Knot surgery formulae for instanton Floer homology I: the main theorem
Abstract.
We prove an integral surgery formula for framed instanton homology for any knot in a -manifold with and . 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 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 , and copies of for any and large . 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 for a closed 3-manifold was introduced by Kronheimer and Mrowka in [17] and has been conjectured to be isomorphic to the hat version of Heegaard Floer homology . 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 -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 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 where 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 ) and leave the general setup to Section 3.3.
First, let us recall the results from [22]. Suppose is a null-homologous knot. Let be the knot complement, and let be the union of two oppositely oriented meridians of the knot on . Let be the corresponding sutured instanton homology introduced by Kronheimer-Mrowka [16], where the minus sign denotes orientation reversal for technical needs (note that and in particular ). A Seifert surface of induces a -grading on . In [22], we constructed a set of differentials on
for any gradings . We then constructed bent complexes
From [22], the homologies of these complexes are related to the Dehn surgeries of as follows:
(1.1) |
(1.2) |
To state the integral surgery formula, we introduce more notations. For , let be identical copies of . Define chain maps
as follows. For ,
Let denote the direct sum of all . 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 is a null-homologous knot. Let , , and be defined as above. Then for any , there exists an isomorphism
as the direct sum of isomorphisms
such that
Remark 1.2.
As an analog of the surgery formula in Heegaard Floer homology, the map is related to the vertical projection map , and the map is related to the map , which is defined by the composition of the horizontal projection and a chain homotopy equivalence between to . 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 . Hence we only state the existence of the isomorphism.
Remark 1.3.
The hypothesis of Theorem 1.1 excludes the case where . 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 . As a result, all related maps are only well-defined up to scalars. When , the maps and have distinct target spaces, namely and . As a result, the scalar ambiguity for individual maps does not influence the dimension of the homology of the mapping cone. However, when , 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 of the dual knot inside the resulting manifold in Theorem 3.22, which is analogous to the results by Eftekhary for knot Floer homology [10, Proposition 1.5] and by Hedden-Levine [13]. In this formula, we may assume 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 to obtain the following exact triangle.
Corollary 1.5 (Generalized surgery exact triangle).
Suppose is a null-homologous knot, and is a fixed non-zero integer. Then for any sufficiently large integer , there exists an exact triangle
(1.3) |
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 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 -manifold is a rational homology sphere. For any integer , there is a natural cobordism from to . From [9, Section 1.2], there exists a decomposition of the cobordism map along basic classes
where denotes the class that satisfies the equality
We make the following conjecture.
Conjecture 1.7.
Suppose is a null-homologous knot. Suppose and with . Let , , , , be defined as above. Then for any , there are commutative diagrams
Remark 1.8.
In Heegaard Floer theory, the large surgery formula in [26, Theorem 4.1] states that the homology of the bent complex is isomorphic to the Heegaard Floer homology of together with a spinc structure specified by . 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 -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 -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)
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)
We study the -surgery on a knot inside . We derive a formula computing the non-zero grading part of with respect to the grading induced by the Seifert surface.
-
(3)
We study non-zero integral surgeries on Boromean knots inside , which gives nontrivial circle bundles over surfaces. In this case the bent complexes and can be computed directly and the maps between them can be fixed with the help of the -action. Moreover, we show for most Seifert fibered manifolds with non-zero orbifold degrees.
-
(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)
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 , where one knot is a twist knot, i.e. the Whitehead double of the unknot.
-
(6)
We study almost L-space knots, i.e., a non-L-space knot such that there exists with (see [6] for the results in Heegaard Floer theory). We prove a genus one almost -space knot is either the figure-eight or the mirror of the knot . We also show that almost -space knots of genus at least 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 -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 coefficients. We write for and for the field with two elements.
A knot is called null-homologous if it represents the trivial homology class in , while it is called rationally null-homologous if it represents the trivial homology class in .
For any oriented 3-manifold , we write for the manifold obtained from by reversing the orientation. For any surface in and any suture , we write and for the same surface and suture in , without reversing their orientations. For a knot in a 3-manifold , we write for the induced knot in with induced orientation, called the mirror knot of . The corresponding balanced sutured manifold is .
2.2. Sutured instanton homology
For any balanced sutured manifold [15, Definition 2.2], Kronheimer-Mrowka [16, Section 7] constructed an isomorphism class of -vector spaces . Later, Baldwin-Sivek [2, Section 9] dealt with the naturality issue and constructed (untwisted and twisted vesions of) projectively transitive systems related to . We will use the twisted version, which we write as 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 -grading on 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 is a knot in a closed 3-manifold . Let and let be a simple closed curve on . Let be the knot complement and let be two oppositely oriented meridians of on . Define
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 and . These data can be fixed by choosing a basepoint in or . For simplicity, we omit those choices in the notations.
From now on, we will suppose is a rationally null-homologous knot and fix some notations. Let be the meridian of and pick a longitude (such that ) to fix a framing of . We will always assume is irreducible, but many results still hold due to the following connected sum formula of sutured instanton homology [18, Remark 1.6]:
Given coprime integers and , let be the suture on consists of two oppositely oriented, simple closed curves of slope , with respect to the chosen framing (i.e. the homology of the curves are ). Pick to be a minimal genus Seifert surface of , with genus . Note that may have multiple components.
Convention.
For a fixed pair as above, we write and . Note that when an orientation of the knot is chosen, the orientation of is induced by the knot. The orientation of is chosen such that and the orientation of is chosen such that . Note that is the order of , i.e., is the minimal positive integer satisfying . When is null-homologous, we always choose the Seifert framing . In such case, we have .
Remark 2.3.
The meanings of and above are different from our previous papers [23, 22]. Before, we used and to denote the meridian of the knot and the preferred framing. In particular, the framing is fixed by [23, Definition 4.2]. Note that in that case, we assume that is connected, and hence it is the same as the homological longitude (with the notation in previous papers, while we use to denote the homological meridian). Also, the numbers and in this paper should be and 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:
When , we write . When , we write . We also write and for the corresponding sutured instanton homologies.
Remark 2.4.
Strictly speaking, the sutures corresponding to and are not identical because the orientations are opposite. Since both sutures are on 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 for both sutures. The same discussion also applies to the relation between and .
