Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Balance with Memory in Signed Networks via Mittag-Leffler Matrix Functions

Yu Tian Nordita, Stockholm University and KTH Royal Institute of Technology, SE-106 91 Stockholm, Sweden (yu.tian@su.se).    Ernesto Estrada Institute of Cross-Disciplinary Physics and Complex Systems, IFISC (UIB-CSIC), Palma de Mallorca, 07122, Spain (estrada@ifisc.uib-csic.es).
Abstract

Structural balance is an important characteristic of graphs/networks where edges can be positive or negative, with direct impact on the study of real-world complex systems. When a network is not structurally balanced, it is important to know how much balance still exists in it. Although several measures have been proposed to characterize the degree of balance, the use of matrix functions of the signed adjacency matrix emerges as a very promising area of research. Here, we take a step forward to using Mittag-Leffler (ML) matrix functions to quantify the notion of balance of signed networks. We show that the ML balance index can be obtained from first principles on the basis of a nonconservative diffusion dynamic, and that it accounts for the memory of the system about the past, by diminishing the penalization that long cycles typically receive in other matrix functions. Finally, we demonstrate the important information in the ML balance index with both artificial signed networks and real-world networks in various contexts, ranging from biological and ecological to social ones.

1 Introduction

The use of matrix functions [31] has represented a significant advance in the development of mathematical models of networks G=(V,E)𝐺𝑉𝐸G=\left(V,E\right)italic_G = ( italic_V , italic_E ) in the last 20 years [7, 21]. In particular, the use of functions of the adjacency matrix A𝐴Aitalic_A of a network, f(A)𝑓𝐴f\left(A\right)italic_f ( italic_A ), has impacted the areas of study of vertex centrality measures [7] as well as our understanding of the navigability of networks [17, 44]. The mathematical roots of these developments come from the fact that (Ak)uvsubscriptsuperscript𝐴𝑘𝑢𝑣\left(A^{k}\right)_{uv}( italic_A start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_u italic_v end_POSTSUBSCRIPT counts the number of walks of length k𝑘kitalic_k connecting the vertices u,vV𝑢𝑣𝑉u,v\in Vitalic_u , italic_v ∈ italic_V, where a walk is a sequence of (not necessarily different) consecutive vertices and edges in the network (see [8, 25, 33] for original sources). Therefore, defining matrix functions of the type f(A)=k=0ckAk𝑓𝐴superscriptsubscript𝑘0subscript𝑐𝑘superscript𝐴𝑘f\left(A\right)=\sum_{k=0}^{\infty}c_{k}A^{k}italic_f ( italic_A ) = ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_A start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT allows to quantify the “importance”, or centrality, of a vertex uV𝑢𝑉u\in Vitalic_u ∈ italic_V, by taking (f(A))uusubscript𝑓𝐴𝑢𝑢\left(f\left(A\right)\right)_{uu}( italic_f ( italic_A ) ) start_POSTSUBSCRIPT italic_u italic_u end_POSTSUBSCRIPT as a counting of all self-returning walks starting at vertex u𝑢uitalic_u, and giving more weight to the smaller than to the longer ones through the constants {ck}subscript𝑐𝑘\{c_{k}\}{ italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT } [22]. Similarly, the term (f(A))uvsubscript𝑓𝐴𝑢𝑣\left(f\left(A\right)\right)_{uv}( italic_f ( italic_A ) ) start_POSTSUBSCRIPT italic_u italic_v end_POSTSUBSCRIPT accounts for the “communicability” capacity between the vertices [18]. Building on the field of Euclidean matrix theory [6, 28, 35], circum-Euclidean distances [1, 47], a.k.a, spherical Euclidean distance, between pairs of vertices can also be obtained by defining (f(A))uu+(f(A))vv(f(A))uvsubscript𝑓𝐴𝑢𝑢subscript𝑓𝐴𝑣𝑣subscript𝑓𝐴𝑢𝑣\left(f\left(A\right)\right)_{uu}+\left(f\left(A\right)\right)_{vv}-\left(f% \left(A\right)\right)_{uv}( italic_f ( italic_A ) ) start_POSTSUBSCRIPT italic_u italic_u end_POSTSUBSCRIPT + ( italic_f ( italic_A ) ) start_POSTSUBSCRIPT italic_v italic_v end_POSTSUBSCRIPT - ( italic_f ( italic_A ) ) start_POSTSUBSCRIPT italic_u italic_v end_POSTSUBSCRIPT for positive-definite matrix functions f(A)𝑓𝐴f\left(A\right)italic_f ( italic_A ) [12] and angles (f(A))uv/(f(A))uu(f(A))vvsubscript𝑓𝐴𝑢𝑣subscript𝑓𝐴𝑢𝑢subscript𝑓𝐴𝑣𝑣\left(f\left(A\right)\right)_{uv}/\sqrt{\left(f\left(A\right)\right)_{uu}\left% (f\left(A\right)\right)_{vv}}( italic_f ( italic_A ) ) start_POSTSUBSCRIPT italic_u italic_v end_POSTSUBSCRIPT / square-root start_ARG ( italic_f ( italic_A ) ) start_POSTSUBSCRIPT italic_u italic_u end_POSTSUBSCRIPT ( italic_f ( italic_A ) ) start_POSTSUBSCRIPT italic_v italic_v end_POSTSUBSCRIPT end_ARG [19] (see also [23]).

The historical background for the use of matrix functions to study networks can be traced back to the work of Katz [33] who proposed (IεA)1superscript𝐼𝜀𝐴1\left(I-\varepsilon A\right)^{-1}( italic_I - italic_ε italic_A ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, with 0<ε<(λ1(A))10𝜀superscriptsubscript𝜆1𝐴10<\varepsilon<\left(\lambda_{1}\left(A\right)\right)^{-1}0 < italic_ε < ( italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_A ) ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT where λ1(A)subscript𝜆1𝐴\lambda_{1}\left(A\right)italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_A ) is the spectral radius of A𝐴Aitalic_A, to define a vertex centrality index, nowadays known as Katz centrality. However, the resolvent of the adjacency matrix (IεA)1superscript𝐼𝜀𝐴1\left(I-\varepsilon A\right)^{-1}( italic_I - italic_ε italic_A ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT is parametric, where the parameter ε𝜀\varepsilonitalic_ε is upper-bounded by the reciprocal of λ1(A)subscript𝜆1𝐴\lambda_{1}\left(A\right)italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_A ). Then when λ1(A)subscript𝜆1𝐴\lambda_{1}\left(A\right)italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_A ) is significantly large, most of the information of the network structure stored in the A𝐴Aitalic_A matrix is making almost no contribution. This has been recently shown in examples where λ1(A)1much-greater-thansubscript𝜆1𝐴1\lambda_{1}\left(A\right)\gg 1italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_A ) ≫ 1, and the resolvent of A𝐴Aitalic_A does not provide reasonable results [15]. Hence, the definitions of subgraph centrality (eA)uusubscriptsuperscript𝑒𝐴𝑢𝑢\left(e^{A}\right)_{uu}( italic_e start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_u italic_u end_POSTSUBSCRIPT [22] and communicability (eA)uvsubscriptsuperscript𝑒𝐴𝑢𝑣\left(e^{A}\right)_{uv}( italic_e start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_u italic_v end_POSTSUBSCRIPT [18] have triggered much recent interest. Another advance of the use of matrix exponential is its interpretation and derivation in different contexts, ranging from coupled quantum harmonic oscillators [20] and compartmental epidemiological models [36], to nonconservative diffusion [11]. Last but not least, we can think of eA=k=0(k!)1Aksuperscript𝑒𝐴superscriptsubscript𝑘0superscript𝑘1superscript𝐴𝑘e^{A}=\sum_{k=0}^{\infty}\left(k!\right)^{-1}A^{k}italic_e start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( italic_k ! ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_A start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT by replacing the factorial for its more general definition based on Euler Gamma functions: Eα,β(A)=k=0(Γ(αk+β))1Aksubscript𝐸𝛼𝛽𝐴superscriptsubscript𝑘0superscriptΓ𝛼𝑘𝛽1superscript𝐴𝑘E_{\alpha,\beta}\left(A\right)=\sum_{k=0}^{\infty}\left(\varGamma\left(\alpha k% +\beta\right)\right)^{-1}A^{k}italic_E start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ( italic_A ) = ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ( roman_Γ ( italic_α italic_k + italic_β ) ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_A start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT, α,β>0𝛼𝛽0\alpha,\beta>0italic_α , italic_β > 0. It retrieves the exponential when α=1𝛼1\alpha=1italic_α = 1 and β=1𝛽1\beta=1italic_β = 1, but in general represents the Mittag-Leffler matrix functions of A𝐴Aitalic_A. The idea of using Eα,β(A)subscript𝐸𝛼𝛽𝐴E_{\alpha,\beta}\left(A\right)italic_E start_POSTSUBSCRIPT italic_α , italic_β end_POSTSUBSCRIPT ( italic_A ) to define centrality and communicability indices was previously developed independently by Arrigo and Durastante [4] and by Estrada [14].

In this work, we take a step forward to using Mittag-Leffler matrix functions to quantify the degree of balance of signed graphs. A signed graph Gssubscript𝐺𝑠G_{s}italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT can have both positive and negative edges [51]. The signs of the edges emerge in various real-world scenarios. For instance, positive signs may represent friendship, collaboration, alliances, etc., while negative ones may represent enemity, hostility, conflicts, etc. in social networks [2, 32]. In voting systems, they may represent whether two voters support the same or different candidates ; in recommendation systems, they can correspond to whether two users recommend the same product, or they have discrepant opinions about the same product. In transcriptional networks, edges represent the action of a transcription factor on one of its target genes, and the sign means activation (+++) or inhibition (--) [46]. Cooperation and competition between species in ecological networks [43] and between products in economic networks [48, 50] can also be assigned to positive and negative edges, respectively.

The important notion of balance can be defined through the sign of cycles, which is the product of the signs of its edges [9, 29]. Specifically, a graph is balanced if and only if all its cycles are positive; otherwise it is unbalanced. If we focus on a signed triangle, it is balanced if either (i) all its edges are positive or (ii) two edges are positive and one is negative. The stability of the first triangle is self-evident, while in the second, the structure indicating that the “enemy of my enemy is my friend” provokes our feeling of stability by the formation of a coalition against the common enemy. The all-negative triangle is clearly unbalanced, the same as the one with only one negative edge. In the latter case, there are clear tensions between the two vertices sharing the negative edge and the one with whom they share positive ones. Think about the tensions in a cycle of friends apart from one couple in conflict. We would expect that the tensions existing between the members decay as its length increases. Hence, Estrada and Benzi has proposed the index K(Gs)=tr[eA]/tr[e|A|]𝐾subscript𝐺𝑠𝑡𝑟delimited-[]superscript𝑒𝐴𝑡𝑟delimited-[]superscript𝑒𝐴K\left(G_{s}\right)=tr\left[e^{A}\right]/tr\left[e^{\left|A\right|}\right]italic_K ( italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) = italic_t italic_r [ italic_e start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ] / italic_t italic_r [ italic_e start_POSTSUPERSCRIPT | italic_A | end_POSTSUPERSCRIPT ] where |A|𝐴\left|A\right|| italic_A | is the entrywise absolute value of A𝐴Aitalic_A, to quantify the degree of balance [16]. In this way, K(C3)0.686𝐾superscriptsubscript𝐶30.686K\left(C_{3}^{-}\right)\approx 0.686italic_K ( italic_C start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ≈ 0.686, K(C4)0.915𝐾superscriptsubscript𝐶40.915K\left(C_{4}^{-}\right)\approx 0.915italic_K ( italic_C start_POSTSUBSCRIPT 4 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ≈ 0.915, and K(C5)0.983𝐾superscriptsubscript𝐶50.983K\left(C_{5}^{-}\right)\approx 0.983italic_K ( italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ≈ 0.983, where Cksuperscriptsubscript𝐶𝑘C_{k}^{-}italic_C start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT denotes the cycle of length k𝑘kitalic_k and one negative edge. Further, K(C10)0.99999947𝐾superscriptsubscript𝐶100.99999947K\left(C_{10}^{-}\right)\approx 0.99999947italic_K ( italic_C start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ≈ 0.99999947, which is very close to balance (where K=1𝐾1K=1italic_K = 1). Is it not the case that the factorial penalization used in the exponential is too heavy and fool us in this case? Here, by completing a close walk of length 10101010, the information contained by the negative edge present in this cycle is almost completely forgotten.

In this paper, we start by showing that the balance index K(Gs)𝐾subscript𝐺𝑠K\left(G_{s}\right)italic_K ( italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) can be obtained from first principles on the basis of a nonconservative (NC) diffusion dynamic taking place on the graph Gssubscript𝐺𝑠G_{s}italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT relative to its underlying unsigned graph. Using this approach, we generalize the NC diffusion on graphs to a temporal-fractional model using Caputo fractional derivative. In this way, we generalize the balance index K(Gs)𝐾subscript𝐺𝑠K(G_{s})italic_K ( italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) to indices based on Mittag-Leffler (ML) matrix functions of A𝐴Aitalic_A. These new indices are derived from first-principles diffusion processes which are temporally non-local. Therefore, the ML balance index accounts for certain memory of the system about the past, by diminishing the penalization that long cycles typically receive. We illustrate our results with the use of some artificial signed graphs, as well as real-world networks representing gene transcription networks, ecological competition between plant species in vast regions of Spain, and social networks in rural villages in Honduras.

2 Preliminaries

Let us consider an undirected connected signed graph Gs=(V,E,ρ)subscript𝐺𝑠𝑉𝐸𝜌G_{s}=\left(V,E,\rho\right)italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = ( italic_V , italic_E , italic_ρ ) where V={1,2,,n}𝑉12𝑛V=\{1,2,\dots,n\}italic_V = { 1 , 2 , … , italic_n } is the vertex set, an edge (i,j)E𝑖𝑗𝐸(i,j)\in E( italic_i , italic_j ) ∈ italic_E is an unordered pair of two distinct nodes in the set V𝑉Vitalic_V, and ρ:EΣ:𝜌𝐸Σ\rho:E\to\varSigmaitalic_ρ : italic_E → roman_Σ, Σ={±1}Σplus-or-minus1\varSigma=\left\{\pm 1\right\}roman_Σ = { ± 1 }, associates each edge with a sign. Let A(G)=An×n𝐴𝐺𝐴superscript𝑛𝑛A\left(G\right)=A\in\mathbb{R}^{n\times n}italic_A ( italic_G ) = italic_A ∈ blackboard_R start_POSTSUPERSCRIPT italic_n × italic_n end_POSTSUPERSCRIPT be the adjacency matrix of G𝐺Gitalic_G. Specifically, if there is no edge between nodes i,j𝑖𝑗i,jitalic_i , italic_j, Aij=0subscript𝐴𝑖𝑗0A_{ij}=0italic_A start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = 0; otherwise, Aij=ρ((i,j))subscript𝐴𝑖𝑗𝜌𝑖𝑗A_{ij}=\rho((i,j))italic_A start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = italic_ρ ( ( italic_i , italic_j ) ) denotes the edge sign. We will also consider the graph where we ignore the edge sign G~~𝐺\tilde{G}over~ start_ARG italic_G end_ARG, and the unsigned adjacency matrix |A|𝐴\left|A\right|| italic_A | where the absolute values are taken entrywise.

2.1 Structural balance

A fundamental notion in the study of signed networks is the so-called structural balance [9, 30]. A signed graph is structurally balanced if and only if there is no cycle with an odd number of negative edges, which can be effectively defined through the following theorem.

Theorem 2.1 (structure theorem for balance [29]).

A signed graph G𝐺Gitalic_G is structurally balanced if and only if there is a bipartition of the node set into V=V1V2𝑉subscript𝑉1subscript𝑉2V=V_{1}\cup V_{2}italic_V = italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∪ italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT with V1subscript𝑉1V_{1}italic_V start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and V2subscript𝑉2V_{2}italic_V start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT being mutually disjoint and one of them being nonempty, s.t. any edge between the two is negative while any edge within each node subset is positive.

There are several indices proposed to quantify the degree of balance, e.g., [13, 16, 24, 27, 29, 34, 45, 49]. One of the first measures based on the walk lengths was proposed by Estrada and Benzi [16],

K(G)Tr(eA)Tr(e|A|)=j=1neλjj=1neμj,𝐾𝐺𝑇𝑟superscript𝑒𝐴𝑇𝑟superscript𝑒𝐴superscriptsubscript𝑗1𝑛superscript𝑒subscript𝜆𝑗superscriptsubscript𝑗1𝑛superscript𝑒subscript𝜇𝑗K\left(G\right)\coloneqq\dfrac{Tr\left(e^{A}\right)}{Tr\left(e^{\left|A\right|% }\right)}=\dfrac{\sum_{j=1}^{n}e^{\lambda_{j}}}{\sum_{j=1}^{n}e^{\mu_{j}}},italic_K ( italic_G ) ≔ divide start_ARG italic_T italic_r ( italic_e start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ) end_ARG start_ARG italic_T italic_r ( italic_e start_POSTSUPERSCRIPT | italic_A | end_POSTSUPERSCRIPT ) end_ARG = divide start_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG , (2.1)

where λjsubscript𝜆𝑗\lambda_{j}italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT and μjsubscript𝜇𝑗\mu_{j}italic_μ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT denote the eigenvalues of A𝐴Aitalic_A and |A|𝐴\left|A\right|| italic_A |, respectively.

