missing
The upsilon invariant at of -braid knots
Abstract.
We provide explicit formulas for the integer-valued smooth concordance invariant for every -braid knot . We determine this invariant, which was defined by Ozsváth, Stipsicz and Szabó [OSS17a], by constructing cobordisms between -braid knots and (connected sums of) torus knots. As an application, we show that for positive -braid knots several alternating distances all equal the sum , where denotes the -genus of . In particular, we compute the alternation number, the dealternating number and the Turaev genus for all positive -braid knots. We also provide upper and lower bounds on the alternation number and dealternating number of every -braid knot which differ by .
Key words and phrases:
3-braids, Upsilon invariant, alternation number, fractional Dehn twist coefficient.1991 Mathematics Subject Classification:
57K10, 57K18, 20F36.1. Introduction
We study knots in the -sphere , i. e. non-empty, connected, oriented, closed smooth -dimensional submanifolds of , considered up to ambient isotopy.
Two knots and are called concordant
if there exists an annulus smoothly and properly embedded in such that and such that the induced orientation on the boundary of the annulus agrees with the orientation of , but is the opposite one on .
Knots up to concordance form a group, the concordance group , with the group operation induced by connected sum.
In [OSS17a], Ozsváth, Stipsicz and Szabó used the Heegaard Floer knot complex to define the invariant of a knot ,
which induces a homomorphism from the knot concordance group to the group of real-valued piecewise linear functions on the interval .
The function evaluated at , , induces a homomorphism . In this article, we will call upsilon of .
A -braid is an element of the braid group on three strands, denoted . The classical presentation of with generators and and relation , the braid relation,
was introduced by Artin [Art25]. A braid word — a word in the generators of and their inverses — defines a diagram for a (geometric) -braid;
the generators and correspond to the geometric -braids given by braid diagrams as in Figure 1. In our figures, braid diagrams will always be oriented from bottom to top.
We denote by the braid , and note that its square (the positive full twist on three strands) generates the center of
[Cho48, Theorem 3].
A -braid knot is a knot that arises as the closure of a -braid .
As our main result, we determine the upsilon invariant for all -braid knots. More precisely, we show the following.
Theorem 1.1.
Let be a braid word in the generators and of for some integers , and for , where . Suppose that the closure of is a knot. Then its upsilon invariant is
It follows from Murasugi’s classification of the conjugacy classes of -braids [Mur74, Proposition 2.1] that indeed all -braid knots — except for the torus knots that are closures of -braids — are covered by Theorem 1.1.
However, for torus knots the invariant can be calculated explicitly by a combinatorial, inductive formula in terms of their Alexander polynomial [OSS17a, Theorem 1.15]; see Equation 12 below.
Hence, we have indeed determined for all -braid knots .
As an application of Theorem 1.1, we show that the following invariants coincide for positive -braid knots — knots that are the closure of positive -braids.
Corollary 1.2.
Let be a knot that is the closure of a positive -braid, i.e. an element of that can be written as a word in the generators and only (no inverses). Then
Here, the alternation number , dealternating number and Turaev genus are different ways of measuring
how far the knot is from being alternating.
The best known among them is certainly the first one: the alternation number of a knot was first defined by Kawauchi [Kaw10] as the minimal Gordian distance of to the set of alternating knots. In Section 5, we will review the precise definition and prove 1.2.
The invariant introduced by Friedl, Livingston and Zentner [FLZ17] is defined as the minimal number of double point singularities in a generically immersed concordance from a knot to an alternating knot. Lastly, denotes the -genus of , the minimal genus of a compact, connected, oriented smooth surface in with oriented boundary the knot .
Two other corollaries of Theorem 1.1 for positive -braid knots are the following.
Corollary 1.3.
Let be a positive -braid knot. Then the minimal such that is the closure of for positive integers , , is .
Corollary 1.4.
If and are concordant knots that are both closures of positive -braids, then the minimal from 1.3 is the same for both and .
3.2 in Section 3 provides a normal form for -braids, the Garside normal form, which is different from Murasugi’s normal form mentioned above (cf. 4.15).
The Garside normal form allows us to read off from a braid word whether it is conjugate to a positive braid word.
In Section 6, we provide formulas for the fractional Dehn twist coefficient for all -braids in Garside normal form; see
6.1.
Proof strategy for Theorem 1.1. A crucial property of the invariant is that it provides a lower bound on the -genus of a knot , the minimal genus of a compact, connected, oriented surface smoothly embedded in the -ball with oriented boundary the knot in : we have
(1) |
for any knot [OSS17a, Theorem 1.11]. Our general strategy to find for any -braid knot will be to construct a cobordism between and another knot for which the value of is known. A cobordism between and is a smoothly and properly embedded oriented surface in with boundary such that the induced orientation on the boundary of agrees with the orientation of and disagrees with the orientation of . We have
(2) |
for any cobordism between and , where denotes the genus of the cobordism; see inequality (15) in Section 4.1. This provides bounds on in terms of and .
We will find such cobordisms for example by algebraic modifications of a braid word representing and by saddle moves corresponding to the addition or deletion of generators from such braid words. We will also repeatedly make use of the trick described in 4.1 in Section 4.1 of looking at cobordisms of genus between and for -braid words and .
To prove Theorem 1.1, we will first determine for all positive -braid knots and then generalize our computations to all -braid knots. This extension was somewhat unexpected for the author since, in contrast, the same method would not work to determine slice-torus invariants [Lew14] like the invariant defined by Ozsváth and Szabó [OS03] or Rasmussen’s invariant [Ras10] for all -braid knots. We will elaborate on this in Section 4.4.2.
Remark 1.5.
As we will only use properties of the upsilon invariant (see Section 2.2) and not its definition, we can similarly determine any concordance homomorphism whose absolute value bounds the -genus of a knot from below and which takes the same value as on torus knots of braid index and . An example is for the concordance invariant constructed by Ballinger [Bal20] from the spectral sequence on Khovanov homology. The invariant defines a concordance homomorphism valued in the even integers which satisfies for any knot [Bal20, Theorem 1.1]. Moreover, it fulfills for the torus knots for any coprime positive integers and [Bal20, p. 22]. The same method we use for the proof of Theorem 1.1 shows that for any -braid knot .
Remark 1.6.
