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

Generating sets for the kauffman skein module of a family of Seifert manifolds

José Román Aranda    Nathaniel Ferguson
Abstract

We study spanning sets for the Kauffman bracket skein module 𝒮(M,(A))\mathcal{S}(M,\mathbb{Q}(A)) of orientable Seifert fibered spaces with orientable base and non-empty boundary. As a consequence, we show that the KBSM of such manifolds is a finitely generated 𝒮(M,(A))\mathcal{S}(\partial M,\mathbb{Q}(A))-module.

\extrafloats

100

Generating sets for the kauffman skein module of a family
of Seifert fibered spaces
José Román Aranda and Nathaniel Ferguson

Skein modules are a useful tool to study 3-manifolds. Roughly speaking, a skein module captures the space of links in a given 3-manifold, modulo certain local (skein) relations between the links. The choice of skein relations must strike a careful balance between providing interesting structure and ensuring that the structure is managable [13]. The most studied skein module is the Kauffman bracket skein module, so named because the skein relations are the same relations used in the construction of the Kauffman bracket polynomial.

Let \mathcal{R} be a ring containing an invertible element AA. The Kauffman bracket skein module of a 3-manifold MM is defined to be the \mathcal{R}-module 𝒮(M,)\mathcal{S}(M,\mathcal{R}) spanned by all framed links in MM, modulo isotopy and the skein relations

(K1): [Uncaptioned image]=A[Uncaptioned image]+A1[Uncaptioned image]\includegraphics[valign={c},scale={.45}]{K+.png}=A\includegraphics[valign={c},scale={.45}]{K_inf.png}+A^{-1}\includegraphics[valign={c},scale={.5}]{K_0.png}           (K2): L[Uncaptioned image]=(A2A2)LL\cup\includegraphics[valign={c},scale={.35}]{U.png}=\left(-A^{2}-A^{-2}\right)L.

Throughout this note, when \mathcal{R} is unspecified, it is assumed that 𝒮(M)=𝒮(M,(A))\mathcal{S}(M)=\mathcal{S}(M,\mathbb{Q}(A)). Since its introduction by Przytycki [12] and Turaev [15], 𝒮(M,)\mathcal{S}(M,\mathcal{R}) has been studied and computed for various 3-manifolds. It is difficult to describe 𝒮(M,)\mathcal{S}(M,\mathcal{R}) for a given 3-manifold, although some results have been found111As summarized in [13] this is not a complete list of 3-manifolds for which 𝒮(M,)\mathcal{S}(M,\mathcal{R}) is known..

  • 𝒮(S3,Z[A±1])=Z[A±1]\mathcal{S}(S^{3},Z[A^{\pm 1}])=Z[A^{\pm 1}].

  • 𝒮(S1×S2,[A±1])\mathcal{S}(S^{1}\times S^{2},\mathbb{Z}[A^{\pm 1}]) is isomorphic to [A±1](i=1[A±1]/(1A2i+4))\mathbb{Z}[A^{\pm 1}]\oplus\left(\bigoplus_{i=1}^{\infty}\mathbb{Z}[A^{\pm 1}]/(1-A^{2i+4})\right) [8].

  • 𝒮(L(p,q),Z[A±1])\mathcal{S}(L(p,q),Z[A^{\pm 1}]) is a free Z[A±1]Z[A^{\pm 1}] module with p/2+1\lfloor p/2\rfloor+1 generators [7, 3].

  • 𝒮(Σ×[0,1],Z[A±1])\mathcal{S}(\Sigma\times[0,1],Z[A^{\pm 1}]) is a free module generated by multicurves in Σ\Sigma [13, 14].

  • 𝒮(Σ×S1,(A))\mathcal{S}(\Sigma\times S^{1},\mathbb{Q}(A)) is a vector space of dimension 22g+1+2g12^{2g+1}+2g-1 if Σ=\partial\Sigma=\emptyset, [4, 2].

In 2019, Gunningham, Jordan and Safronov proved that, for closed 3-manifolds, 𝒮(M,(A))\mathcal{S}(M,\mathbb{C}(A)) is finite dimensional [5]. However, for 3-manifolds with boundary, this problem is still open. In [1], Detcherry asked versions of a finiteness conjecture for the skein module of knot complements and general 3-manifolds (see Section 3 of [1] for a detailed exposition).

Conjecture 1 (Finiteness conjecture for manifolds with boundary [1]).

Let MM be a compact oriented 3-manifold. Then 𝒮(M)\mathcal{S}(M) is a finitely generated 𝒮(M,(A))\mathcal{S}(\partial M,\mathbb{Q}(A))-module.

This paper studies the finiteness conjecture for a large family of Seifert fibered spaces, SFS. Let Σ\Sigma be an orientable surface of genus gg with NN boundary components. Let nn, bb be non-negative integers with N=n+bN=n+b. For each i=1,,ni=1,\dots,n, pick pairs of relatively prime integers (qi,pi)(q_{i},p_{i}) satisfying 0<qi<|pi|0<q_{i}<|p_{i}|. The 3-manifold Σ×S1\Sigma\times S^{1} has torus boundary components with horizontal meridians μiΣ×{pt}\mu_{i}\subset\Sigma\times\{pt\} and vertical longitudes λi={pt}×S1\lambda_{i}=\{pt\}\times S^{1}. Denote by M(g;b,{(qi,pi)}i=1n)M\left(g;b,\{(q_{i},p_{i})\}_{i=1}^{n}\right) the result of Dehn filling the first nn tori of (Σ×S1)\partial\left(\Sigma\times S^{1}\right) with slopes qiμi+piλiq_{i}\mu_{i}+p_{i}\lambda_{i}. Every SFS with orientable base orbifold is of the form M(g;b,{(qi,pi)}i=1n)M\left(g;b,\{(q_{i},p_{i})\}_{i=1}^{n}\right) [6]. The main result of this paper is to establish Conjecture 1 for such SFS.

Theorem 3.10.

Let Σ\Sigma be an orientable surface with non-empty boundary. Then 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) is a finitely generated 𝒮(Σ×S1,(A))\mathcal{S}(\partial\Sigma\times S^{1},\mathbb{Q}(A))-module of rank at most 22g+112^{2g+1}-1.

Theorem 4.1.

Let M=M(g;b,{(qi,pi)}i=1n)M=M\left(g;b,\{(q_{i},p_{i})\}_{i=1}^{n}\right) be an orientable Seifert fibered space with non-empty boundary. Suppose MM has orientable orbifold base. Then, 𝒮(M)\mathcal{S}(M) is a finitely generated 𝒮(M,(A))\mathcal{S}(\partial M,\mathbb{Q}(A))-module of rank at most (22g+11)i=1n(2qi1)(2^{2g+1}-1)\prod_{i=1}^{n}(2q_{i}-1).

The following is a more general formulation of the finiteness conjecture.

Conjecture 2 (Strong finiteness conjecture for manifolds with boundary [1]).

Let MM be a compact oriented 3-manifold. Then there exists a finite collection Σi,,Σk\Sigma_{i},\dots,\Sigma_{k} of essential subsurfaces ΣiM\Sigma_{i}\subset\partial M such that:

  • for each ii, the dimension of H1(Σi,)H_{1}(\Sigma_{i},\mathbb{Q}) is half of H1(M,)H_{1}(\partial M,\mathbb{Q});

  • the skein module 𝒮(M)\mathcal{S}(M) is a sum of finitely many subspaces F1,,FkF_{1},\dots,F_{k}, where FiF_{i} is a finitely generated 𝒮(Σi,(A))\mathcal{S}(\Sigma_{i},\mathbb{Q}(A))-module.

We are able to show this conjecture for a subclass of SFS.

Theorem 4.2.

Seifert fibered spaces of the form M(g;1,{(1,pi)}i=1n)M\left(g;1,\{(1,p_{i})\}_{i=1}^{n}\right) satisfy Conjecture 2. In particular, Conjecture 2 holds for Σg,1×S1\Sigma_{g,1}\times S^{1}.

The techniques in this work are based on the ideas of Detcherry and Wolff in [2]. For simplicity, we set =(A)\mathcal{R}=\mathbb{Q}(A) by default, even though our statements work for any ring \mathcal{R} such that 1A2m1-A^{2m} is invertible for all m>0m>0. It would be interesting to see if the generating sets in this work can be upgraded to verify Conjecture 2 for all Seifert fibered spaces. Although there is no reason to expect the generating sets to be minimal, we wonder if the work of Gilmer and Masmbaum in [4] could yield similar lower bounds.

Outline of the work. The sections in this paper build-up to the proof of Theorems 4.1 and 4.2 in Section 4. Section 1 introduces the arrowed diagrams which describe links in Σ×S1\Sigma\times S^{1}. We show basic relations among arrowed diagrams in Section 1.1. Section 2 proves that 𝒮(Σ0,N×S1)\mathcal{S}(\Sigma_{0,N}\times S^{1}) is generated by boundary parallel diagrams. Section 3 studies the positive genus case 𝒮(Σg,N×S1)\mathcal{S}(\Sigma_{g,N}\times S^{1}); we find a generating set over (A)\mathbb{Q}(A) in Proposition 3.9. In Section 4, we describe global and local relations between links in the skein module of Seifert fibered spaces. We use this to build generating sets in Section 4.2.

Acknowledgments. This work is the result of a course at and funding from Colby College. The authors are grateful to Puttipong Pongtanapaisan for helpful conversations and Scott Taylor for all his valuable advice.

1 Preliminaries

Most of the arguments in this paper will focus on finding relations among links in Σ×S1\Sigma\times S^{1} for some compact orientable surface Σ\Sigma. The main technique is the use of arrow diagrams introduced by Dabkowski and Mroczkowski in [9].
An arrow diagram in Σ\Sigma is a generically immersed 1-manifold in Σ\Sigma with finitely many double points, together with crossing data on the double points, and finitely many arrows in the embedded arcs. Such diagrams describe links in Σ×S1\Sigma\times S^{1} as follows: Write S1=[0,1]/(01)S^{1}=[0,1]/\left(0\sim 1\right). Lift the knot diagram in Σ×{1/2}\Sigma\times\{1/2\} away from the arrows to a union of knotted arcs in Σ×[1/4,3/4]\Sigma\times[1/4,3/4], and interpret the arrows as vertical arcs intersecting Σ×{1}\Sigma\times\{1\} in the positive direction. We can use the surface framing on arrowed diagrams to describe framed links in Σ×S1\Sigma\times S^{1}.

Refer to caption
Figure 1: Example of arrowed diagram.

Arrowed diagrams have been used to study the skein module of Σ0,3×S1\Sigma_{0,3}\times S^{1} [9], prism manifolds [10], the connected sum of two projective spaces [11], and Σg×S1\Sigma_{g}\times S^{1} [2].

Proposition 1.1 ([9]).

Two arrowed diagrams of framed links in Σ×S1\Sigma\times S^{1} correspond to isotopic links if and only if they are related by standard Reidemeister moves R1R^{\prime}_{1}, R2R_{2}, R3R_{3} and the moves

(R4): [Uncaptioned image][Uncaptioned image][Uncaptioned image]\includegraphics[valign={c},scale={.5}]{R4_1.png}\sim\includegraphics[valign={c},scale={.5}]{R4_2.png}\sim\includegraphics[valign={c},scale={.5}]{R4_3.png}\quad\quad\quad\quad (R5): [Uncaptioned image][Uncaptioned image]\includegraphics[valign={c},scale={.5}]{R5_1.png}\sim\includegraphics[valign={c},scale={.5}]{R5_2.png}.

From relation R4R_{4}, we only need to focus on the total number of the arrows between crossings. We will keep track of them by writting a number nn\in\mathbb{Z} next to an arrow. Negative values of nn correspond to |n||n| arrows in the opposite direction.

Throughout this work, a simple arrowed diagram (or arrowed multicurve) will denote an arrowed diagram with no crossings. A simple closed curve in Σ\Sigma will be said to be trivial if it bounds a disk. We will sometimes refer to trivial curves bounding disks disjoint from a given diagram as unknots. Loops parallel to the boundary will not be considered trivial. A simple closed curve will be essential if it does not bound a disk nor is parallel to the boundary in Σ\Sigma.

