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

Signed Heights of Knotoids

Larsen Linov
Abstract.

The height of a knotoid is a measure of how far it is from being a knot. Here we define the positive and negative parts of the height, and we prove that they determine the unsigned height. Some polynomial invariants provide lower bounds for the signed heights. We also study a set of sequences associated to a knotoid.

1. Introduction

The theory of knotoids, introduced by Turaev in [2], is an extension of classical knot theory. Knotoids have been studied recently in [3], [4], [5], and [7], and they have also been used for studying proteins.

1.1. Knotoids and Sign Sequences

A knotoid diagram is an immersion of an interval into S2S^{2} with only transverse double crossings, together with over/under crossing information. A knotoid is a class of knotoid diagrams up to planar isotopy and the Reidemeister moves performed away from the endpoints. (We will provide a geometric definition as well, in Section 2.6.) Knotoids are always considered to be oriented, that is, the endpoints are labelled as the tail v0v_{0} and the head v1v_{1}.

A shortcut for a knotoid diagram KK is an embedded path from v0v_{0} to v1v_{1} that intersects KK transversely and does not intersect the crossings. Of course, every knotoid diagram has many shortcuts. The intersections between KK and a shortcut are signed; see Figure 1. (The endpoints are not considered to be intersections.)

Refer to caption
Figure 1. The signs of intersections with a shortcut.

A pair (K,a)(K,a), where aa is a shortcut for KK, will be called a shortcut diagram. We can connect any two shortcut diagrams for a knotoid by planar isotopy, the Reidemeister moves away from the shortcut, and the three shortcut moves shown in Figure 2.

Refer to caption
Figure 2. The shortcut moves, Types I, II, and III.
Refer to caption
Figure 3. The bifoil, an example of a nontrivial knotoid. The sign sequence of this shortcut diagram is (+)(+).

For each diagram KK and shortcut aa, the height h(K,a)h(K,a) of the pair is the number of intersection points with aa. A sign sequence is a finite sequence with values in {+,}\{+,-\}, and for a shortcut diagram (K,a)(K,a) we define Seq(K,a)\operatorname{Seq}(K,a) to be the sign sequence of length h(K,a)h(K,a) expressing the signs of the intersections between KK and aa in the order they appear when following KK from v0v_{0} to v1v_{1}. The positive (resp. negative) height h±(K,a)h_{\pm}(K,a) is the number of appearances of ++ (resp. -) in Seq(K,a)\operatorname{Seq}(K,a).

These values give rise to natural invariants of knotoids:

Definition 1.1.

For a knotoid kk, the height of kk is the minimum of the heights h(K,a)h(K,a) over all shortcut diagrams representing kk. (This is the complexity in [2].) We define the signed heights h±(k)h_{\pm}(k) similarly.

Definition 1.2.

A sign sequence is attainable for kk if it is Seq(K,a)\operatorname{Seq}(K,a) for some shortcut diagram representing kk. We will care in particular about the minimal attainable sequences, that is, those realizing the height.

Remark.

There are several related theories not considered in this paper. In particular, knotoids on 2\mathbb{R}^{2}, virtual knotoids, and multi-knotoids are interesting generalizations.

1.2. Main Results

There is a simple relationship between the height of a knotoid kk and its signed heights.

Theorem 1.3.

For all kk, h(k)=h+(k)+h(k)h(k)=h_{+}(k)+h_{-}(k).

1.3 reduces questions about the height of a knotoid to questions about its signed heights, which form a height pair (h+,h)(h_{+},h_{-}). This will make it easier to compute the heights of some knotoids. The theorem also implies that all minimal attainable sequences for a knotoid are rearrangements of each other. The next theorem provides another restriction on the set of minimal attainable sequences.

For a sign sequence AA, a left shift move of size nn on AA is the result of deleting nn appearances of (,+)(-,+) as a consecutive subsequence and then inserting nn copies of (+,)(+,-). Similarly, a right shift move deletes copies (+,)(+,-) and adds copies of (,+)(-,+). The deletions and insertions all happen at the same time. For example, a nontrivial shift move on (,,+,+)(-,-,+,+) must be a left shift of size 11, deleting the second and third entries. The possible results after inserting (+,)(+,-) are (+,,,+)(+,-,-,+), (,+,,+)(-,+,-,+), and (,+,+,)(-,+,+,-).

Theorem 1.4.

If AA and AA^{\prime} are two minimal attainable sequences for kk, there is a sequence of minimal attainable sequences

A=A0,A1,An=AA=A_{0},A_{1}\ldots,A_{n}=A^{\prime}

such that each Ai+1A_{i+1} differs from AiA_{i} by a shift move.

The signed heights of a knotoid can be bounded by some polynomial invariants, in particular the Turaev polynomial k\langle\kern-5.16663pt~{}\langle k\rangle\kern-5.16663pt~{}\rangle_{\circ} of [2] and the index polynomial FkF_{k} of [5]. For a nonzero Laurent polynomial p(t)[t±1]p(t)\in\mathbb{Z}[t^{\pm 1}], we will write deg+(p)\deg^{+}(p) for max{maxdeg(p),0}\max\{\operatorname{maxdeg}(p),0\} and deg(p)\deg^{-}(p) for max{mindeg(p),0}\max\{-\operatorname{mindeg}(p),0\}. We also set deg±(0)=0\deg^{\pm}(0)=0. For a Laurent polynomial in multiple variables, the signed degree in a specific variable will be denoted by (for example) degt±\deg_{t}^{\pm}.

Theorem 1.5.

For a knotoid kk, h±(k)deg±(Fk)h_{\pm}(k)\geq\deg^{\pm}(F_{k}).

Theorem 1.6.

For a knotoid kk, 2h±(k)degu(k)2h_{\pm}(k)\geq\deg_{u}^{\mp}(\langle\kern-5.16663pt~{}\langle k\rangle\kern-5.16663pt~{}\rangle_{\circ}).

The index polynomial also gives more specific information about attainable sign sequences.

Theorem 1.7.

Any attainable sequence for kk must have a consecutive subsequence adding up to deg+(Fk)\deg^{+}(F_{k}), and a consecutive subsequence adding up to deg(Fk)-\deg^{-}(F_{k}).

In the theorem above, of course, we treat ++ terms as +1+1 and - as 1-1.

Theorem 1.8.

If kk is a knotoid such that the bounds in 1.5 are equalities, then kk has a unique minimal attainable sign sequence.

1.3. Organization

In Section 2, we give background information on knotoids, including basic operations and the relationship between knotoids and theta-curves. In Section 3, we prove 1.3 and 1.4. Section 4 addresses the behavior of height and sign sequences under knotoid operations. In Section 5, we recall background on nn-writhes and the Turaev polynomial, and we prove Theorems 1.5 through 1.8. Section 6 contains interesting examples and applications to knotoids of low height.

1.4. Acknowledgements

I would like to thank my advisor, Ian Agol, for his thorough and careful feedback. I am also grateful to Kyle Miller for our very helpful conversations.

2. Background on Knotoids and Theta-Curves

2.1. Closures and Knot-Type Knotoids

The first examples of knotoid invariants are the over- and underpass closures. A shortcut diagram for kk gives rise to an oriented knot diagram by incorporating the shortcut into the diagram to create an immersion of a circle into S1S^{1}. By taking the shortcut to pass over (resp. under) the knotoid at each crossing, we obtain a diagram of the overpass (resp. underpass) closure of kk, denoted k+k_{+} (resp. kk_{-}).

Conversely, given a diagram of an oriented knot κ\kappa, and a point on an edge of the diagram, we may obtain a knotoid diagram by deleting an open interval around the chosen point. The resulting knotoid depends only on κ\kappa and is denoted κ\kappa^{\bullet}. By construction, κ\kappa^{\bullet} has height 0 and (κ)±=κ(\kappa^{\bullet})_{\pm}=\kappa. Any knotoid of height 0 may be obtained in this way, and so we have a natural identification of the set of height-0 knotoids with oriented knots. Such knotoids are called knot-type. Knotoids that are not knot-type are proper.

2.2. Multiplication

The set of knotoid types has a natural noncommutative product. Given knotoid diagrams K1K_{1} and K2K_{2}, we may form a diagram K1K2K_{1}K_{2} by deleting small open disks around the head of K1K_{1} and tail of K2K_{2}, then gluing appropriately along the boundaries. The resulting knotoid depends only on the knotoids represented by K1K_{1} and K2K_{2}. It also makes sense to refer to the product of two shortcut diagrams as another shortcut diagram.

Knotoid multiplication is associative. The trivial knotoid, the one that can be drawn without crossings, is an identity. The over/underpass closure operations and the κκ\kappa\mapsto\kappa^{\bullet} operation are monoid homomorphisms.

