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

The spectral density of dense random networks and the breakdown of the Wigner semicircle law

Fernando L. Metz [email protected] Physics Institute, Federal University of Rio Grande do Sul, 91501-970 Porto Alegre, Brazil London Mathematical Laboratory, 18 Margravine Gardens, London W6 8RH, United Kingdom    Jeferson D. Silva Physics Institute, Federal University of Rio Grande do Sul, 91501-970 Porto Alegre, Brazil
Abstract

Although the spectra of random networks have been studied for a long time, the influence of network topology on the dense limit of network spectra remains poorly understood. By considering the configuration model of networks with four distinct degree distributions, we show that the spectral density of the adjacency matrices of dense random networks is determined by the strength of the degree fluctuations. In particular, the eigenvalue distribution of dense networks with an exponential degree distribution is governed by a simple equation, from which we uncover a logarithmic singularity in the spectral density. We also derive a relation between the fourth moment of the eigenvalue distribution and the variance of the degree distribution, which leads to a sufficient condition for the breakdown of the Wigner semicircle law for dense random networks. Based on the same relation, we propose a classification scheme of the distinct universal behaviours of the spectral density in the dense limit. Our theoretical findings should lead to important insights on the mean-field behaviour of models defined on graphs.

I Introduction

A random network or graph is a collection of nodes joined by edges following a probabilistic rule. Random networks are formidable tools to model large assemblies of interacting units, like neurons in the brain, computers and routers in the Internet, or persons forming a friendship network [1]. Motivated by our increasing ability to collect and process vast amounts of empirical data, the theory of random networks has experienced an enormous progress, leading to important insights in physics, biology, and sociology [2]. The implications of the structure of networks to the dynamical processes occurring on them remains a fundamental topic in network theory [3].

Dynamical processes on a random network are to a large extent governed by the spectrum of the corresponding adjacency random matrix. This is a random matrix where each entry equals the strength of the interaction between a pair of nodes. The study of a broad range of problems amounts to linearize a large set of differential equations [4, 5, 6], coupled through a random network, around a stationary state, whose stability and transient dynamics are ultimately determined by the eigenvalue distribution of the adjacency matrix. Examples in this context are the study of the epidemic threshold for the spreading of diseases [7], the synchronization transition in networks of coupled oscillators [8, 9], and the functional stability of large biological systems, such as gene regulatory networks [10, 11], ecosystems [12, 13, 14], and neural networks[15, 16, 17]. The spectrum of the adjacency matrix also contains information about the network structure, since the trace of the adjacency matrix raised to a certain power yields the number of network loops of a given length [18]. Therefore, the problem of how the network architecture influences the spectrum of the adjacency matrix has attracted considerable research efforts [19, 20, 21, 22, 23, 24, 4, 25].

Synthetic models of random networks provide a controlled way to study the role of the network structure on the spectrum of the adjacency matrix. The degree sequence is the most basic tool to characterize the graph structure [1]. The degree of a node counts the number of edges attached to the node, while the degree sequence specifies all network degrees. When a network has an infinitely large number of constituents, it is natural to consider the degree distribution, i.e. the fraction of nodes with a certain degree, instead of dealing with the degree sequence. The configuration model stands out as one of the most fundamental and versatile models of random graphs [26, 27, 1, 28], since it enables the degree distribution to be freely specified, while keeping the pattern of interconnections entirely random. From a practical viewpoint, the configuration model resolves one of the main shortcomings of Erdös- Rényi random graphs [29, 30, 31], namely its Poisson degree distribution, which has little resemblance with the long-tailed degree distributions found in empirical networks [32, 33]. The priceless possibility to fix the degree distribution of a random graph is not only useful to model the structure of empirical networks, but it offers the ideal setting to examine the impact of degree fluctuations on the spectrum of the adjacency matrix.

There has been a significant amount of numeric and analytic work on the spectral properties of adjacency random matrices [34, 19, 35, 20, 21, 36, 37, 38, 22, 24, 23, 39, 40, 4, 25, 41]. A remarkable analytic result is the set of exact and mathematically rigorous equations determining the eigenvalue distribution of the configuration model [36, 37, 38, 42]. These equations form a natural starting point to study the impact of degree fluctuations on network spectra. Unfortunately, aside from a few particular cases [42, 39, 24], these equations can be analytically solved only in the dense limit [36, 37], when the mean degree becomes infinitely large and random networks approach a fully-connected structure.

The analytic solution for the dense limit of the aforementioned equations is at the root of perturbative and non-perturbative approximations for the eigenvalue distribution of large-degree random networks [34, 43, 20]. With the exception of graphs with a power-law degree distribution [19, 35, 21], whose moments are divergent, one expects that the eigenvalue distribution of undirected random networks converges, in the dense limit, to the Wigner semicircle distribution of random matrix theory, reflecting a high level of universality. This expectation has been confirmed numerically [19] and analytically [36, 37] for Erdös-Rényi and regular random graphs. However, the recent work [44] has rigorously shown that the spectrum of the configuration model does depend on the degree distribution in the dense limit. The main theorem in [44] also provides an approach to compute the spectral density for specific degree distributions. In spite of these rigorous results, a more detailed understanding of how the structure of a dense random network influences its spectrum, and the universal status of the Wigner semicircle law, is still lacking.

In this work we study the dense limit of the eigenvalue distribution of undirected networks drawn from the configuration model. We analyze four examples of degree distributions in which all moments are finite and the variances scale differently with the average degree, and we show that the dense limit of the eigenvalue distribution is determined by the degree fluctuations. In particular, we derive an equation yielding the eigenvalue distribution of dense random graphs with an exponential degree distribution [28], from which we unveil the existence of a logarithmic divergence in the spectral density and the absence of sharp spectral edges, i.e., the eigenvalue distribution of exponential random graphs is supported on the entire real line. These are remarkable differences with respect to the Wigner semicircle distribution of random matrix theory [45]. We also discuss how the analytic equations for the spectral density can be derived in two different ways: from the dense limit of the exact resolvent equations in [36, 37, 42], and from the main theorem proved in [44]. Based on an exact calculation of the fourth moment of the eigenvalue distribution, we obtain a sufficient condition for the breakdown of the Wigner semicircle law. This condition is given only in terms of the variance of the degree distribution. We finish the paper by proposing a classification of the different universal behaviours of the eigenvalue distribution of the configuration model in the dense limit, based on an exponent characterizing how the variance of the degree distribution diverges for increasing mean degree.

Our paper is organized as follows. In the next section we define the adjacency matrix of random networks and the spectral density. Section III introduces the configuration model and defines the degree distributions studied in this work. In section IV, we present numeric and analytic results for the dense limit of the spectral density for each example of degree distribution. Section IV explains how the analytic equations for the spectral density are obtained from the dense limit of the resolvent equations, while in section V we derive these equations from the main theorem in reference [44]. The condition for the breakdown of the Wigner semicircle law and the classification of the different universal behaviours are discussed in section VI. We summarize our results and conclude in section VII.

II The adjacency matrix of random networks

Let NN be the total number of network nodes. The set of binary random variables {Cij}i,j=1,,N\{C_{ij}\}_{i,j=1,\dots,N} specifies the network structure: if Cij=1C_{ij}=1, there is an undirected link iji\leftrightarrow j between nodes ii and jj, while Cij=0C_{ij}=0 means this link is absent. We also associate a random weight JijJ_{ij}\in\mathbb{R} to each edge iji\leftrightarrow j, which accounts for the interaction strength between nodes ii and jj. We consider undirected simple random networks that have no self-edges nor multi-edges, in which Cij=CjiC_{ij}=C_{ji}, Jij=JjiJ_{ij}=J_{ji}, and Cii=0C_{ii}=0.

The degree Ki=j=1NCijK_{i}=\sum_{j=1}^{N}C_{ij} of node ii is a random variable that counts the number of nodes attached to ii, while the degree sequence K1,,KNK_{1},\dots,K_{N} provides global information on the fluctuations of the network connectivity. In the limit NN\rightarrow\infty, it is more convenient to work with the degree distribution

pk=limN1Ni=1Nδk,Ki,p_{k}=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{i=1}^{N}\delta_{k,K_{i}}, (1)

where δ\delta is the Kronecker symbol. The average degree cc and the variance σ2\sigma^{2} of pkp_{k} read

c=k=0kpk,σ2=k=0pk(kc)2.c=\sum_{k=0}^{\infty}kp_{k},\qquad\sigma^{2}=\sum_{k=0}^{\infty}p_{k}\left(k-c\right)^{2}.

The degree distribution pkp_{k} is one of the primary quantities to characterize the structure of random networks in the limit NN\rightarrow\infty.

