Splitting maps in link Floer homology and integer points in permutahedra
Abstract.
In this paper, we study the skein exact sequence for links via the exact surgery triangle of link Floer homology and compare it with other skein exact sequences given by Ozsváth and Szabó. As an application, we use the skein exact sequence to study the splitting number and splitting maps for links. In particular, we associate the splitting maps for the torus link to integer points in the -dimensional permutahedron, and obtain the link Floer homology of an -component homology nontrivial unlink in .
1. Introduction
In this paper, we study various maps in Heegaard Floer homology associated to crossing changes in link diagrams. Given such a diagram with a chosen crossing, we can consider three links , and in the three-sphere corresponding to the positive crossing, negative crossing and oriented resolution (see Figure 1). We will always assume that the crossing is between different components of , so that and both have one more component than .
For various technical reasons, we work with the “full” version of the Heegaard Floer homology with two marked points on each link component over , developed in [32, 33]. In particular, for a link with components in , is a -graded module over where all products act by the same operator which we will denote by . Sometimes we will need to work with the completion which is a module over the power series ring .
Our first result describes the crossing change maps in this version of Heegaard Floer homology generalizing the maps in [26, 21, 30] for and .
Theorem 1.1.
Given a crossing between the components and of an oriented link in the three-sphere, for all corresponding to -structures in certain surgery cobordism shown in Figure 3, there are maps
(1) |
satisfying the following equations:
-
(a)
The maps are determined by and :
-
(b)
We have and .
-
(c)
The maps are determined by and :
-
(d)
The maps and compose as follows:
The rest of compositions are determined by these.
See Section 3 for more details and the gradings for all these maps. In [30] Ozsváth and Szabó proved a skein exact triangle for
where is some given bigraded module.
We generalize this as follows.
Theorem 1.2.
Given a crossing between the components and , there is an exact triangle
(2) |
where the map is given by , and is the completion of the module defined in (6).
Theorem 1.1 implies the following:
Corollary 1.3.
We have where is an explicit invertible power series in defined in Lemma 4.5. In particular, the cones of and of are homotopy equivalent.
Note that the map has homological degree and can be defined without completion.
Remark 1.4.
Since we work over the field , the signs here and below are purely for esthetic reasons. However, we expect all the maps to exist for theories with integer coefficients (similar to [1]), and conjecture that (up to an overall normalization) the signs would match. See also Section 6.4 on comparison of the signs with triply graded Khovanov-Rozansky homology.
Next, we study the compositions of crossing change maps. Since we have two essentially different maps and (resp. and ) for a single crossing change, for a sequence of crossing changes we have possible associated maps in Heegaard Floer homology of various degrees, some of which may coincide. We determine the degrees of all such maps in Section 5 and use them to bound splitting numbers for links.
In a striking example, we can take the -component torus link , change crossings between different components from positive to negative and obtain the unlink . In this case, we are able to completely determine all crossing change maps.
Theorem 1.5.
If one chooses either or for each of crossing changes from to , the Alexander degrees of the resulting maps correspond to integer points in the permutahedron . Any two maps of the same degree coincide, and any integer point in corresponds to an injective map which can be described explicitly on generators of .
For example, is a hexagon with 6 vertices and 1 interior point, see Figure 14. To get from to unlink, one needs to change 3 crossings, so there are possible splitting maps. Six of them correspond to the vertices of and two remaining ones coincide and correspond to the interior point of . We generalize Theorem 1.5 to arbitrary L-space links in Section 5.3.
Theorem 1.6.
Suppose that is an L-space link. Then:
a) For any choice of crossing changes and the maps at the crossings, the resulting map is completely determined by its Alexander and Maslov degrees.
b) If, in addition, all crossings between the different components of are positive, the splitting maps are in bijection with the integer points in a certain polytope (see Definition 5.4).
One can also study the compositions of maps from skein exact sequence (2) for crossings in between and . Let be the ideal in generated by determinants
for all possible -element subsets . We denote by the completion of in .
Theorem 1.7.
a) Let be the composition of the maps from Theorem 1.2 over all . Then is injective and its image is the ideal in .
b) Let be the composition of the maps from Corollary 1.3 over all . Then is injective and its image is the ideal in .
Corollary 1.8.
We have as modules over .
Theorem 1.7 can be compared with the main result of [12] where the “-ified” triply graded Khovanov-Rozansky homology (also known as HOMFLY homology) of was computed using a very similar ideal to , see Section 6.4. This suggests a spectral sequence from the “-ified” HOMFLY homology to which we plan to study in a future work. Such a spectral sequence should generalize the spectral sequences for reduced homology studied in [5, 10, 11]
Finally, we can use the above results to compute the Heegaard Floer homology of certain links in .
Theorem 1.9.
Let be the link consisting of parallel copies of inside . Then where
and
Acknowledgments
We are grateful to Daren Chen, Matthew Hedden, Peter Kronheimer, Tye Lidman, Robert Lipshitz, Ciprian Manolescu, Lisa Piccirillo and Ian Zemke for useful discussions. A. A. and E. G. were partially supported by the NSF grant DMS-1928930 while they were in residence at the Simons Laufer Mathematical Sciences Institute (previously known as MSRI) in Berkeley, California, during the Fall 2022 semester. A. A. was also partially supported by NSF grants DMS-2000506 and DMS- 2238103. E. G. was also partially supported by the NSF grant DMS-1760329. B. L. is partially supported by the NSF grant DMS-2203237.
2. Background
2.1. Lattices
We will work with the lattice and its translates. We define a partial order on by
We will denote the basis vectors by . Given a vector , and a set of variables (resp. ), we write
2.2. Variables and gradings
We will be working with links in and the “full” version of Heegaard Floer complex for links, defined in [32]. The coefficients are in . Let be an oriented link with components. Unless stated otherwise, we will assume that each component has exactly two marked points and . The corresponding link homology is a module over the polynomial ring . We let denote the ring in variables satisfying the relations
The actions of on the complex are pairwise homotopic, and the action of on factors through .
Further, define
and .
We denote by the linking number between the components and , and write . Moreover, we let .
The link Floer homology has an Alexander grading valued in the lattice
It also has a homological (or Maslov) grading and an additional grading satisfying
Thanks to the relation between , and , we can determine from the Alexander and Maslov gradings. Note that the differential on the chain complex preserves Alexander multi-grading and changes the Maslov grading by 1. So, for any , let denote the subcomplex of generated by the elements of . The variable decreases by , decreases by and preserves , while the variable increases by , preserves and decreases by . Therefore, the coefficient ring for the subcomplex is the subring and so is an -module.
For example, the homology of the unlink with components has one generator in Alexander degree and Maslov degree 0, and is isomorphic to the ground ring .
Sometimes we will need to work with the completion which is a module over the power series ring .
2.3. Specializing
We will need to compare the above construction of Heegaard Floer homology with more “classical” ones [23, 26, 20]. This is done by specializing in various ways.
