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

On the slice-torus invariant qMq_{M} from 2\mathbb{Z}_{2}-equivariant Seiberg–Witten Floer cohomology

Nobuo Iida Tokyo Institute of Technology, Ookayama, Meguro-ku, Tokyo [email protected] Taketo Sano RIKEN iTHEMS, Wako, Saitama 351-0198, Japan [email protected] Kouki Sato Meijo University, Tempaku, Nagoya 468-8502, Japan [email protected]  and  Masaki Taniguchi Department of Mathematics, Graduate School of Science, Kyoto University, Kitashirakawa Oiwake-cho, Sakyo-ku, Kyoto 606-8502, Japan [email protected]
Abstract.

We show that Iida–Taniguchi’s \mathbb{Z}-valued slice-torus invariant qMq_{M} cannot be realized as a linear combination of Rasmussen’s ss-invariant, Ozsváth–Szabó’s τ\tau-invariant, all of the 𝔰𝔩N\mathfrak{sl}_{N}-concordance invariants (N2N\geq 2), Baldwin–Sivek’s instanton τ\tau-invariant, Daemi–Imori–Sato–Scaduto–Taniguchi’s instanton s~\tilde{s}-invariant and Sano–Sato’s Rasmussen type invariants ss~c\tilde{ss}_{c}.

1. Introduction

Link homology theories have become central to modern knot theory, particularly from a four-dimensional perspective. These include prominent examples such as Khovanov homology, Heegaard Floer homology, monopole Floer homology, and instanton Floer homology. Various link homology theories yield numerous concordance invariants, among which one of the simplest and most extensively studied classes is the family of slice-torus invariants. A slice-torus invariant ([Li04, Le14]) is a real-valued function ff defined on the smooth knot concordance group that satisfies the following properties: for knots K,KK,K^{\prime} in S3S^{3},

  • (i)

    |f(K)|g4(K)|f(K)|\leq g_{4}(K),

  • (ii)

    f(K#K)=f(K)+f(K)f(K\#K^{\prime})=f(K)+f(K^{\prime}), and

  • (iii)

    f(T(p,q))=12(p1)(q1)f(T(p,q))=\frac{1}{2}(p-1)(q-1),

where g4(K)g_{4}(K) denotes the smooth slice genus and T(p,q)T(p,q) is the (p,q)(p,q)-positive torus knot for coprime integers p,q2p,q\geq 2.

Examples of slice-torus invariants arise from several theories, including Heegaard Floer theory [OS03], Khovanov homology theory [Ra10, Lobb09, Wu09, LS14, SS22, Lo12, Le14, LL:2016], instanton Floer theory [GLW19, BS21, DISST22]111 Ghosh–Li–Wong [GLW19] proved Baldwin–Sivek’s instanton tau invariant τ#\tau^{\#} coicides with the concordance invariant τI\tau_{I}\in\mathbb{Z} which comes from the Alexander decomposition of a variant of sutured instanton homology KHI¯\underline{KHI}^{-}. This ensures τ#\tau^{\#} is actually integer-valued. , and Seiberg–Witten Floer theory [IT24]. Once a slice-torus invariant is obtained, it immediately provides a solution to the Milnor conjecture on the slice genus of torus knots and reproves the existence of exotic 4\mathbb{R}^{4} by showing the existence of knots that are topologically slice but not smoothly slice. Moreover, for a large class of knots, including quasipositive and alternating knots, the values of all slice-torus invariants coincide. This observation underlies the conjecture that Ozsváth–Szabó’s τ\tau-invariant and Rasmussen’s ss-invariant are equal.

Notably, in [HO08], Hedden and Ording demonstrated that Rasmussen’s ss-invariant and Ozsváth–Szabó’s τ\tau-invariant are not identical. Similarly, Lewark [Le14] proved the linear independence of τ\tau, ss, and Rasmussen-type invariants sNs_{N} derived from 𝔰𝔩N\mathfrak{sl}_{N}-Khovanov–Rozansky homology theory. See also [MPP07, LS14, LC24, Sc23] for linear independence of Rasmussen invariants with different coefficients. There has also been significant progress in studying the general behavior of slice-torus invariants [Li04, Le14, FLL22, FLL24].

In [IT24], the first and fourth authors introduced a \mathbb{Z}-valued slice-torus invariant qM(K)q_{M}(K) arising from 2\mathbb{Z}_{2}-equivariant Seiberg–Witten theory applied to double branched covering spaces of knots, conjecturally equal to the Heegaard Floer qτq_{\tau}-invariant introduced by Hendricks–Lipshitz–Sarkar [HLS16] with a signature correction term. A natural question is whether qMq_{M} coincides with other known slice-torus invariants. In this paper, we address this question by proving the following:

Theorem 1.1.

Let us denote by ss, τ\tau, sω,αs_{\partial\omega,\alpha}, τ#\tau^{\#}, s~\tilde{s}, and ss~c\tilde{ss}_{c} the Rasmussen invariant [Ra10], the Ozsváth–Szabó τ\tau-invariant [OS03], the 𝔰𝔩N\mathfrak{sl}_{N}-concordance invariant (N2N\geq 2) [Lobb09, Wu09, LL:2016] with any separable potential ω\partial\omega equipped with any root α\alpha, the instanton τ\tau-invariant [BS21], the instanton s~\tilde{s}-invariant [DISST22], and Sano–Sato’s Rasmussen type invariant [SS22] for any PID RR with a prime element cc respectively. Then we have

qM(942)=1 while s(942)=τ(942)=sω,α(942)=τ#(942)=s~(942)=ss~c(942)=0,q_{M}(9_{42})=-1\text{ while }s(9_{42})=\tau(9_{42})=s_{\partial\omega,\alpha}(9_{42})=\tau^{\#}(9_{42})=\tilde{s}(9_{42})=\tilde{ss}_{c}(9_{42})=0,

where the convention of 9429_{42} follows the knotinfo [knotinfo]. In particular, the invariant qMq_{M} cannot be realized as a linear combination of ss, τ\tau, sω,αs_{\partial\omega,\alpha}, τ#\tau^{\#}, s~\tilde{s}, and ss~c\tilde{ss}_{c}.