We can always resolve the crossings of an arrowed diagram via skein relations. Thus, every element in 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) can be written a [A±1]\mathbb{Z}[A^{\pm 1}]-linear combination of arrowed diagrams with no crossings. The following equation will permit us to disregard arrowed unknots, since we can merge them with other loops.

[Uncaptioned image]=[Uncaptioned image]A[Uncaptioned image]+A1[Uncaptioned image]=A1[Uncaptioned image]+A[Uncaptioned image].\includegraphics[valign={c},scale={.4}]{fig_23_1_n.png}=\includegraphics[valign={c},scale={.4}]{fig_23_1_1n-1.png}\quad\implies\quad A\includegraphics[valign={c},scale={.4}]{fig_23_2_n.png}+A^{-1}\includegraphics[valign={c},scale={.4}]{fig_23_2_0n.png}=A^{-1}\includegraphics[valign={c},scale={.4}]{fig_23_2_n-2.png}+A\includegraphics[valign={c},scale={.4}]{fig_23_2_1n-1.png}. (1)

Equation (1) implies Proposition 1.2.

Proposition 1.2 ([2]).

The skein module 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) is spanned by arrowed multicurves containing no trivial component, and by the arrowed multicurves consisting of just one arrowed unknot with some number of boundary parallel arrowed curves.

Definition 1.3 (Dual graph [2]).

Let γΣ\gamma\subset\Sigma be an arrowed multicurve. Let cc be the multicurve consisting of one copy of each isotopy class of separating essential loop in γ\gamma. Let VV be the set of connected components of Σc\Sigma-c. For vVv\in V, denote by Σ(v)Σ\Sigma(v)\subset\Sigma the corresponding connected component of Σc\Sigma-c. Two distinct vertices share an edge (v1,v2)E(v_{1},v_{2})\in E if the subsurfaces Σ(v1)\Sigma(v_{1}) and Σ(v2)\Sigma(v_{2}) have a common boundary component. Define the dual graph of γ\gamma to be the graph Γ(γ)=(V,E)\Gamma(\gamma)=(V,E).

1.1 Relations between skeins

We now study some operations among arrowed multicurves in 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) that change the number of arrows in a controlled way. Although one can observe that all relations happen on a three-holed sphere, we write them separately for didactical purposes.

In practice, a vertical strand will be part of a concentric circle. Lemma 1.4 states that we can ‘pop-out’ the arrows from a loop with the desired sign (of yy and xx) without increasing the number of arrows in the diagrams. Lemma 1.5 states that we can change the sign of the arrows in an unknot at the expense of adding skeins with fewer arrows. Lemma 1.6 allows us to ‘break’ and ‘merge’ the arrows in between two unknots. Lemma 1.7 lets us pass arrows between parallel (or nested), and Lemma 1.8 is an explicit case of Equation (1). The symbol [Uncaptioned image] in Lemmas 1.6 and 1.7 will correspond to any subsurface of surface Σ\Sigma. In practice, [Uncaptioned image] will correspond to a boundary component of Σ\Sigma or an exceptional fiber in Section 4.

Lemma 1.4.

For any m{0}m\in\mathbb{Z}-\{0\},

  1. (i)

    [Uncaptioned image][A±]{[Uncaptioned image]:mx0,y{0,1},y+|x||m|}\includegraphics[valign={c},scale={.55}]{fig_14_m.png}\in\mathbb{Z}[A^{\pm}]\bigg{\{}\quad\includegraphics[valign={c},scale={.55}]{fig_14_yx.png}\quad:mx\geq 0,y\in\{0,1\},y+|x|\leq|m|\bigg{\}}.

  2. (ii)

    [Uncaptioned image][A±]{[Uncaptioned image]:mx0,y{0,1},|y|+|x||m|}\includegraphics[valign={c},scale={.55}]{fig_14_m.png}\in\mathbb{Z}[A^{\pm}]\bigg{\{}\quad\includegraphics[valign={c},scale={.55}]{fig_14_yx.png}\quad:mx\geq 0,y\in\{0,-1\},|y|+|x|\leq|m|\bigg{\}}.

Proof.

Add one arrow pointing upwards at the top end of the arcs in Equation (1) and set n=m1n=m-1. We obtain the following equation

A[Uncaptioned image]+A1[Uncaptioned image]=A1[Uncaptioned image]+A[Uncaptioned image].A\includegraphics[valign={c},scale={.5}]{fig_14_m.png}+A^{-1}\includegraphics[valign={c},scale={.5}]{fig_14_1m-1.png}=A^{-1}\includegraphics[valign={c},scale={.5}]{fig_14_m-2.png}+A\includegraphics[valign={c},scale={.5}]{fig_14_0m-2.png}. (2)

If m>0m>0, we can solve for [Uncaptioned image] and use it inductively to show Part (i). If m<0m<0, we can instead solve for [Uncaptioned image] and set m=m2m^{\prime}=m-2. This new equation can be use to prove Part (i) for m<0m^{\prime}<0.
Part (ii) is similar. Start with Equation (1) with m=nm=n and solve for [Uncaptioned image] to prove Part (ii) for m>0m>0. If m<0m<0, set m=n2m=n-2 in Equation (1). ∎

Lemma 1.5 (Proposition 4.2 of [2]).

Let SkS_{k} be an unknot in Σ\Sigma with kk\in\mathbb{Z} arrows oriented counterclockwise. The following holds for n1n\geq 1,

  1. (i)

    S1=A6S1S_{1}=A^{6}S_{-1}

  2. (ii)

    Sn=A(2n+4)SnS_{-n}=A^{-(2n+4)}S_{n} modulo (A){S0,,Sn1}\mathbb{Q}(A)\{S_{0},\dots,S_{n-1}\}.

  3. (iii)

    Sn=A2n+4SnS_{n}=A^{2n+4}S_{-n} modulo (A){S(n1),,S0}\mathbb{Q}(A)\{S_{-(n-1)},\dots,S_{0}\}.

Lemma 1.6.

Let a,ba,b\in\mathbb{Z} with ab>0ab>0. Then

  1. (i)

    [Uncaptioned image]+A2a|a|[Uncaptioned image][A±1]{[Uncaptioned image]:0ax,0|x|<|a|+|b|}\includegraphics[valign={c},scale={.5}]{fig_16_ab.png}+A^{\frac{2a}{|a|}}\includegraphics[valign={c},scale={.5}]{fig_16_a+b.png}\in\mathbb{Z}[A^{\pm 1}]\big{\{}\includegraphics[valign={c},scale={.5}]{fig_16_x.png}:0\leq ax,0\leq|x|<|a|+|b|\big{\}}.

  2. (ii)

    [Uncaptioned image][A±1]{[Uncaptioned image]:0|x||a|}\includegraphics[valign={c},scale={.5}]{fig_16_a0.png}\in\mathbb{Z}[A^{\pm 1}]\big{\{}\includegraphics[valign={c},scale={.5}]{fig_16_x.png}:0\leq|x|\leq|a|\big{\}}.

Proof.

Suppose first that a,b>0a,b>0. Using R4 we obtain [Uncaptioned image]=[Uncaptioned image]\includegraphics[valign={c},scale={.5}]{merging_1_ax.png}=\includegraphics[valign={c},scale={.5}]{merging_1_a-1x.png}. Thus,

[Uncaptioned image]+A2[Uncaptioned image]=A2[Uncaptioned image]+[Uncaptioned image].\includegraphics[valign={c},scale={.5}]{merging_ax+1.png}+A^{2}\includegraphics[valign={c},scale={.5}]{merging_a+x+1.png}=A^{2}\includegraphics[valign={c},scale={.5}]{merging_a-1x.png}+\includegraphics[valign={c},scale={.5}]{merging_a+x-1.png}. (3)

By setting x=0x=0, the statement follows for b=1b=1 and all a1a\geq 1. For general b1b\geq 1, we proceed by induction on a1a\geq 1 setting x=bx=b in the equation above. The proof of case ab<0ab<0 uses the equation above after the change of variable a=a+1a=-a^{\prime}+1 and x=xx=-x^{\prime}. Part (ii) follows from Equation (1) with n=an=a. ∎

Lemma 1.7.

For all a,ba,b\in\mathbb{Z},

(i) [Uncaptioned image]=A2[Uncaptioned image]+[Uncaptioned image]A2[Uncaptioned image].\displaystyle\text{(i) }\includegraphics[valign={c},scale={.35}]{pushing_p_a_b.png}=A^{2}\includegraphics[valign={c},scale={.35}]{pushing_p_a-1_b+1.png}+\includegraphics[valign={c},scale={.45}]{pushing_p_b-a+2.png}-A^{2}\includegraphics[valign={c},scale={.45}]{pushing_p_b-a.png}.
(ii) [Uncaptioned image]=A2[Uncaptioned image]+[Uncaptioned image]A2[Uncaptioned image].\displaystyle\text{(ii) }\includegraphics[valign={c},scale={.35}]{pushing_p_a_b.png}=A^{-2}\includegraphics[valign={c},scale={.35}]{pushing_p_a+1_b-1.png}+\includegraphics[valign={c},scale={.45}]{pushing_p_b-a-2.png}-A^{-2}\includegraphics[valign={c},scale={.45}]{pushing_p_b-a.png}.
(iii) [Uncaptioned image]=A2[Uncaptioned image]+[Uncaptioned image]A2[Uncaptioned image].\displaystyle\text{(iii) }\includegraphics[valign={c},scale={.35}]{pushing_n_a_b.png}=A^{2}\includegraphics[valign={c},scale={.35}]{pushing_n_a-1_b+1.png}+\includegraphics[valign={c},scale={.35}]{pushing_n_b-a+2.png}-A^{2}\includegraphics[valign={c},scale={.35}]{pushing_n_b-a.png}.
(iv) [Uncaptioned image]=A2[Uncaptioned image]+[Uncaptioned image]A2[Uncaptioned image].\displaystyle\text{(iv) }\includegraphics[valign={c},scale={.35}]{pushing_n_a_b.png}=A^{-2}\includegraphics[valign={c},scale={.35}]{pushing_n_a+1_b-1.png}+\includegraphics[valign={c},scale={.35}]{pushing_n_b-a-2.png}-A^{-2}\includegraphics[valign={c},scale={.35}]{pushing_n_b-a.png}.
Proof.

One can use (K1) on the LHS of each equation to create a new croossing. The result follows from (R4). ∎

Lemma 1.8.

   

[Uncaptioned image]=A4[Uncaptioned image]A2[Uncaptioned image].\includegraphics[valign={c},scale={.55}]{fig_18_a.png}=-A^{4}\includegraphics[valign={c},scale={.55}]{fig_18_b.png}-A^{2}\includegraphics[valign={c},scale={.55}]{fig_18_c.png}.
Proof.

Rotate Equation (1) by 180 degrees and set n=1n=1. ∎

2 Planar case

Fix a planar subsurface ΣΣ\Sigma^{\prime}\subset\Sigma with at least 4 boundary components. The goal of this section is to prove Proposition 2.8 which states that 𝒮(Σ×S1)\mathcal{S}(\Sigma^{\prime}\times S^{1}) is generated by arrowed diagrams with \partial-parallel arrowed curves only. In particular, the dimension of 𝒮(Σ×S1)\mathcal{S}(\Sigma^{\prime}\times S^{1}) as a module over its boundary is one; generated by the empty link.

We will study diagrams in linear pants decompositions. These are pants decompositions for Σ\Sigma^{\prime} with dual graph isomorphic to a line. See Figure 2 for a concrete picture. Linear decompositions have N=|χ(Σ)|2N=|\chi(\Sigma^{\prime})|\geq 2 pairs of pants. By fixing a linear pants decomposition, there is a well-defined notion of left and right ends of Σ\Sigma^{\prime}. We denote the specific curves of a linear pants decomposition as in Figure 2. We think of such decomposition as the planar analogues for the sausague decompositions of positive genus surfaces in [2].

Refer to caption
Figure 2: Linear pants decomposition for spheres with holes.

