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

Crossing change maps in filtered grid homology

Matthew Kendall
Abstract.

We extend the crossing-change maps between grid complexes, defined by OzsvΓ‘th–Szabó–Stipsicz, to filtered grid complexes and give a combinatorial formulation of the Alishahi–Eftekhary 𝔩​(K)\mathfrak{l}(K) invariant.

1. Introduction

Grid homology is a combinatorial theory for knots and links first developed by Manolescu, OzsvΓ‘th, and Sarkar [MOS09], inspired by the ideas of Sarkar and Wang [SW10]. It is isomorphic to knot Floer homology, an invariant for knots in three-manifolds discovered in 2003 by OzsvΓ‘th and SzabΓ³ [OS04] and independently by Rasmussen [Ras03]. Grid homology can be used to prove the Milnor conjecture [Mil68], which states that the unknotting number of the (p,q)(p,q) torus knot is (pβˆ’1)​(qβˆ’1)2\frac{(p-1)(q-1)}{2}. This conjecture was first verified by Kronheimer and Mrowka [KM93] in 1993 using smooth four-manifold topology. Inspired by the proof of Sarkar [Sar11], OzsvΓ‘th, SzabΓ³, and Stipsicz in their book [OSS15] use grid homology to define the Ο„\tau invariant and compute Ο„\tau for torus knots. They use a definition of Ο„\tau relies on pairs of maps of grid complexes whose underlying knots differ by a crossing change. The goal of this note is to show that these crossing change maps by OzsvΓ‘th–Stipsicz–SzabΓ³ [OSS15] can be generalized to filtered grid complexes and used to give a combinatorial formulation of the Alishahi–Eftekhary 𝔩​(K)\mathfrak{l}(K) invariant [AE20], compare also [Ras03].

To a grid diagram 𝔾\mathbb{G} of a knot KβŠ‚S3K\subset S^{3}, the grid complex G​Cβˆ’β€‹(𝔾)GC^{-}(\mathbb{G}) is a chain complex over the polynomial ring 𝔽​[V1,…,Vn]\mathbb{F}[V_{1},\ldots,V_{n}], where the field 𝔽=β„€/2​℀\mathbb{F}=\mathbb{Z}/2\mathbb{Z}. It is equipped with a Maslov and Alexander bigrading. Similarly, the filtered grid complex π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}) is a chain complex over 𝔽​[V1,…,Vn]\mathbb{F}[V_{1},\ldots,V_{n}] equipped with the same Maslov grading, but whose Alexander function on grid states induces a filtration rather than a grading. The filtered grid complex is a refinement of the standard grid complex in the sense that G​Cβˆ’β€‹(𝔾)GC^{-}(\mathbb{G}) is the associated graded object of π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}). Its key property is that the filtered chain homotopy type of π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}) depends only on the underlying knot KK.

Let K+K_{+} be a knot with a distinguished positive crossing and Kβˆ’K_{-} be the knot with the crossing changed. Represent K+K_{+} and Kβˆ’K_{-} by the grid diagrams 𝔾+\mathbb{G}_{+} and π”Ύβˆ’\mathbb{G}_{-} that differ by a cross-commutation of columns, see DefinitionΒ 2.2. We first state the proposition that we wish to refine, compare Section 6.2 of [OSS15].

Proposition 1.1.

There exist chain maps

Cβˆ’:G​Cβˆ’β€‹(𝔾+)β†’G​Cβˆ’β€‹(π”Ύβˆ’)C_{-}:GC^{-}(\mathbb{G}_{+})\to GC^{-}(\mathbb{G}_{-}) and C+:G​Cβˆ’β€‹(π”Ύβˆ’)β†’G​Cβˆ’β€‹(𝔾+)C_{+}:GC^{-}(\mathbb{G}_{-})\to GC^{-}(\mathbb{G}_{+})

where Cβˆ’C_{-} and C+C_{+} is homogeneous of degree (0,0)(0,0) and (βˆ’2,βˆ’1)(-2,-1), such that Cβˆ’βˆ˜C+C_{-}\circ C_{+} and C+∘Cβˆ’C_{+}\circ C_{-} are filtered chain homotopy to multiplication by V1V_{1}.

The filtered analog to PropositionΒ 1.1 that we prove is the following:

Theorem 1.2.

There exist filtered chain maps

π’žβˆ’:π’’β€‹π’žβˆ’β€‹(𝔾+)β†’π’’β€‹π’žβˆ’β€‹(π”Ύβˆ’)​andβ€‹π’ž+:π’’β€‹π’žβˆ’β€‹(π”Ύβˆ’)β†’π’’β€‹π’žβˆ’β€‹(𝔾+)\mathcal{C}_{-}:\mathcal{GC}^{-}(\mathbb{G}_{+})\to\mathcal{GC}^{-}(\mathbb{G}_{-})\;\;\text{and}\;\;\mathcal{C}_{+}:\mathcal{GC}^{-}(\mathbb{G}_{-})\to\mathcal{GC}^{-}(\mathbb{G}_{+})

where π’žβˆ’\mathcal{C}_{-} and π’ž+\mathcal{C}_{+} are homogeneous of degree (0,0)(0,0) and (βˆ’2,βˆ’1)(-2,-1) respectively, such that π’žβˆ’βˆ˜π’ž+\mathcal{C}_{-}\circ\mathcal{C}_{+} and π’ž+βˆ˜π’žβˆ’\mathcal{C}_{+}\circ\mathcal{C}_{-} are filtered chain homotopic to multiplication by V1V_{1}.

Using the fact that G​Cβˆ’β€‹(𝔾)GC^{-}(\mathbb{G}) is the associated graded object of π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}), it is not difficult to show that TheoremΒ 1.2 implies PropositionΒ 1.1, see SectionΒ 3 for details.

Using the crossing change maps in TheoremΒ 1.2, we obtain a knot invariant 𝔩​(K)\mathfrak{l}(K) which is a lower bound on the unknotting number, see TheoremΒ 4.2. Moreover, we show that 𝔩​(K)\mathfrak{l}(K) is equal to the Alishahi–Eftekhary knot invariant see TheoremΒ 5.2. The Alishahi–Eftekhary invariant, given in Definition 3.1 of [AE20], is defined using a generalization of sutured Floer homology [AE15], first developed by JuhΓ‘sz [Juh06]. In particular, this paper provides a combinatorial interpretation of their invariant. In the cases when the filtered chain homotopy type π’’β€‹π’žβˆ’β€‹(K)\mathcal{GC}^{-}(K) of the filtered grid complex is known, computations of 𝔩​(K)\mathfrak{l}(K) are possible, see SectionΒ 4 for more details.

The paper is organized as follows. In SectionΒ 2, we review the necessary constructions from grid homology. In SectionΒ 3, we prove TheoremΒ 1.2. In SectionΒ 4, we define a knot invariant based on the crossing change maps in TheoremΒ 1.2 and prove that the invariant is a lower bound on the unknotting number. In SectionΒ 5, we review the construction of the Alishahi–Eftekhary knot invariant, show the two invariants are equal.

Acknowledgements

The author would like to thank Peter OzsvΓ‘th for his constant guidance and support throughout the project. The author would also like to thank Ollie Thakar and Isabella Khan for helpful discussions.

2. Definitions and constructions

In this section, we review the necessary definitions and constructions from grid homology. We follow OzsvΓ‘th–Szabó–Stipsicz’s book [OSS15].

2.1. Grid diagrams

In this section, we review grid diagrams and an operation relating two grids whose underlying knots differ by a crossing change, called cross-commutation.

Definition 2.1.

A planar grid diagram 𝔾\mathbb{G} is an nΓ—nn\times n grid on the plane with nn of the squares marked with an XX and nn of the squares marked with an OO. The markings are subject to the following rules:

  1. (1)

    Each row has exactly one square marked with an XX and a single square marked with an OO. Each column also has exactly one square marked with an XX and exactly one square marked with an OO.

  2. (2)

    No square is marked with an XX and an OO.

The number nn is called the grid number of 𝔾\mathbb{G}.

A grid diagram 𝔾\mathbb{G} specifies an oriented link LL by the following steps:

  1. (1)

    Draw oriented segments from the XX-marked squares to the OO-marked squares in each column.

  2. (2)

    Draw oriented segments connecting the OO-marked squares to the XX-marked squares in each row such that the vertical segments always cross above the horizontal segments.

Every oriented link in S3S^{3} can be represented by a grid diagram, see Lemma 3.1.3 in [OSS15] for a proof. For example, FigureΒ 1 is a grid diagram for the right-handed trefoil.

Refer to caption
Figure 1. A grid diagram for the right-handed trefoil.

The only operation on grid diagrams we will be using is cross-commutation.

Definition 2.2.

Fix two consecutive columns (resp. rows) in a grid diagram 𝔾\mathbb{G} and let 𝔾′\mathbb{G}^{\prime} be obtained by interchanging those two columns (resp. rows). Suppose that the interiors of their corresponding intervals intersect nontrivially, but neither is contained in the other. Then 𝔾,𝔾′\mathbb{G},\mathbb{G}^{\prime} are said to be related by a cross-commutation.

See FigureΒ 2 for an example of a cross-commutation move.

Refer to caption
Figure 2. A cross-commutation move.

Let K+K_{+} be a knot with a distinguished positive crossing, and let Kβˆ’K_{-} be the same knot with the distinguished crossing changed. Then we can represent K+K_{+} and Kβˆ’K_{-} with grid diagrams 𝔾+\mathbb{G}_{+} and π”Ύβˆ’\mathbb{G}_{-} differing by a cross-commutation of columns.

Before defining the grid complex, we need to discuss the generators of our complex, called grid states, and their properties.

2.2. Grid states and connecting grid states

Consider a toroidal grid diagram 𝔾\mathbb{G} for a knot KK with grid number nn. We can label the horizontal circles 𝜢={Ξ±1,…,Ξ±n}\bm{\alpha}=\{\alpha_{1},\ldots,\alpha_{n}\} and the vertical circles 𝜷={Ξ²1,…,Ξ²n}\bm{\beta}=\{\beta_{1},\ldots,\beta_{n}\}. Define a grid state to be an nn-tuple of points 𝐱={x1,…,xn}\mathbf{x}=\{x_{1},\ldots,x_{n}\} such that each horizontal circle contains exactly one of the elements in 𝐱\mathbf{x} and each vertical circle containe exactly one of the elements in 𝐱\mathbf{x}. Let 𝐒​(𝔾)\mathbf{S}(\mathbb{G}) be the set of grid states for 𝔾\mathbb{G}.

Definition 2.3.

Fix two grid states 𝐱,π²βˆˆπ’β€‹(𝔾)\mathbf{x},\mathbf{y}\in\mathbf{S}(\mathbb{G}). An embedded disk rr in the torus whose boundary is the union of four arcs, each of which lies in some Ξ±j\alpha_{j} or Ξ²j\beta_{j}, is called a rectangle 𝐱\mathbf{x} to 𝐲\mathbf{y} if it satisfies conditions

  • β€’

    At any of the corner points xx of rr, the rectangle contains exactly one of the four quadrants determined by the two intersecting curves at xx.

  • β€’

    All of the corner points of rr belong to 𝐱βˆͺ𝐲\mathbf{x}\cup\mathbf{y}.

  • β€’

    Let βˆ‚ar\partial_{a}r be the part of the boundary of rr belonging to Ξ±1βˆͺβ‹―βˆͺΞ±n\alpha_{1}\cup\cdots\cup\alpha_{n}. Then βˆ‚(βˆ‚ar)=π²βˆ’π±\partial(\partial_{a}r)=\mathbf{y}-\mathbf{x}, where π²βˆ’π±\mathbf{y}-\mathbf{x} is thought of as a formal sum of points.

Denote the set of rectangles from 𝐱\mathbf{x} to 𝐲\mathbf{y} by Rect⁑(𝐱,𝐲)\operatorname{Rect}(\mathbf{x},\mathbf{y}). A rectangle is called an empty rectangle if

𝐱∩int⁑(r)=𝐲∩int⁑(r)=βˆ….\mathbf{x}\cap\operatorname{int}(r)=\mathbf{y}\cap\operatorname{int}(r)=\emptyset.

The set of empty rectangles from 𝐱\mathbf{x} to 𝐲\mathbf{y} is denoted Rect∘⁑(𝐱,𝐲)\operatorname{Rect}^{\circ}(\mathbf{x},\mathbf{y}).

2.3. Unblocked grid complex

For the rest of the paper, we will be working over the field of two elements 𝔽=β„€/2​℀\mathbb{F}=\mathbb{Z}/2\mathbb{Z}.

Definition 2.4.