Remark 1.2.

Note that Baraglia showed that the concordance invariant θ0\theta\in\mathbb{Z}_{\geq 0} from S1×2S^{1}\times\mathbb{Z}_{2}-equivariant Seiberg–Witten theory [Ba22, BH] satisfies

θ(942)=0andθ(942)=1.\theta(9_{42})=0\quad\text{and}\quad\theta(-9_{42})=1.

In general, for any knot KS3K\subset S^{3},

qM(K)θ(K)g4(K)q_{M}(K)\leq\theta(K)\leq g_{4}(K)

holds.

The first inequality was proved by the first and fourth authors in [IT24, Theorem 1.11], while the second inequality was proved by Baraglia in [Ba22, Theorem 1.4]. It is known that g4(942)=1g_{4}(9_{42})=1, so in summary, qMq_{M} and θ\theta provide the optimal 4-ball genus bound for 9429_{42}, whereas the others in Theorem 1.1 do not.

As an immediate corollary, we have:

Corollary 1.3.

Let nn be a non-zero integer. The nn-fold connected sum #n942\#_{n}9_{42} is not a squeezed knot.

Proof.

This follows from the fact that every slice torus invariant takes the same value [FLL22] for squeezed knots. ∎

Remark 1.4.

Note that a refinement of the Rasmussen invariant s+Sq2s_{+}^{\operatorname{Sq}_{2}} introduced by Lipshitz–Sarkar [LS14], which uses Khovanov homotopy type with Steenrod operator, satisfies s+Sq2(942)/2=1s_{+}^{\operatorname{Sq}_{2}}(9_{42})/2=1. This fact has been used to prove 9429_{42} is not squeezed. Corollary 1.3 might have alternative proof by showing s+Sq2(#n942)0s_{+}^{\operatorname{Sq}_{2}}(\#_{n}9_{42})\neq 0.

Structure of the paper: In Section 2, we give a proof of Theorem 1.1. In Section 3, we briefly discuss about the backgrounds of the invariants qMq_{M} and θ\theta and summarize basic properties of them. Also, we shall discuss the values of these invariants for prime knots up to 1010 crossings. We put the tables of these values in Section 4.

Acknowledgement.

We would like to thank Joshua Greene for answering to our question regarding [Gr13]. The first author acknowledges support from JSPS KAKENHI Grant Number 22J00407. The second author acknowledges support from JSPS KAKENHI Grant Numbers 23K12982, RIKEN iTHEMS Program and academist crowdfunding. The fourth author acknowledges partial support from JSPS KAKENHI Grant Number 22K13921.

2. Proof of Theorem 1.1

Refer to caption
Figure 1. Kirby moves which show S13(942)S13(K)S^{3}_{1}(9_{42})\cong S^{3}_{-1}(K) for some knot KK in S3S^{3}
Refer to caption
Figure 2. Kirby moves which show that S13(942)S^{3}_{1}(-9_{42}) bounds a smooth compact contractible 4-manifold

We first prove that the invariants τ\tau, τ\tau^{\sharp} and s~\tilde{s} vanish for 9429_{42}:

Lemma 2.1.

We have

τ(942)=τ#(942)=s~(942)=0.\tau(9_{42})=\tau^{\#}(9_{42})=\tilde{s}(9_{42})=0.
Proof of Lemma 2.1.

We shall use the following statements:

  • For any knot KK, if τ(K)>0\tau(K)>0, then we have d(S13(K))<0d(S^{3}_{1}(K))<0.

  • For any knot KK, if s~(K)>0\tilde{s}(K)>0, then we have h(S13(K))<0h(S^{3}_{1}(K))<0.

  • For any knot KK, if τ#(K)>0\tau^{\#}(K)>0, then we have h(S13(K))<0h(S^{3}_{1}(K))<0.

Here dd denotes the Heegaard Floer correction term and hh denotes the instanton Frøyshov invariant of oriented homology 33-spheres introduced in [Fr02] and the convention follows h(Σ(2,3,5))=1h(\Sigma(2,3,5))=1.

The first claim follows from [HW16], see also [Sa18, Section 2.2]. The second fact is proven in [DISST22, Theorem 1.5] and the third fact is proven in [BS22II, Proposition 9.2]. Therefore, to show τ(942)=τ#(942)=s~(942)=0\tau(9_{42})=\tau^{\#}(9_{42})=\tilde{s}(9_{42})=0, it is sufficient to see

h(S13(942))=h(S13(942))=d(S13(942))=d(S13(942))=0.h(S^{3}_{1}(9_{42}))=h(S^{3}_{1}(-9_{42}))=d(S^{3}_{1}(9_{42}))=d(S^{3}_{1}(-9_{42}))=0.

The Kirby moves described in Figure 2 shows that

S13(942)S13(K)S^{3}_{1}(9_{42})\cong S^{3}_{-1}(K)

for some knot KK in S3S^{3}. So, S13(942)S^{3}_{1}(9_{42}) bounds both of positive definite and negative definite 4-manifolds, hence we have h(S13(942))=d(S13(942))=0h(S^{3}_{1}(9_{42}))=d(S^{3}_{1}(9_{42}))=0. The Kirby moves described in Figure 2 shows that S13(942)S^{3}_{1}(-9_{42}) bounds a smooth compact contractible 44-manifold. So, we see the latter statements. This completes the proof. ∎

Next, we determine Sano–Sato’s concordance invariants for 9429_{42}. The link invariant ss~h\widetilde{ss}_{h} of [SS22] is defined for each non-zero non-invertible element hh in an integral domain RR, and is slice-torus when RR is a PID and hh is prime. [SS21, Theorem 2] states that for the special case (R,h)=(𝔽[H],H)(R,h)=(\mathbb{F}[H],H) where 𝔽\mathbb{F} is any field, the invariant coincides with the Rasmussen invariant s𝔽s^{\mathbb{F}} over 𝔽\mathbb{F}.

