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

Splitting maps in link Floer homology and integer points in permutahedra

Akram Alishahi A. A. :Department of Mathematics, University of Georgia
Athens, GA 30602
akram.alishahi@uga.edu
Eugene Gorsky E. G.: Department of Mathematics, University of California Davis, One Shields Avenue, Davis CA 94702, USA egorskiy@math.ucdavis.edu  and  Beibei Liu B.L.: Department of Mathematics, The Ohio State University, 100 Math Tower, 231 West 18th Avenue, Columbus, OH, 43210, USA bbliumath@gmail.com
Abstract.

In this paper, we study the skein exact sequence for links via the exact surgery triangle of link Floer homology and compare it with other skein exact sequences given by Ozsváth and Szabó. As an application, we use the skein exact sequence to study the splitting number and splitting maps for links. In particular, we associate the splitting maps for the torus link T(n,n)T(n,n) to integer points in the (n1)(n-1)-dimensional permutahedron, and obtain the link Floer homology of an nn-component homology nontrivial unlink in S1×S2S^{1}\times S^{2}.

1. Introduction

In this paper, we study various maps in Heegaard Floer homology associated to crossing changes in link diagrams. Given such a diagram with a chosen crossing, we can consider three links L_+L_{\_}+, L_L_{\_}- and L_0L_{\_}0 in the three-sphere corresponding to the positive crossing, negative crossing and oriented resolution (see Figure 1). We will always assume that the crossing is between different components of L_±L_{\_}{\pm}, so that L_+L_{\_}+ and L_L_{\_}- both have one more component than L_0L_{\_}0.

Refer to captionL_+L_{\_}+L_L_{\_}-L_0L_{\_}0
Figure 1. From left to right: positive crossing, negative crossing and oriented resolution.

For various technical reasons, we work with the “full” version \mathcal{H\!F\!L} of the Heegaard Floer homology with two marked points on each link component over 𝔽=/2\mathbb{F}=\mathbb{Z}/2\mathbb{Z}, developed in [32, 33]. In particular, for a link LL with nn components in S3S^{3}, (L)\mathcal{H\!F\!L}(L) is a n\mathbb{Z}\oplus\mathbb{Z}^{n}-graded module over 𝔽[U_1,,U_n,V_1,,V_n]\mathbb{F}[U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n] where all products U_iV_iU_{\_}iV_{\_}i act by the same operator which we will denote by 𝐔\mathbf{U}. Sometimes we will need to work with the completion 𝓗𝓕𝓛(L)\bm{{\mathcal{H\!F\!L}}}(L) which is a module over the power series ring 𝔽[[U_1,,U_n,V_1,,V_n]]\mathbb{F}[[U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n]].

Our first result describes the crossing change maps in this version of Heegaard Floer homology generalizing the maps in [26, 21, 30] for 𝐻𝐹𝐾^\widehat{\mathit{HFK}} and 𝐻𝐹𝐾\mathit{HFK}^{-}.

Theorem 1.1.

Given a crossing between the components L_iL_{\_}i and L_jL_{\_}j of an oriented link in the three-sphere, for all kk\in\mathbb{Z} corresponding to Spinc\text{Spin}^{c}-structures in certain surgery cobordism shown in Figure 3, there are maps

(1) ψ_k:(L_+)(L_)andϕ_k:(L_)(L_+)\psi_{\_}k:\mathcal{H\!F\!L}(L_{\_}+)\to\mathcal{H\!F\!L}(L_{\_}-)\ \mathrm{and}\ \phi_{\_}k:\mathcal{H\!F\!L}(L_{\_}-)\to\mathcal{H\!F\!L}(L_{\_}+)

satisfying the following equations:

  • (a)

    The maps ψ_k\psi_{\_}k are determined by ψ_0\psi_{\_}0 and ψ_1\psi_{\_}{-1}:

    ψ_k=(V_iU_j)k𝐔k(k1)2ψ_0fork0,ψ_k=(V_jU_i)1k𝐔(k+1)(k+2)2ψ_1fork1.\psi_{\_}k=(V_{\_}iU_{\_}j)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}\psi_{\_}0\quad\mathrm{for}\ k\geq 0,\quad\psi_{\_}k=(V_{\_}jU_{\_}i)^{-1-k}\mathbf{U}^{\frac{(k+1)(k+2)}{2}}\psi_{\_}{-1}\quad\mathrm{for}\ k\leq-1.
  • (b)

    We have V_jψ_0=V_iψ_1V_{\_}j\psi_{\_}0=V_{\_}i\psi_{\_}{-1} and U_iψ_0=U_jψ_1U_{\_}i\psi_{\_}0=U_{\_}j\psi_{\_}{-1}.

  • (c)

    The maps ϕ_k\phi_{\_}k are determined by ϕ_0\phi_{\_}0 and ϕ_1\phi_{\_}1:

    ϕ_k=(U_iU_j)k1𝐔(k1)(k2)2ϕ_1fork1,ϕ_k=(V_iV_j)k𝐔k(k+1)2ϕ_0fork0.\phi_{\_}k=(U_{\_}iU_{\_}j)^{k-1}\mathbf{U}^{\frac{(k-1)(k-2)}{2}}\phi_{\_}1\quad\mathrm{for}\ k\geq 1,\quad\phi_{\_}k=(V_{\_}iV_{\_}j)^{-k}\mathbf{U}^{\frac{k(k+1)}{2}}\phi_{\_}{0}\quad\mathrm{for}\ k\leq 0.
  • (d)

    The maps ψ_k\psi_{\_}k and ϕ_k\phi_{\_}k compose as follows:

    ϕ_0ψ_0=V_i,ϕ_0ψ_1=V_j,ϕ_1ψ_0=U_j,ϕ_1ψ_1=U_i\phi_{\_}0\psi_{\_}0=V_{\_}i,\ \phi_{\_}0\psi_{\_}{-1}=V_{\_}j,\ \phi_{\_}1\psi_{\_}0=U_{\_}j,\ \phi_{\_}1\psi_{\_}{-1}=U_{\_}i
    ψ_0ϕ_0=V_i,ψ_1ϕ_0=V_j,ψ_0ϕ_1=U_j,ψ_1ϕ_1=U_i,\psi_{\_}0\phi_{\_}0=V_{\_}i,\ \psi_{\_}{-1}\phi_{\_}0=V_{\_}j,\ \psi_{\_}0\phi_{\_}1=U_{\_}j,\ \psi_{\_}{-1}\phi_{\_}1=U_{\_}i,

    The rest of compositions are determined by these.

See Section 3 for more details and the gradings for all these maps. In [30] Ozsváth and Szabó proved a skein exact triangle for 𝐻𝐹𝐾\mathit{HFK}^{-}

𝐻𝐹𝐾(L_+)𝐻𝐹𝐾(L_)𝐻𝐹𝐾(L_0)W\rightarrow\mathit{HFK}^{-}(L_{\_}+)\rightarrow\mathit{HFK}^{-}(L_{\_}-)\rightarrow\mathit{HFK}^{-}(L_{\_}0)\otimes W\rightarrow\ldots

where WW is some given bigraded module.

We generalize this as follows.

Theorem 1.2.

Given a crossing between the components L_iL_{\_}i and L_jL_{\_}j, there is an exact triangle

(2) 𝓗𝓕𝓛(L_+)Ψ_ij𝓗𝓕𝓛(L_)H_(𝓒𝓕𝓛(L_0)𝓚)\rightarrow\bm{{\mathcal{H\!F\!L}}}(L_{\_}+)\xrightarrow{\Psi_{\_}{ij}}\bm{{\mathcal{H\!F\!L}}}(L_{\_}-)\rightarrow H_{\_}*(\bm{\mathcal{C\!F\!L}}(L_{\_}0)\otimes\bm{\mathcal{K}})\rightarrow\ldots

where the map Ψ_ij:𝓗𝓕𝓛(L_+)𝓗𝓕𝓛(L_)\Psi_{\_}{ij}:\bm{{\mathcal{H\!F\!L}}}(L_{\_}+)\to\bm{{\mathcal{H\!F\!L}}}(L_{\_}-) is given by Ψ_ij=_k(1)kψ_k\Psi_{\_}{ij}=\sum_{\_}{k\in\mathbb{Z}}(-1)^{k}\psi_{\_}k, and 𝓚\bm{\mathcal{K}} is the completion of the module 𝒦\mathcal{K} defined in (6).

Theorem 1.1 implies the following:

Corollary 1.3.

We have Ψ_ij=τ(ψ_0ψ_1)\Psi_{\_}{ij}=\tau(\psi_{\_}0-\psi_{\_}{-1}) where τ=1+\tau=1+\ldots is an explicit invertible power series in 𝔽[[U_1,,U_n,V_1,,V_n]]\mathbb{F}[[U_{\_}1,\cdots,U_{\_}n,V_{\_}1,\cdots,V_{\_}n]] defined in Lemma 4.5. In particular, the cones of Ψ_ij\Psi_{\_}{ij} and of Ψ_ij0=ψ_0ψ_1\Psi_{\_}{ij}^{0}=\psi_{\_}0-\psi_{\_}{-1} are homotopy equivalent.

Note that the map Ψ_ij0\Psi_{\_}{ij}^{0} has homological degree 0 and can be defined without completion.

Remark 1.4.

Since we work over the field 𝔽=/2\mathbb{F}=\mathbb{Z}/2\mathbb{Z}, the signs here and below are purely for esthetic reasons. However, we expect all the maps to exist for theories with integer coefficients (similar to [1]), and conjecture that (up to an overall normalization) the signs would match. See also Section 6.4 on comparison of the signs with triply graded Khovanov-Rozansky homology.

Next, we study the compositions of crossing change maps. Since we have two essentially different maps ψ_0\psi_{\_}0 and ψ_1\psi_{\_}{-1} (resp. ϕ_0\phi_{\_}0 and ϕ_1\phi_{\_}1) for a single crossing change, for a sequence of rr crossing changes we have 2r2^{r} possible associated maps in Heegaard Floer homology of various degrees, some of which may coincide. We determine the degrees of all such maps in Section 5 and use them to bound splitting numbers for links.

In a striking example, we can take the nn-component torus link T(n,n)T(n,n), change (n2)\binom{n}{2} crossings between different components from positive to negative and obtain the unlink O_nO_{\_}n. In this case, we are able to completely determine all crossing change maps.

Theorem 1.5.

If one chooses either ψ_0\psi_{\_}0 or ψ_1\psi_{\_}{-1} for each of (n2)\binom{n}{2} crossing changes from T(n,n)T(n,n) to O_nO_{\_}n, the Alexander degrees of the resulting maps correspond to integer points in the permutahedron P_nP_{\_}n. Any two maps of the same degree coincide, and any integer point in P_nP_{\_}n corresponds to an injective map (T(n,n))(O_n)\mathcal{H\!F\!L}(T(n,n))\to\mathcal{H\!F\!L}(O_{\_}n) which can be described explicitly on generators of (T(n,n))\mathcal{H\!F\!L}(T(n,n)).

For example, P_3P_{\_}3 is a hexagon with 6 vertices and 1 interior point, see Figure 14. To get from T(3,3)T(3,3) to unlink, one needs to change 3 crossings, so there are 23=82^{3}=8 possible splitting maps. Six of them correspond to the vertices of P_3P_{\_}3 and two remaining ones coincide and correspond to the interior point of P_3P_{\_}3. We generalize Theorem 1.5 to arbitrary L-space links in Section 5.3.

Theorem 1.6.

Suppose that LL is an L-space link. Then:

a) For any choice of crossing changes and the maps ψ_k,ϕ_k\psi_{\_}k,\phi_{\_}k at the crossings, the resulting map F:(L)(split(L))F:\mathcal{H\!F\!L}(L)\to\mathcal{H\!F\!L}(\mathrm{split}(L)) is completely determined by its Alexander and Maslov degrees.

b) If, in addition, all crossings between the different components of LL are positive, the splitting maps are in bijection with the integer points in a certain polytope P_LP_{\_}L (see Definition 5.4).

One can also study the compositions of maps Ψ_ij\Psi_{\_}{ij} from skein exact sequence (2) for crossings in T(n,n)T(n,n) between L_iL_{\_}i and L_jL_{\_}j. Let 𝒥\mathcal{J} be the ideal in (O_n)\mathcal{H\!F\!L}(O_{\_}n) generated by determinants

Δ_S=det(U_1a_1V_1b_1U_1a_nV_1b_nU_na_1V_nb_1U_na_nV_nb_n)\Delta_{\_}S=\det\left(\begin{matrix}U_{\_}1^{a_{\_}1}V_{\_}1^{b_{\_}{1}}&\cdots&U_{\_}1^{a_{\_}n}V_{\_}1^{b_{\_}{n}}\\ \vdots&&\vdots\\ U_{\_}n^{a_{\_}1}V_{\_}n^{b_{\_}{1}}&\cdots&U_{\_}n^{a_{\_}n}V_{\_}n^{b_{\_}{n}}\\ \end{matrix}\right)

for all possible nn-element subsets S={(a_1,b_1),,(a_n,b_n)}_0×_0S=\{(a_{\_}1,b_{\_}1),\ldots,(a_{\_}n,b_{\_}n)\}\subset\mathbb{Z}_{\_}{\geq 0}\times\mathbb{Z}_{\_}{\geq 0}. We denote by 𝓙\bm{\mathcal{J}} the completion of 𝒥\mathcal{J} in 𝓗𝓕𝓛(O_n)\bm{{\mathcal{H\!F\!L}}}(O_{\_}n).

Theorem 1.7.

a) Let Ω:𝓗𝓕𝓛(T(n,n))𝓗𝓕𝓛(O_n)\Omega:\bm{{\mathcal{H\!F\!L}}}(T(n,n))\to\bm{{\mathcal{H\!F\!L}}}(O_{\_}n) be the composition of the maps Ψ_ij\Psi_{\_}{ij} from Theorem 1.2 over all i<ji<j. Then Ω\Omega is injective and its image is the ideal 𝓙\bm{\mathcal{J}} in 𝓗𝓕𝓛(O_n)\bm{{\mathcal{H\!F\!L}}}(O_{\_}n).

b) Let Ω0:(T(n,n))(O_n)\Omega^{0}:\mathcal{H\!F\!L}(T(n,n))\to\mathcal{H\!F\!L}(O_{\_}n) be the composition of the maps Ψ_0ij\Psi^{0}_{\_}{ij} from Corollary 1.3 over all i<ji<j. Then Ω0\Omega^{0} is injective and its image is the ideal 𝒥\mathcal{J} in (O_n)\mathcal{H\!F\!L}(O_{\_}n).

Corollary 1.8.

We have (T(n,n))𝒥\mathcal{H\!F\!L}(T(n,n))\simeq\mathcal{J} as modules over 𝔽[U_1,,U_n,V_1,,V_n]\mathbb{F}[U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n].

Theorem 1.7 can be compared with the main result of [12] where the “yy-ified” triply graded Khovanov-Rozansky homology (also known as HOMFLY homology) of T(n,n)T(n,n) was computed using a very similar ideal to 𝒥\mathcal{J}, see Section 6.4. This suggests a spectral sequence from the “yy-ified” HOMFLY homology to \mathcal{H\!F\!L} which we plan to study in a future work. Such a spectral sequence should generalize the spectral sequences for reduced homology studied in [5, 10, 11]

Finally, we can use the above results to compute the Heegaard Floer homology of certain links in S1×S2S^{1}\times S^{2}.

Theorem 1.9.

Let Z_nZ_{\_}n be the link consisting of nn parallel copies of S1S^{1} inside S1×S2S^{1}\times S^{2}. Then 𝓗𝓕𝓛(S1×S2,Z_n)𝓙/(γ)\bm{{\mathcal{H\!F\!L}}}(S^{1}\times S^{2},Z_{\_}n)\simeq\bm{\mathcal{J}}/(\gamma) where

γ=μ_0_i<j(V_iV_j)+μ_n1_i<j(U_iU_j)+_j=1n2det(U_1jU_11V_1V_1n1jU_njU_n1V_nV_nn1j).\gamma=\mu_{\_}0\prod_{\_}{i<j}(V_{\_}i-V_{\_}j)+\mu_{\_}{n-1}\prod_{\_}{i<j}(U_{\_}i-U_{\_}j)+\sum_{\_}{j=1}^{n-2}\det\left(\begin{matrix}U_{\_}1^{j}&\cdots&U_{\_}1&1&V_{\_}1&\cdots&V_{\_}1^{n-1-j}\\ \vdots&&\vdots&\vdots&\vdots&&\vdots\\ U_{\_}n^{j}&\cdots&U_{\_}n&1&V_{\_}n&\cdots&V_{\_}n^{n-1-j}\end{matrix}\right).

and

μ_0=_k=0(V_1V_n)k𝐔k(k1)2,μ_n1=_k=0(U_1U_n)k𝐔k(k1)2.\mu_{\_}0=\sum_{\_}{k=0}^{\infty}(V_{\_}1\cdots V_{\_}n)^{k}\mathbf{U}^{\frac{k(k-1)}{2}},\quad\mu_{\_}{n-1}=\sum_{\_}{k=0}^{\infty}(U_{\_}1\cdots U_{\_}n)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}.

Acknowledgments

We are grateful to Daren Chen, Matthew Hedden, Peter Kronheimer, Tye Lidman, Robert Lipshitz, Ciprian Manolescu, Lisa Piccirillo and Ian Zemke for useful discussions. A. A. and E. G. were partially supported by the NSF grant DMS-1928930 while they were in residence at the Simons Laufer Mathematical Sciences Institute (previously known as MSRI) in Berkeley, California, during the Fall 2022 semester. A. A. was also partially supported by NSF grants DMS-2000506 and DMS- 2238103. E. G. was also partially supported by the NSF grant DMS-1760329. B. L. is partially supported by the NSF grant DMS-2203237.

2. Background

2.1. Lattices

We will work with the lattice n\mathbb{Z}^{n} and its translates. We define a partial order on n\mathbb{Z}^{n} by

𝐮𝐯u_iv_ifor alli.\mathbf{u}\preceq\mathbf{v}\ \Leftrightarrow\ u_{\_}i\leq v_{\_}i\ \text{for all}\ i.

We will denote the basis vectors by 𝐞_i=(0,0,1,0,,0)\mathbf{e}_{\_}i=(0,\ldots 0,1,0,\ldots,0). Given a vector 𝐤=(k_1,,k_n)n\mathbf{k}=(k_{\_}1,\ldots,k_{\_}n)\in\mathbb{Z}^{n}, and a set of variables U_1,,U_nU_{\_}1,\ldots,U_{\_}n (resp. V_1,V_nV_{\_}1,\ldots V_{\_}n), we write

U𝐤=U_1k_1U_nk_n,V𝐤=V_1k_1V_nk_n.U^{\mathbf{k}}=U_{\_}1^{k_{\_}1}\cdots U_{\_}n^{k_{\_}n},\quad V^{\mathbf{k}}=V_{\_}1^{k_{\_}1}\cdots V_{\_}n^{k_{\_}n}.

2.2. Variables and gradings

We will be working with links in S3S^{3} and the “full” version of Heegaard Floer complex 𝒞\mathcal{C\!F\!L} for links, defined in [32]. The coefficients are in 𝔽=/2\mathbb{F}=\mathbb{Z}/2\mathbb{Z}. Let L=L_1L_nL=L_{\_}1\cup\ldots\cup L_{\_}n be an oriented link with nn components. Unless stated otherwise, we will assume that each component L_iL_{\_}i has exactly two marked points z_iz_{\_}i and w_iw_{\_}i. The corresponding link homology (L)\mathcal{H\!F\!L}(L) is a module over the polynomial ring R=𝔽[U_1,,U_n,V_1,,V_n]R=\mathbb{F}[U_{\_}1,\cdots,U_{\_}n,V_{\_}1,\cdots,V_{\_}n]. We let R_UVR_{\_}{UV} denote the ring in variables U_1,,U_n,V_1,,V_n,𝐔U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n,\mathbf{U} satisfying the relations

U_1V_1==U_nV_n=𝐔.U_{\_}1V_{\_}1=\ldots=U_{\_}nV_{\_}n=\mathbf{U}.

The actions of U_iV_iU_{\_}iV_{\_}i on the complex 𝒞(L)\mathcal{C\!F\!L}(L) are pairwise homotopic, and the action of RR on (L)\mathcal{H\!F\!L}(L) factors through R_UVR_{\_}{UV}.

Further, define

𝒞(L):=𝒞(L)_R𝔽[U_1,U_11,,U_n,U_n1,V_1,V_11,,V_n,V_n1]\mathcal{C\!F\!L}^{\infty}(L):=\mathcal{C\!F\!L}(L)\otimes_{\_}R\mathbb{F}[U_{\_}1,U_{\_}1^{-1},\cdots,U_{\_}n,U_{\_}n^{-1},V_{\_}1,V_{\_}1^{-1},\cdots,V_{\_}n,V_{\_}{n}^{-1}]

and (L):=H_(𝒞(L))\mathcal{H\!F\!L}^{\infty}(L):=H_{\_}*(\mathcal{C\!F\!L}^{\infty}(L)).

We denote by lk(L_i,L_j)\mathrm{lk}(L_{\_}i,L_{\_}j) the linking number between the components L_iL_{\_}i and L_jL_{\_}j, and write _i=_jilk(L_i,L_j)\ell_{\_}i=\sum_{\_}{j\neq i}\mathrm{lk}(L_{\_}i,L_{\_}j). Moreover, we let _L=12(_1,,_n)\ell_{\_}{L}=\dfrac{1}{2}(\ell_{\_}1,\cdots,\ell_{\_}n).

The link Floer homology has an Alexander grading A=(A_1,,A_n)A=(A_{\_}1,\ldots,A_{\_}n) valued in the lattice

_L=n+12(_1,,_n).\mathbb{H}_{\_}L=\mathbb{Z}^{n}+\frac{1}{2}(\ell_{\_}1,\ldots,\ell_{\_}n).

It also has a homological (or Maslov) grading gr_𝐰\textup{gr}_{\_}{{\bf{w}}} and an additional grading gr_𝐳\textup{gr}_{\_}{{\bf{z}}} satisfying

A_1++A_n=12(gr_𝐰gr_𝐳).A_{\_}1+\ldots+A_{\_}n=\frac{1}{2}(\textup{gr}_{\_}{{\bf{w}}}-\textup{gr}_{\_}{{\bf{z}}}).

Thanks to the relation between AA, gr_𝐰\textup{gr}_{\_}{{\bf{w}}} and gr_𝐳\textup{gr}_{\_}{{\bf{z}}}, we can determine gr_𝐳\textup{gr}_{\_}{{\bf{z}}} from the Alexander and Maslov gradings. Note that the differential on the chain complex 𝒞(L)\mathcal{C\!F\!L}(L) preserves Alexander multi-grading and changes the Maslov grading by 1. So, for any 𝐤_L\mathbf{k}\in\mathbb{H}_{\_}{L}, let 𝒞(L,𝐤)\mathcal{C\!F\!L}(L,\mathbf{k}) denote the subcomplex of 𝒞(L)\mathcal{C\!F\!L}(L) generated by the elements of A(x)=𝐤A(x)=\mathbf{k}. The variable U_iU_{\_}i decreases A_iA_{\_}i by 11, decreases gr_𝐰\textup{gr}_{\_}{{\bf{w}}} by 22 and preserves gr_𝐳\textup{gr}_{\_}{{\bf{z}}}, while the variable V_iV_{\_}i increases A_iA_{\_}{i} by 11, preserves gr_𝐰\textup{gr}_{\_}{{\bf{w}}} and decreases gr_𝐳\textup{gr}_{\_}{{\bf{z}}} by 22. Therefore, the coefficient ring for the subcomplex 𝒞(L,𝐤)\mathcal{C\!F\!L}(L,\mathbf{k}) is the subring 𝔽[U_1V_1,U_2V_2,,U_nV_n]\mathbb{F}[U_{\_}1V_{\_}1,U_{\_}2V_{\_}2,\cdots,U_{\_}nV_{\_}n] and so (L,𝐤)\mathcal{H\!F\!L}(L,\mathbf{k}) is an 𝔽[𝐔]\mathbb{F}[\mathbf{U}]-module.

For example, the homology of the unlink with nn components has one generator in Alexander degree (0,,0)(0,\ldots,0) and Maslov degree 0, and is isomorphic to the ground ring R_UVR_{\_}{UV}.

Sometimes we will need to work with the completion 𝓗𝓕𝓛(L)\bm{{\mathcal{H\!F\!L}}}(L) which is a module over the power series ring 𝔽[[U_1,,U_n,V_1,,V_n]]\mathbb{F}[[U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n]].

2.3. Specializing V_iV_{\_}i

We will need to compare the above construction of Heegaard Floer homology with more “classical” ones [23, 26, 20]. This is done by specializing V_iV_{\_}i in various ways.

First, we specialize V_i=1V_{\_}i=1 for all ii and denote the specialized complex by 𝐶𝐹𝐿\mathit{CFL}^{-} following [26]. The specialized complex still has commuting actions of U_iU_{\_}i, which are all homotopic to 𝐔\mathbf{U}. Since gr_𝐰(V_i)=0\textup{gr}_{\_}{{\bf{w}}}(V_{\_}i)=0, the specialized complex has a homological grading given by gr_𝐰\textup{gr}_{\_}{{\bf{w}}}. On the other hand, the Alexander grading becomes Alexander filtration, as follows:

Proposition 2.1.

For all 𝐤_L\mathbf{k}\in\mathbb{H}_{\_}L, there is a bijection between the generators of 𝐶𝐹𝐿\mathit{CFL}^{-} of Alexander grading 𝐤\preceq\mathbf{k} and the generators of 𝒞\mathcal{C\!F\!L} of Alexander grading exactly 𝐤\mathbf{k} i.e. generators of 𝒞(L,𝐤)\mathcal{C\!F\!L}(L,\mathbf{k}):

xV𝐤A(x)x,A(x)𝐤.x\leftrightarrow V^{\mathbf{k}-A(x)}x,\ A(x)\preceq\mathbf{k}.

The span of such generators, denoted by 𝔄(𝐤)=𝔄(L;𝐤)\mathfrak{A}^{-}(\mathbf{k})=\mathfrak{A}^{-}(L;\mathbf{k}), is a subcomplex of 𝐶𝐹𝐿\mathit{CFL}^{-}, and such subcomplexes yield a n\mathbb{Z}^{n}-filtration on 𝐶𝐹𝐿\mathit{CFL}^{-}.

Another specialization is V_i=0V_{\_}i=0 for all ii. Similarly to the above, one immediately verifies that this is equivalent to considering the associated graded complex gr𝐶𝐹𝐿\textup{gr}\mathit{CFL}^{-} with respect to the Alexander filtration.

2.4. Large surgery and L-space links

We recall the large surgery theorem of Manolescu and Ozsváth:

Theorem 2.2 ([20]).

Let L=L_1L_nL=L_{\_}1\cup\ldots\cup L_{\_}n be an nn-component link in the three-sphere. For 𝐝=(d_1,,d_n)n\mathbf{d}=(d_{\_}1,\ldots,d_{\_}n)\in\mathbb{Z}^{n} denote the 33-manifold obtained by performing d_id_{\_}i-surgery on L_iL_{\_}i for all 1in1\leq i\leq n by

Y_𝐝=S_3𝐝(L).Y_{\_}{\mathbf{d}}=S^{3}_{\_}{\mathbf{d}}(L).

Then, for d_i0d_{\_}i\gg 0 and arbitrary 𝐤\mathbf{k} we have an isomorphism of graded 𝔽[𝐔]\mathbb{F}[\mathbf{U}]-modules, up to a grading shift:

𝔄(L;𝐤)𝐶𝐹(Y_𝐝;𝔰_𝐤)\mathfrak{A}^{-}(L;\mathbf{k})\simeq\mathit{CF}^{-}(Y_{\_}{\mathbf{d}};\mathfrak{s}_{\_}{\mathbf{k}})

where 𝔰_𝐤\mathfrak{s}_{\_}{\mathbf{k}} is a Spinc\text{Spin}^{c}-structure on Y_𝐝Y_{\_}{\mathbf{d}} determined by 𝐤\mathbf{k}.

Corollary 2.3.

As a graded 𝔽[𝐔]\mathbb{F}[\mathbf{U}]-module, the homology (L,𝐤)\mathcal{H\!F\!L}(L,\mathbf{k}) splits as a direct sum of one copy of 𝔽[𝐔]\mathbb{F}[\mathbf{U}] and some 𝐔\mathbf{U}-torsion.

The link invariant h(𝐤)h(\mathbf{k}), known as the hh-function, is defined as the 12gr_𝐰-\frac{1}{2}\textup{gr}_{\_}{{\bf{w}}} for the generator of the 𝔽[𝐔]\mathbb{F}[\mathbf{U}]-summand of (L,𝐤)\mathcal{H\!F\!L}(L,\mathbf{k}).

An oriented, connected, closed 3-manifold MM is an L-space if it is a rational homology sphere, and for each Spinc\text{Spin}^{c}-structure 𝔰\mathfrak{s} on MM, one has HF(M,𝔰)𝔽[𝐔]HF^{-}(M,\mathfrak{s})\cong\mathbb{F}[\mathbf{U}]. A link LL is called an L-space link if S_3𝐝(L)S^{3}_{\_}{\mathbf{d}}(L) is an L-space for 𝐝0\mathbf{d}\gg 0. Since Dehn surgery does not depend on the orientations of the link, being an L-space link is independent of the orientations on the components of the link.

Corollary 2.4.

For an L-space link, we have (L,𝐤)=𝔽[𝐔][2h(𝐤)]\mathcal{H\!F\!L}(L,\mathbf{k})=\mathbb{F}[\mathbf{U}][-2h(\mathbf{k})] for all 𝐤\mathbf{k}.

This is a useful way to characterize L-space links [18]. That is, a link LS3L\subset S^{3} is an L-space link if and only if the link Floer homology (L)\mathcal{H\!F\!L}(L) is torsion free as an 𝔽[𝐔]\mathbb{F}[\mathbf{U}]-module.

Example 2.5.

Let O_nO_{\_}n be the unlink with nn components. Then, in Alexander grading 𝐤=(k_1,,k_n)\mathbf{k}=(k_{\_}1,\ldots,k_{\_}n) we have

𝒞(L,𝐤)=_i=1rU_i[k_i]_V_i[k_i]_+𝔽[𝐔],\mathcal{C\!F\!L}(L,\mathbf{k})=\prod_{\_}{i=1}^{r}U_{\_}i^{-[k_{\_}i]_{\_}{-}}V_{\_}i^{[k_{\_}i]_{\_}{+}}\cdot\mathbb{F}[\mathbf{U}],

where [k]_+=max(k,0)[k]_{\_}{+}=\max(k,0) and [k]_=min(k,0)[k]_{\_}{-}=\min(k,0). Note that _i=1rU_i[k_i]_V_i[k_i]_+\prod_{\_}{i=1}^{r}U_{\_}i^{-[k_{\_}i]_{\_}{-}}V_{\_}i^{[k_{\_}i]_{\_}{+}} has homological degree gr_𝐰=2_i=1k[k_i]_\textup{gr}_{\_}{{\bf{w}}}=2\sum_{\_}{i=1}^{k}[k_{\_}i]_{\_}{-}, so h(𝐤)=_i=1k[k_i]_h(\mathbf{k})=-\sum_{\_}{i=1}^{k}[k_{\_}i]_{\_}{-}.

Example 2.6.

The Hopf link T(2,2)T(2,2) with linking number 11 and the negative Hopf link T(2,2)-T(2,2) with linking number 1-1 are both L-space links. In [7], there are explicit computations of \mathcal{H\!F\!L} of these two links using the Heegaard diagrams in Figure 2. The link Floer chain complex of the positive Hopf link T(2,2)T(2,2) is a module over RR given as follows:

a=b=0,c=U_1a+V_2b,d=U_2a+V_1b\partial a=\partial b=0,\quad\partial c=U_{\_}1a+V_{\_}2b,\quad\partial d=U_{\_}2a+V_{\_}1b

where the gradings of a,ba,b are the following:

A(a)=(12,12),gr_𝐰(a)=0,A(b)=(12,12),gr_𝐰(b)=2.A(a)=\left(\frac{1}{2},\frac{1}{2}\right),\ \textup{gr}_{\_}{{\bf{w}}}(a)=0,\quad A(b)=\left(-\frac{1}{2},-\frac{1}{2}\right),\ \textup{gr}_{\_}{{\bf{w}}}(b)=-2.

The gradings of c,dc,d are as follows

gr_𝐰(c)=gr_𝐰(d)=gr_𝐳(c)=gr_𝐳(d)=1.\textup{gr}_{\_}{{\bf{w}}}(c)=\textup{gr}_{\_}{{\bf{w}}}(d)=\textup{gr}_{\_}{{\bf{z}}}(c)=\textup{gr}_{\_}{{\bf{z}}}(d)=-1.

Hence, the full homology of the Hopf link T(2,2)T(2,2) is generated by a,ba,b and can be written as

(T(2,2))=Ra,bU_1a=V_2b,U_2a=V_1b.\mathcal{H\!F\!L}(T(2,2))=\frac{R\langle a,b\rangle}{U_{\_}1a=V_{\_}2b,U_{\_}2a=V_{\_}1b}.

