1 Introduction
A fundamental question for any knot invariant asks which knots it detects, if any. The most famous open version of this question asks whether the Jones polynomial detects the unknot. In this paper, we study the closely related detection question for knot Floer homology and Khovanov homology, as well as for Khovanov–Rozansky’s HOMFLY homology.
Considerable attention has been paid to this question over the last twenty years, and yet we have only managed to prove that these homology theories detect six knots: the unknot [Reference Kronheimer and Mrowka38, Reference Ozsváth and Szabó47], the two trefoils and the figure eight [Reference Baldwin, Dowlin, Levine, Lidman and Sazdanovic1, Reference Baldwin and Sivek4, Reference Ghiggini18] and the two cinquefoils [Reference Baldwin, Hu and Sivek2, Reference Farber, Reinoso and Wang14]. Each of these detection results required substantial new ideas, which have in several cases reverberated far beyond knot detection, but one thing they have in common is that each (save for that of the unknot) relied crucially on the knot in question being fibered. This paper expands the knot detection landscape dramatically. In particular, we prove for the first time that knot Floer homology and Khovanov homology can detect non-fibered knots and that HOMFLY homology detects infinitely many knots.
Our detection results are summarized in the list below. See Figure 1 for diagrams of the knots in this list, which are each non-fibered of Seifert genus one. In particular, $\operatorname {Wh}^\pm (T_{2,3},2)$ is the 2-twisted Whitehead double of the right-handed trefoil with a positive or a negative clasp, respectively, and the $P(-3,3,2n+1)$ are pretzel knots. We prove that
-
○ Knot Floer homology detects $5_2$ and $\operatorname {Wh}^+(T_{2,3},2)$ .
-
○ Knot Floer homology detects membership in each of the sets
$$\begin{align*}\{ 15n_{43522}, \operatorname{Wh}^-(T_{2,3},2) \} \,\textrm{ and }\, \{ P(-3,3,2n+1) \mid n\in\mathbb{Z} \}.\end{align*}$$ -
○ Khovanov homology detects $5_2$ .
-
○ Khovanov homology together with the degree of the Alexander polynomial detects $P(-3,3,2n+1)$ for each $n\in \mathbb {Z}$ .
-
○ HOMFLY homology detects $P(-3,3,2n+1)$ for each $n\in \mathbb {Z}$ .
These new detection results rely on our surprising main result, Theorem 1.2, which gives a complete classification of what we call nearly fibered genus-1 knots in $S^3$ . We motivate and explain Theorem 1.2 below and then state precise versions of the detection results above. We next outline the proof of Theorem 1.2, which combines in novel ways arguments involving sutured manifolds [Reference Gabai16], involutions, the cyclic surgery theorem [Reference Culler, McA, Gordon and Shalen12] and foundational work of Birman and Menasco on braids [Reference Birman and Menasco7, Reference Birman and Menasco8]. Finally, we discuss applications of this theorem to problems in Dehn surgery, which are carried out in our papers [Reference Baldwin and Sivek5, Reference Baldwin and Sivek6]. Perhaps the most striking of these is our proof in [Reference Baldwin and Sivek5] that $0$ -surgery characterizes infinitely many knots, where this was previously only known for the unknot, trefoils and figure eight by Gabai’s celebrated 1987 work on the Property R Conjecture [Reference Gabai17].
1.1 Our results
Recall that knot Floer homology assigns to a knot $K\subset S^3$ a bigraded vector space over $\mathbb {Q}$ ,
where m and a are the Maslov and Alexander gradings, respectively. Letting
knot Floer homology detects the Seifert genus of K by the formula
[Reference Ozsváth and Szabó47]. Moreover, K is fibered if and only if
[Reference Ghiggini18, Reference Ni44]. The knot Floer homology detection results for the unknot, trefoils and figure eight follow readily from these properties, as the first is the only knot of genus zero and the others are the only fibered knots of genus one. Detection for the cinquefoils is substantially more involved [Reference Farber, Reinoso and Wang14] but also hinges on the fact that the cinquefoils are fibered.
We focus in this paper on what we call nearly fibered knots. These are non-fibered knots which are as close as possible, from the knot Floer homology perspective, to being fibered:
Definition 1.1. A knot $K \subset S^3$ is nearly fibered if $\dim \widehat {\mathit {HFK}}(K,g(K); \mathbb {Q}) =2.$
Our main result is the complete classification of genus-1 nearly fibered knots:
Theorem 1.2. If $K \subset S^3$ is a genus-1 nearly fibered knot, then K is one of the knots
shown in Figure 1, or the mirror of one of these knots.
The knot Floer homologies of these knots are displayed for reference in Table 1, with the computations explained in Appendix A. Together with Theorem 1.2 and the symmetry
under taking mirrors, these computations immediately imply the promised detection results for knot Floer homology, stated as Theorems 1.3, 1.4 and 1.5 below. The first of these makes precise our claim that knot Floer homology detects the knots $5_2$ and $\operatorname {Wh}^+(T_{2,3},2)$ :
Theorem 1.3. Let $K\subset S^3$ be a knot, and let $J\in \{5_2,\operatorname {Wh}^+(T_{2,3},2)\}$ . If
as bigraded vector spaces, then $K=J$ .
The next two theorems make precise our claim that knot Floer homology detects membership in each of the sets
Note from Table 1 that knot Floer homology cannot distinguish the knots in either set.
Theorem 1.4. Let $K\subset S^3$ be a knot, and let $J\in \{ 15n_{43522}, \operatorname {Wh}^-(T_{2,3},2) \}$ . If
as bigraded vector spaces, then $K\in \{ 15n_{43522}, \operatorname {Wh}^-(T_{2,3},2) \}$ .
Theorem 1.5. Let $K\subset S^3$ be a knot, and let $J\in \{P(-3,3,2n+1)\mid n\in \mathbb {Z}\}$ . If
as bigraded vector spaces, then $K\in \{P(-3,3,2n+1)\mid n\in \mathbb {Z}\}$ .
As alluded to above, Theorem 1.3 is the first result which shows that knot Floer homology can detect non-fibered knots. We note that it is also the first knot Floer detection result for knots whose Floer homology is not thin (i.e., not supported in a single $\delta = m-a$ grading).
We now turn to our detection results for Khovanov homology. Recall that reduced Khovanov homology also assigns to a knot $K\subset S^3$ a bigraded vector space over $\mathbb {Q}$ ,
where h and q are the homological and quantum gradings, respectively. We use Theorem 1.2 together with Dowlin’s spectral sequence from Khovanov homology to knot Floer homology [Reference Dowlin13] to prove that reduced Khovanov homology detects $5_2$ :
Theorem 1.6. Let $K \subset S^3$ be a knot, and suppose that
as bigraded vector spaces. Then $K = 5_2$ .
As mentioned previously, Theorem 1.6 is the first result showing that Khovanov homology can detect non-fibered knots. Using the same strategy, we can also nearly show for the first time that Khovanov homology detects infinitely many knots:
Theorem 1.7. Let $K \subset S^3$ be a knot, and suppose for some $n\in \mathbb {Z}$ that
as bigraded vector spaces. If in addition the Alexander polynomial $\Delta _K(t)$ has degree 1, then $K=P(-3,3,2n+1)$ .
We expect that $\overline {\mathit {Kh}}(K;\mathbb {Q})$ alone should detect each of these pretzel knots. Indeed, their reduced Khovanov homologies are all 9-dimensional but (unlike their knot Floer homologies) are distinguished by their bigradings. The only remaining obstacle is to show that there are no fibered knots of genus at least two with the same reduced Khovanov homology as one of these pretzels. We are currently unable to show this, which is the reason for the additional Alexander polynomial hypothesis in Theorem 1.7.
However, we can achieve the desired detection result using the reduced version of Khovanov–Rozansky’s HOMFLY homology [Reference Khovanov and Rozansky35]. This theory assigns to a knot $K\subset S^3$ a triply-graded vector space over $\mathbb {Q}$ ,
which determines the HOMFLY polynomial of K. We use the fact that the HOMFLY polynomial encodes the Alexander polynomial, together with recent results of Wang [Reference Wang64], to bypass the obstacle described above and prove for the first time that HOMFLY homology detects infinitely many knots:
Theorem 1.8. Let $K \subset S^3$ be a knot, and suppose for some $n\in \mathbb {Z}$ that
as triply-graded vector spaces. Then $K=P(-3,3,2n+1)$ .
Remark 1.9. Some of the knots in Theorem 1.2 may be more familiar under other names. For instance, $6_1$ is the pretzel knot $P(-3,3,1)$ . The knot $15n_{43522}$ is one of the simplest hyperbolic knots, as tabulated in the census [Reference Champanerkar, Kofman and Mullen10], where it is labeled $k8_{218}$ . The twisted Whitehead doubles $\operatorname {Wh}^+(T_{2,3},2)$ and $\operatorname {Wh}^-(T_{2,3},2)$ appear in the tabulation [Reference Hoste, Thistlethwaite and Weeks28] as the knots $15n_{115646}$ and $16n_{696530}$ , respectively.
We outline our proof of Theorem 1.2 in some detail below. For the reader interested in fewer details, the key new idea is that if K is a genus-1 nearly fibered knot, then the fact that
is small allows us to determine the complement of a genus-1 Seifert surface F for K. This complement is not simply a product $F\times [-1,1]$ since K is not fibered, but work of Juhász [Reference Juhász32] provides us with product annuli that we can use to cut the complement into simpler pieces and identify it anyway. In each case, the complement of F admits an involution which extends over the complement of K, and by taking quotients, we can reduce the classification problem in Theorem 1.2 to a difficult but ultimately solvable question about 3-braids.
1.2 Proof outline
Let $K\subset S^3$ be a genus-1 nearly fibered knot, so that
Let F be a genus-1 Seifert surface for K. Let us identify a closed tubular neighborhood of F with the product $F\times [-1,1]$ , and consider the sutured Seifert surface complement
Then $S^3$ is recovered by gluing this neighborhood back in,
and K is the image of the suture
in this glued manifold. Our strategy is to first identify the complement $S^3(F)$ abstractly, in a way which does not remember its embedding into $S^3$ , and then classify the gluings that recover $S^3$ from this abstract point of view, so as to ultimately determine the knot K.
It will be helpful to consider the following slightly different perspective. Let
in which we identify each circle
with the corresponding $\partial D^2 \times \{t\}$ . Then $M_F$ has two toroidal boundary components and can be viewed as the 3-manifold obtained from $S^3_0(K)$ by removing a neighborhood of the capped off Seifert surface. This manifold contains a distinguished arc
whose complement recovers $S^3(F)$ , where the suture $s(\gamma )$ is identified with a meridian of $\alpha $ .
Work of Juhász [Reference Juhász30] tells us that the sutured Floer homology of $S^3(F)$ has dimension
This dimension is sufficiently small that another theorem of Juhász [Reference Juhász32] guarantees the existence of an essential product annulus A inside of $S^3(F)$ . Because F has genus $1$ , we can guarantee that the components of $\partial A$ are homologically essential in their respective copies of F, or equivalently in the tori of $\partial M_F$ , so by Dehn filling along curves dual to $\partial A$ , we can identify $M_F$ as the complement of a 2-component cable link, in which A is the cabling annulus.
A similar argument shows that the manifold obtained by decomposing $S^3(F)$ along A also contains an essential product annulus B, since such decompositions preserve the dimension of sutured Floer homology. We prove that B separates $S^3(F)\setminus A$ into two pieces and argue based on the dimensions of the sutured Floer homologies of these pieces that the component containing $\gamma $ must be a product sutured manifold. We then use this to show that the arc $\alpha $ in $M_F$ can be isotoped into the cabling annulus A.
It follows that the manifold obtained by cutting $M_F$ open along the cabling annulus A can alternatively be obtained by first removing a neighborhood of $\alpha $ to form $S^3(F)$ and then decomposing $S^3(F)$ along a product disk to remove the rest of the annulus A. Since $S^3(F)$ is a subset of $S^3$ , this cut-open manifold with torus boundary must then be the complement of a knot $C\subset S^3$ , with sutures isotopic to $\partial A$ . Moreover, its sutured Floer homology is also 2-dimensional, since product disk decomposition preserves dimension. Using this, we argue that C is an unknot or a trefoil and conclude the following:
Theorem 5.1. Up to orientation reversal, $M_F$ must be the complement of the $(2,4)$ -cable of either the unknot or the right-handed trefoil, and $\alpha $ is an arc in the cabling annulus.
This then gives us two possibilities (up to orientation reversal) for $S^3(F)$ , which we recall is obtained from $M_F$ by removing a neighborhood of $\alpha $ . The next important observation is that in both cases, there is an involution
which fixes $\gamma $ setwise and restricts to a hyperelliptic involution on the once-punctured tori $R_+(\gamma )$ and $R_-(\gamma )$ , as shown in Figures 7 and 18. The quotient of $S^3(F)$ by this involution is a sutured 3-ball with connected suture. It is natural to identify this quotient 3-ball with the complement of a thickened disk in $S^3$ ,
and the quotient map realizes $S^3(F)$ as the branched double cover of this ball along a tangle $\tau \subset S^3(D^2)$ , as shown in Figures 7 and 19.
As discussed at the beginning, $S^3$ is recovered by gluing $F\times [-1,1]$ to $S^3(F)$ by a map which in particular identifies $\partial F\times [-1,1]$ with $\gamma $ . For any such gluing map $\varphi $ , the facts that the once-punctured torus admits a unique hyperelliptic involution up to isotopy, and that this commutes with $\varphi $ up to isotopy – note that these facts require our assumption that $g(F)=1$ – imply that $\iota $ extends to an involution $\hat \iota $ of the glued manifold
whose restriction to the piece $F\times [-1,1]$ is a hyperelliptic involution on each $F\times \{t\}$ . The quotient map
therefore restricts on this piece to a branched double covering
along some 3-braid
It follows that $Y_\varphi $ is the branched double cover of
along the link $\tau \cup \beta $ . Moreover, K is the lift of the braid axis
in this double cover, as shown in Figures 7 and 8 in the case that $M_F$ is the complement of the $(2,4)$ -cable of the unknot. In particular, $Y_\varphi \cong S^3$ if and only if $\tau \cup \beta $ is an unknot.
This leads to our strategy for identifying K:
-
1. Identify all 3-braids $\beta $ such that $\tau \cup \beta $ is an unknot.
-
2. For each such $\beta $ , lift $\kappa $ to the branched double cover
$$\begin{align*}\Sigma_2(S^3, \tau \cup \beta) \cong \Sigma_2(S^3, U) \cong S^3, \end{align*}$$and this lift $\tilde \kappa $ is the corresponding knot K.
The first step is generally difficult and takes up a lot of room in this paper. Our approach is to find a crossing of $\tau $ whose various resolutions are all relatively simple and then understand surgeries between the branched double covers of these resolutions, making heavy use of the cyclic surgery theorem [Reference Culler, McA, Gordon and Shalen12] throughout. We eventually conclude the following:
Theorem 6.1. If $M_F$ is the complement of the $(2,4)$ -cable of the unknot, then K must be $5_2$ , $15n_{43522}$ or a pretzel knot $P(-3,3,2n+1)$ , up to mirroring.
Theorem 7.1. If $M_F$ is the complement of the $(2,4)$ -cable of the right-handed trefoil, then K must be a twisted Whitehead double $\operatorname {Wh}^{\pm }(T_{2,3},2)$ .
Given that taking the mirror of K corresponds to reversing the orientation of $M_F$ , this completes the proof of Theorem 1.2.