66 . . . . . . [H]\mathbb{Z}[H] 44 . . . . . [H]\mathbb{Z}[H] . 22 . . . . [H]\mathbb{Z}[H] . . 0 . . . [H]2\mathbb{Z}[H]^{2} [H]\mathbb{Z}[H] . . 2-2 . . [H]\mathbb{Z}[H] . . . . 4-4 . [H]\mathbb{Z}[H] . . . . . 6-6 [H]\mathbb{Z}[H] . . . . . . 4-4 3-3 2-2 1-1 0 11 22        66 . . . . . . [H]/(H)\mathbb{Z}[H]/(H) 44 . . . . . . . 22 . . . . [H]/(H)\mathbb{Z}[H]/(H) . . 0 . . . [H]/(H)\mathbb{Z}[H]/(H) [H]\mathbb{Z}[H] . . 2-2 . . . . . . . 4-4 . [H]/(H)\mathbb{Z}[H]/(H) . . . . . 6-6 . . . . . . . 4-4 3-3 2-2 1-1 0 11 22

Table 1. 𝐶𝐵𝑁~(942;)\widetilde{\mathit{CBN}}(9_{42};\mathbb{Z}) and 𝐵𝑁~(942;).\widetilde{\mathit{BN}}(9_{42};\mathbb{Z}).
Lemma 2.2.

ss~h(942;R)=0\widetilde{ss}_{h}(9_{42};R)=0 for any PID RR and prime hRh\in R.

Proof.

From [SS22, Lemma 4.37], we always have

ss~Hss~h\widetilde{ss}_{H}\leq\widetilde{ss}_{h}

where the left-hand side is the (non slice-torus) invariant corresponding to the pair (R,h)=([H],H)(R,h)=(\mathbb{Z}[H],H). Thus it suffices to prove that

ss~H(942)=ss~H(942)=0.\widetilde{ss}_{H}(9_{42})=\widetilde{ss}_{H}(-9_{42})=0.

Using the program yui [YUI] developed in [SS22], the reduced Bar-Natan complex 𝐶𝐵𝑁~(942;)\widetilde{\mathit{CBN}}(9_{42};\mathbb{Z}) of 9429_{42} over \mathbb{Z} (in its simplified form, together with the differential matrices) can be computed by the command

$ ykh ckh 9_42 -t Z -c H -r -d

whose result is displayed in the left side of Table 1. With the differential matrices, its homology 𝐵𝑁~(942;)\widetilde{\mathit{BN}}(9_{42};\mathbb{Z}) can be easily computed as in the right side of Table 1. Note the single free summand [H]\mathbb{Z}[H] in bigrading (0,0)(0,0). This shows that ss~H(942)=0\widetilde{ss}_{H}(9_{42})=0. Similarly, we can show that ss~H(942)=0\widetilde{ss}_{H}(-9_{42})=0. ∎

Next, we determine the 𝔰𝔩N\mathfrak{sl}_{N}-concordance invariants for 9429_{42}. Recall from [LL:2016] that a family of slice-torus invariants sw,αs_{\partial w,\alpha} is given for each choice of a separable potential, i.e. a degree NN monic polynomial w[x]\partial w\in\mathbb{C}[x] having NN distinct roots in \mathbb{C}, together with a choice of a root α\alpha of w\partial w. For any knot KK, there is a spectral sequence starting from the (reduced) 𝔰𝔩N\mathfrak{sl}_{N} Khovanov–Rozansky homology H𝔰𝔩NH_{\mathfrak{sl}_{N}} and converging to the perturbed homology Hw,αH_{\partial w,\alpha} of dimension 11. The invariant is defined as

sw,α=j2(N1)s_{\partial w,\alpha}=\frac{j}{2(N-1)}

where jj is the 𝔰𝔩N\mathfrak{sl}_{N}-quantum grading of the surviving generator of the EE_{\infty} term. In particular, the special case (w,α)=(xN1,1)(\partial w,\alpha)=(x^{N}-1,1) gives the invariant sNs_{N} introduced in [Lobb09, Wu09].

22 . 11 . 11 .
0 11 . 22 . 11
2-2 . 11 . 11 .
a\qa\backslash q 4-4 2-2 0 22 44
Δ=2\Delta=-2
22 . . . . .
0 . . 11 . .
2-2 . . . . .
a\qa\backslash q 4-4 2-2 0 22 44
Δ=0\Delta=0
Table 2. (942)\mathcal{H}(9_{42})
Lemma 2.3.

sw,α(942)=0s_{\partial w,\alpha}(9_{42})=0 for N2N\geq 2 and any choice of (w,α)(\partial w,\alpha).

Proof.

When N=2N=2, it is known that sw,α=ss_{\partial w,\alpha}=s^{\mathbb{Q}} and the statement is proved above. The case for N3N\geq 3 follows from [Chandler-Gorsky:2024, Theorem 3.14] together with the computational result of the reduced HOMFLY-PT homology of 9429_{42}. First, from [LL:2016, Proposition 3.3], sw,αs_{\partial w,\alpha} is invariant under any translation of xx, so we may assume that α=0\alpha=0. Writing w\partial w as

w=xN+aNxN1++a2x\partial w=x^{N}+a_{N}x^{N-1}+\cdots+a_{2}x

with a20a_{2}\neq 0, [Chandler-Gorsky:2024, Theorem 3.14] states that the first differential of the spectral sequence H𝔰𝔩NHw,αH_{\mathfrak{sl}_{N}}\Rightarrow H_{\partial w,\alpha} is given by

dw=aNdN1++a2d1d_{\partial w}=a_{N}d_{N-1}+\cdots+a_{2}d_{1}

where each dkd_{k} is the first differential of the spectral sequence given by Rasmussen in [Rasmussen:2015], starting from the (reduced) HOMFLY-PT homology \mathcal{H} in the E1E_{1} page and converging to the 𝔰𝔩k\mathfrak{sl}_{k} homology H𝔰𝔩kH_{\mathfrak{sl}_{k}}. We claim that for K=942K=9_{42}, we have H𝔰𝔩N\mathcal{H}\cong H_{\mathfrak{sl}_{N}} and dN1==d2=0d_{N-1}=\cdots=d_{2}=0.