The main idea of Proposition 2.8 is to ‘push’ loops parallel to lil_{i} in a linear pants decomposition towards the boundary of Σ\partial\Sigma in both directions. We do this with the help of the Δ\Delta-maps from Definition 2.4; Δ+\Delta_{+} ‘pushes’ loops towards the left and Δ\Delta_{-} towards the right (see Lemma 2.5). This idea is based on Section 3.3 of [2] where the authors concluded a version of Proposition 2.8 for closed surfaces. The following definition helps us to keep track of the arrowed curves in the boundary.

Definition 2.1 (Diagrams in linear pants decompositions).

Fix a linear pants decomposition of Σ\Sigma^{\prime} and integers m0m\geq 0, k0{1,,N1}k_{0}\in\{1,\dots,N-1\}. For each k{0,,N}{k0}k\in\{0,\dots,N\}-\{k_{0}\}, aa\in\mathbb{Z}, and v{0,1,}Nv\in\{0,1,\emptyset\}^{N}, we define the arrowed multicurves Da,vkD^{k}_{a,v} in Σ\Sigma^{\prime} as follows: Da,vkD^{k}_{a,v} has one copy of lkl_{k} with aa arrows, mm copies of lk0l_{k_{0}}with no arrows, and one copy of cic_{i} with viv_{i} arrows if vi=0,1v_{i}=0,1 and no curve cic_{i} if vi=v_{i}=\emptyset. Notice that the positive direction of the arrows of the curves cic_{i} depends on the (left/right) position of cic_{i} with respect to lk0l_{k_{0}}. If k=k0k=k_{0}, we define Da,vk0l{}_{l}D^{k_{0}}_{a,v} (resp. Da,vk0r{}_{r}D^{k_{0}}_{a,v}) as before with the condition that the left-most (resp. right-most) copy of lk0l_{k_{0}} contains aa arrows.

Refer to caption
Figure 3: Definition of Da,vkD^{k}_{a,v}, Da,vk0l{}_{l}D^{k_{0}}_{a,v} and Da,vk0r{}_{r}D^{k_{0}}_{a,v}.
Lemma 2.2 (Lemma 3.11 of [2]).

The following holds for any two parallel curves,

A1[Uncaptioned image]+A[Uncaptioned image]=A[Uncaptioned image]+A1[Uncaptioned image].A^{-1}\includegraphics[valign={c},scale={.4}]{fig_311_a+1_b.png}+A\includegraphics[valign={c},scale={.4}]{fig_311_a-b+1.png}=A\includegraphics[valign={c},scale={.4}]{fig_311_a_b+1.png}+A^{-1}\includegraphics[valign={c},scale={.4}]{fig_311_a-b-1.png}.

In particular, for any aa\in\mathbb{Z}, m0m\geq 0, and v{0,1,}Nv\in\{0,1,\emptyset\}^{N}, we have

Da,vk0lA2maDa,vk0r,{}_{l}D^{k_{0}}_{a,v}\cong A^{2ma}{}_{r}D^{k_{0}}_{a,v},

modulo [A±1]\mathbb{Z}[A^{\pm 1}]-linear combinations of diagrams with fewer non-trivial loops.

Lemma 2.3 allows us to change the location of the aa arrows in the diagram DD at the expense of changing the vector vv. Its proof follows from Proposition 3.5 of [2].

Lemma 2.3.

The following equations hold for k{1,,N1}k\in\{1,\dots,N-1\}.

  1. (i)

    If k>k0k>k_{0}, then

    ADa,(,vk,,)kA1Da+2,(,vk,,)k=ADa+1,(,vk,1,,)k+1ADa,(,vk,0,,)k+1.AD^{k}_{a,(\dots,v_{k},\emptyset,\dots)}-A^{-1}D^{k}_{a+2,(\dots,v_{k},\emptyset,\dots)}=AD^{k+1}_{a+1,(\dots,v_{k},1,\emptyset,\dots)}-AD^{k+1}_{a,(\dots,v_{k},0,\emptyset,\dots)}.
  2. (ii)

    If k=k0k=k_{0}, then

    ADa,(,vk,,)k0rA1Da+2,(,vk,,)k0r=ADa+1,(,vk,1,,)k0+1A1Da,(,vk,0,,)k0+1.A{}_{r}D^{k_{0}}_{a,(\dots,v_{k},\emptyset,\dots)}-A^{-1}{}_{r}D^{k_{0}}_{a+2,(\dots,v_{k},\emptyset,\dots)}=AD^{k_{0}+1}_{a+1,(\dots,v_{k},1,\emptyset,\dots)}-A^{-1}D^{k_{0}+1}_{a,(\dots,v_{k},0,\emptyset,\dots)}.
  3. (iii)

    If k=k0k=k_{0}, then

    ADa+2,(,,vk0,)k0lA1Da,(,,vk0,)k0l=ADa,(,,0,vk0,)k01A1Da+1,(,,1,vk0,)k01.A{}_{l}D^{k_{0}}_{a+2,(\dots,\emptyset,v_{k_{0}},\dots)}-A^{-1}{}_{l}D^{k_{0}}_{a,(\dots,\emptyset,v_{k_{0}},\dots)}=AD^{k_{0}-1}_{a,(\dots,\emptyset,0,v_{k_{0}},\dots)}-A^{-1}D^{k_{0}-1}_{a+1,(\dots,\emptyset,1,v_{k_{0}},\dots)}.
  4. (iv)

    If k<k0k<k_{0}, then

    ADa+2,(,,vk,)kA1Da,(,,vk,)k=ADa,(,,0,vk,)k1A1Da+1,(,,1,vk,)k1.AD^{k}_{a+2,(\dots,\emptyset,v_{k},\dots)}-A^{-1}D^{k}_{a,(\dots,\emptyset,v_{k},\dots)}=AD^{k-1}_{a,(\dots,\emptyset,0,v_{k},\dots)}-A^{-1}D^{k-1}_{a+1,(\dots,\emptyset,1,v_{k},\dots)}.
Definition 2.4 (Δ\Delta-maps).

Following [2], let VΣV^{\partial\Sigma^{\prime}} be the subspace of 𝒮(Σ×S1)\mathcal{S}(\Sigma^{\prime}\times S^{1}) generated by arrowed diagrams with trivial loops and boundary parallel curves in Σ\Sigma^{\prime}. Consider VV to be the formal vector space over (A)\mathbb{Q}(A) spanned by the diagrams Da,vkD^{k}_{a,v}, Da,vk0l{}_{l}D^{k_{0}}_{a,v} and Da,vk0r{}_{r}D^{k_{0}}_{a,v} for all aa\in\mathbb{Z}, v{0,1,}Nv\in\{0,1,\emptyset\}^{N} and k{0,,N}{k0}k\in\{0,\dots,N\}\setminus\{k_{0}\}. Define the linear map s:VVs:V\rightarrow V given by s(Da,vk)=Da+2,vks(D^{k}_{a,v})=D^{k}_{a+2,v} (similarly for Da,vk0l{}_{l}D^{k_{0}}_{a,v} and Da,vk0r{}_{r}D^{k_{0}}_{a,v}). Define the maps Δ,Δ+\Delta_{-},\Delta_{+}, and Δ+,m\Delta_{+,m} by

Δ=AA1s,Δ+=AsA1,Δ+,m=A4m+1sA1.\Delta_{-}=A-A^{-1}s,\quad\Delta_{+}=As-A^{-1},\quad\Delta_{+,m}=A^{4m+1}s-A^{-1}.

Combinations of Δ\Delta-maps, together with Lemmas 2.5 and 2.6, will show that VVΣV\subset V^{\partial\Sigma^{\prime}}.

Lemma 2.5.

Let o(e)o(e) and z(e)z(e) be the number of ones and zeros of a vector e{0,1}ne\in\{0,1\}^{n}.

  1. (i)

    The following equation holds for all 1nk01\leq n\leq k_{0}.

    Δ+n(Da,(,,)k0l)=e{0,1}n(1)o(e)Az(e)o(e)Da+o(e),(,,e,,)k0n,\Delta^{n}_{+}\left({}_{l}D^{k_{0}}_{a,(\dots,\emptyset,\dots)}\right)=\sum_{e\in\{0,1\}^{n}}(-1)^{o(e)}A^{z(e)-o(e)}D^{k_{0}-n}_{a+o(e),(\dots,\emptyset,e,\emptyset,\dots)}, (4)

    where e=(e1,,en)e=(e_{1},\dots,e_{n}) is located so that vk0=env_{k_{0}}=e_{n}.

  2. (ii)

    The following equation holds for all 1nNk01\leq n\leq N-k_{0}.

    Δn(Da,(,,)k0r)=e{0,1}n(1)z(e)Ao(e)z(e)Da+o(e),(,,e,,)k0+n,\Delta^{n}_{-}\left({}_{r}D^{k_{0}}_{a,(\dots,\emptyset,\dots)}\right)=\sum_{e\in\{0,1\}^{n}}(-1)^{z(e)}A^{o(e)-z(e)}D^{k_{0}+n}_{a+o(e),(\dots,\emptyset,e,\emptyset,\dots)}, (5)

    where e=(e1,,en)e=(e_{1},\dots,e_{n}) is located so that vk0=e1v_{k_{0}}=e_{1}.

Proof.

We now prove Equation (4). The proof of Equation (5) is symmetric and it is left to the reader. Lemma 2.3 with k=k0k=k_{0} is the statement of case n=1n=1. We proceed by induction on nn and suppose that Equation (4) holds for some 1nk011\leq n\leq k_{0}-1. Using Lemma 2.3 with k<k0k<k_{0}, we show the inductive step as follows,

Δ+n+1(Da,k0l)=\displaystyle\Delta^{n+1}_{+}\left({}_{l}D^{k_{0}}_{a,\emptyset}\right)= (AsA1)Δ+n(Da,k0l)\displaystyle\left(As-A^{-1}\right)\circ\Delta^{n}_{+}\left({}_{l}D^{k_{0}}_{a,\emptyset}\right)
=\displaystyle= e{0,1}n(1)o(e)A1+z(e)o(e)Da+2+o(e),(,,e,,)k0n\displaystyle\sum_{e\in\{0,1\}^{n}}(-1)^{o(e)}A^{1+z(e)-o(e)}D^{k_{0}-n}_{a+2+o(e),(\dots,\emptyset,e,\emptyset,\dots)}
(1)o(e)A1+z(e)o(e)Da+o(e),(,,e,,)k0n\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad-(-1)^{o(e)}A^{-1+z(e)-o(e)}D^{k_{0}-n}_{a+o(e),(\dots,\emptyset,e,\emptyset,\dots)}
=\displaystyle= e{0,1}n(1)o(e)Az(e)o(e)[ADa+o(e)+2,(,,e,,)k0nA1Da+o(e),(,,e,,)k0n]\displaystyle\sum_{e\in\{0,1\}^{n}}(-1)^{o(e)}A^{z(e)-o(e)}\left[AD^{k_{0}-n}_{a+o(e)+2,(\dots,\emptyset,e,\emptyset,\dots)}-A^{-1}D^{k_{0}-n}_{a+o(e),(\dots,\emptyset,e,\emptyset,\dots)}\right]
=\displaystyle= e{0,1}n(1)o(e)Az(e)o(e)[ADa+o(e),(,,0,e,,)k0n1A1Da+o(e)+1,(,,1,e,,)k0n1]\displaystyle\sum_{e\in\{0,1\}^{n}}(-1)^{o(e)}A^{z(e)-o(e)}\left[AD^{k_{0}-n-1}_{a+o(e),(\dots,\emptyset,0,e,\emptyset,\dots)}-A^{-1}D^{k_{0}-n-1}_{a+o(e)+1,(\dots,\emptyset,1,e,\emptyset,\dots)}\right]
=\displaystyle= e{0,1}n(1)o(e)A1+z(e)o(e)Da+o(e),(,,0,e,,)k0n1\displaystyle\sum_{e\in\{0,1\}^{n}}(-1)^{o(e)}A^{1+z(e)-o(e)}D^{k_{0}-n-1}_{a+o(e),(\dots,\emptyset,0,e,\emptyset,\dots)}
+(1)o(e)+1Az(e)o(e)1Da+o(e)+1,(,,1,e,,)k0n1\displaystyle\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad\quad+(-1)^{o(e)+1}A^{z(e)-o(e)-1}D^{k_{0}-n-1}_{a+o(e)+1,(\dots,\emptyset,1,e,\emptyset,\dots)}
=\displaystyle= e{0,1}n+1(1)o(e)Az(e)o(e)Da+o(e),(,,e,,)k0(n+1).\displaystyle\sum_{e\in\{0,1\}^{n+1}}(-1)^{o(e)}A^{z(e)-o(e)}D^{k_{0}-(n+1)}_{a+o(e),(\dots,\emptyset,e,\emptyset,\dots)}.