We will study the eigenvalue distribution of the N×NN\times N symmetric adjacency matrix 𝑨\bm{A}, with elements defined as

Aij=CijJijc.A_{ij}=\frac{C_{ij}J_{ij}}{\sqrt{c}}. (2)

The adjacency matrix fully encodes the network structure, along with the coupling strengths between adjacent nodes. The empirical spectral measure of 𝑨\bm{A} is given by

ρN(λ)=1Nα=1Nδ(λλα),\rho_{N}(\lambda)=\frac{1}{N}\sum_{\alpha=1}^{N}\delta\left(\lambda-\lambda_{\alpha}\right), (3)

where λ1,,λN\lambda_{1},\dots,\lambda_{N} are the (real) eigenvalues of 𝑨\bm{A}. By introducing the N×NN\times N resolvent matrix

𝑮(z)=(z𝑨)1,\bm{G}(z)=\left(z-\bm{A}\right)^{-1}, (4)

with z=λiηz=\lambda-i\eta on the lower half complex plane, the empirical spectral measure follows from the diagonal elements of 𝑮(z)\bm{G}(z)

ρN(λ)=1πNlimη0+i=1NImGii(z).\rho_{N}(\lambda)=\frac{1}{\pi N}\lim_{\eta\rightarrow 0^{+}}\sum_{i=1}^{N}{\rm Im}G_{ii}(z). (5)

In the limit NN\rightarrow\infty, the empirical mean of ImGii(z){\rm Im}G_{ii}(z) normally converges to its ensemble averaged value, obtained from the distribution of 𝑨\bm{A}, which implies that ρ(λ)=limNρN(λ)\rho(\lambda)=\lim_{N\rightarrow\infty}\rho_{N}(\lambda) is well-defined. Here we will study the spectral density ρ(λ)\rho(\lambda) when the average degree grows to infinity, hence the limit cc\rightarrow\infty is performed after NN\rightarrow\infty. The scaling of the elements AijA_{ij} with the average degree cc in Eq. (2) ensures the spectral density has a finite variance when cc\rightarrow\infty.

III The configuration model of networks

We study random networks with arbitrary degree distributions, known as the configuration model of networks [26, 27, 28, 1], where pkp_{k} is specified at the outset. A single instance of the adjacency matrix of the configuration model is created as follows. First, the degrees K1,,KNK_{1},\dots,K_{N} are drawn independently from a common distribution pkp_{k}. After assigning KiK_{i} stubs of edges to each node ii (i=1,,Ni=1,\dots,N), a pair of stubs is uniformly chosen at random and then connected to form an edge. This last step is repeated on the remainder stubs until there are no stubs left, and the outcome is a particular matching of the stubs with the prescribed random degrees K1,,KNK_{1},\dots,K_{N}. We do not allow for the existence of self-edges and multi-edges in the random network. We set Aij=Jij/cA_{ij}=J_{ij}/\sqrt{c} if there is an edge connecting nodes ii and jj, and Aij=0A_{ij}=0 otherwise. The coupling strengths {Jij}i,j=1,,N\{J_{ij}\}_{i,j=1,\dots,N} are i.i.d. random variables drawn from a common distribution PJP_{J}. We refer to [1] for other properties and subtleties of the configuration model. In the limit NN\rightarrow\infty, the ensemble of adjacency random matrices 𝑨\bm{A} constructed from this procedure is specified by the degree distribution pkp_{k} and the distribution of weights PJP_{J}. This is probably the simplest network model that allows to clearly exploit the influence of degree fluctuations on the spectral properties of 𝑨\bm{A}.

We will present results for regular, Poisson, exponential, and Borel degree distributions. The analytic expression for pkp_{k} in each case, together with the variance σ2\sigma^{2}, is displayed in table 1. The properties of the configuration model with Poisson and exponential degree distributions have been extensively discussed in [28, 1]. The degree distributions in table 1 obey the following properties: (i) pkp_{k} is parametrized solely in terms of the mean degree cc; (ii) all moments of pkp_{k} are finite for c<c<\infty; (iii) for sufficiently large cc, the variances obey σreg2<σpois2<σexp2<σbor2\sigma_{\rm reg}^{2}<\sigma_{\rm pois}^{2}<\sigma_{\rm exp}^{2}<\sigma_{\rm bor}^{2}. These four examples of degree distributions are very convenient to study, in a controllable way, the effect of degree fluctuations on random networks as we increase cc. Indeed, we can compare for c1c\gg 1 the spectral density of networks with the same average degree, but with increasing variances σ2\sigma^{2}. We do not consider here random networks with power-law degree distributions [32, 33], since in this case higher-order moments of pkp_{k} diverge already for finite cc.

Regular Poisson Exponential Borel
pkp_{k} δk,c\delta_{k,c} ecckk!\frac{e^{-c}c^{k}}{k!} 1c+1(cc+1)k\frac{1}{c+1}\left(\frac{c}{c+1}\right)^{k} e(c1)ckk![k(c1)c]k1\frac{e^{-\frac{(c-1)}{c}k}}{k!}\left[\frac{k(c-1)}{c}\right]^{k-1}
σ2\sigma^{2} 0 cc c2+cc^{2}+c c3c2c^{3}-c^{2}
Table 1: The analytic expression for the degree distribution pkp_{k} and the corresponding variance σ2\sigma^{2} in the case of regular, Poisson, exponential, and Borel random networks. The Borel degree distribution is defined for c>1c>1, while the other three distributions are defined for c>0c>0. The Borel degree distribution is supported in k1k\geq 1, the degree distribution of regular graphs is supported at k=ck=c, and both Poisson and exponential degree distributions are supported in k0k\geq 0.

Although the Borel degree distribution is not commonly employed in the study of random networks, the rate of its exponential decay is smaller than the one of the exponential degree distribution with the same cc, which makes the Borel distribution well-suited to our analysis. The Borel distribution, introduced in the context of queuing theory [46, 47], also appears as the distribution of the total progeny in a Galton-Watson branching process with Poisson distributed degrees [48]. Figure 1 illustrates the Borel degree distribution for different average degrees. For fixed kck\ll c, the Borel distribution pkp_{k} attains a finite limit as cc\rightarrow\infty, in contrast to the Poisson and the exponential degree distribution. For k=O(c)k=O(c), the Borel distribution pkp_{k} is proportional to 1/c1/\sqrt{c} for c1c\gg 1. Therefore, Borel networks with large cc contain a finite fraction of nodes with cc-independent degrees and a smaller fraction of nodes with degrees proportional to cc. Note also that pkp_{k} in figure 1 is highly skewed, its mode is located at k=1k=1, and the tail of pkp_{k} decays as lnpk(1c+ln(c1c))k\ln p_{k}\sim\left(\frac{1}{c}+\ln\left(\frac{c-1}{c}\right)\right)k for k1k\gg 1. All network models considered here have exponentially decaying degree distributions.

Refer to caption
Figure 1: The Borel degree distribution pkp_{k} (see table 1) for different values of the average degree cc. The inset depicts the exponential tail of pkp_{k} for large kk. The values of cc in the inset are c=40c=40 (solid line), c=20c=20 (dotted line), c=10c=10 (dashed line), and c=5c=5 (dot-dashed line).

IV The dense limit of the spectral density

The configuration model of networks has a key property: in the limit NN\rightarrow\infty, the set of nodes in the neighborhood of a given node, drawn uniformly from the network, is arranged in a tree-like structure [42]. Nevertheless, the topology of random networks is fundamentally different from a Cayley tree [49], in the sense that boundary nodes are absent from the configuration model and cycles do survive in the limit NN\rightarrow\infty, but their average length scales as lnN\ln N for N1N\gg 1. Roughly speaking, a random network can be seen as a Cayley tree with fluctuating degrees, wrapped onto itself.

In the limit NN\rightarrow\infty, the spectral density is obtained from

ρ(λ)=1πlimη0+Img0d2g𝒫(g)Img,\rho(\lambda)=\frac{1}{\pi}\lim_{\eta\rightarrow 0^{+}}\int_{{\rm Im}g\geq 0}d^{2}g\mathcal{P}(g){\rm Im}g, (6)

where d2gdRegdImgd^{2}g\equiv d{\rm Re}g\,d{\rm Im}g, and 𝒫(g)\mathcal{P}(g) is the joint distribution of the real and imaginary parts of Gii(z)G_{ii}(z). The symbol Img0\int_{{\rm Im}g\geq 0} represents an integral over gg\in\mathbb{C} restricted to the upper half complex plane. The distribution of the resolvent follows from the solution of the coupled equations [36, 37]

𝒫(g)\displaystyle\mathcal{P}(g) =\displaystyle= k=0pkImg0[=1kd2g𝒫cav(g)]\displaystyle\sum_{k=0}^{\infty}p_{k}\int_{{\rm Im}g\geq 0}\left[\prod_{\ell=1}^{k}d^{2}g_{\ell}\mathcal{P}_{{\rm cav}}(g_{\ell})\right] (7)
×\displaystyle\times δ[g(1z1c=1kJ2g)]J1,,Jk,\displaystyle\left\langle\delta\left[g-\left(\frac{1}{z-\frac{1}{c}\sum_{\ell=1}^{k}J_{\ell}^{2}g_{\ell}}\right)\right]\right\rangle_{J_{1},\dots,J_{k}},
𝒫cav(g)\displaystyle\mathcal{P}_{\rm cav}(g) =\displaystyle= k=0kpkcImg0[=1k1d2g𝒫cav(g)]\displaystyle\sum_{k=0}^{\infty}\frac{kp_{k}}{c}\int_{{\rm Im}g\geq 0}\left[\prod_{\ell=1}^{k-1}d^{2}g_{\ell}\mathcal{P}_{{\rm cav}}(g_{\ell})\right] (8)
×\displaystyle\times δ[g(1z1c=1k1J2g)]J1,,Jk1,\displaystyle\left\langle\delta\left[g-\left(\frac{1}{z-\frac{1}{c}\sum_{\ell=1}^{k-1}J_{\ell}^{2}g_{\ell}}\right)\right]\right\rangle_{J_{1},\dots,J_{k-1}},

where J1,,JL\langle\dots\rangle_{J_{1},\dots,J_{L}} denotes the average over the independent interaction strengths J1,,JLJ_{1},\dots,J_{L}, distributed according to PJP_{J}. The quantity 𝒫cav(g)\mathcal{P}_{\rm cav}(g) is the joint distribution of the real and imaginary parts of the resolvent elements Gii(j)(z)G_{ii}^{(j)}(z) on the cavity graph [36, 50], defined as the graph where an arbitrary node jj and all its edges have been removed. Equations (7) and (8) have been derived using both the cavity [36, 50] and the replica method [37] of disordered systems, based on the locally tree-like structure of the configuration model. Equations (6-8) have been rigorously proven in [42] and they are exact in the limit NN\rightarrow\infty, constituting an interesting point of departure to study the dense limit cc\rightarrow\infty of the spectral density ρ(λ)\rho(\lambda).