Remark 1.10. One of the main inspirations for this work and for our approach was a paper by Cantwell and Conlon [Reference Cantwell and Conlon9], who showed (among other things) that if K is either $5_2$ or $P(-3,3,2n+1)$ , then $M_F$ is the complement of the $(2,4)$ -torus link.
1.3 Other applications
One of the strengths of knot Floer homology is its relationship to the Heegaard Floer homology of Dehn surgeries on knots. Indeed, the fact that knot Floer homology detects the unknot can be used to give another proof that the unknot is uniquely characterized by each of its nontrivial Dehn surgeries (this was first proved via different but similar means by Kronheimer–Mrowka–Ozsváth–Szabó in [Reference Kronheimer, Mrowka, Ozsváth and Szabó37]). Likewise, Ozsváth–Szabó used the fact that knot Floer homology detects the trefoils and figure eight to prove that these knots are also characterized by each of their nontrivial surgeries [Reference Ozsváth and Szabó53].
In [Reference Baldwin and Sivek6], we use Theorem 1.2 to prove that Dehn surgeries of nearly all rational slopes uniquely characterize the knot $5_2$ :
Theorem 1.11 [Reference Baldwin and Sivek6].
Let $K \subset S^3$ be a knot, and suppose that r is a rational number for which there is an orientation-preserving homeomorphism
If r is not a positive integer, then $K=5_2$ .
This is the strongest result to date concerning characterizing slopes for any hyperbolic knot other than the figure eight. Note that we cannot hope to extend Theorem 1.11 to all positive integers, since, for example, $S^3_1(5_2) \cong S^3_1(P(-3,3,8)),$ as shown in [Reference Baldwin and Sivek6].
Using Theorem 1.11, we can then determine all of the ways in which the Brieskorn sphere $\Sigma (2,3,11)$ can arise from Dehn surgery on a knot in $S^3$ :
Theorem 1.12 [Reference Baldwin and Sivek6].
Given a knot $K \subset S^3$ and a rational number r, there exists an orientation-preserving homeomorphism
if and only if $(K,r)$ is either $(T_{-2,3},-\frac {1}{2})$ or $(5_2,-1)$ .
We note that similar results were achieved for $\Sigma (2,3,5)$ by Ghiggini in [Reference Ghiggini18], and for $\Sigma (2,3,7)$ by Ozsváth–Szabó in [Reference Ozsváth and Szabó53].
Similarly, the only knots for which 0-surgery was previously known to be characterizing are the unknot, trefoils and figure eight, by a 1987 theorem of Gabai [Reference Gabai17]. (This is an immediate corollary of Gabai’s proof that $S^3_0(K)$ determines the Seifert genus of K as well as whether or not K is fibered.) Combining the case $r=0$ of Theorem 1.11 with the main result of [Reference Baldwin and Sivek5] lets us add the infinitely many knots of Theorem 1.2 to this list.
Theorem 1.13 [Reference Baldwin and Sivek6, Reference Baldwin and Sivek5].
Let $K \subset S^3$ be a genus-1 nearly fibered knot. If for some knot $J \subset S^3$ there is an orientation-preserving homeomorphism
then $J=K$ .
1.4 Coefficients
Every Floer theory and link homology theory in this paper will be considered with coefficients in $\mathbb {Q}$ unless specified otherwise (as in Appendix A). For this reason, we will typically omit the coefficients from our notation for these theories going forward.
1.5 Organization
In §2, we review necessary background on sutured Floer homology. In §3–§5, we classify the possible pairs $(M_F,\alpha )$ , eventually proving Theorem 5.1. In §6, we determine the knots K arising when $M_F$ is the complement of a cabled unknot, proving Theorem 6.1. In §7, we do the same when $M_F$ is the complement of a cabled trefoil, proving Theorem 7.1. This proves Theorem 1.2, and the knot Floer homology detection results in Theorems 1.3, 1.4 and 1.5 follow immediately. In §8, we use Dowlin’s spectral sequence to prove the Khovanov homology detection results in Theorems 1.6 and 1.7. We then apply Theorem 1.7 in §9 to prove the HOMFLY homology detection result in Theorem 1.8. We finish with Appendix A, detailing the computations which appear in Table 1.
2 Sutured Floer homology background
In this section, we briefly review some facts about sutured Floer homology which will be of use in this paper, and establish some notation. See [Reference Gabai16, Reference Juhász29, Reference Juhász30] for more background.
Following Gabai [Reference Gabai16], a sutured manifold is a pair $(M,\gamma )$ , where M is a compact, oriented $3$ -manifold and $\gamma \subset \partial M$ is a union of annuli $A(\gamma )$ and tori $T(\gamma )$ , all of which are pairwise disjoint. We identify an oriented simple closed curve inside each annulus that is isotopic to the core of that annulus and take the sutures $s(\gamma )$ to be their union. We orient the components of $R(\gamma ) = \partial M - \operatorname {int}(\gamma )$ so that their boundary orientations agree with the orientations of $s(\gamma )$ and then let $R_+(\gamma )$ and $R_-(\gamma )$ consist of those components of $R(\gamma )$ whose orientations agree or disagree with the boundary orientation of $\partial M$ , respectively.
Juhász [Reference Juhász29, Definition 2.2] calls $(M,\gamma )$ a balanced sutured manifold if M has no closed components, the subsurfaces $R_+(\gamma )$ and $R_-(\gamma )$ have the same Euler characteristic, and every component of $\partial M$ contains an annulus of $A(\gamma )$ . In this case, the set of tori $T(\gamma )$ must be empty.
Sutured Floer homology, as defined by Juhász in [Reference Juhász29], assigns to a balanced sutured manifold $(M,\gamma )$ a vector space over $\mathbb {Q}$ ,
generalizing the hat version of Heegaard Floer homology. For example, given a knot $K \subset Y$ , we consider the sutured knot complement
whose sutures $s(\gamma _\mu )$ are the union of two oppositely oriented meridians of K. Moreover, given a Seifert surface F for K, we identify a closed tubular neighborhood of F with the product $F\times [-1,1]$ and define the sutured Seifert surface complement by
with suture
and
Then sutured Floer homology recovers the knot Floer homology of K, as well as its summand in the top Alexander grading with respect to F, by
as shown in [Reference Juhász29, Proposition 9.2] and [Reference Juhász30, Theorem 1.5], respectively.
Juhász also proved [Reference Juhász29, Reference Juhász30] that sutured Floer homology detects whether a balanced sutured manifold is taut and whether it is a product, as stated in Theorem 2.1 below. Recall for this theorem that a sutured manifold $(M,\gamma )$ is taut if it is irreducible and if $R(\gamma )$ is incompressible and Thurston norm-minimizing in
It is a product sutured manifold if it is of the form
with $s(\gamma ) = \partial \Sigma \times \{0\}$ , where $\Sigma $ is a compact, oriented surface with no closed components.
Theorem 2.1. Let $(M,\gamma )$ be a balanced sutured manifold.
-
○ If $(M,\gamma )$ is irreducible and not taut, then $\mathit {SFH}(M,\gamma ) \cong 0$ .
-
○ If $(M,\gamma )$ is taut, then $\dim \mathit {SFH}(M,\gamma )\geq 1$ .
-
○ If $(M,\gamma )$ is taut and not a product, then $\dim \mathit {SFH}(M,\gamma )\geq 2$ .
Proof. These claims are [Reference Juhász29, Proposition 9.18] (whose proof is attributed to Yi Ni), [Reference Juhász30, Theorem 1.4] and [Reference Juhász30, Theorem 9.7], respectively.
Remark 2.2. If $K\subset S^3$ is a knot and F is a genus-minimizing Seifert surface for K, then the sutured Seifert surface complement $S^3(F)$ is taut.
Sutured Floer homology behaves well with respect to sutured manifold decompositions
for certain surfaces $S\subset (M,\gamma )$ , as stated precisely in [Reference Juhász30, Theorem 1.3]. In this paper, we will be concerned with decompositions along:
-
○ product disks, which are properly embedded disks
$$\begin{align*}S \subset (M,\gamma) \end{align*}$$such that $\partial S$ meets the sutures $s(\gamma )$ in two points; and -
○ product annuli, which are properly embedded annuli
$$\begin{align*}S \subset (M,\gamma) \end{align*}$$such that $\partial S$ has one component in $R_+(\gamma )$ and the other component in $R_-(\gamma )$ .
The two theorems below state that sutured Floer homology is preserved under product disk decomposition, and under product annulus decomposition with mild additional hypotheses.
Theorem 2.3 [Reference Juhász29, Lemma 9.13].
Let $(M,\gamma )$ be a balanced sutured manifold. If $(M',\gamma ')$ is obtained by decomposing $(M,\gamma )$ along a product disk, then
Theorem 2.4 [Reference Juhász30, Lemma 8.9].
Let $(M,\gamma )$ be a balanced sutured manifold such that $H_2(M) \cong 0.$ Let
be a product annulus where at least one component of $\partial S$ is nonzero in $H_1(R(\gamma ))$ . If $(M',\gamma ')$ is obtained by decomposing $(M,\gamma )$ along S, then
Remark 2.5. These two theorems are closely related to the fact that decompositions along product disks and along product annuli preserve tautness [Reference Gabai16, Lemma 3.12].
We say that a product annulus $S\subset (M,\gamma )$ is essential if it is incompressible and if it is not isotopic to any component of $\gamma $ by an isotopy which keeps $\partial S$ in $R(\gamma )$ at all times. As discussed in §1.2, our proof of Theorem 1.2 relies on finding essential product annuli in the sutured complement of a genus-1 Seifert surface for a nearly fibered knot. Our main source of such annuli will be the following result:
Theorem 2.6. Let $(M,\gamma )$ be a taut balanced sutured manifold with $H_2(M) \cong 0$ , and suppose that $(M,\gamma )$ is not a product $(\Sigma \times [-1,1],\partial \Sigma \times [-1,1])$ in which $\Sigma $ is either an annulus or a pair of pants. If
and
then $(M,\gamma )$ contains an essential product annulus S.
Proof. Since $(M,\gamma )$ is taut and $\dim \mathit {SFH}(M,\gamma )<4$ , [Reference Juhász31, Corollary 2.2] says that $(M,\gamma )$ is horizontally prime (see [Reference Juhász31, Definition 1.7]). If $(M,\gamma )$ is also reduced, meaning that it does not contain an essential product annulus, and if it is not one of the forbidden products, then [Reference Juhász32, Theorem 3] says that
By hypothesis, this is not the case, so since $(M,\gamma )$ is not such a product, it is not reduced. (The products were not excluded in the statement of [Reference Juhász32, Theorem 3], but the proof assumes that there are no essential product disks in $(M,\gamma )$ , which by [Reference Juhász32, Lemma 2.13] holds if and only if $(M,\gamma )$ is not one of these products. See [Reference Ghosh and Li19, Remark 5.10].)
Lastly, we record the following for eventual use in our proof of Theorem 5.1.
Proposition 2.7. Let $K \subset S^3$ be a nontrivial knot, and let
denote the balanced sutured manifold whose sutures $s(\gamma _0)$ are a union of two oppositely oriented Seifert longitudes. Then
Proof. For any balanced sutured manifold $(M,\gamma )$ , a choice of homology orientation for the pair $(M,R_-(\gamma ))$ gives rise to an absolute lift of the relative $\mathbb {Z}/2\mathbb {Z}$ -grading on $\mathit {SFH}(M,\gamma )$ , and therefore to a well-defined Euler characteristic
for each $\mathfrak {s}\in \operatorname {Spin}^c(M,\gamma )$ , as described in [Reference Friedl, Juhász and Rasmussen15]. Fixing an $H_1(M)$ -affine isomorphism
these Euler characteristics can be packaged as an element
of the group ring $\mathbb {Z}[H_1(M)]$ .
Let us write $E_K = S^3 \setminus N(K)$ for convenience. Then
as shown in [Reference Friedl, Juhász and Rasmussen15, Example 8.1], which means that
for each $\mathfrak {s} \in \operatorname {Spin}^c(E_K,\gamma _0)$ . In particular, $\dim \mathit {SFH}(E_K,\gamma _0,\mathfrak {s})$ is always even.
Since K is nontrivial, its complement $E_K$ is irreducible. Thus, if we let
then [Reference Friedl, Juhász and Rasmussen15, Theorem 1.4] tells us that for all $\alpha \in H_2(E_K,\partial E_K;\mathbb {R})$ , we have
where $x^s$ is the sutured Thurston norm on $(E_K,\gamma _0)$ . If $\alpha $ is the class of a Seifert surface for K, with genus $g= g(K) \geq 1$ , then we compute by [Reference Friedl, Juhász and Rasmussen15, Lemma 7.3] that
and since this is nonzero, there must be two different $\operatorname {Spin}^c$ structures $\mathfrak {s}$ on $(E_K,\gamma _0)$ , each pairing differently with $\alpha $ , for which $\mathit {SFH}(E_K,\gamma _0,\mathfrak {s})$ is nonzero. But then $\mathit {SFH}(E_K,\gamma _0)$ has dimension at least two in each of these two $\operatorname {Spin}^c$ structures, so we conclude that
as desired.
3 Nearly fibered knots and essential annuli
Let $K\subset S^3$ be a nearly fibered knot of genus g, as in Definition 1.1. Then
Since this dimension is less than 4, [Reference Juhász31, Theorem 2.3] says that K has a unique genus-g Seifert surface F, up to isotopy. In this section, we will use Theorem 2.6 to study essential product annuli in the sutured Seifert surface complement
The lemma below guarantees the existence of such annuli with nice boundary properties.
Lemma 3.1. Let $K\subset S^3$ be a nearly fibered knot, and let F be a Seifert surface for K of genus $g = g(K)$ . Then there is an essential product annulus A in the sutured manifold
whose boundary components
are not both boundary-parallel in their respective surfaces $R_\pm (\gamma )$ .
Proof. Let us check that the hypotheses of Theorem 2.6 are met. First, note that $S^3(F)$ is not one of the excluded products, since $R_+(\gamma )= F\times \{1\}$ is not an annulus or pair of pants. Next, we have that
by Alexander duality. We also know that $S^3(F)$ is irreducible (in fact, this sutured manifold is taut, per Remark 2.2) and that
by (2.2). Note that $g\geq 1$ since the unknot is not nearly fibered. Therefore,
and so Theorem 2.6 provides an essential product annulus $A\subset (M,\gamma )=S^3(F)$ .
Let us suppose for a contradiction that both boundary components
of A are boundary-parallel in their respective surfaces. We recover the knot complement
from M by gluing $R_+(\gamma )$ to $R_-(\gamma )$ by some homeomorphism, and we can assume that this gluing map sends $A_+$ to $A_-$ since these curves are boundary-parallel in $R_\pm (\gamma )$ , respectively. Then A becomes a torus $T \subset E_K$ which meets F in a boundary-parallel circle.
We first claim that T is incompressible in $E_K$ . Indeed, its fundamental group is spanned by a longitude $\lambda $ of K and the image c of a curve
which is homologically essential in $E_K$ since it is dual to F. If some product $\lambda ^i c^j$ is nullhomotopic in $E_K$ , then its homology class satisfies
so it is a power $\lambda ^i$ of the longitude of K, but then $i=0$ since K is a nontrivial knot in $S^3$ . Therefore, $\lambda ^i c^j$ is nullhomotopic in T as well.
