Social clustering reinforces external influence on the majority opinion model

Niels Van Santen Jan Ryckebusch Luis E. C. Rocha
Abstract

Public opinion is subject to peer interaction via social networks and external pressures from the media, advertising, and other actors. In this paper, we study the interaction between external and peer influence on the stochastic opinion dynamics of a majority vote model. We introduce a model where agents update their opinions based on the combined influence from their local neighbourhood (peers) and from an external actor in the transition rates. In the first model, the external influence is only felt by agents non-aligned with the external actor (”push strategy”). In the second model, agents are affected by external influence, independently of their opinions (”nudging strategy”). In both cases, the external influence increases the possible macroscopic outcomes. These outcomes are determined by the chosen strategy. We also find that the social network structure affects the opinion dynamics, with high social clustering positively reinforcing the external influence while degree heterogeneity weakens it. These findings are relevant to businesses and policy making, hel** to understand how groups of individuals collectively react to external actors.

keywords:
opinion dynamics , complex networks , complex contagion , external influence , peer influence
journal: Physica A: Statistical Mechanics and its Applications
\affiliation

[inst1]organization=Department of Data-Analysis, Ghent University,addressline=Henri Dunantlaan 1, city=Ghent, postcode=9000, country=Belgium

\affiliation

[inst2]organization=Department of Physics and Astronomy, Ghent University,addressline=Proeftuinstraat 86, city=Ghent, postcode=9000, country=Belgium

\affiliation

[inst3]organization=Department of Economics, Ghent University,addressline=Sint-Pietersplein 5, city=Ghent, postcode=9000, country=Belgium

{highlights}

The interaction between peer pressure and an external field leads to more complex behaviour than reinforcing the external opinion.

The intensity and weight of an external field push the population into a greater variation of macroscopic states.

Macroscopic outcomes depend on the alignment of agents to the external field.

Social clustering reinforces the influence of an external field.

Degree heterogeneity and hubs hinder the effect of an external field.

1 Introduction

Statistical physics offers helpful tools to study social systems involving collective and emergent phenomena [1, 2, 3, 4, 5]. In particular, physical models have been used to explore and understand the collective behaviour of opinions in society [6, 7]. Supporting the idea that human behaviour is not wholly unpredictable [8]. Given the universality in order-disorder transitions, it is appropriate to use physics-based models to investigate large-scale regularities as collective effects. Models composed of adaptive agents have been effective at describing complex social behaviour [9, 10, 11]. The rules and constraints of those models, for example, threshold imitation or homophily [12], should be informed by psychology and social science. Opinion dynamics is a multidisciplinary field built on early sociological observations that people tend to conform to each other’s behaviours and opinions. This led to questions about collective behaviour and inspired early models in sociology and economics, such as the Schelling model [13] and Axelrod’s cultural dissemination model [14]. Behaviour and opinions can only spread when there is interaction, implying the importance of the structure of the underlying contact networks [15].

In public opinion formation, next to interpersonal social interaction, also mass media and other external actors play a role [16, 17, 18, 19, 20]; thus opinion dynamics is subject to the interaction of personal predispositions, external media influence and social interaction [21]. We are interested in exploring the interaction between the internal peer pressure towards conforming, and an external influence. In the last decades, many models for opinion dynamics have been created for discrete and continuous opinion spaces or a combination of both [22]. Voting problems have served as early applications. Examples include the well-known Voter Model [23, 24] and its many variations [25]. We focus here on binary opinion models, particularly complex contagions [26], where opinions require reinforcement to become more easily adapted. This idea of reinforcement has been implemented in different ways. For example, by interacting with more than one neighbour [27, 28, 29, 30, 31], via the use of memory [32], by introducing different levels of confidence in the opinion [33], or by repeated interactions [34] and algorithmic personalisation  [35, 36]. We will follow the first approach and have individuals take their entire social neighbourhood into account when updating their opinions.

Several investigations have been conducted to include an external influence, such as the presence of media, in opinion dynamics models. In the binary opinion case, for the Voter Model and some of its extensions [37, 38, 39], the Majority Rule model [40] and the Sznajd Model [41]. Often, at every time step, there is a probability r𝑟ritalic_r with which an agent interacts with its environment, feeling only internal pressure, and a probability 1r1𝑟1-r1 - italic_r they only feel the external media influence in that time step. The model we introduce in this paper makes a different assumption that agents feel both internal and external pressure at every time step. This allows us to understand the effect of opposing and reinforcing interactions simultaneously; an agent does not feel external influence independently from their neighbourhood pressure. This paper also addresses the impact of external influence applied to complex contagion. It proposes asymmetric scenarios where the external influence is felt depending on the (opinion) alignment of the agent with the media. The concurrent influence on complex contagions and the comparison between homogeneous and alignment-based influence have been subject to little investigation in the literature [7].

This paper introduces an opinion dynamics model containing the following properties: (i) complex contagion spread, (ii) concurrent interaction between internal and external pressure, (iii) social network structure, and (iv) stochasticity. The latter is chosen to resemble real-world noise and incomplete information during decision-making. To capture the above properties, we propose a model where, at each time step, a randomly chosen agent updates their opinion with a certain transition rate that results from the interaction of internal and external influence. Internal influence contributes to the transition rate according to the fraction of nearest neighbours disagreeing with the chosen agent. The external influence adds or subtracts from that transition rate based on the disagreement or agreement between the agent and the media, respectively. The media influence will have a strength fixed in time, and the relative weight for internal and external pressures is set by a parameter α𝛼\alphaitalic_α. We test two versions of this model: an asymmetric version where the media only influences agents that disagree with it and a symmetric version where every agent, independently of its own opinion, feels the external media influence such that the media adjusts transition rates in both directions. Both versions enrich the space of possible consensus regimes over the absence of external influence, with the asymmetric version showing the most variation concerning possible final consensus states.