Let 𝔾\mathbb{G} be a toroidal grid diagram with grid number nn representing a knot KK. The unblocked grid complex G​Cβˆ’β€‹(𝔾)GC^{-}(\mathbb{G}) is a free module over 𝔽​[V1,…,Vn]\mathbb{F}[V_{1},\ldots,V_{n}] generated by the grid states 𝐒​(𝔾)\mathbf{S}(\mathbb{G}) equipped with a 𝔽​[V1,…,Vn]\mathbb{F}[V_{1},\ldots,V_{n}]-module endomorphism βˆ‚π•βˆ’\partial_{\mathbb{X}}^{-} sending π±βˆˆπ’β€‹(𝔾)\mathbf{x}\in\mathbf{S}(\mathbb{G}) to

βˆ‚π•βˆ’(𝐱)=βˆ‘π²βˆˆπ’β€‹(𝔾)βˆ‘{r∈Rect∘⁑(𝐱,𝐲)∣rβˆ©π•=βˆ…}V1O1​(r)​⋯​VnOn​(r)⋅𝐲.\partial^{-}_{\mathbb{X}}(\mathbf{x})=\sum_{\mathbf{y}\in\mathbf{S}(\mathbb{G})}\sum_{\{r\in\operatorname{Rect}^{\circ}(\mathbf{x},\mathbf{y})\mid r\cap\mathbb{X}=\emptyset\}}V_{1}^{O_{1}(r)}\cdots V_{n}^{O_{n}(r)}\cdot\mathbf{y}.

The endomorphism βˆ‚π•βˆ’\partial^{-}_{\mathbb{X}} satisfies (βˆ‚π•βˆ’)2=0(\partial^{-}_{\mathbb{X}})^{2}=0, making G​Cβˆ’β€‹(𝔾)GC^{-}(\mathbb{G}) into a complex, see Chapter 4 of [OSS15]. The unblocked grid complex is bigraded with the Maslov grading and the Alexander grading. We define these gradings now and give some properties.

Proposition 2.5.

The Maslov grading M=M𝕆M=M_{\mathbb{O}} is a unique function on grid states 𝐒​(𝔾)\mathbf{S}(\mathbb{G}) uniquely characterized by

  1. (M-1)

    If 𝐱N​W​O\mathbf{x}^{NWO} is the grade state whose components are the upper left corners of squares marked OO, then

    M​(𝐱N​W​O)=0.M(\mathbf{x}^{NWO})=0.
  2. (M-2)

    If 𝐱\mathbf{x} and 𝐲\mathbf{y} can be connected by a rectangle r∈Rect⁑(𝐱,𝐲)r\in\operatorname{Rect}(\mathbf{x},\mathbf{y}), then

    M​(𝐱)βˆ’M​(𝐲)=1βˆ’2​#​(rβˆ©π•†)+2​#​(𝐱∩int⁑(r)).M(\mathbf{x})-M(\mathbf{y})=1-2\#(r\cap\mathbb{O})+2\#(\mathbf{x}\cap\operatorname{int}(r)).

The existence and uniqueness of the Maslov grading is shown in Section 4.3 of [OSS15]. Define M𝕏M_{\mathbb{X}} analogously, replacing all instances of OO with XX.

There is another characterization of MM. Consider the partial ordering on ℝ2\mathbb{R}^{2}, specified by (p1,p2)<(q1,q2)(p_{1},p_{2})<(q_{1},q_{2}) if p1<q1p_{1}<q_{1} and p2<q2p_{2}<q_{2}. For two finite point-sets P,QP,Q, let I​(P,Q)I(P,Q) be the number of pairs p∈Pp\in P and q∈Qq\in Q such that p<qp<q. Let π’₯​(P,Q)=12​(I​(P,Q)+I​(Q,P))\mathcal{J}(P,Q)=\frac{1}{2}(I(P,Q)+I(Q,P)). Extend π’₯\mathcal{J} bilinearly over formal sums and differences of subsets of the plane. The following is [OSS15, Lemma 4.3.5].

Lemma 2.6.
M​(𝐱)=π’₯​(π±βˆ’π•†,π±βˆ’π•†)+1.M(\mathbf{x})=\mathcal{J}(\mathbf{x}-\mathbb{O},\mathbf{x}-\mathbb{O})+1.
Definition 2.7.

The Alexander function on grid states is defined by

A​(𝐱)=12​(M𝕆​(𝐱)βˆ’M𝕏​(𝐱))βˆ’(nβˆ’12).A(\mathbf{x})=\frac{1}{2}(M_{\mathbb{O}}(\mathbf{x})-M_{\mathbb{X}}(\mathbf{x}))-\left(\frac{n-1}{2}\right).

The following property characterizes the Alexander grading up to an additive constant, compare Proposition 4.3.3 of [OSS15].

Proposition 2.8.

The Alexander function is characterized, up to an additive constant, by the following property. For any rectangle r∈Rect⁑(𝐱,𝐲)r\in\operatorname{Rect}(\mathbf{x},\mathbf{y}),

A​(𝐱)βˆ’A​(𝐲)=#​(rβˆ©π•)βˆ’#​(rβˆ©π•†).A(\mathbf{x})-A(\mathbf{y})=\#(r\cap\mathbb{X})-\#(r\cap\mathbb{O}).

Extend the Maslov and Alexander functions to the basis V1k1​⋯​VnknV_{1}^{k_{1}}\cdots V_{n}^{k_{n}} for arbitrary nonnegative integers k1,…,knk_{1},\ldots,k_{n} by defining M​(V1k1​⋯​Vnkn⋅𝐱)=M​(𝐱)βˆ’2β€‹βˆ‘i=1nkiM(V_{1}^{k_{1}}\cdots V_{n}^{k_{n}}\cdot\mathbf{x})=M(\mathbf{x})-2\sum_{i=1}^{n}k_{i} and A​(V1k1​⋯​Vnkn⋅𝐱)=A​(𝐱)βˆ’βˆ‘i=1nkiA(V_{1}^{k_{1}}\cdots V_{n}^{k_{n}}\cdot\mathbf{x})=A(\mathbf{x})-\sum_{i=1}^{n}k_{i}.

In particular, multiplication by ViV_{i} has chain map of degree (βˆ’2,βˆ’1)(-2,-1).

2.4. Filtered grid complexes

Let kk be a field, which in our case will always be β„€/2​℀\mathbb{Z}/2\mathbb{Z}.

Definition 2.9.

A β„€\mathbb{Z}-filtered, β„€\mathbb{Z}-graded chain complex over k​[V1,…,Vn]k[V_{1},\ldots,V_{n}] is kk-module π’ž\mathcal{C} with

  • β€’

    A differential βˆ‚:π’žβ†’π’ž\partial:\mathcal{C}\to\mathcal{C}, i.e. a k​[V1,…,Vn]k[V_{1},\ldots,V_{n}]-module homomorphism with βˆ‚2=0\partial^{2}=0 compatible with the β„€\mathbb{Z}-grading π’ž=⨁dβˆˆβ„€π’žd\mathcal{C}=\bigoplus_{d\in\mathbb{Z}}\mathcal{C}_{d}, in the sense that βˆ‚π’ždβŠ†π’ždβˆ’1\partial\mathcal{C}_{d}\subseteq\mathcal{C}_{d-1}.

  • β€’

    A β„€\mathbb{Z}-filtration β„±sβ€‹π’ž\mathcal{F}_{s}\mathcal{C} for sβˆˆβ„€s\in\mathbb{Z} that exhausts π’ž\mathcal{C}, in the sense that ⋃sβˆˆβ„€β„±sβ€‹π’ž=π’ž\bigcup_{s\in\mathbb{Z}}\mathcal{F}_{s}\mathcal{C}=\mathcal{C}. The filtration is compatible with the β„€\mathbb{Z}-grading, in the sense that if β„±sβ€‹π’žd=β„±sβ€‹π’žβˆ©π’žd\mathcal{F}_{s}\mathcal{C}_{d}=\mathcal{F}_{s}\mathcal{C}\cap\mathcal{C}_{d}, then β„±sβ€‹π’ž=⋃sβˆˆβ„€β„±sβ€‹π’žd\mathcal{F}_{s}\mathcal{C}=\bigcup_{s\in\mathbb{Z}}\mathcal{F}_{s}\mathcal{C}_{d}. The filtration is compatible with the differential, in the sense that βˆ‚(β„±sβ€‹π’ž)βŠ†β„±sβ€‹π’ž\partial(\mathcal{F}_{s}\mathcal{C})\subseteq\mathcal{F}_{s}\mathcal{C}, making β„±sβ€‹π’ž\mathcal{F}_{s}\mathcal{C} into a subcomplex. The filtration is bounded below, in the sense that for any integer dd, there exists an integer ndn_{d} such that β„±sβ€‹π’žd=0\mathcal{F}_{s}\mathcal{C}_{d}=0 for all s≀nds\leq n_{d}.

  • β€’

    There are kk-module endomorphisms Vi:π’žβ†’π’žV_{i}:\mathcal{C}\to\mathcal{C}, which satisfy: Vi,VjV_{i},V_{j} commute for all 1≀i,j≀n1\leq i,j\leq n, βˆ‚Vi=Viβ€‹βˆ‚\partial V_{i}=V_{i}\partial, Vi​(π’žd)βŠ†π’ždβˆ’2V_{i}(\mathcal{C}_{d})\subseteq\mathcal{C}_{d-2}, and Vi​(β„±sβ€‹π’ž)βŠ†β„±sβˆ’1β€‹π’žV_{i}(\mathcal{F}_{s}\mathcal{C})\subseteq\mathcal{F}_{s-1}\mathcal{C}.

Definition 2.10.

Let 𝔾\mathbb{G} be a toroidal grid diagram with grid number nn representing a knot KK. The filtered grid complex π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}) is generated over 𝔽​[V1,…,Vn]\mathbb{F}[V_{1},\ldots,V_{n}] by 𝐒​(𝔾)\mathbf{S}(\mathbb{G}) and has differential

βˆ‚βˆ’π±=βˆ‘π²βˆˆπ’β€‹(𝔾)βˆ‘r∈Rect∘⁑(𝐱,𝐲)V1O1​(r)​⋯​VnOn​(r)⋅𝐲.\partial^{-}\mathbf{x}=\sum_{\mathbf{y}\in\mathbf{S}(\mathbb{G})}\sum_{r\in\operatorname{Rect}^{\circ}(\mathbf{x},\mathbf{y})}V_{1}^{O_{1}(r)}\cdots V_{n}^{O_{n}(r)}\cdot\mathbf{y}.

The Maslov grading is given by M​(𝐱)=dM(\mathbf{x})=d and the Alexander filtration is given by A​(𝐱)=sA(\mathbf{x})=s, in the sense that β„±sβ€‹π’ž\mathcal{F}_{s}\mathcal{C} is the 𝔽\mathbb{F}-span of generators V1k1​⋯​Vnkn⋅𝐱V_{1}^{k_{1}}\cdots V_{n}^{k_{n}}\cdot\mathbf{x} which evaluate to most ss under the Alexander function.

Note that the differential now counts rectangles containing XX-markings. The filtered module π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}) is shown to be a chain complex in Chapter 13 of [OSS15]. The most important property is that the filtered quasi-isomorphism type of π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}) depends only on the underlying knot KK, see Theorem 13.2.9 of [OSS15].

Given a β„€\mathbb{Z}-graded, β„€\mathbb{Z}-filtered chain complex (π’ž,βˆ‚)(\mathcal{C},\partial), the associated graded object is the chain complex gr⁑(π’ž)=⨁d,sβˆˆβ„€(β„±sβ€‹π’žd/β„±sβˆ’1β€‹π’žd)\operatorname{gr}(\mathcal{C})=\bigoplus_{d,s\in\mathbb{Z}}(\mathcal{F}_{s}\mathcal{C}_{d}/\mathcal{F}_{s-1}\mathcal{C}_{d}) with bigrading gr(π’ž)d,s=β„±sπ’žd/β„±sβˆ’1π’žd\operatorname{gr}(\mathcal{C})_{d,s}=\mathcal{F}_{s}\mathcal{C}_{d}/\mathcal{F}_{s-1}\mathcal{C}_{d} and differential gr⁑(βˆ‚)\operatorname{gr}(\partial) induced by βˆ‚:π’žβ†’π’ž\partial:\mathcal{C}\to\mathcal{C}, i.e. gr(βˆ‚)=βˆ‘d,sgr(βˆ‚)d,s\operatorname{gr}(\partial)=\sum_{d,s}\operatorname{gr}(\partial)_{d,s} where gr(βˆ‚)d,s:β„±sπ’žd/β„±sβˆ’1π’ždβ†’β„±sπ’ždβˆ’1/β„±sβˆ’1π’ždβˆ’1\operatorname{gr}(\partial)_{d,s}:\mathcal{F}_{s}\mathcal{C}_{d}/\mathcal{F}_{s-1}\mathcal{C}_{d}\to\mathcal{F}_{s}\mathcal{C}_{d-1}/\mathcal{F}_{s-1}\mathcal{C}_{d-1}. The following is Proposition 13.2.6 of [OSS15].

Proposition 2.11.

The associated graded object of (π’’β€‹π’žβˆ’β€‹(𝔾),βˆ‚βˆ’)(\mathcal{\mathcal{GC}}^{-}(\mathbb{G}),\partial^{-}) is (G​Cβˆ’β€‹(𝔾),βˆ‚π•βˆ’)(GC^{-}(\mathbb{G}),\partial^{-}_{\mathbb{X}}).

