[bib,biblist]nametitledelim,
A link invariant from higher-dimensional Heegaard Floer homology
Abstract.
We define a higher-dimensional analogue of symplectic Khovanov homology. Consider the standard Lefschetz fibration of a -dimensional Milnor fiber of the singularity. We represent a link by a -strand braid, which is expressed as an element of the symplectic mapping class group . We then apply the higher-dimensional Heegaard Floer homology machinery to the pair , where is a collection of unstable manifolds of which are Lagrangian spheres. We prove its invariance under arc slides and Markov stabilizations, which shows that it is a link invariant. This work constitutes part of the author’s PhD thesis.
Key words and phrases:
Higher-dimensional Heegaard Floer homology, Khovanov homology2010 Mathematics Subject Classification:
Primary 53D40; Secondary 57M27.1. Introduction
Many powerful Floer-theoretic invariants of knots and links have emerged over the past two decades. These include Heegaard Floer homology [OS04] and knot Floer homology in dimension 1; symplectic Khovanov homology [SS06] and knot contact homology [EENS13] in dimension 2. Here when we say “dimension ”, we are taking the ambient symplectic manifold to be -dimensional and the Lagrangian submanifolds (if we are talking about Lagrangian intersection Floer thoeries) to be -dimensional. In [Man07] Manolescu also used quiver varieties to define a higher-dimensional analogue of -homologies.
Along similar lines, the aim of this paper is to construct a link invariant using higher-dimensional Heegaard Floer homology, which is defined in [CHT20] as a higher-dimensional analogue of Heegaard Floer homology in a cylindrical setting.
More specifically, in dimension 1, Lipschitz [Lip06] proved the equivalence between Ozsváth and Szabó’s Heegaard Floer homology [OS04] and its cylindrical analogue. In dimension 2, Mak and Smith [MS19] established the equivalence of symplectic Khovanov homology and its cylindrical interpretation. Colin, Honda, and Tian [CHT20] then defined a higher-dimensional analogue of cylindrical Heegaard Floer homology, which helps place the cylindrical symplectic Khovanov homology in a more general framework. In this paper, the ambient manifold is a -dimensional Milnor fiber of the -singularity , extending the case of considered in [CHT20]. Given a link, we consider its -strand braid representation , which corresponds to an element of the symplectic mapping class group . There is a natural collection of Lagrangian spheres by the matching cycle construction between pairs of critical points of . We then apply the higher-dimensional Heegaard Floer homology machinery to the pair to define the link invariant and denote the homology group by .
Though cylindrical versions of Heegaard Floer theories are more convenient for visualizing pseudoholomorphic curves, the original theories defined in the symmetric products have their advantages: In dimension 1, Perutz [Per08] proved that Lagrangians in related by a handle slide are in fact Hamiltonian isotopic for some specific symplectic form, which directly implies the handle slide invariance property without curve counting techniques in [OS04]. In dimension 2, Seidel and Smith [SS06] considered nilpotent slices instead of , which was shown to be a subset of by Manolescu [Man06]. Inside the nilpotent slice, matching cycles as Lagrangians related by arc slides are also Hamiltonian isotopic, which is not obvious in the cylindrical formulation. Mak and Smith [MS19] then showed that the cylindrical version is equivalent to the original symplectic Khovanov homology in nilpotent slices.
However, in higher dimensions, the cylindrical symplectic Khovanov homology with vanishing cycles -tuples of does not have its “original” version; at this moment we do not know how to put the theory inside a nilpotent slice setting.
Another problem is that Hilbert schemes of points in higher dimensions are not smooth, so we need new ways to resolve the singularities along the diagonal of .
One possible solution is to restrict to some smooth stratum of .
For example, the subset of subschemes where each support point is of length at most 3 (at most triple point) is smooth.
However we will not continue the discussion in this paper further.
It would be interesting to study this problem in future.
Therefore, we will adopt the curve counting approach as in [OS04] and [CHT20] to prove the invariance under arc slides and Markov stabilizations.
In Section 2, we begin with a brief review of Section 9 of [CHT20], providing necessary notations, definitions and prerequisite theorems. Then we state the main result. We use a subsection to explain the Morse flow tree theory and its relation to pseudoholomorphic curves originated from [FO97], which is crucial in our proof.
In Section 3, we show the arc slide invariance by counting pseudoholomorphic curves with certain boundary Lagrangians. The idea is to stretch the curve into several parts so that each one is easy to count by elementary model calculation.
In Section 4, we perform a model calculation of pseudoholomorphic quadrilaterals, which will be used in Section 5 and Section 6 for several times.
In Section 5, we translate the Markov stabilization into the gluing of pseudoholomorphic curves, which we count by Morse flow tree arguments instead.
In Section 6, we compute for unknots, Hopf links and trefoils.
Acknowledgements. I would like to thank Ko Honda for countless discussions and introducing this project to me. I also thank Yin Tian for numerous ideas and suggestions, and thank Eilon Reisin-Tzur for his patient revision. I am partially supported by China Postdoctoral Science Foundation 2023T160002.
2. Definitions and main results
2.1. Higher-dimensional analogue of symplectic Khovanov homology
This subsection works as a review of [CHT20], so most proofs and details are omitted.
Let . We consider the standard -dimensional Lefschetz fibration
for a Milnor fiber of the singularity, where the regular fiber is . There are critical values , where , and , . Let
be the restriction of to . For , connect and by straight arcs . Then the matching cycles over are Lagrangian spheres. Let , , and be “half” of , , and , i.e., their restrictions to .
Given a -strand braid , let be the monodromy on which descends to and let be the extension of to by identity.
In this paper we always do cohomology. The variant of the symplectic Khovanov cochain complex is defined as the higher-dimensional Heegaard Floer cochain complex, in the sense of [Lip06] and [CHT20], denoted by . Specifically, a -tuple of intersection points of and is a -tuple where and is some permutation of . Then is the free -module generated by all such -tuples , where the coefficient ring is discussed below.
To define the differential, let be a surface with boundary of Euler characteristic and be the surface with boundary punctures. We start with the split almost complex structure on , and apply a perturbation to achieve transversality, denoted by .
For , let be the moduli space of satisfying
-
(1)
;
-
(2)
;
-
(3)
As tends to (resp. ), tends to (resp. ), where are the projections of to and .
The differential is then defined as
(2.1) |
We write for the cohomology group .
The coefficient ring (power series in and polynomial in ) keeps track of the relative homology class and Euler characteristic of the domain, where
The following lemmas justify the use of coefficient for and for :
Lemma 2.1.
For fixed , and , is finite.
Lemma 2.2.
Suppose is of dimension . For , , where is the number of connected components of ; for , .
Proof.
Denote , , and for the inclusion.
If , and are trivial. By the long exact sequence for relative singular homology, . Since there are critical points of , is generated by matching cycles over arcs connecting pairs of critical values in . is simply generated by collections and . Therefore, is generated by arcs quotient by collapsing arcs corresponding to and . The number of remaining nontrivial arcs is , where is the number of connected components of the braid .
If , and are trivial, so by the relative homology exact sequence. ∎
The following lemma computes the Fredholm index, which is a generalization from [CGH12] and the proof is omitted:
Lemma 2.3.
The Fredholm index of is
where is the Maslov index.
The Floer homology group is well-defined and now we state our main result, which is a higher-dimensional version of Theorem 1.2.1 of [CHT20]:
Theorem 2.4.
For or , is a link invariant, i.e., it is independent of the choice of arcs and Lagrangian thimbles , and is invariant under Markov stabilizations.
Remark 2.6.
The computation of for simple links in Section 6 gives results highly similar to Khovanov homology [Kho00, Wil08]. We will not discuss their relation further in this paper. Instead, we leave it as a conjecture:
Conjecture 2.7.
2.2. Moduli space of Morse gradient trees
If two closed Lagrangian submanifolds of a symplectic manifold are -close to each other, and moreover they are exact Lagrangian isotopic, then by the Lagrangian neighbourhood theorem, one can view as the graph , where . In this case, the count of pseudoholomorphic disks bounding can be reduced to the count of Morse gradient trajectories on , which is much more convenient to compute. More generally, if we have exact Lagrangian isotopic , where for , then we count gradient trees, to be explained below. This alternative method of counting appears several times in Section 3 and 5, so we make the statement precise here. The main reference is [FO97], which generalizes Floer’s correspondence between gradient trajectories and pseudoholomorphic strips [Flo88]. There is also a simpler proof of [FO97] by Iacovino [Iac08], which considers the perturbation of Floer equations.