Let F(G)F(G) be an arbitrary function of the complex random variable GG distributed as 𝒫cav(g)\mathcal{P}_{\rm cav}(g). By defining the average of F(G)F(G)

F(G)=Img0d2g𝒫cav(g)F(g),\langle F(G)\rangle=\int_{{\rm Im}g\geq 0}d^{2}g\mathcal{P}_{\rm cav}(g)F(g),

we can use Eq. (8) and rewrite the above expression as

F(G)=Imq0d2q𝒲cav(q)F(1zq),\langle F(G)\rangle=\int_{{\rm Im}q\geq 0}d^{2}q\mathcal{W}_{\rm cav}(q)F\left(\frac{1}{z-q}\right), (9)

where we introduced the distribution of Q1c=1K1J2GQ^{\prime}\equiv\frac{1}{c}\sum_{\ell=1}^{K-1}J_{\ell}^{2}G_{\ell}

𝒲cav(q)\displaystyle\mathcal{W}_{\rm cav}(q) =\displaystyle= k=0kpkcImg0[=1k1d2g𝒫cav(g)]\displaystyle\sum_{k=0}^{\infty}\frac{kp_{k}}{c}\int_{{\rm Im}g\geq 0}\left[\prod_{\ell=1}^{k-1}d^{2}g_{\ell}\mathcal{P}_{\rm cav}(g_{\ell})\right] (10)
×\displaystyle\times δ(q1c=1k1J2g)J1,,Jk1,\displaystyle\left\langle\delta\left(q-\frac{1}{c}\sum_{\ell=1}^{k-1}J_{\ell}^{2}g_{\ell}\right)\right\rangle_{J_{1},\dots,J_{k-1}},

with the random variable KK denoting the degree of an uniformly chosen node. Following an analogous procedure, we can also rewrite the spectral density in terms of the distribution 𝒲(q)\mathcal{W}(q) of the random variable Q1c=1KJ2GQ\equiv\frac{1}{c}\sum_{\ell=1}^{K}J_{\ell}^{2}G_{\ell}

ρ(λ)=1πlimη0+Imq0d2q𝒲(q)Im(1zq),\rho(\lambda)=\frac{1}{\pi}\lim_{\eta\rightarrow 0^{+}}\int_{{\rm Im}q\geq 0}d^{2}q\mathcal{W}(q){\rm Im}\left(\frac{1}{z-q}\right), (11)

defined as

𝒲(q)\displaystyle\mathcal{W}(q) =\displaystyle= k=0pkImg0[=1kd2g𝒫cav(g)]\displaystyle\sum_{k=0}^{\infty}p_{k}\int_{{\rm Im}g\geq 0}\left[\prod_{\ell=1}^{k}d^{2}g_{\ell}\mathcal{P}_{\rm cav}(g_{\ell})\right] (12)
×\displaystyle\times δ(q1c=1kJ2g)J1,,Jk.\displaystyle\left\langle\delta\left(q-\frac{1}{c}\sum_{\ell=1}^{k}J_{\ell}^{2}g_{\ell}\right)\right\rangle_{J_{1},\dots,J_{k}}.

We note that 𝒲(q)\mathcal{W}(q) and 𝒲cav(q)\mathcal{W}_{\rm cav}(q) are distributions of sums of independent complex random variables containing a random and large number of terms. For the examples of pkp_{k} in table 1, one can show that, in the limit cc\rightarrow\infty, the number of summands in QQ and QQ^{\prime} diverges. Thus, instead of working with the distributions of the resolvent, it is more convenient to extract the cc\rightarrow\infty limit of 𝒲(q)\mathcal{W}(q) and 𝒲cav(q)\mathcal{W}_{\rm cav}(q). Let us introduce the characteristic functions φ(p,t)\varphi(p,t) and φcav(p,t)\varphi_{\rm cav}(p,t) of, respectively, 𝒲(q)\mathcal{W}(q) and 𝒲cav(q)\mathcal{W}_{\rm cav}(q)

φ(p,t)=k=0pkexp[kSc(p,t)],\varphi(p,t)=\sum_{k=0}^{\infty}p_{k}\exp{\left[kS_{c}(p,t)\right]}, (13)
φcav(p,t)=k=0kpkcexp[(k1)Sc(p,t)],\varphi_{\rm cav}(p,t)=\sum_{k=0}^{\infty}\frac{kp_{k}}{c}\exp{\left[(k-1)S_{c}(p,t)\right]}, (14)

with

Sc(p,t)=ln[Img0d2g𝒫cav(g)dxPJ(x)\displaystyle S_{c}(p,t)=\ln\Big{[}\int_{{\rm Im}g\geq 0}d^{2}g\mathcal{P}_{\rm cav}(g)\int_{-\infty}^{\infty}dxP_{J}(x)
×exp(ipx2Regcitx2Imgc)].\displaystyle\times\exp{\left(-\frac{ipx^{2}{\rm Re}g}{c}-\frac{itx^{2}{\rm Im}g}{c}\right)}\Big{]}. (15)

In order to study the dense limit cc\rightarrow\infty, we expand Sc(p,t)S_{c}(p,t) in powers of 1/c1/c, keeping in mind that the moments of 𝒫cav(g)\mathcal{P}_{\rm cav}(g) depend on cc. The leading term

S(p,t)=iJ2Jc(pReG+tImG)S_{\infty}(p,t)=-\frac{i\langle J^{2}\rangle_{J}}{c}\left(p\,{\rm Re}\langle G\rangle_{\infty}+t\,{\rm Im}\langle G\rangle_{\infty}\right) (16)

should yield the dense limit cc\rightarrow\infty of the spectral density. Note that we assumed G\langle G\rangle converges to a well-defined limit G\langle G\rangle_{\infty} when cc\rightarrow\infty. In order to proceed further, we specify the degree distribution pkp_{k} in Eqs. (13) and (14).

IV.1 Regular and Poisson random graphs

Although it is well-known that the dense limit of ρ(λ)\rho(\lambda) converges to the semicircle distribution for regular and Poisson random graphs, it is instructive to illustrate our approach for these simple models, characterized by highly peaked degree distributions around the mean value cc. Substituting the explicit forms of pkp_{k} (see table 1) in Eqs. (13) and (14), and using the asymptotic behaviour of Eq. (16), we obtain

φ(p,t)=φcav(p,t)=eiJ2J(pReG+tImG),\varphi(p,t)=\varphi_{\rm cav}(p,t)=e^{-i\langle J^{2}\rangle_{J}\left(p\,{\rm Re}\langle G\rangle_{\infty}+t\,{\rm Im}\langle G\rangle_{\infty}\right)}, (17)

which promptly leads to the Dirac delta distribution in the complex plane

𝒲(q)=𝒲cav(q)=δ(qJ2JG).\mathcal{W}(q)=\mathcal{W}_{\rm cav}(q)=\delta\left(q-\langle J^{2}\rangle_{J}\langle G\rangle_{\infty}\right). (18)

The fact that the resolvent statistics of regular and Poisson random graphs are both described by the same characteristic function, Eq. (17), already demonstrates that these models exhibit the same universal behaviour for cc\rightarrow\infty. The analytic expression for 𝒲cav(q)\mathcal{W}_{\rm cav}(q) allows to determine G\langle G\rangle_{\infty} through Eq. (9), which fulfils

G=1zJ2JG,\langle G\rangle_{\infty}=\frac{1}{z-\langle J^{2}\rangle_{J}\langle G\rangle_{\infty}}, (19)

while the spectral density simply follows from Eq. (11)

ρ(λ)=1πlimη0+Im(1zJ2JG).\rho(\lambda)=\frac{1}{\pi}\lim_{\eta\rightarrow 0^{+}}{\rm Im}\left(\frac{1}{z-\langle J^{2}\rangle_{J}\langle G\rangle_{\infty}}\right).

By solving the quadratic equation (19), we recover the Wigner semicircle law for the Gaussian ensembles of random matrix theory [45]

ρw(λ)={12πJ2J4J2Jλ2,if |λ|<2J2J0,if |λ|2J2J.\rho_{\rm w}(\lambda)=\begin{cases}\frac{1}{2\pi\langle J^{2}\rangle_{J}}\sqrt{4\langle J^{2}\rangle_{J}-\lambda^{2}},&\text{if }|\lambda|<2\sqrt{\langle J^{2}\rangle_{J}}\\ 0,&\text{if }|\lambda|\geq 2\sqrt{\langle J^{2}\rangle_{J}}.\end{cases} (20)

Figure 2 compares Eq. (20) with numerical results obtained from diagonalizing large adjacency matrices 𝑨\bm{A} with average degree c=100c=100. The correspondence between the numerical data and the theoretical expression is excellent for both regular and Poisson random graphs.

