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

Almost-extreme Khovanov spectra

Federico Cantero Morán and Marithania Silvero [email protected], [email protected] Departamento de Matemáticas, Universidad Autónoma de Madrid and ICMAT, Madrid, Spain Departamento de Matemática Aplicada, Universidad de Sevilla and IMUS, Sevilla, Spain
Abstract.

We introduce a functor from the cube to the Burnside 22-category and prove that it is equivalent to the Khovanov spectrum given by Lipshitz and Sarkar in the almost-extreme quantum grading. We provide a decomposition of this functor into simplicial complexes. This decomposition allows us to compute the homotopy type of the almost-extreme Khovanov spectra of diagrams without alternating pairs.

2010 Mathematics Subject Classification:
Primary 57M25, 55P42
Both authors were supported by the Spanish projects MTM2016-76453-C2, MDM-2014-0445 and FEDER. The first author was partially supported by project PID2019-108936GB-C21. The second author was partially supported by project US-1263032.

1. Introduction

Khovanov homology is a powerful link invariant introduced by Mikhail Khovanov in [Kho00] as a categorification of the Jones polynomial. More precisely, given an oriented diagram DD representing a link LL, he constructed a finite \mathbb{Z}-graded family of chain complexes

\textstyle{\ldots\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Ci,j(D)\textstyle{C^{i,j}(D)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}di\scriptstyle{d_{i}}Ci+1,j(D)\textstyle{C^{i+1,j}(D)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}di+1\scriptstyle{d_{i+1}}Ci+2,j(D)\textstyle{C^{i+2,j}(D)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}\textstyle{\ldots}

whose bigraded homology groups, Khi,j(D)Kh^{i,j}(D), are link invariants. The groups Khi,j(L)Kh^{i,j}(L) are known as Khovanov homology groups of LL, and the indexes ii and jj as homological and quantum gradings, respectively.

In [LS14] Lipshitz and Sarkar refined this invariant to obtain, for each quantum grading jj, a spectrum 𝒳Dj\mathcal{X}_{D}^{j} whose stable homotopy type is a link invariant and whose cohomology is isomorphic to the Khovanov homology of the link, i.e., H(𝒳Dj)Kh,j(L)H^{*}(\mathcal{X}_{D}^{j})\cong Kh^{*,j}(L). Together with Lawson [LLS20] they gave a neat construction of this spectrum as the realization of a 22-functor FjF^{j} from the cube category to the Burnside 22-category (see also [LLS17, LS18]).

Shortly after, another geometric realization of the (minimal) extreme Khovanov homology was given [GMS18] in terms of an independence simplicial complex constructed from the link diagram. This construction was later shown to coincide with the one of Lipshitz and Sarkar [CMS20] as follows: functors from the cube to the Burnside 22-category can be understood as generalizations of simplicial complexes; such a functor is a simplicial complex if it factors through the category of sets and takes values on singletons and the empty set.

On the other hand, the Khovanov spectrum of a link in its maximal extreme quantum grading is Spanier-Whitehead dual to that of its mirror image in minimal grading. This construction was exploded in [PS18] to make explicit computations of the Khovanov spectrum in (maximal) extreme quantum grading.

In 2018 the second author, together with Przytycki applied these techniques one step further, to the almost-extreme quantum grading. They restricted to the study of 11-adequate link diagrams, i.e., those whose 11-resolution contains no chords with both endpoints in the same circle (the extremal homology of these diagrams has been computed in [PS14, SS20, DL20]), and built a pointed semi-simplicial set whose homology is isomorphic to the almost-extreme Khovanov homology of the diagram:

Theorem ([PS20]).

If DD is a 11-adequate link diagram, then

  • (i)

    The functor FDjalmaxF_{D}^{{j_{\mathrm{almax}}}} gives rise to a pointed semi-simplicial set.

  • (ii)

    The realization of FDjalmaxF_{D}^{{j_{\mathrm{almax}}}} has the homotopy type of a wedge of spheres and possibly a copy of a (de)suspension of P2\mathbb{R}P^{2}.

Their construction relates to the functor of Lawson, Lipshitz and Sarkar as follows: the category of pointed sets includes into the Burnside 22-category, and if a functor from the cube to the Burnside 22-category factors through it, then it gives rise to a pointed (augmented) semi-simpicial set.

Simplicial complexes and pointed semi-simplicial sets are simpler objects than functors to the Burnside 22-category, to which many classical tools can be applied, so these results prompt the question

Are there simpler models of the spectra of Lipshitz and Sarkar in the almost-extreme quantum grading?

Unfortunately, it turns out that FDjalmaxF^{j_{\mathrm{almax}}}_{D} gives rise to a pointed semi-simplicial set if and only if DD is 11-adequate. However, in this paper we overcome this problem by introducing a new functor MDM_{D}, whose realization coincides with that of FDjalmaxF^{j_{\mathrm{almax}}}_{D}, and gives rise to an (augmented) pointed semi-simplicial set in a broader number of cases. This is achieved by constructing a natural transformation between MDM_{D} and FDjalmaxF^{{j_{\mathrm{almax}}}}_{D} and proving the following result:

Theorem A.

The natural transformation is a homology isomorphism, and therefore the realizations of MDM_{D} and FDjalmaxF^{{j_{\mathrm{almax}}}}_{D} are homotopy equivalent.

The functor MDM_{D} can be understood as a “change of basis” of FDjalmaxF_{D}^{{j_{\mathrm{almax}}}}: The relevant generators of the Khovanov chain complex in its almost-extreme degree (which is the same as the chain complex of FDjalmaxF^{j_{\mathrm{almax}}}_{D}) are indexed by the circles of the state resolutions of DD, whereas the generators of the chain complex associated to MDM_{D} are indexed by the edges and connected components of certain graph whose vertices are the circles of the state resolutions. A crucial property at the almost-extreme quantum grading is that the relevant involved graphs are forests, and therefore both basis have the same size (as it should be).

The functor MDM_{D} is simpler than FDjF^{j}_{D} in the sense that it takes values in morphisms sending generators to (multiples of) generators, whereas FDjF^{j}_{D} takes values in morphisms sending generators to sums of possibly different generators.

This simplified new basis allows us to decompose MDM_{D} in terms of independence simplicial complexes of simpler link diagrams (Proposition 7.5 and Remark 7.6). The last part of the paper is devoted to make explicit computations using this decomposition.

One of the advantages of MDM_{D} over FDjalmaxF^{j_{\mathrm{almax}}}_{D} is that it gives rise to an augmented pointed semi-simplicial set if and only if the 11-resolution of the diagram contains no alternating pairs (i.e., two chords whose endpoints alternate along the same circle). Moreover, the forementioned decomposition allows us to compute the homotopy type of the realization of MDM_{D} in these cases:

Theorem B.

Let DD be a diagram whose 11-resolution contains no alternating pairs.

  • (i)

    The functor MDM_{D} gives rise to an augmented pointed semi-simplicial set.

  • (ii)

    If DD is not 11-adequate then the realization of MDM_{D} is homotopy equivalent to a wedge of spheres.

In [PS18] the second author conjectured (and proved for several cases) that the extreme Khovanov spectrum is always a wedge of spheres. If this conjecture were true, then the above decomposition would give an upper bound for the cone length of the spectrum 𝒳Djalmax\mathcal{X}^{j_{\mathrm{almax}}}_{D} (Remark 7.8). This bound seems to be far from optimal, as the computations of this paper and [PS20] produce spectra of cone length at most 22, so we rise the following question:

Are there diagrams whose almost-extreme Khovanov spectrum is not homotopy equivalent to a wedge of spheres and possibly (de)suspensions of projective planes?

The structure of the paper is as follows. In Sections 2 and 3 we recall the categorical notions that will be used along the paper. In Section 4 we develop some results on link diagrams which allow, in Section 5, to introduce the functor MDM_{D} and prove Theorem A. In Section 6 we prove the first part of Theorem B. In Section 7 we decompose MDM_{D} in terms of independence simplicial complexes and present three skein sequences which allow us to determine, in Section 8, the homotopy type of MDM_{D} for all links with no alternating pairs, proving the second part of Theorem B.

Acknowledgments

The authors thank Carles Casacuberta for his guidance and advice at early stages of this project.

2. The Burnside 2-category

A 22-category is a higher category that generalizes the notion of category, in the sense that on top the objects and morphisms there are also 22-morphisms. More precisely, a 22-category consists of the data of

  1. (1)

    a collection of objects,

  2. (2)

    for each pair of objects X,YX,Y, a category of morphisms hom(X,Y)\hom(X,Y),

  3. (3)

    for each object XX, an object IdX\mathrm{Id}_{X} of hom(X,X)\hom(X,X) and

  4. (4)

    for each triple of objects X,Y,ZX,Y,Z, a composition functor

    hom(Y,Z)×hom(X,Y)hom(X,Z)\hom(Y,Z)\times\hom(X,Y)\to\hom(X,Z)

which is associative and unital. The objects in hom(X,Y)\hom(X,Y) are called 11-morphisms and the morphisms in hom(X,Y)\hom(X,Y) are called 22-morphisms.

Recall that given two finite sets X,YX,Y, a span from XX to YY is a triple (Q,s,t)(Q,s,t) where QQ is a finite set, and ss and tt are a pair of functions X𝑠Q𝑡YX\overset{s}{\longleftarrow}Q\overset{t}{\longrightarrow}Y.

We say that a span is free (resp. very free) if the source map ss is injective (resp. bijective).

Given two spans X𝑠Q𝑡YX\overset{s}{\longleftarrow}Q\overset{t}{\longrightarrow}Y and XsQtYX\overset{s^{\prime}}{\longleftarrow}Q^{\prime}\overset{t^{\prime}}{\longrightarrow}Y, a fibrewise bijection between them is a bijection τ:QQ\tau\colon Q\to Q^{\prime} such that sτ=ss^{\prime}\circ\tau=s and tτ=tt^{\prime}\circ\tau=t:

Q\textstyle{Q\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}s\scriptstyle{s}t\scriptstyle{t}τ\scriptstyle{\tau}X\textstyle{X}Y\textstyle{Y}Q\textstyle{Q^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}s\scriptstyle{s^{\prime}}t\scriptstyle{t^{\prime}}

The composition of two fibrewise bijections of spans is the composition of bijections, and the identity morphism of a locally finite span is the identity bijection. This defines a category of spans from the set XX to the set YY, which we note by (X,Y)\mathcal{B}(X,Y).

Definition 2.1.

The Burnside 22-category, denoted by \mathcal{B}, is the 22-category whose objects are finite sets, and the category of morphisms from a set XX to a set YY is given by (X,Y)\mathcal{B}(X,Y), the category of spans from XX to YY; in other words, the 11-morphisms are given by spans and 22-morphisms are fibrewise bijections.

Let X1,X2X_{1},X_{2} and X3X_{3} be three sets in \mathcal{B}. Given a span (Q1,s1,t1)(Q_{1},s_{1},t_{1}) in (X1,X2)\mathcal{B}(X_{1},X_{2}) and a span (Q2,s2,t2)(Q_{2},s_{2},t_{2}) in (X2,X3)\mathcal{B}(X_{2},X_{3}), their composition is given by their fibre product (Q1×X2Q2,s,t)(Q_{1}\times_{X_{2}}Q_{2},s,t):

Q1×X2Q2\textstyle{Q_{1}\times_{X_{2}}Q_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}s\scriptstyle{s}t\scriptstyle{t}Q1\textstyle{Q_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}s1\scriptstyle{s_{1}}t1\scriptstyle{t_{1}}Q2\textstyle{Q_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}s2\scriptstyle{s_{2}}t2\scriptstyle{t_{2}}X1\textstyle{X_{1}}X2\textstyle{X_{2}}X3.\textstyle{X_{3}.}

The composition of two fibrewise bijections τ1:Q1Q1\tau_{1}\colon Q_{1}\to Q_{1}^{\prime} and τ2:Q2Q2\tau_{2}\colon Q_{2}\to Q_{2}^{\prime} is also given by their fibre product τ1×τ2:Q1×X2Q2Q1×X2Q2\tau_{1}\times\tau_{2}\colon Q_{1}\times_{X_{2}}Q_{2}\to Q^{\prime}_{1}\times_{X_{2}}Q^{\prime}_{2} as follows

Q1\textstyle{Q_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ1\scriptstyle{\tau_{1}}Q2\textstyle{Q_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ2\scriptstyle{\tau_{2}}X1\textstyle{X_{1}}X2\textstyle{X_{2}}X3\textstyle{X_{3}}Q1\textstyle{Q^{\prime}_{1}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Q2\textstyle{Q^{\prime}_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}      Q1×X2Q2\textstyle{Q_{1}\times_{X_{2}}Q_{2}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}τ1×τ2\scriptstyle{\tau_{1}\times\tau_{2}}X\textstyle{X}Y,\textstyle{Y,}Q1×X2Q2,\textstyle{Q^{\prime}_{1}\times_{X_{2}}Q^{\prime}_{2},\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}

and the identity on a set XX is the span XIdXIdXX\overset{\mathrm{Id}}{\leftarrow}X\overset{\mathrm{Id}}{\to}X.

Remark 2.2.

Note that a span 𝔰=(Q,s,t)\mathfrak{s}=(Q,s,t) from XX to YY is determined by the matrix whose entries are the sets 𝔰x,y:=s1(x)t1(y)\mathfrak{s}_{x,y}:=s^{-1}(x)\cap t^{-1}(y), for every xXx\in X, yYy\in Y. Giving a fibrewise bijection τ\tau from a span 𝔰\mathfrak{s} to a span 𝔰\mathfrak{s}^{\prime} is the same as giving, for each xXx\in X and each yYy\in Y, a bijection τx,y\tau_{x,y} from 𝔰x,y\mathfrak{s}_{x,y} to 𝔰x,y\mathfrak{s}^{\prime}_{x,y}. The composition of 𝔰:XY\mathfrak{s}\colon X\to Y and 𝔰:YZ\mathfrak{s}^{\prime}\colon Y\to Z is the span 𝔰′′\mathfrak{s}^{\prime\prime} with 𝔰x,z′′=yY𝔰x,y×𝔰y,z\mathfrak{s}^{\prime\prime}_{x,z}=\bigcup_{y\in Y}\mathfrak{s}_{x,y}\times\mathfrak{s}^{\prime}_{y,z}. The span 𝔰\mathfrak{s} is determined by the formal sum

𝔰(x)=yY𝔰x,yy,\mathfrak{s}(x)=\sum_{y\in Y}\mathfrak{s}_{x,y}\cdot y,

for each xXx\in X. If a coefficient 𝔰x,y\mathfrak{s}_{x,y} is not specified, we understand that 𝔰x,y={x}\mathfrak{s}_{x,y}=\{x\} (for example, if 𝔰(x)=y+z\mathfrak{s}(x)=y+z, then 𝔰(x)={x}y+{x}z\mathfrak{s}(x)=\{x\}\cdot y+\{x\}\cdot z).

The category of finite sets 𝐒𝐞𝐭\mathbf{Set} can be mapped into the category of pointed finite sets 𝐒𝐞𝐭\mathbf{Set}_{\bullet} by sending XX to X+:=X{}X_{+}:=X\cup\{\mathbf{\ast}\}, and a morphism f:XYf\colon X\to Y to the morphism f+f_{+} that coincides with ff on XX and sends the basepoint of X+X_{+} to the basepoint of Y+Y_{+}. Moreover, the category 𝐒𝐞𝐭\mathbf{Set}_{\bullet} sits inside \mathcal{B} by sending a pointed finite set XX to X{}X\smallsetminus\{\mathbf{\ast}\}, and a morphism f:XYf\colon X\to Y to the span

X{}Xf1()𝑓Y{}.X\smallsetminus\{\mathbf{\ast}\}\hookleftarrow X\smallsetminus f^{-1}(\mathbf{\ast})\overset{f}{\to}Y\smallsetminus\{\mathbf{\ast}\}.

The inclusions 𝐒𝐞𝐭𝐒𝐞𝐭\mathbf{Set}\,\hookrightarrow\,\mathbf{Set}_{\bullet}\,\hookrightarrow\,\mathcal{B} induce equivalences between the following 22-subcategories of \mathcal{B}:

  • The essential image of 𝐒𝐞𝐭\mathbf{Set}_{\bullet} in \mathcal{B} is the 22-subcategory of free spans.

  • The essential image of 𝐒𝐞𝐭\mathbf{Set} in \mathcal{B} is the 22-subcategory of very free spans.

Fix now a commutative ring RR with unit, and let RR-Mod be the category of modules over RR. There is a functor

(2.1) 𝒜:R-Mod\mathcal{A}\colon\mathcal{B}\to\text{$R$-Mod}

sending the finite set XX to the free RR-module RXR\langle X\rangle, and a span 𝔰:XY\mathfrak{s}\colon X\to Y to the homomorphism f:RXRYf\colon R\langle X\rangle\to R\langle Y\rangle given, for every xXx\in X, by

f(x)=yY|𝔰x,y|y.f(x)=\sum_{y\in Y}|\mathfrak{s}_{x,y}|\cdot y.

Note that every pair of spans connected by a 22-morphism are sent to the same homomorphism, and therefore the functor 𝒜\mathcal{A} is well-defined.

3. Cubes of pointed sets and augmented semi-simplicial pointed sets

The nn-dimensional cube 𝟐𝐧\mathbf{2^{n}} is the partially ordered set (poset) whose elements are nn-tuples u=(u1,,un){0,1}nu=(u_{1},\ldots,u_{n})\in\{0,1\}^{n} endowed with the standard partial order so that uvu\geqslant v if uiviu_{i}\geqslant v_{i} for all ii. Following [LLS20], we regard 𝟐𝐧\mathbf{2^{n}} as the category whose objects are the elements of this poset and, for any two such elements u,vu,v, the morphism set Hom(u,v)\operatorname{Hom}(u,v) has a single element φu,v\varphi_{u,v} if uvu\geqslant v, and is empty otherwise. This category has an initial element 1=(1,1,,1)\vec{1}=(1,1,\ldots,1) and a terminal element 0=(0,0,,0)\vec{0}=(0,0,\ldots,0).

The poset map ||:𝟐𝐧|\cdot|\colon\mathbf{2^{n}}\to\mathbb{N} given by |u|=iui|u|=\sum_{i}u_{i} assigns a grading to each vertex in the cube. We write uvu\succ v if u>vu>v and |uv|=1|u-v|=1, and we write uivu\succ_{i}v if, additionally, uu and vv differ in the iith-coordinate, i.e., if ui=1u_{i}=1, vi=0v_{i}=0 and uj=vju_{j}=v_{j} for iji\neq j.

Given nn\in\mathbb{Z}, n1n\geqslant-1, the finite ordinal [n][n] is the linearly ordered set {0<1<<n}\{0<1<\ldots<n\}. The augmented semi-simplicial category, Δinj\Delta_{\mathrm{inj*}}, has as objects these finite ordinals, and as morphisms injective order-preserving maps between them. The inclusion i:[n1][n]\partial_{i}\colon[n-1]\hookrightarrow[n] that forgets the ii-th element is called iith-face map.

Definition 3.1.

Given a category 𝒞\mathcal{C}, let 𝒞op\mathcal{C}^{\mathrm{op}} denote its opposite category, obtained by reversing the morphisms. An nn-dimensional cube of pointed (finite) sets is a functor F:𝟐𝐧𝐒𝐞𝐭F\colon\mathbf{2^{n}}\to\mathbf{Set}_{\bullet}, while an augmented semi-simplicial pointed (finite) set is a functor X:Δinjop𝐒𝐞𝐭X\colon\Delta_{\mathrm{inj*}}^{\mathrm{op}}\to\mathbf{Set}_{\bullet}. As usual, we write XnX_{n} and i\partial_{i} for X([n])X([n]) and X(i)X(\partial_{i}), respectively.

These two categories are related by a functor λ:𝟐𝐧Δinjop\lambda\colon\mathbf{2^{n}}\to\Delta_{\mathrm{inj*}}^{\mathrm{op}} which maps every vertex of the cube uu to the ordinal [|u|1][|u|-1], and every morphism uivu\succ_{i}v to the opposite of the (j<iui)\left(\sum_{j<i}u_{i}\right)th-face map.

Taking left Kan extension along the functor λ:𝟐𝐧Δinjop\lambda\colon\mathbf{2^{n}}\to\Delta_{\mathrm{inj*}}^{\mathrm{op}} defines a functor

(3.1) Λ:𝐒𝐞𝐭𝟐𝐧𝐒𝐞𝐭Δinjop,\Lambda\colon\mathbf{Set}_{\bullet}^{\mathbf{2^{n}}}\longrightarrow\mathbf{Set}_{\bullet}^{\Delta_{\mathrm{inj*}}^{\mathrm{op}}},

whose value on a cube of pointed sets FF is defined explicitely as follows: its set of kk-simplices is Λ(F)k=|u|=k+1F(u)\Lambda(F)_{k}=\coprod_{|u|=k+1}F(u), for every u𝟐𝐧u\in\mathbf{2^{n}}, and the iith-face map i:Λ(F)kΛ(F)k1\partial_{i}\colon\Lambda(F)_{k}\to\Lambda(F)_{k-1} is the union |u|=k+1F(u>u[i])\coprod_{|u|=k+1}F(u>u[i]), where u[i]u[i] is the vertex of the cube 𝟐𝐧\mathbf{2^{n}} obtained by replacing the (i+1)(i+1)th one in uu by a zero.