2 Model

2.1 Networks

A network comprises a set of N𝑁Nitalic_N nodes i𝑖iitalic_i connected by M𝑀Mitalic_M links (i,j)𝑖𝑗(i,j)( italic_i , italic_j ), where we exclude self-links (i,i)𝑖𝑖(i,i)( italic_i , italic_i ) and multiple links between the same node pair. We will restrict ourselves to undirected networks; a link (i,j)𝑖𝑗(i,j)( italic_i , italic_j ) implies a symmetric link (j,i)𝑗𝑖(j,i)( italic_j , italic_i ). The topology is characterised by a symmetric matrix 𝑨𝑨\boldsymbol{A}bold_italic_A, the adjacency matrix, with entries aij=1subscript𝑎𝑖𝑗1a_{ij}=1italic_a start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = 1 if a link exists between node i𝑖iitalic_i and j𝑗jitalic_j, and zero otherwise. Each node i𝑖iitalic_i has a degree ki=jaijsubscript𝑘𝑖subscript𝑗subscript𝑎𝑖𝑗k_{i}=\sum_{j}a_{ij}italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT italic_a start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT representing the number of other nodes j𝑗jitalic_j to which i𝑖iitalic_i is connected. All networks used in this work are connected, i.e. ki>0,isubscript𝑘𝑖0for-all𝑖k_{i}>0,\;\forall iitalic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT > 0 , ∀ italic_i, and have N=1000𝑁1000N=1000italic_N = 1000 nodes with k10delimited-⟨⟩𝑘10\langle k\rangle\approx 10⟨ italic_k ⟩ ≈ 10. The average shortest path length is given by l=1N(N1)ijlijdelimited-⟨⟩𝑙1𝑁𝑁1subscript𝑖𝑗subscript𝑙𝑖𝑗\langle l\rangle=\frac{1}{N(N-1)}\sum_{i\neq j}l_{ij}⟨ italic_l ⟩ = divide start_ARG 1 end_ARG start_ARG italic_N ( italic_N - 1 ) end_ARG ∑ start_POSTSUBSCRIPT italic_i ≠ italic_j end_POSTSUBSCRIPT italic_l start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT, where lijsubscript𝑙𝑖𝑗l_{ij}italic_l start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT is the shortest-path between nodes i𝑖iitalic_i and j𝑗jitalic_j, and the number of local bridges μ𝜇\muitalic_μ are the links whose endpoints have no common neighbours. The average clustering coefficient is cc=1Ni2Miki(ki1)delimited-⟨⟩𝑐𝑐1𝑁subscript𝑖2subscript𝑀𝑖subscript𝑘𝑖subscript𝑘𝑖1\langle cc\rangle=\frac{1}{N}\sum_{i}\frac{2M_{i}}{k_{i}(k_{i}-1)}⟨ italic_c italic_c ⟩ = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG 2 italic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - 1 ) end_ARG, where Misubscript𝑀𝑖M_{i}italic_M start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is the number of links between common neighbours or i𝑖iitalic_i, and the betweenness centrality of node k𝑘kitalic_k is Bk=ijσ(i,k,j)σ(i,j)subscript𝐵𝑘subscript𝑖𝑗𝜎𝑖𝑘𝑗𝜎𝑖𝑗B_{k}=\sum_{ij}\frac{\sigma(i,k,j)}{\sigma(i,j)}italic_B start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT = ∑ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT divide start_ARG italic_σ ( italic_i , italic_k , italic_j ) end_ARG start_ARG italic_σ ( italic_i , italic_j ) end_ARG, where σ(i,j)𝜎𝑖𝑗\sigma(i,j)italic_σ ( italic_i , italic_j ) is the number of shortest-paths between i𝑖iitalic_i and j𝑗jitalic_j, and σ(i,k,j)𝜎𝑖𝑘𝑗\sigma(i,k,j)italic_σ ( italic_i , italic_k , italic_j ) is the number of shortest-paths between i𝑖iitalic_i and j𝑗jitalic_j passing through node k𝑘kitalic_k.

We will use three theoretical network models to study the opinion dynamics and investigate the sensitivity of our results on the different structures observed in real-world social networks. We are particularly interested in studying short path length, the combination of short path length and high clustering, and the presence of high-degree nodes (i.e. hubs). The first model is the Erdős-Rényi (ER) model, where nodes are randomly connected with probability p𝑝pitalic_p. This model exhibits short path lengths ddelimited-⟨⟩𝑑\langle d\rangle⟨ italic_d ⟩ and low (local) clustering ccdelimited-⟨⟩𝑐𝑐\langle cc\rangle⟨ italic_c italic_c ⟩. The second model is the Watts-Strogatz (WS) model, which combines short path lengths with high local clustering. It is constructed from a regular lattice by rewiring edges with a probability q𝑞qitalic_q. According to this probability, the WS model leads to structures intermediate between the regular lattice (q=0𝑞0q=0italic_q = 0) with high clustering and long path lengths and the ER model (q=1𝑞1q=1italic_q = 1) with low clustering and short path lengths. Finally, the Barabási-Albert (BA) preferential attachment growth model, where nodes are added to the network and connected preferentially to existing nodes with higher degrees, is used to obtain a structure with heterogeneous degree distribution. Table 1 shows the mean values of relevant network measures characterising those networks.