First, we specialize for all and denote the specialized complex by following [26]. The specialized complex still has commuting actions of , which are all homotopic to . Since , the specialized complex has a homological grading given by . On the other hand, the Alexander grading becomes Alexander filtration, as follows:
Proposition 2.1.
For all , there is a bijection between the generators of of Alexander grading and the generators of of Alexander grading exactly i.e. generators of :
The span of such generators, denoted by , is a subcomplex of , and such subcomplexes yield a -filtration on .
Another specialization is for all . Similarly to the above, one immediately verifies that this is equivalent to considering the associated graded complex with respect to the Alexander filtration.
2.4. Large surgery and L-space links
We recall the large surgery theorem of Manolescu and Ozsváth:
Theorem 2.2 ([20]).
Let be an -component link in the three-sphere. For denote the -manifold obtained by performing -surgery on for all by
Then, for and arbitrary we have an isomorphism of graded -modules, up to a grading shift:
where is a -structure on determined by .
Corollary 2.3.
As a graded -module, the homology splits as a direct sum of one copy of and some -torsion.
The link invariant , known as the -function, is defined as the for the generator of the -summand of .
An oriented, connected, closed 3-manifold is an L-space if it is a rational homology sphere, and for each -structure on , one has . A link is called an L-space link if is an L-space for . Since Dehn surgery does not depend on the orientations of the link, being an L-space link is independent of the orientations on the components of the link.
Corollary 2.4.
For an L-space link, we have for all .
This is a useful way to characterize L-space links [18]. That is, a link is an L-space link if and only if the link Floer homology is torsion free as an -module.
Example 2.5.
Let be the unlink with components. Then, in Alexander grading we have
where and . Note that has homological degree , so .
Example 2.6.
The Hopf link with linking number and the negative Hopf link with linking number are both L-space links. In [7], there are explicit computations of of these two links using the Heegaard diagrams in Figure 2. The link Floer chain complex of the positive Hopf link is a module over given as follows:
where the gradings of are the following:
The gradings of are as follows
Hence, the full homology of the Hopf link is generated by and can be written as
The link Floer chain complex of the negative Hopf link is the dual complex of , i.e.,
where the gradings of are
Hence, the full homology of can be written as
2.5. Cobordism maps and link TQFT
We first review 3-manifolds with multi-based links and decorated cobordisms between them. A 3-manifold with a multi-based link consists of an oriented closed 3-manifold , an oriented, embedded link together with disjoint collection of basepoints and on such that each component of has at least two basepoints , and the basepoints alternate between those in and those in when one traverses a component of . The basepoints and correspond to the variables and in a polynomial ring where . Then, is defined as a curved complex over .
In this paper, we mainly consider the case that each component of a link has exactly two basepoints, i.e., the link component contains and in and , respectively, and . Furthermore, for simplicity, we will drop and from the notation of a multi-based link if the context is clear.
A coloring of a multi-based link is a map , where is a finite set, considered as the set of colors. Corresponding to the set of colors , a polynomial ring
is defined, which clearly is a -module. For a colored multi-based link
Definition 2.7.
[33, Definition 1.3] A decorated link cobordism from a 3-manifold with a multi-based link to another one consists of a pair such that
-
(1)
is a compact 4-manifold with .
-
(2)
is an oriented, properly embedded surface in , along with a properly embedded 1-manifold in , called dividing arcs. Further, consists of two disjoint (possibly disconnected) subsurfaces, and , such that the intersection of the closures of and is .
-
(3)
.
-
(4)
Each component of (and ) contains exactly one basepoint.
-
(5)
The basepoints are all in and the basepoints are all in .
-
(6)
is equipped with a coloring , i.e. a map , where denotes the set of component of .
To a decorated link cobordism and a structure on , Zemke[32, Theorem A] associated a functorial chain maps
Here, denotes the colorings on obtained by restricting , for . The maps are -equivariant, -filtered, and are invariants up to -equivariant, -filtered chain homotopies.
Another version of functorial maps for decorated cobordism between links have been independently defined by the first author and Eftekhary in [2].
Convention 3.
In this paper, we consider the case that every component of a link has exactly two basepoints (unless when we stabilize them), i.e., the link component contains and in and , respectively, and so . Moreover, mostly we work with special cobordisms that every connected component of is an annulus, decorated with two parallel vertical dividing arcs. More precisely, for , and where each is an annulus with . Further, each consists of two parallel, vertical dividing arcs connecting to and dividing into two rectangles, one containing and another containing basepoints. Finally, our coloring set , which is the codomain of , contains exactly colors, and such that under this identification and are identified with and , respectively. Here, and are the basepoints on . Thus, if we do not emphasis on the basepoints, dividing curves and the colorings, we automatically mean this fixed convention.
For a decorated cobordism as above the cobordism maps are -equivariant and -filtered. The grading changes under the cobordism maps are as follows:
Theorem 2.8.
(Special case of [33, Theorems 1.4 and 2.14]) Suppose is a decorated link cobordism from to . Then,
-
(1)
If and are torsion, then is graded with respect to , and satisfies
-
(2)
If and are torsion, then the map is graded with respect to , and satisfies
-
(3)
Suppose and are null-homologous links, i.e. in for and . Moreover, assume both and are torsion. Then,
where denotes the closure of by adding arbitrary Seifert surfaces of and , and
Theorem 2.9.
Assume that is a decorated link cobordism with as in Convention 3. Then for all and the induced map on homology
is an isomorphism, where is the Alexander multi-degree of .
Proof.
Consider the diagram
where the left (resp. right) vertical arrow is defined by sending (resp. ) to (resp. ). Moreover, is the cobordism map corresponding to and as defined in [27]. Similar to Proposition 2.1, it is easy to see that the vertical maps are chain maps and define an isomorphism between the chain complexes. Moreover, it follows from the definition of the cobordism maps that this diagram commutes. By the proof of [22, Theorem 9.6] the induced map on homology by is an isomorphism and thus the induced map on homology by is an isomorphism as well.
∎
Corollary 2.10.
Assume is a decorated link cobordism with , and and are -space links. If has Alexander multi-degree and homological degree then for all the induced map on homology
is injective and completely determined by its homological degree.
Proof.
Let denote the generator of of homological degree , for . Then
where
∎
3. Surgery maps
3.1. Crossing changes
As shown in Figure 3, one can locally change a positive or negative crossing by performing -surgery on the specified red unknot. In this section, we associate a link cobordism with a simple decoration to these crossing change surgeries and then study the properties of the corresponding cobordism maps.