Recall from [LLS17, Definition 4.1] that a strictly unitary lax 2-functor FF from the cube 𝟐𝐧\mathbf{2^{n}} to the Burnside category \mathcal{B} consists of the data:

  1. (1)

    a finite set F(v)Ob()F(v)\in\operatorname{Ob}(\mathcal{B}), for each v{0,1}nv\in\{0,1\}^{n}.

  2. (2)

    a finite span F(φu,v):F(u)F(v)F(\varphi_{u,v}):F(u)\to F(v) in \mathcal{B}, for each morphism φu,v\varphi_{u,v} in 𝟐𝐧\mathbf{2^{n}}.

  3. (3)

    a 22-isomorphism Fu,v,wF_{u,v,w} from F(φv,w)F(φu,v)F(\varphi_{v,w})\circ F(\varphi_{u,v}) to F(φu,w)F(\varphi_{u,w}), for each decomposition φu,w=φv,wφu,v\varphi_{u,w}=\varphi_{v,w}\circ\varphi_{u,v};

satisfying that, for every u>v>w>zu>v>w>z, the following diagram commutes:

F(φw,z)×F(w)F(φv,w)×F(v)F(φu,v)\textstyle{F(\varphi_{w,z})\times_{F(w)}F(\varphi_{v,w})\times_{F(v)}F(\varphi_{u,v})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Id×Fu,v,w\scriptstyle{\,\operatorname{Id}\times F_{u,v,w}\,}Fv,w,z×Id\scriptstyle{F_{v,w,z}\times\operatorname{Id}}F(φw,z)×F(w)F(φu,w)\textstyle{F(\varphi_{w,z})\times_{F(w)}F(\varphi_{u,w})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fu,w,z\scriptstyle{F_{u,w,z}}F(φv,z)×F(v)F(φu,v)\textstyle{F(\varphi_{v,z})\times_{F(v)}F(\varphi_{u,v})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fu,v,z\scriptstyle{F_{u,v,z}}F(φu,z).\textstyle{F(\varphi_{u,z}).}

In practice, when defining a strictly unitary lax 2-functor, we use the following result, from [LLS17, Section 4]. We keep the notation u\textstyle{u\ignorespaces\ignorespaces\ignorespaces\ignorespaces}v\textstyle{v} for the morphism uvu\succ v:

Lemma 3.2.

A strictly unitary lax 22-functor from the cube category 𝟐𝐧\mathbf{2^{n}} to the Burnside category \mathcal{B} is uniquely determined (up to natural isomorphism) by the following data:

  1. (D1)

    for each vertex v{0,1}nv\in\{0,1\}^{n}, a finite set F(v)Ob()F(v)\in\operatorname{Ob}(\mathcal{B});

  2. (D2)

    for each uvu\succ v, a finite span F(φu,v):F(u)F(v)F(\varphi_{u,v}):F(u)\to F(v) in \mathcal{B};

  3. (D3)

    for each two-dimensional face
    v\textstyle{v\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u\textstyle{u\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w\textstyle{w}v\textstyle{v^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
    of the cube, a 22-morphism

    Fu,v,v,w:F(φv,w)×F(v)F(φu,v)F(φv,w)×F(v)F(φu,v),F_{u,v,v^{\prime},w}\colon F(\varphi_{v,w})\times_{F(v)}F(\varphi_{u,v})\to F(\varphi_{v^{\prime},w})\times_{F(v^{\prime})}F(\varphi_{u,v^{\prime}}),

satisfying the following two conditions:

  1. (C1)

    for every two-dimensional face
    v\textstyle{v\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u\textstyle{u\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w\textstyle{w}v\textstyle{v^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
    , Fu,v,v,w=Fu,v,v,w1,F_{u,v^{\prime},v,w}=F_{u,v,v^{\prime},w}^{-1},

  2. (C2)

    for every three-dimensional face
    v\textstyle{v\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w′′\textstyle{w^{\prime\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}u\textstyle{u\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}v\textstyle{v^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w\textstyle{w^{\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces}z\textstyle{z}v′′\textstyle{v^{\prime\prime}\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}w\textstyle{w\ignorespaces\ignorespaces\ignorespaces\ignorespaces}
    , the following commutes:

    F(φw′′,z)×F(w′′)F(φv,w′′)×F(v)F(φu,v)\textstyle{F(\varphi_{w^{\prime\prime},z})\times_{F(w^{\prime\prime})}F(\varphi_{v,w^{\prime\prime}})\times_{F(v)}F(\varphi_{u,v})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fv,w′′,w,z×Id\scriptstyle{F_{v,w^{\prime\prime},w^{\prime},z}\times\mathrm{Id}}Id×Fu,v,v,w′′\scriptstyle{\mathrm{Id}\times F_{u,v,v^{\prime},w^{\prime\prime}}}F(φw,z)×F(w)F(φv,w)×F(v)F(φu,v)\textstyle{F(\varphi_{w^{\prime},z})\times_{F(w^{\prime})}F(\varphi_{v,w^{\prime}})\times_{F(v)}F(\varphi_{u,v})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Id×Fu,v,v′′,w\scriptstyle{\mathrm{Id}\times F_{u,v,v^{\prime\prime},w^{\prime}}}F(φw′′,z)×F(w′′)F(φv,w′′)×F(v)F(φu,v)\textstyle{F(\varphi_{w^{\prime\prime},z})\times_{F(w^{\prime\prime})}F(\varphi_{v^{\prime},w^{\prime\prime}})\times_{F(v^{\prime})}F(\varphi_{u,v^{\prime}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fv,w′′,w,z×Id\scriptstyle{F_{v^{\prime},w^{\prime\prime},w,z}\times\mathrm{Id}}F(φw,z)×F(w)F(φv′′,w)×F(v′′)F(φu,v′′)\textstyle{F(\varphi_{w^{\prime},z})\times_{F(w^{\prime})}F(\varphi_{v^{\prime\prime},w^{\prime}})\times_{F(v^{\prime\prime})}F(\varphi_{u,v^{\prime\prime}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fv′′,w,w,z×Id\scriptstyle{F_{v^{\prime\prime},w^{\prime},w,z}\times\mathrm{Id}}F(φw,z)×F(w)F(φv,w)×F(v)F(φu,v)\textstyle{F(\varphi_{w,z})\times_{F(w)}F(\varphi_{v^{\prime},w})\times_{F(v^{\prime})}F(\varphi_{u,v^{\prime}})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Id×Fu,v,v′′,w\scriptstyle{\mathrm{Id}\times F_{u,v^{\prime},v^{\prime\prime},w}}F(φw,z)×F(w)F(φv′′,w)×F(v′′)F(φu,v′′).\textstyle{F(\varphi_{w,z})\times_{F(w)}F(\varphi_{v^{\prime\prime},w})\times_{F(v^{\prime\prime})}F(\varphi_{u,v^{\prime\prime}}).}
Definition 3.3.

[LLS17] A natural transformation α:FG\alpha\colon F\to G between two strictly unitary lax 22-functors F,G:𝟐𝐧F,G\colon\mathbf{2^{n}}\to\mathcal{B} is a strictly unitary lax 22-functor H:𝟐𝐧+𝟏H\colon\mathbf{2^{n+1}}\to\mathcal{B} such that H|𝟐𝐧×{1}=FH|_{\mathbf{2^{n}}\times\{1\}}=F and H|𝟐𝐧×{0}=GH|_{\mathbf{2^{n}}\times\{0\}}=G. For every u𝟐𝐧u\in\mathbf{2^{n}}, we denote by αu\alpha_{u} the morphism H(φ(u,1),(u,0))H(\varphi_{(u,1),(u,0)}). For every uv𝟐𝐧u\succ v\in\mathbf{2^{n}}, we denote αu,v\alpha_{u,v} the 22-morphism H(u,1),(u,0),(v,1),(v,0)H_{(u,1),(u,0),(v,1),(v,0)}.

We write 𝒞𝒟\mathcal{C}^{\mathcal{D}} for the category whose objects are strictly unitary lax 22-functors from 𝒟\mathcal{D} to 𝒞\mathcal{C} and whose morphisms are natural transformations between them. When 𝒞=\mathcal{C}=\mathcal{B} and 𝒟=𝟐𝐧\mathcal{D}=\mathbf{2^{n}} the objects are called Burnside cubes.

3.1. Realizations and totalizations

Let 𝒞\mathcal{C} be a model category, as the category of (pointed) topological spaces 𝐓𝐨𝐩\mathbf{Top} (resp. 𝐓𝐨𝐩\mathbf{Top}_{\bullet}), the category of spectra 𝐒𝐩\mathbf{Sp}, or the category Ch(R)\mathrm{Ch}(R) of chain complexes of RR-modules, with RR a commutative ring.

The totalization of a functor F:𝟐𝐧𝒞F\colon\mathbf{2^{n}}\to\mathcal{C} is the object of 𝒞\mathcal{C} given by

TotF=hocofib(hocolim(F|𝟐𝐧{0})F(0)).\operatorname{Tot}F=\mathrm{hocofib}\left(\operatorname*{hocolim}\left(F|_{\mathbf{2^{n}}\smallsetminus\{0\}}\right)\to F(0)\right).

This defines a functor

Tot:𝒞𝟐𝐧𝒞.\operatorname{Tot}\colon\mathcal{C}^{\mathbf{2^{n}}}\longrightarrow\mathcal{C}.

On the other hand, the relative realization of a functor X:Δinjop𝒞X\colon\Delta_{\mathrm{inj*}}^{\mathrm{op}}\to\mathcal{C} is the object in 𝒞\mathcal{C} given by

X=hocofib(hocolim(X|Δinj{[1]})X1),\|X\|=\mathrm{hocofib}\left(\operatorname*{hocolim}\left(X|_{\Delta_{\mathrm{inj*}}\smallsetminus\{[-1]\}}\right)\to X_{-1}\right),

which defines a functor

:𝒞Δinjop𝒞.\|-\|\colon\mathcal{C}^{\Delta_{\mathrm{inj*}}^{\mathrm{op}}}\longrightarrow\mathcal{C}.

A semi-simplicial object X:Δinjop𝒞X\colon\Delta_{\mathrm{inj}}^{\mathrm{op}}\to\mathcal{C} may be seen as an augmented semi-simplicial set by defining X1=X_{-1}=*, the final object of 𝒞\mathcal{C}. Its relative realization is the suspension of the realization of XX:

XΣ|X|.\|X\|\cong\Sigma|X|.

Recall now the functor λ:𝟐𝐧Δinjop\lambda\colon\mathbf{2^{n}}\to\Delta_{\mathrm{inj*}}^{\mathrm{op}} mapping every u𝟐𝐧u\in\mathbf{2^{n}} to [|u|1][|u|-1], and every morphism uivu\succ_{i}v to the opposite of (j<iui)th\left(\sum_{j<i}u_{i}\right)th-face map. Since λ\lambda is cofinal, λ(𝟐𝐧{0})=Δinj{[1]}\lambda(\mathbf{2^{n}}\smallsetminus\{0\})=\Delta_{\mathrm{inj*}}\smallsetminus\{[-1]\} and λ(0)=[1]\lambda(\vec{0})=[-1], we have that

(3.2) Λ(F)TotF.\|\Lambda(F)\|\simeq\operatorname{Tot}F.

3.2. From Burnside cubes to chain complexes

Let RR be a ring, and let Ch(R)\mathrm{Ch}(R) be the category of chain complexes of RR-modules. There is a functor

K:R-ModCh(R),K\colon\mathcal{B}\longrightarrow\text{$R$-Mod}\longrightarrow\mathrm{Ch}(R),

where the first functor is 𝒜\mathcal{A}, defined in (2.1), and the second functor sends an RR-module to the chain complex given by the RR-module concentrated at degree 0.

We define the functor C(;R)C_{*}(-;R) as the composition

C(;R):𝟐𝐧Ch(R)𝟐𝐧TotCh(R),C_{*}(-;R)\colon\mathcal{B}^{\mathbf{2^{n}}}\longrightarrow\mathrm{Ch}(R)^{\mathbf{2^{n}}}\overset{\operatorname{Tot}}{\longrightarrow}\mathrm{Ch}(R),

where the first functor is the result of composing a 22-functor F:𝟐𝐧F\colon\mathbf{2^{n}}\to\mathcal{B} from 𝟐𝐧\mathcal{B}^{\mathbf{2^{n}}} with KK.

Its value at the Burnside cube FF can also be described as the chain complex whose kk-cochains are

Ck(F;R)=u𝟐𝐧,|u|=kRF(u)C_{k}(F;R)=\bigoplus_{u\in\mathbf{2^{n}},|u|=k}R\langle F(u)\rangle

and whose differential is

=|u|=k,|v|=k1(1)j<iui𝒜F(uiv).\partial=\sum_{|u|=k,|v|=k-1}(-1)^{\sum_{j<i}u_{i}}\mathcal{A}\circ F(u\succ_{i}v).

Observe that since R-Mod\mathcal{B}\to\text{$R$-Mod} sends 22-morphisms to identities, a natural transformation between Burnside cubes induces a morphism of chain complexes.

3.3. From Burnside cubes to spectra

In [LLS20], Lawson, Lipshitz and Sarkar gave an explicit construction of the totalization functor for cubes in the Burnside category. They associated, to each cube F:𝟐𝐧F\colon\mathbf{2^{n}}\to\mathcal{B}, a spectrum TotF\operatorname{Tot}F, and to each map f:FGf\colon F\to G of Burnside cubes, a map Totf:TotFTotG\operatorname{Tot}f\colon\operatorname{Tot}F\to\operatorname{Tot}G, both well-defined up to homotopy. Additionally, the homology H(TotF;R)H_{*}(\operatorname{Tot}F;R) was isomorphic to the homology of C(F;R)C_{*}(F;R), and the map induced by Totf\operatorname{Tot}f on homology was the one induced by C(f;R)C_{*}(f;R).

Remark 3.4.

If C(f;R)C_{*}(f;R) is a homology isomorphism then Totf\operatorname{Tot}f is a homotopy equivalence of spectra.

On the other hand, in [CMS20] the authors showed that if a cube F:𝟐𝐧F\colon\mathbf{2^{n}}\to\mathcal{B} factors through some functor F~:𝟐𝐧𝐒𝐞𝐭\tilde{F}\colon\mathbf{2^{n}}\to\mathbf{Set}_{\bullet}, then TotF=ΣTotF~\operatorname{Tot}F=\Sigma^{\infty}\operatorname{Tot}\tilde{F}, where the homotopy colimit is taken in pointed topological spaces. Therefore, by (3.2), we have:

Proposition 3.5.

If a Burnside cube F:𝟐𝐧F\colon\mathbf{2^{n}}\to\mathcal{B} in the Burnside category factors through 𝐒𝐞𝐭\mathbf{Set}_{\bullet}, then the pointed semi-simplicial set Λ(F~)\Lambda(\tilde{F}) satisfies

TotFΣΛ(F~).\operatorname{Tot}F\simeq\Sigma^{\infty}\displaystyle\|\Lambda(\tilde{F})\|.
Remark 3.6.

The spectrum TotF\operatorname{Tot}F is denoted |F||F| in [LLS20]. The chain complex C(F;R)C_{*}(F;R) is denoted TotF\operatorname{Tot}F in [LLS17]. We reserve the notation Tot\operatorname{Tot} for the totalization of a cube, \|\cdot\| for the relative realization of an augmented semi-simplicial object and |||\cdot| for the realization of a semi-simplicial object.

4. Knots and graphs

4.1. States

Let DD be an oriented link diagram with nn ordered crossings {c1<c2<<cn}\{c_{1}<c_{2}<\ldots<c_{n}\}, where n+n_{+} (nn_{-}) of them are positive (negative). A (Kauffman) state of the diagram DD is an assignation of a label, 0 or 11, to each crossing in DD. The order of the crossings induces a bijection between the set of states of DD and the elements of 𝟐𝐧\mathbf{2^{n}} by considering u𝟐𝐧u\in\mathbf{2^{n}} as the state that assigns the label uiu_{i} to cic_{i}.

Smoothing a crossing cic_{i} of DD consists on replacing it by a pair of arcs and a segment connecting them; the way we smooth the crossings depends on its associated label uiu_{i}, as shown in Figure 1. The result of smoothing each crossing of DD according to its label is the chord diagram D(u)D(u), a collection of disjoint circles together with some segments that we call 0- and 11-chords, depending on the value of the associated uiu_{i}. We represent 11-chords as light segments, and 0-chords as dark ones.

Refer to caption
Figure 1. The smoothing of a crossing according to its 0 or 11 label.

We introduce the following notation to refer to the circles and chords of D(u)D(u):

  • Z(u)Z(u) denotes the set of circles of D(u)D(u),

  • E(u)E(u) denotes the set of 0-chords of D(u)D(u),

  • E¯(u)\bar{E}(u) denotes the set of 11-chords of D(u)D(u),

  • ei(u)e_{i}(u) denotes the iith chord in D(u)D(u),

  • 𝒪i(u)\mathcal{O}_{i}(u) denotes the set of circles containing an endpoint of ei(u)e_{i}(u) (thus |𝒪i(u)||\mathcal{O}_{i}(u)| is either equal to 11 or 22). If |𝒪i(u)|=2|\mathcal{O}_{i}(u)|=2 (resp. |𝒪i(u)|=1|\mathcal{O}_{i}(u)|=1), we say that ei(u)e_{i}(u) is a bichord (resp. monochord).

Given a state uu of DD, we define its associated state graph, 𝒢(u)\mathcal{G}(u), as the labelled graph obtained by collapsing each circle of D(u)D(u) to a vertex so that each chord in D(u)D(u) becomes an edge in 𝒢(u)\mathcal{G}(u); each edge inherits a label, 0 or 11, from the associated chord. The circles and chords of D(u)D(u) are in bijection with the vertices and edges of 𝒢(u)\mathcal{G}(u). Therefore, the above notation introduced to refer to the circles and chords of D(u)D(u) will be used to refer to the vertices and edges of 𝒢(u)\mathcal{G}(u). In particular, the set of vertices of 𝒢(u)\mathcal{G}(u) is Z(u)Z(u) and the set of edges labeled by 0 and 11 are E(u)E(u) and E¯(u)\bar{E}(u), respectively.

In this setting, write G(u)G(u) for the subgraph obtained after removing the 11-edges from 𝒢(u)\mathcal{G}(u). See Figure 2 for such an example. We consider loops as length-1 cycles.

Refer to caption
Figure 2. A link diagram DD, the configuration D(u)D(u) corresponding to the state u=(0,1,0,1,1,1,0,0)u=(0,1,0,1,1,1,0,0) and the associated graphs 𝒢(u)\mathcal{G}(u) and G(u)G(u).

Recall that given u,v𝟐𝐧u,v\in\mathbf{2^{n}}, we write uivu\succ_{i}v if both vectors are equal in all but the ithi^{th} coordinate, where ui=1u_{i}=1 and vi=0v_{i}=0. The partial order >> is defined as the transitive closure of the above relation.

If uivu\succ_{i}v, then D(v)D(v) is obtained from D(u)D(u) by doing surgery on D(u)D(u) along the 11-chord ei(u)e_{i}(u) and adding a new chord ei(v)e_{i}(v). Observe that, depending on the cardinality of 𝒪i(u)\mathcal{O}_{i}(u), there are two possible surgeries:

Refer to caption
Figure 3. A merging and a splitting are illustrated in (a)(a) and (b)(b), respectively. The endpoints of some chords have been drawn in order to show the effect of both surgeries in the graphs associated to the states.
  • If |𝒪i(u)|=2|\mathcal{O}_{i}(u)|=2, then |𝒪i(v)|=1|\mathcal{O}_{i}(v)|=1 and D(v)D(v) is obtained from D(u)D(u) by joining two circles into one. The graph G(v)G(v) is obtained from G(u)G(u) by identifying the two vertices in 𝒪i(u)\mathcal{O}_{i}(u) and adding a loop (the edge ei(v)e_{i}(v)) based on the identified vertex, as shown in Figure 3(a)(a). We say that uivu\succ_{i}v a merging.

  • If |𝒪i(u)|=1|\mathcal{O}_{i}(u)|=1, then |𝒪i(v)|=2|\mathcal{O}_{i}(v)|=2 and D(v)D(v) is obtained from D(u)D(u) by splitting one circle into two. In this case, G(u)G(u) is obtained from G(v)G(v) by collapsing the edge ei(v)e_{i}(v), as shown in Figure 3(b)(b). We say that uivu\succ_{i}v is a splitting.

Remark 4.1.

Given two states uu and vv of DD so that u>vu>v, with |u||v|=k|u|-|v|=k, there exist k!k! possible chains

u=u0i1u1i2ikuk=vu=u_{0}\succ_{i_{1}}u_{1}\succ_{i_{2}}\ldots\succ_{i_{k}}u_{k}=v

connecting them. However, the total composition of the one-by-one surgeries induced by each of the possible chains does not depend on the chosen chain.