The link Floer chain complex of the negative Hopf link is the dual complex of 𝒞(T(2,2))\mathcal{C\!F\!L}(T(2,2)), i.e.,

c=d=0,,a=U_1c+U_2d,b=V_1d+V_2c\partial c^{\prime}=\partial d^{\prime}=0,\quad,\partial a^{\prime}=U_{\_}1c^{\prime}+U_{\_}2d^{\prime},\quad\partial b^{\prime}=V_{\_}1d^{\prime}+V_{\_}2c^{\prime}

where the gradings of c,dc^{\prime},d^{\prime} are

gr_𝐰(c)=gr_𝐳(c)=gr_𝐰(d)=gr_𝐳(d)=1,A(c)=(12,12),A(d)=(12,12).\textup{gr}_{\_}{{\bf{w}}}(c^{\prime})=\textup{gr}_{\_}{{\bf{z}}}(c^{\prime})=\textup{gr}_{\_}{{\bf{w}}}(d^{\prime})=\textup{gr}_{\_}{{\bf{z}}}(d^{\prime})=1,\quad A(c^{\prime})=\left(\dfrac{1}{2},-\dfrac{1}{2}\right),\quad A(d^{\prime})=\left(-\dfrac{1}{2},\dfrac{1}{2}\right).

Hence, the full homology of T(2,2)-T(2,2) can be written as

(T(2,2))=Rc,dU_1c=U_2d,V_2c=V_1d.\mathcal{H\!F\!L}(-T(2,2))=\frac{R\langle c^{\prime},d^{\prime}\rangle}{U_{\_}1c^{\prime}=U_{\_}2d^{\prime},V_{\_}2c^{\prime}=V_{\_}1d^{\prime}}.
Refer to caption𝐰_2{\bf{w}}_{\_}2𝐰_1{\bf{w}}_{\_}1𝐳_2{\bf{z}}_{\_}2𝐳_1{\bf{z}}_{\_}1aaccbbdd𝐳_2{\bf{z}}_{\_}2𝐰_1{\bf{w}}_{\_}1𝐰_2{\bf{w}}_{\_}2𝐳_1{\bf{z}}_{\_}1cc^{\prime}aa^{\prime}dd^{\prime}bb^{\prime}
Figure 2. Left: genus 0 Heegaard diagram for T(2,2)T(2,2), Right: genus 0 Heegaard diagram for T(2,2)-T(2,2)

2.5. Cobordism maps and link TQFT

We first review 3-manifolds with multi-based links and decorated cobordisms between them. A 3-manifold with a multi-based link consists of an oriented closed 3-manifold YY, an oriented, embedded link LYL\subset Y together with disjoint collection of basepoints 𝐰{\bf{w}} and 𝐳{\bf{z}} on LL such that each component L_iL_{\_}i of LL has at least two basepoints z_i,w_iz_{\_}i,w_{\_}i, and the basepoints alternate between those in 𝐰{\bf{w}} and those in 𝐳{\bf{z}} when one traverses a component of LL. The basepoints w_iw_{\_}i and z_iz_{\_}i correspond to the variables U_iU_{\_}i and V_iV_{\_}i in a polynomial ring 𝔽[U_𝐰,V_𝐳]=𝔽[U_1,U_2,,U_m,V_1,V_2,,V_m]\mathbb{F}[U_{\_}{{\bf{w}}},V_{\_}{{\bf{z}}}]=\mathbb{F}[U_{\_}1,U_{\_}2,\cdots,U_{\_}m,V_{\_}{1},V_{\_}2,\cdots,V_{\_}m] where m=|𝐰|=|𝐳|m=|{\bf{w}}|=|{\bf{z}}|. Then, 𝒞(L,𝐰,𝐳)\mathcal{C\!F\!L}(L,{\bf{w}},{\bf{z}}) is defined as a curved complex over 𝔽[U_𝐰,V_𝐳]\mathbb{F}[U_{\_}{{\bf{w}}},V_{\_}{{\bf{z}}}].

In this paper, we mainly consider the case that each component of a link has exactly two basepoints, i.e., the link component L_iL_{\_}i contains w_iw_{\_}i and z_iz_{\_}i in 𝐰{\bf{w}} and 𝐳{\bf{z}}, respectively, and 𝔽[U_𝐰,V_𝐳]=R\mathbb{F}[U_{\_}{{\bf{w}}},V_{\_}{{\bf{z}}}]=R. Furthermore, for simplicity, we will drop 𝐰{\bf{w}} and 𝐳{\bf{z}} from the notation of a multi-based link if the context is clear.

A coloring of a multi-based link (L,𝐰,𝐳)(L,{\bf{w}},{\bf{z}}) is a map σ:𝐰𝐳P\sigma:{\bf{w}}\cup{\bf{z}}\rightarrow\mathrm{P}, where P\mathrm{P} is a finite set, considered as the set of colors. Corresponding to the set of colors P={p_1,p_2,,p_k}P=\{p_{\_}1,p_{\_}2,\cdots,p_{\_}k\}, a polynomial ring

_P:=𝔽[X_p_1,X_p_2,,X_p_k]\mathcal{R}^{-}_{\_}P:=\mathbb{F}[X_{\_}{p_{\_}1},X_{\_}{p_{\_}2},\cdots,X_{\_}{p_{\_}k}]

is defined, which clearly is a 𝔽[U_𝐰,V_𝐳]\mathbb{F}[U_{\_}{{\bf{w}}},V_{\_}{{\bf{z}}}]-module. For a colored multi-based link (L,𝐰,𝐳,σ)(L,{\bf{w}},{\bf{z}},\sigma)

𝒞(L,𝐰,𝐳,σ)=𝒞(L,𝐰,𝐳)_𝔽[U_𝐰,V_𝐳]_P\mathcal{C\!F\!L}(L,{\bf{w}},{\bf{z}},\sigma)=\mathcal{C\!F\!L}(L,{\bf{w}},{\bf{z}})\otimes_{\_}{\mathbb{F}[U_{\_}{{\bf{w}}},V_{\_}{{\bf{z}}}]}\mathcal{R}_{\_}{P}^{-}
Definition 2.7.

[33, Definition 1.3] A decorated link cobordism from a 3-manifold with a multi-based link (Y_1,(L_1,𝐰_1,𝐳_1))(Y_{\_}1,(L_{\_}1,{\bf{w}}_{\_}1,{\bf{z}}_{\_}1)) to another one (Y_2,(L_2,𝐰_2,𝐳_2))(Y_{\_}2,(L_{\_}2,{\bf{w}}_{\_}2,{\bf{z}}_{\_}2)) consists of a pair (W,σ)(W,\mathcal{F}^{\sigma}) such that

  1. (1)

    WW is a compact 4-manifold with W=Y_1Y_2\partial W=-Y_{\_}1\sqcup Y_{\_}2.

  2. (2)

    =(Σ,A)\mathcal{F}=(\Sigma,A) is an oriented, properly embedded surface Σ\Sigma in WW, along with a properly embedded 1-manifold AA in Σ\Sigma, called dividing arcs. Further, ΣA\Sigma\setminus A consists of two disjoint (possibly disconnected) subsurfaces, Σ_𝐰\Sigma_{\_}{\bf{w}} and Σ_𝐳\Sigma_{\_}{\bf{z}}, such that the intersection of the closures of Σ_𝐰\Sigma_{\_}{\bf{w}} and Σ_𝐳\Sigma_{\_}{\bf{z}} is AA.

  3. (3)

    Σ=L_1L_2\partial\Sigma=-L_{\_}1\cup L_{\_}2.

  4. (4)

    Each component of L_1AL_{\_}1\setminus A (and L_2AL_{\_}2\setminus A) contains exactly one basepoint.

  5. (5)

    The 𝐰{\bf{w}} basepoints are all in Σ_𝐰\Sigma_{\_}{\bf{w}} and the 𝐳{\bf{z}} basepoints are all in Σ_𝐳\Sigma_{\_}{\bf{z}}.

  6. (6)

    \mathcal{F} is equipped with a coloring σ\sigma, i.e. a map σ:C(ΣA)P\sigma:C(\Sigma\setminus A)\to P, where C(ΣA)C(\Sigma\setminus A) denotes the set of component of ΣA\Sigma\setminus A.

To a decorated link cobordism (W,σ)(W,\mathcal{F}^{\sigma}) and a Spinc\text{Spin}^{c} structure 𝔰\mathfrak{s} on WW, Zemke[32, Theorem A] associated a Spinc\text{Spin}^{c} functorial chain maps

F_W,σ,𝔰:𝒞(Y_1,L_1,𝐰_1,𝐳_1,σ_1,𝔰_Y_1)𝒞(Y_2,L_2,𝐰_2,𝐳_2,σ_2,𝔰_Y_2).F_{\_}{W,\mathcal{F}^{\sigma},\mathfrak{s}}:\mathcal{C\!F\!L}(Y_{\_}1,L_{\_}1,{\bf{w}}_{\_}1,{\bf{z}}_{\_}1,\sigma_{\_}1,\mathfrak{s}\mid_{\_}{Y_{\_}1})\rightarrow\mathcal{C\!F\!L}(Y_{\_}2,L_{\_}2,{\bf{w}}_{\_}2,{\bf{z}}_{\_}2,\sigma_{\_}2,\mathfrak{s}\mid_{\_}{Y_{\_}{2}}).

Here, σ_j\sigma_{\_}j denotes the colorings on L_jL_{\_}j obtained by restricting σ\sigma, for j=1,2j=1,2. The maps are _P\mathcal{R}_{\_}P^{-}-equivariant, P\mathbb{Z}^{P}-filtered, and are invariants up to _P\mathcal{R}_{\_}P^{-}-equivariant, P\mathbb{Z}^{P}-filtered chain homotopies.

Another version of functorial maps for decorated cobordism between links have been independently defined by the first author and Eftekhary in [2].

Convention 3.

In this paper, we consider the case that every component of a link has exactly two basepoints (unless when we stabilize them), i.e., the link component L_iL_{\_}i contains w_iw_{\_}i and z_iz_{\_}i in 𝐰{\bf{w}} and 𝐳{\bf{z}}, respectively, and so 𝔽[U_𝐰,V_𝐳]=R\mathbb{F}[U_{\_}{{\bf{w}}},V_{\_}{{\bf{z}}}]=R. Moreover, mostly we work with special cobordisms that every connected component of Σ\Sigma is an annulus, decorated with two parallel vertical dividing arcs. More precisely, for j=1,2j=1,2, L_j=_i=1nL_i,jL_{\_}j=\coprod_{\_}{i=1}^{n}L_{\_}{i,j} and Σ=_i=1nΣ_i\Sigma=\coprod_{\_}{i=1}^{n}\Sigma_{\_}{i} where each Σ_i\Sigma_{\_}i is an annulus with Σ_i=L_i,1L_i,2\partial\Sigma_{\_}i=-L_{\_}{i,1}\sqcup L_{\_}{i,2}. Further, each A_i=AΣ_iA_{\_}i=A\cap\Sigma_{\_}i consists of two parallel, vertical dividing arcs connecting L_i,1L_{\_}{i,1} to L_i,2L_{\_}{i,2} and dividing Σ_i\Sigma_{\_}i into two rectangles, one containing w_i,1,w_i,2w_{\_}{i,1},w_{\_}{i,2} and another containing z_i,1,z_i,2z_{\_}{i,1},z_{\_}{i,2} basepoints. Finally, our coloring set PP, which is the codomain of σ\sigma, contains exactly 2n2n colors, and _PR\mathcal{R}_{\_}P^{-}\cong R such that under this identification X_σ_j(w_i,j)X_{\_}{\sigma_{\_}j(w_{\_}{i,j})} and X_σ_j(z_i,j)X_{\_}{\sigma_{\_}j(z_{\_}{i,j})} are identified with U_iU_{\_}i and V_iV_{\_}i, respectively. Here, j=1,2j=1,2 and w_i,j,z_i,jw_{\_}{i,j},\ z_{\_}{i,j} are the basepoints on L_i,jL_{\_}{i,j}. Thus, if we do not emphasis on the basepoints, dividing curves and the colorings, we automatically mean this fixed convention.

For a decorated cobordism (W,)(W,\mathcal{F}) as above the cobordism maps F_W,,𝔰F_{\_}{W,\mathcal{F},\mathfrak{s}} are RR-equivariant and 2n\mathbb{Z}^{2n}-filtered. The grading changes under the cobordism maps F_W,,𝔰F_{\_}{W,\mathcal{F},\mathfrak{s}} are as follows:

Theorem 2.8.

(Special case of [33, Theorems 1.4 and 2.14]) Suppose (W,)(W,\mathcal{F}) is a decorated link cobordism from (Y_1,L_1)(Y_{\_}1,L_{\_}1) to (Y_2,L_2)(Y_{\_}2,L_{\_}2). Then,

  1. (1)

    If c_1(𝔰|_Y_1)c_{\_}1(\mathfrak{s}|_{\_}{Y_{\_}{1}}) and c_1(𝔰|_Y_2)c_{\_}{1}(\mathfrak{s}|_{\_}{Y_{\_}{2}}) are torsion, then F_W,,𝔰F_{\_}{W,\mathcal{F},\mathfrak{s}} is graded with respect to gr_𝐰\textup{gr}_{\_}{\bf{w}}, and satisfies

    gr_𝐰(F_W,,𝔰(x))gr_𝐰(x)=c_1(𝔰)22χ(W)3σ(W)4.\textup{gr}_{\_}{\bf{w}}(F_{\_}{W,\mathcal{F},\mathfrak{s}}(x))-\textup{gr}_{\_}{{\bf{w}}}(x)=\dfrac{c_{\_}{1}(\mathfrak{s})^{2}-2\chi(W)-3\sigma(W)}{4}.
  2. (2)

    If c_1(𝔰|_Y_1𝑃𝐷[L_1])c_{\_}{1}(\mathfrak{s}|_{\_}{Y_{\_}{1}}-\mathit{PD}\left[L_{\_}1\right]) and c_1(𝔰|_Y_2𝑃𝐷[L_2])c_{\_}{1}(\mathfrak{s}|_{\_}{Y_{\_}{2}}-\mathit{PD}\left[L_{\_}2\right]) are torsion, then the map F_W,,𝔰F_{\_}{W,\mathcal{F},\mathfrak{s}} is graded with respect to gr_𝐳\textup{gr}_{\_}{{\bf{z}}}, and satisfies

    gr_𝐳(F_W,,𝔰(x))gr_𝐳(x)=(c_1(𝔰)𝑃𝐷[Σ])22χ(W)3σ(W)4.\textup{gr}_{\_}{{\bf{z}}}(F_{\_}{W,\mathcal{F},\mathfrak{s}}(x))-\textup{gr}_{\_}{{\bf{z}}}(x)=\dfrac{(c_{\_}{1}(\mathfrak{s})-\mathit{PD}\left[\Sigma\right])^{2}-2\chi(W)-3\sigma(W)}{4}.
  3. (3)

    Suppose L_1Y_1L_{\_}1\subset Y_{\_}1 and L_2Y_2L_{\_}2\subset Y_{\_}2 are null-homologous links, i.e. [L_i,j]=0[L_{\_}{i,j}]=0 in H_1(Y_j,)H_{\_}1(Y_{\_}{j},\mathbb{Z}) for 1in1\leq i\leq n and j=1,2j=1,2. Moreover, assume both c_1(𝔰|_Y_1)c_{\_}1(\mathfrak{s}|_{\_}{Y_{\_}1}) and c_1(𝔰|_Y_2)c_{\_}1(\mathfrak{s}|_{\_}{Y_{\_}2}) are torsion. Then,

    A_i(F_W,,𝔰(x))A_i(x)=c_1(𝔰),[Σ^_i][Σ^][Σ^_i]2,A_{\_}{i}(F_{\_}{W,\mathcal{F},\mathfrak{s}}(x))-A_{\_}{i}(x)=\dfrac{\left\langle c_{\_}{1}(\mathfrak{s}),\left[\widehat{\Sigma}_{\_}{i}\right]\right\rangle-\left[\widehat{\Sigma}\right]\cdot\left[\widehat{\Sigma}_{\_}{i}\right]}{2},

    where Σ^_i\widehat{\Sigma}_{\_}{i} denotes the closure of Σ_i\Sigma_{\_}i by adding arbitrary Seifert surfaces of L_i,1Y_1L_{\_}{i,1}\subset Y_{\_}1 and L_i,2Y_2L_{\_}{i,2}\subset Y_{\_}2, and [Σ^]=_i=1n[Σ^_i].\left[\widehat{\Sigma}\right]=\sum_{\_}{i=1}^{n}\left[\widehat{\Sigma}_{\_}i\right].

Theorem 2.9.

Assume that (W,):(S3,L_1)(S3,L_2)(W,\mathcal{F}):(S^{3},L_{\_}1)\rightarrow(S^{3},L_{\_}2) is a decorated link cobordism with b_2+(W)=0b_{\_}2^{+}(W)=0 as in Convention 3. Then for all 𝔰\mathfrak{s} and 𝐤_L_1\mathbf{k}\in\mathbb{H}_{\_}{L_{\_}1} the induced map on homology

F_W,,𝔰:(L_1,𝐤)(L_2,𝐤+𝐝)F_{\_}{W,\mathcal{F},\mathfrak{s}}:\mathcal{H\!F\!L}^{\infty}(L_{\_}1,\mathbf{k})\to\mathcal{H\!F\!L}^{\infty}(L_{\_}2,\mathbf{k}+\mathbf{d})

is an isomorphism, where 𝐝\mathbf{d} is the Alexander multi-degree of F_W,,𝔰F_{\_}{W,\mathcal{F},\mathfrak{s}}.

Proof.

Consider the diagram

𝐶𝐹(S3,𝐰_1){\mathit{CF}^{\infty}(S^{3},{\bf{w}}_{\_}1)}𝐶𝐹(S3,𝐰_2){\mathit{CF}^{\infty}(S^{3},{\bf{w}}_{\_}2)}𝒞(L_1,𝐤){\mathcal{C\!F\!L}^{\infty}(L_{\_}1,\mathbf{k})}𝒞(L_2,𝐤+𝐝){\mathcal{C\!F\!L}^{\infty}(L_{\_}2,\mathbf{k}+\mathbf{d})}F_W,𝔰\scriptstyle{F_{\_}{W,\mathfrak{s}}}F_W,,𝔰\scriptstyle{F_{\_}{W,\mathcal{F},\mathfrak{s}}}

where the left (resp. right) vertical arrow is defined by sending x𝐶𝐹(S3,𝐰_1)x\in\mathit{CF}^{\infty}(S^{3},{\bf{w}}_{\_}1) (resp. x𝐶𝐹(S3,𝐰_2)x\in\mathit{CF}^{\infty}(S^{3},{\bf{w}}_{\_}2)) to V𝐤A(x)x𝒞(L_1,𝐤)V^{\mathbf{k}-A(x)}x\in\mathcal{C\!F\!L}^{\infty}(L_{\_}1,\mathbf{k}) (resp. V𝐤+𝐝A(x)x𝒞(L_2,𝐤+𝐝)V^{\mathbf{k}+\mathbf{d}-A(x)}x\in\mathcal{C\!F\!L}^{\infty}(L_{\_}2,\mathbf{k}+\mathbf{d})). Moreover, F_W,𝔰F_{\_}{W,\mathfrak{s}} is the cobordism map corresponding to WW and Σ_𝐰\Sigma_{\_}{{\bf{w}}} as defined in [27]. Similar to Proposition 2.1, it is easy to see that the vertical maps are chain maps and define an isomorphism between the chain complexes. Moreover, it follows from the definition of the cobordism maps that this diagram commutes. By the proof of [22, Theorem 9.6] the induced map on homology by F_W,𝔰F_{\_}{W,\mathfrak{s}} is an isomorphism and thus the induced map on homology by F_W,,𝔰F_{\_}{W,\mathcal{F},\mathfrak{s}} is an isomorphism as well.

Corollary 2.10.

Assume (W,):(S3,L_1)(S3,L_2)(W,\mathcal{F}):(S^{3},L_{\_}1)\rightarrow(S^{3},L_{\_}2) is a decorated link cobordism with b_2+(W)=0b_{\_}2^{+}(W)=0, and L_1L_{\_}1 and L_2L_{\_}2 are LL-space links. If F_W,,𝔰F_{\_}{W,\mathcal{F},\mathfrak{s}} has Alexander multi-degree 𝐝\mathbf{d} and homological degree dd then for all 𝐤_L_1\mathbf{k}\in\mathbb{H}_{\_}{L_{\_}1} the induced map on homology

F_W,,𝔰:(L_1,𝐤)(L_2,𝐤+𝐝)F_{\_}{W,\mathcal{F},\mathfrak{s}}:\mathcal{H\!F\!L}(L_{\_}1,\mathbf{k})\to\mathcal{H\!F\!L}(L_{\_}2,\mathbf{k}+\mathbf{d})

is injective and completely determined by its homological degree.

Proof.

Let z_L_j(𝐤)z_{\_}{L_{\_}j}(\mathbf{k}) denote the generator of (L_j,𝐤)𝔽[𝐔]\mathcal{H\!F\!L}(L_{\_}j,\mathbf{k})\cong\mathbb{F}[\mathbf{U}] of homological degree 2h_L_j(𝐤)-2h_{\_}{L_{\_}j}(\mathbf{k}), for j=1,2j=1,2. Then

F_W,,𝔰(z_L_1(𝐤))=𝐔m(𝐤)z_L_2(𝐤+𝐝).F_{\_}{W,\mathcal{F},\mathfrak{s}}\left(z_{\_}{L_{\_}1}(\mathbf{k})\right)=\mathbf{U}^{m(\mathbf{k})}z_{\_}{L_{\_}2}(\mathbf{k}+\mathbf{d}).

where

m(𝐤)=(d+2h_L_2(𝐤+𝐝)2h_L_1(𝐤))/2.m(\mathbf{k})=-\left(d+2h_{\_}{L_{\_}2}(\mathbf{k}+\mathbf{d})-2h_{\_}{L_{\_}1}(\mathbf{k})\right)/{2}.

3. Surgery maps

3.1. Crossing changes

As shown in Figure 3, one can locally change a positive or negative crossing by performing (1)(-1)-surgery on the specified red unknot. In this section, we associate a link cobordism with a simple decoration to these crossing change surgeries and then study the properties of the corresponding cobordism maps.

Refer to caption++--++1-11-1
Figure 3. Crossing changes: In top (resp. bottom) figure, (1)(-1)-surgery on the red unknot will change the positive (resp. negative) crossing to the negative (resp. positive) crossing.

Suppose L_+=_i=1nL_i,+L_{\_}+=\coprod_{\_}{i=1}^{n}L_{\_}{i,+} is an nn-component link in S3S^{3}, and L_L_{\_}- is the link obtained from L_+L_{\_}+ by changing a positive crossing between different components to a negative crossing. So, L_L_{\_}- will have nn components as well. Denote the component of L_L_{\_}- corresponding to L_i,+L_{\_}{i,+} by L_i,L_{\_}{i,-}. Let WW be the cobordism from S3S^{3} to S3S^{3} obtained by attaching a 22-handle to S3×{1}S^{3}\times\{1\} in S3×[0,1]S^{3}\times[0,1], along the (1)(-1)-framed unknot as in the top of Figure 3. Then, the embedded surface Σ=L_+×[0,1]\Sigma=L_{\_}+\times[0,1] in WW gives a cobordism from L_+L_{\_}+ to L_L_{\_}-. The surface Σ\Sigma consists of nn connected components, all of them annuli. Denote the component of Σ\Sigma that bounds L_i,+-L_{\_}{i,+} and L_i,L_{\_}{i,-} by Σ_i\Sigma_{\_}i. Assume each connected component L_i,+L_{\_}{i,+} of L_+L_{\_}+ contains exactly two basepoints w_i,+w_{\_}{i,+}, z_i,+z_{\_}{i,+}, and denote the corresponding basepoints on L_i,L_{\_}{i,-} by w_i,w_{\_}{i,-} and z_i,z_{\_}{i,-}, respectively. Decorate each Σ_i\Sigma_{\_}i with two parallel and vertical arcs A_iA_{\_}i to divide Σ_i\Sigma_{\_}i into two rectangles, such that one of these rectangles contains the basepoins z_i,±z_{\_}{i,\pm}, and the other one contains w_i,±w_{\_}{i,\pm}. Then, for =(Σ,A=_i=1nA_i)\mathcal{F}=(\Sigma,A=\coprod_{\_}{i=1}^{n}A_{\_}i) colored as in Convention 3, the pair (W,)(W,\mathcal{F}) gives a decorated cobordism from (S3,L_+)(S^{3},L_{\_}+) to (S3,L_)(S^{3},L_{\_}-). Similarly, we define a decorated cobordism from (S3,L_)(S^{3},L_{\_}-) to (S3,L_+)(S^{3},L_{\_}+) using the unknot in the bottom of Figure 3 as well.

Proposition 3.1.

Let (W,):(S3,L_+)(S3,L_)(W,\mathcal{F}):(S^{3},L_{\_}{+})\rightarrow(S^{3},L_{\_}{-}) be the decorated link cobordism induced from attaching a 2-handle along the (1)(-1)-framed unknot KK so that a positive crossing becomes a negative crossing, as above. Suppose that the link components L_i,+,L_j,+L_{\_}{i,+},L_{\_}{j,+} are passing through the unknot KK with i<ji<j. Let 𝔰_k\mathfrak{s}_{\_}k be the Spinc\text{Spin}^{c} structure on WW such that

c_1(𝔰_k),[S2]=2k+1\langle c_{\_}{1}(\mathfrak{s}_{\_}k),[S^{2}]\rangle=2k+1

where [S2][S^{2}] is the generator of H_2(W)H_{\_}{2}(W) corresponding to the attached 2-handle. Define ψ_k:=F_W,,𝔰_k\psi_{\_}k:=F_{\_}{W,\mathcal{F},\mathfrak{s}_{\_}k}, the corresponding cobordism map in link Floer homology. Then

gr_𝐰(ψ_k)=gr_𝐳(ψ_k)=k2k,\textup{gr}_{\_}{{\bf{w}}}(\psi_{\_}k)=\textup{gr}_{\_}{{\bf{z}}}(\psi_{\_}k)=-k^{2}-k,

and

A_i(ψ_k)=A_j(ψ_k)=k+1/2.A_{\_}{i}(\psi_{\_}k)=-A_{\_}{j}(\psi_{\_}k)=k+1/2.

Moreover, A_l(ψ_k)=0A_{\_}{l}(\psi_{\_}k)=0 for all li,jl\neq i,j. In particular, A(ψ_0)=12(𝐞_i𝐞_j)A(\psi_{\_}0)=\dfrac{1}{2}(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j), while A(ψ_1)=12(𝐞_i𝐞_j)A(\psi_{\_}{-1})=-\dfrac{1}{2}(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j) and both ψ_0,ψ_1\psi_{\_}0,\psi_{\_}{-1} have homological grading zero.

Proof.

By a direct computation, χ(W)=1,σ(W)=1\chi(W)=1,\sigma(W)=-1 and

c_1(𝔰_k)2=(c_1(𝔰_k)PD[Σ])2=(2k+1)2.c_{\_}{1}(\mathfrak{s}_{\_}{k})^{2}=(c_{\_}{1}(\mathfrak{s}_{\_}{k})-PD[\Sigma])^{2}=-(2k+1)^{2}.

By Theorem 2.8,

Δgr_𝐰=Δgr_𝐳=(2k+1)2+14=k2k.\Delta\textup{gr}_{\_}{{\bf{w}}}=\Delta\textup{gr}_{\_}{{\bf{z}}}=\dfrac{-(2k+1)^{2}+1}{4}=-k^{2}-k.

For the Alexander grading, observe that [Σ^]=0H_2(W)[\widehat{\Sigma}]=0\in H_{\_}{2}(W). Then

c_1(𝔰_k),[Σ^_i]=(2k+1)[S2],[Σ^_i]=(2k+1)lk(K,L_i).\langle c_{\_}{1}(\mathfrak{s}_{\_}{k}),[\widehat{\Sigma}_{\_}{i}]\rangle=-(2k+1)\langle[S^{2}],[\widehat{\Sigma}_{\_}{i}]\rangle=-(2k+1)\mathrm{lk}(K,L_{\_}i).

Similarly, c_1(𝔰_k),[Σ^_j]=(2k+1)lk(K,L_j)\langle c_{\_}{1}(\mathfrak{s}_{\_}{k}),[\widehat{\Sigma}_{\_}{j}]\rangle=-(2k+1)\mathrm{lk}(K,L_{\_}j). Since, all other components in LL do not interact with KK, by Theorem 2.8, ΔA_l=0\Delta A_{\_}{l}=0 for all li,jl\neq i,j, and

ΔA_i=ΔA_j=k+1/2.\Delta A_{\_}{i}=-\Delta A_{\_}{j}=k+1/2.

Example 3.2.

By Example 2.6, the full homology of the Hopf link T(2,2)T(2,2) has two generators a,ba,b and is given by

(T(2,2))=Ra,baU_1=bV_2,aU_2=bV_1.\mathcal{H\!F\!L}(T(2,2))=\frac{R\langle a,b\rangle}{aU_{\_}1=bV_{\_}2,aU_{\_}2=bV_{\_}1}.

Cobordism maps ψ_k:(T(2,2))(O_2)\psi_{\_}k:\mathcal{H\!F\!L}(T(2,2))\to\mathcal{H\!F\!L}(O_{\_}2) are nonzero by Corollary 2.10, since T(2,2)T(2,2) and O_2O_{\_}2 are LL-space links and WW is a nonpositive definite cobordism. Thus, the grading shifts from Proposition 3.1 will determine ψ_k\psi_{\_}k. Therefore,

ψ_0(a)=V_1,ψ_0(b)=U_2;ψ_1(a)=V_2,ψ_1(b)=U_1.\psi_{\_}0(a)=V_{\_}1,\ \psi_{\_}0(b)=U_{\_}2;\quad\psi_{\_}{-1}(a)=V_{\_}2,\ \psi_{\_}{-1}(b)=U_{\_}1.

In general, we have

ψ_k(a)={V_1k+1U_2k𝐔k(k1)2ifk0V_2kU_11k𝐔(k+1)(k+2)2ifk1,ψ_k(b)={V_1kU_2k+1𝐔k(k1)2ifk0V_21kU_1k𝐔(k+1)(k+2)2ifk1.\psi_{\_}k(a)=\begin{cases}V_{\_}1^{k+1}U_{\_}2^{k}\mathbf{U}^{\frac{k(k-1)}{2}}&\text{if}\ k\geq 0\\ V_{\_}2^{-k}U_{\_}1^{-1-k}\mathbf{U}^{\frac{(k+1)(k+2)}{2}}&\text{if}\ k\leq-1,\end{cases}\psi_{\_}k(b)=\begin{cases}V_{\_}1^{k}U_{\_}2^{k+1}\mathbf{U}^{\frac{k(k-1)}{2}}&\text{if}\ k\geq 0\\ V_{\_}2^{-1-k}U_{\_}1^{-k}\mathbf{U}^{\frac{(k+1)(k+2)}{2}}&\text{if}\ k\leq-1.\end{cases}
Example 3.3.

One can also regard L_+L_{\_}{+} as the 2-component unlink, and L_L_{\_}{-} as the negative Hopf link. By Example 2.6, the full homology of T(2,2)-T(2,2) is generated by c,dc^{\prime},d^{\prime} with the relations:

(T(2,2))=Rc,dcU_1=dU_2,cV_2=dV_1.\mathcal{H\!F\!L}(-T(2,2))=\frac{R\langle c^{\prime},d^{\prime}\rangle}{c^{\prime}U_{\_}1=d^{\prime}U_{\_}2,c^{\prime}V_{\_}2=d^{\prime}V_{\_}1}.

By Corollary 2.10 the cobordism maps ψ_k:(O_2)(T(2,2))\psi_{\_}{k}:\mathcal{H\!F\!L}(O_{\_}2)\rightarrow\mathcal{H\!F\!L}(-T(2,2)) are nonzero and determined by the grading shift. Therefore,

ψ_0(1)=c,ψ_1(1)=d.\psi_{\_}{0}(1)=c^{\prime},\quad\psi_{\_}{-1}(1)=d^{\prime}.

In general, by the grading reasons we have

ψ_k={V_1kU_2k𝐔k(k1)2ψ_0ifk0V_21kU_11k𝐔(k+1)(k+2)2ψ_1ifk1.\psi_{\_}k=\begin{cases}V_{\_}1^{k}U_{\_}2^{k}\mathbf{U}^{\frac{k(k-1)}{2}}\psi_{\_}0&\text{if}\ k\geq 0\\ V_{\_}2^{-1-k}U_{\_}1^{-1-k}\mathbf{U}^{\frac{(k+1)(k+2)}{2}}\psi_{\_}{-1}&\text{if}\ k\leq-1.\end{cases}
Proposition 3.4.

For any link L=L_+L=L_{\_}+, the maps ψ_k\psi_{\_}k have the following properties:

  • (a)

    For k0k\geq 0, we have ψ_k=(V_iU_j)k𝐔k(k1)2ψ_0\psi_{\_}k=(V_{\_}iU_{\_}j)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}\psi_{\_}0.

  • (b)

    For k1k\leq-1, we have ψ_k=(V_jU_i)1k𝐔(k+1)(k+2)2ψ_1.\psi_{\_}k=(V_{\_}jU_{\_}i)^{-1-k}\mathbf{U}^{\frac{(k+1)(k+2)}{2}}\psi_{\_}{-1}.

  • (c)

    We have V_jψ_0=V_iψ_1V_{\_}j\psi_{\_}0=V_{\_}i\psi_{\_}{-1} and U_iψ_0=U_jψ_1U_{\_}i\psi_{\_}0=U_{\_}j\psi_{\_}{-1}.

Proof.

The Proposition holds for the special case that L_+=O_2L_{\_}+=O_{\_}2 and L_=T(2,2)L_{\_}-=-T(2,2) by Example 3.3. We use the functoriality of link Floer homology and the properties of cobordism maps to show that the general case follows from this special case.