We next claim that T is not boundary-parallel. Indeed, if it were, then T and $\partial E_K$ would cobound a thickened torus intersecting F in a properly embedded annulus, in which case cutting $E_K$ back open along F would give a thickened annulus in $(M,\gamma )$ which is the trace of an isotopy between A and $\gamma $ that keeps $\partial A$ in $R(\gamma )$ at all times. But A is essential, which by definition implies that no such isotopy exists – a contradiction.
We have shown that under these circumstances, K must be a satellite knot, and the torus T splits its exterior into two pieces: the exterior $E_C$ of the companion C and the exterior $E_P$ of the pattern $P\subset S^1\times D^2$ . But then T splits the Seifert surface F into two pieces as well, one of which is an annulus in $E_P$ cobounded by the image of $A_\pm $ and the boundary $\partial F$ . This annulus gives an isotopy of the pattern P into T, where it is identified with a longitude of C, so P must be a cable pattern with winding number one. But this means that P is isotopic to the core of $S^1\times D^2$ , so T is boundary-parallel, and we have a contradiction. We conclude that $A_\pm $ cannot both be boundary-parallel, as desired.
3.1 The manifold $M_F$
While Lemma 3.1 applies to nearly fibered knots of any genus, we are especially interested in the genus-1 case. In this setting, we introduce the following construction, as in §1.2, which we will refer to repeatedly throughout the paper.
Definition 3.2. Let F be a genus-1 Seifert surface for a nontrivial knot $K\subset S^3$ . We define
to be the manifold obtained by gluing $D^2\times [-1,1]$ to $S^3(F)$ by a diffeomorphism
which preserves the interval coordinate. The boundary $\partial M_F$ is a disjoint union of two tori,
Let $\alpha $ be the properly embedded arc in $M_F$ given by
Then $(M,\gamma ) = S^3(F)$ is clearly recovered by removing the neighborhood
of $\alpha $ from $M_F$ , with suture $s(\gamma )$ given by the meridian
of the arc $\alpha $ .
As noted in §1.2, $M_F$ can also be described as the manifold obtained from the 0-surgery $S^3_0(K)$ by removing a tubular neighborhood of the torus $\hat {F}$ formed by capping off the Seifert surface F with a disk in the solid surgery torus. This perspective shows the following:
Lemma 3.3. Let F be a genus-1 Seifert surface for a nontrivial knot $K\subset S^3$ . Then the manifold $M_F$ is irreducible, and the tori $T_+$ and $T_-$ are incompressible.
Proof. [Reference Gabai17, Corollary 8.2] says that $S^3_0(K)$ admits a taut foliation with $\hat {F}$ a compact leaf. Cutting open along this leaf then gives a taut foliation on $M_F$ for which $T_\pm $ are compact leaves, from which the lemma follows.
We end this section with the following lemma:
Lemma 3.4. Let F be a genus-1 Seifert surface for a nearly fibered knot $K\subset S^3$ , and let
be the image of the annulus provided by Lemma 3.1 under the inclusion of $S^3(F)$ into $M_F.$ Then the boundary components
are each homologically essential in their respective tori $T_\pm $ .
Proof. Lemma 3.1 says that at least one of the boundary components of A, which we can take to be $A_+$ without loss of generality, is not boundary-parallel in $R_+(\gamma )$ , where
Since $R_+(\gamma )$ is a once-punctured torus in the case at hand, and the torus $T_+$ is obtained by capping off $R_+(\gamma )$ with a disk, it follows that $A_+$ is homologically essential in $T_+$ .
It remains to show that $A_-$ is homologically essential in $T_-$ . If not, then this means that $A_-$ must be boundary-parallel when viewed as a curve in the once-punctured torus $R_-(\gamma )$ . In this case, $A_-$ bounds the disk $D\subset T_-$ which caps off $R_-(\gamma )$ to form $T_-$ . Then the union
is a disk bounded by $T_+$ . Pushing this disk slightly into the interior of $M_F$ gives a compressing disk for $T_+$ . But this contradicts the fact that $T_+$ is incompressible, per Lemma 3.3. It follows that $A_-$ is homologically essential in $T_-$ , completing the proof of the lemma.
This lemma is notable in part for the following consequence, as mentioned in §1.2:
Remark 3.5. It follows from Lemma 3.4 that if F is a genus-1 Seifert surface for a nearly fibered knot, then $M_F$ is the complement of a 2-component cable link in some 3-manifold, with
being the cabling annulus. Indeed, since the curves $A_\pm $ are homologically essential in $T_\pm $ , there are curves $c_\pm \subset T_\pm $ which are homologically dual to $A_\pm $ . Then $M_F$ is the complement
where Y is the closed 3-manifold obtained by Dehn filling the tori $T_\pm $ along the curves $c_\pm $ , and L is the 2-component link given by the union of the cores of the solid tori in this filling. Recall that our eventual goal is to prove that $M_F$ is the complement of 2-component cables of the unknot or trefoils, per Theorem 5.1.
4 On the manifold $M_F$ and the arc $\alpha $
Let $K \subset S^3$ be a nearly fibered knot, with a Seifert surface F of genus 1. Let
be the arc in Definition 3.2 whose complement recovers $S^3(F)$ . Per Remark 3.5, $M_F$ is the complement of a 2-component cable link, with cabling annulus
as provided in Lemma 3.4. By construction, $\alpha $ is disjoint from A. Our goal in this section is to prove that it can be isotoped to lie in this cabling annulus, however. This is a key step toward our eventual classification of $M_F$ and thus $S^3(F)$ in the next section.
Proposition 4.1. Let F be a genus-1 Seifert surface for a nearly fibered knot $K \subset S^3$ . Let
be the annulus provided by Lemma 3.4. Then the arc $\alpha $ admits an isotopy, keeping $\partial \alpha $ in $\partial M_F$ at all times, which carries $\alpha $ to a properly embedded arc in A.
Proposition 4.1 will follow from a combination of several lemmas in this section. To start, note that we can view $M_F$ as a (non-balanced) sutured manifold $(M_F,\gamma _F)$ , where $\gamma _F = A(\gamma _F) \sqcup T(\gamma _F)$ is empty and the two boundary tori $T_\pm $ are oriented so that
Choose an orientation for A and consider the sutured manifold decomposition
along A, illustrated in Figure 2. In particular,
is a balanced sutured manifold with torus boundary, whose sutures $s(\gamma _A)$ are the union of two oppositely oriented curves of the same slope as the boundary components of A.
Since $\alpha $ is disjoint from A in $M_F$ , we can also view $\alpha $ as a properly embedded arc in $M_A$ . From this perspective, we then define the sutured arc complement
pictured on the right side in Figure 2, where
is a neighborhood in $\partial M_\alpha $ of the meridian $\mu _\alpha $ of $\alpha $ .
Note that $(M_\alpha ,\gamma _\alpha )$ can alternatively be obtained from $(M,\gamma ) = S^3(F)$ via sutured manifold decomposition
along the product annulus A (to be precise, the annulus whose image in $M_F$ is A), and the image of $\gamma $ under this decomposition is $N(\mu _\alpha )$ . It follows from Lemma 3.4 that at least one (in fact, both) of the boundary components of
is homologically essential in $R(\gamma )$ . Moreover, we have by Alexander duality as in (3.1) that
The product annulus decomposition in (4.1) therefore preserves sutured Floer homology,
by Theorem 2.4. Since $S^3(F)$ is taut, it follows that $(M_\alpha ,\gamma _\alpha )$ is taut as well (Remark 2.5).
Lemma 4.2. We have $H_2(M_\alpha ;R) \cong 0$ and $H_2(M_A;R) \cong 0$ for any commutative ring R.
Proof. Forgetting about the sutures, note that the Seifert surface complement
can be recovered from $M_\alpha $ by gluing a thickened annulus $N(A')$ along $\gamma _A$ by a map which identifies $\partial A'$ with $s(\gamma _A).$ The Mayer–Vietoris sequence associated to the decomposition
with coefficients in R (which we momentarily suppress for convenience) reads in part
We have that
by Alexander duality, so the leftmost portion of the sequence tells us that $H_2(M_\alpha ;R) \cong 0$ , proving the first claim.
Moreover, the map $H_1( \gamma _A;R) \to H_1(N(A');R)$ sends the class $[s(\gamma _A)]$ to
Since the rightmost map in the sequence is injective, and
it follows that
Note that the meridian $\mu _\alpha $ of $\alpha $ and the sutures $s(\gamma _A)$ cobound the pair of pants $R_+(\gamma _\alpha )\subset \partial M_\alpha .$ It follows that
and therefore that
To prove the second claim, note that $M_A$ is recovered from $M_\alpha $ by gluing back the neighborhood $N(\alpha )$ along the annular neighborhood $N(\mu _\alpha )$ of $\mu _\alpha $ ,
Let us consider the Mayer–Vietoris sequence corresponding to this decomposition. Since
the portion of the sequence beginning at $H_2(M_\alpha ;R) \oplus H_2(N(\alpha );R)$ has the form
with $H_1(N(\mu _\alpha );R)$ generated by the class $[\mu _\alpha ]$ . Then it follows from (4.3) that the rightmost map is injective, and we conclude by exactness that $H_2(M_A;R) \cong 0$ , as desired.
The next lemma provides the product annulus B mentioned in §1.2:
Lemma 4.3. There exists an essential product annulus $B \subset (M_\alpha ,\gamma _\alpha )$ .
Proof. We know that $(M_\alpha ,\gamma _\alpha )$ is a taut balanced sutured manifold, with $H_2(M_\alpha ;\mathbb {Z})=0$ by Lemma 4.2, and its boundary $\partial M_\alpha $ is a connected genus-2 surface. Then
Given the product annulus B from Lemma 4.3, let us denote its boundary circles by
Neither $B_+$ nor $B_-$ bounds a disk in $R(\gamma _\alpha )$ , since B is essential and hence incompressible. It follows that $B_+$ and $B_-$ are each boundary-parallel curves in the pairs of pants $R_+(\gamma _\alpha )$ and $R_-(\gamma _\alpha )$ , respectively. In particular, $B_\pm $ are each isotopic in $\partial M_\alpha $ either to a component of $s(\gamma _A)$ or to the meridian $\mu _\alpha $ of $\alpha $ . We rule out the latter possibility below:
Lemma 4.4. Neither $B_+$ nor $B_-$ is isotopic in $\partial M_\alpha $ to the meridian $\mu _\alpha $ of $\alpha $ .
Proof. Suppose that $B_+$ is isotopic in $\partial M_\alpha $ to $\mu _\alpha $ but $B_-$ is not. From the discussion above, $B_-$ must then be isotopic in $\partial M_\alpha $ to a component of $s(\gamma _A)$ . Recall that $M_A$ is obtained from $M_\alpha $ by gluing back a thickened disk (namely, the neighborhood $N(\alpha )$ ) along a neighborhood of the meridian $\mu _\alpha $ , as in (4.4). It follows that under the inclusion
the boundary component $B_+$ of the annulus B gets capped off with a disk D, so that
is a disk bounded by the curve $B_-\subset \partial M_A$ . This disk then gives rise under the inclusion
to a disk in $M_F$ bounded by the image
But $B_-$ is isotopic in $T_-$ to the boundary component $A_-$ of the annulus A, which by Lemma 3.4 is homologically essential. The fact that this curve bounds a disk in $M_F$ then contradicts the fact that the torus $T_-$ is incompressible, as shown in Lemma 3.3.
Swapping the roles of $B_+$ and $B_-$ leads to the same contradiction, so let us now assume that the curves $B_\pm $ are both are isotopic in $\partial M_\alpha $ to $\mu _\alpha $ . In this case, reversing the decompositions $S^3(K) \stackrel {F}{\leadsto } S^3(F) \stackrel {A}{\leadsto } (M_\alpha ,\gamma _\alpha )$ , we can glue $B_+$ to $B_-$ to turn the annulus B into a closed, embedded surface $\Sigma _B$ in $S^3(K)$ that meets F transversely in a single boundary-parallel curve. Then $\Sigma _B$ must be a torus, since if it were a Klein bottle, it could not embed in $S^3(K) \subset S^3$ ; as a torus in $S^3$ , it must bound a solid torus $V_B$ on one side or the other.
If $V_B \subset S^3$ were contained in the knot complement $S^3(K)$ , then
would be a properly embedded, punctured torus (consisting of F minus a collar neighborhood of its boundary) in the solid torus $V_B$ ; but then it must compress inside $V_B$ and hence in $S^3(K)$ , contradicting the incompressibility of F. Thus, $V_B$ must not lie entirely in $S^3(K)$ , and this means that it must contain $\partial \big (S^3(K)\big ) = \partial N(K)$ as well as the knot K. We now argue exactly as in the proof of Lemma 3.1: the torus $\partial V_B = \Sigma _B$ must be incompressible in $S^3(K)$ , realizing K as a satellite knot, but then the annulus $F \cap V_B$ provides an isotopy from K to its companion knot, so the satellite pattern must have been trivial. This means that $\Sigma _B = \partial V_B$ is boundary-parallel in $S^3(K)$ . Decomposing again along F and then A, we conclude that our original annulus B must have been parallel to an annular neighborhood of $\mu _\alpha $ in $\partial M_\alpha $ . But this contradicts the claim from Lemma 4.3 that B is essential, so we are done.
The proof of Lemma 4.4 in the case where both of $B_\pm $ are isotopic to $\mu _\alpha $ was substantially longer in the original version of this paper; we thank one of the referees for providing the much simpler argument used here.
Lemma 4.5. The annulus B separates $M_A$ , and its oriented boundary meets the torus $\partial M_A$ in a pair of parallel but oppositely oriented essential curves.
Proof. Let us orient B as well as its boundary curves $B_+$ and $B_-$ so that
Recall from Lemma 4.2 that $H_2(M_A) = 0$ . Therefore, the long exact sequence of the pair $(M_A,\partial M_A)$ reads in part
If B is nonseparating in $M_A$ , then it is nonzero in $H_2(M_A,\partial M_A)$ . It then follows from the exact sequence above that the class $[\partial B]$ is nonzero in $H_1(\partial M_A)$ , and hence that
Let us suppose for a contradiction that this is the case.
As discussed before Lemma 4.4, $B_+$ and $B_-$ are each isotopic in $\partial M_\alpha $ either to components of the sutures $s(\gamma _A)$ or to a meridian of the arc $\alpha $ , as unoriented curves. We ruled out the latter possibility in Lemma 4.4. Therefore, when viewed as curves in $\partial M_A$ , $B_\pm $ are each isotopic to components of $s(\gamma _A)$ (and are thus core circles of $R_\pm (\gamma _A)$ ). In particular, $B_+$ and $B_-$ are isotopic to one another as unoriented curves in $\partial M_A$ . Given (4.5), it must therefore be the case that $B_+$ and $B_-$ are parallel, oppositely oriented curves in $\partial M_A$ .
Forgetting their orientation, these curves cobound an annulus in $\partial M_A$ , whose union with B is then a Klein bottle
Since $M_A$ is orientable, the Klein bottle $\Sigma $ must be one-sided and in particular nonseparating. This implies that the mod- $2$ intersection pairing
is nonzero. But this contradicts the fact that $H_2(M_A;\mathbb {Z}/2\mathbb {Z})=0$ , by Lemma 4.2. Therefore, $[B_+] = [B_-]$ , and then B has the desired properties.
Lemma 4.6. The arc $\alpha \subset M_A$ can be isotoped rel endpoints so that it lies in $\partial M_A$ and meets the sutures $s(\gamma _A)$ transversely in a single point.