Suppose is an -component link in , and is the link obtained from by changing a positive crossing between different components to a negative crossing. So, will have components as well. Denote the component of corresponding to by . Let be the cobordism from to obtained by attaching a -handle to in , along the -framed unknot as in the top of Figure 3. Then, the embedded surface in gives a cobordism from to . The surface consists of connected components, all of them annuli. Denote the component of that bounds and by . Assume each connected component of contains exactly two basepoints , , and denote the corresponding basepoints on by and , respectively. Decorate each with two parallel and vertical arcs to divide into two rectangles, such that one of these rectangles contains the basepoins , and the other one contains . Then, for colored as in Convention 3, the pair gives a decorated cobordism from to . Similarly, we define a decorated cobordism from to using the unknot in the bottom of Figure 3 as well.
Proposition 3.1.
Let be the decorated link cobordism induced from attaching a 2-handle along the -framed unknot so that a positive crossing becomes a negative crossing, as above. Suppose that the link components are passing through the unknot with . Let be the structure on such that
where is the generator of corresponding to the attached 2-handle. Define , the corresponding cobordism map in link Floer homology. Then
and
Moreover, for all . In particular, , while and both have homological grading zero.
Proof.
Example 3.2.
Example 3.3.
Proposition 3.4.
For any link , the maps have the following properties:
-
(a)
For , we have .
-
(b)
For , we have
-
(c)
We have and .
Proof.
The Proposition holds for the special case that and by Example 3.3. We use the functoriality of link Floer homology and the properties of cobordism maps to show that the general case follows from this special case.
We stabilize by adding two extra pairs of basepoints and to and , respectively. Denote the stabilized links by . Moreover, we extend the coloring on to a coloring on so that its codomain is and
Note that restricts to a coloring of with codomain , and abusing the notation we denote it by . The isomorphism extends to an isomorphism
by sending and to and , respectively. Under this isomorphism
By [26, Section 6] (or [34, Proposition 5.3]) we have
and under this isomorphism the induced map from to itself by the quasi-stabilization maps (see [32, Section 4.1]) is identity. Note that corresponds to the quasi-stabilization cobordism from to obtained from the product cobordism by adding two dividing arcs on the -th and -th cylinders that split off disks containing and , as in Figure 4.
Next, we construct a decorated cobordism from to by modifying the decoration on , as follows. Add two parallel, vertical dividing arcs to (resp. ) such that they divide (resp. ) into four rectangles. Moreover, each one of them contains exactly one of the pairs , , and (resp., , and ) on its boundary. Clearly, extends to a coloring on . Define
Under the aforementioned isomorphism , the homomorphism induced by the cobordism map from to is equal to .
On the other hand, can be decomposed as a cobordism containing two births from to followed by two band attachments from to as in Figure 5.

On the other hand, one may isotope the attaching circle of the -handle in as in Figure 6. Then, changing the order of -handle attachment and band attachments as in Figure 7, we get a decomposition
where denotes the change of crossing cobordism map from to and is the band attachment cobordism from to .
By [32, Theorem B] for any structure we have
Moreover, for any multipointed colored links and , and under corresponding identifications
where denotes the map for the unlink . So, the claim holds, because equalities hold for the change of crossing maps for the unlink from Example 3.3.
∎
As in the bottom of Figure 3, -surgery on the specified red unknot can change a negative crossing to a positive crossing. Hence, we can also consider the cobordism from to induced by attaching a -handle along this unknot. The embedded surface is a disjoint union of annuli, and each one of them is equipped with two parallel and vertical dividing arcs.
Proposition 3.5.
Let be the decorated link cobordism induced by attaching a 2-handle to the -framed unknot in Figure 3 which changes a negative crossing to a positive crossing. Suppose that the link components are passing through the -framed unknot with . Let be the structure on satisfying that
where is the generator of corresponding to the attached 2-handle. Let be the corresponding map in link Floer homology. Then
and
Moreover, for all . In particular, , and , .
Proof.
The proof is very similar to the one of Proposition 3.1. By the same computation, we get . For the Alexander gradings, note that
The same computation works for . However, for all . Hence, for all such . Note that , so .
∎
Example 3.6.
Example 3.7.
Let and . Then by Corollary 2.10 the cobordism maps are nonzero and determined by the grading shifts. In particular,
In general, we have
Proposition 3.8.
The maps satisfy the following properties:
-
(a)
For , we have .
-
(b)
For , we have
-
(c)
We have and .
Proof.
The proof is very similar to that of Proposition 3.4. By Example 3.6, the claim holds for and . It remains to show that the general case follows from this special case. Let denote the decorated crossing change cobordism from to . Following the notation in the proof of Proposition 3.4, let denote the links obtained from by adding two extra pairs of base points on and , and be the decorated cobordism from to obtained from by adding two pairs of parallel and vertical dividing arcs. Further, denotes the decorated quasi-stabilization cobordisms from to . As in the proof of Proposition 3.4, for the decorated cobordism , composing with a specific isomorphism is equal to identity.
We now still decompose as a cobordism from to followed by two band attachments from to , as in Figure 8.
Changing the order of the -handle attachment in and the band attachments in , we get
where denotes the cobordism from to and is the band attachment cobordism from to , see Figure 9. Hence, for any Spinc structure we have
Here where denotes the map for the unlink . Hence the claim holds for general links.
∎
Proposition 3.9.
Proof.
We prove the equalities in the first row, and the proof for the second row is similar. It is straightforward from Examples 3.3 and 3.7 that the claim holds for and the negative Hopf link, because
The strategy is similar to the proof of Propositions 3.4 and 3.8 and we fix the same notation. To distinguish the cobordisms defining and we use subscripts and , i.e. let be the decorated crossing change cobordism from to and be the decorated crossing change cobordism from to . As before, we denote the links obtained from by adding two extra pairs of base points on and by . Further, we denote the corresponding cobordism from to (resp. to ) obtained from (resp. ) by adding two pairs of parallel and vertical dividing arcs by (resp. ). Moreover, we consider quasi-stabilization cobordisms from to .
Let . For any , denote the structure on whose restriction to and is equal to and , respectively, by . Thus, under the aforementioned isomorphism the cobordism map is equal to .
On the other hand, as depicted in Figure 5, the cobordism can be decomposed as , where is the decorated cobordism from to corresponding to two births, and is defined by attaching two bands. By Figure 10, after an isotopy on the attaching circles of the -handles in and , we may change their order with the band attachments in to get another decomposition
Here, denotes the decorated cobordism from to corresponding to changing a positive crossing to a negative crossing in . Similarly, is the cobordism from to corresponding to changing a negative crossing to a positive crossing in . Thus,
and the claim follows from the special case of and .

∎
3.2. Full twists
In this section, we will apply similar computation as in Proposition 3.5 to get the properties of the cobordism map induced by attaching a -handle along a -framed unknot through -strand braid to get a positive full twist.
Using the similar computation as in Proposition 3.5, we have the following:
Proposition 3.10.
