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

Links in Surfaces and Laplacian Modules

Daniel S. Silver    Susan G. Williams
Abstract

Laplacian matrices of weighted graphs in surfaces SS are used to define module and polynomial invariants of /2\mathbb{Z}/2-homologically trivial links in S×[0,1]S\times[0,1]. Information about virtual genus is obtained.

MSC: 05C10, 57M25

1 Introduction

The Laplacian matrix of a signed graph is a discrete version of the well-studied Laplacian operator of physics. Its spectrum yields insights about the structure of the graph. (For a survey see [12]). For signed graphs embedded in a closed oriented surface SS, information about how the graph winds about SS can be used to define a Laplacian matrix LGL_{G} with homology coefficients (see, for example, [9]). We introduce the module G{\cal L}_{G} presented by LGL_{G}. Its module order is a polynomial ΔG\Delta_{G} over the Laurent polynomial ring Λ=[x1±1,y1±1,,xg±1,yg±1]{\Lambda}=\mathbb{Z}[x_{1}^{\pm 1},y_{1}^{\pm 1},\ldots,x_{g}^{\pm 1},y_{g}^{\pm 1}], where gg is the genus of SS. The polynomial is the determinant of LGL_{G}, but we will see that it can also be computed using a simple skein relation.

When the graph GG is embedded in SS the well-known medial construction associates a checkerboard colored diagram DSD\subset S for a link \ell in the thickened surface S×[0,1]S\times[0,1]. In [7] the authors showed that signature, nullity and determinant, classical link invariants, can be defined for \ell using the usual Laplacian matrix, which can be viewed as a generalization of the Goeritz matrix of a link. Here we consider the Laplacian matrix with homological coefficients. We show that the module G{\cal L}_{G} it presents is unaffected by Reidemeister moves. Invariants of links in the thickened surface S×[0,1]S\times[0,1] are described.

Any link \ell in S×[0,1]S\times[0,1] can be described by a diagram in SS. Adding or deleting “hollow handles” to SS, avoiding DD, might produce a surface of smaller genus. The minimum possible genus is the virtual genus of \ell. In many cases the polynomial ΔG\Delta_{G} can be used to establish that a surface achieves the minimum possible genus.

Acknowledgements

The authors are grateful to Seiichi Kamada, Louis H. Kauffman and Christine Ruey Shan Lee for helpful comments.

2 Laplacian modules and polynomials

A graph GG in a closed, connected oriented surface SS with edge and vertex sets VG,EGV_{G},E_{G}, respectively, is signed if every edge eEGe\in E_{G} is labeled with σe=+1\sigma_{e}=+1 or 1-1. In order to avoid visual clutter, unlabeled edges will be assumed to have sign +1+1.

Let S~\tilde{S} denote the universal abelian covering space of SS. We regard the deck transformation group A(S~)A(\tilde{S}) multiplicatively. It is isomorphic to H1(S~;)H_{1}(\tilde{S};\mathbb{Z}). We fix a symplectic homology basis ={x1,y1,,xg,yg}{\cal B}=\{x_{1},y_{1},\ldots,x_{g},y_{g}\}, and represent its members by oriented closed curves. Each closed curve in SS is labeled by the element of A(S~)A(\tilde{S}) that it determines, a monomial in the variables xi,yix_{i},y_{i}.

Although we work with undirected graphs, it is convenient for the following definition to regard GG as directed. There is a standard way to do this: replace each edge eEGe\in E_{G} (including loops) with a pair of edges joining the same endpoints, each with the same weight as ee but with opposite directions. Any directed path in SS determines an element ϕPA(S~)\phi_{P}\in A(\tilde{S}), read by traveling along the path in the preferred direction, recording yk±1y_{k}^{\pm 1} (resp. xk±1x_{k}^{\pm 1}) as we cross the homology basis curve that is labeled xkx_{k} (resp. yky_{k}), the exponent determined as in Figure 1. In particular, each directed edge eEGe\in E_{G} determines a monomial ϕe\phi_{e}.

Refer to caption
Figure 1: Finding the monomial of a path transverse to GG

Assume that GG is a signed directed graph with vertex set VG={v1,,vn}V_{G}=\{v_{1},\ldots,v_{n}\}. The adjacency matrix AG=(ai,j)A_{G}=(a_{i,j}) of GG (with respect to {\cal B}) is the n×nn\times n matrix with non-diagonal entries ai,ja_{i,j} equal to σeϕe\sum\sigma_{e}\phi_{e}, where the summation ranges over all edges from viv_{i} to vjv_{j}. Diagonal entries of AGA_{G} are zero. Denote by δg=(δi,j){\delta}_{g}=({\delta}_{i,j}) the n×nn\times n diagonal matrix with δi,i=jai,j{\delta}_{i,i}=\sum_{j}a_{i,j}.