We now introduce switching equivalence, which generalizes the idea of balance.

Definition 2.2.

The operation of reversing the signs of all edges connecting a subset SV𝑆𝑉S\subseteq Vitalic_S ⊆ italic_V and its complement is called switching the subset S𝑆Sitalic_S. Two signed configurations ρ,ρ:EΣ:𝜌superscript𝜌𝐸Σ\rho,\rho^{\prime}:E\to\varSigmaitalic_ρ , italic_ρ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT : italic_E → roman_Σ are said to be switching equivalent if there exists SV𝑆𝑉S\subseteq Vitalic_S ⊆ italic_V such that ρsuperscript𝜌\rho^{\prime}italic_ρ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT can be obtained from ρ𝜌\rhoitalic_ρ by switching the subset S𝑆Sitalic_S, denoted by ρρ𝜌superscript𝜌\rho\approx\rho^{\prime}italic_ρ ≈ italic_ρ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT.

Switching equivalence is an equivalence relation on sign configurations of a fixed underlying graph, and the corresponding equivalent classes are called switching classes. Clearly, balanced graphs comprise one switching class. It is also known that the spectra of signed graphs are switching invariant [5, 51].

2.2 Laplacians and Mittag-Leffler matrix function

We consider the signed Laplacian LAsubscript𝐿𝐴L_{A}italic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT as

LA(i,j)={(i,j)E|Aij|i=jAijij.subscript𝐿𝐴𝑖𝑗casessubscript𝑖𝑗𝐸subscript𝐴𝑖𝑗𝑖𝑗subscript𝐴𝑖𝑗𝑖𝑗L_{A}\left(i,j\right)=\left\{\begin{array}[]{cc}\sum_{\left(i,j\right)\in E}% \left|A_{ij}\right|&i=j\\ -A_{ij}&i\neq j.\end{array}\right.italic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_i , italic_j ) = { start_ARRAY start_ROW start_CELL ∑ start_POSTSUBSCRIPT ( italic_i , italic_j ) ∈ italic_E end_POSTSUBSCRIPT | italic_A start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT | end_CELL start_CELL italic_i = italic_j end_CELL end_ROW start_ROW start_CELL - italic_A start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT end_CELL start_CELL italic_i ≠ italic_j . end_CELL end_ROW end_ARRAY (2.2)

It governs the diffusion dynamics by Altafini’s consensus model [3] that we will talk about in more detail later. We also introduce the Lerman-Ghosh Laplacian [26, 37],

Lχ=χIA,subscript𝐿𝜒𝜒𝐼𝐴\displaystyle L_{\chi}=\chi I-A,italic_L start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT = italic_χ italic_I - italic_A , (2.3)

where χ+𝜒superscript\chi\in\mathbb{R}^{+}italic_χ ∈ blackboard_R start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT, and I𝐼Iitalic_I is the identity matrix. The index we will propose in this paper is closely related to the dynamics governed by this Laplacian, and we will show that it shares an important property with the dynamics by the signed Laplacian. Specifically, we will apply the time-fractional Caputo derivative,

Dtαu(t)=1Γ(1α)0tu(τ)(tτ)α𝑑τ,superscriptsubscript𝐷𝑡𝛼𝑢𝑡1Γ1𝛼superscriptsubscript0𝑡superscript𝑢𝜏superscript𝑡𝜏𝛼differential-d𝜏\displaystyle D_{t}^{\alpha}u(t)=\frac{1}{\Gamma(1-\alpha)}\int_{0}^{t}\frac{u% ^{\prime}(\tau)}{(t-\tau)^{\alpha}}d\tau,italic_D start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_u ( italic_t ) = divide start_ARG 1 end_ARG start_ARG roman_Γ ( 1 - italic_α ) end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT divide start_ARG italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_τ ) end_ARG start_ARG ( italic_t - italic_τ ) start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT end_ARG italic_d italic_τ , (2.4)

where u(τ)superscript𝑢𝜏u^{\prime}(\tau)italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_τ ) denotes the usual derivative. We assume that u𝑢uitalic_u is differentiable and the convolution can be defined. Here, 0<α10𝛼10<\alpha\leq 10 < italic_α ≤ 1, 0<t<0𝑡0<t<\infty0 < italic_t < ∞, and Γ(x)Γ𝑥\varGamma\left(x\right)roman_Γ ( italic_x ) is the Euler gamma function. We also recall that a diffusion process is said to be conservative if the number of diffusive particles is constant along the time; otherwise, it is called a nonconservative diffusion [11]. Finally, we introduce the building block of the balance index we will propose, the Mittag-Leffler (ML) function of a matrix, say M𝑀Mitalic_M,

Eα(M)k=0MkΓ(αk+1).subscript𝐸𝛼𝑀superscriptsubscript𝑘0superscript𝑀𝑘Γ𝛼𝑘1E_{\alpha}\left(M\right)\coloneqq\sum_{k=0}^{\infty}\dfrac{M^{k}}{\varGamma% \left(\alpha k+1\right)}.italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_M ) ≔ ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_M start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( italic_α italic_k + 1 ) end_ARG . (2.5)

The study of these matrix functions for networks was previously studied by Arrigo and Durastante [4], and they also proposed to use EαγEα(γM)superscriptsubscript𝐸𝛼𝛾subscript𝐸𝛼𝛾𝑀E_{\alpha}^{\gamma}\coloneqq E_{\alpha}\left(\gamma M\right)italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ≔ italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_γ italic_M ) with γΓ(α+1)𝛾Γ𝛼1\gamma\leq\Gamma(\alpha+1)italic_γ ≤ roman_Γ ( italic_α + 1 ) accounting for fair contribution of walks in graphs. In our implementations, we adopt this suggestion, with γ=Γ(α+1)𝛾Γ𝛼1\gamma=\Gamma(\alpha+1)italic_γ = roman_Γ ( italic_α + 1 ).

3 Motivation

There are 215superscript2152^{15}2 start_POSTSUPERSCRIPT 15 end_POSTSUPERSCRIPT ways to put signs on the edges of the Petersen graph, on which only five (excluding the unsigned one) are essentially different [52] (see Fig. 1). We consider the diffusion dynamics by Altafini’s consensus model [3]. Let u(t)𝑢𝑡u\left(t\right)italic_u ( italic_t ) be the vector representing the state of the vertices in Gssubscript𝐺𝑠G_{s}italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT at time t𝑡titalic_t, with u(0)=u0𝑢0superscript𝑢0u\left(0\right)=u^{0}italic_u ( 0 ) = italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT, and let u˙(t)˙𝑢𝑡\dot{u}\left(t\right)over˙ start_ARG italic_u end_ARG ( italic_t ) be the vector of their time derivatives. Then,

u˙i(t)=(i,j)E|Aij|(ui(t)sgn(Aij)uj(t)),subscript˙𝑢𝑖𝑡subscript𝑖𝑗𝐸subscript𝐴𝑖𝑗subscript𝑢𝑖𝑡sgnsubscript𝐴𝑖𝑗subscript𝑢𝑗𝑡\dot{u}_{i}\left(t\right)=-\sum_{\left(i,j\right)\in E}\left|A_{ij}\right|% \left(u_{i}\left(t\right)-\textnormal{sgn}\left(A_{ij}\right)u_{j}\left(t% \right)\right),over˙ start_ARG italic_u end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) = - ∑ start_POSTSUBSCRIPT ( italic_i , italic_j ) ∈ italic_E end_POSTSUBSCRIPT | italic_A start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT | ( italic_u start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) - sgn ( italic_A start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT ) italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) ) , (3.1)

where sgn()sgn\text{sgn}(\cdot)sgn ( ⋅ ) returns the sign of the value. Hence,

u˙(t)=LAu(t);u(0)=u0.formulae-sequence˙𝑢𝑡subscript𝐿𝐴𝑢𝑡𝑢0superscript𝑢0\dot{u}\left(t\right)=-L_{A}u\left(t\right);u\left(0\right)=u^{0}.over˙ start_ARG italic_u end_ARG ( italic_t ) = - italic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT italic_u ( italic_t ) ; italic_u ( 0 ) = italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT . (3.2)
Refer to caption
Figure 1: The five switching isomorphism types of signed Petersen graph excluding the unsigned one, where solid lines represent positive edges and dashed lines represent negative ones.

We consider the convergence time, tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, at which the state values are sufficiently close to each other, with tolerance 105superscript10510^{-5}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT, i.e., |uv(tc)uu(tc)|<105subscript𝑢𝑣subscript𝑡𝑐subscript𝑢𝑢subscript𝑡𝑐superscript105\left|u_{v}\left(t_{c}\right)-u_{u}\left(t_{c}\right)\right|<10^{-5}| italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) - italic_u start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ( italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) | < 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT, u,vGsfor-all𝑢𝑣subscript𝐺𝑠\forall u,v\in G_{s}∀ italic_u , italic_v ∈ italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. We note that the only difference of graphs in Fig. 1 lies in their sign patterns, and we denote a negative cycle of length k𝑘kitalic_k by Cksuperscriptsubscript𝐶𝑘C_{k}^{-}italic_C start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT. We find that the graph having the most C5superscriptsubscript𝐶5C_{5}^{-}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, graph e), is the graph reaching the consensus in a fastest time, with 12 C5superscriptsubscript𝐶5C_{5}^{-}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT and tc=11subscript𝑡𝑐11t_{c}=11italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 11. The graph having the least C5superscriptsubscript𝐶5C_{5}^{-}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, graph a), is the one that delays the most, with 4 C5superscriptsubscript𝐶5C_{5}^{-}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT and tc=48subscript𝑡𝑐48t_{c}=48italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 48. It is known that consensus will never be reached if a graph is balanced, but instead the dynamics reaches a dissensus state. Therefore, graph a) is more similar to a balanced graph in the sense that it delays more to reach the consensus than graph e). However, this simple arithmetic is broken when we consider that graphs b) and d), both with 6 C5superscriptsubscript𝐶5C_{5}^{-}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, but with the first almost doubling the time for consensus of the second (24242424 versus 14141414). We can then extend the analysis to consider C6superscriptsubscript𝐶6C_{6}^{-}italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, which clearly indicates that graph d) is less similar to a balanced graph than graph b), with 10101010 versus 6666 C6superscriptsubscript𝐶6C_{6}^{-}italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, respectively. Under this kind of semiquantitative analysis, a problem emerges when considering graph c) with tc=22subscript𝑡𝑐22t_{c}=22italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 22, which has 8888 C5superscriptsubscript𝐶5C_{5}^{-}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, more than that of graphs b) and d), but 4 C6superscriptsubscript𝐶6C_{6}^{-}italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT less than that of the previous two graphs. We defer more details to Supplementary Material.

Since the Petersen graphs are cubic, LA(i,i)=3subscript𝐿𝐴𝑖𝑖3L_{A}\left(i,i\right)=3italic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_i , italic_i ) = 3, iVfor-all𝑖𝑉\forall i\in V∀ italic_i ∈ italic_V, which allows us to write LA=Lχ=χIAsubscript𝐿𝐴subscript𝐿𝜒𝜒𝐼𝐴L_{A}=L_{\chi}=\chi I-Aitalic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT = italic_χ italic_I - italic_A, where Lχsubscript𝐿𝜒L_{\chi}italic_L start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT is the Lerman-Ghosh Laplacian (2.3) as introduced in section 2, and χ=3𝜒3\chi=3italic_χ = 3 here. The solution to the Cauchy problem (3.1) is

u(t)=etLχu0=etχetAu0.𝑢𝑡superscript𝑒𝑡subscript𝐿𝜒superscript𝑢0superscript𝑒𝑡𝜒superscript𝑒𝑡𝐴superscript𝑢0u\left(t\right)=e^{-tL_{\chi}}u^{0}=e^{-t\chi}e^{tA}u^{0}.italic_u ( italic_t ) = italic_e start_POSTSUPERSCRIPT - italic_t italic_L start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = italic_e start_POSTSUPERSCRIPT - italic_t italic_χ end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_t italic_A end_POSTSUPERSCRIPT italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT . (3.3)

Then, at a given time t𝑡titalic_t, the concentration at a vertex v𝑣vitalic_v is

uv(t)=j(etχetA)v,juj0.subscript𝑢𝑣𝑡subscript𝑗subscriptsuperscript𝑒𝑡𝜒superscript𝑒𝑡𝐴𝑣𝑗superscriptsubscript𝑢𝑗0u_{v}\left(t\right)=\sum_{j}\left(e^{-t\chi}e^{tA}\right)_{v,j}u_{j}^{0}.italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_e start_POSTSUPERSCRIPT - italic_t italic_χ end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_t italic_A end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_v , italic_j end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT . (3.4)

Suppose that the initial concentration is totally located at the vertex v𝑣vitalic_v, uj0=δj,vsuperscriptsubscript𝑢𝑗0subscript𝛿𝑗𝑣u_{j}^{0}=\delta_{j,v}italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = italic_δ start_POSTSUBSCRIPT italic_j , italic_v end_POSTSUBSCRIPT , where δi,jsubscript𝛿𝑖𝑗\delta_{i,j}italic_δ start_POSTSUBSCRIPT italic_i , italic_j end_POSTSUBSCRIPT is the Kronecker delta, then

uv(t)=etχ(etA)vv.subscript𝑢𝑣𝑡superscript𝑒𝑡𝜒subscriptsuperscript𝑒𝑡𝐴𝑣𝑣u_{v}\left(t\right)=e^{-t\chi}\left(e^{tA}\right)_{vv}.italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) = italic_e start_POSTSUPERSCRIPT - italic_t italic_χ end_POSTSUPERSCRIPT ( italic_e start_POSTSUPERSCRIPT italic_t italic_A end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_v italic_v end_POSTSUBSCRIPT . (3.5)

Then the total concentration remaining at the vertices when the initial concentration has been totally allocated at them is

Tsvuv(t)=etχv(etA)vv=etχTr(etA),subscript𝑇𝑠subscript𝑣subscript𝑢𝑣𝑡superscript𝑒𝑡𝜒subscript𝑣subscriptsuperscript𝑒𝑡𝐴𝑣𝑣superscript𝑒𝑡𝜒𝑇𝑟superscript𝑒𝑡𝐴T_{s}\coloneqq\sum_{v}u_{v}\left(t\right)=e^{-t\chi}\sum_{v}\left(e^{tA}\right% )_{vv}=e^{-t\chi}Tr\left(e^{tA}\right),italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≔ ∑ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) = italic_e start_POSTSUPERSCRIPT - italic_t italic_χ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_e start_POSTSUPERSCRIPT italic_t italic_A end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_v italic_v end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT - italic_t italic_χ end_POSTSUPERSCRIPT italic_T italic_r ( italic_e start_POSTSUPERSCRIPT italic_t italic_A end_POSTSUPERSCRIPT ) , (3.6)

where Tr()𝑇𝑟Tr\left(\cdot\right)italic_T italic_r ( ⋅ ) returns the trace of a matrix. In a similar way, we can ignore the edge sign and consider the underlying graph G~~𝐺\tilde{G}over~ start_ARG italic_G end_ARG,

Tuvuv(t)=etχv(et|A|)vv=etχTr(et|A|).subscript𝑇𝑢subscript𝑣subscript𝑢𝑣𝑡superscript𝑒𝑡𝜒subscript𝑣subscriptsuperscript𝑒𝑡𝐴𝑣𝑣superscript𝑒𝑡𝜒𝑇𝑟superscript𝑒𝑡𝐴T_{u}\coloneqq\sum_{v}u_{v}\left(t\right)=e^{-t\chi}\sum_{v}\left(e^{t\left|A% \right|}\right)_{vv}=e^{-t\chi}Tr\left(e^{t\left|A\right|}\right).italic_T start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ≔ ∑ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) = italic_e start_POSTSUPERSCRIPT - italic_t italic_χ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_e start_POSTSUPERSCRIPT italic_t | italic_A | end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_v italic_v end_POSTSUBSCRIPT = italic_e start_POSTSUPERSCRIPT - italic_t italic_χ end_POSTSUPERSCRIPT italic_T italic_r ( italic_e start_POSTSUPERSCRIPT italic_t | italic_A | end_POSTSUPERSCRIPT ) . (3.7)

A way to account for the “influence” of the edge signs on the diffusion is Ts/Tusubscript𝑇𝑠subscript𝑇𝑢T_{s}/T_{u}italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_T start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT, such that for t=1𝑡1t=1italic_t = 1 we recover the measure K(G)𝐾𝐺K(G)italic_K ( italic_G ) in (2.1). This balance index can be easily generalized by taking any value of t=β𝑡𝛽t=\betaitalic_t = italic_β, K(G,β)=Tr(exp(βA))/Tr(exp(β|A|))𝐾𝐺𝛽𝑇𝑟𝛽𝐴𝑇𝑟𝛽𝐴K(G,\beta)=Tr(\exp(\beta A))/Tr(\exp(\beta\left|A\right|))italic_K ( italic_G , italic_β ) = italic_T italic_r ( roman_exp ( italic_β italic_A ) ) / italic_T italic_r ( roman_exp ( italic_β | italic_A | ) ).

