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

The Tree-Forest Ratio

Oliver Knill Department of Mathematics
Harvard University
Cambridge, MA, 02138
(Date: May 22, 2022)
Abstract.

The number of rooted spanning forests divided by the number of spanning rooted trees in a graph GG with Kirchhoff matrix KK is the spectral quantity τ(G)=det(1+K)/det(K)\tau(G)={\rm det}(1+K)/{\rm det}(K) of GG by the matrix tree and matrix forest theorems. We prove that that under Barycentric refinements, the tree index T(G)=log(det(K))/|G|T(G)=\log({\rm det}(K))/|G| and forest index F(G)=log(det(1+K))/|G|F(G)=\log({\rm det}(1+K))/|G| and so the tree-forest index i=FG=log(τ(G))/|G|i=F-G=\log(\tau(G))/|G| converge to numbers that only depend on the size of the maximal clique in the graph. In the one-dimensional case, all numbers are known: T(G)=0,F(G)=i(G)=2log(ϕ)T(G)=0,F(G)=i(G)=2\log(\phi), where ϕ\phi is the golden ratio. The convergent proof uses the Barycentral limit theorem assuring the Kirchhoff spectrum converges weakly to a measure μ\mu on [0,)[0,\infty) that only depends on dimension of GG. Trees and forests indices are potential values i=U(1)U(0)i=U(-1)-U(0) for the subharmonic function U(z)=0log|xz|dμ(x)U(z)=\int_{0}^{\infty}\log|x-z|\;d\mu(x) defined by the Riesz measure dμ=ΔUd\mu=\Delta U which only depends on the dimension of GG. The potential U(z)U(z) is defined for all zz away from the support of dμd\mu and finite at z=0z=0. Convergence follows from the tail estimate μ[x,]Ceadx\mu[x,\infty]\leq Ce^{-a_{d}x}, where the decay rate ada_{d} only depends on the maximal dimension. With the normalized zeta function ζ(s)=1|V(G)|kλks\zeta(s)=\frac{1}{|V(G)|}\sum_{k}\lambda_{k}^{-s}, we have for all finite graphs of maximal dimension 2\geq 2 the identity i(G)=t=1(1)s+1ζ(s)/si(G)=\sum_{t=1}^{\infty}(-1)^{s+1}\zeta(s)/s. The limiting zeta function ζ(s)=0xs𝑑μ(x)\zeta(s)=\int_{0}^{\infty}x^{-s}d\mu(x) is analytic in ss for s<0s<0. The Hurwitz spectral zeta function ζz(s)=Us(z)=0(xz)s𝑑μ(x)\zeta_{z}(s)=U_{s}(z)=\int_{0}^{\infty}(x-z)^{-s}\;d\mu(x) complements U(z)=0log(xz)𝑑μ(x)U(z)=\int_{0}^{\infty}\log(x-z)\;d\mu(x) and is analytic for zz in +\mathbb{\mathbb{C}}\setminus\mathbb{R}^{+} and for fixed zz in +\mathbb{C}\setminus\mathbb{R}^{+} is an entire function in ss\in\mathbb{CC}.

Key words and phrases:
Trees, Forests, Graphs, Barycentric refinement, Zeta functions, Gap labeling
1991 Mathematics Subject Classification:
15A15, 16Kxx, 05C10, 57M15, 68R10, 05E45

1. In a nutshell

1.1.

We define here the tree forest ratio τ(G)\tau(G) of a connected finite simple graph G=(V,E)G=(V,E) as the number of rooted spanning forests in GG divided by the number of rooted spanning trees in GG. By the tree and forest matrix theorems, this is τ=Det(K+1)/Det(K)\tau={\rm Det}(K+1)/{\rm Det}(K), where Det{\rm Det} is the pseudo determinant and KK the Kirchhoff Laplacian of GG. In other words, it is τ=λ0(1+1/λ)\tau=\prod_{\lambda\neq 0}(1+1/\lambda), where λ\lambda runs over the non-zero eigenvalues of KK. An upper bound is (1+λ2)1)|V|(1+\lambda_{2})^{-1})^{|}V|, where λ2\lambda_{2} is the ground state, the smallest non-zero eigenvalue of KK. An expansion of the logarithm and a basic bound on the product gives log(1+ζ(1))log(τ)=s=1(1)s+1ζ(s1)/sζ(1)\log(1+\zeta(1))\leq\log(\tau)=\sum_{s=1}^{\infty}(-1)^{s+1}\zeta(s-1)/s\leq\zeta(1), where ζ(s)\zeta(s) is the Kirchhoff spectral zeta function of GG.

1.2.

The same functional could be considered for other operators like the connection matrix LL or the Hodge matrix HH of a graph, even so a tree or forest interpretation lacks then. For the connection matrix which has negative eigenvalues in general, we would use L2L^{2} to have no ambiguity when defining zeta functions As we have an equivalent zeta function expression in the Kirchhoff case, the function τ(G)\tau(G) can now make sense even for manifolds GG which do not feature spanning trees or spanning forests. The reason is that the Laplacian on the manifold defines a Minakshisundaram - Pleijel zeta function and a Ray-Singer determinant. For M=𝕋=/M=\mathbb{T}=\mathbb{R}/\mathbb{Z}, where ζ\zeta is the classical Riemann zeta function, we have τG=sinh(π)/π\tau_{G}=\sinh(\pi)/\pi.

1.3.

Potential theory comes in when seeing the logarithm of the tree number as a potential value V(0)=log(Det(L))V(0)=\log({\rm Det}(L)). Also the logarithm of the forest number V(1)=log(Det(1+L))V(-1)=\log({\rm Det(1+L)}) is a potential value of the potential V(z)=log(Det(LzI))=log(xz)𝑑k(x)V(z)=\log({\rm Det}(L-zI))=\int\log(x-z)\;dk(x), where dk=jδλjdk=\sum_{j}\delta_{\lambda_{j}} is a finite pure point measure defined by the eigenvalues of LL. When normalizing dkdk to become a probability measure dμd\mu, then U(z)=log(xz)𝑑μ(x)U(z)=\int\log(x-z)\;d\mu(x) has a subharmonic real part and dμd\mu is the Riesz measure dμ=ΔUd\mu=\Delta U, a probability measure with support on the positive real axes. The Barycentral limit theorem assures that dμnd\mu_{n} converges in the Barycentric limit weakly to a measure dμd\mu which only depends on the maximal dimension of the initial graph. The potential value U(z)U(z) is defined for all zz away from the support of dμd\mu and finite at z=0z=0 because it is there the limiting growth rate of trees which is boxed in by 0 and the growth rate U(1)U(-1) of forests. In the 11-dimensional case, it is the arc-sin distribution on [0,4][0,4] which is the potential theoretical equilibrium measure on that interval.