Definition 2.8.
A ribbon tree is a directed tree which satisfies:
-
(1)
has no degree 2 vertex.
-
(2)
For each internal vertex , there is a cyclic order of the edges attached to , i.e., we have a labeling bijection , where is the set of edges attached to and is the degree of such that there is a unique outgoing edge and it is labeled 0. Therefore, there are two labels for each internal edge coming from the two labeling functions corresponding to the endpoints of the edge and one label for each external edge.
-
(3)
There is a cyclic order of external vertices of such that the root has label 0, i.e., we have a bijection where is the set of external vertices and the one adjacent to the unique outgoing edge is labeled by 0.
-
(4)
There exists an embedding so that at each internal vertex , are in counterclockwise order around . Also, and are in counterclockwise order on . Note that the way we embed is not important.
In the definition univalent vertices are called external and the others are internal; edges adjacent to external vertices are called external and the others are internal.
The edges of will represent gradient flows. Specifically, let be a Riemannian manifold and be -functions on so that is Morse for each (where ). We define the moduli space of gradient trees to consist of pairs where is a ribbon tree and is a continuous function satisfying:
-
(1)
, where is a critical point of .
-
(2)
External edges are identified with except the outgoing one which is identified with ; each internal edge is identified with where is a length function defined on internal edges. Now takes the above intervals to . For each edge of ,
where and are defined with respect to the direction of , i.e. if one looks in the positive direction of on , then (resp. ) is the component of on the left (resp. right) side of .
Figure 1 shows an element , together with an embedding of in . Note that we assign each function to a corresponding region of .
Lemma 2.9 ([FO97]).
Fixing a generic choice of , is a smooth manifold of dimension
(2.2) |
where is the number of boundary vertices and is the dimension of the manifold.
Next we define the moduli space of pseudoholomorphic disks bounding exact Lagrangian submanifolds in . Let be a Riemannian manifold. Let be a generic collection of functions on , and let be the graphs of their differentials. We associate each critical point of with where is small.
We then fix a canonical almost complex structure on associated to the metric on such that
-
(1)
is compatible with the canonical symplectic form on .
-
(2)
maps vertical tangent vectors to horizontal tangent vectors of with respect to .
-
(3)
On the zero section of , for , let .
Definition 2.10.
Let consist of pairs where is the domain with marked points arranged in counterclockwise order on and satisfies
-
(1)
,
-
(2)
,
-
(3)
,
where is the shortest counterclockwise arc between and on . Here we are identifying pairs and which are related by an isomorphism of the domain.
The following theorem relates Morse gradient trees on to pseudoholomorphic disks on :
Theorem 2.11 ([FO97]).
For sufficiently small, there is an oriented diffeomorphism .
Roughly speaking, for each gradient tree , we can construct a pseudoholomorphic curve near inside by some gluing techniques. We should note that the theorem depends on a specific choice of almost complex structure .

