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

Khovanov homology and exotic 44-manifolds

Qiuyu Ren Department of Mathematics, University of California, Berkeley, Berkeley, CA 94720, USA [email protected]  and  Michael Willis Department of Mathematics, Texas A&M University, College Station, TX 77843, USA [email protected]
Abstract.

We show that the Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} skein lasagna module distinguishes the exotic pair of knot traces X1(52)X_{-1}(-5_{2}) and X1(P(3,3,8))X_{-1}(P(3,-3,-8)), an example first discovered by Akbulut. This gives the first analysis-free proof of the existence of exotic compact orientable 44-manifolds. We also present a family of exotic knot traces that seem not directly recoverable from gauge/Floer-theoretic methods. Along the way, we present new explicit calculations of the Khovanov skein lasagna modules, and we define lasagna generalizations of the Lee homology and Rasmussen ss-invariant, which are of independent interest. Other consequences of our work include a slice obstruction of knots in 44-manifolds with nonvanishing skein lasagna module, a sharp shake genus bound for some knots from the lasagna ss-invariant, and a construction of induced maps on Khovanov homology for cobordisms in k2k\mathbb{CP}^{2}.

1. Introduction

The revolutionary work of Donaldson and Freedman in the 19801980s revealed a drastic dichotomy between the topological and smooth category of manifolds in dimension 44. Since then, constant efforts are made to find exotic pairs of 44-manifolds.

A pair of smooth 44-manifolds X1X_{1} and X2X_{2} is said to be exotic if they are homeomorphic but not diffeomorphic. In preferable cases, showing two 44-manifolds are homeomorphic boils down to checking some algebraic topological invariants agree, thanks to Freedman’s fundamental theorems [freedman1982topology]. Showing that they are non-diffeomorphic usually involves showing some gauge-theoretic or Floer-theoretic smooth invariants (Donaldson invariants [donaldson1990polynomial], Seiberg-Witten invariants [seiberg1994monopoles], Heegaard Floer type invariants [ozsvath2004holomorphic], and various variations/generalizations) differ. To date, all detection of compact orientable exotic 44-manifolds depends heavily on analysis, mostly elliptic PDE. We present the first analysis-free detection result.111For a noncompact example, an exotic 4\mathbb{R}^{4} can be constructed from a topologically slice but not smoothly slice knot. Such knots can be detected by Freedman’s result [freedman1982topology] and Rasmussen’s ss-invariant [rasmussen2010khovanov] from Khovanov homology. For nonorientable examples, Kreck [kreck2006some] showed K3#4K3\#\mathbb{RP}^{4} and 11(S2×S2)#411(S^{2}\times S^{2})\#\mathbb{RP}^{4} form an exotic pair using Rochlin’s Theorem, and indeed exotic 4\mathbb{RP}^{4}’s were constructed by Cappell-Shaneson [cappell1976some] (shown to be homeomorphic to 4\mathbb{RP}^{4} later in [hambleton1994nonorientable]).222Exotic surfaces in 44-manifolds were also known to be detectable by Khovanov homology [hayden2021khovanov].

The relevant smooth invariant in our story is the Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} skein lasagna module (Khovanov skein lasagna module for short) defined by Morrison-Walker-Wedrich [morrison2022invariants], which is an extension of a suitable variation of the classical Khovanov homology [khovanov2000categorification] via a skein-theoretic construction. It assigns a tri-graded (by homological degree hh, quantum degree qq, and homology class degree α\alpha below) abelian group

𝒮0,h,q2(X;L)=αH2(X,L)𝒮0,h,q2(X;L;α)\mathcal{S}_{0,h,q}^{2}(X;L)=\bigoplus_{\alpha\in H_{2}(X,L)}\mathcal{S}_{0,h,q}^{2}(X;L;\alpha)

to every pair (X,L)(X,L) of compact oriented smooth 44-manifold and framed oriented link on its boundary. When LL is the empty link we simply write 𝒮02(X)\mathcal{S}_{0}^{2}(X) for the skein lasagna module.

The nn-trace on a knot KK, denoted Xn(K)X_{n}(K), is the 44-manifold obtained by attaching an nn-framed 22-handle to B4B^{4} along KS3=B4K\subset S^{3}=\partial B^{4}. Our main theorem shows that the (1)(-1)-traces on the mirror image of the knot 525_{2} and the Pretzel knot P(3,3,8)P(3,-3,-8), namely

X1:=X1(52),X2:=X1(P(3,3,8)),X_{1}:=X_{-1}(-5_{2}),\quad X_{2}:=X_{-1}(P(3,-3,-8)),

have non-isomorphic skein lasagna module.

Refer to caption     Refer to caption

Figure 1. Kirby diagrams of X1,X2X_{1},X_{2}.
Theorem 1.1.

The graded modules 𝒮02(X1)\mathcal{S}_{0}^{2}(X_{1}) and 𝒮02(X2)\mathcal{S}_{0}^{2}(X_{2}) are non-isomorphic over \mathbb{Q}. Therefore, X1X_{1} and X2X_{2} form an exotic pair of 44-manifolds. In fact, even their interiors are exotic.

The exotic pair X1,X2X_{1},X_{2} was first discovered by Akbulut [akbulut1991exotic, akbulut1991fake]. The fact that X1,X2X_{1},X_{2} are homeomorphic is more or less standard: a sequence of Kirby moves shows X1\partial X_{1} and X2\partial X_{2} are homeomorphic and a suitable extension of Freedman’s Theorem [freedman1982topology, boyer1986simply] then implies X1X_{1} and X2X_{2} are homeomorphic. On the other hand, the fact that X1,X2X_{1},X_{2} are not diffeomorphic was detected by an intricate calculation of Donaldson invariants. Later, the result was reproved by Akbulut-Matveyev [akbulut1997exotic, Theorem 3.2] using a simplified argument which essentially says X1(52)X_{-1}(-5_{2}) is Stein (since the maximal Thurston-Bennequin number of 52-5_{2} is 11) while X1(P(3,3,8))X_{-1}(P(3,-3,-8)) is not (since P(3,3,8)P(3,-3,-8) is slice). This latter proof is elegant, yet still depends heavily on calculation of Seiberg-Witten invariants of Stein surfaces, as well as deep results of Eliashberg [eliashberg1990topological] and Lisca-Matić [lisca1997tight]. In comparison, our proof of Theorem 1.1 depends on the much lighter package of Khovanov homology, which is combinatorial in nature.

Since the Khovanov skein lasagna module over a field of a connected sum or boundary connected sum of 44-manifolds is equal to the tensor product of the Khovanov skein lasagna modules of the components, a fact proved in [manolescu2022skein], X1X_{1} and X2X_{2} remain exotic after (boundary) connected sums with any 44-manifolds WW with nonvanishing skein lasagna module (at least when the support of 𝒮02(W)\mathcal{S}_{0}^{2}(W) is small in a suitable sense). Such tricks lead to many new exotic pairs. We write down some explicit examples in the following corollary to give the reader a flavor of what can be proven. For more choices of such candidates of connected summands one may look into Section 6.

Corollary 1.2.

The 44-manifolds X1aX2bX_{1}^{\natural a}\natural X_{2}^{\natural b} and X1aX2bX_{1}^{\natural a^{\prime}}\natural X_{2}^{\natural b^{\prime}} form an exotic pair if a+b=a+ba+b=a^{\prime}+b^{\prime} and (a,b)(a,b)(a,b)\neq(a^{\prime},b^{\prime}). The exotica remain under connected sums or boundary sums with any copies of S1×S3S^{1}\times S^{3}, S2×D2S^{2}\times D^{2}, 2¯\overline{\mathbb{CP}^{2}}, and the negative E8E8 manifold. The exotica also remains upon further taking the interiors of the pair.

Anubhav Mukherjee informed us that most of these variations seem also achievable via gauge/Floer-theoretic methods and topological arguments. However, below we present a family of exotic knot traces generalizing Theorem 1.1 that seems not directly recoverable from other methods. Let Pn,mP_{n,m} and Qn,mQ_{n,m} be satellite patterns drawn as in Figure 2, and Pn,m(K)P_{n,m}(K), Qn,m(K)Q_{n,m}(K) be the corresponding satellite knots with companion knot KK. These satellite patterns were first studied by Yasui [yasui2015corks]. Let s(K)s(K) denote the Rasmussen ss-invariant of KK.

Theorem 1.3.

If a knot KS3K\subset S^{3} has a slice disk Σ\Sigma in k2\int(B4)k\mathbb{CP}^{2}\backslash int(B^{4}) and n,mn,m are two integers satisfying

s(K)=|[Σ]|[Σ]2,n<[Σ]2,m0,s(K)=|[\Sigma]|-[\Sigma]^{2},\ n<-[\Sigma]^{2},\ m\geq 0,

then Xn(Pn,m(K))X_{n}(P_{n,m}(K)) and Xn(Qn,m(K))X_{n}(Q_{n,m}(K)) form an exotic pair. Their interiors are also exotic.

In the special case when K=UK=U is the unknot, k=0k=0, Σ\Sigma is the trivial slice disk of KK in B4B^{4}, and n=1n=-1, m=0m=0, Theorem 1.3 recovers Theorem 1.1. One can also construct more exotic pairs in the spirit of Corollary 1.2 by taking connected sums of these examples and other standard manifolds appearing in that corollary.

Now we turn to some explicit calculations of 𝒮02\mathcal{S}_{0}^{2} and other applications. Some topological applications we derived can indeed be proved in other ways, but our approach provides the first analysis-free proofs. We also give a proof of Theorem 1.1 in Section 1.5, assuming some results stated.

Throughout this paper, all 44-manifolds are assumed to be smooth, compact, and oriented. All surfaces are oriented and smoothly and properly embedded. In the current section, many statements are presented in a specialized form. In such cases, we indicate the corresponding generalizations in the parentheses after the statement numbers.

1.1. A vanishing result and slice obstruction

For a knot KS3K\subset S^{3}, let TB(K)TB(K) denote its maximal Thurston-Bennequin number. We have the following vanishing criterion for the Khovanov skein lasagna module.

Theorem 1.4.

If the knot trace Xn(K)X_{n}(K) for some knot KS3K\subset S^{3} and framing nTB(K)n\geq-TB(-K) embeds into a 44-manifold XX, then 𝒮02(X;L)=0\mathcal{S}_{0}^{2}(X;L)=0 for any LXL\subset\partial X.

For example, since S2×S2S^{2}\times S^{2} contains an embedded sphere with self-intersection 22, we see the 22-trace of the unknot embeds into S2×S2S^{2}\times S^{2}, implying 𝒮02(S2×S2)=0\mathcal{S}_{0}^{2}(S^{2}\times S^{2})=0. This answers a question of Manolescu (see [sullivan2024kirby, Question 2]), generalizing one main Theorem of Sullivan-Zhang [sullivan2024kirby, Theorem 1.3] and answering [sullivan2024kirby, Question 3] in the affirmative.

The folklore trace embedding lemma says that Xn(K)X_{n}(K) embeds into XX if and only if K-K is (n)(-n)-slice in XX, i.e. there is a framed slice disk in X\int(B4)X\backslash int(B^{4}) of the (n)(-n)-framed knot KS3=B4-K\subset S^{3}=\partial B^{4}. Therefore, the contrapositive of Theorem 1.4 can be phrased as follows.

Corollary 1.5.

Suppose 𝒮02(X;L)0\mathcal{S}_{0}^{2}(X;L)\neq 0 for some LXL\subset\partial X. If a knot KS3K\subset S^{3} is nn-slice in XX, then n>TB(K)n>TB(K).∎

1.2. Lee skein lasagna modules and lasagna ss-invariants

In the construction of the Khovanov skein lasagna module, the relevant TQFT (the Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} homology) enters only at the very last step. We could replace this TQFT by its Lee deformation [lee2005endomorphism], and define an analogous notion of a Lee skein lasagna module of a 44-manifold XX with link LXL\subset\partial X, denoted 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L). It is a \mathbb{Q}-vector space equipped with a homological \mathbb{Z}-grading, a quantum /4\mathbb{Z}/4-grading, a homology class grading by H2(X,L)H_{2}(X,L), and a quantum \mathbb{Z}-filtration (where 0 has filtration degree -\infty), with the caveat that a nonzero vector can have quantum filtration degree -\infty.

We show in Section 4 that the structure of 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L), except for the quantum filtration, is determined by simple algebraic data of (X,L)(X,L). Moreover, we extract numerical invariants from the quantum filtration structure333The authors are aware that the recent paper [morrison2024invariants] by Morrison-Walker-Wedrich has an independent proof of a part of this structural result (except addenda (3) and (4)), stated in a more general form. They extract slightly different invariants from the filtration structure, which gives somewhat less information than ours in the 𝔤𝔩2\mathfrak{gl}_{2} case. Moreover, our invariants generalize Rasmussen ss-invariants for links and enjoy nice functorial properties.. For the purpose of exposition, in the current section we assume LL is the empty link. Let q:𝒮0Lee(X){}q\colon\mathcal{S}_{0}^{Lee}(X)\to\mathbb{Z}\cup\{-\infty\} denote the quantum filtration function.

Theorem 1.6.

(Theorem 4.1) The Lee skein lasagna module of a 44-manifold XX is 𝒮0Lee(X)H2(X)2,\mathcal{S}_{0}^{Lee}(X)\cong\mathbb{Q}^{H_{2}(X)^{2}}, with a basis consisting of canonical generators xα+,αx_{\alpha_{+},\alpha_{-}}, one for each pair α+,αH2(X)\alpha_{+},\alpha_{-}\in H_{2}(X). Moreover,

  1. (1)

    xα+,αx_{\alpha_{+},\alpha_{-}} has homological degree 2α+α-2\alpha_{+}\cdot\alpha_{-};

  2. (2)

    xα+,αx_{\alpha_{+},\alpha_{-}} has homology class degree α:=α++α\alpha:=\alpha_{+}+\alpha_{-};

  3. (3)

    xα+,α+xα,α+x_{\alpha_{+},\alpha_{-}}+x_{\alpha_{-},\alpha_{+}}, xα+,αxα,α+x_{\alpha_{+},\alpha_{-}}-x_{\alpha_{-},\alpha_{+}} have quantum /4\mathbb{Z}/4 degrees α2-\alpha^{2}, α2+2-\alpha^{2}+2, respectively;

  4. (4)

    q(xα+,α)=q(xα,α+)=max(q(xα+,α±xα,α+))q(x_{\alpha_{+},\alpha_{-}})=q(x_{\alpha_{-},\alpha_{+}})=\max(q(x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}})), which also equals min(q(xα+,α±xα,α+))+2\min(q(x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}}))+2 if α+α\alpha_{+}\neq\alpha_{-} in H2(X;)H_{2}(X;\mathbb{Q}).

Definition 1.7.

(Definition 4.6) The lasagna ss-invariant of XX at a homology class αH2(X)\alpha\in H_{2}(X) is

s(X;α):=q(xα,0).s(X;\alpha):=q(x_{\alpha,0}).

The lasagna ss-invariants of XX thus define a smooth invariant s(X;):H2(X){}s(X;\bullet):H_{2}(X)\to\mathbb{Z}\cup\{-\infty\}.

The genus function of a 44-manifold XX, defined by

g(X,α):=min{g(Σ):ΣX,[Σ]=α},g(X,\alpha):=\min\{g(\Sigma)\colon\Sigma\subset X,[\Sigma]=\alpha\},

captures subtle information about the smooth topology of XX. Like the classical ss-invariant [rasmussen2010khovanov, beliakova2008categorification], which provides a lower bound on the slice genus of links, the lasagna ss-invariants provide a lower bound on the genus function of 44-manifolds.

Theorem 1.8.

(Theorem 4.7(6) and Corollaries after) The genus function of a 44-manifold XX has a lower bound given by

g(X;α)s(X;α)+α22.g(X;\alpha)\geq\frac{s(X;\alpha)+\alpha^{2}}{2}.

Unlike the classical ss-invariant, the lasagna ss-invariants are not directly computable from link diagrams. However, we are able to compute them in some special cases. We indicate one special case where Theorem 1.8 does provide useful information.

The nn-shake genus of a knot KK is defined by gshn(K):=g(Xn(K);1)g_{sh}^{n}(K):=g(X_{n}(K);1), where 11 is a generator of H2(Xn(K))H_{2}(X_{n}(K))\cong\mathbb{Z}. Clearly, gshn(K)g4(K)g_{sh}^{n}(K)\leq g_{4}(K), but the inequality may not be strict [akbulut19772, piccirillo2019shake]. Theorem 1.8 specializes to a shake genus bound

gshn(K)s(Xn(K);1)+n2.g_{sh}^{n}(K)\geq\frac{s(X_{n}(K);1)+n}{2}. (1)
Theorem 1.9.
  1. (1)

    If KK is concordant to a positive knot, then for n0n\leq 0 we have s(Xn(K);1)=s(K)n,s(X_{n}(K);1)=s(K)-n, where s(K)s(K) is the Rasmussen ss-invariant of KK. Thus

    gshn(K)=g4(K)=s(K)2,n0.g_{sh}^{n}(K)=g_{4}(K)=\frac{s(K)}{2},\ n\leq 0.
  2. (2)

    If Σk2¯\(int(B4)int(B4))\Sigma\subset k\overline{\mathbb{CP}^{2}}\backslash(int(B^{4})\sqcup int(B^{4})) is a concordance from KK to a positive knot KK^{\prime}, then for n[Σ]2n\leq[\Sigma]^{2} we have s(Xn(K);1)s(K)+[Σ]2+|[Σ]|ns(X_{n}(K);1)\geq s(K^{\prime})+[\Sigma]^{2}+|[\Sigma]|-n. Thus

    gshn(K)s(K)+[Σ]2+|[Σ]|2,n[Σ]2.g_{sh}^{n}(K)\geq\frac{s(K^{\prime})+[\Sigma]^{2}+|[\Sigma]|}{2},\ n\leq[\Sigma]^{2}.
Remark 1.10.

The computation of shake genus in Theorem 1.9(1) also follows from the following two facts:

  1. (i)

    The adjunction inequality for Stein manifolds [lisca1997tight], which implies

    gshn(K)(tb(𝒦)+|rot(𝒦)|+1)/2g_{sh}^{n}(K)\geq(tb(\mathcal{K})+|rot(\mathcal{K})|+1)/2 (2)

    for any Legendrian representative 𝒦\mathcal{K} of KK with tb(𝒦)=n+1tb(\mathcal{K})=n+1. When n<TB(K)n<TB(K), take 𝒦\mathcal{K} to be a stabilization that maximize |rot||rot|, of a max tb representative of KK. Then the right hand side agrees with s(K)/2s(K)/2 for positive knots [tanaka1999maximal, Theorem 2].

  2. (ii)

    Two concordant knots have the same shake genus (by an easy topological argument).

Lasagna ss-invariants also obstruct embeddings between 44-manifolds in the following sense.

Proposition 1.11.

If i:XXi\colon X\hookrightarrow X^{\prime} is an inclusion between two 44-manifolds, then for any class αH2(X)\alpha\in H_{2}(X),

s(X;iα)s(X;α).s(X^{\prime};i_{*}\alpha)\leq s(X;\alpha).

1.3. Nonvanishing result for 22-handlebodies from lasagna ss-invariants

The Khovanov homology of a link admits a spectral sequence to its Lee homology. In particular, in every bidegree, the rank of the Khovanov homology is bounded below by the dimension of the associated graded vector space of the Lee homology. Although we are not able to define a spectral sequence from 𝒮02\mathcal{S}_{0}^{2} to 𝒮0Lee\mathcal{S}_{0}^{Lee}, we do have a rank inequality when XX is a 22-handlebody. Let grVgr_{\bullet}V denote the associated graded vector space of a filtered vector space VV.

Theorem 1.12.

If XX is a 22-handlebody, then in any tri-degree (h,q,α)2×H2(X,L)(h,q,\alpha)\in\mathbb{Z}^{2}\times H_{2}(X,L),

rank(𝒮0,h,q2(X;L;α))dim(grq(𝒮0,hLee(X;L;α))).\mathrm{rank}(\mathcal{S}_{0,h,q}^{2}(X;L;\alpha))\geq\dim(gr_{q}(\mathcal{S}_{0,h}^{Lee}(X;L;\alpha))).

In particular, if the quantum filtration on 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) is not identically -\infty, then 𝒮02(X;L)0\mathcal{S}_{0}^{2}(X;L)\neq 0.

We state a comparison result for Lee skein lasagna modules of 22-handlebodies, which can sometimes be used as a nonvanishing criterion for Khovanov skein lasagna modules in view of Theorem 1.12.

Proposition 1.13.

(Proposition 5.3) Suppose XX (resp. XX^{\prime}) is a 22-handlebody obtained by attaching 22-handles to B4B^{4} along a framed oriented link KS3K\subset S^{3} (resp. KS3K^{\prime}\subset S^{3}). If KK is framed concordant to KK^{\prime}, then 𝒮0Lee(X)𝒮0Lee(X)\mathcal{S}_{0}^{Lee}(X)\cong\mathcal{S}_{0}^{Lee}(X^{\prime}) as tri-graded, quantum filtered vector spaces. Under the identification H2(X)H2(X)H_{2}(X)\cong H_{2}(X^{\prime}) induced by a link concordance, XX and XX^{\prime} have equal lasagna ss-invariants.

1.4. Nonvanishing result for 22-handlebodies from link diagram

Using a diagrammatic argument, in Section 5.4, we prove a nonvanishing result for 22-handlebodies with sufficiently negative framings. For the ease of exposition, here we only state the case for knot traces. Let n(K)n_{-}(K) be the minimal number of negative crossings among all diagrams of a knot KK.

Theorem 1.14.

(Theorem 5.5) Let KK be a knot and X=Xn(K)X=X_{n}(K) be its nn-trace.

  1. (1)

    If n<2n(K)n<-2n_{-}(K), then s(X;0)=0s(X;0)=0 and 𝒮0,0,q2(X;0)=grq(𝒮0,0Lee(X;0))={,q=00,q0.\mathcal{S}_{0,0,q}^{2}(X;0)\otimes\mathbb{Q}=gr_{q}(\mathcal{S}_{0,0}^{Lee}(X;0))=\begin{cases}\mathbb{Q},&q=0\\ 0,&q\neq 0.\end{cases}

  2. (2)

    If KK is positive, then for n0n\leq 0, s(X;0)=0s(X;0)=0, s(X;1)=s(K)ns(X;1)=s(K)-n. For n<0n<0,

    𝒮0,0,q2(X;1)=grq(𝒮0,0Lee(X;1))={,q=s(X;1),s(X;1)20,otherwise.\mathcal{S}_{0,0,q}^{2}(X;1)\otimes\mathbb{Q}=gr_{q}(\mathcal{S}_{0,0}^{Lee}(X;1))=\begin{cases}\mathbb{Q},&q=s(X;1),s(X;1)-2\\ 0,&\text{otherwise}.\end{cases}

1.5. Proof of Theorem 1.1

Since 52-5_{2} is a positive knot with s(52)=2s(-5_{2})=2, by Theorem 1.9(1) we have gsh1(52)=1g_{sh}^{-1}(-5_{2})=1. On the other hand, P(3,3,8)P(3,-3,-8) is slice, thus gsh1(P(3,3,8))=0g_{sh}^{-1}(P(3,-3,8))=0. This shows intX1int\,X_{1} and intX2int\,X_{2} are non-diffeomorphic, without actually calculating 𝒮0(Xi)\mathcal{S}_{0}^{\bullet}(X_{i}).

We still need to argue the Khovanov (and indeed Lee) skein lasagna modules of X1,X2X_{1},X_{2} actually differ. Apply Theorem 1.14(2) to 52-5_{2}, we obtain s(X1;1)=3s(X_{1};1)=3,

𝒮0,0,q2(X1;1)=grq(𝒮0,0Lee(X1;1))={,q=1,30,otherwise.\mathcal{S}_{0,0,q}^{2}(X_{1};1)\otimes\mathbb{Q}=gr_{q}(\mathcal{S}_{0,0}^{Lee}(X_{1};1))=\begin{cases}\mathbb{Q},&q=1,3\\ 0,&\text{otherwise}.\end{cases}

Apply Theorem 1.14(2) to the unknot, and then Proposition 1.13 to P(3,3,8)P(3,-3,-8) and the unknot, we obtain s(X2;1)=1s(X_{2};1)=1, and

𝒮0,0,q2(X2;1)grq(𝒮0,0Lee(X2;1))={,q=±10,otherwise.\mathcal{S}_{0,0,q}^{2}(X_{2};1)\otimes\mathbb{Q}\supset gr_{q}(\mathcal{S}_{0,0}^{Lee}(X_{2};1))=\begin{cases}\mathbb{Q},&q=\pm 1\\ 0,&\text{otherwise}.\end{cases}

where the first containment follows from Theorem 1.12. In particular, 𝒮0,0,12(X1;1)𝒮0,0,12(X2;1)\mathcal{S}_{0,0,-1}^{2}(X_{1};1)\otimes\mathbb{Q}\neq\mathcal{S}_{0,0,-1}^{2}(X_{2};1)\otimes\mathbb{Q}.∎

1.6. Example: Khovanov and Lee skein lasagna module of 2¯\overline{\mathbb{CP}^{2}}

As an example, we note that the recent work by the first author [ren2023lee] has the following implications on the Lee and Khovanov skein lasagna modules for 2¯\overline{\mathbb{CP}^{2}}.

Proposition 1.15.

The lasagna ss-invariants of 2¯\overline{\mathbb{CP}^{2}} are given by s(2¯;α)=|α|s(\overline{\mathbb{CP}^{2}};\alpha)=|\alpha|. The associated graded vector space of the Lee skein lasagna module of 2¯\overline{\mathbb{CP}^{2}} is given by grq𝒮0,(β2α2)/2Lee(2¯;α)=gr_{q}\mathcal{S}_{0,(\beta^{2}-\alpha^{2})/2}^{Lee}(\overline{\mathbb{CP}^{2}};\alpha)=\mathbb{Q} for

βα(mod2),q={α2β22+|β|1±1,β0α22,β=0,\beta\equiv\alpha\pmod{2},\ q=\begin{cases}\frac{\alpha^{2}-\beta^{2}}{2}+|\beta|-1\pm 1,&\beta\neq 0\\ \frac{\alpha^{2}}{2},&\beta=0,\end{cases}

and 0 in other tri-degrees. Moreover, for βα(mod2)\beta\equiv\alpha\pmod{2},

𝒮0,(β2α2)/2,(α2β2)/2+|β|2(2¯;α)=.\mathcal{S}_{0,(\beta^{2}-\alpha^{2})/2,(\alpha^{2}-\beta^{2})/2+|\beta|}^{2}(\overline{\mathbb{CP}^{2}};\alpha)=\mathbb{Z}.

A consequence of this result, together with some properties of lasagna ss-invariants, is the following.

Corollary 1.16.

If a 44-manifold XX admits an embedding i:Xk2¯i\colon X\hookrightarrow k\overline{\mathbb{CP}^{2}} for some kk, then s(X;α)|iα|s(X;\alpha)\geq|i_{*}\alpha|, with equality achieved if XX is a 22-handlebody and α=0\alpha=0. In particular, 𝒮0,0,q2(X;α)0\mathcal{S}_{0,0,q}^{2}(X;\alpha)\neq 0 for q=s(X;α)q=s(X;\alpha), and for q=s(X;α)2q=s(X;\alpha)-2 if α0H2(X;)\alpha\neq 0\in H_{2}(X;\mathbb{Q}).

It is also worth pointing out that Conjecture 6.1 in [ren2023lee] is equivalent to a complete determination of the Khovanov skein lasagna module of 2¯\overline{\mathbb{CP}^{2}} (see Conjecture LABEL:conj:-CP2).

If Σ2\(int(B4)int(B4))\Sigma\subset\mathbb{CP}^{2}\backslash(int(B^{4})\sqcup int(B^{4})) is an oriented cobordism in 2\mathbb{CP}^{2} between two oriented links L0,L1L_{0},L_{1}, the last part of Proposition 1.15 enables us to define an induced map

Kh(Σ):Kh(L0)Kh(L1)Kh(\Sigma)\colon Kh(L_{0})\to Kh(L_{1})

between ordinary Khovanov homology of L0L_{0} and L1L_{1}, at least over a field and up to sign, with bidegree (0,χ(Σ)[Σ]2+|[Σ]|)(0,\chi(\Sigma)-[\Sigma]^{2}+|[\Sigma]|). We carry out this construction in more details in Section LABEL:sbsec:Kh_CP2_cob.

1.7. A conjecture for Stein manifolds

Many similarities between skein lasagna modules and other gauge/Floer-theoretic smooth invariants appear. However, a serious constraint for our computation is that we were not able to produce any 44-manifold with b2+>0b_{2}^{+}>0 that has finite lasagna ss-invariants or nonvanishing Khovanov skein lasagna module. In fact, if one were able to find a simply-connected closed 44-manifold XX with b2+(X)>0b_{2}^{+}(X)>0 that has 𝒮02(X)0\mathcal{S}_{0}^{2}(X)\neq 0, then X#2¯X\#\overline{\mathbb{CP}^{2}} is an exotic closed 44-manifold, as it has nonvanishing 𝒮02\mathcal{S}_{0}^{2} while its standard counterpart (some k2#2¯k\mathbb{CP}^{2}\#\ell\overline{\mathbb{CP}^{2}}) has vanishing 𝒮02\mathcal{S}_{0}^{2}.

To pursue this thread, we propose the following conjecture, which seems plausible and desirable.

Conjecture 1.17.

If KK is a knot and X=Xn(K)X=X_{n}(K) for some n<TB(K)n<TB(K), then s(X;0)=0s(X;0)=0 and s(X;1)=s(K)ns(X;1)=s(K)-n. In particular, 𝒮02(X)0\mathcal{S}_{0}^{2}(X)\neq 0. More generally, any Stein 22-handlebody XX has s(X;0)>s(X;0)>-\infty and nonvanishing Khovanov skein lasagna module.