Refer to caption
Refer to caption
Figure 1. The real and imaginary part of the potential in the 1-dimensional case is seen to the left. The support of the density of states is [0,4][0,4]. In the 2-dimensional case, where the support of the density of states is non-compact, we see gaps and the support of the spectrum could be a Cantor set.

1.4.

A major point we want to make here is to see that the limiting universal measure dμd\mu has an exponentially decaying tail distribution μ[x,)0\mu[x,\infty)\to 0 as xx\to\infty. This implies that both the forest and tree numbers converge in the Barycentral limit because the potential values U(1)U(-1) and U(0)U(0) exist. Since by nature of tree and forest numbers 0U(0)U(1)0\leq U(0)\leq U(-1), the convergence at the bottom of the spectrum z=0z=0 is not a problem. The tail estimate also shows that for zz away from the positive real axes, the Hurwitz zeta function ζz(s)=Us(z)=0infty(xz)s𝑑μ(x)\zeta_{z}(s)=U_{s}(z)=\int_{0}^{infty}(x-z)^{-s}\;d\mu(x) is an entire function in ss. For z=0z=0, we only know that ζ0(s)=ζ(s)\zeta_{0}(s)=\zeta(s) is analytic for s<0s<0.

1.5.

We will use some linear algebra to estimate the integrated density of states near \infty in terms of the vertex degree distribution of the graph. We need a bit more than the Schur-Horn inequality or the Gershgorin circle theorem: we use λk2dk\lambda_{k}\leq 2d_{k}, where dkd_{k} is the kk’th vertex degree. This appears to be a new result we have written down separately and which follows directly from the Cauchy interlace theorem. We have looked at the zeta function ζ(s)\zeta(s) of circular graphs G=CnG=C_{n} in [11] and looked at the limiting behavior of roots and then in [18] for the connection graphs of circular graphs, where the connection Laplacian is invertible [16, 19] and the Zeta function satisfies a functional equation.

1.6.

For small graphs like complete graphs KnK_{n} or the graph complements CnC_{n}^{\prime} of cyclic graphs CnC_{n}, the tree-forest ratio converges to the Euler constant e=2.718e=2.718\dots in the limit nn\to\infty. For KnK_{n}, one can see that directly from the eigenvalues λ0=0,λk=n\lambda_{0}=0,\lambda_{k}=n because τ(G)=(1+1/n)n1e\tau(G)=(1+1/n)^{n-1}\to e. For Barycentric refinements of a graph G=G0G=G_{0} the tree forest index i(Gn)=log(τ(Gn))/|V(Gn)|i(G_{n})=\log(\tau(G_{n}))/|V(G_{n})| of GnG_{n} converges to a universal constant that only depends on dd. Since k=0dfk(Gn1)=|Gn|=fdk=0dfk(Gn)\sum_{k=0}^{d}f_{k}(G_{n-1})=|G_{n}|=f_{d}\leq\sum_{k=0}^{d}f_{k}(G_{n}) we could also take the limit, where |V(G||V(G| is replaced by any fkf_{k} like for example the volume fdf_{d}. The result is a direct consequence of spectral Barycentral limit theorem [14, 15]. In the Barycentric limit, we have log(τ(Gn))/|Gn|0log(1+1/x)𝑑μ(x)\log(\tau(G_{n}))/|G_{n}|\to\int_{0}^{\infty}\log(1+1/x)d\mu(x), where μ\mu is the limiting density of states. To show that this is a finite number, we only need to show that the density of states dkdk decays fast. We needed something stronger than the Gershgorin circle theorem [6, 26] or the Schur inequality in matrix theory [3] and got λkdk\lambda_{k}\leq d_{k}. For d=1d=1, we can compute this limiting constant as limni(Gn)=2log(ϕ)\lim_{n\to\infty}i(G_{n})=2\log(\phi), where ϕ=(1+5)/2\phi=(1+\sqrt{5})/2 is the golden ratio. When looking at Schur for the sum of the first nn eigenvalues and the Cheeger inequality for the ground state λ1\lambda_{1}, it is natural to conjecture that the linear graph with nn vertices and (n1)(n-1) edges strictly maximizes τ(G)\tau(G) among all connected graphs with nn vertices.

1.7.

Having made more experiments with the limiting density of states in the case d=2d=2 in particular, it is numerically indicative that [4,6][4,6] is the largest gap in the Barycentric spectral measure for d=2d=2. If we take a triangle G=K3G=K_{3} and make Barycentric refinements, then after 22 or more Barycentric refinements, exactly half of the positive eigenvalues are in the interval [0,4][0,4] and half of the positive eigenvalues are in [6,)[6,\infty). It is natural to ask whether there is a gap labeling: the integrated density of states 0x𝑑μ(x)\int_{0}^{x}d\mu(x) is a rational number for xx in a gap of the spectrum. But already in the case d=2d=2 for the relatively large gap containing the point 88, we don’t see an obvious small rational number attached. Our best guess for the integrated density of states is 08𝑑k(x)=137/164\int_{0}^{8}dk(x)=137/164. Gap labeling results are prototyped by p-periodic Jacobi matrices, where the integrated density of states is j/pj/p with integer jj. For classes of non-periodic operators like almost periodic operators, the gap labeling works too. For early sources,see [21, 5, 25, 1]. For a recent development interesting for graph theory is the case of periodic Jacobi matrices on universal covers of leaf-free one-dimensional graphs [20], where one observes only point or absolutely continuous spectrum. In that respect, we have absolutely no idea yet what the spectral type of the Barycentric density of states dkdk is in dimension d2d\geq 2. Observing that the potential U(z)U(z) (the analogue of the Lyapunov exponent in the case of Schrödinger operators) appears to be positive almost everywhere on the support of dkdk if the dimension is larger than 11.

1.8.