We always assume that has minimal intersections with , i.e. . When the intersection number is odd, then induces a -grading on . When is even, we need to perform either a positive stabilization or negative stabilization on to induce a -grading, and the two gradings are related by an overall grading shift of . To get rid of stabilizations, we can equivalently regard that, in this case, the surface induces a -grading. We write the graded part of as
with or , 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 when
Proof.
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 , there are two graded bypass exact triangles
where the maps are homogeneous with respect to the homological -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 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 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 is large enough so that and [23, Proposition 4.14 (1)] applies, we have
where we omit since we think about if necessary. Then
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 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 , so the grading shifts of are .
∎
Remark 2.7.
The reason to use balanced sutured manifolds with reversed orientation is because of the above bypass exact triangles.
Remark 2.8.
Corollary 2.9.
For any sufficiently large integer , we have the following properties for restrictions of maps.
-
(1)
The map is an isomorphism when
-
(2)
The map is an isomorphism when
-
(3)
For any grading such that
there is an isomorphism
-
(4)
The map is an isomorphism when
-
(5)
The map is an isomorphism when
-
(6)
For any grading such that
there is an isomorphism
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 to denote one of the bypass maps. For any integer and any positive integer , define
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 , we have the following commutative diagrams up to scalars.
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 is rationally null-homologous. The second diagram is obtained from the first diagram by changing the choice of the framed knot. Explicitly, let be the dual knot corresponding to . Let denote its meridian. Then is a framing of . Applying the first diagram to , we will obtain the second diagram for the original . 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 , we have the following commutative diagrams up to scalars
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 and , there are two graded bypass exact triangles
Lemma 2.14.
For a knot and , there are commutative diagrams up to scalars
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 is a connected non-separating simple closed curve on . We can push into the interior of . For any fixed suture on and a closure of the sutured manifold, the push-off of is inside the closure, which is a closed -manifold. We can then take the framing on induced by the surface and there is an exact triangle associated to the instanton Floer homology of the - - and -surgeries along the push-off of . Since the push-off of is disjoint from , the exact triangle descends to one between corresponding sutured instanton Floer homologies.
According to [3, Section 4], when intersects the suture at two points, the -surgery along the push-off of (with framing induced by ) corresponds to a -handle attachment along . Note that attaching a -handle along will change the -manifold from to , where is the Dehn surgery along with slope specified by . We write
and in particular
Lemma 2.16 ([23, Lemma 3.21]).
For any , we have the following exact triangles.
Proof.
To obtain the first exact triangle, we can take the sutured manifold , and take a meridian . 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 -, -, and -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 -surgery will keep the manifold . The -surgery changes the framing and hence we obtain . The -surgery, as discussed above, gives rise to the manifold which is since is the meridian. Hence we obtain the expected triangle. The second exact triangle in the statement of the lemma is obtained similarly by taking to be a curve on having slope instead of a meridian. ∎
Remark 2.17.
From [3, Section 4], we know the -surgery corresponds to a -handle attachment and a -handle attachment. Hence in the above lemma is indeed , where is the unknot in bounding an embedded disk. By [3, Section 4], a -handle attachment does not change the closure of the balanced sutured manifold, and then there is a canonical identification between and . Hence we can abuse the notations. The same discussion also applies to .
Furthermore, we proved the following properties in [23]. Note that the assumption that 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 , we have the following commutative diagrams up to scalars
Lemma 2.19 ([23, Lemma 4.17, Proposition 4.26, Lemma 4.29 and Proposition 4.32]).
Let and be defined as in Lemma 2.16. Then for any sufficiently large integer , we have the following properties.
-
(1)
The map is zero and is surjective. Moreover, for any grading
the restriction of the map
is an isomorphism.
-
(2)
The map is zero and is injective. Moreover, for any grading
the map
is an isomorphism, where
is the projection.
The following lemma is a special case of Proposition 4.1, which we will prove later.
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 is a framed knot with the meridian and the framing . Fix a knot complement and the suture . We fix a special choice of a (marked odd) closure of following the construction in [16, Formula (18)].
Let be a closed, oriented, connected surface of genus at least . Suppose is a non-separating curve in . Let and let be the meridian and the longitude of (the latter comes from the surface framing). Let
(2.1) |
where is a point different from . We can pick and be a curve intersecting once. Since and , the pair defines an instanton Floer homology in the setting of [16, Section 7.1]. Moreover, forms a marked odd closure in [2, Definition 9.2], which was used in the naturality result [2, Theorem 9.17]. The reason for is to apply the naturality result (c.f. [2, Remark 9.4]).
Similarly, for and , we fix closures and as in (2.1), except replacing the gluing map by and , respectively.
For the sutured manifold , we regard it as by Remark 2.17, where is the unknot and is meridian suture on the unknot complement. Then we apply the above construction to obtain a special closure of . 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 since we only care about the dimension of its framed instanton homology.
After fixing the choices of closures, we can view and 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 . When , the first exact triangle in Lemma 2.16 is trivial for any and hence the maps and 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 for any rational homology sphere and, up to author’s knowledge, there is no known closed oriented -manifold with .
To help us fixing the scalars, suppose the maps and are defined as in the proof of Lemma 2.16, and we define
(2.2) |
We have the following basic properties for these indices.
Lemma 2.21.
Assuming . Suppose and are defined as in Equation (2.2). Then we have
Moreover, we have if and only if and if and only if .
Proof.
The fact that and 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 whenever and thus we know if and only if . The argument for is similar. Finally, by definition we know and hence from the exact triangle we know that
Hence we conclude that . ∎
By Lemma 2.19, we can pick a sufficiently large integer such that . Pick arbitrary representatives of the maps
and we also pick arbitrary representatives of the maps
for all .
Now we explain how to fix the scalars for maps with . Note that we have already chosen a representative for and . From Lemma 2.18, we have
where means commutative up to a non-zero scalar. We can choose a representative of to obtain an equality
We then choose a representative of so that
Next, we pick representatives of the maps inductively, with the base case constructed above, so that
hold for all . 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 more carefully later.
Similarly, we pick the maps inductively to satisfy the commutative diagram
We can choose representatives of , , and in a similar manner.
Furthermore, the representatives of
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 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 and in Lemma 2.16. We choose representatives of the maps inductively so that
is satisfied for all . We pick arbitrary representatives of the map and then pick inductively so that
is satisfied for all . We can then use induction to prove the following two equalities.