Using the program yui-kr [YUIKR] developed in [Nakagane-Sano:2024], the reduced HOMFLY-PT homology of 9429_{42} can be computed by the command

$ ykr 9_42 -f delta

whose result is given in Table 2. Here, the triply graded homology group (942)\mathcal{H}(9_{42}) is sliced by the Δ\Delta-grading introduced in [Chandler-Gorsky:2024], and the structure (i.e. \mathbb{Q}-dimension) of each Δ\Delta-slice is displayed with respect to the (q,a)(q,a)-bigrading. Since (942)\mathcal{H}(9_{42}) has Δ\Delta-thickness 22, it follows that the spectral sequence (942)H𝔰𝔩N(942)\mathcal{H}(9_{42})\Rightarrow H_{\mathfrak{sl}_{N}}(9_{42}) has trivial differentials and hence (942)H𝔰𝔩N(942)\mathcal{H}(9_{42})\cong H_{\mathfrak{sl}_{N}}(9_{42}) (see [Chandler-Gorsky:2024, Corollary 2.14]). Moreover, for each 2kN12\leq k\leq N-1, the first differential dkd_{k} changes the Δ\Delta-grading by 2k22k-2 and the (q,a)(q,a) bigrading by (2k,2)(2k,-2), so we see that dk=0d_{k}=0.

Next, let us consider the spectral sequence H𝔰𝔩1\mathcal{H}\Rightarrow H_{\mathfrak{sl}_{1}} for k=1k=1. The ii-th differential d1(i)d^{(i)}_{1} changes the Δ\Delta-grading by 22i2-2i and the (q,a)(q,a) bigrading by (2i,2i)(2i,-2i). Again from grading reasons, we have d1(i)=0d^{(i)}_{1}=0 when i>1i>1, and also d1=0d_{1}=0 on the single \mathbb{Q}-summand in Δ\Delta-grading 0. This \mathbb{Q}-summand is necessarily the surviving one, and other summands must cancel out by the first differential d1d_{1}.

We conclude that the spectral sequence (942)H𝔰𝔩N(942)Hw,α(942)\mathcal{H}(9_{42})\cong H_{\mathfrak{sl}_{N}}(9_{42})\Rightarrow H_{\partial w,\alpha}(9_{42})\cong\mathbb{Q} with first differential dw=a2d1d_{\partial w}=a_{2}d_{1} collapses after the first page and the quantum grading of the surviving generator is 0. ∎

Now, we only need to see qM(942)=1q_{M}(9_{42})=-1 to prove Theorem 1.1 from Lemma 2.1, Lemma 2.2 and Lemma 2.3.

Proof of Theorem 1.1. This is an immediate consequence of the fact that the 2\mathbb{Z}_{2}-branched cover Σ2(942)\Sigma_{2}(9_{42}) is an L-space, which is observed by Greene in [Gr13]. (However, notice that 9429_{42} is not quasi-alternating.) In [IT24, Theorem 1.10], it is proved that

qM(K)=σ(K)2q_{M}(K)=-\frac{\sigma(K)}{2}

when Σ2(K)\Sigma_{2}(K) is an L-space. Since σ(942)=2\sigma(9_{42})=2 we obtain the desired result. Here our convention of the knot signature σ\sigma follows that of [IT24], i.e. σ(T(2,3))=2\sigma(T(2,3))=-2. ∎

In our argument above, we used the fact that Σ2(942)\Sigma_{2}(9_{42}) is an L-space over 𝔽2\mathbb{F}_{2}. Since details of the proof are not given in [Gr13], we give a brief computation to prove it.

Proposition 2.4.

(Greene [Gr13]). The 3-manifold Σ2(942)\Sigma_{2}(9_{42}) is an L-space over the coefficient 𝔽2\mathbb{F}_{2}.

Refer to caption
Figure 3. A diagram of 9429_{42} from the KnotInfo [knotinfo]. We added names A,B,CA,B,C for three crossings and markings.
Proof of Proposition 2.4.

KnotFolio [knotfolio] is helpful for the authors to check the following argument. As the two smoothings at the crossing AA in the figure 3, we obtain the knot 8198_{19} and the link 7n17n1. Note that the determinants of 8198_{19} and 7n17n1 are 33 and 44 respectively. The not 8198_{19} is nothing but the torus knot T3,4T_{3,4}, and since its branched double cover Σ2(T3,4)=Σ(2,3,4)\Sigma_{2}(T_{3,4})=\Sigma(2,3,4) has a metric with positive scalar curvature (See [milnor19753] for example), it is an L-space. We can check that 7n17n1 is two-fold quasi-alternating (TQA)[SS18] [IsTu24, Section 3] by smoothing at BB and CC, with the marking described by dots in the diagram. More precisely we obtain the following:

