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

Concordance invariant Υ\Upsilon for balanced spatial graphs using grid homology

Hajime Kubota
Abstract.

The Υ\Upsilon invariant is a concordance invariant using knot Floer homology. Földvári[1] gives a combinatorial restructure of it using grid homology. We extend the combinatorial Υ\Upsilon invariant for balanced spatial graphs. Regarding links as spatial graphs, we give an upper and lower bound for the Υ\Upsilon invariant when two links are connected by a cobordism. Also, we show that the combinatorial Υ\Upsilon invariant is a concordance invariant for knots.

1. Introduction

The τ\tau invariant and the Υ\Upsilon invariant are defined by Ozsváth, Szabó [4],[6] using knot Floer homology. The τ\tau invariant and the Υ\Upsilon invariant give homomorphisms from the (smooth) knot concordance group 𝒞\mathcal{C} to \mathbb{Z} and lower bounds for the slice genus and the unknotting number. The τ\tau invariant is known to prove the Milnor conjecture g4(Tp,q)=12(p1)(q1)g_{4}(T_{p,q})=\frac{1}{2}(p-1)(q-1) [7]. The Υ\Upsilon invariant is a family of concordance invariants Υt\Upsilon_{t} defined for every t[0,2]t\in[0,2] and the slope of the Υ\Upsilon invariant at t=0t=0 equals the value of the τ\tau invariant, so the Υ\Upsilon invariant is stronger than the τ\tau invariant. The Υ\Upsilon invariant shows that the subgroup of 𝒞\mathcal{C} generated by topologically slice knots has \mathbb{Z}^{\infty} direct summand [6].

Grid homology is a combinatorial version of knot (link) Floer homology developed by Manolescu, Ozsváth, Szabó, and Thurston [3]. The original definition of knot Floer homology needs gauge theory and pseudo-holomorphic curves. In contrast, grid homology only needs combinatorial procedures such as gazing at a planar figure called a grid diagram (figure 1) and counting some rectangles on the grid diagram.

One of the main research directions of grid homology is to give purely combinatorial proofs for the known properties of knot Floer homology. For example, Sarkar [7] gave a combinatorial reconstruction of the τ\tau invariant, which we denote τgrid\tau^{grid}, using grid homology. Sarkar also gave a purely combinatorial proof that the τgrid\tau^{grid} is a concordance invariant. As an application of it, a combinatorial proof of the Milnor conjecture was obtained. Földvári [1] defined the Υgrid\Upsilon^{grid} invariant using grid homology for t[0,2]t\in[0,2]\cap\mathbb{Q} and evaluated the change of values on crossing changes.

A spatial graph is a smooth embedding f:GS3f\colon G\to S^{3}, where GG is a one-dimensional CW-complex. We assume that spatial graphs are always oriented. A transverse spatial graph is a spatial graph such that for each vertex vv, there is a small disk DS3D\subset S^{3} whose center is vv separating incoming edges and outgoing edges. In this paper, we need one more condition: a transverse spatial graph is balanced if at each vertex, the number of incoming edges equals the number of outgoing edges.

Harvey, O’Donnol [2] extended grid homology to transverse spatial graphs. Then it is natural to explore generalizing of results of knot Floer Homology to transverse spatial graphs. Vance [8] defined the τspatial\tau^{spatial} invariant for balanced spatial graphs as an extension of Sarkar’s τgrid\tau^{grid} invariant. As an application of the τspatial\tau^{spatial}, Vance gave the bounds for the change of the τspatial\tau^{spatial} invariant of two links connected by a link cobordism.

Refer to caption
Figure 1. an example of a balanced spatial graph and a graph grid diagram

In this paper, we first define Υspatial\Upsilon^{spatial} for balanced spatial graphs for every t[0,2]t\in[0,2] as an extension of Földvári’s knot invariant Υgrid\Upsilon^{grid}. Then we give a combinatorial proof that Υspatial\Upsilon^{spatial} (and also Υgrid\Upsilon^{grid}) is a concordance invariant of knots. Our Υspatial\Upsilon^{spatial} contains more information than Földvári’s Υgrid\Upsilon^{grid} because Υgrid\Upsilon^{grid} is defined for [0,2][0,2]\cap\mathbb{Q} but our Υspatial\Upsilon^{spatial} is defined for all [0,2][0,2].

hhTo construct the Υspatial\Upsilon^{spatial} invariant, we use the t-modified chain complex tCFH(g)tCF^{-H}(g) from the grid chain complex in the same way as the way of Ozsváth, Stipsicz, and Szabó [6] and prove that the homology of the t-modified chain complex tHFH(g)tHF^{-H}(g) is (almost) independent of the choice of the graph grid diagram. To define the Υ\Upsilon invariant using knot Floer homology, Ozsváth, Stipsicz, and Szabó used a formal construction of the t-modified chain complex from filtered chain complex CCtC\mapsto C^{t}, and to see the invariance, they showed that filtered chain homotopy equivalent complexes CDC\simeq D are lifted to chain homotopy equivalent complexes CtDtC^{t}\simeq D^{t}. Applying these ideas and considering the connection between the grid chain complexes and the t-modified chain complexes enable us to define Υspatial\Upsilon^{spatial} for all [0,2][0,2] and to ensure the invariance of the Υspatial\Upsilon^{spatial}.

1.1. Main results.

For t[0,2]t\in[0,2], let \mathcal{R} be a certain based ring (see Definition 3.1). Let WtW_{t} be a two-dimensional graded vector space Wt𝔽0𝔽1+tW_{t}\cong\mathbb{F}_{0}\oplus\mathbb{F}_{-1+t}, where 𝔽=/2\mathbb{F}=\mathbb{Z}/2\mathbb{Z} and their indices describe the grading (we call t-grading). For a graded \mathcal{R}-module XX, XaX\llbracket a\rrbracket denotes a shift of XX such that Xad=Xd+aX\llbracket a\rrbracket_{d}=X_{d+a}. Then,

XWtXX1t.X\otimes W_{t}\cong X\oplus X\llbracket 1-t\rrbracket.
Theorem 1.1.

Let g,gg,g^{\prime} be two graph grid diagrams for the same balanced spatial graph. If their grid numbers are n,m(nm)n,m\ (n\geq m), which means that gg has n×nn\times n squares and gg^{\prime} has m×mm\times m squares, then as graded \mathcal{R} modules,

tHFH(g)tHFH(g)Wt(nm).tHF^{-H}(g)\cong tHF^{-H}(g^{\prime})\otimes W_{t}^{\otimes(n-m)}.
Definition 1.2.

Let gg be a graph grid diagram for balanced spatial graph ff. For t[0,1]t\in[0,1], let

Υg(t):=max{grt(x)|xtHFH(g),xishomogeneous,nontorsion},\Upsilon_{g}(t):=\mathrm{max}\{\mathrm{gr}_{t}(x)|x\in tHF^{-H}(g),x\mathrm{\ is\ homogeneous,non-torsion}\},

and for t[1,2]t\in[1,2], let Υg(t)=Υg(2t)\Upsilon_{g}(t)=\Upsilon_{g}(2-t), where ”non-torsion” means that it is non-torsion as an element of \mathcal{R}-module.

Theorem 1.3.

If g,gg,g^{\prime} are two graph grid diagrams for ff,then Υg(t)=Υg(t)\Upsilon_{g}(t)=\Upsilon_{g^{\prime}}(t).

In other words, Υg(t)\Upsilon_{g}(t) is an invariant for balanced spatial graphs. We will denote by Υf(t)\Upsilon_{f}(t).

Let L1,L2L_{1},L_{2} be two oriented links. A genus 𝐠\mathbf{g} link cobordism from L1L_{1} to L2L_{2} is an oriented genus gg surface FF smoothly embedded in S3×[0,1]S^{3}\times[0,1], such that F(S3×{0})=L1F\cap(S^{3}\times\{0\})=L_{1} and F(S3×{1})=L2F\cap(S^{3}\times\{1\})=-L_{2}. Two ll-component links are concordant if they are connected by a cobordism consisting of ll disjoint annuli.

Theorem 1.4.

Let L1,L2L_{1},L_{2} be two links of l1,l2l_{1},l_{2}-components respectively. If there is a genus gg link cobordism from L1L_{1} to L2L_{2}, then

ΥL1(t)tgt(l11)(l1l2)ΥL2(t)ΥL1(t)+tg+t(l21)+(l2l1).\Upsilon_{L_{1}}(t)-tg-t(l_{1}-1)-(l_{1}-l_{2})\leq\Upsilon_{L_{2}}(t)\leq\Upsilon_{L_{1}}(t)+tg+t(l_{2}-1)+(l_{2}-l_{1}).

Especially, if L1,L2L_{1},L_{2} are two knots K1,K2K_{1},K_{2}, then

|ΥK1(t)ΥK2(t)|tg,|\Upsilon_{K_{1}}(t)-\Upsilon_{K_{2}}(t)|\leq tg,

i.e. Υ\Upsilon is a concordance invariant for knots.

Corollary 1.5.

For t[0,1]t\in[0,1],

|ΥK(t)|tgs(K),|\Upsilon_{K}(t)|\leq t\cdot g_{s}(K),

where gs(K)g_{s}(K) is the slice genus of KK.

1.2. Some properties of Υ\Upsilon

Grid homology is a combinatorial version of knot Floer homology, so it is expected that our Υ\Upsilon satisfies the same properties as the original one of Ozsváth, Szabó [4],[6]. Some properties of the original Υ\Upsilon invariant are proved using algebraic techniques, and we show some of them in the same way.

Proposition 1.6.

Υf(0)=0\Upsilon_{f}(0)=0.

Proposition 1.7.

If two links L+L_{+} and LL_{-} differ in a crossing change, then for t[0,1]t\in[0,1],

ΥL+(t)ΥL(t)ΥL+(t)+(2t)\Upsilon_{L_{+}}(t)\leq\Upsilon_{L_{-}}(t)\leq\Upsilon_{L_{+}}(t)+(2-t)

Földvári proved this property when two links are knots and tt is rational, by constructing concrete maps between the t-modified chain complexes (see [1]). In this paper, we show that these maps are induced from some maps on grid chain complexes. Note that this Proposition is weaker than the original property in [6, Proposition 1.10].

For a balanced spatial graph ff, let 𝒰(f)\mathcal{U}(f) be a balanced spatial graph which is the disjoint union of a trivial knot and ff, regarding the trivial knot as a graph consisting of one vertex and one edge.

Proposition 1.8.

For a balanced spatial graph ff,

Υ𝒰(f)(t)=Υf(t).\Upsilon_{\mathcal{U}(f)}(t)=\Upsilon_{f}(t).

For a balanced spatial graph ff and a vertex vv, let f#v𝒪f\#_{v}\mathcal{O} denote the balanced spatial graph obtained by attaching a new unknotted, unlinked edge going from vv to vv.

Proposition 1.9.

For a balanced spatial graph ff,

Υf#v𝒪(t)=Υf(t).\Upsilon_{f\#_{v}\mathcal{O}}(t)=\Upsilon_{f}(t).

1.3. Outline of the paper

In Section 2, we review grid homology for transverse spatial graphs and Vance’s the symmetrized Alexander filtration. In Section 3, we define the combinatorial t-modified chain complex directly. In Section 4, we give an alternative definition of the t-modified chain complex from the original grid chain complex using the idea of Ozsváth, Stipsicz, and Szabó [6] and see the two definitions are equivalent. In Section 5, we give chain homotopy equivalences corresponding to the moves on grid diagrams. Using these chain homotopy equivalences, in Section 6, we prove Theorem 1.1 and Theorem 1.3. In Section 7, we consider link cobordisms on grid homology and give a proof for Theorem 1.4. Finally, in Section 8, we verify Proposition 1.6, 1.7, 1.8, and 1.9.

2. Grid homology for transverse spatial graphs

This section provides an overview of grid homology for transverse spatial graphs. See [2] for details. For grid homology for knots and links, see [3],[5].

A planar graph grid diagram gg (figure 1) is a n×nn\times n grid of squares some of which is decorated with an XX- or OO- (sometimes OO{}^{*}-) marking such that it satisfies the following conditions.

  1. (i)

    There is just one OO or OO{}^{*} on each row and column.

  2. (ii)

    There is at least one XX on each row and column.

  3. (iii)

    OO’s (or OO{}^{*}’s) and XX’s do not share the same square.

We denote the set of OO-markings by 𝕆\mathbb{O} and the set of XX-markings by 𝕏\mathbb{X}. We often use the labeling of markings as {Oi}i=1n\{O_{i}\}_{i=1}^{n} and {Xj}j=1m\{X_{j}\}_{j=1}^{m}.

For any transverse spatial graph ff, we can take graph grid diagrams representing ff. The relation between a spatial graph and representing grid diagram is the following: Drawing horizontal segments from the OO- (or OO{}^{*}-) marking to the XX-markings in each row and vertical segments from the XX-markings to the OO- (or OO{}^{*}-) marking in each column and assuming that the vertical segments always cross above the horizontal segments, then we can recover the spatial graph from the corresponding graph grid diagram. We think that OO^{*}-markings correspond to vertices, OO- and XX-markings to the interior of edges of the transverse spatial graph.