In the case of Jacobi matrices, one is interested in the spectral measures of random operators L(ω)L(\omega) parametrized by a point in a probability space Ω\Omega which in the almost periodic case is a compact topological group with Haar measure. We believe that there is a non-Abelian compact topological group in all dimensions. It is hidden for d2d\geq 2 still. There should be Laplacian on this group such that dkddk_{d} is the density of states of a “random” Laplacian. In the case d=1d=1, where things are Abelian and understood, the group 𝒢1\mathcal{G}_{1} is the dyadic group of integers and the limiting operator is the extension of Hf(n)=2f(n+1)f(n1)Hf(n)=2-f(n+1)-f(n-1) from the dense set \mathbb{Z} to its completion 𝒢1\mathcal{G}_{1}. By Fourier theory, HH is diagonalized and unitarily equivalent to H^f(x)=22cos(πx)=4sin2(πx/2)\hat{H}f(x)=2-2\cos(\pi x)=4\sin^{2}(\pi x/2) on the Prüfer group 𝒢^1\hat{\mathcal{G}}_{1} of all dyadic rationals on 𝕋\mathbb{T}. While dk=1/(π4(4x))dk=1/(\pi\sqrt{4(4-x)}) on [0,4][0,4] is clearly absolutely continuous, the operator HH has dense pure point spectrum as an operator on L2(𝒢1)L^{2}(\mathcal{G}_{1}). Unlike in the Pontryagin pair 𝕋\mathbb{Z}\leftrightarrow\mathbb{T} where the Dirac points δx\delta_{x} in 𝕋\mathbb{T} are distributions and not functions in in L2(𝕋)L^{2}(\mathbb{T}), the Dirac points δxinl2(𝒢^1)\delta_{x}inl^{2}(\hat{\mathcal{G}}_{1}) are actual functions as the Prüfer group is discrete. In the Jacobi case, the operator was diagonal on the compact, non-discrete group of the Pontryagin pair, in the dyadic case, the operator is diagonal on the discrete-non-compact case.

2. Introduction

2.1.

Investigating the relation between spectral data of a geometry and observable quantities like topological invariants is a classical theme in geometry [2]. It was popularized by Mark Kac whether one can “hear the geometry” [7]. Historically, a prototype result is Weyl’s law [27] relating the number of Dirichlet eigenvalues nn of the Dirichlet Laplacian on GG below λ\lambda with the volume |G||G| of a domain GG as nCdλd2|G|n\sim C_{d}\lambda^{\frac{d}{2}}|G| with Cd=|Bd|2πdC_{d}=\frac{|B_{d}|}{2\pi^{d}}. Herman Weyl looked in [27] at the planar case and proved n/λn=|G|/(4π)n/\lambda_{n}=|G|/(4\pi) using the explicit eigenvalues λm,n=π2(m2+n2)/|G|\lambda_{m,n}=\pi^{2}(m^{2}+n^{2})/|G| for a square GG. Kac’s question addresses the problem to what degree the spectral data determine space and how large the set of objects is that are isospectral. Because the zeta function encodes the spectrum, this also means to what degree the spectral zeta function determines the geometry.

2.2.

Spectral questions are also related to inverse problems like to find the isospectral set in classes of geometries. While in one-dimensional cases, one can deform geometry in an isospectral way, in higher dimensions, isospectral rigidity appears. Isospectral sets tend to be discrete as isospectral deformations of Riemannian geometries in general do not exist any more [23]. For finite simple graphs, the question of how spectral data relate to a given invariant is largely parallel to the continuum theory [4] and can be investigated experimentally. A typical spectral result is that the pseudo determinant Det(K){\rm Det}(K) of the Kirchhoff Laplacian of a graph is the number of rooted spanning trees and the determinant of the Fredholm modification is the number of rooted spanning forests. We look here at this ratio. In this context, we can also remind about the spectral properties of connection matrices, where the number of positive eigenvalues minus the number of negative eigenvalues is the Euler characteristic [17].

2.3.

The tree forest ratio is picked up here also because it belongs to a larger quest to investigate functionals on graphs, that is assigning numerical values to a graph. Considering critical points of functionals on geometries has been successful in physics, like for geodesics, area minimizing surfaces, total curvature minimizing In graph theory, one can look at “packing numbers” like the chromatic number or the independence number manifolds. In graph theory, quantities that can be counted are the average simplex cardinality [9], the number fkf_{k} of kk-dimensional simplices or the Euler characteristicχ(G)=k(1)kfk\chi(G)=\sum_{k}(-1)^{k}f_{k} or other length-related functionals [13]. Interesting examples of spectral quantities is the ground state energy, the smallest positive eigenvalue of the Laplacian or then the analytic torsion which involves spectral data of the Hodge Laplacian L=(d+d)2L=(d+d^{*})^{2}.

3. Trees and Forests

3.1.

Given a finite simple graph GG, the collection of rooted spanning trees and the set of rooted spanning forests inside GG are of interest. For forests, “rooted” means that every tree in the forest has an assigned vertex, which is interpreted as the root of the tree. Rooted trees in a forest include also seeds, which are one-point graphs. Each rooted tree can be seen as pointed topological space without closed loops nor triangles. It is a directory organization of the vertices and a stratification of space by giving the distance from the root. For any spanning tree TT or spanning forest, the induced graph is the full space GG because TT spans GG: the vertex set of TT and GG is the same. The ratio of rooted spanning forests and trees is the same than the ratio of spanning forests and trees.

3.2.

By the Kirchhoff matrix tree theorem, the number of rooted spanning trees in GG is Det(K){\rm Det}(K), where KK is the Kirchhoff matrix of GG and Det{\rm Det} is the pseudo determinant of KK, the product of the non-zero eigenvalues of KK. Since every spanning tree has nn possible roots if the graph has nn vertices, the number Det(K)/n{\rm Det}(K)/n is the number of spanning trees in GG.

3.3.

By the Chebotarev-Schamis forest theorem [22, 24, 10], the number of rooted spanning forests in GG is det(1+K){\rm det}(1+K), the Fredholm determinant of the Kirchhoff matrix KK. See [12]. We introduce here the tree-forest ratio

τ(G)=det(1+K)Det(K)\tau(G)=\frac{{\rm det}(1+K)}{{\rm Det}(K)}

and the tree forest index

i(G)=log(τ(G))/|G|i(G)=\log(\tau(G))/|G|\;

where |G||G| is the number of vertices. In general τ(G)0\tau(G)\geq 0 because there are more forests than trees.

3.4.

In dimension larger than 11, both the number of forests as well as the number of trees grow exponentially when doing Barycentric refinements. The reason is that in the two dimensional case already, we can see this geometrically. This is different for one-dimensional graphs, where the number of trees grows polynomially under Barycentric subdivisions while the number of forests grows exponentially. The precise polynomial degree of the tree growth rate in the one-dimensional case depends on the genus b1b_{1} of the graph.

3.5.