Definition 2.1.

The Laplacian matrix LGL_{G} of GG (with respect to {\cal B}) is AGδGA_{G}-{\delta}_{G}. The Laplacian matrix of a signed undirected graph GG is the Laplacian matrix of the associated directed graph.

Henceforth the graphs that we consider will be weighted but undirected.

The entries of LGL_{G} are elements of the Laurent polynomial ring Λ=[x1±1,y1±1,,xg±1,yg±1]{\Lambda}=\mathbb{Z}[x_{1}^{\pm 1},y_{1}^{\pm 1},\ldots,x_{g}^{\pm 1},y_{g}^{\pm 1}]. The graph GG determines a line bundle with a copy of Λ{\Lambda} at each vertex. Adopting the terminology of [9], we call ϕe\phi_{e} the connection of the edge ee.

Definition 2.2.

Let GG be a graph in SS. Its Laplacian polynomial ΔG\Delta_{G} (with respect to {\cal B}) is the determinant of LGL_{G}.

Remark 2.3.

When GG is directed and connections are ignored, the Laplacian matrix LGL_{G} is expressible as FFtFF^{t}, where FF is the incidence matrix of GG and FtF^{t} denotes its transpose. This form has had many applications. For example it commonly used in the proof of the Matrix Tree Theorem (see, for example, [12]). In [5] the authors use it to extend the theory of Laplacian matrices and critical groups from graphs to simplicial complexes. There FF functions as a boundary operator in a chain complex.

We can express the Laplacian matrix LGL_{G}, as defined above, as FFtFF^{t}, but we must define FF and FF^{*} suitably.

We begin by directing the edges of the graph GG arbitrarily. As above, v1,,vnv_{1},\ldots,v_{n} are the vertices of GG. Let e1,,eme_{1},\ldots,e_{m} denote the edges. Then F=(fi,j)F=(f_{i,j}) is an n×mn\times m matrix with entries in the ring [i][x1±1,y1±1,,xg±1,yg±1]\mathbb{Z}[i][x_{1}^{\pm 1},y_{1}^{\pm 1},\ldots,x_{g}^{\pm 1},y_{g}^{\pm 1}]. For each i,ji,j, the entry fi,jf_{i,j} is equal to wjϕ(ej)\sqrt{w_{j}\phi(e_{j})} if eje_{j} has initial vertex viv_{i}; fi,j=wjϕ(ej)f_{i,j}=-\sqrt{w_{j}\phi(e_{j})} if eje_{j} terminates at viv_{i}; otherwise, fi,j=0f_{i,j}=0. Note that since wj=1,1w_{j}=1,-1, the value of wj\sqrt{w_{j}} is either 1 or ii.

The transpose matrix FF^{*} is defined as usual but we replace all connections ϕ(ej)\phi(e_{j}) with their inverses.

Example 2.4.

If GG is a 1-vertex graph, then its Laplacian polynomial has the form

σe(2ϕeϕe1),\sum\sigma_{e}(2-\phi_{e}-\phi_{e}^{-1}), (2.1)

where the summation is over the edges (loops) eEGe\in E_{G}.

Proposition 2.5.

Let ee be a non-loop edge of G. Then

ΔG=ΔGe+σeΔG/e.\Delta_{G}=\Delta_{G\setminus e}+\sigma_{e}\Delta_{G/e}. (2.2)
Proof.

We make use of a result of Forman [6] that expresses ΔG\Delta_{G} in terms of certain subgraphs of GG. A cycle-rooted spanning forest (CRSF) is a subgraph FF containing all the vertices of GG and such that each component has a unique cycle. Then

ΔG=FeEFσecyclesofF(2ϕϕ1),\Delta_{G}=\sum_{F}\prod_{e\in E_{F}}\sigma_{e}\prod_{{\rm cycles\ of\ }F}(2-\phi-\phi^{-1}), (2.3)

where the summation is over all cycle-rooted spanning forests FF of GG, and ϕ±1\phi^{\pm 1} are the connections of the two orientations of the cycle.

For ee a fixed non-loop edge of GG, we partition the CRSFs of GG into those that contain ee and those that do not. These are in one-to-one correspondence with the CRSFs of G/eG/e and GeG\setminus e, respectively. In the second case the weight eEFσe\prod_{e^{\prime}\in E_{F}}\sigma_{e^{\prime}} is unchanged, but in the first case the factor σe\sigma_{e} is lost. ∎

Proposition 2.2 enables us to compute ΔG\Delta_{G} using a skein computation tree. Leaves of the tree are disjoint unions of 1-vertex graphs. The following are easily proved and useful for computation.