Let be the decorated link cobordism obtained by attaching a -handle on the -framed unknot which adds a full twist to the parallel strands as in Figure 11. Let be the structure on satisfying that
where is the generator of corresponding to the attached 2-handle. Let be the corresponding map in link Floer homology. Then
and
for .
Proof.
The computation of is exactly the same as the one of Proposition 3.5. Hence, . For the computation of the Alexander grading, it is also similar to the one of Proposition 3.5, except now for each , we have
∎
Now we consider the following example where and . It is known [13] that is an -space link. We first recall the link Floer homology . For the explicit computation, see [7].
Theorem 3.11 ([7]).
The Heegaard Floer homology has generators, which we denote by subject to the following relations:
(4) |
Here, is any subset of the set of length (so the first equation has relations for each ), and in the second equation , and range from to .
Now we list the explicit gradings of the generators where . The Alexander multi-grading of is
The generator has homological grading
The Maslov grading is obtained from the -function of the torus link , which is computed in [13]. The computation of follows from the relation
Example 3.12.
By Corollary 2.10 the cobordism maps are non-zero and determined by the grading shift formulas from Proposition 3.10. Recall that and . Then
for .
In general, we have
Similar to Proposition 3.8, the maps satisfy the following properties:
Proposition 3.13.
The maps satisfy the following properties:
a) For , we have .
b) For , we have
Proof.
The proof is very similar to the one of Proposition 3.8. As before, we denote (resp. ) as the link obtained from (resp. ) by adding an extra pair of basepoints for each component . Let be the induced decorated cobordism from to induced from the decorated cobordism from to , and be the induced coloring on and as in the proof of Proposition 3.4. We still get the isomorphism
Similarly, we also have .
As before, we let be the decorated cobordism from to , which can be decomposed as a cobordism from to followed by band attachments from to . Hence, where . Now we use the same trick as before to isotope attaching circle of the 2-handle as in Figure 6 so that it encircles the unlink and change the order of the 2-handle attachment and band attachments. Then
where denotes the cobordism obtained by attaching a -handle along -framed unknot from to and denotes the band attachment cobordism from to . Hence,
where denotes the map for the unlink . Since the proposition holds for unlink by Example 3.12, the general case follows.
∎
4. Skein exact sequence
4.1. Surgery exact triangle
Suppose is a link in an integer homology sphere , and is a knot. Let be the decorated link cobordism from to obtained by attaching a two-handle along with framing , and decorated as in Convention 3. Similarly, and denote the cobordisms from to and to , respectively.
Proposition 4.1.
The link cobordism maps form an exact triangle as follows.
Proof.
This is a straightforward generalization of the surgery exact triangle for in [24, Section 9]. We will outline the proof and highlight the differences here. Consider a multi-pointed Heegaard diagram
where , such that
-
•
, and are Heegaard diagrams for the link in -manifolds , and , respectively. So, and consist of basepoints, where is the number of connected components of .
-
•
For any , and are small isotopic translations of such that they intersect transversely in two points and are disjoint from for . Moreover, intersects in two transverse points as well.
-
•
Pairwise intersections of , and are single points with signs . Moreover, is obtained from the juxtaposition of and .
- •
Let be the chain map defined by counting holomorphic triangles as:
Analogously, we define chain maps and .
The Heegaard diagram represents an component unlink in , denoted by . It is straightforward that is a free -module of rank . Moreover, the summand with largest has rank one and so it has a unique generator. The Heegaard diagram has an intersection point denoted by that generates this top degree homology class, called top generator. Specifically, is the intersection point that every element of has nonzero coefficient in at least one or basepoint, for all other intersection points . Top generators and for and , respectively, are defined analogously.
By [29, Lemma 4.4] to show that they form an exact triangle, we need to check that
-
(1)
is chain homotopically trivial by a chain homotopy ,
-
(2)
is a homotopy equivalence,
where indices are cyclic modulo three. Note that we need a version of [29, Lemma 4.4] for chain complexes over , which for example follows from [1, Lemma 3.3].
First, we check condition . Suppose . The proof for and is similar. Then,
where the chain homotopy is and
An argument similar to the proof of [24, Proposition 9.5] implies that and so .
For condition , let be a generic small Hamiltonian isotopic translate of , and be the chain map defined by counting pentagons of Maslov index , analogous to and . Then, gives a chain homotopy between and . By a standard “stretching the neck argument” and following the strategy in [28, Section 2] and [29, Section 4.2] we have
and
Since is a homotopy equivalence (see the proof of [1, Theorem 8.6]) and is invertible, is a homotopy equivalence.
∎
Now let us relate the surgery exact triangle with resolutions. Suppose , is a link in with a fixed positive crossing, and is an unknot as in the top of Figure 3. Then, will be identified with . Next we relate with , where denotes the oriented resolution at the fixed crossing. Observe that , and in still has components, while is an -component link. Note that we can replace the 2-handle attaching to with framing by a 1-handle as in Figure 12.

Lemma 4.2.
The link can be identified with where is the 2-component unlink in consisting of two parallel circles representing the homology class of , and the between and is identified as in Figure 13.

Proof.
The proof is depicted in Figure 12. Specifically, we regard the -handle for the -surgery on as a -handle and then we move the feet of 1-handle along the link . At the end, we get the connected sum of with one component of , colored blue, along with the other component of , colored red, in Figure 13.
∎
Therefore, we have the following theorem:
Theorem 4.3.
Given a local positive crossing of the link components and of a link in the integer homology sphere , there is a skein exact sequence
(5) |
where the map from to is given by , and
Proof.
By Lemma 4.2, the cone of is the homology of the tensor of with some complex corresponding to the unlink in . Moreover, is independent of the pair . We use the special case that and to compute which gives the module . We put the detailed computation of Hopf link in Section 4.2 (equivalently, see (8)).
The complex is given as follows:
(6) |
Note that it is a free resolution of over . Since is a complex of free -modules, we get
∎
Remark 4.4.
In version of Heegaard Floer homology one sets , and the complex breaks into a direct sum of two copies of This explains the appearance of a two-dimensional vector space in [30].
Similarly, for one sets , and the complex breaks into four copies of .
Without loss of generality, we assume that for the rest of the section. Recall that and have components while has components. For all we have chain maps , and one can consider the formal sum
Note that is a non-homogeneous map containing terms of various non-positive homological degrees.
As the non-homogeneous map is hard to deal with, we would like to reduce it to the degree zero piece .
Lemma 4.5.
Let be an arbitrary link in the three-sphere with a fixed positive crossing between its first and second components. For , suppose is the corresponding crossing change map and . Then in homology we have where
(7) |
In particular, is an invertible power series.
Proof.
Corollary 4.6.
The cones of and of are homotopy equivalent.
Remark 4.7.
Note that at and , we get that the invertible factor is
which agrees with [27, Theorem 3.7, Blow-up formula] modulo . That is because in this case is equal to the Ozsváth-Szabó’s cobordism map associated to the blow-up of the product cobordism from to .