We stabilize L_±L_{\_}{\pm} by adding two extra pairs of basepoints w_i,±,z_i,±w_{\_}{i,\pm}^{\prime},z_{\_}{i,\pm}^{\prime} and w_j,±,z_j,±w_{\_}{j,\pm}^{\prime},z_{\_}{j,\pm}^{\prime} to L_i,±L_{\_}{i,\pm} and L_j,±L_{\_}{j,\pm}, respectively. Denote the stabilized links by L_±L^{\prime}_{\_}{\pm}. Moreover, we extend the coloring σ\sigma on L_±L_{\_}{\pm} to a coloring σ\sigma^{\prime} on L_±L^{\prime}_{\_}{\pm} so that its codomain is P=P{p_i,p_j}{P}^{\prime}=P\sqcup\{p^{\prime}_{\_}i,p^{\prime}_{\_}{j}\} and

σ(w_i,±)=p_i,σ(w_j,±)=p_j,σ(z_i,±)=σ(z_i)andσ(z_j,±)=σ(z_j).\sigma^{\prime}(w_{\_}{i,\pm}^{\prime})=p^{\prime}_{\_}i,\quad\sigma^{\prime}(w^{\prime}_{\_}{j,\pm})=p^{\prime}_{\_}j,\quad\sigma^{\prime}(z^{\prime}_{\_}{i,\pm})=\sigma(z_{\_}i)\quad\text{and}\quad\sigma^{\prime}(z^{\prime}_{\_}{j,\pm})=\sigma(z_{\_}j).

Note that σ\sigma^{\prime} restricts to a coloring of L_±L_{\_}{\pm} with codomain P{P}^{\prime}, and abusing the notation we denote it by σ\sigma^{\prime}. The isomorphism _P𝔽[U_1,,U_n,V_1,,V_n]\mathcal{R}_{\_}{P}^{-}\cong\mathbb{F}[U_{\_}1,\cdots,U_{\_}n,V_{\_}1,\cdots,V_{\_}n] extends to an isomorphism

_P𝔽[U_1,,U_n,V_1,,V_n,U_i,U_j]\mathcal{R}_{\_}{P^{\prime}}^{-}\cong\mathbb{F}[U_{\_}1,\cdots,U_{\_}n,V_{\_}1,\cdots,V_{\_}n,U_{\_}i^{\prime},U_{\_}j^{\prime}]

by sending X_p_iX_{\_}{p^{\prime}_{\_}i} and X_p_jX_{\_}{p^{\prime}_{\_}j} to U_iU_{\_}i^{\prime} and U_jU_{\_}j^{\prime}, respectively. Under this isomorphism

(L_±σ)(L_±)_𝔽𝔽[U_i,U_j].\mathcal{H\!F\!L}(L_{\_}{\pm}^{\sigma^{\prime}})\cong\mathcal{H\!F\!L}(L_{\_}{\pm})\otimes_{\_}{\mathbb{F}}\mathbb{F}[U_{\_}i^{\prime},U_{\_}j^{\prime}].

By [26, Section 6] (or [34, Proposition 5.3]) we have

(L_±σ)(L_±σ)/U_iU_i,U_jU_j(L_±)\mathcal{H\!F\!L}(L_{\_}{\pm}^{{}^{\prime}\sigma^{\prime}})\cong\mathcal{H\!F\!L}(L_{\_}{\pm}^{\sigma^{\prime}})/\langle U_{\_}i-U_{\_}i^{\prime},U_{\_}j-U_{\_}j^{\prime}\rangle\cong\mathcal{H\!F\!L}(L_{\_}{\pm})

and under this isomorphism the induced map from (L_±)\mathcal{H\!F\!L}(L_{\_}{\pm}) to itself by the quasi-stabilization maps S_±S_{\_}{\pm} (see [32, Section 4.1]) is identity. Note that S_±S_{\_}{\pm} corresponds to the quasi-stabilization cobordism 𝒞_±\mathcal{C}_{\_}{\pm} from L_±L_{\_}{\pm} to L_±L^{\prime}_{\_}{\pm} obtained from the product cobordism by adding two dividing arcs on the ii-th and jj-th cylinders that split off disks containing w_i,±w_{\_}{i,\pm}^{\prime} and w_j,±w_{\_}{j,\pm}^{\prime}, as in Figure 4.

Refer to captionw_i,±w_{\_}{i,\pm}z_i,±z_{\_}{i,\pm}w_i,±w_{\_}{i,\pm}z_i,±z_{\_}{i,\pm}w_i,±w^{\prime}_{\_}{i,\pm}z_i,±z^{\prime}_{\_}{i,\pm}
Figure 4. The decoration on the component Σ_i\Sigma_{\_}i of the quasi-stabilization cobordism 𝒞_±\mathcal{C}_{\_}{\pm}

Next, we construct a decorated cobordism 𝒞=(W,)\mathcal{C}^{\prime}=(W,\mathcal{F}^{\prime}) from L_+L^{\prime}_{\_}+ to L_L^{\prime}_{\_}- by modifying the decoration on 𝒞\mathcal{C}, as follows. Add two parallel, vertical dividing arcs to Σ_i\Sigma_{\_}i (resp. Σ_j\Sigma_{\_}j) such that they divide Σ_i\Sigma_{\_}i (resp. Σ_j\Sigma_{\_}j) into four rectangles. Moreover, each one of them contains exactly one of the pairs w_i,±w_{\_}{i,\pm}, z_i,±z_{\_}{i,\pm}, w_i,±w^{\prime}_{\_}{i,\pm} and z_i,±z^{\prime}_{\_}{i,\pm} (resp.w_j,±w_{\_}{j,\pm}, z_j,±z_{\_}{j,\pm}, w_j,±w^{\prime}_{\_}{j,\pm} and z_j,±z^{\prime}_{\_}{j,\pm}) on its boundary. Clearly, σ\sigma^{\prime} extends to a coloring on \mathcal{F}^{\prime}. Define

𝒞~=𝒞𝒞_+=𝒞_𝒞.\widetilde{\mathcal{C}}=\mathcal{C}^{\prime}\circ\mathcal{C}_{\_}+=\mathcal{C}_{\_}-\circ\mathcal{C}.

Under the aforementioned isomorphism (L_σ)(L_)\mathcal{H\!F\!L}(L_{\_}-^{{}^{\prime}\sigma^{\prime}})\cong\mathcal{H\!F\!L}(L_{\_}-), the homomorphism induced by the cobordism map F_𝒞~,𝔰_kF_{\_}{\widetilde{\mathcal{C}},\mathfrak{s}_{\_}k} from (L_+)\mathcal{H\!F\!L}(L_{\_}+) to (L_)\mathcal{H\!F\!L}(L_{\_}-) is equal to ψ_k\psi_{\_}k.

On the other hand, 𝒞_+\mathcal{C}_{\_}+ can be decomposed as a cobordism \mathcal{B} containing two births from L_+L_{\_}+ to L_+O_2L_{\_}+\coprod O_{\_}2 followed by two band attachments 𝒞_b\mathcal{C}_{\_}{b} from L_+O_2L_{\_}+\coprod O_{\_}2 to L_+L^{\prime}_{\_}+ as in Figure 5.

Refer to caption\mathcal{B}𝒞_b\mathcal{C}_{\_}{b}
Figure 5. Decomposition of the cobordism 𝒞_+\mathcal{C}_{\_}+ as \mathcal{B} followed by 𝒞_b\mathcal{C}_{\_}b.
Refer to caption
Figure 6.
Refer to caption\mathcal{B}𝒞_O\mathcal{C}_{\_}{O}𝒞_b\mathcal{C}_{\_}{b}
Figure 7.

On the other hand, one may isotope the attaching circle of the 22-handle in 𝒞~\widetilde{\mathcal{C}} as in Figure 6. Then, changing the order of 22-handle attachment and band attachments as in Figure 7, we get a decomposition

𝒞~=𝒞_b𝒞_O\widetilde{\mathcal{C}}=\mathcal{C}_{\_}b\circ\mathcal{C}_{\_}O\circ\mathcal{B}

where 𝒞_O\mathcal{C}_{\_}O denotes the change of crossing cobordism map from L_+O_2L_{\_}+\coprod O_{\_}2 to L_+T(2,2)L_{\_}+\coprod-T(2,2) and 𝒞_b\mathcal{C}_{\_}b is the band attachment cobordism from L_+T(2,2)L_{\_}+\coprod-T(2,2) to L_L^{\prime}_{\_}-.

By [32, Theorem B] for any Spinc\text{Spin}^{c} structure 𝔰_k\mathfrak{s}_{\_}k we have

F_𝒞~,𝔰_k=F_𝒞_bF_𝒞_O,𝔰_kF_.F_{\_}{\widetilde{\mathcal{C}},\mathfrak{s}_{\_}k}=F_{\_}{\mathcal{C}_{\_}b}\circ F_{\_}{\mathcal{C}_{\_}O,\mathfrak{s}_{\_}k}\circ F_{\_}{\mathcal{B}}.

Moreover, (LL)=(L)_𝔽(L)\mathcal{H\!F\!L}(L\coprod L^{\prime})=\mathcal{H\!F\!L}(L)\otimes_{\_}{\mathbb{F}}\mathcal{H\!F\!L}(L^{\prime}) for any multipointed colored links LL and LL^{\prime}, and under corresponding identifications

F_𝒞_O,s_k=Idψ_kOF_{\_}{\mathcal{C}_{\_}O,s_{\_}k}=\mathrm{Id}\otimes\psi_{\_}k^{O}

where ψ_kO\psi_{\_}k^{O} denotes the map ψ_k\psi_{\_}k for the unlink O_2O_{\_}2. So, the claim holds, because equalities hold for the change of crossing maps for the unlink O_2O_{\_}2 from Example 3.3.

As in the bottom of Figure 3, (1)(-1)-surgery on the specified red unknot can change a negative crossing to a positive crossing. Hence, we can also consider the cobordism from (S3,L_)(S^{3},L_{\_}{-}) to (S3,L_+)(S^{3},L_{\_}+) induced by attaching a 22-handle along this unknot. The embedded surface is a disjoint union of nn annuli, and each one of them is equipped with two parallel and vertical dividing arcs.

Proposition 3.5.

Let (W,):(S3,L_)(S3,L_+)(W,\mathcal{F}):(S^{3},L_{\_}{-})\rightarrow(S^{3},L_{\_}{+}) be the decorated link cobordism induced by attaching a 2-handle to the (1)(-1)-framed unknot in Figure 3 which changes a negative crossing to a positive crossing. Suppose that the link components L_i,,L_j,L_{\_}{i,-},L_{\_}{j,-} are passing through the (1)(-1)-framed unknot with i<ji<j. Let 𝔰_k\mathfrak{s}_{\_}k be the Spinc\text{Spin}^{c} structure on WW satisfying that

c_1(𝔰_k),[S2]=2k+1\langle c_{\_}{1}(\mathfrak{s}_{\_}k),[S^{2}]\rangle=2k+1

where [S2][S^{2}] is the generator of H_2(W)H_{\_}{2}(W) corresponding to the attached 2-handle. Let ϕ_k=F_W,,𝔰_k\phi_{\_}k=F_{\_}{W,\mathcal{F},\mathfrak{s}_{\_}k} be the corresponding map in link Floer homology. Then

gr_𝐰(ϕ_k)=k2k,gr_𝐳(ϕ_k)=k2+3k2\textup{gr}_{\_}{{\bf{w}}}(\phi_{\_}k)=-k^{2}-k,\quad\textup{gr}_{\_}{{\bf{z}}}(\phi_{\_}k)=-k^{2}+3k-2

and

A_i(ϕ_k)=A_j(ϕ_k)=k+1/2.A_{\_}{i}(\phi_{\_}k)=A_{\_}{j}(\phi_{\_}k)=-k+1/2.

Moreover, A_l(ϕ_k)=0A_{\_}{l}(\phi_{\_}k)=0 for all li,jl\neq i,j. In particular, A(ϕ_0)=12(𝐞_i+𝐞_j)A(\phi_{\_}0)=\dfrac{1}{2}(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j), A(ϕ_1)=12(𝐞_i+𝐞_j)A(\phi_{\_}1)=-\dfrac{1}{2}(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j) and gr_𝐰(ϕ_0)=0\textup{gr}_{\_}{\bf{w}}(\phi_{\_}0)=0, gr_𝐰(ϕ_1)=2\textup{gr}_{\_}{{\bf{w}}}(\phi_{\_}1)=-2.

Proof.

The proof is very similar to the one of Proposition 3.1. By the same computation, we get gr_𝐰(ϕ_k)=k2k\textup{gr}_{\_}{{\bf{w}}}(\phi_{\_}k)=-k^{2}-k. For the Alexander gradings, note that

A_i(ϕ_k)=c_1(𝔰_k),[Σ^_i][Σ^][Σ^_i]2=2k1+22=k+1/2.A_{\_}i(\phi_{\_}k)=\dfrac{\langle c_{\_}{1}(\mathfrak{s}_{\_}{k}),[\widehat{\Sigma}_{\_}i]\rangle-[\widehat{\Sigma}]\cdot[\widehat{\Sigma}_{\_}{i}]}{2}=\dfrac{-2k-1+2}{2}=-k+1/2.

The same computation works for A_j(ϕ_k)=k+1/2A_{\_}{j}(\phi_{\_}k)=-k+1/2. However, [Σ^_l]=0[\widehat{\Sigma}_{\_}l]=0 for all li,jl\neq i,j. Hence, A_l(ϕ_k)=0A_{\_}l(\phi_{\_}k)=0 for all such ll. Note that gr_𝐰(ϕ_k)gr_𝐳(ϕ_k)=2(A_1(ϕ_k)++A_n(ϕ_k))=2(A_i(ϕ_k)+A_j(ϕ_k))=2(2k+1)\textup{gr}_{\_}{{\bf{w}}}(\phi_{\_}k)-\textup{gr}_{\_}{{\bf{z}}}(\phi_{\_}k)=2(A_{\_}1(\phi_{\_}k)+\cdots+A_{\_}n(\phi_{\_}k))=2(A_{\_}i(\phi_{\_}k)+A_{\_}j(\phi_{\_}k))=2(-2k+1), so gr_𝐳(ϕ_k)=k2+3k2\textup{gr}_{\_}{{\bf{z}}}(\phi_{\_}k)=-k^{2}+3k-2.

Example 3.6.

By Corollary 2.10 cobordism maps ϕ_k:(O_2)(T(2,2))\phi_{\_}k:\mathcal{H\!F\!L}(O_{\_}2)\to\mathcal{H\!F\!L}(T(2,2)) are non-zero and determined by the grading shift formulas from Proposition 3.5. We compute

ϕ_0(1)=aandϕ_1(1)=b.\phi_{\_}0(1)=a\quad\quad\text{and}\quad\quad\phi_{\_}1(1)=b.

In general, we have

ϕ_k(1)={(U_1U_2)k1𝐔(k1)(k2)2bifk1(V_1V_2)k𝐔k(k+1)2aifk0.\phi_{\_}k(1)=\begin{cases}(U_{\_}1U_{\_}2)^{k-1}\mathbf{U}^{\frac{(k-1)(k-2)}{2}}b&\text{if}\ k\geq 1\\ (V_{\_}1V_{\_}2)^{-k}\mathbf{U}^{\frac{k(k+1)}{2}}a&\text{if}\ k\leq 0.\end{cases}
Example 3.7.

Let L_=T(2,2)L_{\_}{-}=-T(2,2) and L_+=O_2L_{\_}{+}=O_{\_}2. Then by Corollary 2.10 the cobordism maps ϕ_k:(T(2,2))(O_2)\phi_{\_}{k}:\mathcal{H\!F\!L}(-T(2,2))\rightarrow\mathcal{H\!F\!L}(O_{\_}2) are nonzero and determined by the grading shifts. In particular,

ϕ_0(c)=V_1,ϕ_1(c)=U_2,ϕ_0(d)=V_2,ϕ_1(d)=U_1.\phi_{\_}{0}(c^{\prime})=V_{\_}1,\quad\phi_{\_}{1}(c^{\prime})=U_{\_}2,\quad\phi_{\_}{0}(d^{\prime})=V_{\_}2,\quad\phi_{\_}{1}(d^{\prime})=U_{\_}1.

In general, we have

ϕ_k(c)={U_1k1U_2k𝐔(k1)(k2)2ifk1V_1k+1V_2k𝐔k(k+1)2ifk0,ϕ_k(d)={U_1kU_2k1𝐔(k1)(k2)2ifk1V_1kV_2k+1𝐔k(k+1)2ifk0.\phi_{\_}k(c^{\prime})=\begin{cases}U_{\_}1^{k-1}U_{\_}2^{k}\mathbf{U}^{\frac{(k-1)(k-2)}{2}}&\text{if}\ k\geq 1\\ V_{\_}1^{-k+1}V_{\_}2^{-k}\mathbf{U}^{\frac{k(k+1)}{2}}&\text{if}\ k\leq 0,\end{cases}\phi_{\_}k(d^{\prime})=\begin{cases}U_{\_}1^{k}U_{\_}2^{k-1}\mathbf{U}^{\frac{(k-1)(k-2)}{2}}&\text{if}\ k\geq 1\\ V_{\_}1^{-k}V_{\_}2^{-k+1}\mathbf{U}^{\frac{k(k+1)}{2}}&\text{if}\ k\leq 0.\end{cases}
Proposition 3.8.

The maps ϕ_k\phi_{\_}k satisfy the following properties:

  • (a)

    For k1k\geq 1, we have ϕ_k=(U_iU_j)k1𝐔(k1)(k2)2ϕ_1\phi_{\_}k=(U_{\_}iU_{\_}j)^{k-1}\mathbf{U}^{\frac{(k-1)(k-2)}{2}}\phi_{\_}1.

  • (b)

    For k0k\leq 0, we have ϕ_k=(V_iV_j)k𝐔k(k+1)2ϕ_0.\phi_{\_}k=(V_{\_}iV_{\_}j)^{-k}\mathbf{U}^{\frac{k(k+1)}{2}}\phi_{\_}{0}.

  • (c)

    We have U_iϕ_0=V_jϕ_1U_{\_}i\phi_{\_}0=V_{\_}j\phi_{\_}1 and U_jϕ_0=V_iϕ_1U_{\_}j\phi_{\_}0=V_{\_}i\phi_{\_}1.

Proof.

The proof is very similar to that of Proposition 3.4. By Example 3.6, the claim holds for L_=O_2L_{\_}-=O_{\_}2 and L_+=T(2,2)L_{\_}+=T(2,2). It remains to show that the general case follows from this special case. Let 𝒞=(W,)\mathcal{C}=(W,\mathcal{F}) denote the decorated crossing change cobordism from L_L_{\_}{-} to L_+L_{\_}+. Following the notation in the proof of Proposition 3.4, let L_±L^{\prime}_{\_}{\pm} denote the links obtained from L_±L_{\_}{\pm} by adding two extra pairs of base points on L_i,±L_{\_}{i,\pm} and L_j,±L_{\_}{j,\pm}, and 𝒞\mathcal{C}^{\prime} be the decorated cobordism from L_L^{\prime}_{\_}{-} to L_+L^{\prime}_{\_}{+} obtained from 𝒞\mathcal{C} by adding two pairs of parallel and vertical dividing arcs. Further, 𝒞_±\mathcal{C}_{\_}{\pm} denotes the decorated quasi-stabilization cobordisms from L_±L_{\_}{\pm} to L_±L^{\prime}_{\_}{\pm}. As in the proof of Proposition 3.4, for the decorated cobordism 𝒞~=𝒞𝒞_\widetilde{\mathcal{C}}=\mathcal{C}^{\prime}\circ\mathcal{C}_{\_}{-}, composing F_𝒞~,𝔰_kF_{\_}{\widetilde{\mathcal{C}},\mathfrak{s}_{\_}{k}} with a specific isomorphism (L_+σ)(L_+)\mathcal{H\!F\!L}(L_{\_}+^{{}^{\prime}\sigma^{\prime}})\cong\mathcal{H\!F\!L}(L_{\_}+) is equal to identity.

We now still decompose 𝒞_\mathcal{C}_{\_}{-} as a cobordism \mathcal{B} from L_L_{\_}{-} to L_O_2L_{\_}-\coprod O_{\_}2 followed by two band attachments 𝒞_b\mathcal{C}_{\_}b from L_O_2L_{\_}-\coprod O_{\_}2 to L_L^{\prime}_{\_}-, as in Figure 8.

Refer to caption\mathcal{B}𝒞_b\mathcal{C}_{\_}{b}
Figure 8. Decomposition of the cobordism 𝒞_\mathcal{C}_{\_}- as \mathcal{B} followed by 𝒞_b\mathcal{C}_{\_}b.

Changing the order of the 22-handle attachment in 𝒞\mathcal{C}^{\prime} and the band attachments in 𝒞_b\mathcal{C}_{\_}b, we get

𝒞~=𝒞_b𝒞_O\widetilde{\mathcal{C}}=\mathcal{C}_{\_}b\circ\mathcal{C}_{\_}O\circ\mathcal{B}

where 𝒞_O\mathcal{C}_{\_}O denotes the cobordism from L_O_2L_{\_}{-}\coprod O_{\_}2 to L_T(2,2)L_{\_}{-}\coprod T(2,2) and 𝒞_b\mathcal{C}_{\_}b is the band attachment cobordism from L_T(2,2)L_{\_}{-}\coprod T(2,2) to L_+L^{\prime}_{\_}{+}, see Figure 9. Hence, for any Spinc structure 𝔰_k\mathfrak{s}_{\_}{k} we have

F_𝒞~,𝔰_k=F_𝒞_bF_𝒞_O,𝔰_kF_.F_{\_}{\widetilde{\mathcal{C}},\mathfrak{s}_{\_}k}=F_{\_}{\mathcal{C}_{\_}{b}}\circ F_{\_}{\mathcal{C}_{\_}{O},\mathfrak{s}_{\_}{k}}\circ F_{\_}{\mathcal{B}}.

Here F_𝒞_O,𝔰_k=Idϕ_kOF_{\_}{\mathcal{C}_{\_}{O},\mathfrak{s}_{\_}{k}}=\mathrm{Id}\otimes\phi_{\_}{k}^{O} where ϕ_kO\phi_{\_}{k}^{O} denotes the map ϕ_k\phi_{\_}k for the unlink O_2O_{\_}2. Hence the claim holds for general links.

Refer to caption\mathcal{B}𝒞_O\mathcal{C}_{\_}{O}𝒞_b\mathcal{C}_{\_}{b}
Figure 9.

Proposition 3.9.

The maps ψ_k\psi_{\_}k and ϕ_k\phi_{\_}k in Proposition 3.1 and Proposition 3.5 compose as follows:

ϕ_0ψ_0=V_i,ϕ_0ψ_1=V_j,ϕ_1ψ_0=U_j,ϕ_1ψ_1=U_i\phi_{\_}0\psi_{\_}0=V_{\_}i,\ \phi_{\_}0\psi_{\_}{-1}=V_{\_}j,\ \phi_{\_}1\psi_{\_}0=U_{\_}j,\ \phi_{\_}1\psi_{\_}{-1}=U_{\_}i
ψ_0ϕ_0=V_i,ψ_1ϕ_0=V_j,ψ_0ϕ_1=U_j,ψ_1ϕ_1=U_i,\psi_{\_}0\phi_{\_}0=V_{\_}i,\ \psi_{\_}{-1}\phi_{\_}0=V_{\_}j,\ \psi_{\_}0\phi_{\_}1=U_{\_}j,\ \psi_{\_}{-1}\phi_{\_}1=U_{\_}i,

The rest of compositions are determined by these.

Proof.

We prove the equalities in the first row, and the proof for the second row is similar. It is straightforward from Examples 3.3 and 3.7 that the claim holds for L_+=O_2L_{\_}+=O_{\_}2 and L_=T(2,2)L_{\_}-=-T(2,2) the negative Hopf link, because

ϕ_0ψ_0(1)=V_1,ϕ_0ψ_1(1)=V_2,ϕ_1ψ_0(1)=U_2,ϕ_1ψ_1(1)=U_1.\phi_{\_}0\psi_{\_}0(1)=V_{\_}1,\ \phi_{\_}0\psi_{\_}{-1}(1)=V_{\_}2,\ \phi_{\_}1\psi_{\_}0(1)=U_{\_}2,\ \phi_{\_}1\psi_{\_}1(1)=U_{\_}1.

The strategy is similar to the proof of Propositions 3.4 and 3.8 and we fix the same notation. To distinguish the cobordisms defining ψ_k\psi_{\_}k and ϕ_k\phi_{\_}k we use subscripts 11 and 22, i.e. let 𝒞_1=(W_1,_1)\mathcal{C}_{\_}1=(W_{\_}1,\mathcal{F}_{\_}1) be the decorated crossing change cobordism from L_+L_{\_}+ to L_L_{\_}- and 𝒞_2=(W_2,_2)\mathcal{C}_{\_}2=(W_{\_}2,\mathcal{F}_{\_}2) be the decorated crossing change cobordism from L_L_{\_}- to L_+L_{\_}+. As before, we denote the links obtained from L_±L_{\_}{\pm} by adding two extra pairs of base points on L_i,±L_{\_}{i,\pm} and L_j,±L_{\_}{j,\pm} by L_±L^{\prime}_{\_}{\pm}. Further, we denote the corresponding cobordism from L_+L_{\_}+^{\prime} to L_L_{\_}-^{\prime} (resp. L_L_{\_}-^{\prime} to L_+L_{\_}+^{\prime}) obtained from 𝒞_1\mathcal{C}_{\_}{1} (resp. 𝒞_2\mathcal{C}_{\_}2) by adding two pairs of parallel and vertical dividing arcs by 𝒞_1\mathcal{C}_{\_}1^{\prime} (resp. 𝒞_2\mathcal{C}_{\_}2^{\prime}). Moreover, we consider quasi-stabilization cobordisms 𝒞_±\mathcal{C}_{\_}{\pm} from L_±L_{\_}{\pm} to L_±L^{\prime}_{\_}{\pm}.

Let 𝒞~=𝒞_2𝒞_1𝒞_+=𝒞_+𝒞_2𝒞_1\widetilde{\mathcal{C}}=\mathcal{C}^{\prime}_{\_}2\circ\mathcal{C}^{\prime}_{\_}1\circ\mathcal{C}_{\_}+=\mathcal{C_{\_}+}\circ\mathcal{C}_{\_}2\circ\mathcal{C}_{\_}1. For any k_1,k_2k_{\_}1,k_{\_}2\in\mathbb{Z}, denote the Spinc\text{Spin}^{c} structure on 𝒞~\widetilde{\mathcal{C}} whose restriction to 𝒞_1\mathcal{C}_{\_}1 and 𝒞_2\mathcal{C}_{\_}2 is equal to 𝔰_k_1\mathfrak{s}_{\_}{k_{\_}1} and 𝔰_k_2\mathfrak{s}_{\_}{k_{\_}2}, respectively, by 𝔰_k_1,k_2\mathfrak{s}_{\_}{k_{\_}1,k_{\_}2}. Thus, under the aforementioned isomorphism (L_+σ)(L_+)\mathcal{H\!F\!L}(L_{\_}{+}^{{}^{\prime}\sigma^{\prime}})\cong\mathcal{H\!F\!L}(L_{\_}+) the cobordism map F_C~,𝔰_k_1,k_2F_{\_}{\widetilde{C},\mathfrak{s}_{\_}{k_{\_}1,k_{\_}2}} is equal to ϕ_k_2ψ_k_1\phi_{\_}{k_{\_}2}\circ\psi_{\_}{k_{\_}1}.

On the other hand, as depicted in Figure 5, the cobordism 𝒞_+\mathcal{C}_{\_}+ can be decomposed as 𝒞_+=𝒞_b\mathcal{C}_{\_}+=\mathcal{C}_{\_}b\circ\mathcal{B}, where \mathcal{B} is the decorated cobordism from L_+L_{\_}+ to L_+O_2L_{\_}+\coprod O_{\_}2 corresponding to two births, and 𝒞_b\mathcal{C}_{\_}b is defined by attaching two bands. By Figure 10, after an isotopy on the attaching circles of the 22-handles in 𝒞_1\mathcal{C}_{\_}1 and 𝒞_2\mathcal{C}_{\_}2, we may change their order with the band attachments in 𝒞_b\mathcal{C}_{\_}b to get another decomposition

𝒞~=𝒞_b𝒞_2O𝒞_1O\widetilde{\mathcal{C}}=\mathcal{C}_{\_}b\circ\mathcal{C}_{\_}2^{O}\circ\mathcal{C}_{\_}1^{O}\circ\mathcal{B}

Here, 𝒞_1O\mathcal{C}_{\_}1^{O} denotes the decorated cobordism from L_+O_2L_{\_}+\coprod O_{\_}2 to L_+T(2,2)L_{\_}+\coprod-T(2,2) corresponding to changing a positive crossing to a negative crossing in O_2O_{\_}2. Similarly, 𝒞_2O\mathcal{C}_{\_}2^{O} is the cobordism from L_+T(2,2)L_{\_}+\coprod-T(2,2) to L_+O_2L_{\_}+\coprod O_{\_}2 corresponding to changing a negative crossing to a positive crossing in T(2,2)-T(2,2). Thus,

F_𝒞~,𝔰_k_1,k_2=F_𝒞_bF_𝒞_2O,𝔰_k_2F_𝒞_1O,𝔰_k_1F_,F_{\_}{\widetilde{\mathcal{C}},\mathfrak{s}_{\_}{k_{\_}1,k_{\_}2}}=F_{\_}{\mathcal{C}_{\_}b}\circ F_{\_}{\mathcal{C}_{\_}2^{O},\mathfrak{s}_{\_}{k_{\_}2}}\circ F_{\_}{\mathcal{C}_{\_}1^{O},\mathfrak{s}_{\_}{k_{\_}1}}\circ F_{\_}{\mathcal{B}},

and the claim follows from the special case of L_+=O_2L_{\_}+=O_{\_}2 and L_=T(2,2)L_{\_}-=-T(2,2).

Refer to caption
Figure 10. Cobordisms 𝒞_1\mathcal{C}_{\_}1 and 𝒞_2\mathcal{C}_{\_}2 are define by attaching 22-handles along the red and the blue unknots, respectively.

3.2. Full twists

In this section, we will apply similar computation as in Proposition 3.5 to get the properties of the cobordism map induced by attaching a 22-handle along a (1)(-1)-framed unknot through nn-strand braid to get a positive full twist.

Refer to captionnn1-1+n+n
Figure 11. (1)(-1)-surgery on the red unknot will add a positive full twist.

Using the similar computation as in Proposition 3.5, we have the following:

Proposition 3.10.

Let (W,):(S3,L)(S3,L¯)(W,\mathcal{F}):(S^{3},L)\rightarrow(S^{3},\bar{L}) be the decorated link cobordism obtained by attaching a 22-handle on the (1)(-1)-framed unknot which adds a full twist to the nn parallel strands as in Figure 11. Let 𝔰_k\mathfrak{s}_{\_}k be the Spinc\text{Spin}^{c} structure on WW satisfying that

c_1(𝔰_k),[S2]=2k+1\langle c_{\_}{1}(\mathfrak{s}_{\_}k),[S^{2}]\rangle=2k+1

where [S2][S^{2}] is the generator of H_2(W)H_{\_}{2}(W) corresponding to the attached 2-handle. Let ϕn_k=F_W,,𝔰_k\phi^{n}_{\_}k=F_{\_}{W,\mathcal{F},\mathfrak{s}_{\_}k} be the corresponding map in link Floer homology. Then

gr_𝐰(ϕn_k)=k2k\textup{gr}_{\_}{{\bf{w}}}(\phi^{n}_{\_}k)=-k^{2}-k

and

A_i(ϕn_k)=k+(n1)/2,A_{\_}{i}(\phi^{n}_{\_}k)=-k+(n-1)/2,

for i=1,2,,ni=1,2,\cdots,n.

Proof.

The computation of gr_𝐰\textup{gr}_{\_}{{\bf{w}}} is exactly the same as the one of Proposition 3.5. Hence, gr_𝐰(ϕn_k)=k2k\textup{gr}_{\_}{{\bf{w}}}(\phi^{n}_{\_}k)=-k^{2}-k. For the computation of the Alexander grading, it is also similar to the one of Proposition 3.5, except now for each i=1,2,,ni=1,2,\cdots,n, we have

A_i(ϕn_k)=c_1(𝔰_k),[Σ^_i][Σ^][Σ^_i]2=2k1+n2=k+(n1)/2.A_{\_}i(\phi^{n}_{\_}k)=\dfrac{\langle c_{\_}{1}(\mathfrak{s}_{\_}{k}),[\widehat{\Sigma}_{\_}i]\rangle-[\widehat{\Sigma}]\cdot[\widehat{\Sigma}_{\_}{i}]}{2}=\dfrac{-2k-1+n}{2}=-k+(n-1)/2.

Now we consider the following example where L=O_nL=O_{\_}n and L¯=T(n,n)\bar{L}=T(n,n). It is known [13] that T(n,n)T(n,n) is an LL-space link. We first recall the link Floer homology (T(n,n))\mathcal{H\!F\!L}(T(n,n)). For the explicit computation, see [7].

Theorem 3.11 ([7]).

The Heegaard Floer homology (T(n,n))\mathcal{HFL}(T(n,n)) has nn generators, which we denote by a_0,,a_n1a_{\_}0,\dots,a_{\_}{n-1} subject to the following relations:

(4) (_iI_kU_i)a_k1=(_j{1,,n}I_kV_j)a_k,U_iV_ia_k=U_jV_ja_k\left(\prod_{\_}{i\in I_{\_}k}U_{\_}i\right)a_{\_}{k-1}=\left(\prod_{\_}{j\in\{1,\cdots,n\}\setminus I_{\_}k}V_{\_}j\right)a_{\_}{k},\quad U_{\_}iV_{\_}ia_{\_}k=U_{\_}jV_{\_}ja_{\_}k

Here, I_kI_{\_}k is any subset of the set {1,,n}\{1,\dots,n\} of length kk (so the first equation has (nk)\binom{n}{k} relations for each kk), and in the second equation i,ji,j, and kk range from 11 to nn.