  • If GG is the union of two disjoint subgraphs G1G_{1} and G2G_{2}, then ΔG=ΔG1ΔG2\Delta_{G}=\Delta_{G_{1}}\Delta_{G_{2}}.

  • If GG has no edges, then ΔG=0\Delta_{G}=0.

  • If GG is a cyclic graph, then ΔG=σ(2ϕϕ1)\Delta_{G}=\sigma(2-\phi-\phi^{-1}), where σ\sigma is the product of the signs σe\sigma_{e} of edges of GG, and ϕ\phi is the connection of the cycle (with either orientation).

Example 2.6.

Consider the theta-graph GG of Figure 2, embedded in the torus. (Recall that unlabeled edges have positive sign.) Deleting the middle, vertical edge ee, produces a 2-vertex cyclic graph GeG\setminus e. Contracting it results in a 1-vertex graph. By equation (2.2) we have

ΔG=(2xy1x1y)+(4xx1yy1)=6xx1yy1xy1x1y.\Delta_{G}=(2-xy^{-1}-x^{-1}y)+(4-x-x^{-1}-y-y^{-1})=6-x-x^{-1}-y-y^{-1}-xy^{-1}-x^{-1}y.

The polynomial can also be computed directly from the Laplacian matrix LGL_{G}.

LG=(31x1y11xy3)L_{G}=\begin{pmatrix}3&-1-x^{-1}-y^{-1}\\ -1-x-y&3\\ \end{pmatrix} (2.4)
Refer to caption
Figure 2: Theta-graph GG embedded in a torus

The Laplacian matrix LGL_{G} determines a homomorphism of the free module Λn{\Lambda}^{n} generated by the vertices of GG. The cokernel Λn/ΛnLG{\Lambda}^{n}/{\Lambda}^{n}L_{G} is a Λ{\Lambda}-module G{\cal L}_{G}, the Laplacian module of GG. Its module order is the Laplacian polynomial ΔG\Delta_{G}. In this context it is well defined only up to multiplication by units in Λ{\Lambda}. However, Forman’s equation (2.3) provides a normal form that is preserved by equation 2.2. (Definition 2.2 also gives the normalized form since it agrees with Forman’s formula.) We will henceforth assume that ΔG\Delta_{G} is normalized in this way, well defined up to multiplication by 1-1.

Remark 2.7.

When G𝕊2G\subset{\mathbb{S}}^{2}, all connections are trivial and the Laplacian module G{\cal L}_{G} is a finite abelian group. If furthermore all edge weights of GG are +1+1, then G{\cal L}_{G} is the direct sum of \mathbb{Z} and the abelian sandpile group of GG. See [3] for details.

3 Applications to links in surfaces

A thickened surface is a product S×[0,1]S\times[0,1], where SS is a closed, connected oriented surface. For pSp\in S, we think of the point (p,1)(p,1) as lying above (p,0)(p,0). A link S×[0,1]\ell\subset S\times[0,1] is a pairwise disjoint collection of finitely many embedded closed curves. It can be described by a link diagram DD: a 4-valent graph |D||D| embedded in SS, called a universe, with hidden-line effect in a neighborhood of each vertex indicating how one strand of \ell passes over another.

Refer to caption
Figure 3: Diagram DD and universe |D||D|

R-equivalence of link diagrams in SS is the equivalence relation generated by Reidemeister moves (See Figure 4.) It is well known that Reidemeister’s original proof in [13] for planar links shows generally that two links in S×[0,1]S\times[0,1] are isotopic if and only if their diagrams are R-equivalent.

A region of a link diagram is a connected component of S|D|S\setminus|D|. A link diagram is cellular if all of its regions are contractible. We can turn any diagram into a cellular diagram with Reidemeister moves of type II. Less obvious is the following.

Proposition 3.1.

If two cellular link diagrams D,DSD,D^{\prime}\subset S are equivalent by Reidemeister moves then they are related by a sequence of cellular diagrams such that each diagram is obtained from the previous one by a single Reidemeister move.

Proof.

Assume that DD and DD^{\prime} are cellular link diagrams that are equivalent by a sequence of Reidemeister moves. The only “bad move” that can destroy the contractibility of a region is the forward direction of a type II move in Figure 4. A bad move opens a channel from a region to itself. Before any bad move is performed, we add a thin finger as in Figure 5, bridging the channel and thereby protecting the region’s contractibility. The tip of the finger is free but it should follow the diagram as it is deformed, staying above the arc below its tip. We perform the sequence Reidemeister moves that brings DD to DD^{\prime}. The fingers will be stretched, twisted, and might even cross over each other but no arc should pass through a finger’s interior. Finally we retract the fingers by type II moves. Since the DD^{\prime} is cellular, non-contractible regions will not appear during this final phase, a claim that is easy to see by imagining the retractions in reverse.