A prime knotoid is one that cannot be written as a nontrivial product. Every knotoid has a unique decomposition of the form

κk1k2kn,\kappa^{\bullet}k_{1}k_{2}\cdots k_{n},

where each kik_{i} is a proper prime knotoid. A knot-type knotoid is prime if and only if the corresponding knot is prime. Two distinct prime knotoids commute if and only if one or both is knot-type ([2]).

2.3. Basic Involutions

For a knotoid kk, the reverse rev(k)\operatorname{rev}(k) is obtained by switching the orientation on a diagram of KK, that is, swapping the labels of the vertices. The mirror image mir(k)\operatorname{mir}(k) is obtained from switching the over/under information on each crossing, and the symmetry operation acts by reversing the orientation of the ambient S2S^{2}. Rotation is the composition of symmetry and mirror image reflection. See Figure 4. The basic involutions generate a group isomorphic to (/2)3(\mathbb{Z}/2\mathbb{Z})^{3}.

Refer to caption
Figure 4. From left to right, the bifoil φ1\varphi_{1}, rev(φ1)\operatorname{rev}(\varphi_{1}), mir(φ1)\operatorname{mir}(\varphi_{1}), sym(φ1)\operatorname{sym}(\varphi_{1}), and rot(φ1)\operatorname{rot}(\varphi_{1}). Only the first two are equivalent.

2.4. Lifting

Given a diagram KK of some knotoid kk and positive nn, we may choose a lift of KK to the nn-fold cover of S2S^{2} branched over v0v_{0} and v1v_{1}. The result is a new diagram K/nK/n with orientation and over/under information inherited from KK, and k/nk/n is a well-defined knotoid. See Figure 5. For all kk, the sequence k/nk/n stabilizes to a knot-type knotoid; we define k/k/\infty to be the corresponding knot.

Refer to caption
Figure 5. Lifts of the spiral knotoid φ2\varphi_{2}: φ2/1=φ2\varphi_{2}/1=\varphi_{2}, φ2/2=φ1\varphi_{2}/2=\varphi_{1}, and φ2/3=(φ2/)=1\varphi_{2}/3=(\varphi_{2}/\infty)^{\bullet}=1. Here we use periodic diagrams: Given a knotoid diagram KK on S2S^{2}, we may delete regular neighborhoods of the endpoints to obtain a diagram on an annulus, with one end on each boundary component. Then on ×I\mathbb{R}\times I we draw the preimage of the diagram under the covering map.

A similar construction is studied in [4]: The entire preimage of KK under the double cover of S2S^{2} branched over the endpoints constitutes a diagram of an unoriented knot. The unoriented knot is an invariant and is called the double branched cover of kk.

2.5. Framings

A framing of a knotoid kk is a class of diagrams of kk up to regular isotopy, that is, up to Reidemeister moves I’, II, and III. Two diagrams of kk are in the same framing class if and only if they have the same writhe.

Similarly, a shortcut framing for kk is a class of shortcut diagrams related by all moves except the Type I shortcut move. Shortcut framings are classified by algebraic intersection number between the main strand KK and the shortcut aa.

Refer to caption
Figure 6. Reidemeister moves, Types I’, I, II, and III.

2.6. Simple Theta-Curves

The theta graph Θ\Theta is the graph with two vertices, v0v_{0} and v1v_{1}, and three oriented edges e0e_{0}, e+e_{+}, and ee_{-} from v0v_{0} to v1v_{1}. A theta-curve θ\theta is an embedding of Θ\Theta into S3S^{3}. Such a curve is simple if (the image of) e+ee_{+}\cup e_{-} is unknotted. A spanning disk DD for a simple theta-curve is a choice of embedded disk with boundary D=e+e\partial D=e_{+}\cup e_{-} such that DD intersects e0e_{0} transversely.

Like knotoids, isotopy classes of simple theta-curves form a monoid: The product of θ1\theta_{1} and θ2\theta_{2} is formed by deleting small balls around v1v_{1} in θ1\theta_{1} and v0v_{0} in θ2\theta_{2}, and gluing the boundaries so that each edge of θ1\theta_{1} is glued to the edge of θ2\theta_{2} with the same label. The resulting theta-curve is well-defined up to isotopy, because the pure mapping class group of a thrice-punctured sphere is trivial. Similarly, we may multiply isotopy classes of pairs (θ,D)(\theta,D).

The height h(θ)h(\theta) of θ\theta is the minimal number of intersections of a spanning disk with e0e_{0}, and the positive and negative heights h±h_{\pm} are the minimal numbers of intersections of those signs. The sign sequence associated to (θ,D)(\theta,D) is the sequence of signs of the intersections of e0e_{0} with DD, in the order along e0e_{0} from v0v_{0} to v1v_{1}. Sequences obtained this way are attainable for θ\theta.

Refer to caption
Figure 7. A positive intersection of e0e_{0} with a spanning disk.

There is a natural map τ\tau from the set of knotoids to the set of isotopy classes of simple theta-curves. Given a shortcut diagram (K,a)(K,a), for a knotoid kk, we may form a simple theta-curve by considering the diagram as lying in a neighborhood of S2S3S^{2}\subset S^{3}, with KK lifting to an embedded path e0e_{0}, and adding edges e+e_{+} and ee_{-} over and under aa. The resulting theta-curve represents τ(k)\tau(k). Note that, in τ(k)\tau(k), the isotopy class of e0e±e_{0}\cup e_{\pm} is k±k_{\pm}.

Theorem 2.1 (Turaev [2]).

The map τ\tau is a monoid isomorphism.

The construction above also yields a correspondence between shortcut diagrams for kk and spanning disks for τ(k)\tau(k): For each such diagram (K,a)(K,a), we may choose DD to be the “vertical” disk between e+e_{+} and ee_{-} in the constructed theta-curve θ\theta. Under this construction, the sign sequence of (θ,D)(\theta,D) is the same as that of (K,a)(K,a). Every isotopy class of spanned embeddings (θ,D)(\theta,D) for τ(k)\tau(k) may be obtained in this way from a shortcut diagram for kk. Thus, heights and attainable sign sequences for kk correspond directly with heights and attainable sequences for τ(k)\tau(k). Isotopy classes of spanned embeddings (θ,D)(\theta,D) for τ(k)\tau(k) correspond to classes of shortcut diagrams of kk under planar isotopy, the Reidemeister moves away from aa, and the Type III shortcut move.

3. Comparing Attainable Sequences

Here we prove Theorems 1.3 and 1.4. But first, we establish that allowing certain self-intersections in spanning disks would not reduce the height of a simple theta-curve.

Lemma 3.1.

Suppose, for some simple theta-curve θ\theta, that φ:ΔS3\varphi:\Delta\to S^{3} is an immersed (not necessarily embedded) disk such that (a) φ(Δ)=e+e\varphi(\partial\Delta)=e_{+}\cup e_{-}, (b) e0e_{0} intersects φ\varphi transversely, and (c) the self-intersections of Δ\Delta under φ\varphi are disjoint circles identified transversely in pairs. Then φ(Δ)\varphi(\Delta) has at least h(θ)h(\theta) intersections with e0e_{0}.

Proof.

There is some finite number of intersecting pairs of circles on Δ\Delta. Any circle of self-intersection has a neighborhood in which φ\varphi looks like the product of a plus sign with a circle. There are two ways of resolving the intersection by smoothing. The resolution that preserves the number of components creates a new immersion of a disk. Note that this resolution might not respect orientation. Replacing φ\varphi by this new immersion, we have reduced the number of self-intersections without changing the number of intersections with e0e_{0}. Proceeding in this fashion shows that there is a spanning disk with the same number of e0e_{0} intersections as φ\varphi. ∎

We will say that two spanning disks D1D_{1} and D2D_{2} for a theta-curve θ\theta are compatible if their interiors are disjoint. Compatible pairs of spanning disks are useful because of the following lemma.

Lemma 3.2.

If D1D_{1} and D2D_{2} are compatible spanning disks for a simple theta-curve θ\theta and h(θ,D1)h(θ,D2)h(\theta,D_{1})\geq h(\theta,D_{2}), then h±(θ,D1)h±(θ,D2)h_{\pm}(\theta,D_{1})\geq h_{\pm}(\theta,D_{2}).

Proof.

Because D1D_{1} and D2D_{2} are compatible, D1D2D_{1}\cup D_{2} is an embedded sphere dividing S3S^{3} into two balls. Note that D1D2D_{1}\cup D_{2} is not naturally oriented, because the orientations of D1D_{1} and D2D_{2} agree on e+ee_{+}\cup e_{-}. Let Σ\Sigma denote the oriented sphere D1(D2)D_{1}\cup(-D_{2}).