Theorem 1.1 and a result of Erle [Erl99] imply that for all -braid knots except when for odd and . Here denotes the classical signature of the knot [Tro62]111We use the standard signature convention that the positive torus knots have negative signatures, e. g. .. In the exceptional cases, we have . This observation improves a result by Feller and Krcatovich who showed that for all -braid knots [FK17, Proposition 4.4]; see also Section 4.4.1.
Organization.
The remainder of this article is organized as follows.
In Section 2, we will provide the necessary background on (positive) braids and the upsilon invariant before providing a normal form for -braids (3.2) that we call Garside normal form in Section 3.
Then in Section 4, after a more detailed outline of our proof strategy (Section 4.1), we will prove Theorem 1.1
first for positive -braid knots (Section 4.2) and afterwards in the general -braid case (Section 4.3). We will prove 1.3 and 1.4
in Section 4.2.
Section 4.4 will provide further context on our results.
Section 5 is concerned with the proof of 1.2 (Section 5.1) and the application of our result about the upsilon invariant to alternating distances of general -braid knots (Section 5.2). In particular, we determine the alternation number of any -braid knot up to an additive error of at most .
Finally, in Section 6, we determine the fractional Dehn twist coefficient for all -braids in Garside normal form.
Acknowledgments. I would like to thank Peter Feller for introducing me to the topic and for all the helpful discussions. Thanks also to Lukas Lewark for lots of useful comments, including during my stay in Regensburg in September 2020, and to Xenia Flamm for her feedback. Finally, I thank the referee for many valuable remarks and improvements. This project is supported by the Swiss National Science Foundation Grant 181199.
2. Preliminaries
We recall important concepts about knots and braids, and also the necessary properties of the upsilon invariant and the knot invariant coming from Heegaard Floer homology.
2.1. Knots and braids
By a fundamental theorem of Alexander [Ale23], every knot in can be represented as the closure of a geometric -braid for some positive integer . An -braid is an element of the braid group on strands, denoted , which is presented by generators and relations
We call a word in the generators of and their inverses a braid word. A braid word defines a diagram for a (geometric) -braid where the generators of the braid group correspond to the geometric -braids given by the braid diagrams in which the -th and -st strands cross once positively. In the following, we will always identify braid words with the corresponding geometric braids, and we suppress if the context is clear.
By gluing the top ends of the (oriented) strands of a geometric braid to the corresponding bottom ends, we get a knot (or link) , called the closure of . If induces a permutation with only one cycle on the ends of its strands, then its closure is a knot and we call it an -braid knot. Note that conjugate braids , denoted by , have isotopic closures . For a more detailed account on braids, we refer the reader to [BB05].
A positive braid is an element of the braid group for some that can be written as a positive braid word with .
A knot is called a positive braid knot if it can be represented as the closure of a positive braid.
The set of positive braid knots contains the sets of (positive) torus knots and algebraic knots, while itself being a subset of the set of positive knots or, more generally, the frequently studied set of (strongly) quasipositive knots.
2.2. The concordance invariants and
In [OS03], Ozsváth and Szabó constructed the knot invariant via the knot filtration on the Heegaard Floer chain complex of ; the latter was also defined independently by Rasmussen [Ras03]. The invariant induces a group homomorphism from the (smooth) knot concordance group to the group of integers and gives a lower bound on the -ball genus : we have for any knot . For the torus knots , where and are coprime positive integers, the invariant recovers the -genus [OS03, Corollary 1.7], namely we have
(4) |
Moreover, it follows from [Liv04, Theorem 4 and Corollary 7] together with Equation 3 above that, for any knot that is the closure of a positive -braid , we have
(5) |
The invariant was defined by Ozsváth, Stipsicz and Szabó in [OSS17a]. We will not recall the definition of via the knot Floer complex since the properties of mentioned below will be enough for our later computations and we will not explicitly use the Heegaard Floer theory behind it. For an overview on the properties of , see the original article [OSS17a] or Livingston’s notes on [Liv17]; see [Hom17] for a survey on Heegaard Floer homology and knot concordance.
For every knot , the knot invariant is a continuous, piecewise linear function with the following properties [OSS17a]:
(6) | |||
(7) | |||
(8) | |||
(9) | |||
(10) |
Here, is the knot obtained by mirroring and reversing its orientation. Its concordance class is the inverse of the class of in the knot concordance group .
It follows from (8)-(10) that induces a homomorphism from the concordance group to the group of real-valued piecewise linear functions on the interval .
For some classes of knots, the invariant can be explicitly computed in terms of classical knot invariants like the signature and the Alexander polynomial.
Proposition 2.1 ([OSS17a, Theorem 1.14]).
We have for all alternating or quasi-alternating knots and all .
3. The Garside normal form for -braids
In this section, we provide a classification result on the conjugacy classes of -braids; see 3.2.
This result is basically due to work of Garside [Gar69]
who gave the first solution to the conjugacy problem for all braid groups , , in 1965.
3.2 might be known to the experts, but since the explicit formulas appear to be missing from the literature,
we will provide them here.
Throughout, we denote the two generators of the braid group by and which are subject to the braid relation . Recall that the braid generates the center of .
Remark 3.1.
Any -braid is conjugate to the same braid with generators and interchanged. More precisely, let for some and integers , be a -braid. Then using and , we have
In 3.2, we will provide a certain standard form for the conjugacy classes of -braids.
Proposition 3.2.
While we will never use it in this article, we note — without proof — the following uniqueness result related to 3.2.
Remark 3.3.
Up to cyclic permutation of the powers in (C) and in (D), respectively, each -braid is conjugate to exactly one of the -braids listed in 3.2. This follows from Garside’s work [Gar69]. In his notation, each of the -braids listed in (A)–(D) in 3.2 is the standard form of a certain element in the (so-called) summit set of . For -braids of the form (C) or (D), the summit set consists of those -braids obtained by cyclic permutation of the powers in (C) and in (D), respectively.
Remark 3.5.
The advantage of the Garside normal form over Murasugi’s normal form for -braids used later in Section 4.3 (see 4.15) is that positive -braids are easier to detect in this normal form: if is a positive -braid, then is conjugate to one of the braids in (A)–(D) with . Since Garside’s solution to the conjugacy problem works for any -braid with , one might hope to generalize an explicit standard form as in 3.2 to -braids for any .