For graphs with multiple connected components, where spanning trees naturally need to be disconnected, one can define the number of rooted spanning trees as the product kDet(Kk)\prod_{k}{\rm Det}(K_{k}), where KkK_{k} is the Kirchhoff matrix block belonging to the kk’the component. For the number of rooted forests, we naturally have the product kDet(Kk+1)\prod_{k}{\rm Det}(K_{k}+1) already because forests do not need to be connected. With this extension of the definition for not necessarily connected graphs GG, the ratio τ(G)=Det(K+1)/Det(K)\tau(G)={\rm Det}(K+1)/{\rm Det}(K) can in general be understood as the tree-forest ratio, whether GG is a connected or a disconnected graph. To make the tree-forest ratio functional defined on all graphs, we define τ(0)=1\tau(0)=1 and i(0)=0i(0)=0 if 0 is the empty graph.

3.6.

Since all rooted spanning trees are also rooted spanning forests, we have τ(G)1\tau(G)\geq 1 and so i(G)=log(τ(G))/|G|0i(G)=\log(\tau(G))/|G|\geq 0. It is easy to check that τ(G)=1\tau(G)=1 happens if and only if GG has no edges. The reason is that already the presence of one single edge produces more forests than trees. For the complete graph K2K_{2} for example, we have τ(K2)=3/2\tau(K_{2})=3/2 because there are 33 rooted forests and 22 rooted trees in K2K_{2}.

3.7.

When taking graph limits with larger and larger trees, the tree forest ratio diverges in general exponentially fast with the diameter. For the cycle graph CnC_{n} for example, there are n2n^{2} rooted spanning trees and nn spanning trees. In the same circular graph CnC_{n}, the number of spanning forests is the alternate Lucas number L(n)L(n) given by the recursion L(n+1)=3L(n)L(n1)+2,L(0)=0,L(1)=1L(n+1)=3L(n)-L(n-1)+2,L(0)=0,L(1)=1. The number of forests grows exponentially, while the number of trees grows polynomially. We have

τ(Cn)=n2/L(n)\tau(C_{n})=n^{2}/L(n)\;

and

log(τ(Cn))/n2log(ϕ)=arccosh(3/2),\log(\tau(C_{n}))/n\to 2\log(\phi)={\rm arccosh}(3/2)\;,

where ϕ\phi is the golden ratio.

3.8.

For any one-dimensional graph GG with |G|=|E||G|=|E| edges, the Barycentral limit measure dμd\mu exists explicitly. The limiting density of states measure is the arcsin distribution. Potential theoretically, the fact that that the number of trees grows polynomially while the number of forests grows exponentially is reflect in 01log(4sin2(πx))𝑑x=0\int_{0}^{1}\log(4\sin^{2}(\pi x))\;dx=0 which is not totally obvious. In terms of the density of states dk(x)=1πx(4x)dk(x)=\frac{1}{\pi\sqrt{x(4-x)}} we have

U(0)=04log(x)πx(4x)𝑑x=0U(0)=\int_{0}^{4}\frac{\log(x)}{\pi\sqrt{x(4-x)}}\;dx=0\;

and

U(1)=04log(1+x)πx(4x)𝑑x=log(ϕ2)=arccosh(3/2)=0.962424.U(-1)=\int_{0}^{4}\frac{\log(1+x)}{\pi\sqrt{x(4-x)}}\;dx=\log(\phi^{2})={\rm arccosh}(3/2)=0.962424\;.

3.9.

Here is an other class, of graphs where we have a chance to compute the limit. One can look at circulant graphs Cn,AC_{n,A}, the Cayley graph of a generating set AV(Cn)A\subset V(C_{n}). If the diameter of the graph is larger than 2, then the tree-forest ratio is going to explode for such graphs The radius of a circulant graph is 22 if the set A+AA+A in p\mathbb{Z}_{p} covers the entire set p\mathbb{Z}_{p}. An extreme case is if A=V(Cn)A=V(C_{n}), in which case we have the complete graph. An other example is the graph complement CnC_{n}^{\prime} of CnC_{n} in which case the diameter is always 22. If the set A+AA+A for the generating set of the Cayley graph does not cover the graph, meaning that there are relations in the finitely generated group with words 33 or longer, then the tree-forest ratio of GnG_{n} diverges exponentially as nn\to\infty. In some cases, we can compute the limit. For the graph complement C¯n\overline{C}_{n} of cyclic graphs CnC_{n}, which are graphs of diameter 22, one has:

Proposition 1.

τ(C¯n)e\tau(\overline{C}_{n})\to e.

Proof.

The eigenvalues of C¯n\overline{C}_{n} are explicitly known as

λk,n=m=2n21cos(2πmkn)=m=2n22sin2(πmkn).\lambda_{k,n}=\sum_{m=2}^{n-2}1-\cos(2\pi m\frac{k}{n})=\sum_{m=2}^{n-2}2\sin^{2}(\pi m\frac{k}{n})\;.

We have 2λ=52+2cos(2πk/n)sin(k(2m3)π/nsin(kπ/n)2\lambda=5-2+2\cos(2\pi k/n)-\frac{\sin(k(2m-3)\pi/n}{\sin(k\pi/n)}. ∎

Proposition 2.

τ(Kn)e\tau(K_{n})\to e.

Proof.

The non-zero eigenvalues are all nn. We have τ(Kn)=Det(1+K)/Det(K)=(1+1/n)n1\tau(K_{n})={\rm Det}(1+K)/{\rm Det}(K)=(1+1/n)^{n-1}. ∎

4. The Kirchhoff spectral Zeta function

4.1.

The Kirchhoff spectral zeta function of a finite simple graph G=(V,E)G=(V,E) is defined as

ζ(s)=λ01λs,\zeta(s)=\sum_{\lambda\neq 0}\frac{1}{\lambda^{s}}\;,

where λ\lambda ranges over the eigenvalues of the Kirchhoff Laplacian KK of GG. We see immediately from the definitions that

τ(G)=λ0(1+1λ).\tau(G)=\prod_{\lambda\neq 0}(1+\frac{1}{\lambda})\;.

When taking the Barycentral limit, we need to normalize the function, usually by scaling with a factor 1/n1/n, where nn are the number of vertices in the graph.

4.2.

It follows from the definition that we can estimate the spectral gap λ1\lambda_{1}, which is the smallest eigenvalue of KK from below.

Proposition 3.

λ11/(τ(G)1)\lambda_{1}\geq 1/(\tau(G)-1).

Proof.