Proof. Lemma 4.5 implies that decomposing $(M_\alpha ,\gamma _\alpha )$ along the product annulus B produces a disconnected balanced sutured manifold
where we have labeled the components so that $(M_2,\gamma _2)$ has two sutures and $(M_3,\gamma _3)$ has three, as depicted in Figure 3.
Indeed, in $M_A$ , the components of $\partial B$ are core circles of the annuli $R_+(\gamma _A)$ and $R_-(\gamma _A)$ , so decomposing $(M_A,\gamma _A)$ along the separating B produces a disjoint union of two sutured manifolds, with two sutures each,
One of these components is disjoint from the arc $\alpha $ , so we label it $(M_2,\gamma _2)$ . We then remove a tubular neighborhood of $\alpha $ from the other component $(M',\gamma ')$ and add a meridional suture $\mu _\alpha $ to get $(M_3,\gamma _3)$ .
Since the components of $\partial B$ are homologically essential in $R(\gamma _A)$ , we have that
by Theorem 2.4. Since the left side is 2-dimensional, per (4.2), it follows that
for some $i\in \{2,3\}$ . Then Theorem 2.1 tells us that the corresponding $(M_i,\gamma _i)$ is a product sutured manifold (note that $(M_i,\gamma _i)$ is taut since $(M_\alpha ,\gamma _\alpha )$ is taut, per Remark 2.5).
Suppose first that $(M_2,\gamma _2)$ is a product sutured manifold. Since $\partial M_2$ is a torus and the sutures $s(\gamma _2)$ consist of two parallel essential curves on this torus, $R_+(\gamma _2)$ is an annulus, and so there is a homeomorphism
But if this is the case, then B could have been isotoped onto the component of $\gamma _\alpha $ which became a component of $\gamma _2$ , by an isotopy keeping $\partial B$ in $R(\gamma _\alpha )$ at all times. This contradicts the fact that B is essential.
It follows that $(M_3,\gamma _3)$ is a product sutured manifold. Since $R_+(\gamma _3)$ is a pair of pants P, we have that
One component of the sutures $s(\gamma _3)$ is a meridian $\mu _\alpha $ of $\alpha $ , and $(M',\gamma ')$ is recovered by gluing back a thickened disk $D^2\times I$ along an annular neighborhood of this meridian. The meridian $\mu _\alpha $ corresponds to a certain boundary component of P. Letting
be the annulus formed by capping off this boundary component with a disk, we then have the identification
where the arc $\alpha \subset M'$ is given by
for some point
as depicted in Figure 4.
The portion of $\partial M'$ which came from the annulus B (i.e., which was in the interior of $M_A$ ) is contained in a tubular neighborhood $N \subset \partial M'$ of one of the two components of $\gamma '$ – let us say the component
Since $\alpha $ is disjoint from B, we can isotope this arc into $\partial M' \setminus N$ while keeping its endpoints fixed, so that it meets the other component
of $\gamma '$ in an arc $\{\mathrm {pt}\} \times [-1,1]$ , as indicated in Figure 4. Gluing $(M',\gamma ')$ and $(M_2,\gamma _2)$ back together to form $(M_A,\gamma _A)$ , this gives an isotopy in $M_A$ which fixes the endpoints of $\alpha $ while carrying $\alpha $ to an arc in $\partial M_A$ which meets the sutures $s(\gamma _A)$ in one point.
With Lemma 4.6 in hand, we may now complete the proof of Proposition 4.1.
Proof of Proposition 4.1.
Viewing $\alpha $ as an arc in $M_A$ , Lemma 4.6 says that we can isotope it rel its endpoints to lie in $\partial M_A$ , so that it meets the sutures $s(\gamma _A)$ transversely in a single point, as shown on the left side of Figure 5.
Recall that $M_A$ was formed from $M_F$ by removing the interior of a tubular neighborhood $A \times [-1,1]$ , where the original cabling annulus A is identified as $A\times \{0\}$ . We can arrange the interval coordinate so that $\alpha \subset A\times \{1\}$ , and then the desired isotopy is simply $\phi _t(x) = (x,1-t)$ for $x\in \alpha $ .
5 Identifying the manifolds $M_F$ and $S^3(F)$
Let $K \subset S^3$ be nearly fibered, with a genus-1 Seifert surface F. According to Proposition 4.1, we can assume that the arc
in Definition 3.2, whose complement recovers $S^3(F)$ , lies in the annulus
of Lemma 3.4; moreover, Remark 3.5 says that A is a cabling annulus. In this section, we use these facts to identify the manifold $M_F$ and hence the sutured Seifert surface complement $S^3(F)$ . Specifically, we prove the following:
Theorem 5.1. Let $K\subset S^3$ be a nearly fibered knot, with genus-1 Seifert surface F. Then, up to possibly replacing K with its mirror, the manifold $M_F$ is the complement of either
-
1. the $(2,4)$ -cable of the unknot in $S^3$ , or
-
2. the $(2,4)$ -cable of the right-handed trefoil in $S^3$ .
In each case, the arc $\alpha $ is a properly embedded arc in the cabling annulus.
Proof. As defined in §4, the manifold $(M_A,\gamma _A)$ is obtained from $(M_F,\gamma _F)$ by decomposing along the cabling annulus A provided in Lemma 3.4,
Recall from Definition 3.2 that
can be recovered from $M_F$ by removing a neighborhood $N(\alpha )$ of the arc $\alpha $ , where the suture $s(\gamma )$ is identified with a meridian $\mu _\alpha $ of $\alpha $ . By Proposition 4.1, we can assume that $\alpha \subset A$ . Therefore, when we remove a neighborhood of $\alpha $ from $M_F$ to form $S^3(F)$ , what remains of the cabling annulus A is a product disk $D\subset S^3(F)$ . Thus, $(M_A,\gamma _A)$ can alternatively be obtained via the product disk decomposition
as indicated in Figure 6.
This shows in particular that $M_A$ is a subset of $S^3$ , as
Since $M_A$ has torus boundary, it follows that $M_A$ can be identified with the complement of a knot $C\subset S^3$ and moreover that we have an identification of sutured manifolds,
where the sutures $s(\gamma _r)$ are a union of two parallel oppositely oriented curves of slope r, with respect to the Seifert framing of C. Furthermore, we have
by Theorem 2.3. It remains to determine the slope r and the knot C.
Suppose first that $r=0$ . Then C is the unknot because otherwise, we would have
by Proposition 2.7, contradicting (5.1). But then $M_F$ is the complement of the $(2,0)$ -cable of the unknot in $S^3$ , which contradicts the fact in Lemma 3.3 that $M_F$ is irreducible.
The above argument shows that $r\neq 0$ . Note that we can identify $(S^3\setminus N(C),\gamma _r)$ as the sutured complement of the core $C' \subset S^3_r(C)$ of r-surgery on C, whose sutures are a union of two oppositely oriented meridians of $C'$ . With this in mind, equation (2.1) becomes
Since $r\neq 0$ , the core $C'$ is rationally nullhomologous in $S^3_r(C)$ . It follows that there is a spectral sequence
leading to the chain of inequalities
We conclude that
as this dimension has the same parity as
It also has the same parity as
which then implies that
Combining (5.2) and (5.3), we have shown that $S^3_r(C)$ is an L-space. Moreover, if $r=p/q$ with $q\geq 0$ and $\gcd (p,q)=1$ , then $|p|=2$ .
We now recall from [Reference Ozsváth and Szabó54, Proposition 9.6] (see [Reference Hom27, §2] for details) that if $C \subset S^3$ is a nontrivial knot, then r-surgery on C can only be an L-space if
Moreover, if we also have that $r>0$ , then C must additionally be fibered [Reference Ghiggini18, Reference Ni44] and strongly quasipositive [Reference Hedden24]. Note that when C is knotted, we have that
for slopes $r=\pm 2/q$ unless $q=1$ , so there are three cases to consider:
-
1. C is an unknot and $r=2/q$ for some odd $q\in \mathbb {Z}$ .
-
2. C is knotted and $r=2$ . Then $S^3_2(C)$ is an L-space, so $g(C) = 1.$ Then C must be the right-handed trefoil since this is the only genus-1, fibered, strongly quasipositive knot in the 3-sphere.
-
3. C is knotted and $r=-2$ . Then $S^3_{-2}(C)$ is an L-space, so again, $g(C) \leq 1$ . But now C must be the left-handed trefoil since its mirror $\overline {C}$ admits a positive L-space surgery and is therefore the right-handed trefoil, as discussed above.
In case (1), it follows that
since any two choices of $\gamma _{2/q}$ are related by a homeomorphism of the solid torus $S^3 \setminus N(U)$ . We conclude that
Similarly, in case (2), we have that
and therefore conclude that
This leaves only case (3), in which
Then we have
But in this case, we can replace K with its mirror $\overline {K}$ , and doing so replaces $M_F$ with $-M_F$ . So again, case (2) applies here, and we are done.
6 The $(2,4)$ -cable of the unknot
In this lengthy section, we determine all knots $K\subset S^3$ which arise from the first case of Theorem 5.1, in which $M_F$ is the complement of the $(2,4)$ -cable of the unknot. Our goal is to prove the following:
Theorem 6.1. Let $K\subset S^3$ be a nearly fibered knot with genus-1 Seifert surface F, and suppose that
Then K is one of the knots
or their mirrors.
The key observation is that under the hypotheses of Theorem 6.1, $M_F$ admits an involution which is rotation by $180^\circ $ about an axis of symmetry containing the arc $\alpha \subset M_F$ . This then gives rise to an involution $\iota $ of the sutured Seifert surface complement
obtained by removing a neighborhood of $\alpha $ from $M_F$ , where $s(\gamma )$ is identified with a meridian $\mu _\alpha $ of $\alpha $ . This involution is depicted on the left side of Figure 7, while the right side illustrates the quotient
which is a sutured 3-ball with connected suture. As suggested by the figure, it is natural to identify this quotient 3-ball with the complement of a thickened disk in $S^3$ ,
and the quotient map realizes $S^3(F)$ as the branched double cover of this ball along a tangle $\tau \subset S^3(D^2)$ , as shown in Figure 7.
Now, under the identification
we can assume that $\iota $ restricts on each $\partial F\times \{t\}\subset \gamma $ to a rotation of $\partial F$ which is independent of t. Recall that $S^3$ is recovered by gluing $F\times [-1,1]$ back into
by a map which in particular identifies
via the identity. Any such gluing map
is then determined by its restrictions to the once-punctured tori
Note that $\iota $ restricts to a hyperelliptic involution on each of the once-punctured tori,
Pulling back the involution $\iota $ via the maps $\varphi _\pm $ then induces hyperelliptic involutions
which agree on the boundary under the canonical identification of these two surfaces. Since once-punctured tori admit unique hyperelliptic involutions up to isotopy, we can extend $\iota _\pm $ to all of $F\times [-1,1]$ by a map restricting to a hyperelliptic involution on each $F\times \{t\}$ .
In summary, we have shown that $\iota $ extends to an involution $\hat \iota $ of the glued manifold
whose restriction to the piece $F\times [-1,1]$ is a hyperelliptic involution on each $F\times \{t\}$ . The quotient map
therefore restricts on this piece to a branched double covering
along some 3-braid
It follows that $Y_\varphi $ is the branched double cover of
along the link $\tau \cup \beta $ . Moreover, K is the lift $\tilde \kappa $ of the braid axis
of $\beta $ in this double cover. Since $Y_\varphi \cong S^3$ if and only if $\tau \cup \beta $ is an unknot, we conclude the following:
Lemma 6.2. Suppose that $K\subset S^3$ is a nearly fibered knot with genus-1 Seifert surface F and that
Then there is a 3-braid $\beta \in B_3$ such that $\tau \cup \beta $ is an unknot in $S^3$ , and such that the lift
of $\kappa $ is isotopic to K.
Figure 8 shows an isotopy of the unknot $U=\tau \cup \beta $ into a simpler form, which we will use in the subsections below.
In the sequel, we will often write $K = K_\beta $ when K arises from a given braid $\beta \in B_3$ in the sense of Lemma 6.2. We write each 3-braid as a word in
where x and y denote positive crossings between the top two strands and the bottom two strands, respectively. The following observations will help simplify our analysis in the following subsections.
Lemma 6.3. Let $r: B_3 \to B_3$ be the map which reverses a braid word, defined recursively by
for any $g\in \{x^{\pm 1},y^{\pm 1}\}$ . If $\beta $ is a 3-braid for which $\tau \cup \beta $ is unknotted, then $\tau \cup r(\beta )$ is also unknotted and $K_\beta \cong K_{r(\beta )}$ .
Proof. We can rotate the diagram of the unknot $U=\tau \cup \beta $ on the right side of Figure 8 about a vertical axis, and this preserves the tangle $\tau $ and the linked curve $\kappa $ while replacing the braid $\beta $ with its reverse $r(\beta )$ . It follows that $\tau \cup r(\beta )$ is also unknotted, since it is isotopic to the unknot U. This isotopy also carries $\kappa $ to itself, so up to isotopy, $\kappa $ must lift to both $K_\beta $ and $K_{r(\beta )}$ in the branched double cover of U; hence, $K_\beta \cong K_{r(\beta )}$ .
Lemma 6.4. Let $m: B_3 \to B_3$ be the map which mirrors a braid word, defined recursively by
for any $g\in \{x^{\pm 1},y^{\pm 1}\}$ . If $\beta $ is a 3-braid for which $\tau \cup \beta $ is unknotted, then $\tau \cup \big (m(\beta )y\big )$ is also unknotted, and $K_{m(\beta )y}$ is the mirror of $K_\beta $ .
Proof. In Figure 9, we perform an isotopy of $U = \tau \cup \big (m(\beta )y\big )$ in the complement of $\kappa $ , and we quickly find ourselves with a mirror image (reflecting across the plane of the page) of the diagram used to recover $K_\beta $ . Thus, if $\tau \cup \beta $ is unknotted, then so is $\tau \cup \big (m(\beta )y\big )$ , and the unknot $\kappa $ for $\tau \cup \big (m(\beta )y\big )$ lifts to the mirror of the lift $K_\beta $ of the corresponding knot in the $\tau \cup \beta $ diagram.
We remark that the mirror of $\beta $ is equal to the reverse of $\beta ^{-1}$ (i.e., $m(\beta ) = r(\beta ^{-1})$ ).
Lemma 6.5. If $\beta \in B_3$ produces an unknot $U=\tau \cup \beta $ , then so does $y^a \beta y^{-a}$ for any $a \in \mathbb {Z}$ , and moreover, $K_{y^a\beta y^{-a}} \cong K_\beta $ .
Proof. It is straightforward to see that $\tau \cup (y\beta y^{-1})$ is isotopic to $\tau \cup \beta $ in the complement of $\kappa $ , so the lemma follows by induction on a.
We now outline the proof of Theorem 6.1.
Proof of Theorem 6.1.
By Lemma 6.2, it suffices to classify the braids $\beta \in B_3$ such that $U = \tau \cup \beta $ is unknotted and to determine $K = K_\beta $ for each of them.
Supposing that U is an unknot, in Subsection 6.1, we will identify an arc (see Figure 10) that lifts to a knot $\gamma $ in the branched double cover $\Sigma _2(U) \cong S^3$ . We will argue via the cyclic surgery theorem [Reference Culler, McA, Gordon and Shalen12] that $\gamma $ must be an unknot or a torus knot, and we will study various surgeries on $\gamma $ which must be lens spaces or connected sums of lens spaces. In Subsections 6.2 and 6.3, we will study the cases $\gamma \cong U$ and $\gamma \cong T_{p,q}$ separately, proving in Propositions 6.12 and 6.13 that $\beta $ must be one of
or
respectively, up to reversal and conjugation by powers of y. Lemmas 6.3 and 6.5 tell us that it is enough to consider these particular braids.
