Random matrix universality for classical transport in composite materials N. Benjamin Murphy, Elena Cherkaev, and Kenneth M. Golden∗ University of Utah, Department of Mathematics, 155 S 1400 E, JWB 233, Salt Lake City, UT 84112, USA Universality of eigenvalue statistics has been observed in random matrices arising in studies of atomic spectra, internet signal dynamics, and even the zeros of the famous Riemann zeta function. Here we consider the flow of electrical current, heat, and electromagnetic waves in two-phase composites. We discover that for a new class of random matrices at the heart of transport phenomena, phase connectedness determines the nature of the system statistics. A striking transition to universality is observed as a percolation threshold is approached, with eigenvalue statistics shifting from weakly correlated toward highly correlated, repulsive behavior of the Gaussian orthogonal ensemble. Moreover, the eigenvectors exhibit behavior similar to Anderson localization in quantum systems, with ”mobility edges” separating extended states. Our findings help explain the collapse of spectral gaps and associated critical behavior near percolation thresholds. The spectral transition is explored in resistor networks, bone, and sea ice structures important in climate modeling. Abbreviations ACM, analytic continuation method; ESD, eigenvalue spacing distribution; GOE, Gaussian orthogonal ensemble; RME, random matrix ensemble; RMT, random matrix theory; RRN, random resistor network Complex systems with a large number of interacting components are ubiquitous in the physical and biological sciences, and have a broad range of applications. As the number of components increases, structure begins to emerge from the underlying randomness of the system, and universal behavior can arise. Random matrix theory (RMT) has been quite successful in modeling universal features found in statistical fluctuations of characteristic quantities of complex systems, such as atomic nuclei [1], biological networks [2], financial time series [3], the departure times for bus systems [4], quantum systems whose classical counterparts are chaotic [1, 5], and mesoscopic conductors [6]. Remarkably, it has been shown that even the (non-random) zeros of the famous Riemann zeta function have universal features which are very accurately captured by RMT [5, 7, 8]. Each of these systems can be described in terms of discrete “spectra” confined to a line. In RMT, the spectra of such systems are modeled by the eigenvalues of ensembles of large random matrices called random matrix ensembles (RME’s). Long and short range correlations of these random eigenvalues are measured in terms of various eigenvalue statistics introduced by Dyson and Mehta [8]. The localization properties of the associated eigenvectors are characterized by quantities such as the inverse participation ratio (IPR) [3] and the Shannon information entropy [9]. In many systems, the spectral statistics are parameter independent when properly scaled [1, 10] and ∗ Electronic address: golden@math.utah.edu fall into two universal categories: uncorrelated Poisson statistics [1] and highly correlated Wigner–Dyson (WD) statistics [1, 11]. The statistical behavior of the spectrum is related to the extent that the eigenfunctions overlap. A key example is the metal-insulator Anderson transition exhibited by noninteracting electrons in a random potential [10, 12]. For small disorder, the electron wave functions are extended structures. Their overlap gives rise to correlated WD energy level statistics with strong level repulsion [12]. However, for large disorder, electrons are typically localized at different points in the sample and “do not talk to each other,” resulting in uncorrelated Poisson level statistics [12]. During transitions from small to large disorder, the wave functions become increasingly localized and an intermediate Poisson-like behavior of level statistics arises [6, 10]. There are other systems which undergo an analogous spectral transition as a system parameter varies. Among them are: the hydrogen atom in a magnetic field [13], random points on a fractal set [14], quantum chaos [15], and complex networks [9]. We demonstrate here that transitions in the connectedness or percolation properties of macroscopic composite media, with microstructural scales spanning many orders of magnitude, can also be characterized by a transition in the statistics of eigenvalues and the delocalization of eigenvectors of a random matrix. This is a new type of random matrix within RMT that depends only on the geometry of the composite medium, and not directly on a probability distribution as usual. While the connectedness driven transition in the statistical behavior of eigenvalues is analogous to that of the Anderson transition and other systems, the delocalization of eigenvectors reveals new subtleties that distinguish the behavior we see from