We have (1+1/λ1)=τ(G)/k2(1+1/λk)τ(G)(1+1/\lambda_{1})=\tau(G)/\prod_{k\geq 2}(1+1/\lambda_{k})\leq\tau(G). So λ11/(τ(G)1)\lambda_{1}\geq 1/(\tau(G)-1). ∎

4.3.

Since h(G)λ2/2h(G)\geq\lambda_{2}/2, where h(G)h(G) is the Cheeger constant we also have

Proposition 4.

h(G)1/2(τ(G)1h(G)\geq 1/2(\tau(G)-1.

4.4.

Also in reverse, a very small spectral gap λ1\lambda_{1} produces a large τ\tau because τ(G)1+1/λ1\tau(G)\geq 1+1/\lambda_{1}.

4.5.

In gteneral, we have for a finite simple graph with (not yet normalized) zeta function:

Proposition 5.

log(τ(G))=k=1(1)k+1ζ(k)k\log(\tau(G))=\sum_{k=1}^{\infty}(-1)^{k+1}\frac{\zeta(k)}{k}.

Proof.

Taking logs gives

log(τ(G))=λ0log(1+1λ).\log(\tau(G))=\sum_{\lambda\neq 0}\log(1+\frac{1}{\lambda})\;.

Using the expansion log(1+x)=xx2/2+x3/3+\log(1+x)=x-x^{2}/2+x^{3}/3+\dots and Fubini, we see

log(τ(G))=ζ(1)ζ(2)/2ζ(3)/3+.\log(\tau(G))=\zeta(1)-\zeta(2)/2-\zeta(3)/3+\dots\;.

4.6.

The Kirchhoff Laplacian KK is one possible choice which comes natural when looking at trees or forests. We could also look at the Hodge Laplacian L=D2L=D^{2} with Dirac operator D=d+dD=d+d^{*} of a graph GG. The later has symmetric spectrum λn,,λ1,0,λ1,λn-\lambda_{n},\dots,-\lambda_{1},0,\lambda_{1},\lambda_{n}. This gives a Hodge zeta function ζL(s)\zeta_{L}(s). Define the Dirac spectral function as

ζD(s)=λ>01λs\zeta_{D}(s)=\sum_{\lambda>0}\frac{1}{\lambda^{s}}

A Dirac tree forest ratio of a graph could now be defined as

τD(G)=Det(1+D)Det(D)\tau_{D}(G)=\frac{{\rm Det}(1+D)}{{\rm Det}(D)}

Because non-zero eigenvalues come in pairs λ,λ\lambda,-\lambda, we have

τD(G)=λ>0(1+1λ)(11λ)=λ>0(11λ2).\tau_{D}(G)=\prod_{\lambda>0}(1+\frac{1}{\lambda})(1-\frac{1}{\lambda})=\prod_{\lambda>0}(1-\frac{1}{\lambda^{2}})\;.

If 11 is an eigenvalue of DD we would exclude that factor. One could also look at the connection Laplacian zeta function.

4.7.

Looking at the zeta function rather than the eigenvalues adds an analytic angle to spectral theory. One can look at a analytic properties of the function ζ(s)\zeta(s) and especially at the places, where ζ(s)\zeta(s) is singular or places where ζ(s)\zeta(s) is zero. The above formula shows that the tree-forest ratio measures a sort of growth rate of ζ(s)\zeta(s) along the real axes.

4.8.

Unlike trees or forests, the spectral zeta function can be considered for manifolds MM, even so the interpretation of trees and forests is no more adequate in the continuum. On compact Riemannian manifolds MM we have an exterior derivative dd and so a Hodge Laplacian L=D2=(d+d)2L=D^{2}=(d+d^{*})^{2}. Define

τ(M)=Det(1+L)/Det(L),\tau(M)={\rm Det}(1+L)/{\rm Det}(L)\;,

where both Det{\rm Det} is the Ray-Singer determinant defined by a zeta regularization of L=(d+d)2L=(d+d^{*})^{2}. One could also look just at the zeta function defined by the scalar Laplacian. We do not see any use yet for the quantity τ(M)\tau(M), the point is just that it can be considered.

4.9.

Let us look for example at the case M=𝕋M=\mathbb{T}, where D=iddxD=i\frac{d}{dx} which has the eigenfunctions einxe^{inx} with eigenvalues n-n. The Dirac zeta function

n>01ns.\sum_{n>0}\frac{1}{n^{s}}\;.

The Laplacian zeta function is n>01n2s\sum_{n>0}\frac{1}{n^{2s}} as the eigenvalues of LL are 1/n21/n^{2}.

4.10.

Proposition 6.

τ(𝕋)=sinh(π)/π\tau(\mathbb{T})=\sinh(\pi)/\pi.

Proof.

We have

τ(M)=n>0(1+1n2)=sinh(π)/π\tau(M)=\prod_{n>0}(1+\frac{1}{n^{2}})=\sinh(\pi)/\pi

using sin(πx)=πxn=1(1x2n2)\sin(\pi x)=\pi x\prod_{n=1}^{\infty}(1-\frac{x^{2}}{n^{2}}) at x=ix=i. ∎

5. The Barycentral limit

5.1.

The quantity T(G)=log(τ(G))T(G)=\log(\tau(G)) is additive with respect to the disjoint union operation T(G+H)=T(G)+T(H)T(G+H)=T(G)+T(H). Let |G||G| denote the number of vertices of GG. Define the tree-forest index

i(G)=log(τ(G))/|G|i(G)=\log(\tau(G))/|G|\;

where |G||G| is the number of vertices. It has a chance to remain finite when taking graph limits. Let GnG_{n} denote the nn’th Barycentric refinement of a graph G=G0G=G_{0}. Now, the growth of the tree forest ratio is dominated by the growth of the forests. The index i(Gn)i(G_{n}) is the difference of the Forest index log(F(Gn))/|Gn|\log(F(G_{n}))/|G_{n}| and the tree index.

Theorem 1 (Tree Forest Universality).

limni(Gn)\lim_{n\to\infty}i(G_{n}) exists for every graph GG and only depends on the maximal dimension dd of GG. For d=1d=1, the limit is 2log(ϕ)2\log(\phi), where ϕ\phi is the golden ratio.

Proof.

We prove this from the Barycental central limit theorem which asserts that there is a (only dimension-dependent measure) μ\mu on [0,)[0,\infty) such that the spectral measures of L(Gn)L(G_{n}) (which are point measures for all nn) converge weakly to μ\mu. Now, we have i(Gn)=0log(1+1/x)𝑑μ(x)i(G_{n})=\int_{0}^{\infty}\log(1+1/x)d\mu(x) which is a difference of two potential values

i(Gn)=0log(1+x)𝑑μ(x)0log(x)𝑑μ(x).i(G_{n})=\int_{0}^{\infty}\log(1+x)d\mu(x)-\int_{0}^{\infty}\log(x)\;d\mu(x)\;.