Now we list the explicit gradings of the generators a_ka_{\_}k where 0kn10\leq k\leq n-1. The Alexander multi-grading of a_ka_{\_}k is

(n12k,n12k,,n12k).\left(\dfrac{n-1}{2}-k,\dfrac{n-1}{2}-k,\cdots,\dfrac{n-1}{2}-k\right).

The generator a_ka_{\_}k has homological grading

(gr_𝐰(a_k),gr_𝐳(a_k))=(k(k+1),k(k+1)n(n1)+2kn).(\textup{gr}_{\_}{{\bf{w}}}(a_{\_}k),\textup{gr}_{\_}{{\bf{z}}}(a_{\_}k))=\left(-k(k+1),-k(k+1)-n(n-1)+2kn\right).

The Maslov grading gr_𝐰\textup{gr}_{\_}{{\bf{w}}} is obtained from the HH-function of the torus link T(n,n)T(n,n), which is computed in [13]. The computation of gr_𝐳\textup{gr}_{\_}{{\bf{z}}} follows from the relation

gr_𝐰gr_𝐳2=A_1+A_2++A_n.\dfrac{\textup{gr}_{\_}{{\bf{w}}}-\textup{gr}_{\_}{{\bf{z}}}}{2}=A_{\_}1+A_{\_}2+\cdots+A_{\_}n.
Example 3.12.

By Corollary 2.10 the cobordism maps ϕn_k:(O_n)(T(n,n))\phi^{n}_{\_}k:\mathcal{H\!F\!L}(O_{\_}n)\to\mathcal{H\!F\!L}(T(n,n)) are non-zero and determined by the grading shift formulas from Proposition 3.10. Recall that A_i(ϕn_k)=(k+n12,,k+n12)A_{\_}i(\phi^{n}_{\_}k)=(-k+\frac{n-1}{2},\cdots,-k+\frac{n-1}{2}) and gr_𝐰(ϕn_k)=k2k\textup{gr}_{\_}{\bf{w}}(\phi^{n}_{\_}k)=-k^{2}-k. Then

ϕn_k(1)=a_k\phi^{n}_{\_}k(1)=a_{\_}{k}

for k=0,1,,n1k=0,1,\cdots,n-1.

In general, we have

ϕn_k(1)={(U_1U_n)k(n1)𝐔(k(n1))(kn)2a_n1ifkn1(V_1V_n)k𝐔k(k+1)2a_0ifk0.\phi^{n}_{\_}k(1)=\begin{cases}(U_{\_}1\cdots U_{\_}n)^{k-(n-1)}\mathbf{U}^{\frac{(k-(n-1))(k-n)}{2}}a_{\_}{n-1}&\text{if}\ k\geq n-1\\ (V_{\_}1\cdots V_{\_}n)^{-k}\mathbf{U}^{\frac{k(k+1)}{2}}a_{\_}0&\text{if}\ k\leq 0.\end{cases}

Similar to Proposition 3.8, the maps ϕ_kn\phi_{\_}k^{n} satisfy the following properties:

Proposition 3.13.

The maps ϕn_k:(L)(L¯)\phi^{n}_{\_}k:\mathcal{H\!F\!L}(L)\rightarrow\mathcal{H\!F\!L}(\bar{L}) satisfy the following properties:

a) For kn1k\geq n-1, we have ϕn_k=(U_1U_n)k(n1)𝐔(k(n1))(kn)2ϕn_n1\phi^{n}_{\_}k=(U_{\_}1\cdots U_{\_}n)^{k-(n-1)}\mathbf{U}^{\frac{(k-(n-1))(k-n)}{2}}\phi^{n}_{\_}{n-1}.

b) For k0k\leq 0, we have ϕn_k=(V_1V_n)k𝐔k(k+1)2ϕn_0.\phi^{n}_{\_}k=(V_{\_}1\cdots V_{\_}n)^{-k}\mathbf{U}^{\frac{k(k+1)}{2}}\phi^{n}_{\_}{0}.

Proof.

The proof is very similar to the one of Proposition 3.8. As before, we denote LL^{\prime} (resp. L¯\bar{L}^{\prime}) as the link obtained from LL (resp. L¯\bar{L}) by adding an extra pair of basepoints w_i,z_iw^{\prime}_{\_}i,z^{\prime}_{\_}i for each component L_iL_{\_}i. Let 𝒞\mathcal{C}^{\prime} be the induced decorated cobordism from LL^{\prime} to L¯\bar{L}^{\prime} induced from the decorated cobordism 𝒞=(W,)\mathcal{C}=(W,\mathcal{F}) from LL to L¯\bar{L}, and σ\sigma^{\prime} be the induced coloring on LL^{\prime} and L¯\bar{L}^{\prime} as in the proof of Proposition 3.4. We still get the isomorphism

(Lσ)((L)𝔽[U_1,,U_n,V_1,,V_n,U_1,,U_n])/U_1U_1,,U_nU_n(L).\mathcal{H\!F\!L}(L^{\prime\sigma^{\prime}})\cong\left(\mathcal{H\!F\!L}(L)\otimes\mathbb{F}[U_{\_}1,\cdots,U_{\_}n,V_{\_}1,\cdots,V_{\_}n,U^{\prime}_{\_}1,\cdots,U^{\prime}_{\_}n]\right)/\langle U_{\_}1-U^{\prime}_{\_}1,\cdots,U_{\_}n-U^{\prime}_{\_}n\rangle\cong\mathcal{H\!F\!L}(L).

Similarly, we also have (L¯σ)(L¯)\mathcal{H\!F\!L}(\bar{L}^{\prime\sigma^{\prime}})\cong\mathcal{H\!F\!L}(\bar{L}).

As before, we let 𝒞_+\mathcal{C}_{\_}+ be the decorated cobordism from LL to LL^{\prime}, which can be decomposed as a cobordism \mathcal{B} from LL to LO_nL\coprod O_{\_}n followed by nn band attachments 𝒞_b\mathcal{C}_{\_}b from LO_nL\coprod O_{\_}n to LL^{\prime}. Hence, F_𝒞~,𝔰_k=ϕn_kF_{\_}{\widetilde{\mathcal{C}},\mathfrak{s}_{\_}k}=\phi^{n}_{\_}k where 𝒞~=𝒞𝒞_+\widetilde{\mathcal{C}}=\mathcal{C}^{\prime}\circ\mathcal{C}_{\_}+. Now we use the same trick as before to isotope attaching circle of the 2-handle as in Figure 6 so that it encircles the unlink O_nO_{\_}n and change the order of the 2-handle attachment and band attachments. Then

𝒞~=𝒞_b𝒞_O_n\widetilde{\mathcal{C}}=\mathcal{C}_{\_}{b}\circ\mathcal{C}_{\_}{O_{\_}{n}}\circ\mathcal{B}

where 𝒞_O_n\mathcal{C}_{\_}{O_{\_}{n}} denotes the cobordism obtained by attaching a 22-handle along (1)(-1)-framed unknot from LO_nL\coprod O_{\_}n to LT(n,n)L\coprod T(n,n) and 𝒞_b\mathcal{C}_{\_}{b} denotes the band attachment cobordism from LT(n,n)L\coprod T(n,n) to L¯\bar{L}. Hence,

F_𝒞_O_n,𝔰_k=Idϕ_kF_{\_}{\mathcal{C}_{\_}{O_{\_}{n}},\mathfrak{s}_{\_}k}=\mathrm{Id}\otimes\phi_{\_}{k}

where ϕ_k\phi_{\_}k denotes the map ϕ_kn\phi_{\_}{k}^{n} for the unlink O_nO_{\_}n. Since the proposition holds for unlink O_nO_{\_}n by Example 3.12, the general case follows.

4. Skein exact sequence

4.1. Surgery exact triangle

Suppose LL is a link in an integer homology sphere YY, and KYLK\subset Y\setminus L is a knot. Let (W_1,_1)(W_{\_}1,\mathcal{F}_{\_}1) be the decorated link cobordism from (Y,L)(Y,L) to (Y_1(K),L)(Y_{\_}{-1}(K),L) obtained by attaching a two-handle along KK with framing 1-1, and decorated as in Convention 3. Similarly, (W_2,_2)(W_{\_}2,\mathcal{F}_{\_}2) and (W_3,_3)(W_{\_}3,\mathcal{F}_{\_}3) denote the cobordisms from (Y_1(K),L)(Y_{\_}{-1}(K),L) to (Y_0(K),L)(Y_{\_}{0}(K),L) and (Y_0(K),L)(Y_{\_}{0}(K),L) to (Y,L)(Y,L), respectively.

Proposition 4.1.

The link cobordism maps F_i=_𝔰Spinc(W_i)F_W_i,_i,𝔰F_{\_}{i}=\sum_{\_}{\mathfrak{s}\in\text{Spin}^{c}(W_{\_}i)}F_{\_}{W_{\_}i,\mathcal{F}_{\_}i,\mathfrak{s}} form an exact triangle as follows.

𝓗𝓕𝓛(Y,L){\bm{{\mathcal{H\!F\!L}}}(Y,L)}𝓗𝓕𝓛(Y_1(K),L){\bm{{\mathcal{H\!F\!L}}}(Y_{\_}{-1}(K),L)}𝓗𝓕𝓛(Y_0(K),L){\bm{{\mathcal{H\!F\!L}}}(Y_{\_}{0}(K),L)}F_1\scriptstyle{F_{\_}{1}}F_2\scriptstyle{F_{\_}{2}}F_3\scriptstyle{F_{\_}{3}}
Proof.

This is a straightforward generalization of the surgery exact triangle for 𝐻𝐹+\mathit{HF}^{+} in [24, Section 9]. We will outline the proof and highlight the differences here. Consider a multi-pointed Heegaard diagram

=(Σ,𝜶,𝜷={β_1,,β_k},𝜸={γ_1,,γ_k},𝜹={δ_1,,δ_k},𝐳,𝐰)\mathcal{H}=\left(\Sigma,\bm{\alpha},\bm{\beta}=\{\beta_{\_}1,\cdots,\beta_{\_}k\},\bm{\gamma}=\{\gamma_{\_}1,\cdots,\gamma_{\_}k\},\bm{\delta}=\{\delta_{\_}1,\cdots,\delta_{\_}k\},{\bf{z}},{\bf{w}}\right)

where k=g+n1k=g+n-1, such that

  • _αβ=(Σ,𝜶,𝜷,𝐳,𝐰)\mathcal{H}_{\_}{\alpha\beta}=(\Sigma,\bm{\alpha},\bm{\beta},{\bf{z}},{\bf{w}}), _αγ=(Σ,𝜶,𝜸,𝐳,𝐰)\mathcal{H}_{\_}{\alpha\gamma}=(\Sigma,\bm{\alpha},\bm{\gamma},{\bf{z}},{\bf{w}}) and _αδ=(Σ,𝜶,𝜹,𝐳,𝐰)\mathcal{H}_{\_}{\alpha\delta}=(\Sigma,\bm{\alpha},\bm{\delta},{\bf{z}},{\bf{w}}) are Heegaard diagrams for the link LL in 33-manifolds YY, Y_1(K)Y_{\_}{-1}(K) and Y_0(K)Y_{\_}0(K), respectively. So, 𝐳{\bf{z}} and 𝐰{\bf{w}} consist of nn basepoints, where nn is the number of connected components of LL.

  • For any 1ik11\leq i\leq k-1, γ_i\gamma_{\_}i and δ_i\delta_{\_}i are small isotopic translations of β_i\beta_{\_}i such that they intersect β_i\beta_{\_}i transversely in two points and are disjoint from β_j\beta_{\_}j for jij\neq i . Moreover, δ_i\delta_{\_}i intersects γ_i\gamma_{\_}i in two transverse points as well.

  • Pairwise intersections of β_k\beta_{\_}k, γ_k\gamma_{\_}k and δ_k\delta_{\_}k are single points with signs #(β_kγ_k)=#(γ_kδ_k)=#(δ_kβ_k)=1\#(\beta_{\_}k\cap\gamma_{\_}k)=\#(\gamma_{\_}k\cap\delta_{\_}k)=\#(\delta_{\_}k\cap\beta_{\_}k)=-1. Moreover, γ_k\gamma_{\_}k is obtained from the juxtaposition of β_k\beta_{\_}k and δ_k\delta_{\_}k.

  • Strongly admissible in the sense of [31, Definition 4.15] which is a multipointed version of [25, Section 8.4.2].

Let F_αβγF_{\_}{\alpha\beta\gamma} be the chain map defined by counting holomorphic triangles as:

F_αβγ:𝓒𝓕𝓛(Σ,𝜶,𝜷,𝐳,𝐰)𝓒𝓕𝓛(Σ,𝜷,𝜸,𝐳,𝐰)𝓒𝓕𝓛(Σ,𝜶,𝜸,𝐳,𝐰)F_αβγ(𝐱𝐱)=_𝐲𝕋_α𝕋_γ_{Ψπ_2(𝐱,𝐱,𝐲)|μ(Ψ)=0}_i=1nU_in_w_i(Ψ)V_in_z_i(Ψ)𝐲\begin{split}&F_{\_}{\alpha\beta\gamma}:\bm{\mathcal{C\!F\!L}}(\Sigma,\bm{\alpha},\bm{\beta},{\bf{z}},{\bf{w}})\otimes\bm{\mathcal{C\!F\!L}}(\Sigma,\bm{\beta},\bm{\gamma},{\bf{z}},{\bf{w}})\to\bm{\mathcal{C\!F\!L}}(\Sigma,\bm{\alpha},\bm{\gamma},{\bf{z}},{\bf{w}})\\ &F_{\_}{\alpha\beta\gamma}(\mathbf{x}\otimes\mathbf{x}^{\prime})=\sum_{\_}{\mathbf{y}\in\mathbb{T}_{\_}{\alpha}\cap\mathbb{T}_{\_}{\gamma}}\sum_{\_}{\{\Psi\in\pi_{\_}2(\mathbf{x},\mathbf{x}^{\prime},\mathbf{y})|\mu(\Psi)=0\}}\prod_{\_}{i=1}^{n}U_{\_}i^{n_{\_}{w_{\_}i}(\Psi)}V_{\_}i^{n_{\_}{z_{\_}i}(\Psi)}\cdot\mathbf{y}\end{split}

Analogously, we define chain maps F_αγδF_{\_}{\alpha\gamma\delta} and F_αδβF_{\_}{\alpha\delta\beta}.

The Heegaard diagram _βγ=(Σ,𝜷,𝜸,𝐳,𝐰)\mathcal{H}_{\_}{\beta\gamma}=(\Sigma,\bm{\beta},\bm{\gamma},{\bf{z}},{\bf{w}}) represents an nn component unlink in #g1(S1×S2)\#^{g-1}(S^{1}\times S^{2}), denoted by O_nO_{\_}n. It is straightforward that 𝓗𝓕𝓛(#g1(S1×S2),O_n)\bm{{\mathcal{H\!F\!L}}}(\#^{g-1}(S^{1}\times S^{2}),O_{\_}n) is a free 𝐑\mathbf{R}-module of rank 2g12^{g-1}. Moreover, the summand with largest gr_𝐰\textup{gr}_{\_}{{\bf{w}}} has rank one and so it has a unique generator. The Heegaard diagram _βγ\mathcal{H}_{\_}{\beta\gamma} has an intersection point denoted by Θ_βγ\Theta_{\_}{\beta\gamma} that generates this top degree homology class, called top generator. Specifically, Θ_βγ\Theta_{\_}{\beta\gamma} is the intersection point that every element of π_2(𝐱,Θ_βγ)\pi_{\_}2(\mathbf{x},\Theta_{\_}{\beta\gamma}) has nonzero coefficient in at least one 𝐳{\bf{z}} or 𝐰{\bf{w}} basepoint, for all other intersection points 𝐱\mathbf{x}. Top generators Θ_γδ\Theta_{\_}{\gamma\delta} and Θ_δβ\Theta_{\_}{\delta\beta} for 𝓒𝓕𝓛(_γδ)\bm{\mathcal{C\!F\!L}}(\mathcal{H}_{\_}{\gamma\delta}) and 𝓒𝓕𝓛(_δβ)\bm{\mathcal{C\!F\!L}}(\mathcal{H}_{\_}{\delta\beta}), respectively, are defined analogously.

Let

f_1()=F_αβγ(Θ_βγ),f_2()=F_αγδ(Θ_γδ)andf_3()=F_αδβ(Θ_δβ).f_{\_}1(\cdot)=F_{\_}{\alpha\beta\gamma}(\cdot\otimes\Theta_{\_}{\beta\gamma}),\quad f_{\_}2(\cdot)=F_{\_}{\alpha\gamma\delta}(\cdot\otimes\Theta_{\_}{\gamma\delta})\quad\text{and}\quad f_{\_}3(\cdot)=F_{\_}{\alpha\delta\beta}(\cdot\otimes\Theta_{\_}{\delta\beta}).

By definition of cobordism maps in [32], for any 1i31\leq i\leq 3 we have

F_i=(f_i)_=_𝔰Spinc(W_i)F_W_i,_i,𝔰.F_{\_}i=(f_{\_}i)_{\_}*=\sum_{\_}{\mathfrak{s}\in\mathrm{Spin}^{c}(W_{\_}i)}F_{\_}{W_{\_}i,\mathcal{F}_{\_}i,\mathfrak{s}}.

By [29, Lemma 4.4] to show that they form an exact triangle, we need to check that

  1. (1)

    f_i+1f_if_{\_}{i+1}\circ f_{\_}i is chain homotopically trivial by a chain homotopy h_ih_{\_}i,

  2. (2)

    f_i+2h_i+h_i+1f_if_{\_}{i+2}\circ h_{\_}i+h_{\_}{i+1}\circ f_{\_}i is a homotopy equivalence,

where indices are cyclic modulo three. Note that we need a version of [29, Lemma 4.4] for chain complexes over 𝐑\mathbf{R}, which for example follows from [1, Lemma 3.3].

First, we check condition (1)(1). Suppose i=1i=1. The proof for i=2i=2 and 33 is similar. Then,

f_2f_1()=F_αγδ(F_αβγ(Θ_βγ)Θ_γδ)F_αβδ(F_βγδ(Θ_βγΘ_γδ)),f_{\_}2\circ f_{\_}1(\cdot)=F_{\_}{\alpha\gamma\delta}(F_{\_}{\alpha\beta\gamma}(\cdot\otimes\Theta_{\_}{\beta\gamma})\otimes\Theta_{\_}{\gamma\delta})\simeq F_{\_}{\alpha\beta\delta}(\cdot\otimes F_{\_}{\beta\gamma\delta}(\Theta_{\_}{\beta\gamma}\otimes\Theta_{\_}{\gamma\delta})),

where the chain homotopy is h_i()=F_αβγδ(Θ_βγΘ_γδ)h_{\_}i(\cdot)=F_{\_}{\alpha\beta\gamma\delta}(\cdot\otimes\Theta_{\_}{\beta\gamma}\otimes\Theta_{\_}{\gamma\delta}) and

F_αβγδ:𝓒𝓕𝓛(_αβ)𝓒𝓕𝓛(_βγ)𝓒𝓕𝓛(_γδ)𝓒𝓕𝓛(_αδ)F_αβγδ(𝐱𝐱𝐱)=_𝐲𝕋_α𝕋_δ_{ϕπ_2(𝐱,𝐱,𝐱,𝐲)|μ(ϕ)=1}_i=1nU_in_w_i(ϕ)V_in_z_i(ϕ)𝐲\begin{split}&F_{\_}{\alpha\beta\gamma\delta}:\bm{\mathcal{C\!F\!L}}(\mathcal{H}_{\_}{\alpha\beta})\otimes\bm{\mathcal{C\!F\!L}}(\mathcal{H}_{\_}{\beta\gamma})\otimes\bm{\mathcal{C\!F\!L}}(\mathcal{H}_{\_}{\gamma\delta})\to\bm{\mathcal{C\!F\!L}}(\mathcal{H}_{\_}{\alpha\delta})\\ &F_{\_}{\alpha\beta\gamma\delta}(\mathbf{x}\otimes\mathbf{x}^{\prime}\otimes\mathbf{x}^{\prime\prime})=\sum_{\_}{\mathbf{y}\in\mathbb{T}_{\_}{\alpha}\cap\mathbb{T}_{\_}{\delta}}\sum_{\_}{\{\phi\in\pi_{\_}2(\mathbf{x},\mathbf{x}^{\prime},\mathbf{x}^{\prime\prime},\mathbf{y})|\mu(\phi)=-1\}}\prod_{\_}{i=1}^{n}U_{\_}i^{n_{\_}{w_{\_}i}(\phi)}V_{\_}i^{n_{\_}{z_{\_}i}(\phi)}\cdot\mathbf{y}\end{split}

An argument similar to the proof of [24, Proposition 9.5] implies that F_βγδ(Θ_βγΘ_γδ)=0F_{\_}{\beta\gamma\delta}(\Theta_{\_}{\beta\gamma}\otimes\Theta_{\_}{\gamma\delta})=0 and so f_2f_10f_{\_}2\circ f_{\_}1\simeq 0.

For condition (2)(2), let 𝜷\bm{\beta}^{\prime} be a generic small Hamiltonian isotopic translate of 𝜷\bm{\beta}, and F_αβγδβF_{\_}{\alpha\beta\gamma\delta\beta^{\prime}} be the chain map defined by counting pentagons of Maslov index 2-2, analogous to F_αβγF_{\_}{\alpha\beta\gamma} and F_αβγδF_{\_}{\alpha\beta\gamma\delta}. Then, F_αβγδβ(Θ_βγΘ_γδΘ_δβ)F_{\_}{\alpha\beta\gamma\delta\beta^{\prime}}(\cdot\otimes\Theta_{\_}{\beta\gamma}\otimes\Theta_{\_}{\gamma\delta}\otimes\Theta_{\_}{\delta\beta^{\prime}}) gives a chain homotopy between f_3h_1+h_2f_1f_{\_}3\circ h_{\_}1+h_{\_}2\circ f_{\_}1 and F_αββ(F_βγδβ(Θ_βγΘ_γδΘ_δβ))F_{\_}{\alpha\beta\beta^{\prime}}(\cdot\otimes F_{\_}{\beta\gamma\delta\beta^{\prime}}(\Theta_{\_}{\beta\gamma}\otimes\Theta_{\_}{\gamma\delta}\otimes\Theta_{\_}{\delta\beta^{\prime}})). By a standard “stretching the neck argument” and following the strategy in [28, Section 2] and [29, Section 4.2] we have

F_βγδβ(Θ_βγΘ_γδΘ_δβ)=_k=0𝐔k(k+1)2Θ_ββF_{\_}{\beta\gamma\delta\beta^{\prime}}(\Theta_{\_}{\beta\gamma}\otimes\Theta_{\_}{\gamma\delta}\otimes\Theta_{\_}{\delta\beta^{\prime}})=\sum_{\_}{k=0}^{\infty}\mathbf{U}^{\frac{k(k+1)}{2}}\Theta_{\_}{\beta\beta^{\prime}}

and

F_αββ(F_βγδβ(Θ_βγΘ_γδΘ_δβ))=(_k=0𝐔k(k+1)2)F_αββ(Θ_ββ)F_{\_}{\alpha\beta\beta^{\prime}}(\cdot\otimes F_{\_}{\beta\gamma\delta\beta^{\prime}}(\Theta_{\_}{\beta\gamma}\otimes\Theta_{\_}{\gamma\delta}\otimes\Theta_{\_}{\delta\beta^{\prime}}))=\left(\sum_{\_}{k=0}^{\infty}\mathbf{U}^{\frac{k(k+1)}{2}}\right)F_{\_}{\alpha\beta\beta^{\prime}}(\cdot\otimes\Theta_{\_}{\beta\beta^{\prime}})

Since F_αββ(Θ_ββ)F_{\_}{\alpha\beta\beta^{\prime}}(\cdot\otimes\Theta_{\_}{\beta\beta^{\prime}}) is a homotopy equivalence (see the proof of [1, Theorem 8.6]) and _k=0𝐔k(k+1)2\sum_{\_}{k=0}^{\infty}\mathbf{U}^{\frac{k(k+1)}{2}} is invertible, f_3h_1+h_2f_1f_{\_}3\circ h_{\_}1+h_{\_}2\circ f_{\_}1 is a homotopy equivalence.

Now let us relate the surgery exact triangle with resolutions. Suppose Y=S3Y=S^{3}, L=L_+L=L_{\_}+ is a link in S3S^{3} with a fixed positive crossing, and KS3LK\subset S^{3}\setminus L is an unknot as in the top of Figure 3. Then, (S3_1(K),L)(S^{3}_{\_}{-1}(K),L) will be identified with (S3,L_)(S^{3},L_{\_}{-}). Next we relate (S3,L_0)(S^{3},L_{\_}0) with (S3_0(K),L)(S^{3}_{\_}{0}(K),L), where L_0L_{\_}0 denotes the oriented resolution at the fixed crossing. Observe that S3_0(K)=S3#(S2×S1)S^{3}_{\_}{0}(K)=S^{3}\#(S^{2}\times S^{1}), and LL in S3_0(K)S^{3}_{\_}{0}(K) still has nn components, while L_0L_{\_}{0} is an (n1)(n-1)-component link. Note that we can replace the 2-handle attaching to KK with framing 0 by a 1-handle as in Figure 12.

Refer to caption
Figure 12.
Lemma 4.2.

The link (S3_0(K),L)(S^{3}_{\_}{0}(K),L) can be identified with (S3#(S2×S1),L_0#Z_2)(S^{3}\#(S^{2}\times S^{1}),L_{\_}0\#Z_{\_}2) where Z_2Z_{\_}2 is the 2-component unlink in S2×S1S^{2}\times S^{1} consisting of two parallel circles representing the homology class of S1S^{1}, and the #\# between L_0L_{\_}0 and Z_2Z_{\_}2 is identified as in Figure 13.

Refer to caption
Figure 13. Special (local) connected sum between L_0S3L_{\_}0\subset S^{3} (in blue) and Z_2S2×S1Z_{\_}2\subset S^{2}\times S^{1} (one component in blue and another component in red)
Proof.

The proof is depicted in Figure 12. Specifically, we regard the 22-handle for the 0-surgery on KK as a 11-handle and then we move the feet of 1-handle along the link LL. At the end, we get the connected sum of L_0L_{\_}0 with one component of Z_2Z_{\_}2, colored blue, along with the other component of Z_2Z_{\_}2, colored red, in Figure 13.

Therefore, we have the following theorem:

Theorem 4.3.

Given a local positive crossing of the link components L_iL_{\_}i and L_jL_{\_}j of a link L_+L_{\_}+ in the integer homology sphere YY, there is a skein exact sequence

(5) (Y,L_+)Ψ(Y,L_)𝛼H_(𝒞(Y,L_0)_RR_0)𝛽(Y,L_+)\rightarrow\mathcal{H\!F\!L}(Y,L_{\_}{+})\xrightarrow{\Psi}\mathcal{H\!F\!L}(Y,L_{\_}{-})\xrightarrow{\alpha}H_{\_}*(\mathcal{C\!F\!L}(Y,L_{\_}0)\otimes_{\_}{R}R_{\_}0)\xrightarrow{\beta}\mathcal{H\!F\!L}(Y,L_{\_}{+})\rightarrow

where the map from (Y,L_+)\mathcal{H\!F\!L}(Y,L_{\_}{+}) to (Y,L_)\mathcal{H\!F\!L}(Y,L_{\_}{-}) is given by Ψ=_k(1)kψ_k\Psi=\sum_{\_}{k\in\mathbb{Z}}(-1)^{k}\psi_{\_}k, and

R_0=𝔽[U_1,,U_n,V_1,,V_n](U_iU_j,V_iV_j).R_{\_}0=\frac{\mathbb{F}[U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n]}{(U_{\_}i-U_{\_}j,V_{\_}i-V_{\_}j)}.
Proof.

By Lemma 4.2, the cone of Ψ\Psi is the homology of the tensor of 𝒞(Y,L_0)\mathcal{C\!F\!L}(Y,L_{\_}0) with some complex ZZ corresponding to the unlink Z_2Z_{\_}2 in S2×S1S^{2}\times S^{1}. Moreover, Z_2Z_{\_}2 is independent of the pair (Y,L_+)(Y,L_{\_}+). We use the special case that Y=S3Y=S^{3} and L_+=T(2,2)L_{\_}+=T(2,2) to compute 𝒞(S2×S1,Z_2)\mathcal{C\!F\!L}(S^{2}\times S^{1},Z_{\_}2) which gives the module 𝒦\mathcal{K}. We put the detailed computation of Hopf link in Section 4.2 (equivalently, see (8)).

The complex 𝒦\mathcal{K} is given as follows:

(6) 𝒦=RRRRU_iU_jV_iV_jV_iV_jU_jU_i\mathcal{K}=\leavevmode\hbox to99.9pt{\vbox to67.8pt{\pgfpicture\makeatletter\hbox{\hskip 49.95216pt\lower-32.15279pt\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\pgfsys@setlinewidth{0.4pt}\pgfsys@invoke{ }\nullfont\hbox to0.0pt{\pgfsys@beginscope\pgfsys@invoke{ }{}{}{}{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{\offinterlineskip{}{}{{{}}{{}}{{}}{{}}}{{{}}}{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-28.2812pt}{-32.15279pt}\pgfsys@invoke{ }\hbox{\vbox{\halign{\pgf@matrix@init@row\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding&&\pgf@matrix@step@column{\pgf@matrix@startcell#\pgf@matrix@endcell}&#\pgf@matrix@padding\cr\hfil\hskip 8.14061pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-3.83507pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${R}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}}}&\hskip 8.14061pt\hfil&\hfil\hskip 32.14058pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-3.83507pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${R}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 8.14061pt\hfil\cr\vskip 18.00005pt\cr\hfil\hskip 8.14061pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-3.83507pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${R}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 8.14061pt\hfil&\hfil\hskip 32.14058pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-3.83507pt}{0.0pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{${R}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 8.14061pt\hfil\cr\vskip 18.00005pt\cr\hfil\hskip 0.0pt\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}&\hskip 0.0pt\hfil\cr}}}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}}{{{{}}}{{}}{{}}{{}}{{}}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} {}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-11.79999pt}{24.15976pt}\pgfsys@lineto{11.40002pt}{24.15976pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{11.6pt}{24.15976pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-12.55675pt}{28.71251pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{U_{\_}i-U_{\_}j}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-20.1406pt}{17.80005pt}\pgfsys@lineto{-20.1406pt}{0.59996pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{-20.1406pt}{0.39998pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{{}{}}}{{}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-47.7994pt}{7.70836pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{V_{\_}i-V_{\_}j}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{20.1406pt}{17.80005pt}\pgfsys@lineto{20.1406pt}{0.59996pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{0.0}{-1.0}{1.0}{0.0}{20.1406pt}{0.39998pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{22.49336pt}{7.70836pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{V_{\_}i-V_{\_}j}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{ {}{}{}}{}{ {}{}{}} {{{{{}}{ {}{}}{}{}{{}{}}}}}{}{{{{{}}{ {}{}}{}{}{{}{}}}}}{{}}{}{}{}{}{}{{{}{}}}{}{{}}{}{}{}{{{}{}}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@setlinewidth{0.39998pt}\pgfsys@invoke{ }{}{}{}{}{{}}{}{}{{}}\pgfsys@moveto{-11.79999pt}{-7.99303pt}\pgfsys@lineto{11.40002pt}{-7.99303pt}\pgfsys@stroke\pgfsys@invoke{ }{{}{{}}{}{}{{}}{{{}}}}{{}{{}}{}{}{{}}{{{}}{{{}}{\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{11.6pt}{-7.99303pt}\pgfsys@invoke{ }\pgfsys@invoke{ \lxSVG@closescope }\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}{{}}}}\hbox{\hbox{{\pgfsys@beginscope\pgfsys@invoke{ }{{}{}{{ {}{}}}{ {}{}} {{}{{}}}{{}{}}{}{{}{}} { }{{{{}}\pgfsys@beginscope\pgfsys@invoke{ }\pgfsys@transformcm{1.0}{0.0}{0.0}{1.0}{-12.55675pt}{-15.12912pt}\pgfsys@invoke{ }\hbox{{\definecolor{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@rgb@stroke{0}{0}{0}\pgfsys@invoke{ }\pgfsys@color@rgb@fill{0}{0}{0}\pgfsys@invoke{ }\hbox{$\scriptstyle{U_{\_}j-U_{\_}i}$} }}\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope}}} \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope \pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope{}{}{}\hss}\pgfsys@discardpath\pgfsys@invoke{\lxSVG@closescope }\pgfsys@endscope\hss}}\lxSVG@closescope\endpgfpicture}}

Note that it is a free resolution of R_0R_{\_}0 over RR. Since 𝒞(Y,L_0)\mathcal{C\!F\!L}(Y,L_{\_}0) is a complex of free RR-modules, we get

H_(𝒞(Y,L_0)𝒦)H_(𝒞(Y,L_0)R_0).H_{\_}*(\mathcal{C\!F\!L}(Y,L_{\_}0)\otimes\mathcal{K})\simeq H_{\_}*(\mathcal{C\!F\!L}(Y,L_{\_}0)\otimes R_{\_}0).

Remark 4.4.