Given a surgery uivu\succ_{i}v, we compare G(u)G(u) and G(v)G(v):

|E(v)|\displaystyle|E(v)| =|E(u)|+1.\displaystyle=|E(u)|+1.
|Z(v)|\displaystyle|Z(v)| ={|Z(u)|+1if uiv is a splitting,|Z(u)|1if uiv is a merging.\displaystyle=\begin{cases}|Z(u)|+1&\text{if $u\succ_{i}v$ is a splitting,}\\ |Z(u)|-1&\text{if $u\succ_{i}v$ is a merging.}\end{cases}
|π0G(v)|\displaystyle|\pi_{0}G(v)| ={|π0G(u)|if uiv is a splitting, |π0G(u)|if uiv is a merging and 𝒪i(u)π0G(u) is not injective,|π0G(u)|1if uiv is a merging and 𝒪i(u)π0G(u) is injective.\displaystyle=\begin{cases}|\pi_{0}G(u)|&\text{if $u\succ_{i}v$ is a splitting, }\\ |\pi_{0}G(u)|&\text{if $u\succ_{i}v$ is a merging and $\mathcal{O}_{i}(u)\to\pi_{0}G(u)$ is not injective,}\\ |\pi_{0}G(u)|-1&\text{if $u\succ_{i}v$ is a merging and $\mathcal{O}_{i}(u)\to\pi_{0}G(u)$ is injective.}\end{cases}
Definition 4.2.

Given a state u𝟐𝐧u\in\mathbf{2^{n}} with |u|=nk|u|=n-k and a chain

(4.1) 1=u0i1u1i2ikuk=u\vec{1}=u_{0}\succ_{i_{1}}u_{1}\succ_{i_{2}}\ldots\succ_{i_{k}}u_{k}=u

connecting uu to the state 1\vec{1}, we define Φ(u)\Phi(u) as the number of loops in {eij(uj)}j=1k\{e_{i_{j}}(u_{j})\}_{j=1}^{k}. In other words, Φ(u)\Phi(u) counts the number of mergings in the chain.

Equalities above allow us to describe G(u)G(u) in terms of G(1)G(\vec{1}) and Φ(u)\Phi(u):

(4.2) |Z(u)|\displaystyle|Z(u)| =|Z(1)|+n|u|2Φ(u),\displaystyle=|Z(\vec{1})|+n-|u|-2\Phi(u),
(4.3) χ(G(u))\displaystyle\chi(G(u)) =χ(G(1))2Φ(u),\displaystyle=\chi(G(\vec{1}))-2\Phi(u),
(4.4) |π0G(1)|Φ(u)\displaystyle|\pi_{0}G(\vec{1})|-\Phi(u) |π0G(u)||π0G(1)|\displaystyle\leqslant|\pi_{0}G(u)|\leqslant|\pi_{0}G(\vec{1})|

In particular, the above relations imply that Φ(u)\Phi(u) does not depend on the chosen chain connecting the states 1\vec{1} and uu.

Lemma 4.3.

Let u𝟐𝐧u\in\mathbf{2^{n}} be a state of a link diagram DD.

  1. (1)

    If Φ(u)=0\Phi(u)=0, then G(u)G(u) is a forest, namely, it consists of a collection of |Z(1)||Z(\vec{1})| contractible components.

  2. (2)

    If Φ(u)=1\Phi(u)=1, then the number of connected components of G(u)G(u), |π0G(u)||\pi_{0}G(u)|, is either |Z(1)||Z(\vec{1})| or |Z(1)|1|Z(\vec{1})|-1. In addition:

    1. (a)

      If |π0G(u)|=|Z(1)|1|\pi_{0}G(u)|=|Z(\vec{1})|-1, then there is one single cycle in G(u)G(u),

    2. (b)

      If |π0G(u)|=|Z(1)||\pi_{0}G(u)|=|Z(\vec{1})|, then G(u)G(u) contains exactly two cycles 𝔠1\mathfrak{c}_{1} and 𝔠2\mathfrak{c}_{2} sharing a common vertex zz. Moreover, in D(u)D(u) the chords corresponding to the edges of 𝔠1\mathfrak{c}_{1} adjacent to the vertex zz alternate with those of 𝔠2\mathfrak{c}_{2} along the circle zz. Therefore the chords corresponding to the edges of 𝔠1\mathfrak{c}_{1} and the chords corresponding to the edges of 𝔠2\mathfrak{c}_{2} lie in different regions of 2z\mathbb{R}^{2}\smallsetminus z.

Proof.

The assertions concerning the homotopy type of G(u)G(u) follow from (4.2) and (4.3) after computing the Euler characteristic of G(u)G(u) as

|π0G(u)|#(cycles in G(u))=χ(G(u))=χ(G(1))2Φ(u)=|Z(1)|2Φ(u).|\pi_{0}G(u)|-\#(\text{cycles in }G(u))=\chi(G(u))=\chi(G(\vec{1}))-2\Phi(u)=|Z(\vec{1})|-2\Phi(u).

If Φ(u)=0\Phi(u)=0, then |π0G(u)|=|π0G(1)|=|Z(1)||\pi_{0}G(u)|=|\pi_{0}G(\vec{1})|=|Z(\vec{1})| by (4.4), and therefore G(u)G(u) contains no cycles. In the case (2a), |π0G(u)|=|Z(1)|1|\pi_{0}G(u)|=|Z(\vec{1})|-1, and therefore G(u)G(u) has a single cycle, whereas in case (2b), |π0G(u)|=|Z(1)||\pi_{0}G(u)|=|Z(\vec{1})|, and therefore there are two cycles in G(u)G(u).

We prove now that when Φ(u)=1\Phi(u)=1 and |π0G(u)|=|π0G(1)||\pi_{0}G(u)|=|\pi_{0}G(\vec{1})|, the chords of 𝔠1\mathfrak{c}_{1} adjacent to the circle zz alternate with those of 𝔠2\mathfrak{c}_{2} along the boundary of zz.

Consider a chain as (4.1), and let ur1iruru_{r-1}\succ_{i_{r}}u_{r} be the states right before and right after the merging is performed, respectively. Write 𝒪ir(ur1)={z1,z2}\mathcal{O}_{i_{r}}(u_{r-1})=\{z_{1},z_{2}\} and define ε\varepsilon as the (possibly empty) set containing those edges eije_{i_{j}} in the chain so that 𝒪ij(ur1)={z1,z2}\mathcal{O}_{i_{j}}(u_{r-1})=\{z_{1},z_{2}\}, for r<jkr<j\leqslant k.

Since Φ(ur1)=0\Phi(u_{r-1})=0, we have just shown that G(ur1)G(u_{r-1}) consists on |Z(1)||Z(\vec{1})| disjoint trees. When passing from ur1u_{r-1} to uru_{r}, vertices z1z_{1} and z2z_{2} are identified and the 11-edge eire_{i_{r}} becomes a 0-loop (i.e., a length-one cycle) in G(ur)G(u_{r}). Figure 4(a)-(b) illustrates this process.

Refer to caption
Figure 4. Diagrams D(ur1)D(u_{r-1}), D(ur)D(u_{r}), D(u)D(u) and the associated graphs illustrating the proof of Lemma 4.3 are shown. Recall that dark (resp. light) chords corresponds to 0-chords (resp. 11-chords). In this case ε={eα,eγ}\varepsilon=\{e_{\alpha},e_{\gamma}\}.

The case (2a)(2a) corresponds to the case when z1z_{1} and z2z_{2} belong to different connected components of G(ur1)G(u_{r-1}). Since no more mergings are possible, no more cycles are created. Moreover, the length of the (unique) cycle 𝔠1\mathfrak{c}_{1} in G(u)G(u) is |ε|+1|\varepsilon|+1 (see Figure 4(c)).

The case (2b)(2b) corresponds to the case when z1z_{1} and z2z_{2} belong to the same connected component: the fact that there is a path connecting z1z_{1} and z2z_{2} in G(ur1)G(u_{r-1}) implies that an additional cycle 𝔠2\mathfrak{c}_{2} is created when identifying both vertices. As illustrated in Figure 4(b)-(c), the condition on the alternacy of the chords holds. The planarity of D(u)D(u) completes the proof. ∎

4.2. Enhanced states

An enhacement of a state u𝟐𝐧u\in\mathbf{2^{n}} is a map xx assigning a label +1+1 or 1-1 to each of the circles in Z(u)Z(u); we note by (u,x)(u,x) the associated enhanced state. Write Z+(u,x)Z_{+}(u,x) and Z(u,x)Z_{-}(u,x) for the subsets of elements of Z(u)Z(u) labeled by +1+1 and 1-1, respectively. Define, for the enhanced state (u,x)(u,x), the integers

h(u,x)=n+|u|,q(u,x)=n+2n+|u|+|Z+(u,x)||Z(u,x)|,h(u,x)=-n_{-}+|u|,\quad\quad q(u,x)=n_{+}-2n_{-}+|u|+|Z_{+}(u,x)|-|Z_{-}(u,x)|,

which are the homological and quantum gradings for Khovanov homology, respectively [LS14].

Let x+x_{+} be the constant enhacement with value +1+1 and, for a given circle zZ(u)z\in Z(u), let xz+x_{z}^{+} be the enhacement assigning a positive label to every circle but zz. Sometimes we will write u+=(u,x+)u_{+}=(u,x_{+}).

Define

jmax(D)=max{q(u,x)(u,x) is an enhanced state of D}j_{\max}(D)=\max\{{q(u,x)\mid(u,x)\mbox{ is an enhanced state of }D}\}

and jalmax(D)=jmax(D)2{j_{\mathrm{almax}}}(D)=j_{\max}(D)-2; we refer to these numbers as extreme and almost-extreme (quantum) gradings for Khovanov homology of the diagram DD. It turns out that

jmax(D)\displaystyle j_{\max}(D) =q(1,x+)\displaystyle=q(\vec{1},x_{+}) and jalmax(D)\displaystyle{j_{\mathrm{almax}}}(D) =q(1,xz+), for any zZ(1).\displaystyle=q(\vec{1},x_{z}^{+}),\,\text{ for any }z\in Z(\vec{1}).
Proposition 4.4.

Let (u,x)(u,x) be an enhanced state of a diagram DD with q(u,x)=jq(u,x)=j. Then,

|Z(u,x)|=jmax(D)j2Φ(u).|Z_{-}(u,x)|=\frac{j_{\max}(D)-j}{2}-\Phi(u).
Proof.

From the definition of quantum grading, it holds that

jmax(D)j\displaystyle j_{\max}(D)-j =n+|Z(1)||u||Z+(u,x)|+|Z(u,x)|\displaystyle=n+|Z(\vec{1})|-|u|-|Z_{+}(u,x)|+|Z_{-}(u,x)|
=n+|Z(1)||u||Z(u)|+2|Z(u,x)|.\displaystyle=n+|Z(\vec{1})|-|u|-|Z(u)|+2|Z_{-}(u,x)|.

Relation (4.2) completes the proof. ∎

Corollary 4.5.

Let (u,x)(u,x) be an enhanced state of DD satisfying q(u,x)=jalmax(D)q(u,x)={j_{\mathrm{almax}}}(D). Then Φ(u)1\Phi(u)\leqslant 1. Moreover,

  1. (1)

    if Φ(u)=0\Phi(u)=0, then x=xz+x=x_{z}^{+} for some zZ(u)z\in Z(u).

  2. (2)

    if Φ(u)=1\Phi(u)=1, then x=x+x=x_{+}.

Since we are interested in studying the almost-extreme Khovanov complex of a link diagram, in the next section we study the characterization of those states uu so that Φ(u)\Phi(u) equals 0 or 11.

4.3. States with Φ(u)=0,1\Phi(u)=0,1

We are interested now in characterizing those states taking part in the almost-extreme Khovanov complex, i.e., given a state u𝟐𝐧u\in\mathbf{2^{n}} of DD, we want to determine whether there exists an enhacement xx so that q(u,x)=jalmax(D)q(u,x)={j_{\mathrm{almax}}}(D), just by looking at D(1)D(\vec{1}).

Definition 4.6.

Given u𝟐𝐧u\in\mathbf{2^{n}} a state of DD and a,b,c,da,b,c,d chords in D(u)D(u), we say that:

  1. (1)

    aa and bb are parallel if they are bichords with their endpoints in the same circles, i.e., 𝒪a(u)=𝒪b(u)={z,z}\mathcal{O}_{a}(u)=\mathcal{O}_{b}(u)=\{z,z^{\prime}\}.

  2. (2)

    a,ba,b form an alternating pair if they are monochords and their endpoints alternate along the same circle, as illustrated in Figure 5 (a)(a) (compare to ladybug configuration from [LS14]).

  3. (3)

    a,b,ca,b,c form an alternating triple if 𝒪a(u)=𝒪b(u)={z,z}\mathcal{O}_{a}(u)=\mathcal{O}_{b}(u)=\{z,z^{\prime}\} and their endpoints alternate with the endpoints of the monochord cc along zz. See Figure 5 (b)(b).

  4. (4)

    a,b,c,da,b,c,d form a mixed alternating pair if the only alternating pairs among them are (a,b),(b,c)(a,b),(b,c) and (c,d)(c,d). See Figure 5 (c)(c).

Refer to caption
Figure 5. Chords forming an alternating pair (a)(a), an alternating triple (b)(b) and a mixed alternating pair (c)(c).

Given v>uv>u states of DD, we write D(v)uD(v)_{u} for the chord diagram having the same circles as D(v)D(v), but only the 11-chords ei(v)e_{i}(v) such that viuiv_{i}\neq u_{i}.

Remark 4.7.

If w>v>uw>v>u are states of DD, then ei(v)D(v)ue_{i}(v)\in D(v)_{u} implies that ei(w)D(w)ue_{i}(w)\in D(w)_{u}.

Proposition 4.8.

Let uu be a state of DD. Then, Φ(u)>0\Phi(u)>0 if and only if D(1)uD(\vec{1})_{u} contains a bichord or an alternating pair, and Φ(u)>1\Phi(u)>1 if and only if D(1)uD(\vec{1})_{u} contains at least one of the following configurations:

  1. (1)

    Two non-parallel bichords.

  2. (2)

    An alternating pair and a bichord.

  3. (3)

    An alternating triple.

  4. (4)

    Two disjoint alternating pairs.

  5. (5)

    A mixed alternating pair.

The proof of the necessary condition (i.e., the implication \Leftarrow) is a case-by-case straightforward checking. We use Lemmas 4.9 and 4.10 in the proof of the sufficient condition.

Lemma 4.9.

Let wv>uw\succ v>u be three states with Φ(w)=Φ(v)\Phi(w)=\Phi(v). Then:

  1. (1)

    If D(v)uD(v)_{u} contains a bichord, then D(w)uD(w)_{u} contains a bichord or an alternating pair.

  2. (2)

    If D(v)uD(v)_{u} contains an alternating pair, then D(w)uD(w)_{u} contains an alternating pair.

  3. (3)

    If D(v)uD(v)_{u} contains an alternating triple, then D(w)uD(w)_{u} contains an alternating triple or a mixed alternating pair.

  4. (4)

    If D(v)uD(v)_{u} contains a mixed alternating pair, then D(w)uD(w)_{u} contains a mixed alternating pair.

  5. (5)

    If D(v)uD(v)_{u} contains an alternating pair and a bichord, then D(w)uD(w)_{u} contains an alternating pair and a bichord, or a mixed alternating pair or two disjoint alternating pairs.

Proof.

Since Φ(w)=Φ(v)\Phi(w)=\Phi(v), wivw\succ_{i}v is a splitting for some monochord ei(w)e_{i}(w). The lemma follows from Remark 4.7 together with the following facts (ijkli\neq j\neq k\neq l):

  • -

    If ej(v)e_{j}(v) is a bichord, then either ej(w)e_{j}(w) is a bichord or (ej(w),ei(w))(e_{j}(w),e_{i}(w)) form an alternating pair.

  • -

    If (ej(v),ek(v))(e_{j}(v),e_{k}(v)) is an alternating pair, then (ej(w),ek(w))(e_{j}(w),e_{k}(w)) is an alternating pair.

  • -

    If (ej(v),ek(v),el(v))(e_{j}(v),e_{k}(v),e_{l}(v)) is an alternating triple, then either (ej(w),ek(w),el(w))(e_{j}(w),e_{k}(w),e_{l}(w)) is an alternating triple, or (ei(w),ej(w),ek(w),el(w))(e_{i}(w),e_{j}(w),e_{k}(w),e_{l}(w)) is a mixed alternating pair.

  • -

    If (ej(v),ek(v),el(v),em(v))(e_{j}(v),e_{k}(v),e_{l}(v),e_{m}(v)) is a mixed alternating pair, then (ej(w),ek(w),(e_{j}(w),e_{k}(w), el(w),em(w))e_{l}(w),e_{m}(w)) is a mixed alternating pair.

  • -

    If (ej(v),ek(v))(e_{j}(v),e_{k}(v)) is an alternating pair and el(v)e_{l}(v) is a bichord, then either (ej(w),ek(w))(e_{j}(w),e_{k}(w)) is an alternating pair and el(w)e_{l}(w) is a bichord or (ei(w),ej(w),(e_{i}(w),e_{j}(w), ek(w),el(w))e_{k}(w),e_{l}(w)) is a mixed alternating pair, or (ei(w),el(w))(e_{i}(w),e_{l}(w)) and (ej(w),ek(w))(e_{j}(w),e_{k}(w)) are disjoint alternating pairs. ∎

Lemma 4.10.

Let wv>uw\succ v>u be three states with Φ(w)=Φ(v)1\Phi(w)=\Phi(v)-1. Then:

  1. (1)

    If D(v)uD(v)_{u} contains a bichord, then D(w)uD(w)_{u} contains two non-parallel bichords.

  2. (2)

    If D(v)uD(v)_{u} contains an alternating pair, then D(w)uD(w)_{u} contains an alternating triple or an alternating pair and a bichord.

Proof.

Since Φ(w)=Φ(v)1\Phi(w)=\Phi(v)-1, wivw\succ_{i}v is a merging for some bichord ei(w)e_{i}(w). The lemma follows from Remark 4.7 together with the following facts (ijkli\neq j\neq k\neq l):

  • -

    If ej(v)e_{j}(v) is a bichord, then ej(w)e_{j}(w) and ei(w)e_{i}(w) are non-parallel bichords.

  • -

    If (ej(v),ek(v))(e_{j}(v),e_{k}(v)) is an alternating pair, then either (ej(w),ek(w))(e_{j}(w),e_{k}(w)) is an alternating pair and ei(w)e_{i}(w) is a bichord, or (ei(w),ej(w),ek(w))(e_{i}(w),e_{j}(w),e_{k}(w)) is an alternating triple. ∎

Proof of the sufficient condition of Proposition 4.8.

Fix a chain

1=u0i1u1i2ikuk=u.\vec{1}=u_{0}\succ_{i_{1}}u_{1}\succ_{i_{2}}\ldots\succ_{i_{k}}u_{k}=u.

Consider first the case Φ(u)>0\Phi(u)>0 and write mm for be the minimal number such that um1imumu_{m-1}\succ_{i_{m}}u_{m} is a merging. Therefore, eim(um1)D(um1)ue_{i_{m}}(u_{m-1})\in D(u_{m-1})_{u} is a bichord and applying Lemma 4.9 recursively, we deduce that D(1)uD(\vec{1})_{u} contains a bichord or an alternating pair.

Consider now the case Φ(u)>1\Phi(u)>1 and write m<pm<p for the first two indices such that um1imumu_{m-1}\succ_{i_{m}}u_{m} and up1ipupu_{p-1}\succ_{i_{p}}u_{p} are mergings. Therefore, Φ(um)=Φ(up1)\Phi(u_{m})=\Phi(u_{p-1}), and applying recursively Lemma 4.9 to the bichord eip(up1)D(up1)ue_{i_{p}}(u_{p-1})\in D(u_{p-1})_{u} we deduce that D(um)uD(u_{m})_{u} contains a bichord or an alternating pair. Apply once Lemma 4.10 and deduce that D(um1)uD(u_{m-1})_{u} contains at least one of the following: two non-parallel bichords or an alternating triple or an alternating pair and a bichord.

Finally, applying in each case recursively Lemma 4.9 we deduce that D(1)uD(\vec{1})_{u} contains at least one of the following configurations:

  1. (1)

    Two non-parallel bichords.

  2. (2)

    An alternating pair and a bichord.

  3. (3)

    An alternating triple.

  4. (4)

    Two disjoint alternating pairs.

  5. (5)

    A mixed alternating pair.∎

4.4. Ladybug set

Let u𝟐𝐧u\in\mathbf{2^{n}} so that Φ(u)=1\Phi(u)=1. If |π0G(u)|=|π0G(1)|1|\pi_{0}G(u)|=|\pi_{0}G(\vec{1})|-1, recall from Lemma 4.3(2a)(2a) that G(u)G(u) contains a unique cycle. We define the ladybug set of uu, (u)\mathcal{L}(u), as the singleton consisting of the connected component of G(u)G(u) containing that cycle.

Consider now the case when |π0G(u)|=|π0G(1)||\pi_{0}G(u)|=|\pi_{0}G(\vec{1})|. Then, by Lemma 4.3(2b)(2b), G(u)G(u) contains two cycles 𝔠1\mathfrak{c}_{1} and 𝔠2\mathfrak{c}_{2} sharing a common vertex zz. Write ai,bia_{i},b_{i} for the edges of 𝔠i\mathfrak{c}_{i} adjacent to zz, for i=1,2i=1,2 (if 𝔠i\mathfrak{c}_{i} is a loop, then ai=bia_{i}=b_{i}).