The part concerning s(X;1)s(X;1) in Conjecture 1.17 implies the nn-shake genus of KK for n<TB(K)n<TB(K) (or more generally n<TB(K)n<TB(K^{\prime}) for some KK^{\prime} concordant to KK) is bounded below by s(K)/2s(K)/2, generalizing Theorem 1.9(1). This seems to be the proper straightening of the bound from the Stein adjunction inequality (2).

As will be clear later, Conjecture 1.17 is equivalent to an assertion on the ss-invariants of cables of KK (or some Legendrian link in the 22-handlebody case) (cf. Proposition 5.4).

1.8. Organization of the paper

In Section 2, we review the theory of Khovanov skein lasagna modules, expanding on a few properties, as well as making some new comments on the skein-theoretic construction. In the short Section 3, we prove Theorem 1.4 by applying a bound on Khovanov homology by Ng [ng2005legendrian], and we give some examples of 44-manifolds with vanishing Khovanov skein lasagna module.

In Section 4, we develop the theory of Lee skein lasagna modules and lasagna ss-invariants, which can be thought of as a lasagna generalization of the classical theory by Lee [lee2005endomorphism] and Rasmussen [rasmussen2010khovanov] for links in S3S^{3}. A generalization of Theorem 1.6, in particular the definition of canonical Lee lasagna generators, is presented in Section 4.1. The behavior of canonical Lee lasagna generators under morphisms is given in Theorem 4.5 in Section 4.2. Building on these results, in Section 4.3 we define and examine the lasagna ss-invariants, generalizing Definition 1.7, Theorem 1.8, and Proposition 1.11.

In Section 5, we make use of a 22-handlebody formula by Manolescu-Neithalath [manolescu2022skein] to prove nonvanishing results for Khovanov and Lee skein lasagna modules of 22-handlebodies. In Section 5.1, we prove Theorem 1.12 which shows that finiteness results for the quantum filtration of Lee skein lasagna modules imply nonvanishing results for Khovanov skein lasagna modules. In Section 5.2 we prove a generalization of Proposition 1.13 and in Section 5.3 we show that lasagna ss-invariants can be expressed in terms of classical ss-invariants of some cable links. Finally, in the somewhat technical Section 5.4, we prove a vast generalization of the diagrammatic nonvanishing criterion Theorem 1.14.

Finally, in Section 6, we give examples and applications, mostly to the theory and tools developed in Section 4 and Section 5. This includes proofs of Corollary 1.2, Theorem 1.3, Theorem 1.9, and an expanded discussion of Section 1.6. Other discussions include a proof that ss-invariants of cables of the Conway knot can obstruct its sliceness, and that the Khovanov skein lasagna module over \mathbb{Q} does not detect potential exotica arising from Gluck twists in S4S^{4}.

Most of our developments are independent to proving Theorem 1.1. The readers only interested in Theorem 1.1 may want to look at its proof in Section 1.5 and delve into the corresponding sections. In particular, to distinguish the smooth structures of the pair X1,X2X_{1},X_{2} in Theorem 1.1, one can read the first half of Section 2, certain statements in Section 4 (mainly Definition 4.6 and Corollary 4.13), Section 5.4, and Section 6.4. The minimal amount of theory needed to actually distinguish their Khovanov skein lasagna modules depends moreover on Section 5.1 and Section 5.2 (but not really on Corollary 4.13 and Section 6.4). Still, the statements in those sections are often much more general than needed for Theorem 1.1, which at times complicate the proofs.

Acknowledgement

We thank Ian Agol, Daren Chen, Kyle Hayden, Sungkyung Kang, Ciprian Manolescu, Marco Marengon, Anubhav Mukherjee, Ikshu Neithalath, Lisa Piccirillo, Mark Powell, Yikai Teng, Joshua Wang, Paul Wedrich, Hongjian Yang for helpful discussions. This work was supported in part by the Simons grant “new structures in low-dimensional topology” through the workshop on exotic 44-manifolds held at Stanford University December 16-18, 2023.

2. Khovanov and Lee skein lasagna modules

The Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} skein lasagna modules were defined by Morrison-Walker-Wedrich [morrison2022invariants]. In the first part of this section we review its construction, cast in a slightly different way, while defining its Lee deformation under the same framework. In Section 2.1, we begin with a general skein-theoretic construction of skein lasagna modules, using an arbitrary (stratified) TQFT for links in S3S^{3}. In Section 2.2, we state a few properties of skein lasagna modules, most of which are essentially proved in [manolescu2022skein]. Only then, in Section 2.3, do we fix the TQFT in consideration to be either the 𝔤𝔩2\mathfrak{gl}_{2} Khovanov-Rozansky homology KhR2KhR_{2} [morrison2022invariants] or its Lee deformation KhRLeeKhR_{Lee} [lee2005endomorphism]. We state a few more computational results for these settings in Section 2.4.

The second part of the section is new, and can be skipped on a first reading. In Section 2.5, we identify the structure of the underlying space of skeins in the construction of Section 2.1. The proof will eventually be useful in Section 4.1 where we calculate the Lee skein lasagna module of a pair (X,L)(X,L). Finally, in Section 2.6, we suggest a few variations of the skein-theoretic construction.

2.1. Construction of skein lasagna modules

444The formulation of our construction was motivated by a conversation with Sungkyung Kang.

Let XX be a 44-manifold and LXL\subset\partial X be a framed oriented link on its boundary. A skein in (X,L)(X,L) or in XX rel LL is a properly embedded framed oriented surface in XX with the interiors of finitely many disjoint 44-balls deleted, whose boundary on X\partial X is LL. That is, a skein in (X,L)(X,L) can be written as

ΣX\i=1kint(Bi), with Σ|X=L,\Sigma\subset X\backslash{\sqcup_{i=1}^{k}int(B_{i})},\text{ with }\partial\Sigma|_{\partial X}=L, (3)

The deleted balls BiB_{i} are called input balls of the skein Σ\Sigma.

We define the category of skeins in (X,L)(X,L), denoted 𝒞(X;L)\mathcal{C}(X;L), as follows. The objects consist of all skeins in XX rel LL. Suppose Σ\Sigma is an object as in (3) and Σ\Sigma^{\prime} is another object with input balls BjB_{j}^{\prime}, j=1,,kj=1,\cdots,k^{\prime}. If BiBj\sqcup B_{i}\subset\sqcup B_{j}^{\prime} and each BjB_{j}^{\prime} either coincides with some BiB_{i} or does not intersect any BiB_{i} on its boundary, then the set of morphisms from Σ\Sigma to Σ\Sigma^{\prime} consists of pairs of an isotopy class of surfaces555We formally allow some components of SS to be curves on the boundaries of those Bi=BjB_{i}=B_{j}^{\prime}. SBj\int(Bi)S\subset\sqcup B_{j}^{\prime}\backslash\sqcup\,int(B_{i}) rel boundary with S|Bi=Σ|BiS|_{\partial B_{i}}=\Sigma|_{\partial B_{i}}, S|Bj=Σ|BjS|_{\partial B_{j}^{\prime}}=\Sigma^{\prime}|_{\partial B_{j}^{\prime}}, and an isotopy class of isotopies from SΣS\cup\Sigma^{\prime}666In order for the surfaces to glue smoothly, one may assume throughout that all embedded surfaces have a standard collar near each boundary, and all isotopies rel boundary respect these collars. to Σ\Sigma rel boundary. We will suppress the isotopy and write [S]:ΣΣ[S]\colon\Sigma\to\Sigma^{\prime} for this morphism. If the input balls Bi,BjB_{i},B_{j}^{\prime} do not satisfy the condition above, there is no morphism from Σ\Sigma to Σ\Sigma^{\prime}. The identity morphism at Σ\Sigma is given by [Σ|Bi][\Sigma|_{\sqcup\partial B_{i}}]. The composition of two morphisms [S]:ΣΣ[S]\colon\Sigma\to\Sigma^{\prime} and [S]:ΣΣ[S^{\prime}]\colon\Sigma^{\prime}\to\Sigma^{\prime\prime} is [SS][S\cup S^{\prime}] (together with the composition of isotopies).

Next we introduce an arbitrary (stratified) TQFT of framed oriented links in S3S^{3}, by which we mean a symmetric monoidal functor ZZ from the category of framed oriented links in S3S^{3} and link cobordisms in S3×IS^{3}\times I up to isotopy rel boundary, to the category of modules over a commutative ring RR, such that Z()=RZ(\emptyset)=R and Z(L)Z(L) is finitely generated for all links LL. Equivalently, ZZ is a functorial link invariant in 3\mathbb{R}^{3} satisfying the conditions in Theorem 2.1 of [morrison2024invariants]. At times we also assume RR is a field and ZZ is involutive, meaning that Z(L)Z(L)Z(-L)\cong Z(L)^{*} for a link LL and Z(Σ)Z(Σ)Z(-\Sigma)\cong Z(\Sigma)^{*} for a cobordism Σ:L0L1\Sigma\colon L_{0}\to L_{1}, where Σ:L1L0-\Sigma\colon-L_{1}\to-L_{0} is the reverse cobordism.

Assign to every object Σ\Sigma of 𝒞(X;L)\mathcal{C}(X;L) as in the notations before the RR-module i=1kZ(Ki)\bigotimes_{i=1}^{k}Z(K_{i}) which we denote by an abuse of notation Z(Σ)Z(\Sigma), where each Ki=Σ|BiK_{i}=\partial\Sigma|_{\partial B_{i}} is a framed link. Assign to every morphism [S]:ΣΣ[S]\colon\Sigma\to\Sigma^{\prime} the morphism Z([S]):Z(Σ)Z(Σ)Z([S])\colon Z(\Sigma)\to Z(\Sigma^{\prime}) induced by the cobordism S:KiKjS\colon\sqcup K_{i}\to\sqcup K_{j}^{\prime}. Then ZZ defines a functor from the category of skeins in (X,L)(X,L) to the category of RR-modules.

The skein lasagna module of (X,L)(X,L) with respect to ZZ, denoted 𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X;L), can be now defined as

𝒮0Z(X;L):=colim(Z:𝒞(X;L)𝐌𝐨𝐝R).\mathcal{S}_{0}^{Z}(X;L):=\mathrm{colim}(Z\colon\mathcal{C}(X;L)\to\mathbf{Mod}_{R}). (4)

More explicitly,

𝒮0Z(X;L)={(Σ,v):ΣX\i=1kint(Bi),vZ(Σ)}/\mathcal{S}_{0}^{Z}(X;L)=\{(\Sigma,v)\colon\Sigma\subset X\backslash\sqcup_{i=1}^{k}int(B_{i}),\ v\in Z(\Sigma)\}/\sim (5)

where \sim is the equivalence relation generated by linearity in vv, and (Σ,v)(Σ,Z([S])(v))(\Sigma,v)\sim(\Sigma^{\prime},Z([S])(v)) for [S]:ΣΣ[S]\colon\Sigma\to\Sigma^{\prime}. A pair (Σ,v)(\Sigma,v) (or just vv if there’s no confusion) as in (5) is called a lasagna filling of (X,L)(X,L).

The skein lasagna module 𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X;L) is an invariant of the pair (X,L)(X,L) up to diffeomorphism. When L=L=\emptyset, we write for short 𝒮0Z(X)\mathcal{S}_{0}^{Z}(X) for 𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X;L). In this case, since all lasagna fillings and their relations appear in the interior of XX, 𝒮0Z(X)\mathcal{S}_{0}^{Z}(X) is an invariant of int(X)int(X).

2.2. Properties of skein lasagna modules

We state some properties of the skein lasagna modules, mostly following Manolescu-Neithalath [manolescu2022skein]. Although their paper dealt with the case when the TQFT ZZ is the Khovanov-Rozansky 𝔤𝔩N\mathfrak{gl}_{N} homology, the proofs remain valid in the more general setup. We remark that the more recent paper [manolescu2023skein] gives a formula for 𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X;L) in terms of a handlebody decomposition of XX. However, we will not be using this result, so we refer interested readers to their paper.

Every object Σ\Sigma of 𝒞(X;L)\mathcal{C}(X;L) represents a homology class in H2(X,L)H_{2}(X,L), whose image under the boundary homomorphism \partial in the long exact sequence

0H2(X)H2(X,L)H1(L)H1(X)0\to H_{2}(X)\to H_{2}(X,L)\xrightarrow{\partial}H_{1}(L)\to H_{1}(X)\to\cdots

is [L][L], the fundamental class of LL. If [S]:ΣΣ[S]\colon\Sigma\to\Sigma^{\prime} is a morphism, then [Σ]=[Σ]H2(X,L)[\Sigma]=[\Sigma^{\prime}]\in H_{2}(X,L). This means 𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X;L) admits a grading by H2L(X):=1([L])H_{2}^{L}(X):=\partial^{-1}([L]), which we write as

𝒮0Z(X;L)=αH2L(X)𝒮0Z(X;L;α).\mathcal{S}_{0}^{Z}(X;L)=\bigoplus_{\alpha\in H_{2}^{L}(X)}\mathcal{S}_{0}^{Z}(X;L;\alpha).

Of course, H2L(X)H_{2}^{L}(X) is nonempty if and only if the image of [L][L] in H1(X)H_{1}(X) is zero, i.e. LL is null-homologous in XX. From now on this will be assumed, since otherwise 𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X;L) is trivially zero as there is no lasagna filling.

Proposition 2.1 (3,43,4-handle attachments, [manolescu2022skein, Proposition 2.1]).

If (X,L)(X^{\prime},L) is obtained from (X,L)(X,L) by attaching a 33-handle (resp. 44-handle) away from LL, then there is a natural surjection (resp. isomorphism) 𝒮0Z(X;L)𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X;L)\to\mathcal{S}_{0}^{Z}(X^{\prime};L).∎

Proposition 2.2 (Connected sum formula, [manolescu2022skein, Theorem 1.4,Corollary 7.3]).

If the base ring RR is a field and ZZ is involutive, then there is a natural isomorphism

𝒮0Z(XX;LL)𝒮0Z(X;L)𝒮0Z(X;L).\mathcal{S}_{0}^{Z}(X\natural X^{\prime};L\sqcup L^{\prime})\cong\mathcal{S}_{0}^{Z}(X;L)\otimes\mathcal{S}_{0}^{Z}(X^{\prime};L^{\prime}).

The same formula holds with the boundary sum replaced by the connected sum or the disjoint union.∎

Suppose (X,L)(X,L) is a pair and YXY\subset X is a properly embedded separating oriented 33-manifold, possibly with transverse intersections with LL in some point set PYP\subset\partial Y. Then YY cuts (X,L)(X,L) into two pairs (X1,T1)(X_{1},T_{1}) and (X2,T2)(X_{2},T_{2}), where X1\partial X_{1} induces the orientation on YY and X2\partial X_{2} induces the reverse orientation, and T1,T2T_{1},T_{2} are framed oriented tangles with endpoints PP. The framing and the orientation of LL restrict to a framing and an orientation of PYP\subset\partial Y.

If T0YT_{0}\subset Y is a framed oriented tangle with T0=P\partial T_{0}=P in the framed oriented sense, then T1T0T_{1}\cup T_{0} is a framed oriented link in X1\partial X_{1} and T0T2-T_{0}\cup T_{2} is one in X2\partial X_{2}. A lasagna filling of (X1,T1T0)(X_{1},T_{1}\cup T_{0}) and a lasagna filling of (X2,T0T2)(X_{2},-T_{0}\cup T_{2}) glue to a lasagna filling of (X,L)(X,L) in a bilinear way, inducing a map

𝒮0Z(X1;T1T0)𝒮0Z(X2;T0T2)𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X_{1};T_{1}\cup T_{0})\otimes\mathcal{S}_{0}^{Z}(X_{2};-T_{0}\cup T_{2})\to\mathcal{S}_{0}^{Z}(X;L) (6)
Proposition 2.3 (Gluing).

The gluing map

T0YT0=P𝒮0Z(X1;T1T0)𝒮0Z(X2;T0T2)𝒮0Z(X;L)\bigoplus_{\begin{subarray}{c}T_{0}\subset Y\\ \partial T_{0}=P\end{subarray}}\mathcal{S}_{0}^{Z}(X_{1};T_{1}\cup T_{0})\otimes\mathcal{S}_{0}^{Z}(X_{2};-T_{0}\cup T_{2})\to\mathcal{S}_{0}^{Z}(X;L)

is a surjection.

Proof.

Every lasagna filling of (X,L)(X,L) can be isotoped to intersect YY transversely in some tangle T0T_{0}. ∎

Write L1=T1T0L_{1}=T_{1}\cup T_{0}. A framed surface SX2S\subset X_{2} with S=T0T2\partial S=-T_{0}\cup T_{2} is a lasagna filling of (X2;T0T2)(X_{2};-T_{0}\cup T_{2}) (without input balls), thus putting its class in the second tensor summand of (6) defines a gluing map

𝒮0Z(X2;S):𝒮0Z(X1;L1)𝒮0Z(X;L).\mathcal{S}_{0}^{Z}(X_{2};S)\colon\mathcal{S}_{0}^{Z}(X_{1};L_{1})\to\mathcal{S}_{0}^{Z}(X;L). (7)

The most useful cases of gluing might be when YY is closed, where P=P=\emptyset, and T0,T1,T2T_{0},T_{1},T_{2} are links. Still more specially, if Xint(X)X\subset int(X^{\prime}) is an inclusion of 44-manifolds, there is an induced map 𝒮0Z(X)𝒮0Z(X)\mathcal{S}_{0}^{Z}(X)\to\mathcal{S}_{0}^{Z}(X^{\prime}) by gluing the exterior of XXX\subset X^{\prime} to XX. For example, the map in Proposition 2.1 is defined this way.

The skein lasagna module with respect to ZZ refines the TQFT ZZ in the following sense.

Proposition 2.4 (Recover TQFT).

For a framed oriented link LS3L\subset S^{3} we have a natural isomorphism 𝒮0Z(B4;L)Z(L)\mathcal{S}_{0}^{Z}(B^{4};L)\cong Z(L). If SS3×IS\subset S^{3}\times I is a cobordism from LL to LL^{\prime}, then the gluing map 𝒮0Z(S3×I;S):𝒮0Z(B4;L)𝒮0Z(B4;L)\mathcal{S}_{0}^{Z}(S^{3}\times I;S)\colon\mathcal{S}_{0}^{Z}(B^{4};L)\to\mathcal{S}_{0}^{Z}(B^{4};L^{\prime}) agrees with Z(S):Z(L)Z(L)Z(S)\colon Z(L)\to Z(L^{\prime}) under the natural isomorphisms.

Proof.

The statement on objects is proved in [morrison2022invariants, Example 5.6] and the statement on morphisms is a simple exercise. ∎

Finally, we remark that extra structures on the TQFT ZZ usually induce structures on the skein lasagna module 𝒮0Z\mathcal{S}_{0}^{Z}. For example, suppose ZZ takes values in RR-modules with a grading by an abelian group AA, so that for a link cobordism SS, Z(S)Z(S) is homogeneous with degree χ(S)a\chi(S)a for some aAa\in A independent of SS. Then the lasagna module 𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X;L) inherits an AA-grading, once we impose an extra grading shift of χ(Σ)a\chi(\Sigma)a in the definition of Z(Σ)Z(\Sigma) for a skein Σ\Sigma. Analogously, a filtration structure on ZZ induces a filtration structure on 𝒮0Z\mathcal{S}_{0}^{Z}. The isomorphisms or morphisms stated or constructed in this section all respect the extra grading or filtration structure, with a warning that the gluing morphism 𝒮0Z(X2;S)\mathcal{S}_{0}^{Z}(X_{2};S) in (7) has degree or filtration degree (χ(S)χ(T0))a(\chi(S)-\chi(T_{0}))a.

Remark 2.5.

In view of the gluing morphism (7) (in the special case Y=\partial Y=\emptyset and T1=T_{1}=\emptyset), one is tempted to define a TQFT for framed oriented links in oriented 33-manifolds and cobordisms between them. Namely, for a 33-manifold YY and link LYL\subset Y, define

Z(Y,L):=X=Y𝒮0Z(X;L).Z(Y,L):=\bigoplus_{\partial X=Y}\mathcal{S}_{0}^{Z}(X;L).

For a cobordism (W,S):(Y,L)(Y,L)(W,S)\colon(Y,L)\to(Y^{\prime},L^{\prime}), define Z(W,S):Z(Y,L)Z(Y,L)Z(W,S)\colon Z(Y,L)\to Z(Y^{\prime},L^{\prime}) by taking the direct sum of the gluing maps 𝒮0Z(X;L)𝒮0Z(XYW;L)\mathcal{S}_{0}^{Z}(X;L)\to\mathcal{S}_{0}^{Z}(X\cup_{Y}W;L^{\prime}).

However, the resulting modules Z(Y,L)Z(Y,L) are usually infinite dimensional. To obtain a finite theory, one has to either restrict the domain category (one extreme of which is to recover back the TQFT ZZ for links in S3S^{3} again), or modulo suitable skein-theoretic relations on Z(Y,L)Z(Y,L). We leave the exploration of these possibilities to future works.

2.3. Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} homology and its Lee deformation

In this paper we will take the TQFT ZZ in the skein lasagna construction to be the Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} homology or its Lee deformation. In this section we review some features of these TQFTs, partly also to fix the renormalization conventions and do some bookkeeping. We assume the reader is familiar with the classical Khovanov homology as well as its Lee deformation. See [khovanov2000categorification, lee2005endomorphism, rasmussen2010khovanov] for relevant backgrounds and [bar2002khovanov] for an accessible overview.

We follow the convention in [morrison2022invariants]. Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} homology is a renormalization of the Khovanov-Rozansky 𝔰𝔩2\mathfrak{sl}_{2} homology defined in [khovanov2008matrix] (and extended to \mathbb{Z}-coefficient in [blanchet2010oriented]) that is sensitive to the framing of a link. Since conventionally most calculations have been done using the classical Khovanov homology [khovanov2000categorification], we write down explicitly the renormalization differences between it and the 𝔤𝔩2\mathfrak{gl}_{2} theory.

Let KhR2KhR_{2} and KhKh denote the Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} homology and the usual Khovanov homology, respectively. Then for a framed oriented link LS3L\subset S^{3}, we have

KhR2h,q(L)Khh,qw(L)(L),KhR_{2}^{h,q}(L)\cong Kh^{h,-q-w(L)}(-L), (8)

where w(L)w(L) is the writhe of LL, and L-L is the mirror image of LL. When we want to emphasize that we are working over a commutative ring RR, write Kh(L;R)Kh(L;R) and KhR2(L;R)KhR_{2}(L;R) for the corresponding homology theories.

A dotted framed oriented cobordism S:L0L1S\colon L_{0}\to L_{1} in S3×IS^{3}\times I induces a map KhR2(S):KhR2(L0)KhR2(L1)KhR_{2}(S)\colon KhR_{2}(L_{0})\to KhR_{2}(L_{1}) of bidegree (0,χ(S)+2#(dots))(0,-\chi(S)+2\#(\text{dots})) which only depends on the isotopy class of SS rel boundary. Let S¯:L0L1\bar{S}\colon-L_{0}\to-L_{1} denote the mirror image of SS. Then the identification (8) is functorial in the sense that Kh(S¯):Kh(L0)Kh(L1)Kh(\bar{S})\colon Kh(-L_{0})\to Kh(-L_{1}) is equal to KhR2(S)KhR_{2}(S) under (8) up to sign. Since the classical Khovanov homology is only known to be functorial in 3\mathbb{R}^{3} up to sign, KhR2KhR_{2} can be thought of as a sign fix and a functorial upgrade of KhKh.

Both the constructions of KhR2KhR_{2} and KhKh depend on the input of the Frobenius algebra V=[X]/(X2)V=\mathbb{Z}[X]/(X^{2})777Strictly speaking, the construction of KhR2KhR_{2} uses the formalism of foams, but for simplicity, our discussion stays in the classical Khovanov formalism throughout. with comultiplication and counit given by

Δ1=X1+1X,ΔX=XX,ϵ1=0,ϵX=1.\Delta 1=X\otimes 1+1\otimes X,\ \Delta X=X\otimes X,\ \epsilon 1=0,\ \epsilon X=1.

Here XX has quantum degree 22 in the KhR2KhR_{2} case and 2-2 in the KhKh case.

The Lee deformation of VV over \mathbb{Q} [lee2005endomorphism] is the Frobenius algebra VLee=[X]/(X21)V_{Lee}=\mathbb{Q}[X]/(X^{2}-1) with

Δ1=X1+1X,ΔX=XX+11,ϵ1=0,ϵX=1.\Delta 1=X\otimes 1+1\otimes X,\ \Delta X=X\otimes X+1\otimes 1,\ \epsilon 1=0,\ \epsilon X=1.

Using VLeeV_{Lee} instead of VV defines the Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} Lee homology and the classical Lee homology, denoted KhRLeeKhR_{Lee} and KhLeeKh_{Lee}, respectively. Since VLeeV_{Lee} is no longer graded but only filtered, the resulting homology theories carry quantum filtration structures, instead of quantum gradings (although a compatible quantum /4\mathbb{Z}/4-grading is still present). We remark that 0KhRLee0\in KhR_{Lee} has quantum filtration degree -\infty while 0KhLee0\in Kh_{Lee} has quantum filtration degree ++\infty, since the two filtrations go in different directions.

Either KhR2KhR_{2} or KhRLeeKhR_{Lee} changes only by a renormalization upon changing the orientation of the link LL. More precisely, if (L,𝔬)(L,\mathfrak{o}) denote the link LL equipped with a possibly different orientation 𝔬\mathfrak{o}, then for {2,Lee}\bullet\in\{2,Lee\},

KhR(L,𝔬)KhR(L)[w(L)w(L,𝔬)2]{w(L,𝔬)w(L)2}.KhR_{\bullet}(L,\mathfrak{o})\cong KhR_{\bullet}(L)\left[\frac{w(L)-w(L,\mathfrak{o})}{2}\right]\left\{\frac{w(L,\mathfrak{o})-w(L)}{2}\right\}. (9)

It is sometimes more convenient to change from the basis {1,X}\{1,X\} of VLeeV_{Lee} to the basis {𝐚=(1+X)/2,𝐛=(1X)/2}\{\mathbf{a}=(1+X)/2,\,\mathbf{b}=(1-X)/2\}. In this new basis,

𝐚𝐛=0,𝐚2=𝐚,𝐛2=𝐛, 1=𝐚+𝐛,X=𝐚𝐛,Δ𝐚=2𝐚𝐚,Δ𝐛=2𝐛𝐛,ϵ𝐚=12,ϵ𝐛=12.\mathbf{a}\mathbf{b}=0,\ \mathbf{a}^{2}=\mathbf{a},\ \mathbf{b}^{2}=\mathbf{b},\ 1=\mathbf{a}+\mathbf{b},\ X=\mathbf{a}-\mathbf{b},\ \Delta\mathbf{a}=2\mathbf{a}\otimes\mathbf{a},\ \Delta\mathbf{b}=-2\mathbf{b}\otimes\mathbf{b},\ \epsilon\mathbf{a}=\tfrac{1}{2},\ \epsilon\mathbf{b}=-\tfrac{1}{2}.

Like the classical story, the Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} Lee homology of a framed oriented link LL has a basis consisting of “(rescaled) canonical generators” x𝔬x_{\mathfrak{o}} [rasmussen2010khovanov, Section 2.3]888Here x𝔬x_{\mathfrak{o}}’s are defined to be the (grading renormalized) canonical generators on the right hand side of equation (1) in [morrison2024invariants] for Σ={±1}\Sigma=\{\pm 1\} (see Proposition 3.12 there), rescaled by 2#L/2(1)#L(𝔬)/22^{\#L/2}(-1)^{\#L_{-}(\mathfrak{o})/2}. More classically, x𝔬x_{\mathfrak{o}} up to sign is Rasmussen’s generator [𝔰𝔬][\mathfrak{s}_{-\mathfrak{o}}] of L-L renormalized to the 𝔤𝔩2\mathfrak{gl}_{2} theory, rescaled by 2(w(D,𝔬)r(D,𝔬)+#L)/22^{(w(D,-\mathfrak{o})-r(D,-\mathfrak{o})+\#L)/2} where w(D,𝔬),r(D,𝔬)w(D,-\mathfrak{o}),r(D,-\mathfrak{o}) are the writhe and the number of Seifert circles of a framed diagram DD of L-L equipped with orientation 𝔬-\mathfrak{o} used to define [𝔰𝔬][\mathfrak{s}_{-\mathfrak{o}}]. , one for each orientation 𝔬\mathfrak{o} of LL as an unoriented link. Having fixed 𝔬\mathfrak{o} (and thus the generator x𝔬x_{\mathfrak{o}}), we write L=L+(𝔬)L(𝔬)L=L_{+}(\mathfrak{o})\cup L_{-}(\mathfrak{o}) as a union of sublinks on which 𝔬\mathfrak{o} agrees (resp. differs) with the orientation of LL. Then x𝔬x_{\mathfrak{o}} has homological degree 2k(L+(𝔬),L(𝔬))=(w(L,𝔬)w(L))/2-2\ell k(L_{+}(\mathfrak{o}),L_{-}(\mathfrak{o}))=(w(L,\mathfrak{o})-w(L))/2, where w(L,𝔬)w(L,\mathfrak{o}) is the writhe of LL with orientation 𝔬\mathfrak{o}. Moreover, x𝔬+x𝔬¯x_{\mathfrak{o}}+x_{\bar{\mathfrak{o}}}, x𝔬x𝔬¯x_{\mathfrak{o}}-x_{\bar{\mathfrak{o}}} have quantum /4\mathbb{Z}/4 degrees w(L)+#L,w(L)+#L+2-w(L)+\#L,\,-w(L)+\#L+2, respectively, where 𝔬¯\bar{\mathfrak{o}} is the reverse orientation of 𝔬\mathfrak{o} (This statement is preserved under disjoint union, cobordisms (in view of (10)), so it suffices to be checked for the two Hopf links).