Lemma 2.6.

For any aa\in\mathbb{Z}, we have Δ+k0(Da,k0l),ΔNk0(Da,k0r)VΣ\Delta^{k_{0}}_{+}\left({}_{l}D^{k_{0}}_{a,\emptyset}\right),\Delta^{N-k_{0}}_{-}\left({}_{r}D^{k_{0}}_{a,\emptyset}\right)\in V^{\partial\Sigma^{\prime}}. Furthermore, Δ+k0(Da,k0l)\Delta^{k_{0}}_{+}\left({}_{l}D^{k_{0}}_{a,\emptyset}\right) is a linear combination of elements of the form Da,(v1,,vk0,,,)0D^{0}_{a^{\prime},(v_{1},\dots,v_{k_{0}},\emptyset,\dots,\emptyset)} and ΔNk0(Da,k0r)\Delta^{N-k_{0}}_{-}\left({}_{r}D^{k_{0}}_{a,\emptyset}\right) is a sum of elements Da,(,,,vk0+1,,vN)ND^{N}_{a^{\prime},(\emptyset,\dots,\emptyset,v_{k_{0}+1},\dots,v_{N})}.

Proof.

Setting n=k0n=k_{0} in Equation (4) yields the condition Δ+k0(Da,k0l)VΣ\Delta^{k_{0}}_{+}\left({}_{l}D^{k_{0}}_{a,\emptyset}\right)\in V^{\partial\Sigma^{\prime}} and the first half of the statement. The second conclusion ΔNk0(Da,k0r)VΣ\Delta^{N-k_{0}}_{-}\left({}_{r}D^{k_{0}}_{a,\emptyset}\right)\in V^{\partial\Sigma^{\prime}} follows by setting n=Nk0n=N-k_{0} in Equation (5). ∎

Proposition 2.7.

Da,vk0l{}_{l}D^{k_{0}}_{a,v} and Da,vk0r{}_{r}D^{k_{0}}_{a,v} lie in VΣV^{\partial\Sigma^{\prime}} for any aa\in\mathbb{Z} and v{0,1,}Nv\in\{0,1,\emptyset\}^{N}.

Proof.

By pushing the boundary parallel curves ‘outside’ Σ\Sigma^{\prime}, it is enough to show the proposition for v=v=\emptyset. Using Lemma 2.2, modulo arrowed multicurves with fewer non-trivial loops, we get that s(Da,k0l)=Da+2,k0lA2m(a+2)Da+2,k0rs\left({}_{l}D^{k_{0}}_{a,\emptyset}\right)={}_{l}D^{k_{0}}_{a+2,\emptyset}\cong A^{2m(a+2)}{}_{r}D^{k_{0}}_{a+2,\emptyset}. Thus,

Δ+(Da,k0l)=\displaystyle\Delta_{+}\left({}_{l}D^{k_{0}}_{a,\emptyset}\right)= As(Da,k0l)A1Da,k0l\displaystyle As\left({}_{l}D^{k_{0}}_{a,\emptyset}\right)-A^{-1}{}_{l}D^{k_{0}}_{a,\emptyset}
\displaystyle\cong A4m+1A2maDa+2,k0rA1A2maDa,k0r\displaystyle A^{4m+1}A^{2ma}{}_{r}D^{k_{0}}_{a+2,\emptyset}-A^{-1}A^{2ma}{}_{r}D^{k_{0}}_{a,\emptyset}
=\displaystyle= A2ma[A4m+1sA1](Da,k0r)\displaystyle A^{2ma}\left[A^{4m+1}s-A^{-1}\right]\left({}_{r}D^{k_{0}}_{a,\emptyset}\right)
=\displaystyle= A2maΔ+,m(Da,k0r).\displaystyle A^{2ma}\Delta_{+,m}\left({}_{r}D^{k_{0}}_{a,\emptyset}\right).

Hence, up to sums of curves with less non-trivial loops in Σ\Sigma^{\prime}, Lemma 2.6 implies

Δ+,mk0(Da,k0r),ΔNk0(Da,k0r)VΣ.\Delta^{k_{0}}_{+,m}\left({}_{r}D^{k_{0}}_{a,\emptyset}\right),\Delta^{N-k_{0}}_{-}\left({}_{r}D^{k_{0}}_{a,\emptyset}\right)\in V^{\partial\Sigma^{\prime}}.

Finally, observe that A1Δ+,m+A4m+1Δ=(A4m+2A2)IdVA^{-1}\Delta_{+,m}+A^{4m+1}\Delta_{-}=\left(A^{4m+2}-A^{-2}\right)Id_{V}. This yields

Da,k0r=IdVN(Da,k0r)=1(A4m+2A2)N(A1Δ+,m+A4m+1Δ)N(Da,k0r).{}_{r}D^{k_{0}}_{a,\emptyset}=Id_{V}^{N}\left({}_{r}D^{k_{0}}_{a,\emptyset}\right)=\frac{1}{\left(A^{4m+2}-A^{-2}\right)^{N}}\left(A^{-1}\Delta_{+,m}+A^{4m+1}\Delta_{-}\right)^{N}\left({}_{r}D^{k_{0}}_{a,\emptyset}\right).

The result follows since Δ+,mk0(Da,k0r)\Delta^{k_{0}}_{+,m}\left({}_{r}D^{k_{0}}_{a,\emptyset}\right) and ΔNk0(Da,k0r)\Delta^{N-k_{0}}_{-}\left({}_{r}D^{k_{0}}_{a,\emptyset}\right) are both elements of VΣV^{\partial\Sigma^{\prime}}. ∎

We are ready to describe an explicit generating set for 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) for any planar surface Σ\Sigma.

Proposition 2.8.

Let Σ\Sigma be a NN-holed sphere with N1N\geq 1. Then 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) is generated by arrowed unknots and \partial-parallel arrowed multicurves.

Proof.

Proposition 2.8 is equivalent to the statement that 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) is generated by arrowed multicurves with dual graphs isomorphic to a point. Let γ\gamma be an arrowed multicurve in Σ\Sigma with Γ(γ){pt}\Gamma(\gamma)\neq\{pt\}. Let e=(v1,v2)e=(v_{1},v_{2}) be a fixed edge of Γ(γ)\Gamma(\gamma), and let ΣΣ\Sigma^{\prime}\subset\Sigma be the subsurface Σ(v1)Σ(v2)\Sigma(v_{1})\cup\Sigma(v_{2}). By Lemma 2.2, up to curves of smaller degree, we can arrange the arrows in the loops corresponding to ee so that only one loop (the closest to Σ(v2)\Sigma(v_{2})) may have arrows. By construction, γΣ\gamma\cap\Sigma^{\prime} has one isotopy class of separating non \partial-parallel curve in Σ\Sigma^{\prime}. Thus, there exists a linear pants decomposition for Σ\Sigma^{\prime} and integers aa\in\mathbb{Z}, k0{1,,|χ(Σ)|1}k_{0}\in\{1,\dots,|\chi(\Sigma^{\prime})|-1\} so that γDa,k0r\gamma\cong{}_{r}D^{k_{0}}_{a,\emptyset} (we focus on the non \partial-parallel components of γΣ\gamma\cap\Sigma^{\prime}). Proposition 2.7 states that Da,k0rVΣ{}_{r}D^{k_{0}}_{a,\emptyset}\in V^{\partial\Sigma^{\prime}}. Therefore, γ\gamma is a (a)\mathbb{Q}(a)-linear combination of arrowed multicurves with dual graphs isomorphic to Γ(γ)/e\Gamma(\gamma)/e; with fewer vertices than Γ(γ)\Gamma(\gamma). ∎

3 Non-planar case

This section further exploits the proofs in [2] to give a finiteness result for 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) for all orientable surfaces with boundary (Proposition 3.9). Throughout this section, Σ\Sigma will be a compact orientable surface of genus g>0g>0 with N>0N>0 boundary components.

3.1 Properties of stable multicurves

Definition 3.1 (Complexity).

Let γ\gamma be an arrowed multicurve. Denote by nn the number of non-separating circles of γ\gamma, mm the number of non-trivial non \partial-parallel separating circles in γ\gamma, and bb the number of vertices in the dual graph of γ\gamma intersecting Σ\partial\Sigma. We define the complexity of a multicurve γ\gamma as (b,n+2m,n+m)\left(b,n+2m,n+m\right) and order them with the lexicographic order. An arrowed multicurve is said to be stable if it is not a linear combination of diagrams with lower complexity.

Proposition 3.2, Proposition 3.3, and Lemma 3.4 restate properties of stable curves from [2]. Fix a stable arrowed multicurve γ\gamma in Σ\Sigma.

Proposition 3.2 (Proposition 3.7 of [2]).

Let Σ=Σ(v)\Sigma^{\prime}=\Sigma(v) be a vertex of Γ\Gamma with |Σ|1|\partial\Sigma^{\prime}|\geq 1 and g(Σ)1g(\Sigma^{\prime})\geq 1. Then γΣ\gamma\cap\Sigma^{\prime} contains at most one non-separating curve.

Proposition 3.3 (From proof of Proposition 3.8 of [2]).

If e=(v,v)e=(v,v^{\prime}) is an edge of Γ\Gamma with g(v)1g(v^{\prime})\geq 1, then the valence of vv is at most two.

Lemma 3.4 (Lemma 3.9 of [2]).

For a vertex vv with g(v)1g(v)\geq 1 and valence two, γΣ(v)\gamma\cap\Sigma(v) contains no non-separating curves.

Now, Proposition 3.5 shows that stable arrowed multicurves satisfy b(γ)=1b(\gamma)=1.

Proposition 3.5.

Stable arrowed multicurves have dual graphs isomorphic to lines. Moreover, they are (A)\mathbb{Q}(A)-linear combinations of arrowed unknots and the two types of arrowed multicurves depicted in Figure 4.

Refer to caption
Figure 4: Type 1 multicurves contain only one isotopy class of non-separating simple curve and type 2 at most two non-separating loops.
Proof.

Suppose b(γ)>1b(\gamma)>1; i.e., there exist two distinct vertices v1,v2Γv_{1},v_{2}\in\Gamma containing boundary components of Σ\Sigma. We will show that γ\gamma is not stable. There exists a path PΣP\subset\Sigma connecting v1v_{1} and v2v_{2}. For each vertex xPx\in P, we define a subsurface Σ(x)Σ(x)\Sigma^{\prime}(x)\subset\Sigma(x) as follows: If Σ(x)\Sigma(x) is planar, define Σ(x):=Σ(x)\Sigma^{\prime}(x):=\Sigma(x). Suppose now that g(x)1g(x)\geq 1 and x{v1,v2}x\notin\{v_{1},v_{2}\}. Proposition 3.2 states that γΣ(x)\gamma\cap\Sigma(x) contains at most one non-separating loop. Thus, we can find a planar surface Σ(x)Σ(x)\Sigma^{\prime}(x)\subset\Sigma(x) disjoint from the non-separating loop such that Σ(x)\partial\Sigma^{\prime}(x) contains the two boundaries of Σ(x)\Sigma(x) participating in the path PP (see Figure 5). Suppose now g(x)1g(x)\geq 1 and x=vix=v_{i}. Using Proposition 3.2 again, we can find a subsurface Σ(x)Σ(x)\Sigma^{\prime}(x)\subset\Sigma(x) with Σ(x)\partial\Sigma^{\prime}(x) containing the Σ(x)Σ\Sigma(x)\cap\partial\Sigma and the one loop of Σ(x)\partial\Sigma(x) participating in the path PP (see Figure 5). Define ΣΣ\Sigma^{\prime}\subset\Sigma to be the connected surface obtained by gluing the subsurfaces Σ(x)\Sigma^{\prime}(x) for all xPx\in P. Since Γ\Gamma is a tree, Σ\Sigma^{\prime} must be planar.