classical Anderson localization. This large family of random matrices arises in the analytic continuation method (ACM) [16–19] for representing transport in composites. The method provides Stieltjes integral representations for the bulk transport coefficients of a two-component ran- 2 A B C D E FIG. 1: Connectedness transitions in composite structures. (A)–(E) Increasingly connected composites from left to right. (A) Realizations of the two-dimensional lattice percolation model, with (black) bond probabilities p = 0.20, p = 0.30, and p = pc = 0.5. (B) cross-sections from X-ray CT volume renderings of the brine phase within a lab-grown sea ice single crystal (H. Eicken), with image brine area fractions of φ = 0.20, φ = 0.51, and φ = 0.70. (C) Melt ponds on the surface of Arctic sea ice (D. K. Perovich), with area fractions φ = 0.09, φ = 0.27, and φ = 0.57. (D) Arctic sea ice pack (D. K. Perovich), with open ocean area fractions φ = 0.06, φ = 0.10, and φ = 0.47. (E) SEM images of osteoporotic (left) and healthy (right) trabecular bone (P. Hansma), with cross-sectional area fractions φ = 0.26 and φ = 0.55. dom medium, such as the effective electrical conductivity ~ involving σ ∗ of a medium immersed in an electric field E, a spectral measure µ of the random matrix [18, 20]. The measure µ exhibits fascinating transitional behavior as a function of system connectivity, which controls critical behavior of σ ∗ near connectedness thresholds. For example, in the case of a random resistor network (RRN) with a low volume fraction p of open bonds, as shown in Fig. 1A, there are spectrum-free regions at the spectral endpoints λ = 0, 1 [21]. However, as p approaches the percolation threshold pc [22, 23] and the system becomes increasingly connected, these spectral gaps shrink and then vanish [21, 24], leading to the formation of δcomponents of µ at the spectral endpoints, precisely [21] when p = pc and p = 1 − pc . This leads to critical behavior of σ ∗ for insulating/conducting and conducting/superconducting systems [21]. This gap behavior of µ has led [21] to a detailed description of these critical transitions in σ ∗ , which is directly analogous to the Lee–Yang–Ruelle–Baker description [25, 26] of the Ising model phase transition in a ferromagnet’s magnetization M . Moreover, using this gap behavior, all of the classical critical exponent scaling relations were recovered [21, 26] without heuristic scaling forms but instead by using the rigorous integral representation for σ ∗ involving µ. Our results here reveal a mechanism for the collapse in the spectral gaps of µ and illustrate that localized and extended eigenvectors of the matrix are in direct corre~ that spondence with components of the electric field E are localized in, and extended throughout the composite medium. In particular, we demonstrate that eigenvalues associated with a disordered state, such as a low volume fraction RRN, are weakly correlated and welldescribed by Poisson-like statistics. However, as the percolation threshold pc is approached and the system develops long range order, the eigenvalues become increasingly correlated and their statistics approach classical WD statistics, causing the eigenvalues to spread out due to increased level repulsion, subsequently forming δcomponents in µ at the spectral endpoints. Correspondingly, the eigenvectors become increasingly extended and those associated with these δ-components are typically highly extended. These regions of extended states are separated from each other by “mobility edges” [1] of lo~ involving calized states. A resolvent representation of E the random matrix provides a one-to-one correspondence between localized (extended) eigenvectors and localized (extended) components of the electric field within the medium. We show that this spectral behavior emerges in a variety of composite systems, such as the brine microstructure of sea ice [27–29], melt ponds on the surface of Arctic sea ice [30], the sea ice pack itself, and porous human bone [31]. Our results indicate that it is pervasive in such macroscopic systems and arises simply from connectedness – at the most basic level of characterizing any physical system with inhomogeneities. The behavior of composite materials exhibiting a critical transition as system parameters are varied is particularly challenging to describe physically, and to predict mathematically. Here, we discuss composites which exhibit critical behavior in transport properties induced by transitions in connectedness or percolation properties of a particular material phase. Lattice and continuum percolation models have been used to study a broad range of disordered materi- 3 als [22, 23]. In the simplest case of the two-dimensional square lattice [22, 23], as shown in Fig. 1A, the bonds are open with probability p and closed with probability 1 − p. Connected sets of open bonds are called open clusters. The average cluster size grows