The following example counts pseudoholomorphic triangles with 3 Lagrangian boundaries, which we will meet again in Section 3.
Example 2.12.
We give an example of a gradient tree with 3 vertices here. Consider functions on so that the domain looks like the left side of Figure 2. We can perturb generically so that is a source (top generator) of , is a sink (bottom generator) of and is a sink (bottom generator) of . The middle of Figure 2 shows a possible perturbation when for illustration. Sources and sinks are denoted by blue arrows. The right side of Figure 2 shows what happens on : is the unique top generator of on , so the flows from form a -family and pass through all the points of except two critical points. is the bottom generator of , so the flow from of is of length 0. Therefore the flow from should pass and there is a unique such flow line. Similarly, there is a unique flow line from to of . To conclude, there is a unique gradient tree. Thus there is a unique pseudoholomorphic disk bounded by the Lagrangians for small by Theorem 2.11.
3. Invariance under arc slides
Given from Section 2.1, consider the arc slide of over : Let be the new set of arcs as in Figure 3. Let be the new tuple of Lagrangians over . For , let (resp. ) be the intersection point of and that lies over (resp. ). Denote and .

The purpose of this section is to prove:
Theorem 3.1.
and are quasi-isomorphic.
Proof.
It suffices to show the quasi-isomorphism between and . One can show the quasi-isomorphism between and similarly.
We define the cochain map by the composition map of the -relation:
where is the product map
Similarly we define the cochain map going back by
The following lemma justifies that , are well-defined:
Lemma 3.2.
For or , is a cocycle in and is a cocycle in .
Proof of Lemma 3.2: For simplicity assume . By Lemma 2.3 and a Maslov index calculation,
(3.1) | ||||
(3.2) | ||||
(3.3) |
where for example, denotes the Fredholm index of curves such that
-
(1)
,
-
(2)
,
-
(3)
As tends to (resp. ), tends to (resp. ), where and are the projections of to and .
We explain (3.1)-(3.3) now. First, the Maslov index of a closed path over and is by definition, which implies (3.2). To get (3.3), observe that the projection to of the curve from to is over the region surrounded by and , which contains two critical values of the Lefschetz fibration. The region surrounded by and can be viewed as the outcome after applying Lagrangian surgery twice on a trivial strip to incorporate the two critical values. By Theorem 55.5 of [FOOO09], each Lagrangian surgery increases the Fredholm index by . Therefore, by comparing with (3.2). Finally, (3.1) follows by adding the Maslov index terms of (3.2) and (3.3).
For , the Fredholm index does not depend on : and .
For , observe that in all 3 cases the domain has 2 punctures and thus is even. In particular, is even and ; If , then , which is not a integer for ; If , then , which is also not a integer for .
Therefore, for or , there is no index 1 curve from , i.e., is a cocycle in . The case of is similar. ∎
Remark 3.3.
Lemma 3.2 does not deal with the case of , where the index could be 1 for some . Therefore, to show is a cocycle, we still need to understand the index 1 pseudoholomorphic curves from to , whose projection to is the thin strip surrounded by and , and also the curves from to , whose projection to is the fat strip surrounded by and .
It remains to show that induces identity on cohomology (with some nonzero coefficient). The composition of and is given by the left-hand side of Figure 4, which is viewed as a degeneration of a family of curves. The right-hand side of Figure 4 shows another degeneration. Observe that the right-hand part of the right degeneration is of index 0. In fact the right degeneration corresponds to the identity map times the count of its left-hand part, which is chain homotopic to .

Now we count the left-hand part of the right degeneration of Figure 4. For convenience, assume , i.e., focus on the arc sliding of over . The moduli space contains curves with boundary condition which maps to . There is another restriction on the right degeneration: passes through and , where and . Passing through a generic is a codimension condition. By Lemma 2.3, if and only if or for .
By Theorem 3.5 below, the count of passing a generic is 1 (mod 2). Thus is cochain homotopic to identity with some nonzero coefficient. The case of is similar. ∎
3.1. Half of curve counting
There is a half version of the counting problem, where the base is shown as Figure 5. The base can be extended to by adding cylindrical ends, denoted by . Let and be the cylindrical completion of and . The asymptotic Reeb chords from to form a -family. We then perturb the contact form so that the -family becomes Morse-Bott and let be the longer and shorter asymptotic Reeb chords. Let . The problem is to count , which is over the cylindrical extension of the region surrounded by .