ER WS BA
kdelimited-⟨⟩𝑘\langle k\rangle⟨ italic_k ⟩ 9.926 ±plus-or-minus\pm± 0.077 10 ±plus-or-minus\pm± 0 9.95 ±plus-or-minus\pm± 0
ccdelimited-⟨⟩𝑐𝑐\langle cc\rangle⟨ italic_c italic_c ⟩ 0.010 ±plus-or-minus\pm± 0.001 0.089 ±plus-or-minus\pm± 0.003 0.040 ±plus-or-minus\pm± 0.003
ddelimited-⟨⟩𝑑\langle d\rangle⟨ italic_d ⟩ 3.267 ±plus-or-minus\pm± 0.009 3.382 ±plus-or-minus\pm± 0.004 2.977 ±plus-or-minus\pm± 0.012
μ/Mdelimited-⟨⟩𝜇𝑀\langle\mu\rangle/M⟨ italic_μ ⟩ / italic_M 0.909 ±plus-or-minus\pm± 0.009 0.558 ±plus-or-minus\pm± 0.007 0.685 ±plus-or-minus\pm± 0.014
Bkdelimited-⟨⟩subscript𝐵𝑘\langle B_{k}\rangle⟨ italic_B start_POSTSUBSCRIPT italic_k end_POSTSUBSCRIPT ⟩ 0.002 ±plus-or-minus\pm± 0.001 0.002 ±plus-or-minus\pm± 0.001 0.002 ±plus-or-minus\pm± 0.007
Table 1: The mean and the standard deviation (±plus-or-minus\pm±) of relevant network measures over ten realisations of the ER, WS, and BA network models with N=1000𝑁1000N=1000italic_N = 1000 nodes.

2.2 Opinion Dynamics

We devise a model with N𝑁Nitalic_N agents (nodes) i=0,1,,N1𝑖01𝑁1i=0,1,\ldots,N-1italic_i = 0 , 1 , … , italic_N - 1. At time step t, each agent has a binary opinion oi(t)𝒪={A,B}subscript𝑜𝑖𝑡𝒪𝐴𝐵o_{i}(t)\in\mathcal{O}=\{A,B\}italic_o start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) ∈ caligraphic_O = { italic_A , italic_B }. We assume that the opinion of agent i𝑖iitalic_i can change as a result of two competing mechanisms: (i) pressure from its social contacts ν(i)𝜈𝑖\nu(i)italic_ν ( italic_i ), i.e. the ratio of neighbours j𝑗jitalic_j of i𝑖iitalic_i having opinion oj(t)oi(t)subscript𝑜𝑗𝑡subscript𝑜𝑖𝑡o_{j}(t)\neq o_{i}(t)italic_o start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) ≠ italic_o start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) (Eq. 1) and (ii) an external fixed source promoting opinion A𝐴Aitalic_A and affecting all agents equally A(B)[0,1]superscript𝐴𝐵01\mathcal{E}^{A(B)}\in[0,1]\subset\mathbb{R}caligraphic_E start_POSTSUPERSCRIPT italic_A ( italic_B ) end_POSTSUPERSCRIPT ∈ [ 0 , 1 ] ⊂ blackboard_R, i.e. an external field representing, for example, mass media, institutions, or nudging.

i(t)=<ij>oj(t)oi(t)kioj(t).subscript𝑖𝑡superscriptsubscriptexpectation𝑖𝑗subscript𝑜𝑗𝑡subscript𝑜𝑖𝑡subscript𝑘𝑖subscript𝑜𝑗𝑡\mathcal{I}_{i}(t)=\frac{\sum_{<ij>}^{o_{j}(t)\neq o_{i}(t)}}{k_{i}}o_{j}(t).caligraphic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) = divide start_ARG ∑ start_POSTSUBSCRIPT < italic_i italic_j > end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_o start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) ≠ italic_o start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) end_POSTSUPERSCRIPT end_ARG start_ARG italic_k start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG italic_o start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( italic_t ) . (1)

Agents with opinion oi(t)=A(B)subscript𝑜𝑖𝑡𝐴𝐵o_{i}(t)=A(B)italic_o start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) = italic_A ( italic_B ) have a transition probability towards opinion B(A)𝐵𝐴B(A)italic_B ( italic_A ) at time t𝑡titalic_t given by WiA(B)B(A)(t)superscriptsubscript𝑊𝑖𝐴𝐵𝐵𝐴𝑡W_{i}^{A(B)\rightarrow B(A)}(t)italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A ( italic_B ) → italic_B ( italic_A ) end_POSTSUPERSCRIPT ( italic_t ). The relative importance of the external field concerning social pressure is regulated via the parameter 0α10𝛼10\leq\alpha\leq 10 ≤ italic_α ≤ 1, reflecting the trust agents have in the external opinion. Larger α𝛼\alphaitalic_α means that agents give more weight to the external field than to pressure from social contacts. We consider two stochastic versions of the model, differing in how agents are subject to the pressure of the external field.

In the first model (hereafter called asymmetric model), the external field only influences agents with an opinion opposite to the field, i.e. agents agreeing with the external field (defined to be opinion A𝐴Aitalic_A) are only influenced by their social contacts (there is no reinforcement). Equations 2 give the transition rates for agents with opinions A𝐴Aitalic_A and B𝐵Bitalic_B. In this case, the external field can promote the transition from BA𝐵𝐴B\rightarrow Aitalic_B → italic_A but not vice-versa. This corresponds to a scenario where the external field acts towards one of the opinions, e.g. an employee has to decide on adopting a method or software and is subject to both “outside” influence (employer suggests using software A𝐴Aitalic_A) and “internal” peer pressure (colleagues recommend A𝐴Aitalic_A or B𝐵Bitalic_B). An employee who uses the default option A𝐴Aitalic_A, as set by the company, does not feel pressure from the employer but only from colleagues using B𝐵Bitalic_B. Meanwhile, when using B𝐵Bitalic_B, the employee feels the influence of colleagues and the employer. In the limiting cases, the model reduces to complex contagion for α=0𝛼0\alpha=0italic_α = 0 whereas for α=1𝛼1\alpha=1italic_α = 1, transitions from BA𝐵𝐴B\rightarrow Aitalic_B → italic_A only depend on the external field. We could also model the case of outside pressure affecting only those with the same opinion as the external field, but that would lead to confirmation bias and, thus, simple reinforcement.