The 11-dimensional case is special in that dμd\mu has compact support on [0,4][0,4] and is the equilibrium measure on that interval. The potential f(z)=log|az|dμ(z)f(z)=\int\log|a-z|\;d\mu(z) exists for every aa\in\mathbb{Z} and is subharmonic and zero exactly on the interval [0,π][0,\pi]. The limiting index i(G)i(G) in 11-dimension is therefore the potential value at a=1a=-1.
In the d=2d=2 or higher dimensional case, the measure μ\mu does no more have compact support. This is related to the fact that the maximal vertex degree becomes unbounded under Barycentric refinements. While in one dimension, the measure μ\mu was absolutely continuous with a density concentrated mostly on the boundary 0,40,4 of the support, in higher dimensions, the measure has a tail continuity both at the lower bound 0 as well as at infinity. At zero we need a mild decay to make sure that 0log(x)𝑑μ(x)\int_{0}^{\infty}\log(x)\;d\mu(x) is finite. At infinity, we need a stronger decay to make sure that 0log(1+x)𝑑μ(x)\int_{0}^{\infty}\log(1+x)d\mu(x) exists. But we know also from potential theory of subharmonic functions and in our case from the combinatorial context that 00log(x)𝑑μ(x)<0log(1+x)𝑑μ(x)0\leq\int_{0}^{\infty}\log(x)\;d\mu(x)<\int_{0}^{\infty}\log(1+x)d\mu(x). It suffices therefore to show that 0log(1+x)𝑑μ(x)\int_{0}^{\infty}\log(1+x)d\mu(x) exists. This amounts to estimate the decay rate of μ\mu at infinity. The fact that there is an exponential decay in dimension 22 or higher in the next section. We are using an estimate λk2dk\lambda_{k}\leq 2d_{k}, a Perron-Frobenius result on the limiting dimension distribution in Barycentric limits as well as an estimation of the vertex cardinality points of GnG_{n} which depends on the generation of a point. A vertex in GnG_{n} has generation kk if it has become a vertex in GnkG_{n-k} and not before. A vertex in GnG_{n} of generation zero is a vertex which has in the previous level Gn1G_{n-1} not been a vertex. Note that if a simplex is a vertex, it is especially again a vertex in the next generation. What happens is that we have an exponential decay of vertex degrees depending on generation. ∎

5.2.

A different proof could be done along the lines of the Barycentral spectral limit theorem without using the spectral measure dμd\mu: if GG has maximal dimension dd then the Barycentric refinement G1G_{1} of a dd-simplex in GG consists of (d+1)!(d+1)! simplices glued. There is a stratification of growth rates of the different simplices depending on dimension. The number of maximal simplices grows exponentially faster than the number of co-dimension-one simplices, we can asymptotically treat every refined nn-simplex as a copy of disjoint (n+1)!(n+1)! simplices plus a gluing correction which as it is lower dimensional will be a definite factor smaller. The limiting tree forest ratio is again Cauchy sequence as we add up a sequence of errors which geometrically decay.

6. Tail decay

6.1.

In this section we prove a proposition on the limiting density of states dkdk in dimension d2d\geq 2.

Proposition 7.

There exists a positive constant cc such that μ([x,))ecx\mu([x,\infty))\leq e^{-cx}.

6.2.

In dimension d=1d=1, the density of states dkdk has compact support [0,4][0,4] and therefore does not need an exponential decay estimate. The lemma of course still applies. Let therefore GG be a graph of dimension d2d\geq 2 and let GnG_{n} be the Barycentric refinements of GG. The vertices of GnG_{n} are the simplices (complete subgraphs) of Gn1G_{n-1} and two are connected, if one is contained in the other. Every vertex xx in GnG_{n} can be assigned a generation γ(x)\gamma(x) defined to be the integer kk, if it has been a vertex in Gn1,,Gnk+1G_{n-1},\dots,G_{n-k+1}. For example, a vertex has generation 11 if it has been a simplex of positive dimension in Gn1G_{n-1}. A point has generation nn if has been a vertex in GG already. Every vertex in GnG_{n} has generation γ(x){1,2,,n}\gamma(x)\in\{1,2,\dots,n\}.

6.3.

The ff-vector of GG is the vector (f0,f1,,fd)(f_{0},f_{1},\dots,f_{d}), where fkf_{k} are the number of kk-dimensional simplices in GG. There are genn=f0(G0){\rm gen}_{n}=f_{0}(G_{0}) vertices of generation nn.
There are genn1=f0(G1)genn{\rm gen}_{n-1}=f_{0}(G_{1})-{\rm gen}_{n} vertices of generation n1n-1.
There are genn2=f0(G2)genngenn1=f0(G2)f0(G1){\rm gen}_{n-2}=f_{0}(G_{2})-{\rm gen}_{n}-{\rm gen}_{n-1}=f_{0}(G_{2})-f_{0}(G_{1}) vertices of generation n2n-2
There are gen1=f0(Gn)genngenn2..gen2{\rm gen}_{1}=f_{0}(G_{n})-{\rm gen}{n}-{\rm gen}_{n-2}..-gen_{2} vertices of generation 11.
Since f0(Gn)f_{0}(G_{n}) grows exponentially at least like (d+1)!n(d+1)!^{n}, we also have in GnG_{n} at least C(d+1)!nkC(d+1)!^{n-k} vertices of generaton kk.

The Barycentric refinement operator AA is a (d+1)×(d+1)(d+1)\times(d+1) upper triangular matrix defined by f(G1)=Af(G)f(G_{1})=Af(G). It is given as Aij=Stirling(i,j)i!A_{ij}={\rm Stirling}(i,j)i!, involving the Stirling numbers of the second kind. The Perron-Frobenius eigenvector of AA is the probability eigenvector of AA belonging to the largest eigenvalue (d+1)!(d+1)!. For example, in the case d=2d=2, when A=[111026006]A=\left[\begin{array}[]{ccc}1&1&1\\ 0&2&6\\ 0&0&6\\ \end{array}\right], the Perron-Frobenius eigenvector is [1/6,1/2,1/3][1/6,1/2,1/3] to the eigenvalue (d+1)!=6(d+1)!=6. This means that for large nn, the graph GnG_{n} has about 1/61/6th of the simplices which are points, about a half which are edges and about 1/31/3rd which are triangles. In the limit dd\to\infty, the Perron Frobenius probability measures appear converge weakly to a point measure around 0.72150.7215, a fact which we could not prove yet.