Refer to caption
Figure 4: Reidmeister moves
Refer to caption
Figure 5: Adding a finger before second Reidemeister move.

A link diagram DD is checkerboard colorable if it is possible to shade certain regions of DD so that whenever two regions share a boundary, exactly one of them is shaded. A diagram with such a shading is checkerboard colored. (We will denote it also by DD in order to avoid excessive notation.) Any checkerboard colorable diagram admits exactly two such shadings. The proof of the following is easy and left to the reader.

Lemma 3.2.

A link diagram DSD\subset S is checkerboard colorable if and only if the inclusion map |D|S|D|\hookrightarrow S induces a trivial homomorphism of homology groups with /2\mathbb{Z}/2 coefficients.

Remark 3.3.

1. Links that satisfy the condition of Lemma 3.2 were called mod-2 almost classical in [1]. The condition is equivalent to the link having a diagram admitting a mod-2 Alexander numbering [4].

2. Any Reidemeister move takes one checkerboard colored diagram to another (while changing the diagrams only locally). If both diagrams are cellular, then their associated graphs are related by the Reidemeister graph moves in Figure 7. (See [17] or [8].)

3. Given a checkerboard colorable diagram for a link S×[0,1]\ell\subset S\times[0,1], the two shaded diagrams D,DD,D^{*} have a homological interpretation: Let NN be a tubular neighborhood of \ell, and consider the section of the long exact sequence of homology groups with /2\mathbb{Z}/2 coefficients that corresponding to the pair NS×[0,1]N\subset S\times[0,1]:

0/2H2(S×[0,1],N)H1(N)0\to\mathbb{Z}/2\to H_{2}(S\times[0,1],N)\to H_{1}(N) (3.1)

The shaded diagrams D,DD,D^{*} can be seen as spanning surfaces of \ell; they represent elements [D],[D]H2(S×[0,1],N)[D],[D*]\in H_{2}(S\times[0,1],N). The classes of meridians m1,,mdm_{1},\ldots,m_{d} of the components of \ell form a basis for H(N)(/2)dH_{(}N)\cong(\mathbb{Z}/2)^{d}. By exactness of the sequence (3.1), the total class m=[m1]++[md]m=[m_{1}]+\ldots+[m_{d}] has precisely two preimages in H2(S×[0,1],N)H_{2}(S\times[0,1],N). Clearly [D][D] and [D][D^{*}] are preimages. Moreover, they are distinct since DD=SD\cup D^{*}=S and hence [D]+[D][D]+[D^{*}] generates H2(S×[0,1])H2(S)/2H_{2}(S\times[0,1])\cong H_{2}(S)\cong\mathbb{Z}/2.

Reidemeister moves can interchange the two homology classes. In fact, if S=𝕊2S={\mathbb{S}}^{2}, then this is always possible [17]. However, for surfaces of higher genus this is not generally the case (see Example 3.8 below.)

A checkerboard colored cellular link diagram DSD\subset S determines a signed embedded graph GSG\subset S, unique up to isotopy, via the “medial graph” construction (see Figure 6). Vertices of GG correspond to shaded regions of DD, while each pair of vertices corresponding to shaded regions meeting at a crossing of DD are joined by an edge ee with weight wew_{e} determined as in Figure 6. From the graph we can reconstruct DD. The other checkerboard coloring of DD determines a dual signed graph GG^{*}. If ee^{*} is an edge of GG^{*} dual to ee, then we=wew_{e^{*}}=-w_{e}.

Refer to caption
Figure 6: Constructing a checkerboard colored link diagram from a medial graph
Refer to caption
Figure 7: Reidmeister moves and Reidemeister graph moves
Proposition 3.4.

Assume that G,GG,G^{\prime} are signed graphs (not necessarily embedded) in a closed, connected orientable surface SS. If GG^{\prime} is the result of applying a Reidemeister graph move to GG, then the Λ{\Lambda}-modules G,G{\cal L}_{G},{\cal L}_{G^{\prime}} are isomorphic.

Proof.

There are three moves to check, with multiple cases depending on shading and crossing signs. We check two representative cases, and leave the rest to the reader.

We examine the first Reidemeister graph move. Vertices v,wv,w in Figure 8(a) contribute the relations

(Ev+1)v=Σv+ϕew,w=ϕe1v,(E_{v}+1)v={\Sigma}_{v}+\phi_{e}w,\quad w=\phi^{-1}_{e}v, (3.2)

where EvE_{v} is the sum of signs of edges incident to vv but not ww (with loops counted twice), Σv{\Sigma}_{v} is the Λ{\Lambda}-linear combination of vertices other than ww that are joined by edges to vv, and ee is the directed edge from vv to ww. We use the second relation to eliminate ww. The first relation becomes (Ev)v=Σv(E_{v})v={\Sigma}_{v}, which is the relation corresponding to the vertex vv in Figure 8(b). Hence ΛG{\Lambda}_{G} is unchanged by the move.