4.2. Skein exact sequence for Hopf link
Next, we compute the skein exact sequence on the chain complex level for the Hopf link. Recall that the full link Floer complex for has generators and the differential
The homology is generated by modulo relations as above.
On the other hand, the link Floer complex for the unknot has generators and the differential . By Example 3.2 we have
and the maps and can be uniquely lifted to chain complex level by setting
Furthermore, for all we have where we follow the notations in Lemma 4.5 and its proof. For concreteness, we can lift the statement of Lemma 4.5 to the level of chain complexes.
Lemma 4.8.
The map is homotopic to , where is defined by (7).
Proof.
We have
Define
then
We define the homotopy by and , and let Then,
Similarly, and we conclude that . ∎
By Lemma 4.8 we can replace the cone of by the cone of which is isomorphic to the following complex:
We can define and change the basis from to :
(8) |
Here we use the fact that the quotient complex is contractible.
Lemma 4.9.
Let , then the cone of is quasi-isomorphic to the cone of up to relabeling the variables.
Proof.
By Example 3.6 we have and for , so
The homology of the cone of is generated by and modulo relations and . Since the coefficients at and in are invertible, the result is generated by modulo relations
(9) | ||||
We claim that
(10) |
where
Indeed, let , then
and , hence
Now by (10) we can rewrite the equations (9) as
which is equivalent to
∎
Remark 4.10.
Note that the above computation is very similar to the one in Lemma 4.5, in particular, is related to (resp. is related to ) by exchanging the variables which corresponds to changing the orientation on a link component.
5. Link splitting maps
5.1. Splitting maps
Let be an arbitrary link in the three-sphere. We can change the crossings between different components in arbitrarily and consider the corresponding maps in Heegaard Floer homology: if we change a positive crossing to a negative one we can use either or , and if we change a negative crossing to a positive one we can use either or . This does not change the topological type of the components and we will denote the components for all such links by . In particular, by such crossing changes we can transform to the split link obtained by the split union of all link components .
More precisely, we consider two links related by such crossing changes, and a chain map obtained as a composition of:
-
•
of maps associated to positive crossings between and ;
-
•
of maps associated to positive crossings between and ;
-
•
of maps associated to negative crossings between and ;
-
•
of maps associated to negative crossings between and .
Note that, in principle, may depend on the order of the crossings and the choice of these crossings (for example, for the Hopf link we can change either one of two crossings to transform it to unlink). Also note that the linking number between and changes by
Nevertheless, we have the following general result:
Lemma 5.1.
Let be two links related by crossing changes as above. Then:
-
(a)
The chain map has Alexander degree
and homological degree
-
(b)
For any , the map , restricted to the -free summand, is injective and determined by its homological degree.
-
(c)
The -functions of and are related by the inequality
Proof.
Part (b) follows from Proposition 3.9. Note that is a composition of crossing change maps. For each crossing change map associated to the link components , by Proposition 3.9, one can choose a corresponding crossing change map such that the composition of these two crossing change maps is one of the monomials . Hence, we can compose the map with another map such that the composition is given by a monomial in variables . Therefore, the map is injective when restricted to the -free part of .
Now part (c) follows from (b): indeed, sends to
which should inject into Hence,
which yields the desired inequality. ∎
If and differ by a single positive crossing change, that is, is obtained by changing a negative crossing of into a positive one, we can recover the comparison between and in item (b) of Theorem 6.20 in [6].
Corollary 5.2.
Let be an -component link in . Suppose is obtained by changing a negative crossing between the components and of into a positive one. Then
for all .
Proof.
Note that is obtained from by changing a negative crossing to a positive crossing, which induces a cobordism map or . So we can make or . By Lemma 5.1, if then
If then
Conversely, can be obtained from by changing a positive crossing to a negative crossing. Then we can make or . By Lemma 5.1 again, if , then
If , then
∎
Note that [6, Theorem 6.20] uses the -function, which is some normalizations of the -function in the following formula:
for . Let . Since is obtained from by changing a negative crossing to a positive crossing between and . Then . Then the inequalities in Corollary 5.2 become
So we recover Theorem 6.20 in [6]. Moreover, we have an extra inequality that . Note that or for all links and all . So the above inequalities imply that or .
5.2. Positive links
Suppose that all crossings between different components of are positive (in particular, this holds if is a positive link). Then there are exactly crossings between and , and we need to change of them to split the components and . We can encode crossing change maps as above, with and
Then Lemma 5.1 simplifies dramatically, and we get the following
Corollary 5.3.
Suppose that all crossings between different components of are positive. Let be a composition of maps of type and maps of type between the components and for all . Define , then has Alexander degree
(11) |
and homological degree zero.
We can visualize the degrees of such maps as follows.
Definition 5.4.
Suppose that is a link such that all crossings between different components are positive, and . We define the link zonotope as the Minkowski sum of the intervals for all .
Note that is an -dimensional polytope contained in the hyperplane
It is centrally symmetric around the point where .
Example 5.5.
For the polytope is a segment with integer points on it.
Example 5.6.
For , the polytope is a hexagon where the opposite sides are parallel to each other and both contain integer points. The vertices of are:
Theorem 5.7.
After a shift by the vector , the Alexander degrees of splitting maps (11) of a positive link correspond to the integer points in the polytope . Conversely, any such integer point corresponds to at least one nontrivial splitting map.
Proof.
By Lemma 5.1 all compositions of splitting maps are nontrivial, and for positive links all such maps have homological degree zero, but we need to understand their Alexander degrees.
By varying and , the terms can have the values
By shifting these by , we get the points
which coincide with the set of integer points on the interval . By adding these degrees over all , we obtain an integer point in , and the overall shift equals
It remains to prove that any integer point in can be obtained as a sum of integer points in the intervals . This follows from the combinatorial results in [4, Section 9]. Indeed, by [4, Lemma 9.1] the polytope can be decomposed into a disjoint union of parallelepipeds of various dimensions labeled by the linearly independent subsets of the set (the edges of each parallelepiped have integer length ), see Figure 15. These parallelepipeds can be themselves decomposed into smaller parallelepipeds with edges of integer length 1. It is easy to see [4, Lemma 9.6] that the linearly independent subsets correspond to forests on vertices, and by [4, Theorem 9.5] the relative volume of each small parallelepiped equals 1. Therefore every vector connecting an integer point inside each large parallelepiped with a vertex can be written as a linear combination of vectors along the edges and the result follows. ∎
Remark 5.8.
In principle, there could be several splitting maps of the same degree. For example, for we can write the point at the center as a sum of integer points on intervals in two ways:
5.3. L-space links
If is an L-space link, then some results simplify significantly.
Theorem 5.9.
Suppose is an L-space link. Then:
-
(a)
For any choice of crossing changes and the maps at the crossings, the resulting map is completely determined by its Alexander and Maslov degrees.