The chords a1,b1,a2,b2a_{1},b_{1},a_{2},b_{2} divide the circle zz into four arcs (see Figure 6). We define the ladybug set (u)\mathcal{L}(u) of D(u)D(u) as the set whose elements are the two arcs of zz that are reached by traveling along any of the chords a1,b1,a2,b2a_{1},b_{1},a_{2},b_{2} until zz, and then taking a right turn.

Refer to caption
Figure 6. An example of D(u)D(u) for the case when Φ(u)=1\Phi(u)=1 and |π0G(u)|=|π0G(1)||\pi_{0}G(u)|=|\pi_{0}G(\vec{1})|. The cycles 𝔠1\mathfrak{c}_{1} and 𝔠2\mathfrak{c}_{2} (of length 4 and 5, respectively) divide zz into four arcs. The ladybug set corresponds to the thickened arcs, i.e. (u)={P,Q}\mathcal{L}(u)=\{P,Q\}.
Remark 4.11.

In the case when |π0G(u)|=|π0G(1)||\pi_{0}G(u)|=|\pi_{0}G(\vec{1})| and ai=bia_{i}=b_{i} for i=1,2i=1,2, the set (u)\mathcal{L}(u) coincides with the right pair giving rise to the celebrated ladybug matching defined in [LS14].

Lemma 4.12.

If u>vu>v, |π0G(v)|=|π0G(1)||\pi_{0}G(v)|=|\pi_{0}G(\vec{1})| and Φ(u)=Φ(v)=1\Phi(u)=\Phi(v)=1, then each arc in (u)\mathcal{L}(u) intersects one of the two arcs in (v)\mathcal{L}(v) and viceversa.

Proof.

Let Λ={ei(u)|uivi}E¯(u)\Lambda=\{e_{i}(u)\,|\,u_{i}\neq v_{i}\}\subset\bar{E}(u). Since Φ(u)=Φ(v)\Phi(u)=\Phi(v), all elements in Λ\Lambda are loops and none pair among them constitute an alternating pair.

By Lemma 4.3(2b)(2b), G(u)G(u) contains two cycles 𝔠1\mathfrak{c}_{1} and 𝔠2\mathfrak{c}_{2} sharing a common vertex zuz_{u}. Write zvz_{v} for the circle in D(v)D(v) with the same property. Then zvz_{v} is obtained from zuz_{u} by performing surgery along the loops in Λ\Lambda with their endpoints in zuz_{u}.

Write (u)={P,Q}\mathcal{L}(u)=\{P,Q\} and let ΛP\Lambda_{P} (resp. ΛQ\Lambda_{Q}) be the subset of loops of Λ\Lambda so that at least one of their endpoints lies in PP (resp. QQ). The disposition of 𝔠1\mathfrak{c}_{1} and 𝔠2\mathfrak{c}_{2} (stated in Lemma 4.3(2b)(2b)) implies that ΛPΛQ=\Lambda_{P}\cap\Lambda_{Q}=\emptyset. Define the following subsets of 2\mathbb{R}^{2}:

p=ei(u)ΛPei(u) and q=ei(u)ΛQei(u).p=\bigcup_{e_{i}(u)\in\Lambda_{P}}e_{i}(u)\quad\mbox{ and }\quad q=\bigcup_{e_{i}(u)\in\Lambda_{Q}}e_{i}(u).

The endpoints of each loop ei(u)ΛPe_{i}(u)\in\Lambda_{P} (resp. ei(u)ΛQe_{i}(u)\in\Lambda_{Q}) separate zuz_{u} into two arcs: write did_{i} for the unique arc in zuz_{u} which is disjoint from QQ (resp. from PP). Define the following subsets of 2\mathbb{R}^{2}:

dP=ei(u)ΛPdi and dQ=ei(u)ΛQdi.d_{P}=\bigcup_{e_{i}(u)\in\Lambda_{P}}d_{i}\quad\mbox{ and }\quad d_{Q}=\bigcup_{e_{i}(u)\in\Lambda_{Q}}d_{i}.

As there are no alternating pairs in Λ\Lambda, we deduce that

dPdQ\displaystyle d_{P}\cap d_{Q} =\displaystyle=\emptyset and (pdP)(qdQ)\displaystyle(p\cup d_{P})\cap(q\cup d_{Q}) =\displaystyle=\emptyset

As a consequence, we can consider two disjoint open subsets A,BA,B of 2\mathbb{R}^{2} such that

PpdP\displaystyle P\cup p\cup d_{P} A\displaystyle\subset A QqdQB.\displaystyle Q\cup q\cup d_{Q}\subset B.

Each of the arcs of (v)\mathcal{L}(v) is obtained by doing surgery on each of the arcs of (u)\mathcal{L}(u). The surgery performed on PP is supported in AA, while the surgery perfomed on QQ is supported in BB. Since AB=A\cap B=\emptyset, it is possible to define PP^{\prime} (resp. QQ^{\prime}) as the arc of (v)\mathcal{L}(v) obtained from PP (resp. QQ). Thus, PP intersects PP^{\prime}, QQ intersects QQ^{\prime} and PQ=QP=P\cap Q^{\prime}=Q\cap P^{\prime}=\emptyset, as desired. ∎

4.5. The bijection ρu,v\rho_{u,v}

Given uivu\succ_{i}v so that Φ(u)=Φ(v)=1\Phi(u)=\Phi(v)=1, then |π0G(u)|=|π0G(v)||\pi_{0}G(u)|=|\pi_{0}G(v)|, and therefore it is possible to define a bijection ρu,v:(u)(v)\rho_{u,v}:\mathcal{L}(u)\to\mathcal{L}(v) in the following way:

  1. (1)

    If |π0G(v)|=|π0G(1)|1|\pi_{0}G(v)|=|\pi_{0}G(\vec{1})|-1, then ρu,v\rho_{u,v} maps the unique element in (u)\mathcal{L}(u) to the unique element in (v)\mathcal{L}(v).

  2. (2)

    If |π0G(v)|=|π0G(1)||\pi_{0}G(v)|=|\pi_{0}G(\vec{1})|, then by Lemma 4.12 each of the two arcs in (u)\mathcal{L}(u) intersects one of the two arcs in (v)\mathcal{L}(v). Write (u)={Pu,Qu}\mathcal{L}(u)=\{P_{u},Q_{u}\} and (v)={Pv,Qv}\mathcal{L}(v)=\{P_{v},Q_{v}\}, labelling the arcs so that PuPvP_{u}\cap P_{v}\neq\emptyset and QuQvQ_{u}\cap Q_{v}\neq\emptyset as subsets of 2\mathbb{R}^{2} (see Figure 7). The bijection ρu,v:(u)(v)\rho_{u,v}:\mathcal{L}(u)\to\mathcal{L}(v) is given by sending PuP_{u} to PvP_{v} and QuQ_{u} to QvQ_{v}. In particular, in those cases when none of the endpoints of the chord ei(u)e_{i}(u) lie in PuP_{u} nor QuQ_{u}, ρu,v\rho_{u,v} becomes the identity.

Refer to caption
Figure 7. Two examples illustrating the bijection ρu,v\rho_{u,v} for the case uivu\succ_{i}v, Φ(u)=Φ(v)=1\Phi(u)=\Phi(v)=1 and |π0G(v)|=|π0G(1)||\pi_{0}G(v)|=|\pi_{0}G(\vec{1})|, with the edge ei(u)e_{i}(u) depicted in light colour. In both cases ρu,v:(u)(v)\rho_{u,v}:\mathcal{L}(u)\to\mathcal{L}(v) maps PuP_{u} to PvP_{v} and QuQ_{u} to QvQ_{v}.

The following lemma follows immediately from Lemma 4.12:

Lemma 4.13.

Let uv,vwu\succ v,v^{\prime}\succ w so that Φ(u)=Φ(w)=1\Phi(u)=\Phi(w)=1. Then the following diagram commutes

(u)\textstyle{\mathcal{L}(u)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρu,v\scriptstyle{\rho_{u,v}}ρu,v\scriptstyle{\rho_{u,v^{\prime}}}(v)\textstyle{\mathcal{L}(v)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρv,w\scriptstyle{\rho_{v,w}}(v)\textstyle{\mathcal{L}(v^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρv,w\scriptstyle{\rho_{v^{\prime},w}}(w).\textstyle{\mathcal{L}(w).}

4.6. The bijection μi(u)\mu_{i}(u)

Given uivu\succ_{i}v with Φ(u)=0\Phi(u)=0, Φ(v)=1\Phi(v)=1 (i.e., a merging), we define a function

μi(u):𝒪i(u)(v)\mu_{i}(u)\colon\mathcal{O}_{i}(u)\to\mathcal{L}(v)

as follows:

  1. (1)

    If |π0G(v)|=|π0G(1)|1|\pi_{0}G(v)|=|\pi_{0}G(\vec{1})|-1, then |(v)|=1|\mathcal{L}(v)|=1 and we define μi(u)\mu_{i}(u) as the unique constant function.

  2. (2)

    If |π0G(v)|=|π0G(1)||\pi_{0}G(v)|=|\pi_{0}G(\vec{1})|, write 𝒪i(u)={z1,z2}\mathcal{O}_{i}(u)=\{z_{1},z_{2}\}, (v)={P,Q}\mathcal{L}(v)=\{P,Q\}, and assume without loss of generality that Pz1P\cap z_{1}\neq\emptyset and Qz2Q\cap z_{2}\neq\emptyset as subsets of 2\mathbb{R}^{2}. Then, we define the bijection μi(u):𝒪i(u)(v)\mu_{i}(u):\mathcal{O}_{i}(u)\to\mathcal{L}(v) by declaring μi(u)(z1)=P\mu_{i}(u)(z_{1})=P, μi(u)(z2)=Q\mu_{i}(u)(z_{2})=Q (see Figure 8).

Refer to caption
Figure 8. An example illustrating the bijection μi(u)\mu_{i}(u) for the case uivu\succ_{i}v, Φ(u)=0\Phi(u)=0, Φ(v)=1\Phi(v)=1 and |π0G(v)|=|π0G(1)||\pi_{0}G(v)|=|\pi_{0}G(\vec{1})|. The edge ei(u)e_{i}(u) is depicted in light colour. The bijection μi(u):𝒪i(u)(v)\mu_{i}(u)\colon\mathcal{O}_{i}(u)\to\mathcal{L}(v) sends z1z_{1} and z2z_{2} to PP and QQ, respectively.

The next result is a consequence of Lemma 4.12.

Lemma 4.14.

Let uivjwu\succ_{i}v\succ_{j}w, ujviwu\succ_{j}v^{\prime}\succ_{i}w. Then, the following squares commute:

  1. (1)

    If Φ(u)=Φ(v)=0\Phi(u)=\Phi(v^{\prime})=0 and Φ(v)=Φ(w)=1\Phi(v)=\Phi(w)=1, then111We defer the definition of hu,v:Z(u)Z(v)h_{u,v}\colon Z(u)\longrightarrow Z(v) some lines until the beginning of Section 5.

    𝒪i(u)\textstyle{\mathcal{O}_{i}(u)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μi(u)\scriptstyle{\mu_{i}(u)}hu,v\scriptstyle{h_{u,v^{\prime}}}(v)\textstyle{\mathcal{L}(v)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρv,w\scriptstyle{\rho_{v,w}}𝒪i(v)\textstyle{\mathcal{O}_{i}(v^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μi(v)\scriptstyle{\mu_{i}(v^{\prime})}(w).\textstyle{\mathcal{L}(w).}
  2. (2)

    If Φ(u)=0\Phi(u)=0 and Φ(v)=Φ(v)=Φ(w)=1\Phi(v)=\Phi(v^{\prime})=\Phi(w)=1, then

    𝒪j(u)=𝒪i(u)\textstyle{\mathcal{O}_{j}(u)=\mathcal{O}_{i}(u)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}μi(u)\scriptstyle{\mu_{i}(u)}μj(u)\scriptstyle{\mu_{j}(u)}(v)\textstyle{\mathcal{L}(v)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρv,w\scriptstyle{\rho_{v,w}}(v)\textstyle{\mathcal{L}(v^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρv,w\scriptstyle{\rho_{v^{\prime},w}}(w).\textstyle{\mathcal{L}(w).}

5. Khovanov functors

In this section we review the functor given by Lipshitz and Sarkar in [LLS17, LLS20] and introduce a new Khovanov functor, giving a natural transformation between them which allows us to prove that the geometric realizations of both functors are homotopy equivalent at the almost-extreme quantum grading (Corollary 5.7). First, we introduce some notation.

Let uu be a state of a diagram DD so that Φ(u)=0\Phi(u)=0. Orient G(u)G(u) by fixing, for each of its edges, one of its two possible orientations. We define, for each eE(u)e\in E(u), the subgraph e+G(u)e_{+}\subset G(u) as the connected component of G(u){e}G(u)\smallsetminus\{e\} towards which ee is pointing. Note that e+e_{+} is well defined, since Lemma 4.3 guarantees that G(u)G(u) has no cycles.

Given any two states u>vu>v of DD, there is an inclusion of the associated graphs G(u)G(v)G(u)\subset G(v), and we define the maps

(5.1) fu,v:π0G(u)π0G(v),gu,v:E(u)E(v);f_{u,v}\colon\pi_{0}G(u)\twoheadrightarrow\pi_{0}G(v),\quad g_{u,v}\colon E(u)\hookrightarrow E(v);

observe that the first one is an isomorphism if Φ(u)=Φ(v)\Phi(u)=\Phi(v), while the second one is always injective. Additionally, if uivu\succ_{i}v, there exists a Burnside morphism

hu,v:Z(u)Z(v)\displaystyle h_{u,v}\colon Z(u)\longrightarrow Z(v)

mapping a circle zZ(u)z\in Z(u) either to itself if z𝒪i(u)z\notin\mathcal{O}_{i}(u), or to 𝒪i(v)\mathcal{O}_{i}(v) if z𝒪i(u)z\in\mathcal{O}_{i}(u).

5.1. The Khovanov functor in almost-maximal grading

The key piece in the construction of the Khovanov spectra in [LLS20] was the Khovanov functor, whose associated stable homotopy type is a link invariant. Using Lemma 3.2, we restate now this functor, adapted to the particular case of the almost-maximal quantum grading.

Given a link diagram DD with nn ordered crossings, consider the functor222When working with different diagrams, we write FDF_{D} to denote the functor FF associated to the link diagram DD.

F:𝟐𝐧F\colon\mathbf{2^{n}}\longrightarrow\mathcal{B}

defined, in a vertex u𝟐𝐧u\in\mathbf{2^{n}}, as333The original functor FF from [LLS20] splits into functors FjF^{j} which associates to each state uu a set u,j\mathcal{E}_{u,j} of enhancements so that q(u,x)=jq(u,x)=j, for every xu,jx\in\mathcal{E}_{u,j}. When particularizing to the case when j=jalmaxj=j_{almax}, Corollary 4.5 implies that if Φ(u)=0\Phi(u)=0, then the set u,j\mathcal{E}_{u,j} can be identified with Z(u)Z(u), whereas if Φ(u)=1\Phi(u)=1 then u,j={u+}\mathcal{E}_{u,j}=\{u_{+}\} . :

F(u)\displaystyle F(u) ={Z(u)if Φ(u)=0,{u+}if Φ(u)=1,if Φ(u)>1.\displaystyle=\begin{cases}Z(u)&\text{if $\Phi(u)=0$},\\ \{u_{+}\}&\text{if $\Phi(u)=1$},\\ \emptyset&\text{if $\Phi(u)>1$}.\\ \end{cases}

On a morphism φu,v\varphi_{u,v}, with uivu\succ_{i}v, the span Fu>v:=F(φu,v):F(u)F(v)F_{u>v}:=F(\varphi_{u,v})\colon F(u)\to F(v) is given, for zZ(u)z\in Z(u) and u+=(u,x+)u_{+}=(u,x_{+}), by

Fu>v(z)\displaystyle F_{u>v}(z) ={hu,v(z)if Φ(v)=0,v+if Φ(v)=1 and z𝒪i(u),if Φ(v)=1 and z𝒪i(u);\displaystyle=\begin{cases}h_{u,v}(z)&\text{if $\Phi(v)=0$},\\ v_{+}&\text{if $\Phi(v)=1$ and $z\in\mathcal{O}_{i}(u)$},\\ \emptyset&\text{if $\Phi(v)=1$ and $z\notin\mathcal{O}_{i}(u)$};\end{cases}
Fu>v(u+)\displaystyle F_{u>v}(u_{+}) ={v+if Φ(v)=1,if Φ(v)=2.\displaystyle=\begin{cases}v_{+}&\text{if $\Phi(v)=1$},\\ \emptyset&\text{if $\Phi(v)=2$}.\end{cases}

Finally, we specify 22-morphisms: Let uivjwu\succ_{i}v\succ_{j}w and ujviwu\succ_{j}v^{\prime}\succ_{i}w. We need to produce a 2-morphism Fu,v,v,wF_{u,v,v^{\prime},w} between the 1-morphisms Fv>wFu>vF_{v>w}\circ F_{u>v} and Fv>wFu>vF_{v^{\prime}>w}\circ F_{u>v^{\prime}}.

Consider first the case when Φ(u)=0\Phi(u)=0. If Φ(u)=Φ(v)=Φ(v)=0\Phi(u)=\Phi(v)=\Phi(v^{\prime})=0, Φ(w)=1\Phi(w)=1, then the chords ei(u)e_{i}(u) and ej(u)e_{j}(u) form an alternating pair attached to some circle z1z_{1}, so 𝒪j(v)=𝒪i(v)\mathcal{O}_{j}(v)=\mathcal{O}_{i}(v) and 𝒪i(v)=𝒪j(v)\mathcal{O}_{i}(v^{\prime})=\mathcal{O}_{j}(v^{\prime}). If zZ(u)z\in Z(u) and zz1z\neq z_{1}, then Fv>wFu>v(z)=F_{v>w}\circ F_{u>v}(z)=\emptyset.

In the case z=z1z=z_{1}, label the circles involved in the mergings and splittings as

{z1}=𝒪i(u)\displaystyle\{z_{1}\}=\mathcal{O}_{i}(u) =𝒪j(u),\displaystyle=\mathcal{O}_{j}(u), {z2,z3}\displaystyle\{z_{2},z_{3}\} =𝒪i(v)=𝒪j(v),\displaystyle=\mathcal{O}_{i}(v)=\mathcal{O}_{j}(v),
{z6}=𝒪i(w)\displaystyle\{z_{6}\}=\mathcal{O}_{i}(w) =𝒪j(w),\displaystyle=\mathcal{O}_{j}(w), {z4,z5}\displaystyle\{z_{4},z_{5}\} =𝒪i(v)=𝒪j(v).\displaystyle=\mathcal{O}_{i}(v^{\prime})=\mathcal{O}_{j}(v^{\prime}).

Therefore,

Fv>wFu>v(z1)\displaystyle F_{v>w}\circ F_{u>v}(z_{1}) =Fv>w(z2+z3)={z2}w++{z3}w+=𝒪j(v)w+,\displaystyle=F_{v>w}(z_{2}+z_{3})=\{z_{2}\}\cdot w_{+}+\{z_{3}\}\cdot w_{+}=\mathcal{O}_{j}(v)\cdot w_{+},
Fv>wFu>v(z1)\displaystyle F_{v^{\prime}>w}\circ F_{u>v^{\prime}}(z_{1}) =Fv>w(z4+z5)={z4}w++{z5}w+=𝒪i(v)w+.\displaystyle=F_{v^{\prime}>w}(z_{4}+z_{5})=\{z_{4}\}\cdot w_{+}+\{z_{5}\}\cdot w_{+}=\mathcal{O}_{i}(v^{\prime})\cdot w_{+}.

Then, the 2-morphism Fu,v,v,wF_{u,v,v^{\prime},w} consists of a bijection between 𝒪j(v)\mathcal{O}_{j}(v) and 𝒪i(v)\mathcal{O}_{i}(v^{\prime}) given by the ladybug matching μi(v)1μj(v):𝒪j(v)(w)𝒪i(v)\mu_{i}(v^{\prime})^{-1}\circ\mu_{j}(v):\mathcal{O}_{j}(v)\leftrightarrow\mathcal{L}(w)\leftrightarrow\mathcal{O}_{i}(v^{\prime}) (see Section 4.6).

If Φ(u)=0\Phi(u)=0 and we are not in the previous situation, then every summand in the formal sums (Fv>wFu>v)(z)(F_{v>w}\circ F_{u>v})(z) and (Fv>wFu>v)(z)(F_{v^{\prime}>w}\circ F_{u>v^{\prime}})(z) has a singleton as coefficient, so there is a unique choice for the 2-morphism Fu,v,v,wF_{u,v,v^{\prime},w}.

In the case when Φ(u)=1\Phi(u)=1, then Fv,wFu,vF_{v,w}\circ F_{u,v} and Fv,wFu,vF_{v^{\prime},w}\circ F_{u,v^{\prime}} are empty if Φ(w)>1\Phi(w)>1; otherwise, both Fv,wFu,vF_{v,w}\circ F_{u,v} and Fv,wFu,vF_{v^{\prime},w}\circ F_{u,v^{\prime}} map u+u_{+} to the formal sum w+w_{+} which has a singleton as coefficient.

Lemma 5.1.