3. Proof of TheoremΒ 1.2

First we show that TheoremΒ 1.2 is a generalization of the OzsvΓ‘th–Stipsicz–SzabΓ³ crossing maps in PropositionΒ 1.1. Recall that (G​Cβˆ’β€‹(𝔾),βˆ‚π•βˆ’)(GC^{-}(\mathbb{G}),\partial^{-}_{\mathbb{X}}) is the associated graded object of (π’’β€‹π’žβˆ’β€‹(𝔾),βˆ‚βˆ’)(\mathcal{GC}^{-}(\mathbb{G}),\partial^{-}), see PropositionΒ 2.11. By functoriality we retrieve two maps

(1) cβˆ’:G​Cβˆ’β€‹(𝔾+)β†’G​Cβˆ’β€‹(π”Ύβˆ’)​and​c+:G​Cβˆ’β€‹(π”Ύβˆ’)β†’G​Cβˆ’β€‹(𝔾+)c_{-}:GC^{-}(\mathbb{G}_{+})\to GC^{-}(\mathbb{G}_{-})\;\;\text{and}\;\;c_{+}:GC^{-}(\mathbb{G}_{-})\to GC^{-}(\mathbb{G}_{+})

such that cβˆ’c_{-} is bigraded and c+c_{+} is homogeneous of degree (βˆ’2,βˆ’1)(-2,-1) such that cβˆ’βˆ˜c+c_{-}\circ c_{+} and c+∘cβˆ’c_{+}\circ c_{-} are chain-homotopic to multiplication by V1V_{1}. Taking homology, we get the statement of PropositionΒ 1.1.

Now we proceed to the proof of TheoremΒ 1.2. We use a similar set up as in Chapter 5 of [OSS15]. Draw both grids 𝔾+\mathbb{G}_{+} and π”Ύβˆ’\mathbb{G}_{-} on the same torus, where the 𝕆\mathbb{O} and 𝕏\mathbb{X} markings are fixed. Start with nn horizontal circles, nβˆ’1n-1 vertical circles, and two additional circles Ξ²i\beta_{i} and Ξ³i\gamma_{i} corresponding to the ii-th circles of 𝔾+\mathbb{G}_{+} and π”Ύβˆ’\mathbb{G}_{-}, such that Ξ²i\beta_{i} and Ξ³i\gamma_{i} intersect in four points. To define our crossing change maps, we need to mark two of these intersection points, which we explain now. The complement of Ξ²iβˆͺΞ³i\beta_{i}\cup\gamma_{i} in the grid torus has five components, four of which are bigons, and each bigon contains exactly one XX or exactly one OO. Since 𝔾+\mathbb{G}_{+} and π”Ύβˆ’\mathbb{G}_{-} differ by a cross-commutation, the two XX marked bigons share a vertex on Ξ²i∩γi\beta_{i}\cap\gamma_{i}; call this vertex tt. Label the two OO-markings such that O1O_{1} is above O2O_{2} in our grid diagram. The bigon containing O2O_{2} and one of the XX-labeled bigons share a vertex which we call ss. These notational choices are shown in FigureΒ 3.

To define the crossing-change maps, we need to count other domains between grid states: pentagons and hexagons.

Refer to caption
Figure 3. Grid diagram for crossing changes.
Definition 3.1.

Fix grid states π±βˆ’βˆˆπ’β€‹(π”Ύβˆ’)\mathbf{x}_{-}\in\mathbf{S}(\mathbb{G}_{-}) and 𝐱+βˆˆπ’β€‹(𝔾+)\mathbf{x}_{+}\in\mathbf{S}(\mathbb{G}_{+}). An embedded disk pp in the torus whose boundary is the union of five arcs, each of which lies in some Ξ±j,Ξ²j\alpha_{j},\beta_{j}, or Ξ³i\gamma_{i}, is called a pentagon from 𝐱+\mathbf{x}_{+} to π²βˆ’\mathbf{y}_{-} if it satisfies the following conditions

  • β€’

    At any of the corner points xx of pp, the pentagon contains exactly one of the four quadrants determined by the two intersecting curves at xx.

  • β€’

    Four of the corners of pp are in 𝐱+βˆͺπ²βˆ’\mathbf{x}_{+}\cup\mathbf{y}_{-}, the other corner point is chosen from the four intersection points of the curves Ξ²i\beta_{i} and Ξ³i\gamma_{i}.

  • β€’

    Let βˆ‚Ξ±p\partial_{\alpha}p be the part of the boundary of pp belonging to Ξ±1βˆͺβ‹―βˆͺΞ±n\alpha_{1}\cup\cdots\cup\alpha_{n}. Then βˆ‚(βˆ‚Ξ±p)=π²βˆ’βˆ’π±+\partial(\partial_{\alpha}p)=\mathbf{y}_{-}-\mathbf{x}_{+}.

Let Pent⁑(𝐱+,π²βˆ’)\operatorname{Pent}(\mathbf{x}_{+},\mathbf{y}_{-}) be the set of pentagons from 𝐱+βˆˆπ’β€‹(𝔾+)\mathbf{x}_{+}\in\mathbf{S}(\mathbb{G}_{+}) to π²βˆ’βˆˆπ’β€‹(π”Ύβˆ’)\mathbf{y}_{-}\in\mathbf{S}(\mathbb{G}_{-}). A pentagon p∈Pent⁑(𝐱+,π²βˆ’)p\in\operatorname{Pent}(\mathbf{x}_{+},\mathbf{y}_{-}) is empty if int⁑(p)∩𝐱+=int⁑(p)βˆ©π²βˆ’=βˆ…\operatorname{int}(p)\cap\mathbf{x}_{+}=\operatorname{int}(p)\cap\mathbf{y}_{-}=\emptyset. Let Pent∘⁑(𝐱+,π²βˆ’)\operatorname{Pent}^{\circ}(\mathbf{x}_{+},\mathbf{y}_{-}) be the set of empty pentagons from 𝐱+\mathbf{x}_{+} to π²βˆ’\mathbf{y}_{-}. Let Pents∘⁑(𝐱+,π²βˆ’)\operatorname{Pent}_{s}^{\circ}(\mathbf{x}_{+},\mathbf{y}_{-}) be the set of empty pentagons from 𝐱+\mathbf{x}_{+} to π²βˆ’\mathbf{y}_{-} with one vertex at ss. For two grid states π±βˆ’βˆˆπ’β€‹(π”Ύβˆ’)\mathbf{x}_{-}\in\mathbf{S}(\mathbb{G}_{-}) and 𝐲+βˆˆπ’β€‹(𝔾+)\mathbf{y}_{+}\in\mathbf{S}(\mathbb{G}_{+}), we define Pent⁑(π±βˆ’,𝐲+),Pent∘⁑(π±βˆ’,𝐲+),Pentt∘⁑(π±βˆ’,𝐲+)\operatorname{Pent}(\mathbf{x}_{-},\mathbf{y}_{+}),\operatorname{Pent}^{\circ}(\mathbf{x}_{-},\mathbf{y}_{+}),\operatorname{Pent}^{\circ}_{t}(\mathbf{x}_{-},\mathbf{y}_{+}) similarly.

Definition 3.2.

Fix grid states 𝐱,π²βˆˆπ’β€‹(𝔾)\mathbf{x},\mathbf{y}\in\mathbf{S}(\mathbb{G}). An embedded disk hh in the torus whose boundary is in the union of the Ξ±j,Ξ²j\alpha_{j},\beta_{j} (for j=1,…,nj=1,\ldots,n) and Ξ³i\gamma_{i} is called a hexagon from 𝐱\mathbf{x} to 𝐲\mathbf{y} if it satisfies the following conditions:

  • β€’

    At any of the six corner points xx of hh, the hexagon contains exactly one of the four quadrants determined by the two intersecting curves at xx.

  • β€’

    Four of the corner points of hh are in 𝐱βˆͺ𝐲\mathbf{x}\cup\mathbf{y}, and the other two corners are aa and bb, where a,ba,b are chosen from the four intersection points of the curves Ξ²i\beta_{i} and Ξ³i\gamma_{i}.

  • β€’

    Let βˆ‚Ξ±h\partial_{\alpha}h be the part of the boundary of hh belonging to Ξ±1βˆͺβ‹―βˆͺΞ±n\alpha_{1}\cup\cdots\cup\alpha_{n}. Then βˆ‚(βˆ‚Ξ±h)=π²βˆ’π±\partial(\partial_{\alpha}h)=\mathbf{y}-\mathbf{x}.

Let Hex⁑(𝐱,𝐲)\operatorname{Hex}(\mathbf{x},\mathbf{y}) be the set of hexagons from 𝐱\mathbf{x} to 𝐲\mathbf{y}. A hexagon h∈Hex⁑(𝐱,𝐲)h\in\operatorname{Hex}(\mathbf{x},\mathbf{y}) is empty if int⁑(h)∩𝐱=int⁑(h)∩𝐲=βˆ…\operatorname{int}(h)\cap\mathbf{x}=\operatorname{int}(h)\cap\mathbf{y}=\emptyset. Let Hex∘⁑(𝐱,𝐲)\operatorname{Hex}^{\circ}(\mathbf{x},\mathbf{y}) be the set of empty hexagons from 𝐱\mathbf{x} to 𝐲\mathbf{y}. Let Hexs,t∘⁑(𝐱,𝐲)\operatorname{Hex}_{s,t}^{\circ}(\mathbf{x},\mathbf{y}) denote the set of empty hexagons with two consecutive corners at ss and tt in the order specified by the orientation of the hexagon. Let Hext,s∘⁑(𝐱,𝐲)\operatorname{Hex}_{t,s}^{\circ}(\mathbf{x},\mathbf{y}) be the analogous set with the order of ss and tt reversed.

Now we define the crossing change maps. Fix arbitrary grid states 𝐱+βˆˆπ’β€‹(𝔾+)\mathbf{x}_{+}\in\mathbf{S}(\mathbb{G}_{+}) and π±βˆ’βˆˆπ’β€‹(π”Ύβˆ’)\mathbf{x}_{-}\in\mathbf{S}(\mathbb{G}_{-}). Recall that Pents∘⁑(𝐱+,π±βˆ’)\operatorname{Pent}_{s}^{\circ}(\mathbf{x}_{+},\mathbf{x}_{-}) is the set of empty pentagons from 𝐱+\mathbf{x}_{+} to π±βˆ’\mathbf{x}_{-} with one vertex at ss, and Pentt∘⁑(π±βˆ’,𝐱+)\operatorname{Pent}_{t}^{\circ}(\mathbf{x}_{-},\mathbf{x}_{+}) is the set of empty pentagons from π±βˆ’\mathbf{x}_{-} to 𝐱+\mathbf{x}_{+} with one vertex at tt. Define the two 𝔽​[V1,…,Vn]\mathbb{F}[V_{1},\ldots,V_{n}]-module maps

π’žβˆ’:π’’β€‹π’žβˆ’β€‹(𝔾+)β†’π’’β€‹π’žβˆ’β€‹(π”Ύβˆ’)​andβ€‹π’ž+:π’’β€‹π’žβˆ’β€‹(π”Ύβˆ’)β†’π’’β€‹π’žβˆ’β€‹(𝔾+)\mathcal{C}_{-}:\mathcal{GC}^{-}(\mathbb{G}_{+})\to\mathcal{GC}^{-}(\mathbb{G}_{-})\;\;\text{and}\;\;\mathcal{C}_{+}:\mathcal{GC}^{-}(\mathbb{G}_{-})\to\mathcal{GC}^{-}(\mathbb{G}_{+})

by counting pentagons either containing the vertex ss or the vertex tt:

(2) π’žβˆ’β€‹(𝐱+)\displaystyle\mathcal{C}_{-}(\mathbf{x}_{+}) =βˆ‘π²βˆ’βˆˆπ’β€‹(π”Ύβˆ’)βˆ‘p∈Pents∘⁑(𝐱+,π²βˆ’)V1O1​(p)​⋯​VnOn​(p)β‹…π²βˆ’,\displaystyle=\sum_{\mathbf{y}_{-}\in\mathbf{S}(\mathbb{G}_{-})}\sum_{p\in\operatorname{Pent}_{s}^{\circ}(\mathbf{x}_{+},\mathbf{y}_{-})}V_{1}^{O_{1}(p)}\cdots V_{n}^{O_{n}(p)}\cdot\mathbf{y}_{-},
(3) π’ž+​(π±βˆ’)\displaystyle\mathcal{C}_{+}(\mathbf{x}_{-}) =βˆ‘π²+βˆˆπ’β€‹(𝔾+)βˆ‘p∈Pentt∘⁑(π±βˆ’,𝐲+)V1O1​(p)​⋯​VnOn​(p)⋅𝐲+.\displaystyle=\sum_{\mathbf{y}_{+}\in\mathbf{S}(\mathbb{G}_{+})}\sum_{p\in\operatorname{Pent}_{t}^{\circ}(\mathbf{x}_{-},\mathbf{y}_{+})}V_{1}^{O_{1}(p)}\cdots V_{n}^{O_{n}(p)}\cdot\mathbf{y}_{+}.