-
(b)
If, in addition, all crossings between the different components of are positive, the splitting maps all have homological degree zero and are in bijection with the integer points in the polytope . Two splitting maps of the same Alexander degree coincide.
Proof.
If is an L-space link then by [18] all its components are L-space knots. In particular, the corresponding split link is an -space link as well.
By Corollary 2.4 this means that and are isomorphic to , and there is a unique nonzero map between two copies of of a given degree. The splitting map is nonzero by Lemma 5.1, so this implies (a).
Part (b) follows from (a) and Theorem 5.7. ∎
5.4. Torsion estimates
Recall that the splitting number is the minimal number of crossings between different components of a link that should be changed to turn into the split link. Let be the number of crossing changes between the -th and -th components, so that the resulting link is the split link.
Consider -dimensional lattice with basis . A point in this lattice parametrizes a monomial in . For each pair we consider the 3-dimensional tetrahedron:
Next, we consider the Minkowski sum
Remark 5.10.
Note that the intersection of with the -dimensional sublattice is the Minkowski sum of segments which is similar to the generalized permutahedra above.
Theorem 5.11.
Suppose that is an -component link where the components are L-space knots. Suppose that is the number of crossing changes between the -th and -th components so that the resulting link is a split link. Then:
-
(a)
For any integer point in the polytope the corresponding monomial in annihilates any torsion element in .
-
(b)
The monomial annihilates any torsion element in .
Proof.
(a) Given an integer point in , we construct two maps
as follows. Each time one changes a positive crossing to a negative crossing between and , we choose either or for and either or for ; if one changes a negative crossing to a positive, the roles of and are switched. Specifically, given nonnegative integers such that (or an integer point in the tetrahedron ) we can arrange the maps so that
-
•
To crossings between we associate
-
•
To crossings between we associate
-
•
To crossings between we associate
-
•
To crossings between we associate
By composing those crossing change maps, we obtain and . By Proposition 3.9 we get
The right hand side defines a monomial corresponding to a point in . Note that the composition does not depend on the order of the crossings, just the number of crossings of each type.
If is a torsion element in , then since there are no torsion elements in , therefore
(b) We can choose , then and
By part (a), the monomial
annihilates any torsion element in .
∎
Corollary 5.12.
Suppose that is a 2-component link with splitting number , and are L-space knots. Then for any such that the monomial annihilates any torsion element in .
Example 5.13.
We consider the following boundary links. Given any knot as in Figure 16, let be the new 2-component link obtained by applying an -twisted Bing doubling to . Observe that is a boundary link with unknotted components. The linking number of is , consisting of a positive crossing and a negative crossing between the link components. So one has to change both of the crossings to split the link, and the splitting number is 2 automatically. Hence, by Corollary 5.12 any monomial where annihilates any torsion element of for all and . In particular, the -torsion and -torsion are bounded by for all and where .
6. Example:
We illustrate all of the above constructions in detail for the torus link with components. By [13] it is an L-space link and its Heegaard Floer homology is given by Theorem 3.11.
6.1. Integer points in permutahedra and splitting maps
The permutahedron is the convex hull of points for all permutations . It is a convex polytope of dimension , and it is easy to see that the center of is at the point . By [4, Theorem 9.4] the permutahedron is the Minkowski sum of segments and hence agrees with the link zonotope .
Following Theorem 5.7 we will be interested in the integer points in . For example, is a segment connecting and . Next, is a hexagon with vertices obtained by permutations of which contains one more integer point at its center, see Figure 14 (right).
For we have a 3-dimensional polytope with 24 vertices corresponding to permutations of . Additionally, it contains 6 permutations of
and 4 permutations of both
In total we get integer points in . In general, it is known that the integer points in correspond to forests on labeled vertices, but we will not need this. We refer to [4, Chapter 9] for more information on permutahedra and integer points in them.
For all we define the vector111The reader might recognize the root system of type .:
Lemma 6.1.
For any integer ,
-
(a)
We have
-
(b)
For any choice of signs define
Then is an integer point in the permutahedron and all integer points in can be obtained this way.
Proof.
Part (a) is clear. Part (b) follows from the description of as a zonotope, that is, Minkowski sum of segments
see [4, Theorems 9.4, 9.5].
∎
Recall that by Theorem 3.11 has generators .
Theorem 6.2.
Let be a choice of signs as above. Choose a minimal sequence of crossings changes that splits . For any , this sequence contains exactly one crossing change between and . Consider the local crossing change map
and compose them to define the splitting map . Then is independent of the crossing change sequence and satisfies the following properties:
-
(a)
The Alexander multi-degree of equals and the homological degree is zero.
-
(b)
sends the generator to and every other generator to some other monomials in .
-
(c)
If then . In other words, the splitting maps for are parametrized by the integer points in the permutahedron .
-
(d)
For the map is defined on generators by the equation:
-
(e)
For any permutation there is a map corresponding to a vertex of . It is obtained from by permuting the indices of and by :
Proof.
Part (a) is clear and (c) follows from Theorem 5.9. Part (b) is immediate from (a) since has Alexander degree and . This agrees with the normalization of the generator of and neither nor change .
Example 6.3.
For we have as in Example 3.2.
Example 6.4.
For , the signs correspond to the point and the map . The maps for other vertices of can be obtained from it by permuting the variables.
The maps for and both correspond to the central point and the corresponding splitting map is given by following:
6.2. Surgery map
In this section, we study the map obtained by composing the surgery maps from Section 4 for any sequence of crossing changes. Specifically, we choose a minimal sequence of crossing changes that splits and we compose the local surgery maps associated to each crossing change to define . Below we will show that (at least on the level of homology) the choice of crossing change sequence does not matter and resulting maps agree.
To specify the link components involved in a crossing change, we will use subscripts for i.e. for a positive crossing between and we denote the local crossing change map by . Topologically, each map corresponds to a cobordism obtained attaching a 2-handle, and their composition corresponds to the composition of cobordisms . We have and . A choice of a -structure on corresponds to a choice of an integer and a map defined as in Section 3, so that . Similarly, a choice of a -structure on corresponds to a choice of a vector in the -dimensional lattice, and
The choices of and correspond, respectively, to choices of binary sequences and in previous section. In Theorem 6.11 below we prove that is injective on homology. To describe its image explicitly, we need to introduce some algebraic notations.
Definition 6.5.
Let be a subset of cardinality , order its elements as . Then we can define the polynomial
Reordering the elements of changes the sign of . Sometimes we will use notation for -tuples where some elements are repeated, in this case .
Definition 6.6.
We define as the ideal generated by for all possible -element subsets .
Remark 6.7.
The polynomials and the ideal in generated by were first introduced by Haiman in his work on Hilbert scheme of points on the plane [14].
The following lemma gives a useful characterization of the ideal , we postpone its proof until Section 6.3. It can be used as an alternative definition of .