Refer to caption
Figure 2: The dense limit cc\rightarrow\infty of the spectral density of regular and Poisson random networks with coupling strengths drawn from PJ(x)=δ(x1)P_{J}(x)=\delta(x-1). The solid line represents the Wigner semicircle distribution, Eq. (20), and the symbols are numerical diagonalization results obtained from 100100 samples of 104×10410^{4}\times 10^{4} adjacency random matrices with c=100c=100.

IV.2 Exponential random graphs

In this subsection we consider the dense limit of exponential random graphs, for which pkp_{k} decays slower than Poisson graphs for k1k\gg 1 (see table 1). Exponential degree distributions have been observed in some examples of empirical networks, such as the North American power grid [51] and the neural network of the worm C. Elegans [52].

Inserting the expressions for S(p,t)S_{\infty}(p,t) and pkp_{k} in Eqs. (13-14) and taking the limit cc\rightarrow\infty, we obtain the asymptotic forms

φ(p,t)\displaystyle\varphi(p,t) =11+ipJ2JReG+itJ2JImG,\displaystyle=\frac{1}{1+ip\langle J^{2}\rangle_{J}{\rm Re}\langle G\rangle_{\infty}+it\langle J^{2}\rangle_{J}{\rm Im}\langle G\rangle_{\infty}},
φcav(p,t)\displaystyle\varphi_{\rm cav}(p,t) =1[1+ipJ2JReG+itJ2JImG]2,\displaystyle=\frac{1}{\left[1+ip\langle J^{2}\rangle_{J}{\rm Re}\langle G\rangle_{\infty}+it\langle J^{2}\rangle_{J}{\rm Im}\langle G\rangle_{\infty}\right]^{2}},

which already show that the dense limit of ρ(λ)\rho(\lambda) is not described by the Wigner semicircle law in this case. The distributions 𝒲(q)\mathcal{W}(q) and 𝒲cav(q)\mathcal{W}_{\rm cav}(q) are the Fourier transforms of their characteristic functions. Defining

𝒦(q,ϵ)\displaystyle\mathcal{K}(q,\epsilon) =dpdt4π2exp(ipReq+itImq)\displaystyle=\int_{-\infty}^{\infty}\frac{dp\,dt}{4\pi^{2}}\exp{\left(ip\,{\rm Re}q+it\,{\rm Im}q\right)}
×[ϵ+ipJ2JReG+itJ2JImG]1,\displaystyle\times\left[\epsilon+ip\langle J^{2}\rangle_{J}{\rm Re}\langle G\rangle_{\infty}+it\langle J^{2}\rangle_{J}{\rm Im}\langle G\rangle_{\infty}\right]^{-1}, (21)

with ϵ1\epsilon\geq 1 an auxiliary parameter, 𝒲(q)\mathcal{W}(q) and 𝒲cav(q)\mathcal{W}_{\rm cav}(q) are obtained from

𝒲(q)=𝒦(q,ϵ=1),𝒲cav(q)=𝒦(q,ϵ)ϵ|ϵ=1.\mathcal{W}(q)=\mathcal{K}(q,\epsilon=1),\qquad\mathcal{W}_{\rm cav}(q)=-\frac{\partial\mathcal{K}(q,\epsilon)}{\partial\epsilon}\Bigg{|}_{\epsilon=1}. (22)

The solution of the integral in Eq. (21) is given by

𝒦(q,ϵ)\displaystyle\mathcal{K}(q,\epsilon) =Θ(Imq)J2JImGδ(ReqImqReGImG)\displaystyle=\frac{\Theta\left({\rm Im}q\right)}{\langle J^{2}\rangle_{J}{\rm Im}\langle G\rangle_{\infty}}\delta\left({\rm Re}q-{\rm Im}q\frac{{\rm Re}\langle G\rangle_{\infty}}{{\rm Im}\langle G\rangle_{\infty}}\right)
×exp(ϵImqJ2JImG),\displaystyle\times\exp{\left(-\frac{\epsilon\,{\rm Im}q}{\langle J^{2}\rangle_{J}{\rm Im}\langle G\rangle_{\infty}}\right)},

leading to the analytic expressions

𝒲(q)\displaystyle\mathcal{W}(q) =Θ(Imq)J2JImGδ(ReqImqReGImG)\displaystyle=\frac{\Theta\left({\rm Im}q\right)}{\langle J^{2}\rangle_{J}{\rm Im}\langle G\rangle_{\infty}}\delta\left({\rm Re}q-{\rm Im}q\frac{{\rm Re}\langle G\rangle_{\infty}}{{\rm Im}\langle G\rangle_{\infty}}\right)
×exp(ImqJ2JImG),\displaystyle\times\exp{\left(-\frac{{\rm Im}q}{\langle J^{2}\rangle_{J}{\rm Im}\langle G\rangle_{\infty}}\right)}, (23)
𝒲cav(q)\displaystyle\mathcal{W}_{\rm cav}(q) =Θ(Imq)Imq[J2JImG]2δ(ReqImqReGImG)\displaystyle=\frac{\Theta\left({\rm Im}q\right){\rm Im}q}{\left[\langle J^{2}\rangle_{J}{\rm Im}\langle G\rangle_{\infty}\right]^{2}}\delta\left({\rm Re}q-{\rm Im}q\frac{{\rm Re}\langle G\rangle_{\infty}}{{\rm Im}\langle G\rangle_{\infty}}\right)
×exp(ImqJ2JImG),\displaystyle\times\exp{\left(-\frac{{\rm Im}q}{\langle J^{2}\rangle_{J}{\rm Im}\langle G\rangle_{\infty}}\right)}, (24)

where Θ()\Theta(\dots) is the Heaviside step function.

As in the previous subsection, Eqs. (23) and (24) depend on G\langle G\rangle_{\infty}, but this quantity can be determined through Eq. (9). Substituting 𝒲cav(q)\mathcal{W}_{\rm cav}(q) in (9) and calculating the integral, we get

G\displaystyle\langle G\rangle_{\infty} =zJ2J2G2exp(zJ2JG)\displaystyle=\frac{z}{\langle J^{2}\rangle_{J}^{2}\langle G\rangle_{\infty}^{2}}\exp{\left(-\frac{z}{\langle J^{2}\rangle_{J}\langle G\rangle_{\infty}}\right)}
×Ei(zJ2JG)1J2JG,\displaystyle\times{\rm Ei}\left(\frac{z}{\langle J^{2}\rangle_{J}\langle G\rangle_{\infty}}\right)-\frac{1}{\langle J^{2}\rangle_{J}\langle G\rangle_{\infty}}, (25)

with Ei(){\rm Ei}(\dots) denoting the complex exponential integral function [53]. The solutions of the above equation yield the dense limit of the averaged resolvent G\langle G\rangle_{\infty} on the cavity graph [36, 50]. The expression for ρ(λ)\rho(\lambda) in terms of G\langle G\rangle_{\infty} follows in an analogous way, i.e., one inserts Eq. (23) in Eq. (11) and calculates the remainder integral

ρ(λ)\displaystyle\rho(\lambda) =1πlimη0+Im[1J2JGexp(zJ2JG)\displaystyle=\frac{1}{\pi}\lim_{\eta\rightarrow 0^{+}}{\rm Im}\Bigg{[}\frac{1}{\langle J^{2}\rangle_{J}\langle G\rangle_{\infty}}\exp{\left(-\frac{z}{\langle J^{2}\rangle_{J}\langle G\rangle_{\infty}}\right)}
×Ei(zJ2JG)],\displaystyle\times{\rm Ei}\left(\frac{z}{\langle J^{2}\rangle_{J}\langle G\rangle_{\infty}}\right)\Bigg{]}, (26)

with z=λiηz=\lambda-i\eta. It is suitable to introduce the dimensionless variable γ(z)zJ2JG\gamma(z)\equiv\frac{z}{\langle J^{2}\rangle_{J}\langle G\rangle_{\infty}}, in terms of which the spectral density is given by

ρ(λ)=1πlimη0+Im(z2+γ2J2Jzγ2J2J),\rho(\lambda)=\frac{1}{\pi}\lim_{\eta\rightarrow 0^{+}}{\rm Im}\left(\frac{z^{2}+\gamma^{2}\langle J^{2}\rangle_{J}}{z\,\gamma^{2}\langle J^{2}\rangle_{J}}\right), (27)

where γ(z)\gamma(z) solves the following equation

J2Jγ=z2γ2eγEi(γ)γ.\langle J^{2}\rangle_{J}\gamma=\frac{z^{2}}{\gamma^{2}e^{-\gamma}{\rm Ei}(\gamma)-\gamma}. (28)

Equations (27) and (28) constitute one of the main results of this paper. In contrast to the resolvent Eq. (19), whose analytic solution yields the Wigner semicircle law, the fixed-point Eq. (28) for the cc\rightarrow\infty limit of exponential graphs has no analytic solution for arbitrary zz. The spectral density obtained from the numerical solutions of Eq. (28) is shown in figure 3-(a), together with direct diagonalization of large adjacency matrices of exponential random graphs with c=100c=100. The agreement between these two different approaches is excellent, confirming the exactness of Eqs. (27) and (28).