The unique canonical generator of the empty link is 11\in\mathbb{Q}. For the 0-framed unknot (U,𝔬)(U,\mathfrak{o}) we have KhRLee(U)=VLee{1}KhR_{Lee}(U)=V_{Lee}\{-1\} as a filtered vector space sitting in homological degree 0, and the canonical generators are given by x𝔬=2𝐚x_{\mathfrak{o}}=2\mathbf{a}, x𝔬¯=2𝐛x_{\bar{\mathfrak{o}}}=-2\mathbf{b}. Via Reidemeister I induced maps, for any nn, the nn-framed unknot (Un,𝔬)(U^{n},\mathfrak{o}) has KhRLee(Un)=VLee{1n}KhR_{Lee}(U^{n})=V_{Lee}\{-1-n\} naturally, and the canonical generators are again given by x𝔬=2𝐚x_{\mathfrak{o}}=2\mathbf{a}, x𝔬¯=2𝐛x_{\mathfrak{\bar{o}}}=-2\mathbf{b}.

The generators x𝔬x_{\mathfrak{o}} behave nicely under cobordisms (cf. Footnote 8 and proof of Theorem 3.3 in [morrison2024invariants] about twisting; see also [rasmussen2005khovanov, Proposition 3.2] for the classical version with sign ambiguity999The equation in [rasmussen2005khovanov, Proposition 3.2] missed a 12\tfrac{1}{2} factor on the exponent on the right hand side.). If S:LLS\colon L\to L^{\prime} is a framed oriented link cobordism and 𝔬\mathfrak{o} is an orientation of LL as an unoriented link, then

KhRLee(S)(x𝔬)=2n(S)𝔒|L=𝔬(1)n(S(𝔒))x𝔒|LKhR_{Lee}(S)(x_{\mathfrak{o}})=2^{n(S)}\sum_{\mathfrak{O}|_{L}=\mathfrak{o}}(-1)^{n(S_{-}(\mathfrak{O}))}x_{\mathfrak{O}|_{L^{\prime}}} (10)

where 𝔒\mathfrak{O} runs over orientations on SS that are compatible with the orientation 𝔬\mathfrak{o} (thus each (S,𝔒)(S,\mathfrak{O}) is an oriented cobordism from (L,𝔬)(L,\mathfrak{o}) to (L,𝔒|L)(L^{\prime},\mathfrak{O}|_{L^{\prime}})), S(𝔒)S_{-}(\mathfrak{O}) is the union of components of SS whose orientation disagree with 𝔒\mathfrak{O}, and

n(Σ)=χ(Σ)+#+Σ#Σ2n(\Sigma)=-\frac{\chi(\Sigma)+\#\partial_{+}\Sigma-\#\partial_{-}\Sigma}{2}\in\mathbb{Z} (11)

for a surface Σ\Sigma with positive boundary +Σ\partial_{+}\Sigma and negative boundary Σ\partial_{-}\Sigma. In particular, n()n(\cdot) is additive under gluing surfaces.

The 𝔤𝔩2\mathfrak{gl}_{2} ss-invariant of LL is defined to be

s𝔤𝔩2(L):=q(x𝔬L)1s_{\mathfrak{gl}_{2}}(L):=q(x_{\mathfrak{o}_{L}})-1

where 𝔬L\mathfrak{o}_{L} is the orientation of LL and q:KhRLee(L){}q\colon KhR_{Lee}(L)\to\mathbb{Z}\sqcup\{-\infty\} is the quantum filtration function. It relates to the classical ss-invariant of oriented links (as defined in [beliakova2008categorification], after [rasmussen2010khovanov]) by

s𝔤𝔩2(L)=s(L)w(L).s_{\mathfrak{gl}_{2}}(L)=-s(-L)-w(L). (12)

2.4. Properties for Khovanov and Lee skein lasagna modules

Letting Z=KhR2Z=KhR_{2} or KhRLeeKhR_{Lee} as the TQFT for the skein lasagna module, we recover the Khovanov and Lee skein lasagna modules, denoted 𝒮02\mathcal{S}_{0}^{2} and 𝒮0Lee\mathcal{S}_{0}^{Lee}, respectively, the first of which agrees with the construction in [morrison2022invariants] (or more concisely in [manolescu2022skein]) for the Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} skein lasagna module. Since KhR2KhR_{2} and KhRLeeKhR_{Lee} allow dots on cobordisms, in the definition of lasagna fillings we may also allow dots on skeins, which does not change the structure of 𝒮02\mathcal{S}_{0}^{2} or 𝒮0Lee\mathcal{S}_{0}^{Lee} but makes the description of certain elements easier. One can think of a dot as an XX-decoration on the unknot in the boundary of an extra deleted local 44-ball near the dot.

It follows from our general discussion in Section 2.1 and Section 2.2 that for a pair (X,L)(X,L) of 44-manifold XX and framed oriented link LXL\subset\partial X, the Khovanov skein lasagna module

𝒮02(X;L)=𝒮0,h,q2(X;L;α)\mathcal{S}_{0}^{2}(X;L)=\bigoplus\mathcal{S}_{0,h,q}^{2}(X;L;\alpha)

is an abelian group with a homological grading by hh\in\mathbb{Z}, a quantum grading by qq\in\mathbb{Z}, and a homology class grading by αH2L(X)\alpha\in H_{2}^{L}(X). Similarly, the Lee skein lasagna module 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) is a \mathbb{Q}-vector space with a homological \mathbb{Z}-grading, a quantum filtration, a compatible quantum /4\mathbb{Z}/4-grading, and a homology class grading by H2L(X)H_{2}^{L}(X). A Khovanov/Lee lasagna filling of (X,L)(X,L) is of the form (Σ,v)(\Sigma,v), where Σ\Sigma is a skein in XX rel LL with some input balls BiB_{i}, and viKhR(Σ|Bi)v\in\otimes_{i}KhR_{\bullet}(\partial\Sigma|_{\partial B_{i}}), {2,Lee}\bullet\in\{2,Lee\}. If vv is homogeneous with homological degree hh and quantum degree (or filtration degree) qq, then the class represented by (Σ,v)(\Sigma,v) in 𝒮0(X;L)\mathcal{S}_{0}^{\bullet}(X;L) has homological degree hh, quantum degree (or filtration degree at most) qχ(Σ)+2#(dots)q-\chi(\Sigma)+2\#(dots), and homology class degree [Σ][\Sigma]. When we work over a ring RR, write 𝒮02(X;L;R)=α𝒮02(X;L;α;R)\mathcal{S}_{0}^{2}(X;L;R)=\oplus_{\alpha}\mathcal{S}_{0}^{2}(X;L;\alpha;R) for the corresponding Khovanov skein lasagna module.

Next we describe a formula of Khovanov and Lee skein lasagna modules for 22-handlebodies, developed in [manolescu2022skein]. From now on till the end of this section, suppose XX is a 22-handlebody obtained by attaching 22-handles along a framed oriented link K=K1KmS3=B4K=K_{1}\cup\cdots\cup K_{m}\subset S^{3}=\partial B^{4} disjoint from a framed oriented link LS3L\subset S^{3}. Then LL can be regarded as a link in X\partial X. A class αH2L(X)\alpha\in H_{2}^{L}(X) intersects the cocore of the 22-handle on KiK_{i} algebraically some αi\alpha_{i} times, and the assignment α(αi)i\alpha\mapsto(\alpha_{i})_{i} gives an isomorphism H2L(X)mH_{2}^{L}(X)\cong\mathbb{Z}^{m}. Write αi+=max(0,αi)\alpha_{i}^{+}=\max(0,\alpha_{i}) and αi=max(0,αi)\alpha_{i}^{-}=\max(0,-\alpha_{i}). Write |||\cdot| for the L1L^{1}-norm on m\mathbb{Z}^{m}.

For k+,k0mk^{+},k^{-}\in\mathbb{Z}_{\geq 0}^{m}, let K(k+,k)LS3K(k^{+},k^{-})\cup L\subset S^{3} denote the framed oriented link obtained from KLK\cup L by replacing each KiK_{i} by ki++kik_{i}^{+}+k_{i}^{-} parallel copies of itself, with orientations on kik_{i}^{-} of them reversed.

Proposition 2.6 (22-handlebody formula).

Let (X,L)(X,L) be given as above. Then

𝒮0(X;L;α)(r0mKhR(K(α++r,α+r)L){|α|2|r|})/\mathcal{S}_{0}^{\bullet}(X;L;\alpha)\cong\left(\bigoplus_{r\in\mathbb{Z}_{\geq 0}^{m}}KhR_{\bullet}(K(\alpha^{+}+r,\alpha^{-}+r)\cup L)\{-|\alpha|-2|r|\}\right)\bigg{/}\sim (13)

where {m}\{m\} denotes the shifting of the quantum degree by mm, {2,Lee}\bullet\in\{2,Lee\}, and \sim is the linear equivalence relation generated by

  1. (i)

    σixx\sigma_{i}\cdot x\sim x for xKhR(K(α++r,α+r)L){|α|2|r|}x\in KhR_{\bullet}(K(\alpha^{+}+r,\alpha^{-}+r)\cup L)\{-|\alpha|-2|r|\} and σiS|αi|+2ri\sigma_{i}\in S_{|\alpha_{i}|+2r_{i}}, where σi\sigma_{i} acts by permuting the cables for KiK_{i} (cf. [grigsby2018annular])101010Strictly speaking, elements in Sαi+2riS_{\alpha_{i}+2r_{i}} may permute cables of KiK_{i} with different orientations, so it does not act on KhR(K(α++r,α+r)L){|α|2|r|}KhR_{\bullet}(K(\alpha^{+}+r,\alpha^{-}+r)\cup L)\{-|\alpha|-2|r|\}. Nevertheless, since changing orientation only affects KhRKhR_{\bullet} by a renormalization, we can first work with K(α++α+2r,0)LK(\alpha^{+}+\alpha^{-}+2r,0)\cup L and then renormalize.;

  2. (ii)

    ϕi(x)x\phi_{i}(x)\sim x for xKhR(K(α++r,α+r)L){|α|2|r|}x\in KhR_{\bullet}(K(\alpha^{+}+r,\alpha^{-}+r)\cup L)\{-|\alpha|-2|r|\} where

    ϕi:KhR(K(α++r,\displaystyle\phi_{i}\colon KhR_{\bullet}(K(\alpha^{+}+r, α+r)L){|α|2|r|}\displaystyle\alpha^{-}+r)\cup L)\{-|\alpha|-2|r|\}\to
    KhR(K(α++r+ei,α+r+ei)L){|α|2|r|2}\displaystyle\,KhR_{\bullet}(K(\alpha^{+}+r+e_{i},\alpha^{-}+r+e_{i})\cup L)\{-|\alpha|-2|r|-2\}

    is the map induced by the dotted annular cobordism that creates two parallel, oppositely oriented cables of KiK_{i} (which is bidegree-preserving), and where eie_{i} is the iith coordinate vector.

Here, when =2\bullet=2, we assume to work over a base ring where 22 is invertible; otherwise, 𝒮02(X;L;α)\mathcal{S}_{0}^{2}(X;L;\alpha) is given by a further quotient of the right hand side of (13).

Proof.

The 𝒮02\mathcal{S}_{0}^{2} version follows from a slight generalization of Theorem 1.1 and Proposition 3.8111111The KhKh in Proposition 3.8 in [manolescu2022skein] should be KhR2KhR_{2} instead. in [manolescu2022skein], stated as Theorem 3.2 in [manolescu2023skein]. The claim for the case when 22 is not invertible follows from the proof of Proposition 3.8 in [manolescu2022skein]. The version for 𝒮0Lee\mathcal{S}_{0}^{Lee} has an identical proof. ∎

Changing the orientation of a link only affects KhRKhR_{\bullet} by a grading shift, given by (9). Therefore, by (13), 𝒮0(X;L)\mathcal{S}_{0}^{\bullet}(X;L) is independent of the orientation of LL (as along as it is null-homologous) up to renormalizations, and 𝒮0(X;L;α)\mathcal{S}_{0}^{\bullet}(X;L;\alpha) only depends on the mod 22 reduction of αH2L(X)H2(X,L)\alpha\in H_{2}^{L}(X)\subset H_{2}(X,L) up to renormalizations. Explicitly, suppose βH2(X,L)\beta\in H_{2}(X,L) has the same mod 22 reduction as αH2L(X)\alpha\in H_{2}^{L}(X) such that βH1(L)\partial\beta\in H_{1}(L) is the fundamental class of LL with a possibly different orientation. Let (L,β)(L,\partial\beta) denote the link LL with orientation given by β\partial\beta. Then βH2(L,β)(X)\beta\in H_{2}^{(L,\partial\beta)}(X), and

𝒮0(X;(L,β);β)𝒮0(X;L;α)[α2β22]{β2α22}.\mathcal{S}_{0}^{\bullet}(X;(L,\partial\beta);\beta)\cong\mathcal{S}_{0}^{\bullet}(X;L;\alpha)\left[\frac{\alpha^{2}-\beta^{2}}{2}\right]\left\{\frac{\beta^{2}-\alpha^{2}}{2}\right\}. (14)

Here when =2\bullet=2 we assume 22 is invertible in the base ring.

The bilinear product on H2L(X)H_{2}^{L}(X) in (14) (similarly for H2(L,β)(X)H_{2}^{(L,\partial\beta)}(X)) is defined by counting intersections of two surfaces in XX with boundary LL, one copy of which is pushed off LL via its framing. In particular, it is independent of the choice of KLS3K\cup L\subset S^{3}.

Example 2.7.

For the 22-handlebody X=S2×D2X=S^{2}\times D^{2}, KK is the 0-framed unknot. It follows from Proposition 2.6 that 𝒮0(S2×D2)\mathcal{S}_{0}^{\bullet}(S^{2}\times D^{2}) can be computed from the Khovanov or Lee homology of unlinks and dotted cobordism maps between them. Explicitly, [manolescu2022skein, Theorem 1.2] calculated that for every αH2(S2×D2)\alpha\in H_{2}(S^{2}\times D^{2}),

𝒮0,h,q2(S2×D2;α)={,h=0,q=2k00,otherwise.\mathcal{S}_{0,h,q}^{2}(S^{2}\times D^{2};\alpha)=\begin{cases}\mathbb{Z},&h=0,\,q=-2k\leq 0\\ 0,&\text{otherwise}.\end{cases}

Similarly, grq𝒮0,hLee(S2×D2;α)gr_{q}\mathcal{S}_{0,h}^{Lee}(S^{2}\times D^{2};\alpha) is given by the same formula with \mathbb{Z} replaced by \mathbb{Q}.

Using Proposition 2.6, [manolescu2022skein, Theorem 1.3] also obtained partial information on Khovanov skein lasagna modules for D2D^{2}-bundles over S2S^{2} with euler number n0n\neq 0, denoted D(n)D(n), namely that

𝒮0,0,2(D(n);0)={,n<0,=00,otherwise.\mathcal{S}_{0,0,*}^{2}(D(n);0)=\begin{cases}\mathbb{Z},&n<0,\ *=0\\ 0,&\text{otherwise.}\end{cases}

For n=±1n=\pm 1, this also yields the corresponding calculation for 2\mathbb{CP}^{2}, 2¯\overline{\mathbb{CP}^{2}}, in view of Proposition 2.1. These results will be improved for n>0n>0 in Example 3.2, and for n<0n<0 in Example 6.1, Section LABEL:sbsec:torus_traces, and Section LABEL:sbsec:-CP^2.

The set of framings of a link LL in an oriented 33-manifold is affine over π0(L)\mathbb{Z}^{\pi_{0}(L)}. If L,LL,L^{\prime} are two framed links with the same underlying link, the writhe difference of them is defined to be w(L,L):=ϵ(LL)w(L^{\prime},L):=\epsilon(L^{\prime}-L) where LLL^{\prime}-L is thought of as an element in π0(L)\mathbb{Z}^{\pi_{0}(L)} and ϵ:π0(L)\epsilon\colon\mathbb{Z}^{\pi_{0}(L)}\to\mathbb{Z} is the augmentation map, which is defined by taking the sum over all coordinates.

Proposition 2.8 (Framing change).

If LXL^{\prime}\subset\partial X is a framed oriented link which agrees with LXL\subset\partial X upon forgetting the framing, then for {2,Lee}\bullet\in\{2,Lee\},

𝒮0(X;L)𝒮0(X;L){w(L,L)}.\mathcal{S}_{0}^{\bullet}(X;L^{\prime})\cong\mathcal{S}_{0}^{\bullet}(X;L)\{-w(L^{\prime},L)\}.
Proof.

For a given lasagna filling of (X,L)(X,L), create new input balls near each component of LL that intersect the skein in disks. Change the framing of LL to LL^{\prime} and push the framing change into the intersection disks. Then the sum of framings on the input unknots is w(L,L)w(L^{\prime},L). Assigning these unknots the Khovanov or Lee class 11 (the element corresponds to 1KhR(U)1\in KhR_{\bullet}(U) under the Reidemeister I induced isomorphisms, where UU is the 0-framed unknot) produces a lasagna filling of (X,L)(X,L^{\prime}). This assignment induces the claimed isomorphism. ∎

Finally, in view of the trick for framing changes in the proof of Proposition 2.8, we can state a slight improvement of the gluing construction (7). In the general setup when gluing (X2,S)(X_{2},S) to (X1,L1)(X_{1},L_{1}) to obtain (X,L)(X,L), in order to get an induced map on skein lasagna modules we needed to assume SS to be framed so that it can be regarded as a skein in X2X_{2} rel L2L_{2}. In the Khovanov or Lee skein lasagna module setup, however, we can drop the framing assumption, at the expense of deleting extra input balls to compensate. Thus, for any (unframed) surface SX2S\subset X_{2} with the previous assumptions, there is a gluing morphism

𝒮0(X2;S):𝒮0(X1;L1)𝒮0(X;L),\mathcal{S}_{0}^{\bullet}(X_{2};S)\colon\mathcal{S}_{0}^{\bullet}(X_{1};L_{1})\to\mathcal{S}_{0}^{\bullet}(X;L), (15)

which has homological degree 0 and quantum (filtration) degree χ(S)+χ(T0)[S]2-\chi(S)+\chi(T_{0})-[S]^{2}. If SS is already framed, this map agrees with the previous construction.

2.5. Structure of space of skeins

The skein lasagna module 𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X;L) decomposes as a direct sum over Ω0(X;L):=π0(𝒞(X;L))\Omega_{\geq 0}(X;L):=\pi_{0}(\mathcal{C}(X;L)), the set of connected components of 𝒞(X;L)\mathcal{C}(X;L). We call Ω0(X;L)\Omega_{\geq 0}(X;L) the space of skeins in XX rel LL. In this section we show that Ω0(X;L)\Omega_{\geq 0}(X;L) is no finer than H2L(X)H_{2}^{L}(X).

Theorem 2.9.

Taking homology defines an isomorphism

Ω0(X;L)H2L(X).\Omega_{\geq 0}(X;L)\xrightarrow{\cong}H_{2}^{L}(X). (16)

We recall that Ω0(X;L)\Omega_{\geq 0}(X;L) is the set of (properly embedded, oriented) framed surfaces ΣX\iint(Bi)\Sigma\subset X\backslash\sqcup_{i}int(B_{i}) with Σ|X=L\partial\Sigma|_{\partial X}=L up to

  1. (a)

    isotopy rel boundary;

  2. (b)

    enclosing some small balls by some large balls and deleting the interior of the large balls, or its converse.

Although

  1. (c)

    isotopy of input balls in XX (that drags surfaces along)

is not directly allowed in the relations, it can be recovered using a combination of (a)(b).

Lemma 2.10.

The relations (a),(b) between skeins are equivalent to the relation of framed cobordisms in X×IX\times I between such skeins away from the tubular neighborhood of a 11-dimensional complex, relative to LXL\subset\partial X.

Proof.

One direction is clear: (a) is clearly contained in the relation of framed cobordisms, and (b) is realized by taking the identity cobordism but gradually enlarging the small deleted balls so that they merge with each other and become the large balls.

For the converse, after an isotopy we can decompose every framed cobordism of surfaces rel LXL\subset\partial X away from a 11-complex into the following elementary moves:

  1. (i)

    Isotopy of the surface and deleted balls rel X\partial X;

  2. (ii)

    Morse move away from deleted balls;

  3. (iii)

    Pushing a Morse critical point into one deleted ball;

  4. (iv)

    Deleting a local ball away from the surface, or its converse;

  5. (v)

    Choosing a path γ\gamma between two deleted balls B1,B2B_{1},B_{2} disjoint from the surface and replacing B1,B2B_{1},B_{2} by a small tubular neighborhood of B1B2γB_{1}\cup B_{2}\cup\gamma, or its converse.

We see immediately that (i) is realized by (a) and (c), while (iii),(iv), and (v) are realized by (b). Since a Morse move is local, we can use (b) twice to delete a local ball and reglue it back to realize (ii). ∎

Proof of Theorem 2.9.

First we show the map is surjective. Any class αH2L(X)\alpha\in H_{2}^{L}(X) can be represented by an oriented embedded surface ΣX\Sigma\subset X with Σ=L\partial\Sigma=L. The obstruction to putting a framing on Σ\Sigma lies in H2(Σ,L)H^{2}(\Sigma,L). Once we delete local 44-balls at points on each component of Σ\Sigma, this obstruction vanishes and we get a skein Σ\Sigma with [Σ]=α[\Sigma]=\alpha.

Next we show the map is injective. Suppose two skeins Σ0\Sigma_{0}, Σ1\Sigma_{1} have the same homology class. By enclosing all input balls by one large ball, we may assume Σ0,Σ1X:=X\int(B4)\Sigma_{0},\Sigma_{1}\in X^{\circ}:=X\backslash int(B^{4}) for some fixed ball B4B^{4} in XX. Since [Σ1][Σ0]=0H2(X×I,X×I)[\Sigma_{1}]-[\Sigma_{0}]=0\in H_{2}(X^{\circ}\times I,\partial X\times I), we can find a properly embedded immersed oriented 33-manifold YX×IY\subset X^{\circ}\times I that cobounds Σ0,Σ1\Sigma_{0},\Sigma_{1}. By transversality we may assume the self-intersection of YY is a properly embedded 11-manifold in X×IX^{\circ}\times I. Deleting a tubular neighborhood along the self intersection makes YY an embedded cobordism between Σ0\Sigma_{0} and Σ1\Sigma_{1} in X×IX\times I with the tubular neighborhood of a 11-complex deleted. The obstruction of putting a compatible framing on YY lies in H2(Y,Y|(X×I))H1(Y,Y\Y|(X×I))H^{2}(Y,\partial Y|_{\partial(X\times I)})\cong H_{1}(Y,\partial Y\backslash\partial Y|_{\partial(X\times I)}), which can be made zero upon a further deletion of the tubular neighborhood of a 11-complex in X×IX^{\circ}\times I. ∎

Remark 2.11.

An alternative proof of Theorem 2.9 can be deduced from a relative and noncompact version of [kirby2012cohomotopy, Theorem 2] applied to (X,L)(X^{\circ},L) together with the fact that a framed link translation (as defined in their paper) can be undone in Ω0(X;L)\Omega_{\geq 0}(X;L) by a neck-cutting. However, for our purpose in Section 4, we need a similar result for the space of a pair of skeins, which can be proved in the same way as we demonstrated, but is inaccessible by directly applying [kirby2012cohomotopy, Theorem 2].

2.6. Variations of the construction

In this section we assume XX to be connected. The skein lasagna module 𝒮0Z(X;L)\mathcal{S}_{0}^{Z}(X;L) is the colimit of ZZ along the category 𝒞0(X;L):=𝒞(X;L)\mathcal{C}_{\geq 0}(X;L):=\mathcal{C}(X;L) of skeins. One can restrict to the subcategory 𝒞1(X;L)𝒞0(X;L)\mathcal{C}_{1}(X;L)\subset\mathcal{C}_{\geq 0}(X;L) of skeins with exactly 11 input ball. The colimit of ZZ along 𝒞1(X;L)\mathcal{C}_{1}(X;L) produces a slightly different version of skein lasagna module, say 𝒮0Z(X;L)\mathcal{S}_{0}^{{}^{\prime}Z}(X;L). It comes with a map

𝒮0Z(X;L)𝒮0Z(X;L)\mathcal{S}_{0}^{{}^{\prime}Z}(X;L)\twoheadrightarrow\mathcal{S}_{0}^{Z}(X;L) (17)

which is surjective since every object in 𝒞0(X;L)\mathcal{C}_{\geq 0}(X;L) has a successor in 𝒞1(X;L)\mathcal{C}_{1}(X;L).

In fact, (17) is an isomorphism whenever XX has no 11-handles. This is because the relevant formulas of 𝒮0Z\mathcal{S}_{0}^{Z} for attaching 2,3,42,3,4-handles (generalized appropriately for 22-handles when ZKhRZ\neq KhR_{\bullet}), developed by Manolescu-Walker-Wedrich [manolescu2023skein], still hold for 𝒮0Z\mathcal{S}_{0}^{{}^{\prime}Z}. In particular, most calculations done in this paper, including the proof of Theorem 1.1, remain valid with 𝒮0\mathcal{S}_{0}^{\bullet} replaced by 𝒮0\mathcal{S}_{0}^{{}^{\prime}\bullet}, {2,Lee}\bullet\in\{2,Lee\}.

However, when H1(X)0H_{1}(X)\neq 0, (17) is usually not an injection. In fact, a relative and noncompact version of [kirby2012cohomotopy, Theorem 2] shows that the relevant space of skeins Ω1(X;L):=π0(𝒞1(X;L))\Omega_{1}(X;L):=\pi_{0}(\mathcal{C}_{1}(X;L)) admits a surjection onto H2L(X)Ω0(X;L)H_{2}^{L}(X)\cong\Omega_{\geq 0}(X;L) by taking homology, whose fiber is usually nontrivial.

We illustrate the problem in the following example. Take Z=KhR2Z=KhR_{2} and let KK be a knot. Then the lasagna filling (K×S1,1)(K\times S^{1},1) of S3×S1S^{3}\times S^{1} (with any framing on K×S1K\times S^{1}) is equivalent to the lasagna filling (,2)(\emptyset,2) by a neck-cutting relation in 𝒮02\mathcal{S}_{0}^{2} [manolescu2022skein, Lemma 7.2]. However, the neck-cutting relation no longer holds for 𝒮02\mathcal{S}_{0}^{{}^{\prime}2}, so (K×S1,1)(K\times S^{1},1) cannot be simplified there in general. More precisely, if the framing on K×S1K\times S^{1} has twists along the KK factor, then K×S1K\times S^{1} and \emptyset do not represent the same element in Ω1(S3×S1)\Omega_{1}(S^{3}\times S^{1}) [kirby2012cohomotopy], so (K×S1,1)(K\times S^{1},1) is not equivalent to (,2)(\emptyset,2) in 𝒮02(S3×S1)\mathcal{S}_{0}^{{}^{\prime}2}(S^{3}\times S^{1}), as the latter represents a class that is nonzero in the quotient 𝒮02(S3×S1)\mathcal{S}_{0}^{2}(S^{3}\times S^{1}) [manolescu2023skein, Corollary 4.2].

Since the proof for the connected sum formula (Proposition 2.2) requires a neck-cutting, we also lose it for 𝒮02\mathcal{S}_{0}^{{}^{\prime}2}. Thus, although 𝒮0Z\mathcal{S}_{0}^{{}^{\prime}Z} is a refinement of 𝒮0Z\mathcal{S}_{0}^{Z}, its computation could be more challenging. We suspect that when XX is simply connected, 𝒞1(X;L)\mathcal{C}_{1}(X;L) is cofinal in 𝒞0(X;L)\mathcal{C}_{\geq 0}(X;L) and thus (17)\eqref{eq:refine_S_0^Z} is an isomorphism.

One may also consider versions of skein lasagna modules with various conditions on framings and orientations. For example, if one drops the assumption that surfaces are framed, one can obtain a skein lasagna module for every input TQFT of unframed oriented links in S3S^{3}. The underlying space of skeins in XX rel LL, no matter whether one allows only one input ball or finitely many (or none), is again isomorphic to H2L(X)H_{2}^{L}(X). The many-balls version of the unframed skein lasagna module for the Khovanov or Lee TQFT seems to agree with its framed counterpart up to renormalizations, in view of the framing-change trick in the proof of Proposition 2.8. In fact, it could be notationally simpler to work with this version throughout the paper, but we follow the existing convention. The structure of the one-ball version of the unframed Khovanov skein lasagna module is less clear.

3. Vanishing criterion and examples

In this section we prove Theorem 1.4 and give a few examples where the theorem applies. The proof is a straightforward application of Ng’s maximal Thurston-Bennequin bound from Khovanov homology [ng2005legendrian] and the 22-handlebody formula (13).

Lemma 3.1 ([ng2005legendrian, Theorem 1]).

For an oriented link LL, the Khovanov homology Khh,q(L)Kh^{h,q}(L) is supported in the region

TB(L)qhTB(L).TB(L)\leq q-h\leq-TB(-L).
Proof of Theorem 1.4.

It suffices to prove the case X=Xn(K)X=X_{n}(K), nTB(K)n\geq-TB(-K), since the general case follows from it and Proposition 2.3 applied to X1=Xn(K)X_{1}=X_{n}(K), X2=X\int(X1)X_{2}=X\backslash int(X_{1}).