Lemma 6.8.
The ideal is generated by the maximal minors corresponding to -tuples of consecutive columns in the matrix
It is clear that all the minors in Lemma 6.8 are of the form for some choices of subsets .
Example 6.9.
For we have two determinants
Example 6.10.
For we have three determinants
Now we are ready to state the main theorem of this section.
Theorem 6.11.
The surgery map is injective and its image coincides with the (completed) ideal . The map does not depend on the order and choices of crossing changes. It particular,
as modules over .
Proof.
For the reader’s convenience, we break the proof into several steps.
Step 1: By Lemma 4.5, each map is proportional, up to an explicit invertible factor , to The factors do not depend on the order of crossing changes, and do not affect the injectivity or the image (which is an -submodule of ), so we can ignore them from now on and focus on .
Step 2: By following the notations of Theorem 6.2, we can then rewrite the composition of as
(12) |
By Theorem 6.2, for any given the order of composition does not matter.
Step 3: The terms in (12) are parametrized by the integer points in the permutahedron . However, some points will appear several times for different choices of , and by Theorem 6.2(c) the corresponding terms in (12) might cancel as long as they have the same Alexander degree. We claim that in fact the terms for all points will cancel except for the vertices of . To show this, consider the generating function
Here, we used that and . On the other hand, we have the Vandermonde determinant
As a conclusion of this step, we can write
Step 4: We are ready to compute the image of or, equivalently, of . Indeed, by Theorem 6.2(d),(e) we get
By Lemma 6.8 these determinants generate the ideal .
Step 5: It remains to prove that is injective. Indeed, is an -space link, so in each Alexander multi-degree we have . By the above, the image of any element of this tower under is a sum of elements in towers in located at the vertices of a permutahedron centered at . Since all these elements appear with nonzero coefficients, their sum is also nonzero. ∎
6.3. Proof of Lemma 6.8
We start with several results which allow us to simplify the determinants . Given a subset , we define and
Note that some elements of may coincide even if all elements of are distinct. The subset is contained in the union of the horizontal strip and the vertical strip , dashed in Figure 17.
Lemma 6.12.
We have for .
Proof.
For all we have
and
∎
Example 6.13.
For we have , see Figure 17. In this case and we have
Let be the -th elementary symmetric function. We have the following analogue of the Pieri rule for Schur functions [19].
Lemma 6.14.
We have
where is obtained by adding to distinct elements of and leaving other elements unchanged.
Proof.
Given a polynomial , we define
For we get Clearly, is linear in and for any symmetric polynomial . Since is symmetric, we get
where
∎
Example 6.15.
We have
The first term vanishes since is repeated twice. By Lemma 6.12 the second term equals
and the third term equals
Therefore
Example 6.16.
Similarly, we have
and all other terms vanish. The first term can be simplified as
Proof of Lemma 6.8.
Recall that is generated by the determinants for arbitrary subsets . We need to prove that it is generated by where
By Lemma 6.12 we have proportional to where all elements of have the form or (that is, is contained in the dashed region in Figure 17), so after reordering its elements we can write
It remains to prove that, in fact, it is sufficient to only consider the “dense” subsets where and . There is a natural partial order on such where if is obtained by “sliding” some elements of left or down. Clearly, “dense” subsets are minimal in this order.
Assume that , and is maximal with this property, that is, bounds the rightmost gap in . This implies that for . Let
In Example 6.15 we have
in Example 6.16 we have
Then by Lemmas 6.14 and 6.12 (see also Examples 6.15 and 6.16) we have
where . Therefore belongs to the ideal generated by and . We can proceed by induction in the above partial order until are dense. Then repeat the same argument swapping and and using a version of Lemma 6.14 for elementary symmetric functions in , this will ensure that are dense as well. ∎
6.4. More on ideal and its cousins
In this section we collect some further facts on the ideal and discuss some analogues of Theorem 6.11 for other homology theories.
Definition 6.17.
A polynomial is called antisymmetric if
Note that over our ground field of characteristic 2, any antisymmetric polynomial is also symmetric. However, for other ground fields there is an important distinction.
Lemma 6.18.
Let . Then any antisymmetric polynomial with coefficients in is a linear combination of minors .
Proof.
Let be an antisymmetric polynomial, and be a monomial in with nonzero coefficient. If and then the transposition fixes this monomial, but since is antisymmetric it must change its sign, contradiction. Therefore all pairs are distinct and the -orbit of this monomial adds up to for . ∎
Informally, we can think of as a characteristic 2 reduction of the ideal generated by antisymmetric polynomials in . Next, we check directly that Theorem 6.11 is compatible with Theorem 3.11. By Lemma 6.8 the ideal is generated by minors and it is sufficient to check the relations between them.
Lemma 6.19.
Proof.
Recall that the determinants are antisymmetrizations of monomials
let us check the relations (4) between them for all possible . Let us first pick , then and
More generally, for an arbitrary define
Let , then
Here we used the relation .
Since the relations (4) are -equivariant, the relations for imply the same relations for . ∎
Next, we study the relations between the ideals corresponding to the link and its sublinks.
Lemma 6.20.
For any subset with let be the ideal generated by the monomial minors in variables . Then .
Proof.
Let , then one can write the minor as follows:
where the sum runs over all -element subsets . Since , we get and . ∎
Remark 6.21.
Topologically, the ideal corresponds to the union of the sublink formed by the components of with unknotted disjoint circles. Since the splitting map for does not depend on the order of crossing changes, we can first split off the components with indices not in , and the splitting map (respectively, ) will factor through the splitting map (resp. ) for the resulting link. Therefore the image of is contained in the image of .
Corollary 6.22.
We have
Proof.
By Lemma 6.8, for a two-component sublink we have , and for all . ∎
The analogues of the ideal and of Theorem 6.11 appeared in triply graded Khovanov-Rozansky homology. Recall that the triply graded homology of the unknot is where is an even and is an odd variable. There is also a skein exact triangle
analogous to the skein exact triangle (5). The map can be used to define a splitting map which is, unfortunately, not injective.
To resolve this problem, second author and Hogancamp introduced in [12] a deformation, or -ification of Khovanov-Rozansky homology . The skein exact triangle can be defined in this deformed theory, and there is a (unique up to homotopy) splitting map . By composing these, one obtains a splitting map
Theorem 6.23 ([12]).
The map is injective, and its image coincides with the ideal generated by antisymmetric polynomials in . Furthermore, the image of coincides with the ideal
Remark 6.24.
In [3] Batson and Seed defined a deformation of Khovanov homology, which was generalized in [9] by Cautis and Kamnitzer to Khovanov-Rozansky homology. Their constructions predate and motivate the construction of , and we commonly refer to them as -ified link homology.
Problem 6.25.
Define the analogues of the splitting map for -ified Khovanov and homologies. Is it possible to describe their images as some determinantal ideals in the homology of unlink?