For the five nonsimilar signed Petersen graphs, although there is a good correlation between K(G)𝐾𝐺K\left(G\right)italic_K ( italic_G ) and tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (Pearson correlation: r20.924superscript𝑟20.924r^{2}\approx 0.924italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≈ 0.924), there is an important inversion in the values of K(G)𝐾𝐺K\left(G\right)italic_K ( italic_G ) for graphs c) and d). Specifically, for K(G)𝐾𝐺K(G)italic_K ( italic_G ), graph c) has value 0.9410.9410.9410.941 while graph d) has value 0.9470.9470.9470.947, but graph c) has larger tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT than graph d); see Supplementary Materials for details. The problem seems to be produced by the differences in the penalization that the cycles of length 5 and 6 receive in the exponential function. To see this, we examine the difference between Tr(Ak)/k!𝑇𝑟superscript𝐴𝑘𝑘Tr\left(A^{k}\right)/k!italic_T italic_r ( italic_A start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) / italic_k ! for graph c) and graph d) for values of 1k101𝑘101\leq k\leq 101 ≤ italic_k ≤ 10; see Fig. 2. We find that the largest contribution is the one of k=5,𝑘5k=5,italic_k = 5 , which is about 0.3330.333-0.333- 0.333, followed by that of k=6,𝑘6k=6,italic_k = 6 , which is 0.20.20.20.2. This reflects the fact that graph c) has more negative cycles of length 5 than d), that d) has more negative cycles of length 6 than c), but that cycles of length 6 are much heavily penalized than those of length 5. We can put it in the following way. If one has to pay $1 for every C5superscriptsubscript𝐶5C_{5}^{-}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT but only $0.1 for each C6superscriptsubscript𝐶6C_{6}^{-}italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, graph c) will have to pay $8.40, while only $7.00 is needed for graph d). But if the penalty for C6superscriptsubscript𝐶6C_{6}^{-}italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT increases to $0.5, then graph c) will need to pay $10 while $11 will be paid by graph d). Therefore, the problem we raise in this paper is how to tune the penalization of longer cycles, such that their contribution to the balance/unbalance of networks becomes more relevant when necessary. We propose to achieve it while keeping the first principles explained before that connect the balance index K(G)𝐾𝐺K\left(G\right)italic_K ( italic_G ) with a (nonconservative) diffusion on graphs.

Refer to caption
Figure 2: Change of the difference of Tr(Ak)/k!𝑇𝑟superscript𝐴𝑘𝑘Tr\left(A^{k}\right)/k!italic_T italic_r ( italic_A start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) / italic_k ! between graphs c) and d) of Fig. 1.

We end up this section by proving that Altafini’s model of consensus on signed graphs is nonconeservative, unless the graph is balanced and the initial vector is the eigenvector corresponding to the eigenvalue 00. We should also notice that when the graph does not contain any negative edge, Altafini’s consensus model is effectively the consensus model with the graph Laplacian and it is conservative.

Proposition 3.1.

The diffusion by Altafini’s consensus model is nonconservative, unless the graph is balanced and the intial vector u0superscript𝑢0u^{0}italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT is the eigenvector corresponding to the smallest eigenvalue 00.

Proof.

The solution to the Altafini’s consensus is

u(t)=etLAu0.𝑢𝑡superscript𝑒𝑡subscript𝐿𝐴superscript𝑢0\displaystyle u(t)=e^{-tL_{A}}u^{0}.italic_u ( italic_t ) = italic_e start_POSTSUPERSCRIPT - italic_t italic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT . (3.8)

Let 0μ1μ2μn0subscript𝜇1subscript𝜇2subscript𝜇𝑛0\leq\mu_{1}\leq\mu_{2}\leq\dots\leq\mu_{n}0 ≤ italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≤ italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≤ ⋯ ≤ italic_μ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT be the eigenvalues of LAsubscript𝐿𝐴L_{A}italic_L start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT, and let ϕisubscriptitalic-ϕ𝑖\phi_{i}italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT be the orthonormal eigenvector associated with μisubscript𝜇𝑖\mu_{i}italic_μ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. Then,

u(t)=etμ1ϕ1ϕ1Tu0+etμ2ϕ2ϕ2Tu0++etμnϕnϕnTu0.𝑢𝑡superscript𝑒𝑡subscript𝜇1subscriptitalic-ϕ1superscriptsubscriptitalic-ϕ1𝑇superscript𝑢0superscript𝑒𝑡subscript𝜇2subscriptitalic-ϕ2superscriptsubscriptitalic-ϕ2𝑇superscript𝑢0superscript𝑒𝑡subscript𝜇𝑛subscriptitalic-ϕ𝑛superscriptsubscriptitalic-ϕ𝑛𝑇superscript𝑢0\displaystyle u(t)=e^{-t\mu_{1}}\phi_{1}\phi_{1}^{T}u^{0}+e^{-t\mu_{2}}\phi_{2% }\phi_{2}^{T}u^{0}+\dots+e^{-t\mu_{n}}\phi_{n}\phi_{n}^{T}u^{0}.italic_u ( italic_t ) = italic_e start_POSTSUPERSCRIPT - italic_t italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT + italic_e start_POSTSUPERSCRIPT - italic_t italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT + ⋯ + italic_e start_POSTSUPERSCRIPT - italic_t italic_μ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT . (3.9)

We know that if a signed graph is unbalanced, μ1>0subscript𝜇10\mu_{1}>0italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > 0. Then,

limtu(t)=𝟎,subscript𝑡𝑢𝑡0\displaystyle\lim_{t\to\infty}u(t)=\mathbf{0},roman_lim start_POSTSUBSCRIPT italic_t → ∞ end_POSTSUBSCRIPT italic_u ( italic_t ) = bold_0 ,

where 𝟎0\mathbf{0}bold_0 is the all-zero vector. Hence, the diffusion is nonconservative.

We now consider the balanced case, where μ1=0subscript𝜇10\mu_{1}=0italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 0 and μ2>0subscript𝜇20\mu_{2}>0italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT > 0. Then,

limtu(t)=ϕ1ϕ1Tu0.subscript𝑡𝑢𝑡subscriptitalic-ϕ1superscriptsubscriptitalic-ϕ1𝑇superscript𝑢0\displaystyle\lim_{t\to\infty}u(t)=\phi_{1}\phi_{1}^{T}u^{0}.roman_lim start_POSTSUBSCRIPT italic_t → ∞ end_POSTSUBSCRIPT italic_u ( italic_t ) = italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT .

Hence limt𝟏Tu(t)=𝟏Tu0subscript𝑡superscript1𝑇𝑢𝑡superscript1𝑇superscript𝑢0\lim_{t\to\infty}\mathbf{1}^{T}u(t)=\mathbf{1}^{T}u^{0}roman_lim start_POSTSUBSCRIPT italic_t → ∞ end_POSTSUBSCRIPT bold_1 start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_u ( italic_t ) = bold_1 start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT if and only if u0=ϕ1superscript𝑢0subscriptitalic-ϕ1u^{0}=\phi_{1}italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. In the case of u0=ϕ1superscript𝑢0subscriptitalic-ϕ1u^{0}=\phi_{1}italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, from Eq. (3.9), we have 𝟏Tu(t)=𝟏Tu0superscript1𝑇𝑢𝑡superscript1𝑇superscript𝑢0\mathbf{1}^{T}u(t)=\mathbf{1}^{T}u^{0}bold_1 start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_u ( italic_t ) = bold_1 start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT, by the orthogonality of eigenvectors. Hence, the diffusion is conservative if and only if the graph is balanced and u0=ϕ1superscript𝑢0subscriptitalic-ϕ1u^{0}=\phi_{1}italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = italic_ϕ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. ∎

4 Main results

4.1 Nonconservative fractional diffusion and balance

We know that Altafini’s dynamics on signed graphs is nonconservative. Let us now consider a more general nonconservative diffusive model on the signed graph based on the Lerman-Ghosh Laplacian Lχsubscript𝐿𝜒L_{\chi}italic_L start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT [26, 37]. To make the process more general, we also replace the standard time derivative u˙(t)˙𝑢𝑡\dot{u}\left(t\right)over˙ start_ARG italic_u end_ARG ( italic_t ) by the time-fractional Caputo derivative Dtαsuperscriptsubscript𝐷𝑡𝛼D_{t}^{\alpha}italic_D start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT as in Eq. (2.4). Hence, the nonconservative diffusion on the signed graph we consider is

Dtαu(t)=Lχu(t);u(0)=u0.formulae-sequencesuperscriptsubscript𝐷𝑡𝛼𝑢𝑡subscript𝐿𝜒𝑢𝑡𝑢0superscript𝑢0D_{t}^{\alpha}u\left(t\right)=-L_{\chi}u\left(t\right);u\left(0\right)=u^{0}.italic_D start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_u ( italic_t ) = - italic_L start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT italic_u ( italic_t ) ; italic_u ( 0 ) = italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT . (4.1)

The solution of Eq. (4.1) is given by

u(t)=Eα(tαLχ)u0,𝑢𝑡subscript𝐸𝛼superscript𝑡𝛼subscript𝐿𝜒superscript𝑢0\displaystyle u(t)=E_{\alpha}(-t^{\alpha}L_{\chi})u^{0},italic_u ( italic_t ) = italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( - italic_t start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_L start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ) italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT , (4.2)

where Eα()subscript𝐸𝛼E_{\alpha}(\cdot)italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( ⋅ ) is the Mittag-Leffler function as in Eq. (2.5). Let us focus again on the concentration at a vertex designated by v𝑣vitalic_v,

uv(t)=j(Eα(tαLχ))vjuj0,subscript𝑢𝑣𝑡subscript𝑗subscriptsubscript𝐸𝛼superscript𝑡𝛼subscript𝐿𝜒𝑣𝑗superscriptsubscript𝑢𝑗0u_{v}\left(t\right)=\sum_{j}\left(E_{\alpha}\left(-t^{\alpha}L_{\chi}\right)% \right)_{vj}u_{j}^{0},italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( - italic_t start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_L start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ) ) start_POSTSUBSCRIPT italic_v italic_j end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT , (4.3)

and if the initial concentration is totally located at the vertex v𝑣vitalic_v, uj0=δj,vsuperscriptsubscript𝑢𝑗0subscript𝛿𝑗𝑣u_{j}^{0}=\delta_{j,v}italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = italic_δ start_POSTSUBSCRIPT italic_j , italic_v end_POSTSUBSCRIPT, we get

uv(t)=(Eα(tαLχ))vv.subscript𝑢𝑣𝑡subscriptsubscript𝐸𝛼superscript𝑡𝛼subscript𝐿𝜒𝑣𝑣u_{v}\left(t\right)=\left(E_{\alpha}\left(-t^{\alpha}L_{\chi}\right)\right)_{% vv}.italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) = ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( - italic_t start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_L start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT ) ) start_POSTSUBSCRIPT italic_v italic_v end_POSTSUBSCRIPT . (4.4)

One main difference between the exponential and the Mittag-Leffler function is that in general Eα(P+Q)Eα(P)Eα(Q)subscript𝐸𝛼𝑃𝑄subscript𝐸𝛼𝑃subscript𝐸𝛼𝑄E_{\alpha}\left(P+Q\right)\neq E_{\alpha}\left(P\right)E_{\alpha}\left(Q\right)italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_P + italic_Q ) ≠ italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_P ) italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_Q ), even when P𝑃Pitalic_P and Q𝑄Qitalic_Q commute [42]. This equality holds in general only when (i) P𝑃Pitalic_P and Q𝑄Qitalic_Q commute and (ii) α=1𝛼1\alpha=1italic_α = 1.

Here, we consider the special case when χ=0𝜒0\chi=0italic_χ = 0, such that

Dtαu(t)=Au(t);u(0)=u0.formulae-sequencesuperscriptsubscript𝐷𝑡𝛼𝑢𝑡𝐴𝑢𝑡𝑢0superscript𝑢0\displaystyle D_{t}^{\alpha}u\left(t\right)=Au\left(t\right);u\left(0\right)=u% ^{0}.italic_D start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_u ( italic_t ) = italic_A italic_u ( italic_t ) ; italic_u ( 0 ) = italic_u start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT . (4.5)

The concentration at vertex v𝑣vitalic_v with uj0=δj,vsuperscriptsubscript𝑢𝑗0subscript𝛿𝑗𝑣u_{j}^{0}=\delta_{j,v}italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = italic_δ start_POSTSUBSCRIPT italic_j , italic_v end_POSTSUBSCRIPT is

uv(t)=(Eα(tαA))vv.subscript𝑢𝑣𝑡subscriptsubscript𝐸𝛼superscript𝑡𝛼𝐴𝑣𝑣\displaystyle u_{v}\left(t\right)=\left(E_{\alpha}\left(t^{\alpha}A\right)% \right)_{vv}.italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) = ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_A ) ) start_POSTSUBSCRIPT italic_v italic_v end_POSTSUBSCRIPT .

Then the total concentration remaining at the vertices when the initial concentration has been totally allocated at them is

T~svuv(t)=v(Eα(tαA))vv=Tr(Eα(tαA)).subscript~𝑇𝑠subscript𝑣subscript𝑢𝑣𝑡subscript𝑣subscriptsubscript𝐸𝛼superscript𝑡𝛼𝐴𝑣𝑣𝑇𝑟subscript𝐸𝛼superscript𝑡𝛼𝐴\tilde{T}_{s}\coloneqq\sum_{v}u_{v}\left(t\right)=\sum_{v}\left(E_{\alpha}% \left(t^{\alpha}A\right)\right)_{vv}=Tr\left(E_{\alpha}\left(t^{\alpha}A\right% )\right).over~ start_ARG italic_T end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≔ ∑ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_A ) ) start_POSTSUBSCRIPT italic_v italic_v end_POSTSUBSCRIPT = italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_A ) ) . (4.6)

Similarly, we can ignore the edge sign and obtain the total concentration in G~~𝐺\tilde{G}over~ start_ARG italic_G end_ARG,

T~uvuv(t)=v(Eα(tα|A|))vv=Tr(Eα(tα|A|)).subscript~𝑇𝑢subscript𝑣subscript𝑢𝑣𝑡subscript𝑣subscriptsubscript𝐸𝛼superscript𝑡𝛼𝐴𝑣𝑣𝑇𝑟subscript𝐸𝛼superscript𝑡𝛼𝐴\tilde{T}_{u}\coloneqq\sum_{v}u_{v}\left(t\right)=\sum_{v}\left(E_{\alpha}% \left(t^{\alpha}\left|A\right|\right)\right)_{vv}=Tr\left(E_{\alpha}\left(t^{% \alpha}\left|A\right|\right)\right).over~ start_ARG italic_T end_ARG start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT ≔ ∑ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) = ∑ start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT | italic_A | ) ) start_POSTSUBSCRIPT italic_v italic_v end_POSTSUBSCRIPT = italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT | italic_A | ) ) . (4.7)

Finally, we summarise the influence of the edge signs on the diffusion as the ratio T~s/T~usubscript~𝑇𝑠subscript~𝑇𝑢\tilde{T}_{s}/\tilde{T}_{u}over~ start_ARG italic_T end_ARG start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / over~ start_ARG italic_T end_ARG start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT, such that for t=1𝑡1t=1italic_t = 1 we have

Kα(Gs)Tr(Eα(A))Tr(Eα(|A|)).subscript𝐾𝛼subscript𝐺𝑠𝑇𝑟subscript𝐸𝛼𝐴𝑇𝑟subscript𝐸𝛼𝐴K_{\alpha}\left(G_{s}\right)\coloneqq\dfrac{Tr\left(E_{\alpha}\left(A\right)% \right)}{Tr\left(E_{\alpha}\left(\left|A\right|\right)\right)}.italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) ≔ divide start_ARG italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_A ) ) end_ARG start_ARG italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( | italic_A | ) ) end_ARG . (4.8)

We note that K(G)𝐾𝐺K\left(G\right)italic_K ( italic_G ) is the particular case when α=1𝛼1\alpha=1italic_α = 1.

