Weibel. An introduction to homological algebra, volume 38 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1994.
  • Zem [19] Ian Zemke. Graph cobordisms and Heegaard Floer homology. ArXiv: 1512.01184, v3, 2019.

References

  • Boa [99] J. Michael Boardman. Conditionally convergent spectral sequences. In Homotopy invariant algebraic structures, volume 239 of Contemporary Mathematics, page 49–84. American Mathematical Society, Providence, RI, 1999.
  • BS [15] John A. Baldwin and Steven Sivek. Naturality in sutured monopole and instanton homology. J. Differ. Geom., 100(3):395–480, 2015.
  • [3] John A. Baldwin and Steven Sivek. A contact invariant in sutured monopole homology. Forum Math. Sigma, 4:e12, 82, 2016.
  • [4] John A. Baldwin and Steven Sivek. Instanton Floer homology and contact structures. Selecta Math. (N.S.), 22(2):939–978, 2016.
  • BS [21] John A. Baldwin and Steven Sivek. Framed instanton homology and concordance. J. Topol., 14(4):1113–1175, 2021.
  • [6] John A. Baldwin and Steven Sivek. Characterizing slopes for 525_{2}. ArXiv:2209.09805, v1, 2022.
  • [7] John A. Baldwin and Steven Sivek. Framed instanton Floer homology and concordance, II. ArXiv:2206.11531, v1, 2022.
  • [8] John A. Baldwin and Steven Sivek. Khovanov homology detects the trefoils. Duke Math. J., 174(4):885–956, 2022.
  • BS [23] John A. Baldwin and Steven Sivek. Instantons and L-space surgeries. J. Eur. Math. Soc. (JEMS), 25(10):4033–4122, 2023.
  • Eft [18] Eaman Eftekhary. Bordered Floer homology and existence of incompressible tori in homology spheres. Compositio Math., 154(6):1222–1268, 2018.
  • Flo [90] Andreas Floer. Instanton homology, surgery, and knots. In Geometry of low-dimensional manifolds, 1 (Durham, 1989), volume 150 of London Math. Soc. Lecture Note Ser., pages 97–114. Cambridge Univ. Press, Cambridge, 1990.
  • GLW [19] Sudipta Ghosh, Zhenkun Li, and C.-M. Michael Wong. Tau invariants in monopole and instanton theories. ArXiv:1910.01758, v3, 2019.
  • HL [21] Mathew Hedden and Adam Simon Levine. A surgery formula for knot Floer homology. arXiv:1901.02488, v2, 2021.
  • Hon [00] Ko Honda. On the classification of tight contact structures I. Geom. Topol., 4:309–368, 2000.
  • Juh [06] András Juhász. Holomorphic discs and sutured manifolds. Algebr. Geom. Topol., 6:1429–1457, 2006.
  • KM [10] Peter B. Kronheimer and Tomasz S. Mrowka. Knots, sutures, and excision. J. Differ. Geom., 84(2):301–364, 2010.
  • KM [11] Peter B. Kronheimer and Tomasz S. Mrowka. Knot homology groups from instantons. J. Topol., 4(4):835–918, 2011.
  • Li [20] Zhenkun Li. Contact structures, excisions and sutured monopole Floer homology. Algebr. Geom. Topol., 20(5):2553–2588, 2020.
  • [19] Zhenkun Li. Gluing maps and cobordism maps in sutured monopole and instanton Floer theories. Algebr. Geom. Topol., 21(6):3019–3071, 2021.
  • [20] Zhenkun Li. Knot homologies in monopole and instanton theories via sutures. J. Symplectic Geom., 19(6):1339–1420, 2021.
  • LPCS [22] Tye Lidman, Juanita Pinzón-Caicedo, and Christopher Scaduto. Framed instanton homology of surgeries on L-space knots. Indiana Univ. Math. J., 71(3):1317–1347, 2022.
  • LY [21] Zhenkun Li and Fan Ye. SU(2) representations and a large surgery formula. ArXiv:2107.11005, v1, 2021.
  • [23] Zhenkun Li and Fan Ye. Instanton Floer homology, sutures, and Heegaard diagrams. J. Topol., 15(1):39–107, 2022.
  • [24] Zhenkun Li and Fan Ye. Knot surgery formulae for instanton Floer homology II: applications. ArXiv:2209.11018, v1, 2022.
  • LY [23] Zhenkun Li and Fan Ye. An enhanced Euler characteristic of sutured instanton homology. Int. Math. Res. Not. IMRN, page rnad066, 04 2023.
  • OS [04] Peter S. Ozsváth and Zoltán Szabó. Holomorphic disks and knot invariants. Adv. Math., 186(1):58–116, 2004.
  • OS [08] Peter S. Ozsváth and Zoltán Szabó. Knot Floer homology and integer surgeries. Algebr. Geom. Topol., 8(1):101–153, 2008.
  • OS [11] Peter S. Ozsváth and Zoltán Szabó. Knot Floer homology and rational surgeries. Algebr. Geom. Topol., 11(1):1–68, 2011.
  • Sar [15] Sucharit Sarkar. Moving basepoints and the induced automorphisms of link Floer homology. Algebr. Geom. Topol., 15(5):2479–2515, 2015.
  • Sca [15] Christopher Scaduto. Instantons and odd Khovanov homology. J. Topol., 8(3):744–810, 2015.
  • Wei [94] Charles A. Weibel. An introduction to homological algebra, volume 38 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1994.
  • Zem [19] Ian Zemke. Graph cobordisms and Heegaard Floer homology. ArXiv: 1512.01184, v3, 2019.