6.5. Unlinks in
As an application of the above results, we can compute the Heegaard Floer homology of certain links in and prove Theorem 1.9. Let be the union of the parallel copies of inside . Recall from Section 3.2 that we have generalized crossing change maps
defined by attaching a 2-handle along the -framed meridian. The index corresponds to the choice of a structure on the corresponding cobordism, see Proposition 3.10. The surgery exact sequence immediately implies the following:
Lemma 6.27.
There is a long exact sequence
where is the sum of over all structures.
Remark 6.28.
As above, if we work with integer coefficients then would acquire some signs in the sum. Instead of guessing the signs, we simply work over .
Define two series
Theorem 6.29.
The homology admits two equivalent descriptions:
b) where is the (completed) determinantal ideal from Theorem 6.11 and
The second part of the theorem implies Theorem 1.9.
Proof.
a) By Lemma 6.27 we can write
Since all are injective on homology and have pairwise different Alexander degrees, is injective as well, and we can write
Now part (a) follows from Example 3.12 and Proposition 3.13.
b) By the proof of Theorem 6.11 the map provides an isomorphism and
in particular
and
Therefore
and the result follows. ∎
Remark 6.30.
It would be interesting to compare this result to the recent work of Kronheimer and Mrowka [16]. Their main result expresses the (deformed) instanton homology of the -component unlink in (with local coefficients) as a quotient of the polynomial ring by a certain determinantal ideal .
References
- [1] A. Alishahi, E. Eftekhary. A refinement of sutured Floer homology. J. Symplectic Geom. 13 (2015), no. 3, 609–743.
- [2] A. Alishahi, E. Eftekhary. Tangle Floer homology and cobordisms between tangles. J. Topol. 13 (2020), no. 4, 1582–1657.
- [3] J. Batson and C. Seed. A link-splitting spectral sequence in Khovanov homology. Duke Math. J. 164.5 (2015), pp. 801–841.
- [4] M. Beck, S. Robins. Computing the continuous discretely. Integer-point enumeration in polyhedra. Second edition. With illustrations by David Austin. Undergraduate Texts in Mathematics. Springer, New York, 2015.
- [5] A. Beliakova, K. Putyra, L.-H. Robert, E. Wagner. A proof of Dunfield-Gukov-Rasmussen Conjecture. arXiv:2210.00878
- [6] M. Borodzik, E. Gorsky. Immersed concordances of links and Heegaard Floer homology. Indiana Univ. Math. J. 67 (2018), no. 3, 1039–1083.
- [7] M. Borodzik, B. Liu, I. Zemke. Lattice homology, formality, and plumbed L–space links. arXiv:2210.15792
- [8] M. Borodzik, B. Liu, I. Zemke. Heegaard Floer homology and plane curves with non-cuspidal singularities. arXiv:2104.13709
- [9] S. Cautis, J.Kamnitzer. Knot homology via derived categories of coherent sheaves IV, coloured links. Quantum Topol. 8 (2017), no. 2, 381–411.
- [10] N. Dowlin. A Categorification of the HOMFLY-PT Polynomial with a Spectral Sequence to Knot Floer Homology, 2017. arXiv:1703.01401.
- [11] A. Gilmore. Invariance and the knot Floer cube of resolutions. Quantum Topol., 7(1):107–183, 2016.
- [12] E. Gorsky, M. Hogancamp. Hilbert schemes and -ification of Khovanov-Rozansky homology. Geom. Topol. 26 (2022), no. 2, 587–678.
- [13] E. Gorsky, J. Hom. Cable links and L–space surgeries. Quantum Topol. 8 (2017), no. 4, 629–666.
- [14] M. Haiman. -Catalan numbers and the Hilbert scheme. Discrete Math. 193 (1998), no. 1–3, 201–224.
- [15] M. Hogancamp, D. E. V. Rose, P. Wedrich. Link splitting deformation of colored Khovanov–Rozansky homology. arXiv: 2107.09590
- [16] P. Kronheimer, T. Morwka. Relations in singular instanton homology. arXiv:2210.07059
- [17] Y. Lekili, T. Perutz. Fukaya categories of the torus and Dehn surgery. Proc. Natl. Acad. Sci. USA 108 (2011), no. 20, 8106–8113.
- [18] Y. Liu. L-space surgeries on links. Quantum Topol. 8 (2017), no. 3, 505–570.
- [19] Macdonald, I. G. Symmetric functions and Hall polynomials. Second edition. With contributions by A. Zelevinsky. Oxford Mathematical Monographs. Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, 1995.
- [20] C. Manolescu, P. Ozsváth. Heegaard Floer homology and integer surgeries on links. arXiv:1011.1317
- [21] P. Ozsváth, Z. Szabó. A cube of resolutions for knot Floer homology. J. Topol. 2 (2009), no. 4, 865–910.
- [22] P. Ozsváth, Z. Szabó. Absolutely graded Floer homologies and intersection forms for four-manifolds with boundary. Adv. in Math. 173 (2003), no. 2, 179–261.
- [23] P. Ozsváth, Z. Szabó. Holomorphic disks and knot invariants. Adv. Math. 186 (2004), no. 1, 58–116.
- [24] P. Ozsváth, Z. Szabó. Holomorphic disks and three-manifold invariants: properties and applications. Annals of Mathematics, 159 (2004), 1159–1245.
- [25] P. Ozsváth, Z. Szabó. Holomorphic disks and topological invariants for closed three-manifolds. Ann. Math. 159 (2004), 1027–1158.
- [26] P. Ozsváth, Z. Szabó. Holomorphic disks, link invariants and the multi-variable Alexander polynomial. Algebr. Geom. Topol. 8 (2008), no. 2, 615–692.
- [27] P. Ozsváth, Z. Szabó. Holomorphic triangles and invariants for smooth four-manifolds. Adv. Math. 202 (2006), no. 2, 326–400.
- [28] P. Ozsváth, Z. Szabó. Lectures on Heegaard Floer homology. Clay Math. Proc. 5 (2007), 29–70.
- [29] P. Ozsváth, Z. Szabó. On the Heegaard Floer homology of branched double-covers. Adv. Math. 194 (2005), no. 1, 1–33.
- [30] P. Ozsváth, Z. Szabó. On the skein exact sequence for knot Floer homology. arXiv:0707.1165.
- [31] I. Zemke. Graph cobordisms and Heegaard Floer homology. arXiv:1512.01184.
- [32] I. Zemke. Link cobordisms and functoriality in link Floer homology. J. Topol. 12 (2019), no. 1, 94–220.
- [33] I. Zemke. Link cobordisms and absolute gradings on link Floer homology. Quantum Topol. 10 (2019), no. 2, 207–323.
- [34] I. Zemke. Quasistabilization and basepoint moving maps in link Floer homology. Algebraic & Geometric Topology 17 (2017), no. 6, 3461 – 3518.