7n1smoothing at BU˙2 and L6n17n1\xrightarrow{\text{smoothing at }B}\dot{U}_{2}\,\text{ and }\,L6n1
L6n1smoothing at CU˙2 and L4a1L6n1\xrightarrow{\text{smoothing at }C}\dot{U}_{2}\,\text{ and }\,L4a1

where U˙2\dot{U}_{2} is the two-component unlink with a dot on each component. Since L4a1L4a1 is a non-split alternating link, which is TQA, L6n1L6n1 and thus 7n17n1 are also TQA. Thus the branched double covering Σ2(7n1)\Sigma_{2}(7n1) is also an LL-space over 𝔽2\mathbb{F}_{2} coefficient by [SS18, Corollary 1]. Now by applying the long exact sequence for the branched double coverings [OS05] for the smoothing

942smoothing at A819 and  7n1,9_{42}\xrightarrow{\text{smoothing at }A}8_{19}\,\text{ and }\,7n1,

we obtain

dim𝔽2HF^(Σ2(942))dim𝔽2HF^(Σ2(819))+dim𝔽2HF^(Σ2(7n1))=3+4=7=|H1(Σ2(942))|.\operatorname{dim}_{\mathbb{F}_{2}}\widehat{HF}(\Sigma_{2}(9_{42}))\leq\operatorname{dim}_{\mathbb{F}_{2}}\widehat{HF}(\Sigma_{2}(8_{19}))+\operatorname{dim}_{\mathbb{F}_{2}}\widehat{HF}(\Sigma_{2}(7n1))=3+4=7=|H_{1}(\Sigma_{2}(9_{42}))|.

On the other hand, for any closed oriented 3-manifold YY,

dim𝔽2HF^(Y)χ(HF^(Y))=|H1(Y;)|\operatorname{dim}_{\mathbb{F}_{2}}\widehat{HF}(Y)\geq\chi(\widehat{HF}(Y))=|H_{1}(Y;\mathbb{Z})|

holds and this implies

dim𝔽2HF^(Σ2(942))|H1(Σ2(942))|.\operatorname{dim}_{\mathbb{F}_{2}}\widehat{HF}(\Sigma_{2}(9_{42}))\geq|H_{1}(\Sigma_{2}(9_{42}))|.

Thus, dim𝔽2HF^(Σ2(942))=|H1(Σ2(942))|\operatorname{dim}_{\mathbb{F}_{2}}\widehat{HF}(\Sigma_{2}(9_{42}))=|H_{1}(\Sigma_{2}(9_{42}))| and therefore Σ2(942)\Sigma_{2}(9_{42}) is an LL-space over 𝔽2\mathbb{F}_{2} as well.

Remark 2.5.

As we discussed in the proof of Lemma 2.1, the slice torus invariants τ\tau, τ#\tau^{\#} and s~\tilde{s} are related to Heegaard Floer correction term dd and instanton Frøyshov invariant hh of surgeries. This is one reason that the equalities

τ=s~=τ#\tau=\tilde{s}=\tau^{\#}

are conjectured in [DISST22]. Our arugment shows the implication

qM(K)>0d(S13(K))<0q_{M}(K)>0\Rightarrow d(S^{3}_{1}(K))<0

does not hold. This is the first slice torus invariant from gauge theory which does not satisfy this kind of implication.

3. qMq_{M} and θ\theta for prime knots with small crossing number

Let us review constructions of invariants qMq_{M} and θ\theta and summarize their basic properties.

Let pp be a prime number. For a knot KK in S3S^{3}, the associated Seiberg–Witten Floer homotopy type introduced by Manolescu [Man03] to its p\mathbb{Z}_{p}-branched cover Σp(K)\Sigma_{p}(K) with unique p\mathbb{Z}_{p}-invariant spin structure

SWF(Σp(K))SWF(\Sigma_{p}(K))

has a S1×pS^{1}\times\mathbb{Z}_{p}-action 222Strictly speaking, S1×pS^{1}\times\mathbb{Z}_{p}-equivariant Seiberg–Witten Floer stable homotopy type which is independent of the Riemannian metric is not formulated, while the equivariant Seiberg–Witten Floer cohomology which is independent of the Riemannian metric is formulated. We use the latter only. . Baraglia–Hekmati [BH, BH2] and Baraglia [Bar] studied its S1×pS^{1}\times\mathbb{Z}_{p}-equivariant cohomology

HS1×p(SWF(Σp(K));𝔽p)H^{*}_{S^{1}\times\mathbb{Z}_{p}}(SWF(\Sigma_{p}(K));\mathbb{F}_{p})

equipped with the module structure over H(B(S1×p);𝔽p)H^{*}(B(S^{1}\times\mathbb{Z}_{p});\mathbb{F}_{p}) and introduced a family of concordance invariants

θ(p)(K)1p10\theta^{(p)}(K)\in\frac{1}{p-1}\mathbb{Z}_{\geq 0}

from the module structure of HS1×p(SWF(Σp(K)))H^{*}_{S^{1}\times\mathbb{Z}_{p}}(SWF(\Sigma_{p}(K))). They reproved and the Milnor conjecture using the 4-genus bound comes from the invariants θ(p)(K)\theta^{(p)}(K). For the computations of θ(p)(K)\theta^{(p)}(K), they have used:

  • the spectral sequence

    E2=H(Bp;H~S1(SWF(Σp(K));𝔽p)HS1×p(SWF(Σp(K));𝔽p),E_{2}=H^{*}(B\mathbb{Z}_{p};\tilde{H}^{*}_{S^{1}}(SWF(\Sigma_{p}(K));\mathbb{F}_{p})\Rightarrow H^{*}_{S^{1}\times\mathbb{Z}_{p}}(SWF(\Sigma_{p}(K));\mathbb{F}_{p}),
  • the isomorphism between the S1S^{1}-equivariant monopole Floer homologies and the Heegaard Floer homology HF+HF^{+} , and

  • the graded root technique to compute the Heegaard Floer cohomology of almost rational 3-manifolds.

The first and fourth authors studied p=2p=2 case and found that 2\mathbb{Z}_{2}-equivariant cohomology (forgetting the S1S^{1} action!) gave a \mathbb{Z}-valued slice-torus invariant qMq_{M} and in particular gave an alternative proof of the Milnor conjecture. The authors did not use Baraglia–Hekmati’s tools mentioned above, but instead used the following methods to show that qMq_{M} is actually a slice-torus invariant.