A toroidal graph grid diagram is a graph grid diagram that we think it as a diagram on the torus obtained by identifying edges in a natural way. We assume that every toroidal diagram is oriented in a natural way.

Harvey and O’Donnol showed that any two graph grid diagrams representing the same spatial graph are connected by a finite sequence of the graph grid moves. The graph grid moves are the following three moves on the torus (See also [2, Section 3.4]):\colon

  • Cyclic permutation (figure 2) permuting the rows or columns cyclically.

  • Commutation (figure 3) permuting two adjacent columns satisfying the following condition; there are vertical line segments LS1,LS2\textrm{LS}_{1},\textrm{LS}_{2} on the torus such that (1) LS1LS2\mathrm{LS}_{1}\cup\mathrm{LS}_{2} contain all the XX’s and OO’s in the two adjacent columns, (2) the projection of LS1LS2\mathrm{LS}_{1}\cup\mathrm{LS}_{2} to a single vertical circle βi\beta_{i} is βi\beta_{i} , and (3) the projection of their endpoints, (LS1)(LS2)\partial(\mathrm{LS}_{1})\cup\partial(\mathrm{LS}_{2}), to a single βi\beta_{i} is precisely two points. Permuting two rows is defined in the same way.

  • (De-)stabilization (figure 4) let gg be an n×nn\times n graph grid diagram and choose an XX-marking. Then gg^{\prime} is called a stabilization of gg if it is an (n+1)×(n+1)(n+1)\times(n+1) graph grid diagram obtained by adding one new row and column next to the XX-marking of gg, moving the XX-marking to next column, and putting new one OO-marking just above the XX-marking and one XX-marking just upper left of the XX-marking. The inverse of stabilization is called destabilization.

Refer to caption
Figure 2. cyclic permutation
Refer to caption
Figure 3. commutation, gray lines are LS1\mathrm{LS}_{1} and LS2\mathrm{LS}_{2}
Refer to caption
Figure 4. stabilization

We write the horizontal circles and vertical circles which separate the torus into squares as 𝜶={αi}i=1n\bm{\alpha}=\{\alpha_{i}\}_{i=1}^{n} and 𝜷={βj}j=1n\bm{\beta}=\{\beta_{j}\}_{j=1}^{n} respectively.

A state 𝐱\mathbf{x} of gg is a bijection 𝜶𝜷\bm{\alpha}\rightarrow\bm{\beta}. We denote by 𝐒(g)\mathbf{S}(g) the set of states of gg. We describe a state as nn points on a toroidal graph grid diagram (figure 5).

Refer to caption
Figure 5. an example of a state and a rectangle

We think that there are n×nn\times n squares on gg that is separated by 𝜶𝜷\bm{\alpha}\cup\bm{\beta}. Fix 𝐱,𝐲S\mathbf{x,y}\in S(g), a domain pp from 𝐱\mathbf{x} to 𝐲\mathbf{y} is a formal sum of the closure of squares satisfying (αp)=𝐲𝐱\partial(\partial_{\alpha}p)=\mathbf{y}-\mathbf{x} and (βp)=𝐱𝐲\partial(\partial_{\beta}p)=\mathbf{x}-\mathbf{y}, where αp\partial_{\alpha}p is the portion of the boundary of pp in the horizontal circles α1αn\alpha_{1}\cup\dots\cup\alpha_{n} and βp\partial_{\beta}p is the portion of the boundary of pp in the vertical ones. A domain pp is positive if the coefficient of any square is non-negative. Here we always assume that any domain is positive. Let π(𝐱,𝐲)\pi(\mathbf{x,y}) denote the set of domains from 𝐱\mathbf{x} to 𝐲\mathbf{y}.

Consider 𝐱,𝐲𝐒(g)\mathbf{x,y\in S}(g) that coincide with n2n-2 points. An rectangle rr from 𝐱\mathbf{x} to 𝐲\mathbf{y} is a domain that satisfies that r\partial r is the union of four segments. A rectangle rr is an empty rectangle if 𝐱Int(r)=𝐲Int(r)=\mathbf{x}\cap\mathrm{Int}(r)=\mathbf{y}\cap\mathrm{Int}(r)=\emptyset. Let Rect(𝐱,𝐲)\mathrm{Rect}^{\circ}(\mathbf{x,y}) be the set of empty rectangles from 𝐱\mathbf{x} to 𝐲\mathbf{y}.

The grid chain complex (CF(g),)(CF^{-}(g),\partial^{-}) is a module over 𝔽[U1,,Un]\mathbb{F}[U_{1},\dots,U_{n}] freely generated by 𝐒(g)\mathbf{S}(g), where 𝔽=/2\mathbb{F}=\mathbb{Z}/2\mathbb{Z} and the UiU_{i}’s are the formal variables corresponding to the OiO_{i}’s in gg. The differential \partial^{-} is defined as counting empty rectangles by

(𝐱)=𝐲𝐒(g)(rRect(𝐱,𝐲)U1O1(r)UnOn(r))𝐲,\partial^{-}(\mathbf{x})=\sum_{\mathbf{y}\in\mathbf{S}(g)}\left(\sum_{r\in\mathrm{Rect}^{\circ}(\mathbf{x,y})}U_{1}^{O_{1}(r)}\cdots U_{n}^{O_{n}(r)}\right)\mathbf{y},

where Oi(r)=1O_{i}(r)=1 if rr contains OiO_{i} and Oi(r)=0O_{i}(r)=0 otherwise for i=1,,ni=1,\dots,n.

There are two gradings for CF(g)CF^{-}(g), the Maslov grading and the Alexander grading. A planar realization of toroidal diagram gg is a planar figure obtained by cutting toroidal diagram gg along αi\alpha_{i} and βj\beta_{j} for some ii and jj, and putting on [0,n)×[0,n)2[0,n)\times[0,n)\in\mathbb{R}^{2} in a natural way. For two points (a1,a2),(b1,b2)2(a_{1},a_{2}),(b_{1},b_{2})\subset\mathbb{R}^{2}, let (a1,a2)<(b1,b2)(a_{1},a_{2})<(b_{1},b_{2}) if a1<b1a_{1}<b_{1} and a2<b2a_{2}<b_{2}. For two sets of finitely points A,B2A,B\subset\mathbb{R}^{2}, let (A,B)\mathcal{I}(A,B) be the number of pairs aA,bBa\in A,b\in B with a<ba<b and let 𝒥(A,B)=((A,B)+(B,A))/2\mathcal{J}(A,B)=(\mathcal{I}(A,B)+\mathcal{I}(B,A))/2. Let mim_{i} be the number of XX-markings in the row caontaining OiO_{i} (i=1,,ni=1,\dots,n). Then for 𝐱𝐒(g)\mathbf{x\in S}(g), the Maslov grading M(𝐱)M(\mathbf{x}) and the Alexander grading A(𝐱)A(\mathbf{x}) are defined by

(2.1) M(𝐱)=𝒥(𝐱𝕆,𝐱𝕆)+1,\displaystyle M(\mathbf{x})=\mathcal{J}(\mathbf{x}-\mathbb{O},\mathbf{x}-\mathbb{O})+1,
(2.2) A(𝐱)=𝒥(𝐱,𝕏i=1nmiOi).\displaystyle A(\mathbf{x})=\mathcal{J}(\mathbf{x},\mathbb{X}-\sum_{i=1}^{n}m_{i}O_{i}).

These two gradings are extended to the whole of CF(g)CF^{-}(g) by

(2.3) M(Ui)=2,A(Ui)=mi(i=1,,n).\displaystyle M(U_{i})=-2,A(U_{i})=-m_{i}\ (i=1,\dots,n).

Note that the Alexander grading here is the same as Vance’s definition, while the original Alexander grading by Harvey and O’Donnol has values in H1(S3f(G))H_{1}(S^{3}-f(G)), where f:GS3f\colon G\rightarrow S^{3} is the transverse spatial graph represented by gg. The Alexander grading here is given from the original one by taking the canonical homomorphism H1(S3f(G))H_{1}(S^{3}-f(G))\rightarrow\mathbb{Z} which sends the generators to 1. Also, note that the Alexander grading is not well-defined as a toroidal diagram, however, relative Alexander grading Arel(𝐱,𝐲)=A(𝐱)A(𝐲)A^{rel}(\mathbf{x,y})=A(\mathbf{x})-A(\mathbf{y}) is well-defined.

It is shown that the differential \partial^{-} drops Maslov grading by 1 and preserves or drops Alexander grading, so (CF(g),)(CF^{-}(g),\partial^{-}) is a Maslov graded, Alexander filtered chain complex [8, Proposition 3.7].

Suppose the OO-markings are labeled so that O1,,OVO_{1},\dots,O_{V} are OO^{*}-markings and OV+1,,OnO_{V+1},\dots,O_{n} are OO-markings. Let 𝒰\mathcal{U} be the minimal subcomplex of CF(g)CF^{-}(g) containing U1CF(g)UVCF(g)U_{1}CF^{-}(g)\cup\dots\cup U_{V}CF^{-}(g). Then (CF^(g),^)(\widehat{CF}(g),\widehat{\partial}) is a Maslov graded, Alexander filtered chain complex over 𝔽\mathbb{F}-vector space obtained by letting CF^(g)=CF(g)/𝒰\widehat{CF}(g)=CF^{-}(g)/\mathcal{U} and ^\widehat{\partial} be the map induced by \partial^{-}.

We denote by {m(g)}m\{\mathcal{F}^{-}_{m}(g)\}_{m\in\mathbb{Z}} (respectively {^m(g)}m\{\widehat{\mathcal{F}}_{m}(g)\}_{m\in\mathbb{Z}}) the Alexander filtration of CF(g)CF^{-}(g) (respectively CF^(g)\widehat{CF}(g)).

Definition 2.1 ([8, Definition3.10]).

For a graph grid diagram gg, define the symmetrized Alexander filtration {^m(g)}m12\{\widehat{\mathcal{F}}_{m}(g)\}_{m\in\frac{1}{2}\mathbb{Z}} to be the absolute Alexander filtration obtained by fixing the relative Alexander grading so that mmax(g)=mmin(g)m_{\mathrm{max}}(g)=-m_{\mathrm{min}}(g), where

mmax(g)\displaystyle m_{\mathrm{max}}(g) =max{m|H(^m(g)/^m1(g))0},\displaystyle=\mathrm{max}\left\{m|H_{*}(\widehat{\mathcal{F}}_{m}(g)/\widehat{\mathcal{F}}_{m-1}(g))\neq 0\right\},
mmin(g)\displaystyle m_{\mathrm{min}}(g) =min{m|H(^m(g)/^m1(g))0}.\displaystyle=\mathrm{min}\left\{m|H_{*}(\widehat{\mathcal{F}}_{m}(g)/\widehat{\mathcal{F}}_{m-1}(g))\neq 0\right\}.
Definition 2.2.

The symmetrized Alexander grading AH:𝐒(g)12A^{H}\colon\mathbf{S}(g)\rightarrow\frac{1}{2}\mathbb{Z} is determined by symmetrized Alexander filtration {^m(g)}m12\{\widehat{\mathcal{F}}_{m}(g)\}_{m\in\frac{1}{2}\mathbb{Z}} so that for 𝐱𝐒(g)\mathbf{x\in S}(g), the value of AH(𝐱)A^{H}(\mathbf{x}) is the maximal filtration level which 𝐱CF^(g)\mathbf{x}\in\widehat{CF}(g) belongs.

Let CFH(g)CF^{-H}(g) be a Maslov graded, symmetrized Alexander filtered chain complex obtained from CF(g)CF^{-}(g) by using the symmetrized Alexander grading AHA^{H} rather than the Alexander grading AA.

The homology of the associated graded object of CFH(g)CF^{-H}(g) is an invariant for balanced spatial graphs [8, Theorem 3.15].

3. The t-modified chain complex

This section provides the definition of the t-modified chain complex tCFH(g)tCF^{-H}(g). The t-modified chain complex here is the extension of one of Földvári.

First of all, we define the based ring \mathcal{R} of tCFH(g)tCF^{-H}(g).

A set AA\subset\mathbb{R} is well-ordered if any subset AAA^{\prime}\subset A has a minimal element.

Definition 3.1 ([6, Definition 3.1]).