Refer to caption
Figure 3: Dense limit of the spectral density ρ(λ)\rho(\lambda) of random networks with an exponential degree distribution and coupling strengths drawn from PJ(x)=δ(x1)P_{J}(x)=\delta(x-1). (a) Comparison between the theoretical results (black solid lines), obtained from the numerical solutions of Eqs. (27) and (28) for η=0\eta=0, and numerical diagonalizations (red circles) of 100100 samples of 104×10410^{4}\times 10^{4} adjacency matrices with c=100c=100. (b) The logarithmic divergence of the spectral density for |λ|0|\lambda|\rightarrow 0. The black solid line is the analytic result of Eq. (31), while the symbols are data derived from the numerical solutions of Eqs. (7) and (8) for c=80c=80 and two values of η\eta: η=104\eta=10^{-4} (brown triangles) and η=105\eta=10^{-5} (green squares). (c) The exponential behaviour of the spectral density for large |λ||\lambda|, obtained from the numerical solutions of Eqs. (27) and (28) for η=0\eta=0

Figure 3-(a) suggests that ρ(λ)\rho(\lambda) diverges at λ=0\lambda=0. To study the singular behaviour of ρ(λ)\rho(\lambda) as |λ|0|\lambda|\rightarrow 0, one needs to understand how γ(z)\gamma(z) vanishes as |z|0|z|\rightarrow 0. With a modest amount of foresight, we make the following ansatz

γ(z)=β1zJ2J+β2(z)z2J2J,|z|0,\gamma(z)=\beta_{1}\frac{z}{\sqrt{\langle J^{2}\rangle_{J}}}+\beta_{2}(z)\frac{z^{2}}{\langle J^{2}\rangle_{J}},\quad|z|\rightarrow 0, (29)

where the function β2(z)\beta_{2}(z) is such that lim|z|0z2β2(z)=0\lim_{|z|\rightarrow 0}z^{2}\beta_{2}(z)=0, and the coefficient β1\beta_{1} is independent of zz. Plugging this assumption for γ(z)\gamma(z) in the right hand side of Eq. (28) and expanding the result in powers of zz up to O(z2)O(z^{2}), we find that the above ansatz is consistent with Eq. (28) if β1\beta_{1} and β2(z)\beta_{2}(z) are

β1=±i,β2(z)=12[E+ln(β1zJ2J)],\beta_{1}=\pm i,\quad\beta_{2}(z)=-\frac{1}{2}\left[E+\ln\left(-\frac{\beta_{1}z}{\sqrt{\langle J^{2}\rangle_{J}}}\right)\right], (30)

with EE representing the Euler-Mascheroni constant. The last step is to substitute Eq. (29) in Eq. (27) and take the limit η0+\eta\rightarrow 0^{+}, which leads to the logarithmic divergence

ρ(λ)=1πJ2J[E+ln(|λ|J2J)]\rho(\lambda)=-\frac{1}{\pi\sqrt{\langle J^{2}\rangle_{J}}}\left[E+\ln\left(\frac{|\lambda|}{\sqrt{\langle J^{2}\rangle_{J}}}\right)\right] (31)

for |λ|0|\lambda|\rightarrow 0. The choice of the negative sign β1=i\beta_{1}=-i ensures that ρ(λ)\rho(\lambda) is non-negative. A divergence in the spectral density of random graphs at λ=0\lambda=0 may appear due to different reasons. For instance, the spectrum of protein-protein interaction networks has a singularity at λ=0\lambda=0 due to the existence of duplicated genes [54]. In the context of solid state physics, the density of states of a tight-binding model for a quantum particle hoping on a random graph displays a divergence at λ=0\lambda=0 due to the presence of localized eigenstates [55].

Figure 3-(b) compares Eq. (31) with the numerical solutions of the distributional Eqs. (7) and (8) for c=80c=80 using a Monte-Carlo method, also known as the population dynamics algorithm [36, 37, 50]. The numerical results lie on the top of the theoretical curve up to a certain |λ||\lambda_{*}|, below which the numerical data for ρ(λ)\rho(\lambda) deviates from the logarithmic behaviour of Eq. (31). As η\eta decreases, |λ||\lambda_{*}| shifts towards smaller values, confirming that the discrepancy between the numerical data and Eq. (31) is an artifact due to the finite values of η\eta employed in the numerical solutions of Eqs. (7) and (8).

As a final important property, we inspect the behaviour of ρ(λ)\rho(\lambda) for |λ|1|\lambda|\gg 1. As illustrated in figure 3-(c), ρ(λ)\rho(\lambda) displays a crossover from an exponential behaviour lnρ(λ)λ\ln\rho(\lambda)\propto-\lambda to a faster decay for increasing |λ||\lambda|, indicating that the spectral density does not have sharp spectral edges, but instead it is supported on the entire real line. This is also consistent with a stability analysis of the fixed-point Eq. (28), which shows that the complex solution for γ(z)\gamma(z) remains stable for |λ|1|\lambda|\gg 1. We point out that a perturbative study of Eq. (28) for |λ|1|\lambda|\gg 1 is not able to capture the analytic form describing the tails of ρ(λ)\rho(\lambda). This should not come as a surprise, as the tails of ρ(λ)\rho(\lambda) are normally caused by rare statistical fluctuations in the graph structure [34, 43].

IV.3 Borel random graphs

As a final example of network model, we present results for random networks with a Borel degree distribution (see table 1). In this case, our approach to derive analytic expressions for φ(p,t)\varphi(p,t) and φcav(p,t)\varphi_{\rm cav}(p,t) is not useful, because the probability generating function of the Borel degree distribution does not have a closed analytic form [48]. Thus, we obtain results through the numerical solution of Eqs. (7) and (8) using the population dynamics algorithm [36, 37, 50]. For Borel random graphs, the ratio σ2/c2\sigma^{2}/c^{2} diverges as cc\rightarrow\infty, which is precisely what renders this model interesting in comparison to Poisson and exponential graphs.

Figure 4 illustrates the evolution of the spectral density for increasing average degree. Similarly to exponential random graphs, the spectral density is not described by the semicircle distribution of random matrix theory, although the values of cc in figure 4 are not large enough to observe the dense limit of ρ(λ)\rho(\lambda). We do not present results for larger values of cc than those in figure 4 because the population dynamics algorithm becomes prohibitively time consuming for increasing cc, due to the large variance of the Borel degree distribution.

Among the distinctive features of figure 4, we note the existence of δ\delta-peaks in ρ(λ)\rho(\lambda). These peaks, often located at the eigenvalues of finite components [56, 37, 23, 41], disappear for cc\rightarrow\infty, when the fraction of nodes belonging to the giant component converges to one [1]. The spectral density in figure 4 displays δ\delta-peaks at the eigenvalues {1/c,1/c}\{-1/\sqrt{c},1/\sqrt{c}\} and {0,2/c,2/c}\{0,-\sqrt{2}/\sqrt{c},\sqrt{2}/\sqrt{c}\} of the adjacency matrices of open chains with, respectively, two and three nodes. One can show that these δ\delta-peaks disappear in the dense limit by computing the fraction Πs(c)\Pi_{s}(c) of nodes that belongs to a finite component with s>1s>1 nodes 111See chapter 13 of [1].. For Poisson, exponential, and Borel degree distributions, we obtain Πs(pois)(c)esc\Pi_{s}^{({\rm pois})}(c)\propto e^{-sc}, Πs(exp)(c)c12s\Pi_{s}^{({\rm exp})}(c)\propto c^{1-2s}, and Πs(borel)(c)c1s\Pi_{s}^{({\rm borel})}(c)\propto c^{1-s} for c1c\gg 1. Consequently, all network models studied in this paper do not have finite clusters in the dense limit. In the case of the Borel degree distribution, ρ(λ)\rho(\lambda) contains δ\delta-peaks even for very large cc because the function Πs(borel)(c)\Pi_{s}^{({\rm borel})}(c) decays slower for c1c\gg 1 when compared to Poisson and exponential networks.

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 4: The spectral density of random networks from the configuration model with a Borel degree distribution (see table 1) for different average degrees cc and coupling strengths drawn from PJ(x)=δ(x1)P_{J}(x)=\delta(x-1). All results are obtained from the numerical solutions of Eqs. (7) and (8) using the population dynamics algorithm with η=103\eta=10^{-3} (subfigures (a), (b), and (c)) and η=104\eta=10^{-4} (subfigure (d)). Subfigures (a), (b), and (c) show the spectral density of Borel random graphs for different cc. Figure (a) also indicates the positive eigenvalues of open chains with one and two edges, and their correspondence with the position of δ\delta-peaks. Figure (d) exhibits results for the tails of ρ(λ)\rho(\lambda). The symbols are results for Borel degree distributions with different cc, while the solid black curve shows the tail of ρ(λ)\rho(\lambda) for exponential dense random graphs as a comparison (see also figure 3). The dashed lines are just a guide to the eye.

We have also inspected the tails of the spectral density of Borel random graphs. According to figure 4-(d), the behaviour of ρ(λ)\rho(\lambda) as a function of λ\lambda is consistent with a power-law decay up to a certain threshold λp\lambda_{p}, above which the data clearly deviates from a straight line. As a matter of fact, we have confirmed that ρ(λ)\rho(\lambda) decreases as η0+\eta\rightarrow 0^{+} for values of λ\lambda larger than those appearing in figure 4-(d). In spite of that, the overall tendency of the data for increasing cc suggests that ρ(λ)\rho(\lambda) does decay as a power-law in the limit cc\rightarrow\infty. The results in figure 4 are insufficient to extract the exponent characterizing the power-law tails of ρ(λ)\rho(\lambda) for cc\rightarrow\infty, since the spectral density has not reached its limiting behaviour for c=40c=40. Taken together, these results indicate that ρ(λ)\rho(\lambda) is supported on the entire real line.