The net number of intersections of e0e_{0} with Σ\Sigma, not including v0v_{0} or v1v_{1}, must be 0, 11, or 1-1. In symbols,

(1) 1h+(D1)h(D1)h+(D2)+h(D2)1.-1\leq h_{+}(D_{1})-h_{-}(D_{1})-h_{+}(D_{2})+h_{-}(D_{2})\leq 1.

(Here we have suppressed θ\theta in the notation.)

By assumption, we have

(2) h+(D1)+h(D1)h+(D2)+h(D2).h_{+}(D_{1})+h_{-}(D_{1})\geq h_{+}(D_{2})+h_{-}(D_{2}).

Combining (2) with each inequality in (1), we obtain 2h+(D1)2h+(D2)12h_{+}(D_{1})-2h_{+}(D_{2})\geq-1 and 2h(D1)2h(D2)12h_{-}(D_{1})-2h_{-}(D_{2})\geq-1. Therefore, h±(D1)h±(D2)h_{\pm}(D_{1})\geq h_{\pm}(D_{2}), as desired. ∎

The next fact provides opportunities to apply 3.2.

Lemma 3.3.

If DD and DD^{\prime} are spanning disks for a simple theta-curve θ\theta and DD realizes the height of θ\theta, then there is a sequence of spanning disks

D=D0,D1,,Dn=DD^{\prime}=D_{0},D_{1},\ldots,D_{n}=D

such that consecutive disks are compatible and the sequence (h(θ,Di))(h(\theta,D_{i})) is nonincreasing.

Proof.

We may choose D1D_{1} such that (θ,D1)(\theta,D_{1}) is isotopic to (θ,D)(\theta,D^{\prime}) and such that DD and D1D_{1} intersect transversely away from C=e+eC=e_{+}\cup e_{-}. Then the intersections consist of CC and a system 𝒞1\mathcal{C}_{1} of disjoint circles and arcs embedded properly in both DD and D1D_{1}. We may also require in our choice of D1D_{1} that none of the intersection curves meet e0e_{0}. For each i1i\geq 1, once we have chosen DiD_{i} we will form Di+1D_{i+1} in such a way that Di+1D_{i+1} has fewer total components of intersection with DD than does DiD_{i}. Let 𝒞i\mathcal{C}_{i} be the system of intersections between DiD_{i} and DD, not including CC.

Case 1: 𝒞i\mathcal{C}_{i} is empty.

If 𝒞i\mathcal{C}_{i} is empty, then DD is compatible with DiD_{i}, so we set n=i+1n=i+1 and Dn=DD_{n}=D. Because DD realizes the height of θ\theta, h(Di)h(D)h(D_{i})\geq h(D).

Case 2: 𝒞i\mathcal{C}_{i} has an arc, but no circles.

If 𝒞i\mathcal{C}_{i} contains an arc, but no circles, we can find an innermost such arc ss on DD. By innermost arc, we mean one for which all other arcs lie on one side of ss in DD. In particular, the endpoints of ss divide CC into two segments tt and tt^{\prime} such that the endpoints of all other curves of 𝒞i\mathcal{C}_{i} lie on tt^{\prime}. Let EE be the disk in DD bounded by S=stS=s\cup t, and EE^{\prime} the disk in DiD_{i} bounded by SS. Since tt contains no endpoints of the arcs in 𝒞i\mathcal{C}_{i}, ss is also innermost in DiD_{i}. In particular, EDi=ED=SE\cap D_{i}=E^{\prime}\cap D=S.

Since (DE)E(D-E)\cup E^{\prime} is an embedded disk with boundary CC, its height is at least h(D)h(D). Therefore, EE^{\prime} has at least as many intersections with e0e_{0} as does EE. Now let Di+1D_{i+1} be the result of slightly perturbing (DiE)E(D_{i}-E^{\prime})\cup E to be compatible with DiD_{i}. Then Di+1D_{i+1} has fewer intersection curves than DiD_{i} with DD, and h(Di+1)h(Di)h(D_{i+1})\leq h(D_{i}).

Case 3: 𝒞i\mathcal{C}_{i} contains a circle.

If there is at least one circle, there is an innermost circle SS in DD. Then SS bounds a disk EDE\subset D with EDi=SE\cap D_{i}=S. Let EE^{\prime} denote the disk in DiD_{i} bounded by SS. In contrast with Case 2, SS is not necessarily innermost in DiD_{i}, so EDE^{\prime}\cap D may be more than just SS.

By 3.1, e0e_{0} has at least as many intersections with (DE)E(D-E)\cup E^{\prime} as with DD, so it intersects EE^{\prime} at least as many times as EE. Therefore, we may proceed as in Case 2. Let Di+1=(DiE)ED_{i+1}=(D_{i}-E^{\prime})\cup E, and perturb it so that it is compatible with DiD_{i}. Then h(Di+1)h(Di)h(D_{i+1})\leq h(D_{i}), and we have reduced the number of intersection curves with DD.

This covers all the cases, so we are done. ∎

3.2 immediately implies that 3.3 can be strengthened as follows.

Lemma 3.4.

If DD and DD^{\prime} are spanning disks for a simple theta-curve θ\theta and DD realizes the height of θ\theta, then there is a sequence of spanning disks

D=D0,D1,,Dn=DD^{\prime}=D_{0},D_{1},\ldots,D_{n}=D

such that consecutive disks are compatible and the sequences (h±(θ,Di))(h_{\pm}(\theta,D_{i})) are both nonincreasing.

We can now prove Theorems 1.3 and 1.4.

Proof of 1.3.

Given a simple theta-curve θ\theta, a spanning disk DD realizing the height, and any other spanning disk DD^{\prime}, 3.4 implies that DD has no greater positive or negative height than DD^{\prime}. Therefore, DD realizes the signed heights, and so,

h(θ)=h(θ,D)=h+(θ,D)+h(θ,D)=h+(θ)+h(θ).h(\theta)=h(\theta,D)=h_{+}(\theta,D)+h_{-}(\theta,D)=h_{+}(\theta)+h_{-}(\theta).

For a knotoid kk, we obtain h(k)=h+(k)+h(k)h(k)=h_{+}(k)+h_{-}(k) by setting θ=τ(k)\theta=\tau(k). ∎

To prove 1.4, we will use another lemma about compatible spanning disks.

Lemma 3.5.

If D1D_{1} and D2D_{2} are compatible spanning disks for a simple theta-curve θ\theta and h(θ,D1)=h(θ,D2)h(\theta,D_{1})=h(\theta,D_{2}), then Seq(θ,D1)\operatorname{Seq}(\theta,D_{1}) and Seq(θ,D2)\operatorname{Seq}(\theta,D_{2}) differ by a shift move.

Proof.

Let Σ\Sigma be the sphere D1(D2)D_{1}\cup(-D_{2}) as in the proof of 3.2. Let BB be the ball in S3S^{3} such that Σ\Sigma is the oriented boundary of BB, and assume that e0{v0,v1}e_{0}-\{v_{0},v_{1}\} “starts” outside of BB. In the overall sequence of intersections of e0e_{0} with Σ\Sigma (not including the endpoints), the signs of the intersections alternate: The odd- and even-index intersections are negative and positive, respectively. Because h(θ,D1)=h(θ,D2)h(\theta,D_{1})=h(\theta,D_{2}), the total number of intersections is even. Each odd-even pair of consecutive intersections has type (+D2,+D1)(+D_{2},+D_{1}), (D1,D2)(-D_{1},-D_{2}), (D1,+D1)(-D_{1},+D_{1}), or (+D2,D2)(+D_{2},-D_{2}). Therefore, Seq(θ,D2)\operatorname{Seq}(\theta,D_{2}) is obtained from Seq(θ,D1)\operatorname{Seq}(\theta,D_{1}) by a left shift move.

If e0e_{0} instead starts in the inside of BB, then Seq(θ,D2)\operatorname{Seq}(\theta,D_{2}) is obtained from Seq(θ,D1)\operatorname{Seq}(\theta,D_{1}) by a right shift move. ∎

Proof of 1.4.

For a simple theta-curve θ\theta, suppose DD and DD^{\prime} are both spanning disks realizing the height of θ\theta. Then in the sequence (Di)(D_{i}) of spanning disks obtained from 3.4, each DiD_{i} realizes the height of θ\theta. Therefore, applying 3.5 to the sequence (Di)(D_{i}) implies that Seq(θ,D)\operatorname{Seq}(\theta,D) and Seq(θ,D)\operatorname{Seq}(\theta,D^{\prime}) are connected among minimal attainable sequences by shift moves. ∎

4. Knotoid Operations and Attainable Sequences

4.1. Signed Heights under the Basic Involutions

The signed heights of knotoids behave in straightforward ways under the basic knotoid involutions.

Proposition 4.1.

For all kk, we have

h±(k)=h±(rev(k))=h±(mir(k))=h(sym(k))=h(rot(k)).h_{\pm}(k)=h_{\pm}(\operatorname{rev}(k))=h_{\pm}(\operatorname{mir}(k))=h_{\mp}(\operatorname{sym}(k))=h_{\mp}(\operatorname{rot}(k)).