-
•
for all , and
-
•
for a non-zero scalar that is independent of with .
We verify the equality for first. The base case is by construction. Assuming we have already established the equality for , from Lemma 2.11 and Lemma 2.18, we have
The argument for is similar, once we take to be the complex number such that
(2.3) |
The remaining issues are summarized as follows:
-
(i)
When choosing representatives of , we use the commutative diagram
However, when , there is no unique choice of and one might worry that different choices of may affect the commutative diagrams in Lemma 2.18.
- (ii)
-
(iii)
We want to get rid of the scalar in Equation 2.3.
We treat these issues in several different cases.
Case 1. . In this case, we know that
for any integer . Indeed, if , we know from Lemma 2.18 that
If , instead of the above equation involving , we have
This implies that inductively we can fix a unique representative of for all .
For any integer with , we have . Then we take an element so that . Then we can solve the scalar as follows.
Hence we conclude that . In particular, we can take . Note so we can take so that . Then we have
This implies that as well. Now for any , we can take such that . we have
Hence we conclude that . In summary, in Case 1, we have the following.
-
•
We can fix a unique representative of for any .
-
•
We have for any .
-
•
We have .
Case 2. . Note that some arguments in Case 1 still apply. We summarize as follows.
-
•
For , we have so there is a unique choice of .
-
•
For , we have so again there is a unique choice of .
-
•
For , we have so in the expression of .
-
•
For , we have so .
To resolve the issue (i), the only nonfixed index is . In case
there is a unique choice of so that
Otherwise, we just fix any representative of .
To resolve issues (ii) and (iii), we rescale the maps according to the grading on . To do this, for an integer and a grading , define
It is straightforward that
Hence we define
(2.4) |
where the map
is the projection. Equivalently, if we have an element , we take
We then need to verify the following two equalities for :
-
•
, and
-
•
.
For the first one, note that increases the grading by from Lemma 2.6. So assume , we have
On the other hand, the map decreases the grading by from Lemma 2.6, and so for ,
Hence we can use instead of to get rid of the scalar . However, changing to , we might break the exact triangle in Lemma 2.16. Hence we need to alter the definition of as well. We define
(2.5) |
and it remains to establish the following exact triangle:
When , the triangle automatically holds. Indeed, for such , we have so ; the fact implies that , so the new exact triangle is exactly the original one in Lemma 2.16.
When , we still have and hence . So the exactness at is from Lemma 2.16. By construction we have for any so we conclude the exactness at . This also implies that
Hence to show the exactness at , it remains to show that
For any , we have
Hence we are done.
Finally, we verify the exact triangle for . Define a homomorphism
as
It is clear that is an isomorphism as its inverse is
and from the construction of in Equation (2.4), we have
From the fact that (we have ), , and that shift the grading by , we conclude that
As a result, we conclude the exactness at . The exactness at holds as above, since we still have
Dimension counting similar to the above argument then implies that
We then verify that
For any , we have
In summary, in Case 2, we did the following (extra things):
-
•
We choose representatives of for all so that
-
•
We define new maps for all so that
(2.6) -
•
We define new maps for all so that we have the following exact triangle for all .
(2.7)
Case 3. or . The situation and argument are similar to those in Case 2. We summarize the differences here:
-
•
In Case 3, the composition could only be zero when . In case the composition is indeed , we choose an arbitrary representative of the map .
- •
Note from Lemma 2.21, we must have , so the above three cases cover all situations.
Finally, we could extend the choice of representatives for all relevant maps for the indices . Note that when , we have that is injective and . Hence we do not need to worry about the issues (i), (ii), and (iii).
Convention.
From now on, we write the maps and simply as and , respectively. From the above discussion, when 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 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 -graded, where we regard the -grading as a homological grading. Many results in this subsection come from properties of the derived category of vector spaces over , for which people usually consider cochain complexes. However, for a -graded space there is no difference between the chain complex and the cochain complex. Hence by saying a complex we mean a -graded (co)chain complex, though all results apply to -graded cochain complexes verbatim.
For a complex and an integer , we write for its grading part (under the natural map ). With this notation, we suppose the differential on sends to . For any integer , we write for the complex obtained from by the grading shift . We write or for the homology of a complex with differential . A -graded vector space is regarded as a complex with the trivial differential.
For a chain map , we write for the mapping cone of , i.e., the complex consisting of the space and the differential
Then there is a long exact sequence
where sends to and sends to . If the differentials of and are trivial, then we know
(2.8) |
Remark 2.22.
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 are -graded vector spaces satisfying the following long exact sequences
Then we have the fourth long exact sequence
such that the following diagram commutes
(2.9) |
where we omit the grading shifts and the notation for the maps . We can also write (2.9) in another form so that there is enough room to write the maps
(2.10) |
The maps and in (2.10) can be written explicitly as follows. By the long exact sequences in the assumption of Lemma 2.23, we know that are chain homotopic to the mapping cones , respectively. Under such homotopies, we can write
and
However, the chain homotopies are not canonical, and hence the maps and are also not canonical. Thus, usually we cannot identify them with other given maps. Fortunately, with an extra -grading, we may identify with for another map .
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 -grading in the following lemma is not the homological grading used before.
Lemma 2.24.
Suppose and are -graded vector spaces and suppose are homogeneous maps with different grading shifts and . Then is isomorphic to for any .
Proof.
For simplicity, we can suppose and . The proof for the general case is similar. For , we write and for grading summands of and , respectively. Suppose is an automorphism of that acts by
Then is an isomorphism between and . ∎
Then we state the lemma that relates the map in Lemma 2.23 to another map constructed explicitly.
Lemma 2.25.
Suppose are -graded vector spaces satisfying the following horizontal exact sequences.
where the shift is for the -grading. Suppose satisfies the two commutative diagrams and suppose satisfies the left commutative diagram. Suppose and are homogeneous with respect to the -grading. Suppose and are sums of homogeneous maps with different grading shifts with respect to the -grading. Moreover, suppose the following diagrams hold up to scalars.
Then there is an isomorphism between and .
Proof.
Since and share the same domain and codomain, it suffices to show that they have the same rank. Fix inner products on and such that we have orthogonal decompositions
By commutativity, we know both and send onto . 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
where has full row rank. Then it suffices to show and have the same row rank.