Let 0\mathbb{R}_{\geq 0} denote the set of nonnegative real numbers. The ring of long power series \mathcal{R} is defined as follows. As an abelian group, \mathcal{R} is the group of formal sums

{αAvα|A0,Aiswellordered},\left\{\sum_{\alpha\in A}v^{\alpha}|A\subset\mathbb{R}_{\geq 0},A\ \mathrm{is\ well-ordered}\right\},

The sum in \mathcal{R} is given by the formula

(αAvα)+(βBvβ)=γC=ABABvγ,\left(\sum_{\alpha\in A}v^{\alpha}\right)+\left(\sum_{\beta\in B}v^{\beta}\right)=\sum_{\gamma\in C=A\cup B\setminus A\cap B}v^{\gamma}\ ,

and the product is given by the formula

(αAvα)(βBvβ)=γA+B#{(α,β)A×B|α+β=γ}vγ,\left(\sum_{\alpha\in A}v^{\alpha}\right)\cdot\left(\sum_{\beta\in B}v^{\beta}\right)=\sum_{\gamma\in A+B}\#\{(\alpha,\beta)\in A\times B|\alpha+\beta=\gamma\}\cdot v^{\gamma},

where

A+B={γ|γ=α+βforsomeαAandβB}.A+B=\left\{\gamma|\gamma=\alpha+\beta\ for\ some\ \alpha\in A\ and\ \beta\in B\right\}.

It is straightforward to check that the above operations are well-defined.

Then we define the t-modified chain complex tCFH(g)tCF^{-H}(g). For a domain pp as a formal sum of squares, let Oi(p)O_{i}(p) denote the coefficient of the square containing OiO_{i}, and let

|𝕆p|:=i=1nOi(p).|\mathbb{O}\cap p|:=\sum_{i=1}^{n}O_{i}(p).

We define |𝕏p||\mathbb{X}\cap p| in the same manner.

Definition 3.2.

For t[0,2]t\in[0,2], the t-modified grid complex tCFH(g)tCF^{-H}(g) is a free module over \mathcal{R} generated by 𝐒(g)\mathbf{S}(g) with the \mathcal{R}-module endmorphism t\partial^{-}_{t} defined by

(3.1) t(𝐱)=𝐲𝐒(g)(rRect(𝐱,𝐲)vt|𝕏r|+2|𝕆r|t(Oi𝕆rmi))𝐲,\partial_{t}^{-}(\mathbf{x})=\sum_{\mathbf{y}\in\mathbf{S}(g)}\left(\sum_{r\in\mathrm{Rect}^{\circ}(\mathbf{x,y})}v^{t|\mathbb{X}\cap r|+2|\mathbb{O}\cap r|-t\left(\sum_{O_{i}\in\mathbb{O}\cap r}m_{i}\right)}\right)\mathbf{y},

where mim_{i} is the number of XX-markings on the same row as OiO_{i} (i=1,,ni=1,\dots,n).

Definition 3.3.

For 𝐱𝐒(g)\mathbf{x\in S}(g) and α>0\alpha>0, the t-grading grt\mathrm{gr}_{t} is,

(3.2) grt(vα𝐱)=M(𝐱)tAH(𝐱)α,\displaystyle\mathrm{gr}_{t}(v^{\alpha}\mathbf{x})=M(\mathbf{x})-tA^{H}(\mathbf{x})-\alpha,

where AHA^{H} is the symmetrized Alexander grading.

Proposition 3.4.

tt=0\partial_{t}^{-}\circ\partial_{t}^{-}=0.

Proof.

For states 𝐱\mathbf{x} and 𝐳\mathbf{z} and a fixed domain pπ(𝐱,𝐳)p\in\pi(\mathbf{x,z}) denote by N(p)N(p) the number of ways to decompose pp as a composite of two empty rectangles r1r2r_{1}*r_{2}. Note that if p=r1r2p=r_{1}*r_{2} for some r1Rect(𝐱,𝐲),r2Rect(𝐲,𝐳)r_{1}\in\mathrm{Rect}^{\circ}(\mathbf{x,y}),r_{2}\in\mathrm{Rect}^{\circ}(\mathbf{y,z}), then

|𝕏p|\displaystyle|\mathbb{X}\cap p| =|𝕏r1|+|𝕏r2|,\displaystyle=|\mathbb{X}\cap r_{1}|+|\mathbb{X}\cap r_{2}|,
|𝕆p|\displaystyle|\mathbb{O}\cap p| =|𝕆r1|+|𝕆r2|,\displaystyle=|\mathbb{O}\cap r_{1}|+|\mathbb{O}\cap r_{2}|,
Oi𝕆pmi\displaystyle\sum_{O_{i}\in\mathbb{O}\cap p}m_{i} =Oi𝕆r1mi+Oi𝕆r2mi.\displaystyle=\sum_{O_{i}\in\mathbb{O}\cap r_{1}}m_{i}+\sum_{O_{i}\in\mathbb{O}\cap r_{2}}m_{i}.

It follows that for 𝐱𝐒(g)\mathbf{x\in S}(g),

(3.3) tt(𝐱)=𝐳𝐒(g)(pπ(𝐱,𝐳)N(p)vt|𝕏p|+2|𝕆p|t(Oi𝕆pmi))𝐳.\partial_{t}^{-}\circ\partial_{t}^{-}(\mathbf{x})=\sum_{\mathbf{z\in S}(g)}\left(\sum_{p\in\pi(\mathbf{x,z})}N(p)v^{t|\mathbb{X}\cap p|+2|\mathbb{O}\cap p|-t\left(\sum_{O_{i}\in\mathbb{O}\cap p}m_{i}\right)}\right)\mathbf{z}.

If #{𝐱(𝐱𝐳)}=4\#\{\mathbf{x\setminus(x\cap z)}\}=4 or #{𝐱(𝐱𝐳)}=3\#\{\mathbf{x\setminus(x\cap z)}\}=3, the same argument in [6, Theorem 3.2] shows that N(p)N(p) is even. If 𝐱=𝐳\mathbf{x=z}, pp is an annulus and N(p)=1N(p)=1. Since r1r_{1} and r2r_{2} are empty, this annulus has a height or width equal to 1. Such an annulus is called a thin annulus. For each 𝐱\mathbf{x}, there are 2n2n thin annuli appearing in (3.3). We can pair annuli that contain the same OO-marking. If (p1,p2)(p_{1},p_{2}) is such a pair, then |p1𝕏|=|p2𝕏||p_{1}\cap\mathbb{X}|=|p_{2}\cap\mathbb{X}| because ff is a balanced spatial graph. So all terms are canceled in pairs (we are working modulo 2). ∎

Proposition 3.5.

The map t\partial^{-}_{t} drops the t-grading by one.

Proof.

By [3, Lemma 2.5] and [8, Lemma 3.5], for rRect(𝐱,𝐲)r\in\mathrm{Rect}^{\circ}(\mathbf{x,y}),

(3.4) M(𝐱)M(𝐲)\displaystyle M(\mathbf{x})-M(\mathbf{y}) =12|r𝕆|,\displaystyle=1-2|r\cap\mathbb{O}|,
(3.5) AH(𝐱)AH(𝐲)\displaystyle A^{H}(\mathbf{x})-A^{H}(\mathbf{y}) =|r𝕏|Oir𝕆mi.\displaystyle=|r\cap\mathbb{X}|-\sum_{O_{i}\in r\cap\mathbb{O}}m_{i}.

If vt|𝕏r|+2|𝕆r|t(Oi𝕆rmi)𝐲v^{t|\mathbb{X}\cap r|+2|\mathbb{O}\cap r|-t\left(\sum_{O_{i}\in\mathbb{O}\cap r}m_{i}\right)}\mathbf{y} appears in t(𝐱)\partial^{-}_{t}(\mathbf{x}), then

grt(vt|𝕏r|+2|𝕆r|t(Oi𝕆rmi)𝐲)=M(𝐲)tAH(𝐲)(t|𝕏r|+2|𝕆r|t(Oi𝕆rmi))=M(𝐱)1+2|r𝕆|t(AH(𝐱)|r𝕏|(Oi𝕆mi))(t|𝕏r|+2|𝕆r|t(Oi𝕆rmi))=M(𝐱)tAH(𝐱)1.\displaystyle\begin{split}&\mathrm{gr}_{t}(v^{t|\mathbb{X}\cap r|+2|\mathbb{O}\cap r|-t\left(\sum_{O_{i}\in\mathbb{O}\cap r}m_{i}\right)}\mathbf{y})\\ &=M(\mathbf{y})-tA^{H}(\mathbf{y})-\left(t|\mathbb{X}\cap r|+2|\mathbb{O}\cap r|-t\left(\sum_{O_{i}\in\mathbb{O}\cap r}m_{i}\right)\right)\\ &=M(\mathbf{x})-1+2|r\cap\mathbb{O}|-t\left(A^{H}(\mathbf{x})-|r\cap\mathbb{X}|-\left(\sum_{O_{i}\in\mathbb{O}}m_{i}\right)\right)\\ &\quad-\left(t|\mathbb{X}\cap r|+2|\mathbb{O}\cap r|-t\left(\sum_{O_{i}\in\mathbb{O}\cap r}m_{i}\right)\right)\\ &=M(\mathbf{x})-tA^{H}(\mathbf{x})-1.\end{split}

Proposition 3.6.

(tCFH(g),t)(tCF^{-H}(g),\partial^{-}_{t}) is a t-graded chain complex over \mathcal{R}.

Proof.

From Proposition 3.4 and Proposition 3.5, we conclude that (tCFH(g),t)(tCF^{-H}(g),\partial^{-}_{t}) is a t-graded chain complex. ∎

Definition 3.7.

Let tCFdH(g)={αtCF(g)|grt(α)=d}tCF^{-H}_{d}(g)=\{\alpha\in tCF^{-}(g)|\mathrm{gr}_{t}(\alpha)=d\}, and we define t-modified graph grid homology of gg to be

tHFdH(g)\displaystyle tHF^{-H}_{d}(g) =Ker(t)tCFdH(g)Im(t)tCFdH(g),\displaystyle=\frac{\mathrm{Ker}(\partial_{t}^{-})\cap tCF^{-H}_{d}(g)}{\mathrm{Im}(\partial_{t}^{-})\cap tCF^{-H}_{d}(g)},
tHFH(g)\displaystyle tHF^{-H}(g) =dtHFdH(g).\displaystyle=\bigoplus_{d}tHF^{-H}_{d}(g).

4. Formal construction of the t-modified chain complex

In this section, we give an alternative definition of the t-modified chain complex using the grid chain complex. See [6, Section 4] for details.

Suppose that CC is a finitely generated, Maslov graded, Alexander filtered chain complex over 𝔽[U]\mathbb{F}[U]. Let 𝐱\mathbf{x} be a generator of CC over 𝔽[U]\mathbb{F}[U], with Maslov grading M(𝐱)M(\mathbf{x}). Since multiplication by UU drops the Maslov grading by 2, elements of Maslov grading M(𝐱)1M(\mathbf{x})-1 are linear combinations of elements of the form UM(𝐲)M(𝐱)+12𝐲U^{\frac{M(\mathbf{y})-M(\mathbf{x})+1}{2}}\mathbf{y}, where 𝐲\mathbf{y} is a generator. Then the differential on CC can be written as

(4.1) (𝐱)=𝐲c𝐱,𝐲UM(𝐲)M(𝐱)+12𝐲,\partial(\mathbf{x})=\sum_{\mathbf{y}}c_{\mathbf{x},\mathbf{y}}\cdot U^{\frac{M(\mathbf{y})-M(\mathbf{x})+1}{2}}\mathbf{y},

where c𝐱,𝐲{0,1}c_{\mathbf{x},\mathbf{y}}\in\{0,1\}.

Definition 4.1 (Definition 4.1,[6]).

For t[0,2]t\in[0,2], suppose that CC is a finitely generated, Maslov graded, Alexander filtered chain complex over 𝔽[U]\mathbb{F}[U], and let \mathcal{R} be the ring of Definition 3.1 (containing 𝔽[U]\mathbb{F}[U] by U=v2U=v^{2}). The formal t-modified chain complex CtC^{t} of CC is defined as follows:\colon

  • As an \mathcal{R}-module, Ct=C𝔽[U]C^{t}=C\otimes_{\mathbb{F}[U]}\mathcal{R}

  • For each generator 𝐱\mathbf{x} of CC, define grt(vα𝐱)=M(𝐱)tA(𝐱)α\mathrm{gr}_{t}(v^{\alpha}\mathbf{x})=M(\mathbf{x})-tA(\mathbf{x})-\alpha

  • Endow the graded module CtC^{t} with a differential

    (4.2) t(𝐱)=𝐲c𝐱,𝐲vgrt(𝐲)grt(𝐱)+1𝐲,\partial_{t}(\mathbf{x})=\sum_{\mathbf{y}}c_{\mathbf{x},\mathbf{y}}\cdot v^{\mathrm{gr}_{t}(\mathbf{y})-\mathrm{gr}_{t}(\mathbf{x})+1}\mathbf{y},

    where c𝐱,𝐲c_{\mathbf{x},\mathbf{y}} are determined by (4.1).

Definition 4.2.

For a graph grid diagram gg, let CFUH(g)CF^{-H}_{U}(g) denote the quotient chain complex CFUH(g)=CFH(g)U1==UnCF^{-H}_{U}(g)=\frac{CF^{-H}(g)}{U_{1}=\dots=U_{n}} as a Maslov graded chain complex over 𝔽[U]\mathbb{F}[U]-module.