More specifically, we can say the following.

Proposition 4.2.

If AA is an attainable sign sequence for kk, then,

  1. (1)

    rev(A)\operatorname{rev}(A) is attainable for rev(k)\operatorname{rev}(k),

  2. (2)

    AA is attainable for mir(k)\operatorname{mir}(k),

  3. (3)

    A-A is attainable for sym(k)\operatorname{sym}(k), and

  4. (4)

    A-A is attainable for rot(k)\operatorname{rot}(k).

where A-A is the result of switching all terms ++\leftrightarrow- in AA and rev(A)\operatorname{rev}(A) is the result of reversing the order.

Proof.

Given a shortcut diagram (K,a)(K,a) for kk, the signs of the crossings between KK and aa are not changed by switching the over/under information of KK or by simultaneously switching the orientations of KK and aa (recall that the orientation for a shortcut is determined by the rest of the diagram). However, changing the orientation on S2S^{2} changes the signs of the intersections. ∎

4.1, together with 1.3, has implications for unsigned heights of knotoids, such as for rotatable knotoids, which are addressed in [4]. A knotoid kk is rotatable if it equals rot(k)\operatorname{rot}(k).

Corollary 4.3.

Every rotatable knotoid has even height.

4.2. Multiplication and Concatenation

In this section we relate the set of attainable sign sequences for a product to the attainable sequences of its factors.

Proposition 4.4.

For any k1k_{1} and k2k_{2}, if A1A_{1} is an attainable sign sequence for k1k_{1} and A2A_{2} is attainable for k2k_{2}, then the concatenation A1A2A_{1}A_{2} is attainable for k1k2k_{1}k_{2}.

Proof.

Given shortcut diagrams for k1k_{1} and k2k_{2}, the sign sequence for the product of the diagrams is the concatenation of the sign sequences for the two original diagrams. ∎

Note that as a particular case of the statement above, if AA is any attainable sequence for a knotoid kk, then the result of appending ++ or - to either end of AA is also attainable for kk, because on any shortcut diagram for kk we can perform a Type I shortcut move around either endpoint.

The next theorem is a converse for 4.4.

Theorem 4.5.

Any minimal attainable sequence for k1k2k_{1}k_{2} is the concatenation of minimal attainable sequences for k1k_{1} and k2k_{2}.

We will prove 4.5 using a modification of original argument appearing in [2] for 4.7 below, which is an immediate corollary.

Corollary 4.6.

For two knotoids k1k_{1} and k2k_{2}, h±(k1k2)=h±(k1)+h±(k2)h_{\pm}(k_{1}k_{2})=h_{\pm}(k_{1})+h_{\pm}(k_{2}).

Corollary 4.7 (Turaev [2]).

For two knotoids k1k_{1} and k2k_{2}, h(k1k2)=h(k1)+h(k2)h(k_{1}k_{2})=h(k_{1})+h(k_{2}).

We will treat knotoids as simple theta-curves, and use the following lemmas:

Lemma 4.8.

Suppose that θ\theta is a simple theta-curve and Δ\Delta is a compact oriented surface (not necessarily connected) embedded in S3S^{3} such that (a) Δ=e+e\partial\Delta=e_{+}\cup e_{-}, (b) e0e_{0} intersects Δ\Delta transversely, and (c) The component of Δ\Delta with boundary is a disk. Then the sequence Seq(θ,Δ)\operatorname{Seq}(\theta,\Delta) of signs of intersections of e0e_{0} with Δ\Delta is attainable for θ\theta.

Proof.

Let DD be the disk component of Δ\Delta.

Case 1: Every closed component of Δ\Delta is a sphere.

If Δ=D\Delta=D, then of course Seq(θ,Δ)=Seq(θ,D)\operatorname{Seq}(\theta,\Delta)=\operatorname{Seq}(\theta,D) is attainable.

If Δ\Delta is not connected, some spherical component HH of Δ\Delta must be “outermost” in the sense that no other sphere separates it from DD. If DD is on the positive side of HH, then we may choose an embedded path from HH to the positive side of DD such that the path does not otherwise intersect Δ\Delta or e0e_{0}. Then we can incorporate HH into DD in an orientation-respecting way by adding an annulus to connect HH to DD and deleting disks in DD and HH around the path’s endpoints. If DD is on the negative side of HH, we do the same but with the negative side of DD. Doing this several times replaces Δ\Delta with a spanning disk and realizes the sign sequence as attainable.

Case 2: General Case.

Each closed component of Δ\Delta separates S3S^{3}, and as in Case 11 we can consider DD to be “outside” of every other component, regardless of their orientations. Call a spherical component of Δ\Delta trivial if it intersects e0e_{0} twice and e0e_{0} is unknotted inside the sphere. Let NN be the set of closed components that are not trivial spheres. If NN is nonempty, consider an innermost element GG of NN. Inside of GG are some number of segments of e0e_{0}. Some of these segments may have trivial spheres attached. Let us delete GG and replace it with several trivial spheres: One sphere is added for each segment of e0e_{0} inside GG, surrounding the segment and all preexisting trivial spheres on that segment. The new spheres may be oriented appropriately so that we have not changed Seq(θ,Δ)\operatorname{Seq}(\theta,\Delta). Repeating this process renders NN empty and reduces us to Case 1. ∎

Lemma 4.9.

For any simple theta-curve θ\theta and knot-type theta-curve κ\kappa^{\bullet}, a sign sequence is attainable for κθ\kappa^{\bullet}\theta if and only if it is attainable for θ\theta.

Proof.

It is immediate that any attainable sequence for θ\theta is attainable for κθ\kappa^{\bullet}\theta.

For the other direction, suppose we have a spanning disk DD for κθ\kappa^{\bullet}\theta. Pick a ball BB such that (a) BB intersects κθ\kappa^{\bullet}\theta only on e0e_{0}, (b) B\partial B intersects DD transversely, and (c) the restriction of e0e_{0} to BB is κ\kappa in the form of a 11-tangle. If e0e_{0} intersects DD inside BB, we may push these intersections to the outside: Choose a subinterval of e0Be_{0}\cap B containing the intersections with DD as well as one of the two endpoints, then delete from BB a regular neighborhood of that interval. Therefore we may choose BB so that e0DBe_{0}\cap D\cap B is empty.

Let b0b_{0} and b1b_{1} be the intersections of e0e_{0} with B\partial B, assigned such that e0e_{0} is oriented from b0b_{0} to b1b_{1}. Each component of DBD\cap B is a genus-0 surface (with boundary) properly embedded in BB, and each component of DBD\cap\partial B is an oriented circle with winding number 0, 11, or 1-1 around B{b0,b1}\partial B-\{b_{0},b_{1}\}. Suppose there is at least one circle with winding number 0. Then there is an innermost such circle. We may cut DD along this circle and fill in two disks on either side of B\partial B to obtain a new spanning surface consisting of a disk and a sphere. Doing these repeatedly, we obtain a surface Δ\Delta as in 4.8 (specifically, as in Case 1) such that Seq(κθ,Δ)=Seq(κθ,D)\operatorname{Seq}(\kappa^{\bullet}\theta,\Delta)=\operatorname{Seq}(\kappa^{\bullet}\theta,D) and and such that every component of ΔB\Delta\cap\partial B has winding number ±1\pm 1. There may now be some spherical components of Δ\Delta contained entirely within BB, but they do not intersect e0e_{0}.

Now, let us label the components of ΔB\Delta\cap B as E1,,EnE_{1},\ldots,E_{n}. Each separates BB into two regions, and each is disjoint from the others and from e0e_{0}. The components of ΔB\Delta\cap\partial B, all concentric circles, have an order based on how they are arranged from b0b_{0} to b1b_{1} and so can be indexed 1,,m1,\ldots,m. For each ii, let s(i)s(i) be the winding number of the iith circle and let c(i)c(i) be the index of its component in ΔB\Delta\cap B. For each j{1,,n}j\in\{1,\ldots,n\}, the sum of the s(i)s(i) over all ii with c(i)=jc(i)=j must be 0, because e0e_{0} does not intersect EjE_{j}. For similar reasons, for all jj and all w1w_{1} and w2w_{2} with w1<w2w_{1}<w_{2} and c(w1)=c(w2)jc(w_{1})=c(w_{2})\neq j, the sum of the s(i)s(i) over all ii with c(i)=jc(i)=j and w1<i<w2w_{1}<i<w_{2} is also 0.