Give KK an orientation, and suppose LS3L\subset S^{3} is a framed oriented link disjoint from KK that gives rise to a link LXn(K)L\subset\partial X_{n}(K). Since nTB(K)-n\leq TB(-K), we can find a Legendrian representative 𝒦\mathcal{K}\cup\mathcal{L} of KL-K\cup-L with Thurston-Bennequin number tb(𝒦)=ntb(\mathcal{K})=-n. Let 𝒦(k+,k)\mathcal{K}(k^{+},k^{-}) denote the Legendrian (k++k)(k^{+}+k^{-})-cable of 𝒦\mathcal{K} with the orientation on kk^{-} of the strands reversed. Then the underlying link type of 𝒦(k+,k)\mathcal{K}(k^{+},k^{-})\cup\mathcal{L} is (K(k+,k)L)-(K(k^{+},k^{-})\cup L) where KK has framing nn. As a framed link with the contact framing, we have

tb(𝒦(k+,k))=w(𝒦(k+,k))=w(K(k+,k)L).tb(\mathcal{K}(k^{+},k^{-})\cup\mathcal{L})=w(\mathcal{K}(k^{+},k^{-})\cup\mathcal{L})=-w(K(k^{+},k^{-})\cup L).

By Lemma 3.1, Kh((K(k+,k)L))Kh(-(K(k^{+},k^{-})\cup L)) is supported in qhw(K(k+,k)L)q-h\geq-w(K(k^{+},k^{-})\cup L). By (8), KhR2(K(k+,k)L)KhR_{2}(K(k^{+},k^{-})\cup L) is supported in q+hw(K(k+,k)L)w(K(k+,k)L)=0q+h\leq w(K(k^{+},k^{-})\cup L)-w(K(k^{+},k^{-})\cup L)=0. Now Proposition 2.6 implies 𝒮02(X;L;α)=0\mathcal{S}_{0}^{2}(X;L;\alpha)=0 for any α\alpha, as KhR2(K(r+α+,r+α)L){|α|2r}KhR_{2}(K(r+\alpha^{+},r+\alpha^{-})\cup L)\{-|\alpha|-2r\} is supported in q+h|α|2rq+h\leq-|\alpha|-2r\to-\infty as rr\to\infty yet the map ϕ\phi in (ii) preserves the grading q+hq+h. ∎

Example 3.2.

The unknot K=UK=U has maximal tbtb number 1-1. The nn-trace on UU is D(n)D(n), the D2D^{2}-bundle over S2S^{2} with euler number nn, which is also a tubular neighborhood of an embedded sphere in a 44-manifold with self-intersection nn. Therefore in this case, Theorem 1.4 says 𝒮02(X;L)=0\mathcal{S}_{0}^{2}(X;L)=0 whenever XX contains an embedded sphere with positive self-intersection. This is an adjunction-type obstruction for 𝒮02\mathcal{S}_{0}^{2} to be nontrivial.

For example, D(n)D(n) with n>0n>0, 2\mathbb{CP}^{2}, S2×S2S^{2}\times S^{2}, or the mirror image of the K3K3 surface have vanishing Khovanov skein lasagna modules.

Example 3.3.

The right handed trefoil 31-3_{1} has TB(31)=1TB(-3_{1})=1. Therefore, the (±1)(\pm 1)-trace on 313_{1} both have vanishing Khovanov skein lasagna modules. The boundary of this 44-manifold is the Poincaré homology sphere Σ(2,3,5)\Sigma(2,3,5) for 1-1, and the mirror image of the Brieskorn sphere Σ(2,3,7)\Sigma(2,3,7) for +1+1.

One way to see embedded traces is through a Kirby diagram of 44-manifold XX. If there is an nn-framed 22-handle attached along some knot KK not going over 11-handles, then the 22-handle together with the 44-ball is a copy of Xn(K)X_{n}(K) inside XX. Consequently, if we see such a 22-handle with nTB(K)n\geq-TB(-K), then 𝒮02(X;L)=0\mathcal{S}_{0}^{2}(X;L)=0 for any LL.

Remark 3.4.

As proved by [manolescu2022skein, Theorem 1.3] (and will be reproved more generally in Section LABEL:sbsec:-CP^2), k2¯k\overline{\mathbb{CP}^{2}} has nonvanishing skein lasagna module. Therefore, we obtain a slice obstruction from Corollary 1.5, which states that if Σ\Sigma is a slice disk in k2¯\int(B4)k\overline{\mathbb{CP}^{2}}\backslash int(B^{4}) for a knot KK, then

TB(K)+[Σ]2<0.TB(K)+[\Sigma]^{2}<0.

In this special case, however, the obstruction is somewhat weaker than the obstruction coming from Rasmussen ss-invariant from [ren2023lee, Corollary 1.5] which implies s(K)[Σ]2|[Σ]|s(K)\leq-[\Sigma]^{2}-|[\Sigma]|. Since TB(K)<s(K)TB(K)<s(K) [plamenevskaya2006transverse, Proposition 4], this implies TB(K)+[Σ]2+|[Σ]|<0TB(K)+[\Sigma]^{2}+|[\Sigma]|<0.

4. Lee skein lasagna modules and lasagna ss-invariants

In this section, we explore the structure of the Lee skein lasagna module for an arbitrary pair (X,L)(X,L) of 44-manifold XX and framed oriented link LXL\subset\partial X. In particular, we prove the statements in Section 1.2 (except Theorem 1.9) in a more general setup.

Recall from Section 2.4 that the Lee skein lasagna module is defined to be the skein lasagna module with KhRLeeKhR_{Lee} as the TQFT input. For a pair (X,L)(X,L), its Lee skein lasagna module 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) is a vector space over \mathbb{Q} with a homological \mathbb{Z}-grading, a quantum /4\mathbb{Z}/4-grading, a homology class grading by H2L(X)H_{2}^{L}(X), and an (increasing) quantum filtration (so that 0 has filtration degree -\infty). Since the filtration degree of an element can decrease under a filtered map, a nonzero element in the colimit 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) may also have filtration degree -\infty. Let q:𝒮0Lee(X;L){}q\colon\mathcal{S}_{0}^{Lee}(X;L)\to\mathbb{Z}\sqcup\{-\infty\} denote the quantum filtration function.

4.1. Structure of Lee skein lasagna module

For a pair (X,L)(X,L), let LiL_{i} be the components of LL and [Li]H1(L)[L_{i}]\in H_{1}(L) be the fundamental class of LiL_{i}. Let :H2(X,L)H1(L)\partial\colon H_{2}(X,L)\to H_{1}(L) be the boundary homomorphism. Define the set of double classes in XX rel LL to be

H2L,×2(X):={(α+,α)H2(X,L)2:α±=iϵi,±[Li],ϵi,±{0,1},ϵi,++ϵi,=1}.H_{2}^{L,\times 2}(X):=\left\{(\alpha_{+},\alpha_{-})\in H_{2}(X,L)^{2}\colon\partial\alpha_{\pm}=\sum_{i}\epsilon_{i,\pm}[L_{i}],\,\epsilon_{i,\pm}\in\{0,1\},\,\epsilon_{i,+}+\epsilon_{i,-}=1\right\}.

Thus for any double class (α+,α)(\alpha_{+},\alpha_{-}) we have α++αH2L(X)\alpha_{+}+\alpha_{-}\in H_{2}^{L}(X).

The purpose of this section is to prove the following generalization of Theorem 1.6.

Theorem 4.1.

The Lee skein lasagna module of (X,L)(X,L) is 𝒮0Lee(X;L)H2L,×2(X)\mathcal{S}_{0}^{Lee}(X;L)\cong\mathbb{Q}^{H_{2}^{L,\times 2}(X)}, with a basis consisting of canonical generators xα+,αx_{\alpha_{+},\alpha_{-}}, one for each double class (α+,α)H2L,×2(X)(\alpha_{+},\alpha_{-})\in H_{2}^{L,\times 2}(X). Moreover,

  1. (1)

    xα+,αx_{\alpha_{+},\alpha_{-}} has homological degree 2α+α-2\alpha_{+}\cdot\alpha_{-};

  2. (2)

    xα+,αx_{\alpha_{+},\alpha_{-}} has homology class degree α:=α++α\alpha:=\alpha_{+}+\alpha_{-};

  3. (3)

    xα+,α+xα,α+x_{\alpha_{+},\alpha_{-}}+x_{\alpha_{-},\alpha_{+}}, xα+,αxα,α+x_{\alpha_{+},\alpha_{-}}-x_{\alpha_{-},\alpha_{+}} have quantum /4\mathbb{Z}/4 degrees α2+#L-\alpha^{2}+\#L, α2+#L+2-\alpha^{2}+\#L+2, respectively;

  4. (4)

    q(xα+,α)=q(xα,α+)=max(q(xα+,α±xα,α+))q(x_{\alpha_{+},\alpha_{-}})=q(x_{\alpha_{-},\alpha_{+}})=\max(q(x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}})), which also equals min(q(xα+,α±xα,α+))+2\min(q(x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}}))+2 if α+α\alpha_{+}\neq\alpha_{-} in H2(X,L;)H_{2}(X,L;\mathbb{Q}) (in particular if LL\neq\emptyset).

Here, the bilinear product on H2(X,L)H_{2}(X,L) is defined using the framing of LL.

When X=B4X=B^{4}, the bilinear product computes the linking numbers of the boundary links and thus in view of Proposition 2.4, Theorem 4.1 recovers classical structural results on KhRLee(L)KhR_{Lee}(L) by Lee [lee2005endomorphism, Proposition 4.3] and Rasmussen [rasmussen2010khovanov, Proposition 2.3,3.3,Lemma 3.5] [beliakova2008categorification, Section 6.1].

We define a double skein in XX rel LL to be a skein Σ\Sigma together with a partition of its components Σ=Σ+Σ\Sigma=\Sigma_{+}\cup\Sigma_{-}. Equivalently, a double skein is a pair (Σ,𝔒)(\Sigma,\mathfrak{O}) where Σ\Sigma is a skein (which is already oriented) and 𝔒\mathfrak{O} is an orientation of Σ\Sigma as an unoriented surface. The equivalence is given by defining Σ+\Sigma_{+} to be the union of components of Σ\Sigma on which 𝔒\mathfrak{O} matches the given orientation of Σ\Sigma. A double skein Σ=Σ+Σ\Sigma=\Sigma_{+}\cup\Sigma_{-} represents a double class ([Σ+],[Σ])H2L,×2(X)([\Sigma_{+}],[\Sigma_{-}])\in H_{2}^{L,\times 2}(X).

The canonical Lee lasagna filling of a double skein (Σ,𝔒)(\Sigma,\mathfrak{O}) is the lasagna filling x(Σ,𝔒)=x(Σ+,Σ)KhRLee(Σ)x(\Sigma,\mathfrak{O})=x(\Sigma_{+},\Sigma_{-})\in KhR_{Lee}(\Sigma) defined as follows. Let BiB_{i} be the input balls for Σ\Sigma and Ki=Σ|BiK_{i}=\partial\Sigma|_{\partial B_{i}} be the input links. Assign KiK_{i} the (rescaled) canonical generator x𝔒|KiKhRLee(Ki)x_{\mathfrak{O}|_{K_{i}}}\in KhR_{Lee}(K_{i}), and put an extra decoration 𝐚\mathbf{a} on each component of Σ+\Sigma_{+} and 𝐛\mathbf{b} on each component of Σ\Sigma_{-} (𝐚,𝐛\mathbf{a},\mathbf{b} are linear combinations of 11 and the dot decoration; cf. Section 2.3 and Section 2.4 for relevant notations). Define x(Σ+,Σ)x(\Sigma_{+},\Sigma_{-}) to be the resulting lasagna filling rescaled by 2n(Σ)(1)n(Σ)2^{-n(\Sigma)}(-1)^{-n(\Sigma_{-})} where n()n(\cdot) is defined as in (11).

Equivalently, the canonical Lee lasagna filling x(Σ,𝔒)=x(Σ+,Σ)x(\Sigma,\mathfrak{O})=x(\Sigma_{+},\Sigma_{-}) can be thought of as obtained by first deleting a local ball on each component of Σ\Sigma, then assigning x𝔒|Kix_{\mathfrak{O}|_{K_{i}}} to each input link KiK_{i} (which now includes an unknot for each component of Σ\Sigma) and rescaling by 2n(Σ)(1)n(Σ)2^{-n(\Sigma^{\circ})}(-1)^{-n(\Sigma^{\circ}_{-})}, where Σ\Sigma^{\circ} is Σ\Sigma with local disks deleted. (Strictly speaking this is not a lasagna filling with skein Σ\Sigma anymore, but on the level of 𝒮0Lee\mathcal{S}_{0}^{Lee} they define the same element. The difference will be insignificant for our discussion.)

Lemma 4.2.

Let Σ\Sigma be a skein in XX rel LL. Every element in KhRLee(Σ)KhR_{Lee}(\Sigma) is equivalent in 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) to a linear combination of canonical Lee lasagna fillings of Σ\Sigma with various double skein structures.

Proof.

Let KiK_{i} be the input links for Σ\Sigma. Every element in KhRLee(Σ)KhR_{Lee}(\Sigma) (possibly with decorations on surfaces) can be written as a linear combination of “pure Lee lasagna fillings,” which are by definition the ones with a decoration x𝔬ix_{\mathfrak{o}_{i}} on each KiK_{i} (for some orientation 𝔬i\mathfrak{o}_{i} of KiK_{i} as an unoriented link), and a decoration of 𝐚\mathbf{a} or 𝐛\mathbf{b} on each component of Σ\Sigma. Now we claim that such a pure Lee lasagna filling is 0 in 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) unless it is a (multiple of a) canonical Lee lasagna filling for some double skein structure on Σ\Sigma.

If a pure Lee lasagna filling does not come from a double skein structure on Σ\Sigma, then the orientation on Σ\Sigma defined by the standard orientation on 𝐚\mathbf{a}-decorated components and the reverse orientation on the 𝐛\mathbf{b}-decorated components is not compatible with 𝔬i\mathfrak{o}_{i} for some ii. Take a component Σk\Sigma_{k} of Σ\Sigma and a knot component KijKiK_{ij}\subset K_{i} where the orientation is incompatible. Think of the 𝐚/𝐛\mathbf{a}/\mathbf{b} decoration on Σk\Sigma_{k} to be a multiple of the Lee canonical generator of a local unknot UU on Σk\partial\Sigma_{k}^{\circ}. Then UU and KijK_{ij} have incompatible orientations. Choose a path γ\gamma in Σk\Sigma_{k}^{\circ} from a point on UU to a point on KijK_{ij}. Then deleting a tubular neighborhood of γ\gamma induces a map KiUKiK_{i}\sqcup U\to K_{i} by connect-summing UU and KijKiK_{ij}\subset K_{i}, which maps the chosen Lee lasagna filling to zero since the corresponding Lee canonical generators on UU and KiK_{i} are induced by incompatible orientations. ∎

If [S]:ΣΣ[S]\colon\Sigma\to\Sigma^{\prime} is a morphism in the category of skeins 𝒞(X;L)\mathcal{C}(X;L), then a double skein structure 𝔒\mathfrak{O} on Σ\Sigma induces one on Σ\Sigma^{\prime} by restricting the orientation, and ([Σ+],[Σ])=([Σ+],[Σ])H2L,×2(X)([\Sigma_{+}],[\Sigma_{-}])=([\Sigma^{\prime}_{+}],[\Sigma^{\prime}_{-}])\in H_{2}^{L,\times 2}(X).

Lemma 4.3.

If Σ=Σ+Σ\Sigma=\Sigma_{+}\cup\Sigma_{-} is a double skein in XX rel LL and [S]:ΣΣ[S]\colon\Sigma\to\Sigma^{\prime} is a morphism in 𝒞(X;L)\mathcal{C}(X;L) inducing the double skein structure Σ+Σ\Sigma^{\prime}_{+}\cup\Sigma^{\prime}_{-} on Σ\Sigma^{\prime}, then

KhRLee([S])(x(Σ+,Σ))=x(Σ+,Σ)+KhR_{Lee}([S])(x(\Sigma_{+},\Sigma_{-}))=x(\Sigma^{\prime}_{+},\Sigma^{\prime}_{-})+\cdots

where the term \cdots is zero in 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L). Moreover, there is no \cdots term if every component of SS has a boundary on some input ball of Σ\Sigma.

Proof.

This is a consequence of (10). More explicitly, we need to calculate the image of x𝔒|Ki\otimes x_{\mathfrak{O}|_{K_{i}}} under the cobordism KhRLee([S]):KhRLee(Ki)KhRLee(Kj)KhR_{Lee}([S])\colon\otimes KhR_{Lee}(K_{i})\to\otimes KhR_{Lee}(K_{j}^{\prime}) and keep track of the extra decorations on surfaces. We may assume, by further decomposing SS, that either every component of SS has some boundary in Σ\Sigma^{\prime}, or SS is a single cap.

We first deal with the first case which is more complicated. To keep track of the decorations on surfaces, suppose a component ΣkΣ\Sigma_{k}\subset\Sigma splits into nk>0n_{k}>0 components in Σ\Sigma^{\prime}, then we rewrite its decoration as 𝐚=𝐚nk\mathbf{a}=\mathbf{a}^{n_{k}} or 𝐛=𝐛nk\mathbf{b}=\mathbf{b}^{n_{k}} and push each copy of 𝐚\mathbf{a} or 𝐛\mathbf{b} into one corresponding component in Σ\Sigma^{\prime}.

By (10),

KhRLee([S])(x𝔒|Ki)= 2n(S)(1)n(S)(x𝔒|Kj)+KhR_{Lee}([S])(\otimes x_{\mathfrak{O}|_{K_{i}}})=\,2^{n(S)}(-1)^{n(S_{-})}(\otimes x_{\mathfrak{O}|_{K_{j}^{\prime}}})+\cdots

where S=S+SS=S_{+}\cup S_{-} is induced from Σ=Σ+Σ\Sigma=\Sigma_{+}\cup\Sigma_{-}, and \cdots is a linear combination of lasagna fillings incompatible with the double skein structure Σ+Σ\Sigma^{\prime}_{+}\cup\Sigma^{\prime}_{-}, thus is zero in 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) by the proof of Lemma 4.2. If every component of SS has a boundary on some input ball, there is only one compatible orientation 𝔒\mathfrak{O} for SS in (10), thus there is no \cdots term.

It now follows that

KhRLee([S])(x(Σ+,Σ))\displaystyle KhR_{Lee}([S])(x(\Sigma_{+},\Sigma_{-}))
=\displaystyle=  2n(Σ)(1)n(Σ)2n(S)(1)n(S)2n(Σ)(1)n(Σ)x(Σ+,Σ)+\displaystyle\,2^{-n(\Sigma)}(-1)^{-n(\Sigma_{-})}2^{n(S)}(-1)^{n(S_{-})}2^{n(\Sigma^{\prime})}(-1)^{n(\Sigma^{\prime}_{-})}x(\Sigma^{\prime}_{+},\Sigma^{\prime}_{-})+\cdots
=\displaystyle= x(Σ+,Σ)+,\displaystyle\,x(\Sigma^{\prime}_{+},\Sigma^{\prime}_{-})+\cdots,

where the additivity of n()n(\cdot) ensures n(Σ)=n(ΣS)=n(Σ)+n(S)n(\Sigma)=n(\Sigma^{\prime}\cup S)=n(\Sigma^{\prime})+n(S) (and similarly for n(Σ)n(\Sigma_{-})).

Next we consider the case when SS is a cap that annihilates an unknot component in some KiK_{i} (with an identity cobordism in other link components, which we may assume to be empty since it contributes an identity map in a tensor summand). Thus Σ=S\Sigma=S is a disk, Σ=\Sigma^{\prime}=\emptyset, and n(Σ)=n(S)=n(Σ)=0n(\Sigma)=n(S)=n(\Sigma^{\prime})=0. If Σ+=Σ\Sigma_{+}=\Sigma, then x(Σ+,Σ)=2𝐚𝐚x(\Sigma_{+},\Sigma_{-})=2\mathbf{a}\otimes\mathbf{a} (one of 𝐚\mathbf{a} is the decoration on Σ\Sigma) is mapped to ϵ(2𝐚𝐚)=1=x(Σ+,Σ)\epsilon(2\mathbf{a}\cdot\mathbf{a})=1=x(\Sigma^{\prime}_{+},\Sigma^{\prime}_{-}). If Σ=Σ\Sigma_{-}=\Sigma, then x(Σ+,Σ)=2𝐛𝐛x(\Sigma_{+},\Sigma_{-})=-2\mathbf{b}\otimes\mathbf{b} is mapped to ϵ(2𝐛𝐛)=1=x(Σ+,Σ)\epsilon(-2\mathbf{b}\cdot\mathbf{b})=1=x(\Sigma^{\prime}_{+},\Sigma^{\prime}_{-}). ∎

Now we are ready to prove Theorem 4.1.

Proof of Theorem 4.1.

We define an augmentation map ε:𝒮0Lee(X;L)H2L,×2(X)\varepsilon\colon\mathcal{S}_{0}^{Lee}(X;L)\to\mathbb{Q}^{H_{2}^{L,\times 2}(X)} by demanding it to map the class represented by a canonical Lee lasagna filling x(Σ+,Σ)x(\Sigma_{+},\Sigma_{-}) to e([Σ+],[Σ])e_{([\Sigma_{+}],[\Sigma_{-}])}, the generator of the ([Σ+],[Σ])([\Sigma_{+}],[\Sigma_{-}])th coordinate of the codomain. Lemma 4.2 and Lemma 4.3 guarantee that ε\varepsilon is well-defined.

We show ε\varepsilon is an isomorphism. Any pair (α+,α)H2L,×2(X)(\alpha_{+},\alpha_{-})\in H_{2}^{L,\times 2}(X) can be represented by properly embedded immersed surfaces Σ±X\Sigma_{\pm}\subset X with Σ±\partial\Sigma_{\pm} forming a partition of LXL\subset X as an oriented unframed link. By transversality we may assume all singularities of Σ=Σ+Σ\Sigma=\Sigma_{+}\cup\Sigma_{-} are transverse double points. Deleting balls around these singularities makes Σ\Sigma embedded in XX with some balls deleted, and after deleting one extra local ball on each component of Σ\Sigma we can put a framing on Σ\Sigma compatibly with LL. Now the class represented by the Lee canonical lasagna filling x(Σ+,Σ)x(\Sigma_{+},\Sigma_{-}) is mapped to e(α+,α)e_{(\alpha_{+},\alpha_{-})}, proving the surjectivity of ε\varepsilon.

We remark that the above proof for surjectivity is a replica of that of (16) with minor changes. We do the same for injectivity but skip the details. It suffices to show any x(Σ+,Σ),x(Σ+,Σ)x(\Sigma_{+},\Sigma_{-}),\,x(\Sigma^{\prime}_{+},\Sigma^{\prime}_{-}) with ([Σ+],[Σ])=([Σ+],[Σ])H2L,×2(X)([\Sigma_{+}],[\Sigma_{-}])=([\Sigma^{\prime}_{+}],[\Sigma^{\prime}_{-}])\in H_{2}^{L,\times 2}(X) represent the same element in 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L). By mimicking the proof of Theorem 2.9, we see that Σ=Σ+Σ\Sigma=\Sigma_{+}\cup\Sigma_{-} and Σ=Σ+Σ\Sigma^{\prime}=\Sigma^{\prime}_{+}\cup\Sigma^{\prime}_{-} are related by morphisms in 𝒞(X;L)\mathcal{C}(X;L) (or their reverses) compatible with the double skein structures. Namely, we can find a zigzag of morphisms in 𝒞(X;L)\mathcal{C}(X;L)

Σ=Σ0Σ1Σ2Σ2k=Σ\Sigma=\Sigma_{0}\leftarrow\Sigma_{1}\rightarrow\Sigma_{2}\leftarrow\cdots\rightarrow\Sigma_{2k}=\Sigma^{\prime}

and double skein structures on each Σi\Sigma_{i} compatible with Σ,Σ\Sigma,\Sigma^{\prime} and the morphisms. Take the corresponding canonical Lee lasagna fillings and apply Lemma 4.3, we obtain a zigzag

x(Σ+,Σ)=x(Σ0+,Σ0)\leftmapstox(Σ1+,Σ1)x(Σ2k+,Σ2k)=x(Σ+,Σ)x(\Sigma_{+},\Sigma_{-})=x(\Sigma_{0+},\Sigma_{0-})\leftmapsto x(\Sigma_{1+},\Sigma_{1-})\mapsto\cdots\mapsto x(\Sigma_{2k+},\Sigma_{2k-})=x(\Sigma^{\prime}_{+},\Sigma^{\prime}_{-})

up to lasagna fillings that are zero in 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L). This proves the injectivity of ε\varepsilon.

The canonical generator xα+,αx_{\alpha_{+},\alpha_{-}} of 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) can now be defined by [x(Σ+,Σ)][x(\Sigma_{+},\Sigma_{-})] for any double skein Σ=Σ+Σ\Sigma=\Sigma_{+}\cup\Sigma_{-} representing (α+,α)(\alpha_{+},\alpha_{-}). It remains to prove the four addenda to the theorem. (2) is clear since [Σ]=[Σ+]+[Σ]=α++α=α[\Sigma]=[\Sigma_{+}]+[\Sigma_{-}]=\alpha_{+}+\alpha_{-}=\alpha. Denote by KiK_{i}’s the input links of Σ\Sigma, which also admit decompositions Ki+KiK_{i+}\cup K_{i-} induced from the decomposition Σ+Σ\Sigma_{+}\cup\Sigma_{-}. The Lee canonical generator x𝔒|Kix_{\mathfrak{O}|_{K_{i}}} has homological degree 2k(Ki+,Ki)-2\ell k(K_{i+},K_{i-}). This proves (1) since k(Ki+,Ki)=[Σ+][Σ]\sum\ell k(K_{i+},K_{i-})=[\Sigma_{+}]\cdot[\Sigma_{-}]. The class xα,α+x_{\alpha_{-},\alpha_{+}} is represented by the reverse double skein Σ=ΣΣ+\Sigma=\Sigma_{-}\cup\Sigma_{+}, which comes from canonical generators x𝔒|Ki¯x_{\overline{\mathfrak{O}|_{K_{i}}}}, which are conjugates of x𝔒|Kix_{\mathfrak{O}|_{K_{i}}}, meaning that they are equal in quantum /4\mathbb{Z}/4 degree w(Ki)+#Ki-w(K_{i})+\#K_{i} and negatives of each other in degree w(Ki)+#Ki+2-w(K_{i})+\#K_{i}+2. The decorations 𝐚,𝐛\mathbf{a},\mathbf{b} (which are equal in degree 0 and negatives of each other in degree 22) also exchange under reversing the double skein structure. Taking into account the extra renormalization factor (1)n()(-1)^{n(\cdot)} and the quantum shift χ(Σ)-\chi(\Sigma), we see that x(Σ+,Σ)x(\Sigma_{+},\Sigma_{-}) and x(Σ,Σ+)x(\Sigma_{-},\Sigma_{+}) agree in quantum /4\mathbb{Z}/4 degree

(w(Ki)+#Ki)χ(Σ)2(n(Σ+)+n(Σ))\displaystyle\sum(-w(K_{i})+\#K_{i})-\chi(\Sigma)-2(n(\Sigma_{+})+n(\Sigma_{-}))
=\displaystyle= (α2+#Σ)χ(Σ)+(χ(Σ)+#L#Σ)\displaystyle\,(-\alpha^{2}+\#\partial_{-}\Sigma)-\chi(\Sigma)+(\chi(\Sigma)+\#L-\#\partial_{-}\Sigma)
=\displaystyle= α2+#L,\displaystyle\,-\alpha^{2}+\#L,

proving (3).

Finally we prove (4) which is more subtle. If q(xα+,α)=q>q(x_{\alpha_{+},\alpha_{-}})=q>-\infty, we can find a Lee lasagna filling (Σ,v)(\Sigma,v) representing xα+,αx_{\alpha_{+},\alpha_{-}} where vKhRLee(Σ)=KhRLee(Ki){χ(Σ)}v\in KhR_{Lee}(\Sigma)=\otimes KhR_{Lee}(K_{i})\{-\chi(\Sigma)\} has q(v)=qq(v)=q and h(v)=h(xα+,α)=2α+αh(v)=h(x_{\alpha_{+},\alpha_{-}})=-2\alpha_{+}\cdot\alpha_{-}. By deleting extra balls if necessary, we may assume every component of Σ\Sigma has a boundary on some input ball. Define a conjugation action ι\iota on KhRLee(Σ)KhR_{Lee}(\Sigma) by 11 on the quantum /4\mathbb{Z}/4 degree α2+#L-\alpha^{2}+\#L part and 1-1 on the other part.

Write v=𝔬v𝔬=𝔬λ𝔬(ix𝔬i)v=\sum_{\mathfrak{o}}v_{\mathfrak{o}}=\sum_{\mathfrak{o}}\lambda_{\mathfrak{o}}(\bigotimes_{i}x_{\mathfrak{o}_{i}}) as a linear combination of Lee canonical generators, where 𝔬=(𝔬i)\mathfrak{o}=(\mathfrak{o}_{i}) runs over orientations of KiK_{i}’s. Each term v𝔬v_{\mathfrak{o}} represents c(𝔬)x[Σ+(𝔒)],[Σ(𝔒)]c(\mathfrak{o})x_{[\Sigma_{+}(\mathfrak{O})],[\Sigma_{-}(\mathfrak{O})]} for some constant c(𝔬)c(\mathfrak{o}) if 𝔬\mathfrak{o} comes from a double skein structure 𝔒\mathfrak{O} on Σ\Sigma, and zero otherwise. In the first case, [ιv𝔬]=c(𝔬)x[Σ],[Σ+][\iota v_{\mathfrak{o}}]=c(\mathfrak{o})x_{[\Sigma_{-}],[\Sigma_{+}]} by (3). In the second case [ιv𝔬]=0[\iota v_{\mathfrak{o}}]=0. Summing up, we have [ιv]=xα,α+[\iota v]=x_{\alpha_{-},\alpha_{+}}, thus q(xα,α+)q(ιv)=q(v)=q(xα+,α)q(x_{\alpha_{-},\alpha_{+}})\leq q(\iota v)=q(v)=q(x_{\alpha_{+},\alpha_{-}}). Now by symmetry q(xα+,α)=q(xα,α+)q(x_{\alpha_{+},\alpha_{-}})=q(x_{\alpha_{-},\alpha_{+}}). By the triangle inequality this also equals max(q(xα+,α±xα,α+))\max(q(x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}})).