V A rigorous result for the spectral density of the configuration model

In this section we state the main theorem of reference [44] and explain how the resolvent equations for cc\rightarrow\infty, obtained in the previous section, are recovered from this theorem. We use the notation of the previous sections and we discuss the theorem in a less technical language, which helps to establish the link between [44] and the results presented in this paper in a more straightforward way.

The theorem concerns the spectral density ρ(λ)\rho(\lambda) of the adjacency matrix 𝑨\bm{A} (see Eq. (2)) of the configuration model in the dense limit. Let ν(κ)\nu(\kappa) be the probability density of the rescaled degrees Kic\frac{K_{i}}{c} (i=1,,Ni=1,\dots,N) in the limit cc\rightarrow\infty

ν(κ)=limck=0pkδ(κkc).\nu(\kappa)=\lim_{c\rightarrow\infty}\sum_{k=0}^{\infty}p_{k}\delta\left(\kappa-\frac{k}{c}\right). (32)

The main result of [44] is the following theorem:

Let {Ki}i=1N\{K_{i}\}_{i=1}^{N} be i.i.d random variables forming a degree sequence of the configuration model such that c=o(N)c=o(N) and Jij=1J_{ij}=1. In the limit NN\rightarrow\infty, if the distribution of Ki/cK_{i}/c converges to a limit distribution ν\nu with a finite second moment, then ρ(λ)=limNρN(λ)\rho(\lambda)=\lim_{N\rightarrow\infty}\rho_{N}(\lambda) is given by νρw\nu\boxtimes\rho_{\rm w}, where ρw\rho_{\rm w} is the Wigner semicircle law.

Thus, the spectral density ρ(λ)\rho(\lambda) of the configuration model with c=o(N)c=o(N) is given by the free multiplicative convolution \boxtimes between the probability density ν\nu of rescaled degrees and the Wigner semicircle law ρw\rho_{\rm w}. The free multiplicative convolution νρw\nu\boxtimes\rho_{\rm w} between the probability measure associated to ρw\rho_{\rm w} and the probability measure corresponding to ν\nu is the unique probability measure associated to ρ\rho such that its SS-transform Sρ(ω)S_{\rho}(\omega) (ω\omega\in\mathbb{C}) fulfills the convolution property Sρ(ω)=Sν(ω)Sρw(ω)S_{\rho}(\omega)=S_{\nu}(\omega)S_{\rm\rho_{w}}(\omega), where Sν(ω)S_{\nu}(\omega) and Sρw(ω)S_{\rm\rho_{w}}(\omega) are the SS-transforms of ν\nu and ρw\rho_{\rm w}, respectively.

The free multiplicative convolution and the SS-transform were introduced by Voiculescu to deal with the multiplication of free non-commuting random variables [58]. The SS-transform finds applications in random matrix theory, since it is an important analytic tool to determine the spectral properties of products of large and independent random matrices from the spectra of the individual matrices [59, 60].

For unbounded measures on +\mathbb{R}_{+} and for measures with zero mean and all moments finite, such as the Wigner semicircle distribution, the SS-transform Sφ(z)S_{\varphi}(z) (zz\in\mathbb{C}) of the corresponding probability density φ\varphi is obtained from [61, 62]

Sφ(z)=(1+z)zωφ1(z),S_{\varphi}(z)=\frac{\left(1+z\right)}{z}\omega_{\varphi}^{-1}(z), (33)

where

ωφ(z)=1z𝒞φ(1/z)1.\omega_{\varphi}(z)=\frac{1}{z}\mathcal{C}_{\varphi}(1/z)-1. (34)

The function ωφ1(z)\omega_{\varphi}^{-1}(z) is the functional inverse of ωφ(z)\omega_{\varphi}(z) with respect to composition and 𝒞φ(z)\mathcal{C}_{\varphi}(z) is the Cauchy transform of φ\varphi

𝒞φ(z)=dλφ(λ)zλ,\mathcal{C}_{\varphi}(z)=\int_{-\infty}^{\infty}\frac{d\lambda\varphi(\lambda)}{z-\lambda}, (35)

where zz lies outside the support of φ(λ)\varphi(\lambda). Instead of working directly with the inverse ωφ1(z)\omega_{\varphi}^{-1}(z) in Eq. (33), it is more convenient to express the SS-transform in terms of the Cauchy transform [63, 60]. Since ωφ1(ωφ(z))=z\omega_{\varphi}^{-1}\left(\omega_{\varphi}(z)\right)=z, we can rewrite Eq. (33) as follows

Sφ[ωφ(z)]=[1+ωφ(z)]ωφ(z)z.S_{\varphi}\left[\omega_{\varphi}(z)\right]=\frac{\left[1+\omega_{\varphi}(z)\right]}{\omega_{\varphi}(z)}z. (36)

Hence, given the SS-transform of a probability density φ\varphi, we are able to uniquely determine φ\varphi.

In the context of random matrix theory, 𝒞ρ(z)\mathcal{C}_{\rho}(z) is the trace of the resolvent matrix

𝒞ρ(z)=limN1Ni=1NGii(z),\mathcal{C}_{\rho}(z)=\lim_{N\rightarrow\infty}\frac{1}{N}\sum_{i=1}^{N}G_{ii}(z), (37)

which, after setting z=λiηz=\lambda-i\eta, gives access to the spectral density ρ\rho through Eq. (5). Combining Eqs. (5), (34), and (37), we can also express the spectral density in terms of ωρ(z)\omega_{\rho}(z)

ρ(λ)=1πlimη0+Im(ωρ(1/z)+1z).\rho(\lambda)=\frac{1}{\pi}\lim_{\eta\rightarrow 0^{+}}{\rm Im}\left(\frac{\omega_{\rho}(1/z)+1}{z}\right). (38)

The above theorem holds for networks where cc scales slowly with NN such that limNcN=0\lim_{N\rightarrow\infty}\frac{c}{N}=0. In the previous section, we have derived results by performing the limit cc\rightarrow\infty after NN\rightarrow\infty, which also implies that limNcN=0\lim_{N\rightarrow\infty}\frac{c}{N}=0. Consequently, with the exception of dense Borel networks for which ν(κ)\nu(\kappa) has a divergent second moment (see table 1), the results for ρ(λ)\rho(\lambda) derived in the previous section should be recovered from the above theorem.

V.1 Regular and Poisson random graphs

Here we apply the theorem to degree distributions pkp_{k} such that ν(κ)=δ(κ1)\nu(\kappa)=\delta(\kappa-1), like regular and Poisson degree distributions. The procedure to derive the spectral density from the theorem comprises three steps: first, we compute the SS-transforms SνS_{\nu} and SρwS_{\rho_{\rm w}}; second, we obtain the SS-transform SρS_{\rho} of ρ(λ)\rho(\lambda) from the convolution Sρ(ω)=Sν(ω)Sρw(ω)S_{\rho}(\omega)=S_{\nu}(\omega)S_{\rho_{\rm w}}(\omega); third, we derive the resolvent equation from SρS_{\rho}.

Let us compute the SS-transform of the Wigner semicircle law ρw(λ)\rho_{\rm w}(\lambda) of Eq. (20) with J2J=1\langle J^{2}\rangle_{J}=1. The Cauchy transform 𝒞ρw(z)\mathcal{C}_{\rho_{\rm w}}(z) of ρw(λ)\rho_{\rm w}(\lambda) solves the algebraic equation [63]

𝒞ρw(z)=1z𝒞ρw(z).\mathcal{C}_{\rho_{\rm w}}(z)=\frac{1}{z-\mathcal{C}_{\rho_{\rm w}}(z)}. (39)

Using Eq. (34) and rewriting Eq. (39) in terms of wρw(z)w_{\rho_{\rm w}}(z), we get

z=±ωρw(z)1+ωρw(z).z=\pm\frac{\sqrt{\omega_{\rho_{\rm w}}(z)}}{1+\omega_{\rho_{\rm w}}(z)}. (40)

Inserting the above expression in Eq. (36) leads to

Sρw(ωρw)=1ωρw,S_{\rho_{\rm w}}\left(\omega_{\rho_{\rm w}}\right)=\frac{1}{\sqrt{\omega_{\rho_{\rm w}}}}, (41)

where we have chosen the positive sign in Eq. (40). Equation (39) for the Cauchy transform of ρw\rho_{\rm w} is obtained from the SS-transform regardless of the choice of sign in Eq. (40).

Now we turn our attention to the SS-transform of ν(κ)=δ(κ1)\nu(\kappa)=\delta(\kappa-1). The Cauchy transform 𝒞ν(z)\mathcal{C}_{\nu}(z) of ν(κ)\nu(\kappa) reads

𝒞ν(z)=1z1.\mathcal{C}_{\nu}(z)=\frac{1}{z-1}. (42)

Combining Eqs. (34) and (42), we get

z=ων(z)1+ων(z),z=\frac{\omega_{\nu}(z)}{1+\omega_{\nu}(z)}, (43)

which leads to

Sν(ω)=1S_{\nu}(\omega)=1 (44)

after substitution in Eq. (36). According to the above theorem, the SS-transform of the spectral density reads

Sρ(ω)=Sν(ω)Sρw(ω)=1ω,S_{\rho}(\omega)=S_{\nu}(\omega)S_{\rho_{\rm w}}(\omega)=\frac{1}{\sqrt{\omega}}, (45)

which immediately implies that ρ(λ)\rho(\lambda) is given by the semicircle distribution, Eq. (20). We conclude that the spectral density of dense random networks for which the distribution of the rescaled degrees Ki/cK_{i}/c converges to a δ\delta-peak is given by Eq. (20).