[LLS17] FF satisfies conditions (C1) and (C2) in Lemma 3.2.

The (strictly unitary lax) 2-functor FF defined above is the Khovanov functor given in [LLS20] at the almost-maximal quantum grading and, if DD has nn_{-} negative crossings, the almost-extreme (maximal) Khovanov spectrum is defined as

(5.2) 𝒳Djalmax=ΣnΣTotFD.\mathcal{X}^{{j_{\mathrm{almax}}}}_{D}=\Sigma^{-n_{-}}\Sigma^{\infty}\operatorname{Tot}F_{D}.

5.2. A new equivalent functor

We introduce now a (strictly unitary lax) 2-functor M:𝟐𝐧M\colon\mathbf{2^{n}}\to\mathcal{B}, with the property that its realization coincides with the realization of FF, as will be show in Corollary 5.7. As before, we start by defining444When working with different diagrams, we write MDM_{D} to denote the functor MM associated to the link diagram DD. MM for the vertices of the cube, corresponding to the states of DD:

M(u)\displaystyle M(u) ={π0G(u)E(u)if Φ(u)=0,u+if Φ(u)=1,if Φ(u)>1.\displaystyle=\begin{cases}\pi_{0}G(u)\cup E(u)&\text{if $\Phi(u)=0$,}\\ u_{+}&\text{if $\Phi(u)=1$,}\\ \emptyset&\text{if $\Phi(u)>1$.}\end{cases}

On a morphism φu,v\varphi_{u,v}, with uivu\succ_{i}v, the span Mu>v:=M(φu,v):M(u)M(v)M_{u>v}:=M(\varphi_{u,v})\colon M(u)\to M(v) is defined either on a connected component CC or an edge ee of G(u)G(u) (if Φ(u)=0\Phi(u)=0), or on the enhanced state u+u_{+} (if Φ(u)=1\Phi(u)=1), as follows:

Mu>v(C)\displaystyle M_{u>v}(C) ={fu,v(C)if Φ(v)=0,((v)fu,v(C))v+if Φ(v)=1;\displaystyle=\begin{cases}f_{u,v}(C)&\text{if $\Phi(v)=0$},\\ (\mathcal{L}(v)\cap f_{u,v}(C))\cdot v_{+}&\text{if $\Phi(v)=1$};\\ \end{cases}
Mu>v(e)\displaystyle M_{u>v}(e) ={gu,v(e)if Φ(v)=0,(μi(u)(e+𝒪i(u))v+if Φ(v)=1;\displaystyle=\begin{cases}g_{u,v}(e)&\text{if $\Phi(v)=0$},\\ (\mu_{i}(u)(e_{+}\cap\mathcal{O}_{i}(u))\cdot v_{+}&\text{if $\Phi(v)=1$};\\ \end{cases}
Mu>v(u+)\displaystyle M_{u>v}(u_{+}) ={v+if Φ(v)=1,if Φ(v)=2.\displaystyle=\begin{cases}v_{+}&\text{if $\Phi(v)=1$},\\ \emptyset&\text{if $\Phi(v)=2$}.\end{cases}
Remark 5.2.

Observe that the expression (v)fu,v(C)\mathcal{L}(v)\cap f_{u,v}(C) equals either (v)\mathcal{L}(v), in the case when 𝒪i(u)C\mathcal{O}_{i}(u)\cap C\neq\emptyset, or the empty set otherwise.

Remark 5.3.

Observe that the evaluation of the expression e+𝒪i(u)e_{+}\cap\mathcal{O}_{i}(u) leads to either the empty set or one or two circles, as illustrated in Figure 9. When e+𝒪i(u)=zZ(u)e_{+}\cap\mathcal{O}_{i}(u)=z\in Z(u), there are two possibilities: if |π0G(v)|=|π0G(1)||\pi_{0}G(v)|=|\pi_{0}G(\vec{1})|, then μi(u)(e+𝒪i(u))\mu_{i}(u)(e_{+}\cap\mathcal{O}_{i}(u)) maps zz to one of the two arcs in (v)\mathcal{L}(v); however, if |π0G(v)|=|π0G(1)|1|\pi_{0}G(v)|=|\pi_{0}G(\vec{1})|-1, then μi(u)(e+𝒪i(u))\mu_{i}(u)(e_{+}\cap\mathcal{O}_{i}(u)) maps zz to the singleton (v)\mathcal{L}(v).

Refer to caption
Figure 9. We illustrate the three cases considered in Remark 5.3. Given D(u)D(u) and the orientation shown in G(u)G(u), e+𝒪1(u)=e_{+}\cap\mathcal{O}_{1}(u)=\emptyset, e+𝒪2(u)={z2}e_{+}\cap\mathcal{O}_{2}(u)=\{z_{2}\} and e+𝒪3(u)={z1,z2}e_{+}\cap\mathcal{O}_{3}(u)=\{z_{1},z_{2}\}.

Finally, given uivjwu\succ_{i}v\succ_{j}w and ujviwu\succ_{j}v^{\prime}\succ_{i}w, we define the 2-morphism Mu,v,v,wM_{u,v,v^{\prime},w} between the 1-morphisms Mv>wMu>vM_{v>w}\circ M_{u>v} and Mv>wMu>vM_{v^{\prime}>w}\circ M_{u>v^{\prime}} as follows.

First, note that Mu,v,v,wM_{u,v,v^{\prime},w} is trivially defined when Φ(u)=Φ(v)\Phi(u)=\Phi(v) or |(w)|=1|\mathcal{L}(w)|=1, since it is a bijection between singletons. Then, we just need to specify the cases when Φ(u)=0\Phi(u)=0, Φ(w)=1\Phi(w)=1 and |(w)|=2|\mathcal{L}(w)|=2, depicted in Figure 10. Moreover, the value of Mv>wMu>vM_{v>w}\circ M_{u>v} and Mv>wMu>vM_{v^{\prime}>w}\circ M_{u>v^{\prime}} on a component Cπ0G(u)C\in\pi_{0}G(u) equals the empty set or a singleton, unless fu,w(C)f_{u,w}(C) is the component of G(w)G(w) containing two cycles; a similar reasoning applies for edges eG(u)e\in G(u), unless ee (or possibly gu,v(e)g_{u,v}(e) or gu,v(e)g_{u,v^{\prime}}(e)) points to the two circles involved in the unique merging. We define Mu,v,v,wM_{u,v,v^{\prime},w} in such components and edges:

Refer to caption
Figure 10. Cases (aa), (bb) and (cc) in the definition of Mu,v,v,wM_{u,v,v^{\prime},w}. In order |(w)|=2|\mathcal{L}(w)|=2, |π0G(w)||\pi_{0}G(w)| must be equal to |π0G(1)||\pi_{0}G(\vec{1})| and therefore the vertices corresponding to the two circles depicted in D(u)D(u) in cases (b)(b) and (c)(c) must belong to the same connected component in G(u)G(u).
  1. (aa)

    If Φ(u)=Φ(v)=Φ(v)=0\Phi(u)=\Phi(v)=\Phi(v^{\prime})=0, Φ(w)=1\Phi(w)=1 and |(w)|=2|\mathcal{L}(w)|=2, then ei(u)e_{i}(u) and ej(u)e_{j}(u) form an alternating pair (see Figure 10(a)) and we have

    Mv>wMu>v(C)=(w)w+\displaystyle M_{v>w}\circ M_{u>v}(C)=\mathcal{L}(w)\cdot w_{+} IdMv>wMu>v(C)=(w)w+,\displaystyle\overset{\mathrm{Id}}{\longleftrightarrow}M_{v^{\prime}>w}\circ M_{u>v^{\prime}}(C)=\mathcal{L}(w)\cdot w_{+},
    Mv>wMu>v(e)=(w)w+\displaystyle M_{v>w}\circ M_{u>v}(e)=\mathcal{L}(w)\cdot w_{+} IdMv>wMu>v(e)=(w)w+.\displaystyle\overset{\mathrm{Id}}{\longleftrightarrow}M_{v^{\prime}>w}\circ M_{u>v^{\prime}}(e)=\mathcal{L}(w)\cdot w_{+}.
  2. (bb)

    If Φ(u)=Φ(v)=0\Phi(u)=\Phi(v^{\prime})=0 and Φ(v)=Φ(w)=1\Phi(v)=\Phi(w)=1 and |(w)|=2|\mathcal{L}(w)|=2, then ei(u)e_{i}(u) is a bichord and ej(u)e_{j}(u) a monochord (see Figure 10(b)) and we have

    Mv>wMu>v(C)=(v)w+\displaystyle M_{v>w}\circ M_{u>v}(C)=\mathcal{L}(v)\cdot w_{+} ρv,wMv>wMu>v(C)=(w)w+,\displaystyle\overset{\rho_{v,w}}{\longleftrightarrow}M_{v^{\prime}>w}\circ M_{u>v^{\prime}}(C)=\mathcal{L}(w)\cdot w_{+},
    Mv>wMu>v(e)=(v)w+\displaystyle M_{v>w}\circ M_{u>v}(e)=\mathcal{L}(v)\cdot w_{+} ρv,wMv>wMu>v(e)=(w)w+.\displaystyle\overset{\rho_{v,w}}{\longleftrightarrow}M_{v^{\prime}>w}\circ M_{u>v^{\prime}}(e)=\mathcal{L}(w)\cdot w_{+}.
  3. (cc)

    If Φ(u)=0\Phi(u)=0 and Φ(v)=Φ(v)=Φ(w)=1\Phi(v)=\Phi(v^{\prime})=\Phi(w)=1 and |(w)|=2|\mathcal{L}(w)|=2, then ei(w)e_{i}(w) and ej(w)e_{j}(w) are parallel bichords (see Figure 10(c)) and we have:

    Mv>wMu>v(C)=(v)w+\displaystyle M_{v>w}\circ M_{u>v}(C)=\mathcal{L}(v)\cdot w_{+} ρv,w1ρv,wMv>wMu>v(C)=(v)w+,\displaystyle\overset{\rho_{v^{\prime},w}^{-1}\rho_{v,w}}{\longleftrightarrow}M_{v^{\prime}>w}\circ M_{u>v^{\prime}}(C)=\mathcal{L}(v^{\prime})\cdot w_{+},
    Mv>wMu>v(e)=(v)w+\displaystyle M_{v>w}\circ M_{u>v}(e)=\mathcal{L}(v)\cdot w_{+} ρv,w1ρv,wMv>wMu>v(e)=(v)w+.\displaystyle\overset{\rho_{v^{\prime},w}^{-1}\rho_{v,w}}{\longleftrightarrow}M_{v^{\prime}>w}\circ M_{u>v^{\prime}}(e)=\mathcal{L}(v^{\prime})\cdot w_{+}.
Lemma 5.4.

The 2-functor MM satisfies conditions (C1) and (C2) in Lemma 3.2 and therefore is well defined.

Proof.

Condition (C1) follows from the definition of MM. Consider an arbitrary three-dimensional face and denote its vertices as follows:

(5.3) 111\textstyle{111\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M111>011\scriptstyle{M_{111>011}}011\textstyle{011\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M011>001\scriptstyle{M_{011>001}}101\textstyle{101\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}001\textstyle{001\ignorespaces\ignorespaces\ignorespaces\ignorespaces}M001>000\scriptstyle{M_{001>000}}110\textstyle{110\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}010\textstyle{010\ignorespaces\ignorespaces\ignorespaces\ignorespaces}100\textstyle{100\ignorespaces\ignorespaces\ignorespaces\ignorespaces}000.\textstyle{000.}

In order to prove condition (C2), we need to show that 22-morphisms in this cube commute. To do so, we study the bijections obtained when we move along the cube. More precisely, starting from M001>000M011>001M111>011M_{001>000}\circ M_{011>001}\circ M_{111>011} we move to M010>000M011>010M111>011M_{010>000}\circ M_{011>010}\circ M_{111>011} and continue moving along the cube following Figure (11) until we reach again M001>000M011>001M111>011M_{001>000}\circ M_{011>001}\circ M_{111>011}.

Refer to caption
Figure 11. Moving along the cube. Paths (a)(a), (b)(b) and (c)(c) represent M001>000M011>001M111>011M_{001>000}\circ M_{011>001}\circ M_{111>011},   M010>000M011>010M111>011M_{010>000}\circ M_{011>010}\circ M_{111>011} and M010>000M110>010M111>110M_{010>000}\circ M_{110>010}\circ M_{111>110}, respectively.

To simplify notation, write ρi\rho_{i} for the morphism ρu,v\rho_{u,v} if uivu\succ_{i}v; then, the commutative diagram of Lemma 4.13 associated to uivjwu\succ_{i}v\succ_{j}w and ujviwu\succ_{j}v^{\prime}\succ_{i}w becomes:

(5.4) (u)\textstyle{\mathcal{L}(u)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρi\scriptstyle{\rho_{i}}ρj\scriptstyle{\rho_{j}}(v)\textstyle{\mathcal{L}(v)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρj\scriptstyle{\rho_{j}}(v)\textstyle{\mathcal{L}(v^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρi\scriptstyle{\rho_{i}}(w).\textstyle{\mathcal{L}(w).}

We just need to study the cases when Φ(111)=0\Phi(111)=0, Φ(000)=1\Phi(000)=1 and |(000)|=2|\mathcal{L}(000)|=2 (otherwise, the coefficients involved in all compositions of 11-morphisms of the edges of the cube are singletons, and therefore the 22-morphisms commute trivially). In principle, up to permutation, there are 8 different situations attending to the value of Φ\Phi in each vertex, depicted in Figure 12. However, it is not hard to check that situations (a)(a), (e)(e) and (f)(f) leads to inconsistency in the chord diagrams, and therefore they are not possible. We study the remaining 5 possible situations:

Refer to caption
Figure 12. Situations (a)(a), (e)(e) and (f)(f) are not possible when considering Φ(111)=0\Phi(111)=0, Φ(000)=1\Phi(000)=1. Each 0 or 11 represent the value of Φ\Phi in each of the vertices of the cube (5.3).
  1. (1)

    Figure 12(b)(b) (i.e., Φ(110)=Φ(101)=Φ(011)=Φ(100)=Φ(010)=0,Φ(001)=1\Phi(110)=\Phi(101)=\Phi(011)=\Phi(100)=\Phi(010)=0,\Phi(001)=1): D(111)000D(111)_{000} contains three monochords, two of them constituting an alternating pair. Coefficients involved in the three-fold compositions when moving around the cube are either singletons or the following:

    (001)ρ3(000)=(000)=(000)=(000)ρ31(001)=(001).\displaystyle\mathcal{L}(001)\overset{\rho_{3}}{\longleftrightarrow}\mathcal{L}(000)=\mathcal{L}(000)=\mathcal{L}(000)=\mathcal{L}(000)\overset{\rho_{3}^{-1}}{\longleftrightarrow}\mathcal{L}(001)=\mathcal{L}(001).

    The composition of the 22-morphisms connecting coefficients is the identity morphism.

  2. (2)

    Figure 12(c)(c) (i.e., Φ(110)=Φ(101)=Φ(011)=Φ(100)=0,Φ(001)=Φ(010)=1\Phi(110)=\Phi(101)=\Phi(011)=\Phi(100)=0,\Phi(001)=\Phi(010)=1): D(111)000D(111)_{000} contains three monochords constituting two alternating pairs. Coefficients involved in the three-fold compositions when moving around the cube are either singletons or the following:

    (001)ρ21ρ3(010)=(010)ρ2(000)=(000)ρ31(001)=(001).\displaystyle\mathcal{L}(001)\overset{\rho_{2}^{-1}\rho_{3}}{\longleftrightarrow}\mathcal{L}(010)=\mathcal{L}(010)\overset{\rho_{2}}{\longleftrightarrow}\mathcal{L}(000)=\mathcal{L}(000)\overset{\rho_{3}^{-1}}{\longleftrightarrow}\mathcal{L}(001)=\mathcal{L}(001).

    The composition of the 22-morphisms connecting coefficients is the identity morphism.

  3. (3)

    Figure 12(d)(d): D(111)000D(111)_{000} contains two monochords not forming an alternating pair and a bichord. Coefficients involved in the three-fold compositions when moving around the cube are either singletons or the following:

    (011)=(011)ρ3(010)ρ2(000)=(000)ρ31(001)ρ21(011).\displaystyle\mathcal{L}(011)=\mathcal{L}(011)\overset{\rho_{3}}{\longleftrightarrow}\mathcal{L}(010)\overset{\rho_{2}}{\longleftrightarrow}\mathcal{L}(000)=\mathcal{L}(000)\overset{\rho_{3}^{-1}}{\longleftrightarrow}\mathcal{L}(001)\overset{\rho_{2}^{-1}}{\longleftrightarrow}\mathcal{L}(011).

    The composition of the 22-morphisms is the following square, which is a particular case of (5.4):

    (011)\textstyle{\mathcal{L}(011)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ3\scriptstyle{\rho_{3}}(001)\textstyle{\mathcal{L}(001)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ21\scriptstyle{\rho_{2}^{-1}}(010)\textstyle{\mathcal{L}(010)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ2\scriptstyle{\rho_{2}}(000).\textstyle{\mathcal{L}(000)\ignorespaces\ignorespaces\ignorespaces\ignorespaces.}ρ31\scriptstyle{\rho_{3}^{-1}}
  4. (4)

    Figure 12(g)(g): D(111)000D(111)_{000} contains one monochord and two parallel bichords not forming an alternating triple. Coefficients involved in the three-fold compositions when moving around the cube are either singletons or the following:

    (011)=(011)ρ3(010)ρ11ρ2(100)ρ31(101)=(101)ρ21ρ1(011).\displaystyle\mathcal{L}(011)=\mathcal{L}(011)\overset{\rho_{3}}{\longleftrightarrow}\mathcal{L}(010)\overset{\rho^{-1}_{1}\rho_{2}}{\longleftrightarrow}\mathcal{L}(100)\overset{\rho_{3}^{-1}}{\longleftrightarrow}\mathcal{L}(101)=\mathcal{L}(101)\overset{\rho^{-1}_{2}\rho_{1}}{\longleftrightarrow}\mathcal{L}(011).

    The composition of 22-morphisms is the boundary of the union of two squares of the form (5.4) along an edge, which commutes:

    (011)\textstyle{\mathcal{L}(011)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ3\scriptstyle{\rho_{3}}(101)\textstyle{\mathcal{L}(101)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ1\scriptstyle{\rho_{1}}(001)\textstyle{\mathcal{L}(001)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ21\scriptstyle{\rho_{2}^{-1}}(010)\textstyle{\mathcal{L}(010)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ2\scriptstyle{\rho_{2}}(100)\textstyle{\mathcal{L}(100)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ31\scriptstyle{\rho_{3}^{-1}}(000).\textstyle{\mathcal{L}(000)\ignorespaces\ignorespaces\ignorespaces\ignorespaces.}ρ11\scriptstyle{\rho_{1}^{-1}}
  5. (5)

    Figure 12(h)(h): D(111)000D(111)_{000} contains three parallel bichords. Coefficients involved in the three-fold compositions when moving around the cube are either singletons or the following:

    (011)=(011)ρ11ρ3(110)=(110)ρ31ρ2(101)=(101)ρ21ρ1(011).\displaystyle\mathcal{L}(011)=\mathcal{L}(011)\overset{\rho_{1}^{-1}\rho_{3}}{\longleftrightarrow}\mathcal{L}(110)=\mathcal{L}(110)\overset{\rho^{-1}_{3}\rho_{2}}{\longleftrightarrow}\mathcal{L}(101)=\mathcal{L}(101)\overset{\rho_{2}^{-1}\rho_{1}}{\longleftrightarrow}\mathcal{L}(011).

    The composition of 22-morphisms is the boundary of the union of three squares of the form (5.4) along their common edges, which commutes:

    (011)\textstyle{\mathcal{L}(011)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ3\scriptstyle{\rho_{3}}(101)\textstyle{\mathcal{L}(101)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ1\scriptstyle{\rho_{1}}(001)\textstyle{\mathcal{L}(001)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ21\scriptstyle{\rho_{2}^{-1}}(110)\textstyle{\mathcal{L}(110)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ2\scriptstyle{\rho_{2}}(010)\textstyle{\mathcal{L}(010)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ11\scriptstyle{\rho_{1}^{-1}}(100)\textstyle{\mathcal{L}(100)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}ρ31\scriptstyle{\rho_{3}^{-1}}

Once we have defined MM, we will prove that its geometric realization is equivalent to that of the Khovanov functor FF. To do so, we introduce a natural transformation γ:MF\gamma\colon M\to F and show that the induced homomorphism γ:C(M;)C(F;)\gamma_{*}\colon C_{*}(M;\mathbb{Z})\to C_{*}(F;\mathbb{Z}) is in fact an isomorphism (see Section 3.2).

Given a state u𝟐𝐧u\in\mathbf{2^{n}}, we set the natural natural transformation γ\gamma as follows:

γu(C)\displaystyle\gamma_{u}(C) =zCz,\displaystyle=\sum_{z\in C}z,
γu(e)\displaystyle\gamma_{u}(e) =ze+z,\displaystyle=\sum_{z\in e_{+}}z,
γu(u+)\displaystyle\gamma_{u}(u_{+}) =u+.\displaystyle=u_{+}.