WiAB(t)=iB(t)WiBA(t)=(1α)iA(t)+αA.superscriptsubscript𝑊𝑖𝐴𝐵𝑡superscriptsubscript𝑖𝐵𝑡superscriptsubscript𝑊𝑖𝐵𝐴𝑡1𝛼superscriptsubscript𝑖𝐴𝑡𝛼superscript𝐴\begin{split}W_{i}^{A\rightarrow B}(t)&=\mathcal{I}_{i}^{B}(t)\\ W_{i}^{B\rightarrow A}(t)&=(1-\alpha)\mathcal{I}_{i}^{A}(t)+\alpha\mathcal{E}^% {A}.\end{split}start_ROW start_CELL italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A → italic_B end_POSTSUPERSCRIPT ( italic_t ) end_CELL start_CELL = caligraphic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_t ) end_CELL end_ROW start_ROW start_CELL italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B → italic_A end_POSTSUPERSCRIPT ( italic_t ) end_CELL start_CELL = ( 1 - italic_α ) caligraphic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_t ) + italic_α caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT . end_CELL end_ROW (2)

In the second model (hereafter called symmetric model), all agents may be influenced by the external field (Eqs. 3). This version will not only facilitate that agents with opposing opinions switch to the external opinion (i.e. through nudging opinion A𝐴Aitalic_A by controlling α𝛼\alphaitalic_α) but will also hinder agents who agree with the external field from changing their opinion (i.e. reinforcing opinion A𝐴Aitalic_A to counter-balance social pressure).

WiAB(t)=max[0,(1α)B(t)αA]WiBA(t)=(1α)iA(t)+αAsuperscriptsubscript𝑊𝑖𝐴𝐵𝑡max01𝛼superscript𝐵𝑡𝛼superscript𝐴superscriptsubscript𝑊𝑖𝐵𝐴𝑡1𝛼superscriptsubscript𝑖𝐴𝑡𝛼superscript𝐴\begin{split}W_{i}^{A\rightarrow B}(t)&=\text{max}[0,(1-\alpha)\mathcal{I}^{B}% (t)-\alpha\mathcal{E}^{A}]\\ W_{i}^{B\rightarrow A}(t)&=(1-\alpha)\mathcal{I}_{i}^{A}(t)+\alpha\mathcal{E}^% {A}\end{split}start_ROW start_CELL italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A → italic_B end_POSTSUPERSCRIPT ( italic_t ) end_CELL start_CELL = max [ 0 , ( 1 - italic_α ) caligraphic_I start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT ( italic_t ) - italic_α caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ] end_CELL end_ROW start_ROW start_CELL italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B → italic_A end_POSTSUPERSCRIPT ( italic_t ) end_CELL start_CELL = ( 1 - italic_α ) caligraphic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ( italic_t ) + italic_α caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT end_CELL end_ROW (3)

These two versions of the model are simulated computationally using agent-based modelling. All simulations start with an initial configuration where each agent i𝑖iitalic_i has an opinion A𝐴Aitalic_A or B𝐵Bitalic_B chosen uniformly. At each time step t𝑡titalic_t, an agent i𝑖iitalic_i is chosen uniformly. The fraction of disagreeing social contacts (Eq. 1) is calculated and used to determine the opinion flip rate Wi(t)subscript𝑊𝑖𝑡W_{i}(t)italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ), based on social pressure, the external field, and α𝛼\alphaitalic_α. Agent i𝑖iitalic_i changes its opinion with probability Wi(t)subscript𝑊𝑖𝑡W_{i}(t)italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) and keeps it with probability 1Wi(t)1subscript𝑊𝑖𝑡1-W_{i}(t)1 - italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ). This procedure is repeated for T𝑇Titalic_T time steps (Fig. 1), chosen long enough for the system to reach the stationary state (TER40×104subscript𝑇𝐸𝑅40superscript104T_{ER}\approx 40\times 10^{4}italic_T start_POSTSUBSCRIPT italic_E italic_R end_POSTSUBSCRIPT ≈ 40 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT; TWS80×104subscript𝑇𝑊𝑆80superscript104T_{WS}\approx 80\times 10^{4}italic_T start_POSTSUBSCRIPT italic_W italic_S end_POSTSUBSCRIPT ≈ 80 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT; TBA60×104subscript𝑇𝐵𝐴60superscript104T_{BA}\approx 60\times 10^{4}italic_T start_POSTSUBSCRIPT italic_B italic_A end_POSTSUBSCRIPT ≈ 60 × 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT). Because time-dependent measures only depend on other time-dependent measures at the same time step, the explicit time dependence will be omitted for the sake of brevity.

Refer to caption
Figure 1: An example of the transition rates with external influence A=0.7superscript𝐴0.7\mathcal{E}^{A}=0.7caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 0.7, where iB=0.6superscriptsubscript𝑖𝐵0.6\mathcal{I}_{i}^{B}=0.6caligraphic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT = 0.6 (iA=0.4superscriptsubscript𝑖𝐴0.4\mathcal{I}_{i}^{A}=0.4caligraphic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 0.4) and α=0.5𝛼0.5\alpha=0.5italic_α = 0.5.