Theorem 3.4.
mod for generic .
Proof.
For , let , where is the cylindrical extension of and is the unit disk in .
Recall that is the projection. The map has degree 2 (resp. 1) over region A (resp. B) of Figure 5, which are connected components of with extension. There are two possible branching behaviors: Type has a branch point that maps to the interior of region A; Type is more obscure which has two switch points on the boundary of region A instead of a branch point. Denote Type for the case on and Type for on . Choose as the point closer to the puncture (resp. ) on that maps to (resp. ).
There is another codimension-1 constraint. Denote the preimage of under as , and we will often leave this notation out if there is no ambiguity. A possible arrangement of points on is shown as Figure 6. Since the projection of to is a double branched cover and under this projection, , , are mapped to , respectively, the deck transformation requires the involution:
(3.4) |

The counting strategy uses the fact that the mod 2 count does not depend on , which allows us to stretch the curve by letting tend to some limit:
-
(1)
Let ;
-
(2)
Let the width between ;
-
(3)
Choose such that .
1. Denote . We first observe that as , either or . If this is not true, then by Gromov compactness there exists a limiting curve which contradicts the involution condition 3.4. We refer the reader to [CHT20] for details.
2. Write for the curves satisfying or . The case of is shown as the left of Figure 7, where the limiting 2-level curve contains . The case of is similar. We push off the bottom singular point a little so that and have clean -intersection, which does not change the moduli space of by [FOOO09].

The region of contains a singular point of a Lefschetz fibration, which can be viewed as the outcome after a Lagrangian surgery from the trivial region without singular points. By Theorem 55.5 of [FOOO09], the moduli space of is diffeomorphic to . Therefore the evaluation map of at sweeps inside the -family of Reeb chords, of which the homology class vanishes. The result is mod 2.
3. The remaining case is the curve satisfying and . The stretched limiting curves are shown as the middle and the right of Figure 7. We expect that such curves exist and contribute 1 (mod 2) to .
First we treat Type . is uniquely determined since there is a unique gradient trajectory from to passing through . Denote the bottom-left Reeb chord of by . We show that the pseudoholomorphic triangle exists uniquely: The Lefschetz fibration around is the trivial one . We can perturb the Lagrangians in the fiber direction. The case of is shown as Figure 8, where there exists a unique pseudoholomorphic triangle.

More generally, the case of is done by the gradient tree argument. Put the perturbed Lagrangians as Figure 8 on and extend to , where we view as a equator. After a further small perturbation, we consider the moduli space of Morse gradient trees with 3 vertices . Viewing all vertices as sources of gradient flow, we check that , .
By Lemma 2.9, the dimension of this moduli space is 0 and it contains a unique gradient tree which corresponds to a single pseudoholomorphic triangle by Theorem 2.11. Note that this is almost the same as Example 2.12.
It remains to consider , which is a pseudoholomorphic disk with a slit and a singular point inside. The singular point can be viewed as coming from a Lagrangian surgery [FOOO09], which increases the Fredholm index by . Now the moduli space of passing through a generic is of dimension . Consider the evaluation map of on its top-left end, where the image lies in a -family of Reeb chords. The gluing condition on both ends says should take the value of . All we need is that intersects at a unique point. We will show in Type , sweeps half of and the other half is dealt with by Type . This is done by a model calculation of explicit pseudoholomorphic curves in Step 3’ below.
3’. A model calculation. All notations are limited to this step.
We replace the base of in Type by a standard one, i.e., the unit disk in with a slit . The Lefschetz fibration over the unit disk is
(3.5) |
with a critical value at . It is however more convenient to think of the case first, and the Lefschetz fibration is
(3.6) |
where . Let be the Clifford torus over and let be the Lagrangian thimble over . is a clean -intersection over .
We consider curves with boundary on : let (2 stands for ) be the moduli space of holomorphic disks
with standard complex structure satisfying
-
(1)
and ;
-
(2)
has degree 1 over and degree 0 otherwise;
-
(3)
.

Condition (3) is essentially the same as that passes through in Step 3. It is not hard to see that is homeomorphic to a line segment where consists of two curves and . Figure 9 gives a schematic description of , from the top row of to the bottom row of , where the right-hand column changes the coordinates to .
We then consider the evaluation map. The top-left end of in Type is translated to for small . Define as the projection of the intersection between and , which is shown by red dots in the column of Figure 9. Clearly is a homeomorphism between and .
For general , define similar to , where in condition (1) are still the Lagrangian vanishing cycles over the unit circle and . Condition (3) is modified to
-
(3’)
.
The moduli space in coordinate is viewed as the slice of that restricts to on . Observe that for , the coordinates are symmetric. Thus we can recover from by a symmetric rotation of -coordinate. Each corresponds to a -family of curves in :
where and for . Figure 10 shows recovered from the first row of Figure 9.