Next, for each uivu\succ_{i}v, we define a 22-morphism γu,v:Fu>vγuMu>vγv\gamma_{u,v}:F_{u>v}\circ\gamma_{u}\to M_{u>v}\circ\gamma_{v}, i.e., γu,v\gamma_{u,v} makes the following diagram commute:

M(u)\textstyle{M(u)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γu\scriptstyle{\gamma_{u}}Mu>v\scriptstyle{M_{u>v}}F(u)\textstyle{F(u)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}Fu>v\scriptstyle{F_{u>v}}M(v)\textstyle{M(v)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γv\scriptstyle{\gamma_{v}}F(v).\textstyle{F(v).}

If Φ(u)=0\Phi(u)=0 and Φ(v)=0\Phi(v)=0, γu,v\gamma_{u,v} is defined as the identity:

Fu>vγu(C)\displaystyle F_{u>v}\circ\gamma_{u}(C) =zChu,v(z)Idzfu,v(C)z=γvMu>v(C),\displaystyle=\sum_{z\in C}h_{u,v}(z)\overset{\mathrm{Id}}{\longleftrightarrow}\sum_{z\in f_{u,v}(C)}z\,=\,\gamma_{v}\circ M_{u>v}(C),
Fu>vγu(e)\displaystyle F_{u>v}\circ\gamma_{u}(e) =ze+hu,v(z)Idzgu,v(e)+z=γvMu>v(e).\displaystyle=\sum_{z\in e_{+}}h_{u,v}(z)\overset{\mathrm{Id}}{\longleftrightarrow}\sum_{z\in g_{u,v}(e)_{+}}z\,=\,\gamma_{v}\circ M_{u>v}(e).

In the case when Φ(u)=0\Phi(u)=0 and Φ(v)=1\Phi(v)=1 we define γu,v\gamma_{u,v} as μi(u)\mu_{i}(u):

Fu>vγu(C)\displaystyle F_{u>v}\circ\gamma_{u}(C) =(𝒪i(u)C)v+μi(u)((v)fu,v(C))v+=γvMu>v(C),\displaystyle=(\mathcal{O}_{i}(u)\cap C)\cdot v_{+}\overset{\mu_{i}(u)}{\longleftrightarrow}(\mathcal{L}(v)\cap f_{u,v}(C))\cdot v_{+}=\gamma_{v}\circ M_{u>v}(C),
Fu>vγu(e)\displaystyle F_{u>v}\circ\gamma_{u}(e) =(𝒪i(u)e+)v+μi(u)(μi(u)(e+𝒪i(u))v+=γvMu>v(e).\displaystyle=(\mathcal{O}_{i}(u)\cap e_{+})\cdot v_{+}\overset{\mu_{i}(u)}{\longleftrightarrow}(\mu_{i}(u)(e_{+}\cap\mathcal{O}_{i}(u))\cdot v_{+}=\gamma_{v}\circ M_{u>v}(e).

And when Φ(u)=1=Φ(v)\Phi(u)=1=\Phi(v), γu,v\gamma_{u,v} is defined as the identity:

Fu>vγu(u+)\displaystyle F_{u>v}\circ\gamma_{u}(u_{+}) =v+Idv+=γvMu>v(u+).\displaystyle=v_{+}\overset{\mathrm{Id}}{\longleftrightarrow}v_{+}\,=\,\gamma_{v}\circ M_{u>v}(u_{+}).
Lemma 5.5.

The natural transformation γu\gamma_{u} is well defined.

Proof.

We need to prove that γu\gamma_{u} satisfies conditions (C1) and (C2) in Lemma 3.2. Condition (C1) follows from definition of γu\gamma_{u}. Next, we prove (C2) by showing that 22-morphisms in the following cube commutes, similar as we did in proof of Lemma 5.4:

M(u)\textstyle{M(u)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γu\scriptstyle{\gamma_{u}}M(v)\textstyle{M(v)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γv\scriptstyle{\gamma_{v}}M(v)\textstyle{M(v^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γv\scriptstyle{\gamma_{v}^{\prime}}M(w)\textstyle{M(w)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}γw\scriptstyle{\gamma_{w}}F(u)\textstyle{F(u)\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F(v)\textstyle{F(v)\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F(v)\textstyle{F(v^{\prime})\ignorespaces\ignorespaces\ignorespaces\ignorespaces}F(w).\textstyle{F(w).}

The commutativity is clear when all coefficients are singletons, i.e., it holds unless Φ(u)=0\Phi(u)=0, Φ(w)=1\Phi(w)=1 and |(w)|=2|\mathcal{L}(w)|=2; thus, we have to study the three following situations555The case Φ(u)=0,Φ(v)=0,Φ(v)=1,Φ(w)=1\Phi(u)=0,\Phi(v)=0,\Phi(v^{\prime})=1,\Phi(w)=1 is symmetric to (5.6)., corresponding again to the cases illustrated in Figure 10:

(5.5) Φ(u)\displaystyle\Phi(u) =0,\displaystyle=0, Φ(v)\displaystyle\Phi(v) =0,\displaystyle=0, Φ(v)\displaystyle\Phi(v^{\prime}) =0,\displaystyle=0, Φ(w)\displaystyle\Phi(w) =1;\displaystyle=1;
(5.6) Φ(u)\displaystyle\Phi(u) =0,\displaystyle=0, Φ(v)\displaystyle\Phi(v) =1,\displaystyle=1, Φ(v)\displaystyle\Phi(v^{\prime}) =0,\displaystyle=0, Φ(w)\displaystyle\Phi(w) =1;\displaystyle=1;
(5.7) Φ(u)\displaystyle\Phi(u) =0,\displaystyle=0, Φ(v)\displaystyle\Phi(v) =1,\displaystyle=1, Φ(v)\displaystyle\Phi(v^{\prime}) =1,\displaystyle=1, Φ(w)\displaystyle\Phi(w) =1.\displaystyle=1.

Now, for each of the three cases above, we have to study the bijections obtained when we move along the cube: starting from γwMv>wMu>v\gamma_{w}\circ M_{v>w}\circ M_{u>v}, we move to Fv>wγvMu>vF_{v>w}\circ\gamma_{v}\circ M_{u>v}, and continue moving along the cube following Figure 11, until we reach again γwMv>wMu>v\gamma_{w}\circ M_{v>w}\circ M_{u>v}.

We just need to consider the action of those compositions on the component Cπ0G(u)C\in\pi_{0}G(u) such that fu,w(C)π0G(w)f_{u,w}(C)\in\pi_{0}G(w) contains two cycles (otherwise, the value of Fv>wγvMu>vF_{v>w}\circ\gamma_{v}\circ M_{u>v} is trivial, as explained in the proof of Lemma 5.4).

Consider the case (5.5), where edges ei(u)e_{i}(u) and ej(u)e_{j}(u) form an alternating pair (see Figure 10(a)(a)). Then Fv>wγvMu>v(C)F_{v>w}\circ\gamma_{v}\circ M_{u>v}(C) equals (w)w+\mathcal{L}(w)\cdot w_{+}, and we obtain the following values when moving along the cube as in Figure 11:

(w)w+μj(v)1𝒪j(v)w+=𝒪j(v)w+μi(v)1μj(v)𝒪i(v)w+=𝒪i(v)w+\mathcal{L}(w)\cdot w_{+}\overset{\mu_{j}(v)^{-1}}{\longleftrightarrow}\mathcal{O}_{j}(v)\cdot w_{+}=\mathcal{O}_{j}(v)\cdot w_{+}\overset{\mu_{i}(v^{\prime})^{-1}\mu_{j}(v)}{\longleftrightarrow}\mathcal{O}_{i}(v^{\prime})\cdot w_{+}=\mathcal{O}_{i}(v^{\prime})\cdot w_{+}
μi(v)(w)w+=(w)w+.\overset{\mu_{i}(v^{\prime})}{\longleftrightarrow}\mathcal{L}(w)\cdot w_{+}=\mathcal{L}(w)\cdot w_{+}.

We focus now in case (5.6), where ej(u)e_{j}(u) is a loop and ei(u)e_{i}(u) a bichord (Figure 10(b)(b) shows an example of this situation). In this case Fv>wγvMu>v(C)F_{v>w}\circ\gamma_{v}\circ M_{u>v}(C) equals (v)w+\mathcal{L}(v)\cdot w_{+}, and we get:

(v)w+=(v)w+μi(u)1𝒪i(u)w+hu,v𝒪i(v)w+=𝒪i(v)w+\mathcal{L}(v)\cdot w_{+}=\mathcal{L}(v)\cdot w_{+}\overset{\mu_{i}(u)^{-1}}{\longleftrightarrow}\mathcal{O}_{i}(u)\cdot w_{+}\overset{h_{u,v^{\prime}}}{\longleftrightarrow}\mathcal{O}_{i}(v^{\prime})\cdot w_{+}=\mathcal{O}_{i}(v^{\prime})\cdot w_{+}
μi(v)(w)w+ρv,w1(v)w+.\overset{\mu_{i}(v^{\prime})}{\longleftrightarrow}\mathcal{L}(w)\cdot w_{+}\overset{\rho_{v,w}^{-1}}{\longleftrightarrow}\mathcal{L}(v)\cdot w_{+}.

Consider the third case (5.7), where ei(u)e_{i}(u) and ej(u)e_{j}(u) are parallel bichords, so 𝒪i(u)=𝒪j(u)\mathcal{O}_{i}(u)=\mathcal{O}_{j}(u) (see Figure 10(c)(c)). Then Fv>wγvMu>v(C)=(v)w+F_{v>w}\circ\gamma_{v}\circ M_{u>v}(C)=\mathcal{L}(v)\cdot w_{+}, and we obtain:

(v)w+=(v)w+μi(u)1𝒪i(u)w+=𝒪j(u)w+μj(u)(v)w+=(v)w+\mathcal{L}(v)\cdot w_{+}=\mathcal{L}(v)\cdot w_{+}\overset{\mu_{i}(u)^{-1}}{\longleftrightarrow}\mathcal{O}_{i}(u)\cdot w_{+}=\mathcal{O}_{j}(u)\cdot w_{+}\overset{\mu_{j}(u)}{\longleftrightarrow}\mathcal{L}(v^{\prime})\cdot w_{+}=\mathcal{L}(v^{\prime})\cdot w_{+}
ρv,w1ρv,w(v)w+.\overset{\rho_{v,w}^{-1}\rho_{v^{\prime},w}}{\longleftrightarrow}\mathcal{L}(v)\cdot w_{+}.

Hence, showing the commutativity of the cube reduces to proving the following equalities in the corresponding situations:

μi(v)μi(v)1μj(v)μj(v)1\displaystyle\mu_{i}(v^{\prime})\circ\mu_{i}(v^{\prime})^{-1}\circ\mu_{j}(v)\circ\mu_{j}(v)^{-1} =Id,\displaystyle=\mathrm{Id},
ρv,w1μi(v)hu,vμi(u)1\displaystyle\rho_{v,w}^{-1}\circ\mu_{i}(v^{\prime})\circ h_{u,v^{\prime}}\circ\mu_{i}(u)^{-1} =Id,\displaystyle=\mathrm{Id},
ρv,w1ρv,wμj(u)μi(u)1\displaystyle\rho_{v,w}^{-1}\circ\rho_{v^{\prime},w}\circ\mu_{j}(u)\circ\mu_{i}(u)^{-1} =Id.\displaystyle=\mathrm{Id}.

The first line holds on the nose, whereas the second and third are a consequence of the commutativity of the following squares, proved in Lemma 4.14:

Proposition 5.6.

γ:C(M;)C(F;)\gamma_{*}\colon C_{*}(M;\mathbb{Z})\to C_{*}(F;\mathbb{Z}) is an isomorphism of chain complexes.

Proof.

Given a state u𝟐𝐧u\in\mathbf{2^{n}} with Φ(u)=0\Phi(u)=0 and a vertex zZ(u)z\in Z(u) in G(u)G(u), write CzC_{z} for the component of G(u)G(u) containing zz and write Ez(u)E_{z}(u) for the set of edges incident to zz. For each edge eEz(u)e\in E_{z}(u), set

σ(e,z)\displaystyle\sigma(e,z) ={0if ze+,1if ze+,\displaystyle=\begin{cases}0&\mbox{if }z\in e_{+},\\ 1&\mbox{if }z\notin e_{+},\end{cases} σ¯(e,z)\displaystyle\bar{\sigma}(e,z) ={1if ze+,0if ze+.\displaystyle=\begin{cases}1&\mbox{if }z\in e_{+},\\ 0&\mbox{if }z\notin e_{+}.\end{cases}

The inverse of γ\gamma_{*} is:

γ1(z)\displaystyle\gamma_{*}^{-1}(z) =eEz(u)(1)σ(e,z)e+(1eEz(u)σ¯(e,z))Cz,\displaystyle=\sum_{e\in E_{z}(u)}(-1)^{\sigma(e,z)}e+\left(1-\sum_{e\in E_{z}(u)}\bar{\sigma}(e,z)\right)C_{z},
γ1(u+)\displaystyle\gamma_{*}^{-1}(u_{+}) =u+.\displaystyle=u_{+}.\qed

The above proposition together with Remark 3.4 yield to the following result:

Corollary 5.7.

The spectra TotF\operatorname{Tot}F and TotM\operatorname{Tot}M are homotopy equivalent.

6. Pointed semi-simplicial sets

Definition 6.1.

A link diagram DD is 11-adequate (resp. 0-adequate) if 𝒢(1)\mathcal{G}(\vec{1}) (resp. 𝒢(0)\mathcal{G}(\vec{0})) contains no loops. DD is said to be adequate if it is both 0-adequate and 11-adequate. If DD is either 0-adequate or 11-adequate, it is called semiadequate. A link is said to be (semi)adequate if it admits a (semi)adequate diagram.

Proposition 6.2.

The functor FF factors through 𝐒𝐞𝐭\mathbf{Set}_{\bullet} if and only if the diagram DD is 11-adequate. The functor MM factors through 𝐒𝐞𝐭\mathbf{Set}_{\bullet} if and only if D(1)D(\vec{1}) contains no alternatig pairs.

Proof.

From the definition of FF it follows that Fu>vF_{u>v} is not a function of pointed sets if and only if Φ(u)=Φ(v)=0\Phi(u)=\Phi(v)=0. This situation is avoided for all possible states u>vu>v if and only if Φ(v)>0\Phi(v)>0 for every state v1v\neq\vec{1}, or equivalently, if and only if 𝒢(1)\mathcal{G}(\vec{1}) contains no loops.

Now, looking at the definition of MM, we get that Mu>vM_{u>v} is not a function of pointed sets if and only if Φ(u)=0\Phi(u)=0, Φ(v)=1\Phi(v)=1 and |(v)|=2|\mathcal{L}(v)|=2. This situation is avoided for all possible states u>vu>v if and only if D(1)D(\vec{1}) contains no alternating pairs. ∎

Combining the functor Λ\Lambda from (3.1) with the definition in (5.2) we get the following results:

Corollary 6.3.

If DD is 1-adequate, then 𝒳Djalmax\mathcal{X}_{D}^{j_{\mathrm{almax}}} is an iterated desuspension of the augmented semisimplicial pointed set Λ(F)\Lambda(F) (cf. [PS20]).

Corollary 6.4.

If D(1)D(\vec{1}) has no alternating pairs, then 𝒳Djalmax\mathcal{X}_{D}^{j_{\mathrm{almax}}} is an iterated desuspension of the augmented semisimplicial pointed set Λ(M)\Lambda(M).

7. Breaking up MDM_{D}

In this section we introduce skein sequences for MDM_{D} (Section 7.3). To do so, we decompose MDM_{D} into subposets of the cube 𝟐𝐧\mathbf{2^{n}} (Section 7.2), and compare one of them with the simplicial complex IDI_{D} introduced in [GMS18] for extreme Khovanov homology.

7.1. The simplicial complex IDI_{D}

Given a link diagram DD, in [GMS18] authors introduce a simplicial complex whose associated cohomology complex coincides with the Khovanov homology of the link in the minimal quantum grading. We restate that construction in terms of the maximal quantum grading jmaxj_{\max}.

Definition 7.1.

Let DD be a link diagram. Its associated Lando graph GL(D)G_{L}(D) is constructed from D(1)D(\vec{1}) by considering a vertex for every monochord, and an edge joining two vertices if the endpoints of the corresponding monochords alternate along the same circle. We define the independence complex666The independence complex is defined in terms of a graph. Hence, given a graph GG we define its associated independence complex IGI_{G} as described in Definition 7.1. associated to DD, IDI_{D}, as the simplicial complex whose set of vertices is the same as the set of vertices of GL(D)G_{L}(D) and σ=(e1e2ek)\sigma=(e_{1}e_{2}\cdots e_{k}) is a simplex in IDI_{D} if and only if the vertices e1,e2,,eke_{1},e_{2},\ldots,e_{k} are independent in GL(D)G_{L}(D), i.e., there are not edges in GL(D)G_{L}(D) between these vertices.

Note that GL(D)G_{L}(D) can be thought as the disjoint union of the Lando graphs arising from each of the chord diagrams in D(1)D(\vec{1}) after removing all bichords. Moreover, if GL(D)=GL(D1)GL(D2)G_{L}(D)=G_{L}(D_{1})\sqcup G_{L}(D_{2}), then ID=ID1ID2I_{D}=I_{D_{1}}*I_{D_{2}}, the join of both simplicial complexes.

Theorem.

[GMS18] Let LL be an oriented link represented by a diagram DD with pp positive crossings. Then

Khi,jmax(D)H~pi1(ID).Kh^{i,j_{\max}}(D)\approx\tilde{H}_{p-i-1}(I_{D}).

The poset of faces of the independence complex IDI_{D} is precisely the subposet of the cube 𝟐𝐧\mathbf{2^{n}} of those states uu for which Φ(u)=0\Phi(u)=0.

Given a set of vertices {v1,,vn}\{v_{1},\ldots,v_{n}\}, each simplicial complex on these vertices gives rise to a downwards closed subposet of 𝟐𝐧\mathbf{2^{n}} (i.e., if a state uu belongs to the subposet and u>vu>v, then vv belongs to the subposet too): its poset of faces. Conversely, every downwards closed subposet of 𝟐𝐧\mathbf{2^{n}} is the poset of faces of some simplicial complex.

Let 𝐒𝐒𝐞𝐭\mathbf{S}\subset\mathbf{Set} be the full subcategory on \emptyset and a singleton, and let 𝐒𝐩𝐒𝐞𝐭\mathbf{S_{p}}\subset\mathbf{Set}_{\bullet} be the full subcategory on the basepoint and S0S^{0}. Downwards closed subposets of the cube are in bijection with functors 𝟐𝐧𝐒\mathbf{2^{n}}\to\mathbf{S}. Subposets of the cube are in bijection with functors 𝟐𝐧𝐒𝐩\mathbf{2^{n}}\to\mathbf{S_{p}} (see discussion in the final section of [CMS20], for example). The realization of a subposet of the cube is the desuspension of the totalization of its associated functor 𝟐𝐧𝐒𝐩𝐒𝐞𝐭\mathbf{2^{n}}\to\mathbf{S_{p}}\subset\mathbf{Set}_{\bullet}. Thus, the realization of a simplicial complex XX coincides with the realization of its poset of faces.

Given u𝟐𝐧u\in\mathbf{2^{n}}, we write u¯\bar{u} for the state of 𝟐𝐧\mathbf{2^{n}} satisfying u¯iui\bar{u}_{i}\neq u_{i}, for 1in1\leqslant i\leqslant n.

Definition 7.2.

The categorical dual of a downwards (upwards) closed subposet X𝟐𝐧X\subset\mathbf{2^{n}} is the upwards (downwards) closed subposet X𝟐𝐧𝐒𝐩X^{*}\subset\mathbf{2^{n}}\to\mathbf{S_{p}} given by uXu\in X^{*} if and only if u¯X\bar{u}\in X.

The complement of a downwards (upwards) closed subposet X𝟐𝐧X\subset\mathbf{2^{n}} is the upwards (downwards) closed subposet X^𝟐𝐧\hat{X}\subset\mathbf{2^{n}} given by uX^u\in\hat{X} if and only if uXu\notin X.

The complement of the categorical dual of a downwards closed set XX is again a downwards closed subposet of the cube whose associated simplicial complex is the Alexander dual of XX. In general |X^|Σ|X||\hat{X}|\simeq\Sigma|X| if XX is downwards closed and |X||X^{*}| is Spanier-Whitehead dual to |X||X|. Observe that

  1. (1)

    If |X||A||B||X|\simeq|A|\vee|B|, then |X||A||B||X^{*}|\simeq|A^{*}|\vee|B^{*}|,

  2. (2)

    If |X|Sk|X|\simeq S^{k}, then |X|Snk2|X^{*}|\simeq S^{n-k-2}, where nn is the dimension of the cube.

Given a link diagram DD, the upwards closed subposet of the cube given by those states uu such that Φ(u)=0\Phi(u)=0, which we denote XDX_{D}, corresponds to the functor Fjmax:𝟐𝐧𝐒𝐞𝐭F^{j_{\max}}\colon\mathbf{2^{n}}\to\mathbf{Set}_{\bullet}\subset\mathcal{B}.

Corollary 7.3.

XDX_{D} is the categorical dual of the poset of faces of IDI_{D}. As a consequence if   |ID|Sk|I_{D}|\simeq S^{k}, then |XD|Snk2|X_{D}|\simeq S^{n-k-2}.

