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

Graph Complexity and Link Colorings

Daniel S. Silver    Susan G. Williams Both authors are grateful for the support of the Simons Foundation.
Abstract

The (torsion) complexity of a finite signed graph is defined to be the order of the torsion subgroup of the abelian group presented by its Laplacian matrix. When GG is dd-periodic (i.e., GG has a free d\mathbb{Z}^{d}-action by graph automorphisms with finite quotient) the Mahler measure of its Laplacian polynomial is the growth rate of the complexity of finite quotients of GG. Any 1-periodic plane graph GG determines a link C\ell\cup C with unknotted component CC. In this case the Laplacian polynomial of GG is related to the Alexander polynomial of the link. Lehmer’s question, an open question about the roots of monic integral polynomials, is equivalent to a question about the complexity growth of signed 1-periodic graphs that are not necessarily embedded.

MSC: 05C10, 37B10, 57M25, 82B20

1 Motivation

Consider a locally finite graph without multiple edges or isolated vertices. Let pp be an odd prime. We can attempt to form a “pp-coloring” of the graph, assigning elements (“colors”) of /p\mathbb{Z}/p to the vertices in such a way that the color of any vertex is the average of the colors of the adjacent vertices. Obviously, the graph can be colored trivially or “monochromatically,” that is, with any single color. But nontrivial pp-colorings might exist as well. It is clear that the set of all pp-colorings of the graph is a vector space under vertex-wise addition and scalar multiplication.

Now consider a plane diagram of a knot or link. (As a knot is a link with only one component, henceforth we will use the more general term.) We can try to form a “pp-coloring,” assigning colors to the arcs in such a way that the color of any overcrossing arc is the average of the colors of its two undercrossing arcs (see Section 4). Again we can color trivially, but nontrivial pp-colorings can exist too. The set of the all pp-colorings of the diagram is a vector space under arc-wise addition and scalar multiplication.

There is a strong relationship between pp-colorings of plane graphs and pp-colorings of link diagrams [27, 33]. We review the relationship in Section 3. By substituting the continuous palette 𝕋=/{\mathbb{T}}=\mathbb{R}/\mathbb{Z}, which contains /p\mathbb{Z}/p as a subgroup, and replacing the plane with a compact orientable surface SS, we enter the world of algebraic dynamical systems [28]. When SS is an annulus, we obtain a version of Lehmer’s unanswered question about roots of monic integral polynomials, which we formulate in terms of signed graphs.

Acknowledgements. It is the authors’ pleasure to thank Abhijit Champanerkar, Eriko Hironaka, Matilde Lalin and Chris Smythe for helpful comments and suggestions.

2 Coloring and complexity of finite graphs

Let GG be a graph, not necessarily planar, with vertex set V(G)={v1,,vn}V(G)=\{v_{1},\ldots,v_{n}\} and edge set E(G)={e1,,em}E(G)=\{e_{1},\ldots,e_{m}\}. The graph is allowed to have loops and multiple edges, but no isolated vertices. Each edge eE(G)e\in E(G) is labeled with a sign σe=±1\sigma_{e}=\pm 1. All graphs that we consider are assumed to have signed edges. The graph is unsigned if every σe=+1\sigma_{e}=+1.

The signed adjacency matrix of GG is the n×nn\times n matrix A=(ai,j)A=(a_{i,j}) such that ai,ja_{i,j} is the sum of the signs of all edges between viv_{i} and vjv_{j}, with loops counted twice. Define the signed degree matrix δ=(δi,j){\delta}=({\delta}_{i,j}) to be the n×nn\times n diagonal matrix with δi,i{\delta}_{i,i} equal to the sum of signs of edges incident on viv_{i}. Again, loops contribute twice.

An integer matrix presents an abelian group in which columns (resp. rows) correspond to the generators (resp. relations) of the group. Equivalently, the group is the cokernel of the matrix when regarded as a linear transformation of free abelian groups.

Definition 2.1.

The Laplacian matrix of a finite graph GG is LG=δAL_{G}={\delta}-A. The abelian group presented by LGL_{G} is the Laplacian group of GG, denoted by G{\cal L}_{G}. The (torsion) complexity κG{\kappa}_{G} is the order of the torsion subgroup TGT{\cal L}_{G}.

We note that in computing the integer matrix LGL_{G} we may ignore loops, since they contribute equally to AA and δ{\delta}. However, this will not be the case for the extended definition we make in Section 5.

When we regard the entries of LGL_{G} modulo pp, the rows represent the relations needed to pp-color the graph. (Here we are extending the notion of pp-coloring to signed graphs.) Hence the space of pp-colorings of GG is isomorphic to the null space of LGL_{G}. Nontrivial pp-colorings exist precisely when its dimension is greater than 1. Note that the abelian group of pp-colorings of GG is isomorphic to G/p{\cal L}_{G}\otimes\mathbb{Z}/p.

Returning to integer coefficients, the nullity of the Laplacian matrix LGL_{G} is equal to 1 whenever GG is connected and unsigned (see, for example, Lemma 13.1.1 of [12]). In this case the Laplacian group G{\cal L}_{G} decomposes as the direct sum of \mathbb{Z} and the torsion subgroup TGT{\cal L}_{G}. The Matrix Tree Theorem (c.f. [35]) implies that κG=|TG|{\kappa}_{G}=|T{\cal L}_{G}| is equal to the number of spanning trees of GG.

More generally, we define tree complexity τG\tau_{G} of a connected graph GG by

τG=|TeE(T)σe|,\tau_{G}=\bigg{|}\sum_{T}\prod_{e\in E(T)}\sigma_{e}\bigg{|},

where the summation is taken over all spanning trees of GG. If GG not connected, then we define τG\tau_{G} to be the product of the tree complexities of its connected components. Again by [35], we have τG=κG\tau_{G}={\kappa}_{G} if and only if τG\tau_{G} is nonzero; for connected GG this common value is equal to the absolute value of any (n1)×(n1)(n-1)\times(n-1) principal minor of LGL_{G}. However, the following example shows that τG\tau_{G} can vanish, whereas κG{\kappa}_{G} is positive by definition.

Example 2.2.

Consider the connected graph GG in Figure 1. Unlabeled edges here and throughout will be assumed to have sign +1+1. The Laplacian matrix LGL_{G} is square of size 88. A routine calculation shows that any principal 7×77\times 7 minor of LGL_{G} vanishes, and hence τG=0\tau_{G}=0. However, the absolute value of the greatest common divisors of the 6×66\times 6 minors of LGL_{G} is 9, and so κG=9{\kappa}_{G}=9.

We can go further by computing the Smith Normal Form of LGL_{G}. We then see that G2/3/3{\cal L}_{G}\cong\mathbb{Z}^{2}\oplus\mathbb{Z}/3\oplus\mathbb{Z}/3. The vector space of pp-colorings has dimension 2 for all primes p3p\neq 3, while the space of 33-colorings of GG has dimension 4.

Refer to caption
Figure 1: Graph GG with τG=0\tau_{G}=0 and κG=9{\kappa}_{G}=9

3 Link diagrams and Tait graphs

A link \ell in 3\mathbb{R}^{3} (or equivalently 𝕊3{\mathbb{S}}^{3}) is a set of smoothly embedded, pairwise disjoint simple closed curves. Any link can be described by a diagram DD, a 4-valent plane graph with a hidden-line device in a neighborhood of each vertex indicating how one strand of the link passes over another.

We will make use of a few other related terms. An arc of DD is a maximal connected subset. Following [14], we refer to the underlying graph of DD as a universe of \ell, denoted here by |D||D|. Finally, a region of DD is a connected component of 2|D|\mathbb{R}^{2}\setminus|D|.

As usual, isotopic links are regarded as the same. It is well known that two links are isotopic if and only if a diagram of one can be transformed into a diagram of the other by a finite sequence of local “Reidemeister moves” as in Figure 2 (see [5] or [18] for details). Consequently, any quantity that is determined by a diagram and is unchanged by Reidemeister moves is a link invariant.

Refer to caption
Figure 2: Reidemeister moves

We can obtain a plane graph GG from a link diagram DD by the following familiar procedure. First we checkerboard shade DD, shading some of the regions so that every edge meets a single shaded region. There are two checkerboard shadings of DD, but for the sake of definiteness we will choose the one in which the unbounded region is unshaded. We construct a graph GG with a vertex in each shaded region of DD and an edge through each crossing, joining the vertices of the regions on both sides. The sign of the edge is determined by the type of crossing, as in Figure 3. Such a graph is often called a “Tait graph,” in honor of Peter Guthrie Tait, a 19th century pioneer of knot theory.

Refer to caption
Figure 3: Constructing a graph from a link diagram

For some link diagrams it can happen that the Tait graph has isolated vertices. We avoid this by always assuming that |D||D| is connected, a condition that we can assume without loss of generality by applying Reidemeister moves to DD.

For a plane graph GG we can reverse the above procedure to obtain a link digram DD. For this we use the medial construction, replacing each edge of GG by a pair of arc segments that run parallel to the edge except at the middle, where they cross, as in Figure 4. We join the segments near vertices in the obvious way without creating any additional crossings.

Figure 7 shows the diagram of a 3-component link \ell obtained from the graph of Example 2.2. According to [8] the link was introduced by J. Milnor. Original interest in the link came from the fact that it was the first non-alternating boundary link discovered with zero Alexander polynomial, a fact that we will not use here. We will return to the link in the next section.

Refer to caption
Figure 4: Constructing a link diagram from a plane graph

4 Coloring link diagrams

Assume that DD is a diagram of a link \ell. Let pp be a prime. A (Fox) pp-coloring of DD is an assignment of elements of /p\mathbb{Z}/p, called colors, to the arcs of DD such that twice the color of any overcrossing arc is equal to the sum of the colors of its undercrossing arcs, as in Figure 5. A pp-coloring is trivial if all assigned colors are the same.