The new evaluation map is defined as the -tuple of coordinates which projects to . Therefore, is homeomorphic to . One can check that is a homeomorphism to its image, which is half of the vanishing cycle over .
In case is not regular, we apply small perturbation of . One can show that for any , mod . Therefore the argument of Step 3 still works. This finishes Step 3’.
3.2. Curve counting
The goal of this section is to count the full version of pseudoholomorphic annuli from to over the region in Figure 3 and prove the following theorem:
Theorem 3.5.
mod for generic .
Proof. The strategy is the same as the proof of Theorem 3.4: We stretch the curve into several levels by choosing some extreme , and then use the restriction of domain involution and gluing conditions to find a unique (mod ) curve.
We closely follow the proof of Theorem 9.3.7 of [CHT20] and some details are omitted. As before, we write for an element in and let be its projection to .
The main idea is to stretch the base in direction as Figure 11: Let and , and . The region bounded by is split into 3 parts: , and . The mod count is independent of . We choose close to , close to and the thin strip between with width .

The curve has degree 2 over and degree 1 over . The types of branching behaviors are denoted by , , , , , . Take for example: 0 means the number of interior branch points is 0; means one pair of switch points is over ; means one pair of switch points is over . Denote (if exist) interior branch points by and switch points by . We assume , and .
. Suppose . Take a sequence of so that with . In the limit , the thin strip tends to a slit, the limiting curve splits into , where is a gradient trajectory from and is a gradient trajectory to , and is a pseudoholomorphic annulus. Figure 12 describes the limiting procedure in the case of Type . In fact we will show that Type is the only nontrivial case.

. We claim that the limiting slit is long enough, i.e., for and ,
(3.7) |
(3.8) |
which are the two endpoints of the slit.
If (3.8) is not true, i.e., , then the part of in region has no slit, which can be viewed as the outcome of a Lagrangian surgery on a trivial pseudoholomorphic disk. Similar to Step 2 in the proof of Theorem 3.4, the moduli space of pseudoholomorphic disks passing through a generic is diffeomorphic to , of which the evaluation map at the cylindrical end has a -intersection with the -family of Reeb chords. The evaluation map vanishes at homology level, which contributes 0 (mod 2) to the curve count. The argument for (3.7) is similar.
. We claim that for and small, if (3.7) and (3.8) hold, the mod contribution of Type is 0. The reason is that if one considers the involution condition
(3.9) |
for a pseudoholomorphic annulus, there is a constraint on the position of on the slit in . One can refer to [CHT20] for detailed discussion that all but Type contradict with (3.9).
. It remains to consider Type . Assuming (3.7) and (3.8) are satisfied, we claim that the contribution of Type is 1 (mod 2).
Although there are other possible arrangements of on the slit, we just consider the case in Figure 12 for illustration. As shown in Figure 13, is the gluing of two regions: with and with , which can be viewed as two pseudoholomorphic disks similar to Step 3’ of the previous section. The conditions that is close to and that sits on the slit will be translated to an explicit model calculation in Step 4’ below. For suitable choices of and , we will show that the space of two disks with is homeomorphic to a line segment . Let be the set of glued from the -family of .

The involution condition (3.9) determines a unique curve inside . In conclusion, mod . We have proved Theorem 3.5 modulo the model calculation below:
4’. A model calculation. We use similar notations as in Step 3’ of the previous section. Define for the space of over the left side of Figure 13 and for those of over the right side. Both are viewed as maps from to over the unit circle with one slit . Consider the evaluation maps
which are defined below, corresponding to in Figure 13. The gluing condition is .
As usual, we first consider the case of with Lefschetz fibration (3.6).
The moduli space is defined as holomorphic disks passing through over . Note that is the same as in Step 3’ of the previous section. Let . Define the evaluation map as
where is the -coordinate of the point that projects to . Figure 14 gives a schematic description of , homeomorphic to a line segment, and its evaluation maps denoted by red and violet dots. We also show the maps in coordinates with .

Since in coordinate , define for . Then , the moduli space for , is viewed as the slice of that restricts to on . Now is a symmetric rotation of which contains
where , . Thus is homeomorphic to .
Next we consider of and first let . We put the constraint that passes through for some , which corresponds to sitting on the slit in Figure 13. Let be the -coordinate of the point that projects to and define the map
where and are switched because we want to identify of with of .

Observe that is of dimension 2. Figure 15 describes some of the curves inside . in equals in . Thus for , if is viewed as the slice of with , then contains curves of
where , .
For , we have the following observation:
Claim 3.6.
For and , the evaluation map of and with images in is described in Figure 16. is the blue line segment and is the 2-dimensional pink region. Their intersection is a line segment .