4.2 How global balance is accounted for

A closed walk (CW) is said to be positive (negative) if the product of the signs of all its composing edges is positive (negative). Let Mk(i,i)=(Ak)iisubscript𝑀𝑘𝑖𝑖subscriptsuperscript𝐴𝑘𝑖𝑖M_{k}\left(i,i\right)=\left(A^{k}\right)_{ii}italic_M start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_i , italic_i ) = ( italic_A start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_i italic_i end_POSTSUBSCRIPT be the total “number” of CWs of length k𝑘kitalic_k starting at vertex i𝑖iitalic_i, then

Mk(i,i)=μk+(i,i)μk(i,i),subscript𝑀𝑘𝑖𝑖superscriptsubscript𝜇𝑘𝑖𝑖superscriptsubscript𝜇𝑘𝑖𝑖M_{k}\left(i,i\right)=\mu_{k}^{+}\left(i,i\right)-\mu_{k}^{-}\left(i,i\right),italic_M start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_i , italic_i ) = italic_μ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_i , italic_i ) - italic_μ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_i , italic_i ) , (4.9)

where μk+(i,i)superscriptsubscript𝜇𝑘𝑖𝑖\mu_{k}^{+}\left(i,i\right)italic_μ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_i , italic_i ) is the number of positive CWs of length k𝑘kitalic_k starting at i𝑖iitalic_i, and μk(i,i)superscriptsubscript𝜇𝑘𝑖𝑖\mu_{k}^{-}\left(i,i\right)italic_μ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_i , italic_i ) is the same for negative CWs [10]. Obviously,

Tr(Eα(A))=i=1nk=0Mk(i,i)Γ(αk+1)=Tr(Eα+(A))Tr(Eα(A)),𝑇𝑟subscript𝐸𝛼𝐴superscriptsubscript𝑖1𝑛superscriptsubscript𝑘0subscript𝑀𝑘𝑖𝑖Γ𝛼𝑘1𝑇𝑟superscriptsubscript𝐸𝛼𝐴𝑇𝑟superscriptsubscript𝐸𝛼𝐴\begin{split}Tr\left(E_{\alpha}\left(A\right)\right)=\sum_{i=1}^{n}\sum_{k=0}^% {\infty}\dfrac{M_{k}\left(i,i\right)}{\varGamma\left(\alpha k+1\right)}=Tr% \left(E_{\alpha}^{+}\left(A\right)\right)-Tr\left(E_{\alpha}^{-}\left(A\right)% \right),\end{split}start_ROW start_CELL italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_A ) ) = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG italic_M start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ( italic_i , italic_i ) end_ARG start_ARG roman_Γ ( italic_α italic_k + 1 ) end_ARG = italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_A ) ) - italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ) ) , end_CELL end_ROW (4.10)

where Tr(Eα±(A))=i=1nk=0μk±(i,i)/Γ(αk+1)𝑇𝑟superscriptsubscript𝐸𝛼plus-or-minus𝐴superscriptsubscript𝑖1𝑛superscriptsubscript𝑘0superscriptsubscript𝜇𝑘plus-or-minus𝑖𝑖Γ𝛼𝑘1Tr\left(E_{\alpha}^{\pm}\left(A\right)\right)=\sum_{i=1}^{n}\sum_{k=0}^{\infty% }\mu_{k}^{\pm}\left(i,i\right)/\varGamma\left(\alpha k+1\right)italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT ( italic_A ) ) = ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_μ start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT ( italic_i , italic_i ) / roman_Γ ( italic_α italic_k + 1 ) are the positive and negative contributions to Tr(Eα(A))𝑇𝑟subscript𝐸𝛼𝐴Tr\left(E_{\alpha}\left(A\right)\right)italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_A ) ). We note that they are not the same as Tr(Eα(A±))𝑇𝑟subscript𝐸𝛼superscript𝐴plus-or-minusTr\left(E_{\alpha}\left(A^{\pm}\right)\right)italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_A start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT ) ) where A±superscript𝐴plus-or-minusA^{\pm}italic_A start_POSTSUPERSCRIPT ± end_POSTSUPERSCRIPT are the adjacency matrices only for positive and negative edges of Gssubscript𝐺𝑠G_{s}italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, respectively. Similarly,

Tr(Eα(|A|))=Tr(Eα+(A))+Tr(Eα(A)).𝑇𝑟subscript𝐸𝛼𝐴𝑇𝑟superscriptsubscript𝐸𝛼𝐴𝑇𝑟superscriptsubscript𝐸𝛼𝐴\begin{split}Tr\left(E_{\alpha}\left(\left|A\right|\right)\right)&=Tr\left(E_{% \alpha}^{+}\left(A\right)\right)+Tr\left(E_{\alpha}^{-}\left(A\right)\right)% \end{split}.start_ROW start_CELL italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( | italic_A | ) ) end_CELL start_CELL = italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_A ) ) + italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ) ) end_CELL end_ROW . (4.11)

Hence,

Kα(Gs)=Tr(Eα+(A))Tr(Eα(A))Tr(Eα+(A))+Tr(Eα(A)).subscript𝐾𝛼subscript𝐺𝑠𝑇𝑟superscriptsubscript𝐸𝛼𝐴𝑇𝑟superscriptsubscript𝐸𝛼𝐴𝑇𝑟superscriptsubscript𝐸𝛼𝐴𝑇𝑟superscriptsubscript𝐸𝛼𝐴K_{\alpha}\left(G_{s}\right)=\dfrac{Tr\left(E_{\alpha}^{+}\left(A\right)\right% )-Tr\left(E_{\alpha}^{-}\left(A\right)\right)}{Tr\left(E_{\alpha}^{+}\left(A% \right)\right)+Tr\left(E_{\alpha}^{-}\left(A\right)\right)}.italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) = divide start_ARG italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_A ) ) - italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ) ) end_ARG start_ARG italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_A ) ) + italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ) ) end_ARG . (4.12)

We now understand Kα(Gs)subscript𝐾𝛼subscript𝐺𝑠K_{\alpha}(G_{s})italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) through its two different terms. Let us recall that a trivial CW is a walk starting at and ending at the same vertex but not involving any cycle in the graph. Hence, any trivial CW is always positive. Therefore, Tr(Eα+(A))𝑇𝑟superscriptsubscript𝐸𝛼𝐴Tr\left(E_{\alpha}^{+}\left(A\right)\right)italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_A ) ) accounts for all trivial CWs and nontrivial positive CWs. In a nontrivial positive CW, there can be any number of balanced cycles, and also even number of unbalanced cycles. We can understand it as follows. Consider a negative triangle with sign pattern +,+,+,+,-+ , + , - for edges (A,B),(B,C),(A,C)𝐴𝐵𝐵𝐶𝐴𝐶(A,B),(B,C),(A,C)( italic_A , italic_B ) , ( italic_B , italic_C ) , ( italic_A , italic_C ), respectively. Then, a voting system on this triangle will end up in contradictions after one round of information passing. For example, if A votes Y(N), then B will vote Y(N), and C will also vote Y(N), but then A will need to vote N(Y) since it is in conflict with C, contradicting its initial vote. However, if the number of rounds is even such contradictions disappear, eliminating the tension in the system. In closing, the term Tr(Eα+(A))𝑇𝑟superscriptsubscript𝐸𝛼𝐴Tr\left(E_{\alpha}^{+}\left(A\right)\right)italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_A ) ) accounts for all CWs in the signed graph that involves no tensions from the perspective of balance. This necessarily leads to the fact that all tensions are encoded in Tr(Eα(A))𝑇𝑟superscriptsubscript𝐸𝛼𝐴Tr\left(E_{\alpha}^{-}\left(A\right)\right)italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ) ). Indeed, any negative CW necessarily contains a negative cycle, which by definition is unbalanced. Therefore, the difference Tr(Eα+(A))Tr(Eα(A))𝑇𝑟superscriptsubscript𝐸𝛼𝐴𝑇𝑟superscriptsubscript𝐸𝛼𝐴Tr\left(E_{\alpha}^{+}\left(A\right)\right)-Tr\left(E_{\alpha}^{-}\left(A% \right)\right)italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_A ) ) - italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ) ) accounts for the magnitude of “tensions” existing in the signed graph in terms of balance, such that Kα(Gs)subscript𝐾𝛼subscript𝐺𝑠K_{\alpha}\left(G_{s}\right)italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) will be 00 if the balanced and unbalanced contributions are equal, and will be 1111 if there are no unbalanced contributions.

4.3 How memory is accounted for

We first show that the time-fractional Caputo derivative accounts for the memory of the system about its past. We start by writing

Dtαu(t)=1Γ(1α)0t[1(tτ)α]u(τ)𝑑τ=1Γ(1α)0tw(τ)u(τ)𝑑τ,superscriptsubscript𝐷𝑡𝛼𝑢𝑡1Γ1𝛼superscriptsubscript0𝑡delimited-[]1superscript𝑡𝜏𝛼superscript𝑢𝜏differential-d𝜏1Γ1𝛼superscriptsubscript0𝑡𝑤𝜏superscript𝑢𝜏differential-d𝜏\begin{split}D_{t}^{\alpha}u\left(t\right)=\dfrac{1}{\varGamma\left(1-\alpha% \right)}\int_{0}^{t}\left[\dfrac{1}{\left(t-\tau\right)^{\alpha}}\right]u^{% \prime}\left(\tau\right)d\tau=\dfrac{1}{\varGamma\left(1-\alpha\right)}\int_{0% }^{t}w\left(\tau\right)u^{\prime}\left(\tau\right)d\tau,\end{split}start_ROW start_CELL italic_D start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_u ( italic_t ) = divide start_ARG 1 end_ARG start_ARG roman_Γ ( 1 - italic_α ) end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT [ divide start_ARG 1 end_ARG start_ARG ( italic_t - italic_τ ) start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT end_ARG ] italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_τ ) italic_d italic_τ = divide start_ARG 1 end_ARG start_ARG roman_Γ ( 1 - italic_α ) end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT italic_w ( italic_τ ) italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_τ ) italic_d italic_τ , end_CELL end_ROW (4.13)

where w(τ)=1/(tτ)α𝑤𝜏1superscript𝑡𝜏𝛼w\left(\tau\right)=1/\left(t-\tau\right)^{\alpha}italic_w ( italic_τ ) = 1 / ( italic_t - italic_τ ) start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT is used to indicate that u(τ)superscript𝑢𝜏u^{\prime}\left(\tau\right)italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_τ ) is integrated in a weighted way that w(τ0)𝑤𝜏0w\left(\tau\rightarrow 0\right)italic_w ( italic_τ → 0 ) is significantly smaller than w(τt)𝑤𝜏𝑡w\left(\tau\rightarrow t\right)italic_w ( italic_τ → italic_t ). Odibat [40] has proved that

Dtαu(t)=C(u,h,α)EC(u,h,α),superscriptsubscript𝐷𝑡𝛼𝑢𝑡𝐶𝑢𝛼subscript𝐸𝐶𝑢𝛼D_{t}^{\alpha}u\left(t\right)=C\left(u,h,\alpha\right)-E_{C}\left(u,h,\alpha% \right),italic_D start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT italic_u ( italic_t ) = italic_C ( italic_u , italic_h , italic_α ) - italic_E start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( italic_u , italic_h , italic_α ) , (4.14)

where EC(u,h,α)𝒪(h2)subscript𝐸𝐶𝑢𝛼𝒪superscript2E_{C}\left(u,h,\alpha\right)\leq\mathcal{O}\left(h^{2}\right)italic_E start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( italic_u , italic_h , italic_α ) ≤ caligraphic_O ( italic_h start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) is the error term, and

C(u,h,α)=h1αΓ(3α)[((k1)2α(k+α2)k1α)u(0)remote past+j=1k1((kj+1)2α2(kj)2α+(kj1)2α)u(tj)recent past+u(t)present],𝐶𝑢𝛼superscript1𝛼Γ3𝛼delimited-[]remote pastsuperscript𝑘12𝛼𝑘𝛼2superscript𝑘1𝛼superscript𝑢0recent pastsuperscriptsubscript𝑗1𝑘1superscript𝑘𝑗12𝛼2superscript𝑘𝑗2𝛼superscript𝑘𝑗12𝛼superscript𝑢subscript𝑡𝑗presentsuperscript𝑢𝑡\begin{split}C\left(u,h,\alpha\right)&=\dfrac{h^{1-\alpha}}{\varGamma\left(3-% \alpha\right)}\left[\underset{\textnormal{remote past}}{\underbrace{\left(% \left(k-1\right)^{2-\alpha}-\left(k+\alpha-2\right)k^{1-\alpha}\right)u^{% \prime}\left(0\right)}}\right.\\ &\left.+\underset{\textnormal{recent past}}{\underbrace{\sum_{j=1}^{k-1}\left(% \left(k-j+1\right)^{2-\alpha}-2\left(k-j\right)^{2-\alpha}+\left(k-j-1\right)^% {2-\alpha}\right)u^{\prime}\left(t_{j}\right)}}+\underset{\textnormal{present}% }{\underbrace{u^{\prime}\left(t\right)}}\right],\end{split}start_ROW start_CELL italic_C ( italic_u , italic_h , italic_α ) end_CELL start_CELL = divide start_ARG italic_h start_POSTSUPERSCRIPT 1 - italic_α end_POSTSUPERSCRIPT end_ARG start_ARG roman_Γ ( 3 - italic_α ) end_ARG [ underremote past start_ARG under⏟ start_ARG ( ( italic_k - 1 ) start_POSTSUPERSCRIPT 2 - italic_α end_POSTSUPERSCRIPT - ( italic_k + italic_α - 2 ) italic_k start_POSTSUPERSCRIPT 1 - italic_α end_POSTSUPERSCRIPT ) italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( 0 ) end_ARG end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL + underrecent past start_ARG under⏟ start_ARG ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_k - 1 end_POSTSUPERSCRIPT ( ( italic_k - italic_j + 1 ) start_POSTSUPERSCRIPT 2 - italic_α end_POSTSUPERSCRIPT - 2 ( italic_k - italic_j ) start_POSTSUPERSCRIPT 2 - italic_α end_POSTSUPERSCRIPT + ( italic_k - italic_j - 1 ) start_POSTSUPERSCRIPT 2 - italic_α end_POSTSUPERSCRIPT ) italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) end_ARG end_ARG + underpresent start_ARG under⏟ start_ARG italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) end_ARG end_ARG ] , end_CELL end_ROW (4.15)

where the interval [0,t]0𝑡\left[0,t\right][ 0 , italic_t ] has been subdivided into k𝑘kitalic_k subintervals [tj,tj+1]subscript𝑡𝑗subscript𝑡𝑗1\left[t_{j},t_{j+1}\right][ italic_t start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT , italic_t start_POSTSUBSCRIPT italic_j + 1 end_POSTSUBSCRIPT ] for j=0,,k𝑗0𝑘j=0,\ldots,kitalic_j = 0 , … , italic_k of equal length h=t/k𝑡𝑘h=t/kitalic_h = italic_t / italic_k. The term C(u,h,α)𝐶𝑢𝛼C\left(u,h,\alpha\right)italic_C ( italic_u , italic_h , italic_α ) confirms that differently from the standard time derivative which considers only the present, the Caputo one takes into account the “remote past” and “recent past” together with the “present” state of the evolution of the function u()𝑢u\left(\cdot\right)italic_u ( ⋅ ). Additionally, the time-fractional Caputo derivative gives smaller weight to the remote past, and such weight increases as we approach to the contribution of the present, which receives the largest weight.

Let us now see the special case of α=1𝛼1\alpha=1italic_α = 1 and how memory could be incorporated while changing α𝛼\alphaitalic_α. We note that C(u,h,α=1)=u(t)𝐶𝑢𝛼1superscript𝑢𝑡C\left(u,h,\alpha=1\right)=u^{\prime}\left(t\right)italic_C ( italic_u , italic_h , italic_α = 1 ) = italic_u start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ). The solution of the NC diffusion (4.5) with uj0=δj,vsuperscriptsubscript𝑢𝑗0subscript𝛿𝑗𝑣u_{j}^{0}=\delta_{j,v}italic_u start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT = italic_δ start_POSTSUBSCRIPT italic_j , italic_v end_POSTSUBSCRIPT is given by uv(t)=(etA)vvsubscript𝑢𝑣𝑡subscriptsuperscript𝑒𝑡𝐴𝑣𝑣u_{v}\left(t\right)=\left(e^{tA}\right)_{vv}italic_u start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT ( italic_t ) = ( italic_e start_POSTSUPERSCRIPT italic_t italic_A end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT italic_v italic_v end_POSTSUBSCRIPT, i.e., the exponential of the adjacency matrix. For t=1𝑡1t=1italic_t = 1, we know that

eA=I+A+A22!++Akk!+,superscript𝑒𝐴𝐼𝐴superscript𝐴22superscript𝐴𝑘𝑘e^{A}=I+A+\dfrac{A^{2}}{2!}+\ldots+\dfrac{A^{k}}{k!}+\ldots,italic_e start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = italic_I + italic_A + divide start_ARG italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 ! end_ARG + … + divide start_ARG italic_A start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG start_ARG italic_k ! end_ARG + … , (4.16)