Remark 3.6.
Proof of 3.2.
The proof will follow from the following claim.
We first deduce 3.2 from this claim. To that end, let be any -braid. If is a positive braid, we are done by 1. If not, then can be written in the form where is a negative integer and a positive -braid [Gar69, Theorem 5]. In fact, inserting if is odd, we can assume to be of the form for some and a positive -braid . The proposition then easily follows using the claim for . It remains to prove 1.
Proof of 1: A positive -braid has the form for integers , , . If all the or all the are , then (possibly using 3.1) is conjugate to for some and we are in case (A) for . Possibly after conjugation and reduction of , we can thus assume that all of the integers are non-zero. If applies for all , then is of the form in (C) for . If , i. e. for integers , and or , then (possibly using 3.1) is conjugate to a braid of the form in (B).
It remains to consider the case where and at least one of the or is . In that case — if necessary after conjugation — contains as a subword and is thus conjugate to for some positive -braid . Now, let be maximal with the property that is conjugate to for some positive -braid . Then, possibly after conjugation of , the braid word must be one of the following:
(13) | |||||
Indeed, using 3.1, up to conjugation these are the only possible words such that does not contain any additional as a subword. Note that can be the empty word, which is covered by the first case in (3) for . Further, note that
(14) |
for any , , , . It follows from a case by case analysis of the cases in (3), using (14) and taking the parity of into account, that any positive -braid is conjugate to one of the -braids in (A)–(D) with .
This concludes the proof of 3.2. ∎
4. The upsilon invariant of -braid knots
In this section, we prove Theorem 1.1. Along the way, we compute the invariant for positive -braid knots in Garside normal form (4.2) and prove 1.3 and 1.4.
4.1. Methodology
We first recall inequality (2) from the introduction — which will be repeatedly used in Section 4 — in more generality.
The cobordism distance between two knots and is defined as the -genus of the connected sum of and the inverse of . Equivalently, the cobordism distance could be defined as the minimal genus of a smoothly and properly embedded oriented surface in with boundary such that the induced orientation on the boundary of agrees with the orientation of and disagrees with the orientation of . Suppose the genus of a cobordism between two knots and is . We then have , so by the properties (8)-(10) of from Section 2.2 we get
(15) |
for all .
This provides bounds on in terms of and .
We now give an example for the cobordisms we will use later on.
Example 4.1.
Among other things, we will frequently use the following trick the author first saw in [FK17, Example 4.5]. Let be a -braid such that is a knot. Consider the -braid for some . Then is also a knot and there is a cobordism between and the connected sum of genus . This cobordism can be realized by two saddle moves (-handle attachments) of the form shown in Figure 2(b), performed in the two circled regions of Figure 2(a). One of them is used to add a generator to the braid to obtain the braid word and the other is used to transform the closure of this new braid word into a connected sum of and . Recall that our braid diagrams are oriented from bottom to top.

Using by Equation 11 and that the genus of the cobordism is , by (15) for , we have
(16) |
which provides the lower bound on .
4.2. The upsilon invariant of positive -braid knots
In this section, we determine the invariant for all positive -braid knots.
By 3.2 and 3.6, positive -braid knots are either the torus knots for and which have braid representatives of Garside normal form (B), or closures of positive -braids of Garside normal form (C) or (D) (cf. 3.4). The following proposition thus proves Theorem 1.1 for all positive -braid knots.
Proposition 4.2.
Let be a positive -braid such that is a knot. Then
Remark 4.3.
In fact, the formulas from 4.2 also give the correct upsilon invariant in terms of the Garside normal form of a -braid representative of a knot if is the closure of any -braid in Garside normal form (C) or (D), not necessarily a positive one. This follows from Theorem 1.1 (proved in the next section) and the observations of Section 4.4.3.
Recall that for the torus knots of braid index , we know the invariant by Equation 12. In the following, we will determine the invariant for all knots that are closures of positive -braids of Garside normal form (C) or (D).
We first provide an upper bound on for positive -braid knots and . The following inequality (17) in 4.4 could also be shown using the dealternating number and a result of Abe and Kishimoto [AK10, Lemma 2.2], whereas the main work for the upper bound on for the knots in the second and third case in 4.2 will be to rewrite the braid words representing these knots. We use the approach below since it will also give bounds on the minimal cobordism distance between any positive -braid knot and an alternating knot; see 4.14.
Lemma 4.4.
Let be a positive -braid, where and , , are integers such that is a knot. Then
(17) |
Proof.
We claim that there is a cobordism of genus
(18) |
between and the connected sum
where are chosen such that is a connected sum of torus knots (rather than links), i. e. such that , , are all odd, and . This cobordism can be realized by saddle moves as follows. Following the schematic in Figure 3, we add generators by saddle moves and additionally perform saddle moves of the form shown in Figure 2(b) in the green circled regions of Figure 3. In Figure 3, a box on the left labeled or stands for the positive braid or , respectively. The Euler characteristic of the cobordism is . Since is connected and — as and are knots — has two boundary components, the genus of is as claimed.
By (15), we get for all , hence
(19) |
By Equations 8 and 11 from Section 2.2, we have
The claim follows, since by Equation 3, we have
The following two lemmas improve the bound from 4.4 for knots that are closures of positive -braids of Garside normal form (C) or (D), respectively.
Lemma 4.5.
Let for some , and for such that is a knot. Then
Proof of 4.5.
Let and note that using Equation 3, we have
(21) |
If , then is conjugate to
and , so . By 4.4, we get
For , using , we have
We have and by (21). Again, 4.4 implies
which proves the claim of the lemma. ∎
Lemma 4.6.
Let for some , and for such that is a knot. Then
In the proof, we will need the following statement about positive -braids.
Lemma 4.7.
In , we have
(22) |
Proof.
Starting with the left-hand side we have
which proves Equation 22 for . We now show by induction that
(23) |
which implies the lemma for all . For , we have
Assuming that (23) is true for some , we get
using the induction hypothesis in the second to last equality. ∎
Proof of 4.6.
Let . If , then by Equation 3 and 4.4 we have
For , using and 4.7, we have
Note that and by Equation 3, we have
Again by 4.4, we get
We will now focus on and prove 4.2 by showing that the upper bounds on from 4.5 and 4.6 for are also lower bounds. We will need the following observation used in [FK17, Example 4.5] about -braids, which we prove here for completeness.