Now, to show that Seq(κθ,Δ)\operatorname{Seq}(\kappa^{\bullet}\theta,\Delta) is attainable for θ\theta, we will create a spanning of θ\theta by deleting and replacing the interior of BB. Let BB^{\prime} be a standard 33-ball, and choose an orientation-respecting identification BB\partial B^{\prime}\cong\partial B. Let e0e_{0}^{\prime} be an unknotted strand properly embedded in BB^{\prime} from b0b_{0} to b1b_{1}. Now consider a partition of the components of ΔB\Delta\cap\partial B into pairs such that (a) paired components have opposite winding numbers, (b) paired components come from the same component of ΔB\Delta\cap B, and (c) for i1<i2<i3<i4i_{1}<i_{2}<i_{3}<i_{4}, we do not have i1i_{1} paired to i3i_{3} and i2i_{2} to i4i_{4}. (A simple induction argument shows this is possible.) Now, we connect each pair of components with an unknotted annulus disjoint from e0e_{0}^{\prime} and disjoint from the other annuli. See Figure 8.

Refer to caption
Figure 8. A cross section of BB^{\prime}. We can obtain BB^{\prime} by rotating around the center axis, which represents e0e_{0}^{\prime}. In this example, ΔB\Delta\cap B had three components, indicated by the three types of curved arrow.

Now we glue BB^{\prime} along B\partial B to the closure of the complement of BB to obtain a spanning surface Δ\Delta^{\prime} for θ\theta, with Seq(θ,Δ)=Seq(κθ,D)\operatorname{Seq}(\theta,\Delta^{\prime})=\operatorname{Seq}(\kappa^{\bullet}\theta,D). Because we replaced each component of ΔB\Delta\cap B with several annuli, we have not created any higher-genus components by replacing Δ\Delta with Δ\Delta^{\prime}. Therefore Δ\Delta^{\prime} is a union of a disk with spheres, and so by 4.8, Seq(κθ,D)\operatorname{Seq}(\kappa^{\bullet}\theta,D) is attainable for θ\theta. ∎

Proof of 4.5.

Given a spanning disk DD for a product θ1θ2\theta_{1}\theta_{2} realizing its height, we wish for there to be an embedded sphere decomposing (θ1θ2,D)(\theta_{1}\theta_{2},D) as a product of spannings of θ1\theta_{1} and θ2\theta_{2}. Such a sphere does not exist in general, but we will assume that θ1\theta_{1} and θ2\theta_{2} have no knot-type factors, and this will be sufficient by 4.9.

By construction, there is a sphere Σ\Sigma decomposing θ1θ2\theta_{1}\theta_{2} as a product of θ1\theta_{1} and θ2\theta_{2}. Necessarily, the two vertices of θ1θ2\theta_{1}\theta_{2} lie on opposite sides of Σ\Sigma, and each edge intersects Σ\Sigma once transversely. We may assume DD to intersect Σ\Sigma transversely as well. Then the intersection of DD and Σ\Sigma consists of a line segment and possibly several circles. If the number of circles is 0, then Σ\Sigma cuts DD into two disks, which are spanning disks for θ1\theta_{1} and θ2\theta_{2}, so we are done.

If there are some circles, we may pick one which is innermost in DD. This bounds a disk EE in DD which does not otherwise intersect DD or Σ\Sigma. It also separates Σ\Sigma into two disks Σ1\Sigma_{1} and Σ2\Sigma_{2}. Since EE sits on one side of Σ\Sigma and cuts that side into two parts, one part contains a vertex and the other does not. We may assign the labels Σ1\Sigma_{1} and Σ2\Sigma_{2} in such a way that Σ1E\Sigma_{1}\cup E is a sphere that separates the vertices of θ\theta and Σ2E\Sigma_{2}\cup E is a sphere with both vertices on one side. Let BB denote the ball with boundary Σ2E\Sigma_{2}\cup E that doesn’t contain the vertices.

Each of e+e_{+} and ee_{-} must have its one intersection with Σ\Sigma on Σ1\Sigma_{1}, as it cannot intersect EE. The intersection of e0e_{0} with Σ\Sigma may be on either Σ1\Sigma_{1} or Σ2\Sigma_{2}, but regardless, e0e_{0} cannot intersect EE more times than it intersects Σ2\Sigma_{2}, by 3.1. Therefore, e0e_{0} either intersects Σ1\Sigma_{1} once and not Σ2\Sigma_{2} or EE, or it intersects Σ2\Sigma_{2} and EE once each but not Σ1\Sigma_{1}. In the latter case, there is a 11-tangle inside of BB, but by our assumption of no knot-type factors, the tangle is unknotted.

Let Σ\Sigma^{\prime} be the sphere formed by pushing Σ2\Sigma_{2} through BB and past EE, so Σ\Sigma^{\prime} is a slight perturbation of Σ1E\Sigma_{1}\cup E and there are fewer circular intersections of DD with Σ\Sigma^{\prime} than with Σ\Sigma. Since BB either does not intersect e0e_{0} or contains only an unknotted segment between Σ2\Sigma_{2} and EE, (θ,Σ)(\theta,\Sigma^{\prime}) is isotopic to (θ,Σ)(\theta,\Sigma), so Σ\Sigma^{\prime} still decomposes θ\theta as θ1θ2\theta_{1}\theta_{2}.

Repeating the above steps yields a sphere intersecting e0e_{0} once and DD in only an interval, so it decomposes (θ1θ2,D)(\theta_{1}\theta_{2},D) as a product of (θ1,D1)(\theta_{1},D_{1}) and (θ2,D2)(\theta_{2},D_{2}) as desired. Then Seq(θ1θ2,D)\operatorname{Seq}(\theta_{1}\theta_{2},D) is the concatenation of Seq(θ1,D1)\operatorname{Seq}(\theta_{1},D_{1}) with Seq(θ2,D2)\operatorname{Seq}(\theta_{2},D_{2}). Since Seq(θ1θ2,D)\operatorname{Seq}(\theta_{1}\theta_{2},D) is minimal, each Seq(θi,Di)\operatorname{Seq}(\theta_{i},D_{i}) is also minimal. ∎

4.3. Signed Heights under Lifting

For a shortcut diagram (K,a)(K,a) of a knotoid kk, there are nn lifts of aa to a shortcut for K/nK/n. The total number of positive/negative intersections of K/nK/n with all such lifts is equal to h±(K,a)h_{\pm}(K,a). Of course, that amount must be at least nn times the minimal number of positive/negative intersections with each of the nn lifts of aa.

Proposition 4.10.

For all kk and nn, nh±(k/n)h±(k)nh_{\pm}(k/n)\leq h_{\pm}(k).

Furthermore, we can obtain attainable sign sequences for k/nk/n from attainable sequences for kk in the following way. Given a sign sequence AA of length rr, for each i{0,,r}i\in\{0,\ldots,r\} let pA(i)p_{A}(i) be the sum of the terms of AA from indices 11 to ii. For i{1,,r}i\in\{1,\ldots,r\}, let qA(i)q_{A}(i) be the maximum of pA(i1)p_{A}(i-1) and pA(i)p_{A}(i). Then for x/nx\in\mathbb{Z}/n\mathbb{Z}, let AxA^{x} be the subsequence of AA consisting of only the terms from indices ii with qA(i)x(modn)q_{A}(i)\equiv x\pmod{n}. Given (K,a)(K,a), we may label the nn lifts of aa as a1,,ana^{1},\ldots,a^{n} in such a way that they increment counterclockwise around the lift of v0v_{0}, and the initial direction of K/nK/n is between ana^{n} and a1a^{1}. Then the iith intersection of KK with aa lifts to an intersection of K/nK/n with aq(i)a^{q(i)}, so for each xx, Seq(K/n,ax)=Seq(K,a)x\operatorname{Seq}(K/n,a^{x})=\operatorname{Seq}(K,a)^{x}. This implies the following.

Proposition 4.11.

For every attainable sequence AA for kk, each AxA^{x} is attainable for k/nk/n.

5. Bounds on Signed Height

5.1. Writhes

Given a crossing cc in a knotoid diagram KK, there is a unique resolution of cc that respects orientation. This resolution creates an oriented diagram with two components, a loop LL and an interval KK^{\prime} with the same endpoints as KK. The winding number of LL around the twice-punctured sphere is called the intersection index of cc, denoted Ind(c)\operatorname{Ind}(c). The index is equal to the intersection number of LL with any shortcut, or with KK^{\prime}. Note that the index of a crossing doesn’t depend on any over/under information. If a crossing has index nn, it will be called an nn-crossing.

Definition 5.1.

For nonzero nn, the nn-writhe Jn(K)J_{n}(K) of KK is half the sum of the signs of the nn-crossings.

Theorem 5.2 (Kim–Im–Lee [5]).

For nonzero nn, the nn-writhe is a knotoid invariant.

Remark.