Refer to caption
Figure 5: Fox coloring relation

We denote the vector space of pp-colorings of DD by Colp(D){\rm Col}_{p}(D) and refer to it as the pp-coloring space of the diagram. It is a vector space over /p\mathbb{Z}/p, with addition and scalar multiplication performed arc-wise. A routine exercise shows that diagrams differing by a Reidemeister move have isomorphic pp-coloring spaces. Hence Colp(D){\rm Col}_{p}(D) is an invariant of the link \ell, and so we write Colp(){\rm Col}_{p}(\ell) and speak of the vector space as the pp-coloring space of the link. (For more information about this popular construction of knot theory, there are many excellent expositions such as [20].)

Fox envisioned pp-colorings as homomorphisms from the link group π=π1(3)\pi_{\ell}=\pi_{1}(\mathbb{R}^{3}\setminus\ell) onto the dihedral group

D2p=τ,ατ2=1,αp=1,ατ=τα1.D_{2p}={\langle}\tau,\alpha\mid\tau^{2}=1,\alpha^{p}=1,\alpha\tau=\tau\alpha^{-1}{\rangle}.

He described the correspondence using the Wirtinger presentation of π\pi_{\ell}, in which generators (resp. relations) are identified with arcs (resp. crossings) of a diagram DD. Given a pp-coloring of DD we obtain a homomorphism from π\pi_{\ell} to D2pD_{2p} by sending a Wirtinger generator of an arc colored by k/pk\in\mathbb{Z}/p to the element ταk\tau\alpha^{k} of D2pD_{2p}.

If instead of the Wirtinger presentation, we use the Dehn presentation of π\pi_{\ell}, a presentation in which generators (resp. relations) are identified with bounded regions (resp. crossings), then a pp-coloring becomes an assignment of colors to the bounded regions of DD. We assign 0 to the unbounded region. The condition at each crossing that corresponds to the Fox pp-coloring condition appears in Figure 6. Details can be found in [32]. We will call such an assignment of colors to the regions a Dehn pp-coloring of the diagram. The collection of all Dehn pp-colorings of a diagram is vector space under region-wise addition and scalar multiplication, isomorphic to the space of Fox pp-colorings.

Refer to caption
Figure 6: Dehn pp-coloring condition

Assume that DD is a link diagram arising from a plane graph GG by the medial construction. Any pp-coloring of GG determines Dehn and Fox pp-colorings of DD. To see this, first assign the colors of the vertices of GG to the associated shaded regions of the link diagram. Then use the Dehn coloring relations to determine uniquely the colors of the unshaded regions. This is possible since the unbounded region is already labeled (with 0). Uniqueness follows from the observation that if we determine the color of some unshaded region and then follow a simple closed path around a vertex, determining the colors of successive unshaded regions along the way, then when we return to the initial unshaded region, the Laplace relation forces us to arrive at the same color with which we began. (More details can be found in [32].) Finally, assign to each arc of DD the sum of the colors of the regions on both sides. It is easy to verify that we obtain in this way is well defined Fox pp-coloring of the diagram.

Conversely, any Fox pp-coloring of a link diagram determines a pp-coloring of the associated Tait graph. The Dehn color of any region is the sum of the colors of the arcs that we cross traveling along any path to the unbounded region. The graph coloring is simply a restriction to the shaded regions.

The two processes above are inverses of each other. Hence the vector spaces of pp-colorings of GG and DD are isomorphic.

As an example, consider the diagram of Milnor’s boundary link as in appears in Figure 7. The Fox 3-coloring displayed corresponds to the 3-coloring of the graph G in Figure 1.

Refer to caption
Figure 7: 33-colored diagram of Milnor’s boundary link

Every finite cyclic group is contained in the compact abelian “circle group” 𝕋=/{\mathbb{T}}=\mathbb{R}/\mathbb{Z} as a subgroup. If we replace /p\mathbb{Z}/p by 𝕋{\mathbb{T}}, then the coloring vector spaces for graphs and link diagrams become compact abelian groups. (This extension of Fox pp-coloring was introduced in [29].) The Laplace group G{\cal L}_{G} tensored with 𝕋{\mathbb{T}} consists of |TG||T{\cal L}_{G}| tori each of dimension equal to the nullity of LGL_{G}. A similar description of the circle-coloring group Col𝕋(){\rm Col}_{\mathbb{T}}(\ell) applies. In the following sections we will explore this idea for infinite graphs and links with symmetries. The result is a rich structure that brings algebraic dynamics into our story.

5 Periodic graphs and Laplacian modules

A (signed) graph GG is dd-periodic if it admits a cofinite free d\mathbb{Z}^{d}-action by automorphisms that preserves the signs of edges. By cofinite we mean the quotient graph G¯\overline{G} is finite, while an action is free if the stabilizer of any edge or vertex is trivial. Such a graph GG is locally finite, and finite if and only if d=0d=0.

We regard d,d1,\mathbb{Z}^{d},d\geq 1, as the multiplicative abelian group freely generated by x1,,xdx_{1},\ldots,x_{d}. We denote the Laurent polynomial ring [d]=[x1±1,,xd±1]\mathbb{Z}[\mathbb{Z}^{d}]=\mathbb{Z}[x_{1}^{\pm 1},\ldots,x_{d}^{\pm 1}] by d{\cal R}_{d}. As an abelian group d{\cal R}_{d} is generated freely by monomials x𝐧=x1n1xdndx^{\bf n}=x_{1}^{n_{1}}\ldots x_{d}^{n_{d}}, where 𝐧=(n1,,nd)d{\bf n}=(n_{1},\cdots,n_{d})\in\mathbb{Z}^{d}. We represent (0,,0)(0,\ldots,0) by 𝟎{\bf 0}. For notational convenience, when d=1d=1, we replace x1x_{1} by xx.

The vertex set V(G)V(G) and the edge set E(G)E(G) consist of finitely many vertex orbits {v1,𝐧𝐧d},,{vn,𝐧𝐧d}\{v_{1,{\bf n}}\mid{\bf n}\in\mathbb{Z}^{d}\},\ldots,\{v_{n,{\bf n}}\mid{\bf n}\in\mathbb{Z}^{d}\} and signed edge orbits {e1,𝐧𝐧d},,{em,𝐧𝐧d}\{e_{1,{\bf n}}\mid{\bf n}\in\mathbb{Z}^{d}\},\ldots,\{e_{m,{\bf n}}\mid{\bf n}\in\mathbb{Z}^{d}\}, respectively. The d\mathbb{Z}^{d}-action is determined by

x𝐧vi,𝐧=vi,𝐧+𝐧,x𝐧ej,𝐧=ej,𝐧+𝐧,x^{{\bf n}^{\prime}}\cdot v_{i,{\bf n}}=v_{i,{\bf n}+{\bf n}^{\prime}},\quad\quad x^{{\bf n}^{\prime}}\cdot e_{j,{\bf n}}=e_{j,{\bf n}+{\bf n}^{\prime}}, (5.1)

where 1in, 1jm1\leq i\leq n,\ 1\leq j\leq m and 𝐧,𝐧d{\bf n},{\bf n}^{\prime}\in\mathbb{Z}^{d}. (When GG is embedded in some Euclidean space with d\mathbb{Z}^{d} acting by translation, it is usually called a lattice graph. Such graphs arise frequently in physics, for example in studying crystal structures.)

When d>1d>1 we can think of GG as covering a finite graph G¯\overline{G} in the dd-torus 𝕋d=d/d{\mathbb{T}}^{d}=\mathbb{R}^{d}/\mathbb{Z}^{d}. When d=1d=1, GG covers a finite graph G¯\overline{G} in the annulus 𝔸=I×𝕊1\mathbb{A}=I\times{\mathbb{S}}^{1}. In either case the cardinality |V(G¯)||V(\overline{G})| is equal to the number nn of vertex orbits of GG, while |E(G¯)||E(\overline{G})| is the number mm of edge orbits. The projection map is given by vi,𝐧viv_{i,{\bf n}}\mapsto v_{i} and ej,𝐧eje_{j,{\bf n}}\mapsto e_{j}.

The Laplacian matrix of a dd-periodic graph GG is defined to be the n×nn\times n-matrix LG=δAL_{G}={\delta}-A, where now A=(ai,j)A=(a_{i,j}) is the adjacency d{\cal R}_{d}-matrix with each entry ai,ja_{i,j} equal to the sum of monomials σex𝐧\sigma_{e}x^{\bf n} for each edge eE(G¯)e\in E(\overline{G}) between vi,𝟎v_{i,\bf{0}} and vj,𝐧v_{j,{\bf n}}. Loops contribute two diagonal terms that are inverse monomials. The signed degree matrix δ=(δi,j){\delta}=({\delta}_{i,j}) is defined as in Section 2.

The matrix LGL_{G} presents a finitely generated d{\cal R}_{d}-module, the Laplacian module of GG, denoted by G{\cal L}_{G}. The Laplacian (determinant) polynomial ΔG\Delta_{G} is the determinant of LGL_{G}. When d=0d=0, d={\cal R}_{d}=\mathbb{Z} and these definitions reduce to the ones in Section 2. Examples appear below; additional examples can be found in [16, 32]. The reader should be aware that in graph theory literature the term “Laplacian polynomial” is often used for the characteristic polynomial of the integral Laplacian matrix.

6 Coloring periodic graphs

Let GG be a dd-periodic graph. The collection of all 𝕋{\mathbb{T}}-colorings of GG is the Pontryagin dual group ^G=Hom(G,𝕋)\widehat{\cal L}_{G}={\rm Hom}({\cal L}_{G},{\mathbb{T}}). Elements are functions f:V(G)𝕋f:V(G)\to{\mathbb{T}} that assign to each vertex vi,𝐧V(G)v_{i,{\bf n}}\in V(G) a color f(vi,𝐧)𝕋f(v_{i,{\bf n}})\in{\mathbb{T}} such that the Laplacian condition (corresponding to the iith row of LGL_{G}) is satisfied:

δi,if(vi,𝐧)=eσef(vi,𝐧),{\delta}_{i,i}f(v_{i,{\bf n}})=\sum_{e}\sigma_{e}f(v_{i^{\prime},{\bf n}^{\prime}}), (6.1)

where δi,i{\delta}_{i,i} is the signed degree of vi,𝐧v_{i,{\bf n}}, and the summation is taken over all edges ee that connect vi,𝐧v_{i,{\bf n}} with some vi,𝐧V(G)v_{i^{\prime},{\bf n}^{\prime}}\in V(G), loops contributing twice.