Lemma 4.8.
In , we have
(24) |
Proof.
We prove the first statement by induction. For , the equality is clearly true. For , using and , we have
We now assume that (24) is true for some . Using the induction hypothesis and the equality for , we get
Lemma 4.9.
Let for some , and for such that is a knot. Then
Proof of 4.9.
Let . From 4.5, it follows directly that so we are left to show that . To that end, consider
where we used . Note that . Now, define
and note that is a knot. By assumption, we have . There is a cobordism between and the connected sum of genus by using two saddle moves similar to the two saddle moves illustrated in Figure 2 from 4.1. Similarly as in (16) from 4.1, we have . In order to find a lower bound for , note that there is a cobordism between and the torus knot of genus . Here we think of as the closure of the braid word , which is equal to as -braids by 4.8. The cobordism between and can thus be realized by
saddle moves corresponding to the deletion of the same number of generators and from the braid word to obtain . Hence the Euler characteristic of the cobordism is . Since is connected and has two boundary components (as and are knots), the genus of is indeed . Now, by (15) and Equation 12, we have
It follows that
as claimed, hence the statement of the lemma. ∎
Lemma 4.10.
Let for some , , and for such that is a knot. Then
Proof of 4.10.
The proof uses similar ideas as the proof of 4.9. Let . By 4.6, we have , so it remains to show that . To that end, we consider
Note that . We define
Then is a knot and by assumption we have . There is a cobordism between and of genus by using two saddle moves similar to the cobordism considered in 4.1 and in the proof of 4.9, hence . To find a lower bound for , we observe that there is a cobordism between the knot and the knot , where
Using Equation 24 from 4.8 for , in , we have
We thus have , so the closure of is the torus knot with by Equation 12. The cobordism between and can be realized by
saddle moves corresponding to the deletion of the same number of generators and from the braid word to obtain . By a similar Euler characteristic argument as in the proofs of 4.4 and 4.9, the genus of this cobordism is . Note that here we used and . Now, by (15), we have
Lemma 4.11.
Let for some , , for . Suppose that and for some and that is a knot. Then
Proof.
We proceed similar as in the proof of 4.10, but here we will look at a different cobordism to obtain a lower bound for . The steps of the proof are exactly the same until then, so we consider
and define
Again, we have Now, in order to find a lower bound for , we observe that there is a cobordism between and the knot , where
We find the cobordism by the deletion of generators from the braid word to obtain , where we use the assumptions and . In fact, the cobordism can be realized by
saddle moves, so its genus is . Using by 4.8, we have
Note that by our assumptions on , and , we have , and , so has the form of the braid words considered in 4.9. We thus have
By (15), we have
Proof of 4.2.
Before we proceed with the general case where the knot is given as the closure of any -braid, let us prove the following corollaries of our results in this section.
Corollary 4.12 (1.3).
Let be a knot that is the closure of a positive -braid. Then
is minimal among all integers such that is the closure of a positive -braid for integers , .
Proof.
By 4.4 we have
whenever is the closure of a positive -braid for integers , , . It remains to show that we can always find a positive braid representative for of the form with . We will use 3.2. In fact, if is the closure of a positive braid of the form in (C) with , then by Equation 3 applied to , 4.10 and 4.11. Moreover, we have
by the proof of 4.6; these give the desired braid representatives for . Furthermore, if is represented by a positive braid of the form in (D) with , then by Equation 3 and 4.9, and we have
by the proof of 4.5. Finally, if for and , then by Equation 4 and Equation 12, we have and and are represented by the positive -braids and respectively, by 4.8 and 4.7. ∎
Corollary 4.13 (1.4).
If and are concordant knots that are both closures of positive -braids, then the minimal from 4.12 is the same for both and .
Proof.
If and are concordant, then their -genus and their upsilon invariants are equal. So by Equation 3 from Section 2.1 and by 4.12, positive -braids with closures and , respectively, will have the same minimal . ∎
Remark 4.14.
Let denote the minimal genus of a cobordism between a knot and an alternating knot, i. e. the cobordism distance . By [FLZ17, Theorem 8], we have for any knot . It thus follows from our results in this section that
for any knot that is the closure of a positive -braid in Garside normal form (C) or (D), where is an integer depending on . The lower bound uses 4.2 and Equation 5 from Section 2.2; see also the proof of 4.12. The upper bound follows from the proofs of 4.5 and 4.6, see also the proof of 4.4. Note that for most positive -braid knots, we have , so we do not get an equality.
A shorter proof of 4.4 without cobordisms follows from a result of Abe and Kishimoto on the dealternating number of positive -braid knots. Indeed, we have
The definitions of the dealternating number and the alternation number of a knot and more details on the inequalities used here will be provided in Section 5.
4.3. Proof of Theorem 1.1
It remains to show Theorem 1.1 when is the closure of a not necessarily positive -braid. We first recall a result of Murasugi, which implies that indeed all -braid knots except for the torus knots of braid index are covered by Theorem 1.1.
Let be a -braid. Then, by [Mur74, Proposition 2.1], is conjugate to one and only one of the -braids
(a) | ||||
(b) | ||||
(c) |
Remark 4.16.
If in case (c), the braid word for integers and , , gives rise to an alternating braid diagram. If is a knot, by 2.1 we thus have in that case and the statement of Theorem 1.1 follows directly from a result by Erle on the signature of -braid knots.
Proposition 4.17 ([Erl99, Theorem 2.6]).
Let for integers , and for such that is a knot. Then
We still need to show Theorem 1.1 when is the closure of a -braid in Murasugi normal form (c) with . The proof will follow from the following two lemmas.
Lemma 4.18.
Let for some and for such that is a knot. Then
Lemma 4.19.
Let for some , and for such that is a knot. Then
Proof of Theorem 1.1.
For , the statement of the theorem follows directly from 4.18 and 4.19. If , the knot is represented by the braid word with and accordingly we have
Using that by Equation 9 from Section 2.2, this implies the claim. ∎
The remainder of this section is devoted to prove the above lemmas.
Proof of 4.18.