For general , we have the same result:
Lemma 3.7.
For and , the intersection between and is still .
Proof of Lemma 3.7. Suppose and satisfy where the 3rd to -th coordinates are not all zero. Denote and .
Observe that and then . From and in Figure 15 we see that (in coordinate ) must lie in of Figure 16. For such curves (as () in Figure 15), and . Since and , the consequence is that , which is a contradiction. ∎
Now we go back to the curve counting problem. We want to pick a single curve from the -family of and then glue it to get a unique for each .
First we show the position of determines uniquely in : Consider the slit in of Figure 12. In the limit , so . If we fix , then passes through for some , corresponding to fixing a hypersurface in , whose intersection with is the dotted arc in Figure 16. The dotted arc intersects the blue line at a single point, which determines . Moreover, the length of the slit in and are determined and thus and are fixed. Finally is fixed by the involution of .
Consider then for large . From the previous paragraph will fix a unique in . By Implicit Function Theorem, it will fix a unique as well, which is close to . Then observe that the distance between and is a monotone function of : As increases, the slit gets longer, moves left and moves right. Therefore leaves and approaches on . The involution of determines a unique and thus a unique .
This finishes the proof of Theorem 3.5. ∎
4. A model calculation of quadrilaterals
We make a model calculation which will be used in Section 5 and 6. Consider the trivial fibration and Lagrangian submanifolds , , , , where is the zero section of . We further modify to by a Hamiltonian perturbation in the fiber direction so that they intersect transversely. Specifically, we choose the restriction of Euclidean metric on and identify with . Choose Morse functions on each with 2 critical points and all of the critical points are disjoint (as the right of Figure 17). We can then rescale these Morse functions so that the difference of each pair is still Morse with 2 critical points:

Lemma 4.1.
For small enough , the difference of each pair of functions in is Morse with 2 critical points.
Proof. For small enough , is a small perturbation of . Since Morse condition is -stable, is still Morse with 2 critical points. By a simple induction the proof is finished. ∎
Denote , the gradients of which correspond to the fiber projection of . Let over be the top and bottom critical points of , .
Now we compute the differentials of , which is generated by 8 elements and , where denotes a check or hat.
Lemma 4.2.

Proof.
Let be a pseudoholomorphic disk with positive ends and negative ends . Suppose the complex structure is split, then its projection to is a degree 1 map over , which fixes the cross ratio of the 4 punctures on .
Then we consider the projection of to the fiber direction , denoted by . By the construction above are graphical near , with respect to Morse functions , . The domain of the Morse moduli space is shown in Figure 19, where inner edges are ignored and arrows denote the direction of .

Viewing all boundary vertices as sources of gradient flow, observe that is of Morse index if and 0 if ; is of Morse index if and if . By Lemma 2.9, we can further perturb , such that all Morse gradient trees we consider are transversely cut out, and
(4.1) |
Therefore, the Morse moduli space with respect to the arrows in Figure 18 is of and the case of gradient tree on from to is shown in Figure 20.

By Theorem 2.11, the moduli space of pseudoholomorphic curves is diffeomorphic to the Morse moduli space, so we can think of gradient trees instead of pseudoholomorphic disks. Taking the base direction into consideration, one checks that
(4.2) |
For example, we still consider the case of Figure 20: The two gradient trees are parametrized by the length of their inner edges. As the inner length tends to zero, the left and right gradient trees tend to the same one. As the inner edge of the left one tends to the bottom generator of , its length tends to infinity and approaches . Similarly, as the inner edge of the right one tends to the top generator of , its length tends to infinity and approaches . Since the cross ratio on the domain is fixed by the base direction, the result is that the algebraic count of is one. This verifies and the arrows in Figure 18.
5. Invariance under Markov stabilization
A Markov stabilization is shown in Figure 21: is a -strand braid which intersects along . On the base , is viewed as an element of , which restricts to identity near . Without loss of generality, we construct a positive Markov stabilization between and : Let be an arc from to which is disjoint from other , perform a positive half twist along , then we get a -strand braid given by .
Now we consider the fiber and Lagrangians. Let be the standard Lefschetz fibration with regular fiber and critical values and be its restriction to . Let denote the Lagrangian thimble over . Let be an element of which descends to and be its extension to by identity. Finally, let be the Dehn twist along the Lagrangian sphere over .


The proof of invariance under Markov stabilization is the same as Theorem 9.4.2 of [CHT20], and we briefly restate its proof here:
Theorem 5.1.
and are isomorphic cochain complexes for specific choices of almost complex structure and and after a Hamiltonian isotopy.
Proof.
We directly construct a homomorphism and show it is a cochain isomorphism. The notations are as in Figure 22.
Consider the -tuples in : It may or may not contain . Similarly, the -tuples in may contain either or . In fact there is a linear isomorphism:
For convenience of gluing below, we put and in a position so that they go over the same arc near . Let and be small. The projections and are written as and . The main requirement is that be a common neck, for example, set . Refer to Figure 22 for details.
We compare the differential of and . If goes from to , then is in bijection with that goes from to since the strip from to is trivial. Similarly, that goes from to is in bijection with that goes from to . There are no curves from to and no curves from to . The nontrivial case is that, if goes from to , then it is in bijection with that goes from to , where comes from by replacing the end containing by the shaded region in Figure 22. The bijection comes from Lemma 4.2 which says that the curve over the shaded region has algebraic count 1. The gluing details are omitted. ∎
6. Examples
In this section we consider some simple links and compute their cohomology groups in the sense of Theorem 2.4, with coefficient ring when and when , where is of degree . We assume that is of characteristic 2 for simplicity.
6.1. Unknots
Figure 23 shows the 2-strand braid representation of an unknot. In the Morse-Bott family of Reeb chords, are viewed as longer Reeb chords (top generators) and are viewed as shorter Reeb chords (bottom generators). The only possible differential in counts quadrilaterals with at positive ends and at negative ends. The projection of to has degree 1 over the region bounded by the loop and degree 0 over its complement.