as p increases. When the system size L tends to infinity, there is a critical probability pc , 0 < pc < 1, called the percolation threshold, where an infinite cluster of open bonds first appears. In dimension d = 2, pc = 1/2, and in d = 3, pc ≈ 0.2488 [22]. Now consider transport through the associated RRN, where the bonds are assigned electrical conductivities σ1 with probability p and σ2 with probability 1 − p. The effective conductivity σ ∗ exhibits two types of critical behavior. First, when 0 < σ1 < ∞ and σ2 → 0, the system is insulating, σ ∗ = 0, for p < pc and is conducting, σ ∗ > 0, for p > pc . Second, when σ1 → ∞ and 0 < σ2 < ∞, the conductive system becomes superconducting, σ ∗ → +∞, as p → p− c . Sea ice is a complex composite consisting of pure ice with sub-millimeter scale brine inclusions, as shown in Fig. 1B, whose volume fraction φ, geometry, and connectedness vary significantly with temperature T . The brine microstructure displays a percolation threshold at a critical brine volume fraction φc ≈ 5% in columnar sea ice [27], which corresponds to a critical temperature Tc ≈ −5◦ C for a typical bulk salinity of 5 ppt. This threshold acts as an on–off switch for fluid flow through sea ice, and is known as the rule of fives. It leads to critical behavior of fluid flow, where sea ice is effectively impermeable to fluid transport for φ < φc , yet is permeable for φ > φc , with the permeability as a function of φ above the 5% threshold described by the universal critical exponent for lattices in three dimensions [27–29]. Fluid flow through sea ice mediates a broad range of physical and biological processes in the polar marine environment [28, 29], including brine drainage, nutrient replenishment for algal communities in the brine phase, snow-ice formation, and the evolution of melt ponds on the surface of Arctic sea ice [30]. Melt ponds (Fig. 1C) determine sea ice albedo in the Arctic, a key parameter in climate modeling. Despite its importance, it remains a major source of uncertainty in climate models. In fact, the lack of inclusion of melt ponds in previous generations of climate models is believed to partially account for the inadequacy of these models to predict the dramatic rate of melting of the summer Arctic sea ice pack. The results here advance our understanding of the effective or homogenized properties of the ice pack, and help provide a path toward more rigorously incorporating sea ice into climate models. Human bone also displays a complex, porous microstructure, as shown in Fig. 1E. The strength of bone and its ability to resist fracture depend strongly on the quality of the connectedness of the hard, solid phase. Osteoporotic trabecular bone can become more disconnected and remaining connections can become more tenuous or fragile [31]. The spectral techniques [32, 33] which have arisen from the ACM provide important methods for analyzing microstructural transitions in bone and its biophysical properties. I. MATHEMATICAL METHODS Random matrices arise naturally in the ACM for representing transport in composites [18, 20]. This method provides Stieltjes integral representations for the effective parameters of two-component composite media, such as electrical conductivity and permittivity, magnetic permeability, and thermal conductivity. The integral representations involve a spectral measure µ associated with a family of random matrices, which depend only on the composite geometry [20]. A remarkable feature of this method is that once the spectral measure is found for a given composite geometry, by spectral coupling of the governing equations [34], the effective electrical, magnetic, and thermal transport properties are all determined by µ. Consider the effective parameter problem for twocomponent conductive media [18, 20]. The electromagnetic transport properties of the composite are governed by the quasi-static limit [18] of Maxwell’s equations ~ ×E ~ = 0, ∇ ~ · J~ = 0. ∇ (1) ~ and J~ are the random electric field and current Here, E ~ and σ denotes the density, which are related by J~ = σ E, electrical conductivity of the locally isotropic, stationary random medium. In the case of a two-component medium with (complex-valued) component conductivities σ1 and σ2 we write σ = σ1 χ1 + σ2 χ2 , (2) where χ1 is the characteristic function of medium one, taking the value 1 in medium one and 0 otherwise, with χ2 = 1 − χ1 . The effective conductivity matrix σ ∗ is defined by [18] ~ hJ~ i = σ ∗ hEi, ~ =E ~ 0. hEi (3) ~0 = Here, h·i denotes ensemble averaging and the vector E E0~ek has magnitude E0 and direction ~ek , taken to be the kth standard basis vector for some k = 1, . . . , d, which defines the d-dimensional coordinate system. Equivalently, the effective conductivity σ ∗ may be defined [18, 20] ~ = in terms of the energy (power) density : hJ~ · Ei ~0 · E ~ 0 = σ ∗ E 2 . For simplicity, we focus on the diagσ∗ E kk 0 ∗ ∗ onal coefficient σkk