Note that CFUH(g)CF^{-H}_{U}(g) is a Maslov graded chain complex but not an Alexander filtered chain complex because the drop in the Alexander grading of each UiU_{i} differs by (2.3). However, Definition 4.1 works for CFUH(g)CF^{-H}_{U}(g) because each generator has its Alexander grading. Therefore we can define the formal t-modified chain complex (CFUH(g))t(CF^{-H}_{U}(g))^{t}.

Proposition 4.3.

(CFUH(g))t(CF^{-H}_{U}(g))^{t} (applying Definition 4.1) is isomorphic to tCFH(g)tCF^{-H}(g) (from Definition 3.2) as graded chain complexes.

Proof.

Identifying the generators and their gradings is natural, so we only need to check that (3.1) and (4.2) are equivalent. In other words, we will check that for 𝐱,𝐲𝐒(g)\mathbf{x,y\in S}(g),

(4.3) c𝐱,𝐲=1#(Rect(𝐱,𝐲))=1,c_{\mathbf{x,y}}=1\Leftrightarrow\#(\mathrm{Rect}^{\circ}(\mathbf{x,y}))=1,

and

(4.4) grt(𝐲)grt(𝐱)+1=t|𝕏r|+2|𝕆r|t(Oi𝕆rmi).\mathrm{gr}_{t}(\mathbf{y})-\mathrm{gr}_{t}(\mathbf{x})+1=t|\mathbb{X}\cap r|+2|\mathbb{O}\cap r|-t(\sum_{O_{i}\in\mathbb{O}\cap r}m_{i}).

We will consider the case of #(Rect(𝐱,𝐲))=2\#(\mathrm{Rect}^{\circ}(\mathbf{x,y}))=2. Let r,rr,r^{\prime} be the two rectangles from 𝐱\mathbf{x} to 𝐲\mathbf{y}. Then U1O1(r)UnOn(r)𝐲U_{1}^{O_{1}(r)}\cdots U_{n}^{O_{n}(r)}\mathbf{y} and U1O1(r)UnOn(r)𝐲U_{1}^{O_{1}(r^{\prime})}\cdots U_{n}^{O_{n}(r^{\prime})}\mathbf{y} appear in (𝐱)\partial^{-}(\mathbf{x}). Because the numbers of OO- and OO^{*}-markings the two rectangles contain are the same, the quotient map CFH(g)CFUH(g)CF^{-H}(g)\to CF^{-H}_{U}(g) sends these two terms to the same element (that is UM(𝐲)M(𝐱)+12𝐲U^{\frac{M(\mathbf{y})-M(\mathbf{x})+1}{2}}\mathbf{y}), so they are canceled modulo 2. Therefore we get c𝐱,𝐲=0c_{\mathbf{x,y}}=0.

Also it is clear that #(Rect(𝐱,𝐲))=0c𝐱,𝐲=0\#(\mathrm{Rect}^{\circ}(\mathbf{x,y}))=0\Rightarrow c_{\mathbf{x,y}}=0 and that #(Rect(𝐱,𝐲))=1c𝐱,𝐲=1\#(\mathrm{Rect}^{\circ}(\mathbf{x,y}))=1\Rightarrow c_{\mathbf{x,y}}=1. Therefore (4.3) is proved.

We have (4.4) directly from (3.4) and (3.5). ∎

Next, we introduce the following proposition playing an important role in this paper.

Proposition 4.4 ([6, Proposition 4.4]).

Let f:CCf\colon C\rightarrow C^{\prime} be a Maslov graded, Alexander filtered chain map between chain complexes over 𝔽[U]\mathbb{F}[U]. There is a corresponding graded chain map ft:Ct(C)tf^{t}\colon C^{t}\rightarrow(C^{\prime})^{t}, with the following properties:\colon

  • If f:CCf\colon C\rightarrow C^{\prime} and g:CC′′g\colon C^{\prime}\rightarrow C^{\prime\prime} are two Maslov graded, Alexander filtered chain maps, then

    (gf)t=gtft(g\circ f)^{t}=g^{t}\circ f^{t}
  • If f,g:CCf,g\colon C\rightarrow C^{\prime} are chain homotopic to each other, then ftf^{t} and gtg^{t} are chain homotopic to each other. In particular, filtered chain homotopy equivalent complexes are transformed by the construction CCtC\mapsto C^{t} into homotopy equivalent complexes.

Again note that this proposition can be applied as CFUH(g)(CFUH(g))tCF^{-H}_{U}(g)\mapsto(CF^{-H}_{U}(g))^{t} even if CFUH(g)CF^{-H}_{U}(g) is just Maslov graded chain complex with AHA^{H} grading only for generators. Similar to (4.1), Maslov graded chain map f:CFUH(g)CFUH(g)f\colon CF^{-H}_{U}(g)\rightarrow CF^{-H}_{U}(g^{\prime}) can be written as

f(𝐱)=𝐲c𝐱,𝐲UM(𝐲)M(𝐱)2𝐲,f(\mathbf{x})=\sum_{\mathbf{y}}c_{\mathbf{x},\mathbf{y}}\cdot U^{\frac{M(\mathbf{y})-M(\mathbf{x})}{2}}\mathbf{y},

then the corresponding graded chain map is

ft(𝐱)=𝐲c𝐱,𝐲vgrt(𝐲)grt(𝐱)𝐲.f^{t}(\mathbf{x})=\sum_{\mathbf{y}}c_{\mathbf{x},\mathbf{y}}\cdot v^{\mathrm{gr}_{t}(\mathbf{y})-\mathrm{gr}_{t}(\mathbf{x})}\mathbf{y}.

This construction gives induced chain homotopy equivalence for the t-modified chain complexes.

5. Chain homotopy equivalences for grid chain complexes

This section provides the filtered chain homotopy equivalences for grid chain complexes connected by each graph grid move. These filtered chain homotopy equivalences are lifted into chain homotopy equivalences for the t-modified chain complexes by applying Proposition 4.4. Chain homotopy equivalences for a cyclic permutation and a commutation are already known. We mainly observe the case of stabilization.

5.1. cyclic permutation

Proposition 5.1.

If gg and gg^{\prime} are connected by a single8 cyclic permutation, then as Maslov graded, Alexander filtered chain complexes, CFH(g)CF^{-H}(g) and CFH(g)CF^{-H}(g^{\prime}) are chain homotopy equivalent.

Proof.

There is a natural bijection c:𝐒(g)𝐒(g)c\colon\mathbf{S}(g)\rightarrow\mathbf{S}(g^{\prime}). By [8, Section 3.3 and Theorem 3.15], cc induces an isomorphism of Maslov graded, symmetrized Alexander filtered chain complexes c:CFH(g)CFH(g)c\colon CF^{-H}(g)\rightarrow CF^{-H}(g^{\prime}). As c:𝐒(g)𝐒(g)c\colon\mathbf{S}(g)\rightarrow\mathbf{S}(g^{\prime}) is a bijection, there is an isomorphism c1:CFH(g)CFH(g)c^{-1}\colon CF^{-H}(g^{\prime})\rightarrow CF^{-H}(g) induced by c1c^{-1} such that c1c=idc^{-1}\circ c=\mathrm{id} and cc1=idc\circ c^{-1}=\mathrm{id}. ∎

5.2. commutation

Proposition 5.2.

If gg and gg^{\prime} are connected by a single commutation, then as Maslov graded, Alexander filtered chain complexes, CFH(g)CF^{-H}(g) and CFH(g)CF^{-H}(g^{\prime}) are chain homotopy equivalent.

Proof.

By [8, Lemma 3.21] and [3, Proposition3.2], there are chain homotopic equivalences pβγ:CF(g)CF(g)p^{\prime}_{\beta\gamma}\colon CF^{-}(g)\rightarrow CF^{-}(g^{\prime}) and pγβ:CF(g)CF(g)p^{\prime}_{\gamma\beta}\colon CF^{-}(g^{\prime})\rightarrow CF^{-}(g). Since [8, Theorem 3.15], they preserve the symmetrized Alexander filtration. ∎

5.3. stabilization

Assume that gg^{\prime} is obtained from gg by a single stabilization and that CFH(g)CF^{-H}(g) is a chain complex over 𝔽[U2,,Un]\mathbb{F}[U_{2},\dots,U_{n}] and CFH(g)CF^{-H}(g^{\prime}) is a chain complex over 𝔽[U1,,Un]\mathbb{F}[U_{1},\dots,U_{n}]. Number the OO-markings of gg^{\prime} so that O1O_{1} is the new one and O2O_{2} is the OO-marking in the row just below O1O_{1}. We will also assume that X1X_{1} lies in the same row as O1O_{1} and X2X_{2} is the XX-marking just below O1O_{1}. We denote by cc the intersection point of the new horizontal and vertical circles in gg^{\prime} (Figure 6). Note that there may be more XX-markings in the row that includes O2O_{2} (in which case O2O_{2} is an OO^{*}-marking).

Refer to caption
Figure 6. labeling of stabilized graph grid diagram

For a Maslov graded, Alexander filtered chain complex XX, let X=Xa,bX^{\prime}=X\llbracket a,b\rrbracket denote the shifted complex such that sXd=s+bXd+a\mathcal{F}_{s}X^{\prime}_{d}=\mathcal{F}_{s+b}X_{d+a}, where sXd\mathcal{F}_{s}X_{d} is the submodule of XX whose grading is dd and filtration level is ss.

Considering the point cc, we will decompose the set of states 𝐒(g)\mathbf{S}(g^{\prime}) as the disjoint union 𝐈(g)𝐍(g)\mathbf{I}(g^{\prime})\cup\mathbf{N}(g^{\prime}), where 𝐈(g)={𝐱𝐒(g)|c𝐱}\mathbf{I}(g^{\prime})=\{\mathbf{x\in S}(g^{\prime})|c\in\mathbf{x}\} and 𝐍(g)={𝐱𝐒(g)|c𝐱}\mathbf{N}(g^{\prime})=\{\mathbf{x\in S}(g^{\prime})|c\notin\mathbf{x}\}. This decomposition gives a decomposition of tCF(g)=INtCF^{-}(g^{\prime})=I\oplus N as an \mathcal{R}-module, where II and NN are the spans of 𝐈(g)\mathbf{I}(g^{\prime}) and 𝐍(g)\mathbf{N}(g^{\prime}) respectively.

There is a natural bijection between 𝐈(g)\mathbf{I}(g^{\prime}) and 𝐒(g)\mathbf{S}(g), given by

e:𝐈(g)𝐒(g),𝐱{c}𝐱.e\colon\mathbf{I}(g^{\prime})\to\mathbf{S}(g),\ \mathbf{x}\cup\{c\}\mapsto\mathbf{x}.

Let (CFH(g)[U1],)(CF^{-H}(g)[U_{1}],\partial) be a chain complex defined by CFH(g)[U1]=CFH(g)𝔽[U2,,Un]𝔽[U1,,Un]CF^{-H}(g)[U_{1}]=CF^{-H}(g)\otimes_{\mathbb{F}[U_{2},\dots,U_{n}]}\mathbb{F}[U_{1},\dots,U_{n}] with the differential =id\partial=\partial^{-}\otimes\mathrm{id}. Consider a chain complex Cone(U1U2:CFH(g)[U1]CFH(g)[U1])\mathrm{Cone}(U_{1}-U_{2}\colon CF^{-H}(g)[U_{1}]\to CF^{-H}(g)[U_{1}]).

For the stabilization invariance, we will prove the following proposition.

Proposition 5.3.

CFH(g)CF^{-H}(g^{\prime}) and Cone(U1U2:CFH(g)[U1]CFH(g)[U1])\mathrm{Cone}(U_{1}-U_{2}\colon CF^{-H}(g)[U_{1}]\to CF^{-H}(g)[U_{1}]) are chain homotopy equivalent complexes.

This can be proved in the same way as the proof of stabilization invariance of grid homology for knots and links written in [5, Sections 13.3.2 and 13.4.2]. Because we are working on Alexander filtered version, the domains which appear in the chain maps can contain any XX-markings of 𝕏\mathbb{X}, in other words, they are independent of the XX-markings. Thus we can use the same domains as in [5, Definitions 13.3.7 and 13.4.2] to show this Proposition. We construct chain homotopy equivalence using the filtered destabilization map 𝒟\mathcal{D} ([5, Definition 13.3.10]), the filtered stabilization map 𝒮\mathcal{S} ([5, Definition 13.4.3]) and chain homotopy 𝒦\mathcal{K} (Definition 5.10). Note that these maps are the answer to [5, Exercise 13.4.4].

Definition 5.4 ([5, Definition 13.3.7]).