Next we consider the third Reidemeister graph move. The vertices v,w1,w2,w3v,w_{1},w_{2},w_{3} in Figure 9(a) contribute the relations (where symbols have meaning similar to those in the previous case):

  1. v=ϕv,w1w1ϕv,w2w2ϕv,w3w3-v=\phi_{v,w_{1}}w_{1}-\phi_{v,w_{2}}w_{2}-\phi_{v,w_{3}}w_{3}

  2. (Ew1+1)w1=Σw1+ϕv,w11v(E_{w_{1}}+1)w_{1}={\Sigma}_{w_{1}}+\phi^{-1}_{v,w_{1}}v

  3. (Ew21)w2=Σw2ϕv,w21v(E_{w_{2}}-1)w_{2}={\Sigma}_{w_{2}}-\phi^{-1}_{v,w_{2}}v

  4. (Ew31)w3=Σw3ϕv,w31v(E_{w_{3}}-1)w_{3}={\Sigma}_{w_{3}}-\phi^{-1}_{v,w_{3}}v

We use the first relation to eliminate vv. The three remaining relations become:

  1. (Ew1+2)w1=Σw1+ϕw1,w2w2+ϕw1,w3w3(E_{w_{1}}+2)w_{1}={\Sigma}_{w_{1}}+\phi_{w_{1},w_{2}}w_{2}+\phi_{w_{1},w_{3}}w_{3}

  2. Ew2w2=Σw2+ϕw2,w1w1ϕw2,w3w3E_{w_{2}}w_{2}={\Sigma}_{w_{2}}+\phi_{w_{2},w_{1}}w_{1}-\phi_{w_{2},w_{3}}w_{3}

  3. Ew3w3=Σw3+ϕw3,w1w1ϕw3,w2w2,E_{w_{3}}w_{3}={\Sigma}_{w_{3}}+\phi_{w_{3},w_{1}}w_{1}-\phi_{w_{3},w_{2}}w_{2},

which are the relations corresponding to w1,w2,w3w_{1},w_{2},w_{3} in Figure 9(b). ∎

Refer to caption
Figure 8: Graph transformation corresponding to first Reidemeister move
Refer to caption
Figure 9: Graph transformation corresponding to third Reidemeister move
Corollary 3.5.

Assume that GG is a signed embedded graph associated to a cellular checkerboard colored diagram of a link S×[0,1]\ell\subset S\times[0,1]. The pair {G,G}\{{\cal L}_{G},{\cal L}_{G^{*}}\} is an invariant of \ell.

Proof.

Assume GG is associated to DD. If DD^{\prime} is another cellular checkerboard colored diagram, then by Proposition 3.1 there is a sequence of Reidemeister moves taking DD^{\prime} to DD and such that each intermediate diagram is cellular. The moves transform GG^{\prime} to either GG or GG^{*} (see Remark [CBC]). Proposition 3.4 implies that G{\cal L}_{G^{\prime}} is isomorphic to G{\cal L}_{G} or G{\cal L}_{G^{*}}. ∎

Recall that we take the normalized Laplacian polynomial ΔG\Delta_{G} is well defined up to multiplication by 1-1.

Corollary 3.6.

Assume that GG is a signed embedded graph associated to a cellular checkerboard colored diagram of a link S×[0,1]\ell\subset S\times[0,1]. The pair {ΔG,ΔG}\{\Delta_{G},\Delta_{G^{*}}\} is an invariant of \ell.

Remark 3.7.

1. Like G{\cal L}_{G}, the polynomial ΔG\Delta_{G} depends on the symplectic basis we chose for H1(S;)H_{1}(S;\mathbb{Z}). If we regard ΔG\Delta_{G} up to possible change of symplectic basis, then {ΔG,ΔG}\{\Delta_{G},\Delta_{G^{*}}\} is an invariant of \ell up to automorphisms of SS.

2. When S=𝕊2S={\mathbb{S}}^{2}, the matrix LGL_{G} is a Goeritz matrix of \ell (see [16]). If we regard \ell as a link in the 3-sphere, then G{\cal L}_{G} is isomorphic to H1(M2;)\mathbb{Z}\oplus H_{1}(M_{2};\mathbb{Z}), where M2M_{2} is the 2-fold cover of 𝕊3{\mathbb{S}}^{3} branched over \ell.