After classifying these braids, we devote Subsection 6.4 to determining the knot $K_\beta $ for each of
These cases occupy Propositions 6.21, 6.22 and 6.23, respectively, and they recover the knots $5_2$ , $P(-3,3,2n+1)$ and $15n_{43522}$ . The only remaining braids are
and
but then Lemma 6.4 says that the corresponding $K_\beta $ are the mirrors of knots which we already found, so the proof is complete.
The remainder of this lengthy section is devoted to proving the results cited in the proof of Theorem 6.1.
6.1 Resolutions and the 3-braid $\beta $
In Figure 10, we take a fixed crossing (indicated by a dashed arc) of the unknot diagram from Figure 8 and modify it in several ways, changing the crossing to produce a new knot $L^\beta $ and also resolving the crossing in two different ways to produce the links $L^\beta _0$ and $L^\beta _1$ . It is clear from the diagrams that $L^\beta $ is a two-bridge knot and that $L^\beta _0 \cong \widehat {\beta }$ and $L^\beta _1\cong \widehat {\beta y^{-1}}$ are both closures of 3-braids.
The dashed arc on the left side of Figure 10 lifts to a simple closed curve $\gamma $ in the branched double cover $\Sigma _2(U) \cong S^3$ . Then the Montesinos trick [Reference Montesinos40] says that $\Sigma _2(L^\beta )$ can be realized as a half-integral surgery on $\gamma $ :
(Indeed, the branch loci U and $L^\beta $ agree outside a neighborhood of the indicated arc, so $\Sigma _2(L^\beta )$ and $\Sigma _2(U)$ agree outside the branched double cover of that neighborhood, which in either case is a solid torus. This says that $\Sigma _2(\beta )$ comes from some surgery on $\gamma $ in $\Sigma _2(U) \cong S^3$ , and then it must be half-integral because the peripheral curves in $S^3 \setminus N(\gamma )$ whose fillings produce $\Sigma _2(U)$ and $\Sigma _2(L^\beta )$ have distance two in $\partial N(\gamma )$ .) Similarly, the $0$ - and $1$ -resolutions of that crossing correspond to consecutive integral surgeries on $\gamma $ , which are each distance-1 from the $\frac {2n+1}{2}$ -surgery corresponding to the crossing change: that is,
To see that $\Sigma _2(L^\beta _0)$ and $\Sigma _2(L^\beta _1)$ are homeomorphic to $S^3_n(\gamma )$ and $S^3_{n+1}(\gamma )$ , respectively, and not vice versa, we note that the ordered triple $(\Sigma _2(L^\beta ), \Sigma _2(L^\beta _1), \Sigma _2(L^\beta _0))$ forms a surgery triad [Reference Ozsváth and Szabó52, Proposition 2.1], meaning that these three manifolds are all Dehn fillings of $S^3\setminus N(\gamma )$ along oriented curves $\alpha , \alpha _1, \alpha _0 \subset \partial N(\gamma )$ such that
(Note that following [Reference Ozsváth and Szabó52, Figure 1], their ‘ $L_0$ ’ and ‘ $L_1$ ’ are our $L^\beta _1$ and $L^\beta _0$ .) Up to reversing the orientation of all three curves simultaneously, we can assume that $\alpha =(2n+1)\mu +2\lambda $ , where $\mu $ and $\lambda $ are a meridian and longitude of $\gamma $ and $\partial N(\gamma )$ is oriented so that $\mu \cdot \lambda =-1$ , and then there is no way to choose signs for $\alpha _1 = \pm (n\mu +\lambda )$ and $\alpha _0 = \pm ((n+1)\mu +\lambda )$ so that (6.3) is satisfied. However,
does satisfy (6.3), so $\Sigma _2(L^\beta _0)$ and $\Sigma _2(L^\beta _1)$ must correspond to n- and $(n+1)$ -surgeries in that order as claimed.
From this discussion, we immediately deduce the following.
Lemma 6.6. The knot $\gamma \subset \Sigma _2(U) \cong S^3$ is either an unknot or a nontrivial torus knot.
Proof. Since $L^\beta $ is a 2-bridge knot, we know that $\Sigma _2(L^\beta )$ is a lens space. But the cyclic surgery theorem [Reference Culler, McA, Gordon and Shalen12] says that a non-integral surgery on $\gamma \subset S^3$ can only produce a lens space if $\gamma $ is an unknot or a nontrivial torus knot $T_{p,q}$ .
We will handle each of the two possible outcomes of Lemma 6.6 separately in the following subsections. The remainder of this subsection is devoted to some computations that will prove useful in that work.
To set the stage, we cut the given 2-bridge diagram of $L^\beta $ along a pair of vertical lines passing just by $\beta $ on either side. Taking the double cover branched over each piece of $L^\beta $ in turn gives a genus-1 Heegaard splitting of $\Sigma _2(L^\beta )$ , illustrated in Figure 11.
The solid tori $S^1\times D^2$ on either side of this splitting would be glued together to form $S^1\times S^2$ if the braid $\beta $ were trivial. But in general, the effect of gluing the middle $T^2\times I$ to either $S^1\times D^2$ is to reparametrize its boundary: the braid generators x and y act as positive Dehn twists along essential curves in $S^1\times S^1$ , which we have labeled $c_x$ and $c_y$ and oriented in Figure 12. Gluing after this reparametrization produces the desired Heegaard splitting of $\Sigma _2(L^\beta )$ .
The braid generators x and y act on the homology of the leftmost $S^1\times D^2$ by
Equivalently, we can view them as fixing that $S^1\times D^2$ , but acting on the rightmost $S^1\times D^2$ by the inverse of the above action:
Thus, if we fix the ordered basis $([c_x],[c_y])$ , then the (left) action of $B_3$ on the rightmost $H_1(\partial (S^1\times D^2)) \cong \mathbb {Z}^2$ is given by a homomorphism
defined by
One can verify that this is well-defined since $\rho (xyx)=\rho (yxy) = \left (\begin {smallmatrix} 0 & 1 \\ -1 & 0 \end {smallmatrix}\right )$ , and that $\rho (\Delta ^2) = -I$ , where $\Delta ^2 = (xyx)^2 = (xy)^3$ is the full twist which generates the center of $B_3$ .
Lemma 6.7. The kernel of $\rho $ is generated by $\Delta ^4$ .
Proof. If $w \in B_3$ satisfies $\rho (w) = I$ , then the same is true for every conjugate of w, and Murasugi [Reference Murasugi42] showed that w is conjugate to one of
-
1. $\Delta ^{2d} xy^{-a_1} xy^{-a_2} \cdots xy^{-a_n}$ , where all $a_i$ are nonnegative and at least one is positive;
-
2. $\Delta ^{2d} y^m$ for some $m\in \mathbb {Z}$ ; or
-
3. $\Delta ^{2d} x^my^{-1}$ where $m=-1,-2,-3$ .
In the second and third cases, we compute that
so the only such braids in the kernel are $\Delta ^{4d} = \Delta ^{2\cdot 2d} y^0$ . For the first case, we have
and a straightforward induction on $n \geq 1$ shows that its entries are nonnegative integers and that the top right entry is strictly positive. In particular, it cannot be $\pm I$ since it is not diagonal, so
is not the identity either. We conclude that $\rho (w) = I$ if and only if w is conjugate to some power of $\Delta ^4$ , and then it must actually be that power of $\Delta ^4$ since $\Delta ^2$ is central.
Lemma 6.8. If the representation (6.4) satisfies
then we have $\Sigma _2(L^\beta ) \cong S^3_{b/d}(U)$ .
Proof. The curve $c_y$ bounds a disk in the rightmost $S^1\times D^2$ of Figure 11, so then $\beta \cdot [c_y] = b[c_x] + d[c_y]$ bounds a disk in the rightmost $(T^2\times I) \cup (S^1\times D^2)$ . Thus, we can obtain the branched double cover of $L^\beta $ by Dehn filling the leftmost $S^1\times D^2$ along $b[c_x]+d[c_y]$ . Thinking of the left $S^1\times D^2$ as the complement of an unknot in $S^3$ , the oriented curves $c_x$ and $c_y$ correspond to a meridian and longitude of that unknot, respectively, so this amounts to a Dehn filling of slope $\frac {b}{d}$ .
Lemma 6.9. We have $\operatorname {tr} \rho (\beta ) = 2 \pm |H_1(\Sigma _2(L^\beta _0);\mathbb {Z})|$ , where we define $|H_1|=0$ if $H_1$ is infinite.
Proof. Inspecting the diagram for $L_0^\beta \cong \widehat \beta $ in Figure 10, we see that its branched double cover admits an open book decomposition whose binding is the lift of the braid axis; the pages are punctured tori (i.e., the double cover of a disk with three branch points), and the monodromy acts on the homology of the pages by $\rho (\beta )$ . It follows that
Thus, if this order is finite, then it equals $\left |\det (\rho (\beta )-I)\right |$ , and otherwise, $\det (\rho (\beta )-I) = 0$ . Writing $\rho (\beta )=\left (\begin {smallmatrix}a&b\\c&d\end {smallmatrix}\right )$ with $ad-bc=1$ , we compute this order up to sign as
which is equal to $2-\operatorname {tr}(\rho (\beta ))$ , so $\operatorname {tr}(\rho (\beta )) = 2 \pm |H_1(\Sigma _2(L_0^\beta ))|$ , as claimed.
According to (6.1) and (6.2), we have some $n\in \mathbb {Z}$ such that
so $|H_1(\Sigma _2(L^\beta _0))| = |n|$ , and we can write the conclusion of Lemma 6.9 more simply as
Lemma 6.10. Let $\beta $ be a 3-braid such that the link $L^\beta $ of Figure 10 satisfies
Let $\bar {q}$ be any integer with $q\cdot \bar {q} \equiv 1\ \pmod {p}$ , and write
for some $r\in \mathbb {Z}$ . Then either
or
where $d\in \mathbb {Z}$ and $e\in \{0,1\}$ .
Proof. Suppose that we have
Then by Lemma 6.8 and the classification of lens spaces up to orientation-preserving homeomorphism, we must have
for some $k\in \mathbb {Z}$ . In this case, since $\det \rho (\beta ) = 1$ , we know that $\rho (\beta )$ must have the form
for some integers k and $\ell $ . These matrices factor exactly as in (6.5) and (6.6), completing the proof.
In either case of Lemma 6.10, we have
which by Lemma 6.9 is equal to $2 \pm |H_1(\Sigma _2(L^\beta _0);\mathbb {Z})|$ . In other words, we must have
which will be useful in the following subsections.
6.2 The case $\gamma =U$
For now, we suppose that the curve $\gamma \subset \Sigma _2(U) \cong S^3$ from Subsection 6.1 is unknotted. We recall from (6.1) that $\Sigma _2(L^\beta ) \cong S^3_{(2n+1)/2}(U)$ for some integer n. Thus, in Lemma 6.10, we can take
This gives
in the first case, and
in the second.
In each of (6.9) and (6.10), the braid $\beta $ is uniquely determined up to the value of $d\in \mathbb {Z}$ since Lemma 6.7 says that $\Delta ^4$ generates $\ker (\rho )$ . In fact, we can disregard the braids arising from (6.10) because up to conjugation by powers of y, they are all obtained by reversing the braids from (6.9): we have
It follows by Lemmas 6.3 and 6.5 that every knot K with $M_F \cong S^3 \setminus N(T_{2,4})$ and $\gamma $ unknotted has the form $K \cong K_\beta $ , where
is one of the braids in (6.9). Recalling that $\Delta ^2$ generates the center of $B_3$ , we can now rewrite them as
with Lemma 6.5 in mind.
Lemma 6.11. Suppose that $\beta $ is a 3-braid of the form (6.11) and that its closure $L_0^\beta = \hat \beta $ has branched double cover $S^3_n(U)$ . Then the following must be true:
-
○ If $n\neq \pm 1$ , then $6(2d+e) = (k+\ell )\pm 1$ .
-
○ If $n = \pm 1$ , then $6(2d+e) + n - (k+\ell ) \in \{-2,0,2\}$ .
Proof. Since $S^3_n(U)$ is a lens space, Hodgson and Rubinstein [Reference Hodgson and Rubinstein26] proved that it is the branched double cover of exactly one link in $S^3$ , which we know to be the $(2,n)$ torus link. Thus, $\hat \beta \cong T_{2,n}$ , and so Birman and Menasco [Reference Birman and Menasco8] proved that up to conjugacy, we have
Now we can read from (6.11) that $\beta $ has exponent sum
where $\varepsilon : B_3 \to \mathbb {Z}$ is the homomorphism defined by $\varepsilon (x)=\varepsilon (y)=1$ . This exponent sum is invariant under conjugation, so $\varepsilon (\beta )$ must also be equal to $n\pm 1$ if $n\neq \pm 1$ and one of $2,0,-2$ otherwise. The lemma follows immediately.
Proposition 6.12. Let $\beta \in B_3$ be a braid for which $U=\tau \cup \beta $ is unknotted and the curve $\gamma \subset \Sigma _2(U) \cong S^3$ is also unknotted. Up to reversal, there is some integer $a\in \mathbb {Z}$ such that $y^a \beta y^{-a}$ is one of the 3-braids
Proof. As discussed above, it suffices to consider $\beta $ as in (6.9). We fix $n\in \mathbb {Z}$ so that $\Sigma _2(L^\beta )$ , $\Sigma _2(L^\beta _0)$ , and $\Sigma _2(L^\beta _1)$ are all surgeries on $\gamma $ of slopes $\frac {2n+1}{2}$ , n and $n+1$ , respectively, as guaranteed by (6.1) and (6.2). Then in particular, $\Sigma _2(L^\beta _0) \cong S^3_n(U)$ , with first homology of order $|n|$ , so Lemma 6.9 now says that
for $\beta $ as in (6.9). After multiplying through by $(-1)^e$ , we have four cases, where in each case, we can determine the value of $e\in \{0,1\}$ from the sign of the constant term $(-1)^e\cdot 2$ . These cases are as follows:
Case 1: $n+3 + (k+\ell )(2n+1) = n + 2$ , so $e=0$ .
This simplifies to $(k+\ell )(2n+1) = -1$ , so $(k+\ell ,n)$ is either $(1,-1)$ or $(-1,0)$ . Then (6.11) becomes
respectively, where we have simplified the first braid using the relation $x^{-1}y^{-1}x = yx^{-1}y^{-1}$ . Lemma 6.11 says that $d=0$ in each case, so now (6.11) becomes
for some $a\in \mathbb {Z}$ .
Case 2: $n+3 + (k+\ell )(2n+1) = -n + 2$ , so $e=0$ .
After rearranging, we get
and $2n+1$ is nonzero, so we must have $k+\ell =-1$ . Now we apply Lemma 6.11 to see that if $n\neq \pm 1$ , then $12d = -1\pm 1$ , while if $n=\pm 1$ , then $12d + (\pm 1) - (-1) \in \{-2,0,2\}$ . Thus, in either case, $d=0$ , and so (6.11) becomes
Case 3: $n+3 + (k+\ell )(2n+1) = n - 2$ , so $e=1$ .