We regard G{\cal L}_{G} with the discrete topology. Endowed with the compact-open topology, ^G\widehat{\cal L}_{G} is a compact space (see, for example, Section 2 of [19]). It admits a d\mathbb{Z}^{d}-action by automorphisms. Such an action is a homomorphism 𝐬:𝐧𝐬𝐧{\bf s}:{\bf n}\mapsto{\bf s}_{{\bf n}} from d\mathbb{Z}^{d} to the automorphism group of ^G\widehat{\cal L}_{G}.

We denote ^G\widehat{\cal L}_{G} with its d\mathbb{Z}^{d}-action by Col𝕋,d(G){\rm Col}_{{\mathbb{T}},\mathbb{Z}^{d}}(G). It is an example of a dynamical system known as a d\mathbb{Z}^{d}-shift. By the Pontryagin Duality Theorem we recover the Laplace module by taking the dual of Col𝕋,d(G){\rm Col}_{{\mathbb{T}},\mathbb{Z}^{d}}(G). We will say more about the dynamical properties of ^G\widehat{\cal L}_{G} in Section 9.

In the case of a finite plane graph and its associated link diagram (d=0d=0), their isomorphic groups of 𝕋{\mathbb{T}}-colorings are dual respectively to the Laplacian group of the graph and the abelian core group of the link. The latter group is generated by the arcs of the diagram with relations given by the Fox coloring condition of Figure 5; it is well known to be isomorphic to the direct sum of the first homology group of the 2-fold branched cover of the link and an infinite cyclic group. (For more about such dynamical systems see [26, 19] or [29]. Information about the core group can be found in [30].)

7 Computing the Laplacian polynomial

A cycle-rooted spanning forest (CRSF) of G¯\overline{G} is a subgraph of G¯\overline{G} containing all of V¯\overline{V} such that each connected component has exactly as many vertices as edges and therefore has a unique cycle. The connection ϕ\phi of an oriented cycle is its homology class in H1(𝕋d;)dH_{1}({\mathbb{T}}^{d};\mathbb{Z})\cong\mathbb{Z}^{d}. See [15] for details.

The following is a consequence of the main theorem of [11]. It is made explicit in Theorem 5.2 of [15].

Theorem 7.1.

[15] Let GG be a dd-periodic graph. Its Laplacian polynomial has the form

ΔG=F¯e¯E(F¯)σe¯CyclesofF¯(2ϕϕ1),\Delta_{G}=\sum_{\overline{F}}\ \prod_{\overline{e}\in E(\overline{F})}\sigma_{\overline{e}}\prod_{\rm Cycles\ of\ \overline{F}}(2-\phi-\phi^{-1}), (7.1)

where the sum is over all cycle-rooted spanning forests F¯\overline{F} of G¯\overline{G}, and ϕ,ϕ1\phi,\phi^{-1} are the connections of the two orientations of the cycle.

A dd-periodic graph need not be connected. In fact, it can have countably many connected components. Nevertheless, the number of d\mathbb{Z}^{d}-orbits of components, henceforth called component orbits, is necessarily finite.

Proposition 7.2.

If GG is a dd-periodic graph with component orbits G1,,GtG_{1},\ldots,G_{t}, then ΔG=ΔG1ΔGt\Delta_{G}=\Delta_{G_{1}}\cdots\Delta_{G_{t}}.

Proof.

After suitable relabeling, the Laplacian matrix for GG is a block diagonal matrix with diagonal blocks equal to the Laplacian matrices for G1,,GtG_{1},\ldots,G_{t}. The result follows immediately. Alternatively, it can be deduced from Theorem 7.1. ∎

Proposition 7.3.

Let GG a dd-periodic graph. If GG contains a finite component, then its Laplacian polynomial ΔG\Delta_{G} is identically zero. The converse statement is true if GG is unsigned.

Proof.

If GG contains a finite component, then some component orbit GiG_{i} consists of finite components. We have ΔGi=0\Delta_{G_{i}}=0 by Theorem 7.1, since all cycles of Gi¯\overline{G_{i}} represent trivial homology classes and hence have vanishing connection. By Proposition 7.2, ΔG\Delta_{G} is identically zero.

Conversely, assume GG is unsigned and every component is infinite. Each component of G¯\overline{G} must contain a nontrivial cycle. We can extend this collection of cycles to a cycle rooted spanning forest FF with no additional cycles. The corresponding summand in Theorem 7.1 has positive constant coefficient. Since every summand has nonnegative constant coefficient, ΔG\Delta_{G} is not identically zero.

8 Plane 1-periodic graphs and links in solid tori

When a plane graph GG is 1- or 2-periodic, the medial construction in Section 4 produces a diagram DD of an infinite link \ell. It has a finite quotient diagram D¯\overline{D} modulo the \mathbb{Z}- or 2\mathbb{Z}^{2}-action induced by the action on GG.

For d=1d=1 we regard D¯\overline{D} in an annulus 𝔸\mathbb{A}. It describes a link ¯=¯1¯μ\overline{\ell}=\overline{\ell}_{1}\cup\cdots\cup\overline{\ell}_{\mu} in a solid unknotted torus VV. The complement (intV)¯({\rm int}\ V)\setminus\overline{\ell} is homeomorphic to 𝕊3^{\mathbb{S}}^{3}\setminus\hat{\ell}, where ^=¯C\hat{\ell}=\overline{\ell}\cup C is the link formed by the union of ¯\overline{\ell} with a meridian CC of VV. The meridian acquires an orientation induced by the infinite cyclic action on DD. It is easy to see that every link with an unknotted component that has even linking number with the rest of the link arises in this way.

The following result relates the Laplacian polynomial ΔG\Delta_{G} to the Alexander polynomial Δ^\Delta_{\hat{\ell}}.

Theorem 8.1.

Let GG be a plane 1-periodic graph and ^\hat{\ell} the encircled link ¯C\overline{\ell}\cup C. Then

ΔG(x)=(x1)Δ^(1,,1,x),\Delta_{G}(x)\ {\buildrel\cdot\over{=}}\ (x-1)\ \Delta_{\hat{\ell}}(-1,\ldots,-1,x),

where ={\buildrel\cdot\over{=}} indicates equality up to multiplication by units in [x±1]\mathbb{Z}[x^{\pm 1}].

Proof.

The argument at the end of section 6 shows that the Laplacian group of a finite plane graph is isomorphic to the abelian core group of the associated link. The same argument can be applied to any 1-periodic graph GG and its associated link \ell. (Either of the two unbounded regions of the link diagram can be used as the base region as we pass from from vertex colorings to Fox colorings via Dehn colorings.)

Let DD be the diagram of \ell obtained from GG by the medial construction. We regard DD as lying in a strip I×2I\times\mathbb{R}\subset\mathbb{R}^{2}, with R=I×[0,1]R=I\times[0,1] a fundamental domain for the \mathbb{Z}-action. Then DD meets RR in a tangle diagram D0D_{0}. We label the arcs of D0D_{0} meeting the “top,” R×{1}R\times\{1\}, by a1,,ana_{1},\ldots,a_{n} and those meeting the “bottom,” R×{0}R\times\{0\}, by a1,,ana^{\prime}_{1},\ldots,a^{\prime}_{n}. (It can happen that some aia_{i} and aja^{\prime}_{j} are identical.) Let b,c,b,c,\ldots be labels for the remaining arcs of D0D_{0}.

Define BB to be the quotient of the free abelian group on a1,,an,a1,,an,b,ca_{1},\ldots,a_{n},a^{\prime}_{1},\ldots,a^{\prime}_{n},b,c\ldots by the Fox relations (Figure 5) of the crossings in D0D_{0}. Let UU be the free abelian on u1,,unu_{1},\ldots,u_{n}, and f:UBf:U\to B (resp. g:UBg:U\to B) the homomorphisms mapping each uiu_{i} to aia_{i} (resp. uiu_{i} to aia^{\prime}_{i}). The Laplacian module has the form

UBUBU\cdots\oplus_{U}B\oplus_{U}B\oplus_{U}\cdots (8.1)

with identical amalgamations BgUfB.B\ {\buildrel g\over{\leftarrow}}\ U\ {\buildrel f\over{\rightarrow}}\ B. The module action of xx merely shifts each summand BB one place to the right. Thus the Laplacian module G{\cal L}_{G} is the cokernel of the square matrix AA with columns corresponding to the arcs of D0D_{0} and rows recording the Fox relations as well as the relations xa1=a1,,xan=anxa^{\prime}_{1}=a_{1},\ldots,xa^{\prime}_{n}=a_{n}.

We claim that the matrix AA is an Alexander matrix of the link ¯C\bar{\ell}\cup C with xx corresponding to a meridian mm of CC while the meridianal variables of \ell are set equal to 1-1 (cf. [5]). To see this, consider Figure 8. The Alexander matrix has columns corresponding to a1,,an,a1,,an,b,ca_{1},\ldots,a_{n},a^{\prime}_{1},\ldots,a^{\prime}_{n},b,c\ldots and the meridian mm. There are nn rows corresponding to the crossing relations in D0D_{0}, and additional rows for relations m+xai=aimm+xa^{\prime}_{i}=a_{i}-m, where i=1,,ni=1,\ldots,n. (The relations, which can be determined by Fox calculus or by considering the appropriate infinite cyclic cover, are unaffected by the directions of the arcs of D0D_{0}. The n1n-1 arcs in the back of CC correspond to generators defined by the relations that arise from n1n-1 of the nn crossings in the back; the last relation is redundant and can be ignored. Hence the extra generators and relations can be disregarded.) We delete the column corresponding to mm in order to obtain an Alexander matrix. The rows that we have added now correspond to the relations xai=aixa^{\prime}_{i}=a_{i}. The result is the matrix AA. The Alexander polynomial of ¯C\bar{\ell}\cup C is the determinant of AA divided by x1x-1.

Refer to caption
Figure 8: Detail of ¯C\bar{\ell}\cup C

Proposition 8.2.

Let GG be a plane 1-periodic graph, DD its associated link diagram, and \ell the associated link. The Laplacian polynomial ΔG\Delta_{G} is nonzero if and only if the Laplacian module G{\cal L}_{G} is a torsion module.