By the exactness at and , we know the restriction of on is an isomorphism between and and the restriction of on is an isomorphism between and . By commutativity, we know that both and send to and
for some . Since and are homogeneous, there exist induced -gradings on and . The maps and 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 and on 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 is a framed rationally null-homologous knot, and we adopt the notations introduced in Section 2.2. Define
and write as the restriction of on . From Lemma 2.13 and Lemma 2.6, we can verify that shifts grading by , and then the integral surgery formula can be stated as follow.
Theorem 3.1.
Suppose is a fixed integer such that . Then for any sufficiently large integer , there exists an exact triangle
Hence we have an isomorphism
Remark 3.2.
Let and represent the meridian and the longitude of the knot , respectively. Then, is equivalent to the fact that is not isotopic to a connected component of the boundary of the Seifert surface. Specifically, if is null-homologous, we must have .
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) |
To obtain the dotted exact triangle, we need to establish the following three exact triangles:
(3.2) |
(3.3) |
(3.4) |
and establish the following commutative diagram:
(3.5) |
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 . We also require the following two extra diagrams to commute, where the maps other than come from (3.1).
(3.6) |
(3.7) |
Indeed, by applying Lemma 2.25, it suffices to prove some weaker commutative diagrams involving 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) |
where
The reader can compare (3.8) with (2.9) and (2.10). We omit the term corresponding to 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) |
follows from the second exact triangle in Lemma 2.16. Though the map may not be the same as the sum , we can use the following proposition and Lemma 2.24 (another special case of Proposition 4.1) to identify with . Here we use the assumption that .
Proposition 3.3.
Suppose is the map in Lemma 2.16. For any integer , there exist scalars such that
The exactness at
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 and . Then, for any satisfying the equation
the following sequence is exact
Remark 3.5.
The exactness at the direct summand for the second exact sequence (the one involving ) might not be as clear from Proposition 3.4. Explicitly, we apply the proposition to the dual knot corresponding to with framing and .
The exactness at and in the second exact sequence of (3.8)
(3.10) |
will also be proved by diagram chasing. We can explicitly construct the map by the composition of bypass maps
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 is constructed as above. For any , the following sequence is exact
Remark 3.7.
In the proof of [22, Theorem 3.23], we obtained a long exact sequence
by the octahedral lemma. However, we did not know the two maps involving 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 is because
and
where the right hand sides of the equations are independent of scalars.
The exactness at and in the third exact sequence of (3.8)
(3.11) |
is harder to prove since the map cannot be constructed by bypass maps. We expect that there are many equivalent constructions of 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 is large. See Section 7.2 for more details.
Proposition 3.9.
Suppose and suppose is a large integer. For any , there exists a map such that the following sequence is exact
Remark 3.10.
In the first arXiv version of this paper, we only proved Proposition 3.9 for knots in because we had to use the fact that and has an orientation-reversing involution. The construction of for knots in general -manifolds is inspired by the original proof for and the proof in Section 7 is a generalization of the previous proof.
Remark 3.11.
Then we consider the commutative diagrams mentioned in Section 3.1. By Lemma 2.6 and Lemma 2.12, we have
which verifies the commutative diagram in the assumption of the octahedral axiom.
Define
By Lemma 2.13 and Lemma 2.14 with , we have
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.
3.3. Reformulation by bent complexes
In this subsection, we restate Theorem 3.1 using the language of bent complexes introduced in [22]. Suppose is a rationally null-homologous knot in a closed 3-manifold . 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 together, we obtain the following diagram:
(3.12) |
where the -grading shift of is for any . From (3.12), we constructed in [22, Section 3.4] two spectral sequences and from to , where is roughly
(3.13) |
The composition with the inverse map is well-defined on the -th page, and the independence of (and hence in (3.13)) follows from Lemma 2.12. The -grading shift of is . By fixing an inner product on , we then lifted those spectral sequences to two differentials and on such that
In such way, the inverses of are also well-defined, which we will use freely later.
Then we propose an integral surgery formula for using differentials and on . To state the formula, we introduce the following notations.
Definition 3.13 ([22, Construction 3.27 and Definition 5.12]).
For any integer , define the complexes
Furthermore, define
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 are finite. Then, for any sufficiently large integer satisfying
we have
In such case, and are identities.
By splitting the diagram (3.12) into -gradings, we can calculate homologies of the complexes defined in Definition 3.13.
Proposition 3.15.
Suppose and is a grading. Fix an inner product on . If , then there exists a canonical isomorphism
If , then there exists a canonical isomorphism
Proof.
The proof mirrors that of [22, Lemma 5.13]. Following the notation in [22, (3.9) and (3.10)], if
then (the corresponding grading summand of ) 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 , 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 . The other statement holds for the same reason. ∎
Definition 3.16 ([22, Construction 3.27 and Definition 5.12]).
For any integer , define the bent complex
where for any element ,
Define
by
where . Define
by putting together for all . We also use the same notation for the induced map on homology.
Remark 3.17.
Now, we state the integral surgery formula in the above setup.
Theorem 3.18.
Suppose is a fixed integer such that . Then there exists an isomorphism
as the direct sum of isomorphisms
so that
Proof.
According to Remark 3.17, we only need to consider the maps for less than a fixed integer. For such values of , we can apply the following proposition.
Proposition 3.19 ([22, Proposition 3.28]).
Fix such that . For any large integer , fix inner products on and . Then there exist such that the following diagram commutes
where the maps , defined in Section 3.1, factor through .
Remark 3.20.
From the calculation in [22, Remark 3.29] (we replace and there by and , and note that there is a typo about sign in the first arXiv version of [22]), the difference of the grading shifts is
Note that the notations in this paper and [22] are different (c.f. Remark 2.3).
Then we can construct the isomorphism
for by identifying both and with for a sufficiently large . A priori, this isomorphism depends on inner products on
For other , we can take any isomorphism since the choice does not affect the computation of the mapping cone.
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 action on sutured instanton homology, where is the ambient manifold of the knot. This action vanishes on so vanishes on the bent complex. But it is nonvanishing on and and we use this information to realize the computation. This issue for the bent complex might be resolved by introducing some -pages for differentials and such that the action is nontrivial on -pages.