In 𝐻𝐹𝐿\mathit{HFL}^{-} version of Heegaard Floer homology one sets V_i=V_j=0V_{\_}i=V_{\_}j=0, and the complex 𝒦\mathcal{K} breaks into a direct sum of two copies of 𝔽[U_1,,U_n]U_iU_j𝔽[U_1,,U_n].\mathbb{F}[U_{\_}1,\ldots,U_{\_}n]\xrightarrow{U_{\_}i-U_{\_}j}\mathbb{F}[U_{\_}1,\ldots,U_{\_}n]. This explains the appearance of a two-dimensional vector space in [30].

Similarly, for 𝐻𝐹𝐿^\widehat{\mathit{HFL}} one sets U_i=U_j=V_i=V_j=0U_{\_}i=U_{\_}j=V_{\_}i=V_{\_}j=0, and the complex 𝒦\mathcal{K} breaks into four copies of 𝔽\mathbb{F}.

Without loss of generality, we assume that i=1,j=2i=1,j=2 for the rest of the section. Recall that L_+L_{\_}+ and L_L_{\_}- have nn components while L_0L_{\_}0 has (n1)(n-1) components. For all kk\in\mathbb{Z} we have chain maps ψ_k:(Y,L_+)(Y,L_)\psi_{\_}k:\mathcal{H\!F\!L}(Y,L_{\_}{+})\to\mathcal{H\!F\!L}(Y,L_{\_}{-}), and one can consider the formal sum

Ψ=_k(1)kψ_k:𝓗𝓕𝓛(Y,L_+)𝓗𝓕𝓛(Y,L_)\Psi=\sum_{\_}{k\in\mathbb{Z}}(-1)^{k}\psi_{\_}k:\bm{{\mathcal{H\!F\!L}}}(Y,L_{\_}{+})\to\bm{{\mathcal{H\!F\!L}}}(Y,L_{\_}{-})

Note that Ψ\Psi is a non-homogeneous map containing terms of various non-positive homological degrees.

As the non-homogeneous map Ψ\Psi is hard to deal with, we would like to reduce it to the degree zero piece Ψ0=ψ_0ψ_1\Psi^{0}=\psi_{\_}0-\psi_{\_}{-1}.

Lemma 4.5.

Let L=L_+L=L_{\_}+ be an arbitrary link in the three-sphere with a fixed positive crossing between its first and second components. For kk\in\mathbb{Z}, suppose ψ_k:𝓗𝓕𝓛(L_+)𝓗𝓕𝓛(L_)\psi_{\_}k:\bm{{\mathcal{H\!F\!L}}}(L_{\_}{+})\to\bm{{\mathcal{H\!F\!L}}}(L_{\_}{-}) is the corresponding crossing change map and Ψ=_k(1)kψ_k\Psi=\sum_{\_}{k\in\mathbb{Z}}(-1)^{k}\psi_{\_}k. Then in homology we have Ψ=τ(ψ_0ψ_1)\Psi=\tau(\psi_{\_}{0}-\psi_{\_}{-1}) where

(7) τ=_k0(1)k[(V_1U_2)k+(V_1U_2)k1(U_1V_2)+(U_1V_2)k]𝐔k(k1)2+_k1(1)k[(V_1U_2)k1+(V_1U_2)k2(U_1V_2)+(U_1V_2)k1]𝐔k(k1)2+1=1+.\tau=\sum_{\_}{k\geq 0}(-1)^{k}\left[(V_{\_}1U_{\_}2)^{k}+(V_{\_}1U_{\_}2)^{k-1}(U_{\_}1V_{\_}2)\ldots+(U_{\_}1V_{\_}2)^{k}\right]\mathbf{U}^{\frac{k(k-1)}{2}}+\\ \sum_{\_}{k\geq 1}(-1)^{k}\left[(V_{\_}1U_{\_}2)^{k-1}+(V_{\_}1U_{\_}2)^{k-2}(U_{\_}1V_{\_}2)\ldots+(U_{\_}1V_{\_}2)^{k-1}\right]\mathbf{U}^{\frac{k(k-1)}{2}+1}=1+\ldots.

In particular, τ\tau is an invertible power series.

Proof.

First, we introduce notations

A_k=(V_1U_2)k𝐔k(k1)2,B_k=(U_1V_2)k𝐔k(k1)2,A_{\_}k=(V_{\_}1U_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}},\ B_{\_}k=(U_{\_}1V_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}},
C_k=(V_1U_2)k1++(U_1V_2)k1=_i=0k1(V_1U_2)i(U_1V_2)k1i.C_{\_}k=(V_{\_}1U_{\_}2)^{k-1}+\ldots+(U_{\_}1V_{\_}2)^{k-1}=\sum_{\_}{i=0}^{k-1}(V_{\_}1U_{\_}2)^{i}(U_{\_}1V_{\_}2)^{k-1-i}.

Clearly,

A_kB_k=(V_1U_2U_1V_2)C_k𝐔k(k1)2A_{\_}k-B_{\_}k=(V_{\_}1U_{\_}2-U_{\_}1V_{\_}2)C_{\_}k\mathbf{U}^{\frac{k(k-1)}{2}}

and so

A_k+U_1V_2C_k𝐔k(k1)2=B_k+V_1U_2C_k𝐔k(k1)2=C_k+1𝐔k(k1)2,A_{\_}k+U_{\_}1V_{\_}2C_{\_}k\mathbf{U}^{\frac{k(k-1)}{2}}=B_{\_}k+V_{\_}1U_{\_}2C_{\_}k\mathbf{U}^{\frac{k(k-1)}{2}}=C_{\_}{k+1}\mathbf{U}^{\frac{k(k-1)}{2}},

and

τ=_k0(1)k(C_k+1+C_k𝐔)𝐔k(k1)2=_k0(1)k(B_k+(V_1U_2+V_1U_1)C_k𝐔k(k1)2).\tau=\sum_{\_}{k\geq 0}(-1)^{k}(C_{\_}{k+1}+C_{\_}k\mathbf{U})\mathbf{U}^{\frac{k(k-1)}{2}}=\sum_{\_}{k\geq 0}(-1)^{k}\left(B_{\_}k+(V_{\_}1U_{\_}2+V_{\_}1U_{\_}1)C_{\_}k\mathbf{U}^{\frac{k(k-1)}{2}}\right).

By Lemma 3.4 parts (a) and (b) we have ψ_k=A_kψ_0,ψ_1k=B_kψ_1\psi_{\_}k=A_{\_}k\psi_{\_}0,\ \psi_{\_}{-1-k}=B_{\_}k\psi_{\_}{-1} for k0k\geq 0, and therefore

Ψ=_k0(1)kA_kψ_0_k0(1)kB_kψ_1=\Psi=\sum_{\_}{k\geq 0}(-1)^{k}A_{\_}k\psi_{\_}0-\sum_{\_}{k\geq 0}(-1)^{k}B_{\_}k\psi_{\_}{-1}=
_k0(1)kB_k(ψ_0ψ_1)+_k0(1)k(A_kB_k)ψ_0=\sum_{\_}{k\geq 0}(-1)^{k}B_{\_}k(\psi_{\_}0-\psi_{\_}{-1})+\sum_{\_}{k\geq 0}(-1)^{k}(A_{\_}k-B_{\_}k)\psi_{\_}{0}=
_k0(1)kB_k(ψ_0ψ_1)+_k0(1)k𝐔k(k1)2C_k(V_1U_2U_1V_2)ψ_0.\sum_{\_}{k\geq 0}(-1)^{k}B_{\_}k(\psi_{\_}0-\psi_{\_}{-1})+\sum_{\_}{k\geq 0}(-1)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}C_{\_}k(V_{\_}1U_{\_}2-U_{\_}1V_{\_}2)\psi_{\_}{0}.

By Lemma 3.4 part (c) we have

V_2ψ_0=V_1ψ_1,U_1ψ_0=U_2ψ_1,V_{\_}2\psi_{\_}0=V_{\_}1\psi_{\_}{-1},\ U_{\_}1\psi_{\_}0=U_{\_}2\psi_{\_}{-1},

so

(V_1U_2+V_1U_1)(ψ_0ψ_1)=V_1U_2ψ_0V_1U_2ψ_1+V_1U_1ψ_0V_1U_1ψ_1=(V_{\_}1U_{\_}2+V_{\_}1U_{\_}1)(\psi_{\_}0-\psi_{\_}{-1})=V_{\_}1U_{\_}2\psi_{\_}0-V_{\_}1U_{\_}2\psi_{\_}{-1}+V_{\_}1U_{\_}1\psi_{\_}0-V_{\_}1U_{\_}1\psi_{\_}{-1}=
V_1U_2ψ_0V_1U_1ψ_0+V_1U_1ψ_0V_2U_1ψ_0=(V_1U_2V_2U_1)ψ_0.V_{\_}1U_{\_}2\psi_{\_}0-V_{\_}1U_{\_}1\psi_{\_}{0}+V_{\_}1U_{\_}1\psi_{\_}0-V_{\_}2U_{\_}1\psi_{\_}{0}=(V_{\_}1U_{\_}2-V_{\_}2U_{\_}1)\psi_{\_}0.

Therefore

Ψ=_k0(1)kB_k(ψ_0ψ_1)+_k0(1)k𝐔k(k1)2C_k(V_1U_2+V_1U_1)(ψ_0ψ_1)=τ(ψ_0ψ_1).\Psi=\sum_{\_}{k\geq 0}(-1)^{k}B_{\_}k(\psi_{\_}0-\psi_{\_}{-1})+\sum_{\_}{k\geq 0}(-1)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}C_{\_}k(V_{\_}1U_{\_}2+V_{\_}1U_{\_}1)(\psi_{\_}0-\psi_{\_}{-1})=\tau(\psi_{\_}0-\psi_{\_}{-1}).

Corollary 4.6.

The cones of Ψ\Psi and of ψ_0ψ_1\psi_{\_}0-\psi_{\_}{-1} are homotopy equivalent.

Remark 4.7.

Note that at V_1=V_2=1V_{\_}1=V_{\_}2=1 and U_1=U_2=𝐔U_{\_}1=U_{\_}2=\mathbf{U}, we get that the invertible factor τ\tau is

τ=_k0(1)k((k+1)𝐔k𝐔k(k1)2+k𝐔k1𝐔k(k1)2+1)=_k0(1)k(2k+1)𝐔k(k+1)2\tau=\sum_{\_}{k\geq 0}(-1)^{k}\left((k+1)\mathbf{U}^{k}\cdot\mathbf{U}^{\frac{k(k-1)}{2}}+k\mathbf{U}^{k-1}\cdot\mathbf{U}^{\frac{k(k-1)}{2}+1}\right)=\sum_{\_}{k\geq 0}(-1)^{k}(2k+1)\mathbf{U}^{\frac{k(k+1)}{2}}

which agrees with [27, Theorem 3.7, Blow-up formula] modulo 22. That is because in this case ψ_k\psi_{\_}k is equal to the Ozsváth-Szabó’s cobordism map associated to the blow-up of the product cobordism (S3×[0,1])#2¯\left(S^{3}\times[0,1]\right)\#\overline{\mathbb{CP}^{2}} from S3S^{3} to S3S^{3}.

4.2. Skein exact sequence for Hopf link

Next, we compute the skein exact sequence on the chain complex level for the Hopf link. Recall that the full link Floer complex for H=T(2,2)H=T(2,2) has generators a,b,c,da,b,c,d and the differential

(c)=U_1aV_2b,(d)=U_2aV_1b.\partial(c)=U_{\_}1a-V_{\_}2b,\ \partial(d)=U_{\_}2a-V_{\_}1b.

The homology is generated by a,ba,b modulo relations U_1a=V_2b,U_2a=V_1bU_{\_}1a=V_{\_}2b,U_{\_}2a=V_{\_}1b as above.

On the other hand, the link Floer complex for the unknot has generators 1,ξ1,\xi and the differential (ξ)=(U_1V_1U_2V_2)\partial(\xi)=(U_{\_}1V_{\_}1-U_{\_}2V_{\_}2). By Example 3.2 we have

ψ_0(a)=V_1,ψ_0(b)=U_2,ψ_1(a)=V_2,ψ_1(b)=U_1.\psi_{\_}0(a)=V_{\_}1,\ \psi_{\_}0(b)=U_{\_}2,\ \psi_{\_}{-1}(a)=V_{\_}2,\ \psi_{\_}{-1}(b)=U_{\_}1.

and the maps ψ_0\psi_{\_}0 and ψ_1\psi_{\_}{-1} can be uniquely lifted to chain complex level by setting

ψ_0(c)=ξ,ψ_0(d)=0,ψ_1(c)=0,ψ_1(d)=ξ.\psi_{\_}0(c)=\xi,\ \psi_{\_}0(d)=0,\psi_{\_}{-1}(c)=0,\ \psi_{\_}{-1}(d)=-\xi.

Furthermore, for all k0k\geq 0 we have ψ_k=A_kψ_0,ψ_1k=B_kψ_1,\psi_{\_}k=A_{\_}k\psi_{\_}0,\ \psi_{\_}{-1-k}=B_{\_}k\psi_{\_}{-1}, where we follow the notations in Lemma 4.5 and its proof. For concreteness, we can lift the statement of Lemma 4.5 to the level of chain complexes.

Lemma 4.8.

The map Ψ:=_k(1)kψ_k\Psi:=\sum_{\_}{k\in\mathbb{Z}}(-1)^{k}\psi_{\_}k is homotopic to τ(ψ_0ψ_1)\tau(\psi_{\_}0-\psi_{\_}{-1}), where τ\tau is defined by (7).

Proof.

We have

Ψ=_k(1)kψ_k=_k0(1)k(A_kψ_0B_kψ_1).\Psi=\sum_{\_}{k\in\mathbb{Z}}(-1)^{k}\psi_{\_}k=\sum_{\_}{k\geq 0}(-1)^{k}(A_{\_}k\psi_{\_}0-B_{\_}k\psi_{\_}{-1}).

Define

h_a=V_1_k0(1)kC_k𝐔k(k1)2,h_b=U_1_k0(1)kC_k𝐔k(k1)2h_{\_}a=V_{\_}1\sum_{\_}{k\geq 0}(-1)^{k}C_{\_}k\mathbf{U}^{\frac{k(k-1)}{2}},\ h_{\_}b=-U_{\_}1\sum_{\_}{k\geq 0}(-1)^{k}C_{\_}k\mathbf{U}^{\frac{k(k-1)}{2}}

then

τ=_k0(1)kA_k+U_1h_aV_2h_b=_k0(1)kB_k+U_2h_aV_1h_b.\tau=\sum_{\_}{k\geq 0}(-1)^{k}A_{\_}k+U_{\_}1h_{\_}a-V_{\_}2h_{\_}b=\sum_{\_}{k\geq 0}(-1)^{k}B_{\_}k+U_{\_}2h_{\_}a-V_{\_}1h_{\_}b.

We define the homotopy hh by h(a)=h_aξ,h(b)=h_bξh(a)=h_{\_}a\xi,h(b)=h_{\_}b\xi and h(c)=h(d)=0h(c)=h(d)=0, and let Ψ~=Ψ+h+h.\widetilde{\Psi}=\Psi+\partial h+h\partial. Then,

Ψ~(c)=_k0(1)kA_kξ+h(U_1aV_2b)=τξ\widetilde{\Psi}(c)=\sum_{\_}{k\geq 0}(-1)^{k}A_{\_}k\xi+h(U_{\_}1a-V_{\_}2b)=\tau\xi
Ψ~(d)=_k0(1)kB_kξ+h(U_2aV_1b)=τξ\widetilde{\Psi}(d)=\sum_{\_}{k\geq 0}(-1)^{k}B_{\_}k\xi+h(U_{\_}2a-V_{\_}1b)=\tau\xi
Ψ~(a)=V_1_k0(1)kA_kV_2_k0(1)kB_k+(h_aξ)=\widetilde{\Psi}(a)=V_{\_}1\sum_{\_}{k\geq 0}(-1)^{k}A_{\_}k-V_{\_}2\sum_{\_}{k\geq 0}(-1)^{k}B_{\_}k+\partial(h_{\_}a\xi)=
V_1_k0(1)kA_kV_2_k0(1)kB_k+(U_1V_1U_2V_2)h_a=(V_1V_2)τV_{\_}1\sum_{\_}{k\geq 0}(-1)^{k}A_{\_}k-V_{\_}2\sum_{\_}{k\geq 0}(-1)^{k}B_{\_}k+(U_{\_}1V_{\_}1-U_{\_}2V_{\_}2)h_{\_}a=(V_{\_}1-V_{\_}2)\tau

Similarly, Ψ~(b)=τ(U_2U_1)\widetilde{\Psi}(b)=\tau(U_{\_}2-U_{\_}1) and we conclude that Ψ~=τ(ψ_0ψ_1)\widetilde{\Psi}=\tau(\psi_{\_}0-\psi_{\_}{-1}). ∎

By Lemma 4.8 we can replace the cone of Ψ\Psi by the cone of ψ_0ψ_1\psi_{\_}0-\psi_{\_}{-1} which is isomorphic to the following complex:

ξ{\xi}c{c}a{a}1{1}d{d}b{b}U_1V_1U_2V_2\scriptstyle{U_{\_}1V_{\_}1-U_{\_}2V_{\_}2}U_1\scriptstyle{U_{\_}1}V_2\scriptstyle{-V_{\_}2}1\scriptstyle{1}V_1V_2\scriptstyle{V_{\_}1-V_{\_}2}V_1\scriptstyle{-V_{\_}1}U_2\scriptstyle{U_{\_}2}1\scriptstyle{1}U_2U_1\scriptstyle{U_{\_}2-U_{\_}1}

We can define ξ=cd\xi^{\prime}=c-d and change the basis from (c,d)(c,d) to (ξ,d)(\xi^{\prime},d):

(8) ξ{\xi}ξ{\xi^{\prime}}a{a}1{1}d{d}b{b}U_1V_1U_2V_2\scriptstyle{U_{\_}1V_{\_}1-U_{\_}2V_{\_}2}U_1U_2\scriptstyle{U_{\_}1-U_{\_}2}V_1V_2\scriptstyle{V_{\_}1-V_{\_}2}V_1V_2\scriptstyle{V_{\_}1-V_{\_}2}V_1\scriptstyle{-V_{\_}1}U_2\scriptstyle{U_{\_}2}1\scriptstyle{1}U_2U_1\scriptstyle{U_{\_}2-U_{\_}1}

\simeq ξ{\xi^{\prime}}a{a}b{b}1{1}U_1U_2\scriptstyle{U_{\_}1-U_{\_}2}V_1V_2\scriptstyle{V_{\_}1-V_{\_}2}V_1V_2\scriptstyle{V_{\_}1-V_{\_}2}U_2U_1\scriptstyle{U_{\_}2-U_{\_}1}

Here we use the fact that the quotient complex d1ξd\xrightarrow{1}\xi is contractible.

Lemma 4.9.

Let Φ=_k(1)kϕ_k:𝓒𝓕𝓛(O_2)𝓒𝓕𝓛(T(2,2))\Phi=\sum_{\_}{k}(-1)^{k}\phi_{\_}k:\bm{\mathcal{C\!F\!L}}(O_{\_}2)\to\bm{\mathcal{C\!F\!L}}(T(2,2)), then the cone of Φ\Phi is quasi-isomorphic to the cone of Ψ\Psi up to relabeling the variables.

Proof.

By Example 3.6 we have ϕ_k(1)=(V_1V_2)k𝐔k(k1)2a\phi_{\_}{-k}(1)=(V_{\_}1V_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}a and ϕ_1+k(1)=(U_1U_2)k𝐔k(k1)2b\phi_{\_}{1+k}(1)=(U_{\_}1U_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}b for k0k\geq 0, so

Φ(1)=a_k0(1)k(V_1V_2)k𝐔k(k1)2b_k0(1)k(U_1U_2)k𝐔k(k1)2.\Phi(1)=a\sum_{\_}{k\geq 0}(-1)^{k}(V_{\_}1V_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}-b\sum_{\_}{k\geq 0}(-1)^{k}(U_{\_}1U_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}.

The homology of the cone of Φ\Phi is generated by aa and bb modulo relations U_1a=V_2b,U_2a=V_1bU_{\_}1a=V_{\_}2b,U_{\_}2a=V_{\_}1b and Φ(1)=0\Phi(1)=0. Since the coefficients at aa and bb in Φ(1)\Phi(1) are invertible, the result is generated by aa modulo relations

(9) aV_2_k0(1)k(V_1V_2)k𝐔k(k1)2\displaystyle aV_{\_}2\sum_{\_}{k\geq 0}(-1)^{k}(V_{\_}1V_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}} =aU_1_k0(1)k(U_1U_2)k𝐔k(k1)2,\displaystyle=aU_{\_}1\sum_{\_}{k\geq 0}(-1)^{k}(U_{\_}1U_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}},
aV_1_k0(1)k(V_1V_2)k𝐔k(k1)2\displaystyle aV_{\_}1\sum_{\_}{k\geq 0}(-1)^{k}(V_{\_}1V_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}} =aU_2_k0(1)k(U_1U_2)k𝐔k(k1)2\displaystyle=aU_{\_}2\sum_{\_}{k\geq 0}(-1)^{k}(U_{\_}1U_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}

We claim that

(10) V_2_k0(1)k(V_1V_2)k𝐔k(k1)2U_1_k0(1)k(U_1U_2)k𝐔k(k1)2=(V_2U_1)τV_{\_}2\sum_{\_}{k\geq 0}(-1)^{k}(V_{\_}1V_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}-U_{\_}1\sum_{\_}{k\geq 0}(-1)^{k}(U_{\_}1U_{\_}2)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}=(V_{\_}2-U_{\_}1)\tau^{\prime}

where

τ=_k0(1)k((U_1U_2)k+(U_1U_2)k1(V_1V_2)++(V_1V_2)k)𝐔k(k1)2+_k0(1)k((U_1U_2)k1+(U_1U_2)k2(V_1V_2)+(V_1V_2)k1)𝐔k(k1)2+1.\tau^{\prime}=\sum_{\_}{k\geq 0}(-1)^{k}\left((U_{\_}1U_{\_}2)^{k}+(U_{\_}1U_{\_}2)^{k-1}(V_{\_}1V_{\_}2)+\ldots+(V_{\_}1V_{\_}2)^{k}\right)\mathbf{U}^{\frac{k(k-1)}{2}}+\\ \sum_{\_}{k\geq 0}(-1)^{k}\left((U_{\_}1U_{\_}2)^{k-1}+(U_{\_}1U_{\_}2)^{k-2}(V_{\_}1V_{\_}2)\ldots+(V_{\_}1V_{\_}2)^{k-1}\right)\mathbf{U}^{\frac{k(k-1)}{2}+1}.

Indeed, let C_k=_i=0k1(V_1V_2)i(U_1U_2)k1iC^{\prime}_{\_}{k}=\sum_{\_}{i=0}^{k-1}(V_{\_}1V_{\_}2)^{i}(U_{\_}1U_{\_}2)^{k-1-i}, then

V_2(V_1V_2)kU_1(U_1U_2)k=(V_2U_1)(V_1V_2)kU_1(U_1U_2V_1V_2)C_kV_{\_}2(V_{\_}1V_{\_}2)^{k}-U_{\_}1(U_{\_}1U_{\_}2)^{k}=(V_{\_}2-U_{\_}1)(V_{\_}1V_{\_}2)^{k}-U_{\_}1(U_{\_}1U_{\_}2-V_{\_}1V_{\_}2)C^{\prime}_{\_}k

and U_1U_2V_1V_2=(U_1V_2)(U_2+V_1)U_{\_}1U_{\_}2-V_{\_}1V_{\_}2=(U_{\_}1-V_{\_}2)(U_{\_}2+V_{\_}1), hence

(V_2U_1)[(V_1V_2)k+U_1(U_2+V_1)C_k]=(V_2U_1)[C_k+1+𝐔C_k].(V_{\_}2-U_{\_}1)\left[(V_{\_}1V_{\_}2)^{k}+U_{\_}1(U_{\_}2+V_{\_}1)C^{\prime}_{\_}k\right]=(V_{\_}2-U_{\_}1)\left[C^{\prime}_{\_}{k+1}+\mathbf{U}C^{\prime}_{\_}k\right].

Now by (10) we can rewrite the equations (9) as

a(V_2U_1)τ=a(V_1U_2)τ=0a(V_{\_}2-U_{\_}1)\tau^{\prime}=a(V_{\_}1-U_{\_}2)\tau^{\prime}=0

which is equivalent to

a(V_2U_1)=a(V_1U_2)=0.a(V_{\_}2-U_{\_}1)=a(V_{\_}1-U_{\_}2)=0.

Remark 4.10.

Note that the above computation is very similar to the one in Lemma 4.5, in particular, τ\tau^{\prime} is related to τ\tau (resp. C_kC^{\prime}_{\_}k is related to C_kC_{\_}k) by exchanging the variables V_2U_2V_{\_}2\leftrightarrow U_{\_}2 which corresponds to changing the orientation on a link component.

5. Link splitting maps

5.1. Splitting maps

Let L=L_1L_nL=L_{\_}1\cup\ldots\cup L_{\_}n be an arbitrary link in the three-sphere. We can change the crossings between different components in LL arbitrarily and consider the corresponding maps in Heegaard Floer homology: if we change a positive crossing to a negative one we can use either ψ_1\psi_{\_}{-1} or ψ_0\psi_{\_}0, and if we change a negative crossing to a positive one we can use either ϕ_0\phi_{\_}0 or ϕ_1\phi_{\_}1. This does not change the topological type of the components and we will denote the components for all such links by L_iL_{\_}i. In particular, by such crossing changes we can transform LL to the split link split(L)\mathrm{split}(L) obtained by the split union of all link components L_iL_{\_}i.

More precisely, we consider two links L,LL,L^{\prime} related by such crossing changes, and a chain map F:(L)(L)F:\mathcal{H\!F\!L}(L)\to\mathcal{H\!F\!L}(L^{\prime}) obtained as a composition of:

  • P_ij1P_{\_}{ij}^{-1} of maps ψ_1\psi_{\_}{-1} associated to positive crossings between L_iL_{\_}i and L_jL_{\_}j;

  • P_ij0P_{\_}{ij}^{0} of maps ψ_0\psi_{\_}{0} associated to positive crossings between L_iL_{\_}i and L_jL_{\_}j;

  • N_ij0N_{\_}{ij}^{0} of maps ϕ_0\phi_{\_}{0} associated to negative crossings between L_iL_{\_}i and L_jL_{\_}j;

  • N_ij1N_{\_}{ij}^{1} of maps ϕ_1\phi_{\_}{1} associated to negative crossings between L_iL_{\_}i and L_jL_{\_}j.

Note that, in principle, FF may depend on the order of the crossings and the choice of these crossings (for example, for the Hopf link we can change either one of two crossings to transform it to unlink). Also note that the linking number between L_iL_{\_}i and L_jL_{\_}j changes by

lk_L(L_i,L_j)lk_L(L_i,L_j)=N_ij1+N_ij0P_ij0P_ij1.\mathrm{lk}_{\_}{L^{\prime}}(L_{\_}i,L_{\_}j)-\mathrm{lk}_{\_}L(L_{\_}i,L_{\_}j)=N_{\_}{ij}^{1}+N_{\_}{ij}^{0}-P_{\_}{ij}^{0}-P_{\_}{ij}^{-1}.

Nevertheless, we have the following general result:

Lemma 5.1.

Let L,LL,L^{\prime} be two links related by crossing changes as above. Then:

  • (a)

    The chain map F:(L)(L)F:\mathcal{H\!F\!L}(L)\to\mathcal{H\!F\!L}(L^{\prime}) has Alexander degree

    A(F)=_i<j[12(P_ij0P_ij1)(𝐞_i𝐞_j)+12(N_ij0N_ij1)(𝐞_i+𝐞_j)]A(F)=\sum_{\_}{i<j}\left[\frac{1}{2}(P_{\_}{ij}^{0}-P_{\_}{ij}^{-1})(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j)+\frac{1}{2}(N_{\_}{ij}^{0}-N_{\_}{ij}^{1})(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j)\right]

    and homological degree

    gr_𝐰(F)=2_i<jN_ij1.\textup{gr}_{\_}{{\bf{w}}}(F)=-2\sum_{\_}{i<j}N_{\_}{ij}^{1}.
  • (b)

    For any 𝐤_L\mathbf{k}\in\mathbb{H}_{\_}L, the map F:(L,𝐤)(L,𝐤+A(F))F:\mathcal{H\!F\!L}(L,\mathbf{k})\to\mathcal{H\!F\!L}(L^{\prime},\mathbf{k}+A(F)), restricted to the 𝔽[𝐔]\mathbb{F}[\mathbf{U}]-free summand, is injective and determined by its homological degree.

  • (c)

    The hh-functions of LL and LL^{\prime} are related by the inequality

    h_L(𝐤)+_i<jN_ij1h_L(𝐤+A(F)).h_{\_}L(\mathbf{k})+\sum_{\_}{i<j}N_{\_}{ij}^{1}\geq h_{\_}{L^{\prime}}(\mathbf{k}+A(F)).
Proof.

Part (a) is clear from the degrees of maps ψ_1,ψ_0,ϕ_0,ϕ_1\psi_{\_}{-1},\psi_{\_}0,\phi_{\_}0,\phi_{\_}1 computed in Propositions 3.1 and 3.5.

Part (b) follows from Proposition 3.9. Note that FF is a composition of crossing change maps. For each crossing change map ϕ_0,ϕ_1,ψ_0,ψ_1\phi_{\_}0,\phi_{\_}1,\psi_{\_}0,\psi_{\_}{-1} associated to the link components L_i,L_jL_{\_}i,L_{\_}j, by Proposition 3.9, one can choose a corresponding crossing change map such that the composition of these two crossing change maps is one of the monomials U_i,U_j,V_i,V_jU_{\_}i,U_{\_}j,V_{\_}i,V_{\_}j. Hence, we can compose the map FF with another map F:(L)(L)F^{\prime}:\mathcal{H\!F\!L}(L^{\prime})\to\mathcal{H\!F\!L}(L) such that the composition FFF^{\prime}\circ F is given by a monomial in variables U_1,,U_n,V_1,,V_nU_{\_}1,\cdots,U_{\_}n,V_{\_}1,\cdots,V_{\_}n. Therefore, the map F:(L,𝐤)(,𝐤+A(F))F:\mathcal{H\!F\!L}(L,\mathbf{k})\rightarrow\mathcal{H\!F\!L}(\mathcal{L}^{\prime},\mathbf{k}+A(F)) is injective when restricted to the 𝔽[𝐔]\mathbb{F}[\mathbf{U}]-free part of (L,𝐤)\mathcal{H\!F\!L}(L,\mathbf{k}).

Now part (c) follows from (b): indeed, FF sends 𝔽[𝐔][2h_L(𝐤)]\mathbb{F}[\mathbf{U}][-2h_{\_}L(\mathbf{k})] to

𝔽[𝐔][2h_L(𝐤)2_i<jN_ij1]\mathbb{F}[\mathbf{U}]\left[-2h_{\_}L(\mathbf{k})-2\sum_{\_}{i<j}N_{\_}{ij}^{1}\right]

which should inject into 𝔽[𝐔][2h_L(𝐤+A(F))].\mathbb{F}[\mathbf{U}][-2h_{\_}{L^{\prime}}(\mathbf{k}+A(F))]. Hence,

2h_L(𝐤)2_i<jN_ij12h_L(𝐤+A(F))-2h_{\_}L(\mathbf{k})-2\sum_{\_}{i<j}N_{\_}{ij}^{1}\leq-2h_{\_}{L^{\prime}}(\mathbf{k}+A(F))

which yields the desired inequality. ∎

If LL and LL^{\prime} differ by a single positive crossing change, that is, LL^{\prime} is obtained by changing a negative crossing of LL into a positive one, we can recover the comparison between h_Lh_{\_}L and h_Lh_{\_}L^{\prime} in item (b) of Theorem 6.20 in [6].

Corollary 5.2.

Let LL be an nn-component link in S3S^{3}. Suppose LL^{\prime} is obtained by changing a negative crossing between the components L_iL_{\_}i and L_jL_{\_}j of LL into a positive one. Then

max{h_L(𝐤+12(𝐞_i+𝐞_j)),h_L(𝐤12(𝐞_i+𝐞_j))1}h_L(𝐤)min{h_L(𝐤±12(𝐞_i𝐞_j))}\max\{h_{\_}{L^{\prime}}(\mathbf{k}+\dfrac{1}{2}(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j)),h_{\_}{L^{\prime}}(\mathbf{k}-\dfrac{1}{2}(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j))-1\}\leq h_{\_}{L}(\mathbf{k})\leq\min\{h_{\_}{L^{\prime}}(\mathbf{k}\pm\dfrac{1}{2}(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j))\}

for all 𝐤_L\mathbf{k}\in\mathbb{H}_{\_}L.

Proof.

Note that LL^{\prime} is obtained from LL by changing a negative crossing to a positive crossing, which induces a cobordism map ϕ_0\phi_{\_}{0} or ϕ_1\phi_{\_}{1}. So we can make N0_ij=1N^{0}_{\_}{ij}=1 or N1_ij=1N^{1}_{\_}{ij}=1. By Lemma 5.1, if N0_ij=1N^{0}_{\_}{ij}=1 then

h_L(𝐤)h_L(𝐤+12(𝐞_i+𝐞_j)).h_{\_}{L}(\mathbf{k})\geq h_{\_}{L^{\prime}}(\mathbf{k}+\dfrac{1}{2}(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j)).

If N1_ij=1N^{1}_{\_}{ij}=1 then

h_L(𝐤)+1h_L(𝐤12(𝐞_i+𝐞_j)).h_{\_}{L}(\mathbf{k})+1\geq h_{\_}{L^{\prime}}(\mathbf{k}-\dfrac{1}{2}(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j)).