Proof.

The proposition follows from basic facts of commutative algebra. Since G{\cal L}_{G} has a square matrix presentation, ΔG\Delta_{G} generates annihilator ideal of G{\cal L}_{G}. ∎

Proposition 8.3.

Let GG be a plane 1-periodic graph, DD its associated link diagram, and \ell the associated link. The following are equivalent.

  1. 1.

    The link \ell has closed components.

  2. 2.

    The group of 2-colorings of DD is infinite.

  3. 3.

    The Laplacian polynomial ΔG{\Delta}_{G} reduced modulo 2 is identically zero.

Proof.

Any 2-coloring of DD assigns a single color to every arc corresponding to a component of \ell. If the link has no closed components, then it has only finitely many components, and conversely. The equivalence of the first two statements follows.

Regard G{\cal L}_{G} as an abelian group. The vector space of 2-colorings of DD is isomorphic to G/2{\cal L}_{G}\otimes\mathbb{Z}/2. Its dimension is the degree of the mod 2 reduction of ΔG\Delta_{G}, provided the reduced polynomial is nonzero; otherwise the dimension is infinite. However, the dimension is also equal to the number of components of \ell. Hence the first and third statements are equivalent. ∎

The next proposition characterizes the leading coefficient of the Laplacian polynomial of a plane 1-periodic graph in terms of 𝕋{\mathbb{T}}-colorings of its associated link. We will use the notation in the proof of Theorem 8.1. Note that ΔG(x)=ΔG(x1)\Delta_{G}(x)=\Delta_{G}(x^{-1}), by Theorem 7.1.

Proposition 8.4.

Let GG be a plane 1-periodic graph, DD its associated link diagram, and D0RD_{0}\subset R a tangle diagram representing a fundamental region of DD. Suppose ΔG\Delta_{G} is nonzero. If we assign 0𝕋0\in{\mathbb{T}} to the arcs at the top of D0D_{0}, then the number of extensions to 𝕋{\mathbb{T}}-colorings of D0D_{0} is equal to the absolute value of the leading coefficient of ΔG\Delta_{G}.

Proof.

The 𝕋{\mathbb{T}}-colorings of DD that assign 0 to arcs labeled a1,,ana_{1},\ldots,a_{n} are the elements of the dual group of the quotient B/f(U)B/f(U). The group B/f(U)B/f(U) is the cokernel of the specialized Alexander matrix AA constructed in the proof of Theorem 8.1 with variable xx set equal to 0. The determinant of AA is the order of B/f(U)B/f(U) as well as the absolute values of both the trailing and leading coefficients of ΔG\Delta_{G}. ∎

Proposition 8.5.

Let GG be a plane 1-periodic graph, DD its associated link diagram, and D0RD_{0}\subset R a tangle diagram representing a fundamental region of DD. Suppose that the coefficients of ΔG\Delta_{G} are coprime. Then any two distinct 𝕋{\mathbb{T}}-colorings of D0D_{0} with the same color assignments of the top arcs have different color assignments of the bottom arcs.

Proof.

We prove the proposition by contradiction. Assume that there exist two 𝕋{\mathbb{T}}-colorings of D0D_{0} with the same color assignments to the top arcs and also the same assignments to the bottom. We subtract to get a nontrivial 𝕋{\mathbb{T}}-coloring with all arcs on both top and bottom colored trivially. By duality, the group quotient B¯=B/(f(U)+g(U))\overline{B}=B/(f(U)+g(U)) must be nonzero. Consider the quotient module ¯G\overline{\cal L}_{G} of G{\cal L}_{G} described by (8.1) with all elements of xif(U)x^{i}f(U) and xig(U)x^{i}g(U) set equal to 0; it is a direct sum of countably many copies of B¯\overline{B} with the module action of xx given by translation. An integral matrix presenting B¯\overline{B} as an abelian group also presents ¯G\overline{\cal L}_{G} as module over 1=[x,x1]{\cal R}_{1}=\mathbb{Z}[x,x^{-1}]. The group must be torsion since G{\cal L}_{G} is, and its 0th-characteristic polynomial Δ¯G\overline{\Delta}_{G} must divide ΔG\Delta_{G}. Since Δ¯G\overline{\Delta}_{G} is a nonzero constant and the coefficients of ΔG\Delta_{G} are coprime, Δ¯G=±1\overline{\Delta}_{G}=\pm 1. Hence B¯\overline{B} is a trivial group, a contradiction. ∎

The following corollary follows form Propositions 8.4 and 8.5.

Corollary 8.6.

Assume that GG is a plane 1-periodic graph, DD its associated link diagram, and D0RD_{0}\subset R a tangle diagram representing a fundamental region of DD. Assume that ΔG\Delta_{G} is monic. Given any color assignment to the top arcs of D0D_{0} that extends to a 𝕋{\mathbb{T}}-coloring of D0D_{0}, the colors of the bottom arcs are uniquely determined.

Example 8.7.

Consider the 1-periodic graph GG and its associated tangle diagram D0D_{0} in Figure 9. It is easy to see that ΔG(x)=3x+63x1\Delta_{G}(x)=-3x+6-3x^{-1}, which has leading coefficient 3-3. The group BB of the associated tangle has generators a0,a1,b0,b1,ca_{0},a_{1},b_{0},b_{1},c and relations 2b0=a0+c,2c=b0+a1,2a1=c+b12b_{0}=a_{0}+c,2c=b_{0}+a_{1},2a_{1}=c+b_{1}.

Refer to caption
Figure 9: 1-peridodic graph GG and associated tangle diagram D0D_{0}

Since the subgroup f(U)f(U) is generated by a0,a1a_{0},a_{1}, the quotient B/f(U)B/f(U) is generated by b0,b1,cb_{0},b_{1},c, with relations 2b0=c,2c=b0,0=c+b12b_{0}=c,2c=b_{0},0=c+b_{1}. If we choose a color γ𝕋\gamma\in{\mathbb{T}} for the arc of D0D_{0} labeled cc, then the bottom arcs, labeled b0,b1b_{0},b_{1}, must receive colors 2γ,γ2\gamma,-\gamma, respectively. Moreover, the assignment is a 𝕋{\mathbb{T}}-coloring of D0D_{0} provided that 3γ=03\gamma=0. Hence γ=0,1/3,2/3\gamma=0,1/3,2/3 (mod 1), and there are exactly three 𝕋{\mathbb{T}}-colorings of D0D_{0} with the top arcs colored 0, as expected from Proposition 8.4. The three 𝕋{\mathbb{T}}-colorings appear in Figure 10.

Refer to caption
Figure 10: 𝕋{\mathbb{T}}-colorings with top arcs colored trivially
Example 8.8.

The closure ¯\overline{\ell} of any 2n2n-braid arises from graph G¯\overline{G} embedded in the annulus 𝔸\mathbb{A}, since we can checkerboard shade a diagram of ¯\overline{\ell} in 𝔸\mathbb{A} so that the border regions are unshaded.

The graph G¯\overline{G} lifts to a 1-periodic graph GG in the plane. Consider its Laplacian polynomial ΔG\Delta_{G}. We recall that the Burau representation associates to each generator σi\sigma_{i} of the 2n2n-braid group a block diagonal matrix

Ii1(1tt10)I2ni1,I_{i-1}\oplus\begin{pmatrix}1-t&t\\ 1&0\end{pmatrix}\oplus I_{2n-i-1},

where IkI_{k} denotes the k×kk\times k identity matrix. Setting t=1t=-1 produces presentation matrix for the group BB in the proof of Theorem 8.1. The Laplacian polynomial of GG is the characteristic polynomial of the Burau matrix of the braid with t=1t=-1.

9 Complexity growth of periodic graphs

When GG is a dd-periodic graph with quotient G¯\overline{G}, we can consider the intermediate covering graphs GΛG_{\Lambda} in d/Λ\mathbb{R}^{d}/{\Lambda}, where Λd{\Lambda}\subset\mathbb{Z}^{d} is a subgroup of Λ{\Lambda} having finite index. In this section we see that the growth of the torsion complexity κGΛ{\kappa}_{G_{\Lambda}} as the index of Λ{\Lambda} goes to infinity is determined by the Mahler measure of the Laplacian polynomial ΔG\Delta_{G}.

We begin by reviewing the Mahler measure of polynomials.

Definition 9.1.

The Mahler measure of a nonzero polynomial f(x1,,xd)df(x_{1},\ldots,x_{d})\in{\cal R}_{d} is

M(f)=exp0101log|f(e2πiθ1,,e2πiθd)|dθ1dθd.M(f)=\exp\int_{0}^{1}\ldots\int_{0}^{1}\log|f(e^{2\pi i\theta_{1}},\ldots,e^{2\pi i\theta_{d}})|d\theta_{1}\cdots d\theta_{d}.
Remark 9.2.

(1) The integral in Definition 9.1 can be singular, but nevertheless it converges. (See [10] for two different proofs.) If u1,,udu_{1},\ldots,u_{d} is another basis for d\mathbb{Z}^{d}, then f(u1,,ud)f(u_{1},\ldots,u_{d}) has the same logarithmic Mahler measure as f(x1,,xd)f(x_{1},\ldots,x_{d}).

(2) If f,gdf,g\in{\cal R}_{d}, then M(fg)=M(f)M(g)M(fg)=M(f)M(g). Moreover, M(f)=1M(f)=1 if and only if ff is a unit or a unit times a product of 1-variable cyclotomic polynomials, each evaluated at a monomial of d{\cal R}_{d} (see [26]).

(3) When d=1d=1, Jensen’s formula shows that M(f)M(f) can be described in a simple way. If f(x)=csxs+c1x+c0f(x)=c_{s}x^{s}+\cdots c_{1}x+c_{0}, c0cs0c_{0}c_{s}\neq 0, cc is the leading coefficient of ff, then

M(f)=|cs|i=1smax{|λi|,1},M(f)=|c_{s}|\prod_{i=1}^{s}\max\{|\lambda_{i}|,1\},

where λ1,,λs\lambda_{1},\ldots,\lambda_{s} are the roots of ff.

Theorem 9.3.