For 𝐱𝐒(g)\mathbf{x\in S}(g^{\prime}) and 𝐲𝐈(g)\mathbf{y\in I}(g^{\prime}), a positive domain pπ(𝐱,𝐲)p\in\pi(\mathbf{x,y}) is said to be of type 𝒊𝑳\bm{iL} if it is trivial or it satisfies the following conditions:\colon

  • At each corner point in 𝐱𝐲{c}\mathbf{x\cup y}\setminus\{c\}, at least three of the four adjoining squares have the local multiplicity zero.

  • Three of four squares attaching at the corner point cc have the local multiplicity kk and at the southwest square meeting cc the local multiplicity is k1k-1.

  • 𝐲\mathbf{y} has 2k+12k+1 components that are not in 𝐱\mathbf{x}.

And, a positive domain pπ(𝐱,𝐲)p\in\pi(\mathbf{x,y}) is said to be of type 𝒊𝑹\bm{iR} if it is trivial or it satisfies the following conditions:\colon

  • At each corner point in 𝐱𝐲{c}\mathbf{x\cup y}\setminus\{c\}, at least three of the four adjoining squares have the local multiplicity zero.

  • Three of four squares attaching at the corner point cc have the local multiplicity kk and at the southeast square meeting cc the local multiplicity is k+1k+1.

  • 𝐲\mathbf{y} has 2k+22k+2 components that are not in 𝐱\mathbf{x}.

The set of domains of type iLiL (respectively type iRiR) from 𝐱\mathbf{x} to 𝐲\mathbf{y} is denoted πiL(𝐱,𝐲)\pi^{iL}(\mathbf{x,y}) (respectively πiR(𝐱,𝐲)\pi^{iR}(\mathbf{x,y})).

Refer to caption
Figure 7. Examples of πiL(𝐱,𝐲)\pi^{iL}(\mathbf{x,y}) and πiR(𝐱,𝐲)\pi^{iR}(\mathbf{x,y})
Refer to caption
Figure 8. Examples of πoL(𝐱,𝐲)\pi^{oL}(\mathbf{x,y}) and πoR(𝐱,𝐲)\pi^{oR}(\mathbf{x,y})
Definition 5.5 ([5, Definition 13.4.2]).

For 𝐱𝐒(g)\mathbf{x\in S}(g^{\prime}) and 𝐲𝐈(g)\mathbf{y\in I}(g^{\prime}), a positive domain pπ(𝐱,𝐲)p\in\pi(\mathbf{x,y}) is said to be of type 𝒐𝑳\bm{oL} if it is trivial or it satisfies the following conditions:\colon

  • At each corner point in 𝐱𝐲{c}\mathbf{x\cup y}\setminus\{c\}, at least three of the four adjoining squares have local multiplicity zero.

  • Three of four squares attaching at the corner point cc have the local multiplicity kk and at the northwest square meeting cc the local multiplicity is k1k-1.

  • 𝐲\mathbf{y} has 2k+12k+1 components that are not in 𝐱\mathbf{x}.

And, a positive domain pπ(𝐱,𝐲)p\in\pi(\mathbf{x,y}) is said to be of type 𝒐𝑹\bm{oR} if it is trivial or it satisfies the following conditions:\colon

  • At each corner point in 𝐱𝐲{c}\mathbf{x\cup y}\setminus\{c\}, at least three of the four adjoining squares have local multiplicity zero.

  • Three of four squares attaching at the corner point cc have the local multiplicity kk and at the northeast square meeting cc the local multiplicity is k+1k+1.

  • 𝐲\mathbf{y} has 2k+22k+2 components that are not in 𝐱\mathbf{x}.

The set of domains of type oLoL and oRoR from 𝐱\mathbf{x} to 𝐲\mathbf{y} is denoted πoL(𝐱,𝐲)\pi^{oL}(\mathbf{x,y}) and πoR(𝐱,𝐲)\pi^{oR}(\mathbf{x,y}) respectively.

Refer to caption
Figure 9. Examples of πK\pi^{K}

Let βi\beta_{i} be the vertical circle containing the point cc.

Definition 5.6.

For 𝐱,𝐲𝐍(g)\mathbf{x,y\in N}(g^{\prime}), a positive domain pπ(𝐱,𝐲)p\in\pi(\mathbf{x,y}) is said to be of type 𝑲𝟏\bm{K1} if it satisfies the following conditions:\colon

  • #{(𝐱𝐲)(𝐱𝐲)}4\#\{\mathbf{(x\cup y)\setminus(x\cap y)}\}\in 4\mathbb{N}.

  • Only one point of (𝐱𝐲)(𝐱𝐲)\mathbf{(x\cup y)\setminus(x\cap y)} on βj\beta_{j} is contained in the interior of pp, and at the other points, three of the four adjoining squares have local multiplicity zero.

  • pp satisfies that O1(p)=X2(p)=X1(p)+1O_{1}(p)=X_{2}(p)=X_{1}(p)+1, where Xi(p)X_{i}(p) is the coefficient of pp at the square containing XiX_{i} (i=1,2)(i=1,2).

A positive domain pπ(𝐱,𝐲)p\in\pi(\mathbf{x,y}) is said to be of type 𝑲𝟐\bm{K2} if it satisfies the following conditions:\colon

  • #{(𝐱𝐲)(𝐱𝐲)}4\#\{\mathbf{(x\cup y)\setminus(x\cap y)}\}\in 4\mathbb{N}.

  • Let m=#{𝐱𝐲(𝐱𝐲)}m=\#\{\mathbf{x\cup y\setminus(x\cap y)}\}. Then, (m4)/4(m-4)/4 points of (𝐱𝐲)(𝐱𝐲)\mathbf{(x\cup y)\setminus(x\cap y)} are contained in the interior of pp, and at the other points three of the four adjoining squares have local multiplicity zero.

  • pp satisfies that O1(p)=X2(p)=X1(p)+1O_{1}(p)=X_{2}(p)=X_{1}(p)+1

A positive domain pπ(𝐱,𝐲)p\in\pi(\mathbf{x,y}) is said to be of type 𝑲𝟑\bm{K3} if it satisfies the following conditions:\colon

  • #{(𝐱𝐲)(𝐱𝐲)}4\#\{\mathbf{(x\cup y)\setminus(x\cap y)}\}\in 4\mathbb{N}.

  • Let m=#{𝐱𝐲(𝐱𝐲)}m=\#\{\mathbf{x\cup y\setminus(x\cap y)}\}. Then, (m4)/4(m-4)/4 points of (𝐱𝐲)(𝐱𝐲)\mathbf{(x\cup y)\setminus(x\cap y)} are contained in the interior of pp, and at the other points three of the four adjoining squares have local multiplicity zero.

  • One of the vertical segments of (p)\partial(p) is on the vertical circle βj+1\beta_{j+1} and it goes up.

  • pp satisfies that O1(p)=X2(p)=X1(p)O_{1}(p)=X_{2}(p)=X_{1}(p)

The set of domains of type K1K1, K2K2, and K3K3 from 𝐱\mathbf{x} to 𝐲\mathbf{y} is denoted πK(𝐱,𝐲)\pi^{K}(\mathbf{x,y}) .

Note that however there are some domains that are both type K1K1 and type K2K2, the proof holds.

Definition 5.7.

The \mathcal{R}-module homomorphisms 𝒟iL:CFH(g)CFH(g)[U1]1,1\mathcal{D}^{iL}\colon CF^{-H}(g^{\prime})\to CF^{-H}(g)[U_{1}]\llbracket 1,1\rrbracket and 𝒟iR:CFH(g)CFH(g)[U1]\mathcal{D}^{iR}\colon CF^{-H}(g^{\prime})\to CF^{-H}(g)[U_{1}] are given by

(5.1) 𝒟iL(𝐱)=𝐲𝐈(g)(pπiL(𝐱,𝐲)U2O2(p)U3O3(p)UnOn(p))e(𝐲),\displaystyle\mathcal{D}^{iL}(\mathbf{x})=\sum_{\mathbf{y\in I}(g^{\prime})}\left(\sum_{p\in\pi^{iL}(\mathbf{x,y})}U_{2}^{O_{2}(p)}U_{3}^{O_{3}(p)}\dots U_{n}^{O_{n}(p)}\right)\cdot e(\mathbf{y}),
(5.2) 𝒟iR(𝐱)=𝐲𝐈(g)(pπiR(𝐱,𝐲)U2O2(p)U3O3(p)UnOn(p))e(𝐲).\displaystyle\mathcal{D}^{iR}(\mathbf{x})=\sum_{\mathbf{y\in I}(g^{\prime})}\left(\sum_{p\in\pi^{iR}(\mathbf{x,y})}U_{2}^{O_{2}(p)}U_{3}^{O_{3}(p)}\dots U_{n}^{O_{n}(p)}\right)\cdot e(\mathbf{y}).

Then, 𝒟:CFH(g)Cone(U1U2)\mathcal{D}\colon CF^{-H}(g^{\prime})\to\mathrm{Cone}(U_{1}-U_{2}) is defined by

𝒟(𝐱)=(𝒟iL(𝐱),𝒟iR(𝐱)).\mathcal{D}(\mathbf{x})=(\mathcal{D}^{iL}(\mathbf{x}),\mathcal{D}^{iR}(\mathbf{x})).
Definition 5.8.

The \mathcal{R}-module homomorphisms 𝒮oL:CFH(g)[U1]1,1CFH(g)\mathcal{S}^{oL}\colon CF^{-H}(g)[U_{1}]\llbracket 1,1\rrbracket\to CF^{-H}(g^{\prime}) and 𝒮oR:CFH(g)[U1]CFH(g)\mathcal{S}^{oR}\colon CF^{-H}(g)[U_{1}]\to CF^{-H}(g^{\prime}) are given by

(5.3) 𝒮oL(𝐱)=𝐲𝐒(g)(pπoL(𝐞𝟏(𝐱),𝐲)U2O2(p)U3O3(p)UnOn(p))𝐲,\displaystyle\mathcal{S}^{oL}(\mathbf{x})=\sum_{\mathbf{y\in S}(g^{\prime})}\left(\sum_{p\in\pi^{oL}(\mathbf{e^{-1}(x),y})}U_{2}^{O_{2}(p)}U_{3}^{O_{3}(p)}\dots U_{n}^{O_{n}(p)}\right)\cdot\mathbf{y},
(5.4) 𝒮oR(𝐱)=𝐲𝐒(g)(pπoR(𝐞𝟏(𝐱),𝐲)U2O2(p)U3O3(p)UnOn(p))𝐲.\displaystyle\mathcal{S}^{oR}(\mathbf{x})=\sum_{\mathbf{y\in S}(g^{\prime})}\left(\sum_{p\in\pi^{oR}(\mathbf{e^{-1}(x),y})}U_{2}^{O_{2}(p)}U_{3}^{O_{3}(p)}\dots U_{n}^{O_{n}(p)}\right)\cdot\mathbf{y}.

Then, 𝒮:Cone(U1U2)CFH(g)\mathcal{S}\colon\mathrm{Cone}(U_{1}-U_{2})\to CF^{-H}(g^{\prime}) is defined by

𝒮(𝐱)=(𝒮oL(𝐱),𝒮oR(𝐱)).\mathcal{S}(\mathbf{x})=(\mathcal{S}^{oL}(\mathbf{x}),\mathcal{S}^{oR}(\mathbf{x})).

Let \partial^{\prime} denote the differential of CFH(g)CF^{-H}(g^{\prime}).

Proposition 5.9.

The maps 𝒟,𝒮\mathcal{D},\mathcal{S} are chain maps.

Proof.

From [5, Lemma 13.3.13], 𝒟\mathcal{D} is a chain map by pairing domains of 𝒟+Cone𝒟\mathcal{D}\circ\partial^{\prime}+\partial_{Cone}\circ\mathcal{D}.

Consider the bijection θ:𝐒(g)𝐒(g)\theta\colon\mathbf{S}(g^{\prime})\to\mathbf{S}(g^{\prime}) determined by the 180180^{\circ} rotation of the horizontal circle containing cc. Obviously θ\theta induces a natural bijection θ(𝐱,𝐲):π(x,y)π(θ(𝐲),θ(𝐱))\theta(\mathbf{x,y})\colon\pi\mathbf{(}{x,y})\to\pi(\mathbf{\theta(y),\theta(x)}) for each 𝐱,𝐲𝐒(g)\mathbf{x,y\in S}(g^{\prime}). Then a composite domain ds(dπ(𝐱,𝐲),sπ(𝐲,𝐳))d*s\ (d\in\pi(\mathbf{x,y}),\ s\in\pi(\mathbf{y,z})) appears in 𝒟+Cone𝒟\mathcal{D}\circ\partial^{\prime}+\partial_{Cone}\circ\mathcal{D} if and only if sd(sπ(θ(𝐳),θ(𝐲)),dπ(θ(𝐲),θ(𝐱)))s^{\prime}*d^{\prime}\ (s^{\prime}\in\pi(\mathbf{\theta(z),\theta(y)}),\ d^{\prime}\in\pi(\mathbf{\theta(y),\theta(x)})) appears in 𝒮Cone+𝒮\mathcal{S}\circ\partial_{Cone}+\partial^{\prime}\circ\mathcal{S}. Therefore we can also pair all domains of 𝒮Cone+𝒮\mathcal{S}\circ\partial_{Cone}+\partial^{\prime}\circ\mathcal{S}. ∎

Definition 5.10.