Here Oi​(p)O_{i}(p) is 11 if pp contains OiO_{i} and 0 otherwise, where O1,…,OnO_{1},\ldots,O_{n} are the 𝕆\mathbb{O} markings.

For fixed grid states π±Β±βˆˆπ”Ύ+\mathbf{x}_{\pm}\in\mathbb{G}_{+} and π²βˆ“βˆˆπ”Ύβˆ’\mathbf{y}_{\mp}\in\mathbb{G}_{-}, a domain ψ\psi from 𝐱±\mathbf{x}_{\pm} to π²βˆ’\mathbf{y}_{-} is a formal sum of closures of regions in the complement of the Ξ±j,Ξ²j\alpha_{j},\beta_{j}, and Ξ³i\gamma_{i} in the grid torus, taken with integral multiplicities, such that βˆ‚Ξ±Οˆ\partial_{\alpha}\psi, the portion of the boundary in Ξ±1βˆͺβ‹―βˆͺΞ±n\alpha_{1}\cup\cdots\cup\alpha_{n}, satisfies βˆ‚(βˆ‚Ξ±Οˆ)=π²βˆ“βˆ’π±Β±\partial(\partial_{\alpha}\psi)=\mathbf{y}_{\mp}-\mathbf{x}_{\pm}. Let π​(𝐱±,π²βˆ“)\pi(\mathbf{x}_{\pm},\mathbf{y}_{\mp}) be the set of domains from 𝐱±\mathbf{x}_{\pm} to π²βˆ“\mathbf{y}_{\mp}.

Lemma 3.3.

π’žβˆ’\mathcal{C}_{-} and π’ž+\mathcal{C}_{+} are chain maps.

Proof.

This proof follows Lemma 5.1.4 in [OSS15] with some extra cases.

To show that π’žβˆ’\mathcal{C}_{-} is a chain map, since we are working over 𝔽=β„€/2​℀\mathbb{F}=\mathbb{Z}/2\mathbb{Z}, it is enough to show that βˆ‚βˆ’βˆ˜π’žβˆ’+π’žβˆ’βˆ˜βˆ‚βˆ’\partial^{-}\circ\mathcal{C}_{-}+\mathcal{C}_{-}\circ\partial^{-} is identically zero. This expression can be written as

(4) βˆ‚βˆ’βˆ˜π’žβˆ’+π’žβˆ’βˆ˜βˆ‚βˆ’=βˆ‘π³βˆ’βˆˆπ’β€‹(𝔾)βˆ‘ΟˆβˆˆΟ€β€‹(𝐱+,π³βˆ’)N​(ψ)β‹…V1O1​(ψ)​⋯​VnOn​(ψ)β‹…π³βˆ’,\partial^{-}\circ\mathcal{C}_{-}+\mathcal{C}_{-}\circ\partial^{-}=\sum_{\mathbf{z}_{-}\in\mathbf{S}(\mathbb{G})}\sum_{\psi\in\pi(\mathbf{x}_{+},\mathbf{z}_{-})}N(\psi)\cdot V_{1}^{O_{1}(\psi)}\cdots V_{n}^{O_{n}(\psi)}\cdot\mathbf{z}_{-},

where N​(ψ)N(\psi) is the number of ways to decompose ψ\psi as either rβˆ—pr*p or pβ€²βˆ—rβ€²p^{\prime}*r^{\prime}, where r,rβ€²r,r^{\prime} are empty rectangles and p,pβ€²p,p^{\prime} are empty pentagons. There are three cases of ΟˆβˆˆΟ€β€‹(𝐱+,π³βˆ’)\psi\in\pi(\mathbf{x}_{+},\mathbf{z}_{-}):

  • (P-1)

    𝐱+βˆ–(𝐱+βˆ©π³βˆ’)\mathbf{x}_{+}\setminus(\mathbf{x}_{+}\cap\mathbf{z}_{-}) consists of 44 points. In this case, there are two decompositions of ψ\psi with the same underlying rectangle and pentagon, only differing in the grid states they connect. See case (P-1) in FigureΒ 4. Thus, N​(ψ)=2N(\psi)=2, so there are no contribution of terms in this case.

    Refer to caption
    Figure 4. Domain decompositions. The black dots are contained in 𝐱+\mathbf{x}_{+} and the white dots are contained in π³βˆ’\mathbf{z}_{-}.
  • (P-2)

    𝐱+βˆ–(𝐱+βˆ©π³βˆ’)\mathbf{x}_{+}\setminus(\mathbf{x}_{+}\cap\mathbf{z}_{-}) consists of 33 points. There are two cases to consider here: either all of the local multiplicites of ψ\psi are 0 and 11, or some local multiplicity is 22. In the first case, ψ\psi has seven corners, one of them being a 270∘270^{\circ} corner. Cutting this corner in two directions gives two different decompositions of ψ\psi as a rectangle and a pentagon. In the second case, ψ\psi has a 270∘270^{\circ} corner at aa, and cutting it in two ways gives two decompositions of ψ\psi into a rectangle and a pentagon. See case (P-2) in FigureΒ 4. In all cases, N​(ψ)=2N(\psi)=2, so there are no contribution of terms in this case.

  • (P-3)

    𝐱+βˆ–(𝐱+βˆ©π³βˆ’)\mathbf{x}_{+}\setminus(\mathbf{x}_{+}\cap\mathbf{z}_{-}) consists of 11 point. In this case, π³βˆ’\mathbf{z}_{-} is the unique grid state which agrees with 𝐱+\mathbf{x}_{+} in all but the component Ξ²i\beta_{i}, and the domain ψ\psi goes around the torus, either horizontally or vertically. In the horizontal case, the domain ψ\psi is a horizontal thin annulus minus a small triangle, which can be decomposed in two ways. See case (P-3)(h) FigureΒ 4.

    In the vertical case, the decomposition is unique. Luckily, we can pair the off domains ψ\psi depending on whether their support lies between Ξ²iβˆ’1\beta_{i-1} and Ξ²i\beta_{i}, or between Ξ²i\beta_{i} and Ξ²i+1\beta_{i+1}.

    There are two thin annular regions A1A_{1} and A2A_{2} which have three corners: one corner at ss, another corner is at 𝐱+∩βi\mathbf{x}_{+}\cap\beta_{i} and another corner is at π±βˆ’\mathbf{x}_{-}. There are four combinatorial types of A1A_{1} depending on which bigon 𝐱+∩βi\mathbf{x}_{+}\cap\beta_{i} lies on. We show the four decompositions in FigureΒ 5.

    Refer to caption
    Figure 5. Vertical annulus depending on location of 𝐱+∩βi\mathbf{x}_{+}\cap\beta_{i}.

    The annulus A1A_{1} can be decomposed uniquely as a rectangle and a pentagon in an order determined by the position of 𝐱+∩βiβˆ’1\mathbf{x}_{+}\cap\beta_{i-1} relative to 𝐱+∩βi\mathbf{x}_{+}\cap\beta_{i}. See FigureΒ 6 for the decomposition of A1A_{1} when 𝐱∩βi\mathbf{x}\cap\beta_{i} lies on the O2O_{2} bigon. Similarly, the A2A_{2} annulus can be uniquely decomposed as the sum of a rectangle and pentagon. Since the A1A_{1} and A2A_{2} annuli cross the same XX and OO markings, their contributions to equationΒ (4) cancel.

    Refer to caption
    Figure 6. The decomposition of annulus A1A_{1} depending on location of 𝐱+∩βi\mathbf{x}_{+}\cap\beta_{i}. The darker gray region is the first in the decomposition.

This shows that π’žβˆ’\mathcal{C}_{-} is a chain map. A similar argument can be used to show that π’ž+\mathcal{C}_{+} is a chain map. ∎

Fix two β„€\mathbb{Z}-filtered, β„€\mathbb{Z}-graded chain complexes π’ž\mathcal{C} and π’žβ€²\mathcal{C}^{\prime} over 𝔽​[V1,…,Vn]\mathbb{F}[V_{1},\ldots,V_{n}]. Call a chain map f:π’žβ†’π’žβ€²f:\mathcal{C}\to\mathcal{C}^{\prime} homogeneous of degree (m,t)(m,t) if f​(π’žd)βŠ†π’žd+mf(\mathcal{C}_{d})\subseteq\mathcal{C}_{d+m} and f​(β„±sβ€‹π’ž)βŠ†β„±s+tβ€‹π’žβ€²f(\mathcal{F}_{s}\mathcal{C})\subseteq\mathcal{F}_{s+t}\mathcal{C}^{\prime}.

Lemma 3.4.

π’žβˆ’\mathcal{C}_{-} is homogeneous of degree (0,0)(0,0) and π’ž+\mathcal{C}_{+} is homogeneous of degree (βˆ’2,βˆ’1)(-2,-1).

Proof.

There is a one-to-one correspondence

I:𝐒​(π”Ύβˆ’)→𝐒​(𝔾+)I:\mathbf{S}(\mathbb{G}_{-})\to\mathbf{S}(\mathbb{G}_{+})

called the nearest-point map that sends a grid state π³βˆ’βˆˆπ’β€‹(𝔾)\mathbf{z}_{-}\in\mathbf{S}(\mathbb{G}) to the unique grid state 𝐳+=I​(π³βˆ’)\mathbf{z}_{+}=I(\mathbf{z}_{-}) that agrees with π³βˆ’\mathbf{z}_{-} in all but one component (cf. Lemma 5.1.3 of [OSS15]).

To calculate the grading changes, we associate rectangles to each pentagon via the nearest point map. Fix a grid state 𝐱+βˆˆπ’β€‹(𝔾+)\mathbf{x}_{+}\in\mathbf{S}(\mathbb{G}_{+}). For each grid state π²βˆ’βˆˆπ’β€‹(π”Ύβˆ’)\mathbf{y}_{-}\in\mathbf{S}(\mathbb{G}_{-}) such that there exists a pentagon p∈Pents∘⁑(𝐱+,π²βˆ’)p\in\operatorname{Pent}_{s}^{\circ}(\mathbf{x}_{+},\mathbf{y}_{-}), we consider the associated rectangle r=r​(p)∈Rect∘⁑(𝐱+,𝐲+)r=r(p)\in\operatorname{Rect}^{\circ}(\mathbf{x}_{+},\mathbf{y}_{+}). Define the following constants depending on 𝐱+,π²βˆ’βˆˆπ’β€‹(π”Ύβˆ’)\mathbf{x}_{+},\mathbf{y}_{-}\in\mathbf{S}(\mathbb{G}_{-}):

Ξ”r​(𝐱+,π²βˆ’)\displaystyle\Delta_{r}(\mathbf{x}_{+},\mathbf{y}_{-}) =#​(pβˆ©π•†)βˆ’#​(rβˆ©π•†),\displaystyle=\#(p\cap\mathbb{O})-\#(r\cap\mathbb{O}),
Ξ”M​(π²βˆ’)\displaystyle\Delta_{M}(\mathbf{y}_{-}) =M​(π²βˆ’)βˆ’M​(𝐲+),\displaystyle=M(\mathbf{y}_{-})-M(\mathbf{y}_{+}),
Ξ”A​(π²βˆ’)\displaystyle\Delta_{A}(\mathbf{y}_{-}) =A​(π²βˆ’)βˆ’A​(𝐲+).\displaystyle=A(\mathbf{y}_{-})-A(\mathbf{y}_{+}).

First, note that Ξ”r​(𝐱+,π²βˆ’)\Delta_{r}(\mathbf{x}_{+},\mathbf{y}_{-}) only depends on the location of π²βˆ’βˆ©Ξ²i\mathbf{y}_{-}\cap\beta_{i}: the location of 𝐱+∩βi\mathbf{x}_{+}\cap\beta_{i} does not change the difference #​(pβˆ©π•†)βˆ’#​(rβˆ©π•†)\#(p\cap\mathbb{O})-\#(r\cap\mathbb{O}). So we can drop the dependence on 𝐱+\mathbf{x}_{+} and write Ξ”r​(π²βˆ’)=Ξ”r​(𝐱+,π²βˆ’)\Delta_{r}(\mathbf{y}_{-})=\Delta_{r}(\mathbf{x}_{+},\mathbf{y}_{-}).