of the matrix σ ∗ and set σ ∗ = σkk . The key step in the method is obtaining the following Stieltjes integral representation for σ ∗ [16–18] Z 1 dµ(λ) ∗ σ = σ2 (1 − F (s)), F (s) = , (4) 0 s−λ which follows from a resolvent representation of the electric field (in material phase 1) [20] ~ = sE0 (sI − χ1 Γχ1 )−1 χ1~ek . χ1 E (5) 4 B A 0.3 0.4 0.3 0.2 0.2 0.1 0.1 0 0 -0.1 -0.1 FIG. 2: The projection matrix Γ. The matrix Γ with periodic boundary conditions for (A) 2D and (B) 3D networks with system sizes L = 6 and L = 3, respectively, with corresponding matrix size N = Ld d. Notice that the lengths of the repeated structures are multiples of the system size L. Here, s = 1/(1 − σ1 /σ2 ), −F (s) plays the role of the effective electric susceptibility, µ is a spectral measure associated with the random operator χ1 Γχ1 [18], and −1 ~ ~ Γ = −∇(−∆) ∇· is a projection onto curl-free fields, based on convolution with the free-space Green’s function for the Laplacian ∆ = ∇2 . In this way, the ACM determines a homogeneous medium which behaves macroscopically and energetically like a given inhomogeneous medium. Moreover, parameter information in s and E0 is separated from the geometric complexity of the system, which is encoded in the spectral measure µ. Geometric information about the composite is incorporated into the positive Stieltjes meaR1 sure µ in equation (4) via its moments, µn = 0 λn dµ(λ). For example, the mass µ0 of the measure is the volume fraction p of medium one, i.e., µ0 = hχ1 i = p. We may think of the measure dµ(λ) as m(λ) dλ for some density m(λ) which is allowed to have δ-components. This characterization of µ is more transparent in the setting of a finite RRN, where the random operator χ1 Γχ1 can be represented by a real-symmetric random matrix of size N × N , where N = Ld d [20]. In this case, χ1 is a diagonal matrix with 1’s and 0’s along the diagonal, corresponding to conductive bonds with conductivity σ1 and σ2 , respectively, and Γ is a (non-random) projection matrix [20]. In this case, the spectral measure µ can be calculated directly from the eigenvalues λj , j = 1, . . . , N , and eigenvectors ~vj of the matrix χ1 Γχ1 , and is given by a weighted sum of δ-measures N X dµ(λ) = hmj δ(λ − λj ) i dλ = m(λ) dλ, (6) j=1 where h·i denotes ensemble averaging, mj = [~vj · êk ]2 , and êk is a lattice basis vector [20]. The matrix Γ encapsulates the lattice topology and transport characteristics of the resistor network and displays a rich geometric, banded structure, as shown in Fig. 2. The action of the matrix χ1 in χ1 Γχ1 is to randomly zero out each row and column of Γ that corresponds to diagonal components of χ1 satisfying [χ1 ]jj = 0. The spectral weights mj associated with this null space of χ1 Γχ1 are identically zero [20] and do not contribute to the sum in (6). The only eigenvectors that contribute to this sum are associated with random N1 × N1 submatrices of Γ corresponding to the N1 components of χ1 satisfying [χ1 ]jj = 1, where N1 ≈ pN . The discretized sea ice and bone composite structures displayed in Fig. 1 were converted to 2D binary fluid/ice and bone/marrow representations (the ice pack images in (D) are the converted versions). The geometry and the component connectivity of these binary images can be described in terms of high density two-component resistor networks, with fluid and bone corresponding to component one, and ice and marrow corresponding to component two. In this way, the matrix χ1 Γχ1 was obtained for these composite structures. In the next section, we show that its associated eigenvalues and eigenvectors undergo a connectedness driven transition that is analogous to that of the Anderson transition [1, 10, 12]. II. RESULTS In RMT, long and short range correlations of eigenvalues in the bulk of the spectrum [10] are measured in terms of various eigenvalue statistics. In order for the fluctuation properties of the eigenvalues about the mean density ρ(λ) to be compared to the predictions of RMT, the spectrum has to be unfolded [1, 3, 10]. It is nontrivial to unfold spectra associated with the RRN for small volume fractions p, due to prominent “geometric” resonances in ρ(λ) [20, 24], which has limited our analysis of these systems in the dilute limit, p 1. The nearest neighbor eigenvalue spacing distribution (ESD) P (z) is the observable most commonly used to study short-range correlations [1]. For highly correlated WD spectra, such as that exhibited by the realsymmetric matrices of the Gaussian orthogonal ensemble (GOE) [1, 11], the ESD is accurately approximated by P (z) ≈ (πz/2) exp(−πz 2 /2), known as Wigner’s surmise [1, 6], which illustrates the phenomenon of eigenvalue repulsion, vanishing linearly in the limit of small spacings z. In contrast, the ESD for uncorrelated Poisson spectra, P (z) = exp(−z), allows for level degeneracy. In Fig. 3 we display the behavior of the ESD for eigenvalues of the matrix χ1 Γχ1 , which