Our convention differs from [5] by a factor of 22; they omit the word “half” in 5.1. Under our convention, the nn-writhe is still an integer: Any knotoid diagram can be turned into a diagram for the trivial knotoid by switching the signs of crossings such that each “late” strand passes over each “early” strand. Each such switch changes the nn-writhe by an integer, and the nn-writhe of the trivial knotoid is 0, so all nn-writhes of all knotoids are integers. However, what we say here does not apply in general to virtual knotoids, which are considered in [5] alongside classical knotoids.

The following are immediate consequences of the definition of nn-writhe:

Proposition 5.3.

For a knotoid kk, we have the following:

  1. (1)

    Jn(rev(k))=Jn(k)J_{n}(\operatorname{rev}(k))=J_{n}(k)

  2. (2)

    Jn(mir(k))=Jn(k)J_{n}(\operatorname{mir}(k))=-J_{n}(k)

  3. (3)

    Jn(sym(k))=Jn(k)J_{n}(\operatorname{sym}(k))=-J_{-n}(k)

  4. (4)

    Jn(rot(k))=Jn(k)J_{n}(\operatorname{rot}(k))=J_{-n}(k)

Proposition 5.4.

For knotoids k1k_{1} and k2k_{2}, Jn(k1k2)=Jn(k1)+Jn(k2)J_{n}(k_{1}k_{2})=J_{n}(k_{1})+J_{n}(k_{2}).

The nn-writhes of a knotoid can be encoded in the coefficients of a polynomial. The index polynomial for kk is

Fk(t)=n0Jn(k)(tn1)[t,t1].F_{k}(t)=\sum_{n\neq 0}J_{n}(k)(t^{n}-1)\in\mathbb{Z}[t,t^{-1}].

This is closely related to its similarly-named predecessor, the affine index polynomial of [3], defined by

Pk(t)=csign(c)(tw(c)1),P_{k}(t)=\sum_{c}\operatorname{sign}(c)(t^{w(c)}-1),

where w(c)w(c) is sign(c)ssgn(c)Ind(c)\operatorname{sign}(c)\operatorname{ssgn}(c)\operatorname{Ind}(c), and ssgn(c)\operatorname{ssgn}(c) is as shown in Figure 9. Note that w(c)w(c) differs from Ind(c)\operatorname{Ind}(c) only by sign. The affine index polynomial satisfies Pk(t)=Pk(t1)P_{k}(t)=P_{k}(t^{-1}) for all kk ([3]), so it is related to the index polynomial by the formula

(3) Pk(t)=Fk(t)+Fk(t1).P_{k}(t)=F_{k}(t)+F_{k}(t^{-1}).

The t±nt^{\pm n} coefficient Jn(k)+Jn(k)J_{n}(k)+J_{-n}(k) of the affine index polynomial of kk equals the natural linking number of consecutive components in a periodic diagram for k/nk/n (see Figure 5).

Refer to caption
Figure 9. Sequential signs of crossings.

The degree of the affine index polynomial was shown to be a lower bound for the height of a knotoid in [3]. Because of the relationship in eq. 3, this is equivalent to Proposition 3.12 of [5]. 1.5, together with 1.3, is an improvement on this bound in the case that deg+(Fk)\deg^{+}(F_{k}) and deg(Fk)\deg^{-}(F_{k}) are both positive.

We now prove Theorems 1.5, 1.8, and 1.7.

Proof of 1.7.

For nonzero nn, if Jn(k)0J_{n}(k)\neq 0, any shortcut diagram for kk must have an nn-crossing cc. Then the segment of KK starting and ending at cc has, algebraically, nn intersections with the shortcut aa, so the signs in the corresponding segment of Seq(K,a)\operatorname{Seq}(K,a) add up to nn. ∎

Proof of 1.5.

If Jn(k)0J_{n}(k)\neq 0, then as above, in any attainable sign sequence for kk there is a consecutive subsequence with sum nn. Therefore, for positive nn there must be at least nn appearances of ++, and for negative nn there are at least n-n appearances of -. This proves that the positive/negative height of kk is bounded below by the positive/negative degree of FkF_{k}. ∎

Proof of 1.8.

Suppose that h±(k)=deg±(Fk)h_{\pm}(k)=\deg^{\pm}(F_{k}). A minimal attainable sign sequence contains h+(k)h_{+}(k) copies of ++ and h(k)h_{-}(k) copies of -, and by 1.7, the terms of the same sign must all be consecutive. Therefore, any minimal attainable sign sequence is one of (+,,+,,,)(+,\ldots,+,-,\ldots,-) or (,,,+,,+)(-,\ldots,-,+,\ldots,+). Call these two sequences A1A_{1} and A2A_{2}, respectively. To show that only one of these can be attainable, we consider several cases.

Case 1: h+(k)h_{+}(k) or h(k)h_{-}(k) is 0.

If one of the signed heights is zero, then all of the terms are the same sign, and A1=A2A_{1}=A_{2}.

Case 2: h+(k),h(k)>1h_{+}(k),h_{-}(k)>1.

In this case, A1A_{1} and A2A_{2} are not related by a shift move. There are no other minimal attainable sequences, so by 1.4, they cannot both be attainable.

Case 3: h+(k)h_{+}(k) or h(k)h_{-}(k) is 11 and neither is 0.

Supposing that A1A_{1} and A2A_{2} are both attainable, they are the only minimal attainable sequences. By 3.4, there are compatible spanning disks D1D_{1} and D2D_{2} for τ(k)\tau(k) such that Seq(τ(k),Di)=Ai\operatorname{Seq}(\tau(k),D_{i})=A_{i} for i=1,2i=1,2.

Since D1D2D_{1}\cup D_{2} is an embedded sphere in S3S^{3} and e+ee_{+}\cup e_{-} is an embedded circle in D1D2D_{1}\cup D_{2}, we may pick an embedded sphere Σ\Sigma such that Σ\Sigma intersects e+ee_{+}\cup e_{-} at v0v_{0} and v1v_{1} only, and such that D1D_{1} and D2D_{2} each intersect Σ\Sigma in an interval. Then we may slide e0e_{0} down onto Σ\Sigma and obtain a diagram KK for kk with two compatible shortcuts a1a_{1} and a2a_{2} corresponding respectively to D1D_{1} and D2D_{2} (see Figure 10). We have a region EE in S2S^{2} bounded by a1a2a_{1}\cup a_{2} such that there are h(k)h(k) segments of KK in EE, one entering and leaving by a1a_{1}, one entering and leaving by a2a_{2}, and the rest crossing from one side to the other. Without loss of generality, we may assume that, starting from v0v_{0}, KK intersects a2a_{2} before a1a_{1}. Then h(k)h_{-}(k) must be 11, and the intersections come in the order

(a2,+a2,,+a1,a1),(-a_{2},+a_{2},\ldots,+a_{1},-a_{1}),

where the “\ldots” consists of h+(k)1h_{+}(k)-1 consecutive copies of (+a1,+a2)(+a_{1},+a_{2}). Since any crossing cc of KK lies either in EE or the complement of EE, the loop on KK from cc to cc has an even total number of intersections with a1a_{1} and a2a_{2}. Therefore, if that loop includes the negative intersection with a1a_{1}, it also includes at least one positive a1a_{1} intersection, so if we measure the index of cc by intersections of the loop with a1a_{1}, the index is nonnegative.

Refer to caption
Figure 10. Compatible spanning disks correspond to compatible shortcuts.

Since all crossings have nonnegative index, all negative writhes are 0, contradicting the assumption that h(k)=deg(Fk)h_{-}(k)=\deg^{-}(F_{k}). This proves 1.8. ∎

5.2. The Turaev Polynomial

A state of a diagram KK is a function from the set of crossings to {,+}\{-,+\}. For each state ss, the ss-smoothing of KK is given by smoothing each crossing according to Figure 11. The sum of s(c)s(c) over all crossings is denoted n(s)n(s), and the number of embedded circles in the diagram after smoothing by ss is (s)\ell(s). (There is also one embedded interval, which is not counted.) The bracket polynomial of KK is then

K=sAn(s)(A2A2)(s)[A,A1].\langle K\rangle=\sum_{s}A^{n(s)}(-A^{2}-A^{-2})^{\ell(s)}\in\mathbb{Z}[A,A^{-1}].

The bracket polynomial is invariant under Reidemeister moves I’, II, and III, so it is a framed knotoid invariant. A Reidemeister I move changes the bracket polynomial by a factor of A3-A^{-3}, so the normalized bracket polynomial defined by

K=(A)3wr(K)K\langle K\rangle_{\circ}=(-A)^{-3\operatorname{wr}(K)}\langle K\rangle

in [2] is an unframed invariant.

Refer to caption
Figure 11. Smoothings in the bracket polynomial.