For simplicity, let Ξ”M​(𝐱+,π²βˆ’)\Delta_{M}(\mathbf{x}_{+},\mathbf{y}_{-}) be the difference M​(V1O1​(p)​⋯​VnOn​(p)β‹…π²βˆ’)βˆ’M​(𝐱+)M(V_{1}^{O_{1}(p)}\cdots V_{n}^{O_{n}(p)}\cdot\mathbf{y}_{-})-M(\mathbf{x}_{+}), which is the change in Maslov grading. Similarly define Ξ”A​(𝐱+,π²βˆ’)\Delta_{A}(\mathbf{x}_{+},\mathbf{y}_{-}) to be the difference in Alexander grading. Then we can rewrite the change in Maslov grading in terms of these constants:

Ξ”M​(𝐱+,π²βˆ’)\displaystyle\Delta_{M}(\mathbf{x}_{+},\mathbf{y}_{-}) =M​(V1O1​(p)​⋯​VnOn​(p)β‹…π²βˆ’)βˆ’M​(𝐱+)\displaystyle=M(V_{1}^{O_{1}(p)}\cdots V_{n}^{O_{n}(p)}\cdot\mathbf{y}_{-})-M(\mathbf{x}_{+})
=M​(π²βˆ’)βˆ’2​#​(pβˆ©π•†)βˆ’M​(𝐱+)\displaystyle=M(\mathbf{y}_{-})-2\#(p\cap\mathbb{O})-M(\mathbf{x}_{+})
=(M​(𝐲+)+Ξ”M​(π²βˆ’))βˆ’2​#​(pβˆ©π•†)βˆ’M​(𝐱+)\displaystyle=(M(\mathbf{y}_{+})+\Delta_{M}(\mathbf{y}_{-}))-2\#(p\cap\mathbb{O})-M(\mathbf{x}_{+})
=M​(𝐲+)+Ξ”M​(π²βˆ’)βˆ’2​(Ξ”r​(π²βˆ’)+#​(rβˆ©π•†))βˆ’M​(𝐱+)\displaystyle=M(\mathbf{y}_{+})+\Delta_{M}(\mathbf{y}_{-})-2(\Delta_{r}(\mathbf{y}_{-})+\#(r\cap\mathbb{O}))-M(\mathbf{x}_{+})
(Property (M-2)) =βˆ’(1βˆ’2​#​(rβˆ©π•†))+Ξ”M​(π²βˆ’)βˆ’2​Δr​(π²βˆ’)βˆ’2​#​(rβˆ©π•†)\displaystyle=-(1-2\#(r\cap\mathbb{O}))+\Delta_{M}(\mathbf{y}_{-})-2\Delta_{r}(\mathbf{y}_{-})-2\#(r\cap\mathbb{O})
=Ξ”M​(π²βˆ’)βˆ’2​Δr​(π²βˆ’)βˆ’1.\displaystyle=\Delta_{M}(\mathbf{y}_{-})-2\Delta_{r}(\mathbf{y}_{-})-1.

Similarly, we write the change in Alexander grading in terms of these constants:

Ξ”A​(𝐱+,π²βˆ’)\displaystyle\Delta_{A}(\mathbf{x}_{+},\mathbf{y}_{-}) =A​(V1O1​(p)​⋯​VnOn​(p)β‹…π²βˆ’)βˆ’A​(𝐱+)\displaystyle=A(V_{1}^{O_{1}(p)}\cdots V_{n}^{O_{n}(p)}\cdot\mathbf{y}_{-})-A(\mathbf{x}_{+})
=A​(π²βˆ’)βˆ’#​(pβˆ©π•†)βˆ’A​(𝐱+)\displaystyle=A(\mathbf{y}_{-})-\#(p\cap\mathbb{O})-A(\mathbf{x}_{+})
=(A​(𝐲+)+Ξ”A​(π²βˆ’))βˆ’(#​(rβˆ©π•†)+Ξ”r​(π²βˆ’))βˆ’A​(𝐱+)\displaystyle=(A(\mathbf{y}_{+})+\Delta_{A}(\mathbf{y}_{-}))-(\#(r\cap\mathbb{O})+\Delta_{r}(\mathbf{y}_{-}))-A(\mathbf{x}_{+})
=(#​(rβˆ©π•†)βˆ’#​(rβˆ©π•))+Ξ”A​(π²βˆ’)βˆ’#​(rβˆ©π•†)βˆ’Ξ”r​(π²βˆ’)\displaystyle=(\#(r\cap\mathbb{O})-\#(r\cap\mathbb{X}))+\Delta_{A}(\mathbf{y}_{-})-\#(r\cap\mathbb{O})-\Delta_{r}(\mathbf{y}_{-})
(PropositionΒ 2.8) =Ξ”A​(π²βˆ’)βˆ’Ξ”r​(π²βˆ’)βˆ’#​(rβˆ©π•).\displaystyle=\Delta_{A}(\mathbf{y}_{-})-\Delta_{r}(\mathbf{y}_{-})-\#(r\cap\mathbb{X}).

Thus, it remains to evaluate Ξ”r​(π²βˆ’),Ξ”M​(π²βˆ’)\Delta_{r}(\mathbf{y}_{-}),\Delta_{M}(\mathbf{y}_{-}), and Ξ”A​(π²βˆ’)\Delta_{A}(\mathbf{y}_{-}) based on the location of π²βˆ’\mathbf{y}_{-}. To do so, we perform casework based on the location of component ii of π²βˆ’\mathbf{y}_{-}. We split the nn points on the iith circle into four special markings 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D}, as shown on the right of FigureΒ 3. The second, third, and fourth columns of TableΒ 1 show the change in Ξ”r​(π²βˆ’)\Delta_{r}(\mathbf{y}_{-}). Note that the change Ξ”r​(π²βˆ’)\Delta_{r}(\mathbf{y}_{-}) does not depend on whether we have a right or left rectangle.

To compute the change in the Maslov grading, which is the fourth column of TableΒ 1, we use the equation Ξ”M​(𝐱+,π²βˆ’)=Ξ”M​(π²βˆ’)βˆ’2​Δr​(π²βˆ’)βˆ’1\Delta_{M}(\mathbf{x}_{+},\mathbf{y}_{-})=\Delta_{M}(\mathbf{y}_{-})-2\Delta_{r}(\mathbf{y}_{-})-1. The change in the Alexander grading is the fifth column, which for rows labeled 𝐀,𝐁,𝐃\mathbf{A},\mathbf{B},\mathbf{D} is computed using the upper bound Ξ”A​(𝐱+,π²βˆ’)≀ΔA​(π²βˆ’)βˆ’Ξ”r​(π²βˆ’)\Delta_{A}(\mathbf{x}_{+},\mathbf{y}_{-})\leq\Delta_{A}(\mathbf{y}_{-})-\Delta_{r}(\mathbf{y}_{-}) and for row 𝐂\mathbf{C}, we note that #​(rβˆ©π•)β‰₯1\#(r\cap\mathbb{X})\geq 1, so we get a stronger upper bound of Ξ”A​(𝐱+,π²βˆ’)≀ΔA​(π²βˆ’)βˆ’Ξ”r​(π²βˆ’)βˆ’1\Delta_{A}(\mathbf{x}_{+},\mathbf{y}_{-})\leq\Delta_{A}(\mathbf{y}_{-})-\Delta_{r}(\mathbf{y}_{-})-1.

Ξ”r​(π²βˆ’)\Delta_{r}(\mathbf{y}_{-}) Ξ”M​(π²βˆ’)\Delta_{M}(\mathbf{y}_{-}) Ξ”A​(π²βˆ’)\Delta_{A}(\mathbf{y}_{-}) Ξ”M​(𝐱+,π²βˆ’)\Delta_{M}(\mathbf{x}_{+},\mathbf{y}_{-}) Ξ”A​(𝐱+,π²βˆ’)\Delta_{A}(\mathbf{x}_{+},\mathbf{y}_{-})
=#​(pβˆ©π•†)βˆ’#​(rβˆ©π•†)=\#(p\cap\mathbb{O})-\#(r\cap\mathbb{O}) =M​(π²βˆ’)βˆ’M​(𝐲+)=M(\mathbf{y}_{-})-M(\mathbf{y}_{+}) =A​(π²βˆ’)βˆ’A​(𝐲+)=A(\mathbf{y}_{-})-A(\mathbf{y}_{+})
A βˆ’1-1 βˆ’1-1 βˆ’1-1 0 ≀0\leq 0
B 0 11 0 0 ≀0\leq 0
C 0 11 11 0 ≀0\leq 0
D 0 11 0 0 ≀0\leq 0
Table 1. Local changes depending on whether π²βˆ’βˆ©Ξ³i\mathbf{y}_{-}\cap\gamma_{i} lies on 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D}.

Now we show the computation for the change Ξ”M​(π²βˆ’)=M​(π²βˆ’)βˆ’M​(𝐲+)\Delta_{M}(\mathbf{y}_{-})=M(\mathbf{y}_{-})-M(\mathbf{y}_{+}). Let π•†βˆ’,π•βˆ’\mathbb{O}_{-},\mathbb{X}_{-} denote the set of OO’s and XX’s in π”Ύβˆ’\mathbb{G}_{-}, and let 𝕆+,𝕏+\mathbb{O}_{+},\mathbb{X}_{+} be the set of OO’s and XX’s in 𝔾+\mathbb{G}_{+}.

π’₯​(π²βˆ’,π²βˆ’)βˆ’π’₯​(𝐲+,𝐲+)\displaystyle\mathcal{J}(\mathbf{y}_{-},\mathbf{y}_{-})-\mathcal{J}(\mathbf{y}_{+},\mathbf{y}_{+}) =0,\displaystyle=0,
π’₯​(π²βˆ’,π•†βˆ’)βˆ’π’₯​(𝐲+,𝕆+)\displaystyle\mathcal{J}(\mathbf{y}_{-},\mathbb{O}_{-})-\mathcal{J}(\mathbf{y}_{+},\mathbb{O}_{+}) ={1ifΒ π²βˆ’βˆ©Ξ³iβˆˆπ€,0if 𝐲∩γi∈𝐁βˆͺ𝐂βˆͺ𝐃,\displaystyle=\begin{cases}1&\text{if $\mathbf{y}_{-}\cap\gamma_{i}\in\mathbf{A}$,}\\ 0&\text{if $\mathbf{y}\cap\gamma_{i}\in\mathbf{B}\cup\mathbf{C}\cup\mathbf{D}$,}\end{cases}
π’₯​(π•†βˆ’,π•†βˆ’)βˆ’π’₯​(𝕆+,𝕆+)\displaystyle\mathcal{J}(\mathbb{O}_{-},\mathbb{O}_{-})-\mathcal{J}(\mathbb{O}_{+},\mathbb{O}_{+}) =1.\displaystyle=1.

Now using LemmaΒ 2.6, we compute

Ξ”M​(π²βˆ’)=M​(π²βˆ’)βˆ’M​(𝐲+)\displaystyle\Delta_{M}(\mathbf{y}_{-})=M(\mathbf{y}_{-})-M(\mathbf{y}_{+}) =π’₯​(π²βˆ’βˆ’π•†βˆ’,π²βˆ’βˆ’π•†βˆ’)βˆ’π’₯​(𝐲+βˆ’π•†+,𝐲+βˆ’π•†+)\displaystyle=\mathcal{J}(\mathbf{y}_{-}-\mathbb{O}_{-},\mathbf{y}_{-}-\mathbb{O}_{-})-\mathcal{J}(\mathbf{y}_{+}-\mathbb{O}_{+},\mathbf{y}_{+}-\mathbb{O}_{+})
={βˆ’1ifΒ π²βˆ’βˆ©Ξ³iβˆˆπ€,1if 𝐲∩γi∈𝐁βˆͺ𝐂βˆͺ𝐃.\displaystyle=\begin{cases}-1&\text{if $\mathbf{y}_{-}\cap\gamma_{i}\in\mathbf{A}$,}\\ 1&\text{if $\mathbf{y}\cap\gamma_{i}\in\mathbf{B}\cup\mathbf{C}\cup\mathbf{D}$}.\end{cases}

To compute Ξ”A​(π²βˆ’)=A​(π²βˆ’)βˆ’A​(𝐲+)\Delta_{A}(\mathbf{y}_{-})=A(\mathbf{y}_{-})-A(\mathbf{y}_{+}), we compute M𝕏​(π²βˆ’)βˆ’M𝕏​(𝐲+)M_{\mathbb{X}}(\mathbf{y}_{-})-M_{\mathbb{X}}(\mathbf{y}_{+}) and use DefinitionΒ 2.7 of the Alexander function. Following the same computation as above, we find that

M𝕏​(π²βˆ’)βˆ’M𝕏​(𝐲+)={βˆ’1ifΒ π²βˆ’βˆ©Ξ³iβˆˆπ‚,1if 𝐲∩γiβˆˆπ€βˆͺ𝐁βˆͺ𝐃.M_{\mathbb{X}}(\mathbf{y}_{-})-M_{\mathbb{X}}(\mathbf{y}_{+})=\begin{cases}-1&\text{if $\mathbf{y}_{-}\cap\gamma_{i}\in\mathbf{C}$,}\\ 1&\text{if $\mathbf{y}\cap\gamma_{i}\in\mathbf{A}\cup\mathbf{B}\cup\mathbf{D}$}.\end{cases}

Then

Ξ”A​(π²βˆ’)=A​(π²βˆ’)βˆ’A​(𝐲+)\displaystyle\Delta_{A}(\mathbf{y}_{-})=A(\mathbf{y}_{-})-A(\mathbf{y}_{+}) =12​((M​(π²βˆ’)βˆ’M​(𝐲+))βˆ’(M𝕏​(π²βˆ’)βˆ’M𝕏​(𝐲+)))\displaystyle=\frac{1}{2}((M(\mathbf{y}_{-})-M(\mathbf{y}_{+}))-(M_{\mathbb{X}}(\mathbf{y}_{-})-M_{\mathbb{X}}(\mathbf{y}_{+})))
={βˆ’1ifΒ π²βˆ’βˆ©Ξ³iβˆˆπ€,0ifΒ π²βˆ’βˆ©Ξ³i∈𝐁βˆͺ𝐃,1ifΒ π²βˆ’βˆ©Ξ³iβˆˆπ‚.\displaystyle=\begin{cases}-1&\text{if $\mathbf{y}_{-}\cap\gamma_{i}\in\mathbf{A}$},\\ 0&\text{if $\mathbf{y}_{-}\cap\gamma_{i}\in\mathbf{B}\cup\mathbf{D}$},\\ 1&\text{if $\mathbf{y}_{-}\cap\gamma_{i}\in\mathbf{C}$}.\end{cases}

This proves that π’žβˆ’\mathcal{C}_{-} is a β„€\mathbb{Z}-graded, β„€\mathbb{Z}-filtered chain map.