7.2. Decomposing MDM_{D}

Given a state uu of a diagram DD, recall that we write D(1)uD(\vec{1})_{u} for the chord diagram having the same circles as D(1)D(\vec{1}) and those chords ei(1)e_{i}(\vec{1}) so that ui1u_{i}\neq 1. Given a crossing cc in DD, we write D[c=0]D[c=0] (resp. D[c=1]D[c=1]) for the link diagram obtained after smoothing cc following a 0-label (resp. 11-label).

Definition 7.4.

Given a link diagram DD with nn crossings, we consider the following subposets of the cube 𝟐𝐧\mathbf{2^{n}}:

  • XDX_{D}: is the subposet of 𝟐𝐧\mathbf{2^{n}} consisting of those states uu so that Φ(u)=0\Phi(u)=0 (as defined in previous section).

  • XDe=XD[e=0]X_{D}^{e}=X_{D[e=0]}, for any monochord eD(1)e\in D(\vec{1}).

  • YDY_{D}: is the subposet of 𝟐𝐧\mathbf{2^{n}} consisting of those states uu so that Φ(u)=1\Phi(u)=1 and D(1)uD(\vec{1})_{u} contains an alternating pair.

  • ZDbZ_{D}^{b}: is the subposet of 𝟐𝐧\mathbf{2^{n}} consisting of those states uu so that Φ(u)=1\Phi(u)=1 and D(1)uD(\vec{1})_{u} contains a bichord parallel to a given bichord bb of D(1)D(\vec{1}).

Observe that each state (admitting an enhacement) in the almost-extreme complex of DD belongs to one and only one of the previous subcubes XD,YDX_{D},Y_{D} or ZDZ_{D} (see Corollary 4.5 and Proposition 4.8).

Proposition 7.5.

The functor MDM_{D} can be realized as the following cofibre sequence

(7.1) Σ1(Z(1)XDeNXDe)YD[b]BZDbMD,\displaystyle\Sigma^{-1}\left(\bigvee_{Z(\vec{1})}X_{D}\vee\bigvee_{e\in N}X_{D}^{e}\right)\longrightarrow\displaystyle Y_{D}\vee\bigvee_{[b]\in B}Z^{b}_{D}\longrightarrow M_{D},

where NN is the set of monochords in G(1)G(\vec{1}) and BB the set of classes of parallel bichords.

The decomposition follows because the second map is levelwise injective and the suspension of the leftmost term is precisely the quotient functor of this second map.

Remark 7.6.

A Mayer-Vietoris spectral sequence allows to compute both posets YDY_{D} and ZDbZ_{D}^{b} in terms of some posets XDX_{D^{\prime}}, where DD^{\prime} are partially smoothed link diagrams of DD. We explain this below.

The poset YDY_{D} can be covered with the subposets XD[e=0,f=0]X_{D[e=0,f=0]} where {e,f}\{e,f\} runs along the set of all alternating pairs in D(1)D(\vec{1}). The kk-fold intersections AkA_{k} of this covering are

Ak={e1,,ek+1}XD[e1=0,,ek+1=0],A_{k}=\bigvee_{\{e_{1},\ldots,e_{k+1}\}}X_{D[e_{1}=0,\ldots,e_{k+1}=0]},

where {e1,,ek+1}\{e_{1},\ldots,e_{k+1}\} are monochords attached to the same circle in such a way that they can be divided into two non-empty sets H1H_{1} and H2H_{2} so that all monochords in H1H_{1} alternate with all monochords in H2H_{2} and viceversa. Similarly, the poset ZDbZ_{D}^{b} can be covered with the subposets XD[b=0]X_{D[b=0]} where bb runs along the set of bichords of D(1)D(\vec{1}), and the kk-fold intersections BkB_{k} of this covering are

Bk={b1,,bk}XD[b1=0,,bk=0],B_{k}=\bigvee_{\{b_{1},\ldots,b_{k}\}}X_{D[b_{1}=0,\ldots,b_{k}=0]},

where {b1,,bk}\{b_{1},\ldots,b_{k}\} are bichords parallel to bb.

Thus, the (first page of the) Mayer-Vietoris spectral sequences associated to these coverings are:

Hp({e1,,eq+1}XD[e1=0,,eq+1=0])\displaystyle H^{p}\left(\bigvee_{\{e_{1},\ldots,e_{q+1}\}}X_{D[e_{1}=0,\ldots,e_{q+1}=0]}\right) Hp+q(YD),\displaystyle\Rightarrow H^{p+q}(Y_{D}),
Hp({b1,,bq}XD[b1=0,,bq=0])\displaystyle H^{p}\left(\bigvee_{\{b_{1},\ldots,b_{q}\}}X_{D[b_{1}=0,\ldots,b_{q}=0]}\right) Hp+q(ZDb).\displaystyle\Rightarrow H^{p+q}(Z^{b}_{D}).
Example 7.7.

Consider the torus knot T(3,q)T(3,q) and let D=D(3,q)D=D_{(3,q)} be its standard diagram. Since D(1)D(\vec{1}) contains no bichords, ZDbZ_{D}^{b} is trivial. We will combine Remark 7.6 together with Corollary 7.3 to compute the realization of subposets in (7.1). More precisely, we will express XDX_{D}, XDeX_{D}^{e} and YDY_{D} (via AkA_{k}) as the duals of independence complexes of some Lando graphs. Write CnC_{n} for the cycle graph of nn vertices and LnL_{n} the path of length nn.

  • -

    GL(D)=C2qG_{L}(D)=C_{2q};

  • -

    GL(D[e=0])=L2q4G_{L}(D[e=0])=L_{2q-4}, for any monochord ee in D(1)D(\vec{1});

  • -

    GL(D[e1=0,e2=0])=C2q2G_{L}(D[e_{1}=0,e_{2}=0])=C_{2q-2} when e1e_{1} and e2e_{2} are two consecutive monochords in D(1)D(\vec{1}). Note that alternating pairs consist precisely in consecutive monochords;

  • -

    GL(D[e1=0,e2=0,e3=0])=L2q6G_{L}(D[e_{1}=0,e_{2}=0,e_{3}=0])=L_{2q-6} when e1,e2,e3e_{1},e_{2},e_{3} are three consecutive monochords in D(1)D(\vec{1}). Note that the previous condition is equivalent to require that e1,e2,e3e_{1},e_{2},e_{3} can be divided into two subsets H1H_{1} and H2H_{2} such that all monochords in each subset alternate with all monochords in the other subset.

Note that there are no nn-tuples of monochords e1,,ene_{1},\ldots,e_{n} which can be divided into two subsets H1H_{1} and H2H_{2} satisfying the condition above when n>3n>3.

Summarizing we have:

XD\displaystyle X_{D} =IC2q\displaystyle=I^{*}_{C_{2q}} XDe\displaystyle X_{D}^{e} =IL2q4\displaystyle=I^{*}_{L_{2q-4}} Ak\displaystyle A_{k} ={2qIC2q2if k=1,2qIL2q6if k=2,if k>2.\displaystyle=\begin{cases}\bigvee_{2q}I^{*}_{C_{2q-2}}&\text{if $k=1$},\\ \bigvee_{2q}I^{*}_{L_{2q-6}}&\text{if $k=2$},\\ \emptyset&\text{if $k>2$.}\end{cases}

The homotopy types of the above complexes were computed in [PS18, Corollary 3.4 and Proposition 3.9]:

|ICn|\displaystyle|I_{C_{n}}| {Sk1if n=k±1,Sk1Sk1if n=3k.\displaystyle\simeq\begin{cases}S^{k-1}&\text{if $n=k\pm 1$},\\ S^{k-1}\vee S^{k-1}&\text{if $n=3k$}.\end{cases} |ILn|\displaystyle|I_{L_{n}}| {Skif n=3k+1,3k+2,if n=3k.\displaystyle\simeq\begin{cases}S^{k}&\text{if $n=3k+1,3k+2$},\\ *&\text{if $n=3k$}.\end{cases}

This allows to compute the almost-extreme Khovanov spectra of T(3,q)T(3,q) in terms of well-known independence complexes of cycles and paths, with no need to apply induction on qq.

Remark 7.8.

The cone-length c(𝒳)c(\mathcal{X}) of a spectrum 𝒳\mathcal{X} is the least nn such that there is a sequence of cofibre sequences

𝒴i𝒳i𝒳i+1\mathcal{Y}_{i}\to\mathcal{X}_{i}\to\mathcal{X}_{i+1}

for i=0,,n1i=0,\ldots,n-1, such that 𝒳0\mathcal{X}_{0} is contractible, 𝒴i\mathcal{Y}_{i} is a wedge of spheres and 𝒳n=𝒳\mathcal{X}_{n}=\mathcal{X}. In [PS18] it was conjectured that |XD||X_{D}| is homotopy equivalent to a wedge of spheres for any diagram DD. If this were true, then Remark 7.6 would imply that the cone length of TotMD\operatorname{Tot}M_{D} is bounded above by the maximum of the following numbers:

  1. (1)

    the maximum number of parallel bichords plus one in D(1)D(\vec{1}).

  2. (2)

    the maximum number of monochords in D(1)D(\vec{1}) that can be partitioned into two disjoint subsets H1,H2H_{1},H_{2} such that all chords in H1H_{1} alternate with all chords in H2H_{2} and viceversa.

7.3. Skein sequences

Let aa be a crossing in a link diagram DD. Observe that if aa is a monochord in D(1)D(\vec{1}) we get the skein short exact sequence777Compare to skein short exact sequences from [Vir04].

(7.2) MD[a=1]MD[a=0]MD,M_{D[a=1]}\longrightarrow M_{D[a=0]}\longrightarrow M_{D},

whereas if aa is a bichord in D(1)D(\vec{1}), then the skein short exact sequence becomes

(7.3) MD[a=1]XD[a=0]MD.M_{D[a=1]}\longrightarrow X_{D[a=0]}\longrightarrow M_{D}.

Moreover, when aa is a monochord in D(1)D(\vec{1}) we also have the following sequence:

(7.4) XD[a=1]XD[a=0]XD.X_{D[a=1]}\longrightarrow X_{D[a=0]}\longrightarrow X_{D}.

To verify the three above sequences note that the second map is an inclusion whose quotient is the suspension of the first spectrum.

Definition 7.9.

Let DD be a diagram and let aa be a monochord in D(1)D(\vec{1}). We say that aa is:

  1. (1)

    22-free if it is not part of any alternating pair in D(1)D(\vec{1}).

  2. (2)

    33-free if it is not part of any alternating triple in D(1)D(\vec{1}).

  3. (3)

    free if it is 22-free and 33-free.

  4. (4)

    bb-free if it is not part of any alternating triple involving the bichord bb in D(1)D(\vec{1}).

Lemma 7.10.

Let DD be a diagram and let aa be a monochord in D(1)D(\vec{1}).

  1. (1)

    If aa is 22-free, then XD,YDX_{D},Y_{D} and all XDeX_{D}^{e}, with eae\neq a a monochord, are contractible; if, additionally, D(1)D(\vec{1}) contains another 22-free monochord, then XDaX^{a}_{D} is contractible too.

  2. (2)

    If aa does not form an alternating triple with a bichord bb (i.e., aa is bb-free), then ZDbZ^{b}_{D} is contractible.

Proof.

The leftmost map in skein sequence (7.4) along the crossing associated to aa is identity, hence XDX_{D} is contractible. A similar reasoning works for XDeX_{D}^{e} when eae\neq a.

In general, the skein sequence (7.2) does not respect the decomposition in (7.1); however, when aa is 22-free, it respects the term YDY_{D} and when aa is bb-free it respects ZDbZ_{D}^{b}. In both cases the maps YD[a=1]YD[a=0]Y_{D[a=1]}\to Y_{D[a=0]} and ZD[a=1]bZD[a=0]bZ^{b}_{D[a=1]}\to Z^{b}_{D[a=0]} are the identity. ∎

Corollary 7.11.

If D(1)D(\vec{1}) contains at least two 2-free monochords, then

MD[b]BZDb.M_{D}\simeq\bigvee_{[b]\in B}Z^{b}_{D}.
Proof.

The result is direct from cofibre sequence (7.1) and Lemma 7.10(1). ∎

Corollary 7.12.

If DD has a free monochord ee, then MDXD[e=0]M_{D}\simeq X_{D[e=0]}.

Proof.

Lemma 7.10 implies that the only possibly non-contractible subposet in (7.1) is XDe=XD[e=0]X_{D}^{e}=X_{D[e=0]}. ∎

Lemma 7.13.

Let aa be a 2-free monochord in D(1)D(\vec{1}) dividing a circle into two regions so that both of them contain at least one 2-free monochord (we say that D(1)D(\vec{1}) contains nested monochords). Then MDΣMD[a=1]M_{D}\simeq\Sigma M_{D[a=1]}.

Proof.

We use skein sequence (7.2) and show that MD[a=0]M_{D[a=0]} is contractible. First, notice that the circle in the statement is splitted into two circles z1z_{1} and z2z_{2} when changing the 11-label of aa for a 0-label, each of them attached to at least a 22-free monochord that we call cc and dd, respectively. Lemma 7.10(1) implies that the the only possibly non-contractible spaces when applying cofibre sequence (7.1) to D[a=0]D_{[a=0]} are ZD[a=0]bZ_{D[a=0]}^{b}.

Now, since aa is 2-free in D(1)D(\vec{1}), there is no bichord connecting z1z_{1} and z2z_{2} in D[a=0](1)D[a=0](\vec{1}). Therefore, for any bichord bb in D[a=0](1)D[a=0](\vec{1}) at least one of the monochords cc or dd are bb-free, so Lemma 7.10(2) implies that ZD[a=0]bZ_{D[a=0]}^{b} is contractible for all bichords bb, thus MD[a=0]M_{D[a=0]} is contractible. ∎

Definition 7.14.

Two parallel bichords aa and bb of a chord diagram D(1)D(\vec{1}) are equivalent if there is no monochord ee so that a,ba,b and ee constitute an alternating triple.

Lemma 7.15.

If DD has two equivalent bichords a,ba,b, then MDΣMD[a=1]M_{D}\simeq\Sigma M_{D[a=1]}.

Proof.

Since aa and bb are equivalent, the chord bb is 22-free in the chord diagram D[a=0](1)D[a=0](\vec{1}), so Lemma 7.10(1) implies that XD[a=0]X_{D[a=0]} is contractible. The statement follows from the skein sequence (7.3) along the chord aa. ∎

8. Diagrams with no alternating pairs

In this section we determine the homotopy type of MDM_{D} (thus, that of the Khovanov spectrum given by Lipshitz and Sarkar in [LS14] for the almost-extreme quantum grading) for diagrams DD so that D(1)D(\vec{1}) contains no alternating pairs. Observe that this determines their Khovanov homology groups at almost-extreme quantum degree.

8.1. Diagrams with no monochords

This case was studied in [PS20]: If DD is 11-adequate (i.e., D(1)D(\vec{1}) has no monochords), then MDM_{D} is homotopy equivalent to a wedge of spheres and possibly a suspension of the projective plane. See [PS20, Theorem 1.1] for the details.

8.2. Diagrams with one monochord

Proposition 8.1.

Let DD be a link diagram with nn crossings so that D(1)D(\vec{1}) contains a single monochord attached to a circle zz, and let kk be the number of circles connected to zz along an alternating triple. Then,

|MD|k1Sn3.|M_{D}|\simeq\bigvee_{k-1}S^{n-3}.
Proof.

Write ee for the monochord in D(1)D(\vec{1}). By Lemma 7.10(1), the decomposition in (7.1) becomes

Σ1XDe[b]BZDbMD.\Sigma^{-1}X^{e}_{D}\longrightarrow\bigvee_{[b]\in B}Z^{b}_{D}\longrightarrow M_{D}.

The subposet XDeX^{e}_{D} has a single element in degree n1n-1, thus |XDe|Sn2|X^{e}_{D}|\simeq S^{n-2}. By Lemma 7.10(2), ZDbZ^{b}_{D} is contractible unless bb forms an alternating triple with ee. Therefore we may assume, up to suspension, that all bichords are attached to zz and form an alternating triple with ee. Moreover, using Lemma 7.15 we can assume, up to suspension, that circles are connected to zz by exactly two bichords.

Therefore, for each such bichord bb, the subposet ZDbZ^{b}_{D} has five elements in M-shape, with two elements in degree n1n-1 and three in degree n2n-2, thus |ZDb|Sn3|Z^{b}_{D}|\simeq S^{n-3}, and the above decomposition becomes

Sn3[b]BSn3|MD|.S^{n-3}\longrightarrow\bigvee_{[b]\in B}S^{n-3}\longrightarrow|M_{D}|.

The map to the wedge of spheres is a diagonal map. This concludes the proof. ∎

8.3. Diagrams with 2-free monochords

Let DD be a diagram whose associated chord diagram D(1)D(\vec{1}) contains more than one monochord, all of them 2-free. Recall from Corollary 7.11 that in this situation MD[b]BZDbM_{D}\simeq\bigvee_{[b]\in B}Z^{b}_{D}, thus we can restrict to the independent study of each connected pair of circles when computing ZDbZ_{D}^{b}.

At this point we can make some simplifications in D(1)D(\vec{1}): Lemmas 7.13 and 7.15 allow us to remove nested monochords and equivalent bichords when computing MDM_{D} (in exchange of taking suspensions). Moreover, by Lemma 7.10(2) we can assume that D(1)D(\vec{1}) contains no bb-free monochords for any bichord bb. These simplifications motivate the following definition:

Definition 8.2.

A diagram DD is simple if the associated chord diagram D(1)D(\vec{1}) contains at least two monochords, all of them 2-free, and satisfies the following conditions:

  1. (1)

    It contains exactly two circles;

  2. (2)

    It contains no nested monochords;

  3. (3)

    There are no bb-free monochords for any bichord bb;

  4. (4)

    There are no equivalent bichords.

By definition, if DD is simple then D(1)D(\vec{1}) can be isotoped in S2S^{2} so that monochords and bichords lie in different regions (see Figure 13). Moreover, since D(1)D(\vec{1}) contains no nested monochords, a circle with nn monochords is divided into nn half-disks (each of them bounded by a monochord and an arc of the circle) and an additional region (bounded by the nn monochords and nn arcs of the circle) that we call polygon.

Given a monochord ee in D(1)D(\vec{1}), we write ded_{e} for the half-disk bounded by ee888If ee is the only monochord attached to the circle, then there are two possible options for ded_{e}; in this case choose an option minimizing |de||d_{e}|, without loss of generality., and |de||d_{e}| for the number of bichords having one of their endpoints in ded_{e}. In simple diagrams |de|1|d_{e}|\geqslant 1 for any monochord ee.

In the proofs of Lemmas 8.3, 8.7 and 8.8 we use several results from [PS18, Section 3].

Refer to caption
Figure 13. Two chord diagrams corresponding to simple diagrams. Grey regions correspond to half-disks bounded by monochords. In (a)(a) |de|=3|d_{e}|=3, while |de|=1|d_{e^{\prime}}|=1 for every monochord ee^{\prime} in (b)(b).
Lemma 8.3.

Let DD be a simple diagram whose associated chord diagram D(1)D(\vec{1}) contains a monochord ee so that |de|=1|d_{e}|=1. Then |MD||M_{D}| is either contractible or homotopy equivalent to a sphere.

Proof.

Let bb be the only bichord with and endpoint in ded_{e} and consider the skein sequence (7.3) along bb:

MD[b=1]XD[b=0]MD.M_{D[b=1]}\longrightarrow X_{D[b=0]}\longrightarrow M_{D}.
Refer to caption
Figure 14. The 1\vec{1}-resolutions corresponding to DD, D[b=0]D[b=0] and D[b=0,e=0]D[b=0,e=0] illustrating the proof of Lemma 8.3 are shown in (a)(a), (b)(b) and (c)(c), respectively. Regions RR and RR^{\prime} are shaded green and grey, respectively.

The monochord ee is free in D[b=1](1)D[b=1](\vec{1}), thus by Lemma 7.10 ZD[b=1]Z_{D[b=1]} is contractible, thus MD[b=1]M_{D[b=1]} is contractible (by Corollary 7.11) and MDXD[b=0]M_{D}\simeq X_{D[b=0]}.

Now, D[b=0](1)D[b=0](\vec{1}) has a single circle and its inner part is separated into two regions RR and RR^{\prime} by the monochord ee. See Figure 14(b). Moreover:

  1. (1)

    monochords in D(1)D(\vec{1}) become internal monochords in D[b=0](1)D[b=0](\vec{1});

  2. (2)

    monochords attached to each of the circles in D(1)D(\vec{1}) become monochords contained in each of the regions R,RR,R^{\prime} in D[b=0](1)D[b=0](\vec{1});

  3. (3)

    bichords in D(1)D(\vec{1}) become external monochords in D[b=0](1)D[b=0](\vec{1}) having one endpoint in RR and the other one in RR^{\prime}.

Consider now the skein sequence (7.4) along the monochord ee

XD[b=0,e=1]XD[b=0,e=0]XD[b=0].X_{D[b=0,e=1]}\longrightarrow X_{D[b=0,e=0]}\longrightarrow X_{D[b=0]}.

Condition (3) above implies that there are no alternating pairs in D[b=0,e=0](1)D[b=0,e=0](\vec{1}) (see Figure 14(c)), so the middle term in the above sequence is contractible and we have XD[b=0]ΣXD[b=0,e=1]X_{D[b=0]}\simeq\Sigma X_{D[b=0,e=1]}.