Refer to caption
Figure 10: The dual GG^{*} of a theta graph

Let D𝕊2D\subset{\mathbb{S}}^{2} be a checkerboard colored link diagram, and let DD^{*} denote the dual checkerboard colored diagram. By stretching an outer arc of DD and pulling over the rest of the diagram and then around the sphere to its original position, we see that DD and DD^{*} are equivalent under Reidemeister moves. (This fact was first observed in [11], where a more formal argument appears.) Hence G{\cal L}_{G} and G{\cal L}_{G^{*}} are isomorphic.

The following example shows that for diagrams in surfaces of higher genus, G{\cal L}_{G} and G{\cal L}_{G^{*}} need not be isomorphic.

Example 3.8.

The Laplacian module of the graph GG in Figure 2 has a presentation

Gv1,v23v1=(1+x+y)v2,3v2=(1+x1+y1)v1.{\cal L}_{G}\cong\langle v_{1},v_{2}\mid 3v_{1}=(1+x+y)v_{2},\quad 3v_{2}=(1+x^{-1}+y^{-1})v_{1}\rangle.

The dual graph GG^{*}, which appears in Figure 10, has a single vertex. Its Laplacian module is

Gv6v=(x+x1+y+y1+xy1+x1y)v.{\cal L}_{G^{*}}\cong{\langle}v\mid 6v=(x+x^{-1}+y+y^{-1}+xy^{-1}+x^{-1}y)v\rangle.

To see that the modules are not isomorphic, we reduce the ring Λ{\Lambda} to \mathbb{Z}, setting x,yx,y equal to 11. (More rigorously, we treat \mathbb{Z} as a trivial right Λ{\Lambda}-module and pass to the tensor product modules ΛG\mathbb{Z}\otimes_{\Lambda}{\cal L}_{G} and ΛG\mathbb{Z}\otimes_{\Lambda}{\cal L}_{G^{*}}.) The resulting abelian groups are /3\mathbb{Z}\oplus\mathbb{Z}/3 and \mathbb{Z}, respectively. Since they are not isomorphic, neither are G{\cal L}_{G} and G{\cal L}_{G^{*}}.

The Laplacian polynomials ΔG\Delta_{G} and ΔG\Delta_{G^{*}} in Example 3.8 are the same. This holds generally for signed graphs in the torus [15], a consequence of the fact that null-homologous curves in the torus are contractible. The following example shows that equality does not hold for surfaces of higher genus.

Example 3.9.

Consider the signed graph GG embedded in the surface SS of genus 2 in Figure 11. With respect to the indicated symplectic basis for H1(S;)H_{1}(S;\mathbb{Z}) the Laplacian module G{\cal L}_{G} has presentation matrix

(5xx1yy1115uu1vv1)\begin{pmatrix}5-x-x^{-1}-y-y^{-1}&-1\\ -1&5-u-u^{-1}-v-v^{-1}\\ \end{pmatrix} (3.3)

and

ΔG=245(x+x1+y+y1+u+u1+v+v1)+\Delta_{G}=24-5(x+x^{-1}+y+y^{-1}+u+u^{-1}+v+v^{-1})+
ux+u1x1+ux1+u1x+vx+v1x1+vx1+v1x+ux+u^{-1}x^{-1}+ux^{-1}+u^{-1}x+vx+v^{-1}x^{-1}+vx^{-1}+v^{-1}x+
uy+u1y1+uy1+u1y+vy+v1y1+vy1+v1y.uy+u^{-1}y^{-1}+uy^{-1}+u^{-1}y+vy+v^{-1}y^{-1}+vy^{-1}+v^{-1}y.

The dual graph GG^{*} has a single vertex, loops (suitably directed) with connections x,y,u,vx,y,u,v , and a fifth loop (transverse to ee) that is null-homologous and so has trivial connection. Hence

ΔG=8xx1yy1uu1vv1.\Delta_{G^{*}}=8-x-x^{-1}-y-y^{-1}-u-u^{-1}-v-v^{-1}.
Refer to caption
Figure 11: Graph GG in a surface of genus 2 such that ΔG\Delta_{G} and ΔG\Delta_{G^{*}} are different
Example 3.10.

The theta graph in Example 3.8 corresponds to a 3-component link 1\ell_{1} in the torus. Placing the weight 1-1 on one of the three edges results in three links 1,2,3\ell_{1},\ell_{2},\ell_{3}. The links and their graphs G1,G2,G3G_{1},G_{2},G_{3} are shown in Figure 12. The corresponding Laplacian polynomials with respect to the indicated symplectic basis are:

  • ΔG1=2+(x+x1)+(y+y1)(x1y+xy1)\Delta_{G_{1}}=-2+(x+x^{-1})+(y+y^{-1})-(x^{-1}y+xy^{-1})

  • ΔG2=2+(x+x1)(y+y1)+(x1y+xy1)\Delta_{G_{2}}=-2+(x+x^{-1})-(y+y^{-1})+(x^{-1}y+xy^{-1})

  • ΔG3=2(x+x1)+(y+y1)+(x1y+xy1)\Delta_{G_{3}}=-2-(x+x^{-1})+(y+y^{-1})+(x^{-1}y+xy^{-1})

Consequently, the links are pairwise non-isotopic.