Lemma 1.

There is a c1>0c_{1}>0 such that for all k, at least c1nk|Gn|c_{1}^{n-k}|G_{n}| vertices in GnG_{n} are generation kk vertices.

This means that the generation γ\gamma as a random variable on the finite set of vertices of GnG_{n} asymptotically has an exponential distribution. There very few vertices of large generation and very many of small generation. Now we will see that on large generation vertices, there is an upper bound on the eigenvalues which is exponential too in the generation.

6.4.

If S(x)S(x) is the unit sphere of a vertex xx, then the k’th Barycentric refinement Sk(x)S_{k}(x) is the unit sphere of xx in GkG_{k}. Because |Sk(x)||S_{k}(x)| is the vertex degree of xx, the vertex degree of generation kk vertices is at least |Sk(x)|d!k|S_{k}(x)|\sim d!^{k}.

Lemma 2.

There exists c2>0c_{2}>0 such that maximally c1k|Gn|c_{1}^{k}|G_{n}| vertices have degree larger than c2kc_{2}^{k}.

6.5.

It follows from the Gershrogin circle theorem that there less than c1k|Gn|c_{1}^{k}|G_{n}| vertices for which the largest eigenvalue is larger than 2c2k2c_{2}^{k}. Now c1=1/(d+1)!c_{1}=1/(d+1)! and c2=d!c_{2}=d!. Since this implies that the error area decays exponentially for x1x\to 1, was also have the function itself decay exponentially.

7. Questions

7.1.

Which graphs with nn vertices have the maximal tree-forest ratio? The minimum is 11. For small nn we see that the linear graph is the maximum. It is pretty safe to conjecture that the linear graph is always the maximum:

7.2.

What is the average tree-forest ratio on Erdoes-Renyi spaces? We can also look more generally at the potential average on Erdoes-Renyi spaces which produces a density of states measure for each nn and pp. And then we can ask for which choice of pnp_{n} does the density of states average dkn,p(n)dk_{n,p(n)} converge weakly in the limit nn\to\infty.

7.3.

What happens with graph operations like various products or joins? For disjoint unions, we have the product of determinants so that τ(G+H)=τ(G)τ(H)\tau(G+H)=\tau(G)\tau(H). Under graph complements, there is no clear relation yet. Also the Zykov join operation does not produce obvious relations. What happens with the tree-forest ratio when taking Shannon products? We do not see any (at least obvious) relation between the tree number yet.

7.4.

For connection Laplacians, we know that the zeta functions multiply. because the connection Laplacian of a product of graphs tensor. LGH=LGLHL_{G*H}=L_{G}\otimes L_{H}. We therefore have for the connection Laplacians:

log(ρ(GH))=k(1)k+1ζG(k)ζH(k)/k.\log(\rho(G*H))=\sum_{k}(-1)^{k+1}\zeta_{G}(k)\zeta_{H}(k)/k\;.

7.5.

What is the relation between the Hodge Laplacian and connection Laplacian tree-forest ratio? In order to define a tree-forest ratio from the connection Laplacian itself, we have to look at the spectrum of the square. We can for any positive energized complex look at τL(G)=det(L+1)/det(L)\tau_{L}(G)=\det(L+1)/\det(L) which makes sense as LL is always invertible.

7.6.

We should reiterate that in the case d2d\geq 2, we have not even a proof yeat that there are gaps in the spectrum. This should be easy to do as in the case d=2d=2 we see a prominent gap [4,6][4,6] associated to the integrated density of states 1/21/2. While we know now something about the decay of dkdk at infinity, we do not have an asymptotic near 0 or the end of gaps. As for 0, the Schur estimate does not help as as we have always a certain percentage of vertices with minimal vertex degree (d+1)!(d+1)!. That means that we can (using Shur) only prove that there exists a constant cc such that dk([0,x])cxdk([0,x])\leq cx for small xx. As we know U(0)0U(0)\geq 0 due to its relation to the growth rate of trees, we know that U(z)U(z) is finite in a neighborhood of z=0z=0. It is natural to conjecture that the potential U(z)<U(z)<\infty for all zz\in\mathbb{Z}. This would imply that the logarithmic capacity log|zw|dk(z)𝑑k(w)-\int\int\log|z-w|dk(z)dk(w) is finite in all dimensions dd. We know that the value is 0 in the case d=0d=0, where U(z)=0U(z)=0 on the support [0,4][0,4] of dkdk. Having finite logarithmic capacity especially implies that the Lebesgue decomposition of dk=dkpp+dksc+dkacdk=dk_{pp}+dk_{sc}+dk_{ac} has no pure point part.

7.7.

Fourier theory could be an approach to determine the nature of the spectrum dμd\mu. Since the measure μ\mu has exponential decay at infinity, all Fourier coefficients dμ^(t)=0eitx𝑑μ(x)\hat{d\mu}(t)=\int_{0}^{\infty}e^{itx}\;d\mu(x) exist. In order to determine the spectral type, one could also focus on a compact part of the spectrum like for example, [0,4][0,4]. Now, we can look at Fourier series of dμd\mu instead rather than the Fourier transform. For example, by a theorem of Wiener [8], if 1nk=0n1μ^(k)2\frac{1}{n}\sum{k=0}^{n-1}\hat{\mu}(k)^{2} converges to zero for nn\to\infty, then dμd\mu has no point spectrum.

8. Code

8.1.

The following Mathematica code plots the potential function (the analog of the Lyapunov exponent), and the integrated density of states (the analog of the rotation number) as well as the spectrum of the Barycentric measure dkdk in the planar case. The code is optimized for the two-dimensional case so that we do not have to use general costly clique finding algorithms.