We first consider the case where and . Using and
we have
Now, we claim that there is a cobordism of genus between the closure of and the connected sum
where we choose such that is a connected sum of torus knots, i. e. such that , , are all odd; and . This cobordism can be realized using saddle moves as follows. On the one hand, we add generators and generators to the braid word , on the other hand, we perform saddle moves of the form as the saddle moves used in the proof of 4.4 to get a connected sum of torus knots. The Euler characteristic of is . Since is connected and has two boundary components (as and are knots), the genus of is as claimed. By Equations 8 and 11, we have
and by (15), we get
If and , then
and similarly as above, there is a cobordism of genus between the closure of and the connected sum
where we choose such that is a connected sum of torus knots and . The claim follows also in this case from Equations (8) and (11), and the intequality in (15).
It remains to show the claim when . In that case, using , we have
If , then is conjugate to and if , then using Equation 20 from Section 4.2, we have
In both cases, there is a cobordism of genus between the closure of and the connected sum
where we choose such that is a connected sum of torus knots and . Using (8), (11), and (15) again, the claim follows. ∎
We will need the following two technical lemmas for the proof of 4.19.
Lemma 4.20.
Let for some , and integers such that or , and or , for any . Moreover, assume that is a knot. Then
where denotes the cardinality of the set .
Lemma 4.21.
Let for some , and integers such that or for any and or for any . Moreover, assume that is a knot. Then
For the proof of 4.20 and 4.21, we refer the reader to the very end of this section; we will first prove 4.19 using these lemmas.
Proof of 4.19.
Let be the number of exponents of with and let for be the set of indices such that if and only if . For all , we rewrite the subword of using as
Note that if , then After rewriting for all , the braid is conjugate to for some -braid which is of the form
where and where the and fulfill the assumptions of 4.20 and 4.21, respectively, i. e. where or and or for any . The number of negative exponents in equals the number of negative exponents in , so
If is even, by 4.20, we get
Similarly, if is odd, the claim follows from 4.21. ∎
Proof of 4.20.
We will modify the braid word in steps, where each step corresponds to one of the exponents , , of . In every step, we will either just conjugate (if the corresponding exponent is positive) or perform a cobordism of genus between the closure of or and the connected sum for some — similarly as the cobordism described in 4.1 and used in the proofs of 4.9, 4.10 and 4.11. We now describe these steps in more detail. First, let and define
so that for some (note that we assumed or ). Here, if , we choose such that is even and is a knot. Second, let such that is even if , and define
so that for some . Inductively, for any , we let
so that
for some integers . Here we choose such that is even if . Moreover, for , we let such that is even, and define
and we define similarly as . Inductively, after steps, we get the positive -braid
with
for all ; so that . By 4.2, we have
Now, note that if for some , then there is a cobordism of genus between and by using two saddle moves, where , so similarly as in (16) from 4.1, we have
Similarly, if for some , then In addition, if , then we have , and if , then . We conclude
Proof of 4.21.
The strategy of the proof is the same as in the proof of 4.20. Here, we need steps corresponding to the exponents of . The steps are similar as in the proof of 4.20, the only change is that we multiply by a power of if , and by a power of if (since and ). Thus, starting with , after steps we obtain the positive -braid
with
By 4.9, we have
Since the steps we performed have similar effects on as the ones in the proof of 4.20, we get
4.4. Further discussion of Theorem 1.1
In this section, we provide some further context on our main result. In particular, in Section 4.4.2 we will discuss why it might be surprising that our proof strategy works for all -braid knots.
4.4.1. Comparison of upsilon and the classical signature
By Theorem 1.1 and 4.17, we have
(25) |
for any knot that is the closure of a -braid for integers , and for . Computations of the signature for torus knots (and links) of braid index , first done by Hirzebruch, Murasugi and Shinora [Mur74, Proposition 9.1, pp. 34-35], together with Equation 12 from Section 2.2 imply that the equality in (25) is in fact true for all -braid knots except for the cases that for odd or for odd . In the exceptional cases, we have . As mentioned in the introduction, this improves the inequality for all -braid knots in [FK17, Proposition 4.4].
It was shown in [OSS17b, Theorem 1.2] that gives a lower bound on the nonorientable smooth -genus of a knot , denoted , the minimal first Betti number of a nonorientable surface in that meets the boundary along . The similarity of the invariant and the classical signature on -braid knots described above clearly does not lead to a good lower bound on .
However, the equality for most -braid knots is actually no great surprise when noting that in fact must be true for all -braid knots for the following reason. It is not hard to see that for every -braid knot , there is a nonorientable band move to a -bridge knot , which is alternating [Goo72]. This implies that the nonorientable cobordism distance between and is bounded from above by . On the other hand, using that and induce homomorphisms (see Section 2.2 and [Mur65]), the inequality implies that
where we used by 2.1.
Note that a similar argument shows that for all -braid knots , using two nonorientable band moves to transform into a -bridge link, which is also alternating.
4.4.2. On the proof technique
As mentioned in the introduction, it came as a surprise to the author that our proof strategy works not only for positive -braid knots, but for all -braid knots. Let us make this more precise.
The proofs in Section 4.2 and Section 4.3 imply, for any -braid knot , the existence of cobordisms and of genus and between and (connected sums of) torus knots and , respectively, such that
and
For example, for knots that are closures of positive -braids of Garside normal form (D), the proof of 4.5 shows the existence of such a cobordism for as in the proof of 4.4; and the existence of such a cobordism between and follows from the proof of 4.9.
The same strategy would work to determine the concordance invariants and for all positive -braid knots . Indeed, every positive -braid knot can be realized as the slice of a cobordism between the unknot and a torus knot of braid index such that [FLL22, Proposition 4.1]. However, in contrast, there are -braid knots where this strategy provably fails to determine and . A concrete example is the -braid knot — the closure of [LM21] — which is not squeezed [FLL22, Example 3.1]. This means that every cobordism between two connected sums of torus knots and that has as a slice satisfies .
4.4.3. Comparison of the normal forms for -braids
An algorithm described in [BM93, Section 7] as Schreier’s solution to the conjugacy problem [Sch24] can be used to convert -braids in Garside normal form (cf. 3.4) to -braids in Murasugi normal form (cf. 4.15): If is a -braid of Garside normal form (C), then
and if is of Garside normal form (D), then
In addition, it is easy to see how -braids of Garside normal form (A) or (B) are conjugate to braids of Murasugi normal form (a) or (b).