correspond to RRN, melt pond, and Arctic sea ice pack composite structures displayed in Fig. 1. This figure demonstrates that for sparsely connected systems, the behavior of the ESD’s are well described by weakly correlated Poissonlike statistics [10]. The ESD’s increase linearly from zero at short separation but the initial slope is steeper, implying less level repulsion. As the systems become increasingly connected, and long range order is established, the ESD’s transition toward highly correlated WD statistics with high level repulsion. The blue dash-dot curve displayed in Fig. 3 is the ESD for Poisson spectra, while the green dashed curve is the ESD for the GOE. For the 2D and 3D RRN, the eigenvalue density ρ(λ, p) displays the 5 A Arctic sea ice pack 1 Σ2(L) P(z) 0.8 A p=0.47 p=0.10 p=0.06 0.6 0.4 0.2 0 B 0 1 2 3 z 4 0 1 2 p=0.09 3 z 4 0 1 2 3 z 4 B 0.6 0.4 0.2 0 C 0 1 2 1 P(z) 4 0 1 2 3 z 4 0 1 2 3 z 4 p=0.1 p=0.9 p=0.3 p=0.7 p=0.5 0.6 0.4 0.2 0 0 1 2 3 z 4 0 1 2 3 4 0 1 2 3 z 4 3D random resistor network 1 p=0.05 p=0.95 0.8 P(z) z 2D random resistor network 0.8 D 3 p=0.2488 p=0.4 p=0.5 p=0.6 p=0.7512 p=0.13 p=0.87 0.6 0.4 Arctic pack ice 0 1.2 0.8 0.4 0 10 20 L 30 0 Trabecular bone 10 10 20 L 30 0 20 L 30 0 2D RRN 10 Brine inclusions φ=0.20 φ=0.51 φ=0.70 p=0.1 p=0.3 p=0.5 p=0.7 p=0.9 p=0.26 p=0.55 0 Arctic melt ponds p=0.09 p=0.27 p=0.57 p=0.06 p=0.10 p=0.47 2 1.6 p=0.57 p=0.27 ∆3(L) 0.8 P(z) 1 Arctic melt ponds 30 25 20 15 10 5 0 10 20 L 30 3D RRN p=0.05 p=0.13 p=0.2488 p=0.5 p=0.7512 p=0.87 p=0.95 20 L 30 0 10 20 L 30 FIG. 4: Long-range eigenvalue correlations. (A) The eigenvalue number variance Σ2 (L ) and (B) the spectral rigidity ∆3 (L ) for Poisson (blue dash-dot) and Wigner–Dyson (WD) (green dash) spectra are displayed along with those of sea ice brine inclusions, Arctic melt ponds and pack ice, trabecular bone microstructure and random resistor networks (RRN). As the systems become increasingly connected, these long-range eigenvalue statistics transition from the linear Poisson-like behavior toward a logarithmic WD behavior. 0.2 0 0 1 2 3 z 4 0 1 2 3 z 4 0 1 2 3 z 4 FIG. 3: Short-range eigenvalue correlations. The eigenvalue spacing distributions (ESD’s) for Poisson (blue dashdot) and Wigner–Dyson (WD) (green dashed) spectra are displayed in (A)–(D), along with that of (A) the Arctic sea ice pack, (B) Arctic melt ponds, (C) 2D random resistor network (RRN), and (D) 3D RRN. As the systems become increasingly connected, the initial slopes of the ESD’s progressively decrease, indicating an increase in level repulsion and shortrange correlations. symmetry ρ(λ, p) = ρ(1 − λ, 1 − p) in the bulk of the spectrum. This is reflected in the ESD’s by the symmetry P (z, p) = P (z, 1 − p), as shown for the 2D and 3D RRN in Fig. 3C and Fig. 3D. The ESD’s displayed in Fig. 3D for the 3-dimensional percolation model suggest that the GOE limit is attained for all pc ≤ p ≤ 1 − pc . The ESD contains information about the spectrum which involves short scales (a few mean spacings) [1, 10]. Long-range correlations are measured by quantities such as the eigenvalue number variance Σ2 (L ), in intervals of length L (not to be confused with the system size L), and the closely related spectral rigidity ∆3 (L ) [1]. For uncorrelated Poisson spectra, these long range statistics are linear, with Σ2 (L ) = L and ∆3 (L ) = L/15, as displayed in blue color with dash-dot line style in Fig. 4. In contrast, the strong correlations of WD spectra make the spectrum more rigid [10] so that Σ2 (L ) and ∆3 (L ) grow only logarithmically [1, 8]. The green dashed curves of Fig. 4 are numerical computations of the exact solutions of Σ2 (L ) and ∆3 (L ) for WD spectrum [1]. In Fig. 4, we also display the behavior of these longrange eigenvalue statistics for the matrix χ1 Γχ1 , corre- sponding to the composite structures displayed in Fig. 1. For sparsely connected systems, these statistics exhibit linear Poisson-like behavior away from the origin with slope less than their Poisson counterparts. This linear behavior has been attributed to exponentially decaying correlations of eigenvalues [10]. These statistics transition with increasing connectedness toward logarithmic WD behavior typical of the GOE, which has quadratically decaying eigenvalue correlations [10]. Similar to the ESD’s in Fig. 3, for the RRN these statistics also display the symmetry Σ2 (L, p) = Σ2 (L, 1 − p) and ∆3 (L, p) = ∆3 (L, 1 − p). Moreover, Fig. 3D and Fig. 4B suggest that the GOE limit is attained by the short and long range statistics for the 3-dimensional RRN for all pc ≤ p ≤ 1 − pc . They appear to overlie the GOE limit almost exactly for all p values tested in the range pc ≤ p ≤ 1 − pc . With this in mind, we can paint a heuristic analogy with Anderson localization, where low disorder corresponds to extended states and WD statistics. When disorder exceeds a critical level, the states localize and the eigenvalues become