Assume first that bb connects ded_{e} with the region that we call polygon, then we claim that the Lando graph GG of D[b=0,e=1]D[b=0,e=1], contains no cycles (i.e., it is a forest). In order to prove that, we proceed by contradiction. Suppose that CnC_{n} is a cycle of length nn in GG, with vertices e1,e2,,ene_{1},e_{2},\ldots,e_{n} numbered so that e1e_{1} is an internal chord in the region RR. Conditions (2) and (3) imply that monochords e4k+1e_{4k+1} and e4k+3e_{4k+3} lie in RR and RR^{\prime}, respectively. Moreover, the absence of equivalent bichords in D(1)D(\vec{1}) implies the absence of cycles of length four in GG, i.e., n8n\geqslant 8. The monochord e4e_{4} separates the external region into two subregions S,SS,S^{\prime}. Since there are no cycles of length 44, e2e_{2} and e6e_{6} lie one in each of these two subregions, say e2Se_{2}\in S and e6Se_{6}\in S^{\prime}. Moreover, as the internal monochords are non-nested, all internal monochords but e3e_{3} and e5e_{5} have both endpoints in either SS or SS^{\prime}. Therefore

  1. (1)

    the endpoints of e1e_{1} lie in SS (because e2e_{2} lies in SS) and

  2. (2)

    the endpoints of e2k+1e_{2k+1} lie in SS^{\prime} if k3k\geqslant 3 (because e2ke_{2k} lies in SS^{\prime} if k3k\geqslant 3).

In particular, both endpoints of en1e_{n-1} lie in SS^{\prime}. This is a contradiction, since ene_{n} is a external monochord connecting e1e_{1} and en1e_{n-1}. Therefore GG contains no cycles, and by [PS18, Corollary 3.7] its associated independence complex is either contractible or a sphere. Corollary 7.3 completes the proof.

Assume now that bb has its endpoints in two half-discs ded_{e} and dfd_{f}, for some monochord ff, and consider the skein sequence (7.4) along ff. We get that XD[b=0,e=1,f=0]X_{D[b=0,e=1,f=0]} is contractible, since all monochords in RR^{\prime} are 2-free. Therefore, XD[b=0,e=1]ΣXD[b=0,e=1,f=1]X_{D[b=0,e=1]}\simeq\Sigma X_{D[b=0,e=1,f=1]}. The latter is the same as XD[b=0,e=1]X_{D^{\prime}[b=0,e=1]}, where D=D[f=1]D^{\prime}=D[f=1] is a diagram where ee connects ded_{e} with the polygon, which was studied in the previous case. ∎

Remark 8.4.

Since simple diagrams do not contain equivalent bichords, if |de|4|d_{e}|\geqslant 4 for some monochord ee, then there exists a monochord ff with |df|=1|d_{f}|=1 (dfd_{f} is connected to ded_{e} by a bichord). In other words, if |de|>1|d_{e}|>1 for all monochords ee, then |de|3|d_{e}|\leqslant 3 for all monochords ee.

Definition 8.5.

A simple diagram DD so that 2|de|32\leqslant|d_{e}|\leqslant 3 for every monochord ee of D(1)D(\vec{1}) is called super-simple.

Refer to caption
Figure 15. The three possible situations for a half-disk ded_{e} in the 1-resolution of a super-simple diagram are illustrated in Figures (a), (b) and (c). Situation (d) is not allowed in semi-simple diagrams.

If DD is super-simple, every half-disk ded_{e} in D(1)D(\vec{1}) satisfies one of the three following conditions:

  1. (1)

    |de|=2|d_{e}|=2, and both bichords connect ded_{e} with two contiguous half-disks in the other circle (see Figure 15(a));

  2. (2)

    |de|=2|d_{e}|=2, and one bichord connects ded_{e} with a half-disk in the other circle and the other one connects ded_{e} with the polygon in the other circle (see Figure 15(b));

  3. (3)

    |de|=3|d_{e}|=3, and two bichords connect ded_{e} with two contiguous half-disks in the other circle, and the third one is placed between them and connects ded_{e} with the polygon of the other circle (see Figure 15(c)).

The next statement follows from Lemmas 8.7 and 8.8.

Proposition 8.6.

Let DD be a super-simple diagram. Then |MD||M_{D}| is homotopy equivalent to a wedge of spheres.

Lemma 8.7.

Let DD be a super-simple diagram so that D(1)D(\vec{1}) contains at least one bichord connecting a half-disk with the region called polygon. Then |MD||M_{D}| is either contractible or homotopy equivalent to a sphere.

Proof.

Let bb be a bichord connecting a half-disk ded_{e} in a circle z1z_{1} with the polygon of the circle z2z_{2}, for some monochord ee. See Figure 16(a)(a). Since DD is super-simple, there exists at least a bichord aa connecting ded_{e} with dfd_{f} for a monochord ff attached to z2z_{2}. We distinguish two cases, depending on whether z2z_{2} contains at least two monochords or it contains just a single monochord.

Case 1: Assume that the chord diagram D(1)D(\vec{1}) contains more than one monochord attached to z2z_{2}, and consider the skein exact sequence (7.3) along aa:

MD[a=1]XD[a=0]MD.M_{D[a=1]}\longrightarrow X_{D[a=0]}\longrightarrow M_{D}.
Refer to caption
Figure 16. The 1\vec{1}-resolutions corresponding to DD, D[a=0,f=1]D[a=0,f=1] and D[a=0,f=0]D[a=0,f=0] illustrating the proof of Lemma 8.7 are shown in (a),(b)(a),(b), and (c)(c), respectively. Dotted lines represent those chords that may or may not be in the diagrams.

We will show that XD[a=0]X_{D[a=0]} is contractible. To do so, we apply the skein sequence (7.4) along the monochord ff:

XD[a=0,f=1]XD[a=0,f=0]XD[a=0].X_{D[a=0,f=1]}\longrightarrow X_{D[a=0,f=0]}\longrightarrow X_{D[a=0]}.

Figure 16(b)(b) represents D[a=0,f=1](1)D[a=0,f=1](\vec{1}), where bb is a free monochord. Moreover, any monochord different from ff attached to z2z_{2} becomes 2-free in D[a=0,f=0](1)D[a=0,f=0](\vec{1}), as illustrated in Figure 16(c)(c). Therefore by Lemma 7.10(1) XD[a=0,f=1]X_{D[a=0,f=1]} and XD[a=0,f=0]X_{D[a=0,f=0]} are contractible, and so is XD[a=0]X_{D[a=0]}.

As a consequence, MDΣMD[a=1]M_{D}\simeq\Sigma M_{D[a=1]}. If |de|=2|d_{e}|=2 in D(1)D(\vec{1}), then |de|=1|d_{e}|=1 in D[a=1](1)D[a=1](\vec{1}), and the statment holds by Lemma 8.3. If |de|=3|d_{e}|=3 in D(1)D(\vec{1}), then |de|=2|d_{e}|=2 in D[a=1](1)D[a=1](\vec{1}) and it contains a bichord cc connecting ded_{e} with dhd_{h} for a monochord hh (contiguous to ff) attached to the circle z2z_{2}. Applying the skein exact sequence (7.3) along cc and repeating the same reasoning as before leads to MDΣ2MD[a=1,c=1]M_{D}\simeq\Sigma^{2}M_{D[a=1,c=1]}. Since |de|=1|d_{e}|=1 in D[a=1,c=1](1)D[a=1,c=1](\vec{1}), Lemma 8.3 completes the proof for this case.

Case 2: Assume that there is just one monochord ff attached to the circle z2z_{2} in D(1)D(\vec{1}). Then, since DD is super-simple, D(1)D(\vec{1}) is as one of the six chord diagrams depicted in Figure 17 (each picture leads to two possible chord diagrams, depending on whether it contains the bichord pp or not). In order to compute MDM_{D} for each of these situations, we use skein sequence (7.3) together with Corollary 7.3. We also use [PS18, Lemma 3.2, Corollary 3.3] to compute IDI_{D}:

  • -

    If D(1)D(\vec{1}) is as in Figure 17(a)(a) with pp: we consider the skein sequence (7.3) along the bichord cc, and get that |XD[c=0]|S2|X_{D[c=0]}|\simeq S^{2}. Next, we show that MD[c=1]M_{D[c=1]} is contractible by applying the skein sequence (7.3) to D[c=1]D[c=1] along the bichord aa (we use the fact that chord ff is free in D[c=1,a=1](1)D_{[c=1,a=1]}(\vec{1})). Therefore, |MD||XD[c=0]|S2|M_{D}|\simeq|X_{D[c=0]}|\simeq S^{2}.

  • -

    If D(1)D(\vec{1}) is as in Figure 17(a)(a) without pp: we proceed as in the previous case, and get that both MD[c=1]M_{D[c=1]} and XD[c=0]X_{D[c=0]} are contractible, and so is MDM_{D}.

  • -

    If D(1)D(\vec{1}) is as in Figure 17(b)(b) with pp: we consider the skein sequence (7.3) along the bichord aa, and get that |XD[a=0]|S4|X_{D[a=0]}|\simeq S^{4}. Next, we show that MD[a=1]M_{D[a=1]} is contractible by applying the skein sequence (7.3) to D[a=1]D[a=1] along the bichord bb (we use the fact that chord ee is free in D[a=1,b=1](1)D_{[a=1,b=1]}(\vec{1})). Therefore, |MD||XD[a=0]|S4|M_{D}|\simeq|X_{D[a=0]}|\simeq S^{4}.

  • -

    If D(1)D(\vec{1}) is as in Figure 17(b)(b) without pp: the procedure is analogous to the previous one, the only difference is that |XD[a=0]|S3|X_{D[a=0]}|\simeq S^{3} and therefore |MD|S3|M_{D}|\simeq S^{3}.

  • -

    If D(1)D(\vec{1}) is as in Figure 17(c)(c) with pp: we consider the skein sequence (7.3) along the bichord aa, and get that XD[a=0]X_{D[a=0]} is contractible. Next, we show that MD[a=1]M_{D[a=1]} is also contractible by applying the skein sequence (7.3) to D[a=1]D[a=1] along the bichord bb (we use the fact that chord ee is free in D[a=1,b=1](1)D_{[a=1,b=1]}(\vec{1})). Therefore, MDM_{D} is contractible.

  • -

    If D(1)D(\vec{1}) is as in Figure 17(c)(c) without pp: we proceed as in the previous case, and get that XD[a=0]X_{D[a=0]} is contractible and |MD[a=1]|S3|M_{D[a=1]}|\simeq S^{3}. Therefore, |MD|Σ|MD[a=1]|S4|M_{D}|\simeq\Sigma|M_{D[a=1]}|\simeq S^{4}. ∎

Refer to caption
Figure 17. The six possible chord diagrams illustrating Case 2 of the proof of Lemma 8.7.
Lemma 8.8.

Let DD be a super-simple diagram so that all bichords in D(1)D(\vec{1}) connect two half-disks (i.e., there are no bichords with an endpoint in the region called polygon). Then, |MD||M_{D}| is homotopy equivalent to a wedge of spheres. More precisely, if we write 2n2n for the number of monochords in DD, then

|MD|{S8n31S8n31if n0 (mod 3),S8n+131if n1 (mod 3),S8n+232if n2 (mod 3).|M_{D}|\simeq\begin{cases}S^{\frac{8n}{3}-1}\vee S^{\frac{8n}{3}-1}&\text{if }n\equiv 0\text{ (mod 3)},\\ S^{\frac{8n+1}{3}-1}&\text{if }n\equiv 1\text{ (mod 3)},\\ S^{\frac{8n+2}{3}-2}&\text{if }n\equiv 2\text{ (mod 3)}.\end{cases}
Proof.

First, notice that since all bichords in D(1)D(\vec{1}) connect two half-disks, the number of monochords equals the number of bichords. We label the chords as follows (see Figure 18(a)(a)): we write e1,,ene_{1},\ldots,e_{n} (resp. f1,,fn)f_{1},\ldots,f_{n}) for the monochords attached to the circle z1z_{1} (resp. z2z_{2}), and aia_{i} (resp. bib_{i}) for the bichord having its endpoints in the half-disks deid_{e_{i}} and dfid_{f_{i}} (resp. dei+1d_{e_{i+1}} and dfid_{f_{i}}), for 1in1\leqslant i\leqslant n.

Consider the skein sequence (7.3) along the bichord a1a_{1}:

(8.1) MD[a1=1]XD[a1=0]MD.M_{D[a_{1}=1]}\longrightarrow X_{D[a_{1}=0]}\longrightarrow M_{D}.

We study now the homotopy type of XD[a1=0]X_{D[a_{1}=0]}. Figure 18(b)(b) shows the chord diagram D[a1=0](1)D[a_{1}=0](\vec{1}), whose associated Lando graph G=GL(D[a1=0])G=G_{L}(D[a_{1}=0]) is as depicted in Figure 18(c). Now, since vertex e1e_{1} dominates999Following [PS18], given two vertices vv and ww in a graph, we say that vv dominates ww if the adjacent vertices to ww are also adjacent to vv. e2e_{2} and vertex f1f_{1} dominates f2f_{2}, it follows from [PS18, Lemma 3.2] that the independence complex associated to GLG_{L} is homotopy equivalent to the independence complex associated to the graph Ge1f1G-e_{1}-f_{1}, which is a path of length 4n44n-4, L4n4L_{4n-4}, whose independence complex follows from [PS18, Corollary 3.4]:

|IG||IGe1f1|=|IL4n4|{if n1mod3,S4n43otherwise,|I_{G}|\simeq|I_{G-e_{1}-f_{1}}|=|I_{L_{4n-4}}|\simeq\begin{cases}*&\text{if }n\equiv 1\mod 3,\\ S^{\lfloor\frac{4n-4}{3}\rfloor}&\text{otherwise},\end{cases}

and by Corollary 7.3 we get:

|XD[a1=0]|{if n1mod3,S4n34n43otherwise.|X_{D[a_{1}=0]}|\simeq\begin{cases}*&\text{if }n\equiv 1\mod 3,\\ S^{4n-3-\lfloor\frac{4n-4}{3}\rfloor}&\text{otherwise}.\end{cases}
Refer to caption
Figure 18. The chord diagram D(1)D(\vec{1}) satisfying conditions in Lemma 8.8 is shown in (a)(a). The chord diagram and the Lando graph associated to D[a1=0]D[a_{1}=0] are shown in (b)(b) and (c)(c), respectively.

Next, we compute the homotopy type of MD[a1=1]M_{D[a_{1}=1]}. Consider the skein sequence (7.3) along the bichord b1b_{1}:

MD[a1=1,b1=1]XD[a1=1,b1=0]MD[a1=1].M_{D[a_{1}=1,b_{1}=1]}\longrightarrow X_{D[a_{1}=1,b_{1}=0]}\longrightarrow M_{D[a_{1}=1]}.
Refer to caption
Figure 19. The chord diagram associated to D[a1=1,b1=1]D[a_{1}=1,b_{1}=1] is shown in (a)(a). The chord diagram and the Lando graph associated to D[a1=1,b1=0]D[a_{1}=1,b_{1}=0] are shown in (b)(b) and (c)(c), respectively.

Figures 19(a)(a) and (b)(b) represent the chord diagrams associated to D[a1=1,b1=1]D[a_{1}=1,b_{1}=1] and D[a1=1,b1=0]D[a_{1}=1,b_{1}=0], respectively. Notice that f1f_{1} is free and e1e_{1} is 22-free in D[a1=1,b1=1](1)D[a_{1}=1,b_{1}=1](\vec{1}), hence MD[a1=1,b1=1]M_{D[a_{1}=1,b_{1}=1]} is contractible by Lemma 7.10 and MD[a1=1]XD[a1=1,b1=0]M_{D[a_{1}=1]}\simeq X_{D[a_{1}=1,b_{1}=0]}. The Lando graph HH associated to D[a1=1,b1=0]D[a_{1}=1,b_{1}=0] is depicted in Figure 19(c)(c), and since vertex f1f_{1} dominates f2f_{2} and vertex e2e_{2} dominates e3e_{3}, we have the following equivalence relations:

|IH||IHe2f1|=|IL4n5|{if n2mod3,S4n53otherwise,|I_{H}|\simeq|I_{H-e_{2}-f_{1}}|=|I_{L_{4n-5}}|\simeq\begin{cases}*&\text{if }n\equiv 2\mod 3,\\ S^{\lfloor\frac{4n-5}{3}\rfloor}&\text{otherwise},\end{cases}

and by Corollary 7.3 we get:

|MD[a1=1]||XD[a1=1,b1=0]|{if n2mod3,S4n44n53otherwise.|M_{D[a_{1}=1]}|\simeq|X_{D[a_{1}=1,b_{1}=0]}|\simeq\begin{cases}*&\text{if }n\equiv 2\mod 3,\\ S^{4n-4-\lfloor\frac{4n-5}{3}\rfloor}&\text{otherwise}.\end{cases}

Finally, we compute |MD||M_{D}|:

  • -

    If n0n\equiv 0 (mod 33), then the skein sequence (8.1) becomes:

    S8n32S8n31|MD|,S^{\frac{8n}{3}-2}\longrightarrow S^{\frac{8n}{3}-1}\longrightarrow|M_{D}|,

    so |MD|S8n31S8n31.|M_{D}|\simeq S^{\frac{8n}{3}-1}\vee S^{\frac{8n}{3}-1}.

  • -

    If n1n\equiv 1 (mod 33), then the skein sequence (8.1) becomes:

    S8n+132|MD|,S^{\frac{8n+1}{3}-2}\longrightarrow\ast\longrightarrow|M_{D}|,

    so |MD|S8n+131|M_{D}|\simeq S^{\frac{8n+1}{3}-1}.

  • -

    If n2n\equiv 2 (mod 33), then the skein sequence (8.1) becomes:

    S8n+232|MD|,\ast\longrightarrow S^{\frac{8n+2}{3}-2}\longrightarrow|M_{D}|,

    so |MD|S8n+232|M_{D}|\simeq S^{\frac{8n+2}{3}-2}.

Theorem.

Let DD be a diagram so that the associated chord diagram D(1)D(\vec{1}) contains more than one monochord, all of them 2-free . Then |MD||M_{D}| is homotopy equivalent to a wedge of spheres.

Proof.

Since all monochords are 22-free, then MD[b]BZDbM_{D}\simeq\bigvee_{[b]\in B}Z^{b}_{D}, by Corollary 7.11, so we consider each pair of discs in D(1)D(\vec{1}) independently. Moreover, we can remove nested monochords and equivalent bichords when computing MDM_{D} at the expense of taking suspensions (Lemmas 7.13 and 7.15). In addition, we can assume that D(1)D(\vec{1}) contains no bb-free monochords for any bichord bb (otherwise, ZDbZ_{D}^{b} is contractible by Lemma 7.10(2)). Hence, we just need to prove the statement for simple diagrams. Lemma 8.3 and Proposition 8.6 complete the proof. ∎

References

  • [CMS20] F. Cantero Morán and M. Silvero, Extreme Khovanov spectra, Rev. mat. iberoam. 36 (2020), no. 3, 661–670.
  • [DL20] Oliver T. Dasbach and Adam M. Lowrance, Extremal Khovanov homology of Turaev genus one links, Fund. Math. 250 (2020), no. 1, 63–99.
  • [GMS18] J. González-Meneses, P. M. G. Manchón, and M. Silvero, A geometric description of the extreme Khovanov cohomology, Proceedings of the Royal Society of Edinburgh: Section A Mathematics 148 (2018), no. 3, 541–557.
  • [Kho00] M. Khovanov, A categorification of the Jones polynomial, Duke Math. J. 101 (2000), no. 3, 359–426.
  • [LLS17] T. Lawson, R. Lipshitz, and S. Sarkar, The cube and the Burnside category, Categorification in geometry, topology, and physics, Contemp. Math., vol. 684, Amer. Math. Soc., Providence, RI, 2017, pp. 63–85.
  • [LLS20] by same author, Khovanov homotopy type, Burnside category, and products, Geom. Topol. 24 (2020), 623–745.
  • [LS14] R. Lipshitz and S. Sarkar, A Khovanov stable homotopy type, J. Amer. Math. Soc. 27 (2014), no. 4, 983–1042.
  • [LS18] Robert Lipshitz and Sucharit Sarkar, Spatial refinements and Khovanov homology, Proc. Int. Cong. of Math. 1 (2018), 1151–1172.
  • [PS14] Józef H. Przytycki and Radmila Sazdanović, Torsion in Khovanov homology of semi-adequate links, Fund. Math. 225 (2014), no. 1, 277–304.
  • [PS18] J. H. Przytycki and M. Silvero, Homotopy type of circle graph complexes motivated by extreme Khovanov homology, J. Algebr. Comb. 48 (2018), 119–156.
  • [PS20] by same author, Geometric realization of the almost-extreme Khovanov homology of semiadequate links, Geom. Dedicata 204 (2020), no. 1, 387–401.
  • [SS20] Radmila Sazdanovic and Daniel Scofield, Extremal khovanov homology and the girth of a knot, arXiv:2003.05074v1, 2020.
  • [Vir04] O. Viro, Khovanov homology, its definitions and ramifications, Fundam. Math. 184 (2004), 317–342.