We use different measures to quantify the consensus formation in the opinion dynamics model across scales [2]. The opinions oi=Asubscript𝑜𝑖𝐴o_{i}=Aitalic_o start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_A and oi=Bsubscript𝑜𝑖𝐵o_{i}=Bitalic_o start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = italic_B are modeled by spin variables σi=+1subscript𝜎𝑖1\sigma_{i}=+1italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = + 1 and σi=1subscript𝜎𝑖1\sigma_{i}=-1italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = - 1. A magnetization-like measure is used to quantify the level of global consensus m=1Niσi𝑚1𝑁subscript𝑖subscript𝜎𝑖m=\frac{1}{N}\sum_{i}\sigma_{i}italic_m = divide start_ARG 1 end_ARG start_ARG italic_N end_ARG ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT, which is the fraction of the total population holding an opinion (A𝐴Aitalic_A or B𝐵Bitalic_B). The values lie in the interval [1,1]11[-1,1][ - 1 , 1 ], where the extremes m=1𝑚1m=1italic_m = 1 and m=1𝑚1m=-1italic_m = - 1 resemble the total population holding opinion A𝐴Aitalic_A and B𝐵Bitalic_B (full consensus), respectively.

The local consensus is measured using the so-called density of interfaces (Eq. 4), where Npsubscript𝑁𝑝N_{p}italic_N start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the total number of nearest neighbour pairs (i.e. ṡocial contacts (i,j)𝑖𝑗(i,j)( italic_i , italic_j )) and Nasubscript𝑁𝑎N_{a}italic_N start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT the number of such pairs with neighbours having different opinions (i.e. oiojsubscript𝑜𝑖subscript𝑜𝑗o_{i}\neq o_{j}italic_o start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≠ italic_o start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT for (i,j)𝑖𝑗(i,j)( italic_i , italic_j )). The case na=1/2subscript𝑛𝑎12n_{a}=1/2italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = 1 / 2 corresponds to the fully mixed state, and na=0subscript𝑛𝑎0n_{a}=0italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = 0 indicates complete order.

na=NaNp,subscript𝑛𝑎subscript𝑁𝑎subscript𝑁𝑝n_{a}=\frac{N_{a}}{N_{p}},italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT = divide start_ARG italic_N start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG , (4)

The two measures are redundant in the case of full consensus but not in a mixed state. In the latter, nasubscript𝑛𝑎n_{a}italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT can be used to gain additional information about how the opinions are locally ordered, i.e. to measure opinion clusters. These measures provide no information about how the population relates to the external opinion. An energy-like measure is used to measure the combined social stress in the population due to local disagreement and disagreement with the opinion of the external field (Eq. 5). This quantity is not conserved and thus will not be referred to as energy but social stress SEsubscript𝑆𝐸S_{E}italic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT.

SE=(1α)<ij>σiσjαiσi.subscript𝑆𝐸1𝛼subscriptexpectation𝑖𝑗subscript𝜎𝑖subscript𝜎𝑗𝛼subscript𝑖subscript𝜎𝑖S_{E}=-(1-\alpha)\sum_{<ij>}\sigma_{i}\sigma_{j}-\alpha\mathcal{E}\sum_{i}% \sigma_{i}.italic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT = - ( 1 - italic_α ) ∑ start_POSTSUBSCRIPT < italic_i italic_j > end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT - italic_α caligraphic_E ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT . (5)

Social stress increases when two contacts disagree, and agents disagree with the external field opinion. An algorithm based on breadth-first search [42] is used to count distinct opinion clusters in the network. An opinion cluster (ΩosubscriptΩ𝑜\Omega_{o}roman_Ω start_POSTSUBSCRIPT italic_o end_POSTSUBSCRIPT) is thus here defined as a group of agents with the same opinion oisubscript𝑜𝑖o_{i}italic_o start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT for which there exists a path between any two agents in the group without having to go through an agent of the opposite opinion (Fig. 2).

Refer to caption
Figure 2: An example network highlighting the different opinion clusters. Six distinct opinion clusters are identified, two for opinion “red” and four for opinion “green”.

For a given network model and set of parameters, averages are computed using 100 points (10 realizations for each of the ten starting configurations).

3 Results

3.1 Consensus Formation

We start by studying consensus formation and how the external influence affects the stationary population opinion. Without loss of generality, we assign opinion A𝐴Aitalic_A to the external field. We study the consensus formation by analysing how the strength of the external field \mathcal{E}caligraphic_E affects the prevalence of opinion PAsubscript𝑃𝐴P_{A}italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT for a fixed α𝛼\alphaitalic_α, which can be interpreted as the amount of trust placed in the external actor relative to one’s peers. At t=0𝑡0t=0italic_t = 0, both opinions are distributed uniformly at random over the population such that the prevalence of both opinions is approximately equal, PA(0)PB(0)0.5subscript𝑃𝐴0subscript𝑃𝐵00.5P_{A}(0)\approx P_{B}(0)\approx 0.5italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( 0 ) ≈ italic_P start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ( 0 ) ≈ 0.5. Figure 3 shows the temporal evolution of PA(t)subscript𝑃𝐴𝑡P_{A}(t)italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_t ) for different levels of the external field in both versions of the model.