The \mathcal{R}-module homomorphisms 𝒦:CFH(g)CFH(g)\mathcal{K}\colon CF^{-H}(g^{\prime})\to CF^{-H}(g^{\prime}) is given by

(5.5) 𝒦(𝐱)=𝐲𝐒(g)(pπK(𝐱,𝐲)U2O2(p)U3O3(p)UnOn(p))𝐲.\displaystyle\mathcal{K}(\mathbf{x})=\sum_{\mathbf{y\in S}(g^{\prime})}\left(\sum_{p\in\pi^{K}(\mathbf{x,y})}U_{2}^{O_{2}(p)}U_{3}^{O_{3}(p)}\dots U_{n}^{O_{n}(p)}\right)\cdot\mathbf{y}.
Definition 5.11.

The complexity of a domain is one if it is the trivial domain and, otherwise, is the number of horizontal segments in its boundary. Let us denote by k(p)k(p) the complexity of pp.

Proposition 5.12.

The above homomorphisms satisfy that

(5.6) 𝒟𝒮=id,\mathcal{D}\circ\mathcal{S}=\mathrm{id},\\
(5.7) 𝒮𝒟+𝒦+𝒦=id.\mathcal{S}\circ\mathcal{D}+\partial^{\prime}\circ\mathcal{K}+\mathcal{K}\circ\partial^{\prime}=\mathrm{id}.
Proof.

The map DiLD^{iL} (respectively SoLS^{oL}) can be decomposed as DiL=D1iL+D>1iLD^{iL}=D^{iL}_{1}+D^{iL}_{>1} (respectively SoL=S1oL+S>1oLS^{oL}=S^{oL}_{1}+S^{oL}_{>1}), where the subscript represents the restriction on the complexity of the domains. Then we can draw the following diagram:

CFH(g)[U1][1,1]\textstyle{CF^{-H}(g)[U_{1}][1,1]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}U1U2\scriptstyle{U_{1}-U_{2}}S1oL\scriptstyle{S^{oL}_{1}}S>1oL\scriptstyle{S^{oL}_{>1}}CFH(g)[U1]\textstyle{CF^{-H}(g)[U_{1}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SoR\scriptstyle{S^{oR}}I\textstyle{\mathrm{I}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D1iL\scriptstyle{D^{\mathrm{iL}}_{1}}N\textstyle{\mathrm{N}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D>1iL\scriptstyle{D^{\mathrm{iL}}_{>1}}DiR\scriptstyle{D^{\mathrm{iR}}}CFH(g)[U1][1,1]\textstyle{CF^{-H}(g)[U_{1}][1,1]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}U1U2\scriptstyle{U_{1}-U_{2}}CFH(g)[U1]\textstyle{CF^{-H}(g)[U_{1}]}

Note that the complexities of domains of πiL\pi^{iL} and πoL\pi^{oL} are odd, and ones of πiL\pi^{iL} and πoL\pi^{oL} are even.

It is convenient to write the horizontal circle crossing the point cc as αi\alpha_{i} and the vertical circle crossing the point cc as βj\beta_{j}.

Let us first examine the equation (5.6). To see the equation (5.6), we will count domains of the left side of (5.6) and check that domains except for the trivial ones are canceled in modulo 2.

Since we have obviously D1iLS1oL=idD^{iL}_{1}\circ S^{oL}_{1}=\mathrm{id}, all we need to check is DiRSoR=idD^{iR}\circ S^{oR}=\mathrm{id}, 𝒟S>1oL=0\mathcal{D}\circ S^{oL}_{>1}=0, and 𝒟SoR=0\mathcal{D}\circ S^{oR}=0. Let sds*d be a composite domain with sπoR(𝐱,𝐲)πoL(𝐱,𝐲),dπiR(𝐲,𝐳)πiL(𝐲,𝐳)s\in\pi^{oR}(\mathbf{x,y})\cup\pi^{oL}(\mathbf{x,y}),\ d\in\pi^{iR}(\mathbf{y,z})\cup\pi^{iL}(\mathbf{y,z}) appearing in DiRSoRD^{iR}\circ S^{oR} , 𝒟S>1oL\mathcal{D}\circ S^{oL}_{>1}, or 𝒟SoR\mathcal{D}\circ S^{oR}.

  1. (1)

    If k(s)=k(d)=2k(s)=k(d)=2, then sds*d appears in DiRSoRD^{iR}\circ S^{oR} and we get 𝐱=𝐳\mathbf{x=z}.

  2. (2)

    Otherwise, consider two points y1=𝐲αiy_{1}=\mathbf{y}\cap\alpha_{i} and y2=𝐲βjy_{2}=\mathbf{y}\cap\beta_{j}. We see that y1,y2𝐳y_{1},y_{2}\notin\mathbf{z} since c𝐳c\in\mathbf{z}. Then dd must overlay ss. If k(s)=3k(s)=3 and k(d)=3k(d)=3, at some of the corner points of d,sd,s, all adjoining squares have a local multiplicity greater than 0. Therefore such a domain cannot be taken.

  3. (3)

    If k(s)>3k(s)>3 or k(d)>3k(d)>3, then there are some points of the initial state as a point that is not a corner in the interior of dd or ss. Again, such a domain cannot be taken.

These arguments prove that DiRSoR=idD^{iR}\circ S^{oR}=\mathrm{id}, 𝒟S>1oL=0\mathcal{D}\circ S^{oL}_{>1}=0, and 𝒟SoR=0\mathcal{D}\circ S^{oR}=0.

To see the equation (5.7), consider the following diagram:\colon

I\textstyle{\mathrm{I}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D1iL\scriptstyle{D^{\mathrm{iL}}_{1}}N\textstyle{\mathrm{N}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}D>1iL\scriptstyle{D^{\mathrm{iL}}_{>1}}DiR\scriptstyle{D^{\mathrm{iR}}}CFH(g)[U1][1,1]\textstyle{CF^{-H}(g)[U_{1}][1,1]\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}U1U2\scriptstyle{U_{1}-U_{2}}S1oL\scriptstyle{S^{oL}_{1}}S>1oL\scriptstyle{S^{oL}_{>1}}CFH(g)[U1]\textstyle{CF^{-H}(g)[U_{1}]\ignorespaces\ignorespaces\ignorespaces\ignorespaces}SoR\scriptstyle{S^{oR}}I\textstyle{\mathrm{I}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}N\textstyle{\mathrm{N}}

For 𝐱𝐒(g))\mathbf{x\in S}(g^{\prime})), let p=ds(dπ(𝐱,𝐲),sπ(𝐲,𝐳))p=d*s\ (d\in\pi(\mathbf{x,y}),\ s\in\pi(\mathbf{y,z})) be a composite domain appearing in 𝒮𝒟\mathcal{S}\circ\mathcal{D}. Now we will observe that each pp appears also in 𝒦+𝒦\partial^{\prime}\circ\mathcal{K}+\mathcal{K}\circ\partial^{\prime} or id\mathrm{id}. We have five cases (figure 13-13):\colon

Refer to caption
Figure 10. Examples of the case (ii)
Refer to caption
Figure 11. Examples of the case (iii)
Refer to caption
Figure 12. Examples of the case (iv)
Refer to caption
Figure 13. Examples of the case (v)

  1. (i)

    If k(d)=k(s)=1k(d)=k(s)=1, then the composite domain dsd*s must be trivial. This domain also appears in idCFH(g)\mathrm{id}_{CF^{-H}(g^{\prime})}.

  2. (ii)

    If k(d)=1k(d)=1 or k(s)=1k(s)=1, then we can take a small square rr from pp by cutting pp along βj\beta_{j} and another composition p=(pr)rp=(p-r)*r or p=r(pr)p=r*(p-r) appearing in 𝒦\partial^{\prime}\circ\mathcal{K} or 𝒦\mathcal{K}\circ\partial^{\prime}. In this case, prp-r is of type K1K1.

  3. (iii)

    If k(d)>1k(d)>1 and k(s)>1k(s)>1 and pp does not contain any vertical thin annulus, then there is a 270270^{\circ} corner at one point xlx_{l} of (𝐱𝐲)(𝐱𝐲)\mathbf{(x\cup y)\setminus(x\cap y)}. Cutting pp along the horizontal or vertical circle containing xlx_{l} gives another decomposition using the domain of type K2K2.

  4. (iv)

    If k(d)>1k(d)>1 and k(s)>1k(s)>1 and pp has the vertical thin annulus pp^{\prime} containing O1O_{1}, then consider the domain ppp-p^{\prime}. This domain connects the same states as pp and can be decomposed using domains of type K3K3. If it is decomposed as pp=rkp-p=r*k or pp=krp-p=k*r, the rectangle rr does not contain O1O_{1}, so they are canceled in modulo 2.

  5. (v)

    Finally, we need to check that the remaining domains in 𝒦+𝒦\partial^{\prime}\circ\mathcal{K}+\mathcal{K}\circ\partial^{\prime} are canceled each other. Let p=rkp=r*k or p=krp=k*r be the remaining composite domain. Let mm be the number of corners that dd and ss share, in other words, let m=#{((𝐱𝐲)(𝐱𝐲))((𝐲𝐳)(𝐲𝐳))}m=\#\{\mathbf{((x\cup y)\setminus(x\cap y))}\cap\mathbf{((y\cup z)\setminus(y\cap z))}\}. Then we have three cases:\colon m=0,1,3m=0,1,3.

    1. (a)

      If m=0m=0, then pp can be decomposed as p=rkp=r*k and p=krp=k*r.

    2. (b)

      If m=1m=1, then pp has a 270270^{\circ} corner at one point of (𝐱𝐲)(𝐱𝐲)\mathbf{(x\cup y)\setminus(x\cap y)}. Likely the case (iii), cutting at the 270270^{\circ} corner in two different directions gives the two decompositions of pp.

    3. (c)

      If m=3m=3, then pp contains a vertical thin annulus pp^{\prime} containing O1O_{1}, then two domains pp and ppp-p^{\prime} are canceled. The domain of type 𝑲𝟏\bm{K1} or 𝑲𝟐\bm{K2} is used for the decomposition of pp, and the domain of type 𝑲𝟑\bm{K3} of pp^{\prime}.

6. Proof of the main theorem 1.1 and theorem 1.3

Proof of Theorem 1.1.

If gg^{\prime} is obtained from gg by a single cyclic permutation or commutation, Proposition 5.1 and 5.2 provide CFH(g)CFH(g)CF^{-H}(g)\simeq CF^{-H}(g^{\prime}). This chain homotopy equivalent gives induced chain homotopy equivalent CFUH(g)CFUH(g)CF^{-H}_{U}(g)\simeq CF^{-H}_{U}(g^{\prime}) as Maslov graded chain complexes over 𝔽[U]\mathbb{F}[U]-module. Proposition 4.4 and 4.3 provides tCFH(g)tCFH(g)tCF^{-H}(g)\simeq tCF^{-H}(g^{\prime}).

If gg^{\prime} is obtained from gg by a single stabilization, then CFH(g)Cone(U1U2)CF^{-H}(g^{\prime})\simeq\mathrm{Cone}(U_{1}-U_{2}) by Proposition 5.3. By identifying all UiU_{i}’s, Cone(U1U2)\mathrm{Cone}(U_{1}-U_{2}) turns into CFH(g)CFH(g)1,1CF^{-H}(g)\oplus CF^{-H}(g)\llbracket 1,1\rrbracket because U1U_{1} is homogeneous of degree (2,1)(-2,-1). Then we obtain induced chain homotopy CFUH(g)CFH(g)CFH(g)1,1CF^{-H}_{U}(g^{\prime})\simeq CF^{-H}(g)\oplus CF^{-H}(g)\llbracket 1,1\rrbracket. Proposition 4.4 and 4.3 provides tCFH(g)tCFH(g)WttCF^{-H}(g^{\prime})\simeq tCF^{-H}(g)\oplus W_{t}. ∎

Proof of Theorem 1.3.

For t[0,1]t\in[0,1], the grading shift 1t\llbracket 1-t\rrbracket from WtW_{t} does not affect the value of the Υ\Upsilon because 1t01-t\geq 0. ∎

7. Link cobordisms with grid homology

In this section, we observe graph grid diagrams for two links connected by a link cobordism. The basic idea is developed by Sarkar [7].

First, we will use the extended grid diagram which has an XX-marking and an OO^{*}-marking in the same square. The square containing both an XX-marking and an OO^{*}-marking represents an unknotted, unlinked component. It is known that the homology of these grid diagrams is also invariant. See [5, Section 8.4] for details about extended grid diagrams. A grid diagram representing a link is called tight if there is exactly one OO^{*}-marking in each link component.