This simplifies to $(k+\ell )(2n+1) = -5$ , so $(k+\ell ,n)$ is one of $(5,-1)$ , $(-5,0)$ , $(1,-3)$ or $(-1,2)$ . In each of these cases, equation (6.11) and Lemma 6.11 give us
respectively. The third and fourth braids are ruled out by Lemma 6.11 because there is no such $d\in \mathbb {Z}$ , whereas the first and second braids must have $d=0$ and $d=-1$ , respectively. Thus, in the first case, we have
while we can rearrange the second case to get
Thus, up to conjugation by powers of y, the possible braids in this case are
Case 4: $n+3 + (k+\ell )(2n+1) = -n - 2$ , so $e=1$ .
This condition is equivalent to
and $2n+1$ is odd so it must be $\pm 1$ ; hence, $(k+\ell ,n)$ is either $(-5,0)$ or $(3,-1)$ . The first of these already appeared in case 3, leading to
In the second case, equation (6.11) becomes
while Lemma 6.11 says that $6(2d+1)+(-1)-3 \in \{-2,0,2\}$ ; hence, $d=0$ . Thus,
Thus, in this case, the possible braids all have the form
We now combine the lists of braids enumerated in (6.12), (6.13), (6.14) and (6.15) to see that for some $a\in \mathbb {Z}$ , the braid $y^a \beta y^{-a}$ must be one of
But we can eliminate $xy^{-1}$ and $x^{-1}y^{-1}$ from this list because filling the tangle $\tau $ in with either of these produces a right-handed trefoil, as shown in Figure 13.
The braid x is also redundant because if $y^a \beta y^{-a} = x$ , then
belongs to the family $x^n y^{-1} xy$ . Thus, we can remove it, and we are now left with exactly the list of braids promised in this proposition.
6.3 The case where $\gamma $ is a torus knot
In this subsection, we will suppose that $\gamma \cong T_{p,q}$ for some p and q. Our goal is to prove the following.
Proposition 6.13. Suppose that $\beta \in B^3$ is a 3-braid for which $U = \tau \cup \beta $ is unknotted, and the curve $\gamma \subset S^3$ is a nontrivial knot. Then for some $a\in \mathbb {Z}$ , we have
up to braid reversal.
We recall from Subsection 6.1 that $L^\beta $ is a 2-bridge link, so that
is a lens space. The only half-integral lens space surgeries on $\gamma \cong T_{p,q}$ are those of slopes $pq \pm \frac {1}{2}$ [Reference Moser41], so we must have $n+\frac {1}{2} \in \{pq-\frac {1}{2}, pq+\frac {1}{2}\}$ ; hence, exactly one of
occurs. These surgeries were determined by Moser [Reference Moser41, Proposition 3.2], though we follow the notational conventions of Gordon [Reference McA20, Corollary 7.4].
We now observe that whether $n=pq-1$ or $n=pq$ , we have found a 3-braid $\beta ' \in \{\beta ,\beta y^{-1}\}$ whose closure has branched double cover
which is not prime. A theorem of Kim and Tollefson [Reference Kim and Tollefson36] says that the link $\widehat {\beta '}$ is therefore a nontrivial connected sum
where $\Sigma _2(L_1) \cong S^3_{p/q}(U)$ and $\Sigma _2(L_2) \cong S^3_{q/p}(U)$ . Now since $L_1 \# L_2$ has braid index at most $3$ and the summands $L_i$ are nontrivial, the ‘braid index theorem’ of Birman and Menasco [Reference Birman and Menasco7] shows that $L_1$ and $L_2$ are each closures of 2-braids. Thus, we can write
where a and b cannot be $\pm 1$ or $0$ because the branched double covers are nontrivial rational homology spheres, and hence are neither $S^3$ nor $S^1\times S^2$ . Then we have
In particular, this is only possible if $|p| = |a|$ and $|q| = |b|$ , and if, moreover,
Lemma 6.14. Let $P,Q \geq 2$ be coprime positive integers satisfying
Then $P = Q \pm 1$ .
Proof. Write $P = kQ \pm 1$ , where $P,Q \geq 2$ implies that $k \geq 1$ . If $k \geq 2$ , then we have $P \geq 2Q-1$ , so either $P=3$ and then $Q=2$ (hence $P=Q+1$ ), or $P> 3$ and then we have
(The last two inequalities are equivalent to $P \geq 2Q-1$ and $P> 3$ , respectively.) But if $1 < Q < P-1$ , then we cannot possibly have $Q \equiv \pm 1\ \pmod {P}$ , so there are no other solutions with $k \geq 2$ , and thus, we must have $P=Q\pm 1$ .
Lemma 6.14 says that for $\gamma \cong T_{p,q}$ , if we write $P=|p|$ and $Q=|q|$ , then $P = Q \pm 1$ , and (6.16) and (6.17) tell us that either
is the branched double cover of a 3-braid, depending on whether $n=pq-1$ or $n=pq$ , respectively. Reversing orientation if exactly one of p and q is negative replaces that 3-braid with its mirror, which is still a 3-braid, and the surgered manifold is then
so in any case, we see that one of
is the branched double cover of a 3-braid. This gives us strong restrictions on P and Q by the following result of Murasugi.
Proposition 6.15 [Reference Murasugi43, Proposition 7.2].
Let $L_{r/s}$ be the 2-bridge link with branched double cover $L(r,s) = S^3_{r/s}(U)$ , where $0<s<r$ and s is odd. Then $L_{r/s}$ has braid index $2$ if and only if $s=1$ , and it has braid index $3$ if and only if either
-
1. there are integers $c,d> 0$ such that $(r,s) = (2cd+3c+3d+4, 2c+3)$ , or
-
2. there are $c,d>0$ such that $(r,s) = (2cd+c+d+1, 2c+1)$ .
Remark 6.16. We note that in the first and second cases of Proposition 6.15, we have
respectively, so if $L_{r/s}$ has braid index $3$ , then s divides either $2r+1$ or $2r-1$ . In particular, if the braid index is at most $3$ , then we can draw the same conclusion, since braid index $2$ implies $s=1$ .
Putting all of this together, we can now show the following.
Lemma 6.17. Suppose that the link $\tau \cup \beta $ is unknotted and that $\gamma $ is not an unknot. Then $\gamma $ or its mirror must be one of the torus knots $T_{2,3}$ , $T_{3,4}$ or $T_{4,5}$ , and $(\gamma ,n)$ must be one of the pairs indicated in Table 2.
Proof. Lemma 6.6 says that $\gamma \cong T_{p,q}$ for some p and q, and we have argued that if $P=|p|$ and $Q=|q|$ , then $P=Q\pm 1$ ; without loss of generality, we write $P=Q+1 \geq 3$ . We consider each parity of P separately and determine in each case which lens space $S^3_{(PQ\pm 1)/Q^2}(U)$ must arise as the branched double cover of a 3-braid. Up to orientation, we know that the corresponding $S^3_{(pq\pm 1)/q^2}(U)$ is either n-surgery (i.e., $\Sigma _2(L_0^\beta )$ ) or $(n+1)$ -surgery (i.e., $\Sigma _2(L_1^\beta )$ ) on $\gamma $ , so the value of n follows immediately, and then the precise lens spaces are determined by the relations
and the relations $S^3_{r/s}(U) \cong S^3_{r/(s+kr)}(U)$ and $S^3_{r/s}(U) \cong -S^3_{-r/s}(U)$ for all $r,s,k$ .
Case 1: P is odd. Then Q is even, so if $\epsilon = \pm 1$ , then
and $Q+\epsilon $ is odd. According to Murasugi’s result, and in particular Remark 6.16, it follows that
divides one of
hence, it also divides
Thus, s must be either $1$ or $3$ . Then $2 \leq Q = s-\epsilon $ says that $(P,Q)$ is either $(3,2)$ or $(5,4)$ . We determine the following possibilities:
-
○ If $s=1$ , then $(P,Q)=(3,2)$ and $\epsilon = -1$ , so the lens space in question is $L(5,4)$ .
-
○ If $s=3$ and $\epsilon =+1$ , then $(P,Q) = (3,2)$ , and the lens space is $L(7,4)$ .
-
○ If $s=3$ and $\epsilon =-1$ , then $(P,Q) = (5,4)$ , and the lens space is $L(19,16)$ .
Case 2: P is even. Then Q is odd, so if $\epsilon =\pm 1$ , then
(with $Q^2$ odd) arises as the branched double cover of a 3-braid closure if
This is equivalent to $Q^2$ dividing $2Q + (2\epsilon \pm 1) \leq 2Q+3$ , but given that Q is odd and $Q \geq 2$ , we have $Q^2> 2Q+3$ unless $Q=3$ . So $(P,Q) = (4,3)$ and $\epsilon =1$ , and the lens space in question must be $L(13,9)$ .
This completes the identification of the lens spaces in question when $\gamma = T_{p,q}$ and $p,q$ are both positive. If one of p and q is negative, then we can apply the same argument to the mirror of $\gamma $ to determine the value of $-n$ and the proposition follows.
In fact, we can rule out most of the pairs $(\gamma ,n)$ appearing in Lemma 6.17 as well.
Lemma 6.18. If $\gamma $ is a nontrivial torus knot, then $(\gamma ,n,k+\ell ,e)$ is either
Proof. In Table 3, we tabulate the possible pairs $(\gamma ,n)$ from Lemma 6.17, together with
-
○ the corresponding lens spaces
$$\begin{align*}\Sigma_2(L^\beta) \cong S^3_{(2n+1)/2}(\gamma) \cong L(p,q) := S^3_{p/q}(U) \end{align*}$$for some integers p and q; -
○ the integers p and q, as well as $\bar {q}$ such that $q\cdot \bar {q} \equiv 1\ \pmod {p}$ ; and
-
○ the resulting trace of $\rho (\beta )$ , as determined by (6.7), given that Lemma 6.10 says that $\rho (\beta )$ must have one of the two forms (6.5) or (6.6).
The lens spaces $\Sigma _2(L^\beta )$ in Table 3 are determined by the formulas
which again follow from [Reference Moser41] or [Reference McA20].
Lemma 6.9 tells us that $\operatorname {tr} \rho (\beta ) = 2\pm n$ , so we inspect Table 3 to see whether this is possible. We have
and the computations for $(T_{-2,3},-7)$ , $(T_{-3,4},-13)$ and $(T_{-4,5},-20)$ are identical, so there is no solution in any of these cases. This leaves only
with solution $(k+\ell ,e) = (-2,1)$ and
with solution $(k+\ell ,e) = (-1,0)$ .
Proposition 6.19. If $\gamma = T_{-2,3}$ , then up to reversal, there is some integer a such that
Proof. In this case, we have $(n,k+\ell ,e) = (-6,-1,0)$ and $\Sigma _2(L^\beta ) = L(11,3)$ by Lemma 6.18, so we can write
by (6.5) and (6.6). We compute that
and since $\ker (\rho )$ is generated by $\Delta ^4$ , it follows that
for some $d\in \mathbb {Z}$ . These two families of braids are reverses of each other since
so we need only consider the first family – namely,
In order to determine d, we recall that the link $L^\beta _1$ from Figure 10 is the closure of $\beta y^{-1}$ , and by (6.2), we have
As a lens space, this must be the branched double cover of a unique knot [Reference Hodgson and Rubinstein26], so we have $\widehat {\beta y^{-1}} \cong T_{2,5}$ . Then Birman and Menasco’s classification theorem from [Reference Birman and Menasco8] says that $\beta y^{-1}$ must be conjugate to $x^5 y^{\pm 1}$ , so that $\beta $ has exponent sum
However, we can read off the explicit form for $\beta $ above that $\varepsilon (\beta ) = 12d + 5$ , so we must have $d=0$ .
Proposition 6.20. If $\gamma = T_{2,3}$ , then up to reversal, there is some integer a such that
Proof. In this case, we have $(n,k+\ell ,e) = (5,-2,1)$ and $\Sigma _2(L^\beta ) = L(11,8)$ by Lemma 6.18, so we can write
by (6.5) and (6.6). We compute that
so now since $\Delta ^4$ generates $\ker (\rho )$ , we have
for some $d\in \mathbb {Z}$ .
In order to determine d, we note that the braid closure $\widehat \beta = L^\beta _0$ satisfies
so $L^\beta _0 \cong T_{-2,5}$ since every lens space is the branched double cover of a unique knot [Reference Hodgson and Rubinstein26]. Then $\beta $ must be conjugate to either $x^{-5}y$ or $x^{-5}y^{-1}$ [Reference Birman and Menasco8], so its exponent sum is $\varepsilon (\beta ) = -5 \pm 1$ . But in either of the above families, we have $\varepsilon (\beta ) = 12d+8$ , so in fact, $d=-1$ . Moreover, if we reverse the second family above, then we get
so up to reversal, it suffices to consider only the family of braids
We can simplify this somewhat by writing
and so
as claimed.
We can now complete the main result of this subsection.
Proof of Proposition 6.13.
Lemmas 6.17 and 6.18 tell us that if $\gamma $ is knotted, then it must be a trefoil. If it is a left-handed trefoil, then Proposition 6.19 says that up to reversal, $\beta $ is conjugate to $x^3 y^{-1} x^2 y$ by some power of y. Otherwise, it is a right-handed trefoil, so by Proposition 6.20, either $\beta $ or its reverse is conjugate to $x^{-3} y x^{-2}$ by some power of y.
6.4 Some knots arising from specific braids
In this subsection, we consider several families of 3-braids $\beta $ that arise in Propositions 6.12 and 6.13, producing unknots when inserted into the tangle $\tau $ of Figure 8. We will determine the corresponding nearly fibered knots $K = K_\beta $ which arise as lifts of $\kappa $ to $\Sigma _2(U) \cong S^3$ . The results are summarized in Table 4; the proofs in each case occupy Propositions 6.21, 6.22 and 6.23, respectively.
Proposition 6.21. The family of braids $\beta = y^a x^{-1} y^{-a}$ produces $K_\beta \cong 5_2$ .
Proof. By Lemma 6.5, it suffices to take $a=0$ , so $\beta = x^{-1}$ . We insert this into the tangle $\tau $ from Figure 8, apply an isotopy so that $U = \tau \cup \beta $ bounds a planar disk and $\kappa $ winds around it, and then cut $\kappa $ open along that disk and glue two copies together to construct the lift $K_\beta = \tilde \kappa $ . This process is illustrated in Figure 14, where we isotope $U \cup \kappa $ into a convenient position and then take the branched double cover with respect to U at the last step; the resulting diagram of $\tilde \kappa $ is isotopic to $5_2$ , as claimed.
Proposition 6.22. The braids $\beta = y^a x^n y^{-1} x y^{1-a}$ produce $K_\beta \cong P(-3,3,2n+1)$ .
Proof. Again, by Lemma 6.5, we need only consider $\beta = y x^n y^{-1} x$ . In Figure 15, we insert this braid into $\tau \sqcup \kappa $ and perform an isotopy so that the unknot $U = \tau \cup \beta $ clearly bounds a disk, and then in Figure 16, we use this to lift $\kappa $ to the knot $K_\beta = \tilde \kappa $ in the branched double cover $\Sigma _2(U) \cong S^3$ . In the end, we are left with a diagram of $P(3,-3,2n+1)$ , which is isotopic to $P(-3,3,2n+1)$ .
Proposition 6.23. The braids $\beta = y^a x^3 y^{-1} x^2 y^{1-a}$ produce $K_\beta \cong 15n_{43522}$ , possibly up to mirroring.
Proof. In this case, Lemma 6.5 says that we need only consider $\beta = x^3y^{-1}x^2 y$ , as shown in Figure 17. We can repeat the same procedure as in Propositions 6.21 and 6.22 to find $K_\beta $ , but this is not very enlightening because we find it hard to identify 15-crossing knots from their diagrams.