  • Some methods used by Kronhimer [Kr97] and Daemi–Scaduto [DS19][DS23] for singular instanton theory based on Freedman–Quinn’s work [FQ90] on normally immersed surface in four-manifolds. This was used to show

    rank𝔽2[Q]H~2(SWF(Σ2(K));𝔽2)=1\operatorname{rank}_{\mathbb{F}_{2}[Q]}\widetilde{H}^{*}_{\mathbb{Z}_{2}}(SWF(\Sigma_{2}(K));\mathbb{F}_{2})=1

    and the ”cobordism inequality”

    qM(K1)qM(K0)+g(S)q_{M}(K_{1})\leq q_{M}(K_{0})+g(S)

    for a smooth and orientable surface cobordism S:K0K1S:K_{0}\to K_{1} in [0,1]×S3[0,1]\times S^{3}, which immediately implies the concordance invariance and the 4-ball genus bound for qMq_{M}.

  • The homotopical transverse knot invariant

    Ψ2(K):S0ΣSWF(Σ2(K))\Psi_{2}(K):S^{0}\to\Sigma^{\text{\textbullet}}SWF(-\Sigma_{2}(K))

    introduced in [IT24], its gluing property and non-vanishing property, which is nothing but the equivariant version of the stable homotopy version of the monopole contact invariant [IT20] and its corresponding properties previously developed by the authors mainly in [IT20]. This is used to show the ”adjunction equality”

    qM(S)=g(S)q_{M}(\partial S)=g(S)

    for symplectic surface SS in D4D^{4} with transverse knot boundary, which in particular implies the computation for torus knots. A key point in the proof of the ”adjunction equality” is to prove that the relative invariant BFΣ2(S)(1)BF^{*}_{\Sigma_{2}(S)}(1) for such a symplectic surface SS is non-QQ-torsion and non-divisible by QQ, and thus attains the ”bottom of the QQ tower”.

Based on these methods, in [IT24, Theorem 1.16], the authors proved

rank𝔽2[Q]H~2(SWF(Σ2(K));𝔽2)=1,\operatorname{rank}_{\mathbb{F}_{2}[Q]}\widetilde{H}^{*}_{\mathbb{Z}_{2}}(SWF(\Sigma_{2}(K));\mathbb{F}_{2})=1,

where QQ is the degree one variable so that 𝔽2[Q]=H(B2;𝔽2)\mathbb{F}_{2}[Q]=H^{*}(B\mathbb{Z}_{2};\mathbb{F}_{2}). In addition, the module H~2(SWF(Σ2(K));𝔽2)\widetilde{H}^{*}_{\mathbb{Z}_{2}}(SWF(\Sigma_{2}(K));\mathbb{F}_{2}) has an absolute \mathbb{Q}-grading.

Definition 3.1.

From this module, we define a concordance invariant

qM(K):=min{i|xH~2i(SWF(Σ2(K));𝔽2),Qnx0 for all n0}34σ(K)q_{M}(K):=\min\{i|x\in\widetilde{H}^{i}_{\mathbb{Z}_{2}}(SWF(-\Sigma_{2}(K));\mathbb{F}_{2}),\,Q^{n}x\neq 0\text{ for all }n\geq 0\}-\frac{3}{4}\sigma(K)\in\mathbb{Z}

for a given knot KK in S3S^{3}.

Remark 3.2.

Conjectually, the invariant qMq_{M} is equal to Hendricks–Lipshitz–Sarkar’s qτq_{\tau} invariant [HLS16] with a signature correction term, defined using 2\mathbb{Z}_{2}-equivariant Heegaard Floer homology of Σ2(K)\Sigma_{2}(K). A Heegaard Floer counterpart of the transverse knot invariant in [IT24] has been considered in [Ka18] by Kang.

From now on, we give the values of qMq_{M} and θ\theta for all 85 prime knots with crossing number 9\leq 9 and explain the current situation for those with crossing number 1010. First of all, we summarize the basic properties of qMq_{M} and θ=θ(p=2)\theta=\theta^{(p=2)}, which are proven in [Ba22, IT24].

Theorem 3.3.

The invariants qMq_{M} and θ\theta have the following properties:

  1. (1)

    For any knot KS3K\subset S^{3}, we have integers

    qM(K) and θ(K)0,q_{M}(K)\in\mathbb{Z}\text{ and }\theta(K)\in\mathbb{Z}_{\geq 0},

    which do not depend on orientations of KK. Also, these are concordance invariants.

  2. (2)

    For oriented knots K0,K1S3K_{0},K_{1}\subset S^{3}, we have

    qM(K0#K1)=qM(K0)+qM(K1) and θ(K0#K1)θ(K0)+θ(K1).q_{M}(K_{0}\#K_{1})=q_{M}(K_{0})+q_{M}(K_{1})\text{ and }\theta(K_{0}\#K_{1})\leq\theta(K_{0})+\theta(K_{1}).
  3. (3)

    For any knot KS3K\subset S^{3}, we have

    qM(K)θ(K)g4(K) and 12σ(K)θ(K).q_{M}(K)\leq\theta(K)\leq g_{4}(K)\text{ and }-\frac{1}{2}\sigma(K)\leq\theta(K).
  4. (4)