If GG is a signed dd-periodic graph with nonzero Laplacian polynomial ΔG\Delta_{G}, then

lim supΛ1|d/Λ|logκGΛ=logM(ΔG),\limsup_{\langle{\Lambda}\rangle\to\infty}\frac{1}{|\mathbb{Z}^{d}/{\Lambda}|}\log{\kappa}_{G_{\Lambda}}=\log M(\Delta_{G}), (9.1)

where Λ{\Lambda} ranges over all finite-index subgroups of d\mathbb{Z}^{d}, and Λ\langle{\Lambda}\rangle denotes the minimum length of a nonzero vector in Λ{\Lambda}. When d=1d=1, the limit superior can be replaced by an ordinary limit.

We call this limit the complexity growth rate of GG, and denote it by γG\gamma_{G}. Its relationship to the thermodynamic limit or bulk limit defined for a wide class of unsigned lattice graphs is discussed in [16], and also below in Section 11.

Remark 9.4.

(1) The condition Λ\langle{\Lambda}\rangle\to\infty ensures that fundamental region of Λ{\Lambda} grows in all directions.

(2) If GG is unsigned, κGΛ=τGΛ{\kappa}_{G_{\Lambda}}=\tau_{G_{\Lambda}} for every Λ{\Lambda}. In this case, Theorem 9.3 is proven in [21] with the limit superior replaced by ordinary limit.

(3) When d=1d=1, the finite-index subgroups Λ{\Lambda} are simply /r\mathbb{Z}/r\mathbb{Z}, for r>0r>0. In this case, we write GrG_{r} instead of GΛG_{\Lambda}.

(4) When d>1d>1, a recent result of V. Dimitrov [9] asserts that the limit superior in Theorem 9.3 is equal to the ordinary limit along sequences of sublattices Λ{\Lambda} of the form NdN\cdot\mathbb{Z}^{d}, where NN is a positive integer.

Refer to caption
Figure 11: 1-Periodic graph GG with τGr=0\tau_{G_{r}}=0 for all r1r\geq 1

Before proving Theorem 9.3 we give an example that demonstrates the need for defining graph complexity as we do.

Example 9.5.

Consider the 1-periodic graph GG in Figure 11. Generators for the Laplacian module are indicated. The Laplacian matrix is

LG=(01x11x11x0111120x102),L_{G}=\begin{pmatrix}0&1-x^{-1}&1&-x^{-1}\\ 1-x&0&1&-1\\ 1&1&-2&0\\ -x&-1&0&2\end{pmatrix},

and ΔG(x)=9(x2+x1).\Delta_{G}(x)=9(x-2+x^{-1}).

The quotient G2G_{2} is the finite graph in Example 2.2. The Laplacian matrix of any GrG_{r} can be described as a block matrix obtained from LGL_{G} by replacing xx by the companion (permutation) matrix for xr1x^{r}-1, and any scalar cc by cIrcI_{r} (see [29]). It is conjugate to the diagonal block matrix Diag[LG|x=1,,LG|x=ζr1]{\rm Diag}[L_{G}|_{x=1},\ldots,L_{G}|_{x=\zeta^{r-1}}], where ζ\zeta is a primitive rrth root of unity. The matrix LG|x=1L_{G}|_{x=1} is the 4×44\times 4 Laplacian matrix of G¯\overline{G},

LG¯=(0011001111201102),L_{\overline{G}}=\begin{pmatrix}0&0&1&-1\\ 0&0&1&-1\\ 1&1&-2&0\\ -1&-1&0&2\end{pmatrix},

which has nullity 2. Hence the tree complexity τGr\tau_{G_{r}} vanishes for every rr. Nevertheless, by Theorem 9.3 the (torsion) complexity κGr{\kappa}_{G_{r}} is nontrivial and has exponential growth rate equal to 99. One can verify directly that the Laplacian subgroup Gr{\cal L}_{G_{r}} is isomorphic to 2×(/3r1)2.\mathbb{Z}^{2}\times(\mathbb{Z}/3^{r-1}\mathbb{Z})^{2}.

We proceed with the proof of Theorem 9.3.

Proof.

The proof that we present is a direct application of a theorem of D. Lind, K. Schmidt and T. Ward (see [19] or Theorem 21.1 of [26]). We review the ideas for the reader’s convenience.

Recall that the Laplacian module G{\cal L}_{G} is the finitely generated module over the ring d{\cal R}_{d} with presentation matrix equal to the n×nn\times n Laplacian matrix LGL_{G}, and its Pontryagin dual group ^G\widehat{\cal L}_{G} is Hom(G,𝕋){\rm Hom}({\cal L}_{G},{\mathbb{T}}). The module actions of x1,,xdx_{1},\ldots,x_{d} determine commuting homeomorphisms 𝐬1,,𝐬d{\bf s}_{1},\ldots,{\bf s}_{d} of ^G\widehat{\cal L}_{G}. Explicitly, (𝐬jρ)(a)=ρ(xja)({\bf s}_{j}\rho)(a)=\rho(x_{j}a) for every aGa\in{\cal L}_{G}. Consequently, Γ^G\widehat{{\Gamma}}_{G} has a d\mathbb{Z}^{d}-action 𝐬:dAut(^G){\bf s}:\mathbb{Z}^{d}\to\text{Aut}(\widehat{\cal L}_{G}).

The pair (^G,σ)(\widehat{\cal L}_{G},\sigma) is an algebraic dynamical system, well defined up to topological conjugacy (that is, up to a homeomorphism of ^G\widehat{\cal L}_{G} respecting the d\mathbb{Z}^{d} action). In particular its periodic point structure is well defined.

Topological entropy h(𝐬)h({\bf s}) is a well-defined quantity associated to (^G,𝐬)(\widehat{\cal L}_{G},{\bf s}), a measure of complexity of the d\mathbb{Z}^{d}-action 𝐬{\bf s}. We refer the reader to [19] or [26] for the definition.

For any subgroup Λ{\Lambda} of d\mathbb{Z}^{d}, a Λ{\Lambda}-periodic point is a member of ^G\widehat{\cal L}_{G} that is fixed by every element of Λ{\Lambda}. The set of Λ{\Lambda}-periodic points is a finitely generated abelian group isomorphic to the Pontryagin dual group Hom(G/ΛG),𝕋){\rm Hom}({\cal L}_{G}/{\Lambda}{\cal L}_{G}),{\mathbb{T}}).

The group G/ΛG{\cal L}_{G}/{\Lambda}{\cal L}_{G} is the Laplacian module of the quotient graph GΛG_{\Lambda}. As a finitely generated abelian group, it decomposes as βΛT(G/ΛG)\mathbb{Z}^{\beta_{\Lambda}}\oplus T({\cal L}_{G}/{\Lambda}{\cal L}_{G}), where βΛ\beta_{\Lambda} is the rank of G/ΛG{\cal L}_{G}/{\Lambda}{\cal L}_{G} and T()T(\cdots) denotes the (finite) torsion subgroup. The Pontryagin dual group consists of PΛ=|T(G/ΛG)|P_{\Lambda}=|T({\cal L}_{G}/{\Lambda}{\cal L}_{G})| tori each of dimension βΛ\beta_{\Lambda}. By Theorem 21.1 of [26], the topological entropy h(𝐬)h({\bf s}) is:

h(𝐬)=lim supΛ1|d/Λ|logPΛ=lim supΛ1|d/Λ|logκΛ.h({\bf s})=\limsup_{\langle{\Lambda}\rangle\to\infty}\frac{1}{|\mathbb{Z}^{d}/{\Lambda}|}\log P_{\Lambda}=\limsup_{\langle{\Lambda}\rangle\to\infty}\frac{1}{|\mathbb{Z}^{d}/{\Lambda}|}\log{\kappa}_{\Lambda}.

Since the matrix LGL_{G} that presents G{\cal L}_{G} is square, h(𝐬)h({\bf s}) can be computed also as the logarithm of the Mahler measure M(detLG)M(\det L_{G}) (see Example 18.7(1) of [26]). The determinant of LGL_{G} is, by definition, the Laplacian polynomial ΔG\Delta_{G}. Hence the proof is complete. ∎

10 Lehmer’s question

In [17] D.H. Lehmer asked the following question.

Question 10.1.

Do there exist integral polynomials with Mahler measures arbitrarily close but not equal to 1?

Lehmer discovered the polynomial x10+x9x7x6x5x4x3+x+1,x^{10}+x^{9}-x^{7}-x^{6}-x^{5}-x^{4}-x^{3}+x+1, which has Mahler measure equal to 1.176281.17628.... Despite great effort including extensive computer-aided searches [3, 4, 22, 23, 25], no smaller value greater than 1 has been found, and Lehmer’s question remains unanswered.

Topological and geometric perspectives of Lehmer’s question have been found [13]. In [31] we showed that Lehmer’s question is equivalent to a question about Alexander polynomials of fibered hyperbolic knots in the lens spaces L(n,1),n>0L(n,1),n>0. (Lens spaces arose from the need to consider polynomials f(x)f(x) with f(1)=n1f(1)=n\neq 1.) Here we present another, more elementary equivalence, in terms of graph complexity.

An integer polynomial f(x)f(x) is reciprocal if xdegff(x1)=f(x)x^{{\deg}f}f(x^{-1})=f(x). We will say that a Laurent polynomial f(x)1f(x)\in{\cal R}_{1} is palindromic if f(x1)=f(x)f(x^{-1})=f(x). Any reciprocal polynomial becomes palindromic after it is multiplied by xjx^{j} or xj(x+1)x^{j}(x+1), for suitable jj. In [34] C. Smyth proved that any irreducible integral non-reciprocal polynomial other than xx or x1x-1 has Mahler measure at least as large as the real root of x3x1x^{3}-x-1 (approximately 1.324). Since Mahler measure is multiplicative, it suffices to restrict our attention to palindromic Laurent polynomials when investigating Lehmer’s question.

Proposition 10.2.

A polynomial Δ(x)\Delta(x) is the Laplacian polynomial of a 1-periodic graph if and only if it has the form (x2+x1)f(x)(x-2+x^{-1})f(x), where f(x)f(x) is a palindromic polynomial.

Proof.