Refer to caption
Figure 3: The prevalence PA(t)subscript𝑃𝐴𝑡P_{A}(t)italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ( italic_t ) of opinion A𝐴Aitalic_A on the ER, WS, and BA network models (See Methods) for the (a) asymmetric model with A=0superscript𝐴0\mathcal{E}^{A}=0caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 0, (b) asymmetric model with A=0.5superscript𝐴0.5\mathcal{E}^{A}=0.5caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 0.5, (c) asymmetric model with A=1superscript𝐴1\mathcal{E}^{A}=1caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 1 (d) symmetric model with A=0superscript𝐴0\mathcal{E}^{A}=0caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 0, (e) symmetric model with A=0.5superscript𝐴0.5\mathcal{E}^{A}=0.5caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 0.5, and (f) symmetric model with A=1superscript𝐴1\mathcal{E}^{A}=1caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 1. The parameter α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 is fixed. The shaded area represents the standard error.

The dynamics lead to a stationary state where the fraction of agents following the external opinion A𝐴Aitalic_A plateaus below the maximum value, except for the WS network, where the dynamics need longer times to reach stationarity. The plateau values are higher for the symmetric than for the asymmetric version. For A=0.5superscript𝐴0.5\mathcal{E}^{A}=0.5caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 0.5 in the asymmetric model and A=0superscript𝐴0\mathcal{E}^{A}=0caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 0 in the symmetric model, the prevalence of the opinions remains balanced, except by the increasing prevalence observed for opinion A𝐴Aitalic_A in the WS network, likely due to reinforcement. The prevalence of opinion A𝐴Aitalic_A on the WS networks are relatively larger, whereas the prevalence is relatively lower for the BA networks. Generally, increasing Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT leads to increasing PAsubscript𝑃𝐴P_{A}italic_P start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT, but there is a difference between both models for low Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT. The asymmetric version allows for parameter combinations where the opposing opinion becomes most prevalent on average, whereas this is not possible with the symmetric version. These differences are more significant at the stationary state for various combinations of parameters (Figs. 4-5).

Refer to caption
Figure 4: Results of the asymmetric model for various parameter combinations. (a) global consensus m𝑚mitalic_m and ER model, (b) local consensus nasubscript𝑛𝑎n_{a}italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT and ER model, (c) social stress SEsubscript𝑆𝐸S_{E}italic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and ER model, (d) global consensus m𝑚mitalic_m and WS model, (e) local consensus nasubscript𝑛𝑎n_{a}italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT and WS model, (f) social stress SEsubscript𝑆𝐸S_{E}italic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and WS model, (g) global consensus m𝑚mitalic_m and BA model, (h) local consensus nasubscript𝑛𝑎n_{a}italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT and BA model, (i) social stress SEsubscript𝑆𝐸S_{E}italic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and BA model, for different values of the model parameters α𝛼\alphaitalic_α and Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT.

For the asymmetric model, the stationary state results show two regions of high (average) global consensus and order for all network structures. Agents move more towards global consensus when they are subjected to low or high Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT external field, with much trust in the external source (α>0.5𝛼0.5\alpha>0.5italic_α > 0.5), than when they are subject to moderate levels of external pressure (A0.5superscript𝐴0.5\mathcal{E}^{A}\approx 0.5caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT ≈ 0.5). The asymmetric model allows each opinion to become the majority opinion in the population, depending on α𝛼\alphaitalic_α-Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT combinations. The effect of strong consensus formation in the case of high trust in the external field but low field intensity is due to this model’s asymmetry in transition rates (Eq. 2). In this region with large α𝛼\alphaitalic_α, the asymmetric model is subject to two different influences: WiAB=iBsuperscriptsubscript𝑊𝑖𝐴𝐵superscriptsubscript𝑖𝐵W_{i}^{A\rightarrow B}=\mathcal{I}_{i}^{B}italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A → italic_B end_POSTSUPERSCRIPT = caligraphic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B end_POSTSUPERSCRIPT and WiBA=Asuperscriptsubscript𝑊𝑖𝐵𝐴superscript𝐴W_{i}^{B\rightarrow A}=\mathcal{E}^{A}italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B → italic_A end_POSTSUPERSCRIPT = caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT (eq. 2). Thus, the region of high α𝛼\alphaitalic_α and low Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT is driven by local peer pressure alone, while the region with large α𝛼\alphaitalic_α and Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT is driven by an interaction of peer and external pressure (Fig. 4). When the agents are only subject to social pressure (α=0𝛼0\alpha=0italic_α = 0), the population ends up in a mixed state. This is due to the averaging over multiple stochastic realisations where various levels of consensus are reached for both opinions. Another region with, on average, low global order is visible for values of α0𝛼0\alpha\neq 0italic_α ≠ 0 and medium Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT. The higher local disorder and social stress in this area compared to the α0𝛼0\alpha\approx 0italic_α ≈ 0 area could indicate that this average global disorder results from actual disordered states instead of averaging over more or less ordered states in both opinion directions. This area appears as a disordered transition region between going from one majority opinion to the other by increasing or decreasing the intensity of the external field.

Refer to caption
Figure 5: Results of various parameter combinations for the symmetric model. (a) global consensus m𝑚mitalic_m and ER model, (b) local consensus nasubscript𝑛𝑎n_{a}italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT and ER model, (c) social stress SEsubscript𝑆𝐸S_{E}italic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and ER model, (d) global consensus m𝑚mitalic_m and WS model, (e) local consensus nasubscript𝑛𝑎n_{a}italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT and WS model, (f) social stress SEsubscript𝑆𝐸S_{E}italic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and WS model, (g) global consensus m𝑚mitalic_m and BA model, (h) local consensus nasubscript𝑛𝑎n_{a}italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT and BA model, (i) social stress SEsubscript𝑆𝐸S_{E}italic_S start_POSTSUBSCRIPT italic_E end_POSTSUBSCRIPT and BA model, for different values of the model parameters α𝛼\alphaitalic_α and Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT.