de-correlated. We view the 3D RRN with pc ≤ p ≤ 1−pc to be “ordered” with extended states and WD statistics. As p decreases, the disorder − or blockages to the flow − increases, and the eigenstates localize (see below) and the eigenvalue repulsion diminishes. The eigenvectors ~vn , n = 1, . . . , N1 , associated with the random N1 × N1 sub-matrices of Γ, discussed above, also undergo a connectedness driven transition in their localization properties. Two commonly used quantities which measure the localization of the eigenvector ~vn are the inverse participation ratio (IPR) In [3] and the Shan- 6 non entropy Sn [9], defined as Localized Mode x10 - 12 Extended Mode x10 - 2 Log Electric Field A 8 -2 N1 X In = [vnj ]4 , [vnj ]2 ln[vnj ]2 , (7) -8 0 0 j=1 where vnj is the jth component of the eigenvector ~vn . Eigenvectors of matrices in the GOE are known to be highly extended and independent of the distribution of the eigenvalues [35]. In this case, the IPR and entropy are given by IGOE = 3/N1 [3] and SGOE ≈ ln(N1 /2.07) [9], respectively. Associated with the Sn is the eigenvector localization length ` defined as -14 (8) where we denote by hSi and hIi ensemble averaging over all values of the Sn and In . The meaning of In and Sn can be illustrated by two limiting cases (i) a normalized vector with only one component vnj = 1 has In = 1 and Sn = 0, √ whereas (ii) a vector with identical components vnj = 1/ N1 has In = 1/N1 and Sn = ln N1 . When all of the vectors in an ensemble are of type (i) then ` ≈ 2.07, while ` ≈ 2.07N1 when all are of type (ii). If hSi = SGOE then ` = N1 . In the matrix setting, the electric field in equation (5) has the following eigenvector expansion ~ = sE0 χ1 E N1 X [(s − λn )−1 (~vn · χ1~ek )] ~vn . (9) n=1 This provides a direct link between localized eigenvectors ~ that have significant magni~vn and eigen-modes of χ1 E tudes in only a few resistors, while extended eigenvectors correspond to electric field components that extend throughout the network. Displayed in Fig. 5A, from left ~ and examples of to right, is the total electric field χ1 E ~ for a random localized and extended eigen-modes of χ1 E realization of the 2D RRN with p = pc . III. DISCUSSION Our results displayed in Fig. 5B–D show that the eigenvectors ~vn associated with 2D and 3D RRN undergo a fascinating delocalization as p increases and the system becomes increasingly connected. For example, Fig. 5B displays the IPR In of the 3D RRN as a function of the index n of the eigenvector ~vn , with increasing index corresponding to increasing eigenvalue magnitude λn . This figure shows that for p pc , the eigenvectors are typically localized, with values In of IPR quite different from that of the GOE, identified by the red horizontal lines. Moreover, they have an oscillatory behavior that, when plotted as a function of λn , follows the peaks and valleys of “geometric” resonances exhibited by the eigenvalue density ρ(λ) for small p [20, 24], with more localized regions corresponding to lower density. This indicates that there are significant correlations between -1 -4 IPR vs vector index n B 10 0 p=0.05 p=0.35 p=0.7512 10 -1 10 -2 10 -3 50 ` = N1 exp[−(SGOE − hSi)], 1 4 In j=1 Sn = − N1 X 150 n 250 400 800 1200 1600 n Inverse Participation Ratio C 2D 〈I〉 3D 〈I〉 2D IGOE 3D IGOE 0.12 0.08 D 1000 2000 3000 n Entropy and localization length 0.8 0.4 2D 〈S〉/SGOE 3D 〈S〉/SGOE 2D l/lGOE 3D l/lGOE 0.04 0 0 0 0.2 0.4 0.6 0.8 p 1 0 0.2 0.4 0.6 0.8 p 1 FIG. 5: Delocalization of eigenvectors. (A) Displayed ~ (in log scale) from left to right are the full electric field χ1 E ~ and examples of localized and extended eigen-modes of χ1 E (in linear scale) for a realization of the 2D random resistor network (RRN), with a system size L = 50 and volume fraction p = pc = 1/2. The corresponding localized and extended eigenvectors have inverse participation ratios (IPR’s) In ≈ 0.081 and In ≈ 0.002, and the DC values of the component conductivities are those for silver and silicon at 20 ◦ C, σ1 = 6.3×107 S/m and σ2 = 1.56×10−3 S/m. (B) The IPR’s of eigenvectors ~vn associated with realizations of the 3D RRN plotted versus eigenvector index n, with L = 12 and increasing values of p from left to right, with 1 − pc ≈ 0.7512. The vertical lines define the δ-components of the spectral measure µ at the left and right spectral endpoints, where the associated eigenvalues λn satisfy λn . 