Instead, we ask SnapPy [Reference Culler, Dunfield, Goerner and Weeks11] to do the hard work for us: we give it the link $U\cup \kappa $ on the left side of Figure 17, do a $(2,0)$ -Dehn filling of U (i.e., an orbifold Dehn filling of U with meridional slope, so that U has cone angle $\pi $ ), and then look at the double covers of the result that are not themselves orbifolds. SnapPy can produce triangulations of these, and it identifies one of them as the complement of $15n_{43522}$ , so this must be $K_\beta $ .
Remark 6.24. SnapPy looks for isometries between a given pair of hyperbolic manifolds by first attempting to produce a canonical triangulation of each and then comparing the resulting triangulations combinatorially. Thus, when it succeeds, as in the proof of Proposition 6.23, the result is certifiably true: it has found identical triangulations of each, and it does not need any numerical approximation to verify that the triangulations agree.
As explained at the beginning of this section, this completes the proof of Theorem 6.1.
7 The $(2,4)$ -cable of the trefoil
In this section, we determine all knots $K\subset S^3$ which arise from the second case of Theorem 5.1, in which $M_F$ is the complement of the $(2,4)$ -cable of the right-handed trefoil. Our goal is to prove the following:
Theorem 7.1. Let $K\subset S^3$ be a nearly fibered knot with genus-1 Seifert surface F, and suppose that
Then K is one of the twisted Whitehead doubles
Just as in Section 6, we observe that under the hypotheses of Theorem 7.1, the sutured Seifert surface complement $S^3(F)$ admits an involution $\iota $ , illustrated in Figure 18, realizing this complement as the branched double cover of a sutured $3$ -ball along a tangle $\tau $ , as shown in Figure 19. The exact same reasoning as in the previous section then implies the following analogue of Lemma 6.2:
Lemma 7.2. Suppose that $K\subset S^3$ is a nearly fibered knot with genus-1 Seifert surface F and that
Then there is a tangle $\tau $ and a 3-braid $\beta \in B_3$ , depicted in Figure 20, such that $\tau \cup \beta $ is an unknot in $S^3$ , and such that the lift
of the pictured curve $\kappa $ is isotopic to K.
With Lemma 7.2 at hand, we are left to determine which braids $\beta $ cause $\tau \cup \beta $ to be unknotted. Supposing that it is indeed an unknot U, we choose a crossing in Figure 21, indicated by a red dashed arc, and produce two link diagrams $L^\beta $ and $L^\beta _0$ by changing that crossing and by taking its $0$ -resolution, respectively. We can see in Figure 21 that
where $\tau _{-1/4}$ and $\tau _{-1/7}$ are tangle diagrams differing only in the circled rational sub-tangles, having $-4$ and $-7$ half-twists, respectively.
Lemma 7.3. If $\tau \cup \beta $ is an unknot, then so are $\tau _{-1/4}\cup \beta $ and $\tau _{-1/7}\cup \beta $ .
Proof. Just as in Section 6, the Montesinos trick tells us that there is a curve $\gamma \subset \Sigma _2(U) \cong S^3$ and an integer $n\in \mathbb {Z}$ such that
Since $\Sigma _2(L^\beta )$ arises as non-integral surgery on a knot $\gamma \subset S^3$ , it must be irreducible [Reference McA and Luecke21]. But we also know that
and if this is irreducible, then the second summand must be $S^3$ , so then $\tau _{-1/4}\cup \beta $ must be unknotted [Reference Waldhausen63].
Now that we have $\Sigma _2(L^\beta ) \cong L(3,2) \cong S^3_{3/2}(U)$ arising from a non-integral surgery on $\gamma $ , of slope $\frac {2n+1}{2}$ , we know that $\gamma $ must be an unknot or a torus knot [Reference Culler, McA, Gordon and Shalen12]. In fact, it cannot be a nontrivial torus knot since otherwise, no surgery would produce a lens space of order 3 [Reference Moser41]. So $\gamma $ is an unknot, and then we must have $\frac {2n+1}{2} = \frac {3}{2}$ , or $n=1$ . But in this case, we have
so again by [Reference Waldhausen63], we can conclude that $\tau _{-1/7} \cup \beta \cong L^\beta _0$ is an unknot.
Lemma 7.4. If $\tau \cup \beta $ is an unknot, then the link $\tau _{1/0}\cup \beta $ depicted in Figure 22 is an unknot, and the 3-braid closure $\hat \beta $ is a 2-component unlink.
Proof. We take the tangles $\tau _{-1/4}$ and $\tau _{-1/7}$ in Figure 21 and replace their circled twist regions with rational tangles of slopes $\frac {1}{0}$ or $\frac {0}{1}$ to get the tangles $\tau _{1/0}$ and $\tau _{0/1}$ depicted in Figure 22, observing that
Lemma 7.3 says that $\tau _{-1/4} \cup \beta $ and $\tau _{-1/7} \cup \beta $ are both unknotted, so their branched double covers satisfy
In particular, if we remove the circled rational subtangles from either unknot, then the branched double cover of what remains is a knot complement $S^3 \setminus N(L)$ , and it has two different Dehn fillings (corresponding to the rational tangles of slopes $-\frac {1}{4}$ and $-\frac {1}{7}$ ) which both produce $S^3$ . Then L must be an unknot [Reference McA and Luecke22, Theorem 2], and the fillings that produce $\Sigma _2(\tau _{-1/4}\cup \beta )$ and $\Sigma _2(\tau _{-1/7}\cup \beta )$ must have slopes $\frac {1}{n}$ and $\frac {1}{n-3}$ for some $n\in \mathbb {Z}$ .
It follows that if we replace these rational tangles with one of slope $\frac {1}{0}$ , then this corresponds to a Dehn filling of $S^3 \setminus N(L)$ of slope $\frac {1}{n+4}$ , and then
since L is unknotted. We apply Waldhausen’s result [Reference Waldhausen63] once again to see that $\tau _{1/0} \cup \beta $ is an unknot.
Similarly, if we instead use the rational tangle that produces $\tau _{0/1} \cup \beta $ , then the corresponding Dehn filling of $S^3 \setminus N(L)$ is at distance one from both the $\frac {1}{n}$ - and $\frac {1}{n-3}$ -fillings, so it must have slope $\frac {0}{1}$ . In other words, we have shown that
But the only link in $S^3$ with branched double cover $S^1\times S^2$ is the two-component unlink [Reference Tollefson62], so this determines $\hat \beta $ up to isotopy.
We can now apply methods from Section 6 to determine all of the possible braids $\beta $ to which Lemma 7.2 might apply.
Proposition 7.5. If $\tau \cup \beta $ is unknotted, where $\tau $ is the tangle shown in Figure 20, then
for some $a\in \mathbb {Z}$ .
Proof. Lemma 7.4 tells us that the knot $\tau _{1/0} \cup \beta $ on the left side of Figure 22 is an unknot, with branched double cover $S^3$ . Using the representation $\rho : B_3 \to SL_2(\mathbb {Z})$ from (6.4), which was defined by
we apply Lemma 6.10 with $(p,q,\bar {q},r)=(1,1,1,0)$ to see that
for some integers $e\in \{0,1\}$ and $d,k,\ell $ . (Note that the knot labeled $L^\beta $ in Lemma 6.10, as depicted in Figure 10, is our $\tau _{1/0}\cup \beta $ and that the two cases (6.5) and (6.6) of Lemma 6.10 coincide since $q=\bar {q}$ .) In fact, we recall from Lemma 6.7 that $\ker (\rho )$ is generated by $\Delta ^4$ , so we must have
Now we use the other conclusion of Lemma 7.4 – namely, that the 3-braid closure $\hat \beta $ is a 2-component unlink. Viewing this as the $(2,0)$ -torus link, Birman and Menasco [Reference Birman and Menasco8] proved that $\beta $ must be conjugate to either y or $y^{-1}$ , so that its exponent sum is $\pm 1$ and
But we can also compute that
Thus, $(k+\ell ,e)$ is either $(0,0)$ or $(-4,1)$ .
Suppose first that $(k+\ell ,e) = (0,0)$ . Then $\beta = \Delta ^{4d} y^{\ell } x y^{-\ell }$ for some integer d. In this case, its exponent sum is $12d+1$ , and since this is equal to $\pm 1$ , we must have $d=0$ .
In the remaining case, we have $(k+\ell ,e)=(-4,1)$ , so $\beta = \Delta ^{4d+2} y^{\ell } \cdot y^4x \cdot y^{-\ell }$ for some d. The exponent sum is $12d+11 = \pm 1$ , so then $d=-1$ , and we have
We now use the braid relation $xyx=yxy$ to see that
and since $\Delta ^2$ is central, it follows that
or
This completes the proof.
We now determine the knots $K_\beta $ that arise in Lemma 7.2.
Lemma 7.6. Suppose that K satisfies the hypotheses of Lemma 7.2, and write $K = K_\beta $ where K arises as the lift of the curve $\kappa $ in the branched double cover of the unknot $U = \tau \cup \beta $ . Then K is isotopic to either $K_x$ or $K_{x^{-1}}$ .
Proof. By Proposition 7.5, we know that $\beta = y^a x^\epsilon y^{-a}$ , where $a\in \mathbb {Z}$ and $\epsilon =\pm 1$ . These are illustrated in Figure 23, where we have started with a slight isotopy of the unknot $U = \tau \cup \beta $ from Figure 21. The bottom of Figure 23 makes it clear that up to isotopy, the knot $K_\beta $ only depends on $\epsilon $ and the parity of a because the tangle relation
lets us identify the links $U\cup \kappa $ for $\beta =y^{a+2}x^\epsilon y^{-(a+2)}$ and for $\beta =y^ax^\epsilon y^{-a}$ up to isotopy. Thus, we need only consider the cases $a=0$ and $a=1$ .
Starting from the bottom of Figure 23, we simplify part of the corresponding diagrams by an isotopy in Figures 24 and 25, corresponding to $a=0$ and $a=1$ , respectively. In Figure 26, we further isotope the diagrams for each $U \cup \kappa $ , starting from the simplifications in Figures 24 and 25, and we see that the corresponding links for $\beta =x$ and $\beta =yxy^{-1}$ are isotopic to each other in a way which carries U to U and $\kappa $ to $\kappa $ , as are the links for $\beta =x^{-1}$ and $\beta =yxy^{-1}$ . It follows that
for all $a\in \mathbb {Z}$ since it is true for $a=1$ and since for fixed $\epsilon $ , the knot $K_{y^ax^\epsilon y^{-a}}$ depends only on the parity of a. Thus, every $K_\beta $ must be isotopic to either $K_x$ or $K_{x^{-1}}$ , as claimed.
Proposition 7.7. We have $K_x \cong \operatorname {Wh}^+(T_{2,3},2).$
Proof. We take the link with components $U = \tau \cup \beta $ (where $\beta =x$ ) and $\kappa $ from the top row of Figure 26 and isotope it into a convenient position in the first half of Figure 27. Having done so, in the remainder of Figure 27, we then take the branched double cover with respect to the unknot U, lifting $\kappa $ to the knot $\tilde \kappa = K_x$ as we do so, and then isotope it further until it is recognizable as the 2-twisted, positively clasped Whitehead double of $T_{2,3}$ .
Proposition 7.8. We have $K_{x^{-1}} \cong \operatorname {Wh}^-(T_{2,3},2).$
Proof. Just as in Proposition 7.7, we take the link with components $U = \tau \cup \beta $ and $\kappa $ , this time with $\beta =x^{-1}$ , as pictured in the third row of Figure 26. In Figure 28, we carry out an isotopy, take the branched double cover with respect to the unknot U, and then lift $\kappa $ to the knot $\tilde \kappa = K_{x^{-1}}$ , which we recognize after further isotopy as the 2-twisted, negatively clasped Whitehead double of $T_{2,3}$ .
We can now finish the proof of Theorem 7.1 and then conclude Theorem 1.2.
Proof of Theorem 7.1.
We apply Lemma 7.2, according to which K is the lift of $\kappa $ in the branched double cover of the unknot $U = \tau \cup \beta $ . Although there are infinitely many such $\beta $ (see Proposition 7.5), Lemma 7.6 says that in fact K must arise from this construction for either $\beta = x$ or $\beta = x^{-1}$ . In the case $\beta =x$ , Proposition 7.7 says that $K \cong \operatorname {Wh}^+(T_{2,3},2)$ , and if instead we have $\beta =x^{-1}$ , then $K \cong \operatorname {Wh}^-(T_{2,3},2)$ by Proposition 7.8. This completes the proof.
Proof of Theorem 1.2.
Letting F be a genus-1 Seifert surface for K, we proved in Theorem 5.1 that up to replacing K with its mirror, the manifold $M_F$ must be the complement of the $(2,4)$ -cable of either the unknot or the right-handed trefoil. In the unknot case, Theorem 6.1 says that K is one of
for some $n\in \mathbb {Z}$ . Likewise, in the trefoil case, Theorem 7.1 tells us that K is either
Thus, either K or its mirror must be one of the knots listed above.
8 Detection results for Khovanov homology
Our goal in this section is to prove the detection results for reduced Khovanov homology stated in Theorems 1.6 and 1.7. We will do so after establishing some preliminary results. We continue to work with coefficients in $\mathbb {Q}$ throughout this section.
Recall that both reduced Khovanov homology and knot Floer homology admit bigradings, which can be collapsed to a single $\delta $ -grading, defined for these two theories by
respectively. We say that either invariant is thin if it is supported in a unique $\delta $ -grading. Given a knot $K\subset S^3$ , Dowlin’s spectral sequence [Reference Dowlin13]
from reduced Khovanov homology to knot Floer homology respects the $\delta $ -gradings on either side, up to an overall shift. This implies the following:
Lemma 8.1. Let $K\subset S^3$ be a knot for which $\overline {\mathit {Kh}}(K)$ is thin. Then $\widehat {\mathit {HFK}}(K)$ is thin and
Proof. Suppose that $\overline {\mathit {Kh}}(K)$ is thin. Then the fact that Dowlin’s spectral sequence respects the $\delta $ -grading up to an overall shift, together with the symmetry [Reference Ozsváth and Szabó48]
implies that $\widehat {\mathit {HFK}}(K)$ is also thin. Recall that the graded Euler characteristics of reduced Khovanov homology and knot Floer homology recover the Jones and Alexander polynomials, respectively [Reference Khovanov33, Reference Ozsváth and Szabó48]:
Supposing that $\overline {\mathit {Kh}}(K)$ and $\widehat {\mathit {HFK}}(K)$ are supported in $\delta $ -gradings $\delta _1$ and $\delta _2$ , respectively, it follows that
and thus,
as claimed.
The next result pertains to the geography of knot Floer homology. For this result, recall that for any knot $K\subset S^3$ , there are two differentials on knot Floer homology,
where $\xi ^i$ and $ \omega ^i$ are, respectively, sums of maps of the form
Indeed, given a doubly-pointed Heegaard diagram for the knot $K\subset S^3$ ,
the differential $\partial $ in the Heegaard Floer complex
is a sum $\partial = d_0 + d_1,$ where $d_0$ counts those disks that avoid the basepoint z, and $d_1$ counts the rest. Then
and $\xi $ is the differential on this homology induced by $d_1$ . The map $\omega $ is defined in the same way but with the roles of z and w swapped. It follows from the definition that the homology with respect to either differential recovers the Heegaard Floer homology of $S^3$ ,
Furthermore, the components $\xi ^1$ and $\omega ^1$ anticommute. (This follows from Ozsváth–Szabó’s original construction of $\mathit {CFK}^\infty (K)$ in [Reference Ozsváth and Szabó48]; it is also stated explicitly in [Reference Baldwin, Levine and Sarkar3, Equation (3.7)] where our $\xi ^1$ and $\omega ^1$ correspond to their $\Psi ^p$ and $\Omega ^p$ .)