3.4. A formula for instanton knot homology
The third exact sequence (3.11) implies
for any sufficiently large integer . Since there are two copies , we can always regard the grading shifts of the maps as different ones by rescaling the grading of the first summand from to and the second summand from to . Hence we do not need the assumption 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 [10, Proposition 1.5] (see also Hedden-Levine’s work [13]). Since can be any integer, we replace by .
Theorem 3.22.
Suppose . Define
Then there exists an isomorphism
such that
Proof.
As mentioned before, we have
for any sufficiently large integer .
Since bypass maps are homogeneous, the above mapping cone splits into -gradings (or -gradings). Hence we can use it to calculate . By Lemma 2.6, the corresponding spaces are
From Proposition 3.15 with , by fixing an inner product on , we know that
and
Since is fixed, when is sufficiently large, we know that any with nontrivial (i.e. by Lemma 2.5) satisfies the above inequalities. By Proposition 3.15 again (fixing an inner product on ) and Remark 3.14, for sufficiently large, we know that
for such . By unpackaging the construction of differentials and in [22, Section 3.4], we know that the restrictions of maps and on the corresponding gradings coincide with the maps induced by the inclusions and under the canonical isomorphisms, respectively.
For , let
be the isomorphism obtained from identifying both spaces to the corresponding grading summand of . Note that it depends on inner products on and . For other , we can take any isomorphism since the choice does not affect the computation of the mapping cone. Then we know that
∎
4. Dehn surgery and bypass maps
Suppose is a balanced sutured manifold and is a connected simple closed curve that intersects the suture twice. There are two natural bypass arcs associated to , each of which intersects the suture at three points and induces a bypass triangle (c.f. [8, Section 4])
where and 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 be obtained from by attaching a contact -handle along . From [4, Section 3.3], it has been shown that a closure of coincides with a closure of the sutured manifold obtained from by -surgery along with respect to the surface framing. Hence there is also a surgery exact triangle (c.f. [23, Lemma 3.21])
The map is related to the bypass maps as follows:
Proposition 4.1.
There exist , such that
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 be a tubular neighborhood of . Push the interior of into the interior of to make it a properly embedded surface. By a standard argument in [14], we can assume that a collar of is equipped with a product contact structure such that is (isotopic to) the dividing set, is a Legendrian curve, is in the contact collar, and is a convex surface with Legendrian boundary that separates a standard contact neighborhood of off . The convex decomposition of along yields two pieces
where is diffeomorphic to and is the contact neighborhood of . It is straightforward to check that, after rounding the corners, the contact structure near the boundary of is still a product contact structure with being a convex boundary. Let be the dividing set on . Also, after rounding the corners, with the contact structure on , we can suppose is a convex surface with dividing set being the union of two connected simple closed curves on of slope . When viewing as the complement of an unknot in , the dividing set coincides with the suture , so from now on we call it . By the construction of the gluing map in [19], there exists a map
As in [19], the map comes from attaching contact handles to to recover the gluing along . From [20, Proposition 1.4], we know that
Note that and are equipped with the product contact structure near their boundaries. From the functoriality of the contact gluing map in [19], we know that is an isomorphism. Now both the -surgery along a push off of and the bypass attachments can be thought of as happening in the piece . Note that the result of both -surgery and the bypass attachments for is . Hence we have the following commutative diagram.
(4.1) |
where denotes the surgery map for the manifold and is the gluing map obtained by attaching the same set of contact handles as . A similar commutative diagram holds when replacing and by and
in (4.1), respectively.
Since is an isomorphism, to obtain a relation between and , it suffices to understand the relation between and . From [20, Proposition 1.4], we know that
Moreover, the meridian disk of induces a grading on and we have
with
Let
be a generator. In [20, Section 4.3] it is shown that
are non-zero. Also, when viewing as the complement of the unknot , there is an exact triangle
(4.2) |
as in Lemma 2.16. Comparing the dimensions of the spaces in (4.2), we have and is injective. From the fact that , we know from [12, Corollary 3.5] that
By the exactness in (4.2), we have and then is not in , i.e., it is a linear combination of generators of . Hence we know that there are such that
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 is the complement of a knot and consists of two meridians of the knot, which we denote by , we can choose to be a curve on of slope . Then we have a surgery triangle:
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 -manifold that represents the second Stiefel-Whitney class of the relevant -bundle; see [5, Section 2.2] for more details. Floer’s original exact triangle, on the other hand, yields a different triangle
where denotes a meridian of the knot. Note the difference between and is that they come from the same cobordism but the -bundles over the cobordism are different. The local argument to prove Proposition 4.1 works for both and . Hence there exists non-zero complex numbers such that
where the maps
are the two related bypass maps. When , 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 ,
However, when , the two bypass maps both preserves gradings, making the coefficients significant, i.e., and 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 .
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
for any and .
Proposition 5.1.
Given and , for any satisfying the equation
the following sequence is exact
Proof.
For simplicity, we only prove the proposition for . The proof for any general is similar (replacing all below by and modifying the notations for bypass maps). Also, we only prove the case when
The proof for general scalars can be obtained similarly.
We prove the proposition by induction on . 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 . The proposition reduces to
The commutative diagram in Lemma 2.11 implies
We then prove
Suppose
Then we have
By exactness, there exists such that . Then
By exactness, there exists such that
Let Then
and
which concludes the proof for .
Suppose the proposition holds for . We prove it also holds for . The proof is similar to the case for . Again by Lemma 2.11, we have
Then we prove
Suppose
Then we have
By exactness, there exists such that . By a similar reason, there exists such that . The goal is to prove
for some modifications and of and as for in the case of . Then the induction hypothesis will imply that there exists such that
Hence we will have
This will conclude the proof for .
Now we start to construct . We have
By exactness, there exists such that
Let Then
and
Then we start to construct . We have
By exactness, there exists such that
Let Then
and
Then we have the following commutative diagrams
By the induction hypothesis, there exists such that
which concludes the proof for . ∎
Remark 5.2.
By similar arguments, we can prove that the following sequence is exact for any
where the scalars satisfies the equality .
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 with framing and setting will recover Proposition 3.6.
Proposition 5.3.
Suppose
Then for any , the following sequence is exact
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 first. We have
Hence
Then we prove
Suppose
By exactness, there exists such that . Then we have
By exactness, there exists such that . Thus, we have , which concludes the proof for the exactness at .