In general, the symmetrized Alexander grading is not canonical because it needs to calculate the homology of CF(g)CF^{-}(g). However, if the balanced spatial graph is a link, we can take a tight grid diagram representing it, and the symmetrized Alexander grading is written explicitly: from [5, Section 8.2],

(7.1) AH(𝐱)=𝒥(𝐱,𝕏𝕆)12𝒥(𝕏,𝕏)+12𝒥(𝕆,𝕆)nl2,A^{H}(\mathbf{x})=\mathcal{J}(\mathbf{x},\mathbb{X}-\mathbb{O})-\frac{1}{2}\mathcal{J}(\mathbb{X},\mathbb{X})+\frac{1}{2}\mathcal{J}(\mathbb{O},\mathbb{O})-\frac{n-l}{2},

where ll is the number of link components.

According to Sarkar [7] and Vance [8], two tight grid diagrams representing two links connected by a link cobordism are connected by a finite sequence of link-grid moves. These moves are commutations, (de-)stabilizations, births, XX-saddles, OO-saddles, and deaths.

A grid diagram gg^{\prime} is obtained from gg by a birth (figure 16) if adding one row and column to gg and putting an XX-marking and XX-marking in the square which is the intersection of the new row and column. A grid move death is the inverse move of a birth. The move birth (respectively death) represents a birth (respectively a death) on link cobordism. These moves are link-grid move (4)(4) (respectively (7)(7)) in [7].

A grid diagram gg^{\prime} is obtained from gg by an XX-saddle (figure 16) if gg has a 2×22\times 2 small squares with two XX-markings at tow-left and bottom-right, and gg^{\prime} is obtained from gg by deleting these two markings and putting new ones at top-right and bottom-left. The move XX-saddle represents a saddle move on link cobordism. These moves are link-grid move (5)(5) in [7].

A grid diagram gg^{\prime} is obtained from gg by OO-saddle (figure 16) if gg has a 2×22\times 2 small squares with one OO^{*}-marking at tow-left and one OO-marking at bottom-right, and gg^{\prime} is obtained from gg by deleting these two markings and putting new two OO^{*}-markings at top-right and bottom-left. The move OO-saddle represents a split move on link cobordism. These moves are link-grid move (6)(6) in [7].

Refer to caption
Figure 14. A birth in gg produces gg^{\prime}, a death in gg^{\prime} produces gg.
Refer to caption
Figure 15. An XX-saddle in gg produces gg^{\prime}
Refer to caption
Figure 16. An OO-saddle in gg produces gg^{\prime}

Using grid homology for links, Sarkar [7] evaluated the maximum changes of Maslov and Alexander grading on each link-grid move. We will check that we can define the appropriate maps also on the t-modified chain complexes and that the changes of t-grading are the same as Sarkar’s evaluation with MtAM-t\cdot A.

In order to prove theorem 1.4, we evaluate the change of grading on each link-grid move. It is convenient to introduce an alternative Alexander grading A:A^{\prime}\colon

Definition 7.1 ([8, Definition 4.2]).

For 𝐱𝐒(g)\mathbf{x\in S}(g),

A(𝐱)=𝒥(𝐱,𝕏𝕆)12𝒥(𝕏,𝕏)+12𝒥(𝕆,𝕆)n12.A^{\prime}(\mathbf{x})=\mathcal{J}(\mathbf{x},\mathbb{X}-\mathbb{O})-\frac{1}{2}\mathcal{J}(\mathbb{X},\mathbb{X})+\frac{1}{2}\mathcal{J}(\mathbb{O},\mathbb{O})-\frac{n-1}{2}.

The symmetrized Alexander grading AHA^{H} can be obtained from AA^{\prime} by adding l12\frac{l-1}{2}.

In this section, we are thinking about chain complexes tCF(g)tCF^{-}(g) with t-grading using AA^{\prime} rather than AHA^{H}, we set grt(𝐱)=M(𝐱)tA(𝐱)\mathrm{gr}_{t}(\mathbf{x})=M(\mathbf{x})-tA^{\prime}(\mathbf{x}).

For H(tCF(g))=tHF(g)H(tCF^{-}(g))=tHF^{-}(g), we can define Υg(t)\Upsilon^{\prime}_{g}(t) in the same way as Υg(t)\Upsilon_{g}(t). It is clear that

(7.2) Υg(t)=Υg(t)l12t.\Upsilon_{g}(t)=\Upsilon^{\prime}_{g}(t)-\frac{l-1}{2}t.

7.1. births and deaths

Proposition 9.2 will imply that there are isomorphisms

D:tHF(g)tHF(g)tHF(g)1,\displaystyle D^{\prime}\colon tHF^{-}(g^{\prime})\to tHF^{-}(g)\oplus tHF^{-}(g)\llbracket 1\rrbracket,
S:tHF(g)tHF(g)1tHF(g).\displaystyle S^{\prime}\colon tHF^{-}(g)\oplus tHF^{-}(g)\llbracket 1\rrbracket\to tHF^{-}(g^{\prime}).

These maps preserve t-grading if we use symmetrized Alexander grading AHA^{H}, so DD^{\prime} shifts t-grading by 12t-\frac{1}{2}t and SS^{\prime} by 12t\frac{1}{2}t. We can write these maps as D=(H((𝒟iR)t),H((𝒟iL)t))D^{\prime}=(H((\mathcal{D}^{iR})^{t}),H((\mathcal{D}^{iL})^{t})) and S=(H((𝒮oR)t),H((𝒮oL)t))S^{\prime}=(H((\mathcal{S}^{oR})^{t}),H((\mathcal{S}^{oL})^{t})) using the notations in subsection 5.3 and Proposition 4.4.

As we regard H(𝒟iL)H(\mathcal{D}^{iL}) as the map into tHF(g)tHF^{-}(g) rather than tHF(g)1tHF^{-}(g)\llbracket 1\rrbracket, H(𝒟iL)H(\mathcal{D}^{iL}) is surjective map and shifts t-grading by 112t1-\frac{1}{2}t. Then the maximum shift of t-grading of the homogeneous, non-torsion element in homology by a death is 112t1-\frac{1}{2}t, which implies Υg(t)Υg(t)+112t\Upsilon^{\prime}_{g}(t)\geq\Upsilon^{\prime}_{g^{\prime}}(t)+1-\frac{1}{2}t.

On the other hand, the maximum change by a birth is 12t\frac{1}{2}t. Then we get Υg(t)Υg(t)+12t\Upsilon^{\prime}_{g^{\prime}}(t)\geq\Upsilon^{\prime}_{g}(t)+\frac{1}{2}t.

7.2. XX-saddles and OO-saddles

Before observing the saddles, we prepare an algebraic lemma. Let AA be a \mathcal{R}-module. Then the torsion submodule of AA is

Tors(A)={aA|thereisanonzeropwithpa=0}.\mathrm{Tors}(A)=\{a\in A|\mathrm{there\ is\ a\ non-zero}\ p\in\mathcal{R}\ \mathrm{with}\ p\cdot a=0\}.
Lemma 7.2.

Let A,BA,B be two \mathcal{R}-modules. If α:AB\alpha\colon A\to B and β:BA\beta\colon B\to A are two module maps with the property that βα=vs\beta\circ\alpha=v^{s} for some s0s\geq 0, then α\alpha induces an injective map from A/Tors(A)A/\mathrm{Tors}(A) into B/Tors(B)B/\mathrm{Tors}(B).

Proof.

If α(a)Tors(B)\alpha(a)\in\mathrm{Tors}(B), then there is a non-zero element pp\in\mathcal{R} with pα(a)=0p\cdot\alpha(a)=0. Then β(pα(a))=vspa=0\beta(p\cdot\alpha(a))=v^{s}p\cdot a=0, so aTors(A)a\in\mathrm{Tors}(A). ∎

Then we observe XX-saddles and OO-saddles.

Proposition 7.3.

If gg^{\prime} is obtained by an XX-saddle, there are \mathcal{R}-module maps

σ:tHF(g)tHF(g),\displaystyle\sigma\colon tHF^{-}(g)\to tHF^{-}(g^{\prime}),
μ:tHF(g)tHF(g).\displaystyle\mu\colon tHF^{-}(g^{\prime})\to tHF^{-}(g).

with the following properties:\colon

  • σ\sigma shifts t-grading by 12t-\frac{1}{2}t,

  • μ\mu shifts t-grading by 12t-\frac{1}{2}t,

  • μσ=vt\mu\circ\sigma=v^{t},

  • σμ=vt\sigma\circ\mu=v^{t}.

Proof.

Let cc be the point in the center of the 2×22\times 2 squares as in Figure 16. Define σ:tCF(g)tCF(g)\sigma\colon tCF^{-}(g)\to tCF^{-}(g^{\prime}) and μ:tCF(g)tCF(g)\mu\colon tCF^{-}(g)\to tCF^{-}(g^{\prime}) by

σ(𝐱)={𝐱(c𝐱)vt𝐱(c𝐱).\sigma(\mathbf{x})=\begin{cases}\mathbf{x}&(c\in\mathbf{x})\\ v^{t}\cdot\mathbf{x}&(c\notin\mathbf{x}).\end{cases} and μ(𝐱)={vt𝐱(c𝐱)𝐱(c𝐱).\mu(\mathbf{x})=\begin{cases}v^{t}\cdot\mathbf{x}&(c\in\mathbf{x})\\ \mathbf{x}&(c\notin\mathbf{x}).\end{cases}

It is straightforward to see that these two maps preserve Maslov grading and increase Alexander grading by 12\frac{1}{2}, so they shift t-grading by 12t-\frac{1}{2}t. Obviously, both μσ\mu\circ\sigma and σμ\sigma\circ\mu are multiplication by vtv^{t}. It is also clear that both σ\sigma and μ\mu are chain maps. Think of the maps on homology induced by them. ∎

By Lemma 7.2, Proposition 7.3 means that the shift of t-grading of the homogeneous, non-torsion element in homology by a XX-saddle is 12t-\frac{1}{2}t

Proposition 7.4.

If gg^{\prime} is obtained by an OO-saddle, there are \mathcal{R}-module maps

σ:tHF(g)tHF(g),\displaystyle\sigma\colon tHF^{-}(g)\to tHF^{-}(g^{\prime}),
μ:tHF(g)tHF(g).\displaystyle\mu\colon tHF^{-}(g^{\prime})\to tHF^{-}(g).

with the following properties:\colon

  • σ\sigma shifts t-grading by 1+12t-1+\frac{1}{2}t,

  • μ\mu shifts t-grading by 1+12t-1+\frac{1}{2}t,

  • μσ=v2t\mu\circ\sigma=v^{2-t},

  • σμ=v2t\sigma\circ\mu=v^{2-t}.

Proof.

We can prove this in the same way as Proposition 7.3.; All we need is to consider σ:tCF(g)tCF(g)\sigma\colon tCF^{-}(g)\to tCF^{-}(g^{\prime}) and μ:tCF(g)tCF(g)\mu\colon tCF^{-}(g)\to tCF^{-}(g^{\prime}) as

σ(𝐱)={𝐱(c𝐱)v2t𝐱(c𝐱).\sigma(\mathbf{x})=\begin{cases}\mathbf{x}&(c\in\mathbf{x})\\ v^{2-t}\cdot\mathbf{x}&(c\notin\mathbf{x}).\end{cases} and μ(𝐱)={v2t𝐱(c𝐱)𝐱(c𝐱).\mu(\mathbf{x})=\begin{cases}v^{2-t}\cdot\mathbf{x}&(c\in\mathbf{x})\\ \mathbf{x}&(c\notin\mathbf{x}).\end{cases}

By definitions, these two maps drop Maslov grading by 11 and Alexander grading by 12\frac{1}{2}, so they shift t-grading by 1+12t-1+\frac{1}{2}t. It is clear that μσ\mu\circ\sigma and σμ\sigma\circ\mu are multiplication by v2tv^{2-t}.

Again using Lemma 7.2, Proposition 7.4 means that the shift of t-grading of the homogeneous, non-torsion element in homology by an OO-saddle is 1+12t-1+\frac{1}{2}t.

8. Proof of the main Theorem 1.4

We verify Theorem 1.4 using the same method as Sarkar [7] and Vance [8].

Proposition 8.1 ([7, Theorem 4.1]).

If g1,g2g_{1},g_{2} are two tight grid diagrams representing l1,l2l_{1},l_{2}-component links L1,L2L_{1},L_{2}, respectively, and if there is a link cobordism with bb births, ss saddles, and dd deaths, then there is a sequence of link-grid moves connecting from g1g_{1} to g2g_{2}, such that there are exactly bb births, sd+l1l2s-d+l_{1}-l_{2} XX-saddles, dl1+l2d-l_{1}+l_{2} OO-saddles, and dd deaths, and these happen in this order.

Proof of Theorem 1.4.