There is also a two-variable version of the bracket polynomial, called the extended bracket polynomial in [2] or the Turaev polynomial as in [8]. For a shortcut diagram (K,a)(K,a), let a(K)a(K) denote the algebraic height

h+(K,a)h(K,a),h_{+}(K,a)-h_{-}(K,a),

and for any state, let a(s)a(s) be the algebraic height of the interval component of the ss-smoothing of KK, with its natural orientation. Then the Turaev polynomial of (K,a)(K,a) is

K,a=sAn(s)ua(s)(A2A2)(s)[A±1,u±1].\langle\kern-5.16663pt~{}\langle K,a\rangle\kern-5.16663pt~{}\rangle=\sum_{s}A^{n(s)}u^{a(s)}(-A^{2}-A^{-2})^{\ell(s)}\in\mathbb{Z}[A^{\pm 1},u^{\pm 1}].

This is an invariant of knotoids with both a framing and shortcut framing. The normalized version

K,a=(A)3wr(K)ua(K)K,a\langle\kern-5.16663pt~{}\langle K,a\rangle\kern-5.16663pt~{}\rangle_{\circ}=(-A)^{-3\operatorname{wr}(K)}u^{-a(K)}\langle\kern-5.16663pt~{}\langle K,a\rangle\kern-5.16663pt~{}\rangle

is a knotoid invariant and always takes values in [A±2,u±2]\mathbb{Z}[A^{\pm 2},u^{\pm 2}].

The height of a knotoid satisfies 2h(k)degu+(k)+degu(k)2h(k)\geq\deg_{u}^{+}(\langle\kern-5.16663pt~{}\langle k\rangle\kern-5.16663pt~{}\rangle_{\circ})+\deg_{u}^{-}(\langle\kern-5.16663pt~{}\langle k\rangle\kern-5.16663pt~{}\rangle_{\circ}) ([2]). 1.6 does not improve this bound on the overall height but is the equivalent statement for the signed heights.

Proof of 1.6.

Fix a shortcut diagram (K,a)(K,a) representing a knotoid kk. For any state ss, the ss-smoothing of KK only has as many intersections with aa as KK does. In particular, the interval component of the smoothing has no more than

h+(K,a)+h(K,a)h_{+}(K,a)+h_{-}(K,a)

positive or negative intersections with aa, so we have

2h+(K,a)a(s)a(K)2h(K,a).-2h_{+}(K,a)\leq a(s)-a(K)\leq 2h_{-}(K,a).

Therefore, the uu exponents of k\langle\kern-5.16663pt~{}\langle k\rangle\kern-5.16663pt~{}\rangle_{\circ} are no more than 2h(k)2h_{-}(k) and no less than 2h+(k)-2h_{+}(k). ∎

A categorification of the Turaev polynomial, the triply-graded winding homology, is defined in [8]. The corresponding Poincaré polynomial is denoted Wk(t,A,u)W_{k}(t,A,u) and satisfies

Wk(1,A,u)=kW_{k}(-1,A,u)=\langle\kern-5.16663pt~{}\langle k\rangle\kern-5.16663pt~{}\rangle_{\circ}

for every kk. The winding homology is the homology of a chain complex in which each generator is given a uu-grading a(s)a(K)a(s)-a(K) for some state ss, so in addition to 1.6 we may also say that

2h±(k)degu(Wk(t,A,u)).2h_{\pm}(k)\geq\deg_{u}^{\mp}(W_{k}(t,A,u)).

6. Knotoids with Low Height

6.1. Knotoids of Height One

3.4 allows us to characterize knotoids of height 11 using tangles. Let BB be the unit ball in 3\mathbb{R}^{3} with labelled points N=(0,1,0)N=(0,1,0), E=(1,0,0)E=(1,0,0), S=(0,1,0)S=(0,-1,0), and W=(1,0,0)W=(-1,0,0). Suppose we are given a 22-tangle RR in BB with a strand connecting NN to EE, and a strand connecting WW to SS. Then we may form a knotoid R¯\overline{R} as in Figure 12. This knotoid has h+(R¯)1h_{+}(\overline{R})\leq 1 and h(R¯)=0h_{-}(\overline{R})=0. Let TT be the set of (isotopy classes of) such tangles RR such that (a) no ball inside BB contains a nontrivial 11-tangle and (b) RR is not the trivial tangle formed by two straight line segments.

Refer to caption
Figure 12. Forming a knotoid diagram from a 22-tangle diagram RR.

A knotoid of height 11 is prime if and only if it has no knot-type factor. Height-11 knotoids each have height pair (1,0)(1,0) or (0,1)(0,1), and the two types are in bijective correspondence via rotation. Let UU be the set of prime knotoids with height pair (1,0)(1,0).

Theorem 6.1.

The map RR¯R\mapsto\overline{R} is a bijection TUT\to U.

Proof.

For any RTR\in T, there is a spanning disk D0D_{0} for τ(R¯)\tau(\overline{R}) such that Seq(τ(R¯),D0)=(+)\operatorname{Seq}(\tau(\overline{R}),D_{0})=(+) and such that RR can be recovered by deleting a regular neighborhood of D0D_{0} and using the appropriate identification S3ν(D0)BS^{3}-\nu(D_{0})\cong B. Suppose D1D_{1} is another spanning disk for τ(R¯)\tau(\overline{R}) such that D0D_{0} and D1D_{1} are compatible and D1D_{1} also has sign sequence (+)(+). Then on one side of the sphere D0D1D_{0}\cup D_{1} is a 11-tangle that, by condition (a) of the definition of TT, is unknotted. Therefore, (τ(R¯),D1)(\tau(\overline{R}),D_{1}) is isotopic to (τ(R¯),D0)(\tau(\overline{R}),D_{0}).

Suppose τ(R¯)\tau(\overline{R}) has height 0 (and therefore is not in UU). Then by 3.4 and the previous paragraph, there is a 0-height spanning disk DD compatible with D0D_{0}. Then D0DD_{0}\cup D splits τ(R¯)\tau(\overline{R}) into two 11-tangles, which must both be trivial, contradicting condition (b) of the definition of TT. Therefore, τ(R¯)\tau(\overline{R}) has height 11. For all RTR\in T, no ball intersecting τ(R¯)\tau(\overline{R}) may contain a nontrivial 11-tangle, and so R¯\overline{R} is in UU.

Knowing that τ(R¯)\tau(\overline{R}) has height 11, 3.4 now implies that no other RR^{\prime} has R¯=R¯\overline{R^{\prime}}=\overline{R}, so the map is injective.

For any knotoid kUk\in U, we may obtain a RTR\in T with R¯=k\overline{R}=k by finding a spanning disk DD for τ(k)\tau(k) with height 11 and deleting a regular neighborhood of DD. Since kk is prime and has height 11, RR satisfies conditions (a) and (b). ∎

6.2. Knotoids of Height Two

Consider the following two examples of knotoids with height 22.

Example 6.2.

The Kinoshita knotoid ω\omega, shown in Figure 13, is notable for being a nontrivial knotoid with trivial overpass and underpass closures. The diagram shown has a shortcut with sign sequence (+,)(+,-), and ω\omega satisfies Fω(t)=t12+tF_{\omega}(t)=t^{-1}-2+t. Therefore, h+(ω)=h(ω)=1h_{+}(\omega)=h_{-}(\omega)=1, and by 1.8, (+,)(+,-) is the only minimal attainable sign sequence for ω\omega.

The Kinoshita knotoid satisfies rev(ω)=rot(ω)\operatorname{rev}(\omega)=\operatorname{rot}(\omega). Note that neither the index polynomial nor the Turaev polynomial distinguishes rot(ω)\operatorname{rot}(\omega) from ω\omega. However, by 4.2, the only minimal attainable sequence for rot(ω)\operatorname{rot}(\omega) is (,+)(-,+), so ω\omega is not rotatable.

Refer to caption
Figure 13. The Kinoshita knotoid.
Example 6.3.

Let kk be the knotoid shown in Figure 14. The periodic diagram shown has shortcuts realizing (+,)(+,-) and (,+)(-,+) as attainable sign sequences. The index polynomial is 1t1-t, showing that h+(k)=1h_{+}(k)=1. A lower bound of 11 for h(k)h_{-}(k) is provided by the Turaev polynomial: the u2u^{2} coefficient is A2+2A6A10-A^{-2}+2A^{-6}-A^{-10}. Therefore, (+,)(+,-) and (,+)(-,+) are both minimal.

The information above gives us an easy way of showing that kk is prime: Since kk_{-} is trivial, kk has no knot-type factor, so to be composite it would have to be a product of two proper knotoids. One would have to have height pair (1,0)(1,0), and the other (0,1)(0,1), but then by 4.5, only one of (+,)(+,-) or (,+)(-,+) would be attainable for kk.

Refer to caption
Figure 14. A knotoid of height pair (1,1)(1,1) with (+,)(+,-) and (,+)(-,+) attainable. This knotoid is reversible.