Conversely, LL can be obtained from LL^{\prime} by changing a positive crossing to a negative crossing. Then we can make P0_ij=1P^{0}_{\_}{ij}=1 or P1_ij=1P^{-1}_{\_}{ij}=1. By Lemma 5.1 again, if P0_ij=1P^{0}_{\_}{ij}=1, then

h_L(𝐤)h_L(𝐤+12(𝐞_i𝐞_j)).h_{\_}{L^{\prime}}(\mathbf{k})\geq h_{\_}{L}(\mathbf{k}+\dfrac{1}{2}(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j)).

If P1_ij=1P^{-1}_{\_}{ij}=1, then

h_L(𝐤)h_L(𝐤12(𝐞_i𝐞_j)).h_{\_}{L^{\prime}}(\mathbf{k})\geq h_{\_}{L}(\mathbf{k}-\dfrac{1}{2}(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j)).

Note that [6, Theorem 6.20] uses the JJ-function, which is some normalizations of the hh-function in the following formula:

J_L(𝐦)=h_L(𝐦+12(_1,,_n))J_{\_}{L}(\mathbf{m})=h_{\_}{L}(\mathbf{m}+\frac{1}{2}(\ell_{\_}1,\cdots,\ell_{\_}n))

for 𝐦n\mathbf{m}\in\mathbb{Z}^{n}. Let _L=12(_1,,_n)\mathbf{\ell}_{\_}{L}=\frac{1}{2}(\ell_{\_}1,\cdots,\ell_{\_}n). Since LL^{\prime} is obtained from LL by changing a negative crossing to a positive crossing between L_iL_{\_}i and L_jL_{\_}j. Then _L=_L+12(𝐞_i+𝐞_j)\mathbf{\ell}_{\_}{L^{\prime}}=\ell_{\_}L+\dfrac{1}{2}(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j). Then the inequalities in Corollary 5.2 become

max{J_L(𝐦),J_L(𝐦𝐞_i𝐞_j)1}J_L(𝐦)min{J_L(𝐦𝐞_i),J_L(𝐦𝐞_j)}.\max\{J_{\_}{L^{\prime}}(\mathbf{m}),J_{\_}{L^{\prime}}(\mathbf{m}-\mathbf{e}_{\_}i-\mathbf{e}_{\_}j)-1\}\leq J_{\_}{L}(\mathbf{m})\leq\min\{J_{\_}{L^{\prime}}(\mathbf{m}-\mathbf{e}_{\_}i),J_{\_}{L^{\prime}}(\mathbf{m}-\mathbf{e}_{\_}j)\}.

So we recover Theorem 6.20 in [6]. Moreover, we have an extra inequality that J_L(𝐦𝐞_i𝐞_j)1J_L(𝐦)J_{\_}{L^{\prime}}(\mathbf{m}-\mathbf{e}_{\_}i-\mathbf{e}_{\_}j)-1\leq J_{\_}{L}(\mathbf{m}). Note that h_L(𝐤)=h_L(𝐤+𝐞_i)h_{\_}{L}(\mathbf{k})=h_{\_}{L}(\mathbf{k}+\mathbf{e}_{\_}i) or h_L(𝐤)=h_L(𝐤+𝐞_i)+1h_{\_}{L}(\mathbf{k})=h_{\_}{L}(\mathbf{k}+\mathbf{e}_{\_}i)+1 for all links LL and all ii. So the above inequalities imply that J_L(𝐦)=J_L(𝐦)J_{\_}{L}(\mathbf{m})=J_{\_}{L^{\prime}}(\mathbf{m}) or J_L(𝐦)=J_L(𝐦)+1J_{\_}{L}(\mathbf{m})=J_{\_}{L^{\prime}}(\mathbf{m})+1.

5.2. Positive links

Suppose that all crossings between different components of LL are positive (in particular, this holds if LL is a positive link). Then there are exactly 2lk(L_i,L_j)2\mathrm{lk}(L_{\_}i,L_{\_}j) crossings between L_iL_{\_}i and L_jL_{\_}j, and we need to change lk(L_i,L_j)\mathrm{lk}(L_{\_}i,L_{\_}j) of them to split the components L_iL_{\_}i and L_jL_{\_}j. We can encode crossing change maps as above, with N_ij0=N_ij1=0N_{\_}{ij}^{0}=N_{\_}{ij}^{1}=0 and

P_ij1+P_ij0=lk(L_i,L_j)=_ij.P_{\_}{ij}^{-1}+P_{\_}{ij}^{0}=\mathrm{lk}(L_{\_}i,L_{\_}j)=\ell_{\_}{ij}.

Then Lemma 5.1 simplifies dramatically, and we get the following

Corollary 5.3.

Suppose that all crossings between different components of LL are positive. Let F:(L)(split(L))F:\mathcal{H\!F\!L}(L)\to\mathcal{H\!F\!L}(\mathrm{split}(L)) be a composition of P_ij1P_{\_}{ij}^{-1} maps of type ψ_1\psi_{\_}{-1} and P_ij0P_{\_}{ij}^{0} maps of type ψ_0\psi_{\_}0 between the components L_iL_{\_}i and L_jL_{\_}j for all i<ji<j. Define ε_ij=P_ij0P_ij1\varepsilon_{\_}{ij}=P_{\_}{ij}^{0}-P_{\_}{ij}^{-1}, then FF has Alexander degree

(11) A(F)=_i<j12ε_ij(𝐞_i𝐞_j)A(F)=\sum_{\_}{i<j}\frac{1}{2}\varepsilon_{\_}{ij}(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j)

and homological degree zero.

We can visualize the degrees of such maps as follows.

Definition 5.4.

Suppose that LL is a link such that all crossings between different components are positive, and _ij=lk(L_i,L_j)0\ell_{\_}{ij}=\mathrm{lk}(L_{\_}i,L_{\_}j)\geq 0. We define the link zonotope P_LP_{\_}L as the Minkowski sum of the intervals [_ij𝐞_i,_ij𝐞_j][\ell_{\_}{ij}\mathbf{e}_{\_}i,\ell_{\_}{ij}\mathbf{e}_{\_}j] for all i<ji<j.

Note that P_LP_{\_}L is an (n1)(n-1)-dimensional polytope contained in the hyperplane

{_i=1nx_i=_i<j_ij}r.\left\{\sum_{\_}{i=1}^{n}x_{\_}i=\sum_{\_}{i<j}\ell_{\_}{ij}\right\}\subset\mathbb{R}^{r}.

It is centrally symmetric around the point 12(_1,,_n)\frac{1}{2}(\ell_{\_}1,\ldots,\ell_{\_}n) where _i=_ji_ij\ell_{\_}i=\sum_{\_}{j\neq i}\ell_{\_}{ij}.

Example 5.5.

For n=2n=2 the polytope P_LP_{\_}L is a segment [_12𝐞_1,_12𝐞_2][\ell_{\_}{12}\mathbf{e}_{\_}1,\ell_{\_}{12}\mathbf{e}_{\_}2] with _12+1\ell_{\_}{12}+1 integer points on it.

Example 5.6.

For n=3n=3, the polytope P_LP_{\_}L is a hexagon where the opposite sides are parallel to each other and both contain _ij+1\ell_{\_}{ij}+1 integer points. The vertices of P_LP_{\_}L are:

(0,_12,_13+_23),(_12,0,_13+_23),(_12+_13,0,_23),(0,\ell_{\_}{12},\ell_{\_}{13}+\ell_{\_}{23}),(\ell_{\_}{12},0,\ell_{\_}{13}+\ell_{\_}{23}),(\ell_{\_}{12}+\ell_{\_}{13},0,\ell_{\_}{23}),
(_12+_13,_23,0),(_13,_12+_23,0),(0,_12+_23,_13).(\ell_{\_}{12}+\ell_{\_}{13},\ell_{\_}{23},0),(\ell_{\_}{13},\ell_{\_}{12}+\ell_{\_}{23},0),(0,\ell_{\_}{12}+\ell_{\_}{23},\ell_{\_}{13}).
\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet𝐞_1𝐞_2\mathbf{e}_{\_}1-\mathbf{e}_{\_}2𝐞_2𝐞_3\mathbf{e}_{\_}2-\mathbf{e}_{\_}3𝐞_1𝐞_3\mathbf{e}_{\_}1-\mathbf{e}_{\_}3(0,2,7)(2,0,7)(6,0,3)(6,3,0)(4,5,0)(0,5,4)\bullet\bullet\bullet\bullet\bullet\bullet\bullet(0,1,2)(1,0,2)(2,0,1)(2,1,0)(1,2,0)(0,2,1)
Figure 14. Polytopes for 3-component positive links with (_12,_23,_13)=(2,3,4)(\ell_{\_}{12},\ell_{\_}{23},\ell_{\_}{13})=(2,3,4) (left) and (1,1,1)(1,1,1) (right). These are Minkowski sums of intervals [(2,0,0),(0,2,0)]+[(0,3,0),(0,0,3)]+[(4,0,0),(0,0,4)][(2,0,0),(0,2,0)]+[(0,3,0),(0,0,3)]+[(4,0,0),(0,0,4)] and [(1,0,0),(0,1,0)]+[(0,1,0),(0,0,1)]+[(1,0,0),(0,0,1)][(1,0,0),(0,1,0)]+[(0,1,0),(0,0,1)]+[(1,0,0),(0,0,1)], respectively.
Theorem 5.7.

After a shift by the vector 12(_1,,_n)\frac{1}{2}(\ell_{\_}1,\ldots,\ell_{\_}n), the Alexander degrees of splitting maps (11) of a positive link correspond to the integer points in the polytope P_LP_{\_}L. Conversely, any such integer point corresponds to at least one nontrivial splitting map.

Proof.

By Lemma 5.1 all compositions of splitting maps are nontrivial, and for positive links all such maps have homological degree zero, but we need to understand their Alexander degrees.

By varying P_ij0P_{\_}{ij}^{0} and P_ij1P_{\_}{ij}^{-1}, the terms 12ε_ij(𝐞_i𝐞_j)\frac{1}{2}\varepsilon_{\_}{ij}(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j) can have the values

12_ij(𝐞_i𝐞_j),12(_ij2)(𝐞_i𝐞_j),,12_ij(𝐞_i𝐞_j).\frac{1}{2}\ell_{\_}{ij}(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j),\frac{1}{2}(\ell_{\_}{ij}-2)(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j),\ldots,-\frac{1}{2}\ell_{\_}{ij}(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j).

By shifting these by 12_ij(𝐞_i+𝐞_j)\frac{1}{2}\ell_{\_}{ij}(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j), we get the points

_ij𝐞_i,(_ij1)𝐞_i+𝐞_j,,_ij𝐞_j\ell_{\_}{ij}\mathbf{e}_{\_}i,\ (\ell_{\_}{ij}-1)\mathbf{e}_{\_}i+\mathbf{e}_{\_}j,\ldots,\ell_{\_}{ij}\mathbf{e}_{\_}j

which coincide with the set of integer points on the interval [_ij𝐞_i,_ij𝐞_j][\ell_{\_}{ij}\mathbf{e}_{\_}i,\ell_{\_}{ij}\mathbf{e}_{\_}j]. By adding these degrees over all i<ji<j, we obtain an integer point in P_LP_{\_}L, and the overall shift equals

_i<j12_ij(𝐞_i+𝐞_j)=12(_1,,_n).\sum_{\_}{i<j}\frac{1}{2}\ell_{\_}{ij}(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j)=\frac{1}{2}(\ell_{\_}1,\ldots,\ell_{\_}n).

It remains to prove that any integer point in P_LP_{\_}L can be obtained as a sum of integer points in the intervals [_ij𝐞_i,_ij𝐞_j][\ell_{\_}{ij}\mathbf{e}_{\_}i,\ell_{\_}{ij}\mathbf{e}_{\_}j]. This follows from the combinatorial results in [4, Section 9]. Indeed, by [4, Lemma 9.1] the polytope P_LP_{\_}L can be decomposed into a disjoint union of parallelepipeds of various dimensions labeled by the linearly independent subsets of the set {𝐞_i𝐞_j:i<j}\{\mathbf{e}_{\_}i-\mathbf{e}_{\_}j\ :\ i<j\} (the edges of each parallelepiped have integer length _ij\ell_{\_}{ij}), see Figure 15. These parallelepipeds can be themselves decomposed into smaller parallelepipeds with edges of integer length 1. It is easy to see [4, Lemma 9.6] that the linearly independent subsets correspond to forests on nn vertices, and by [4, Theorem 9.5] the relative volume of each small parallelepiped equals 1. Therefore every vector connecting an integer point inside each large parallelepiped with a vertex can be written as a linear combination of vectors along the edges and the result follows. ∎

\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet𝐞_1𝐞_2\mathbf{e}_{\_}1-\mathbf{e}_{\_}2𝐞_2𝐞_3\mathbf{e}_{\_}2-\mathbf{e}_{\_}3𝐞_1𝐞_3\mathbf{e}_{\_}1-\mathbf{e}_{\_}3(0,2,7)(2,0,7)(6,0,3)(6,3,0)(4,5,0)(0,5,4)
Figure 15. A decomposition of P_LP_{\_}L into parallelepipeds.
Remark 5.8.

In principle, there could be several splitting maps of the same degree. For example, for (_12,_23,_13)=(1,1,1)(\ell_{\_}{12},\ell_{\_}{23},\ell_{\_}{13})=(1,1,1) we can write the point (1,1,1)(1,1,1) at the center as a sum of integer points on intervals in two ways:

(1,1,1)=(1,0,0)+(0,1,0)+(0,0,1)=(0,1,0)+(0,0,1)+(1,0,0).(1,1,1)=(1,0,0)+(0,1,0)+(0,0,1)=(0,1,0)+(0,0,1)+(1,0,0).

5.3. L-space links

If LL is an L-space link, then some results simplify significantly.

Theorem 5.9.

Suppose LL is an L-space link. Then:

  • (a)

    For any choice of crossing changes and the maps ψ_k,ϕ_k\psi_{\_}k,\phi_{\_}k at the crossings, the resulting map F:(L)(split(L))F:\mathcal{H\!F\!L}(L)\to\mathcal{H\!F\!L}(\mathrm{split}(L)) is completely determined by its Alexander and Maslov degrees.

  • (b)

    If, in addition, all crossings between the different components of LL are positive, the splitting maps all have homological degree zero and are in bijection with the integer points in the polytope P_LP_{\_}L. Two splitting maps of the same Alexander degree coincide.

Proof.

If LL is an L-space link then by [18] all its components L_iL_{\_}i are L-space knots. In particular, the corresponding split link split(L)\mathrm{split}(L) is an LL-space link as well.

By Corollary 2.4 this means that (L,𝐤)\mathcal{H\!F\!L}(L,\mathbf{k}) and (split(L),𝐤+A(F))\mathcal{H\!F\!L}(\mathrm{split}(L),\mathbf{k}+A(F)) are isomorphic to 𝔽[𝐔]\mathbb{F}[\mathbf{U}], and there is a unique nonzero map between two copies of 𝔽[𝐔]\mathbb{F}[\mathbf{U}] of a given degree. The splitting map is nonzero by Lemma 5.1, so this implies (a).

Part (b) follows from (a) and Theorem 5.7. ∎

5.4. Torsion estimates

Recall that the splitting number sp(L)sp(L) is the minimal number of crossings between different components of a link LL that should be changed to turn LL into the split link. Let n_ijn_{\_}{ij} be the number of crossing changes between the ii-th and jj-th components, so that the resulting link is the split link.

Consider (2n)(2n)-dimensional lattice 2n\mathbb{Z}^{2n} with basis 𝐞_1,𝐟_1,𝐞_2,𝐟_2,,𝐞_n,𝐟_n\mathbf{e}_{\_}1,\mathbf{f}_{\_}1,\mathbf{e}_{\_}2,\mathbf{f}_{\_}2,\ldots,\mathbf{e}_{\_}n,\mathbf{f}_{\_}n. A point in this lattice parametrizes a monomial in U_1,V_1,U_2,V_2,,U_n,V_nU_{\_}1,V_{\_}1,U_{\_}2,V_{\_}2,\ldots,U_{\_}n,V_{\_}n. For each pair i<ji<j we consider the 3-dimensional tetrahedron:

T_ij:={a𝐞_i+b𝐟_i+c𝐞_j+d𝐟_j:a+b+c+d=n_ij,a,b,c,d0}.T_{\_}{ij}:=\left\{a\mathbf{e}_{\_}i+b\mathbf{f}_{\_}i+c\mathbf{e}_{\_}j+d\mathbf{f}_{\_}j\ :\ a+b+c+d=n_{\_}{ij},\ a,b,c,d\geq 0\right\}.

Next, we consider the Minkowski sum

T=_i<jT_ij2n.T=\sum_{\_}{i<j}T_{\_}{ij}\subset\mathbb{Z}^{2n}.
Remark 5.10.

Note that the intersection of TT with the nn-dimensional sublattice span{𝐞_1,𝐞_2,,𝐞_n}\mathrm{span}\{\mathbf{e}_{\_}1,\mathbf{e}_{\_}2,\cdots,\mathbf{e}_{\_}n\} is the Minkowski sum of segments [n_ij𝐞_i,n_ij𝐞_j][n_{\_}{ij}\mathbf{e}_{\_}i,n_{\_}{ij}\mathbf{e}_{\_}j] which is similar to the generalized permutahedra P_LP_{\_}L above.

Theorem 5.11.

Suppose that LL is an nn-component link where the components are L-space knots. Suppose that n_ijn_{\_}{ij} is the number of crossing changes between the ii-th and jj-th components so that the resulting link is a split link. Then:

  • (a)

    For any integer point in the polytope T2nT\subset\mathbb{Z}^{2n} the corresponding monomial in U_i,V_iU_{\_}i,V_{\_}i annihilates any torsion element in (L)\mathcal{H\!F\!L}(L).

  • (b)

    The monomial 𝐔_i<jn_ij2\mathbf{U}^{\sum_{\_}{i<j}\left\lceil\frac{n_{\_}{ij}}{2}\right\rceil} annihilates any torsion element in (L)\mathcal{H\!F\!L}(L).

Proof.

(a) Given an integer point in TT, we construct two maps

F:(L)(split(L)),F:(split(L))(L)F:\mathcal{H\!F\!L}(L)\rightarrow\mathcal{H\!F\!L}(\mathrm{split}(L)),\ F^{\prime}:\mathcal{H\!F\!L}(\mathrm{split}(L))\rightarrow\mathcal{H\!F\!L}(L)

as follows. Each time one changes a positive crossing to a negative crossing between L_iL_{\_}i and L_jL_{\_}j, we choose either ψ_0\psi_{\_}{0} or ψ_1\psi_{\_}{-1} for FF and either ϕ_0\phi_{\_}{0} or ϕ_1\phi_{\_}{1} for FF^{\prime}; if one changes a negative crossing to a positive, the roles of FF and FF^{\prime} are switched. Specifically, given nonnegative integers a_ij,b_ij,c_ij,d_ija_{\_}{ij},b_{\_}{ij},c_{\_}{ij},d_{\_}{ij} such that a_ij+b_ij+c_ij+d_ij=n_ija_{\_}{ij}+b_{\_}{ij}+c_{\_}{ij}+d_{\_}{ij}=n_{\_}{ij} (or an integer point in the tetrahedron T_ijT_{\_}{ij}) we can arrange the maps so that

  • To a_ija_{\_}{ij} crossings between L_i,L_jL_{\_}i,L_{\_}j we associate ψ_1,ϕ_1\psi_{\_}{-1},\phi_{\_}1

  • To b_ijb_{\_}{ij} crossings between L_i,L_jL_{\_}i,L_{\_}j we associate ψ_0,ϕ_0\psi_{\_}{0},\phi_{\_}0

  • To c_ijc_{\_}{ij} crossings between L_i,L_jL_{\_}i,L_{\_}j we associate ψ_0,ϕ_1\psi_{\_}{0},\phi_{\_}1

  • To d_ijd_{\_}{ij} crossings between L_i,L_jL_{\_}i,L_{\_}j we associate ψ_1,ϕ_0\psi_{\_}{-1},\phi_{\_}0

By composing those crossing change maps, we obtain FF and FF^{\prime}. By Proposition 3.9 we get

FF=_i<j(U_ia_ijV_ib_ijU_jc_ijV_jd_ij).F^{\prime}\circ F=\prod_{\_}{i<j}\left(U_{\_}i^{a_{\_}{ij}}V_{\_}i^{b_{\_}{ij}}U_{\_}j^{c_{\_}{ij}}V_{\_}{j}^{d_{\_}{ij}}\right).

The right hand side defines a monomial corresponding to a point in _i<jT_ij=T\sum_{\_}{i<j}T_{\_}{ij}=T. Note that the composition FFF^{\prime}\circ F does not depend on the order of the crossings, just the number of crossings of each type.

If xx is a torsion element in (L)\mathcal{H\!F\!L}(L), then F(x)=0F(x)=0 since there are no torsion elements in (split(L))\mathcal{H\!F\!L}(\mathrm{split}(L)), therefore

_i<jU_ia_ijV_ib_ijU_jc_ijV_jd_ijx=0.\prod_{\_}{i<j}U_{\_}i^{a_{\_}{ij}}V_{\_}i^{b_{\_}{ij}}U_{\_}j^{c_{\_}{ij}}V_{\_}{j}^{d_{\_}{ij}}\cdot x=0.

(b) We can choose a_ij=b_ij=n_ij2,c_ij=d_ij=0a_{\_}{ij}=b_{\_}{ij}=\left\lceil\frac{n_{\_}{ij}}{2}\right\rceil,c_{\_}{ij}=d_{\_}{ij}=0, then a_ij+b_ijn_ija_{\_}{ij}+b_{\_}{ij}\geq n_{\_}{ij} and

(U_ia_ijV_ib_ijU_jc_ijV_jd_ij)=𝐔n_ij2.\left(U_{\_}i^{a_{\_}{ij}}V_{\_}i^{b_{\_}{ij}}U_{\_}j^{c_{\_}{ij}}V_{\_}{j}^{d_{\_}{ij}}\right)=\mathbf{U}^{\left\lceil\frac{n_{\_}{ij}}{2}\right\rceil}.

By part (a), the monomial

_i<j(U_ia_ijV_ib_ijU_jc_ijV_jd_ij)=𝐔_i<jn_ij2\prod_{\_}{i<j}\left(U_{\_}i^{a_{\_}{ij}}V_{\_}i^{b_{\_}{ij}}U_{\_}j^{c_{\_}{ij}}V_{\_}{j}^{d_{\_}{ij}}\right)=\mathbf{U}^{\sum_{\_}{i<j}\left\lceil\frac{n_{\_}{ij}}{2}\right\rceil}

annihilates any torsion element in (L)\mathcal{H\!F\!L}(L).

Corollary 5.12.

Suppose that L=L_1L_2L=L_{\_}1\cup L_{\_}2 is a 2-component link with splitting number sp(L)=nsp(L)=n, and L_1,L_2L_{\_}1,L_{\_}2 are L-space knots. Then for any a,b,c,d0a,b,c,d\geq 0 such that a+b+c+d=na+b+c+d=n the monomial U_1aV_1bU_2cV_2dU_{\_}1^{a}V_{\_}1^{b}U_{\_}2^{c}V_{\_}2^{d} annihilates any torsion element in (L)\mathcal{H\!F\!L}(L).

Example 5.13.

We consider the following boundary links. Given any knot KK as in Figure 16, let B(K,n)B(K,n) be the new 2-component link obtained by applying an nn-twisted Bing doubling to KK. Observe that B(K,n)B(K,n) is a boundary link with unknotted components. The linking number of B(K,n)B(K,n) is 0, consisting of a positive crossing and a negative crossing between the link components. So one has to change both of the crossings to split the link, and the splitting number is 2 automatically. Hence, by Corollary 5.12 any monomial U_1aV_1bU_2cV_2dU_{\_}1^{a}V_{\_}1^{b}U_{\_}2^{c}V_{\_}2^{d} where a+b+c+d=2a+b+c+d=2 annihilates any torsion element of (B(K,n))\mathcal{H\!F\!L}(B(K,n)) for all KK and nn. In particular, the U_iU_{\_}i-torsion and V_iV_{\_}i-torsion are bounded by 22 for all KK and nn where i=1,2i=1,2.

Refer to captionKKK,nK,n
Figure 16. The left figure is KK and right figure is the nn-twisted Bing doubling of KK, B(K,n)B(K,n).

6. Example: T(n,n)T(n,n)

We illustrate all of the above constructions in detail for the torus link T(n,n)T(n,n) with nn components. By [13] it is an L-space link and its Heegaard Floer homology is given by Theorem 3.11.

6.1. Integer points in permutahedra and splitting maps

The permutahedron P_nP_{\_}n is the convex hull of points (σ(n)1,,σ(1)1)(\sigma(n)-1,\ldots,\sigma(1)-1) for all permutations σS_n\sigma\in S_{\_}n. It is a convex polytope of dimension n1n-1, and it is easy to see that the center of P_nP_{\_}n is at the point (n12,,n12)\left(\frac{n-1}{2},\ldots,\frac{n-1}{2}\right). By [4, Theorem 9.4] the permutahedron P_nP_{\_}n is the Minkowski sum of segments [𝐞_i,𝐞_j][\mathbf{e}_{\_}i,\mathbf{e}_{\_}j] and hence agrees with the link zonotope P_T(n,n)P_{\_}{T(n,n)}.

Following Theorem 5.7 we will be interested in the integer points in P_nP_{\_}n. For example, P_2P_{\_}2 is a segment connecting (1,0)(1,0) and (0,1)(0,1). Next, P_3P_{\_}3 is a hexagon with vertices obtained by permutations of (2,1,0)(2,1,0) which contains one more integer point (1,1,1)(1,1,1) at its center, see Figure 14 (right).

For n=4n=4 we have a 3-dimensional polytope with 24 vertices corresponding to permutations of (3,2,1,0)(3,2,1,0). Additionally, it contains 6 permutations of

(2,2,1,1)=12(3,1,2,0)+12(1,3,0,2),(2,2,1,1)=\frac{1}{2}(3,1,2,0)+\frac{1}{2}(1,3,0,2),

and 4 permutations of both

(2,2,2,0)=12(3,2,1,0)+12(1,2,3,0),(3,1,1,1)=12(3,2,1,0)+12(3,0,1,2).(2,2,2,0)=\frac{1}{2}(3,2,1,0)+\frac{1}{2}(1,2,3,0),\quad(3,1,1,1)=\frac{1}{2}(3,2,1,0)+\frac{1}{2}(3,0,1,2).

In total we get 38=24+6+4+438=24+6+4+4 integer points in P_4P_{\_}4. In general, it is known that the integer points in P_nP_{\_}n correspond to forests on nn labeled vertices, but we will not need this. We refer to [4, Chapter 9] for more information on permutahedra and integer points in them.

For all i<ji<j we define the vector111The reader might recognize the root system of type A_n1A_{\_}{n-1}.:

α_ij=(0,,0,12,0,0,12,0,,0)=12(𝐞_i𝐞_j).\alpha_{\_}{ij}=\left(0,\ldots,0,\frac{1}{2},0\ldots,0,-\frac{1}{2},0,\ldots,0\right)=\frac{1}{2}(\mathbf{e}_{\_}i-\mathbf{e}_{\_}j).
Lemma 6.1.

For any integer n>0n>0,

  • (a)

    We have

    (n12,,n12)+_i<jα_ij=(n1,n2,,1,0).\left(\frac{n-1}{2},\ldots,\frac{n-1}{2}\right)+\sum_{\_}{i<j}\alpha_{\_}{ij}=(n-1,n-2,\ldots,1,0).
  • (b)

    For any choice of signs ε=(ε_ij)_i<j{±1}(n2)\varepsilon=(\varepsilon_{\_}{ij})_{\_}{i<j}\in\{\pm 1\}^{\binom{n}{2}} define

    p_ε=(n12,,n12)+_i<jε_ijα_ijp_{\_}{\varepsilon}=\left(\frac{n-1}{2},\ldots,\frac{n-1}{2}\right)+\sum_{\_}{i<j}\varepsilon_{\_}{ij}\alpha_{\_}{ij}

    Then p_εp_{\_}{\varepsilon} is an integer point in the permutahedron P_nP_{\_}n and all integer points in P_nP_{\_}n can be obtained this way.

Proof.

Part (a) is clear. Part (b) follows from the description of P_nP_{\_}n as a zonotope, that is, Minkowski sum of segments

[𝐞_i,𝐞_j]=12(𝐞_i+𝐞_j)+[α_ij,α_ij],[\mathbf{e}_{\_}i,\mathbf{e}_{\_}j]=\frac{1}{2}(\mathbf{e}_{\_}i+\mathbf{e}_{\_}j)+[\alpha_{\_}{ij},-\alpha_{\_}{ij}],

see [4, Theorems 9.4, 9.5].

Recall that by Theorem 3.11 (T(n,n))\mathcal{H\!F\!L}(T(n,n)) has nn generators a_0,,a_n1a_{\_}0,\ldots,a_{\_}{n-1}.

Theorem 6.2.

Let ε=(ε_ij)_i<j\varepsilon=(\varepsilon_{\_}{ij})_{\_}{i<j} be a choice of signs as above. Choose a minimal sequence of crossings changes that splits T(n,n)T(n,n). For any 1i<jn1\leq i<j\leq n, this sequence contains exactly one crossing change between L_iL_{\_}i and L_jL_{\_}j. Consider the local crossing change map

Ψ_ε_ij={ψ_0ifε_ij=1ψ_1ifε_ij=1.\Psi_{\_}{\varepsilon_{\_}{ij}}=\begin{cases}\psi_{\_}{0}\ &\text{if}\ \varepsilon_{\_}{ij}=1\\ \psi_{\_}{-1}\ &\text{if}\ \varepsilon_{\_}{ij}=-1.\end{cases}

and compose them to define the splitting map Ω_ε:(T(n,n))(O_n)\Omega_{\_}{\varepsilon}:\mathcal{H\!F\!L}(T(n,n))\to\mathcal{H\!F\!L}(O_{\_}n). Then Ω_ε\Omega_{\_}{\varepsilon} is independent of the crossing change sequence and satisfies the following properties:

  • (a)

    The Alexander multi-degree of Ω_ε\Omega_{\_}{\varepsilon} equals _i<jε_ijα_ij\sum_{\_}{i<j}\varepsilon_{\_}{ij}\alpha_{\_}{ij} and the homological degree gr_𝐰(Ω_ε)\textup{gr}_{\_}{{\bf{w}}}(\Omega_{\_}{\varepsilon}) is zero.

  • (b)

    Ω_ε\Omega_{\_}{\varepsilon} sends the generator a_0(T(n,n))a_{\_}0\in\mathcal{H\!F\!L}(T(n,n)) to Vp_εV^{p_{\_}{\varepsilon}} and every other generator a_ka_{\_}k to some other monomials in R_UVR_{\_}{UV}.

  • (c)

    If p_ε=p_εp_{\_}{\varepsilon}=p_{\_}{\varepsilon^{\prime}} then Ω_ε=Ω_ε\Omega_{\_}{\varepsilon}=\Omega_{\_}{\varepsilon^{\prime}}. In other words, the splitting maps for T(n,n)T(n,n) are parametrized by the integer points in the permutahedron P_nP_{\_}n.

  • (d)

    For ε=(+1,,+1)\varepsilon=(+1,\ldots,+1) the map Ω_𝟏=Ω_+1,,+1\Omega_{\_}{\bf 1}=\Omega_{\_}{+1,\ldots,+1} is defined on generators by the equation:

    Ω_𝟏(a_k)=V_1n1kV_2n2kV_n1kU_n+1kU_nk.\Omega_{\_}{\bf 1}(a_{\_}k)=V_{\_}1^{n-1-k}V_{\_}2^{n-2-k}\cdots V_{\_}{n-1-k}U_{\_}{n+1-k}\cdots U_{\_}n^{k}.
  • (e)

    For any permutation σS_n\sigma\in S_{\_}n there is a map Ω_σ\Omega_{\_}{\sigma} corresponding to a vertex of P_nP_{\_}n. It is obtained from Ω_𝟏\Omega_{\_}{\bf 1} by permuting the indices of U_iU_{\_}i and V_iV_{\_}i by σ\sigma:

    Ω_σ(a_k)=V_σ(1)n1kV_σ(2)n2kV_σ(n1k)U_σ(n+1k)U_σ(n)k.\Omega_{\_}{\sigma}(a_{\_}k)=V_{\_}{\sigma(1)}^{n-1-k}V_{\_}{\sigma(2)}^{n-2-k}\cdots V_{\_}{\sigma(n-1-k)}U_{\_}{\sigma(n+1-k)}\cdots U_{\_}{\sigma(n)}^{k}.
Proof.

Part (a) is clear and (c) follows from Theorem 5.9. Part (b) is immediate from (a) since a_0a_{\_}0 has Alexander degree (n12,,n12)\left(\frac{n-1}{2},\ldots,\frac{n-1}{2}\right) and gr_𝐰(a_0)=0\textup{gr}_{\_}{{\bf{w}}}(a_{\_}0)=0. This agrees with the normalization of the generator of (O_n)\mathcal{H\!F\!L}(O_{\_}n) and neither Ω_ε\Omega_{\_}{\varepsilon} nor V_iV_{\_}i change gr_𝐰\textup{gr}_{\_}{{\bf{w}}}.