Let g1,g2g_{1},g_{2} be two grid diagrams represents L1,L2L_{1},L_{2}, respectively. By Proposition 8.1, there is a sequence of link-grid moves taking g1g_{1} to g2g_{2} such that bb births, sd+l1l2s-d+l_{1}-l_{2} XX-saddles, dl1+l2d-l_{1}+l_{2} OO-saddles, and dd deaths happen in this order. Take a homogeneous, non-torsion element αtHF(g1)\alpha\in tHF^{-}(g_{1}) whose t-grading is Υg1(t)\Upsilon^{\prime}_{g_{1}}(t). Composing maps in the previous section, we get a map that sends α\alpha onto tHF(g2)tHF^{-}(g_{2}). By adding up the t-grading shifts of each link-grid move, then we get

Υg1(t)+b12t+(sd+l1l2)(12t)+(dl1+l2)(1+12t)+d(112t)\displaystyle\Upsilon^{\prime}_{g_{1}}(t)+b\cdot\frac{1}{2}t+(s-d+l_{1}-l_{2})\cdot\left(-\frac{1}{2}t\right)+(d-l_{1}+l_{2})\cdot\left(-1+\frac{1}{2}t\right)+d\cdot\left(1-\frac{1}{2}t\right)
Υg2(t),\displaystyle\leq\Upsilon^{\prime}_{g_{2}}(t),

so

Υg1(t)l112t(12(sbd)+1l1+l22)t(l11)t+l1l2\displaystyle\Upsilon^{\prime}_{g_{1}}(t)-\frac{l_{1}-1}{2}t-\left(\frac{1}{2}(s-b-d)+1-\frac{l_{1}+l_{2}}{2}\right)t-(l_{1}-1)t+l_{1}-l_{2}
Υg2(t)l212t.\displaystyle\leq\Upsilon^{\prime}_{g_{2}}(t)-\frac{l_{2}-1}{2}t.

Use (7.2) and g=12(sbd)+1l1l22g=\frac{1}{2}(s-b-d)+1-\frac{l_{1}-l_{2}}{2}, then

ΥL1(t)tgt(l11)(l1l2)ΥL2(t)\Upsilon_{L_{1}}(t)-tg-t(l_{1}-1)-(l_{1}-l_{2})\leq\Upsilon_{L_{2}}(t)

We can once get the other inequality if we reverse the direction of link cobordism, so we see that

ΥL1(t)tgt(l11)(l1l2)ΥL2(t)ΥL1(t)+tg+t(l21)+(l2l1).\Upsilon_{L_{1}}(t)-tg-t(l_{1}-1)-(l_{1}-l_{2})\leq\Upsilon_{L_{2}}(t)\leq\Upsilon_{L_{1}}(t)+tg+t(l_{2}-1)+(l_{2}-l_{1}).

9. Proof of some properties of the Υ\Upsilon invariant

9.1. The value of Υ\Upsilon at t=0t=0

Proof of Proposition 1.6.

When t=0t=0, the t-modified chain complex tCFH(g)tCF^{-H}(g) is independent of the XX-markings because the differential is

t(𝐱)=𝐲𝐒(g)(rRect(𝐱,𝐲)v2|𝕆r|)𝐲,\partial_{t}^{-}(\mathbf{x})=\sum_{\mathbf{y}\in\mathbf{S}(g)}\left(\sum_{r\in\mathrm{Rect}^{\circ}(\mathbf{x,y})}v^{2|\mathbb{O}\cap r|}\right)\mathbf{y},

and t-grading is

grt(vα𝐱)=M(𝐱)α.\mathrm{gr}_{t}(v^{\alpha}\mathbf{x})=M(\mathbf{x})-\alpha.

Also, we do not need to distinguish between OO-markings and OO^{*}-markings. Let nn be a size of gg. Remove all XX-markings of gg and put nn XX-markings so that the new tight grid diagram gg^{\prime} represents an unknot. Then HFH(g)𝔽[U]HF^{-H}(g^{\prime})\cong\mathbb{F}[U] in grading zero and tCFH(g)tCFH(g)tCF^{-H}(g)\cong tCF^{-H}(g^{\prime}) as chain complexes.

Refer to caption
Figure 17. Getting grid diagram representing an unknot

By Proposition 4.4, the universal coefficient theorem, and [5, Lemma 14.1.11], we have tHFH(g)H((CFH(g)U1==Un)𝔽[U]0)0W0n1tHF^{-H}(g^{\prime})\cong H\left((\frac{CF^{-H}(g^{\prime})}{U_{1}=\dots=U_{n}})\otimes_{\mathbb{F}[U]}\mathcal{R}_{0}\right)\cong\mathcal{R}_{0}\otimes W_{0}^{n-1}, where 0\mathcal{R}_{0} is the t-graded module isomorphic to \mathcal{R} in grading zero and W0𝔽0𝔽1W_{0}\cong\mathbb{F}_{0}\oplus\mathbb{F}_{-1} (in other words, t=0t=0 for WtW_{t}). Similar to the proof of Theorem 1.3, the grading shift 1\llbracket 1\rrbracket does not affect the value of the Υ\Upsilon. Therefore Υg(0)=Υg(0)=0\Upsilon_{g}(0)=\Upsilon_{g^{\prime}}(0)=0. ∎

9.2. Crossing change

Proposition 9.1.

If two graph grid diagrams g+g_{+} and gg_{-} represent two links L+L_{+} and LL_{-} that differ in a crossing change, then there are \mathcal{R} maps Ct:tHFH(g+)tHFH(g)C^{t}_{-}\colon tHF^{-H}(g_{+})\to tHF^{-H}(g_{-}) and C+t:tHFH(g)tHFH(g+)C^{t}_{+}\colon tHF^{-H}(g_{-})\to tHF^{-H}(g_{+}), with the following properties:\colon

  • CC_{-} is graded,

  • C+C_{+} shifts t-grading by 2+t-2+t,

  • CC+=v2tC_{-}\circ C_{+}=v^{2-t},

  • C+C=v2tC_{+}\circ C_{-}=v^{2-t},

Proof.

Combine [5, Proposition 6.1.1] and Proposition 4.4. ∎

Proof of Proposition 1.7.

We can show this in the same argument as [1, Theorem 5.10] by using maps of Proposition 9.1. ∎

9.3. Adding an unknot

Proposition 9.2.

If g,gg,g^{\prime} be two graph grid diagrams as in Figure 16, then as graded \mathcal{R}-modules,

(9.1) tHFH(g)tHFH(g)tHFH(g)1,tHF^{-H}(g^{\prime})\cong tHF^{-H}(g)\oplus tHF^{-H}(g)\llbracket 1\rrbracket,

where 1\llbracket 1\rrbracket is t-grading shift by 11.

Proof.

The basic idea is the same as in [5, Section 8.4], in other words, we can use the same maps as the maps for stabilization invariance.

We assume that CFH(g)CF^{-H}(g) is a chain complex over 𝔽[U2,,Un]\mathbb{F}[U_{2},\dots,U_{n}] and CFH(g)CF^{-H}(g^{\prime}) is one over 𝔽[U1,,Un]\mathbb{F}[U_{1},\dots,U_{n}].

We will construct a chain homotopy equivalence

(9.2) CFH(g)CFH(g)[U1]CFH(g)[U1]1,0.CF^{-H}(g^{\prime})\simeq CF^{-H}(g)[U_{1}]\oplus CF^{-H}(g)[U_{1}]\llbracket 1,0\rrbracket.

As in Figure 16, we assume that gg^{\prime} has 2×22\times 2 squares such that the top-right square contains both one OO-marking and one XX-marking and the bottom-left square has one OO- or OO^{*}-marking. By definition, gg^{\prime} is strictly not a graph grid diagram but we can consider about CFH(g)CF^{-H}(g^{\prime}) in the same manner (See [5, Lemma 8.4.2] for detail). We denote by c the intersection point of the new horizontal and vertical circles in gg^{\prime} Under these settings, we can use the same maps 𝒟\mathcal{D}, 𝒮\mathcal{S}, and 𝒦\mathcal{K} as in Definition 5.7, 5.8, and 5.10 respectively for CFH(g)CF^{-H}(g) and CFH(g)CF^{-H}(g^{\prime}). Direct computations show that the grading changes of these maps are different from the maps of stabilization invariance in Section 5.3:\colon 𝒟iR\mathcal{D}^{iR} is bigraded, 𝒟iL\mathcal{D}^{iL} shifts Maslov grading by 11, 𝒟oR\mathcal{D}^{oR} is bigraded, and 𝒟oL\mathcal{D}^{oL} shifts Maslov grading by 1-1. Counting domains are independent of markings, so 𝒟\mathcal{D}, 𝒮\mathcal{S} are chain homotopy equivalences.

Applying Proposition 4.4 to the induced chain homotopy equivalence CFUH(g)CFUH(g)CFUH(g)1,0CF^{-H}_{U}(g^{\prime})\simeq CF^{-H}_{U}(g)\oplus CF^{-H}_{U}(g)\llbracket 1,0\rrbracket from (9.2), we get (9.1).

Proof of 1.8.

It is immediately from (9.1) in Proposition 9.2. Similar to the proof of Theorem 1.3, the grading shift 1\llbracket 1\rrbracket does not affect the value of the Υ\Upsilon. ∎

9.4. Wedge sum of an unknot

Proposition 9.3.

If g,gg,g^{\prime} be two graph grid diagrams as in Figure 18, then as graded \mathcal{R}-modules,

(9.3) tHFH(g)tHFH(g)Wt.tHF^{-H}(g^{\prime})\cong tHF^{-H}(g)\otimes W_{t}.
Proof.

The same argument as Proposition 5.3 works even if there is one extra OO^{*}-marking in the 2×22\times 2 block because the domains that appear in 𝒟\mathcal{D}, 𝒮\mathcal{S}, and 𝒦\mathcal{K} is independent of the extra OO^{*}-marking. We can show that there is a chain homotopy equivalence

CFH(g)Cone(U1U2:CFH(g)[U1]CFH(g)[U1]),CF^{-H}(g^{\prime})\simeq\mathrm{Cone}(U_{1}-U_{2}\colon CF^{-H}(g)[U_{1}]\to CF^{-H}(g)[U_{1}]),

where we regard CFH(g)CF^{-H}(g^{\prime}) as 𝔽[U1,,Un]\mathbb{F}[U_{1},\dots,U_{n}]-module and CFH(g)CF^{-H}(g) as 𝔽[U2,,Un]\mathbb{F}[U_{2},\dots,U_{n}]-module.

Refer to caption
Figure 18. grid diagram representing ff, f#c𝒪f\#_{c}\mathcal{O}

Applying Proposition 4.4 to the induced chain homotopy equivalence CFUH(g)CFUH(g)CFUH(g)1,1CF^{-H}_{U}(g^{\prime})\simeq CF^{-H}_{U}(g)\oplus CF^{-H}_{U}(g)\llbracket 1,1\rrbracket by the relation U1==UnU_{1}=\dots=U_{n}, we get (9.3). ∎

Proof of 1.9.

It is immediately from Proposition 9.3. Similar to the proof of Theorem 1.3 and 1.8, the grading shift 1t\llbracket 1-t\rrbracket from WtW_{t} does not affect the value of the Υ\Upsilon for t[0,1]t\in[0,1] because 1t01-t\geq 0. ∎

Acknowledgements

I would like to express my sincere gratitude to my supervisor, Tetsuya Ito, for useful discussions and corrections. I sincerely thank the anonymous reviewers for their valuable feedback. This work was supported by JST, the establishment of university fellowships towards the creation of science technology innovation, Grant Number JPMJFS2123.

References

  • [1] Viktória Földvári, The knot invariant Υ\Upsilon using grid homologies, J. Knot Theory Ramifications 30 (2021), no. 7, Paper No. 2150051, 26. MR 4321933
  • [2] Shelly Harvey and Danielle O’Donnol, Heegaard Floer homology of spatial graphs, Algebr. Geom. Topol. 17 (2017), no. 3, 1445–1525. MR 3677933
  • [3] Ciprian Manolescu, Peter Ozsváth, Zoltán Szabó, and Dylan Thurston, On combinatorial link Floer homology, Geom. Topol. 11 (2007), 2339–2412. MR 2372850
  • [4] Peter Ozsváth and Zoltán Szabó, Knot Floer homology and the four-ball genus, Geom. Topol. 7 (2003), 615–639. MR 2026543
  • [5] Peter S. Ozsváth, András I. Stipsicz, and Zoltán Szabó, Grid homology for knots and links, Mathematical Surveys and Monographs, vol. 208, American Mathematical Society, Providence, RI, 2015. MR 3381987
  • [6] by same author, Concordance homomorphisms from knot Floer homology, Adv. Math. 315 (2017), 366–426. MR 3667589
  • [7] Sucharit Sarkar, Grid diagrams and the Ozsváth-Szabó tau-invariant, Math. Res. Lett. 18 (2011), no. 6, 1239–1257. MR 2915478
  • [8] Katherine Vance, Tau invariants for balanced spatial graphs, J. Knot Theory Ramifications 29 (2020), no. 9, 2050066, 29. MR 4152956