The Laplacian polynomial Δ(x)\Delta(x) of any 11-periodic graph is palindromic. This follows from the fact that the transpose of LGL_{G} is LGL_{G} with x replaced by x1x^{-1}. Since the row-sums of LGL_{G} become zero when we set x=1x=1, x1x-1 divides Δ(x)\Delta(x). (Both observations follow also from Theorem 7.1.) Palindromicity requires that the multiplicity of x1x-1 be even. Hence Δ(x)\Delta(x) has the form (x2+x1)f(x)(x-2+x^{-1})f(x), where f(x)f(x) is palindromic.

In order to see the converse assertion, consider any polynomial of the form p(x)=(x2+x1)f(x)p(x)=(x-2+x^{-1})f(x), where f(x)f(x) is palindromic. Then p(x)p(x) is also palindromic. Clearly, we can write p(x)p(x) as a constant plus a sum of terms ±(xs2+xs)\pm(x^{s}-2+x^{-s}); but the constant must be 0 since p(1)=0p(1)=0. Then p(x)p(x) is the Laplacian polynomial of a 1-periodic graph, constructed as in the following example. ∎

Example 10.3.

Multiplying Lehmer’s polynomial f(x)=x10+x9x7x6x5x4x3+x+1f(x)=x^{10}+x^{9}-x^{7}-x^{6}-x^{5}-x^{4}-x^{3}+x+1 by the unit x5x^{-5} and then by x2+x1x-2+x^{-1} yields x6x5x4+x2+x2x4x5+x6,x^{6}-x^{5}-x^{4}+x^{2}+x^{-2}-x^{-4}-x^{-5}+x^{-6}, which in turn can be written as

(x22+x2)(x42+x4)(x52+x5)+(x62+x6).(x^{2}-2+x^{-2})-(x^{4}-2+x^{-4})-(x^{5}-2+x^{-5})+(x^{6}-2+x^{-6}).

This is the Laplacian polynomial of a 1-periodic graph GG. The quotient graph G¯\overline{G} is easily described. It has a single vertex, two edges with sign +1+1 and two with 1-1. The (+1)(+1)-signed edges wind twice and six times, respectively, around the annulus in the direction corresponding to xx. The (1)(-1)-signed edges wind four and five times, respectively, in the opposite direction.

Theorem 10.4.

Lehmer’s question is equivalent to the following. Given ϵ>0\epsilon>0, does there exist a 1-periodic graph GG such that

1<limr(τGr)1/r<1+ϵ?1<\lim_{r\to\infty}(\tau_{G_{r}})^{1/r}<1+\epsilon?
Proof.

When investigating Lehmer’s question it suffices to consider polynomials of the form (x2+x1)f(x)(x-2+x^{-1})f(x), where f(x)f(x) is palindromic and irreducible. By Proposition 10.2 any such polynomial is realized as the Laplacian polynomial of a 1-periodic graph GG with a single vertex orbit. As in Example 9.5 the Laplacian matrix LGrL_{G_{r}} of any finite quotient GrG_{r} can be obtained from (x2+x1)f(x)(x-2+x^{-1})f(x) by substituting for xx the companion matrix for xr1x^{r}-1. Hence the nullity of LGrL_{G_{r}} is 1 provided that f(x)f(x) is not a cyclotomic polynomial (multiplied by a unit), a condition that we can assume without loss of generality. Hence κGr=τGr\kappa_{G_{r}}=\tau_{G_{r}} for each rr (see discussion following Definition 2.1.) Theorem 9.3 completes the proof.

Remark 10.5.

(1) The closure of the 16-braid (σ1σ2)2(σ1σ2σ3σ4)4(σ1σ2σ15)5(\sigma_{1}\sigma_{2})^{2}(\sigma_{1}\sigma_{2}\sigma_{3}\sigma_{4})^{4}(\sigma_{1}\sigma_{2}\cdots\sigma_{15})^{5} is a link ¯\overline{\ell} associated with a plane graph G¯\overline{G} in the annulus (see Example 8.8). The Laplacian polynomial ΔG\Delta_{G} is

x7(x1)(x2+1)(x10x9x6+x5x4x+1).x^{-7}(x-1)(x^{2}+1)(x^{10}-x^{9}-x^{6}+x^{5}-x^{4}-x+1).

Its Mahler measure is 1.35098\ldots. This is the smallest Mahler measure greater than 1 that we have yet found for any plane graph.

(2) The conclusion of Theorem 10.4 does not hold if we restrict ourselves to unsigned graphs. By Theorem 11.7 below, the Mahler measure of the Laplacian polynomial of any 1-periodic graph with all edge signs equal to 1 is at least 2.

(3) If a 1-periodic graph GG as in Example 10.3 can be found with M(DG)M(D_{G}) less than Lehmer’s value 1.176281.17628..., then by results of [24] some edge of G¯\overline{G} must wind around the annulus at least 29 times.

The cyclic 5-fold cover of the graph in Example 10.3 contains the complete graph on 5 vertices, and hence it is nonplanar. If the answer to the following question is yes, then Lehmer’s question is equivalent to a question about determinant density of links (see Remark 11.3(3)).

Question 10.6.

Is Theorem 10.4 still true if we require that the graphs GG be planar?

We conclude this section with a result that will be used in the next section, but holds for signed as well as unsigned graphs. It concerns complexity growth of a dd-periodic graph that is a union of disjoint dd^{\prime}-periodic graphs for some d<dd^{\prime}<d.

Suppose HH is a subgraph of a dd-periodic graph GG consisting of one or more connected components of GG, such that the orbit of HH under d\mathbb{Z}^{d} is all of GG. Let Γ<d\Gamma<\mathbb{Z}^{d} be the stabilizer of HH. Then Γd\Gamma\cong\mathbb{Z}^{d^{\prime}} for some d<dd^{\prime}<d, and its action on HH can be regarded as a cofinite free action of d\mathbb{Z}^{d^{\prime}}. Consider the limit

γH=limΛ1|Γ/Λ|logκHΛ\gamma_{H}=\lim_{{\langle}{\Lambda}{\rangle}\to\infty}\frac{1}{|\Gamma/{\Lambda}|}\log{\kappa}_{H_{\Lambda}}

where Λ{\Lambda} ranges over finite-index subgroups of Γ\Gamma.

Lemma 10.7.

Under the above conditions we have γG=γH\gamma_{G}=\gamma_{H}.

Proof.

Let Λ{\Lambda} be any finite-index subgroup of d\mathbb{Z}^{d}. Then HH is invariant under ΛΓ{\Lambda}\cap{\Gamma}. The image of HH in the quotient graph GΛG_{\Lambda} is isomorphic to HΛΓH_{{\Lambda}\cap{\Gamma}}.

Note that the quotient H¯\overline{H} of HH by the action of Γ{\Gamma} is isomorphic to G¯\overline{G}, since the d\mathbb{Z}^{d} orbit of HH is all of GG. Since GΛG_{\Lambda} is a |d/Λ||\mathbb{Z}^{d}/{\Lambda}|-fold cover of G¯\overline{G} and HΛΓH_{{\Lambda}\cap{\Gamma}} is a |Γ/(ΛΓ)||{\Gamma}/({\Lambda}\cap{\Gamma})|-fold cover of H¯\overline{H}, GΛG_{\Lambda} comprises k=|d/Λ|/|Γ/(ΛΓ)|k=|\mathbb{Z}^{d}/{\Lambda}|/|{\Gamma}/({\Lambda}\cap{\Gamma})| mutually disjoint translates of a graph that is isomorphic to HΛΓH_{{\Lambda}\cap{\Gamma}}. Hence κGΛ=κHΛΓk{\kappa}_{G_{\Lambda}}={\kappa}_{H_{{\Lambda}\cap{\Gamma}}}^{k} and

1|d/Λ|logκGΛ=1|Γ/(ΛΓ)|logκHΛΓ.\frac{1}{|\mathbb{Z}^{d}/{\Lambda}|}\log{\kappa}_{G_{\Lambda}}=\frac{1}{|{\Gamma}/({\Lambda}\cap{\Gamma})|}\log{\kappa}_{H_{{\Lambda}\cap{\Gamma}}}.

Since ΛΓ{\langle}{\Lambda}\cap{\Gamma}{\rangle}\to\infty as Λ{\langle}{\Lambda}{\rangle}\to\infty, we have γG=γH\gamma_{G}=\gamma_{H}. ∎

11 Complexity growth of unsigned periodic graphs

It is natural to ask whether the Mahler measure of Laplacian polynomials of signed graphs differs in appreciable ways from unsigned graphs. Proposition 11.7 answers emphatically yes.

Throughout the section GG denotes an unsigned dd-periodic graph. For this case, the complexity growth rate γG\gamma_{G} is also the growth rate of the number of spanning trees of finite quotients GΛG_{\Lambda}. Thus contracting or deleting an edge orbit of GG will not increase γG\gamma_{G}.

Denote by R=R(Λ)R=R({\Lambda}) a fundamental domain of Λ{\Lambda}. Let G|RG|_{R} be the full unsigned subgraph of GG on vertices vi,𝐧,𝐧Rv_{i,{\bf n}},\ {\bf n}\in R. We denote by R\ell_{R} the corresponding medial link.

If G|RG|_{R} is connected for each RR, then {τGΛ}\{\tau_{G_{\Lambda}}\} and {τG|R}\{\tau_{G|_{R}}\} have the same exponential growth rates. (See Theorem 7.10 of [16] for a short, elementary proof. A more general result is Corollary 3.8 of [21].) The bulk limit is defined by γG/|V(G¯)|\gamma_{G}/|V(\overline{G})|.

Example 11.1.

The dd-dimensional grid graph 𝔾d\mathbb{G}_{d} is the unsigned graph with vertex set d\mathbb{Z}^{d} and single edges connecting each pair of vertices of distance 1. Its Laplacian polynomial is

Δ(𝔾d)=2dx1x11xdxd1.\Delta(\mathbb{G}_{d})=2d-x_{1}-x_{1}^{-1}-\cdots-x_{d}-x_{d}^{-1}.