which means that walks taking a large number of steps are so heavily penalized by 1/(k!)1𝑘1/\left(k!\right)1 / ( italic_k ! ) that they are almost forgotten. Let us consider again an unbalanced cycle of 10101010 vertices and only one negative edge, C10superscriptsubscript𝐶10C_{10}^{-}italic_C start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT. Here, we truncate the expressions (4.16) and e|A|superscript𝑒𝐴e^{\left|A\right|}italic_e start_POSTSUPERSCRIPT | italic_A | end_POSTSUPERSCRIPT at a given value k𝑘kitalic_k, denoted by eA(k)superscript𝑒𝐴𝑘e^{A}\left(k\right)italic_e start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_k ) and e|A|(k)superscript𝑒𝐴𝑘e^{\left|A\right|}\left(k\right)italic_e start_POSTSUPERSCRIPT | italic_A | end_POSTSUPERSCRIPT ( italic_k ), respectively. Then, for any k<10𝑘10k<10italic_k < 10, we have that Tr(eA(k))=Tr(e|A|(k)).𝑇𝑟superscript𝑒𝐴𝑘𝑇𝑟superscript𝑒𝐴𝑘Tr(e^{A}\left(k\right))=Tr(e^{\left|A\right|}\left(k\right)).italic_T italic_r ( italic_e start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_k ) ) = italic_T italic_r ( italic_e start_POSTSUPERSCRIPT | italic_A | end_POSTSUPERSCRIPT ( italic_k ) ) . Therefore, any penalization cksubscript𝑐𝑘c_{k}italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT in f(A)=k=0ckAk𝑓𝐴superscriptsubscript𝑘0subscript𝑐𝑘superscript𝐴𝑘f\left(A\right)=\sum_{k=0}^{\infty}c_{k}A^{k}italic_f ( italic_A ) = ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT italic_A start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT that makes c10Tr(|A|10)0subscript𝑐10𝑇𝑟superscript𝐴100c_{10}Tr(\left|A\right|^{10})\approx 0italic_c start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT italic_T italic_r ( | italic_A | start_POSTSUPERSCRIPT 10 end_POSTSUPERSCRIPT ) ≈ 0 will lead to Tr(f(A(C10)))Tr(f(|A(C10)|))𝑇𝑟𝑓𝐴superscriptsubscript𝐶10𝑇𝑟𝑓𝐴superscriptsubscript𝐶10Tr\left(f\left(A\left(C_{10}^{-}\right)\right)\right)\approx Tr\left(f\left(% \left|A\left(C_{10}^{-}\right)\right|\right)\right)italic_T italic_r ( italic_f ( italic_A ( italic_C start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ) ) ≈ italic_T italic_r ( italic_f ( | italic_A ( italic_C start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) | ) ). This is exactly what happens with ck=1/k!subscript𝑐𝑘1𝑘c_{k}=1/k!italic_c start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = 1 / italic_k !, where Tr(exp(A(C10)))22.7958𝑇𝑟𝐴superscriptsubscript𝐶1022.7958Tr\left(\exp\left(A\left(C_{10}^{-}\right)\right)\right)\approx 22.7958italic_T italic_r ( roman_exp ( italic_A ( italic_C start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ) ) ≈ 22.7958 and Tr(exp(|A(C10)|))22.7959𝑇𝑟𝐴superscriptsubscript𝐶1022.7959Tr\left(\exp\left(\left|A\left(C_{10}^{-}\right)\right|\right)\right)\approx 2% 2.7959italic_T italic_r ( roman_exp ( | italic_A ( italic_C start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) | ) ) ≈ 22.7959 leading to K1(C10)0.99999947subscript𝐾1superscriptsubscript𝐶100.99999947K_{1}\left(C_{10}^{-}\right)\approx 0.99999947italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_C start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) ≈ 0.99999947. That is, the index has almost completely forgotten that the graph contains a negative edge. However, the extra freedom introduced in α𝛼\alphaitalic_α allows us to incorporate the information in the past in an appropriate manner. For C10superscriptsubscript𝐶10C_{10}^{-}italic_C start_POSTSUBSCRIPT 10 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 makes that even the remote past receives some weight in the navigation of the diffusive particles, remembering the presence of the negative edge, with K0.50.98290619subscript𝐾0.50.98290619K_{0.5}\approx 0.98290619italic_K start_POSTSUBSCRIPT 0.5 end_POSTSUBSCRIPT ≈ 0.98290619. Such memory can further take effect by dropping α𝛼\alphaitalic_α, which may be considered as the memory effect parameter, e.g., K0.250.109subscript𝐾0.250.109K_{0.25}\approx 0.109italic_K start_POSTSUBSCRIPT 0.25 end_POSTSUBSCRIPT ≈ 0.109.

We now proceed to find the analytical expression for the degree of balance with memory for unbalanced cycles, i.e., cycles with an odd number of negative edges. We start by proving that unbalanced cycles share the same spectrum, independently of the exact number of negative edges.

Proposition 4.1.

There are two switching classes for signed cycles of length n𝑛nitalic_n, one corresponding to balance and the other corresponding to unbalance.

Proof.

For balanced cycles of length n𝑛nitalic_n, we know that they form a switching class. We then prove that all unbalanced signed cycles form one switching class. For an unbalanced signed cycle of length n𝑛nitalic_n, denoted by Cnsubscript𝐶𝑛C_{n}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT, if we randomly remove one edge e𝑒eitalic_e, it becomes a tree Tnsubscript𝑇𝑛T_{n}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT. We know that every signed tree is balanced, hence Tnsubscript𝑇𝑛T_{n}italic_T start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is balanced and is switching equivalent to the all-positive configuration. Edge e𝑒eitalic_e necessarily breaks the balance structure, since otherwise Cnsubscript𝐶𝑛C_{n}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is balanced. Hence, only one edge violates the balance structure, and Cnsubscript𝐶𝑛C_{n}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT is then switching equivalent to the unbalanced cycles of length n𝑛nitalic_n and one negative edge. Hence, all unbalanced signed cycles of length n𝑛nitalic_n form one switching class. ∎

Corollary 4.2.

All unbalanced signed cycles of length n𝑛nitalic_n share the same eigenvalues, i.e., they are cospectral.

As a consequence of the previous results we will focus on the analytical study of unbalanced cycles as a general class.

Definition 4.3 ([38]).

Let Eα(z)subscript𝐸𝛼𝑧E_{\alpha}\left(z\right)italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_z ) be the Mittag-Leffler function of z.𝑧z.italic_z . Then, we define the following integral:

ν,α(z)1π0πcos(νθ)Eα(zcosθ)𝑑θ,ν.formulae-sequencesubscript𝜈𝛼𝑧1𝜋superscriptsubscript0𝜋𝜈𝜃subscript𝐸𝛼𝑧𝜃differential-d𝜃𝜈\mathcal{E}_{\nu,\alpha}\left(z\right)\coloneqq\frac{1}{\pi}\int_{0}^{\pi}\cos% \left(\nu\theta\right)E_{\alpha}\left(z\cos\theta\right)d\theta,\>\nu\in% \mathbb{Z}.caligraphic_E start_POSTSUBSCRIPT italic_ν , italic_α end_POSTSUBSCRIPT ( italic_z ) ≔ divide start_ARG 1 end_ARG start_ARG italic_π end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT roman_cos ( italic_ν italic_θ ) italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_z roman_cos italic_θ ) italic_d italic_θ , italic_ν ∈ blackboard_Z . (4.17)
Remark 4.4.

Notice that ν,α=1(z)=1π0πezcosθcos(νθ)𝑑θIν(z)subscript𝜈𝛼1𝑧1𝜋superscriptsubscript0𝜋superscript𝑒𝑧𝜃𝜈𝜃differential-d𝜃subscript𝐼𝜈𝑧\mathcal{E}_{\nu,\alpha=1}\left(z\right)=\frac{1}{\pi}\int_{0}^{\pi}e^{z\cos% \theta}\cos\left(\nu\theta\right)d\theta\eqqcolon I_{\nu}\left(z\right)caligraphic_E start_POSTSUBSCRIPT italic_ν , italic_α = 1 end_POSTSUBSCRIPT ( italic_z ) = divide start_ARG 1 end_ARG start_ARG italic_π end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT italic_e start_POSTSUPERSCRIPT italic_z roman_cos italic_θ end_POSTSUPERSCRIPT roman_cos ( italic_ν italic_θ ) italic_d italic_θ ≕ italic_I start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_z ) is the modified Bessel function of the first kind. The fractional modified Bessel function of the first kind can be calculated by using the following result.

Lemma 4.5 ([38]).

Let ν,α(z)subscript𝜈𝛼𝑧\mathcal{E}_{\nu,\alpha}(z)caligraphic_E start_POSTSUBSCRIPT italic_ν , italic_α end_POSTSUBSCRIPT ( italic_z ) be the fractional modified Bessel function of the first kind of z𝑧zitalic_z with fractional parameter α𝛼\alphaitalic_α and ν𝜈\nu\in\mathbb{Z}italic_ν ∈ blackboard_Z. Then,

ν,α(z)=k=0(2k+ν)!Γ(α(2k+ν)+1)k!(k+ν)!(z2)2k+ν.subscript𝜈𝛼𝑧superscriptsubscript𝑘02𝑘𝜈Γ𝛼2𝑘𝜈1𝑘𝑘𝜈superscript𝑧22𝑘𝜈\mathcal{E}_{\nu,\alpha}(z)=\sum_{k=0}^{\infty}\frac{\left(2k+\nu\right)!}{% \varGamma\left(\alpha\left(2k+\nu\right)+1\right)k!\left(k+\nu\right)!}\left(% \frac{z}{2}\right)^{2k+\nu}.caligraphic_E start_POSTSUBSCRIPT italic_ν , italic_α end_POSTSUBSCRIPT ( italic_z ) = ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT divide start_ARG ( 2 italic_k + italic_ν ) ! end_ARG start_ARG roman_Γ ( italic_α ( 2 italic_k + italic_ν ) + 1 ) italic_k ! ( italic_k + italic_ν ) ! end_ARG ( divide start_ARG italic_z end_ARG start_ARG 2 end_ARG ) start_POSTSUPERSCRIPT 2 italic_k + italic_ν end_POSTSUPERSCRIPT . (4.18)
Theorem 4.6.

Let Cnsuperscriptsubscript𝐶𝑛C_{n}^{-}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT be the cycle graph with an odd number of negative edges and 0<α10𝛼10<\alpha\leq 10 < italic_α ≤ 1. Then,

Kαγ(Cn)=k=1nEα(2γcos(((2k+1)π)n))k=1nEα(2γcos((2kπ)n)),superscriptsubscript𝐾𝛼𝛾superscriptsubscript𝐶𝑛superscriptsubscript𝑘1𝑛subscript𝐸𝛼2𝛾2𝑘1𝜋𝑛superscriptsubscript𝑘1𝑛subscript𝐸𝛼2𝛾2𝑘𝜋𝑛K_{\alpha}^{\gamma}\left(C_{n}^{-}\right)=\dfrac{\sum_{k=1}^{n}E_{\alpha}\left% (2\gamma\cos\left(\dfrac{\left(\left(2k+1\right)\pi\right)}{n}\right)\right)}{% \sum_{k=1}^{n}E_{\alpha}\left(2\gamma\cos\left(\dfrac{\left(2k\pi\right)}{n}% \right)\right)},italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ( italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) = divide start_ARG ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 2 italic_γ roman_cos ( divide start_ARG ( ( 2 italic_k + 1 ) italic_π ) end_ARG start_ARG italic_n end_ARG ) ) end_ARG start_ARG ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 2 italic_γ roman_cos ( divide start_ARG ( 2 italic_k italic_π ) end_ARG start_ARG italic_n end_ARG ) ) end_ARG , (4.19)

where KαγTr(Eα(γA))/Tr(Eα(γ|A|))superscriptsubscript𝐾𝛼𝛾𝑇𝑟subscript𝐸𝛼𝛾𝐴𝑇𝑟subscript𝐸𝛼𝛾𝐴K_{\alpha}^{\gamma}\coloneqq Tr\left(E_{\alpha}\left(\gamma A\right)\right)/Tr% \left(E_{\alpha}\left(\gamma\left|A\right|\right)\right)italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT ≔ italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_γ italic_A ) ) / italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_γ | italic_A | ) ) is a even more general form of the index Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, and γ𝛾\gammaitalic_γ is a positive constant, and

limn1nk=1nEα(2γcos((2kπ)n))=0,α(2γ).𝑛1𝑛superscriptsubscript𝑘1𝑛subscript𝐸𝛼2𝛾2𝑘𝜋𝑛subscript0𝛼2𝛾\underset{n\rightarrow\infty}{\lim}\frac{1}{n}\sum_{k=1}^{n}E_{\alpha}\left(2% \gamma\cos\left(\dfrac{\left(2k\pi\right)}{n}\right)\right)=\mathcal{E}_{0,% \alpha}\left(2\gamma\right).start_UNDERACCENT italic_n → ∞ end_UNDERACCENT start_ARG roman_lim end_ARG divide start_ARG 1 end_ARG start_ARG italic_n end_ARG ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 2 italic_γ roman_cos ( divide start_ARG ( 2 italic_k italic_π ) end_ARG start_ARG italic_n end_ARG ) ) = caligraphic_E start_POSTSUBSCRIPT 0 , italic_α end_POSTSUBSCRIPT ( 2 italic_γ ) . (4.20)
Proof.

We can obtain Eq. (4.19) by the eigenvalues of cycles of size n𝑛nitalic_n and those of the same size and one negative edge [39]. Then for the limit, since for k=1,2,,n𝑘12𝑛k=1,2,\dots,nitalic_k = 1 , 2 , … , italic_n, the angles 2kπ/n2𝑘𝜋𝑛2k\pi/n2 italic_k italic_π / italic_n uniformly cover the interval [0,2π]02𝜋[0,2\pi][ 0 , 2 italic_π ], we can write

limn(1nk=1nEα(2γcos(2kπn)))=(12π02πEα(2γcosϑ)𝑑ϑ),𝑛1𝑛superscriptsubscript𝑘1𝑛subscript𝐸𝛼2𝛾2𝑘𝜋𝑛12𝜋superscriptsubscript02𝜋subscript𝐸𝛼2𝛾italic-ϑdifferential-ditalic-ϑ\underset{n\rightarrow\infty}{\lim}\left(\dfrac{1}{n}\sum_{k=1}^{n}E_{\alpha}% \left(2\gamma\cos\left(\dfrac{2k\pi}{n}\right)\right)\right)=\left(\frac{1}{2% \pi}\int_{0}^{2\pi}E_{\alpha}\left(2\gamma\cos\vartheta\right)d\vartheta\right),start_UNDERACCENT italic_n → ∞ end_UNDERACCENT start_ARG roman_lim end_ARG ( divide start_ARG 1 end_ARG start_ARG italic_n end_ARG ∑ start_POSTSUBSCRIPT italic_k = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 2 italic_γ roman_cos ( divide start_ARG 2 italic_k italic_π end_ARG start_ARG italic_n end_ARG ) ) ) = ( divide start_ARG 1 end_ARG start_ARG 2 italic_π end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_π end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 2 italic_γ roman_cos italic_ϑ ) italic_d italic_ϑ ) , (4.21)

where

(12π02πEα(2γcosϑ)𝑑ϑ)=(1π0πEα(2γcosϑ)𝑑ϑ)=α,0(2γ).12𝜋superscriptsubscript02𝜋subscript𝐸𝛼2𝛾italic-ϑdifferential-ditalic-ϑ1𝜋superscriptsubscript0𝜋subscript𝐸𝛼2𝛾italic-ϑdifferential-ditalic-ϑsubscript𝛼02𝛾\left(\frac{1}{2\pi}\int_{0}^{2\pi}E_{\alpha}\left(2\gamma\cos\vartheta\right)% d\vartheta\right)=\left(\frac{1}{\pi}\int_{0}^{\pi}E_{\alpha}\left(2\gamma\cos% \vartheta\right)d\vartheta\right)=\mathcal{E}_{\alpha,0}\left(2\gamma\right).( divide start_ARG 1 end_ARG start_ARG 2 italic_π end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_π end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 2 italic_γ roman_cos italic_ϑ ) italic_d italic_ϑ ) = ( divide start_ARG 1 end_ARG start_ARG italic_π end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( 2 italic_γ roman_cos italic_ϑ ) italic_d italic_ϑ ) = caligraphic_E start_POSTSUBSCRIPT italic_α , 0 end_POSTSUBSCRIPT ( 2 italic_γ ) . (4.22)

We cannot apply the same approximation as in the proof of (4.20) to the numerator of Kαγsuperscriptsubscript𝐾𝛼𝛾K_{\alpha}^{\gamma}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_γ end_POSTSUPERSCRIPT, because approximating the numerator to the denominator for very large n𝑛nitalic_n largely depends on the values of α𝛼\alphaitalic_α. For instance, for α=1𝛼1\alpha=1italic_α = 1 when n=10𝑛10n=10italic_n = 10 the difference between the two terms is of the order of 106superscript10610^{-6}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT and it drops to 1015superscript101510^{-15}10 start_POSTSUPERSCRIPT - 15 end_POSTSUPERSCRIPT for n=20.𝑛20n=20.italic_n = 20 . However, for α=0.25𝛼0.25\alpha=0.25italic_α = 0.25, this difference is of the order of 107superscript10710^{7}10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT for n=10𝑛10n=10italic_n = 10 and remains of the order of 106superscript10610^{6}10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT for n=20,𝑛20n=20,italic_n = 20 , and of 103superscript10310^{3}10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT for n=40.𝑛40n=40.italic_n = 40 . Therefore, because the denominator can be approximated by α,0(2γ)subscript𝛼02𝛾\mathcal{E}_{\alpha,0}\left(2\gamma\right)caligraphic_E start_POSTSUBSCRIPT italic_α , 0 end_POSTSUBSCRIPT ( 2 italic_γ ), we have that for α=1𝛼1\alpha=1italic_α = 1, the balance index approaches 1111 for relatively small unbalanced cycles, while this is far from being the case for α=0.25𝛼0.25\alpha=0.25italic_α = 0.25. This is visually clear when we examine the change of Kα(Cn)subscript𝐾𝛼subscript𝐶𝑛K_{\alpha}\left(C_{n}\right)italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) as a function of both n𝑛nitalic_n and α𝛼\alphaitalic_α (note that γ=Γ(α+1)𝛾Γ𝛼1\gamma=\varGamma\left(\alpha+1\right)italic_γ = roman_Γ ( italic_α + 1 ) throughout the paper so we ignore the superscript); see Fig. 3 (left). Specifically, for values of α𝛼\alphaitalic_α close to 1111, the values of Kα(Cn)subscript𝐾𝛼subscript𝐶𝑛K_{\alpha}\left(C_{n}\right)italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) are close to 1111 for almost all cycles with size n10𝑛10n\geq 10italic_n ≥ 10. At the other extreme when α𝛼\alphaitalic_α is close to 00, the values of the balance index are extremely low for almost every cycle with n20𝑛20n\leq 20italic_n ≤ 20.