In the symmetric model, the resistance to moving away from the opinion of the external field significantly reduces the possible final states, which are qualitatively similar for every network model (Fig. 5). All measures point to high levels of consensus and order in the large α𝛼\alphaitalic_α and large Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT upper half of the parameter space. The average fully mixed state at α=0𝛼0\alpha=0italic_α = 0 is still visible and appears for A=0superscript𝐴0\mathcal{E}^{A}=0caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 0, regardless of α𝛼\alphaitalic_α. For a fixed α>0𝛼0\alpha>0italic_α > 0 and A>0superscript𝐴0\mathcal{E}^{A}>0caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT > 0, there is generally a monotonous increase in consensus visible with increasing external field and trust, respectively. Thus, higher external pressure results in higher levels of consensus as well as a reduction in the variance around the stationary state (Fig. 3 (d - f)). Global consensus is now only possible towards the external field opinion, in contrast with the asymmetric model, where both forces can define consensus on average. For α>0𝛼0\alpha>0italic_α > 0 and >00\mathcal{E}>0caligraphic_E > 0, the global consensus and local ordering levels are significantly higher than in the asymmetric model, as reflected in the low levels of social stress attained. This restraining effect of the external field is present in all network structures; the difference can be attributed to how much the structure facilitates the spread of opinion. For the WS networks, high levels of consensus are reached for a larger region of the parameter space. In contrast, BA networks show signs of stronger resistance towards opinion spread, requiring a stronger external field. In the WS model, agents reinforce each other due to the redundancy in the social structure (i.e. high clustering), leading to values of isubscript𝑖\mathcal{I}_{i}caligraphic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT larger than 0.50.50.50.5. Therefore, even a small influence of the external field is amplified due to social reinforcement. On the other hand, the hubs (high-degree nodes) in the BA network act as a type of bottleneck. Hubs can affect a large part of the network, composed mainly by low degree nodes, kee** isubscript𝑖\mathcal{I}_{i}caligraphic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT close to 0.5. If a hub is chosen to update its opinion, the chance that all its neighbours agree on a specific opinion is relatively low because they are not exposed to reinforcement, leading to i0.5similar-tosubscript𝑖0.5\mathcal{I}_{i}\sim 0.5caligraphic_I start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ∼ 0.5 for hubs. The poor clustering of the BA network structure hinders the creation of a majority opinion and resists changing towards the external force.

3.2 Opinion Clusters

Figures 6-9 show the evolution of the number |Ω|Ω|\Omega|| roman_Ω | and average size Ωdelimited-⟨⟩Ω\langle\Omega\rangle⟨ roman_Ω ⟩ of the opinion clusters for different levels of the external field Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT. In the asymmetric version (Figs. 6-7), the majority of opinions B𝐵Bitalic_B and A𝐴Aitalic_A at respectively low and high Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT is reflected in the number and size of their respective clusters. When an opinion reaches a majority, its opinion clusters grow and decrease in number. In contrast, for an opinion that declines in its number of followers, the clusters become more numerous but smaller. As an opinion loses ground and agents acting as bridges in their cluster switch to the opposite opinion, the clusters fracture, and agents of the losing opinion become more isolated as the opinion keeps losing ground. The opposite is visible when an opinion grows in popularity; new followers combine nearby clusters into larger ones.

Refer to caption
Figure 6: The temporal evolution of the number of opinion clusters |Ω|Ω|\Omega|| roman_Ω | with (a-c) opinion A𝐴Aitalic_A and (d-f) opinion B𝐵Bitalic_B, for the asymmetric model. Different values of the external field Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT are combined with fixed α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 (equal weight to peers and external influence). Shaded areas represent the standard error. These results are averaged over 25 realisations of a given network/parameter combination.
Refer to caption
Figure 7: The temporal evolution of the average size of opinion clusters with opinion A𝐴Aitalic_A, ΩAdelimited-⟨⟩subscriptΩ𝐴\langle\Omega_{A}\rangle⟨ roman_Ω start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ⟩ (a-c), and opinion B, ΩBdelimited-⟨⟩subscriptΩ𝐵\langle\Omega_{B}\rangle⟨ roman_Ω start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT ⟩ (d-f), for the asymmetric model. Different values of the external field Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT are combined with fixed α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 (equal weight to peers and external influence). Shaded areas represent the standard error. These results are averaged over 25 realisations of a given network/parameter combination.
Refer to caption
Figure 8: The temporal evolution of the number of opinion clusters |Ω|Ω|\Omega|| roman_Ω | with (a-c) opinion A𝐴Aitalic_A and (d-f) opinion B𝐵Bitalic_B, for the symmetric model. Different values of the external field Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT are combined with fixed α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 (equal weight to peers and external influence). Shaded areas represent the standard error. These results are averaged over 25 realisations of a given network/parameter combination.
Refer to caption
Figure 9: The temporal evolution of the average size of opinion clusters Ωdelimited-⟨⟩Ω\langle\Omega\rangle⟨ roman_Ω ⟩ with (a-c) opinion A𝐴Aitalic_A and (d-f) opinion B𝐵Bitalic_B, for the symmetric model. Different values of the external field Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT are combined with fixed α=0.5𝛼0.5\alpha=0.5italic_α = 0.5 (equal weight to peers and external influence). Shaded areas represent the standard error. These results are averaged over 25 realisations of a given network/parameter combination.