10−14 and λn & 1 − 10−14 , respectively, while the horizontal lines mark the IPR value IGOE = 3/N1 for the Gaussian orthogonal ensemble (GOE) with matrix size N1 ≈ pN , where N = Ld d. (C) The ensemble averaged IPR hIi as a function of p, displaying transitional behavior at the percolation thresholds, pc = 1/2 for 2D and pc ≈ 0.2488 for 3D. (D) The normalized, ensemble averaged entropy hSi/SGOE and associated localization length `/`GOE as a function of p, where SGOE ≈ ln(N1 /2.07) and `GOE = N1 . Panels (B)–(D) demonstrate that, as p increases and the systems become increasingly connected, the eigenvectors become progressively extended, on average, with decreasing hIi and increasing hSi/SGOE and `/`GOE . the eigenvalues and eigenvectors, in contrast with the GOE. As p → p− c , gaps in the spectrum about the spectral endpoints shrink and then vanish [20, 24], while the values In of the IPR continually decrease, approaching the GOE limit, as shown in Fig. 5B. As p surpasses pc and 1 − pc , δ-components form in the spectral measure µ at λ = 0 and λ = 1, respectively [21]. Numerically, the δ-component at λ = 0 manifests itself as 7 a large number of eigenvalues with magnitude . 10−14 , followed by an abrupt change of magnitude & 10−4 , with no eigenvalues in the interval (10−14 , 10−4 ), and similarly for λ = 1. The locations of these changes in magnitude are identified by red vertical lines in Fig. 5B, demonstrating that the eigenvectors associated with these δcomponents are typically more extended than the others, with values In of the IPR closer to the GOE limit. The entropy Sn has a similar behavior. In Fig. 5C and 5D the p dependence of hIi, hSi/SGOE , and `/`GOE are displayed. As p increases and the system becomes increasingly connected, hIi decreases, with transitional behavior at the percolation threshold pc , while hSi/SGOE and `/`GOE increase. Each indicate that the eigenvectors, hence the eigen-modes of the electric ~ become progressively extended throughout the field χ1 E network with increasing system connectedness. Fig. 5B shows that this average delocalization is largely due to the formation of the δ-components in µ. Fig. 5B also shows that regions of extended states are separated from one another by a series of “mobility edges” with a sudden increase in the number of localized eigenvectors. This behavior in the eigenvectors of the random matrix χ1 Γχ1 is analogous to that of Anderson localization, where mobility edges mark the characteristic energies of the metal/insulator transition [1]. Remarkably, the mobility edges in Fig. 5B are precisely at the locations of the δ-components, identified by red vertical lines, which control critical behavior of transport in insulator/conductor and conductor/superconductor systems. However, the delocalization behavior of the spectral endpoints displayed in Fig. 5B is different from that of Anderson localization [1], further demonstrating that the behavior we see in χ1 Γχ1 is new to RMT. We have demonstrated that the eigenvalues and eigen- vectors associated with the random matrix χ1 Γχ1 , corresponding to various composite structures, have a statistical behavior that is analogous to, but distinctly different from that of the Anderson transition. The eigenvalues of χ1 Γχ1 shift from weakly correlated Poisson-like statistics toward highly correlated WD statistics as a function of disorder (connectedness). Correspondingly, the eigenvectors undergo a delocalization, with highly extended states appearing at the spectral edges, separated by mobility edges of localized states. Moreover, the delocalization of eigenvectors corresponds to a delocalization of ~ extending the transport field, such as the electric field E, throughout the composite medium near global connectedness thresholds. The disorder driven transition in the statistical behavior of the eigenvalues also accounts for the gap behavior of the spectral measure µ, underlying exact integral representations for the effective transport coefficients of the composite medium, such as effective conductivity σ ∗ , which, in turn, governs their critical behavior via the formation of δ-components of the measure at the spectral endpoints. This provides a novel way of understanding and characterizing disorder driven transitions in the effective transport properties of composite media, and opens a new chapter in the application of RMT to the analysis of complex, macroscopic systems. [1] T. Guhr, A. M uller-Groeling, and H. A. Weidenmüller. Random-matrix theories in quandom physics: Common concepts. Phys. Rept., 299(4):189–425, 1998. [2] F. Luo, Y. Yang, J. Zhong, H. Gao, L. Khan, D. Thompson, and J. Zhou. Constructing gene co-expression networks and predicting functions of unknown genes by random matrix theory. BMC Bioinformatics, 8(1), 2007. [3] V. Plerou, P. Gopikrishnan, B. Rosenow, L. A. N. Amaral, T. Guhr, and H. E. Stanley. Random matrix approach to cross correlations in financial data. Phys. Rev. E, 65:066126, Jun 2002. [4] M. Krbálek and P. Seba. The statistical properties of the city transport in cuernavaca (mexico) and random matrix ensembles. J. Phys. A: Math. Gen., 33:L229?L23, 2000. [5] T. Kriecherbauer, J. Marklof, and A. Soshnikov. Random matrices and quantum chaos. Proceedings of the National Academy of Sciences, 98(19):10531–10532, 2001. [6] A. D. Stone, P. A. Mello, K. A. Muttalib, and J. L. Pichard. Random Matrix Theory and Maximum Entropy Models for Disordered Conductors, chapter 9, pages 