Finally, suppose α+α\alpha_{+}\neq\alpha_{-} in H2(X,L;)H_{2}(X,L;\mathbb{Q}) and q(xα+,α)=q>q(x_{\alpha_{+},\alpha_{-}})=q>-\infty, we prove min(q(xα+,α±xα,α+))=q2\min(q(x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}}))=q-2. Since xα+,α±xα,α+x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}} have different /4\mathbb{Z}/4 grading, we know min(q(xα+,α±xα,α+))q2\min(q(x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}}))\leq q-2. To show the reverse inequality we resort to the following linear algebra lemma.

Lemma 4.4.

Let WnW\subset\mathbb{Q}^{n} be a linear subspace. Suppose E1EmE_{1}\sqcup\cdots\sqcup E_{m} is a partition of the cube E={±1}nnE=\{\pm 1\}^{n}\subset\mathbb{Q}^{n} such that each EiE_{i} is contained in a coset of WW, and E1=E2WE_{1}=-E_{2}\not\subset W. Suppose λ:E\lambda\colon E\to\mathbb{Q} satisfies

ϵEiλ(ϵ)={1,i=1±1,i=20,i>2.\sum_{\epsilon\in E_{i}}\lambda(\epsilon)=\begin{cases}1,&i=1\\ \pm 1,&i=2\\ 0,&i>2.\end{cases}

Then there exist a1,,ana_{1},\cdots,a_{n}\in\mathbb{Q} such that

k=1nϵEiλ(ϵ)ϵkak={1,i=11,i=20,i>2.\sum_{k=1}^{n}\sum_{\epsilon\in E_{i}}\lambda(\epsilon)\epsilon_{k}a_{k}=\begin{cases}1,&i=1\\ \mp 1,&i=2\\ 0,&i>2.\end{cases}
Proof.

Pick any =(a1,,an)W(n)\ell=(a_{1},\cdots,a_{n})\in W^{\perp}\subset(\mathbb{Q}^{n})^{*} with (E1)={1}\ell(E_{1})=\{1\}. ∎

We now finish the proof of (4). Choose a lasagna filling (Σ,v)(\Sigma,v) representing xα+,α±xα,α+x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}} with q(v)=q(xα+,α±xα,α+)q(v)=q(x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}}), h(v)=2α+αh(v)=-2\alpha_{+}\cdot\alpha_{-}, where ±=+\pm=+ or - is the sign that minimizes q(xα+,α±xα,α+)q(x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}}). Assume as before every component of Σ\Sigma has a boundary on an input ball. Then we can rewrite

v=𝔒λ𝔒x(Σ+(𝔒),Σ(𝔒))+v=\sum_{\mathfrak{O}}\lambda_{\mathfrak{O}}x(\Sigma_{+}(\mathfrak{O}),\Sigma_{-}(\mathfrak{O}))+\cdots (18)

where 𝔒\mathfrak{O} runs over orientations of Σ\Sigma with [Σ+(𝔒)][Σ(𝔒)]=α+α[\Sigma_{+}(\mathfrak{O})]\cdot[\Sigma_{-}(\mathfrak{O})]=\alpha_{+}\cdot\alpha_{-}, and \cdots denotes terms that do not come from a double skein structure on Σ\Sigma (in particular they represent the zero class in 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L)). Let Σ1,,Σn\Sigma_{1},\cdots,\Sigma_{n} be the components of Σ\Sigma. Identify an orientation 𝔒\mathfrak{O} on Σ\Sigma as a vector 𝔒E={±1}nn\mathfrak{O}\in E=\{\pm 1\}^{n}\subset\mathbb{Q}^{n}, with 𝔒i=1\mathfrak{O}_{i}=1 if and only if 𝔒\mathfrak{O} agrees with the orientation of Σ\Sigma on Σi\Sigma_{i}; in this way the coefficients λ𝔒\lambda_{\mathfrak{O}} in (18) define a function λ:E\lambda:E\to\mathbb{Q}. Partition EE into E1EmE_{1}\sqcup\cdots\sqcup E_{m} according to the value of [Σ+(𝔒)]H2(X,L)[\Sigma_{+}(\mathfrak{O})]\in H_{2}(X,L). Then each EiE_{i} is contained in a coset of W={(k1,,kn)n:iki[Σi]=0H2(X,L;)}nW=\{(k_{1},\cdots,k_{n})\in\mathbb{Q}^{n}\colon\sum_{i}k_{i}[\Sigma_{i}]=0\in H_{2}(X,L;\mathbb{Q})\}\subset\mathbb{Q}^{n}. Assume E1,E2E_{1},E_{2} are the parts with [Σ+(𝔒)]=α+,α[\Sigma_{+}(\mathfrak{O})]=\alpha_{+},\alpha_{-}, respectively. Then E1,E2E_{1},E_{2} satisfy the condition in Lemma 4.4 (they are not contained in WW as α+αH2(X,L;)\alpha_{+}\neq\alpha_{-}\in H_{2}(X,L;\mathbb{Q})). Also, the condition [v]=xα+,α±xα,α+[v]=x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}} is exactly the condition on λ:E\lambda\colon E\to\mathbb{Q} in Lemma 4.4. Let a1,,ana_{1},\cdots,a_{n} be given by the conclusion of the lemma applied to our data (W,E1Em,λ)(W,E_{1}\sqcup\cdots\sqcup E_{m},\lambda). Now consider the operator L=k=1nakXkL=\sum_{k=1}^{n}a_{k}X_{k} on KhRLee(Σ)KhR_{Lee}(\Sigma), where XkX_{k} is a dot operation on Σk\Sigma_{k}. Since X𝐚=𝐚X\mathbf{a}=\mathbf{a} and X𝐛=𝐛X\mathbf{b}=-\mathbf{b}, it follows that

[Lv]=[k=1n𝔒λ𝔒𝔒kak]=xα+,αxα+,α.[Lv]=[\sum_{k=1}^{n}\sum_{\mathfrak{O}}\lambda_{\mathfrak{O}}\mathfrak{O}_{k}a_{k}]=x_{\alpha_{+},\alpha_{-}}\mp x_{\alpha_{+},\alpha_{-}}.

Since the operator LL has filtered degree 22, we conclude that

q=q(xα+,αxα,α+)q(xα+,α±xα,α+)+2=min(q(xα+,α,xα,α+))+2,q=q(x_{\alpha_{+},\alpha_{-}}\mp x_{\alpha_{-},\alpha_{+}})\leq q(x_{\alpha_{+},\alpha_{-}}\pm x_{\alpha_{-},\alpha_{+}})+2=\min(q(x_{\alpha_{+},\alpha_{-}},x_{\alpha_{-},\alpha_{+}}))+2,

as desired. ∎

4.2. Canonical Lee lasagna generators under morphisms

Recall that the induced map on Lee homology of a link cobordism is determined by its actions on Lee canonical generators, given by (10). In this section we prove a lasagna generalization of this formula. In its most general form, the relevant cobordism map we need to examine is (15).

We recall the notations and explain the setup before stating the generalization to (10). The 44-manifold X=X1YX2X=X_{1}\cup_{Y}X_{2} is obtained by gluing X1,X2X_{1},X_{2} along some common part Y,Y¯Y,\overline{Y} of their boundaries, where YY is a 33-manifold possibly with boundary. Two framed oriented tangles T1X1\YT_{1}\subset\partial X_{1}\backslash Y and T0YT_{0}\subset Y glue to a framed oriented link L1X1L_{1}\subset\partial X_{1}, and SX2S\subset X_{2} is a (not necessarily framed) surface with an oriented link boundary S=T0T2=L2X2\partial S=-T_{0}\cup T_{2}=L_{2}\subset\partial X_{2} where T2T_{2} is a tangle in X2\Y¯\partial X_{2}\backslash\overline{Y}. Then L=T1T2L=T_{1}\cup T_{2} is a framed oriented link in X\partial X. The gluing of (X2,S)(X_{2},S) to (X1,L1)(X_{1},L_{1}) along (Y,T0)(Y,T_{0}) induces the cobordism map

𝒮0Lee(X2;S):𝒮0Lee(X1;L1)𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X_{2};S)\colon\mathcal{S}_{0}^{Lee}(X_{1};L_{1})\to\mathcal{S}_{0}^{Lee}(X;L) (19)

as in (15), which has homological degree 0 and quantum filtration degree χ(S)+χ(T0)[S]2-\chi(S)+\chi(T_{0})-[S]^{2}. We recall that by deleting local balls on SS (one for each component) and putting a compatible framing, the punctured surface SS^{\circ} is a skein in X2X_{2} rel L2L_{2} with a natural lasagna filling by assigning 11 to all boundary unknots, and (19) is defined by putting the class of this lasagna filling in the second tensor summand of the more general gluing map (6).

A class αH2L1(X1)\alpha\in H_{2}^{L_{1}}(X_{1}) and a class βH2L2(X2)\beta\in H_{2}^{L_{2}}(X_{2}) glue to a class αβH2L(X)\alpha*\beta\in H_{2}^{L}(X). Similarly, a double class (α+,α)H2L1,×2(X1)(\alpha_{+},\alpha_{-})\in H_{2}^{L_{1},\times 2}(X_{1}) and a double class (β+,β)H2L2,×2(X2)(\beta_{+},\beta_{-})\in H_{2}^{L_{2},\times 2}(X_{2}) glue to a double class (α+,α)(β+,β)=(α+β+,αβ)H2L,×2(X)(\alpha_{+},\alpha_{-})*(\beta_{+},\beta_{-})=(\alpha_{+}*\beta_{+},\alpha_{-}*\beta_{-})\in H_{2}^{L,\times 2}(X) if they are compatible, i.e. if α+,β+\partial\alpha_{+},\partial\beta_{+} (and thus α,β\partial\alpha_{-},\partial\beta_{-}) agree on T0T_{0}, and do not glue if they are incompatible.

The punctured surface SS^{\circ} represents the class [S]H2L2(X2)[S]\in H_{2}^{L_{2}}(X_{2}). Thus the cobordism map (19) decomposes into direct sums of maps 𝒮0Lee(X1;L1;α)𝒮0Lee(X;L;α[S])\mathcal{S}_{0}^{Lee}(X_{1};L_{1};\alpha)\to\mathcal{S}_{0}^{Lee}(X;L;\alpha*[S]).

Theorem 4.5.

The cobordism map (19) is determined by

𝒮0Lee(X2;S)(xα+,α)=2m(S,L1)𝔒(1)m(S(𝔒),L1)xα+[S+(𝔒)],α[S(𝔒)].\mathcal{S}_{0}^{Lee}(X_{2};S)(x_{\alpha_{+},\alpha_{-}})=2^{m(S,L_{1})}\sum_{\mathfrak{O}}(-1)^{m(S_{-}(\mathfrak{O}),L_{1-})}x_{\alpha_{+}*[S_{+}(\mathfrak{O})],\alpha_{-}*[S_{-}(\mathfrak{O})]}. (20)

Here the sum runs over double skein structures on SS compatible with (α+,α)(\alpha_{+},\alpha_{-}), m(S,L1)m(S,L_{1}) is an integral constant depending only on the topological type of the input pair, which equals n(S)n(S) if SL1S\cap L_{1} is a link, thought of as the negative boundary of SS, and L1L_{1-} is the sublink of L1L_{1} with fundamental class α\partial\alpha_{-}.

Proof.

This is almost tautological from the construction. Let Σ=Σ+Σ\Sigma=\Sigma_{+}\cup\Sigma_{-} be a double skein in X1X_{1} rel L1L_{1} representing (α+,α)(\alpha_{+},\alpha_{-}). Then xα+,αx_{\alpha_{+},\alpha_{-}} is represented by the canonical Lee lasagna filling x(Σ+,Σ)x(\Sigma_{+},\Sigma_{-}) on the skein Σ\Sigma, defined by decorating each input link KiK_{i} by x𝔒|Kix_{\mathfrak{O}|_{K_{i}}} and each component of Σ+,Σ\Sigma_{+},\Sigma_{-} by 𝐚,𝐛\mathbf{a},\mathbf{b}, respectively, with an overall scaling factor 2n(Σ)(1)n(Σ)2^{-n(\Sigma)}(-1)^{-n(\Sigma_{-})}. Gluing in SS produces a skein ΣS\Sigma\cup S^{\circ} in XX rel LL, which has two types of components: the ones that intersect Σ\Sigma and the ones that do not. The lasagna filling also extends to ΣS\Sigma\cup S^{\circ}. The components of the first type already have some 𝐚\mathbf{a} or 𝐛\mathbf{b} decorations from Σ\Sigma. If there is one such component with both 𝐚\mathbf{a} and 𝐛\mathbf{b} decorations, then it represents the zero class in 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L), while there is no orientation on this component compatible with (α+,α)(\alpha_{+},\alpha_{-}), making the right hand side of (20) zero as well. Otherwise, every component of the first type has only 𝐚\mathbf{a} or 𝐛\mathbf{b} decorations that multiply to 𝐚\mathbf{a} or 𝐛\mathbf{b}, which also determine a unique orientation on the component compatible with (α+,α)(\alpha_{+},\alpha_{-}). Each of the components of the second type has two compatible orientations, with a decoration by 1=𝐚+𝐛1=\mathbf{a}+\mathbf{b}. The extra unknots around the punctures of SS^{\circ} can be thought of as assigned 𝐚\mathbf{a} or 𝐛\mathbf{b} for the components of the first type, and 𝐚+𝐛\mathbf{a}+\mathbf{b} for those of the second type.

Keeping track of the constants, we see 𝒮0Lee(X2;S)(xα+,α)\mathcal{S}_{0}^{Lee}(X_{2};S)(x_{\alpha_{+},\alpha_{-}}) is equal to the sum

2n(ΣS)n(Σ)𝔒(1)n(ΣS(𝔒))n(Σ)xα+[S+(𝔒)],α[S(𝔒)].2^{n(\Sigma\cup S)-n(\Sigma)}\sum_{\mathfrak{O}}(-1)^{n(\Sigma_{-}\cup S_{-}(\mathfrak{O}))-n(\Sigma_{-})}x_{\alpha_{+}*[S_{+}(\mathfrak{O})],\alpha_{-}*[S_{-}(\mathfrak{O})]}.

If T0T_{0} is a link then n(ΣS)n(Σ)=n(S)n(\Sigma\cup S)-n(\Sigma)=n(S) by the additivity of n()n(\cdot). In general,

n(ΣS)n(Σ)=χ(S)χ(T0)+#L#L12=χ(S)χ(SL1)+#(SΔL1)#L12n(\Sigma\cup S)-n(\Sigma)=-\frac{\chi(S)-\chi(T_{0})+\#L-\#L_{1}}{2}=-\frac{\chi(S)-\chi(\partial S\cap L_{1})+\#(\partial S\Delta L_{1})-\#L_{1}}{2}

only depends on the topological type of (S,L1)(S,L_{1}). It is an integer since both n(ΣS)n(\Sigma\cup S) and n(Σ)n(\Sigma) are. Similarly for the exponent on 1-1. ∎

4.3. Lasagna ss-invariants and genus bounds

Let (X,L)(X,L) be a pair as before. In view of Theorem 4.1 we make the following definition.

Definition 4.6.

The lasagna ss-invariant of (X,L)(X,L) at a double class (α+,α)H2L,×2(X)(\alpha_{+},\alpha_{-})\in H_{2}^{L,\times 2}(X) is

s(X;L;α+,α):=q(xα+,α)2α+α{}.s(X;L;\alpha_{+},\alpha_{-}):=q(x_{\alpha_{+},\alpha_{-}})-2\alpha_{+}\cdot\alpha_{-}\in\mathbb{Z}\sqcup\{-\infty\}.

The lasagna ss-invariant of (X,L)(X,L) at a class αH2L(X)\alpha\in H_{2}^{L}(X) is

s(X;L;α):=s(X;L;α,0)=q(xα,0){}.s(X;L;\alpha):=s(X;L;\alpha,0)=q(x_{\alpha,0})\in\mathbb{Z}\sqcup\{-\infty\}.

When L=L=\emptyset, we often drop it from the notation of lasagna ss-invariants.

When XX is a 22-handlebody, the isomorphism (14) respects the canonical Lee lasagna generators in the sense that xβ+,β𝒮0Lee(X;(L,β);β)x_{\beta_{+},\beta_{-}}\in\mathcal{S}_{0}^{Lee}(X;(L,\partial\beta);\beta) is identified (up to renormalizations) with xα+,αS0Lee(X;L;α)x_{\alpha_{+},\alpha_{-}}\in S_{0}^{Lee}(X;L;\alpha) for the unique pair (α+,α)H2L,×2(X)(\alpha_{+},\alpha_{-})\in H_{2}^{L,\times 2}(X) with α+α=β+β\alpha_{+}-\alpha_{-}=\beta_{+}-\beta_{-}. In particular,

s(X;L;α+,α)=s(X;(L,(α+α));α+α).s(X;L;\alpha_{+},\alpha_{-})=s(X;(L,\partial(\alpha_{+}-\alpha_{-}));\alpha_{+}-\alpha_{-}). (21)

Thus, the double class version of lasagna ss-invariants in Definition 4.6 can be recovered from the single class version. When XX is not a 22-handlebody, however, it is not clear that (14) holds in a filtered sense for Lee skein lasagna modules, so we do not have this conclusion.

We prove a few properties of lasagna ss-invariants.

Theorem 4.7.

The lasagna ss-invariants satisfy the following properties.

  1. (0)

    (Empty manifold) s(;;0)=0s(\emptyset;\emptyset;0)=0.

  2. (1)

    (Recover classical ss) If * denotes the unique element in H2L(B4)H_{2}^{L}(B^{4}), then

    s(B4;L;)=s𝔤𝔩2(L)+1=s(L)w(L)+1,s(B^{4};L;*)=s_{\mathfrak{gl}_{2}}(L)+1=-s(-L)-w(L)+1,

    where ss is the classical ss-invariant defined by Rasmussen [rasmussen2010khovanov] and Beliakova-Wehrli [beliakova2008categorification].

  3. (2)

    (Symmetries) s(X;L;α+,α)=s(X;L;α,α+)=s(X;Lr;α+,α)s(X;L;\alpha_{+},\alpha_{-})=s(X;L;\alpha_{-},\alpha_{+})=s(X;L^{r};-\alpha_{+},-\alpha_{-}), where LrL^{r} denote the orientation reversal of LL.

  4. (3)

    (Parity) s(X;L;α+,α)(α++α)2+#L(mod2)s(X;L;\alpha_{+},\alpha_{-})\equiv(\alpha_{+}+\alpha_{-})^{2}+\#L\pmod{2} (by convention -\infty has arbitrary parity).

  5. (4)

    (Connected sums) If (X,L=L1L2)(X,L=L_{1}\sqcup L_{2}) is a boundary sum, a connected sum, or the disjoint union of (X1,L1)(X_{1},L_{1}) and (X2,L2)(X_{2},L_{2}), and (αi,+,αi,)H2Li,×2(Xi)(\alpha_{i,+},\alpha_{i,-})\in H_{2}^{L_{i},\times 2}(X_{i}), i=1,2i=1,2, then

    s(X;L;α1,++α2,+,α1,+α2,)=s(X1;L1;α1,+,α1,)+s(X2;L2;α2,+,α2,).s(X;L;\alpha_{1,+}+\alpha_{2,+},\alpha_{1,-}+\alpha_{2,-})=s(X_{1};L_{1};\alpha_{1,+},\alpha_{1,-})+s(X_{2};L_{2};\alpha_{2,+},\alpha_{2,-}).
  6. (5)

    (Reduced connected sum) If (X,L=L1#K1,K2L2)(X,L=L_{1}\#_{K_{1},K_{2}}L_{2}) is a (reduced) boundary sum of (X1,L1)(X_{1},L_{1}) and (X2,L2)(X_{2},L_{2}) performed along framed oriented intervals on components KiLiK_{i}\subset L_{i}, i=1,2i=1,2, and (αi,+,αi,)H2Li,×2(Xi)(\alpha_{i,+},\alpha_{i,-})\in H_{2}^{L_{i},\times 2}(X_{i}) are double classes compatible with the boundary sum, then

    s(X;L1#K1,K2L2;α1,+α2,+,α1,α2,)=s(X1;L1;α1,+,α1,)+s(X2;L2;α2,+,α2,)1.s(X;L_{1}\#_{K_{1},K_{2}}L_{2};\alpha_{1,+}*\alpha_{2,+},\alpha_{1,-}*\alpha_{2,-})=s(X_{1};L_{1};\alpha_{1,+},\alpha_{1,-})+s(X_{2};L_{2};\alpha_{2,+},\alpha_{2,-})-1.
  7. (6)

    (Genus bound) If (X2,L2)(X_{2},L_{2}) is glued to (X1,L1)(X_{1},L_{1}) along some (Y,T0)(Y,T_{0}) to obtain (X,L)(X,L), and (unframed) SX2S\subset X_{2} is properly embedded with S=L2\partial S=L_{2}, as in the setup of (19), such that every component of SS has a boundary on T0T_{0}, then for any compatible partition S=S+SS=S_{+}\cup S_{-} and double class (α+,α)H2L1,×2(X1)(\alpha_{+},\alpha_{-})\in H_{2}^{L_{1},\times 2}(X_{1}),

    s(X;L;α+[S+],α[S])s(X1;L1;α+,α)χ(S)+χ(T0)[S]2.s(X;L;\alpha_{+}*[S_{+}],\alpha_{-}*[S_{-}])\leq s(X_{1};L_{1};\alpha_{+},\alpha_{-})-\chi(S)+\chi(T_{0})-[S]^{2}.

The special case of Theorem 4.7(5) when X1=X2=B4X_{1}=X_{2}=B^{4}, in view of (1), recovers the connected sum formula

s(L1#K1,K2L2)=s(L1)+s(L2)s(L_{1}\#_{K_{1},K_{2}}L_{2})=s(L_{1})+s(L_{2})

for ss-invariant of links in S3S^{3}, a fact that is proved only recently [manolescu2023generalization, Theorem 7.1].

Since we performed gluing in its most general form, Theorem 4.7(6) has a rather complicated look. We state some special cases as corollaries, roughly in decreasing order of generality. Note that for 22-handlebodies, by (21), the double class versions of the genus bounds are equivalent to the single class versions.

Corollary 4.8.

Suppose (X1,L1L0)(X_{1},L_{1}\sqcup L_{0}), (X2,L0L2)(X_{2},-L_{0}\sqcup L_{2}) are two pairs that can be glued along some common boundary (Y,L0)(Y,L_{0}) where YY is a closed 33-manifold (in particular L1,L2YL_{1},L_{2}\not\subset Y). Suppose (unframed) SX2S\subset X_{2} has S=L0L2\partial S=-L_{0}\sqcup L_{2} and each component of SS has a boundary on L0-L_{0}, then for any compatible partition S=S+SS=S_{+}\cup S_{-} and double class (α+,α)H2L1L0,×2(X1)(\alpha_{+},\alpha_{-})\in H_{2}^{L_{1}\sqcup L_{0},\times 2}(X_{1}),

s(X1YX2;L1L2;α+[S+],α[S])s(X1;L1L0;α+,α)χ(S)[S]2.s(X_{1}\cup_{Y}X_{2};L_{1}\sqcup L_{2};\alpha_{+}*[S_{+}],\alpha_{-}*[S_{-}])\leq s(X_{1};L_{1}\sqcup L_{0};\alpha_{+},\alpha_{-})-\chi(S)-[S]^{2}.

In particular, for any class αH2L1L0(X1)\alpha\in H_{2}^{L_{1}\sqcup L_{0}}(X_{1}),

s(X1YX2;L1L2;α[S])s(X1;L1L0;α)χ(S)[S]2.s(X_{1}\cup_{Y}X_{2};L_{1}\sqcup L_{2};\alpha*[S])\leq s(X_{1};L_{1}\sqcup L_{0};\alpha)-\chi(S)-[S]^{2}.
Proof.

This is the special case of Theorem 4.7(6) when YY is closed, so T0=L0T_{0}=L_{0} with χ(T0)=0\chi(T_{0})=0. ∎

Remark 4.9.

Although Theorem 4.7(6) is more general than Corollary 4.8, it is implied by the latter. This is because we can replace (X2,S)(X_{2},S) by its union with a collar neighborhood of (X1,L1)(\partial X_{1},L_{1}). Nevertheless, the proof of Theorem 4.7(6) is only notationally more complicated. The same remark applies to Theorem 4.5.

Corollary 4.10.

Suppose (W,S):(Y1,L1)(Y2,L2)(W,S)\colon(Y_{1},L_{1})\to(Y_{2},L_{2}) is an unframed cobordism between framed oriented links in closed oriented 33-manifolds such that each component of SS has a boundary on L1L_{1}, and XX is a 44-manifold that bounds Y1Y_{1}. Then for any compatible partition S=S+SS=S_{+}\cup S_{-} and double class (α+,α)H2L1,×2(X)(\alpha_{+},\alpha_{-})\in H_{2}^{L_{1},\times 2}(X),

s(XY1W;L2;α+[S+],α[S+])s(X;L1;α+,α)χ(S)[S]2.s(X\cup_{Y_{1}}W;L_{2};\alpha_{+}*[S_{+}],\alpha_{-}*[S_{+}])\leq s(X;L_{1};\alpha_{+},\alpha_{-})-\chi(S)-[S]^{2}.

In particular, for any class αH2L1(X)\alpha\in H_{2}^{L_{1}}(X),

s(XYW;L2;α[S])s(X;L1;α)χ(S)[S]2.s(X\cup_{Y}W;L_{2};\alpha*[S])\leq s(X;L_{1};\alpha)-\chi(S)-[S]^{2}.
Proof.

This is the special case of Corollary 4.8 when X1=Y\partial X_{1}=Y. ∎

In the special case of Corollary 4.10 when SS3×IS\subset S^{3}\times I is a link cobordism between links L1,L2S3L_{1},L_{2}\subset S^{3}, each component of which has a boundary on L1L_{1}, and when X=B4X=B^{4}, in view of Proposition 2.4 and Theorem 4.7(1), we have s(L1)s(L2)χ(S)s(-L_{1})\leq s(-L_{2})-\chi(S). Of course we can reverse the orientations to obtain

s(L1)s(L2)χ(S),s(L_{1})\leq s(L_{2})-\chi(S),

which recovers the classical genus bound for link cobordisms from ss-invariants [beliakova2008categorification, (7)]121212Note (7) in [beliakova2008categorification] incorrectly missed the condition about components of SS having boundaries on ends..

Proof of Proposition 1.11.

This is the special case of the single class version of Corollary 4.10 when L1=L2=S=L_{1}=L_{2}=S=\emptyset. ∎

Corollary 4.11.

Suppose (X,L)(X,L) is a pair and S+,SXS_{+},S_{-}\subset X are disjoint properly embedded surfaces with S+S=L\partial S_{+}\cup\partial S_{-}=L, then

2g(S+)+2g(S)s(X;L;[S+],[S])+[S+]2+[S]2#L.2g(S_{+})+2g(S_{-})\geq s(X;L;[S_{+}],[S_{-}])+[S_{+}]^{2}+[S_{-}]^{2}-\#L.

In particular, if SXS\subset X is properly embedded with S=L\partial S=L, then

2g(S)s(X;L;[S])+[S]2#L.2g(S)\geq s(X;L;[S])+[S]^{2}-\#L.
Proof.

Delete a local ball on each component of S+S_{+} and SS_{-}, regard the rest of XX and S±S_{\pm} as a cobordism and apply Corollary 4.10, we get

s(X;L;[S+],[S])s((B4)(#S++#S);U(#S++#S);)χ(S)[S]2,s(X;L;[S_{+}],[S_{-}])\leq s((B^{4})^{\sqcup(\#S_{+}+\#S_{-})};U^{\sqcup(\#S_{+}+\#S_{-})};*)-\chi(S^{\circ})-[S]^{2},

where s((B4)(#S++#S);U(#S++#S);)=(#S++#S)s(B4;U;)=#S++#Ss((B^{4})^{\sqcup(\#S_{+}+\#S_{-})};U^{\sqcup(\#S_{+}+\#S_{-})};*)=(\#S_{+}+\#S_{-})s(B^{4};U;*)=\#S_{+}+\#S_{-} by Theorem 4.7(1)(4). The statement follows by simplification. ∎

In the special case X=B4X=B^{4} and L=KS3L=K\subset S^{3} is a knot, the single class version of Corollary 4.11 recovers Rasmussen’s slice genus bound for knots [rasmussen2010khovanov, Theorem 1].

Corollary 4.12.

Suppose XX is a 44-manifold, and S+,SXS_{+},S_{-}\subset X are disjoint closed embedded surfaces, then

2g(S+)+2g(S)s(X;[S+],[S])+[S+]2+[S]2.2g(S_{+})+2g(S_{-})\geq s(X;[S_{+}],[S_{-}])+[S_{+}]^{2}+[S_{-}]^{2}.

In particular, if SXS\subset X is an embedded surface, then

2g(S)s(X;[S])+[S]2.2g(S)\geq s(X;[S])+[S]^{2}.
Proof.

This is the special case of Corollary 4.11 when L=L=\emptyset. ∎

The genus function of a 44-manifold XX, g(X;):H2(X)0g(X;\bullet)\colon H_{2}(X)\to\mathbb{Z}_{\geq 0}, is defined by

g(X;α)=min{g(Σ):ΣX is a closed embedded surface with [X]=α}.g(X;\alpha)=\min\{g(\Sigma)\colon\Sigma\subset X\text{ is a closed embedded surface with }[X]=\alpha\}.
Corollary 4.13.