The argument to show that π’ž+\mathcal{C}_{+} is homogeneous of degree (βˆ’2,βˆ’1)(-2,-1) is similar. In this case, we define the constants

Ξ”r​(𝐲+)\displaystyle\Delta_{r}(\mathbf{y}_{+}) =#​(pβˆ©π•†)βˆ’#​(rβˆ©π•†),\displaystyle=\#(p\cap\mathbb{O})-\#(r\cap\mathbb{O}),
Ξ”M​(𝐲+)\displaystyle\Delta_{M}(\mathbf{y}_{+}) =M​(𝐲+)βˆ’M​(π²βˆ’),\displaystyle=M(\mathbf{y}_{+})-M(\mathbf{y}_{-}),
Ξ”A​(𝐲+)\displaystyle\Delta_{A}(\mathbf{y}_{+}) =A​(𝐲+)βˆ’A​(π²βˆ’),\displaystyle=A(\mathbf{y}_{+})-A(\mathbf{y}_{-}),

and similar computations give that

Ξ”M​(𝐱+,π²βˆ’)\displaystyle\Delta_{M}(\mathbf{x}_{+},\mathbf{y}_{-}) =M​(V1O1​(p)​⋯​VnOn​(p)⋅𝐲+)βˆ’M​(π±βˆ’)=Ξ”M​(𝐲+)βˆ’2​Δr​(𝐲+)βˆ’1,\displaystyle=M(V_{1}^{O_{1}(p)}\cdots V_{n}^{O_{n}(p)}\cdot\mathbf{y}_{+})-M(\mathbf{x}_{-})=\Delta_{M}(\mathbf{y}_{+})-2\Delta_{r}(\mathbf{y}_{+})-1,
Ξ”A​(𝐱+,π²βˆ’)\displaystyle\Delta_{A}(\mathbf{x}_{+},\mathbf{y}_{-}) =A​(V1O1​(p)​⋯​VnOn​(p)⋅𝐲+)βˆ’A​(π±βˆ’)=Ξ”A​(𝐲+)βˆ’Ξ”r​(π²βˆ’)βˆ’#​(rβˆ©π•).\displaystyle=A(V_{1}^{O_{1}(p)}\cdots V_{n}^{O_{n}(p)}\cdot\mathbf{y}_{+})-A(\mathbf{x}_{-})=\Delta_{A}(\mathbf{y}_{+})-\Delta_{r}(\mathbf{y}_{-})-\#(r\cap\mathbb{X}).

The table recording each of the above constants depending on the location of 𝐲+∩βi\mathbf{y}_{+}\cap\beta_{i} is shown in TableΒ 2. For rows A, B, D, we note that #​(rβˆ©π•)β‰₯1\#(r\cap\mathbb{X})\geq 1, so we get a stronger upper bound of Ξ”A​(π±βˆ’,𝐲+)≀ΔA​(𝐲+)βˆ’Ξ”r​(𝐲+)βˆ’1\Delta_{A}(\mathbf{x}_{-},\mathbf{y}_{+})\leq\Delta_{A}(\mathbf{y}_{+})-\Delta_{r}(\mathbf{y}_{+})-1.

Ξ”r​(𝐲+)\Delta_{r}(\mathbf{y}_{+}) Ξ”M​(𝐲+)\Delta_{M}(\mathbf{y}_{+}) Ξ”A​(𝐲+)\Delta_{A}(\mathbf{y}_{+}) Ξ”M​(π±βˆ’,𝐲+)\Delta_{M}(\mathbf{x}_{-},\mathbf{y}_{+}) Ξ”A​(π±βˆ’,𝐲+)\Delta_{A}(\mathbf{x}_{-},\mathbf{y}_{+})
=#​(pβˆ©π•†)βˆ’#​(rβˆ©π•†)=\#(p\cap\mathbb{O})-\#(r\cap\mathbb{O}) =M​(𝐲+)βˆ’M​(π²βˆ’)=M(\mathbf{y}_{+})-M(\mathbf{y}_{-}) =A​(𝐲+)βˆ’A​(π²βˆ’)=A(\mathbf{y}_{+})-A(\mathbf{y}_{-})
A 11 11 11 βˆ’2-2 β‰€βˆ’1\leq-1
B 0 βˆ’1-1 0 βˆ’2-2 β‰€βˆ’1\leq-1
C 0 βˆ’1-1 βˆ’1-1 βˆ’2-2 β‰€βˆ’1\leq-1
D 0 βˆ’1-1 0 βˆ’2-2 β‰€βˆ’1\leq-1
Table 2. Local changes depending on whether 𝐲+∩βi\mathbf{y}_{+}\cap\beta_{i} lies on 𝐀,𝐁,𝐂,𝐃\mathbf{A},\mathbf{B},\mathbf{C},\mathbf{D}.

This concludes the proof. ∎

The remaining part of the proof of TheoremΒ 1.2 closely follows the proof of the Proof of Proposition 6.1.1 and Lemma 5.1.6 in [OSS15].

Proof of TheoremΒ 1.2.

By LemmaΒ 3.3 and LemmaΒ 3.4, π’žβˆ’\mathcal{C}_{-} is a β„€\mathbb{Z}-graded, β„€\mathbb{Z}-filtered chain map and π’ž+\mathcal{C}_{+} is a homogeneous chain map of degree (βˆ’2,βˆ’1)(-2,-1). Now it remains the verify that π’žβˆ’βˆ˜π’ž+\mathcal{C}_{-}\circ\mathcal{C}_{+} and π’ž+βˆ˜π’žβˆ’\mathcal{C}_{+}\circ\mathcal{C}_{-} are filtered chain-homotopic to multiplication by V1V_{1}.

Define the homotopy operators

β„‹βˆ’:π’’β€‹π’žβˆ’β€‹(π”Ύβˆ’)β†’π’’β€‹π’žβˆ’β€‹(π”Ύβˆ’)​and​ℋ+:π’’β€‹π’žβˆ’β€‹(𝔾+)β†’π’’β€‹π’žβˆ’β€‹(𝔾+)\mathcal{H}_{-}:\mathcal{GC}^{-}(\mathbb{G}_{-})\to\mathcal{GC}^{-}(\mathbb{G}_{-})\;\;\text{and}\;\;\mathcal{H}_{+}:\mathcal{GC}^{-}(\mathbb{G}_{+})\to\mathcal{GC}^{-}(\mathbb{G}_{+})

by counting certain hexagons (see DefinitionΒ 3.2) from a given point:

β„‹βˆ’β€‹(π±βˆ’)\displaystyle\mathcal{H}_{-}(\mathbf{x}_{-}) =βˆ‘π²βˆ’βˆˆπ’β€‹(π”Ύβˆ’)βˆ‘h∈Hexs,t∘⁑(π±βˆ’,π²βˆ’)V1O1​(h)​⋯​VnOn​(h)β‹…π²βˆ’,\displaystyle=\sum_{\mathbf{y}_{-}\in\mathbf{S}(\mathbb{G}_{-})}\sum_{h\in\operatorname{Hex}_{s,t}^{\circ}(\mathbf{x}_{-},\mathbf{y}_{-})}V_{1}^{O_{1}(h)}\cdots V_{n}^{O_{n}(h)}\cdot\mathbf{y}_{-},
β„‹+​(𝐱+)\displaystyle\mathcal{H}_{+}(\mathbf{x}_{+}) =βˆ‘π²+βˆˆπ’β€‹(𝔾+)βˆ‘h∈Hext,s∘⁑(𝐱+,𝐲+)V1O1​(h)​⋯​VnOn​(h)⋅𝐲+.\displaystyle=\sum_{\mathbf{y}_{+}\in\mathbf{S}(\mathbb{G}_{+})}\sum_{h\in\operatorname{Hex}_{t,s}^{\circ}(\mathbf{x}_{+},\mathbf{y}_{+})}V_{1}^{O_{1}(h)}\cdots V_{n}^{O_{n}(h)}\cdot\mathbf{y}_{+}.

We claim that β„‹+\mathcal{H}_{+} (resp. β„‹βˆ’\mathcal{H}_{-}) is a chain homotopy operator from π’ž+βˆ˜π’žβˆ’\mathcal{C}_{+}\circ\mathcal{C}_{-} (resp. π’žβˆ’βˆ˜π’ž+\mathcal{C}_{-}\circ\mathcal{C}_{+}) to V1V_{1}. It is clear that β„‹+\mathcal{H}_{+} and β„‹βˆ’\mathcal{H}_{-} are 𝔽​[V1,…,Vn]\mathbb{F}[V_{1},\ldots,V_{n}]-module homomorphisms.

Now we show that β„‹+\mathcal{H}_{+} is homogeneous of degree (βˆ’1,0)(-1,0), i.e. β„‹+​(β„±sβ€‹π’žd)βŠ†β„±sβ€‹π’žd+1\mathcal{H}_{+}(\mathcal{F}_{s}\mathcal{C}_{d})\subseteq\mathcal{F}_{s}\mathcal{C}_{d+1}. For an empty hexagon h∈Hexs,t∘⁑(𝐱+,𝐲+)h\in\operatorname{Hex}_{s,t}^{\circ}(\mathbf{x}_{+},\mathbf{y}_{+}), there is a corresponding empty rectangle rr from 𝐱+\mathbf{x}_{+} to 𝐲+\mathbf{y}_{+} that contains one more XX-marking that hh and has the same number of OO-markings as hh. Then we can compute

M​(V1O1​(h)​⋯​V1On​(h)⋅𝐲+)βˆ’M​(𝐱+)\displaystyle M(V_{1}^{O_{1}(h)}\cdots V_{1}^{O_{n}(h)}\cdot\mathbf{y}_{+})-M(\mathbf{x}_{+}) =M​(𝐲+)βˆ’M​(𝐱+)βˆ’2​#​(hβˆ©π•†)\displaystyle=M(\mathbf{y}_{+})-M(\mathbf{x}_{+})-2\#(h\cap\mathbb{O})
(Property (M-2)) =M​(𝐲+)βˆ’M​(𝐱+)βˆ’2​#​(rβˆ©π•†)=βˆ’1.\displaystyle=M(\mathbf{y}_{+})-M(\mathbf{x}_{+})-2\#(r\cap\mathbb{O})=-1.

Now we compute the difference in Alexander gradings:

A​(V1O1​(h)​⋯​V1On​(h)⋅𝐲+)βˆ’A​(𝐱+)=A​(𝐲+)βˆ’A​(𝐱+)βˆ’#​(hβˆ©π•†)=βˆ’#​(hβˆ©π•)≀0.\displaystyle A(V_{1}^{O_{1}(h)}\cdots V_{1}^{O_{n}(h)}\cdot\mathbf{y}_{+})-A(\mathbf{x}_{+})=A(\mathbf{y}_{+})-A(\mathbf{x}_{+})-\#(h\cap\mathbb{O})=-\#(h\cap\mathbb{X})\leq 0.

We can similarly show that β„‹βˆ’\mathcal{H}_{-} is homogeneous of degree (βˆ’1,0)(-1,0).

It remains to show the β„‹+\mathcal{H}_{+} and β„‹βˆ’\mathcal{H}_{-} satisfy the homotopy formulas

(5) βˆ‚βˆ’βˆ˜β„‹++β„‹+βˆ˜βˆ‚βˆ’\displaystyle\partial^{-}\circ\mathcal{H}_{+}+\mathcal{H}_{+}\circ\partial^{-} =π’ž+βˆ˜π’žβˆ’βˆ’V1,\displaystyle=\mathcal{C}_{+}\circ\mathcal{C}_{-}-V_{1},
(6) βˆ‚βˆ’βˆ˜β„‹βˆ’+β„‹βˆ’βˆ˜βˆ‚βˆ’\displaystyle\partial^{-}\circ\mathcal{H}_{-}+\mathcal{H}_{-}\circ\partial^{-} =π’žβˆ’βˆ˜π’ž+βˆ’V1.\displaystyle=\mathcal{C}_{-}\circ\mathcal{C}_{+}-V_{1}.