Refer to caption
Figure 5: Building the subsurface Σ\Sigma^{\prime}.

By construction γΣ\gamma\cap\Sigma^{\prime} can be thought of as an element of 𝒮(Σ×S1)\mathcal{S}(\Sigma^{\prime}\times S^{1}). Proposition 2.8 states that γΣ\gamma\cap\Sigma^{\prime} can be written as (a)\mathbb{Q}(a)-linear combination of arrowed diagrams with only trivial and \partial-parallel curves in Σ\Sigma^{\prime}. In particular, γ\gamma can be written as a linear combination of arrowed diagrams γ\gamma^{\prime} in Σ\Sigma with b(γ)<b(γ)b(\gamma^{\prime})<b(\gamma), and so γ\gamma is not stable.

Let γ\gamma be an stable arrowed multicurve. Since b(γ)=1b(\gamma)=1, there is a unique vertex x0Γx_{0}\in\Gamma with ΣΣ(x0)\partial\Sigma\subset\Sigma(x_{0}). Notice that any vertex vΓv\in\Gamma of valence two either has positive genus or is equal to x0x_{0}. This assertion, together with Proposition 3.3, implies that Γ\Gamma is isomorphic to a line where every vertex different than x0x_{0} has positive genus.

The graph Γ{x0}\Gamma\setminus\{x_{0}\} is the disjoint union of at most two linear graphs Γ1\Gamma_{1} and Γ2\Gamma_{2}; Γi\Gamma_{i} might be empty. For each Γi\Gamma_{i}\neq\emptyset, the subsurface Σ(Γi)\Sigma(\Gamma_{i}) is a surface of positive genus with one boundary component. If each Γi\Gamma_{i} has at most one vertex then γ\gamma looks like curves in Figure 4 and the proposition follows. Suppose then that Γi\Gamma_{i} has two or more vertices and pick an edge ee of Γi\Gamma_{i}. By Proposition 3.2 and Lemma 3.4, γΣ(e)\gamma\cap\Sigma(e) contains at most one isotopy class of non-separating curves. Denote such curve by α\alpha; observe that α\alpha is empty unless ee is has an endpoint on a leaf of Γ\Gamma. Let Σ′′\Sigma^{\prime\prime} be the complement of an open neighborhood of α\alpha in Σ(e)\Sigma(e). By construction, γΣ′′\gamma\cap\Sigma^{\prime\prime} contains one isotopy class of non-trivial separating curves in Σ′′\Sigma^{\prime\prime}. By Lemma 3.12 of [2] we can ‘push’ the separating arrowed loops in γΣ′′\gamma\cap\Sigma^{\prime\prime} towards the boundary of Σ′′\Sigma^{\prime\prime}. Thus, we can write γ\gamma as a linear combination of diagrams with dual graph Γ/e\Gamma/e. We can repeat this process until we obtain only summands with each Γi\Gamma_{i} having at most one vertex. ∎

Refer to caption
Figure 6: One needs to apply Proposition 3.12 of [2] twice for diagrams of type 2.
Proposition 3.6.

Let Σ\Sigma be an orientable surface of genus g>0g>0 and N>0N>0 boundary components. Then 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) is generated by arrowed unknots and arrowed multicurves with \partial-parallel components and at most one non-separating simple closed curve.

Proof.

Using Proposition 3.12 of [2] with Σ\Sigma^{\prime} being the shaded surfaces in Figures 4 and 6, we obtain that the 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) is generated by arrowed diagrams as in Figure 6 where l+l=n1l+l^{\prime}=n_{1} and m,n1,n20m,n_{1},n_{2}\geq 0. Observe that, ignoring the mm curves, the ll and ll^{\prime} curves are parallel. Also observe that, by Lemma 2.2, we can still pass arrows among the ll and ll^{\prime} curves modulo linear combinations of diagrams of the same type with lower n1n_{1} but higher mm. Thus, if we only focus on the complexity n1+n2n_{1}+n_{2}, we can follow the proof of Proposition 3.16 in [2] and conclude that 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) is generated by arrowed diagrams with n1+n21n_{1}+n_{2}\leq 1.

The rest of this proof focuses on making m=0m=0. In order to do this, we combine techniques in Section 2 of this paper and Proposition 3.12 of [2].

Case 1: n1+n2=0n_{1}+n_{2}=0. Fix m0m\geq 0. Let cc be a separating curve cutting Σ\Sigma into a sphere with N+1N+1 holes and one connected surface of genus g>0g>0 with one boundary component. The diagrams in this case contain only boundary parallel curves and copies of cc. Define VmΣV^{\partial\Sigma}_{m} to be the formal vector space defined by such pictures with at most mm parallel separating curves. For each aa\in\mathbb{Z}, define the diagram Dar{}_{r}D_{a} (resp. Dal{}_{l}D_{a}) to be given by m+1m+1 copies of cc, mm of which have no arrows and where the closest to the positive genus surface (resp. to the holed sphere) has aa arrows. By Lemma 2.2, in order to conclude this case, we only need to check DarVmΣ{}_{r}D_{a}\in V^{\partial\Sigma}_{m}.

Define Δ+\Delta_{+}, Δ\Delta_{-} and Δ+,m\Delta_{+,m} as in Section 2. First, observe that Lemma 2.6 implies that Δ+N1(Dal)VmΣ\Delta^{N-1}_{+}\left({}_{l}D_{a}\right)\in V^{\partial\Sigma}_{m}. Using the computation in the proof of Proposition 2.7, we conclude that Δ+,mN1(Dar)VmΣ\Delta^{N-1}_{+,m}\left({}_{r}D_{a}\right)\in V^{\partial\Sigma}_{m}. On the other hand, Lemma 3.14 of [2] gives us Δ2g(Dar)VmΣ\Delta^{2g}_{-}\left({}_{r}D_{a}\right)\in V^{\partial\Sigma}_{m}. Hence,

Dar=IdV2g+N1(Dar)=1(A4m+2A2)2g+N1(A1Δ+,m+A4m+1Δ)2g+N1(Dar)VmΣ.{}_{r}D_{a}=Id_{V}^{2g+N-1}\left({}_{r}D_{a}\right)=\frac{1}{\left(A^{4m+2}-A^{-2}\right)^{2g+N-1}}\left(A^{-1}\Delta_{+,m}+A^{4m+1}\Delta_{-}\right)^{2g+N-1}\left({}_{r}D_{a}\right)\in V^{\partial\Sigma}_{m}.

Case 2: n1+n2=1n_{1}+n_{2}=1. Fix m0m\geq 0. The diagrams in this case contain boundary parallel curves, some copies of cc, and exactly one non-separating curve denoted by α\alpha. Define VmΣV^{\partial\Sigma}_{m} to be the formal vector space defined by such pictures with at most mm copies of cc. For aa\in\mathbb{Z}, define Dal{}_{l}D_{a}, Dar{}_{r}D_{a} as in Case 1 with the addition of one copy of α\alpha. In order to conclude this case, it is enough to show DarVmΣ{}_{r}D_{a}\in V^{\partial\Sigma}_{m}.

Suppose that α\alpha has xx\in\mathbb{Z} arrows. For a,ba,b\in\mathbb{Z}, define Ea,br{}_{r}E_{a,b} and Ea,bl{}_{l}E_{a,b} to be mm copies of cc with no arrows and three copies of α\alpha with arrows arranged as in Figure 7. We can define the map ss on the diagrams Ea,br{}_{r}E_{a,b} and Ea,bl{}_{l}E_{a,b} by s(Ea,b)=Ea+1,b+1s({}_{*}E_{a,b})={}_{*}E_{a+1,b+1}. This way, the maps Δ\Delta_{-}, Δ+\Delta_{+}, Δ+,m\Delta_{+,m} are defined on the diagrams DaD_{a} and Ea,bE_{a,b}. Define Δ,1=AA3s\Delta_{-,1}=A-A^{-3}s. Using Lemma 2.2, up to linear combinations of diagrams in VmΣV^{\partial\Sigma}_{m}, we obtain the following:

Δ(Ea,0r)=\displaystyle\Delta_{-}\left({}_{r}E_{a,0}\right)= AEa,0rA1Ea+1,1r\displaystyle A{}_{r}E_{a,0}-A^{-1}{}_{r}E_{a+1,1}
\displaystyle\cong A2x+1Ea,0lA2(x1)1Ea+1,1l\displaystyle A^{2x+1}{}_{l}E_{a,0}-A^{2(x-1)-1}{}_{l}E_{a+1,1}
=\displaystyle= A2x[AEa,0lA3Ea+1,1l]\displaystyle A^{2x}\left[A{}_{l}E_{a,0}-A^{-3}{}_{l}E_{a+1,1}\right]
=\displaystyle= A2xΔ,1(Ea,0l).\displaystyle A^{2x}\Delta_{-,1}\left({}_{l}E_{a,0}\right).
Refer to caption
Figure 7: Ea,b,xl{}_{l}E_{a,b,x} and Ea,b,xr{}_{r}E_{a,b,x}.

Lemmas 3.13 and 3.14 of [2] give us that Δ(Ea,0r)VmΣ\Delta_{-}\left({}_{r}E_{a,0}\right)\in V^{\partial\Sigma}_{m} and Δ2g1(Dar)=Δ+2g1(Ea,0l)\Delta^{2g-1}_{-}({}_{r}D_{a})=\Delta^{2g-1}_{+}({}_{l}E_{a,0}). The first equation implies that Δ,1(Ea,0l)VmΣ\Delta_{-,1}\left({}_{l}E_{a,0}\right)\in V^{\partial\Sigma}_{m}. This implication, together with the second equation and the fact that the Δ\Delta-maps commute, lets us conclude that Δ,1Δ2g1(Dar)VmΣ\Delta_{-,1}\circ\Delta^{2g-1}_{-}({}_{r}D_{a})\in V^{\partial\Sigma}_{m}.

Finally, notice that the argument in Case 1 implies that Δ+,mN1(Dar)VmΣ\Delta^{N-1}_{+,m}\left({}_{r}D_{a}\right)\in V^{\partial\Sigma}_{m}. We also have the following relations between Δ\Delta-maps.

A4m+3Δ,1+A1Δ+,m=(A4m+4A2)IdVA^{4m+3}\Delta_{-,1}+A^{-1}\Delta_{+,m}=\left(A^{4m+4}-A^{-2}\right)Id_{V}
A4m+1Δ+A1Δ+,m=(A4m+2A2)IdVA^{4m+1}\Delta_{-}+A^{-1}\Delta_{+,m}=\left(A^{4m+2}-A^{-2}\right)Id_{V}
IdV=1(A4m+4A2)N(A4m+2A2)2g1(A4m+3Δ,1+A1Δ+,m)N(A4m+1Δ+A1Δ+,m)2g1Id_{V}=\frac{1}{(A^{4m+4}-A^{-2})^{N}(A^{4m+2}-A^{-2})^{2g-1}}\left(A^{4m+3}\Delta_{-,1}+A^{-1}\Delta_{+,m}\right)^{N}\circ\left(A^{4m+1}\Delta_{-}+A^{-1}\Delta_{+,m}\right)^{2g-1}

When expanding the last expression, we see that every summand has a factor of the form Δ,1Δ2g1\Delta_{-,1}\circ\Delta^{2g-1}_{-} or Δ+,mN1\Delta^{N-1}_{+,m}. Hence, by evaluating Dar{}_{r}D_{a}, we obtain DarVmΣ{}_{r}D_{a}\in V^{\partial\Sigma}_{m} as desired. ∎

3.2 A generating set for 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1})

To conclude the proof of finiteness for the Kauffman Bracket Skein Module of trivial S1S^{1}-bundles over surfaces with boundary, this section studies relations among non-separating simple closed curves.