5. On alternating distances of -braid knots
In this section, we prove 1.2 from the introduction and provide lower and upper bounds on the alternation number and dealternating number of any -braid knot which differ by .
5.1. Alternating distances of positive -braid knots
We will prove the following proposition.
Proposition 5.1.
Remark 5.2.
Some of the cases in 5.1 have already been proved by other authors. Indeed, Feller, Pohlmann and Zentner used the observation (27) below to show that for all , [FPZ18, Theorem 1.1]. The upper bound they used was provided by [Kan10, Theorem 8]; in fact, the equality had already been shown by Kanenobu in half of the cases, namely when is even. Moreover, Abe and Kishimoto [AK10, Theorem 3.1] showed that if is a knot that is the closure of a positive -braid of the form in (C). However, to the best of this author’s knowledge, it is new that for all positive -braid knots . Recall that for all positive -braid knots by Equation 5 from Section 2.1.
Before we prove 5.1, let us provide the necessary definitions and background. The Gordian distance between two knots and is the minimal number of crossing changes needed to transform a diagram of into a diagram of , where the minimum is taken over all diagrams of [Mur85]. The alternation number of a knot is defined as the minimal Gordian distance of the knot to the set of alternating knots [Kaw10], i. e.
The dealternating number of a knot is defined via a more diagrammatic approach [ABB+92]: it is the minimal number such that has a diagram that can be turned into an alternating diagram by crossing changes. It follows from the definitions that
(26) |
for any knot and if and only if is alternating. Note that there are families of knots for which the difference between the alternation number and the dealternating number becomes arbitrarily large [Low15, Theorem 1.1].
In the proof of 5.1, we will use that
(27) |
for any knot . In fact, for all alternating knots , we have
(28) |
for any (see [OS03, Theorem 1.4], [Ras10, Theorem 3] and [OSS17a, Theorem 1.14]), where denotes Rasmussen’s concordance invariant from Khovanov homology [Ras10]. It follows from [Abe09, Theorem 2.1] — which builds on ideas of Livingston [Liv04, Corollary 3] — that the absolute value of the difference of any two of the invariants in (28) is a lower bound on . It was first observed in [FPZ18] that the upsilon invariant fits very well in this context (see also [FLZ17, Lemma 8]).
Another main ingredient of our proof of 5.1 is the inequality
(29) |
for any positive -braid with integers and for [AK10, Lemma 2.2].
Proof of 5.1.
Let be a knot that is the closure of a positive -braid of the form in (C) or (D) from 3.2 with . We claim that then
(30) |
which implies the statement of the proposition for these knots. The two equalities in (30) directly follow from our computations of in 4.2 and Equation 5 applied to . The first two inequalities are direct consequences of the inequalities (27) and (26). Finally, the last inequality follows from inequality (29) applied to the particular braid representatives of considered in the proof of 4.12.
For torus knots of braid index , the statement follows analogously. More precisely, if for and , then by Equations 4 and 12, we have In addition, the inequality in (29) applied to the particular braid representatives of considered in the proof of 4.12 implies that . ∎
From 5.1, it is easy to deduce that the alternating positive -braid knots are precisely the unknot and the connected sums of two torus knots of braid index for . This was already known; in fact, the stronger statement is true that the only prime alternating positive braid knots are the torus knots of braid index [Baa13, Corollary 3]. Note that by [Mor79] (see also [BM93, Corollary 7.2]), the only composite -braid knots are the connected sums for .
By [Abe09, Theorem 1.1], the only torus knots with alternation number are the torus knots and . A knot with dealternating number is called almost alternating.
Corollary 5.3.
A positive -braid knot is almost alternating if and only if it is one of the torus knots and or it is represented by a braid of the form
for some integers .
Proof.
This follows directly from 5.1. ∎
Remark 5.4.
In particular, the seven positive -braid knots with crossing number (cf. [LM21]) are all almost alternating.
Remark 5.5.
Our results imply that the Turaev genus equals the alternation number for all positive -braid knots. Indeed, let be a knot that is the closure of a positive braid of the form in (C) or (D) with . Then we have
(31) |
where denotes the Turaev genus of the knot . The Turaev genus of a knot is another alternating distance [Low15], which was first defined in [DFK+08] as the minimal genus of a Turaev surface , where the minimum is taken over all diagrams of . The Turaev surface is a closed orientable surface embedded in associated to the diagram . It is formed by building the natural cobordism between the circles in the two extreme Kauffman states (the all--state and the all--state) of the diagram via adding saddles for each crossing of , and then capping off the boundary components with disks. More details on the definition can be found e. g. in a survey by Champanerkar and Kofman [CK14].
Remark 5.6.
In [FLZ17], Friedl, Livingston and Zentner introduce the invariant , the minimal number of double point singularities in a generically immersed concordance from a knot to an alternating knot. In the case that the alternating knot is the unknot, this is the well studied invariant called the -dimensional clasp number [Shi74]. A sequence of crossing changes in a diagram of a knot leading to a diagram of an alternating knot realizes an immersed concordance from to where any crossing change gives rise to a double point singularity in the concordance. We thus have for any knot , which resembles the inequality between the -dimensional clasp number and the unknotting number of . Moreover, we have for any knot [FLZ17, Theorem 18], so 5.1 implies for all positive -braid knots .
We are now ready to prove 1.2 from the introduction.
5.2. Bounds on the alternation number of general -braid knots
In the following, we turn our attention to -braid knots in general, which are not necessarily the closure of positive -braids. We will use that
(32) |
for any knot , which follows from [Abe09, Theorem 2.1], see also equation (28) from Section 5.1. Rasmussen’s invariant was computed for all -braid knots in Murasugi normal form (cf. 4.15) by Greene.222These computations were generalized to all links that are closures of -braids in [Mar19].
Corollary 5.7.
Let for some , and for such that is a knot. Then
Proof of 5.7.
The lower bound on the alternation number follows from the inequality (32), Theorem 1.1 and the values of the invariant for [Gre14, Proposition 2.4], namely
Moreover, it follows from [AK10, Theorem 2.5] that . ∎
Remark 5.8.