(Theorem 1.8) The genus function of a 44-manifold XX has a lower bound given by

g(X;α)s(X;α)+α22.g(X;\alpha)\geq\frac{s(X;\alpha)+\alpha^{2}}{2}.
Proof.

This is a reformulation of the second part of Corollary 4.12. ∎

We return to Theorem 4.7. It is an easy consequence of the hard work that we have done in previous sections.

Proof of Theorem 4.7.
  1. (0)

    x0,0𝒮0Lee()x_{0,0}\in\mathcal{S}_{0}^{Lee}(\emptyset) is represented by the Lee lasagna filling 11 on the empty skein, which has filtration degree 0.

  2. (1)

    x,0𝒮0Lee(B4;L)x_{*,0}\in\mathcal{S}_{0}^{Lee}(B^{4};L) is represented by the Lee lasagna filling on the skein L×IS3×I=B4\int(12B4)L\times I\subset S^{3}\times I=B^{4}\backslash int(\tfrac{1}{2}B^{4}) with decoration x𝔬Lx_{\mathfrak{o}_{L}} on the input link LL. Under the identification 𝒮0Lee(B4;L)KhRLee(L)\mathcal{S}_{0}^{Lee}(B^{4};L)\cong KhR_{Lee}(L) given by Proposition 2.4, x,0x_{*,0} is equal to the Lee canonical generator x𝔬Lx_{\mathfrak{o}_{L}}, which has filtration degree s𝔤𝔩2(L)+1=s(L)w(L)+1s_{\mathfrak{gl}_{2}}(L)+1=-s(-L)-w(L)+1 by (12).

  3. (2)

    The first equality is a consequence of Theorem 4.1(4). Reversing the orientation of all skeins defines an isomorphism 𝒮0Lee(X;L)𝒮0Lee(X;Lr)\mathcal{S}_{0}^{Lee}(X;L)\cong\mathcal{S}_{0}^{Lee}(X;L^{r}) which negates the homology class grading but preserves other gradings and the quantum filtration, and exchanges the canonical Lee lasagna generators. This proves the second equality.

  4. (3)

    This is a consequence of Theorem 4.1(3).

  5. (4)

    This is a consequence of Proposition 2.2 and the fact that the canonical Lee lasagna generators behave tensorially under the various sum operations (as can be seen on the canonical Lee lasagna filling level).

  6. (6)

    This is a consequence of Theorem 4.5. Under the condition on the components of SS and the compatibility assumption, the right hand side of (20) has exactly one term, which has S±(𝔒)=S±S_{\pm}(\mathfrak{O})=S_{\pm}. The statement follows from the fact that 𝒮0Lee(X2;S)\mathcal{S}_{0}^{Lee}(X_{2};S) has filtered degree χ(S)+χ(T0)[S]2-\chi(S)+\chi(T_{0})-[S]^{2} (cf. the comment after (19)).

  7. (5)

    There is a framed saddle cobordism (X×I,S)(\partial X\times I,S) from (X,L=L1#K1,K2L2)(\partial X,L=L_{1}\#_{K_{1},K_{2}}L_{2}) to (X,L1L2)(\partial X,L_{1}\sqcup L_{2}), where X\partial X is bounded by X=X1X2X=X_{1}\natural X_{2}. Thus Corollary 4.10 (which is a corollary of (6)) and (4) implies

    s(X;L;α1,+α2,+,α1,α2,)s(X1;L1;α1,+,α1,)+s(X2;L2;α2,+,α2,)1.s(X;L;\alpha_{1,+}*\alpha_{2,+},\alpha_{1,-}*\alpha_{2,-})\geq s(X_{1};L_{1};\alpha_{1,+},\alpha_{1,-})+s(X_{2};L_{2};\alpha_{2,+},\alpha_{2,-})-1.

    To prove the reverse inequality, by Theorem 4.1(4), we can choose signs ϵi=±1\epsilon_{i}=\pm 1, i=1,2i=1,2, such that

    q(xαi,+,αi,+ϵixαi,,αi,+)=s(Xi;Li;αi,+,αi,)2,i=1,2.q(x_{\alpha_{i,+},\alpha_{i,-}}+\epsilon_{i}x_{\alpha_{i,-},\alpha_{i,+}})=s(X_{i};L_{i};\alpha_{i,+},\alpha_{i,-})-2,\ i=1,2.

    By Theorem 4.5, the reverse SrS^{r} of SS induces 𝒮0Lee(X×I;Sr):𝒮0Lee(X;L1L2)𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(\partial X\times I;S^{r})\colon\mathcal{S}_{0}^{Lee}(X;L_{1}\sqcup L_{2})\to\mathcal{S}_{0}^{Lee}(X;L) that maps (xα1,+,α1,+ϵ1xα1,,α1,+)(xα2,+,α2,+ϵ2xα2,,α2,+)(x_{\alpha_{1,+},\alpha_{1,-}}+\epsilon_{1}x_{\alpha_{1,-},\alpha_{1,+}})\otimes(x_{\alpha_{2,+},\alpha_{2,-}}+\epsilon_{2}x_{\alpha_{2,-},\alpha_{2,+}}) to a nonzero multiple of xα1,+α2,+,α1,α2,±xα1,α2,,α1,+α2,+x_{\alpha_{1,+}*\alpha_{2,+},\alpha_{1,-}*\alpha_{2,-}}\pm x_{\alpha_{1,-}*\alpha_{2,-},\alpha_{1,+}*\alpha_{2,+}}. Since 𝒮0Lee(X×I;Sr)\mathcal{S}_{0}^{Lee}(\partial X\times I;S^{r}) has filtered degree 11, it follows that

    s(X;L;α1,+α2,+,α1,α2,)\displaystyle\,s(X;L;\alpha_{1,+}*\alpha_{2,+},\alpha_{1,-}*\alpha_{2,-})
    \displaystyle\leq q(xα1,+α2,+,α1,α2,±xα1,α2,,α1,+α2,+)+2\displaystyle\,q(x_{\alpha_{1,+}*\alpha_{2,+},\alpha_{1,-}*\alpha_{2,-}}\pm x_{\alpha_{1,-}*\alpha_{2,-},\alpha_{1,+}*\alpha_{2,+}})+2
    \displaystyle\leq q(xα1,+,α1,+ϵ1xα1,,α1,+)+q(xα2,+,α2,+ϵ2xα2,,α2,+)+3\displaystyle\,q(x_{\alpha_{1,+},\alpha_{1,-}}+\epsilon_{1}x_{\alpha_{1,-},\alpha_{1,+}})+q(x_{\alpha_{2,+},\alpha_{2,-}}+\epsilon_{2}x_{\alpha_{2,-},\alpha_{2,+}})+3
    =\displaystyle= s(X1;L1;α1,+,α1,)+s(X2;L2;α2,+,α2,)1.\displaystyle\,s(X_{1};L_{1};\alpha_{1,+},\alpha_{1,-})+s(X_{2};L_{2};\alpha_{2,+},\alpha_{2,-})-1.\qed

5. Nonvanishing criteria for 22-handlebodies

When XX is a 22-handlebody, the Khovanov or Lee skein lasagna module of a pair (X,L)(X,L) has a nice formula (13). In this section, we make use of this 22-handlebody formula to provide some explicit criteria for 𝒮02(X;L)\mathcal{S}_{0}^{2}(X;L) or 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) to be nonvanishing, where by convention the latter means the filtration function on 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) is not identically -\infty (equivalently, the lasagna ss-invariants of (X,L)(X,L) are not all -\infty). By our discussions in Section 3 and Section 4, these lead to smooth genus bounds for second homology classes of 44-manifolds. This section is an expansion of Section 1.3 and Section 1.4.

Throughout the section, we assume that XX is the 22-handlebody obtained by attaching 22-handles to B4B^{4} along a framed link K=K1KmS3K=K_{1}\cup\cdots K_{m}\subset S^{3}, and LXL\subset\partial X is given by a framed oriented link in S3S^{3} disjoint from KK, as in the setup of (13).

5.1. Rank inequality between Khovanov and Lee skein lasagna modules

In this section we prove Theorem 1.12, which states that the rank of 𝒮02(X;L)\mathcal{S}_{0}^{2}(X;L) is bounded below by the dimension of gr(𝒮0Lee(X;L))gr(\mathcal{S}_{0}^{Lee}(X;L)) in every tri-degree (h,q,α)(h,q,\alpha). In particular, if 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) is nonvanishing, so is 𝒮02(X;L)\mathcal{S}_{0}^{2}(X;L). By Theorem 1.4, this has the following consequence.

Corollary 5.1.

If a knot trace Xn(K)X_{n}(K) embeds into a 44-manifold XX for some knot KK and framing nTB(K)n\geq-TB(-K), then 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) is vanishing for any LXL\subset\partial X. In particular, the lasagna ss-invariants of (X,L)(X,L) are all -\infty.

Proof.

This is a consequence of Theorem 1.4 and Theorem 1.12 when X=Xn(K)X=X_{n}(K). The general case follows from this and Proposition 2.3. ∎

Again, by taking the contrapositive and using the trace embedding lemma one obtains the following slice obstruction.

Corollary 5.2.

Suppose some class in 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) has finite filtration degree. If KS3K\subset S^{3} is nn-slice in XX, then n>TB(K)n>TB(K).∎

We turn to Theorem 1.12. We will only use the 22-handlebody condition in a mild way, namely that the 22-handlebody formula (13) expresses 𝒮0(X;L)\mathcal{S}_{0}^{\bullet}(X;L) as a filtered colimit instead of an arbitrary colimit (4).

Proof of Theorem 1.12.

We first rewrite (13) into a filtered colimit. The group S(r):=S|α1|+2r1××S|αm|+2rmS(r):=S_{|\alpha_{1}|+2r_{1}}\times\cdots\times S_{|\alpha_{m}|+2r_{m}} acts on KhR(K(r))KhR_{\bullet}(K(r)), where K(r):=K(α++r,α+r)LK(r):=K(\alpha^{+}+r,\alpha^{-}+r)\cup L. Then (13) for =Lee\bullet=Lee can be rewritten as

𝒮0Lee(X;L;α)=colimr0mKhRLee(K(r))S(r){|α|2|r|},\mathcal{S}_{0}^{Lee}(X;L;\alpha)=\mathrm{colim}_{r\in\mathbb{Z}_{\geq 0}^{m}}KhR_{Lee}(K(r))^{S(r)}\{-|\alpha|-2|r|\}, (22)

where AS(r)A^{S(r)} denotes the subgroup of AA fixed by S(r)S(r), and the morphisms are given by dotted annular creation maps composed with symmetrization. A similar formula holds for 𝒮02(X;L;)\mathcal{S}_{0}^{2}(X;L;\mathbb{Q}).

Now for a fixed triple (h,q,α)(h,q,\alpha), pick any linearly independent vectors y1,,yn𝒮0,h,qLee(X;L;α)y_{1},\cdots,y_{n}\in\mathcal{S}_{0,h,\leq q}^{Lee}(X;L;\alpha) whose images in grq(𝒮0,hLee(X;L;α))gr_{q}(\mathcal{S}_{0,h}^{Lee}(X;L;\alpha)) are also linearly independent. Since 𝒮0Lee(X;L;α)\mathcal{S}_{0}^{Lee}(X;L;\alpha) is expressed as a filtered colimit, we can choose representatives v1,,vnKhRLeeh(K(r))S(r){|α|2|r|}v_{1},\cdots,v_{n}\in KhR_{Lee}^{h}(K(r))^{S(r)}\{-|\alpha|-2|r|\} for y1,,yny_{1},\cdots,y_{n} with q(vi)=qq(v_{i})=q (q()q(\cdot) denotes the filtration function).

Let (CKhR2(K(r);),d)(CKhR_{2}(K(r);\mathbb{Q}),d) and (CKhRLee(K(r)),dLee)(CKhR_{Lee}(K(r)),d_{Lee}) be the Khovanov-Rozansky 𝔤𝔩2\mathfrak{gl}_{2} and Lee cochain complexes for K(r)K(r), where CKhRLee(K(r))=CKhR2(K(r);)CKhR_{Lee}(K(r))=CKhR_{2}(K(r);\mathbb{Q}) and dLeed_{Lee} is the sum of dd and another term that decreases the quantum degree. We further lift v1,,vnv_{1},\cdots,v_{n} to the chain level to Lee cocycles a1,,anCKhRLeeh(K(r)){|α|2|r|}a_{1},\cdots,a_{n}\in CKhR_{Lee}^{h}(K(r))\{-|\alpha|-2|r|\} with ai=ci+ϵia_{i}=c_{i}+\epsilon_{i} where cic_{i} is homogeneous of quantum degree qq, and q(ϵi)<qq(\epsilon_{i})<q. Since 0=dLeeai=dci+ϵ0=d_{Lee}a_{i}=dc_{i}+\epsilon where q(ϵ)<qq(\epsilon)<q, we see cic_{i} is a Khovanov cocycle.

We claim that any nontrivial linear combination of cic_{i}’s represents a nonzero element in 𝒮0,h,q2(X;L;)\mathcal{S}_{0,h,q}^{2}(X;L;\mathbb{Q}), which would imply rank(𝒮0,h,q2(X;L))n\mathrm{rank}(\mathcal{S}_{0,h,q}^{2}(X;L))\geq n, proving the theorem. Let c=λicic=\sum\lambda_{i}c_{i} where not all λi\lambda_{i} are zero. To prove cc represents a nonzero element, it suffices to check its homology class survives under an arbitrary morphism KhR2(K(r);)S(r){|α|2|r|}KhR2(K(r);)S(r){|α|2|r|}KhR_{2}(K(r);\mathbb{Q})^{S(r)}\{-|\alpha|-2|r|\}\to KhR_{2}(K(r^{\prime});\mathbb{Q})^{S(r^{\prime})}\{-|\alpha|-2|r^{\prime}|\} in the filtered system. Since this map is a linear combination of induced maps by link cobordisms, it is induced by a map Φ:CKhR2(K(r);){|α|2|r|}CKhR2(K(r);){|α|2|r|}\Phi\colon CKhR_{2}(K(r);\mathbb{Q})\{-|\alpha|-2|r|\}\to CKhR_{2}(K(r^{\prime});\mathbb{Q})\{-|\alpha|-2|r^{\prime}|\} on the chain level. Similarly KhRLee(K(r))S(r)KhRLee(K(r))S(r)KhR_{Lee}(K(r))^{S(r)}\to KhR_{Lee}(K(r^{\prime}))^{S(r^{\prime})} is induced by some ΦLee:CKhRLee(K(r)){|α|2|r|}CKhRLee(K(r)){|α|2|r|}\Phi_{Lee}\colon CKhR_{Lee}(K(r))\{-|\alpha|-2|r|\}\to CKhR_{Lee}(K(r^{\prime}))\{-|\alpha|-2|r^{\prime}|\}, where we can assume CKhRLee(CKhR_{Lee}( K(r))CKhR2(K(r);)K(r^{\prime}))\cong CKhR_{2}(K(r^{\prime});\mathbb{Q}) and ΦLee\Phi_{Lee} is a sum of Φ\Phi (which is homogeneous) and some another term that decreases the quantum degree. It follows that a=λiaia=\sum\lambda_{i}a_{i} satisfies ΦLeea=Φc+ϵ\Phi_{Lee}a=\Phi c+\epsilon for some ϵ\epsilon with q(ϵ)<qq(\epsilon)<q. Now, suppose for contrary that Φc=db\Phi c=db is a Khovanov coboundary. Then ΦLeea\Phi_{Lee}a is homologous to ΦLeeadLeeb\Phi_{Lee}a-d_{Lee}b which has filtration degree less than qq. Thus the element y=iλiyiS0Lee(X;L)y=\sum_{i}\lambda_{i}y_{i}\in S_{0}^{Lee}(X;L), which was represented by [a]KhRLee(K(r))S(r)[a]\in KhR_{Lee}(K(r))^{S(r)}, is also represented by [ΦLeea]KhRLee(K(r))S(r)[\Phi_{Lee}a]\in KhR_{Lee}(K(r^{\prime}))^{S(r^{\prime})} with filtration level less than qq, implying q(y)<qq(y)<q, contradicting the choices of y1,,yny_{1},\cdots,y_{n}. ∎

5.2. Comparison results for Lee skein lasagna modules

In this section, we prove two comparison results for Lee skein lasagna modules of 22-handlebodies.

Recall that (X,L)(X,L) is the pair that arises from KLS3K\cup L\subset S^{3} after attaching 2-handles along KK. Let (X,L)(X^{\prime},L^{\prime}) be another such pair from KLK^{\prime}\cup L^{\prime}. Suppose there is a concordance C:KKC\colon K\to K^{\prime}, that is, a genus 0 component-preserving framed link cobordism in S3×IS^{3}\times I (thus topologically CC is a disjoint union of annuli). Then the surgery on CC gives a homology cobordism WW between the surgery on KK and the surgery on KK^{\prime}. In particular, WW (and thus CC) induces an isomorphism φC:H(X)H(X)\varphi_{C}\colon H_{*}(\partial X)\cong H_{*}(\partial X^{\prime}). Moreover, X=XXWX^{\prime}=X\cup_{\partial X}W. If AWA\subset W is a concordance between LL and LL^{\prime}, then (W,A)(W,A) induces an isomorphism H2(X,L)H2(X,L)H_{2}(X,L)\cong H_{2}(X^{\prime},L^{\prime}) that restricts to ϕA:H2L(X)H2L(X)\phi_{A}\colon H_{2}^{L}(X)\cong H_{2}^{L^{\prime}}(X^{\prime}).

Proposition 5.3.

Let (X,L)(X,L) and (X,L)(X^{\prime},L^{\prime}) be obtained from KL,KLS3K\cup L,K^{\prime}\cup L^{\prime}\subset S^{3} as above, and let C:KKC\colon K\to K^{\prime} be a concordance that induces the homology cobordism W:XXW\colon\partial X\to\partial X^{\prime}.

  1. (1)

    If AWA\subset W is a framed concordance between L,LL,L^{\prime}, then 𝒮0Lee(X;L)𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L)\cong\mathcal{S}_{0}^{Lee}(X^{\prime};L^{\prime}) as tri-graded (by ×/4×H2L(X)\mathbb{Z}\times\mathbb{Z}/4\times H_{2}^{L}(X), where H2L(X)H2L(X)H_{2}^{L}(X)\cong H_{2}^{L^{\prime}}(X^{\prime}) via ϕA\phi_{A}), quantum filtered vector spaces. Moreover, s(X;L;α)=s(X;L;ϕA(α))s(X;L;\alpha)=s(X^{\prime};L^{\prime};\phi_{A}(\alpha)) for all αH2L(X)\alpha\in H_{2}^{L}(X).

  2. (2)

    If LL has components L1,,LkL_{1},\cdots,L_{k}, LL^{\prime} has components L1,,LkL_{1}^{\prime},\cdots,L_{k}^{\prime}, and φC([Li])=[Li]\varphi_{C}([L_{i}])=[L_{i}^{\prime}] for all ii, then 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) is nonvanishing if and only if 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X^{\prime};L^{\prime}) is nonvanishing.

Proof.
  1. (1)

    Gluing (W,A)(W,A) to (X,L)(X,L) defines a filtered (degree 0) map 𝒮0Lee(W;A):𝒮0Lee(X;L)𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(W;A)\colon\mathcal{S}_{0}^{Lee}(X;L)\to\mathcal{S}_{0}^{Lee}(X^{\prime};L^{\prime}). Turning (W,A)(W,A) upside down gives another filtered map 𝒮0Lee(X;L)𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X^{\prime};L^{\prime})\to\mathcal{S}_{0}^{Lee}(X;L). By Theorem 4.5, the two maps interchange the corresponding canonical Lee lasagna generators, thus are inverses to each other. The statement follows.

  2. (2)

    By symmetry, it suffices to assume the filtration of 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X;L) is identically -\infty and prove the same for 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X^{\prime};L^{\prime}).
    Under the given condition, we can find framed immersed cobordisms Si:LiLiS_{i}\colon L_{i}\to L_{i}^{\prime} in WW that intersect each other transversely. After deleting some local balls, we can regard S=S1SkS=S_{1}\cup\cdots\cup S_{k} as a skein in WW rel LL-L\sqcup L^{\prime}. For any double class (α+,α)H2L,×2(X)(\alpha_{+},\alpha_{-})\in H_{2}^{L,\times 2}(X), there is a unique compatible double skein structure S=S+SS=S_{+}\cup S_{-}. Gluing (W,S)(W,S) with the canonical Lee lasagna filling x(S+,S)x(S_{+},S_{-}) to (X,L)(X,L) maps xα+,α𝒮0Lee(X;L)x_{\alpha_{+},\alpha_{-}}\in\mathcal{S}_{0}^{Lee}(X;L) to a nonzero multiple of xα+[S+],α[S]𝒮0Lee(X;L)x_{\alpha_{+}*[S_{+}],\alpha_{-}*[S_{-}]}\in\mathcal{S}_{0}^{Lee}(X^{\prime};L^{\prime}). Since q(xα+,α)=q(x_{\alpha_{+},\alpha_{-}})=-\infty, we see q(xα+[S+],α[S])=q(x_{\alpha_{+}*[S_{+}],\alpha_{-}*[S_{-}]})=-\infty as well. Since all canonical Lee lasagna generators of (X,L)(X^{\prime},L^{\prime}) arise this way, we conclude that the filtration function on 𝒮0Lee(X;L)\mathcal{S}_{0}^{Lee}(X^{\prime};L^{\prime}) is identically -\infty.∎

5.3. Lasagna ss-invariant and ss-invariants of cables

In this section we show that the lasagna ss-invariants of (X,L)(X,L) are determined by the classical ss-invariants of cables of KLK\cup L. Recall that as in the setup of (13), the choice KLS3K\cup L\subset S^{3} gives an identification H2L(X)mH_{2}^{L}(X)\cong\mathbb{Z}^{m}.

Proposition 5.4.

For any αH2L(X)m\alpha\in H_{2}^{L}(X)\cong\mathbb{Z}^{m},

s(X;L;α)=\displaystyle s(X;L;\alpha)= limrm(s𝔤𝔩2(K(α++r,α+r)L)2|r|)|α|+1\displaystyle\,\lim_{r\to\infty^{m}}(s_{\mathfrak{gl}_{2}}(K(\alpha^{+}+r,\alpha^{-}+r)\cup L)-2|r|)-|\alpha|+1
=\displaystyle= limrm(s((K(α++r,α+r)L))+2|r|)w(K(α+,α)L)|α|+1.\displaystyle\,-\,\lim_{r\to\infty^{m}}(s(-(K(\alpha^{+}+r,\alpha^{-}+r)\cup L))+2|r|)-w(K(\alpha^{+},\alpha^{-})\cup L)-|\alpha|+1.

Moreover, the quantity within the limit sign in the first (resp. second) line is noninceasing (resp. nondecreasing) in rr.

Proof.

Since the morphisms in the 22-handlebody formula for =Lee\bullet=Lee are filtered, the quantity within the limit sign in the first line is nonincreasing in rr. In particular, the limit exists (which could be -\infty). The 𝔤𝔩2\mathfrak{gl}_{2} ss-invariant relates to the classical ss-invariant via (12). Since w(K(α++r,α+r)L)w(K(\alpha^{+}+r,\alpha^{-}+r)\cup L) is independent of rr, the claims on the second equality follow from the first. Thus it remains to prove the first equality.

We describe explicitly the isomorphism (13). Delete a 44-ball B4B^{4} slightly smaller than the 0-handle of XX and put K(r):=K(α++r,α+r)LK(r):=K(\alpha^{+}+r,\alpha^{-}+r)\cup L on its boundary. Cap off the sublink K(α++r,α+r)K(\alpha^{+}+r,\alpha^{-}+r) by some parallel copies of cores of the 22-handles of various orientations corresponding to (α++r,α+r)(\alpha^{+}+r,\alpha^{-}+r), and put a cylindrical surface L×IL\times I connecting the inner link LB4L\subset\partial B^{4} and the outer link LXL\subset\partial X. This gives a skein in (X,L)(X,L) with input link K(r)K(r). Putting decorations on K(r)K(r) defines KhRLee(K(r))𝒮0Lee(X;L)KhR_{Lee}(K(r))\to\mathcal{S}_{0}^{Lee}(X;L). Taking the direct sum over rr defines a map that descends to the isomorphism (13). It follows that the Lee canonical filling on this skein as a double skein (given by its own orientation) is given up to a scalar by the Lee canonical generator of K(r)K(r) (with its own orientation) decorated on the input ball.

As in the proof of Theorem 1.12, (13) for =Lee\bullet=Lee can be rewritten as (22), where S(r)=S|α1|+2r1×S|αm|+2rmS(r)=S_{|\alpha_{1}|+2r_{1}}\times S_{|\alpha_{m}|+2r_{m}}. Label the cables of KiK_{i} in K(r)K(r) by 1,2,,|αi|+2ri1,2,\cdots,|\alpha_{i}|+2r_{i}. Then the orientations of K(r)K(r) compatible with LL can be identified with tuples (X1,,Xm)(X_{1},\cdots,X_{m}) where each XiX_{i} is a subset of [|αi|+2ri]:={1,2,,|αi|+2ri}[|\alpha_{i}|+2r_{i}]:=\{1,2,\cdots,|\alpha_{i}|+2r_{i}\}, indicating which strands are given the same orientation as KiK_{i}. Those that give rise to possible orientations of K(r)K(r) compatible with the homology class degree α\alpha correspond to such tuples with #Xi=αi++ri\#X_{i}=\alpha_{i}^{+}+r_{i}. Then the canonical Lee lasagna generator xα,0𝒮0Lee(X;L;α)x_{\alpha,0}\in\mathcal{S}_{0}^{Lee}(X;L;\alpha) under (22) is represented by

xinv(r):=#Xi=αi++rix(X1,,Xm)KhRLee(K(r))S(r)x^{inv}(r):=\sum_{\#X_{i}=\alpha_{i}^{+}+r_{i}}x_{(X_{1},\cdots,X_{m})}\in KhR_{Lee}(K(r))^{S(r)}

with a {|α|2|r|}\{-|\alpha|-2|r|\} shift. Define s𝔤𝔩2inv(K(r)):=q(xinv(r))s_{\mathfrak{gl}_{2}}^{inv}(K(r)):=q(x^{inv}(r)). Then by (22) we have

s(X;L;α)=limrm(s𝔤𝔩2inv(K(r))|α|2|r|).s(X;L;\alpha)=\lim_{r\to\infty^{m}}(s_{\mathfrak{gl}_{2}}^{inv}(K(r))-|\alpha|-2|r|). (23)

It remains to show the right hand side of (23) is unchanged if we replace s𝔤𝔩2inv(K(r))s_{\mathfrak{gl}_{2}}^{inv}(K(r)) by s𝔤𝔩2(K(r))+1s_{\mathfrak{gl}_{2}}(K(r))+1, the latter of which is equal to q(x(X1,,Xm))q(x_{(X_{1},\cdots,X_{m})}) where x(X1,,Xm)x_{(X_{1},\cdots,X_{m})} is any term in the summation that defines xinv(r)x^{inv}(r).

Since xinv(r)x^{inv}(r) is a sum of x(X1,,Xm)x_{(X_{1},\cdots,X_{m})} and some S(r)S(r)-translates of it (which all have the same filtration degree), we have q(xinv(r))q(x(X1,,Xm))q(x^{inv}(r))\leq q(x_{(X_{1},\cdots,X_{m})}), thus

limrm(s𝔤𝔩2inv(K(r))|α|2|r|)limrm(s𝔤𝔩2(K(r))+1|α|2|r|).\lim_{r\to\infty^{m}}(s_{\mathfrak{gl}_{2}}^{inv}(K(r))-|\alpha|-2|r|)\leq\lim_{r\to\infty^{m}}(s_{\mathfrak{gl}_{2}}(K(r))+1-|\alpha|-2|r|).

Now suppose the right hand side is not -\infty, so the value stabilizes to some limit MM for rr sufficiently large. We claim that for sufficiently large rr, we have q(xinv(r))=q(x(X1,,Xm))=M+|α|+2|r|q(x^{inv}(r))=q(x_{(X_{1},\cdots,X_{m})})=M+|\alpha|+2|r|, from which the statement follows.

To this end, we follow the argument in [ren2023lee, Section 4]. By looking at the S(r)S(r)-action on the canonical generators, we see

KhRLee(K(r))i=1m𝒫([|αi|+2ri])KhR_{Lee}(K(r))\cong\bigotimes_{i=1}^{m}\mathbb{Q}\,\mathcal{P}([|\alpha_{i}|+2r_{i}])

as an S(r)S(r)-representation, where 𝒫\mathcal{P} denotes the power set operation on a set. Each tensor summand can be further decomposed as

𝒫([|αi|+2ri])=\displaystyle\mathbb{Q}\,\mathcal{P}([|\alpha_{i}|+2r_{i}])= j=0|αi|+2ri{X[|αi|+2ri]:#X=j}\displaystyle\,\bigoplus_{j=0}^{|\alpha_{i}|+2r_{i}}\mathbb{Q}\{X\subset[|\alpha_{i}|+2r_{i}]\colon\#X=j\}
=\displaystyle= j=0|αi|+2ri(k=0min(j,|αi|+2rij)(|αi|+2rik,k))\displaystyle\,\bigoplus_{j=0}^{|\alpha_{i}|+2r_{i}}\left(\bigoplus_{k=0}^{\min(j,|\alpha_{i}|+2r_{i}-j)}(|\alpha_{i}|+2r_{i}-k,k)\right)

into irreducible S|αi|+2riS_{|\alpha_{i}|+2r_{i}}-representations. Here (a,b)(a,b) denote the irreducible Sa+bS_{a+b}-representation over \mathbb{Q} defined by the two-row Young diagram (a,b)(a,b). Since the S(r)S(r)-action on KhRLee(K(r))KhR_{Lee}(K(r)) respects the filtration structure, all nonzero elements in each irreducible piece VV in the decomposition of S(r)S(r) have the same filtration degree, which is denoted q(V)q(V).