369– 448. Elsevier Science Publishers, Amsterdam, Nether- lands, 1991. [7] F. Mezzadri and N. C. Snaith. Recent Perspectives in Random Matrix Theory and Number Theory. Cambridge University Press, Cambridge, UK, 2005. [8] M. L. Mehta. Random Matrices. Elsevier, Amsterdam, third edition, 2004. [9] J. A. Méndez-Bermúdez, A. Alcazar-López, A. J. Martı́nez-Mendoza, Francisco A. Rodrigues, and Thomas K. DM. Peron. Universality in the spectral and eigenfunction properties of random networks. Phys. Rev. E, 91:032122, Mar 2015. [10] C. M. Canali. Model for a random-matrix description of the energy-level statistics of disordered systems at the Anderson transition. Phys. Rev. B, 53(7):3713–3730, 1996. [11] P. Deift. Random Matrix Theory: Invariant Ensembles and Universality. Courant Institute of Mathematical Sciences, New York, NY, 2009. [12] V. E. Kravtsov and K. A. Muttalib. New class of random matrix ensembles with multifractal eigenvectors. Phys. Rev. Lett., 79(10):1913–1916, 1997. Acknowledgments We gratefully acknowledge support from the Division of Mathematical Sciences and the Division of Polar Programs at the U.S. National Science Foundation (NSF) through Grants DMS-1009704, ARC0934721, DMS-0940249, and DMS-1413454. We are also grateful for support from the Office of Naval Research (ONR) through Grants N00014-13-10291 and N0001412-10861. Finally, we would like to thank the NSF Math Climate Research Network (MCRN) for their support of this work. 8 [13] D. Delande. Les Houches, école d’été de physique théorique. In M. J. Giannoni, A. Voros, and J. ZinnJustin, editors, Chaos et Physique Quantique—Chaos and Quantum Physics = Les Houches, session LII, 1–31 août 1989, pages 665–726. North-Holland, Amsterdam, 1991, 1991. [14] J. Sakhr and J. M. Nieminen. Poisson-to-Wigner crossover transition in the nearest-neighbor statistics of random points on fractals. Phys. Rev. E, 72:045204, Oct 2005. [15] T. H. Seligman, J. J. M. Verbaarschot, and M. R. Zirnbauer. Spectral fluctuation properties of Hamiltonian systems: the transition region between order and chaos. J. Phys. A: Math. Gen., 18:2751–2770, 1985. [16] D. J. Bergman. Exactly solvable microscopic geometries and rigorous bounds for the complex dielectric constant of a two-component composite material. Phys. Rev. Lett., 44:1285–1287, 1980. [17] G. W. Milton. Bounds on the complex dielectric constant of a composite material. Appl. Phys. Lett., 37:300–302, 1980. [18] K. M. Golden and G. Papanicolaou. Bounds for effective parameters of heterogeneous media by analytic continuation. Commun. Math. Phys., 90:473–491, 1983. [19] G. W. Milton. Theory of Composites. Cambridge University Press, Cambridge, 2002. [20] N. B. Murphy, E. Cherkaev, C. Hohenegger, and K. M. Golden. Spectral measure computations for composite materials. Commun. Math. Sci., 13(4):825–862, 2015. [21] N. B. Murphy and K. M. Golden. The Ising model and critical behavior of transport in binary composite media. J. Math. Phys., 53:063506 (25pp.), 2012. [22] D. Stauffer and A. Aharony. Introduction to Percolation Theory. Taylor and Francis, London, second edition, 1992. [23] S. Torquato. Random Heterogeneous Materials: Microstructure and Macroscopic Properties. SpringerVerlag, New York, 2002. [24] T. Jonckheere and J. M. Luck. Dielectric resonances of binary random networks. J. Phys. A: Math. Gen., 31:3687–3717, 1998. [25] G. A. Baker. Quantitative Theory of Critical Phenomena. Academic Press, New York, 1990. [26] K. M. Golden. Critical behavior of transport in lattice and continuum percolation models. Phys. Rev. Lett., 78(20):3935–3938, 1997. [27] K. M. Golden, S. F. Ackley, and V. I. Lytle. The percolation phase transition in sea ice. Science, 282:2238–2241, 1998. [28] K. M. Golden, H. Eicken, A. L. Heaton, J. Miner, D. Pringle, and J. Zhu. Thermal evolution of permeability and microstructure in sea ice. Geophys. Res. Lett., 34:L16501 (6 pages and issue cover), 2007. [29] K. M. Golden. Climate change and the mathematics of transport in sea ice. Notices Amer. Math. Soc., 56(5):562–584 and issue cover, 2009. [30] C. Hohenegger, B. Alali, K. R. Steffen, D. K. Perovich, and K. M. Golden. Transition in the fractal geometry of Arctic melt ponds. The Cryosphere, 6(5):1157–1162, 2012. [31] J. Kabel, A. Odgaard, B. van Rietbergen, and R. Huiskes. Connectivity and the elastic properties of cancellous bone. Bone, 24(2):115–120, 1999. [32] K. M. Golden, N. B. Murphy, and E. Cherkaev. Spectral analysis and connectivity of porous microstructures in bone. J. Biomech., 44(2):337–344, 2011. [33] E. Cherkaev and C. Bonifasi-Lista. Characterization of structure and properties of bone by spectral measure method. J. Biomech., 44(2):345–351, 2011. [34] E. Cherkaev. Inverse homogenization for evaluation of effective properties of a mixture. Inverse Problems, 17(4):1203–1218, 2001. [35] S. N. Evangelou. A numerical study of sparse random matrices. J. Stat. Phys., 69(1-2):361–383, 1992.