Then we prove the exactness at . Similarly by exactness, we have
Suppose . If , then by the exactness, we know . If , then by the exactness, there exists such that
Then we know
Thus, we have
which concludes the proof for the exactness at . ∎
6. Some technical constructions
6.1. Filtrations
In this subsection, we study some filtrations on and that will be important in later sections. We continue to adopt conventions from Section 2.3. In particular, is a rationally null-homologous knot and is a rational Seifert surface of .
Lemma 6.1.
The maps in Lemma 2.16 lead to a filtration on : for a sufficiently large integer ,
Proof.
Lemma 6.2.
For any , the map induces an isomorphism
Proof.
Lemma 6.3.
For any , the maps induce isomorphisms
Proof.
We only prove the lemma for positive bypasses. The proof for the negative bypasses is similar. Let . By Lemma 2.6 and Lemma 2.12, we have
Hence we know
Since , the map is injective on . To show it is surjective as well, pick . Note implies that there exists such that . Lemma 2.12 then implies that
As a result, . ∎
Corollary 6.4.
-
(1)
For any , there is a canonical isomorphism
-
(2)
For sufficiently large , there exists a (noncanonical) isomorphism
Definition 6.5.
For any integer and any grading , define the map as the restriction
where is the map from Lemma 2.16.
Lemma 6.6.
Suppose is small enough such that (c.f. Lemma 2.19). Then for any integer and any grading , we have
where
is the projection.
Proof.
We only prove the lemma for positive bypasses and the proof for negative bypasses is similar. First, suppose
Pick such that
Taking , we know from Lemma 2.6 that from Lemma 2.18 that and from Lemma 2.12 that As a result, we conclude .
Second, suppose is non-zero. Pick such that
By Lemma 2.5 and Lemma 2.6, the fact that implies that
(6.1) |
Pick a sufficiently large integer and then take
By Lemma 2.18 we have
Note that the grading of equals to
(6.2) |
Combining 6.1 and 6.2, we obtain
Note that we pick to be a sufficiently large integer. In particular, we can assume
and
Thus is in this range as well and then Lemma 2.19 implies that is injective on the grading . Hence . Then the following Lemma 6.7 applies to and there exists such that
Thus by Lemma 2.12,
∎
Lemma 6.7.
Suppose and . Suppose such that
Then there exists such that
6.2. Tau invariants in a general -manifold
Definition 6.8.
We will prove the independence of these invariants about later in Lemma 6.12.
Remark 6.9.
Here we fix the knot and define the tau invariants for a homogeneous element . 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.
Lemma 6.11.
We have the following properties.
-
(1)
Suppose are two integers and are two gradings such that there exist and with
Then there exists an integer such that
i.e. when we send and into the same with by bypass maps, then the difference of the expected gradings of the images is divisible by (the grading shifts of the bypass maps are ).
-
(2)
Suppose we have an integer , a grading , and an element . Then for any integer and grading such that there exists an integer with
there exists an element such that
-
(3)
Suppose and for we have a grading and an element such that ,…, are linearly independent. Then the element
is homogeneous if and only if for any , we have
Proof.
(1). Take a sufficiently large integer. For , take to be the unique grading such that there exists an integer with
Take
From Lemma 2.18 we know that
By Lemma 2.19, we know that and in particular, . As a result, we can take then it is straightforward to verify that
(3). The proof is similar to that of (1). ∎
Lemma 6.12.
For a homogeneous element , we have the following.
-
(1)
and hence are well-defined. (i.e. they are independent of the choice of the large integer .)
-
(2)
We have .
-
(3)
For any integer and grading , the following two statements are equivalent.
-
(a)
There exists such that .
-
(b)
We have and there exists such that and
-
(a)
-
(4)
We have
Proof.
(1). Suppose is a homogeneous element. Then by definition there exists for some integer and grading such that
Then for sufficiently large , we can take
and from Lemma 2.18 implies that
and hence exists.
To show the value of is independent of as long as it is sufficiently large, a combination of Lemma 2.5 and Lemma 2.6 implies that the map
is an isomorphism for any . Then Lemma 2.18 implies that is well-defined. The argument for is similar.
(2). It follows directly from Lemma 6.11 part (1).
(3). We first establish the following claim.
Claim. There exists an element
such that
Proof of the claim.
Suppose is sufficiently large and
such that . Note that the existence of follows from the definition of . Let
It follows from Lemma 2.6 that
From Lemma 2.18 we know that
By Lemma 2.19 this implies that
Hence Lemma 6.7 applies and there exists such that
Again Lemma 2.6 implies that is in the grading
and Lemma 2.18 implies
∎
Now if an integer and a grading 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 such that . From the above claim, we already know that there exists
such that
Hence Lemma 6.11 part (1) implies that there exists such that
If , we can take a sufficiently large and
It follows from Lemma 2.6 that
Then Lemma 2.18 implies that
which contradicts the definition of in Definition 6.8. Similarly if we can take
which would be an element contradicting the definition of . When we have so there is always a contradiction by the above argument. This concludes (b).
(4). It follows from the definition of and Lemma 2.19 that is an isomorphism when restricted to the direct sum of consecutive middle gradings of when is large. ∎
Lemma 6.13.
For any we have that
6.3. A basis for framed instanton homology
We pick a basis for as follows. First
To construct the set , first, let if . By Lemma 2.19 this means for all small enough . Write
We pick the set inductively. Note that we have taken for with . Suppose we have already constructed the set that consists of homogeneous elements and is a basis of , we pick the set such that consists of homogeneous elements with , and the set
forms a basis of . Note that Lemma 6.13 implies that exists and
For any such that , define maps
as follows: for any , since is homogeneous and , we can pick
by Lemma 6.12 part (3) such that . Then define
.
Lemma 6.14.
Suppose such that .
-
(1)
The maps are all well-defined.
-
(2)
We have for some scalar .
-
(3)
Elements in are linearly independent.
-
(4)
forms a basis for .
-
(5)
For any we have
-
(6)
We have
Proof.
(1). We only work with and the arguments for are similar. Suppose there are such that , where . Then
and by Lemma 6.6 we have
Here is a small enough integer. As a result,
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 would possibly arise.