By considering fundamental domains in the universal cover of SS, one can readily see that there is an order-3 automorphism ff of the SS sending xx to y1y^{-1} and yy to xy1xy^{-1}, where xx and yy are generators of π1(S;)\pi_{1}(S;\mathbb{Z}), and mapping the diagram of 1\ell_{1} (resp. 2\ell_{2}) to that of 2\ell_{2} (resp. 3\ell_{3}). Moreover, ff induces a change of symplectic basis of H1(S;)H_{1}(S;\mathbb{Z}) mapping xy,yxyx\mapsto-y,y\mapsto x-y that transforms ΔG1\Delta_{G_{1}} (resp. ΔG2\Delta_{G_{2}}) to ΔG2\Delta_{G_{2}} (resp. ΔG3\Delta_{G_{3}}).

Refer to caption
Figure 12: 3-component links 1,2,3\ell_{1},\ell_{2},\ell_{3} and their graphs G1,G2,G3G_{1},G_{2},G_{3} (left to right)

4 Virtual links and genus

A virtual link \ell can be regarded as a link diagram DD defined up to Reidemeister moves, surface automorphisms and addition or deletion of hollow handles. Handle addition is 0-surgery, replacing 𝕊0×𝔻2SD{\mathbb{S}}^{0}\times\mathbb{D}^{2}\subset S\setminus D with 𝔻1×𝕊1\mathbb{D}^{1}\times{\mathbb{S}}^{1}, thereby increasing the genus of the surface. The handle deletion is 1-surgery along a simple closed curve CSDC\subset S\setminus D, replacing C×𝔻1C\times\mathbb{D}^{1} with 𝔻2×𝕊0\mathbb{D}^{2}\times{\mathbb{S}}^{0}; when CC is essential and nonseparating, the surgery decreases the genus of the surface. The virtual genus vg()vg\ (\ell) is the minimum possible genus g(S)g\ (S) a surface SS that can be obtained in this way.

The information in the polynomial ΔG\Delta_{G} can often establish that g(S)=vg()g(S)=vg(\ell). Our approach is similar to that of [2]. Assume that g(S)>vg()g(S)>vg(\ell). Then a theorem of G. Kuperberg [10] implies that after Reidemeister moves S|D|S\setminus|D| contains an essential nonseparating simple closed curve CC upon which we can perform surgey to reduce the genus of SS. If we orient CC and extend its homology class to a symplectic basis for H(S;)H(S;\mathbb{R}), then this basis element will not appear among the coefficients of ΔG\Delta_{G}. Contrapositively, if the coefficients of ΔG\Delta_{G} span H1(S;)H_{1}(S;\mathbb{R}), then g(S)=vg()g(S)=vg(\ell). We make this more precise in the following. The same conclusion holds for the dual graph GG^{*}.

Definition 4.1.

The symplectic rank rks(ΔG)rk_{s}(\Delta_{G}) is the rank of the submodule of H1(S;)H_{1}(S;\mathbb{R}) generated by the summands of ΔG\Delta_{G}.

Theorem 4.2.

Assume that GSG\subset S is a signed embedded graph associated to a checkerboard colorable link S×[0,1]\ell\subset S\times[0,1], and GG^{*} is the dual graph. If either rks(ΔG)rk_{s}(\Delta_{G}) or rks(ΔG)rk_{s}(\Delta_{G^{*}}) is equal to twice the genus of SS, then g(S)=vg()g(S)=vg(\ell).

Example 4.3.

Consider the virtual link k,l,m\ell_{k,l,m} in Figure 12, where k,l,mk,l,m denote the number of positive or negative twists at the indicated site. The link 1,1,1\ell_{1,1,1} corresponds to the graph in Figure 2. The graph of the k,l,m\ell_{k,l,m} is obtained from GG by subdividing edges and appending signs 1-1 if the associated integer is negative. The dual graph Γ{\Gamma}^{*} can be obtained from the graph of Figure 10 by replacing edges with multiple edges and appropriate signs. Hence

ΔG=2(k+l+m)k(x+x1)l(y+y1)m(x1y+xy1).\Delta_{G^{*}}=2(k+l+m)-k(x+x^{-1})-l(y+y^{-1})-m(x^{-1}y+xy^{-1}).

By Theorem 4.2, the virtual genus of k,l,m\ell_{k,l,m} is equal to 1.

Refer to caption
Figure 13: Virtual links k,l,m\ell_{k,l,m}
Example 4.4.