V.2 Exponential random graphs

For random graphs with an exponential degree distribution, ν(κ)\nu(\kappa) reads

ν(κ)={eκ,if κ0,0,if κ<0.\nu(\kappa)=\begin{cases}e^{-\kappa},&\text{if }\kappa\geq 0,\\ 0,&\text{if }\kappa<0.\end{cases} (46)

The transform Sρw(ω)S_{\rho_{\rm w}}(\omega) is given by Eq. (41), hence we just need to obtain Sν(ω)S_{\nu}(\omega). We combine the Cauchy transform of Eq. (46)

𝒞ν(z)=0𝑑κeκ(zκ)=ezEi(z)\mathcal{C}_{\nu}(z)=\int_{0}^{\infty}d\kappa\frac{e^{-\kappa}}{\left(z-\kappa\right)}=e^{-z}{\rm Ei}(z) (47)

with Eq. (34), obtaining

ων(z)=e1/zzEi(1/z)1.\omega_{\nu}(z)=\frac{e^{-1/z}}{z}{\rm Ei}(1/z)-1. (48)

In order to derive an explicit expression for Sν(ων)S_{\nu}(\omega_{\nu}) from Eq. (36), we have to invert analytically the above equation and find z=ϕ(ων)z=\phi(\omega_{\nu}). The analytic form of ϕ(ων)\phi(\omega_{\nu}) is usually obtained when the Cauchy transform has a polynomial form [63], as in the case of the Wigner semicircle distribution. Unfortunately, here this is not the case, but we can still obtain from Eq. (48) a self-consistent equation for the inverse function ϕ(ων)\phi(\omega_{\nu})

ϕ(ων)=1ωνexp[1ϕ(ων)]Ei[1ϕ(ων)]ϕ(ων)ων,\phi\left(\omega_{\nu}\right)=\frac{1}{\omega_{\nu}}\exp{\left[-\frac{1}{\phi(\omega_{\nu})}\right]}{\rm Ei}\left[\frac{1}{\phi\left(\omega_{\nu}\right)}\right]-\frac{\phi\left(\omega_{\nu}\right)}{\omega_{\nu}}, (49)

while the SS-transform follows from Eq. (36)

Sν(ων)=(1+ων)ωνϕ(ων),S_{\nu}(\omega_{\nu})=\frac{\left(1+\omega_{\nu}\right)}{\omega_{\nu}}\phi(\omega_{\nu}), (50)

with ων=ων(z)\omega_{\nu}=\omega_{\nu}(z). The solution of Eq. (49) determines Sν(ων)S_{\nu}(\omega_{\nu}) through Eq. (50).

Now we are ready to recover the equations determining ρ(λ)\rho(\lambda) of exponential random graphs using the theorem of [44]. The SS-transform of ρ(λ)\rho(\lambda) follows from the convolution

Sρ(ωρ)=Sν(ωρ)Sρw(ωρ)=(1+ωρ)ωρωρϕ(ωρ).S_{\rho}(\omega_{\rho})=S_{\nu}(\omega_{\rho})S_{\rho_{\rm w}}(\omega_{\rho})=\frac{\left(1+\omega_{\rho}\right)}{\omega_{\rho}\sqrt{\omega_{\rho}}}\phi(\omega_{\rho}). (51)

From Eq. (36) we conclude that ϕ(ωρ)\phi(\omega_{\rho}) must also fulfill

zωρ=ϕ(ωρ)z\sqrt{\omega_{\rho}}=\phi(\omega_{\rho}) (52)

for zz outside the support of ρ(λ)\rho(\lambda). Substituting the above relation in Eq. (49), we find an equation that determines ωρ(1/z)\omega_{\rho}(1/z)

ωρ(1/z)z\displaystyle\frac{\sqrt{\omega_{\rho}(1/z)}}{z} =1ωρ(1/z)exp[zωρ(1/z)]Ei[zωρ(1/z)]\displaystyle=\frac{1}{\omega_{\rho}(1/z)}\exp{\left[-\frac{z}{\sqrt{\omega_{\rho}(1/z)}}\right]}{\rm Ei}\left[\frac{z}{\sqrt{\omega_{\rho}(1/z)}}\right]
1zωρ(1/z).\displaystyle-\frac{1}{z\sqrt{\omega_{\rho}(1/z)}}. (53)

The solution for ωρ(1/z)\omega_{\rho}(1/z) at z=λiηz=\lambda-i\eta allows to compute the spectral density using Eq. (38). Setting

γ(z)=zωρ(1/z),\gamma(z)=\frac{z}{\sqrt{\omega_{\rho}(1/z)}}, (54)

we conclude that Eqs. (38) and (53) are identical to Eqs. (27) and (28), respectively. We also identify ωρ(1/z)=G\sqrt{\omega_{\rho}(1/z)}=\langle G\rangle_{\infty} as the averaged resolvent on the cavity graph.

VI The fourth moment of the eigenvalue distribution

The results of sections IV and V show that the dense limit of ρ(λ)\rho(\lambda) depends on the degree distribution of the configuration model, and the scope of the Wigner semicircle law is limited to graphs where pkp_{k} becomes highly concentrated around the mean degree for cc\rightarrow\infty. In this section we derive a simple equation that reveals the influence of degree fluctuations on the spectral density, giving an exact condition for the breakdown of the Wigner semicircle law. We end up this section by proposing a classification of the different universal behaviours of the dense limit of ρ(λ)\rho(\lambda) for the configuration model of networks.

It is natural to characterize the statistics of random variables by computing the ratio between their moments. In the case of locally tree-like random networks, the odd moments of ρ(λ)\rho(\lambda) are zero and the simplest dimensionless parameter of this type is the kurtosis KN(c)K_{N}(c)

KN(c)=𝑑λλ4ρN(λ)[𝑑λλ2ρN(λ)]2=NTr𝑨4(Tr𝑨2)2,K_{N}(c)=\frac{\int_{-\infty}^{\infty}d\lambda\lambda^{4}\rho_{N}(\lambda)}{\left[\int_{-\infty}^{\infty}d\lambda\lambda^{2}\rho_{N}(\lambda)\right]^{2}}=\frac{N{\rm Tr}\bm{A}^{4}}{\left({\rm Tr}\bm{A}^{2}\right)^{2}}, (55)

where we used the definition of the empirical spectral measure, Eq. (3). When Jij=ci,jJ_{ij}=\sqrt{c}\,\,\forall i,j, the trace Tr𝑨n{\rm Tr}\bm{A}^{n} (n=2,3,n=2,3,\dots) is the total number of closed walks of length nn in the graph, where the length of a walk between nodes ii and jj is the number of edges that a walker traverses when going from one node to the other [18]. The relation between the moments of ρN(λ)\rho_{N}(\lambda) and Tr𝑨n{\rm Tr}\bm{A}^{n} is very important in spectral graph theory, as it connects the eigenvalue statistics with the network structural properties [18]. The limit NN\rightarrow\infty must be taken before the cc\rightarrow\infty limit.

The trace of the second power of the adjacency matrix 𝑨\bm{A} reads

Tr𝑨2=1ci=1NjiJij2,{\rm Tr}\bm{A}^{2}=\frac{1}{c}\sum_{i=1}^{N}\sum_{j\in\partial_{i}}J_{ij}^{2}, (56)

where i\partial_{i} represents the set of nodes connected to ii. For NN\rightarrow\infty, we obtain

limN1NTr𝑨2=1cjiJij2J,K=J2J,\lim_{N\rightarrow\infty}\frac{1}{N}{\rm Tr}\bm{A}^{2}=\frac{1}{c}\left\langle\sum_{j\in\partial_{i}}J_{ij}^{2}\right\rangle_{J,K}=\langle J^{2}\rangle_{J}, (57)

with J,K\left\langle\dots\right\rangle_{J,K} denoting the ensemble average over the degrees and the coupling strengths. The trace of the fourth power can be written as

c2Tr𝑨4\displaystyle c^{2}{\rm Tr}\bm{A}^{4} =2i=1Nj,riJij2Jri2i=1NjiJij4\displaystyle=2\sum_{i=1}^{N}\sum_{j,r\in\partial_{i}}J_{ij}^{2}J_{ri}^{2}-\sum_{i=1}^{N}\sum_{j\in\partial_{i}}J_{ij}^{4}
+i=1NjikjirkjCirJijJjkJkrJri,\displaystyle+\sum_{i=1}^{N}\sum_{j\in\partial_{i}}\sum_{k\in\partial_{j}\setminus i}\sum_{r\in\partial_{k}\setminus j}C_{ir}J_{ij}J_{jk}J_{kr}J_{ri}, (58)

where ji\partial_{j}\setminus i is the set of nodes adjacent to jj, except for node iji\in\partial_{j}. In the limit NN\rightarrow\infty, the configuration model has a locally tree-like structure, cycles of length four are rare, and hence the last term on the right hand side of Eq. (58) gives only a subleading contribution, which can be neglected for NN\rightarrow\infty, yielding the expression

limN1NTr𝑨4=2c2K(K1)KJ2J2+1cJ4J.\lim_{N\rightarrow\infty}\frac{1}{N}{\rm Tr}\bm{A}^{4}=\frac{2}{c^{2}}\langle K(K-1)\rangle_{K}\langle J^{2}\rangle_{J}^{2}+\frac{1}{c}\langle J^{4}\rangle_{J}. (59)

Thus, we obtain an exact analytic expression for the NN\rightarrow\infty limit of the kurtosis

K(c)=1cJ4JJ2J2+2c(c1)+2σ2c2,K_{\infty}(c)=\frac{1}{c}\frac{\langle J^{4}\rangle_{J}}{\langle J^{2}\rangle_{J}^{2}}+\frac{2}{c}(c-1)+2\frac{\sigma^{2}}{c^{2}}, (60)

valid for arbitrary degree distributions with first and second moments finite. Note that K(c)K_{\infty}(c) is invariant under a rescaling of the adjacency matrix elements. Equation (60) shows how the kurtosis of the eigenvalue distribution is linked to the degree fluctuations.