Without loss of generality, suppose αi0\alpha_{i}\geq 0 for all ii (otherwise reverse the orientations of some KiK_{i}’s). Then αi+=αi\alpha_{i}^{+}=\alpha_{i}, αi=0\alpha_{i}^{-}=0. The canonical generator x(X1,,Xm)KhRLee(K(r))x_{(X_{1},\cdots,X_{m})}\in KhR_{Lee}(K(r)) lies in the S(r)S(r)-subrepresentation (which is optimally small)

i=1m{X[αi+2ri]:#X=αi+r}=k1=0r1km=0rmi=1m(αi+2riki,ki),\bigotimes_{i=1}^{m}\mathbb{Q}\{X\subset[\alpha_{i}+2r_{i}]\colon\#X=\alpha_{i}+r\}=\bigoplus_{k_{1}=0}^{r_{1}}\cdots\bigoplus_{k_{m}=0}^{r_{m}}\bigotimes_{i=1}^{m}(\alpha_{i}+2r_{i}-k_{i},k_{i}), (24)

while the element xinv(r)x^{inv}(r) lies in the subrepresentation

i=1m(αi+2ri,0).\bigotimes_{i=1}^{m}(\alpha_{i}+2r_{i},0).

Thus q(x(X1,,Xm))=maxk(q(i=1m(αi+2riki,ki)))q(x_{(X_{1},\cdots,X_{m})})=\max_{k}(q(\otimes_{i=1}^{m}(\alpha_{i}+2r_{i}-k_{i},k_{i}))) (this uses the fact that these representations are pairwise nonisomorphic) and q(xinv(r))=q(i=1m(αi+2ri,0))q(x^{inv}(r))=q(\otimes_{i=1}^{m}(\alpha_{i}+2r_{i},0)).

Suppose for the contrary, q(xinv(r))<q(x(X1,,Xm))q(x^{inv}(r))<q(x_{(X_{1},\cdots,X_{m})}). Then there is a summand i=1m(αi+2riki,ki)\otimes_{i=1}^{m}(\alpha_{i}+2r_{i}-k_{i},k_{i}) in (24) with some kt0k_{t}\neq 0 that has q(i=1m(αi+2riki,ki))=q(x(X1,,Xm))=M+|α|+2|r|q(\otimes_{i=1}^{m}(\alpha_{i}+2r_{i}-k_{i},k_{i}))=q(x_{(X_{1},\cdots,X_{m})})=M+|\alpha|+2|r|. The argument in [ren2023lee, Section 4.3] shows that q(i=1m(αi+2rikiδit,kiδit))=q(i=1m(αi+2riki,ki))q(\otimes_{i=1}^{m}(\alpha_{i}+2r_{i}-k_{i}-\delta_{it},k_{i}-\delta_{it}))=q(\otimes_{i=1}^{m}(\alpha_{i}+2r_{i}-k_{i},k_{i})) where i=1m(αi+2rikiδit,kiδit)\otimes_{i=1}^{m}(\alpha_{i}+2r_{i}-k_{i}-\delta_{it},k_{i}-\delta_{it}) is the corresponding S(ret)S(r-e_{t})-irreducible summand of KhRLee(K(ret))KhR_{Lee}(K(r-e_{t})) (ete_{t} is the ttth coordinate vector). However, this implies the corresponding canonical generator for KhRLee(K(ret))KhR_{Lee}(K(r-e_{t})) has

q(x(X1,,Xt\{},,Xm))\displaystyle q(x_{(X_{1},\cdots,X_{t}\backslash\{*\},\cdots,X_{m})})\geq q(i=1m(αi+2rikiδit,kiδit))\displaystyle\,q(\otimes_{i=1}^{m}(\alpha_{i}+2r_{i}-k_{i}-\delta_{it},k_{i}-\delta_{it}))
=\displaystyle= q(x(X1,,Xm))=M+|α|+2|r|>M+|α|+2|ret|.\displaystyle\,q(x_{(X_{1},\cdots,X_{m})})=M+|\alpha|+2|r|>M+|\alpha|+2|r-e_{t}|.

This contradicts rr being sufficiently large (so that lim(s𝔤𝔩2(K())+1|α|2||)=M\lim(s_{\mathfrak{gl}_{2}}(K(\cdot))+1-|\alpha|-2|\cdot|)=M already stabilizes at retr-e_{t}). The proof is complete. ∎

5.4. Diagrammatic nonvanishing result

For simplicity, in this section we assume L=L=\emptyset. We give a diagrammatic criterion to guarantee the nonvanishingness of 𝒮02(X)\mathcal{S}_{0}^{2}(X) and 𝒮0Lee(X)\mathcal{S}_{0}^{Lee}(X).

As before, assume XX is a 22-handlebody obtained by attaching 22-handles to B4B^{4} along a framed oriented link KS3K\subset S^{3} with components K1,,KmK_{1},\cdots,K_{m} and framings p1,,pmp_{1},\cdots,p_{m}. Let p=(p1,,pm)p=(p_{1},\cdots,p_{m}) and PP be the diagonal matrix with entries p1,,pmp_{1},\cdots,p_{m}. Choose a link diagram DD of KK as an unframed link. Let NN denote the crossing matrix of DD, where NiiN_{ii} is the number of crossings of KiK_{i} in diagram DD, and NijN_{ij} is a half the number of crossings between KiK_{i} and KjK_{j} in DD for iji\neq j. Let w=(w1,,wm)w=(w_{1},\cdots,w_{m}) be the writhe vector of KK in DD, and WW be the diagonal matrix with entries w1,,wmw_{1},\cdots,w_{m}. Finally, for f,nmf,n\in\mathbb{Z}^{m}, let KfK^{f} denote the knot KK with framing ff, and Kf(n)K^{f}(n) denote the nn-cable of KK with framing ff, obtained by replacing KiK_{i} by an fif_{i}-framed |ni||n_{i}|-cable, with orientation reversed if ni<0n_{i}<0. Then, any Kw(n)K^{w}(n) has a natural diagram induced from DD by taking blackboard-framed cables.

Theorem 5.5.

In the notations above, suppose a class αH2(X)m\alpha\in H_{2}(X)\cong\mathbb{Z}^{m} satisfies

  1. (i)

    α\alpha maximizes h(α):=αT(PW+N)αh(\alpha^{\prime}):=\alpha^{\prime T}(P-W+N)\alpha^{\prime} among all vectors αα+2m\alpha^{\prime}\in\alpha+2\mathbb{Z}^{m} (in particular, PW+NP-W+N is nonpositive);

  2. (ii)

    Kw(α)K^{w}(\alpha^{\prime}) has no negative crossings in the diagram induced from DD for any αα+2m\alpha^{\prime}\in\alpha+2\mathbb{Z}^{m} with h(α)=h(α)h(\alpha^{\prime})=h(\alpha).

Then

  1. (1)

    For {2,Lee}\bullet\in\{2,Lee\}, 𝒮0,>0(X;α)=0\mathcal{S}_{0,>0}^{\bullet}(X;\alpha)=0;

  2. (2)

    The dimension of 𝒮0,0Lee(X;α)\mathcal{S}_{0,0}^{Lee}(X;\alpha) is equal to the number of maximal points of hh on α+2m\alpha+2\mathbb{Z}^{m}. The rank of 𝒮0,02(X;α)\mathcal{S}_{0,0}^{2}(X;\alpha) is bounded above by this number;

  3. (3)

    s(X;α)=s(Kp(α))w(Kp(α))|α|+1s(X;\alpha)=-s(-K^{p}(\alpha))-w(K^{p}(\alpha))-|\alpha|+1. In particular, XX has nonvanishing Khovanov and Lee skein lasagna modules.

Remark 5.6.

It is possible to allow a boundary link LXL\subset\partial X, but we refrained from doing so to avoid making the statement too complicated.

Corollary 5.7.

In the notations of Theorem 5.5, if PW+NP-W+N is negative definite, then s(X;0)=0s(X;0)=0 and 𝒮0,0,q2(X;0)=grq(𝒮0,0Lee(X;0))={,q=00,q0.\mathcal{S}_{0,0,q}^{2}(X;0)\otimes\mathbb{Q}=gr_{q}(\mathcal{S}_{0,0}^{Lee}(X;0))=\begin{cases}\mathbb{Q},&q=0\\ 0,&q\neq 0.\end{cases}

Proof.

Conditions (i)(ii) in Theorem 5.5 are satisfied for α=0\alpha=0, which is the unique maximal point of hh. Thus (3) implies s(X;0)=s()+1=0s(X;0)=-s(\emptyset)+1=0. By definition of ss, this implies dimgr0(𝒮0,0Lee(X;0))1\dim gr_{0}(\mathcal{S}_{0,0}^{Lee}(X;0))\geq 1. Now (2) and Theorem 1.12 imply the claims on 𝒮0,0(X;0)\mathcal{S}_{0,0}^{\bullet}(X;0). ∎

Proof of Theorem 1.14.

For X=Xn(K)X=X_{n}(K), in the notation of Theorem 5.5 we have m=1m=1, PW+N=n+2n(D)P-W+N=n+2n_{-}(D). Thus (1) follows from Corollary 5.7.

For (2), take DD to be a positive diagram of KK. Then for n0n\leq 0, (i)(ii) in Theorem 5.5 are satisfied for both α=0,1\alpha=0,1. Thus s(X;0)=s()+1=0s(X;0)=-s(\emptyset)+1=0 and s(X;1)=s(K)n1+1=s(K)ns(X;1)=-s(-K)-n-1+1=s(K)-n by Theorem 5.5(3). For n<0n<0, α=±1\alpha=\pm 1 are the unique maximal points of hh on 1+21+2\mathbb{Z}, thus Theorem 5.5(2) implies that the rank of 𝒮0,02(X;1)\mathcal{S}_{0,0}^{2}(X;1) is at most 22. Now the claims on 𝒮0,0(X;1)\mathcal{S}_{0,0}^{\bullet}(X;1) follows from Theorem 1.6(4) and Theorem 1.12. ∎

The rest of the section is devoted to proving Theorem 5.5.

For a link LL, let P(L)(t,q)P(L)(t,q) be the Poincaré polynomial of KhR2(L)KhR_{2}(L). Define

KhR~2(L):=KhR2(L)[12w(L)]{12w(L)}\widetilde{KhR}_{2}(L):=KhR_{2}(L)[\tfrac{1}{2}w(L)]\{-\tfrac{1}{2}w(L)\}

to be the renormalized homology of LL, which is independent of the orientation of LL (cf. (9)). Let P~(L)(t,q)\tilde{P}(L)(t,q) be the Poincaré polynomial of KhR~2(L)\widetilde{KhR}_{2}(L) (which is a Laurent polynomial in t1/2,q1/2t^{1/2},q^{1/2}).

For two (Laurent) polynomials P,QP,Q in variables t1/2,q1/2t^{1/2},q^{1/2}, we write PQP\leq Q if every coefficient of PP is less than or equal to that of QQ. For two vectors a,bma,b\in\mathbb{Z}^{m}, write aba\leq b if aibia_{i}\leq b_{i} for all ii.

Lemma 5.8 (Comparison lemma).

For n0mn\in\mathbb{Z}_{\geq 0}^{m} and pmp\in\mathbb{Z}^{m} with pwp\leq w, we have

P~(Kp(n))0nn2|nnRn,n,wpP~(Kw(n))\tilde{P}(K^{p}(n))\leq\sum_{\begin{subarray}{c}0\leq n^{\prime}\leq n\\ 2|n-n^{\prime}\end{subarray}}R_{n,n^{\prime},w-p}\tilde{P}(K^{w}(n^{\prime})) (25)

for some polynomials Rn,n,wp[t±1/2,q±1/2]R_{n,n^{\prime},w-p}\in\mathbb{Z}[t^{\pm 1/2},q^{\pm 1/2}] independent of KK with

Rn,n,wp(t,q)=ti(wipi)n2i2qi(wipi)(n2i+2ni)2j=1m(rj=0njnj2dim(nj+nj2+rj,njnj2rj)q2rj)+R_{n,n^{\prime},w-p}(t,q)=t^{-\sum_{i}\frac{(w_{i}-p_{i})n^{\prime 2}_{i}}{2}}q^{\sum_{i}\frac{(w_{i}-p_{i})(n^{\prime 2}_{i}+2n^{\prime}_{i})}{2}}\prod_{j=1}^{m}\left(\sum_{r_{j}=0}^{\frac{n_{j}-n^{\prime}_{j}}{2}}\dim\left(\frac{n_{j}+n^{\prime}_{j}}{2}+r_{j},\frac{n_{j}-n^{\prime}_{j}}{2}-r_{j}\right)q^{2r_{j}}\right)+\cdots

where \cdots are terms with lower tt-degrees, and (a,b)(a,b) denotes the irreducible Sa+bS_{a+b}-representation given by the Young diagram (a,b)(a,b).

Moreover, if (25) is sharp at the highest tt-degree of the left hand side, then for any i=1,,mi=1,\cdots,m with ni2n_{i}\geq 2, the morphism on KhR2KhR_{2} induced by the dotted annular creation map Kp(n2ei)Kp(n)K^{p}(n-2e_{i})\to K^{p}(n)131313Really, the dotted annular creation map Kp(n2ei)Kp(nei,ei)K^{p}(n-2e_{i})\to K^{p}(n-e_{i},e_{i}) on KhR2KhR_{2} composed with the isomorphism that changes the orientation of the reversed KiK_{i} strand. is injective in the highest nontrivial homological degree of KhR2(Kp(n))KhR_{2}(K^{p}(n)).

For reader’s convenience, we note that for r,n0r,n\in\mathbb{Z}_{\geq 0}, rn/2r\leq n/2,

dim(nr,r)=(nr)(nr1).\dim(n-r,r)=\binom{n}{r}-\binom{n}{r-1}.

Lemma 5.8 is a consequence of an inductive argument pioneered by Stošić [stovsic2009khovanov]. Although it might have been implicit in the proof of Theorem 1 there, or in more details in [tagami2013maximal], for completeness we sketch a proof in Appendix LABEL:sec:append. For our purpose, we do not really need the quantum shifting information in the lemma.

Lemma 5.9.

For n0mn\in\mathbb{Z}_{\geq 0}^{m}, the maximal tt-degree of P~(Kw(n))\tilde{P}(K^{w}(n)) is at most 12nTNn\tfrac{1}{2}n^{T}Nn.

Proof.

For any framed link diagram DD of LL, the maximal tt-degree of P(L)P(L) is at most the number of negative crossings of DD, and thus the maximal tt-degree of P~(L)\tilde{P}(L) is at most one half the total number of crossings of DD. For DD the standard diagram of Kw(n)K^{w}(n), this number is 12nTNn\tfrac{1}{2}n^{T}Nn. ∎

Proof of Theorem 5.5.

Note that adding positive kinks to DD does not affect the conditions and statements, but increases ww. Thus we can assume that wpw\geq p and DD has no 0-framed unknot component. We also assume α0m\alpha\in\mathbb{Z}_{\geq 0}^{m}, as the general case is only notationally more complicated (and can indeed be deduced from the special case by reversing the orientations of some KiK_{i}’s). Thus α+=α\alpha^{+}=\alpha, α=0\alpha^{-}=0. Write N=N++NN=N_{+}+N_{-} to separate contributions coming from positive/negative crossings.

Recall that the 22-handlebody formula for XX and α0m\alpha\in\mathbb{Z}_{\geq 0}^{m} (13) can be rewritten as

𝒮0(X;α)=colimr0mKhR(Kp(α+r,r))Sα+2r{|α|2|r|},\mathcal{S}_{0}^{\bullet}(X;\alpha)=\mathrm{colim}_{r\in\mathbb{Z}_{\geq 0}^{m}}KhR_{\bullet}(K^{p}(\alpha+r,r))^{S_{\alpha+2r}}\{-|\alpha|-2|r|\}, (26)

where Sn=Sn1××SnmS_{n}=S_{n_{1}}\times\cdots\times S_{n_{m}} for n0mn\in\mathbb{Z}_{\geq 0}^{m}. Here if =2\bullet=2 we assume to work over \mathbb{Q}.

Step 1: Isolating the highest tt-degrees in P~\tilde{P}.

In this first step we prove that, for any n0mn\in\mathbb{Z}_{\geq 0}^{m}, nα(mod2)n\equiv\alpha\pmod{2},

P~(Kp(n)))=0nn2|nnh(n)=h(α)Rn,n,wpP~(Kw(n))+\tilde{P}(K^{p}(n)))=\sum_{\begin{subarray}{c}0\leq n^{\prime}\leq n\\ 2|n-n^{\prime}\\ h(n^{\prime})=h(\alpha)\end{subarray}}R_{n,n^{\prime},w-p}\tilde{P}(K^{w}(n^{\prime}))+\cdots (27)

where the highest nontrivial tt-degree in each summand is 12h(α)\tfrac{1}{2}h(\alpha), and \cdots denotes terms with lower tt-degrees. Moreover, if P~Lee(Kp(n))(t,q)\tilde{P}_{Lee}(K^{p}(n))(t,q) denotes the corresponding renormalized Poincaré polynomial for gr(KhRLee(Kp(n)))gr(KhR_{Lee}(K^{p}(n))), then the highest tt-degree parts of P~(Kp(n))\tilde{P}(K^{p}(n)) and P~Lee(Kp(n))\tilde{P}_{Lee}(K^{p}(n)) are equal.

In the comparison lemma 5.8, the highest tt-degree of the nn^{\prime} term on the right hand side is at most

12nT(PW)n+12nTNn=12nT(PW+N)n\tfrac{1}{2}n^{\prime T}(P-W)n^{\prime}+\tfrac{1}{2}n^{\prime T}Nn^{\prime}=\tfrac{1}{2}n^{\prime T}(P-W+N)n^{\prime}

by Lemma 5.9, which is strictly less than h(α)h(\alpha) if h(n)h(α)h(n^{\prime})\neq h(\alpha) by assumption (i). On the other hand, if h(n)=h(α)h(n^{\prime})=h(\alpha), then by assumption (ii) Kw(n)K^{w}(n^{\prime}) has a positive diagram induced by DD with some number s(n)s(n^{\prime}) of connected components (which only depends on the indices of the nonzero coordinates of nn^{\prime} as DD has no 0-framed unknot component). Then KhR20(Kw(n))KhR_{2}^{0}(K^{w}(n^{\prime})) (as well as KhRLee0(Kw(n))KhR_{Lee}^{0}(K^{w}(n^{\prime}))) has rank 2s(n)2^{s(n^{\prime})} [khovanov2003patterns, Proposition 6.1]. After renormalizing, we see that Rn,n,wpP~(Kw(n))R_{n,n^{\prime},w-p}\tilde{P}(K^{w}(n^{\prime})) has maximal tt-degree 12h(α)\tfrac{1}{2}h(\alpha), and in this tt-degree the polynomial has rank (i.e. the evaluation at q=1q=1)

2s(n)j=1m(rj=0njnj2dim(nj+nj2+rj,njnj2rj))=2s(n)(nnn2)2^{s(n^{\prime})}\prod_{j=1}^{m}\left(\sum_{r_{j}=0}^{\frac{n_{j}-n^{\prime}_{j}}{2}}\dim\left(\frac{n_{j}+n^{\prime}_{j}}{2}+r_{j},\frac{n_{j}-n^{\prime}_{j}}{2}-r_{j}\right)\right)=2^{s(n^{\prime})}\binom{n}{\frac{n-n^{\prime}}{2}}

where for a,b0ma,b\in\mathbb{Z}_{\geq 0}^{m} we write (ab)=i(aibi)\binom{a}{b}=\prod_{i}\binom{a_{i}}{b_{i}}. Now it follows from Lemma 5.8 that the highest tt-degree of P~(Kp(n))\tilde{P}(K^{p}(n)) is at most 12h(α)\tfrac{1}{2}h(\alpha), and in this tt-degree it has rank at most

0nn2|nnh(n)=h(α)2s(n)(nnn2).\sum_{\begin{subarray}{c}0\leq n^{\prime}\leq n\\ 2|n-n^{\prime}\\ h(n^{\prime})=h(\alpha)\end{subarray}}2^{s(n^{\prime})}\binom{n}{\frac{n-n^{\prime}}{2}}.

On the other hand, the rank of P~Lee(Kp(n))\tilde{P}_{Lee}(K^{p}(n)) at t=12h(α)t=\tfrac{1}{2}h(\alpha) is equal to the number of orientations of Kp(n)K^{p}(n) with writhe h(α)h(\alpha). Dividing the count into algebraic cables of KpK^{p}, the rank is again

nnn2|nnnT(PW+N+N)n=αT(PW+N)α(nnn2)=nnn2|nnh(n)=h(α)(nnn2)=0nn2|nnh(n)=h(α)2s(n)(nnn2),\displaystyle\sum_{\begin{subarray}{c}-n\leq n^{\prime}\leq n\\ 2|n-n^{\prime}\\ n^{\prime T}(P-W+N_{+}-N_{-})n^{\prime}\\ =\alpha^{T}(P-W+N)\alpha\end{subarray}}\binom{n}{\frac{n-n^{\prime}}{2}}=\sum_{\begin{subarray}{c}-n\leq n^{\prime}\leq n\\ 2|n-n^{\prime}\\ h(n^{\prime})=h(\alpha)\end{subarray}}\binom{n}{\frac{n-n^{\prime}}{2}}=\sum_{\begin{subarray}{c}0\leq n^{\prime}\leq n\\ 2|n-n^{\prime}\\ h(n^{\prime})=h(\alpha)\end{subarray}}2^{s(n^{\prime})}\binom{n}{\frac{n-n^{\prime}}{2}},

where the first equality follows from conditions (i) and (ii). The second equality is true because in view of (ii), for n0n^{\prime}\geq 0, changing signs of coordinates of nn^{\prime} maintains h(n)h(n^{\prime}) if and only if, for each connected component of the diagram of Kw(n)K^{w}(n^{\prime}), we either change all of the corresponding coordinates of nn^{\prime} or none of them. Since P~(Kp(n))\tilde{P}(K^{p}(n)) is bounded below by P~Lee(Kp(n))\tilde{P}_{Lee}(K^{p}(n)) as there is a spectral sequence from KhR2KhR_{2} to KhRLeeKhR_{Lee}, we must have the equality (27) as well as its version for P~Lee(Kp(n))\tilde{P}_{Lee}(K^{p}(n)) as claimed.

Step 2: Unrenormalization and proofs of (1) and (2).

For n=α+2rn=\alpha+2r, condition (ii) ensures that the writhe of Kw(n)K^{w}(n^{\prime}) is same as nTNnn^{\prime T}Nn^{\prime}, allowing us to unrenormalize (27) to give

P(Kp(α+r,r))=0nn2|nnh(n)=h(α)(tq1)12nT(WP)nRn,n,wpP(Kw(n))+,P(K^{p}(\alpha+r,r))=\sum_{\begin{subarray}{c}0\leq n^{\prime}\leq n\\ 2|n-n^{\prime}\\ h(n^{\prime})=h(\alpha)\end{subarray}}(tq^{-1})^{\tfrac{1}{2}n^{\prime T}(W-P)n^{\prime}}R_{n,n^{\prime},w-p}P(K^{w}(n^{\prime}))+\cdots,

where the maximal tt-degree of each term on the right hand side is 0. The same holds for the Lee version. In view of (26), this proves (1).

For (2), by Theorem 4.1(1)(2), 𝒮0,0Lee(X;α)\mathcal{S}_{0,0}^{Lee}(X;\alpha) has a basis consists of xα+,αx_{\alpha_{+},\alpha_{-}} with α±H2(X)\alpha_{\pm}\in H_{2}(X), α++α=α\alpha_{+}+\alpha_{-}=\alpha, α+α=0\alpha_{+}\cdot\alpha_{-}=0. Such generators are in one-one-correspondence with αα+2m\alpha^{\prime}\in\alpha+2\mathbb{Z}^{m} with α2=α2\alpha^{\prime 2}=\alpha^{2}, where the correspondence is given by α:=α+α\alpha^{\prime}:=\alpha_{+}-\alpha_{-}. The bilinear form on H2(X)H_{2}(X) can be rewritten as β2=βT(PW+N2N)β\beta^{2}=\beta^{T}(P-W+N-2N_{-})\beta. Thus by conditions (i) and (ii), we see that such α\alpha^{\prime} are exactly those with h(α)=h(α)h(\alpha^{\prime})=h(\alpha), proving the statement for 𝒮0,0Lee\mathcal{S}_{0,0}^{Lee}.

Write (26) in homological degree 0 for =Lee\bullet=Lee as 𝒮0,0Lee(X;α)=colimrVr\mathcal{S}_{0,0}^{Lee}(X;\alpha)=\mathrm{colim}_{r}V_{r} for short. Then by Step 1 and the naturality of the Lee spectral sequence, we know 𝒮0,02(X;α;)=colimrgr(Vr)\mathcal{S}_{0,0}^{2}(X;\alpha;\mathbb{Q})=\mathrm{colim}_{r}gr(V_{r}), which equals the subspace of gr(colimrVr)gr(\mathrm{colim}_{r}V_{r}) of finite degree elements together with 0. In particular, the rank of 𝒮0,02(X;α)\mathcal{S}_{0,0}^{2}(X;\alpha) is bounded above by the dimension of 𝒮0,0Lee(X;α)\mathcal{S}_{0,0}^{Lee}(X;\alpha).

Step 3: Decomposition of KhRLeeKhR_{Lee} and proof of (3).

As we have already seen once in Step 1 (with n=α+2rn=\alpha+2r, and renormalization), KhRLee0(Kp(α+r,r))KhR_{Lee}^{0}(K^{p}(\alpha+r,r)) is generated by canonical generators coming from orientations of Kp(α+r,r)K^{p}(\alpha+r,r) with writhe h(α)h(\alpha), and we can divide these orientations according to the their algebraic cable type.

More precisely, for nnn-n\leq n^{\prime}\leq n with 2|nn2|n-n^{\prime}, h(n)=h(α)h(n^{\prime})=h(\alpha), let

KhRLee,n0(Kp(α+r,r))KhRLee0(Kp(α+r,r))KhR_{Lee,n^{\prime}}^{0}(K^{p}(\alpha+r,r))\subset KhR_{Lee}^{0}(K^{p}(\alpha+r,r))

be the subspace spanned by generators whose orientations make Kp(n)K^{p}(n) algebraic nn^{\prime}-cables of KK. This subspace is an SnS_{n}-subrepresentation.

We will be interested in n=αn^{\prime}=\alpha. As in [ren2023lee, Proposition 4.2], we know

KhRLee,α0(Kp(α+r,r))0sr(ns,s)KhR_{Lee,\alpha}^{0}(K^{p}(\alpha+r,r))\cong\bigoplus_{0\leq s\leq r}(n-s,s) (28)

as SnS_{n}-representations, where for a,b0ma,b\in\mathbb{Z}_{\geq 0}^{m}, aba\geq b, (a,b)(a,b) denotes the Sa+bS_{a+b}-representation given by the tensor product of the Sai+biS_{a_{i}+b_{i}}-representations (ai,bi)(a_{i},b_{i}). Similarly,

KhRLee,α0(Kp(α+r+ei,r+ei))0sr+ei(n+2eis,s)KhR_{Lee,\alpha}^{0}(K^{p}(\alpha+r+e_{i},r+e_{i}))\cong\bigoplus_{0\leq s\leq r+e_{i}}(n+2e_{i}-s,s) (29)

as Sn+2eiS_{n+2e_{i}}-representations.

By the argument in [ren2023lee, Section 4.3], for every 0sr0\leq s\leq r, the subrepresentation (ns,s)(n-s,s) in (28) and the subrepresentation (ns+ei,s+ei)(n-s+e_{i},s+e_{i}) in (29) have the same filtration degree. On the other hand, the dotted annular creation cobordism map restricts to a filtered degree 22 map

ϕi:KhRLee,α0(Kp(α+r,r))KhRLee,α0(Kp(α+r+ei,r+ei)).\phi_{i}\colon KhR_{Lee,\alpha}^{0}(K^{p}(\alpha+r,r))\to KhR_{Lee,\alpha}^{0}(K^{p}(\alpha+r+e_{i},r+e_{i})).

By Step 1, the associated graded map of ϕi\phi_{i} is exactly the corresponding dotted annular creation map on KhR2KhR_{2}, which is injective by the conclusions of Step 1 and of Lemma 5.8. Thus ϕi\phi_{i} increases the filtration degree of every nonzero element by exactly 22. Inductively from r=0r=0 (where KhRLee,α0(Kp(α))=(α,0)KhR_{Lee,\alpha}^{0}(K^{p}(\alpha))=(\alpha,0) in filtration degree s𝔤𝔩2(Kp(α))+1s_{\mathfrak{gl}_{2}}(K^{p}(\alpha))+1 by definition), we see that the filtration degree of (n,0)(n,0) in (28) is s𝔤𝔩2(Kp(α))+1+2|r|s_{\mathfrak{gl}_{2}}(K^{p}(\alpha))+1+2|r|. Consequently by (13) (cf. (23)) we have

s(X;α)=limrm(s𝔤𝔩2(Kp(α))+1+2|r||α|2|r|)=s(Kp(α))w(Kp(α))|α|+1>.s(X;\alpha)=\lim_{r\to\infty^{m}}(s_{\mathfrak{gl}_{2}}(K^{p}(\alpha))+1+2|r|-|\alpha|-2|r|)=-s(-K^{p}(\alpha))-w(K^{p}(\alpha))-|\alpha|+1>-\infty.