Refer to caption Refer to caption
Figure 3: (left) Contour plot of the values of Kα(Cn)subscript𝐾𝛼subscript𝐶𝑛K_{\alpha}\left(C_{n}\right)italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ) for unbalanced cycles with n𝑛nitalic_n vertices for values of 0.1α1.0.1𝛼10.1\leq\alpha\leq 1.0.1 ≤ italic_α ≤ 1 . (right) Difference of the balance index Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT for graph c) versus that for graph d) in Fig. 1 for values of 0.2α10.2𝛼10.2\leq\alpha\leq 10.2 ≤ italic_α ≤ 1.

4.4 Properties of the balance index with memory

We start with the range of the balance index we have proposed.

Theorem 4.7.

The index Kα(G)subscript𝐾𝛼𝐺K_{\alpha}\left(G\right)italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_G ) is bounded as

0Kα(G)1,0subscript𝐾𝛼𝐺10\leq K_{\alpha}\left(G\right)\leq 1,0 ≤ italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_G ) ≤ 1 , (4.23)

where the upper bound is reached if and only if the signed graph G𝐺Gitalic_G is balanced.

Proof.

It is clear from Eq. (4.12) that Kα(G)1subscript𝐾𝛼𝐺1K_{\alpha}(G)\leq 1italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_G ) ≤ 1. We now examine the lower bound. Let {λj(A)}={λj+(A)}{λj(A)}subscript𝜆𝑗𝐴superscriptsubscript𝜆𝑗𝐴superscriptsubscript𝜆𝑗𝐴\left\{\lambda_{j}\left(A\right)\right\}=\left\{\lambda_{j}^{+}\left(A\right)% \right\}\cup\left\{\lambda_{j}^{-}\left(A\right)\right\}{ italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_A ) } = { italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_A ) } ∪ { italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ) } be the eigenvalues of A𝐴Aitalic_A, where {λj+(A)}superscriptsubscript𝜆𝑗𝐴\left\{\lambda_{j}^{+}\left(A\right)\right\}{ italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_A ) } and {λj(A)}superscriptsubscript𝜆𝑗𝐴\left\{\lambda_{j}^{-}\left(A\right)\right\}{ italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ) } are the sets of nonnegative and negative eigenvalues of A𝐴Aitalic_A, respectively. Clearly, Eα(λj+(A))0subscript𝐸𝛼superscriptsubscript𝜆𝑗𝐴0E_{\alpha}\left(\lambda_{j}^{+}\left(A\right)\right)\geq 0italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT ( italic_A ) ) ≥ 0. We then consider negative eigenvalues λj(A)superscriptsubscript𝜆𝑗𝐴\lambda_{j}^{-}\left(A\right)italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ), and Eα(|λj(A)|)subscript𝐸𝛼superscriptsubscript𝜆𝑗𝐴E_{\alpha}\left(-\left|\lambda_{j}^{-}\left(A\right)\right|\right)italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( - | italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ) | ). We note that

(1)kdkEα(x)dxk0,superscript1𝑘superscript𝑑𝑘subscript𝐸𝛼𝑥𝑑superscript𝑥𝑘0\left(-1\right)^{k}\dfrac{d^{k}E_{\alpha}\left(-x\right)}{dx^{k}}\geq 0,( - 1 ) start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT divide start_ARG italic_d start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( - italic_x ) end_ARG start_ARG italic_d italic_x start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT end_ARG ≥ 0 , (4.24)

for all x𝑥xitalic_x and for 0α10𝛼10\leq\alpha\leq 10 ≤ italic_α ≤ 1 [41]. Hence, Eα(|λj(A)|)0subscript𝐸𝛼superscriptsubscript𝜆𝑗𝐴0E_{\alpha}\left(-\left|\lambda_{j}^{-}\left(A\right)\right|\right)\geq 0italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( - | italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ( italic_A ) | ) ≥ 0 and

Tr(Eα(A))=j=1nEα(λj(A))0.𝑇𝑟subscript𝐸𝛼𝐴superscriptsubscript𝑗1𝑛subscript𝐸𝛼subscript𝜆𝑗𝐴0Tr\left(E_{\alpha}\left(A\right)\right)=\sum_{j=1}^{n}E_{\alpha}\left(\lambda_% {j}\left(A\right)\right)\geq 0.italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_A ) ) = ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_A ) ) ≥ 0 . (4.25)

We note that |A|𝐴\left|A\right|| italic_A | is a nonnegative matrix, and so is any power of |A|𝐴\left|A\right|| italic_A |, thus Eα(|A|)subscript𝐸𝛼𝐴E_{\alpha}\left(\left|A\right|\right)italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( | italic_A | ). Hence, Tr(Eα(|A|))0𝑇𝑟subscript𝐸𝛼𝐴0Tr\left(E_{\alpha}\left(\left|A\right|\right)\right)\geq 0italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( | italic_A | ) ) ≥ 0, and then 0Kα(G)0subscript𝐾𝛼𝐺0\leq K_{\alpha}\left(G\right)0 ≤ italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_G ).

It is clear from Eq. (4.12) that Kα=1subscript𝐾𝛼1K_{\alpha}=1italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT = 1 if and only if Tr(Eα(A))=0𝑇𝑟subscriptsuperscript𝐸𝛼𝐴0Tr\left(E^{-}_{\alpha}\left(A\right)\right)=0italic_T italic_r ( italic_E start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_A ) ) = 0, if and only if there is no negative closed walks of any length involving any vertices, if and only if there is no negative cycles, i.e., the graph is balanced. While for the lower bound, we require Tr(Eα(A))=Tr(Eα+(A))𝑇𝑟subscriptsuperscript𝐸𝛼𝐴𝑇𝑟subscriptsuperscript𝐸𝛼𝐴Tr\left(E^{-}_{\alpha}\left(A\right)\right)=Tr\left(E^{+}_{\alpha}\left(A% \right)\right)italic_T italic_r ( italic_E start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_A ) ) = italic_T italic_r ( italic_E start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_A ) ), which can only happen when the number of negative (unbalanced) closed walks is sufficiently large in a signed graph. ∎

We now consider again the signed Petersen graphs, specifically the two labelled as c) and d) in Fig. 1. We recall that although graph d) reaches consensus at a time significantly smaller than graph d), the first has a larger value of K1subscript𝐾1K_{1}italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, due to the heavy penalization to walks of relatively large sizes, imposed by the exponential (see section 3). We now consider Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT as a function of α𝛼\alphaitalic_α, between these two graphs; see Fig. 3 (right). Specifically, at α=1𝛼1\alpha=1italic_α = 1, the graph d) is more balanced than graph c), corresponding to a negative value of the difference between Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT of c) minus that of d). This negative difference becomes larger when α𝛼\alphaitalic_α drops from 1111, reaching a minimum at about α=0.84.𝛼0.84\alpha=0.84.italic_α = 0.84 . However, after this point, the trend reverses towards positive values, reaching the maximum at around α=0.5.𝛼0.5\alpha=0.5.italic_α = 0.5 . At this value of α𝛼\alphaitalic_α, the penalization of longer cycles is not as heavy, since the larger number of negative hexagons in d) overcome the larger number of negative pentagons in c). If we now correlate the values of K0.5subscript𝐾0.5K_{0.5}italic_K start_POSTSUBSCRIPT 0.5 end_POSTSUBSCRIPT versus tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the squared Pearson correlation coefficient has value 0.9810.9810.9810.981, which clearly contrasts with the one of 0.9240.9240.9240.924 for K1subscript𝐾1K_{1}italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, implying that K0.5subscript𝐾0.5K_{0.5}italic_K start_POSTSUBSCRIPT 0.5 end_POSTSUBSCRIPT provides a better indicator of balance in terms of the convergence of the diffusion (K0.5subscript𝐾0.5K_{0.5}italic_K start_POSTSUBSCRIPT 0.5 end_POSTSUBSCRIPT for the five signed Petersen graphs of Fig. 1, from a) to e), are: 0.38780.38780.38780.3878; 0.19730.19730.19730.1973; 0.15140.15140.15140.1514; 0.13020.13020.13020.1302; 0.07870.07870.07870.0787).

To gain more insights about the significance of the use of memory to account for balance in signed graphs, let us further explore the changes. When α𝛼\alphaitalic_α drops from 1.01.01.01.0 to about 0.80.80.80.8, the balance index of the signed Petersen graph d) increases relative to that of graph c). This can be explained by the fact that these graphs are triangle and quadrilateral free, and the smallest cycle is of length five. As we drop initially the value of α𝛼\alphaitalic_α from 1111, the contribution of C5superscriptsubscript𝐶5C_{5}^{-}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT increases, and because graph c) has more of these cycles than d), it is less balanced relative to d). However, as we continue dropping α𝛼\alphaitalic_α, the contribution of C6superscriptsubscript𝐶6C_{6}^{-}italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT growth significantly. In this case, graph d) overcomes graph c) in the number of C6superscriptsubscript𝐶6C_{6}^{-}italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, which make c) more balanced than d) after some critical value and the difference reaches a maximum at around α=0.5𝛼0.5\alpha=0.5italic_α = 0.5. Below this value of the memory parameter α𝛼\alphaitalic_α, the longer cycles, namely C8superscriptsubscript𝐶8C_{8}^{-}italic_C start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT and C9superscriptsubscript𝐶9C_{9}^{-}italic_C start_POSTSUBSCRIPT 9 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, makes their contribution. In this case, graph c) overcomes d) in the number of C8superscriptsubscript𝐶8C_{8}^{-}italic_C start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT, but d) contains a bit more C9superscriptsubscript𝐶9C_{9}^{-}italic_C start_POSTSUBSCRIPT 9 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT than c); see Supplementary Material for details. The effect of these longer unbalanced cycles is a further decay of the balance index of both graphs for α<0.5𝛼0.5\alpha<0.5italic_α < 0.5.

5 Examples of applications

5.1 Gene regulatory networks

We first consider the gene regulatory networks of Saccharomyses cerevisiae (yeast) and of Bacillus subtilis, previously studied as signed undirected graphs by Soranzo et al. [46]. We maintain the undirected versions of these networks, and consider only their giant connected components. The balance index at α=1𝛼1\alpha=1italic_α = 1 indicates that the network of S. cerevisiae is slightly more balanced than that of B. subtilis, with K10.933subscript𝐾10.933K_{1}\approx 0.933italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 0.933 versus K10.643subscript𝐾10.643K_{1}\approx 0.643italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 0.643, respectively. We note that the difference between the values of K1subscript𝐾1K_{1}italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT for both networks is smaller than 0.30.30.30.3 and both are not far from 1111, which implies that there is no significant difference in their degree of balance and that they form relatively balanced systems. However, if we allow for an increment of the memory in the system by dropping α𝛼\alphaitalic_α, then the results change significantly. As can be seen in Fig. 4 (left), the difference in balance between the two gene regulatory networks increases up to 0.90.90.90.9 (of a maximum of 1.01.01.01.0) when α𝛼\alphaitalic_α drops from 1.01.01.01.0 to 0.620.620.620.62, where K0.620.928subscript𝐾0.620.928K_{0.62}\approx 0.928italic_K start_POSTSUBSCRIPT 0.62 end_POSTSUBSCRIPT ≈ 0.928 for S. cerevisiae and K0.620.014subscript𝐾0.620.014K_{0.62}\approx 0.014italic_K start_POSTSUBSCRIPT 0.62 end_POSTSUBSCRIPT ≈ 0.014 for B. subtilis. That is, while the gene regulatory network of yeast is highly balanced, the one of B. subtilis is extremely unbalanced.

Refer to caption Refer to caption Refer to caption
Figure 4: (left) Plot of the difference in balance Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT as a function of α𝛼\alphaitalic_α for the gene regulatory networks of yeast and B. subtilis. (middle) Plot of the percentages of negative cycles Cnsuperscriptsubscript𝐶𝑛C_{n}^{-}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT for 3n113𝑛113\leq n\leq 113 ≤ italic_n ≤ 11 for the gene regulatory networks of yeast and B. subtilis. The percentages are calculated as 100Cn/(Cn++Cn).100superscriptsubscript𝐶𝑛superscriptsubscript𝐶𝑛superscriptsubscript𝐶𝑛100\cdot C_{n}^{-}/\left(C_{n}^{+}+C_{n}^{-}\right).100 ⋅ italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT / ( italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT + italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT ) . (right) Plot of the nonmonotonic change of Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT as a function of α𝛼\alphaitalic_α for the gene regulatory network of yeast.

This difference is mainly due to the fact that the network of B. subtilis has a large number of relatively large unbalanced cycles. We observe that although Cnsuperscriptsubscript𝐶𝑛C_{n}^{-}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT grows exponentially fast in both networks, it grows faster for the network of B. subtilis; see Supplementary Material for details. This can be implied from Fig. 4 (middle) where we visualise the percentages of negative cycles of increasing lengths. It can be seen that the network of yeast has relatively more unbalanced triangles and pentagons but significantly less percentage of negative squares than the one of B. subtilis. This explains why both networks have comparable values of the balance index when α𝛼\alphaitalic_α is close to one, i.e., the memory of the system is relatively low although the one of yeast is slightly more balanced than the one of B. subtilis. However, when cycles of longer lengths (n6𝑛6n\geq 6italic_n ≥ 6) are taken into account, the gene regulatory network of B. subtilis has systematically more percentage of unbalanced cycles than the one of yeast. This clearly explains why the network of B. subtilis is significantly less balanced than the one of yeast for relatively low values of α𝛼\alphaitalic_α, i.e., the memory of the system increases.

Regarding the change of Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT with respect to the drop of the memory parameter α𝛼\alphaitalic_α in signed graphs, another interesting characteristic is the possibility of nonmonotonicity; see the case of the gene regulatory network of yeast in Fig. 4 (right). Specifically, for the network of yeast, Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT increases when α𝛼\alphaitalic_α drops from 1.01.01.01.0 to about 0.750.750.750.75, and then decays very quickly for values α<0.75𝛼0.75\alpha<0.75italic_α < 0.75. In practical terms, this means that there is an “optimal” value of the memory that maximizes the degree of balance of this network, and that such value is different from α=1𝛼1\alpha=1italic_α = 1. The structural explanation for this nonmonotonicity can also be found in the plot in Fig. 4 (middle). We observe that there is a significant drop in the percentage of negative squares in this network, which contributes to increasing balance when we drop α𝛼\alphaitalic_α from 1.01.01.01.0 to about 0.750.750.750.75. However, as value of α𝛼\alphaitalic_α decays beyond 0.750.750.750.75, the longer negative cycles become more important, and the global balance of the network quickly decays.

5.2 Spatial ecological networks

We now study a series of 31313131 signed networks representing patterns of spatial (co)occurrence of plants in four major locations in Spain, specifically, Cabo de Gata-Nijar National Park (36.77N, –2.11W), Monegros (41.65N, –0.71W), Sierra de Guara (42.27N, 0.18W) and Ordesa-Monte Perdido National Park (42.63N, –0.11W). The vertices of these networks represent plant species and two vertices form an edge in the graph if the corresponding plants has a spatial association, which was calculated by comparing the number of times that the two species appeared at the same point on the transects. Two plant species share a positive edge if they appeared associated in close region of space, while negative associations correspond to plants appearing separated at a significant distance in space [43]. Therefore, patterns of signed cycles appear in these networks. The meaning of these patterns is self-explained, where a fully positive triangle, for instance, indicates that the three plants have certain type of cooperative relations that allow them to coexist in the same spatial region. A fully negative triangle indicates competitive interactions between the three species that avoids their coexistence in the same location.