An alternative way to prove the upper bound on in 5.7 for follows from our observations in the proof of 4.18. In fact, the braid diagrams given by the braid representatives of considered in that proof can easily be transformed into alternating diagrams by crossing changes: it is enough to change the positive crossings corresponding to the single generators in to negative crossings; we obtain generators in the corresponding braid words which then correspond to alternating braid diagrams.
Remark 5.9.
If is represented by a -braid of Garside normal form (C) or (D) (see 3.4), then using the observations in Section 4.4.3, 5.7 implies
(33) | |||||
By 5.1, the lower bound in (33) is sharp whenever is the closure of a positive -braid of Garside normal form (C) or (D). However, there are examples where the upper bound in (33) is sharp. The two easiest such examples in terms of crossing number are the non-alternating knots and , which are represented by the -braids (cf. [LM21])
respectively. The lower bound on the alternation number from (33) is in both cases. Indeed, by [Bal08, Theorem 8.6] both knots are quasialternating, so all the invariants from equation (28) are equal [Bal08, Proposition 1.4], [MO08], [OSS17a].
Remark 5.10.
In a similar fashion as 5.7, the Turaev genus of all -braid knots was determined up to an additive error of at most 1 by Lowrance in [Low11, Proposition 4.15] using his computation of the Khovanov width for these knots. More precisely, we have
for any knot that is represented by for some , and for .
6. The fractional Dehn twist coefficient of -braids in Garside normal form
In this section, we compute the fractional Dehn twist coefficient of any -braid in Garside normal form (cf. 3.4).
The fractional Dehn twist coefficient is a homogeneous quasimorphism on the braid group that assigns to any -braid a rational number . Here, a quasimorphism on a group is any map such that
where is called the defect of . A quasimorphism is called homogeneous if for all and . Any homogeneous quasimorphism is invariant under conjugation, so is invariant under the conjugacy class of .
The fractional Dehn twist coefficient first appeared in [GO89] in a different language. It can be defined for mapping classes of general surfaces with boundary, where we here view braids as mapping classes of the times punctured closed disk. Malyutin defined the fractional Dehn twist coefficient , , for all braid groups and showed that its defect is if and if [Mal04, Theorem 6.3]. We refer the reader to [Mal04] for a more detailed account.
Corollary 6.1.
Remark 6.2.
In the proof of 6.1, we will use that the fractional Dehn twist coefficient of any -braid is completely determined by the writhe and the homogenized upsilon invariant of : we have
(34) |
for any -braid . The invariant is another real-valued homogeneous quasimorphism on the braid group which can be defined as
More generally, Brandenbursky [Bra11, Theorem 2.6] showed that a homogeneous quasimorphism can be assigned to any concordance homomorphism that is bounded above by a constant multiple of the -genus. We refer the reader to [Bra11] or [FH19, Appendix A] for more details on homogenized concordance invariants.
Proposition 6.3.
Let be a -braid. Then
Proof of 6.3.
We will use that if and commute [FH19, Lemma A.1]. In particular, for any -braid and any , we have
(35) |
Moreover, by the definition of , Equation 12 and the homogeneity of , we have
(36) |
We will now compute for the positive -braids of the form (A)–(D), i. e. assuming in (A)–(D). The statement of 6.3 will then follow from (35) and (36).
Proof of 6.1.
This follows directly from 6.3, Equation 34, and a straightforward calculation of the writhe of the braids in (A)–(D). ∎
Remark 6.4.
References
- [ABB+92] Colin C. Adams, Jeffrey F. Brock, John Bugbee, Timothy D. Comar, Keith A. Faigin, Amy M. Huston, Anne M. Joseph, and David Pesikoff. Almost alternating links. Topology Appl., 46(2):151–165, 1992.
- [Abe09] Tetsuya Abe. An estimation of the alternation number of a torus knot. J. Knot Theory Ramifications, 18(3):363–379, 2009.
- [AK10] Tetsuya Abe and Kengo Kishimoto. The dealternating number and the alternation number of a closed 3-braid. J. Knot Theory Ramifications, 19(9):1157–1181, 2010.
- [Ale23] J.W. Alexander. A lemma on systems of knotted curves. Proceedings of the National Academy of Sciences of the United States of America, 9(3):93–95, March 1923.
- [Art25] E. Artin. Theorie der Zöpfe. Abhandlungen aus dem Mathematischen Seminar der Universität Hamburg, 4:47–72, 1925.
- [Baa13] Sebastian Baader. Positive braids of maximal signature. Enseign. Math., 59(3-4):351–358, 2013.
- [Bal08] John A. Baldwin. Heegaard Floer homology and genus one, one-boundary component open books. J. Topol., 1(4):963–992, 2008.
- [Bal20] William Ballinger. Concordance invariants from the spectral sequence on khovanov homology. arXiv e-prints, page arXiv:2004.10807, April 2020.
- [BB05] Joan S. Birman and Tara E. Brendle. Braids: a survey. In Handbook of knot theory, pages 19–103. Elsevier B. V., Amsterdam, 2005.
- [Ben83] Daniel Bennequin. Entrelacements et équations de Pfaff. In Third Schnepfenried geometry conference, Vol. 1 (Schnepfenried, 1982), volume 107 of Astérisque, pages 87–161. Soc. Math. France, Paris, 1983.
- [BM93] Joan S. Birman and William W. Menasco. Studying links via closed braids. III. Classifying links which are closed -braids. Pacific J. Math., 161(1):25–113, 1993.
- [Bra11] Michael Brandenbursky. On quasi-morphisms from knot and braid invariants. J. Knot Theory Ramifications, 20(10):1397–1417, 2011.
- [Cho48] Wei-Liang Chow. On the algebraical braid group. Annals of Mathematics, 49(3):654–658, 1948.
- [CK14] Abhijit Champanerkar and Ilya Kofman. A survey on the Turaev genus of knots. Acta Math. Vietnam., 39(4):497–514, 2014.
- [DFK+08] Oliver T. Dasbach, David Futer, Efstratia Kalfagianni, Xiao-Song Lin, and Neal W. Stoltzfus. The Jones polynomial and graphs on surfaces. J. Combin. Theory Ser. B, 98(2):384–399, 2008.
- [DL11] Oliver T. Dasbach and Adam M. Lowrance. Turaev genus, knot signature, and the knot homology concordance invariants. Proc. Amer. Math. Soc., 139(7):2631–2645, 2011.