Lemma 3.7.

Any arrowed non-separating simple closed curve in Σ\Sigma can be written as follows in 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1})

(AA1)[Uncaptioned image]=A[Uncaptioned image]A1[Uncaptioned image].\left(A-A^{-1}\right)\includegraphics[valign={c},scale={.4}]{fig_37_n.png}=A\includegraphics[valign={c},scale={.4}]{fig_37_1n-1.png}-A^{-1}\includegraphics[valign={c},scale={.4}]{fig_37_0n.png}.
Proof.

Using the R5 relation, we obtain [Uncaptioned image]=[Uncaptioned image]\includegraphics[valign={c},scale={.35}]{fig_37_a.png}=\includegraphics[valign={c},scale={.35}]{fig_37_b.png}. Thus,

A[Uncaptioned image]A1[Uncaptioned image]=A[Uncaptioned image]A1[Uncaptioned image].A\includegraphics[valign={c},scale={.4}]{fig_37_n.png}-A^{-1}\includegraphics[valign={c},scale={.4}]{fig_37_n-2.png}=A\includegraphics[valign={c},scale={.4}]{fig_37_1n-1.png}-A^{-1}\includegraphics[valign={c},scale={.4}]{fig_37_0n.png}.

Proposition 4.1 of [2] states that non-separating curves with nn and n2n-2 arrows are the same in 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}). Thus, the result follows. ∎

Remark 3.8.

[Application of Lemma 3.7] Let γ\gamma be a non-separating simple closed curve in Σ\Sigma and let cΣc\in\partial\Sigma. Let γ~\widetilde{\gamma} be an arrowed diagram with one copy of γ\gamma and some copies of cc; think of γ\gamma to be ‘on the right side’ of cc. Lemma 3.7 states that, at the expense of adding more copies of cc and arrows, γ\gamma is a linear combination of two diagrams where γ\gamma is on the other side of cc.

Proposition 3.9.

Let Σ\Sigma be an orientable surface of genus g>0g>0 and N>0N>0 boundary components. Let DΣD\subset\Sigma be a (N+1)(N+1)-holed sphere containing Σ\partial\Sigma, and let \mathcal{F} be a collection of 22g12^{2g}-1 non-separating simple closed curves in ΣD\Sigma-D such that each curve in \mathcal{F} represents a unique non-zero element of H1(ΣD;/2)H_{1}(\Sigma-D;\mathbb{Z}/2\mathbb{Z}). Let \mathcal{B} be the collection {γα,Uα}\{\gamma\cup\alpha,U\cup\alpha\}, where γ\gamma is a curve in \mathcal{F} zero or one arrow, UU is an arrowed unknot, and α\alpha is any collection of boundary parallel arrowed circles. Then \mathcal{B} is a generating set for 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) over (A)\mathbb{Q}(A).

Proof.

By Proposition 3.6, we only need to focus on the non-separating curves. Let γ~\widetilde{\gamma} be a non-separating simple closed curve in Σ\Sigma. After using Lemma 3.7 repeatedly, we can write γ~\widetilde{\gamma} as a linear combination of arrowed diagrams of the form γα\gamma\cup\alpha where γ\gamma is a non-separating curve in ΣD\Sigma-D and α\alpha is a collection of boundary parallel curves. Observe that the work on Section 5 of [2] holds for surfaces with connected boundary since generators for π1(Sg,)\pi_{1}(S_{g},*) and Mod(Sg)Mod(S_{g}) also work for Sg,1S_{g,1}. Thus, by Proposition 5.5 of [2], two non-separating curves γ1,γ2ΣD\gamma_{1},\gamma_{2}\subset\Sigma-D with the same number of arrows are equal in 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) if [γ1]=[γ2][\gamma_{1}]=[\gamma_{2}] in H1(ΣD;/2)H_{1}(\Sigma-D;\mathbb{Z}/2\mathbb{Z}). The conditions on the number of arrows for non-separating curves follows from Propositions 4.1 of [2]. ∎

Theorem 3.10.

Let Σ\Sigma be an orientable surface with non-empty boundary. Then 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) is a finitely generated 𝒮(Σ×S1,(A))\mathcal{S}(\partial\Sigma\times S^{1},\mathbb{Q}(A))-module of rank at most 22g+112^{2g+1}-1.

Proof.

As a module over 𝒮(M,(A))\mathcal{S}(\partial M,\mathbb{Q}(A)), we can overlook \partial-parallel subdiagrams. Proposition 3.9 implies that 𝒮(M)\mathcal{S}(M) is generated by the empty diagram and diagrams in \mathcal{F} with at most one arrow. ∎

4 Seifert Fibered Spaces

Seifert manifolds with orientable base orbifold can be built as Dehn fillings of Σ×S1\Sigma\times S^{1} where Σ\Sigma is a compact orientable surface. A result of Przytycki [13] implies that their Kauffman bracket skein modules are isomorphic to the quotient of 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) by a submodule generated by curves in Σ×S1\partial\Sigma\times S^{1} bounding disks after the fillings. In this section, we use these new relations to show the finiteness conjectures for a large family of Seifert fibered spaces. For details on the notation see next subsection.

Theorem 4.1.

Let M=M(g;b,{(qi,pi)}i=1n)M=M\left(g;b,\{(q_{i},p_{i})\}_{i=1}^{n}\right) be an orientable Seifert fibered space with non-empty boundary. Suppose MM has orientable orbifold base. Then, 𝒮(M)\mathcal{S}(M) is a finitely generated 𝒮(M,(A))\mathcal{S}(\partial M,\mathbb{Q}(A))-module of rank at most (22g+11)i=1n(2qi1)(2^{2g+1}-1)\prod_{i=1}^{n}(2q_{i}-1).

Theorem 4.2.

Seifert fibered spaces of the form M(g;1,{(1,pi)}i=1n)M\left(g;1,\{(1,p_{i})\}_{i=1}^{n}\right) satisfy Conjecture 2. In particular, Conjecture 2 holds for Σg,1×S1\Sigma_{g,1}\times S^{1}.

4.1 Links in Seifert manifolds

Let Σ\Sigma be a compact orientable surface of genus g0g\geq 0 with N0N\geq 0 boundary components. Fix non-negative integers nn, bb with N=n+bN=n+b. Denote the boundary components of Σ\Sigma by 1,,N\partial_{1},\dots,\partial_{N} and the isotopy class of a circle fiber in Σ×S1\Sigma\times S^{1} by λ={pt}×S1\lambda=\{pt\}\times S^{1}. For each i=1,,ni=1,\dots,n, let (qi,pi)(q_{i},p_{i}) be pairs of relatively prime integers satisfying 0<qi<|pi|0<q_{i}<|p_{i}|. Let M(g;b,{(qi,pi)}i=1n)M\left(g;b,\{(q_{i},p_{i})\}_{i=1}^{n}\right) be the result of gluing nn solid tori to Σ×S1\Sigma\times S^{1} in such way that the curve pi[λ]+qi[i]H1(i×S1)p_{i}[\lambda]+q_{i}[\partial_{i}]\in H_{1}(\partial_{i}\times S^{1}) bounds a disk. In summary, Σ\Sigma is the base orbifold of the Seifert manifold, nn counts the number of exceptional fibers, and is bb the number of boundary components of the 3-manifold.

Let MM be an orientable Seifert manifold with orientable orbifold base. It is a well known fact that MM is homeomorphic to some M(g;b,{(qi,pi)}i=1n)M\left(g;b,\{(q_{i},p_{i})\}_{i=1}^{n}\right) [6]. Links in MM can be isotoped to lie inside Σ×S1\Sigma\times S^{1}. Thus, we can represent links in MM as arrowed diagrams in Σ\Sigma with some extra Reidemester moves. By Proposition 2.2 of [13] and Proposition 3.9, 𝒮(M)\mathcal{S}(M) is generated by the family of simple diagrams ={γα,Uα}\mathcal{B}=\{\gamma\cup\alpha,U\cup\alpha\}.

Definition 4.3.

Let DD\in\mathcal{B}. Let li0l_{i}\geq 0 be the number of parallel copies of i\partial_{i} in DD. Let εi0\varepsilon_{i}\geq 0 be the number of arrows (regardless of orientation) among all components of DD parallel to i\partial_{i}. If DD contains an unknot UU, denote by u0u\geq 0 the number of arrows in UU. If DD contains a non-separating loop, let u=0u=0. The absolute arrow sum of DD is the total number of arrows among its separating loops s:=u+iεis:=u+\sum_{i}\varepsilon_{i}. DD is standard if 0εi10\leq\varepsilon_{i}\leq 1 for every i=1,,Ni=1,\dots,N; and such arrows (if exist) lie in the loop furthest from the boundary.

Lemma 4.4.

Every diagram DD in \mathcal{B} is a [A±1]\mathbb{Z}[A^{\pm 1}]-linear combination of standard diagrams DD^{\prime} satisfying sss^{\prime}\leq s and lili,il_{i}^{\prime}\leq l_{i},\forall i.

Proof.

Follows from Proposition 4.1 of [2] and Lemmas 1.4, 1.6, and 1.7. ∎

The rest of this section is devoted to understand how the quantities ss and lil_{i} behave under certain relations in \mathcal{B}. We use Lemma 4.4 implicitly to rewrite any relation in terms of standard diagrams with bounded sums ss and ll.

Remark 4.5 (Moving arrows).

We think of Lemma 1.7 as a set of moves that change the arrows between consecutive circles at the expense of adding ‘debris’ terms. Observe that |ba+2||a|+|b||b-a+2|\leq|a|+|b| as long as b<0b<0 or a>0a>0. In particular, the debris terms in the equations of Lemma 1.7 parts (i) and (iii) will have absolute arrow sums bounded above by the LHS whenever b<0b<0 or a>0a>0. The same happens with parts (ii) and (iv) when b>0b>0 or a<0a<0. This can be summarized as follows: “We can move arrows between consecutive nested loops without increasing the arrow sum nor lil_{i}.”

4.1.1 Local moves around an exceptional fiber

Fix an index i=1,,ni=1,\dots,n. By construction, there is a loop βi\beta_{i} in the torus i×S1\partial_{i}\times S^{1} bounding a disk BiB_{i} in MM; βi\beta_{i} homologous to (pi[λ]+qi[i])H1(i×S1,)\left(p_{i}[\lambda]+q_{i}[\partial_{i}]\right)\in H_{1}(\partial_{i}\times S^{1},\mathbb{Z}). Following [10], we can slide arcs in Σ×S1\Sigma\times S^{1} over the disk BiB_{i} and get new Reidemeister moves for arrowed projections in Σ×S1\Sigma\times S^{1}. We obtain a new move, denoted by Ω(qi,pi)\Omega(q_{i},p_{i}) (see Figure 8).

Refer to caption
Figure 8: Ω(qi,pi)\Omega(q_{i},p_{i}) is obtained by drawing qiq_{i} concentric circles and pip_{i} arrows equidistributed. Notice that the orientation of the arrows in the RHS is determined by the sign of pip_{i}.

We can perform the Ω(qi,pi)\Omega(q_{i},p_{i})-move on an unknot near the boundary i\partial_{i} and resolve the qi1q_{i}-1 crossings with K1 relations. Since 0<qi<|pi|0<q_{i}<|p_{i}|, there is only one state with orientations of the arrows not cancelling. This unique state has exactly qiq_{i} concentric loops while the other states have strictly fewer loops and no more than |pi|2|p_{i}|-2 arrows. We then obtain an equation in 𝒮(M)\mathcal{S}(M) called the Ω(qi,pi)\Omega(q_{i},p_{i})-relation. Figure 9 shows a concrete example of this equation.

Remark 4.6 (The Ω(qi,pi)\Omega(q_{i},p_{i})-relation).