Since we are working over 𝔽=β„€/2​℀\mathbb{F}=\mathbb{Z}/2\mathbb{Z}, to show equationΒ (5), it will be more convenient to show

(7) βˆ‚βˆ’βˆ˜β„‹++β„‹+βˆ˜βˆ‚βˆ’+π’ž+βˆ˜π’žβˆ’=V1.\partial^{-}\circ\mathcal{H}_{+}+\mathcal{H}_{+}\circ\partial^{-}+\mathcal{C}_{+}\circ\mathcal{C}_{-}=V_{1}.

The left side ofΒ (7) can be expanded as

(βˆ‚βˆ’βˆ˜β„‹++β„‹+βˆ˜βˆ‚βˆ’+π’ž+βˆ˜π’žβˆ’)​(𝐱+)=βˆ‘π³+βˆˆπ’β€‹(𝔾)βˆ‘ΟˆβˆˆΟ€β€‹(𝐱+,𝐳+)N​(ψ)β‹…V1O1​(ψ)​⋯​VnOn​(ψ)⋅𝐳+,(\partial^{-}\circ\mathcal{H}_{+}+\mathcal{H}_{+}\circ\partial^{-}+\mathcal{C}_{+}\circ\mathcal{C}_{-})(\mathbf{x}_{+})=\sum_{\mathbf{z}_{+}\in\mathbf{S}(\mathbb{G})}\sum_{\psi\in\pi(\mathbf{x_{+}},\mathbf{z}_{+})}N(\psi)\cdot V_{1}^{O_{1}(\psi)}\cdots V_{n}^{O_{n}(\psi)}\cdot\mathbf{z}_{+},

where N​(ψ)N(\psi) is the number of ways of decomposing ψ\psi as either:

  • β€’

    ψ=rβˆ—h\psi=r*h, where rr is an empty rectangle and hh is an empty hexagon,

  • β€’

    ψ=hβˆ—r\psi=h*r, where hh is an empty hexagon and rr is an empty rectangle,

  • β€’

    ψ=pβˆ—pβ€²\psi=p*p^{\prime}, where pp is an empty pentagon from 𝔾+\mathbb{G}_{+} to π”Ύβˆ’\mathbb{G}_{-} and pβ€²p^{\prime} is a pentagon from π”Ύβˆ’\mathbb{G}_{-} to 𝔾+\mathbb{G}_{+}.

There are three cases of ΟˆβˆˆΟ€β€‹(𝐱+,𝐳+)\psi\in\pi(\mathbf{x}_{+},\mathbf{z}_{+}).

  • (H-1)

    𝐱+βˆ–(𝐱+∩𝐳+)\mathbf{x}_{+}\setminus(\mathbf{x}_{+}\cap\mathbf{z}_{+}) consists of 44 points. The two decompositions of ψ\psi are r1βˆ—h1r_{1}*h_{1} and h2βˆ—r2h_{2}*r_{2}, where r1r_{1} and r2r_{2} are rectangles with the same support and h1h_{1} and h2h_{2} are hexagons with the same support. See case (H-1) of FigureΒ 7. Therefore, N​(ψ)=2N(\psi)=2.

  • (H-2)

    𝐱+βˆ–(𝐱+∩𝐳+)\mathbf{x}_{+}\setminus(\mathbf{x}_{+}\cap\mathbf{z}_{+}) consists of 33 points. In this case, ψ\psi has eight corners. Either seven of the corners are 90∘90^{\circ} and one is 270∘270^{\circ}, or five are 90∘90^{\circ} and three are 270∘270^{\circ}. In the first case, cutting at the 270∘270^{\circ} corner gives two decompositions of ψ\psi, and in the second case, the at one of the corners labeled ss or tt we can cut in two ways. See case (H-2) of FigureΒ 7. In all cases, N​(ψ)=2N(\psi)=2.

    Refer to caption
    Figure 7. Hexagon domain decompositions. The black dots are contained in 𝐱+\mathbf{x}_{+} and the white dots are contained in 𝐳+\mathbf{z}_{+}.
  • (H-3)

    𝐱=𝐳\mathbf{x}=\mathbf{z}. In this case, ψ\psi is supported inside an annulus between Ξ²i\beta_{i} and one of the consecutive vertical circles Ξ²iβˆ’1\beta_{i-1} and Ξ²i+1\beta_{i+1}. In this case, N​(ψ)=1N(\psi)=1 and decomposes uniquely into one of rectangle-hexagon, hexagon-rectangle, or pentagon-pentagon, depending on the placement of 𝐱+\mathbf{x}_{+}, see the next paragraph. In each case, ψ\psi contains O1O_{1}, so the left side ofΒ (7) contributes V1​𝐱+V_{1}\mathbf{x}_{+}. This agrees with the right side ofΒ (7).

    To describe the decomposition of ψ\psi more specifically, we perform casework on the placement of 𝐱+∩βi\mathbf{x}_{+}\cap\beta_{i}. If 𝐱+∩βi\mathbf{x}_{+}\cap\beta_{i} is on the short arc between ss and tt, then the annulus to the east of Ξ²i\beta_{i} and west of Ξ²i+1\beta_{i+1} has a unique decomposition. If 𝐱+∩βi\mathbf{x}_{+}\cap\beta_{i} is not on the shorter arc connecting ss and tt, then the annulus to the west of Ξ²i\beta_{i} and to the east of Ξ²iβˆ’1\beta_{i-1} has a unique decomposition (cf. pages 119 and 120 of [OSS15]).

The verification of the homotopy formula for β„‹βˆ’\mathcal{H}_{-} inΒ (6) works similarly. ∎

4. Crossing-change invariant

In this section, we obtain a knot invariant from the crossing change maps in TheoremΒ 1.2 and show that it is a lower bound on the unknotting number.

Let KK and Kβ€²K^{\prime} be two knots, which for our purposes are always in S3S^{3}. The Gordian distance u​(K,Kβ€²)u(K,K^{\prime}) between KK and Kβ€²K^{\prime} is the minimum number of crossing changes required to change KK into Kβ€²K^{\prime}. We can assume there is a planar diagram KK such that after switching mm crossings, it is isotopic to Kβ€²K^{\prime}. Label the unknotting sequence by K=K0,K1,…,Kmβˆ’1,Km=Kβ€²K=K_{0},K_{1},\ldots,K_{m-1},K_{m}=K^{\prime}, where m=u​(K,Kβ€²)m=u(K,K^{\prime}) and each consecutive pair of knots differs by a crossing change. For every i=0,…,mi=0,\ldots,m, we can represent KiK_{i} by a grid 𝔾i\mathbb{G}_{i} such that every grid has the same size, and for i=0,…,mβˆ’1i=0,\ldots,m-1, the grids 𝔾i\mathbb{G}_{i} and 𝔾i+1\mathbb{G}_{i+1} differ by a cross-commutation of columns (cf. [OSS15], Section 6.2). By TheoremΒ 1.2, there exist maps

𝔣i+:π’’β€‹π’žβˆ’β€‹(𝔾i)β†’π’’β€‹π’žβˆ’β€‹(𝔾i+1)​and​𝔣iβˆ’:π’’β€‹π’žβˆ’β€‹(𝔾i+1)β†’π’’β€‹π’žβˆ’β€‹(𝔾i)\mathfrak{f}_{i}^{+}:\mathcal{GC}^{-}(\mathbb{G}_{i})\to\mathcal{GC}^{-}(\mathbb{G}_{i+1})\;\;\text{and}\;\;\mathfrak{f}_{i}^{-}:\mathcal{GC}^{-}(\mathbb{G}_{i+1})\to\mathcal{GC}^{-}(\mathbb{G}_{i})

such that 𝔣i+βˆ˜π”£iβˆ’\mathfrak{f}_{i}^{+}\circ\mathfrak{f}_{i}^{-} and 𝔣iβˆ’βˆ˜π”£i+\mathfrak{f}_{i}^{-}\circ\mathfrak{f}_{i}^{+} are filtered chain-homotopic to multiplication by V1V_{1}. So, there exist maps

𝔣+:π’’β€‹π’žβˆ’β€‹(𝔾0)β†’π’’β€‹π’žβˆ’β€‹(𝔾m)​andβ€‹π”£βˆ’:π’’β€‹π’žβˆ’β€‹(𝔾m)β†’π’’β€‹π’žβˆ’β€‹(𝔾0)\mathfrak{f}^{+}:\mathcal{GC}^{-}(\mathbb{G}_{0})\to\mathcal{GC}^{-}(\mathbb{G}_{m})\;\;\text{and}\;\;\mathfrak{f}^{-}:\mathcal{GC}^{-}(\mathbb{G}_{m})\to\mathcal{GC}^{-}(\mathbb{G}_{0})

such that 𝔣i+βˆ˜π”£iβˆ’\mathfrak{f}_{i}^{+}\circ\mathfrak{f}_{i}^{-} and π”£βˆ’βˆ˜π”£+\mathfrak{f}^{-}\circ\mathfrak{f}^{+} are filtered chain-homotopic to multiplication by V1mV_{1}^{m}.

We make the following definition.

Definition 4.1.

Let KK and Kβ€²K^{\prime} be two knots, and let 𝔾\mathbb{G} and 𝔾′\mathbb{G}^{\prime} be grids of the same size representing KK and Kβ€²K^{\prime} respectively. Consider all pairs of maps

𝔣+:π’’β€‹π’žβˆ’β€‹(𝔾)β†’π’’β€‹π’žβˆ’β€‹(𝔾′)​andβ€‹π”£βˆ’:π’’β€‹π’žβˆ’β€‹(𝔾′)β†’π’’β€‹π’žβˆ’β€‹(𝔾)\mathfrak{f}^{+}:\mathcal{GC}^{-}(\mathbb{G})\to\mathcal{GC}^{-}(\mathbb{G}^{\prime})\;\;\text{and}\;\;\mathfrak{f}^{-}:\mathcal{GC}^{-}(\mathbb{G}^{\prime})\to\mathcal{GC}^{-}(\mathbb{G})

such that 𝔣+\mathfrak{f}^{+} and π”£βˆ’\mathfrak{f}^{-} are homogeneous filtered chain maps, and 𝔣+βˆ˜π”£βˆ’\mathfrak{f}^{+}\circ\mathfrak{f}^{-} and π”£βˆ’βˆ˜π”£+\mathfrak{f}^{-}\circ\mathfrak{f}^{+} are filtered chain-homotopic to multiplication by V1mV_{1}^{m} for some positive integer mm. Let 𝔩​(K,Kβ€²)\mathfrak{l}(K,K^{\prime}) the minimal mm over all pairs 𝔣+,π”£βˆ’\mathfrak{f}^{+},\mathfrak{f}^{-} satisfying the above conditions. Let 𝔩​(K)=𝔩​(K,U)\mathfrak{l}(K)=\mathfrak{l}(K,U), where UU is the unknot.

By definition, the unknotting number u​(K)=u​(K,U)u(K)=u(K,U). The above discussion implies

Theorem 4.2.

𝔩​(K,Kβ€²)≀u​(K,Kβ€²)\mathfrak{l}(K,K^{\prime})\leq u(K,K^{\prime}) and 𝔩​(K)≀u​(K)\mathfrak{l}(K)\leq u(K).

Remark.

The invariant 𝔩​(K)\mathfrak{l}(K) can be computed for knots such that the chain homotopy type of the filtered grid complex is known, such as alternating knots. Compare Section 14.2 of [OSS15].

5. Comparison with Alishahi–Eftekhary knot invariant

Alishahi and Eftekhary define an invariant 𝔩A​E​(K)\mathfrak{l}_{AE}(K), which is a lower bound on the unknotting number u​(K)u(K), and an upper bound on the concordance invariant Ξ½+​(K)\nu^{+}(K) and also an upper bound on 𝔱^​(K)\widehat{\mathfrak{t}}(K), where 𝔱^​(K)\widehat{\mathfrak{t}}(K) is the maximum order of UU-torsion in knot Floer homology HFKβˆ’β€‹(K)\mathrm{HFK}^{-}(K), compare Definition 3.1 of [AE20]. We copy the definition of the Aliashahi–Eftekhary invariant in DefinitionΒ 5.1. For further properties of 𝔩​(K)\mathfrak{l}(K), see Theorem 1.1 and Corollary 1.2 of [AE20]. The goal of this section is to show the two definitions of 𝔩\mathfrak{l} coincide: for any two knot KβŠ‚S3K\subset S^{3}, 𝔩​(K)=𝔩A​E​(K)\mathfrak{l}(K)=\mathfrak{l}_{AE}(K).

The definition of 𝔩A​E​(K)\mathfrak{l}_{AE}(K) is in terms of a knot Floer complex CF​(K)\mathrm{CF}(K), which is obtained from a sutured manifold in their construction [AE15], a refinement of JuhΓ‘sz’s construction [Juh06]. The knot chain complex CF​(K)\mathrm{CF}(K) is a module over 𝔽​[u,w]\mathbb{F}[u,w] is β„€\mathbb{Z}–bigraded, with Maslov and Alexander gradings as defined in [OS04]. The complex CF⁑(K)\operatorname{CF}(K) is the chain homotopy type the complex CF⁑(β„‹)\operatorname{CF}(\mathcal{H}), where β„‹\mathcal{H} is a Heegaard diagram of the sutured manifold associated to the knot. The sutured manifold associated to the knot is the complement of a neighborhood of a knot in S3S^{3}, and it has two sutures on the boundary torus oriented in opposite directions.