Proposition 6.1.
is freely generated by and , where the difference of grading between these two generators is (mod ):
generators | grading |
---|---|
Proof.
Similar to the proof of Lemma 4.2, we check that the pseudoholomorphic disk with at positive ends and at negative ends is of Fredholm index and Maslov index . Therefore, and are both cocycles and hence generators of . ∎
6.2. Hopf links
We then consider the 2-strand braid representation of a left-handed Hopf link as the left side of Figure 24. ‘’, ‘’ and ‘’ denote the corresponding regions on the base . For example, we use ‘’ to represent the union of region and .

Lemma 6.2.
contains the following (mod 2) differential relations:
(6.1) | ||||
(6.2) | ||||
(6.3) | ||||
(6.4) |
Proof.
Note that should be viewed as top generators at positive ends or as bottom generators at negative ends. Specifically they give no constraints on the Morse-Bott family.
All nontrivial index 1 curves with count 1 (mod 2) are listed as follows, where we label the regions with positive weights after projection to the base :
regions | ends of generators |
---|---|
To see this, after projection to the base, the domain with positive weights is bounded by one of the following loops:
- (1)
-
(2)
. This corresponds to region ‘’, where is a quadrilateral with at positive ends and or at negative ends. This is similar to (1). Therefore, and we get (6.2).
- (3)
-
(4)
. This corresponds to region ‘’, where is a quadrilateral with at positive ends and at negative ends. are viewed as top generators and there are two critical points inside the domain. We check that for at negative ends; for and at negative ends; for at negative ends. Therefore there is no such with index 1.
∎
Corollary 6.3.
If we set , then is freely generated by the following generators with the corresponding relative grading (mod ):
generators | grading |
---|---|
The computation for the right-handed Hopf link is similar. As shown in Figure 24, the braid representation of the right-handed Hopf link is the mirror of the left-handed one. Note that there is a bijection of curves between these two Hopf links: Each curve in the left-handed moduli space corresponds to a curve in the right-handed one, where the positive and negative ends are exchanged, as well as the checks and hats. As a consequence, the grading by Maslov indices is also reversed. Specifically, the right-handed Hopf link satisfies:
Lemma 6.4.
contains the following (mod 2) differential relations:
Corollary 6.5.
If we set , then is freely generated by the following generators with the corresponding relative grading (mod ):
generators | grading |
---|---|
6.3. Trefoils
The left side of Figure 25 shows the 2-strand braid representation of a left-handed trefoil, where to denote the corresponding regions on the base.

The curve counting for trefoils is more interesting than unknots and Hopf links. We first do a model calculation. Recall the notations in Step 3’ of the proof of Theorem 3.4. We consider curves with boundary on : Let (2 stands for ) be the moduli space of holomorphic disks
with standard complex structure satisfying
-
(1)
and ;
-
(2)
has degree 1 over and degree 0 otherwise;
-
(3)
.
It is easy to see that is homeomorphic to a line segment where consists of two curves and . Figure 26 gives a schematic description of in the coordinates . For a given , define the evaluation map
as the projection of the intersection between and , which is shown by violet dots in the column of Figure 26.

Note that in -coordinates equals in -coordinates. For general , define similar to , where in condition (1) are still the Lagrangian vanishing cycles over the unit circle and . Condition (3) is modified to
-
(3’)
.
The moduli space in coordinate is viewed as the slice of that restricts to on . Observe that for , the coordinates are symmetric. Thus we can recover from by a symmetric rotation of -coordinate. Each corresponds to a -family of curves in :
where and for .
The new evaluation map is defined as the -tuple of coordinates which projects to . Therefore, is homeomorphic to . One can check that is a homeomorphism to its image. As increases from to , the ratio of area of increases from 0 to 1.
Lemma 6.6.
contains the following (mod 2) differential relations:
Proof.
All nontrivial index 1 curves with count 1 (mod 2) are listed as follows:
regions | ends of generators |
---|---|
, | |
, | |
, | |
, | |
, | |
First consider the region ‘’. The generators are viewed as point constraints and impose no constraints on the Morse-Bott family. There are two possible slits: going down from or going to the right from . If the slit goes down, we can apply the model of Figure 26, where corresponds to the red dot and corresponds to the violet dot over for some . If are close on the base, then is close to . Therefore, takes almost all of and hence mod . Note that the disparity between and prevents from being close on the fiber direction (with respect to the symplectic parallel transport away from ), which contributes to the complement . However, even if are not close on the base, the count remains the same: As the slit goes to the right from , are getting closer on the fiber direction, and this part takes care of the complement . The case of ‘’, ‘’ and ‘’ are similar.