The Ω(qi,pi)\Omega(q_{i},p_{i})-relation lets us write a diagram with qiq_{i} concentric loops and |pi||p_{i}| arrows arranged in a particular way as a [A±]\mathbb{Z}[A^{\pm}]-linear combination of diagrams with 0li<qi0\leq l_{i}<q_{i} concentric circles and 0εi<|pi|0\leq\varepsilon_{i}<|p_{i}| arrows (see Figure 9). The LHS has |pi||p_{i}| arrows oriented in the same direction depending on the sign of pip_{i}; counterclockwise if pi>0p_{i}>0 and clockwise otherwise. Notice that the condition 0<qi<|pi|0<q_{i}<|p_{i}| implies that every parallel loop in the LHS has at least one arrow.

The special arrangement of arrows in the LHS of the Ω(qi,pi)\Omega(q_{i},p_{i})-relation is important and depends on the pair (qi,pi)(q_{i},p_{i}). In practice, we rearrange the arrows around the outer qiq_{i} copies of i\partial_{i} to match with the LHS of the Ω(qi,pi)\Omega(q_{i},p_{i})-relation. Lemma 4.7 uses this idea in a particular setup.

Refer to caption=A2Refer to captionA2Refer to captionA2Refer to captionA4Refer to caption\includegraphics[valign={c},scale={.45}]{Omega_221.png}=A^{2}\includegraphics[valign={c},scale={.45}]{Omega_0.png}-A^{2}\includegraphics[valign={c},scale={.45}]{Omega_21.png}-A^{2}\includegraphics[valign={c},scale={.45}]{Omega_01.png}-A^{4}\includegraphics[valign={c},scale={.45}]{Omega_1.png}

Figure 9: Ω(3,5)\Omega(3,5)-relation.
Lemma 4.7.

The following equation in 𝒮(M)\mathcal{S}(M) relates identical diagrams outside a neighborhood of 1\partial_{1}. Let DD\in\mathcal{B} and x|p1|x\geq|p_{1}|. Suppose that l1q1l_{1}\geq q_{1}, the loop furthest from 1\partial_{1} has xx arrows with the same orientation as in the LHS of the Ω(q1,p1)\Omega(q_{1},p_{1})-relation, and no other loop in DD parallel to 1\partial_{1} has arrows. Then DD is a sum of diagrams DD^{\prime}\in\mathcal{B} with l1<l1l^{\prime}_{1}<l_{1} and at most xx arrows.

Proof.

Rearrange the arrows to prepare for the Ω(q1,p1)\Omega(q_{1},p_{1})-relation using Lemma 1.7. Remark 4.5 explains that the debris terms in this procedure will have arrow sum at most xx and l1<l1l^{\prime}_{1}<l_{1}. After performing the Ω(q1,p1)\Omega(q_{1},p_{1})-move, we obtain diagrams with lesser loops l1<l1l^{\prime}_{1}<l_{1}. Observe that the lower arrow sum is explained due to at least one pair of arrows getting cancelled; this always happens since 0<q1<|p1|0<q_{1}<|p_{1}|. In particular, we lose at least two arrows when performing the move. ∎

4.1.2 Global relations

We now discuss relations among elements in \mathcal{B} of the form UαU\cup\alpha. Lemma 4.8 permits us to add new loops around each i\partial_{i} all of which have one arrow of the same direction. This move is valid as long as we have enough arrows on the unknot UU; i.e. u4g+2Nu\geq 4g+2N. The debris terms are [A±1]\mathbb{Z}[A^{\pm 1}]-linear combinations of standard diagrams with fewer arrow sum and lili+1l^{\prime}_{i}\leq l_{i}+1. This move is key to prove Theorem 4.2.

Consider the decomposition 𝒫+\mathcal{P}_{+} of Σ\Sigma described in Figure 10. Set 0\partial_{0} to be the left-most unknot in 𝒫+\mathcal{P}_{+} oriented counterclockwise. As we did in Definition 2.1, if viv_{i}\in\mathbb{Z} we will draw one copy of i\partial_{i} with viv_{i} arrows oriented as in 𝒫+\mathcal{P}_{+}, and do nothing if vi=v_{i}=\emptyset. For v({})N+1v\in\left(\mathbb{Z}\cup\{\emptyset\}\right)^{N+1}, denote by EvE_{v} the diagram obtained by drawing i\partial_{i} with viv_{i} arrows on it. For example, E(b,,,)E_{(b,\emptyset,\dots,\emptyset)} corresponds to the arrowed unknot SbS_{b}.

Refer to caption
Figure 10: 𝒫+\mathcal{P}_{+} induces linear pants decompositions on Σ′′\Sigma^{\prime\prime} and sausage decompositions on Σ\Sigma^{\prime}.

We define the Δ\Delta-maps from Definition 2.4 on the family of diagrams EvE_{v} with exactly one of v0v_{0} and vNv_{N} being empty. If v0=v_{0}=\emptyset and vNv_{N}\in\mathbb{Z}, define s(Ev)=E(v0,,vN1,vN+2)s(E_{v})=E_{(v_{0},\dots,v_{N-1},v_{N}+2)}. If v0v_{0}\in\mathbb{Z} and vN=v_{N}=\emptyset, define s(Ev)=E(v0+2,,vN1,vN)s(E_{v})=E_{(v_{0}+2,\dots,v_{N-1},v_{N})}.

Lemma 4.8.

Let a0a\geq 0. The following equation in 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}) holds modulo [A±1]\mathbb{Z}[A^{\pm 1}]-linear combinations of standard diagrams E(,,,a)E_{(\emptyset,\dots,\emptyset,a^{\prime})} and E(a′′,1,1,,1,)E_{(a^{\prime\prime},1,1,\dots,1,\emptyset)} with aa^{\prime}, a′′+(N1)a^{\prime\prime}+(N-1) integers in [0,a+4g+2N2)[0,a+4g+2N-2).

E(,,,a+4g+2N2)(1)N1A4g2N+4E(a+4g+N1,1,1,,1,).E_{(\emptyset,\dots,\emptyset,a+4g+2N-2)}\cong(-1)^{N-1}A^{-4g-2N+4}E_{(a+4g+N-1,1,1,\dots,1,\emptyset)}.
Proof.

Observe first that 𝒫+\mathcal{P}_{+} induces a linear pants decomposition on Σ′′\Sigma^{\prime\prime} as in Section 2. Here, a copy of N\partial_{N} with xx\in\mathbb{Z} arrows, E(,,,x)E_{(\emptyset,\dots,\emptyset,x)}, corresponds to the diagram Dx,(,,)N1D^{N-1}_{x,(\emptyset,\dots,\emptyset)}. Equation (4) of Lemma 2.5 with n=k0=N1n=k_{0}=N-1 states the following

Δ+N1(Da,(,,)N1)=e{0,1}N1(1)o(e)Az(e)o(e)Da+o(e),e0.\Delta^{N-1}_{+}\left(D^{N-1}_{a,(\emptyset,\dots,\emptyset)}\right)=\sum_{e\in\{0,1\}^{N-1}}(-1)^{o(e)}A^{z(e)-o(e)}D^{0}_{a+o(e),e}\quad.

For any xx\in\mathbb{Z} and v{0,1}N1v\in\{0,1\}^{N-1}, the diagram Dx,v0D^{0}_{x,v} contains a copy of the curve cc (see Figure 10) with xx arrows. Now, observe that 𝒫+\mathcal{P}_{+} also induces a sausage decomposition of Σ\Sigma^{\prime} (see [2]). Using the notation in Section 3.3 of [2], the part of the diagram Dx,v0D^{0}_{x,v} inside the subsurface ΣΣ\Sigma^{\prime}\subset\Sigma is denoted by Dx2gD^{2g}_{x}. Proposition 3.13 of [2] implies the equation Δ+2g(Dx2g)=Δ2g(Dx0)\Delta^{2g}_{+}(D^{2g}_{x})=\Delta^{2g}_{-}(D^{0}_{x}), where Dx0D^{0}_{x} is a copy of the left-most unknot 0\partial_{0} (red loop) in 𝒫+\mathcal{P}_{+} with xx arrows. Putting everything together, we obtain the following relation in 𝒮(Σ×S1)\mathcal{S}(\Sigma\times S^{1}):

Δ+2g+N1(E(,,,a))=e{0,1}N1(1)o(e)Az(e)o(e)Δ2g(E(a+o(e),e1,,eN,)).\Delta^{2g+N-1}_{+}(E_{(\emptyset,\dots,\emptyset,a)})=\sum_{e\in\{0,1\}^{N-1}}(-1)^{o(e)}A^{z(e)-o(e)}\Delta^{2g}_{-}(E_{(a+o(e),e_{1},\dots,e_{N},\emptyset)}).

The result follows by taking the summads on each side with the most number of arrows. ∎

4.2 Proofs of Theorems 4.1 and 4.2

Recall that 𝒮(M)\mathcal{S}(M) is generated by all standard diagrams, and such diagrams are filtered by the complexity (s,ili)(s,\sum_{i}l_{i}) in lexicographic order. Here, s=u+iεis=u+\sum_{i}\varepsilon_{i} is the absolute arrow sum and ili\sum_{i}l_{i} is the number of boundary parallel loops. Throughout the argument we will have debris terms with lower complexity (s,ili)(s^{\prime},\sum_{i}l^{\prime}_{i}); on each of those terms, we can perform a series of combinations of Lemmas 1.4, 1.5, 1.6, and 1.7 in order to write them in terms of standard diagrams with complexities s′′ss^{\prime\prime}\leq s^{\prime} and li′′lil^{\prime\prime}_{i}\leq l^{\prime}_{i}.

Let DD\in\mathcal{B} be a diagram. Suppose that DD is of the form D=γαD=\gamma\cup\alpha, where γ\gamma is an non-separating simple closed curve with at most one arrow and α\alpha is a collection of arrowed boundary parallel loops. We can rewrite DD in 𝒮(M)\mathcal{S}(M) as D=1(A2+A2)(DU)D=\frac{-1}{(A^{2}+A^{-2})}(D\cup U) where UU is a small unknot with no arrows. Proposition 4.9 focuses on the subdiagram UαU\cup\alpha near a fixed boundary component.

Proposition 4.9.

Let DD\in\mathcal{B} be a standard diagram with li0qi0l_{i_{0}}\geq q_{i_{0}} for some i0{1,,n}{i_{0}}\in\{1,\dots,n\}. Then DD is a linear combination of some standard diagrams DD^{\prime} identical to DD outside a neighborhood of i0\partial_{i_{0}}, satisfying

li0<li0 and u+εi02(u+|pi0|).l^{\prime}_{i_{0}}<l_{i_{0}}\text{ and }u^{\prime}+\varepsilon^{\prime}_{i_{0}}\leq 2(u+|p_{i_{0}}|).
Proof.

For simplicity, set i0=1{i_{0}}=1. We assume that p1>0p_{1}>0 so that the arrows in the LHS of the Ω(q1,p1)\Omega(q_{1},p_{1})-relation are oriented counterclockwise; the other case is analogous. We can assume that if ε1=1\varepsilon_{1}=1, then the orientation of the arrow in the loop furtherst from 1\partial_{1} agrees with the LHS of the Ω(q1,p1)\Omega(q_{1},p_{1})-relation. This is true since Lemma 1.8 lets us flip the orientation at the expense of having one debris diagram with l1=l1l^{\prime}_{1}=l_{1}, ε1=0\varepsilon^{\prime}_{1}=0, and u=u+1u^{\prime}=u+1.

Denote by DxD_{x} the standard diagram in \mathcal{B}, identical to DD away from a neighborhood of 1\partial_{1} with l1l_{1} copies of 1\partial_{1}, having xx arrows oriented counterclockwise in the loop furtherst from 1\partial_{1}. Recall that SaS_{a} denotes a small unknot with aa\in\mathbb{Z} arrows oriented counterclockwise. We have that D=Dε1SuD=D_{\varepsilon_{1}}\cup S_{u}, where the disjoint union of the diagrams is made so that SuS_{u} lies inside a small disk away from the diagram DxD_{x}.