T3[s_]:=Module[{v=VertexList[s]},Union[Partition[Flatten[Union[
Table[z=EdgeList[Subgraph[s,Complement[VertexList[
NeighborhoodGraph[s,v[[k]]]],{v[[k]]}]]]; Table[Sort[
{v[[k]],z[[j,1]],z[[j,2]]}],{j,Length[z]}],{k,Length[v]}]]],3]]];
T2[s_]:=Module[{e=EdgeList[s]},
Table[Sort[{e[[k,1]],e[[k,2]]}],{k,Length[e]}]];
T1[s_]:=Module[{v=VertexList[s]},Table[{v[[k]]},{k,Length[v]}]];
Whitney2D[s_]:=Sort[Map[Sort,Union[T3[s],T2[s],T1[s]]]];
GraphFromComplex[G_]:=Module[{n=Length[G],v,e={}},Do[x=Sort[G[[k]]];
If[Length[x]==3,e=Union[e,{x->{x[[1]]},x->{x[[2]]},x->{x[[3]]},
x->Sort[{x[[1]],x[[2]]}],x->Sort[{x[[1]],x[[3]]}],
x->Sort[{x[[2]],x[[3]]}]}]];
If[Length[x]==2,e=Union[e,{x->{x[[1]]},x->{x[[2]]}}]],{k,Length[G]}];
rules=Table[G[[k]]->k,{k,n}]; UndirectedGraph[Graph[e /. rules]]];
Barycentric2D[s_]:=GraphFromComplex[Whitney2D[s]];
Barycentric2D[s_,k_]:=Module[{s1=s},Do[s1=Barycentric2D[s1],{k}];s1];
s=Barycentric2D[CompleteGraph[3],6];
L=Sort[Eigenvalues[1.0*KirchhoffMatrix[s]]];
U[z_]:=Sum[Log[Abs[z-L[[k]]]],{k,Length[L]}]/Length[L];
V[x_]:=Sum[If[L[[k]]<x,1,0],{k,Length[L]}]/Length[L];
f[x_]:={Red,Point[{x,0}]};
W = Graphics[Map[f,L], PlotRange->{{-1,10},{-1,7}}];
Show[{W,Plot[{3*U[x],6*V[x]},{x,-2,10},PlotPoints ->1000]}]
{TreeIndex=U[0],ForestIndex=U[-1],TreeForestIndex=TreeIndex/ForestIndex}

8.2.

Refer to caption
Figure 2. The output of the above code shows the potential Re(U(x+0i)){\rm Re}(U(x+0i)) and the integrated density of states Im(U(x+0i){\rm Im}(U(x+0i) of the complex potential U(z)=log(zw)𝑑k(w)U(z)=\int_{\mathbb{C}}\log(z-w)\;dk(w) of the Barycentric 2-dimensional density of states dkdk. We see the 6th Barycentric refinement G6G_{6} for G=G0=K3G=G_{0}=K_{3}. The prominent gap [4,6][4,6] is very clear numerically, but not theoretically established.
Refer to caption
Figure 3. The 5’th Barycentric refinement G5G_{5} of G0=K3G_{0}=K_{3} has 3937 vertices, 11712 edges and 7776 triangles. The next version G6G_{6} used in the above spectral picture already has 23425=3937+11712+777623425=3937+11712+7776 vertices.

References

  • [1] J. Bellissard, A.Bovier, and J.-M.Ghez. Gap labelling theorems for one dimensional discrete Schrödinger operators. Rev. Math. Phys, 4:1–37, 1992.
  • [2] M. Berger. A Panoramic View of Riemannian Geometry. Springer, 2003.
  • [3] A.E. Brouwer and W.H. Haemers. Spectra of graphs. Springer, 2012.
  • [4] F. Chung. Spectral graph theory, volume 92 of CBMS Reg. Conf. Ser. AMS, 1997.
  • [5] H.L. Cycon, R.G.Froese, W.Kirsch, and B.Simon. Schrödinger Operators—with Application to Quantum Mechanics and Global Geometry. Springer-Verlag, 1987.
  • [6] S. Gershgorin. Über die Abgrenzung der Eigenwerte einer Matrix. Bulletin de l’Academie des Sciences de l’URSS, 6:749–754, 1931.
  • [7] M. Kac. Can one hear the shape of a drum? Amer. Math. Monthly, 73:1–23, 1966.
  • [8] Y. Katznelson. An introduction to harmonic analysis. John Wiley and Sons, Inc, New York, 1968.
  • [9] O. Knill. The average simplex cardinality of a finite abstract simplicial complex. https://arxiv.org/abs/1905.02118, 1999.
  • [10] O. Knill. Counting rooted forests in a network.
    http://arxiv.org/abs/1307.3810, 2013.
  • [11] O. Knill. The zeta function for circular graphs.
    http://arxiv.org/abs/1312.4239, 2013.
  • [12] O. Knill. Cauchy-Binet for pseudo-determinants. Linear Algebra Appl., 459:522–547, 2014.
  • [13] O. Knill. Characteristic length and clustering.
    http://arxiv.org/abs/1410.3173, 2014.
  • [14] O. Knill. The graph spectrum of barycentric refinements.
    http://arxiv.org/abs/1508.02027, 2015.
  • [15] O. Knill. Universality for Barycentric subdivision.
    http://arxiv.org/abs/1509.06092, 2015.
  • [16] O. Knill. On Fredholm determinants in topology.
    https://arxiv.org/abs/1612.08229, 2016.
  • [17] O. Knill. One can hear the Euler characteristic of a simplicial complex.
    https://arxiv.org/abs/1711.09527, 2017.
  • [18] O. Knill. An elementary Dyadic Riemann hypothesis.
    https://arxiv.org/abs/1801.04639, 2018.
  • [19] O. Knill. The energy of a simplicial complex. Linear Algebra and its Applications, 600:96–129, 2020.
  • [20] J. Breuer N. Avni and B. Simon. Periodic jacobi matrices on trees. https://arxiv.org/abs/1911.02612, 2020.
  • [21] L. Pastur and A.Figotin. Spectra of Random and Almost-Periodic Operators, volume 297. Springer-Verlag, Berlin–New York, Grundlehren der mathematischen Wissenschaften edition, 1992.
  • [22] P.Chebotarev and E. Shamis. Matrix forest theorems. arXiv:0602575, 2006.
  • [23] P.v.Moerbeke. About isospectral deformations of discrete Laplacians. In Lecture Notes in Mathematics, volume 755. Springer, 1978.
  • [24] E.V. Shamis P.Yu, Chebotarev. A matrix forest theorem and the measurement of relations in small social groups. Avtomat. i Telemekh., 9:125–137, 1997.
  • [25] J.Lacroix R. Carmona. Spectral Theory of Random Schrödinger Operators. Birkhäuser, 1990.
  • [26] R.S. Varga. Gershgorin and His Circles. Springer Series in Computational Mathematics 36. Springer, 2004.
  • [27] H. Weyl. Asymptotische verteilung der eigenwerte. Nachrichten von der Königlichen Gesellschaft der Wissenschaften, pages 110–117, 1911.