When $\widehat {\mathit {HFK}}(K)$ is thin, we have that $\xi = \xi ^1$ and $\omega = \omega ^1$ according to the grading shifts in (8.3) and (8.4). In particular,
Moreover, in this case, the two homology groups in (8.5) are supported in Alexander gradings $\tau (K)$ and $-\tau (K)$ , respectively, where $\tau (K)$ is the Ozsváth–Szabó tau invariant [Reference Ozsváth and Szabó46]. With this background in place, we may now prove the following:
Lemma 8.2. Let $K\subset S^3$ be a knot of genus $g\geq 1$ for which $\widehat {\mathit {HFK}}(K)$ is thin. Then
If in addition K is fibered with $|\tau (K)|<g$ , then this is a strict inequality.
Proof. Suppose that $g\geq 1$ and $\widehat {\mathit {HFK}}(K)$ is thin. Then $\xi = \xi ^1$ and $\omega = \omega ^1$ and $\xi \omega = -\omega \xi $ . If
then we have also that
by conjugation symmetry. The complex $(\widehat {\mathit {HFK}}(K),\xi )$ , given by
then has nontrivial homology in both of the Alexander gradings g and $-g$ , meaning that
a contradiction. This proves the first claim.
Now suppose that K is also fibered, and assume for a contradiction that $|\tau (K)|<g$ but
The fact that $\tau (K)\neq \pm g$ implies that the complexes $(\widehat {\mathit {HFK}}(K),\xi )$ and $(\widehat {\mathit {HFK}}(K),\omega )$ both have trivial homology in Alexander grading g. This implies that the components
of $\xi $ and $\omega $ are both nontrivial, and hence so is their composition, since
Letting x be a generator of $\widehat {\mathit {HFK}}(K,g)$ , this shows that $\omega (\xi (x)) \neq 0$ . However, $\xi (\omega (x)) = \xi (0)=0$ , which contradicts the fact that $\xi \omega = -\omega \xi .$
We now prove Theorem 1.6, which states that reduced Khovanov homology detects $5_2$ .
Proof of Theorem 1.6.
Suppose that
as bigraded vector spaces. Note that $\overline {\mathit {Kh}}(5_2)$ is thin since $5_2$ is alternating [Reference Lee39]. It then follows from Lemma 8.1 that the knot Floer homology of K is thin and that
Let $g\geq 1$ be the genus of K, and let us first suppose that K is not fibered. Then
Together with the fact from Lemma 8.2 that
and the fact that the total dimension is 7, this implies that $g=1$ and the sequence
In particular, K is a nearly fibered knot of genus 1, and it follows from Theorem 1.2 and Table 1 that K is either $5_2$ or $\overline {5_2}$ . But reduced Khovanov homology distinguishes $5_2$ from its mirror, so we have that $K = 5_2$ , as desired.
Finally, let us suppose for a contradiction that K is fibered. First, note that
Indeed, if $|\tau (K)| = g$ instead, then either K or its mirror is strongly quasipositive [Reference Hedden24, Theorem 1.2]. In this case, [Reference Plamenevskaya56, Proposition 4] implies that Rasmussen’s invariant [Reference Rasmussen58] satisfies $s(K) = \pm 2g$ . Since $\overline {\mathit {Kh}}(K)$ is thin, it is supported in the $\delta $ -grading
as argued at the end of [Reference Baldwin, Dowlin, Levine, Lidman and Sazdanovic1, Proof of Theorem 1]. Since $\overline {\mathit {Kh}}(5_2)$ is supported in $\delta $ -grading 1, it follows that $g=1$ . Then K is a fibered knot of genus 1, and hence a trefoil or the figure eight, but this violates our assumption that
The strict inequality in (8.6) therefore holds.
It then follows from Lemma 8.2 that
The fact that $\dim \widehat {\mathit {HFK}}(K)=7$ then implies that either $g=1$ , which cannot happen (since K is not a trefoil or the figure eight, as discussed above), or else $g>1$ and
Let us assume the latter holds. Note in this case that if $g>2$ , then the complex $(\widehat {\mathit {HFK}}(K),\xi )$ must have nontrivial homology in both Alexander gradings $g-1$ and $1-g$ , meaning that
a contradiction. Therefore, $g=2$ and
The complexes $(\widehat {\mathit {HFK}}(K),\xi )$ and $(\widehat {\mathit {HFK}}(K),\omega )$ therefore take the forms
and
respectively, where the subscripts indicate the Alexander grading. The fact that
implies that the homologies of these complexes are trivial in Alexander gradings $\pm 2$ . This implies that the components $\xi _2,\xi _{-1},\omega _{-2},$ and $\omega _1$ are all nontrivial. Moreover, $\xi _1$ and $\xi _0$ cannot both be nontrivial, as this would imply that their composition is nontrivial, which would violate $\xi ^2 = 0$ . Let us assume without loss of generality that
Let x be an element of $\widehat {\mathit {HFK}}(K,{-1})$ for which $\xi _{-1}(x)\neq 0$ . Then
while
contradicting the fact that $\omega \xi = -\xi \omega $ . We have therefore ruled out the possibility that K is fibered, completing the proof of Theorem 1.6.
Remark 8.3. One can use a similar argument to prove the slightly stronger result that if $\overline {\mathit {Kh}}(K)$ is 7-dimensional and supported in a unique $\delta $ -grading d, then, up to taking mirrors, either $K=5_2$ , or else $d=3$ and
as bigraded vector spaces. Though relatively straightforward, proving this takes quite a bit of room, so we do not pursue it here.
Finally, we prove Theorem 1.7, which states that reduced Khovanov homology together with the degree of the Alexander polynomial detects each pretzel knot $P(-3,3,2n+1)$ .
Proof of Theorem 1.7.
Suppose that
as bigraded vector spaces, and that $\Delta _K(t)$ has degree one. Then K is not fibered. Starkston proved [Reference Starkston60, Theorem 4.1] that the reduced Khovanov homology of this pretzel is thin. It then follows from Lemma 8.1 that the knot Floer homology of K is thin, and that
Since $\widehat {\mathit {HFK}}(K)$ is thin and $\Delta _K(t)$ has degree one, we conclude from (8.2) and the genus detection (1.1) that $g(K)=1$ . Since K is not fibered, we have that
Together with the fact from Lemma 8.2 that
and the fact that the total dimension is 9, this implies that the sequence
But in the latter case, we would have
which would imply that $\Delta _K(1) = \pm 3$ , but $\Delta _K(1) = 1$ for any knot $K\subset S^3$ . Therefore,
and hence K is nearly fibered of genus 1. The fact that $\widehat {\mathit {HFK}}(K)$ is thin and 9-dimensional then means, by Theorem 1.2 and Table 1, that K must be a pretzel knot $P(-3,3,2m+1)$ for some $m\in \mathbb {Z}$ (the mirror of any such pretzel is another such pretzel). But
for $m\neq n$ , by [Reference Starkston60, Theorem 4.1] or the more general [Reference Hedden and Watson25, Theorem 3.2]. We conclude that $K = P(-3,3,2n+1),$ as desired.
9 Detection results for HOMFLY homology
As mentioned in §1.1, reduced HOMFLY homology, defined by Khovanov–Rozansky in [Reference Khovanov and Rozansky35], assigns to a knot $K\subset S^3$ a triply-graded vector space over $\mathbb {Q}$ ,
which determines the HOMFLY polynomial of K by the relation
Our goal in this section is to prove Theorem 1.8, which says that reduced HOMFLY homology detects each pretzel knot $P(-3,3,2n+1)$ . We begin with the following computation:
Lemma 9.1. We have $\dim \bar {H}(P(-3,3,2n+1)) = 9$ for all $n \in \mathbb {Z}$ .
In order to prove this lemma, let us first recall that Khovanov–Rozansky also defined for each integer $N\geq 1$ a reduced $\mathfrak {sl}_N$ homology theory [Reference Khovanov and Rozansky34], which assigns to a knot $K\subset S^3$ a bigraded vector space over $\mathbb {Q}$ ,
Khovanov homology is related to the $\mathfrak {sl}_2$ theory by the following change in gradings:
Rasmussen proved in [Reference Rasmussen59, Theorem 2] that there is a spectral sequence which starts at $\bar {H}(K)$ and converges to $ \bar {H}_N(K),$ for each $N\geq 1$ . Moreover, when this spectral sequence collapses at the first page, as it does for N sufficiently large, the reduced HOMFLY homology determines the $\mathfrak {sl}_N$ theory [Reference Rasmussen59, Theorem 1] by
In particular, $\dim \bar {H}(K) = \dim \bar {H}_N(K)$ for $N\gg 0$ .
Proof of Lemma 9.1.
Let us write
for convenience. First, note that $K_0$ is the 2-bridge knot $6_1$ . It therefore follows from [Reference Rasmussen57, Theorem 1] that $K_0$ is N-thin for all $N> 4$ , which implies by [Reference Rasmussen57, Corollary 4.3] that
Next, observe that $K_0$ can be obtained via band surgery on the 2-stranded pretzel link $P(-3,3)$ , which is a split link (in fact, a 2-component unlink), as shown in Figure 29. Each $K_n$ can then be obtained from $K_0$ by adding n full twists to that band, so a theorem of Wang [Reference Wang64, Proposition 1.7] says that for any $N \geq 2$ , the dimension
is independent of n. Thus, for any $n\in \mathbb {Z}$ , the above computation for $K_0$ tells us that
and hence that $\dim \bar {H}(K_n) = 9$ , as desired.
With this computation in hand, we may now prove Theorem 1.8.
Proof of Theorem 1.8.
Suppose that
as triply-graded vector spaces. Then K has the same HOMFLY polynomial as $P(-3,3,2n+1)$ . Since the HOMFLY polynomial specializes to the Alexander polynomial, we have that
In particular,
Since we also know from the computation in Lemma 9.1 that
it follows that the spectral sequence from $\bar {H}(K)$ to $\bar {H}_2(K)$ must collapse at the first page. Therefore, $\bar {H}(K)$ determines $\bar {H}_2(K)$ as in (9.2). In particular, it follows that
as bigraded vector spaces. Then we have by (9.1) that
as bigraded vector spaces. Since K has the same Alexander polynomial and reduced Khovanov homology as $P(-3,3,2n+1)$ , Theorem 1.7 says that $K=P(-3,3,2n+1)$ .
A Computations of knot Floer homology
In this appendix, we explain the knot Floer homology calculations recorded in Table 1. The computation for $5_2$ follows from the fact that it is alternating [Reference Ozsváth and Szabó45, Theorem 1.3]. For the pretzel knots $P(-3,3,2n+1)$ , we apply [Reference Ozsváth and Szabó51, Theorem 1.3] (but see also [Reference Hedden and Watson25, Theorem 1]). For the twisted Whitehead doubles, Hedden [Reference Hedden23, Theorem 1.2] computed their knot Floer homology over $\mathbb {Z}/2\mathbb {Z}$ , but his results work over arbitrary fields. This leaves only the knot $15n_{43522}$ , which will occupy the remainder of this appendix.
Proposition A.1. We have that
where the subscripts denote Maslov gradings.
To start, we can carry out the same computation with coefficients in a finite field using a program by Zoltán Szabó [Reference Szabó61], and over $\mathbb {F}=\mathbb {Z}/2\mathbb {Z}$ , we find that
Proposition A.1 will then follow from the universal coefficient theorem if we can show that $\widehat {\mathit {HFK}}(15n_{43522};\mathbb {Z})$ has no 2-torsion.
Suppose, for a contradiction, that there is 2-torsion in some Alexander grading a. Then by the universal coefficient theorem, it must contribute $\mathbb {F}$ summands to consecutive homological (i.e., Maslov) gradings of $\widehat {\mathit {HFK}}(15n_{43522},a;\mathbb {F})$ . By inspection, it can only possibly contribute $\mathbb {F}_{(-1)}\oplus \mathbb {F}_{(0)}$ to $\widehat {\mathit {HFK}}(15n_{43522},0;\mathbb {F})$ , and therefore
That is,
as bigraded vector spaces. Since this knot Floer homology is thin, we have that
up to filtered chain homotopy equivalence [Reference Petkova55, Lemma 5]. Since the complex $\mathit {CFK}^\infty (K)$ determines [Reference Ozsváth and Szabó48] the Heegaard Floer homology of n-surgery on a knot $K\subset S^3$ for integers
it follows that
We will use this together with the following lemma to get a contradiction.
Lemma A.2. If $K \subset S^3$ is a knot of genus at least 2, then $\dim \widehat {\mathit {HF}}(S^3_{\pm 1}(K);\mathbb {Q}) \geq 5$ .
Proof. By the surgery exact triangles
and
it suffices to show that $\dim \widehat {\mathit {HF}}(S^3_0(K);\mathbb {Q}) \geq 6$ .
Let $\mathfrak {s}_i \in \operatorname {Spin}^c(S^3_0(K))$ be the $\operatorname {Spin}^c$ structure with
where $\hat \Sigma \subset S^3_0(K)$ is a capped-off Seifert surface for K. Then according to [Reference Ozsváth and Szabó48, Corollary 4.5] and the way in which knot Floer homology detects the genus $g = g(K)$ , which is at least 2, we have
Likewise,
by the conjugation symmetry of Heegaard Floer homology. Furthermore, $\mathit {HF}^+(S^3_0(K),\mathfrak {s}_0;\mathbb {Q})$ is nontrivial because $\mathfrak {s}_0$ is torsion (see [Reference Ozsváth and Szabó50, §10.6]).
We now recall from [Reference Ozsváth and Szabó49, Proposition 2.1] that $\widehat {\mathit {HF}}(Y,\mathfrak {s})$ is nonzero if and only if $\mathit {HF}^+(Y,\mathfrak {s})$ is nonzero, so we have shown that
for each $i=g-1,0,1-g$ . In fact, each of these $\operatorname {Spin}^c$ summands has Euler characteristic zero [Reference Ozsváth and Szabó49, Proposition 5.1] and hence even dimension, so the total dimension of $\widehat {\mathit {HF}}(S^3_0(K))$ must be at least $2+2+2=6$ , as claimed.
Proof of Proposition A.1.
Supposing otherwise, we have already argued that
We now observe the following coincidences in SnapPy [Reference Culler, Dunfield, Goerner and Weeks11]:
In other words, if ${\mathtt{K15n43522}}$ , and denote each of $15n_{43522}$ , $8_{20}$ and $9_{42}$ with the fixed chirality given by SnapPy (which may or may not be mirror to their usual chiralities), then we have
But $8_{20}$ and $9_{42}$ both have genus 2, so we can apply Lemma A.2 to conclude that
and we have a contradiction.
Acknowledgements
We thank Anna Beliakova, Nathan Dunfield, Matt Hedden, Jen Hom, Siddhi Krishna, Zhenkun Li, Tye Lidman, Nikolai Saveliev, Josh Wang and Fan Ye for helpful and interesting conversations related to this work. We are grateful to the referees for all of their feedback and, in particular, for a substantial simplification to the proof of Lemma 4.4. We also note that Zhenkun and Fan independently discovered some of the results in §3–§5.
Competing interest
The authors have no competing interest to declare.
Funding statement
JAB was supported by NSF FRG Grant DMS-1952707.