We give the average values of Kα=1subscript𝐾𝛼1K_{\alpha=1}italic_K start_POSTSUBSCRIPT italic_α = 1 end_POSTSUBSCRIPT and Kα=0.8subscript𝐾𝛼0.8K_{\alpha=0.8}italic_K start_POSTSUBSCRIPT italic_α = 0.8 end_POSTSUBSCRIPT for the networks in each of the four major locations in Table 1. We also reproduce the values of the mean temperature and precipitation of those regions as reported by Saiz et al. [43]. The results for Kα=1delimited-⟨⟩subscript𝐾𝛼1\left\langle K_{\alpha=1}\right\rangle⟨ italic_K start_POSTSUBSCRIPT italic_α = 1 end_POSTSUBSCRIPT ⟩ are qualitatively similar to those in [43], where an index of balance R𝑅Ritalic_R based on triangles only was used. These results lead to the fact that the balance in those places of higher temperature and lower precipitation is bigger than in those where the temperature is low and the precipitation is high. Both Kα=1delimited-⟨⟩subscript𝐾𝛼1\left\langle K_{\alpha=1}\right\rangle⟨ italic_K start_POSTSUBSCRIPT italic_α = 1 end_POSTSUBSCRIPT ⟩ and R𝑅Ritalic_R identify Monegros as the site with the largest balance and Sierra de Guara as the one more out of balance. However, when we increase the memory of the system by considering a lower value of α𝛼\alphaitalic_α, e.g., Kα=0.8delimited-⟨⟩subscript𝐾𝛼0.8\left\langle K_{\alpha=0.8}\right\rangle⟨ italic_K start_POSTSUBSCRIPT italic_α = 0.8 end_POSTSUBSCRIPT ⟩, a swap on the values of balance of Cabo de Gata and Monegros appears;see Table 1.

location Temp. (Csuperscript𝐶{}^{\circ}Cstart_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPT italic_C) Prec. (mm) Kα=1delimited-⟨⟩subscript𝐾𝛼1\left\langle K_{\alpha=1}\right\rangle⟨ italic_K start_POSTSUBSCRIPT italic_α = 1 end_POSTSUBSCRIPT ⟩ Kα=0.8delimited-⟨⟩subscript𝐾𝛼0.8\left\langle K_{\alpha=0.8}\right\rangle⟨ italic_K start_POSTSUBSCRIPT italic_α = 0.8 end_POSTSUBSCRIPT ⟩
Cabo de Gata 24 328 0.335±0.265plus-or-minus0.3350.2650.335\pm 0.2650.335 ± 0.265 0.165±0.192plus-or-minus0.1650.1920.165\pm 0.1920.165 ± 0.192
Monegros 21 360 0.376±0.221plus-or-minus0.3760.2210.376\pm 0.2210.376 ± 0.221 0.129±0.130plus-or-minus0.1290.1300.129\pm 0.1300.129 ± 0.130
Sierra de Guara 17 927 0.0883±0.033plus-or-minus0.08830.0330.0883\pm 0.0330.0883 ± 0.033 0.0061±0.0042plus-or-minus0.00610.00420.0061\pm 0.00420.0061 ± 0.0042
Ordesa-Monte Perdido 11 1485 0.0985±0.072plus-or-minus0.09850.0720.0985\pm 0.0720.0985 ± 0.072 0.014±0.016plus-or-minus0.0140.0160.014\pm 0.0160.014 ± 0.016
Table 1: Values of the balance degree indices Kα=1delimited-⟨⟩subscript𝐾𝛼1\left\langle K_{\alpha=1}\right\rangle⟨ italic_K start_POSTSUBSCRIPT italic_α = 1 end_POSTSUBSCRIPT ⟩ and Kα=0.8delimited-⟨⟩subscript𝐾𝛼0.8\left\langle K_{\alpha=0.8}\right\rangle⟨ italic_K start_POSTSUBSCRIPT italic_α = 0.8 end_POSTSUBSCRIPT ⟩ averaged for all the networks in the four main geographic locations studied here. The values of the mean temperature and precipitation in those regions are reported as in Saiz et al. [42].

We visualise the average and standard deviations of Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT for 0.4α10.4𝛼10.4\leq\alpha\leq 10.4 ≤ italic_α ≤ 1 for the 31313131 signed ecological networks grouped in the four major sites under study in Fig. 5 (left). The crossing between the balance rankings of Cabo de Gata and Monegros occurs around α=0.8𝛼0.8\alpha=0.8italic_α = 0.8. Furthermore, if we try to explain the degree of balance of these sites by considering a single parameter like the precipitation – while noticing the risks of doing any correlation for only four points – we observe some interesting patterns. A power-law fitting of the type: Kα=1P2.116similar-todelimited-⟨⟩subscript𝐾𝛼1superscript𝑃2.116\left\langle K_{\alpha=1}\right\rangle\sim P^{-2.116}⟨ italic_K start_POSTSUBSCRIPT italic_α = 1 end_POSTSUBSCRIPT ⟩ ∼ italic_P start_POSTSUPERSCRIPT - 2.116 end_POSTSUPERSCRIPT gives a correlation coefficient of r0.88𝑟0.88r\approx 0.88italic_r ≈ 0.88. Similarly, RP1.428similar-to𝑅superscript𝑃1.428R\sim P^{-1.428}italic_R ∼ italic_P start_POSTSUPERSCRIPT - 1.428 end_POSTSUPERSCRIPT with r0.89𝑟0.89r\approx 0.89italic_r ≈ 0.89. However, when memory effects take place, we obtain: Kα=0.8P3.156similar-todelimited-⟨⟩subscript𝐾𝛼0.8superscript𝑃3.156\left\langle K_{\alpha=0.8}\right\rangle\sim P^{-3.156}⟨ italic_K start_POSTSUBSCRIPT italic_α = 0.8 end_POSTSUBSCRIPT ⟩ ∼ italic_P start_POSTSUPERSCRIPT - 3.156 end_POSTSUPERSCRIPT with r0.94𝑟0.94r\approx 0.94italic_r ≈ 0.94. That is, the memory effects increase the amount of variance in the index explained from 77%percent7777\%77 % with Kα=1delimited-⟨⟩subscript𝐾𝛼1\left\langle K_{\alpha=1}\right\rangle⟨ italic_K start_POSTSUBSCRIPT italic_α = 1 end_POSTSUBSCRIPT ⟩ to 88%percent8888\%88 % with Kα=0.8delimited-⟨⟩subscript𝐾𝛼0.8\left\langle K_{\alpha=0.8}\right\rangle⟨ italic_K start_POSTSUBSCRIPT italic_α = 0.8 end_POSTSUBSCRIPT ⟩.

Refer to caption Refer to caption
Figure 5: (left) Change of the average balance Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT for 0.4α10.4𝛼10.4\leq\alpha\leq 10.4 ≤ italic_α ≤ 1 of the signed ecological networks in the four major Spanish locations studied. (right) Plot of the change of the Pearson correlation coefficient r2superscript𝑟2r^{2}italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of the balance indices with memory Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT for 0.4α10.4𝛼10.4\leq\alpha\leq 10.4 ≤ italic_α ≤ 1 versus the amount of precipitation in mm on the four major locations studied in this work.

The question of the existence of an optimal value for the memory effect remains. We obtain the power-law correlation between the balance indices Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT for 0.4α10.4𝛼10.4\leq\alpha\leq 10.4 ≤ italic_α ≤ 1 and the mean precipitation in the corresponding main locations. The correlation coefficient increases when α𝛼\alphaitalic_α drops from 1111 up to 0.60.60.60.6, and then it decays very quickly; see Fig. 5 (right). This implies that memory effects in ecological systems may have an optimum. However, more research in this area is needed to obtain more conclusive insights about this important question, and we leave it to future work.

5.3 Social networks in rural villages

Finally, we consider a set of social networks constructed from the data of 24696246962469624696 people aged 12121212 to 93939393 years in geographically isolated villages in western Honduras [32]. The vertices of these networks represent residents within each village, and they are connected by a positive (negative) edge if either of them identify the other as a friend (an enemy), while if one identify the other as a friend while the other identify the one as an enemy, we connect them by a negative edge. We note that the case that they have no opinion of each other is also allowed. By design, the networks are solely within-village networks, and we select 11111111 of them for our analysis, labelled as A𝐴Aitalic_A up to K𝐾Kitalic_K. Cycles of various lengths can frequently occur in such social networks, positive or negative. Corresponding to balance theory [9], positive cycles indicate that the residents can be partitioned into one or two communities without conflicts inside each community, while negative cycles indicate the existence of conflicts of the relationships between residents.

Refer to caption
Figure 6: Plot of the balance index Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT as a function for α𝛼\alphaitalic_α for the 11111111 social networks of rural villages.

We consider the balance index Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT for 0.4α10.4𝛼10.4\leq\alpha\leq 10.4 ≤ italic_α ≤ 1 for the 11111111 signed social networks; see Fig. 6. We find that all networks are not completely balanced: all 11111111 networks reaches the maximum value of the balance index Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT at α=1𝛼1\alpha=1italic_α = 1, and Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT quickly decreases as α𝛼\alphaitalic_α deviates from 1111, where the values are almost 00 for all networks at α=0.5𝛼0.5\alpha=0.5italic_α = 0.5. For example, village D has the maximum index value in all 11111111 networks at α=1𝛼1\alpha=1italic_α = 1, with K10.723subscript𝐾10.723K_{1}\approx 0.723italic_K start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 0.723, which implies that the network is close to being balanced. However, it becomes less than 0.50.50.50.5 at α=0.8𝛼0.8\alpha=0.8italic_α = 0.8, and continue decreasing as we increase the memory effect through parameter α𝛼\alphaitalic_α. These imply the abundance of long negative cycles in these social networks, which is consistent with the results in [32], such as the homophily of negative relationships.

Refer to caption Refer to caption
Figure 7: Plot of the truncated index Mrsubscript𝑀𝑟M_{r}italic_M start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT as a function for r𝑟ritalic_r for the social networks of villages D𝐷Ditalic_D and E𝐸Eitalic_E when α=1𝛼1\alpha=1italic_α = 1 (left) and α=0.6𝛼0.6\alpha=0.6italic_α = 0.6 (right)

Specifically, we observe a clear crossing between the change of the index values of village D and that of village E as α𝛼\alphaitalic_α deviates from 1111. In order to understand the differences between the balance of the villages D𝐷Ditalic_D and E𝐸Eitalic_E, we start by defining the following truncated series, Mrsk=0rTr(Ak)/Γ(αk+1)superscriptsubscript𝑀𝑟𝑠superscriptsubscript𝑘0𝑟𝑇𝑟superscript𝐴𝑘Γ𝛼𝑘1M_{r}^{s}\coloneqq\sum_{k=0}^{r}Tr\left(A^{k}\right)/\varGamma\left(\alpha k+1\right)italic_M start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT ≔ ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT italic_T italic_r ( italic_A start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) / roman_Γ ( italic_α italic_k + 1 ) which accounts for the signed contributions of the different spectral moments; Mruk=0rTr(|A|k)/Γ(αk+1)superscriptsubscript𝑀𝑟𝑢superscriptsubscript𝑘0𝑟𝑇𝑟superscript𝐴𝑘Γ𝛼𝑘1M_{r}^{u}\coloneqq\sum_{k=0}^{r}Tr\left(\left|A\right|^{k}\right)/\varGamma% \left(\alpha k+1\right)italic_M start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_u end_POSTSUPERSCRIPT ≔ ∑ start_POSTSUBSCRIPT italic_k = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_r end_POSTSUPERSCRIPT italic_T italic_r ( | italic_A | start_POSTSUPERSCRIPT italic_k end_POSTSUPERSCRIPT ) / roman_Γ ( italic_α italic_k + 1 ) which accounts for the total contribution of closed walks, and Mr=Mrs/Mrusubscript𝑀𝑟superscriptsubscript𝑀𝑟𝑠superscriptsubscript𝑀𝑟𝑢M_{r}=M_{r}^{s}/M_{r}^{u}italic_M start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = italic_M start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT / italic_M start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_u end_POSTSUPERSCRIPT. Obviously, Mr=Mrs/Mrt=Tr(Eα(A))/Tr(Eα(|A|))subscript𝑀𝑟superscriptsubscript𝑀𝑟𝑠superscriptsubscript𝑀𝑟𝑡𝑇𝑟subscript𝐸𝛼𝐴𝑇𝑟subscript𝐸𝛼𝐴M_{r\rightarrow\infty}=M_{r\rightarrow\infty}^{s}/M_{r\rightarrow\infty}^{t}=% Tr\left(E_{\alpha}\left(A\right)\right)/Tr\left(E_{\alpha}\left(\left|A\right|% \right)\right)italic_M start_POSTSUBSCRIPT italic_r → ∞ end_POSTSUBSCRIPT = italic_M start_POSTSUBSCRIPT italic_r → ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_s end_POSTSUPERSCRIPT / italic_M start_POSTSUBSCRIPT italic_r → ∞ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT = italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_A ) ) / italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( | italic_A | ) ) recovers the balance index Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT. We start by truncating the series at r=3𝑟3r=3italic_r = 3 which is where the first signed cycles appear, and then continue increasing r𝑟ritalic_r. First, we plot the results for the two graphs when α=1𝛼1\alpha=1italic_α = 1 in the left of Fig. 7. Hence, the network of village E𝐸Eitalic_E (purple circles) appears to be more balanced than the one of village D𝐷Ditalic_D if we truncate the sum of spectral moments below r=8𝑟8r=8italic_r = 8. Indeed, a simple index based only on triangles indicates that E𝐸Eitalic_E is more balanced than D𝐷Ditalic_D. At about r=8𝑟8r=8italic_r = 8 the cumulative sum of moments for graph D𝐷Ditalic_D become larger than that of graph E𝐸Eitalic_E, indicating that now the former graph is more balanced. The reason for this swap in the balance order is not directly caused by the larger number of longer signed cycles in one over the other, as we have seen in previous examples, but due to the fact that for graph D𝐷Ditalic_D the ratio of the cumulative sum of moments of length smaller than 8888 is smaller than that for the graph E𝐸Eitalic_E. However, increasing this sum to higher-order moments makes it bigger for graph D𝐷Ditalic_D than to graph E𝐸Eitalic_E. For example, the ratio M7(D)2336/28500.8197subscript𝑀7𝐷233628500.8197M_{7}\left(D\right)\approx 2336/2850\approx 0.8197italic_M start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT ( italic_D ) ≈ 2336 / 2850 ≈ 0.8197 which is smaller than M7(E)4442/53940.8236subscript𝑀7𝐸444253940.8236M_{7}\left(E\right)\approx 4442/5394\approx 0.8236italic_M start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT ( italic_E ) ≈ 4442 / 5394 ≈ 0.8236. However, M8(D)(2336+239)/(2850+393)0.7941subscript𝑀8𝐷233623928503930.7941M_{8}\left(D\right)\approx(2336+239)/(2850+393)\approx 0.7941italic_M start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT ( italic_D ) ≈ ( 2336 + 239 ) / ( 2850 + 393 ) ≈ 0.7941 is smaller than M8(E)(4442+815)/(5394+1278)0.7880subscript𝑀8𝐸4442815539412780.7880M_{8}\left(E\right)\approx(4442+815)/(5394+1278)\approx 0.7880italic_M start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT ( italic_E ) ≈ ( 4442 + 815 ) / ( 5394 + 1278 ) ≈ 0.7880, which is independent of the fact that (239/393)<(815/1278)2393938151278(239/393)<(815/1278)( 239 / 393 ) < ( 815 / 1278 ),but depending on the rates on which the numerator and denominator of the M7(D)subscript𝑀7𝐷M_{7}\left(D\right)italic_M start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT ( italic_D ) and M7(E)subscript𝑀7𝐸M_{7}\left(E\right)italic_M start_POSTSUBSCRIPT 7 end_POSTSUBSCRIPT ( italic_E ) growth by the addition of the individual terms.

This effect previously seen for α=1𝛼1\alpha=1italic_α = 1 disappears when we increases the memory effect; see the right of Fig. 7. Specifically, penalizing less the walks increases the difference in balance in favor of graph E𝐸Eitalic_E relative to D𝐷Ditalic_D. Therefore, we may consider the fact that the factorial penalization points out to graph D𝐷Ditalic_D as more balance than E𝐸Eitalic_E as an artifact of this type of penalization, which indeed is solved when the memory of the system increases. This emphasize the importance of selecting an optimal memory effect parameter α𝛼\alphaitalic_α to understand the level of balance of the signed social networks.