(3). We only work with and the arguments for are similar. Suppose
Suppose there exists such that
Pick such that . Then we have
As a result, there exists such that
Note that, from Lemma 2.12, we know
so as a result there exists such that
Hence by Lemma 2.18 we have
Since form a basis of , the sum cannot be in except for all .
(4). For , pick such that . Then by definition
Now we can compute
and by Lemma 2.12
Hence
Then (4) follows from (3), Lemma 6.2, and .
(5). It follows directly from the construction of and Lemma 2.6.
Convention.
We can define
such that
and the new maps satisfy all properties in Lemma 6.14 except (2). We will use to denote in latter sections.
7. The map in the third exact triangle
In this section, we construct the map 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 for maps.
Proposition 7.1.
Suppose is fixed and is sufficiently large. Then there is an exact triangle
where two of the maps are already constructed
Proposition 7.2.
Suppose is fixed and is sufficiently large. Suppose is constructed in Proposition 7.1. Then, there are two commutative diagrams up to scalars.
7.1. Characterizations of the kernel and the image
Before constructing , we characterize the spaces and . These results will motivate the construction of to ensure that
Since and are constructed using bypass maps, it suffices to consider their restrictions on each grading.
Lemma 7.3.
Suppose is fixed and is sufficiently large. Let
be the projection. Then we have
Proof.
We need to apply Lemma 2.20. Following conventions in Section 2.3, we have
(7.1) |
Suppose . Pick and such that
where . When is sufficiently large, we know from Lemma 2.19 that
In particular, from Lemma 2.18
Since the maps are homogeneous, we know that
which implies that .
Next, suppose . We take and we will pick for all such that
We will use the notation to denote an element in . Recall that from Lemma 2.6, the grading shifts of are . Take
Since we know that
(7.2) |
Hence from Lemma 6.7, there exists
such that
Then we can take
We can apply the same argument and use Lemma 6.7 to find
such that send them to corresponding elements in . Repeating this argument, we can obtain elements
such that , , , and for any we have
Note that we obtain the above for essentially from the fact that as in Equation (7.2). However, so we have as well. A similar argument as above then yields
Together with , we obtain for all .
It is then straightforward to check that
∎
Lemma 7.4.
Suppose is a homogeneous element and
where and for . Let be an integer, be a grading and be a sufficiently large integer. For an element such that , the following is true.
-
(1)
We have
-
(2)
We have if and only if for any , at least one of the following inequalities holds
-
(3)
If then there exists such that , , and
Proof.
(1). We only demonstrate the proof of the result for and the proof for is similar. First, We make the following two claims.
Claim 1. For any homogeneous elements (not necessarily elements in ) and such that is also homogeneous, if then
To prove Claim 1, let be sufficiently large. From Lemma 6.11 part (3) we know that
(7.3) |
Assume . Let
We claim that there exist
such that
We prove only the existence of , and the argument for the existence of is similar. By Definition 6.8, we know that
Taking
we know that
(7.4) |
Equation (7.3) implies that . The definition of makes sure that . The fact that is sufficiently large and Lemma 6.12 part (4) implies that . Hence Lemma 6.12 part (3) implies the existence of such that
Now the existence of and implies that
which contradicts the definition of .
Claim 2. Suppose are pairwise distinct elements in such that
Suppose
and suppose it is homogeneous. Then .
To prove Claim 2, assume that . Without loss of generality, assume that and
Then a similar argument as in the proof of Claim 1 implies that
Note that we have assumed . Hence by Definition 6.8, , which contradicts the construction of the set .
Now we prove part (1). Suppose are pairwise distinct elements in . Let
We want to show that
To do this, relabel the elements if necessary such that
Since is homogeneous, from Lemma 6.11 part (3), we know that the sum
is also homogeneous for any . Applying Claim 2, we conclude that
Hence we can apply Claim 1 repeatedly to conclude that
(2). If , then there exists and such that
By assumption
with and homogeneous. By Lemma 2.18 we have
Since forms a basis for , we can write
where . Then for any , at least one of and is non-zero. Since both and are homogeneous, from part (1) we know
Conversely, suppose for any at least one of the following inequalities holds
We need to show . We deal with three cases.
Case 1. The grading satisfies
We want to argue that
is surjective and hence conclude that . To do this, note that is the composition of maps for . With the assumption of Case 1, we have
(Note that since is sufficiently large this is a very loose inequality.) Then Corollary 2.9 part (2) applies and we conclude that is an isomorphism for all . Hence we conclude Case 1.
Case 2. The grading satisfies
The argument is similar to that for Case 1, except for using instead of .
Case 3. If the grading satisfies
Under the assumption of Case 3, Lemma 2.19 part (1) implies that is injective when restricted to .
Now, for each , if we have
we claim that there exists such that
If instead
we claim that there exists such that
We will verify the existence of or in a moment, but for now let be the sum of all ’s and be the sum of all ’s. Then from Lemma 2.18 it is straightforward to check that
Since in Case 3 the restriction of on is injective, we conclude that
It remains to show that the desired or exists. We only prove the existence of and the argument for is similar. Now assume that
This implies that
The hypothesis of the Lemma and the definition of in Definition 6.8, together with Lemma 6.11 part (3) imply that
As a result, there exists an integer such that
Note that, from Definition 6.8, we know
As a result, we have
The assumption in Case 3 and Lemma 6.12 part (4) then implies that . Hence Lemma 6.12 part (3) implies the existence of .
(3). If , then part (2) means that there exists some such that
Note that, by Lemma 6.11 part (3), we must have
By direct calculation, we have
Then we can choose with as desired. ∎
7.2. The construction of the map
Since and are homogeneous, we can construct 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 preserves the gradings. From Lemma 2.5, for any grading with
we have . From Corollary 2.9, we know either or is surjective onto for such grading . Thus, on such grading , the zero map satisfies the exactness for (though we still have to verify the commutativity in Proposition 7.2).
On the other hand, from Lemma 2.19, the restriction of on the consecutive middle gradings is an isomorphism. In particular, when , it is an isomorphism when restricted to each middle grading. Also from Lemma 7.3, it seems that the definition of on should involve . However, if we simply take
as the definition, the current techniques fall short of demonstrating exactness and commutativity.