    For any transverse knot 𝒯(S3,ξstd)\mathcal{T}\subset(S^{3},\xi_{std}), we have the Bennequin type inequality

    sl(𝒯)2qM(𝒯)1,sl(\mathcal{T})\leq 2q_{M}(\mathcal{T})-1,

    where ξstd\xi_{std} is the unique tight contact structure on S3S^{3}. Moreover, if 𝒯\mathcal{T} is a boundary of a connected symplectic surface S(D4,ωstd)S\subset(D^{4},\omega_{std}), the equality

    sl(𝒯)=2qM(𝒯)1=2g(S)1sl(\mathcal{T})=2q_{M}(\mathcal{T})-1=2g(S)-1

    holds, where ωstd\omega_{std} denotes the standard symplectic structure on D4D^{4} and slsl denotes the self-linking number. As a consequence, for the quasipositive knot KK, we have

    qM(K)=θ(K)=g4(K).q_{M}(K)=\theta(K)=g_{4}(K).
  5. (5)

    For a knot KK such that the double brached covering Σ2(K)\Sigma_{2}(K) is an LL-space over 𝔽2\mathbb{F}_{2} coefficient,

    qM(K)=θ(K)=σ(K)2q_{M}(K)=\theta(K)=-\frac{\sigma(K)}{2}

    holds. In particular, this relation holds for all quasialternating knots, since the branched double covering of a quasialternating knot is an LL-space [OS05].

  6. (6)

    If d(Σ2(K))<σ(K)d(\Sigma_{2}(K))<-\sigma(K) and σ(K)0\sigma(K)\leq 0, then we have

    1+12σ(K)θ(K),1+\frac{1}{2}\sigma(K)\leq\theta(K),

    where the d(Σ2(K))d(\Sigma_{2}(K)) denotes the dd-invariant with the unique spin structure on Σ2(K)\Sigma_{2}(K).

The tables in Section 4 give the values of qMq_{M} and θ\theta for all 85 prime knots with crossing number 9\leq 9. Notice that the values of θ\theta for the knots with the opposite chirality are not calculated. We compute these for a choice of chirality of KK such that σ(K)0\sigma(K)\leq 0. The values of σ/2-\sigma/2 and g4g_{4} are quoted from the knotinfo. The values of qMq_{M} and θ\theta can be seen as follows. Among the 85 prime knots with crossing number 9\leq 9, which are listed above, all but

819,942,9468_{19},9_{42},9_{46}

are quasialternating and thus satisfy qM=θ=σ/2q_{M}=\theta=-\sigma/2. Since 9469_{46} is slice, qM=θ=σ/2=0q_{M}=\theta=-\sigma/2=0.

For 8198_{19}, qM=θ=σ/2q_{M}=\theta=-\sigma/2 still holds since Σ2(819)=Σ(2,3,4)\Sigma_{2}(8_{19})=\Sigma(2,3,4) has a positive scalar curvature metric and thus an LL-space as explained in the previous section. We could also use the fact that 819=T3,48_{19}=T_{3,4} is quasipositive and thus qM=θ=g4=sl¯+12q_{M}=\theta=g_{4}=\frac{\overline{sl}+1}{2}. As stated by Greeene and seen in this paper, Σ2(942)\Sigma_{2}(9_{42}) is also an L-space over 𝔽2\mathbb{F}_{2} and thus qM=θ=σ/2q_{M}=\theta=-\sigma/2 holds for 9429_{42} as well.

Remark 3.4.

In this remark, we explain the circumstance for the 165 prime knots with corssing number 10, which can be read from the knot info. Among them, all but

10124,10128,10132,10136,10139,10140,10145,10152,10153,10154,1016110_{124},10_{128},10_{132},10_{136},10_{139},10_{140},10_{145},10_{152},10_{153},10_{154},10_{161}

are quasialternating and thus satsify qM=θ=σ/2q_{M}=\theta=-\sigma/2. Among the remaining 11 knots,

10124,10128,10140,10145,10152,10154,1016110_{124},10_{128},10_{140},10_{145},10_{152},10_{154},10_{161}

are quasipositive and thus satisfy qM=θ=g4=sl¯+12q_{M}=\theta=g_{4}=\frac{\overline{sl}+1}{2}. The remaining 4-knots are

10132,10136,10139,10153.10_{132},10_{136},10_{139},10_{153}.

Since 1015310_{153} is slice, we have qM(10136)=θ(10136)=0q_{M}(10_{136})=\theta(10_{136})=0. As for 1013910_{139}, by [FLL22, Lemma 2.16], 1013910_{139} is a squeezed knot, so qM=s/2=τ=4q_{M}=s/2=\tau=4 holds for 1013910_{139}. Since g4(10139)=4g_{4}(10_{139})=4 is also known, we know θ(10139)=4\theta(10_{139})=4. For 1013210_{132}, we have σ/2=0-\sigma/2=0 and g4=1g_{4}=1, so we know 0θ(10132)10\leq\theta(10_{132})\leq 1 and qM(10132)θ(10132)1q_{M}(10_{132})\leq\theta(10_{132})\leq 1 but the authors could not determine these. For 1013610_{136}, σ/2=g4=1-\sigma/2=g_{4}=1 holds, thus θ(10136)=1\theta(10_{136})=1 but qM(10136)q_{M}(10_{136}) is unknown.

Remark 3.5.

In all prime knots in which we know the values of invariants qMq_{M} and θ\theta in the tables,

|σ(K)|/2|qM(K)| and qM(K)=θ(K)|\sigma(K)|/2\leq|q_{M}(K)|\text{ and }q_{M}(K)=\theta(K)

hold. However, neither of these do not hold in general for a knot KS3K\subset S^{3}. Set K0=T(3,11)##10T(2,3)K_{0}=-T(3,11)\#\#_{10}T(2,3), where K-K denotes the concordance inverse of KK. We have

qM(T(3,11))=g4(T(3,11))=10,σ(T(3,11))=16,qM(T(2,3))=g4(T(2,3))=1,σ(T(2,3))=2.q_{M}(T(3,11))=g_{4}(T(3,11))=10,\ \sigma(T(3,11))=-16,\ q_{M}(T(2,3))=g_{4}(T(2,3))=1,\ \sigma(T(2,3))=-2.