In general, a knotoid of height 22 falls into one of five categories based on whether its set of minimal attainable sequences is {(+,+)}\{(+,+)\}, {(,)}\{(-,-)\}, {(+,)}\{(+,-)\}, {(,+)}\{(-,+)\}, or {(+,),(,+)}\{(+,-),(-,+)\}. We will further divide the last category into two subcategories.

By 3.4, if kk is a knotoid of height 22 and both (+,)(+,-) and (,+)(-,+) are attainable sequences, then there are compatible spanning disks D1D_{1} and D2D_{2} respectively realizing those two sequences as attainable for τ(k)\tau(k).

Theorem 6.4.

Suppose kk is a knotoid as above. Then exactly one of the following is true.

  1. (a)

    The disks D1D_{1} and D2D_{2} can be chosen in such a way that both of the intersections of e0e_{0} with D1D_{1} come before the intersections with D2D_{2}.

  2. (b)

    The disks D1D_{1} and D2D_{2} can be chosen in such a way that both of the intersections of e0e_{0} with D2D_{2} come before the intersections with D1D_{1}.

To prove 6.4, we will use a particular notion of splitting for 22-tangles, analogous to splitting of links: Suppose BB be a ball with four labelled points NENE, SESE, NWNW, and SWSW on B\partial B, and CC is a fixed choice of isotopy class of circles on B{NE,SE,NW,SW}\partial B-\{NE,SE,NW,SW\} separating NENE and SESE on one side from NWNW and SWSW on the other. A circle in CC will be called a splitting circle. Then a 22-tangle in BB will be called split with respect to CC if there is a properly embedded disk (a splitting disk), disjoint from the strands of tangle, whose boundary is a splitting circle.

Lemma 6.5.

Suppose TT is a 22-tangle formed from two other tangles RR and SS in the way shown in Figure 15. Take CRC_{R} and CTC_{T} to be the classes of (BRBS)\partial(B_{R}\cap B_{S}) on BRB_{R} and BTB_{T}, respectively. Then TT is split if and only if RR is split.

Refer to caption
Figure 15. A 22-tangle TT formed as a sum of two other 22-tangles with particular patterns. The ambient ball BTB_{T} for TT is the union BRBSB_{R}\cup B_{S} of the ambient balls for RR and SS. They are attached in such a way that BRBSB_{R}\cap B_{S} is a disk.
Proof.

Of course, if RR is split then TT is split. Conversely, suppose we have a splitting disk DBTD\subset B_{T} for TT. Choose DD in such a way that DD is in general position and D\partial D is disjoint from BRB_{R}. The intersection DBRD\cap B_{R} must separate the strands of RR from each other. If any components of DBRBSD\cap B_{R}\cap B_{S} bound disks in the punctured surface BRBSB_{R}\cap B_{S}, those components can be removed by cutting and capping. Also, no component of DBRBSD\cap B_{R}\cap B_{S} may separate the NENE point of RR from the SESE point of RR in BRBSB_{R}\cap B_{S}, because the circle must be nullhomotopic in BTTB_{T}-T. Then there must be an odd number of remaining components of DBRBSD\cap B_{R}\cap B_{S}, and they must all be separating circles for RR. An innermost such circle in DD would bound a splitting disk for RR. ∎

For a simple theta-curve θ\theta and spanning disk DD with Seq(θ,D)=(+,)\operatorname{Seq}(\theta,D)=(+,-), deleting a regular neighborhood of DD creates a 33-tangle of the pattern shown in Figure 16. This 33-tangle is well-defined up to simultaneous braiding on the left and right, and the strands can be labelled as the first, second, and third strands based on the order they appear on e0e_{0}. Let PDP_{D} be the 22-tangle formed by deleting the first strand, and QDQ_{D} the 22-tangle formed by deleting the third. We will call DD (a)-split if PDP_{D} is nonsplit and QDQ_{D} is split, where splitting is indicated by the jagged line. Conversely, DD will be called (b)-split if PDP_{D} is split and QDQ_{D} is nonsplit. For spanning disks DD with sequence (,+)(-,+), we can form PDP_{D} and QDQ_{D} in a similar way, but we use the opposite convention for (a)- and (b)-splitting: DD is (a)-split if PDP_{D} is split and QDQ_{D} is nonsplit.

Note that (a)- and (b)-splitting only apply to pairs (θ,D)(\theta,D) with h±(θ,D)=1h_{\pm}(\theta,D)=1, and no disk may be both (a)-split and (b)-split.

Refer to caption
Figure 16. A 33-tangle corresponding to a spanning disk with sequence (+,)(+,-). The splitting condition is determined by the jagged line.
Lemma 6.6.

If kk is a knotoid such that condition (a) from 6.4 is true, then every minimal spanning disk for τ(k)\tau(k) is (a)-split. If instead (b) is true, every minimal spanning disk is (b)-split.

Proof.

Suppose that (a) is true of kk. Then kk may be drawn as in Figure 17, and the tangle HH must be nonsplit, as otherwise kk would have height 0. The 33-tangles corresponding to D1D_{1} and D2D_{2} are each formed by adding one copy of HH with one copy of XX in the appropriate order. By 6.5, D1D_{1} and D2D_{2} are both (a)-split. Furthermore, by another application of 6.5, if DD and DD^{\prime} are any two compatible spanning disks such that one is (a)-split, then the other is (a)-split as well. Then 3.4 implies that all minimal spanning disks are (a)-split.

Refer to caption
Figure 17. A knotoid for which (a) is true, decomposed into tangles HH and XX.

The same reasoning shows that if (b) is true of kk, then all minimal spanning disks of kk are (b)-split. ∎

Proof of 6.4.

First we show that (a) or (b) is true. Suppose we have any choice of D1D_{1} and D2D_{2}. By the same reasoning as in the proof of 3.5, a positive intersection of e0e_{0} with D1D_{1} must be followed by a negative D1D_{1} intersection or positive D2D_{2} intersection, and a negative D2D_{2} intersection must be followed by a positive D2D_{2} intersection or negative D1D_{1} intersection. Therefore, the overall sequence of intersections is either (+D1,D1,D2,+D2)(+D_{1},-D_{1},-D_{2},+D_{2}) or (D2,+D2,+D1,D1)(-D_{2},+D_{2},+D_{1},-D_{1}).

That (a) and (b) cannot both be true follows from 6.6. ∎

Example 6.7.

Consider the knotoid kk shown in Figure 18. The index polynomial is 0, but the Turaev polynomial tells us that the positive and negative heights are both 11. The spanning disk corresponding to the marked shortcut is neither (a)-split nor (b)-split, because the corresponding tangles PP and QQ are both split. Therefore, kk is neither type (a) nor type (b), so (+,)(+,-) is its only minimal attainable sequence.

Refer to caption
Figure 18. A knotoid for which (+,)(+,-) is the only minimal attainable sign sequence. It is unlike the Kinoshita knotoid in that its index polynomial is trivial.

We now have a partition of the set of height-22 knotoids into six categories: Type (a), type (b), and four categories for knotoids that each have only one minimal attainable sequence. For any knotoid kk with height 22, the rotation rot(k)\operatorname{rot}(k) is in a different category from kk. Together with 4.3, this implies the following corollary.

Corollary 6.8.

No proper knotoid with height below 44 is rotatable.

The author does not know if any proper rotatable knotoids exist. In [4] it is shown that a knotoid cannot be rotatable if its double branched cover (see Section 2.4) is hyperbolic.

References

  • [1] Shin’ichi Kinoshita “On Elementary Ideals of Polyhedra in the 3-Sphere” In Pacific Journal of Mathematics 42.1, 1972, pp. 89–98
  • [2] Vladimir Turaev “Knotoids” In Osaka Journal of Mathematics 49, 2012, pp. 195–223
  • [3] Neslihan Gügümcü and Louis Kauffman “New Invariants of Knotoids” In European Journal of Combinatorics 65, 2017, pp. 186–229
  • [4] Agnese Barbensi, Dorothy Buck, Heather Harrington and Marc Lackenby “Double Branched Covers of Knotoids”, 2018 URL: https://arxiv.org/abs/1811.09121
  • [5] Sera Kim, Young Ho Im and Sunho Lee “A Family of Polynomial Invariants for Knotoids” In Journal of Knot Theory and Its Ramifications 27.11, 2018
  • [6] Dimos Goundaroulis, Julien Dorier and Andrzej Stasiak “A Systematic Classification of Knotoids on the Plane and on the Sphere”, 2019 URL: https://arxiv.org/abs/1902.07277
  • [7] Philipp Korablev and Vladimir Tarkaev “A Relation Between the Crossing Number and the Height of a Knotoid”, 2020 URL: https://arxiv.org/abs/2009.02718
  • [8] Deniz Kutluay “Winding Homology of Knotoids”, 2020 URL: https://arxiv.org/abs/2002.07871