When d=2d=2, it is a plane graph. The graphs links R\ell_{R} are indicated in Figure 12 for Λ=x12,x22{\Lambda}=\langle x_{1}^{2},x_{2}^{2}\rangle on left and Λ=x13,x23{\Lambda}=\langle x_{1}^{3},x_{2}^{3}\rangle on right.

Refer to caption
Figure 12: Graphs (𝔾2)R(\mathbb{G}_{2})_{R} and associated links, Λ=x12,x22{\Lambda}=\langle x_{1}^{2},x_{2}^{2}\rangle and x13,x23\langle x_{1}^{3},x_{2}^{3}\rangle

The determinant of a link \ell, denoted here by det(){\rm det}(\ell), is the absolute value of its 1-variable Alexander polynomial evaluated at 1-1. It follows from the Mayberry-Mott theorem [2] that if \ell is an alternating link that arises by the medial construction from a finite plane graph, edge signs ±1\pm 1 allowed, then det(){\rm det}(\ell) is equal to the tree complexity of the graph (see Appendix A.4 in [5]). The following corollary is an immediate consequence of Theorem 9.3. It has been proven independently by Champanerkar and Kofman [6].

Corollary 11.2.

Let GG be a connected dd-periodic unsigned plane graph, d=1d=1 or 22. Then

limΛ1|d/Λ|logdet(R)=γΔG.\lim_{\langle{\Lambda}\rangle\to\infty}\frac{1}{|\mathbb{Z}^{d}/{\Lambda}|}\log{\rm det}(\ell_{R})=\gamma_{\Delta_{G}}.
Remark 11.3.

(1) We regard the limit in that statement of Corollary 11.2 as a determinant density of the collection of links {R}\{\ell_{R}\}. There are other ways to define it (e.g., dividing by the number of crossings of the diagram for R\ell_{R}).

(2) In [7] the authors consider as well more general sequences of links. When G=𝔾2G=\mathbb{G}_{2}, their results imply that:

limΛ2πc(R)logdet(R)=voct,\lim_{\langle{\Lambda}\rangle\to\infty}\frac{2\pi}{c(\ell_{R})}\log{\rm det}(\ell_{R})=v_{oct},

where c(R)c(\ell_{R}) is the number of crossings of R\ell_{R} and voct3.66386v_{oct}\approx 3.66386 is the volume of the regular ideal octohedron.

(3) If Question 10.6 has an affirmative answer then Lehmer’s question becomes a question about link determinants.

Grid graphs are the simplest unsigned dd-periodic graphs, as the following theorem shows.

Theorem 11.4.

If GG is an unsigned connected dd-periodic graph, then γGγ𝔾d\gamma_{G}\geq\gamma_{\mathbb{G}_{d}}.

Asymptotic results about the Mahler measure of certain families of polynomials have been obtained elsewhere. However, the graph theoretic methods that we employ to prove Theorem 11.4 are different from techniques used previously.

Proof.

Consider the case in which GG has a single vertex orbit. Then for some u1,,umdu_{1},\ldots,u_{m}\in\mathbb{Z}^{d}, with mdm\geq d, the edge set E(G)E(G) consists of edges from vv to uivu_{i}\cdot v for each vVv\in V and i=1,,mi=1,\ldots,m. Since GG is connected, we can assume after relabeling that u1,,udu_{1},\ldots,u_{d} generate a finite-index subgroup of d\mathbb{Z}^{d}. Let GG^{\prime} be the d\mathbb{Z}^{d}-invariant subgraph of GG with edges from vv to uivu_{i}\cdot v for each vVv\in V and i=1,,di=1,\ldots,d. Then GG^{\prime} is the orbit of a subgraph of GG that is isomorphic to 𝔾d\mathbb{G}_{d}, and so by Lemma 10.7, γ(𝔾d)=γ(G)γ(G)\gamma(\mathbb{G}_{d})=\gamma(G^{\prime})\leq\gamma(G).

We now consider a connected graph GG having vertex families v1,𝐧,,vn,𝐧v_{1,{\bf n}},\ldots,v_{n,{\bf n}}, where n>1n>1. Since GG is connected, there exists an edge ee joining v1,𝟎v_{1,{\bf 0}} to some vi,𝐧v_{i,{\bf n}}. Contract the edge orbit de\mathbb{Z}^{d}\cdot e to obtain a new graph GG^{\prime} having cofinite free d\mathbb{Z}^{d}-symmetry and complexity growth rate no greater than that of GG. Repeat the procedure with the remaining vertex families so that only v1,𝐧v_{1,{\bf n}} remains. The proof in the previous case of a graph with a single vertex orbit now applies. ∎

Remark 11.5.

The conclusion of Theorem 11.4 does not hold without the hypothesis that GG is connected. Consider the 2-periodic graph GG obtained from 𝔾2\mathbb{G}_{2} by removing all vertical edges, so that GG consists of countably many copies of 𝔾1\mathbb{G}_{1} . Then γG=γ𝔾1=0\gamma_{G}=\gamma_{\mathbb{G}_{1}}=0 while γ𝔾2>0\gamma_{\mathbb{G}_{2}}>0.

The following lemma, needed for the proof Proposition 11.7, is of independent interest.

Lemma 11.6.

The sequence of complexity growth rates γΔ𝔾d\gamma_{\Delta_{\mathbb{G}_{d}}} is nondecreasing.

Proof.

Consider the grid graph 𝔾d\mathbb{G}_{d}. Deleting all edges in parallel to the ddth coordinate axis yields a subgraph GG consisting of countably many mutually disjoint translates of 𝔾d1\mathbb{G}_{d-1}. By Lemma 10.7, γ𝔾d1=γGγ𝔾d\gamma_{\mathbb{G}_{d-1}}=\gamma_{G}\leq\gamma_{\mathbb{G}_{d}}. ∎

Doubling each edge of 𝔾1\mathbb{G}_{1} results in a graph with Laplacian polynomial 2(x2+x1)2(x-2+x^{-1}), which has Mahler measure 2M(x2+x1)=22M(x-2+x^{-1})=2. We show that this graph realizes the minimum nonzero complexity growth rate.

Proposition 11.7.

(Complexity Growth Rate Gap) Let GG be any unsigned dd-periodic graph. If γG0\gamma_{G}\neq 0, then

γGlog2.\gamma_{G}\geq\log 2.

Although Δ𝔾d\Delta_{\mathbb{G}_{d}} is relatively simple, the task of computing its Mahler measure is not. It is well known and not difficult to see that γ𝔾dlog2d\gamma_{\mathbb{G}_{d}}\leq\log 2d. We will use a theorem of N. Alon [1] to show that γ𝔾d\gamma_{\mathbb{G}_{d}} approaches log2d\log 2d asymptotically.

Theorem 11.8.

(1) γ𝔾dlog2d\gamma_{\mathbb{G}_{d}}\leq\log 2d, for all d1d\geq 1.
(2) limdγ𝔾dlog2d=0.\lim_{d\to\infty}\gamma_{\mathbb{G}_{d}}-\log 2d=0.

Proof of Proposition 11.7. By Lemma 10.7 it suffices to consider a connected dd-periodic graph GG with γG\gamma_{G} nonzero. Note that γ𝔾1=0\gamma_{\mathbb{G}_{1}}=0 while γ𝔾21.165\gamma_{\mathbb{G}_{2}}\approx 1.165 is greater than log2\log 2. By Theorem 11.4 and Lemma 11.6 we can assume that d=1d=1.

If GG has an orbit of parallel edges, we see easily that γGlog2\gamma_{G}\geq\log 2. Otherwise, we proceed as in the proof of Theorem 11.4, contracting edge orbits to reduce the number of vertex orbits without increasing the complexity growth rate. If at any step we obtain an orbit of parallel edges, we are done; otherwise we will obtain a graph GG^{\prime} with a single vertex orbit and no loops. If GG^{\prime} is isomorphic to 𝔾1\mathbb{G}_{1} then GG must be a tree; but then γG=0\gamma_{G}=0, contrary to our hypothesis. So GG^{\prime} must have at least two edge orbits. Deleting excess edges, we may suppose GG^{\prime} has exactly two edge orbits.

The Laplacian polynomial ΔG\Delta_{G^{\prime}} has the form 4xrxrxsxs4-x^{r}-x^{-r}-x^{s}-x^{-s}, for some positive integers r,sr,s. Reordering the vertex set of GG^{\prime}, we can assume without loss of generality that r=1r=1. The following calculation is based on an idea suggested to us by Matilde Lalin.

logM(ΔG)=01log|42cos(2πθ)2cos(2πsθ)|dθ\log M(\Delta_{G^{\prime}})=\int_{0}^{1}\log|4-2\cos(2\pi\theta)-2\cos(2\pi s\theta)|\ d\theta
=01log|2(1cos(2πθ))+2(1cos(2πsθ))|dθ=\int_{0}^{1}\log|2(1-\cos(2\pi\theta))+2(1-\cos(2\pi s\theta))|\ d\theta
=01log(4sin2(πθ)+4sin2(πsθ))𝑑θ.=\int_{0}^{1}\log\bigg{(}4\sin^{2}(\pi\theta)+4\sin^{2}(\pi s\theta)\bigg{)}\ d\theta.

Using the inequality (u2+v2)2uv(u^{2}+v^{2})\geq 2uv, for any nonnegative u,vu,v, we have:

logM(ΔG)01log(8|sin(πθ)||sin(πsθ)|)𝑑θ\log M(\Delta_{G^{\prime}})\geq\int_{0}^{1}\log\bigg{(}8|\sin(\pi\theta)|\ |\sin(\pi s\theta)|\bigg{)}\ d\theta
=log8+01log|sin(πθ)|dθ+01log|sin(πsθ)|dθ=\log 8+\int_{0}^{1}\log|\sin(\pi\theta)|\ d\theta+\int_{0}^{1}\log|\sin(\pi s\theta)|\ d\theta
=log8+01log1cos(2πθ)2dθ+01log1cos(2πsθ)2dθ=\log 8+\int_{0}^{1}\log\sqrt{\frac{1-\cos(2\pi\theta)}{2}}\ d\theta+\int_{0}^{1}\log\sqrt{\frac{1-\cos(2\pi s\theta)}{2}}\ d\theta
=log8+0112log(22cos(2πθ)4)𝑑θ+0112log(22cos(2πsθ)4)𝑑θ=\log 8+\int_{0}^{1}\frac{1}{2}\log\bigg{(}\frac{2-2\cos(2\pi\theta)}{4}\bigg{)}\ d\theta+\int_{0}^{1}\frac{1}{2}\log\bigg{(}\frac{2-2\cos(2\pi s\theta)}{4}\bigg{)}\ d\theta
=log8+12m(2xx1)12log4+12m(2xsxs)12log4=\log 8+\frac{1}{2}m(2-x-x^{-1})-\frac{1}{2}\log 4+\frac{1}{2}m(2-x^{s}-x^{-s})-\frac{1}{2}\log 4
=3log2+0log2+0log2=log2.=3\log 2+0-\log 2+0-\log 2=\log 2.