Now by the additivity of qMq_{M} and the signature under the connected sum operation, we have qM(K0)=0q_{M}(K_{0})=0 σ(K0)=4\sigma(K_{0})=-4. Thus, 2=|σ(K0)|/2>|qM(K0)|=02=|\sigma(K_{0})|/2>|q_{M}(K_{0})|=0. Moreover, we have σ(K)/2θ(K)\sigma(K)/2\leq\theta(K), so

0=qM(K0)<σ(K0)/2=2θ(K0).0=q_{M}(K_{0})<-\sigma(K_{0})/2=2\leq\theta(K_{0}).

Furthermore, by considering #nK0\#_{n}K_{0}, we can see that the difference θ(K)qM(K)\theta(K)-q_{M}(K) can be arbitrarily large.

Question 3.6.

Find qM(10132)q_{M}(10_{132}), θ(10132)\theta(10_{132}), and qM(10136)q_{M}(10_{136}). As detailed in (3.4), these are the only undetermined values among the prime knots with the crossing number 10\leq 10. Since 1013210_{132} and 1013610_{136} are Montesinos knots, the authors expect that these can be computed by using the Seifert fibered structures on the double-branched coverings.

4. Tables

KK σ(K)/2-\sigma(K)/2 qM(K)q_{M}(K) θ(K)\theta(K) g4(K)g_{4}(K) (quasi)alternating positivity(BPPSQQPBP\subset P\subset SQ\subset QP)
010_{1} 0 0 0 0 alt BP
313_{1} 11 11 11 11 alt BP
414_{1} 0 0 0 11 alt -
515_{1} 11 11 11 11 alt BP
525_{2} 11 11 11 11 alt P
616_{1} 0 0 0 0 alt -
626_{2} 11 11 11 11 alt -
636_{3} 0 0 0 11 alt -
717_{1} 33 33 33 33 alt BP
727_{2} 11 11 11 11 alt P
737_{3} 22 22 22 22 alt P
747_{4} 11 11 11 11 alt P
757_{5} 22 22 22 22 alt P
767_{6} 11 11 11 11 alt -
777_{7} 0 0 0 11 alt -
818_{1} 0 0 0 11 alt -
828_{2} 22 22 22 22 alt -
838_{3} 0 0 0 0 alt -
848_{4} 11 11 11 11 alt -
858_{5} 22 22 22 22 alt -
868_{6} 11 11 11 11 alt -
878_{7} 11 11 11 11 alt -
888_{8} 0 0 0 0 alt -
898_{9} 11 11 11 11 alt -
8108_{10} 11 11 11 11 alt -
8118_{11} 11 11 11 11 alt -
8128_{12} 0 0 0 11 alt -
8138_{13} 0 0 0 11 alt -
8148_{14} 11 11 11 11 alt -
8158_{15} 22 22 22 22 alt P
8168_{16} 11 11 11 11 alt -
8178_{17} 0 0 0 11 alt -
8188_{18} 0 0 0 11 alt -
8198_{19} 33 33 33 33 non-q.alt! BP
8208_{20} 0 0 0 0 q.alt QP
8218_{21} 11 11 11 11 q.alt -
KK σ(K)/2-\sigma(K)/2 qM(K)q_{M}(K) θ(K)\theta(K) g4(K)g_{4}(K) (quasi)alternating positivity (BPPSQQPBP\subset P\subset SQ\subset QP)
919_{1} 44 44 44 44 alt BP
929_{2} 11 11 11 11 alt P
939_{3} 33 33 33 33 alt P
949_{4} 22 22 22 22 alt P
959_{5} 11 11 11 11 alt P
969_{6} 33 33 33 33 alt P
979_{7} 22 22 22 22 alt P
989_{8} 11 11 11 11 alt -
999_{9} 33 33 33 33 alt P
9109_{10} 22 22 22 22 alt P
9119_{11} 22 22 22 22 alt -
9129_{12} 11 11 11 11 alt P
9139_{13} 22 22 22 22 alt -
9149_{14} 0 0 0 11 alt -
9159_{15} 11 11 11 11 alt -
9169_{16} 33 33 33 33 alt P
9179_{17} 11 11 11 11 alt -
9189_{18} 22 22 22 22 alt P
9199_{19} 0 0 0 11 alt -
9209_{20} 22 22 22 22 alt -
9219_{21} 11 11 11 11 alt -
9229_{22} 11 11 11 11 alt -
9239_{23} 22 22 22 22 alt P
9249_{24} 0 0 0 11 alt -
9259_{25} 11 11 11 11 alt -
9269_{26} 11 11 11 11 alt -
9279_{27} 0 0 0 0 alt -
9289_{28} 11 11 11 11 alt -
9299_{29} 11 11 11 11 alt -
9309_{30} 0 0 0 11 alt -
9319_{31} 11 11 11 11 alt -
9329_{32} 11 11 11 11 alt -
9339_{33} 0 0 0 11 alt -
9349_{34} 0 0 0 11 alt -
9359_{35} 11 11 11 11 alt P
9369_{36} 22 22 22 22 alt -
9379_{37} 0 0 0 11 alt -
9389_{38} 22 22 22 22 alt -
9399_{39} 11 11 11 11 alt -
9409_{40} 11 11 11 11 alt -
9419_{41} 0 0 0 0 alt -
9429_{42} 11 11 11 11 non-q.alt! -
9439_{43} 22 22 22 22 q.alt -
9449_{44} 0 0 0 11 q.alt -
9459_{45} 11 11 11 11 q.alt -
9469_{46} 0 0 0 0 non-q.alt! -
9479_{47} 11 11 11 11 q.alt -
9489_{48} 11 11 11 11 q.alt -
9499_{49} 22 22 22 22 q.alt P

(alt=alternating,  q.alt=quasi alternating,  BP=braid positive, P=positive, SQ=strongly quasipositive,  QP=quasipositive)

References