We resolve this issue by introducing an isomorphism
and define
(7.5) |
The construction of 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 . In this case so up to a scalar we have . In this special case indeed we could prove the exactness and commutativity without explicitly writing down the isomorphism as follows.
We first define the map on the basis
of 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 small enough such that Corollary 2.9 and Lemma 2.19 apply. For any , there exists a grading such that there exists with
Note that, from the above equality, we know that
Note that, except for , the rest of the terms are bounded, so when is small enough. Similarly,
so we have when 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
Then by Lemma 6.7, there exists such that
(7.6) |
Let
From Lemma 2.19, we know
is an isomorphism. Hence we define
The following diagram might be helpful for understanding the construction of . (We write , , and .)
Remark 7.6.
For a general -manifold , our construction of is noncanonical since there are many choices such as the basis and the element for each . However, one could still ask whether we could simply pick or not. If we take , 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 to bypass this conjecture.
Conjecture 7.7.
For any , and any integer , we have
where
and
is the projection.
Lemma 7.8.
We have the following.
-
(1)
Suppose and , , are chosen as above. Suppose are two integers such that . Then (a) if and only if (b) and (in particular, we have ).
-
(2)
The map is an isomorphism.
-
(3)
For an element , an integer and a grading , the following two statements are equivalent.
-
(a)
We have
-
(b)
We have and there exists such that and
-
(a)
-
(4)
Suppose for an integer and a grading we have such that for all , then ,…, are linearly independent.
-
(5)
Suppose . For any such that , we have
Proof.
(1). First, when , from the construction of , we know that
The last equality is from Lemma 6.14 part (6). Similarly, if , we know from Lemma 2.11 that
Next, we need to show that when and . Again, from Lemma 2.11 we have
(2). Suppose where . We order the elements such that
Let , be the data associated to as above. Since
for any , and by Lemma 2.19, the map
is an isomorphism, in order to show that is an isomorphism, it suffices to show that ,…, are linearly independent.
Now suppose there are complex numbers ,…, such that
Our goal is to show that all are zero. The idea is to apply various maps to filter out different indices by part (1) of the lemma.
Applying the map , from the construction of , the order of and part (1) of the lemma, we know
where the summation in the second line is over all with
Note that, from Lemma 6.14 part (2) and the convention after the lemma, we know . From Lemma 6.14 again, we know that are linearly independent, and as a result all relevant must be zero. Suppose is the smallest index in the rest. By our choice of , the element has the largest among the rest of the . Hence we can apply the map to filter out with smaller . Repeating this argument, we could prove that all must be zero.
(3). Let , , , and be constructed as above. We first prove that (b) (a). Note that, when constructing the isomorphism , from Corollary 2.9 and Lemma 2.18 we can take and that will lead to the same as , . (Note that, by construction, passing from to will increase by .)
Remark 7.9.
Note that the main goal of the current paper is to derive an integral surgery formula. For that is sufficiently large, we already know a large surgery as in [22]. When is small enough, we can pass to for the mirror of the knot. As a result, instead of changing for particular , we could assume a universal bound for all the integers that we care about and make universally small.
As a result, we can always assume that is small enough compared with any given . Now recall by construction
and by the assumption in (b) we have
We can assume that is small enough such that . Take . Note that, by construction we have , so from Lemma 2.6 we know that
As a result, we conclude from the definition of and Lemma 2.18 that
Then it is straightforward to verify that and . As a result, we conclude from part (1) that
Next we show that (a) (b). Again assume that is small enough compared with the given . Then there exists such that there exists with
By Lemma 2.18, we know that
From the construction of and Lemma 2.19 we know only if , in which case
Hence implies that
by part (1). Taking , it is then straightforward to check that
(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 of as in Section 6.3.
Proof of Proposition 7.1.
We will verify the exactness at each space of the triangle.
The exactness at . This follows from Proposition 5.1.
The exactness at . From Lemma 7.3 and the construction of in (7.5), we know that . Now pick an arbitrary
Since is an isomorphism, we can assume that
where and . From Lemma 7.8 part (3), we know that this implies that for any , we have and there exists such that
Now, for sufficiently large, we have . Taking
it is straightforward to verify that when is sufficiently large, we have
Hence by Lemma 6.12 part (3), there exists such that . As a result, it is straightforward to check that
The exactness at . Suppose and
with and .
First, if , then from Lemma 7.4 part (2), we know that for any , we have
If we write
for some then the inequality
implies that
Note that the last equality uses the definition of in Definition 6.8. Similarly, we can compute that
implies that
In summary, implies that for all , either or . Hence from Lemma 7.8 part (3), we know that
for all and as a result, .
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 of as in Section 6.3.
Lemma 7.10.
Suppose and is a grading. Suppose such that
with and for all . Then for any , there exists such that
Proof.
Proof of Proposition 7.2.
We only prove the first commutative diagram
The other is similar. Note that at the end of Section 6.3, we introduce new notations of to remove the scalars. Then the second commutative diagram only holds up to a scalar.
First, note that the maps from to and both factor through . As a result, we only need to prove the following commutative diagram for sufficiently large .
Now suppose . Write
with and for . We want to first establish an identity
(7.7) |
and then show that the other composition has exactly the same expression.
From Lemma 7.10, we know for any , there exists such that
(7.8) |
Taking and , we can apply Lemma 6.12 part (3) to find an element
such that
It is then straightforward to check that
(7.9) |
Write
From Lemma 2.18 we know that
As a result, by Lemma 6.6,
Note that, unless , we have
by the exactness. As a result,
This verifies Equation (7.7) if we show that
To verify this last equality, assume that . Then from Lemma 2.12 and the exactness of the bypass maps we have
(7.10) |
Now we deal with . Since , Lemma 2.19 implies that
Hence
where is defined as in (7.9). Note that, by definition we have , so from Lemma 2.6, we know
Note that, by (7.9) and Lemma 2.18, we know that
Hence
We write
Comparing the above formula with (7.8), we know
Note that, by construction, , which means . Hence from Lemma 7.8 we know
if and only if , i.e., . Also when from Lemma 7.8 part (5) we know
Note that we could focus on indices such that This is because if an index makes , then on one hand it does not contribute to since the corresponding summand is , on the other hand, we have hence per Equation (7.7) it does not contribute to , either. Also, we know from Lemma 7.8 part (3) that when we must have . As a result, we know
∎
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 . 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 . 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.