Our proof of Theorem 11.8 depends on the following result of Alon.

Theorem 11.9.

[1] If GG is a finite connected ρ\rho-regular unsigned graph, then

τG[ρ(1ϵ(ρ))]|V(G)|,\tau_{G}\geq[\rho(1-\epsilon(\rho))]^{|V(G)|},

where ϵ(ρ)\epsilon(\rho) is a nonnegative function with ϵ(ρ)0\epsilon(\rho)\to 0 as ρ\rho\to\infty.

Proof of Theorem 11.8. (1) The integral representing the logarithm of the Mahler measure of Δ𝔾d\Delta_{\mathbb{G}_{d}} can be written

0101log|2di=1d2cos(2πθi)|dθ1dθd\int_{0}^{1}\cdots\int_{0}^{1}\log\bigg{|}2d-\sum_{i=1}^{d}2\cos(2\pi\theta_{i})\bigg{|}d\theta_{1}\cdots d\theta_{d}
=log2d+0101log|1+i=1dcos(2πθi)d|dθ1dθd=\log 2d+\int_{0}^{1}\cdots\int_{0}^{1}\log\bigg{|}1+\sum_{i=1}^{d}\frac{\cos(2\pi\theta_{i})}{d}\bigg{|}d\theta_{1}\cdots d\theta_{d}
=log2d+0101k=1(1)kk(i=1dcos(2πθi)d)kdθ1dθd.=\log 2d+\int_{0}^{1}\cdots\int_{0}^{1}-\sum_{k=1}^{\infty}\frac{(-1)^{k}}{k}\bigg{(}\frac{\sum_{i=1}^{d}\cos(2\pi\theta_{i})}{d}\bigg{)}^{k}d\theta_{1}\cdots d\theta_{d}.

By symmetry, odd powers of kk in the summation contribute zero to the integration. Hence

log(Δ𝔾d)=log2d0101k=112k(i=1dcos(2πθi)d)2kdθ1dθdlog2d.\log(\Delta_{\mathbb{G}_{d}})=\log 2d-\int_{0}^{1}\cdots\int_{0}^{1}\sum_{k=1}^{\infty}\frac{1}{2k}\bigg{(}\frac{\sum_{i=1}^{d}\cos(2\pi\theta_{i})}{d}\bigg{)}^{2k}d\theta_{1}\cdots d\theta_{d}\leq\log 2d.

(2) Let Λ{\Lambda} be a finite-index subgroup of d\mathbb{Z}^{d}. Consider the quotient graph (𝔾d)Λ(\mathbb{G}_{d})_{\Lambda}. The cardinality of its vertex set is |d/Λ||\mathbb{Z}^{d}/{\Lambda}|. The main result of [1], cited above as Theorem 11.9, implies that

τ(𝔾d)Λ=((2d)(1μ(d)))|d/Λ|,\tau_{(\mathbb{G}_{d})_{\Lambda}}=\bigg{(}(2d)(1-\mu(d))\bigg{)}^{|\mathbb{Z}^{d}/{\Lambda}|},

where μ\mu is a nonnegative function such that limdμ(d)=0\lim_{d\to\infty}\mu(d)=0. Hence

limd(1|d/Λ|logτ(𝔾d)Λlog2d)=limdlog(1μ(d))=0.\lim_{d\to\infty}\bigg{(}\frac{1}{|\mathbb{Z}^{d}/{\Lambda}|}\log\tau_{(\mathbb{G}_{d})_{\Lambda}}-\log 2d\bigg{)}=\lim_{d\to\infty}\log(1-\mu(d))=0.

Theorem 9.3 completes the proof. ∎

Remark 11.10.

One can evaluate logM(Δ(𝔾d))\log M(\Delta(\mathbb{G}_{d})) numerically and obtain an infinite series representing γ𝔾dlog2d\gamma_{\mathbb{G}_{d}}-\log 2d. However, showing rigorously that the sum of the series approaches zero as dd goes to infinity appears to be difficult. (See [28], p. 16 for a heuristic argument.)

References

  • [1] N. Alon, The number of spanning trees in regular graphs, in: Random Structures and Algorithms, Vol. 1, No. 2 (1990), 175–181.
  • [2] A. Bott and J.P. Mayberry, Matrices and trees, in Economic Activity Analysis (O. Morgenstern, ed.), 391–400, Wiley, New York, 1954.
  • [3] D.W. Boyd, Reciprocal polynomials having small measure. Math. Comp. 35 (1980), 1361–1377.
  • [4] D. Boyd, Speculations concerning the range of Mahler’s measure, Canad. Math. Bull. 24 (1981), 453–469.
  • [5] G. Burde, H. Zieschang, Knots, 2nd ed., Walter de Gruyter and Co., Berlin, 2003.
  • [6] A. Champanerkar I. Kofman, Determinant density and biperiodic alternating links, New York J. Math. 22 (2016) 891–906.
  • [7] A. Champanerkar, I. Kofman and J.S. Purcell, Geometrically and diagrammatically maximal knots, J. Lond. Math. Soc. (2) 94 (2016), no. 3, 883–908.
  • [8] D.S. Cochran, Links with zero Alexander polynomial, Ph.D. dissertation, Dartmouth College, 1980.
  • [9] V. Dimitrov, Convergence to the Mahler measure and the distribution of periodic points for algebraic d\mathbb{Z}^{d}-actions, preprint,  2016 arXiv:1611.04664v1
  • [10] G. Everest and T. Ward, Heights of Polynomials and Entropy in Algebraic Dynamics, Springer-Verlag, London 1999.
  • [11] R. Forman, Determinants of Laplacians on graphs, Topology, 32 (1993), 35–46.
  • [12] C. Godsil, G. Royle, Algebraic Graph Theory, Springer Verlag, 2001.
  • [13] E. Hironaka, Lehmer’s problem, McKay’s correspondence, and 2,3,7 in: Topics in algebraic and noncommutative geometry (Luminy/Annapolis, MD, 2001), 123–138, Contemp. Math., 324, Amer. Math. Soc., Providence, RI, 2003.
  • [14] L.H. Kauffman, Knots and Physics, Third Edition, World Scientific, Singapore, 2001.
  • [15] R. Kenyon, Spanning forests and the vector bundle Laplacian, Annals of Probability 39 (2011), 1983–2017.
  • [16] K.R. Lamey, D.S. Silver and S.G. Williams, Vertex-colored graphs, bicycle spaces and Mahler measure, J. Knot Theory and its Ramifications, 25 (2016), no. 6, 1650033, 22 pp.
  • [17] D.H. Lehmer, Factorization of certain cyclotomic functions. Annals of Math. 34 (1933), 461–479.
  • [18] W.B. Lickorish, An introduction to knot theory, Springer-Verlag, Berlin - New York, 1974.
  • [19] D.A. Lind, K. Schmidt and T. Ward, Mahler measure and entropy for commuting automorphisms of compact groups, Inventiones Math. 101 (1990), 593–629.
  • [20] C. Livingston, Knot Theory, Carus Math. Monographs 24, Math. Assoc. Amer., Washington D.C., 1993.
  • [21] R. Lyons, Asymptotic enumeration of spanning trees, Probability and Computing 14 (2005), 491–522.
  • [22] J. McKee and C. Smyth, Integer symmetric matrices of small spectral radius and small Mahler measure. Int. Math. Res. Not. IMRN 2012, no. 1, 102–136.
  • [23] M.J. Mossinghoff, Polynomials with small Mahler measure. Math. Comp. 67 (1998), 1697–1705.
  • [24] M.J. Mossinghoff, Minimal Mahler measures, Experiment. Math. 17 (2008), no. 4, 451–458.
  • [25] G.A. Ray, A locally parameterized version of Lehmer’s problem. In Mathematics of Computation 1943–1993: A Half-Century of Computational Mathematics (W. Gautschi, ed.), Proc. Symp. Appl. Math. 48 (1994), 573–576.
  • [26] K. Schmidt, Dynamical Systems of Algebraic Origin, Birkhäuser, Basel, 1995.
  • [27] D.S. Silver, L. Traldi and S.G. Williams, Goeritz and Seifert matrices from Dehn presentations, Osaka J. Math. 57 (2020), 663–677.
  • [28] R. Shrock and F.Y. Wu, Spanning trees on graphs and lattices in 𝒟\cal D dimensions, J. Phys. A: Math. Gen. 33 (2000), 3881–3902.
  • [29] D.S. Silver and S.G. Williams, Coloring link diagrams with a continuous palette, Topology 39 (2000), 1225–1237.
  • [30] D.S. Silver and S.G. Williams, Crowell’s derived group and twisted polynomials, J. of Knot Theory Ramifications, 15 (2006), 1079–1094.
  • [31] D.S. Silver and S.G. Williams, Lehmer’s question, knots and surface dynamics, Math. Proc. Camb. Phil. Soc. (2007), 143, 649–661.
  • [32] D.S. Silver and S.G. Williams, Group presentations for links in thickened surfaces, preprint, 2020, arXiv:2005.01576.
  • [33] D.S. Silver and S.G. Williams, Group presentations for links in thickened surfaces, preprint, 2020, arXiv:2005.01576.
  • [34] C.J. Smyth, On the product of conjugates outside the unit circle of an algebraic integer, Bulletin of the London Math. Soc. 3 (1971), 169–175,
  • [35] W.T. Tutte, Graph Theory, Addison-Wesley, Reading, Mass., 1984.

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