Merge the arrows on UU with the outer loop around 1\partial_{1} using Lemma 1.6. Thus, DD is a linear combination of diagrams DxD_{x} with no unknots (U=U=\emptyset). If ε1=1\varepsilon_{1}=1, we get diagrams with 0xu+ε10\leq x\leq u+\varepsilon_{1}, and if ε1=0\varepsilon_{1}=0, we obtain diagrams with 0|x|u0\leq|x|\leq u. We focus on each DxD_{x}. Use the relation around 1\partial_{1}

[Uncaptioned image]=[Uncaptioned image][Uncaptioned image]=A2[Uncaptioned image]A4[Uncaptioned image]\includegraphics[valign={c},scale={.45}]{47_a.png}=\includegraphics[valign={c},scale={.45}]{47_b.png}\quad\implies\quad\includegraphics[valign={c},scale={.55}]{47_x.png}=-A^{2}\includegraphics[valign={c},scale={.55}]{47_x+11.png}-A^{4}\includegraphics[valign={c},scale={.55}]{47_x+2.png} (6)

to write DxD_{x} as a linear combination of Dx+1S1D_{x+1}\cup S_{1} and Dx+2D_{x+2}. Thus, at the expense of getting a cluster of 1-arrowed unknots S±1S_{\pm 1}, we can increase/decrease the arrows in the outermost loop around 1\partial_{1}. Hence, the original diagram DD is a linear combination of diagrams of the form Dx(yS1)D_{x}\cup\left(\cup_{y}S_{1}\right) where xp1x\geq p_{1}, y0y\geq 0 and x+y2(u+p1)x+y\leq 2(u+p_{1}). To see the upper bound for x+yx+y, observe that if we start with DuD_{-u}, one might need to add a copy of S1S_{1} (u+p1)(u+p_{1}) times in order to reach xp1x\geq p_{1}. Lemma 4.7 implies that each D±x(yS1)D_{\pm x}\cup\left(\cup_{y}S_{1}\right) is a linear combination of diagrams with l1<l1l^{\prime}_{1}<l_{1} and at most x+yx+y arrows. After making such diagrams standard and merging the arrowed unknots, we obtain diagrams with l1<l1l^{\prime}_{1}<l_{1} and u+ε12(u+p1)u^{\prime}+\varepsilon^{\prime}_{1}\leq 2(u+p_{1}) as desired. ∎

Proof of Theorem 4.1.

Let M=M(g;b,{(qi,pi)}i=1n)M=M\left(g;b,\{(q_{i},p_{i})\}_{i=1}^{n}\right) be a Seifert fibered space with non-empty boundary. Proposition 3.9 and Lemma 4.4 imply that 𝒮(M)\mathcal{S}(M) is generated over (A)\mathbb{Q}(A) by standard diagrams in \mathcal{B}. Furthermore, it follows from Lemmas 1.6, 1.7, and 1.8 that it is enough to consider standard diagrams with all arrows on separating loops oriented counterclockwise. Notice that the standard condition allow us to overlook the numbers ln+jl_{n+j} for j=1,,bj=1,\dots,b since they correspond to coefficients of the ring 𝒮(M,(A))\mathcal{S}(\partial M,\mathbb{Q}(A)).

Divide the collection \mathcal{B} into two sets ns={γα}\mathcal{B}_{ns}=\{\gamma\cup\alpha\} and U={Uα}\mathcal{B}_{U}=\{U\cup\alpha\}. Proposition 4.1 of [2] implies that arrowed non-separating simple closed curves are equal in 𝒮(M)\mathcal{S}(M) if they are the same loop and have the same number of arrows modulo 2. Thus, using Proposition 4.9, we obtain that (A)ns\mathbb{Q}(A)\cdot\mathcal{B}_{ns} is generated by standard diagrams D=γαD=\gamma\cup\alpha with 0li<qi0\leq l_{i}<q_{i} for i=1,,ni=1,\dots,n and all arrows in copies of \partial-parallel loops oriented counterclockwise. Hence, (A)ns\mathbb{Q}(A)\cdot\mathcal{B}_{ns} is generated as a 𝒮(M,(A))\mathcal{S}(\partial M,\mathbb{Q}(A))-module by a set of cardinality

rns(22g+12)i=1n(2qi1).r_{ns}\leq(2^{2g+1}-2)\prod_{i=1}^{n}(2q_{i}-1).

Proposition 4.9 implies that (A)U\mathbb{Q}(A)\cdot\mathcal{B}_{U} is generated over (A)\mathbb{Q}(A) by standard diagrams satisfying 0li<qi0\leq l_{i}<q_{i} for all i=1,,ni=1,\dots,n. Therefore, since UU can be pushed towards the boundary, (A)U\mathbb{Q}(A)\cdot\mathcal{B}_{U} is generated over 𝒮(M,(A))\mathcal{S}(\partial M,\mathbb{Q}(A)) by a finite set of cardinality

rUi=1n(2qi1),r_{U}\leq\prod_{i=1}^{n}(2q_{i}-1),

Hence, 𝒮(M)\mathcal{S}(M) is a finitely generated 𝒮(M,(A))\mathcal{S}(\partial M,\mathbb{Q}(A))-module. ∎

Proof of Theorem 4.2.

Let λM\lambda\subset\partial M be a S1S^{1}-fiber and let μN=N×{pt}\mu_{N}=\partial_{N}\times\{pt\} be a meridian of M\partial M. For i=1,,ni=1,\dots,n, the Ω(1,pi)\Omega(1,p_{i})-move turns loops parallel to i\partial_{i} into arrowed unknots. Thus, Proposition 3.9, Lemma 4.4, and Equation (1) imply that 𝒮(M)\mathcal{S}(M) is generated over (A)\mathbb{Q}(A) by standard diagrams in ={γα,Uα}\mathcal{B}=\{\gamma\cup\alpha,U\cup\alpha\} with no parallel loops around the exceptional fibers. In particular, α\alpha only contains loops around N\partial_{N}. Hence (A)ns\mathbb{Q}(A)\cdot\mathcal{B}_{ns} is generated over (A)[μN]\mathbb{Q}(A)[\mu_{N}] by elements of the form γ\gamma and γα\gamma\cup\alpha where γ\gamma\in\mathcal{F} has at most one arrow and α\alpha is a copy of N\partial_{N} with one arrow.

Let UαUU\cup\alpha\in\mathcal{B}_{U} and suppose that UU has u0u\neq 0 arrows. Using Equation (6) of Proposition 4.9, we can assume that the loop of α\alpha furthest to the boundary has at least one arrow. Then, using Lemmas 1.5 and 1.6, we can write any diagram in U\mathcal{B}_{U} as a (A)\mathbb{Q}(A)-linear combination of diagrams with only \partial-parallel curves and such that the loop furthest to N\partial_{N} has x0x\geq 0 arrows oriented clockwise. In other words, (A)U=(A)U,μNkαx|k,x0\mathbb{Q}(A)\cdot\mathcal{B}_{U}=\mathbb{Q}(A)\langle U,\mu_{N}^{k}\cdot\alpha_{x}|k,x\geq 0\rangle, where αx\alpha_{x} denotes a copy of μN\mu_{N} with xx arrows.

We will see that it is enough to consider 0x<4g+2n0\leq x<4g+2n. Take μNkαx\mu_{N}^{k}\cdot\alpha_{x} with k0k\geq 0 and x4g+2nx\geq 4g+2n. By Lemma 4.8, μNkαx\mu_{N}^{k}\cdot\alpha_{x} is a [A±1]\mathbb{Z}[A^{\pm 1}]-linear combination of diagrams of the form UμNkU\cup\mu_{N}^{k} and μNkαy\mu_{N}^{k}\cdot\alpha_{y} with 0y<x0\leq y<x. We can proceed as in the previous paragraph and write the diagrams UμNkU\cup\mu_{N}^{k} as [A±1]\mathbb{Z}[A^{\pm 1}]-linear combinations of μNmax(0,k1)αx\mu_{N}^{\max(0,k-1)}\cdot\alpha_{x^{\prime}} for some x0x^{\prime}\geq 0. Hence, (A)U=(A)U,μNkαx|0k,0x<4g+2n\mathbb{Q}(A)\cdot\mathcal{B}_{U}=\mathbb{Q}(A)\langle U,\mu_{N}^{k}\cdot\alpha_{x}|0\leq k,0\leq x<4g+2n\rangle.

To end the proof, consider F1F_{1} the subspace generated by ns{μNkαx|0x<4g+2n}\mathcal{B}_{ns}\cup\{\mu_{N}^{k}\cdot\alpha_{x}|0\leq x<4g+2n\} over (A)\mathbb{Q}(A), and F2F_{2} the (A)\mathbb{Q}(A)-subspace generated by arrowed unknots. By Proposition 3.9, 𝒮(M)=F1+F2\mathcal{S}(M)=F_{1}+F_{2}. Let Σ1\Sigma_{1} and Σ2\Sigma_{2} be neighborhoods of μN\mu_{N} and λ\lambda in M\partial M, respectively. We have shown that F1F_{1} is a 𝒮(Σ1,(A))\mathcal{S}(\Sigma_{1},\mathbb{Q}(A))-module of rank at most 2(22g+12)+4g+2n2(2^{2g+1}-2)+4g+2n. Also, since every arrowed unknot can be pushed inside a neighborhood of Σ2\Sigma_{2}, F2F_{2} is generated over 𝒮(Σ2,(A))\mathcal{S}(\Sigma_{2},\mathbb{Q}(A)) by the empty link. So F2F_{2} is a 𝒮(Σ2,(A))\mathcal{S}(\Sigma_{2},\mathbb{Q}(A))-module of rank at most one. ∎

References

  • [1] Renaud Detcherry. Infinite families of hyperbolic 3-manifolds with finite-dimensional skein modules. Journal of the London Mathematical Society, 2020.
  • [2] Renaud Detcherry and Maxime Wolff. A basis for the kauffman skein module of the product of a surface and a circle, 2020.
  • [3] Charles Frohman and Răzvan Gelca. Skein modules and the noncommutative torus. Trans. Amer. Math. Soc., 352(10):4877–4888, 2000.
  • [4] Patrick M. Gilmer and Gregor Masbaum. On the skein module of the product of a surface and a circle. Proc. Amer. Math. Soc., 147(9):4091–4106, 2019.
  • [5] Sam Gunningham, David Jordan, and Pavel Safronov. The finiteness conjecture for skein modules. arXiv preprint arXiv:1908.05233, 2019.
  • [6] Allen Hatcher. Notes on basic 3-manifold topology, 2007.
  • [7] Jim Hoste and Józef H. Przytycki. The (2,)(2,\infty)-skein module of lens spaces; a generalization of the Jones polynomial. J. Knot Theory Ramifications, 2(3):321–333, 1993.
  • [8] Jim Hoste and Józef H. Przytycki. The Kauffman bracket skein module of S1×S2S^{1}\times S^{2}. Math. Z., 220(1):65–73, 1995.
  • [9] M. Mroczkowski and M. K. Dabkowski. KBSM of the product of a disk with two holes and S1S^{1}. Topology Appl., 156(10):1831–1849, 2009.
  • [10] Maciej Mroczkowski. Kauffman bracket skein module of a family of prism manifolds. J. Knot Theory Ramifications, 20(1):159–170, 2011.
  • [11] Maciej Mroczkowski. Kauffman bracket skein module of the connected sum of two projective spaces. J. Knot Theory Ramifications, 20(5):651–675, 2011.
  • [12] Józef H. Przytycki. Skein modules of 33-manifolds. Bull. Polish Acad. Sci. Math., 39(1-2):91–100, 1991.
  • [13] Józef H. Przytycki. Fundamentals of Kauffman bracket skein modules. Kobe J. Math., 16(1):45–66, 1999.
  • [14] Adam S. Sikora and Bruce W. Westbury. Confluence theory for graphs. Algebr. Geom. Topol., 7:439–478, 2007.
  • [15] V. G. Turaev. The Conway and Kauffman modules of a solid torus. Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), 167(Issled. Topol. 6):79–89, 190, 1988.

José Román Aranda, University of Iowa

email: [email protected]
   

Nathaniel Ferguson, Colby College

email: [email protected]