- [Erl99] Dieter Erle. Calculation of the signature of a 3-braid link. Kobe J. Math., 16(2):161–175, 1999.
- [Fel16] Peter Feller. Optimal cobordisms between torus knots. Communications in Analysis and Geometry, 24(5):993–1025, 2016.
- [FH19] Peter Feller and Diana Hubbard. Braids with as many full twists as strands realize the braid index. Journal of Topology, 12(4):1069–1092, May 2019.
- [FK17] Peter Feller and David Krcatovich. On cobordisms between knots, braid index, and the upsilon-invariant. Mathematische Annalen, 369(1-2):301–329, Mar 2017.
- [FLL22] Peter Feller, Lukas Lewark, and Andrew Lobb. Squeezed knots. arXiv preprint arXiv:2202.12289, 2022.
- [FLZ17] Stefan Friedl, Charles Livingston, and Raphael Zentner. Knot concordances and alternating knots. Michigan Math. J., 66(2):421–432, 2017.
- [FPZ18] Peter Feller, Simon Pohlmann, and Raphael Zentner. Alternation numbers of torus knots with small braid index. Indiana University Mathematics Journal, 67(2):645–655, 2018.
- [Gar69] F. A. Garside. The braid group and other groups. The Quarterly Journal of Mathematics, 20(1):235–254, 01 1969.
- [GO89] David Gabai and Ulrich Oertel. Essential laminations in -manifolds. Ann. of Math. (2), 130(1):41–73, 1989.
- [Goo72] R. E. Goodrick. Two bridge knots are alternating knots. Pacific J. Math., 40:561–564, 1972.
- [Gre14] Joshua Evan Greene. Donaldson’s theorem, Heegaard Floer homology, and knots with unknotting number one. Adv. Math., 255:672–705, 2014.
- [HKK+21] Diana Hubbard, Keiko Kawamuro, Feride Ceren Kose, Gage Martin, Olga Plamenevskaya, Katherine Raoux, Linh Truong, and Hannah Turner. Braids, fibered knots, and concordance questions, 2021. arXiv:2004.07445.
- [Hom17] Jennifer Hom. A survey on Heegaard Floer homology and concordance. J. Knot Theory Ramifications, 26(2):1740015, 24, 2017.
- [Kan10] Taizo Kanenobu. Upper bound for the alternation number of a torus knot. Topology Appl., 157(1):302–318, 2010.
- [Kaw10] Akio Kawauchi. On alternation numbers of links. Topology Appl., 157(1):274–279, 2010.
- [KM93] P. B. Kronheimer and T. S. Mrowka. Gauge theory for embedded surfaces. I. Topology, 32(4):773–826, 1993.
- [Lew14] Lukas Lewark. Rasmussen’s spectral sequences and the -concordance invariants. Adv. Math., 260:59–83, 2014.
- [Liv04] Charles Livingston. Computations of the Ozsváth-Szabó knot concordance invariant. Geom. Topol., 8:735–742, 2004.
- [Liv17] Charles Livingston. Notes on the knot concordance invariant upsilon. Algebr. Geom. Topol., 17(1):111–130, 2017.
- [LM21] Charles Livingston and Allison H. Moore. Knotinfo: Table of knot invariants. URL: knotinfo.math.indiana.edu, June 14, 2021.
- [Low11] Adam Lowrance. The Khovanov width of twisted links and closed 3-braids. Comment. Math. Helv., 86(3):675–706, 2011.
- [Low15] Adam M. Lowrance. Alternating distances of knots and links. Topology Appl., 182:53–70, 2015.
- [Mal04] A. V. Malyutin. Writhe of (closed) braids. Algebra i Analiz, 16(5):59–91, 2004.
- [Mar19] Gage Martin. Annular rasmussen invariants: Properties and 3-braid classification, 2019. arXiv:1909.09245.
- [MO08] Ciprian Manolescu and Peter Ozsváth. On the Khovanov and knot Floer homologies of quasi-alternating links. In Proceedings of Gökova Geometry-Topology Conference 2007, pages 60–81. Gökova Geometry/Topology Conference (GGT), Gökova, 2008.
- [Mor79] H. R. Morton. Closed braids which are not prime knots. Math. Proc. Cambridge Philos. Soc., 86(3):421–426, 1979.
- [Mur65] Kunio Murasugi. On a certain numerical invariant of link types. Trans. Amer. Math. Soc., 117:387–422, 1965.
- [Mur74] Kunio Murasugi. On closed -braids. American Mathematical Society, Providence, R.I., 1974. Memoirs of the American Mathmatical Society, No. 151.
- [Mur85] Hitoshi Murakami. Some metrics on classical knots. Math. Ann., 270(1):35–45, 1985.
- [OS03] Peter Ozsváth and Zoltán Szabó. Knot Floer homology and the four-ball genus. Geom. Topol., 7:615–639, 2003.
- [OSS17a] P. Ozsváth, A. I. Stipsicz, and Z. Szabó. Concordance homomorphisms from knot Floer homology. Advances in Mathematics, 315:366–426, Jul 2017.
- [OSS17b] Peter S. Ozsváth, András I. Stipsicz, and Zoltán Szabó. Unoriented knot Floer homology and the unoriented four-ball genus. Int. Math. Res. Not. IMRN, (17):5137–5181, 2017.
- [Ras03] J. A. Rasmussen. Floer homology and knot complements. ProQuest LLC, Ann Arbor, MI, 2003. Thesis (Ph.D.)–Harvard University.
- [Ras10] Jacob Rasmussen. Khovanov homology and the slice genus. Invent. Math., 182(2):419–447, 2010.
- [Rud93] Lee Rudolph. Quasipositivity as an obstruction to sliceness. Bull. Amer. Math. Soc. (N.S.), 29(1):51–59, 1993.
- [Sch24] Otto Schreier. Über die Gruppen . Abh. Math. Sem. Univ. Hamburg, 3(1):167–169, 1924.
- [Shi74] Tetsuo Shibuya. Some relations among various numerical invariants for links. Osaka Math. J., 11:313–322, 1974.
- [Tro62] H. F. Trotter. Homology of group systems with applications to knot theory. Annals of Mathematics, 76(3):464–498, 1962.