If a diagram D^S\hat{D}\subset S of a virtual link ^\hat{\ell} is not checkerboard colorable, then information about virtual genus of ^\hat{\ell} might be obtained by considering a “doubled diagram” DD that describes a satellite link \ell with companion ^\hat{\ell}. By [14] the virtual genus of \ell is the same as that of ^\hat{\ell}.

Figure 14 (a) is a diagram of a “virtual trefoil.” Figure 14 (b) is a diagram DD of a satellite knot \ell. The reader can check that the Laplacian polynomial ΔG\Delta_{G} of the signed graph GG associated to DD is 4(x+x1)3(y+y1)+2(xy1+x1y)4-(x+x^{-1})-3(y+y^{-1})+2(xy^{-1}+x^{-1}y). Since the symplectic rank rks(ΔG)rk_{s}(\Delta_{G}) is 2, Theorem 4.2 implies the known result that the virtual trefoil has genus 1.

Refer to caption
Figure 14: Virtual trefoil and satellite
Example 4.5.

Consider the 3-component virtual link \ell described by the graph GG in Figure 15(a). It is easy to see that ΔG=0\Delta_{G}=0. Hence Theorem 4.2 gives no information about vg()vg(\ell) in this case. The link can be drawn as in Figure 15 (b).

We give a direct proof that vg()=2vg(\ell)=2. Lift the link diagram to the universal (abelian) cover S~\tilde{S}. There we see countably many unknots in the plane. If \ell has genus 0, then, after Reidemeister moves, we can find an essential simple closed curve in the torus SS that misses the diagram. Perform the Reidemeister moves equivariantly in S~\tilde{S} and consider a single component of the lifted curve. The component separates the plane. However, any two circles of the lifted diagram D~\tilde{D} are part of a chain of consecutively-linked circles. Hence D~\tilde{D} lies on only one side of the separating curve, which is not possible. Hence vg()=1vg(\ell)=1.

Refer to caption
Figure 15: Graph GG and associated virtual link \ell

References

  • [1] H.U. Boden, R. Gaudreau, E. Harper, A.J. Nicas and L. White, Virtual knot groups and almost classical knots, Fund. Math. 238 (2017), no. 2, 101–142.
  • [2] J.S. Carter, D.S. Silver and S.G. Williams, Invariants of links in thickened surfaces, Algebraic and Geometric Topology, 14 (2013), 1377 – 1394.
  • [3] S. Corry and D. Perkinson, Divisors and sandpiles, an introduction to chip-firing, Amer. Math. Soc., Providence RI, 2018.
  • [4] J.S. Carter, M. Elhamdadi, M. Saito and S.G. Williams, Virtual knot invariants from group biquandles and their cocycles, Journal of Knot Theory and its Ramifications, 18 (2009), 957–972.
  • [5] A.M. Duval, C.J.  Klivans, and J.L. Martin, Critical groups of simplicial complexes. Ann. Comb. 17 (2013), no. 1, 53–70.
  • [6] R. Forman, Determinants of Laplacians on graphs, Topology, 32 (1993), 35–46.
  • [7] Y.H. Im, K. Lee, and S.Y. Lee, Signature, nullity and determinant of checkerboard colorable virtual links, J. Knot Theory and its Ramifications, 19 (2010), 1093–1114.
  • [8] L.H. Kauffman, Formal Knot Theory, Enlarged edition, Dover Publications, Mineola, NY, 2006.
  • [9] R. Kenyon, Spanning forests and the vector bundle Laplacian, Annals of Probability 39 (2011), 1983–2017.
  • [10] G. Kuperberg, What is a virtual link?, Alg. Geom. Top. 3 (2003), 587–591.
  • [11] S. Kinoshita and T. Yajima, On the graphs of knots, Osaka J. Math. 9 (1957), 156–163.
  • [12] J.J. Molitierno, Applications of Combinatorial Matrix Theory to Laplacian Matrices of Graphs, CRC Press, Boca Raton, FL, 2012.
  • [13] K. Reidemeister, Knotentheorie, Chelsea Publishing, New York, 1947.
  • [14] D.S. Silver and S.G. Williams, Virtual genus of satellite links, Journal of Knot Theory and its Ramifications, 22 (2013), 1350008 (4 pages).
  • [15] D.S. Silver and S.G. Williams, Knot group presentations of Wirtinger and Dehn, in preparation.
  • [16] D.S. Silver, L. Traldi and S.G. Williams, Goeritz and Seifert matrices from Dehn presentations, Osaka J. Math., in press.
  • [17] T. Yajima and S. Kinoshita, On the graphs of knots, Osaka J. Math. 9 (1957), 155–163.

Department of Mathematics and Statistics,
University of South Alabama
Mobile, AL 36688 USA
Email:
[email protected]
[email protected]