In particular, 𝒮0,0,s(X;α)2(X;α)0\mathcal{S}_{0,0,s(X;\alpha)}^{2}(X;\alpha)\neq 0 by Theorem 1.12. ∎

Remark 5.10.
  1. (1)

    We expect that in the notation of Theorem 5.5, assuming no component of DD is a 0-framed unknot, then

    𝒮0,02(X;α)=α0m2|ααh(α)=h(α)KhR20(Kw(α)){i(wipi)(αi2+2αi)2w(Kw(α))2|α|}.\quad\quad\quad\mathcal{S}_{0,0}^{2}(X;\alpha)=\bigoplus_{\begin{subarray}{c}\alpha^{\prime}\in\mathbb{Z}_{\geq 0}^{m}\\ 2|\alpha-\alpha^{\prime}\\ h(\alpha^{\prime})=h(\alpha)\end{subarray}}KhR_{2}^{0}(K^{w}(\alpha^{\prime}))\left\{\sum_{i}\frac{(w_{i}-p_{i})(\alpha_{i}^{\prime 2}+2\alpha_{i}^{\prime})}{2}-\frac{w(K^{w}(\alpha^{\prime}))}{2}-|\alpha^{\prime}|\right\}. (30)

    In fact, an expanded argument of the proof above shows that the graded rank of the left hand side is bounded above by that of the right hand side. However, we have not been able to work out a proof of (30). We remark that over \mathbb{Q}, (30) is equivalent to adding to the last claim of Lemma 5.8 the conclusion that the symmetrized dotted annular creation map gives an injection

    KhR2hmax(Kp(n2ei);)Sn2eiKhR2hmax(Kp(n);)Sn.KhR_{2}^{h_{max}}(K^{p}(n-2e_{i});\mathbb{Q})^{S_{n-2e_{i}}}\hookrightarrow KhR_{2}^{h_{max}}(K^{p}(n);\mathbb{Q})^{S_{n}}.

    where hmaxh_{max} is the maximal nontrivial homological degree of the right hand side.

  2. (2)

    In particular, in the setup of Theorem 1.14(1) or (2) when n<0n<0, one can easily show the validity of (30). Therefore, the conclusions in Theorem 1.14 for 𝒮0,02\mathcal{S}_{0,0}^{2} can be stated over \mathbb{Z} instead.

6. Nonvanishing examples and applications

In this section we present some explicit calculations and applications for Khovanov and Lee skein lasagna modules and lasagna ss-invariants of some 22-handlebodies (with empty boundary link). Roughly the first half of this section utilizes the tools developed in Section 5, while the second half utilizes results in a paper of the first author [ren2023lee].

6.1. Knot traces of some small knots

Knot Framing 4-4 3-3 2-2 1-1 0 11 22 33 44 55 66
UU N N N N N V V V V V V
313_{1} N ? ? V V V V V V V V
31-3_{1} N N N N N ? ? ? ? ?V V
414_{1} N N N N N ?V ?V V V V V
Table 1. Vanishing/Nonvanishing of Khovanov skein lasagna modules of some knot traces. An entry N/V means the corresponding knot trace has nonvanishing/vanishing Khovanov skein lasagna module. An entry ?V means it has vanishing Lee skein lasagna module (meaning the filtration function is identically -\infty).
Example 6.1 (Unknot).

By Theorem 1.14, for n0n\leq 0, the D2D^{2}-bundle D(n)D(n) over S2S^{2} with Euler number nn has s(D(n);0)=0s(D(n);0)=0, s(D(n);1)=ns(D(n);1)=-n. In particular, 𝒮0,02(D(n);α)0\mathcal{S}_{0,0}^{2}(D(n);\alpha)\neq 0 for α=0,1\alpha=0,1.

For n<0n<0, more explicitly we have 𝒮0,02(D(n);0)=\mathcal{S}_{0,0}^{2}(D(n);0)=\mathbb{Z} concentrating in degree 0 and 𝒮0,02(D(n);1)=2\mathcal{S}_{0,0}^{2}(D(n);1)=\mathbb{Z}^{2} concentrating in degrees n,n2n,n-2 (cf. Remark 5.10(2)).

On the other hand, for n>0n>0, Example 3.2 shows 𝒮02(D(n))=0\mathcal{S}_{0}^{2}(D(n))=0 and consequently all lasagna ss-invariants of D(n)D(n) are -\infty.

For other examples we only mention the vanishingness/nonvanishingness of the Khovanov skein lasagna modules of the knot traces. For small knots, Corollary 1.16 (which will be proved in Section LABEL:sbsec:-CP^2) usually provides more effective nonvanishing bounds than Theorem 1.14. We emphasize however that Corollary 1.16 does not provide an explicit ss-invariant in most cases.

Example 6.2 (Trefoils).

The left-handed trefoil, or 313_{1}, has 33 negative crossings. Therefore 1.14(1) shows Xn(31)X_{n}(3_{1}) has nonvanishing Khovanov skein lasagna module for n<6n<-6.

We can do better as follows. The knot 313_{1} can be unknotted by changing a crossing, which is realized by performing a +1+1 twist on two parallelly oriented strands. This means 313_{1} is (4)(-4)-slice in 2\mathbb{CP}^{2}, thus it is nn-slice in some k2k\mathbb{CP}^{2} for every n4n\leq-4. It follows that Xn(31)X_{n}(3_{1}) embeds in some k2¯k\overline{\mathbb{CP}^{2}} for n4n\leq-4. By Corollary 1.16, Xn(31)X_{n}(3_{1}) has nonvanishing Khovanov skein lasagna module for n4n\leq-4.

On the other hand, TB(31)=1TB(-3_{1})=1, so Xn(31)X_{n}(3_{1}) has vanishing Khovanov skein lasagna module for n1n\geq-1 by Theorem 1.4. It would be interesting to know whether Xn(31)X_{n}(3_{1}) has nonvanishing Khovanov skein lasagna module for n=3,2n=-3,-2.

The right-handed trefoil 31-3_{1} is positive. By either Theorem 1.14 or Corollary 1.16 we know Xn(31)X_{n}(-3_{1}) has nonvanishing Khovanov skein lasagna module for n0n\leq 0. On the other hand, since TB(31)=6TB(3_{1})=-6, Xn(31)X_{n}(-3_{1}) has vanishing Khovanov skein lasagna module for n6n\geq 6. In this case we have a larger gap of unknown. We remark that Proposition 6.8 will imply that X5(31)X_{5}(-3_{1}) has vanishing Lee skein lasagna module, as there is a framed concordance in 2¯\overline{\mathbb{CP}^{2}} from the 11-framed unknot to the 55-framed 31-3_{1}.

Example 6.3 (Figure 88).

The standard diagram of the figure 88 knot, or 414_{1}, has 22 negative crossings. Theorem 1.14(1) shows Xn(41)X_{n}(4_{1}) has nonvanishing Khovanov skein lasagna module for n<4n<-4, while Corollary 1.16 shows so for n0n\leq 0 as 414_{1} is 0-slice in 2\mathbb{CP}^{2}. On the other hand, since 41=41-4_{1}=4_{1} and TB(41)=3TB(4_{1})=-3, we know Xn(41)X_{n}(4_{1}) has vanishing Khovanov skein lasagna module for n3n\geq 3 by Theorem 1.4.

It would be interesting to know whether Xn(41)X_{n}(4_{1}) has nonvanishing Khovanov skein lasagna module for n=1,2n=1,2. Again, Proposition 6.8 will imply they both have vanishing Lee skein lasagna modules.

6.2. Nonpositive definite plumbings and generalizations

Example 6.4 (Nonpositive plumbing tree).

Let XX be a nonpositive definite plumbing tree of spheres, i.e. a plumbing of D2D^{2}-bundles over 22-spheres along a tree, whose intersection form is nonpositive definite. Then we can orient the unknots in the underlying Kirby diagram so that all crossings are positive. In particular, in the notation of Theorem 5.5, PW+N=PW+N+NP-W+N=P-W+N_{+}-N_{-} is the intersection form of XX, which is nonpositive. Therefore Theorem 5.5(3) implies s(X;0)=0s(X;0)=0 and XX has nonvanishing Khovanov skein lasagna module.

Example 6.5 (Traces on alternating chainmail links).

Let GG be a chainmail graph without loops and with positive edge weights and nonpositive vertex weights, and KK be its associated chainmail link, in the sense of [agol2023chainmail, Section 2]. Then KK has a standard positive diagram (as an unframed link), and a nonpositive framing matrix which is equal to PW+NP-W+N for its standard diagram. This means the trace XX on KK has nonvanishing Khovanov skein lasagna module. This generalizes the previous example.

Example 6.6 (Branched double cover of definite surfaces of alternating links).

Suppose LS3L\subset S^{3} is a nonsplit alternating link and SB4S\subset B^{4} is an unoriented surface bounding LL with negative definite Gordon–Litherland pairing [gordon1978signature]. Greene [greene2017alternating] showed every choice of SS comes from one checkerboard surface of an alternating diagram of LL (and conversely one of the checkerboard surface of every alternating diagram is negative definite, which is much easier).

The branched double cover of SS is a 44-manifold Σ(S)\Sigma(S) bounding the branched double cover Σ(L)\Sigma(L) of LL. It is also considered by Ozsváth-Szabó [ozsvath2005heegaard] (denoted XLX_{L} in Section 3 there), and from the description there (see also [greene2013spanning, Section 3.1]) we realize that Σ(S)\Sigma(S) is a special case of Example 6.5, hence it has nonvanishing Khovanov skein lasagna module. In particular, Σ(L)\Sigma(L) bounds a 22-handlebody with nonvanishing Khovanov skein lasagna module. Since Σ(L)¯=Σ(L)\overline{\Sigma(L)}=\Sigma(-L) we see Σ(L)\Sigma(L) bounds such 22-handlebodies on both sides.

Example 6.7.

One can further generalize Example 6.5 by connect-summing some components of KK with 0-framed positive knots. The same argument shows the resulting trace has nonvanishing Khovanov skein lasagna module.

6.3. More exotica from connected sums

In this section we prove Corollary 1.2. One can show the relevant pairs of manifolds have nonisomorphic Khovanov skein lasagna modules, but we give a simpler argument by showing they have different lasagna ss-invariants.

Proof of Corollary 1.2.

Write W1:=X1aX2bW_{1}:=X_{1}^{\natural a}\natural X_{2}^{\natural b} and W2:=X1aX2bW_{2}:=X_{1}^{\natural a^{\prime}}\natural X_{2}^{\natural b^{\prime}}, where a+b=a+b=na+b=a^{\prime}+b^{\prime}=n and (a,b)(a,b)(a,b)\neq(a^{\prime},b^{\prime}).

Recall from Section 1.5 that s(X1;1)=3s(X_{1};1)=3, s(X2;1)=1s(X_{2};1)=1, where 11 is a generator of H2(Xi)H_{2}(X_{i}). Since we did not specify the sign of 11, we equally have s(X1;1)=3s(X_{1};-1)=3, s(X2;1)=1s(X_{2};-1)=1. By the connected sum formula (Theorem 4.7(4)),

s(W1;(ϵ1,,ϵn))=3a+bs(W_{1};(\epsilon_{1},\cdots,\epsilon_{n}))=3a+b

for any ϵi=±1\epsilon_{i}=\pm 1. If there were a diffeomorphism W1W2W_{1}\cong W_{2}, then by considering the intersection form we see it sends each class (ϵ1,,ϵn)(\epsilon_{1},\cdots,\epsilon_{n}) to some class (ϵ1,,ϵn)(\epsilon^{\prime}_{1},\cdots,\epsilon^{\prime}_{n}), ϵi=±1\epsilon_{i}^{\prime}=\pm 1. However, s(W2;(ϵ1,,ϵn))=3a+b3a+bs(W_{2};(\epsilon^{\prime}_{1},\cdots,\epsilon^{\prime}_{n}))=3a^{\prime}+b^{\prime}\neq 3a+b, a contradiction.

By (a Lee version of) [manolescu2023skein, Corollary 4.2] we know s(S1×S3;0)=0s(S^{1}\times S^{3};0)=0. Since H2(S1×S3)=0H_{2}(S^{1}\times S^{3})=0 the above argument remains valid under connected sums with S1×S3S^{1}\times S^{3}. Since s(2¯;1)=1s(\overline{\mathbb{CP}^{2}};1)=1 by Example 6.1 for n=1n=-1 and Proposition 2.1, the argument remains valid under further connected sums with 2¯\overline{\mathbb{CP}^{2}}.

By a direct calculation (or Theorem 5.5 applied to the 0-framed unknot) we know s(S2×D2;α)=0s(S^{2}\times D^{2};\alpha)=0 for all α\alpha. Any diffeomorphism k(S1×S3)#m2¯#W1(S2×D2)ck(S1×S3)#m2¯#W2(S2×D2)ck(S^{1}\times S^{3})\#m\overline{\mathbb{CP}^{2}}\#W_{1}\natural(S^{2}\times D^{2})^{\natural c}\cong k(S^{1}\times S^{3})\#m\overline{\mathbb{CP}^{2}}\#W_{2}\natural(S^{2}\times D^{2})^{\natural c} sends each class (ϵ1,,ϵm+n,0c)(\epsilon_{1},\cdots,\epsilon_{m+n},0^{c}) to some class (ϵ1,,ϵm+n,α1,,αc)(\epsilon_{1}^{\prime},\cdots,\epsilon_{m+n}^{\prime},\alpha_{1},\cdots,\alpha_{c}). Again, the ss-invariant of the latter class is not equal to the ss-invariant of the former class, a contradiction.

By Example 6.4 we know the negative E8E8 manifold has s(E8;0)=0s(E8;0)=0. Any diffeomorphism between the above pair with dd further copies of E8E8 added sends each (ϵ1,,ϵm+n,0c,0d)(\epsilon_{1},\cdots,\epsilon_{m+n},0^{c},0^{d}) to (ϵ1,,ϵm+n,α1,,αc,0d)(\epsilon_{1}^{\prime},\cdots,\epsilon_{m+n}^{\prime},\alpha_{1},\cdots,\alpha_{c},0^{d}), whose ss-invariants still differ, leading to a contradiction.

Finally, since the lasagna ss-invariants of a 44-manifold XX are invariants of the interior of XX (cf. the end of Section 2.1), the interiors of any of the above pairs are not diffeomorphic. ∎

6.4. Nonpositive shake genus of positive knots

We prove Theorem 1.9(1).

Proof of Theorem 1.9(1).

By Theorem 1.14(2), if KK^{\prime} is a positive knot and n0n\leq 0, then s(Xn(K);1)=s(K)ns(X_{n}(K^{\prime});1)=s(K^{\prime})-n. By Proposition 1.13, if KK is a knot concordant to KK^{\prime}, then s(Xn(K);1)=s(Xn(K);1)s(X_{n}(K);1)=s(X_{n}(K^{\prime});1). Thus Equation (1) implies gshn(K)s(K)/2g_{sh}^{n}(K)\geq s(K^{\prime})/2. On the other hand, Rasmussen’s classical bound gives s(K)/2g4(K)s(K^{\prime})/2\geq g_{4}(K^{\prime}). Since g4(K)=g4(K)gshn(K)g_{4}(K^{\prime})=g_{4}(K)\geq g_{sh}^{n}(K), all these quantities are equal. ∎

6.5. Another comparison result for lasagna ss-invariants

In this section we make use the adjunction inequality for ss-invariants [ren2023lee, Corollary 1.4] to deduce a comparison result for lasagna ss-invariants in the spirit of Proposition 5.3. Then we prove Theorem 1.9(2). For simplicity, assume the boundary link LL is empty. Instead of working on the 44-manifold level as in Section 5.2 (which is more invariant), we work on the cabled links level directly.

As in the setup of Section 5.2, assume XX is a 22-handlebody obtained by attaching 22-handles to a framed link K=K1KmS3=B4K=K_{1}\cup\cdots K_{m}\subset S^{3}=\partial B^{4}, and XX^{\prime} is one obtained by attaching handles to K=K1KmK^{\prime}=K_{1}^{\prime}\cup\cdots\cup K_{m}^{\prime}.

Proposition 6.8.

Suppose C:KKC\colon K\to K^{\prime} is a framed concordance in k2¯k\overline{\mathbb{CP}^{2}} that restricts to concordances Ci:KiKiC_{i}\colon K_{i}\to K_{i}^{\prime}. Then

s(X;α)s(X;α)|[C(α)]|s(X^{\prime};\alpha)\leq s(X;\alpha)-|[C(\alpha)]|

for any αH2(X)H2(X)m\alpha\in H_{2}(X)\cong H_{2}(X^{\prime})\cong\mathbb{Z}^{m}, where [C(α)]=iαi[Ci]H2(k2¯)k[C(\alpha)]=\sum_{i}\alpha_{i}[C_{i}]\in H_{2}(k\overline{\mathbb{CP}^{2}})\cong\mathbb{Z}^{k}.

Proof.

The adjunction inequality for the ss-invariant [ren2023lee, Corollary 1.4] in the 𝔤𝔩2\mathfrak{gl}_{2} renormalization states that if Σ:LL\Sigma\colon L\to L^{\prime} is a framed cobordism in k2¯k\overline{\mathbb{CP}^{2}}, such that each component of Σ\Sigma has a boundary on LL, then

s𝔤𝔩2(L)s𝔤𝔩2(L)χ(Σ)|[Σ]|.s_{\mathfrak{gl}_{2}}(L^{\prime})\leq s_{\mathfrak{gl}_{2}}(L)-\chi(\Sigma)-|[\Sigma]|. (31)

Now, the (α++r,α+r)(\alpha^{+}+r,\alpha^{-}+r)-cable of CC gives a concordance between K(α++r,α+r)K(\alpha^{+}+r,\alpha^{-}+r) and K(α++r,α+r)K^{\prime}(\alpha^{+}+r,\alpha^{-}+r) with homology class [C(α)][C(\alpha)], which by (31) implies

s𝔤𝔩2(K(α++r,α+r))s𝔤𝔩2(K(α++r,α+r))|[C(α)]|.s_{\mathfrak{gl}_{2}}(K^{\prime}(\alpha^{+}+r,\alpha^{-}+r))\leq s_{\mathfrak{gl}_{2}}(K(\alpha^{+}+r,\alpha^{-}+r))-|[C(\alpha)]|.

The statement now follows from Proposition 5.4. ∎

Now we prove Theorem 1.9(2).

Proof of Theorem 1.9(2).

Putting a framing on Σ\Sigma makes it a framed cobordism from (K,n)(K,n) to (K,n[Σ]2)(K^{\prime},n-[\Sigma]^{2}). It follows from Proposition 6.8 that s(Xn(K);1)s(Xn[Σ]2(K);1)+|[Σ]|s(X_{n}(K);1)\geq s(X_{n-[\Sigma]^{2}}(K^{\prime});1)+|[\Sigma]|. If n[Σ]2n\leq[\Sigma]^{2}, by Theorem 1.14 we know s(Xn[Σ]2(K);1)=s(K)n+[Σ]2s(X_{n-[\Sigma]^{2}}(K^{\prime});1)=s(K^{\prime})-n+[\Sigma]^{2}, thus the statement follows. ∎

6.6. Yasui’s family of knot trace pairs

Yasui [yasui2015corks, Figure 10] defined two families of satellite patterns Pn,mP_{n,m}, Qn,mQ_{n,m}, n,mn,m\in\mathbb{Z} (see Figure 2). He showed that for any n,mn,m and any knot KS3K\subset S^{3}, the two manifolds Xn(Pn,m(K))X_{n}(P_{n,m}(K)) and Xn(Qn,m(K))X_{n}(Q_{n,m}(K)), namely the nn-traces on the two corresponding satellite knots, are homeomorphic. Moreover, he showed that if there exists a Legendrian representative 𝒦\mathcal{K} of KK with

tb(𝒦)+|rot(𝒦)|1=2g4(K)2,ntb(𝒦),m0,tb(\mathcal{K})+|rot(\mathcal{K})|-1=2g_{4}(K)-2,\ n\leq tb(\mathcal{K}),\ m\geq 0, (32)

then Xn(Pn,m(K))X_{n}(P_{n,m}(K)) and Xn(Qn,m(K))X_{n}(Q_{n,m}(K)) are not diffeomorphic, thus form an exotic pair [yasui2015corks, Theorem 4.1]. In fact, under these conditions, the exotica were detected by a discrepancy of shake slice genera: gshn(Pn,m(K))=g4(K)+1>gshn(Qn,m(K))g_{sh}^{n}(P_{n,m}(K))=g_{4}(K)+1>g_{sh}^{n}(Q_{n,m}(K)). Thus, the interiors of such pairs are also exotic. The case K=UK=U, n=1n=-1, m=0m=0 yields the exotic pair X1(52)X_{-1}(-5_{2}) and X1(P(3,3,8))X_{-1}(P(3,-3,-8)) in Theorem 1.1.

m-mnn     m-m2-2nn

Figure 2. Satellite patterns Pn,mP_{n,m} (left) and Qn,mQ_{n,m} (right). The boxes denote full twists. The shaded band gives rise to a concordance between Qn,mQ_{n,m} and the identity pattern.

Using lasagna ss-invariants, we present a different family (as stated in Theorem 1.3) of Yasui’s knot trace pairs that are exotic. We recall that our hypothesis on (K,n,m)(K,n,m) is that there exists a slice disk Σ\Sigma of KK in some k2k\mathbb{CP}^{2} such that

s(K)=|[Σ]|[Σ]2,n<[Σ]2,m0.s(K)=|[\Sigma]|-[\Sigma]^{2},\ n<-[\Sigma]^{2},\ m\geq 0. (33)

Before giving the proof, we compare and discuss the conditions imposed on KK by Yasui (32) and by us (33).

The adjunction inequality for the ss-invariant [ren2023lee, Corollary 1.5] applied to the slice disk Σ\Sigma implies that s(K)|[Σ]|[Σ]2s(K)\geq|[\Sigma]|-[\Sigma]^{2}. Therefore the first condition in (33) is that this adjunction inequality attains equality for Σ\Sigma. Similarly, the first condition in (32) is that the slice-Bennequin inequality attains equality for 𝒦\mathcal{K}. These two conditions are satisfied on somewhat different families of knots. For example, as remarked by Yasui, his condition on KK is satisfied for all positive torus knots. On the other hand, our condition on KK is satisfied for all negative torus knots (cf. Lemma 6.10). Our condition also has the advantage of being closed under concordances and connected sums of knots.

Proof of Theorem 1.3.

As observed by Yasui, adding a band as indicated in Figure 2 and capping off the unknot component give a concordance between Qn,mQ_{n,m} and the identity satellite pattern. Therefore, Qn,m(K)Q_{n,m}(K) is concordant to KK, thus by Proposition 1.13 and Proposition 5.4 we have

s(Xn(Qn,m(K));1)=s(Xn(K);1)\displaystyle\,s(X_{n}(Q_{n,m}(K));1)=s(X_{n}(K);1)
=\displaystyle= limr(s(K(r+1,r))+2r)n1+1s(K(1,0))n=s(K)n.\displaystyle-\lim_{r\to\infty}(s(-K(r+1,r))+2r)-n-1+1\leq-s(-K(1,0))-n=s(K)-n. (34)

Deleting a local ball in B4B^{4} near Σ\Sigma, turning the cobordism upside down and reversing the ambient orientation, we obtain a concordance Σ:KU-\Sigma^{\circ}\colon K\to U in k2¯k\overline{\mathbb{CP}^{2}}. Applying the satellite operation Pn,mP_{n,m} gives a framed concordance C1=Pn,m(Σ):(Pn,m(K),n)(Pn+[Σ]2,m(U),n+[Σ]2)C_{1}=P_{n,m}(-\Sigma^{\circ})\colon(P_{n,m}(K),n)\to(P_{n+[\Sigma]^{2},m}(U),n+[\Sigma]^{2}) in k2¯k\overline{\mathbb{CP}^{2}} with [C1]=[Σ][C_{1}]=[-\Sigma^{\circ}] since Pn,mP_{n,m} has winding number 11. Since n<[Σ]2n<-[\Sigma]^{2}, there is a framed concordance C2:(Pn+[Σ]2,m(U),n+[Σ]2)(P1,m(U),1)C_{2}\colon(P_{n+[\Sigma]^{2},m}(U),n+[\Sigma]^{2})\to(P_{-1,m}(U),-1) in k12¯k_{1}\overline{\mathbb{CP}^{2}} obtained by adding k1k_{1} positive twists along the meridian of the satellite solid torus, which has [C2]=(1,,1)k1=H2(k12¯)[C_{2}]=(1,\cdots,1)\in\mathbb{Z}^{k_{1}}=H_{2}(k_{1}\overline{\mathbb{CP}^{2}}), where k1=1n[Σ]2k_{1}=-1-n-[\Sigma]^{2}. Similarly, there is a null-homologous framed concordance C3:(P1,m(U),1)(P1,0(U),1)=(52,1)C_{3}\colon(P_{-1,m}(U),-1)\to(P_{-1,0}(U),-1)=(-5_{2},-1) in k22¯k_{2}\overline{\mathbb{CP}^{2}} where k2=mk_{2}=m (note that the two strands being twisted m-m times in Pn,mP_{n,m} are oppositely oriented). Now Proposition 6.8 applied to the concordance C1C2C3:(Pn,m(K),n)(52,1)C_{1}\cup C_{2}\cup C_{3}\colon(P_{n,m}(K),n)\to(-5_{2},-1) yields

s(Xn(Pn,m(K));1)s(X1(52);1)+|[C1]|+|[C2]|+|[C3]|\displaystyle s(X_{n}(P_{n,m}(K));1)\geq s(X_{-1}(-5_{2});1)+|[C_{1}]|+|[C_{2}]|+|[C_{3}]|
=\displaystyle=  3+|[Σ]|+k1+0=|[Σ]|[Σ]2n+2=s(K)n+2.\displaystyle\,3+|[\Sigma]|+k_{1}+0=|[\Sigma]|-[\Sigma]^{2}-n+2=s(K)-n+2. (35)

Together, (6.6) and (6.6) imply that s(Xn(Pn,m(K));1)>s(Xn(Qn,m(K));1)s(X_{n}(P_{n,m}(K));1)>s(X_{n}(Q_{n,m}(K));1), proving the theorem. ∎

Remark 6.9.

Let CKC\cup K be a two-component link in S3S^{3}, where CC is an unknot and k(C,K)=1\ell k(C,K)=1. Regarding CC as a dotted 11-handle and attaching an nn-framed 22-handle along Pn,m(K)P_{n,m}(K) (resp. Qn,m(K)Q_{n,m}(K)) yields a Mazur manifold ZPZ_{P} (resp. ZQZ_{Q}). The manifolds ZPZ_{P} and ZQZ_{Q} are homeomorphic, and Hayden-Mark-Piccirillo [hayden2021exotic, Theorem 2.7] showed that if CKC\cup K satisfies the (very mild) assumptions that

  • KK is not the meridian of CC,

  • every self homeomorphism of ZP\partial Z_{P} preserves the JSJ torus (with orientation) given by the image of the satellite torus of Pn,m(K)P_{n,m}(K) in ZP\partial Z_{P},

then ZPZ_{P} being diffeomorphic to ZQZ_{Q} implies that Xn(Pn,m(K))X_{n}(P_{n,m}(K)) is diffeomorphic to Xn(Qn,m(K))X_{n}(Q_{n,m}(K)). Therefore, by choosing suitable CKC\cup K, Theorem 1.3 also yields various families of exotic Mazur manifolds.

As an example, we present a family of knots KK for which Theorem 1.3 applies.

Lemma 6.10.

For every negative torus knot Tp,q-T_{p,q}, there exists a slice disk Σ\Sigma in some k2k\mathbb{CP}^{2} with |[Σ]|[Σ]2=(p1)(q1)=s(Tp,q)|[\Sigma]|-[\Sigma]^{2}=-(p-1)(q-1)=s(-T_{p,q}).

Proof.

We induct on p+qp+q. For p=q=1p=q=1, Tp,q=U-T_{p,q}=U is the unknot and we take Σ\Sigma to be the trivial slice disk in B4B^{4}.

For the induction step, assume p>q1p>q\geq 1. There is a positive twist along qq strands in Tp,q-T_{p,q} that converts it into Tpq,q-T_{p-q,q}. This gives a concordance C:Tpq,qTp,qC\colon-T_{p-q,q}\to-T_{p,q} in 2\mathbb{CP}^{2} with [C]=q[1][C]=q[\mathbb{CP}^{1}]. By induction hypothesis, there is a slice disk Σ\Sigma^{\prime} of Tpq,q-T_{p-q,q} in k2k^{\prime}\mathbb{CP}^{2} with |[Σ]|[Σ]2=(pq1)(q1)|[\Sigma^{\prime}]|-[\Sigma^{\prime}]^{2}=-(p-q-1)(q-1). Then Σ=ΣC\Sigma=\Sigma^{\prime}\cup C is a slice disk of Tp,q-T_{p,q} in (k+1)2(k^{\prime}+1)\mathbb{CP}^{2} with |[Σ]|[Σ]2=(p1)(q1)|[\Sigma]|-[\Sigma]^{2}=-(p-1)(q-1). ∎

6.7. Conway knot cables can obstruct its sliceness

Recently, Piccirillo [piccirillo2020conway] famously proved that the Conway knot, denoted ConwayConway, is not slice. Building on her proof, we show the following.

Proposition 6.11.

Let Conway(n+,n)Conway(n^{+},n^{-}) denote the 0-framed (n++n)(n^{+}+n^{-})-cable of the Conway knot, with the orientation on nn^{-} of the strands reversed. Then for some n>0n>0 we have

s(Conway(n+1,n))>2n=s(U(2n+1)).s(Conway(n+1,n))>-2n=s(U^{\sqcup(2n+1)}).

Here UmU^{\sqcup m} is the mm-component unlink.

If ConwayConway were slice, then any cable of it is concordant to the corresponding cable of the unknot, implying that they have the same ss-invariant. Therefore, Proposition 6.11 shows that, if people had enough computer power, we could have obstructed the sliceness of the Conway knot by calculating t