6 A useful approximation

We know that

Tr(Eα(γA))=j=1nEα(γλj).𝑇𝑟subscript𝐸𝛼𝛾𝐴superscriptsubscript𝑗1𝑛subscript𝐸𝛼𝛾subscript𝜆𝑗Tr\left(E_{\alpha}\left(\gamma A\right)\right)=\sum_{j=1}^{n}E_{\alpha}\left(% \gamma\lambda_{j}\right).italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_γ italic_A ) ) = ∑ start_POSTSUBSCRIPT italic_j = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_γ italic_λ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) . (6.1)

Let λ1(m1)>λ2(m2)>>λr(mr)superscriptsubscript𝜆1subscript𝑚1superscriptsubscript𝜆2subscript𝑚2superscriptsubscript𝜆𝑟subscript𝑚𝑟\lambda_{1}^{\left(m_{1}\right)}>\lambda_{2}^{\left(m_{2}\right)}>\ldots>% \lambda_{r}^{\left(m_{r}\right)}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT > italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_m start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT > … > italic_λ start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ( italic_m start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT ) end_POSTSUPERSCRIPT be the distinct eigenvalues of A𝐴Aitalic_A together with their multiplicities misubscript𝑚𝑖m_{i}italic_m start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT. For α𝛼\alphaitalic_α relatively low, the function Eα(z)subscript𝐸𝛼𝑧E_{\alpha}\left(z\right)italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_z ) grows extremely fast with the values z𝑧zitalic_z. Therefore, for relatively small values of α𝛼\alphaitalic_α the difference λ1>λ2subscript𝜆1subscript𝜆2\lambda_{1}>\lambda_{2}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT is magnified by Eα(λ1)Eα(λ2),very-much-greater-thansubscript𝐸𝛼subscript𝜆1subscript𝐸𝛼subscript𝜆2E_{\alpha}\left(\lambda_{1}\right)\ggg E_{\alpha}\left(\lambda_{2}\right),italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ⋙ italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) , which implies that

limα0Tr(Eα(γA))=m1Eα(γλ1).𝛼0𝑇𝑟subscript𝐸𝛼𝛾𝐴subscript𝑚1subscript𝐸𝛼𝛾subscript𝜆1\underset{\alpha\rightarrow 0}{\lim}Tr\left(E_{\alpha}\left(\gamma A\right)% \right)=m_{1}E_{\alpha}\left(\gamma\lambda_{1}\right).start_UNDERACCENT italic_α → 0 end_UNDERACCENT start_ARG roman_lim end_ARG italic_T italic_r ( italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_γ italic_A ) ) = italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_γ italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) . (6.2)

If the eigenvalues of |A|𝐴\left|A\right|| italic_A | are μ1>μ2μnsubscript𝜇1subscript𝜇2subscript𝜇𝑛\mu_{1}>\mu_{2}\geq\ldots\geq\mu_{n}italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT > italic_μ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≥ … ≥ italic_μ start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT we will have that

K~αlimα0Kα(Gs)=m1Eα(γλ1)Eα(γμ1).subscript~𝐾𝛼𝛼0subscript𝐾𝛼subscript𝐺𝑠subscript𝑚1subscript𝐸𝛼𝛾subscript𝜆1subscript𝐸𝛼𝛾subscript𝜇1\tilde{K}_{\alpha}\coloneqq\underset{\alpha\rightarrow 0}{\lim}K_{\alpha}\left% (G_{s}\right)=\dfrac{m_{1}E_{\alpha}\left(\gamma\lambda_{1}\right)}{E_{\alpha}% \left(\gamma\mu_{1}\right)}.over~ start_ARG italic_K end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ≔ start_UNDERACCENT italic_α → 0 end_UNDERACCENT start_ARG roman_lim end_ARG italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_G start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ) = divide start_ARG italic_m start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_γ italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) end_ARG start_ARG italic_E start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_γ italic_μ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) end_ARG . (6.3)

We experimentally verify the goodness of this approximation in some of the networks we have studied previously, specifically the signed networks of the rural villages. In Fig. 8 (left), we plot the relative error in the balance index with memory Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT when approximated by K~αsubscript~𝐾𝛼\tilde{K}_{\alpha}over~ start_ARG italic_K end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT for the 11111111 signed networks of the rural villages. We observe that for values of α𝛼\alphaitalic_α close to 1111, the relative error is relatively large for most of the networks, with values of up to 0.70.70.70.7 for village K. However, the error significantly drops when α𝛼\alphaitalic_α changes systematically to 00, and in particular, for α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 the average relative error for the 11111111 networks is 0.0710.0710.0710.071 with only network D having a relatively large error of about 0.30.30.30.3. When α=0.4𝛼0.4\alpha=0.4italic_α = 0.4 all networks display error below 0.10.10.10.1, where most of them have extremely low errors, e.g., below 1010.superscript101010^{-10}.10 start_POSTSUPERSCRIPT - 10 end_POSTSUPERSCRIPT .

Refer to caption Refer to caption
Figure 8: (left) Plot of the relative error in the approximation of Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT by K~αsubscript~𝐾𝛼\tilde{K}_{\alpha}over~ start_ARG italic_K end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT. (right) Plot of the values of α𝛼\alphaitalic_α for which the relative error in the approximation drops below 0.10.10.10.1, αcsubscript𝛼𝑐\alpha_{c}italic_α start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, versus the relative spectral gap, together with the fitted regression line.

As in the derivation, the main driver for this approximation is the spectral gap of the adjacency matrix of the signed graph, i.e., λ1λ2subscript𝜆1subscript𝜆2\lambda_{1}-\lambda_{2}italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. Here we obtain the value of α𝛼\alphaitalic_α for which the relative error in the approximation of Kαsubscript𝐾𝛼K_{\alpha}italic_K start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT by K~αsubscript~𝐾𝛼\tilde{K}_{\alpha}over~ start_ARG italic_K end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT drops below 0.10.10.10.1 (for illustrative purposes), denoted by αcsubscript𝛼𝑐\alpha_{c}italic_α start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. In Fig. 8 (right), we plot the values of αcsubscript𝛼𝑐\alpha_{c}italic_α start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for every network as a function of the relative spectral gap (λ1λ2)/λ1subscript𝜆1subscript𝜆2subscript𝜆1\left(\lambda_{1}-\lambda_{2}\right)/\lambda_{1}( italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) / italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. We observe that increasing the spectral gap makes the approximation works better even for relatively large values of α𝛼\alphaitalic_α (Pearson correlation: 0.9420.9420.9420.942). In contrast, for those networks like the one of village D, where λ16.333subscript𝜆16.333\lambda_{1}\approx 6.333italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≈ 6.333 and λ26.258subscript𝜆26.258\lambda_{2}\approx 6.258italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ≈ 6.258, it is hard to converge even for relatively low values of α𝛼\alphaitalic_α. However, the trend holds where the relative error of this approximation is significantly lower for values of α𝛼\alphaitalic_α relatively low than that for the value of α=1𝛼1\alpha=1italic_α = 1, where the balance index corresponds to the exponential.

Acknowledgments

We would like to acknowledge Dr. H. Saiz and Prof. C. Altafini for sharing datasets used in this work. E.E. acknowledges support from the Maria de Maeztu project CEX2021-001164-M funded by the MCIN/AEI/10.13039/501100011033. Y.T. is funded by the Wallenberg Initiative on Networks and Quantum Information (WINQ).

References

  • [1] A. Alfakih. Euclidean distance matrices and their applications in rigidity theory. Springer, 2018.
  • [2] C. Altafini. Dynamics of opinion forming in structurally balanced social networks. PLoS ONE, 7(6):1–9, 2012.
  • [3] C. Altafini. Consensus problems on networks with antagonistic interactions. IEEE Transactions on Automatic Control, 58(4):935–946, 2013.
  • [4] Francesca Arrigo and Fabio Durastante. Mittag–leffler functions and their applications in network science. SIAM Journal on Matrix Analysis and Applications, 42(4):1581–1601, 2021.
  • [5] F. Atay and S. Liu. Cheeger constants, structural balance, and spectral clustering analysis for signed graphs. Discrete Math., 343(1):111616, 2020.
  • [6] R. Balaji and R. Bapat. On euclidean distance matrices. Linear Algebra Appl., 424(1):108–117, 2007.
  • [7] Michele Benzi and Paola Boito. Matrix functions in network analysis. GAMM-Mitteilungen, 43(3):e202000012, 2020.
  • [8] C. Berge. The theory of graphs. Courier Corporation, 2001.
  • [9] Dorwin Cartwright and Frank Harary. Structural balance: a generalization of heider’s theory. Psychological review, 63(5):277, 1956.
  • [10] F. Diaz-Diaz and E. Estrada. Signed graphs in data sciences via communicability geometry. arXiv preprint arXiv:2403.07493, 2024.
  • [11] E. Estrada. Conservative vs. non-conservative diusion towards a target in a networked environment. In The Target Problem. Springer, Berlin, 2024.
  • [12] Ernesto Estrada. The communicability distance in graphs. Linear Algebra and its Applications, 436(11):4317–4328, 2012.
  • [13] Ernesto Estrada. Rethinking structural balance in signed social networks. Discrete Applied Mathematics, 268:70–90, 2019.
  • [14] Ernesto Estrada. The many facets of the estrada indices of graphs and networks. SeMA Journal, 79(1):57–125, 2022.
  • [15] Ernesto Estrada. Communicability cosine distance: similarity and symmetry in graphs/networks. Computational and Applied Mathematics, 43(1):49, 2024.
  • [16] Ernesto Estrada and Michele Benzi. Walk-based measure of balance in signed networks: Detecting lack of balance in social networks. Physical Review E, 90(4):042802, 2014.
  • [17] Ernesto Estrada, Jesús Gómez-Gardeñes, and Lucas Lacasa. Network bypasses sustain complexity. Proceedings of the National Academy of Sciences, 120(31):e2305001120, 2023.
  • [18] Ernesto Estrada and Naomichi Hatano. Communicability in complex networks. Physical Review E, 77(3):036111, 2008.
  • [19] Ernesto Estrada and Naomichi Hatano. Communicability angle and the spatial efficiency of networks. SIAM Review, 58(4):692–715, 2016.
  • [20] Ernesto Estrada, Naomichi Hatano, and Michele Benzi. The physics of communicability in complex networks. Physics reports, 514(3):89–119, 2012.
  • [21] Ernesto Estrada and Desmond J Higham. Network properties revealed through matrix functions. SIAM review, 52(4):696–714, 2010.
  • [22] Ernesto Estrada and Juan A Rodriguez-Velazquez. Subgraph centrality in complex networks. Physical Review E, 71(5):056103, 2005.
  • [23] Ernesto Estrada, MG Sanchez-Lirola, and José Antonio De La Peña. Hyperspherical embedding of graphs and networks in communicability spaces. Discrete Applied Mathematics, 176:53–77, 2014.
  • [24] G. Facchetti, G. Iacono, and C. Altafini. Computing global structural balance in large-scale signed social networks. Proc. Natl. Acad. Sci., 108(52):20953–20958, 2011.
  • [25] L. Festinger. The analysis of sociograms using matrix algebra. Hum. Relat., 2(2):153–158, 1949.
  • [26] Rumi Ghosh, Kristina Lerman, Tawan Surachawala, Konstatin Voevodski, and Shanghua Teng. Non-conservative diffusion and its application to social network analysis. Journal of Complex Networks, 12(1):cnae006, 2024.
  • [27] P. Giscard, P. Rochet, and R. Wilson. Evaluating balance on social networks from their simple cycles. J. Complex Netw., 5(5):750–775, 05 2017.
  • [28] J. Gower. Properties of euclidean and non-euclidean distance matrices. Linear Algebra Appl., 67:81–97, 1985.
  • [29] F. Harary. On the notion of balance of a signed graph. Michigan Math. J., 2(2):143–146, 1953.
  • [30] F. Heider. Attitudes and cognitive organization. J. Psychol., 21(1):107–112, 1946.
  • [31] Nicholas J Higham. Functions of matrices: theory and computation. SIAM, 2008.
  • [32] Alexander Isakov, James H Fowler, Edoardo M Airoldi, and Nicholas A Christakis. The structure of negative social ties in rural village networks. Sociological science, 6:197, 2019.
  • [33] Leo Katz. A new status index derived from sociometric analysis. Psychometrika, 18(1):39–43, 1953.
  • [34] A. Kirkley, G. Cantwell, and M. Newman. Balance in signed networks. Phys. Rev. E, 99:012320, Jan 2019.
  • [35] N. Krislock and H. Wolkowicz. Euclidean distance matrices and applications. Springer, 2012.
  • [36] Chul-Ho Lee, Srinivas Tenneti, and Do Young Eun. Transient dynamics of epidemic spreading and its mitigation on large networks. In Proceedings of the twentieth ACM international symposium on mobile ad hoc networking and computing, pages 191–200, 2019.
  • [37] Kristina Lerman and Rumi Ghosh. Network structure, topology, and dynamics in generalized models of synchronization. Physical Review E, 86(2):026108, 2012.
  • [38] Andrés Martín and Ernesto Estrada. Fractional-modified bessel function of the first kind of integer order. Mathematics, 11(7):1630, 2023.
  • [39] A. Mathai and T. Zalavsky. On adjacency matrices and descriptors of signed cycle graphs. J. Comb. Inf. Syst. Sci., 37(2-4):369–382, 2012.
  • [40] Zaid Odibat. Approximations of fractional integrals and caputo fractional derivatives. Applied Mathematics and Computation, 178(2):527–533, 2006.
  • [41] H Polard. The completely monotonic character of the mittag-leffler function. Bull. Am. Math. Soc, 52:908–910, 1948.
  • [42] Amir Sadeghi and João R Cardoso. Some notes on properties of the matrix mittag-leffler function. Applied Mathematics and Computation, 338:733–738, 2018.
  • [43] Hugo Saiz, Jesús Gómez-Gardeñes, Paloma Nuche, Andrea Girón, Yolanda Pueyo, and Concepción L Alados. Evidence of structural balance in spatial ecological networks. Ecography, 40(6):733–741, 2017.
  • [44] Caio Seguin, Martijn P Van Den Heuvel, and Andrew Zalesky. Navigation of brain networks. Proceedings of the National Academy of Sciences, 115(24):6297–6302, 2018.
  • [45] R. Singh and B. Adhikari. Measuring the balance of signed networks and its application to sign prediction. J. Stat. Mech. Theory Exp., 2017(6):063302, jun 2017.
  • [46] Nicola Soranzo, Fahimeh Ramezani, Giovanni Iacono, and Claudio Altafini. Decompositions of large-scale biological systems based on dynamical properties. Bioinformatics, 28(1):76–83, 2012.
  • [47] P Tarazaga, T. Hayden, and J. Wells. Circum-euclidean distance matrices and faces. Linear Algebra Appl., 232:77–96, 1996.
  • [48] Y. Tian. Role Extraction, Dynamics, and Optimisation on Networks. PhD thesis, University of Oxford, October 2022. Available at https://ora.ox.ac.uk/objects/uuid:8145297d-3f88-4d67-9c34-575beb1a4c6c.
  • [49] Y. Tian and R. Lambiotte. Spreading and structural balance on signed networks. SIAM J. Appl. Dyn. Syst., 23(1):50–80, 2024.
  • [50] Y. Tian, S. Lautz, A. Wallis, and R. Lambiotte. Extracting complements and substitutes from sales data: a network perspective. EPJ Data Sci., 10(1):45, 2021.
  • [51] Thomas Zaslavsky. Signed graphs. Discrete Applied Mathematics, 4(1):47–74, 1982.
  • [52] Thomas Zaslavsky. Six signed petersen graphs, and their automorphisms. Discrete Mathematics, 312(9):1558–1583, 2012.

Appendix A Tables

We give more details of the results from the signed Petersen graphs in Tables 2 and 3.

graph tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT C5superscriptsubscript𝐶5C_{5}^{-}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT C6superscriptsubscript𝐶6C_{6}^{-}italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT K(G)𝐾𝐺K\left(G\right)italic_K ( italic_G )
a 48 4 4 0.968
b 24 6 6 0.951
c 22 8 4 0.941
d 14 6 10 0.947
e 11 12 0 0.919
Table 2: Values of the time of consensus tcsubscript𝑡𝑐t_{c}italic_t start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT using Altafini’s model [3] for the graphs in Fig. 1 as well as the values of the number of negative cycles of length 5 C5superscriptsubscript𝐶5C_{5}^{-}italic_C start_POSTSUBSCRIPT 5 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT and 6 C6superscriptsubscript𝐶6C_{6}^{-}italic_C start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT in those graphs. The values of the balance degree index K(G)𝐾𝐺K\left(G\right)italic_K ( italic_G ) are also given for each graph.
graph c graph d
n𝑛nitalic_n Cn+superscriptsubscript𝐶𝑛C_{n}^{+}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT Cnsuperscriptsubscript𝐶𝑛C_{n}^{-}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT Cn+superscriptsubscript𝐶𝑛C_{n}^{+}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT Cnsuperscriptsubscript𝐶𝑛C_{n}^{-}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT
5 4 8 6 6
6 6 4 0 10
8 7 8 15 0
9 12 8 10 10
Table 3: Values of the number of positive Cn+superscriptsubscript𝐶𝑛C_{n}^{+}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and negative Cnsuperscriptsubscript𝐶𝑛C_{n}^{-}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT cycles of length 5n85𝑛85\leq n\leq 85 ≤ italic_n ≤ 8 for the signed Petersen graphs c) and d) of Fig. 1.

We summarise the exact numbers of positive and negative cycles in the gene regulatory networks in Table 4.

S. cereviciae B. subtillus
n𝑛nitalic_n Cn+superscriptsubscript𝐶𝑛C_{n}^{+}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT Cnsuperscriptsubscript𝐶𝑛C_{n}^{-}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT Cn+superscriptsubscript𝐶𝑛C_{n}^{+}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT Cnsuperscriptsubscript𝐶𝑛C_{n}^{-}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT
3 35 27 129 75
4 1174 114 1366 547
5 152 135 1780 1031
6 3855 763 8003 4975
7 1875 1119 20645 14261
8 34321 13332 72722 48309
9 29473 16740 179137 122709
10 271954 128469 547246 364806
11 476800 279258 1443443 1002857
Table 4: Values of the number of positive Cn+superscriptsubscript𝐶𝑛C_{n}^{+}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT and negative Cnsuperscriptsubscript𝐶𝑛C_{n}^{-}italic_C start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT cycles of length 3n113𝑛113\leq n\leq 113 ≤ italic_n ≤ 11 for the gene regulatory networks of yeast and B. subtilis.