Let us now establish some general conclusions about the dense limit of ρ(λ)\rho(\lambda). For cc\rightarrow\infty, K(c)K_{\infty}(c) behaves as

limcK(c)=Kw(1+limcσ2c2),\lim_{c\rightarrow\infty}K_{\infty}(c)=K_{\rm w}\left(1+\lim_{c\rightarrow\infty}\frac{\sigma^{2}}{c^{2}}\right), (61)

where Kw=2K_{\rm w}=2 is the kurtosis of the Wigner semicircle distribution, Eq. (20). The kurtosis of the distribution ρw(λ)\rho_{\rm w}(\lambda) follows from its moments, which are given by the Catalan numbers [64]. Equation (61) unveils the central role of the degree fluctuations to the cc\rightarrow\infty limit of ρ(λ)\rho(\lambda). It follows that

limcσ2c2>0\lim_{c\rightarrow\infty}\frac{\sigma^{2}}{c^{2}}>0 (62)

is a sufficient condition for the breakdown of the Wigner semicircle law. In the previous section we have shown that ρ(λ)\rho(\lambda) converges to the semicircle distribution when the distribution ν\nu of rescaled degrees is a Dirac δ\delta-peak, which is fully consistent with Eq. (62). One naturally expects that the eigenvalue statistics of network models that fulfil condition (62) is not described by the traditional results of random matrix theory.

Based on Eq. (61), we can also put forward a classification of the different universal behaviours of ρ(λ)\rho(\lambda) in the dense limit. Let us consider weighted undirected networks, defined by the adjacency matrix of Eq. (2), in which the variance of pkp_{k} scales as σ2=Hcα\sigma^{2}=Hc^{\alpha} for large cc, with H>0H>0 and α\alpha arbitrary parameters. For this broad class of network models, the different universal behaviours of ρ(λ)\rho(\lambda) can be classified in terms of the exponent α\alpha that quantifies the strength of the degree fluctuations. For α<2\alpha<2, the relative variance σ2/c2\sigma^{2}/c^{2} vanishes for cc\rightarrow\infty, which is a strong indication that pkp_{k} is highly peaked around cc. Thus, we obtain limcK(c)=2\lim_{c\rightarrow\infty}K_{\infty}(c)=2 for α<2\alpha<2, and it is reasonable to conjecture that the dense limit of ρ(λ)\rho(\lambda) is given by the Wigner semicircle law. Regular and Poisson random graphs are the main examples of this class of random networks characterized by α<2\alpha<2. For α>2\alpha>2, the limit limcK(c)\lim_{c\rightarrow\infty}K_{\infty}(c) diverges and we expect ρ(λ)\rho(\lambda) decays as a power-law ρ(λ)|λ|β\rho(\lambda)\sim|\lambda|^{-\beta} for |λ|1|\lambda|\gg 1, with an exponent 3<β<53<\beta<5. The lower bound on β\beta is due to the finite variance of ρ(λ)\rho(\lambda) in the dense limit (see Eq. (57)). Borel random graphs are characterized by α>2\alpha>2. Finally, network models with α=2\alpha=2 have a finite kurtosis limcK(c)=2(1+H)>Kw\lim_{c\rightarrow\infty}K_{\infty}(c)=2(1+H)>K_{\rm w}, whose precise value is determined by the prefactor HH. Exponential random networks belong to this later class where α=2\alpha=2 and H=1H=1. The results presented in section IV are entirely consistent with this classification scheme.

VII Final remarks

Traditional random matrix theory deals with the spectral properties of large random matrices with independent and identically distributed elements, providing a theoretical framework to address the universal properties of large interacting systems [65]. It is widely known that the eigenvalue distribution of undirected random networks strongly deviates from the Wigner semicircle law of random matrix theory in the sparse regime [34, 56, 19, 20, 36, 37, 41], i.e., when the average degree cc is finite. As cc grows to infinity, random networks gradually become more fully-connected and one may expect that random matrix theory describes well the spectral properties of dense random networks. We have shown in this paper that this is generally not the case.

We have studied the spectral density of the adjacency matrix of networks drawn from the configuration model with four distinct degree distributions. Our main conclusion is that the dense limit of the spectral density is governed by the strength of the degree fluctuations. It turns out that the semicircle distribution of random matrix theory is recovered only when the degree distribution becomes, in the dense limit, highly concentrated around the mean degree. We have also derived an exact relation between the fourth moment of the eigenvalue distribution and the variance of the degree distribution, from which a sufficient condition for the breakdown of the Wigner semicircle law follows. We point out that the degree distributions considered in this work have exponentially decaying tails, implying that all moments of the degree distribution are finite (the moments diverge only in the dense limit).

From the results derived in this work, one expects that, in general, the circular law of random matrix theory [38, 4] should also fail in describing the spectral density of directed random networks in the dense limit. Following the techniques discussed in sections IV and VI, the study of the eigenvalue distribution of directed networks in the dense limit is just around the corner.

Among the results in section IV, we highlight the analytic equation for the averaged resolvent (see Eq. (25)), which determines the spectral density of exponential random graphs in the dense limit. This equation has no analytic solution, in contrast to the analogous Eq. (19) yielding the Wigner semicircle distribution. We have derived important features of the spectral density of dense exponential graphs, such as the existence of a logarithmic singularity around the origin and the absence of sharp spectral edges. In the case of dense random networks with a Borel degree distribution, we have studied the spectral density by solving numerically the exact Eqs. (7) and (8) for large values of the average degree. The results in figure 4 indicate that the spectral density of dense Borel networks is also supported on the entire real line, exhibiting power-law tails for large eigenvalues. Taken together, these results reveal remarkable differences with respect to the Wigner semicircle law.

We have shown how the analytic results of section IV are recovered from the main theorem in reference [44]. While the results of section IV follow from the equations for the distribution of the resolvent, the theorem in [44] is proved through a series of techniques, including tools from free probability theory. These two approaches are fundamentally different and we hope that our work stimulates research towards a rigorous proof of their equivalence for arbitrary degree distributions. We also remark that, although the theorem in [44] is a compelling result about the spectral density, the approach of section IV is also very interesting, since it allows to compute the distribution of the resolvent, the distribution of the self-energy [66], and the inverse participation ratio [50] in the dense limit. All these quantities are important, for instance, to study the localization properties of eigenvectors.

Based on Eq. (61) for the fourth moment of the eigenvalue distribution, we have put forward a classification scheme of the different universal behaviours of the spectral density in the dense regime. The classification holds for network models in which the variance σ2\sigma^{2} of the degree distribution scales as σ2cα\sigma^{2}\propto c^{\alpha} for c1c\gg 1. Networks with α<2\alpha<2 exhibit weak degree fluctuations and we have conjectured that the spectral density converges to the Wigner semicircle distribution for cc\rightarrow\infty. Networks with α>2\alpha>2 display strong degree fluctuations and the spectral density is characterized by a divergent fourth moment and power-law tails for cc\rightarrow\infty. Finally, the spectral density of dense networks with α=2\alpha=2 has a finite fourth moment, whose value is larger than in the case of the Wigner semicircle distribution. The results for the specific degree distributions in section IV are entirely consistent with this classification scheme. In light of this, it would be interesting to design a network model with a degree distribution that has finite moments and interpolates among the different classes.

Overall, our results shed light on the fundamental role of the degree fluctuations in the dense limit of random networks. In this context, it would be interesting to inspect the role of degree fluctuations in the local spectral properties [67, 68, 69] of dense random networks, since the local statistics of the spectrum usually exhibits a higher level of universality in comparison to the global statistics. The condition for the breakdown of the Wigner semicircle law, Eq. (62), should also apply beyond the realm of network spectra, indicating whether the degree fluctuations are strong enough to cause the failure of classic mean-field models on dense networks, such as the Curie-Weiss model [49] and the Sherrington-Kirkpatrick model [70]. Thus, we expect our results will stimulate research towards a better understanding of the dense limit of various models defined on random graphs, including ferromagnetic and spin-glass models [71], social dynamics [72], neural networks [73], and synchronization [8].

acknowledgements

We thank an anonymous referee for pointing out reference [44], which has led to section V. FLM thanks London Mathematical Laboratory and CNPq/Brazil for financial support. JDS acknowledges a fellowship from the London Mathematical Laboratory.

References