Part (d) also follows from Theorem 5.9 since we can compare the Alexander and Maslov degrees on both sides. Indeed,

A(Ω_𝟏(a_k))=_i<jα_ij+A(a_k)=_i<jα_ij+(n12,,n12)(k,,k)=(n1k,n2k,,1k,k).\begin{split}A(\Omega_{\_}{\bf 1}(a_{\_}k))&=\sum_{\_}{i<j}\alpha_{\_}{ij}+A(a_{\_}k)=\sum_{\_}{i<j}\alpha_{\_}{ij}+\left(\frac{n-1}{2},\ldots,\frac{n-1}{2}\right)-(k,\ldots,k)\\ &=(n-1-k,n-2-k,\ldots,1-k,-k).\end{split}

Here the last equation follows from Lemma 6.1(a). Furthermore,

gr_𝐰(Ω_𝟏(a_k))=gr_𝐰(a_k)=k(k+1),\textup{gr}_{\_}{{\bf{w}}}(\Omega_{\_}{\bf 1}(a_{\_}k))=\textup{gr}_{\_}{{\bf{w}}}(a_{\_}k)=-k(k+1),

while

gr_𝐰(V_1n1kV_2n2kV_n1kU_n+1kU_nk)=2(1++k)=k(k+1).\textup{gr}_{\_}{{\bf{w}}}\left(V_{\_}1^{n-1-k}V_{\_}2^{n-2-k}\cdots V_{\_}{n-1-k}U_{\_}{n+1-k}\cdots U_{\_}n^{k}\right)=-2(1+\ldots+k)=-k(k+1).

Part (e) follows from (d) by permuting the variables. ∎

Example 6.3.

For n=2n=2 we have Ω_𝟏(a_0)=V_1,Ω_𝟏(a_1)=U_2\Omega_{\_}{\bf 1}(a_{\_}0)=V_{\_}1,\Omega_{\_}{\bf 1}(a_{\_}1)=U_{\_}2 as in Example 3.2.

Example 6.4.

For n=3n=3, the signs ε=(+1,+1,+1)\varepsilon=(+1,+1,+1) correspond to the point (2,1,0)P_3(2,1,0)\in P_{\_}3 and the map Ω_𝟏(a_0)=V_12V_2,Ω_𝟏(a_1)=V_1U_3,Ω_𝟏(a_2)=U_2U_32\Omega_{\_}{\bf 1}(a_{\_}0)=V_{\_}1^{2}V_{\_}2,\ \Omega_{\_}{\bf 1}(a_{\_}1)=V_{\_}1U_{\_}3,\ \Omega_{\_}{\bf 1}(a_{\_}2)=U_{\_}2U_{\_}3^{2}. The maps for other vertices of P_3P_{\_}3 can be obtained from it by permuting the variables.

The maps for (ε_12=+1,ε_13=1,ε_23=+1)(\varepsilon_{\_}{12}=+1,\varepsilon_{\_}{13}=-1,\varepsilon_{\_}{23}=+1) and (ε_12=1,ε_13=+1,ε_23=1)(\varepsilon_{\_}{12}=-1,\varepsilon_{\_}{13}=+1,\varepsilon_{\_}{23}=-1) both correspond to the central point (1,1,1)P_3(1,1,1)\in P_{\_}3 and the corresponding splitting map is given by following:

Ω_ε(a_0)=V_1V_2V_3,Ω_ε(a_1)=𝐔,Ω_ε(a_2)=U_1U_2U_3.\Omega_{\_}{\varepsilon}(a_{\_}0)=V_{\_}1V_{\_}2V_{\_}3,\ \Omega_{\_}{\varepsilon}(a_{\_}1)=\mathbf{U},\ \Omega_{\_}{\varepsilon}(a_{\_}2)=U_{\_}1U_{\_}2U_{\_}3.

6.2. Surgery map

In this section, we study the map Ω:𝓗𝓕𝓛(T(n,n))𝓗𝓕𝓛(O_n)\Omega:\bm{{\mathcal{H\!F\!L}}}(T(n,n))\to\bm{{\mathcal{H\!F\!L}}}(O_{\_}n) obtained by composing the surgery maps Ψ\Psi from Section 4 for any sequence of crossing changes. Specifically, we choose a minimal sequence of crossing changes that splits T(n,n)T(n,n) and we compose the local surgery maps Ψ\Psi associated to each crossing change to define Ω\Omega. Below we will show that (at least on the level of homology) the choice of crossing change sequence does not matter and resulting maps agree.

To specify the link components involved in a crossing change, we will use subscripts for Ψ\Psi i.e. for a positive crossing between L_iL_{\_}i and L_jL_{\_}j we denote the local crossing change map by Ψ_ij\Psi_{\_}{ij}. Topologically, each map Ψ_ij\Psi_{\_}{ij} corresponds to a cobordism W_ijW_{\_}{ij} obtained attaching a 2-handle, and their composition Ω\Omega corresponds to the composition WW of cobordisms W_ijW_{\_}{ij}. We have H_2(W_ij)=H_{\_}2(W_{\_}{ij})=\mathbb{Z} and H_2(W)=(n2)H_{\_}2(W)=\mathbb{Z}^{\binom{n}{2}}. A choice of a Spinc\text{Spin}^{c}-structure on W_ijW_{\_}{ij} corresponds to a choice of an integer m_ijm_{\_}{ij} and a map ψ_m_ij\psi_{\_}{m_{\_}{ij}} defined as in Section 3, so that Ψ_ij=_m_ij(1)m_ijψ_m_ij\Psi_{\_}{ij}=\sum_{\_}{m_{\_}{ij}\in\mathbb{Z}}(-1)^{m_{\_}{ij}}\psi_{\_}{m_{\_}{ij}}. Similarly, a choice of a Spinc\text{Spin}^{c}-structure on WW corresponds to a choice of a vector (m_ij)_i<j(m_{\_}{ij})_{\_}{i<j} in the (n2)\binom{n}{2}-dimensional lattice, and

Ω=_(m_ij)(n2)_i<j(1)m_ijψ_m_ij.\Omega=\sum_{\_}{(m_{\_}{ij})\in\mathbb{Z}^{\binom{n}{2}}}\prod_{\_}{i<j}(-1)^{m_{\_}{ij}}\psi_{\_}{m_{\_}{ij}}.

The choices of m_ij=0m_{\_}{ij}=0 and m_ij=1m_{\_}{ij}=-1 correspond, respectively, to choices of binary sequences ε_ij=1\varepsilon_{\_}{ij}=1 and ε_ij=1\varepsilon_{\_}{ij}=-1 in previous section. In Theorem 6.11 below we prove that Ω\Omega is injective on homology. To describe its image explicitly, we need to introduce some algebraic notations.

Definition 6.5.

Let S_02S\subset\mathbb{Z}_{\_}{\geq 0}^{2} be a subset of cardinality nn, order its elements as s_1=(a_1,b_1),,s_n=(a_n,b_n)s_{\_}1=(a_{\_}1,b_{\_}1),\ldots,s_{\_}n=(a_{\_}n,b_{\_}n). Then we can define the polynomial

Δ_S=_σS_n(1)σU_1a_σ(1)V_1b_σ(1)U_na_σ(n)V_nb_σ(n)=det(U_1a_1V_1b_1U_1a_nV_1b_nU_na_1V_nb_1U_na_nV_nb_n).\Delta_{\_}S=\sum_{\_}{\sigma\in S_{\_}n}(-1)^{\sigma}U_{\_}1^{a_{\_}{\sigma(1)}}V_{\_}1^{b_{\_}{\sigma(1)}}\cdots U_{\_}n^{a_{\_}{\sigma(n)}}V_{\_}n^{b_{\_}{\sigma(n)}}=\det\left(\begin{matrix}U_{\_}1^{a_{\_}1}V_{\_}1^{b_{\_}1}&\cdots&U_{\_}1^{a_{\_}n}V_{\_}1^{b_{\_}n}\\ \vdots&\ddots&\vdots\\ U_{\_}n^{a_{\_}1}V_{\_}n^{b_{\_}1}&\cdots&U_{\_}n^{a_{\_}n}V_{\_}n^{b_{\_}n}\\ \end{matrix}\right).

Reordering the elements of SS changes the sign of Δ_S\Delta_{\_}S. Sometimes we will use notation Δ_S\Delta_{\_}S for nn-tuples SS where some elements are repeated, in this case Δ_S=0\Delta_{\_}S=0.

Definition 6.6.

We define 𝒥(O_n)=R_UV\mathcal{J}\subset\mathcal{H\!F\!L}(O_{\_}n)=R_{\_}{UV} as the ideal generated by Δ_S\Delta_{\_}S for all possible nn-element subsets SS.

Remark 6.7.

The polynomials Δ_S\Delta_{\_}S and the ideal in [U_1,,U_n,V_1,,V_n]\mathbb{C}[U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n] generated by Δ_S\Delta_{\_}S were first introduced by Haiman in his work on Hilbert scheme of points on the plane [14].

The following lemma gives a useful characterization of the ideal 𝒥\mathcal{J}, we postpone its proof until Section 6.3. It can be used as an alternative definition of 𝒥\mathcal{J}.

Lemma 6.8.

The ideal 𝒥\mathcal{J} is generated by the nn maximal minors corresponding to nn-tuples of consecutive columns in the matrix

(U_1n1U_1n2U_11V_1V_1n2V_1n1U_nn1U_nn2U_n1V_nV_nn2V_nn1)\left(\begin{matrix}U_{\_}1^{n-1}&U_{\_}1^{n-2}&\cdots&U_{\_}1&1&V_{\_}1&\cdots&V_{\_}1^{n-2}&V_{\_}1^{n-1}\\ \vdots&\vdots&&\vdots&\vdots&\vdots&&\vdots&\vdots\\ U_{\_}n^{n-1}&U_{\_}n^{n-2}&\cdots&U_{\_}n&1&V_{\_}n&\cdots&V_{\_}n^{n-2}&V_{\_}n^{n-1}\\ \end{matrix}\right)

It is clear that all the minors in Lemma 6.8 are of the form Δ_S\Delta_{\_}S for some choices of subsets SS.

Example 6.9.

For n=2n=2 we have two determinants

det(U_11U_21)=U_1U_2,det(1V_11V_2)=V_2V_1.\det\left(\begin{matrix}U_{\_}1&1\\ U_{\_}2&1\end{matrix}\right)=U_{\_}1-U_{\_}2,\det\left(\begin{matrix}1&V_{\_}1\\ 1&V_{\_}2\end{matrix}\right)=V_{\_}2-V_{\_}1.
Example 6.10.

For n=3n=3 we have three determinants

det(U_12U_11U_22U_21U2_3U_31),det(U_11V_1U_21V_2U_31V_3),det(1V_1V_121V_2V_221V_3V_32).\det\left(\begin{matrix}U_{\_}1^{2}&U_{\_}1&1\\ U_{\_}2^{2}&U_{\_}2&1\\ U^{2}_{\_}3&U_{\_}3&1\end{matrix}\right),\det\left(\begin{matrix}U_{\_}1&1&V_{\_}1\\ U_{\_}2&1&V_{\_}2\\ U_{\_}3&1&V_{\_}3\end{matrix}\right),\ \det\left(\begin{matrix}1&V_{\_}1&V_{\_}1^{2}\\ 1&V_{\_}2&V_{\_}2^{2}\\ 1&V_{\_}3&V_{\_}3^{2}\end{matrix}\right).

Now we are ready to state the main theorem of this section.

Theorem 6.11.

The surgery map Ω:𝓗𝓕𝓛(T(n,n))𝓗𝓕𝓛(O_n)\Omega:\bm{{\mathcal{H\!F\!L}}}(T(n,n))\to\bm{{\mathcal{H\!F\!L}}}(O_{\_}n) is injective and its image coincides with the (completed) ideal 𝓙\bm{\mathcal{J}}. The map does not depend on the order and choices of crossing changes. It particular,

(T(n,n))𝒥\mathcal{H\!F\!L}(T(n,n))\simeq\mathcal{J}

as modules over 𝔽[U_1,,U_n,V_1,,V_n]\mathbb{F}[U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n].

Proof.

For the reader’s convenience, we break the proof into several steps.

Step 1: By Lemma 4.5, each map Ψ_ij=_m_ij(1)m_ijψ_m_ij\Psi_{\_}{ij}=\sum_{\_}{m_{\_}{ij}\in\mathbb{Z}}(-1)^{m_{\_}{ij}}\psi_{\_}{m_{\_}{ij}} is proportional, up to an explicit invertible factor τ_ij\tau_{\_}{ij}, to Ψ_ij0=ψ_0ψ_1.\Psi_{\_}{ij}^{0}=\psi_{\_}{0}-\psi_{\_}{-1}. The factors τ_ij\tau_{\_}{ij} do not depend on the order of crossing changes, and do not affect the injectivity or the image (which is an 𝔽[U_1,,U_n,V_1,,V_n]\mathbb{F}[U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n]-submodule of (O_n)\mathcal{H\!F\!L}(O_{\_}n)), so we can ignore them from now on and focus on m_ij{0,1}m_{\_}{ij}\in\{0,-1\}.

Step 2: By following the notations of Theorem 6.2, we can then rewrite the composition of Ψ_ij0=ψ_0ψ_1\Psi_{\_}{ij}^{0}=\psi_{\_}{0}-\psi_{\_}{-1} as

(12) Ω0=_ε{±1}(n2)sgn(ε)Ω_ε,sgn(ε)=_i<jε_ij.\Omega^{0}=\sum_{\_}{\varepsilon\in\{\pm 1\}^{\binom{n}{2}}}\mathrm{sgn}(\varepsilon)\Omega_{\_}{\varepsilon},\quad\mathrm{sgn}(\varepsilon)=\prod_{\_}{i<j}\varepsilon_{\_}{ij}.

By Theorem 6.2, for any given ε\varepsilon the order of composition does not matter.

Step 3: The terms in (12) are parametrized by the integer points in the permutahedron P_nP_{\_}n. However, some points will appear several times for different choices of ε\varepsilon, and by Theorem 6.2(c) the corresponding terms in (12) might cancel as long as they have the same Alexander degree. We claim that in fact the terms for all points will cancel except for the vertices of P_nP_{\_}n. To show this, consider the generating function

_εsgn(ε)Vp_ε=_εsgn(ε)V_1n12V_nn12V_i<jε_ijα_ij=_ε_i<jε_ijV_iV_jVε_ijα_ij=_i<j(V_iV_j).\sum_{\_}{\varepsilon}\mathrm{sgn}(\varepsilon)V^{p_{\_}{\varepsilon}}=\sum_{\_}{\varepsilon}\mathrm{sgn}(\varepsilon)V_{\_}1^{\frac{n-1}{2}}\cdots V_{\_}n^{\frac{n-1}{2}}V^{\sum_{\_}{i<j}\varepsilon_{\_}{ij}\alpha_{\_}{ij}}=\sum_{\_}{\varepsilon}\prod_{\_}{i<j}\varepsilon_{\_}{ij}\sqrt{V_{\_}iV_{\_}j}V^{\varepsilon_{\_}{ij}\alpha_{\_}{ij}}=\prod_{\_}{i<j}(V_{\_}i-V_{\_}j).

Here, we used that V_i=V_iV_jVα_ijV_{\_}i=\sqrt{V_{\_}iV_{\_}j}V^{\alpha_{\_}{ij}} and V_j=V_iV_jVα_ij-V_{\_}j=-\sqrt{V_{\_}iV_{\_}j}V^{-\alpha_{\_}{ij}}. On the other hand, we have the Vandermonde determinant

_i<j(V_iV_j)=±det(1V_1V_1n11V_nV_nn1)=±_σS_n(1)σV_1σ(1)1V_nσ(n)1.\prod_{\_}{i<j}(V_{\_}i-V_{\_}j)=\pm\det\left(\begin{matrix}1&V_{\_}1&\cdots&V_{\_}1^{n-1}\\ \vdots&\vdots&&\vdots\\ 1&V_{\_}n&\cdots&V_{\_}n^{n-1}\end{matrix}\right)=\pm\sum_{\_}{\sigma\in S_{\_}n}(-1)^{\sigma}V_{\_}1^{\sigma(1)-1}\cdots V_{\_}n^{\sigma(n)-1}.

As a conclusion of this step, we can write

Ω0=_εsgn(ε)Ω_ε=±_σS_n(1)sgn(σ)Ω_σ.\Omega^{0}=\sum_{\_}{\varepsilon}\mathrm{sgn}(\varepsilon)\Omega_{\_}{\varepsilon}=\pm\sum_{\_}{\sigma\in S_{\_}n}(-1)^{\mathrm{sgn}(\sigma)}\Omega_{\_}{\sigma}.

Step 4: We are ready to compute the image of Ω\Omega or, equivalently, of Ω0\Omega^{0}. Indeed, by Theorem 6.2(d),(e) we get

Ω0(a_j)=±_σS_n(1)sgn(σ)Ω_σ(a_j)=±_σS_n(1)sgn(σ)V_σ(1)n1jV_σ(2)n2jV_σ(n1j)U_σ(n+1j)U_σ(n)j=±det(U_1jU_11V_1V_1n1jU_njU_n1V_nV_nn1j),j=0,,n1.\begin{split}\Omega^{0}(a_{\_}j)&=\pm\sum_{\_}{\sigma\in S_{\_}n}(-1)^{\mathrm{sgn}(\sigma)}\Omega_{\_}{\sigma}(a_{\_}j)=\pm\sum_{\_}{\sigma\in S_{\_}n}(-1)^{\mathrm{sgn}(\sigma)}V_{\_}{\sigma(1)}^{n-1-j}V_{\_}{\sigma(2)}^{n-2-j}\cdots V_{\_}{\sigma(n-1-j)}U_{\_}{\sigma(n+1-j)}\cdots U_{\_}{\sigma(n)}^{j}\\ &=\pm\det\left(\begin{matrix}U_{\_}1^{j}&\cdots&U_{\_}1&1&V_{\_}1&\cdots&V_{\_}1^{n-1-j}\\ \vdots&&\vdots&\vdots&\vdots&&\vdots\\ U_{\_}n^{j}&\cdots&U_{\_}n&1&V_{\_}n&\cdots&V_{\_}n^{n-1-j}\end{matrix}\right),\quad j=0,\ldots,n-1.\end{split}

By Lemma 6.8 these determinants generate the ideal 𝒥\mathcal{J}.

Step 5: It remains to prove that Ω0\Omega^{0} is injective. Indeed, T(n,n)T(n,n) is an LL-space link, so in each Alexander multi-degree 𝐤\mathbf{k} we have (T(n,n),𝐤)𝔽[𝐔]\mathcal{H\!F\!L}(T(n,n),\mathbf{k})\cong\mathbb{F}[\mathbf{U}]. By the above, the image of any element of this tower under Ω0\Omega^{0} is a sum of elements in n!n! towers in (O_n)\mathcal{H\!F\!L}(O_{\_}n) located at the vertices of a permutahedron centered at 𝐤\mathbf{k}. Since all these elements appear with nonzero coefficients, their sum is also nonzero. ∎

6.3. Proof of Lemma 6.8

We start with several results which allow us to simplify the determinants Δ_S\Delta_{\_}S. Given a subset S={(a_1,b_1),,(a_n,b_n)}S=\{(a_{\_}1,b_{\_}1),\ldots,(a_{\_}n,b_{\_}n)\}, we define m_i=min(a_i,b_i)m_{\_}i=\min(a_{\_}i,b_{\_}i) and

S~={(a_1m_1,b_1m_1),,(a_nm_n,b_nm_n)}.\widetilde{S}=\left\{(a_{\_}1-m_{\_}1,b_{\_}1-m_{\_}1),\ldots,(a_{\_}n-m_{\_}n,b_{\_}n-m_{\_}n)\right\}.

Note that some elements of S~\widetilde{S} may coincide even if all elements of SS are distinct. The subset S~\widetilde{S} is contained in the union of the horizontal strip {b=0}\{b=0\} and the vertical strip {a=0}\{a=0\}, dashed in Figure 17.

Lemma 6.12.

We have Δ_S=𝐔NΔ_S~\Delta_{\_}S=\mathbf{U}^{N}\Delta_{\_}{\widetilde{S}} for N=m_1++m_nN=m_{\_}1+\ldots+m_{\_}n.

Proof.

For all i,ji,j we have

U_ja_iV_jb_i=U_ja_im_iV_jb_im_i(U_jV_j)m_i=U_ja_im_iV_jb_im_i𝐔m_iU_{\_}j^{a_{\_}i}V_{\_}j^{b_{\_}i}=U_{\_}j^{a_{\_}i-m_{\_}i}V_{\_}j^{b_{\_}i-m_{\_}i}(U_{\_}jV_{\_}j)^{m_{\_}i}=U_{\_}j^{a_{\_}i-m_{\_}i}V_{\_}j^{b_{\_}i-m_{\_}i}\mathbf{U}^{m_{\_}i}

and

Δ_S=det(U_1a_1V_1b_1U_1a_nV_1b_nU_na_1V_nb_1U_na_nV_nb_n)=det(U_1a_1m_1V_1b_1m_1𝐔m_1U_1a_nm_nV_1b_nm_n𝐔m_nU_na_1m_1V_nb_1m_1𝐔m_1U_na_nm_nV_nb_nm_n𝐔m_n)=\Delta_{\_}S=\det\left(\begin{matrix}U_{\_}1^{a_{\_}1}V_{\_}1^{b_{\_}1}&\cdots&U_{\_}1^{a_{\_}n}V_{\_}1^{b_{\_}n}\\ \vdots&\ddots&\vdots\\ U_{\_}n^{a_{\_}1}V_{\_}n^{b_{\_}1}&\cdots&U_{\_}n^{a_{\_}n}V_{\_}n^{b_{\_}n}\\ \end{matrix}\right)=\det\left(\begin{matrix}U_{\_}1^{a_{\_}1-m_{\_}1}V_{\_}1^{b_{\_}1-m_{\_}1}\mathbf{U}^{m_{\_}1}&\cdots&U_{\_}1^{a_{\_}n-m_{\_}n}V_{\_}1^{b_{\_}n-m_{\_}n}\mathbf{U}^{m_{\_}n}\\ \vdots&\ddots&\vdots\\ U_{\_}n^{a_{\_}1-m_{\_}1}V_{\_}n^{b_{\_}1-m_{\_}1}\mathbf{U}^{m_{\_}1}&\cdots&U_{\_}n^{a_{\_}n-m_{\_}n}V_{\_}n^{b_{\_}n-m_{\_}n}\mathbf{U}^{m_{\_}n}\\ \end{matrix}\right)=
𝐔m_1++m_ndet(U_1a_1m_1V_1b_1m_1U_1a_nm_nV_1b_nm_nU_na_1m_1V_nb_1m_1U_na_nm_nV_nb_nm_n)=𝐔m_1++m_nΔ_S~.\mathbf{U}^{m_{\_}1+\ldots+m_{\_}n}\det\left(\begin{matrix}U_{\_}1^{a_{\_}1-m_{\_}1}V_{\_}1^{b_{\_}1-m_{\_}1}&\cdots&U_{\_}1^{a_{\_}n-m_{\_}n}V_{\_}1^{b_{\_}n-m_{\_}n}\\ \vdots&\ddots&\vdots\\ U_{\_}n^{a_{\_}1-m_{\_}1}V_{\_}n^{b_{\_}1-m_{\_}1}&\cdots&U_{\_}n^{a_{\_}n-m_{\_}n}V_{\_}n^{b_{\_}n-m_{\_}n}\\ \end{matrix}\right)=\mathbf{U}^{m_{\_}1+\ldots+m_{\_}n}\Delta_{\_}{\widetilde{S}}.

Example 6.13.

For S={(0,0),(1,2),(3,5),(6,4)}S=\{(0,0),(1,2),(3,5),(6,4)\} we have S~={(0,0),(0,1),(0,2),(2,0)}\widetilde{S}=\{(0,0),(0,1),(0,2),(2,0)\}, see Figure 17. In this case N=8N=8 and we have

Δ_S=det(1U_1V_12U_13V_15U_16V_141U_2V_22U_23V_25U_26V_241U_3V_32U_33V_35U_36V_341U_4V_42U_43V_45U_46V_44)=𝐔8det(1V_1V_12U_121V_2V_22U_221V_3V_32U_321V_4V_42U_42)=𝐔8Δ_S~.\Delta_{\_}S=\det\left(\begin{matrix}1&U_{\_}1V_{\_}1^{2}&U_{\_}1^{3}V_{\_}1^{5}&U_{\_}1^{6}V_{\_}1^{4}\\ 1&U_{\_}2V_{\_}2^{2}&U_{\_}2^{3}V_{\_}2^{5}&U_{\_}2^{6}V_{\_}2^{4}\\ 1&U_{\_}3V_{\_}3^{2}&U_{\_}3^{3}V_{\_}3^{5}&U_{\_}3^{6}V_{\_}3^{4}\\ 1&U_{\_}4V_{\_}4^{2}&U_{\_}4^{3}V_{\_}4^{5}&U_{\_}4^{6}V_{\_}4^{4}\\ \end{matrix}\right)=\mathbf{U}^{8}\det\left(\begin{matrix}1&V_{\_}1&V_{\_}1^{2}&U_{\_}1^{2}\\ 1&V_{\_}2&V_{\_}2^{2}&U_{\_}2^{2}\\ 1&V_{\_}3&V_{\_}3^{2}&U_{\_}3^{2}\\ 1&V_{\_}4&V_{\_}4^{2}&U_{\_}4^{2}\\ \end{matrix}\right)=\mathbf{U}^{8}\Delta_{\_}{\widetilde{S}}.
011223344556601122334455\bullet\bullet\bullet\bullet\bullet\bullet\bullet\bullet
Figure 17. The sets SS and S~\widetilde{S}

Let e_k(U)=_i_1<<i_kU_i_1U_i_ke_{\_}k(U)=\sum_{\_}{i_{\_}1<\ldots<i_{\_}k}U_{\_}{i_{\_}1}\cdots U_{\_}{i_{\_}k} be the kk-th elementary symmetric function. We have the following analogue of the Pieri rule for Schur functions [19].

Lemma 6.14.

We have

e_k(U)Δ_S=_SΔ_Se_{\_}k(U)\Delta_{\_}S=\sum_{\_}{S^{\prime}}\Delta_{\_}{S^{\prime}}

where SS^{\prime} is obtained by adding (1,0)(1,0) to kk distinct elements of SS and leaving other elements unchanged.

Proof.

Given a polynomial f(U_1,,U_n,V_1,,V_n)f(U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n), we define

Alt(f)=_σS_n(1)sgn(σ)f(U_σ(1),,U_σ(n),V_σ(1),,V_σ(n)).\mathrm{Alt}(f)=\sum_{\_}{\sigma\in S_{\_}n}(-1)^{\mathrm{sgn}(\sigma)}f(U_{\_}{\sigma(1)},\ldots,U_{\_}{\sigma(n)},V_{\_}{\sigma(1)},\ldots,V_{\_}{\sigma(n)}).

For S={(a_1,b_1),(a_n,b_n)}S=\{(a_{\_}1,b_{\_}1),\cdots(a_{\_}n,b_{\_}n)\} we get Δ_S=Alt(U_1a_1V_1b_1U_na_nV_nb_n).\Delta_{\_}S=\mathrm{Alt}\left(U_{\_}1^{a_{\_}1}V_{\_}1^{b_{\_}1}\cdots U_{\_}n^{a_{\_}n}V_{\_}n^{b_{\_}n}\right). Clearly, Alt(f)\mathrm{Alt}(f) is linear in ff and Alt(fg)=Alt(f)g\mathrm{Alt}(fg)=\mathrm{Alt}(f)g for any symmetric polynomial gg. Since e_k(U)e_{\_}k(U) is symmetric, we get

e_k(U)Δ_S=e_k(U)Alt(U_1a_1V_1b_1U_na_nV_nb_n)=Alt(e_k(U)U_1a_1V_1b_1U_na_nV_nb_n)=e_{\_}k(U)\Delta_{\_}S=e_{\_}k(U)\mathrm{Alt}\left(U_{\_}1^{a_{\_}1}V_{\_}1^{b_{\_}1}\cdots U_{\_}n^{a_{\_}n}V_{\_}n^{b_{\_}n}\right)=\mathrm{Alt}\left(e_{\_}k(U)U_{\_}1^{a_{\_}1}V_{\_}1^{b_{\_}1}\cdots U_{\_}n^{a_{\_}n}V_{\_}n^{b_{\_}n}\right)=
_i_1<<i_kAlt(U_i_1U_i_kU_1a_1V_1b_1U_na_nV_nb_n)=_SΔ_S\sum_{\_}{i_{\_}1<\ldots<i_{\_}k}\mathrm{Alt}\left(U_{\_}{i_{\_}1}\cdots U_{\_}{i_{\_}k}\cdot U_{\_}1^{a_{\_}1}V_{\_}1^{b_{\_}1}\cdots U_{\_}n^{a_{\_}n}V_{\_}n^{b_{\_}n}\right)=\sum_{\_}{S^{\prime}}\Delta_{\_}{S^{\prime}}

where S={(a_1,b_1),(a_n,b_n)},S^{\prime}=\{(a^{\prime}_{\_}1,b^{\prime}_{\_}1),\cdots(a^{\prime}_{\_}n,b^{\prime}_{\_}n)\},

(a_i_s,b_i_s)=(a_i_s,b_i_s)+(1,0),(a_j,b_j)=(a_j,b_j)j{i_1,,i_k}.\quad(a^{\prime}_{\_}{i_{\_}s},b^{\prime}_{\_}{i_{\_}s})=(a_{\_}{i_{\_}s},b_{\_}{i_{\_}s})+(1,0),\quad(a^{\prime}_{\_}j,b^{\prime}_{\_}j)=(a_{\_}j,b_{\_}j)\quad j\notin\{i_{\_}1,\ldots,i_{\_}k\}.

Example 6.15.

We have

e_1(U)Δ_{(0,0),(0,1),(0,2),(1,0)}=Δ_{(1,0),(0,1),(0,2),(1,0)}+Δ_{(0,0),(1,1),(0,2),(1,0)}+e_{\_}1(U)\Delta_{\_}{\{(0,0),(0,1),(0,2),(1,0)\}}=\Delta_{\_}{\{(1,0),(0,1),(0,2),(1,0)\}}+\Delta_{\_}{\{(0,0),(1,1),(0,2),(1,0)\}}+
Δ_{(0,0),(0,1),(1,2),(1,0)}+Δ_{(0,0),(0,1),(0,2),(2,0)}.\Delta_{\_}{\{(0,0),(0,1),(1,2),(1,0)\}}+\Delta_{\_}{\{(0,0),(0,1),(0,2),(2,0)\}}.

The first term vanishes since (1,0)(1,0) is repeated twice. By Lemma 6.12 the second term equals

Δ_{(0,0),(1,1),(0,2),(1,0)}=𝐔Δ_{(0,0),(0,0),(0,2),(1,0)}=0\Delta_{\_}{\{(0,0),(1,1),(0,2),(1,0)\}}=\mathbf{U}\Delta_{\_}{\{(0,0),(0,0),(0,2),(1,0)\}}=0

and the third term equals

Δ_{(0,0),(0,1),(1,2),(1,0)}=𝐔Δ_{(0,0),(0,1),(0,1),(1,0)}=0.\Delta_{\_}{\{(0,0),(0,1),(1,2),(1,0)\}}=\mathbf{U}\Delta_{\_}{\{(0,0),(0,1),(0,1),(1,0)\}}=0.

Therefore

e_1(U)Δ_{(0,0),(0,1),(0,2),(1,0)}=Δ_{(0,0),(0,1),(0,2),(2,0)}.e_{\_}1(U)\Delta_{\_}{\{(0,0),(0,1),(0,2),(1,0)\}}=\Delta_{\_}{\{(0,0),(0,1),(0,2),(2,0)\}}.
Example 6.16.

Similarly, we have

e_2(U)Δ_{(0,0),(0,1),(0,2),(2,0),(3,0))}=Δ_{(1,0),(1,1),(0,2),(2,0),(3,0))}+e_{\_}2(U)\Delta_{\_}{\{(0,0),(0,1),(0,2),(2,0),(3,0))\}}=\Delta_{\_}{\{(1,0),(1,1),(0,2),(2,0),(3,0))\}}+
+Δ_{(1,0),(0,1),(0,2),(2,0),(4,0))}+Δ_{(0,0),(0,1),(0,2),(3,0),(4,0))}+\Delta_{\_}{\{(1,0),(0,1),(0,2),(2,0),(4,0))\}}+\Delta_{\_}{\{(0,0),(0,1),(0,2),(3,0),(4,0))\}}

and all other terms vanish. The first term can be simplified as

Δ_{(1,0),(1,1),(0,2),(2,0),(3,0))}=𝐔Δ_{(1,0),(0,0),(0,2),(2,0),(3,0))}.\Delta_{\_}{\{(1,0),(1,1),(0,2),(2,0),(3,0))\}}=\mathbf{U}\Delta_{\_}{\{(1,0),(0,0),(0,2),(2,0),(3,0))\}}.
Proof of Lemma 6.8.

Recall that 𝒥\mathcal{J} is generated by the determinants Δ_S\Delta_{\_}S for arbitrary subsets SS. We need to prove that it is generated by Δ_S_k\Delta_{\_}{S_{\_}k} where

S_k={(0,0),(1,0),,(k1,0),(0,1),,(0,nk)},1kn.S_{\_}k=\left\{(0,0),(1,0),\ldots,(k-1,0),(0,1),\ldots,(0,n-k)\right\},\quad\quad 1\leq k\leq n.

By Lemma 6.12 we have Δ_S\Delta_{\_}{S} proportional to Δ_S~\Delta_{\_}{\widetilde{S}} where all elements of S~\widetilde{S} have the form (a,0)(a,0) or (0,b)(0,b) (that is, S~\widetilde{S} is contained in the dashed region in Figure 17), so after reordering its elements we can write

S~={(a_1,0),,(a_k,0),(0,b_k+1),(0,b_n)},0a_1<<a_k, 0<b_k+1<<b_n.\widetilde{S}=\{(a_{\_}1,0),\ldots,(a_{\_}k,0),(0,b_{\_}{k+1})\ldots,(0,b_{\_}{n})\},\quad 0\leq a_{\_}1<\ldots<a_{\_}k,\ \ \ 0<b_{\_}{k+1}<\ldots<b_{\_}n.