Next we consider the region ‘’. We stretch the region ‘’ as depicted in Figure 27. The dashed line is viewed as a gluing condition of a point constraint of both sides. There are two different cases:
-
(1)
. On the left side of Figure 27, and are point constraints, which fixes a unique holomorphic disk by similar reasons as case ‘’. Note that the slit going to the left from cannot be too long (across the dashed line): If the slit is long, then it cuts the left part into a trivial quadrilateral and a disk region with a branch point inside. However, the (mod 2) count of a holomorphic disk with a branch point inside is 0.
The left disk then fixes the point constraint of the dashed line. The right part of Figure 27 now has two point constraints: and the dashed line. We then apply the model of Figure 26 again, where corresponds to the red dot and the dashed line corresponds to the violet dot. Similar to the discussion of ‘’, there exists a unique holomorphic disk. Note that the slit going to the left from cannot be too long for the same reason as above. To conclude, the (mod 2) count is 1.
The case of is similar and the (mod 2) count is 1.
-
(2)
. Similar to (1), and fix a unique holomorphic disk on the left side of Figure 27 and gives the dashed line a point constraint. The dashed line and are two point constraints on the right side. If the slit from goes down, we apply the model of Figure 26 where corresponds to the red dot and the dashed line corresponds to the violet dot. Note that sweeps approximately of . One can check that the slit going to the left from also sweeps of (by symplectic parallel transport), and it cancels out with the contribution from . As a result, the (mod 2) count is 0.
The case of is similar and the (mod 2) count is 0.
The remaining cases are similar to the discussion of Hopf links and we omit the details.
∎
Corollary 6.7.
If we set , then is freely generated by the following generators with the corresponding relative grading (mod ):
generators | grading |
---|---|
The computation for the right-handed trefoil is similar and the proof is omitted:
Lemma 6.8.
contains the following (mod 2) differential relations:
Corollary 6.9.
If we set , then is freely generated by the following generators with the corresponding relative grading (mod ):
generators | grading |
---|---|
References
- [CGH12] Vincent Colin, Paolo Ghiggini and Ko Honda “The equivalence of Heegaard Floer homology and embedded contact homology via open book decompositions I”, 2012 arXiv:1208.1074 [math.GT]
- [CHT20] Vincent Colin, Ko Honda and Yin Tian “Applications of higher-dimensional Heegaard Floer homology to contact topology”, 2020 arXiv:2006.05701 [math.SG]
- [EENS13] Tobias Ekholm, John B Etnyre, Lenhard Ng and Michael G Sullivan “Knot contact homology” In Geometry & Topology 17.2 Mathematical Sciences Publishers, 2013, pp. 975–1112
- [Flo88] Andreas Floer “Morse theory for Lagrangian intersections” In J. Differential Geom. 28.3, 1988, pp. 513–547 URL: http://projecteuclid.org/euclid.jdg/1214442477
- [FO97] Kenji Fukaya and Yong-Geun Oh “Zero-loop open strings in the cotangent bundle and Morse homotopy” In Asian J. Math. 1.1, 1997, pp. 96–180 DOI: 10.4310/AJM.1997.v1.n1.a5
- [FOOO09] Kenji Fukaya, Yong-Geun Oh, Hiroshi Ohta and Kaoru Ono “Lagrangian intersection Floer theory: anomaly and obstruction. Part II” 46, AMS/IP Studies in Advanced Mathematics American Mathematical Society, Providence, RI, 2009, pp. i–xii and 397–805 DOI: 10.1090/crmp/049/07
- [Iac08] Vito Iacovino “A simple proof of a theorem of Fukaya and Oh”, 2008 arXiv:0812.0129 [math.SG]
- [Kho00] Mikhail Khovanov “A categorification of the Jones polynomial” In Duke Math. J. 101.3, 2000, pp. 359–426 DOI: 10.1215/S0012-7094-00-10131-7
- [Lip06] Robert Lipshitz “A cylindrical reformulation of Heegaard Floer homology” In Geom. Topol. 10, 2006, pp. 955–1096 DOI: 10.2140/gt.2006.10.955
- [Man06] Ciprian Manolescu “Nilpotent slices, Hilbert schemes, and the Jones polynomial” In Duke Math. J. 132.2, 2006, pp. 311–369 DOI: 10.1215/S0012-7094-06-13224-6
- [Man07] Ciprian Manolescu “Link homology theories from symplectic geometry” In Adv. Math. 211.1, 2007, pp. 363–416 DOI: 10.1016/j.aim.2006.09.007
- [MS19] Cheuk Yu Mak and Ivan Smith “Fukaya-Seidel categories of Hilbert schemes and parabolic category ”, 2019 arXiv:1907.07624 [math.SG]
- [OS04] Peter Ozsváth and Zoltán Szabó “Holomorphic disks and topological invariants for closed three-manifolds” In Ann. of Math. (2) 159.3, 2004, pp. 1027–1158 DOI: 10.4007/annals.2004.159.1027
- [Per08] Timothy Perutz “Hamiltonian handleslides for Heegaard Floer homology”, 2008 arXiv:0801.0564 [math.SG]
- [SS06] Paul Seidel and Ivan Smith “A link invariant from the symplectic geometry of nilpotent slices” In Duke Math. J. 134.3, 2006, pp. 453–514 DOI: 10.1215/S0012-7094-06-13432-4
- [Wil08] Brandon Williams “Computations in Khovanov Homology” In University of Illinois at Chicago, Chicago, Illinois, 2008