If a knot Kβ€²βŠ‚S3K^{\prime}\subset S^{3} is obtained from the knot KβŠ‚S3K\subset S^{3} by a crossing change, Alishahi–Eftekhary prove, that there exist homogeneous chain maps

π”£βˆ’:CF​(K)β†’CF​(Kβ€²)​and​𝔣+:CF​(Kβ€²)β†’CF​(K)\mathfrak{f}^{-}:\mathrm{CF}(K)\to\mathrm{CF}(K^{\prime})\;\;\text{and}\;\;\mathfrak{f}^{+}:\mathrm{CF}(K^{\prime})\to\mathrm{CF}(K)

which preserve the Maslov grading, such that 𝔣+βˆ˜π”£βˆ’\mathfrak{f}^{+}\circ\mathfrak{f}^{-} and π”£βˆ’βˆ˜π”£+\mathfrak{f}^{-}\circ\mathfrak{f}^{+} are chain homotopic to multiplication by ww, see Theorem 2.3 of [AE20].

We define the Alishahi–Eftekhary knot invariant below.

Definition 5.1.

Given two knots K,Kβ€²K,K^{\prime}, consider all pairs of homogeneous maps

π”£βˆ’:CF​(K)β†’CF​(Kβ€²)​and​𝔣+:CF​(Kβ€²)β†’CF​(K)\mathfrak{f}^{-}:\mathrm{CF}(K)\to\mathrm{CF}(K^{\prime})\;\;\text{and}\;\;\mathfrak{f}^{+}:\mathrm{CF}(K^{\prime})\to\mathrm{CF}(K)

which are homogeneous, preserve the Maslov grading, and 𝔣+βˆ˜π”£βˆ’\mathfrak{f}^{+}\circ\mathfrak{f}^{-} and π”£βˆ’βˆ˜π”£+\mathfrak{f}^{-}\circ\mathfrak{f}^{+} are chain homotopic to multiplication by wmw^{m}. Then 𝔩A​E​(K,Kβ€²)\mathfrak{l}_{AE}(K,K^{\prime}) is the minimum nonnegative integer mm such that there exist maps π”£βˆ’,𝔣+\mathfrak{f}^{-},\mathfrak{f}^{+} satisfying the previous conditions. If UU is the unknot, let 𝔩A​E​(K)=𝔩A​E​(K,U)\mathfrak{l}_{AE}(K)=\mathfrak{l}_{AE}(K,U).

Our goal in this section is to show our invariant 𝔩​(K)\mathfrak{l}(K) from DefinitionΒ 4.1 corresponds to the Alishahi–Eftekhary 𝔩\mathfrak{l}-invariant:

Theorem 5.2.

For any two knots K,Kβ€²K,K^{\prime}, 𝔩​(K,Kβ€²)=𝔩A​E​(K,Kβ€²)\mathfrak{l}(K,K^{\prime})=\mathfrak{l}_{AE}(K,K^{\prime}). In particular, 𝔩​(K)=𝔩A​E​(K)\mathfrak{l}(K)=\mathfrak{l}_{AE}(K).

Proof.

Fix two diagrams 𝔾\mathbb{G} and 𝔾′\mathbb{G}^{\prime} of size nn for KK and Kβ€²K^{\prime} respectively. First we show that 𝔩​(K,Kβ€²)\mathfrak{l}(K,K^{\prime}) only depends on the filtered chain homotopy type of π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}) and π’’β€‹π’žβˆ’β€‹(𝔾′)\mathcal{GC}^{-}(\mathbb{G}^{\prime}). Fix two maps

𝔣+:π’’β€‹π’žβˆ’β€‹(𝔾)β†’π’’β€‹π’žβˆ’β€‹(𝔾′)​andβ€‹π”£βˆ’:π’’β€‹π’žβˆ’β€‹(𝔾′)β†’π’’β€‹π’žβˆ’β€‹(𝔾)\mathfrak{f}^{+}:\mathcal{GC}^{-}(\mathbb{G})\to\mathcal{GC}^{-}(\mathbb{G}^{\prime})\;\;\text{and}\;\;\mathfrak{f}^{-}:\mathcal{GC}^{-}(\mathbb{G}^{\prime})\to\mathcal{GC}^{-}(\mathbb{G})

such that for m=𝔩​(K,Kβ€²)m=\mathfrak{l}(K,K^{\prime}), we have 𝔣+βˆ˜π”£βˆ’β‰ƒV1m\mathfrak{f}^{+}\circ\mathfrak{f}^{-}\simeq V_{1}^{m} and π”£βˆ’βˆ˜π”£+≃V1m\mathfrak{f}^{-}\circ\mathfrak{f}^{+}\simeq V_{1}^{m}, where ≃\simeq is filtered chain homotopy equivalence. Let π’ž\mathcal{C} be a filtered chain complex in the filtered quasi-isomorphism class of π’’β€‹π’žβˆ’β€‹(𝔾′)\mathcal{GC}^{-}(\mathbb{G^{\prime}}). Since π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}) is a chain complex freely generated over a polynomial ring, filtered quasi-isomorphism and filtered chain homotopy are equivalent relations on filtered chain complexes, see for example Appendix A.8 of [OSS15]. So there are maps Ο†:π’’β€‹π’žβˆ’β€‹(𝔾′)β†’π’ž\varphi:\mathcal{GC}^{-}(\mathbb{G^{\prime}})\to\mathcal{C} and ψ:π’žβ†’π’’β€‹π’žβˆ’β€‹(𝔾)\psi:\mathcal{C}\to\mathcal{GC}^{-}(\mathbb{G}) such that Ο†βˆ˜Οˆβ‰ƒidπ’ž\varphi\circ\psi\simeq\operatorname{id}_{\mathcal{C}} and Οˆβˆ˜Ο†β‰ƒidπ’’β€‹π’žβˆ’β€‹(𝔾)\psi\circ\varphi\simeq\operatorname{id}_{\mathcal{GC}^{-}(\mathbb{G})}. In summary, we have the diagram

π’’β€‹π’žβˆ’β€‹(𝔾){\mathcal{GC}^{-}(\mathbb{G})}π’’β€‹π’žβˆ’β€‹(𝔾′){\mathcal{GC}^{-}(\mathbb{G}^{\prime})}π’ž.{\mathcal{C}.}𝔣+\scriptstyle{\mathfrak{f}^{+}}π”£βˆ’\scriptstyle{\mathfrak{f}^{-}}Ο†\scriptstyle{\varphi}ψ\scriptstyle{\psi}

Consider 𝔀+=Ο†βˆ˜π”£+\mathfrak{g}^{+}=\varphi\circ\mathfrak{f}^{+} and π”€βˆ’=π”£βˆ’βˆ˜Οˆ\mathfrak{g}^{-}=\mathfrak{f}^{-}\circ\psi. Then 𝔀+βˆ˜π”€βˆ’=(Ο†βˆ˜π”£+)∘(π”£βˆ’βˆ˜Οˆ)β‰ƒΟ†βˆ˜V1m∘ψ=V1mβˆ˜Ο†βˆ˜Οˆβ‰ƒV1m\mathfrak{g}^{+}\circ\mathfrak{g}^{-}=(\varphi\circ\mathfrak{f}^{+})\circ(\mathfrak{f}^{-}\circ\psi)\simeq\varphi\circ V_{1}^{m}\circ\psi=V_{1}^{m}\circ\varphi\circ\psi\simeq V_{1}^{m} and π”€βˆ’βˆ˜π”€βˆ’=(π”£βˆ’βˆ˜Οˆ)∘(Ο†βˆ˜π”£+)≃𝔣+βˆ˜π”£βˆ’β‰ƒV1m\mathfrak{g}^{-}\circ\mathfrak{g}^{-}=(\mathfrak{f}^{-}\circ\psi)\circ(\varphi\circ\mathfrak{f}^{+})\simeq\mathfrak{f}^{+}\circ\mathfrak{f}^{-}\simeq V_{1}^{m}. Similarly, replacing π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}) with any filtered chain complex in its filtered quasi-isomorphism class does not change mm. Therefore, 𝔩​(K,Kβ€²)\mathfrak{l}(K,K^{\prime}) only depends on the filtered chain homotopy type of π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}) and π’’β€‹π’žβˆ’β€‹(𝔾′)\mathcal{GC}^{-}(\mathbb{G}^{\prime}).

Now we show that the filtered chain homotopy class π’’β€‹π’žβˆ’β€‹(K)\mathcal{GC}^{-}(K) agrees with the chain homotopy class CF⁑(K)\operatorname{CF}(K). We have an isomorphism of filtered chain complexes π’’β€‹π’žβˆ’β€‹(𝔾)β‰…CFKβˆ’,βˆ—β‘(β„‹)\mathcal{GC}^{-}(\mathbb{G})\cong\operatorname{CFK}^{-,*}(\mathcal{H}), where β„‹\mathcal{H} is the Heegaard diagram corresponding to 𝔾\mathbb{G}, and both complexes are over the polynomial ring 𝔽​[V1,…,Vn]\mathbb{F}[V_{1},\ldots,V_{n}], see Theorem 16.4.1 in [OSS15]. By Proposition 2.3 of [MOS09], the filtered chain homotopy type of the 2​n2n-basepoint knot Floer complex CFKβˆ’,βˆ—β‘(β„‹)\operatorname{CFK}^{-,*}(\mathcal{H}) agrees with the filtered chain homotopy type of the standard knot Floer complex CFKβˆ’β‘(K)\operatorname{CFK}^{-}(K), a filtered module over 𝔽​[U]\mathbb{F}[U]. If we instead interpret CFKβˆ’β‘(K)\operatorname{CFK}^{-}(K) as a bigraded complex over the two-variable ring 𝔽​[u,v]\mathbb{F}[u,v] by letting the second variable be the filtration level, we get the knot Floer complex CF⁑(K)\operatorname{CF}(K).

By definition 𝔩A​E​(K,Kβ€²)\mathfrak{l}_{AE}(K,K^{\prime}) depends only on CF⁑(K)\operatorname{CF}(K) and CF⁑(Kβ€²)\operatorname{CF}(K^{\prime}), and as we showed above, 𝔩​(K,Kβ€²)\mathfrak{l}(K,K^{\prime}) depends only on the filtered chain homotopy classes π’’β€‹π’žβˆ’β€‹(K)\mathcal{GC}^{-}(K) and π’’β€‹π’žβˆ’β€‹(Kβ€²)\mathcal{GC}^{-}(K^{\prime}). The correspondence of filtered chain homotopy class π’’β€‹π’žβˆ’β€‹(𝔾)\mathcal{GC}^{-}(\mathbb{G}) and chain homotopy class CF⁑(K)\operatorname{CF}(K) implies 𝔩​(K,Kβ€²)=𝔩A​E​(K,Kβ€²)\mathfrak{l}(K,K^{\prime})=\mathfrak{l}_{AE}(K,K^{\prime}). ∎

References

  • [AE15] AkramΒ S. Alishahi and Eaman Eftekhary. A refinement of sutured Floer homology. J. Symplectic Geom., 13(3):609–743, 2015.
  • [AE20] Akram Alishahi and Eaman Eftekhary. Knot Floer homology and the unknotting number. Geom. Topol., 24(5):2435–2469, 2020.
  • [Juh06] AndrΓ‘s JuhΓ‘sz. Holomorphic discs and sutured manifolds. Algebr. Geom. Topol., 6:1429–1457, 2006.
  • [KM93] P.Β B. Kronheimer and T.Β S. Mrowka. Gauge theory for embedded surfaces. I. Topology, 32(4):773–826, 1993.
  • [Mil68] John Milnor. Singular points of complex hypersurfaces. Annals of Mathematics Studies, No. 61. Princeton University Press, Princeton, N.J.; University of Tokyo Press, Tokyo, 1968.
  • [MOS09] Ciprian Manolescu, Peter OzsvΓ‘th, and Sucharit Sarkar. A combinatorial description of knot Floer homology. Ann. of Math. (2), 169(2):633–660, 2009.
  • [OS04] Peter OzsvΓ‘th and ZoltΓ‘n SzabΓ³. Holomorphic disks and knot invariants. Adv. Math., 186(1):58–116, 2004.
  • [OSS15] PeterΒ S. OzsvΓ‘th, AndrΓ‘sΒ I. Stipsicz, and ZoltΓ‘n SzabΓ³. Grid homology for knots and links, volume 208 of Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 2015.
  • [Ras03] JacobΒ Andrew Rasmussen. Floer homology and knot complements. ProQuest LLC, Ann Arbor, MI, 2003. Thesis (Ph.D.)–Harvard University.
  • [Sar11] Sucharit Sarkar. Grid diagrams and the OzsvΓ‘th-SzabΓ³ tau-invariant. Math. Res. Lett., 18(6):1239–1257, 2011.
  • [SW10] Sucharit Sarkar and Jiajun Wang. An algorithm for computing some Heegaard Floer homologies. Ann. of Math. (2), 171(2):1213–1236, 2010.