It remains to prove that, in fact, it is sufficient to only consider the “dense” subsets where a_i=i1a_{\_}i=i-1 and b_i=ikb_{\_}i=i-k. There is a natural partial order on such S~\widetilde{S} where S~S~\widetilde{S}^{\prime}\prec\widetilde{S}^{\prime\prime} if S~\widetilde{S}^{\prime} is obtained by “sliding” some elements of S~\widetilde{S}^{\prime\prime} left or down. Clearly, “dense” subsets are minimal in this order.

Assume that 0<a_j10<a_{\_}{j}-1, (a_j1,0)S~(a_{\_}j-1,0)\notin\widetilde{S} and jj is maximal with this property, that is, a_ja_{\_}j bounds the rightmost gap in S~\widetilde{S}. This implies that a_i=a_j+ija_{\_}{i}=a_{\_}{j}+i-j for j+1ikj+1\leq i\leq k. Let

S_1~={(a_1,0),,(a_j1,0),,(a_k1,0),(0,b_k+1),(0,b_n)}.\widetilde{S_{\_}1}=\{(a_{\_}1,0),\ldots,(a_{\_}j-1,0),\ldots,(a_{\_}k-1,0),(0,b_{\_}{k+1})\ldots,(0,b_{\_}{n})\}.

In Example 6.15 we have

S~={(0,0),(0,1),(0,2),(2,0)},a_j=2andS_1~={(0,0),(0,1),(0,2),(1,0)},\widetilde{S}=\{(0,0),(0,1),(0,2),(2,0)\},\ a_{\_}j=2\ \mathrm{and}\ \widetilde{S_{\_}1}=\{(0,0),(0,1),(0,2),(1,0)\},

in Example 6.16 we have

S~={(0,0),(0,1),(0,2),(3,0),(4,0)},a_j=3andS_1~={(0,0),(0,1),(0,2),(2,0),(3,0)}.\widetilde{S}=\{(0,0),(0,1),(0,2),(3,0),(4,0)\},\ a_{\_}j=3\ \mathrm{and}\ \widetilde{S_{\_}1}=\{(0,0),(0,1),(0,2),(2,0),(3,0)\}.

Then by Lemmas 6.14 and 6.12 (see also Examples 6.15 and 6.16) we have

e_kj+1(U)Δ_S~_1=Δ_S~+_SΔ_S,Δ_S=𝐔NΔ_S~e_{\_}{k-j+1}(U)\Delta_{\_}{\widetilde{S}_{\_}1}=\Delta_{\_}{\widetilde{S}}+\sum_{\_}{S^{\prime}}\Delta_{\_}{S^{\prime}},\quad\Delta_{\_}{S^{\prime}}=\mathbf{U}^{N}\Delta_{\_}{\widetilde{S^{\prime}}}

where S~S~\widetilde{S^{\prime}}\prec\widetilde{S}. Therefore Δ_S~\Delta_{\_}{\widetilde{S}} belongs to the ideal generated by Δ_S~_1\Delta_{\_}{\widetilde{S}_{\_}1} and Δ_S~\Delta_{\_}{\widetilde{S^{\prime}}}. We can proceed by induction in the above partial order until a_ja_{\_}j are dense. Then repeat the same argument swapping U_iU_{\_}i and V_iV_{\_}i and using a version of Lemma 6.14 for elementary symmetric functions in V_iV_{\_}i, this will ensure that b_jb_{\_}j are dense as well. ∎

6.4. More on ideal 𝒥\mathcal{J} and its cousins

In this section we collect some further facts on the ideal 𝒥\mathcal{J} and discuss some analogues of Theorem 6.11 for other homology theories.

Definition 6.17.

A polynomial f(U_1,,U_n,V_1,,V_n)f(U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n) is called antisymmetric if

f(U_σ(1),,U_σ(n),V_σ(1),,V_σ(n))=(1)σf(U_1,,U_n,V_1,,V_n).f(U_{\_}{\sigma(1)},\ldots,U_{\_}{\sigma(n)},V_{\_}{\sigma(1)},\ldots,V_{\_}{\sigma(n)})=(-1)^{\sigma}f(U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n).

Note that over our ground field 𝔽\mathbb{F} of characteristic 2, any antisymmetric polynomial is also symmetric. However, for other ground fields there is an important distinction.

Lemma 6.18.

Let char𝐤2\mathrm{char}\ \mathbf{k}\neq 2. Then any antisymmetric polynomial with coefficients in 𝐤\mathbf{k} is a linear combination of minors Δ_S\Delta_{\_}S.

Proof.

Let ff be an antisymmetric polynomial, and U_1a_1V_1b_1U_na_nV_nb_nU_{\_}1^{a_{\_}1}V_{\_}1^{b_{\_}1}\cdots U_{\_}n^{a_{\_}n}V_{\_}n^{b_{\_}n} be a monomial in ff with nonzero coefficient. If a_i=a_ja_{\_}i=a_{\_}j and b_i=b_jb_{\_}i=b_{\_}j then the transposition (ij)(i\ j) fixes this monomial, but since ff is antisymmetric it must change its sign, contradiction. Therefore all pairs (a_1,b_1),,(a_n,b_n)(a_{\_}1,b_{\_}1),\ldots,(a_{\_}n,b_{\_}n) are distinct and the S_nS_{\_}n-orbit of this monomial adds up to Δ_S\Delta_{\_}S for S={(a_1,b_1),,(a_n,b_n)}S=\{(a_{\_}1,b_{\_}1),\ldots,(a_{\_}n,b_{\_}n)\}. ∎

Informally, we can think of 𝒥\mathcal{J} as a characteristic 2 reduction of the ideal generated by antisymmetric polynomials in (O_n)\mathcal{H\!F\!L}(O_{\_}n). Next, we check directly that Theorem 6.11 is compatible with Theorem 3.11. By Lemma 6.8 the ideal 𝒥\mathcal{J} is generated by nn minors Ω0(a_k)\Omega^{0}(a_{\_}k) and it is sufficient to check the relations between them.

Lemma 6.19.

The determinants

Ω0(a_j)=±det(U_1jU_11V_1V_1n1jU_njU_n1V_nV_nn1j)\Omega^{0}(a_{\_}j)=\pm\det\left(\begin{matrix}U_{\_}1^{j}&\cdots&U_{\_}1&1&V_{\_}1&\cdots&V_{\_}1^{n-1-j}\\ \vdots&&\vdots&\vdots&\vdots&&\vdots\\ U_{\_}n^{j}&\cdots&U_{\_}n&1&V_{\_}n&\cdots&V_{\_}n^{n-1-j}\end{matrix}\right)

satisfy the relations (4) (up to signs).

Proof.

Recall that the determinants Ω0(a_k)\Omega^{0}(a_{\_}k) are antisymmetrizations of monomials

Ω_𝟏(a_k)=V_1n1kV_n1kU_n+1kU_nk,Ω_𝟏(a_k+1)=V_1n2kV_n2kU_nkU_nk+1,\Omega_{\_}{\bf 1}(a_{\_}k)=V_{\_}1^{n-1-k}\cdots V_{\_}{n-1-k}U_{\_}{n+1-k}\cdots U_{\_}n^{k},\ \Omega_{\_}{\bf 1}(a_{\_}{k+1})=V_{\_}1^{n-2-k}\cdots V_{\_}{n-2-k}U_{\_}{n-k}\cdots U_{\_}n^{k+1},

let us check the relations (4) between them for all possible II. Let us first pick I={nk,,n}I=\{n-k,\ldots,n\}, then I¯={1,,nk1}\overline{I}=\{1,\ldots,n-k-1\} and

Ω_𝟏(a_k)U_I=(V_1n1kV_n1kU_n+1kU_nk)×U_nkU_n=\Omega_{\_}{\bf 1}(a_{\_}k)U_{\_}I=\left(V_{\_}1^{n-1-k}\cdots V_{\_}{n-1-k}U_{\_}{n+1-k}\cdots U_{\_}n^{k}\right)\times U_{\_}{n-k}\cdots U_{\_}n=
(V_1n2kV_n2kU_nkU_nk+1)×V_1V_nk1=Ω_𝟏(a_k+1)V_I¯.\left(V_{\_}1^{n-2-k}\cdots V_{\_}{n-2-k}U_{\_}{n-k}\cdots U_{\_}n^{k+1}\right)\times V_{\_}1\cdots V_{\_}{n-k-1}=\Omega_{\_}{\bf 1}(a_{\_}{k+1})V_{\_}{\overline{I}}.

More generally, for an arbitrary II define

I_1=I{1,,n1k},I_2=I{nk,,n},I_{\_}1=I\cap\{1,\ldots,n-1-k\},\ I_{\_}2=I\cap\{n-k,\ldots,n\},
I¯_1={1,,n1k}I,I¯_2={nk,,n}I.\overline{I}_{\_}1=\{1,\ldots,n-1-k\}\setminus I,\ \overline{I}_{\_}2=\{n-k,\ldots,n\}\setminus I.

Let X_k=V_1n2kV_n2kU_n+1kU_nkX_{\_}k=V_{\_}1^{n-2-k}\cdots V_{\_}{n-2-k}U_{\_}{n+1-k}\cdots U_{\_}n^{k}, then

Ω_𝟏(a_k)U_I=X_kV_{1,,n1k}U_I=X_k(V_I_1V_I¯_1)(U_I_1U_I_2)=X_kV_I¯_1U_I_2(V_I_1U_I_1)=\Omega_{\_}{\bf 1}(a_{\_}k)U_{\_}I=X_{\_}kV_{\_}{\{1,\ldots,n-1-k\}}U_{\_}I=X_{\_}k\left(V_{\_}{I_{\_}1}V_{\_}{\overline{I}_{\_}1}\right)\left(U_{\_}{I_{\_}1}U_{\_}{I_{\_}2}\right)=X_{\_}kV_{\_}{\overline{I}_{\_}1}U_{\_}{I_{\_}2}\left(V_{\_}{I_{\_}1}U_{\_}{I_{\_}1}\right)=
X_kV_I¯_1U_I_2(V_I¯_2U_I¯_2)=X_k(V_I¯_1V_I¯_2)(U_I_2U_I¯_2)=X_kV_I¯U_{nk,,n}=Ω_𝟏(a_k+1)V_I¯.X_{\_}kV_{\_}{\overline{I}_{\_}1}U_{\_}{I_{\_}2}\left(V_{\_}{\overline{I}_{\_}2}U_{\_}{\overline{I}_{\_}2}\right)=X_{\_}k\left(V_{\_}{\overline{I}_{\_}1}V_{\_}{\overline{I}_{\_}2}\right)\left(U_{\_}{I_{\_}2}U_{\_}{\overline{I}_{\_}2}\right)=X_{\_}kV_{\_}{\overline{I}}U_{\_}{\{n-k,\ldots,n\}}=\Omega_{\_}{\bf 1}(a_{\_}{k+1})V_{\_}{\overline{I}}.

Here we used the relation V_I_1U_I_1=V_I¯_2U_I¯_2V_{\_}{I_{\_}1}U_{\_}{I_{\_}1}=V_{\_}{\overline{I}_{\_}2}U_{\_}{\overline{I}_{\_}2}.

Since the relations (4) are S_nS_{\_}n-equivariant, the relations for Ω_𝟏(a_k)\Omega_{\_}{\bf 1}(a_{\_}k) imply the same relations for Ω0(a_k)\Omega^{0}(a_{\_}k). ∎

Next, we study the relations between the ideals corresponding to the link T(n,n)T(n,n) and its sublinks.

Lemma 6.20.

For any subset I{1,,n}I\subset\{1,\ldots,n\} with |I|2|I|\geq 2 let 𝒥_I\mathcal{J}_{\_}I be the ideal generated by the monomial minors in variables U_i,V_i,iIU_{\_}i,V_{\_}i,i\in I. Then 𝒥𝒥_I\mathcal{J}\subset\mathcal{J}_{\_}I.

Proof.

Let S={(a_1,b_1),,(a_n,b_n)}S=\{(a_{\_}1,b_{\_}1),\ldots,(a_{\_}n,b_{\_}n)\}, then one can write the minor Δ_S\Delta_{\_}S as follows:

Δ_S=_L±det(U_ia_jV_ib_j)_iI,jLdet(U_ia_jV_ib_j)_iI,jL\Delta_{\_}S=\sum_{\_}{L}\pm\det(U_{\_}i^{a_{\_}j}V_{\_}i^{b_{\_}j})_{\_}{i\in I,j\in L}\det(U_{\_}i^{a_{\_}j}V_{\_}i^{b_{\_}j})_{\_}{i\notin I,j\notin L}

where the sum runs over all |I||I|-element subsets L{1,,n}L\subset\{1,\ldots,n\}. Since det(U_ia_jV_ib_j)_iI,jL𝒥_I\det(U_{\_}i^{a_{\_}j}V_{\_}i^{b_{\_}j})_{\_}{i\in I,j\in L}\in\mathcal{J}_{\_}I, we get Δ_S𝒥_I\Delta_{\_}S\in\mathcal{J}_{\_}I and 𝒥𝒥_I\mathcal{J}\subset\mathcal{J}_{\_}I. ∎

Remark 6.21.

Topologically, the ideal 𝒥_I\mathcal{J}_{\_}I corresponds to the union of the sublink _I\mathcal{L}_{\_}I formed by the components L_i,iIL_{\_}i,i\in I of T(n,n)T(n,n) with n|I|n-|I| unknotted disjoint circles. Since the splitting map for T(n,n)T(n,n) does not depend on the order of crossing changes, we can first split off the components with indices not in II, and the splitting map Ω\Omega (respectively, Ω0\Omega^{0}) will factor through the splitting map Ω_I\Omega_{\_}I (resp. Ω_I0\Omega_{\_}I^{0}) for the resulting link. Therefore the image of Ω\Omega is contained in the image of Ω_I\Omega_{\_}I.

Corollary 6.22.

We have

𝒥_i<j(U_iU_j,V_iV_j)(O_n)=𝔽[U_1,,U_n,V_1,,V_n](U_iV_i=U_jV_j,ij).\mathcal{J}\subset\bigcap_{\_}{i<j}(U_{\_}i-U_{\_}j,V_{\_}i-V_{\_}j)\subset\mathcal{H\!F\!L}(O_{\_}n)=\frac{\mathbb{F}[U_{\_}1,\ldots,U_{\_}n,V_{\_}1,\ldots,V_{\_}n]}{(U_{\_}iV_{\_}i=U_{\_}jV_{\_}j,\ i\neq j)}.
Proof.

By Lemma 6.8, for a two-component sublink L_iL_jL_{\_}i\cup L_{\_}j we have 𝒥_ij=(U_iU_j,V_iV_j)\mathcal{J}_{\_}{ij}=(U_{\_}i-U_{\_}j,V_{\_}i-V_{\_}j), and 𝒥𝒥_ij\mathcal{J}\subset\mathcal{J}_{\_}{ij} for all i<ji<j. ∎

The analogues of the ideal 𝒥\mathcal{J} and of Theorem 6.11 appeared in triply graded Khovanov-Rozansky homology. Recall that the triply graded homology of the unknot is HHH(O_1)=[x,θ]\mathrm{HHH}(O_{\_}1)=\mathbb{C}[x,\theta] where xx is an even and θ\theta is an odd variable. There is also a skein exact triangle

HHH(L_+){\mathrm{HHH}(L_{\_}+)}HHH(L_){\mathrm{HHH}(L_{\_}-)}(HHH(L_0)HHH(L_0)){\left(\mathrm{HHH}(L_{\_}0)\to\mathrm{HHH}(L_{\_}0)\right)}Ψ_ijHHH\scriptstyle{\Psi_{\_}{ij}^{\mathrm{HHH}}}

analogous to the skein exact triangle (5). The map Ψ_ijHHH\Psi_{\_}{ij}^{\mathrm{HHH}} can be used to define a splitting map ΩHHH:HHH(T(n,n))HHH(O_n)=[x_1,,x_n,θ_1,,θ_n]\Omega^{\mathrm{HHH}}:\mathrm{HHH}(T(n,n))\to\mathrm{HHH}(O_{\_}n)=\mathbb{C}[x_{\_}1,\ldots,x_{\_}n,\theta_{\_}1,\ldots,\theta_{\_}n] which is, unfortunately, not injective.

To resolve this problem, second author and Hogancamp introduced in [12] a deformation, or yy-ification of Khovanov-Rozansky homology HY(L)\mathrm{HY}(L). The skein exact triangle can be defined in this deformed theory, and there is a (unique up to homotopy) splitting map Ψ_ijHY:HY(L_+)HY(L_)\Psi_{\_}{ij}^{\mathrm{HY}}:\mathrm{HY}(L_{\_}+)\to\mathrm{HY}(L_{\_}-). By composing these, one obtains a splitting map

ΩHY:HY(T(n,n))HY(O_n)=[x_1,,x_n,y_1,,y_n,θ_1,,θ_n].\Omega^{\mathrm{HY}}:\mathrm{HY}(T(n,n))\to\mathrm{HY}(O_{\_}n)=\mathbb{C}[x_{\_}1,\ldots,x_{\_}n,y_{\_}1,\ldots,y_{\_}n,\theta_{\_}1,\ldots,\theta_{\_}n].
Theorem 6.23 ([12]).

The map Ω_HY\Omega_{\_}{\mathrm{HY}} is injective, and its image coincides with the ideal 𝒥HY\mathcal{J}^{\mathrm{HY}} generated by antisymmetric polynomials in HY(O_n)\mathrm{HY}(O_{\_}n). Furthermore, the image of Ω_HY\Omega_{\_}{\mathrm{HY}} coincides with the ideal

_i<j𝒥HY_ij=_i<j(x_ix_j,y_iy_j,θ_iθ_j)[x_1,,x_n,y_1,,y_n,θ_1,,θ_n].\bigcap_{\_}{i<j}\mathcal{J}^{\mathrm{HY}}_{\_}{ij}=\bigcap_{\_}{i<j}(x_{\_}i-x_{\_}j,y_{\_}i-y_{\_}j,\theta_{\_}i-\theta_{\_}j)\subset\mathbb{C}[x_{\_}1,\ldots,x_{\_}n,y_{\_}1,\ldots,y_{\_}n,\theta_{\_}1,\ldots,\theta_{\_}n].
Remark 6.24.

Recently Hogancamp, Rose and Wedrich [15] studied yy-ification and the splitting maps for colored triply graded homology, and obtained similar ideals. See [15, Conjecture 8.12,Theorem 9.33] for more details.

In [3] Batson and Seed defined a deformation of Khovanov homology, which was generalized in [9] by Cautis and Kamnitzer to 𝔰𝔩(N)\mathfrak{sl}(N) Khovanov-Rozansky homology. Their constructions predate and motivate the construction of HY\mathrm{HY}, and we commonly refer to them as yy-ified link homology.

Problem 6.25.

Define the analogues of the splitting map Ω_Kh,Ω_𝔰𝔩(N):T(n,n)O_n\Omega_{\_}{Kh},\Omega_{\_}{\mathfrak{sl}(N)}:T(n,n)\to O_{\_}n for yy-ified Khovanov and 𝔰𝔩(N)\mathfrak{sl}(N) homologies. Is it possible to describe their images as some determinantal ideals in the homology of unlink?

Problem 6.26.

The main result of [5] (following the earlier work in [10, 11]) defines a spectral sequence from the reduced version of HHH\mathrm{HHH} to 𝐻𝐹𝐾^\widehat{\mathit{HFK}} for knots. Is it possible to extend this to a spectral sequence from HY\mathrm{HY} to \mathcal{H\!F\!L} for arbitrary links?

6.5. Unlinks in S1×S2S^{1}\times S^{2}

As an application of the above results, we can compute the Heegaard Floer homology of certain links in S1×S2S^{1}\times S^{2} and prove Theorem 1.9. Let Z_nZ_{\_}n be the union of the nn parallel copies of S1S^{1} inside S1×S2S^{1}\times S^{2}. Recall from Section 3.2 that we have generalized crossing change maps

ϕ_kn:(O_n)(T(n,n))\phi_{\_}k^{n}:\mathcal{H\!F\!L}(O_{\_}n)\to\mathcal{H\!F\!L}(T(n,n))

defined by attaching a 2-handle along the (1)(-1)-framed meridian. The index kk corresponds to the choice of a Spinc\text{Spin}^{c} structure on the corresponding cobordism, see Proposition 3.10. The surgery exact sequence immediately implies the following:

Lemma 6.27.

There is a long exact sequence

𝓗𝓕𝓛(S3,O_n)Φ𝓗𝓕𝓛(S3,T(n,n))𝓗𝓕𝓛(S1×S2,Z_n)𝓗𝓕𝓛(S3,O_n)\rightarrow\bm{{\mathcal{H\!F\!L}}}(S^{3},O_{\_}n)\xrightarrow{\Phi}\bm{{\mathcal{H\!F\!L}}}(S^{3},T(n,n))\rightarrow\bm{{\mathcal{H\!F\!L}}}(S^{1}\times S^{2},Z_{\_}n)\rightarrow\bm{{\mathcal{H\!F\!L}}}(S^{3},O_{\_}n)\rightarrow

where Φ=_kϕ_kn\Phi=\sum_{\_}{k\in\mathbb{Z}}\phi_{\_}k^{n} is the sum of ϕ_kn\phi_{\_}k^{n} over all Spinc\text{Spin}^{c} structures.

Remark 6.28.

As above, if we work with integer coefficients then ϕ_nk\phi_{\_}n^{k} would acquire some signs in the sum. Instead of guessing the signs, we simply work over 𝔽\mathbb{F}.

Define two series

μ_0=_k=0(V_1V_n)k𝐔k(k1)2,μ_n1=_k=0(U_1U_n)k𝐔k(k1)2.\mu_{\_}0=\sum_{\_}{k=0}^{\infty}(V_{\_}1\cdots V_{\_}n)^{k}\mathbf{U}^{\frac{k(k-1)}{2}},\quad\mu_{\_}{n-1}=\sum_{\_}{k=0}^{\infty}(U_{\_}1\cdots U_{\_}n)^{k}\mathbf{U}^{\frac{k(k-1)}{2}}.
Theorem 6.29.

The homology 𝓗𝓕𝓛(S1×S2,Z_n)\bm{{\mathcal{H\!F\!L}}}(S^{1}\times S^{2},Z_{\_}n) admits two equivalent descriptions:

a)

𝓗𝓕𝓛(S1×S2,Z_n)𝓗𝓕𝓛(T(n,n))(μ_0a_0+a_1++a_n2+μ_n1a_n1)\bm{{\mathcal{H\!F\!L}}}(S^{1}\times S^{2},Z_{\_}n)\simeq\frac{\bm{{\mathcal{H\!F\!L}}}(T(n,n))}{(\mu_{\_}0a_{\_}0+a_{\_}1+\ldots+a_{\_}{n-2}+\mu_{\_}{n-1}a_{\_}{n-1})}

where a_ia_{\_}i are the generators of 𝓗𝓕𝓛(T(n,n))\bm{{\mathcal{H\!F\!L}}}(T(n,n)) from Theorem 3.11.

b) 𝓗𝓕𝓛(S1×S2,Z_n)𝓙/(γ)\bm{{\mathcal{H\!F\!L}}}(S^{1}\times S^{2},Z_{\_}n)\simeq\bm{\mathcal{J}}/(\gamma) where 𝓙\bm{\mathcal{J}} is the (completed) determinantal ideal from Theorem 6.11 and

γ=μ_0_i<j(V_iV_j)+μ_n1_i<j(U_iU_j)+_j=1n2det(U_1jU_11V_1V_1n1jU_njU_n1V_nV_nn1j).\gamma=\mu_{\_}0\prod_{\_}{i<j}(V_{\_}i-V_{\_}j)+\mu_{\_}{n-1}\prod_{\_}{i<j}(U_{\_}i-U_{\_}j)+\sum_{\_}{j=1}^{n-2}\det\left(\begin{matrix}U_{\_}1^{j}&\cdots&U_{\_}1&1&V_{\_}1&\cdots&V_{\_}1^{n-1-j}\\ \vdots&&\vdots&\vdots&\vdots&&\vdots\\ U_{\_}n^{j}&\cdots&U_{\_}n&1&V_{\_}n&\cdots&V_{\_}n^{n-1-j}\end{matrix}\right).

The second part of the theorem implies Theorem 1.9.

Proof.

a) By Lemma 6.27 we can write

𝓗𝓕𝓛(S1×S2,Z_n)Cone[𝓗𝓕𝓛(S3,O_n)Φ𝓗𝓕𝓛(S3,T(n,n))]\bm{{\mathcal{H\!F\!L}}}(S^{1}\times S^{2},Z_{\_}n)\simeq\mathrm{Cone}\left[\bm{{\mathcal{H\!F\!L}}}(S^{3},O_{\_}n)\xrightarrow{\Phi}\bm{{\mathcal{H\!F\!L}}}(S^{3},T(n,n))\right]

Since all ϕ_kn\phi_{\_}k^{n} are injective on homology and have pairwise different Alexander degrees, Φ=_kϕ_kn\Phi=\sum_{\_}{k\in\mathbb{Z}}\phi_{\_}k^{n} is injective as well, and we can write

𝓗𝓕𝓛(S1×S2,Z_n)𝓗𝓕𝓛(S3,T(n,n))/Im(Φ).\bm{{\mathcal{H\!F\!L}}}(S^{1}\times S^{2},Z_{\_}n)\simeq\bm{{\mathcal{H\!F\!L}}}(S^{3},T(n,n))/\mathrm{Im}(\Phi).

Now part (a) follows from Example 3.12 and Proposition 3.13.

b) By the proof of Theorem 6.11 the map Ω0\Omega^{0} provides an isomorphism (S3,T(n,n))𝒥\mathcal{H\!F\!L}(S^{3},T(n,n))\simeq\mathcal{J} and

Ω0(a_i)=±det(U_1jU_11V_1V_1n1jU_njU_n1V_nV_nn1j),0in1,\Omega^{0}(a_{\_}i)=\pm\det\left(\begin{matrix}U_{\_}1^{j}&\cdots&U_{\_}1&1&V_{\_}1&\cdots&V_{\_}1^{n-1-j}\\ \vdots&&\vdots&\vdots&\vdots&&\vdots\\ U_{\_}n^{j}&\cdots&U_{\_}n&1&V_{\_}n&\cdots&V_{\_}n^{n-1-j}\end{matrix}\right),\quad 0\leq i\leq n-1,

in particular

Ω0(a_0)=±det(1V_1V_1n11V_nV_nn1)=±_i<j(V_iV_j)\Omega^{0}(a_{\_}0)=\pm\det\left(\begin{matrix}1&V_{\_}1&\cdots&V_{\_}1^{n-1}\\ \vdots&\vdots&&\vdots\\ 1&V_{\_}n&\cdots&V_{\_}n^{n-1}\end{matrix}\right)=\pm\prod_{\_}{i<j}(V_{\_}i-V_{\_}j)

and

Ω0(a_n1)=±det(U_1n1U_11U_nn1U_n1)=±_i<j(U_iU_j).\Omega^{0}(a_{\_}{n-1})=\pm\det\left(\begin{matrix}U_{\_}1^{n-1}&\cdots&U_{\_}1&1\\ \vdots&&\vdots&\vdots\\ U_{\_}n^{n-1}&\cdots&U_{\_}n&1\end{matrix}\right)=\pm\prod_{\_}{i<j}(U_{\_}i-U_{\_}j).

Therefore

Ω0(μ_0a_0+a_1++a_n2+μ_n1a_n1)=γ\Omega^{0}(\mu_{\_}0a_{\_}0+a_{\_}1+\ldots+a_{\_}{n-2}+\mu_{\_}{n-1}a_{\_}{n-1})=\gamma

and the result follows. ∎

Remark 6.30.

It would be interesting to compare this result to the recent work of Kronheimer and Mrowka [16]. Their main result expresses the (deformed) instanton homology I(Z_n,Γ)I(Z_{\_}n,\Gamma) of the nn-component unlink Z_nZ_{\_}n in S1×S2S^{1}\times S^{2} (with local coefficients) as a quotient of the polynomial ring by a certain determinantal ideal 𝒥_inst\mathcal{J}_{\_}{inst}.

References

  • [1] A. Alishahi, E. Eftekhary. A refinement of sutured Floer homology. J. Symplectic Geom. 13 (2015), no. 3, 609–743.
  • [2] A. Alishahi, E. Eftekhary. Tangle Floer homology and cobordisms between tangles. J. Topol. 13 (2020), no. 4, 1582–1657.
  • [3] J. Batson and C. Seed. A link-splitting spectral sequence in Khovanov homology. Duke Math. J. 164.5 (2015), pp. 801–841.
  • [4] M. Beck, S. Robins. Computing the continuous discretely. Integer-point enumeration in polyhedra. Second edition. With illustrations by David Austin. Undergraduate Texts in Mathematics. Springer, New York, 2015.
  • [5] A. Beliakova, K. Putyra, L.-H. Robert, E. Wagner. A proof of Dunfield-Gukov-Rasmussen Conjecture. arXiv:2210.00878
  • [6] M. Borodzik, E. Gorsky. Immersed concordances of links and Heegaard Floer homology. Indiana Univ. Math. J. 67 (2018), no. 3, 1039–1083.
  • [7] M. Borodzik, B. Liu, I. Zemke. Lattice homology, formality, and plumbed L–space links. arXiv:2210.15792
  • [8] M. Borodzik, B. Liu, I. Zemke. Heegaard Floer homology and plane curves with non-cuspidal singularities. arXiv:2104.13709
  • [9] S. Cautis, J.Kamnitzer. Knot homology via derived categories of coherent sheaves IV, coloured links. Quantum Topol. 8 (2017), no. 2, 381–411.
  • [10] N. Dowlin. A Categorification of the HOMFLY-PT Polynomial with a Spectral Sequence to Knot Floer Homology, 2017. arXiv:1703.01401.
  • [11] A. Gilmore. Invariance and the knot Floer cube of resolutions. Quantum Topol., 7(1):107–183, 2016.
  • [12] E. Gorsky, M. Hogancamp. Hilbert schemes and yy-ification of Khovanov-Rozansky homology. Geom. Topol. 26 (2022), no. 2, 587–678.
  • [13] E. Gorsky, J. Hom. Cable links and L–space surgeries. Quantum Topol. 8 (2017), no. 4, 629–666.
  • [14] M. Haiman. t,qt,q-Catalan numbers and the Hilbert scheme. Discrete Math. 193 (1998), no. 1–3, 201–224.
  • [15] M. Hogancamp, D. E. V. Rose, P. Wedrich. Link splitting deformation of colored Khovanov–Rozansky homology. arXiv: 2107.09590
  • [16] P. Kronheimer, T. Morwka. Relations in singular instanton homology. arXiv:2210.07059
  • [17] Y. Lekili, T. Perutz. Fukaya categories of the torus and Dehn surgery. Proc. Natl. Acad. Sci. USA 108 (2011), no. 20, 8106–8113.
  • [18] Y. Liu. L-space surgeries on links. Quantum Topol. 8 (2017), no. 3, 505–570.
  • [19] Macdonald, I. G. Symmetric functions and Hall polynomials. Second edition. With contributions by A. Zelevinsky. Oxford Mathematical Monographs. Oxford Science Publications. The Clarendon Press, Oxford University Press, New York, 1995.
  • [20] C. Manolescu, P. Ozsváth. Heegaard Floer homology and integer surgeries on links. arXiv:1011.1317
  • [21] P. Ozsváth, Z. Szabó. A cube of resolutions for knot Floer homology. J. Topol. 2 (2009), no. 4, 865–910.
  • [22] P. Ozsváth, Z. Szabó. Absolutely graded Floer homologies and intersection forms for four-manifolds with boundary. Adv. in Math. 173 (2003), no. 2, 179–261.
  • [23] P. Ozsváth, Z. Szabó. Holomorphic disks and knot invariants. Adv. Math. 186 (2004), no. 1, 58–116.
  • [24] P. Ozsváth, Z. Szabó. Holomorphic disks and three-manifold invariants: properties and applications. Annals of Mathematics, 159 (2004), 1159–1245.
  • [25] P. Ozsváth, Z. Szabó. Holomorphic disks and topological invariants for closed three-manifolds. Ann. Math. 159 (2004), 1027–1158.
  • [26] P. Ozsváth, Z. Szabó. Holomorphic disks, link invariants and the multi-variable Alexander polynomial. Algebr. Geom. Topol. 8 (2008), no. 2, 615–692.
  • [27] P. Ozsváth, Z. Szabó. Holomorphic triangles and invariants for smooth four-manifolds. Adv. Math. 202 (2006), no. 2, 326–400.
  • [28] P. Ozsváth, Z. Szabó. Lectures on Heegaard Floer homology. Clay Math. Proc. 5 (2007), 29–70.
  • [29] P. Ozsváth, Z. Szabó. On the Heegaard Floer homology of branched double-covers. Adv. Math. 194 (2005), no. 1, 1–33.
  • [30] P. Ozsváth, Z. Szabó. On the skein exact sequence for knot Floer homology. arXiv:0707.1165.
  • [31] I. Zemke. Graph cobordisms and Heegaard Floer homology. arXiv:1512.01184.
  • [32] I. Zemke. Link cobordisms and functoriality in link Floer homology. J. Topol. 12 (2019), no. 1, 94–220.
  • [33] I. Zemke. Link cobordisms and absolute gradings on link Floer homology. Quantum Topol. 10 (2019), no. 2, 207–323.
  • [34] I. Zemke. Quasistabilization and basepoint moving maps in link Floer homology. Algebraic & Geometric Topology 17 (2017), no. 6, 3461 – 3518.