This process is also observed in the symmetric version (Figs. 8-9), where low |ΩA|subscriptΩ𝐴|\Omega_{A}|| roman_Ω start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT | for every Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT indicates that indeed the media opinion A𝐴Aitalic_A does not, on average, become the minority opinion in the symmetric version. Fragmentation of the minority opinion clusters and conglomeration of the majority opinion clusters also occur in the symmetric version. Still, this effect seems more extreme than the asymmetric version. The symmetric version shows higher |Ω|Ω|\Omega|| roman_Ω | and lower Ωdelimited-⟨⟩Ω\langle\Omega\rangle⟨ roman_Ω ⟩ for a minority cluster and visa versa for a majority cluster. The values for |Ω|Ω|\Omega|| roman_Ω | and Ωdelimited-⟨⟩Ω\langle\Omega\rangle⟨ roman_Ω ⟩ approach those at A=1superscript𝐴1\mathcal{E}^{A}=1caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 1 for lower values of Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT in the symmetric version, showing that here a strong impact of the external opinion is already present for median values of Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT.

For both model versions, |Ω|Ω|\Omega|| roman_Ω | and Ωdelimited-⟨⟩Ω\langle\Omega\rangle⟨ roman_Ω ⟩ display differences for the different network structures. These differences are most pronounced in the regimes that lead to an |m|>0𝑚0|m|>0| italic_m | > 0, i.e. where, on average, we have a majority and minority opinion. We notice that the BA and ER topologies have opinion communities that evolve similarly concerning their average size but not when it comes to their numbers. When there is fragmentation of a minority opinion, this is more extreme in BA than in any other network topology. The asymmetrical degree distribution and hub structure of the BA network could lead to the many low-degree nodes being isolated in their single or small opinion clusters. For WS, cluster size and number evolution differ from ER and BA. When fragmentation occurs, |Ω|Ω|\Omega|| roman_Ω | first peaks, after which it declines to settle at a lower value. This effect can be explained by the fact that much higher values of m𝑚mitalic_m are reached for WS. So, the initial rise is fragmentation, as seen for BA and ER. Still, the following decline is because WS ends up with fewer agents that adhere to the minority opinion and, thus, fewer opinion communities. This is also seen by looking at the values of ΩAdelimited-⟨⟩subscriptΩ𝐴\langle\Omega_{A}\rangle⟨ roman_Ω start_POSTSUBSCRIPT italic_A end_POSTSUBSCRIPT ⟩ for WS, which are significantly more extreme than other network topologies.

4 Discussion

The asymmetric model results in three regimes that can be mapped in the αA𝛼superscript𝐴\alpha-\mathcal{E}^{A}italic_α - caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT parameter space (Fig. 10(a)). In the first regime B, the population consensus moves towards the opinion opposite to the external field. This regime is for high α𝛼\alphaitalic_α-low Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT combinations. The second regime A/B, showing average low consensus and order with high social stress, is at intermediate values of Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT or low values of α𝛼\alphaitalic_α. A third regime A of average high global and local consensus is formed in favour of the external field opinion. It is located in the parameter space’s high α𝛼\alphaitalic_α-high Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT part. This picture is recurrent across all studied network models with differences in the stationary state values and the sizes of and transition regions between the different regimes (Figs. 3-4). The average stationary consensus in the A/B regime is qualitatively similar to the situation where agents ignore the external influence, i.e. when α=0𝛼0\alpha=0italic_α = 0. For a fixed α>0𝛼0\alpha>0italic_α > 0, increasing the strength of the media opinion results in breaking up majority opinion clusters (regime B) into smaller ones, increasing nasubscript𝑛𝑎n_{a}italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT or the number of disagreeing nearest neighbour pairs. This leads to fracturing of those clusters, increasing social stress at first (regime A/B), decreasing again when enough agents take the media opinion. Large clusters of opinion are formed (regime A) (Figs. 3-7). The existence of regime B for the asymmetric version can be understood from eq. 2 where WiBAsuperscriptsubscript𝑊𝑖𝐵𝐴W_{i}^{B\rightarrow A}italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_B → italic_A end_POSTSUPERSCRIPT is small for high α𝛼\alphaitalic_α and small Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT. In contrast, WiABsuperscriptsubscript𝑊𝑖𝐴𝐵W_{i}^{A\rightarrow B}italic_W start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A → italic_B end_POSTSUPERSCRIPT is unaffected, meaning that the asymmetry results in the internal pressure away from the field being felt the strongest in that region.

Refer to caption
(a)
Refer to caption
(b)
Figure 10: Qualitative diagram of the macroscopic population opinion of the (a) asymmetric and (b) symmetric models at the stationary state.

The symmetric model exhibits two regimes in the αA𝛼superscript𝐴\alpha-\mathcal{E}^{A}italic_α - caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT parameter space (Fig. 10(b)). A disordered regime A/B presents itself at low α𝛼\alphaitalic_α and low Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT and for α=0𝛼0\alpha=0italic_α = 0 or A=0superscript𝐴0\mathcal{E}^{A}=0caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT = 0. An ordered regime A with high consensus in favour of the external opinion emerges for higher values of α𝛼\alphaitalic_α and \mathcal{E}caligraphic_E. Fixing a positive α𝛼\alphaitalic_α and increasing Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT leads to a shift from a state where (on average) consensus towards both opinions is possible (with a significant variation on the ordering – regime A/B), to a state where holding a different opinion to the external field is unlikely, with the emergence of consensus and order (regime A). This pattern is observed for all the network structures, with the main difference being that for WS networks the regime A is larger. The WS case shows the most different patterns in both directions. This results from the larger social clustering and fewer bridges concerning the ER and BA networks. These topological characteristics facilitate the spread of complex contagions, which can amplify the effect of the external opinion. The difference in the structure also leads to the higher prevalence of the majority opinions and the more extreme fragmentation of the minority opinion clusters. The exit probability for the voter model on degree heterogeneous networks (BA model) strongly favours the opinion held by large degree vertices [43], implying that high degree nodes (i.e. hubs) play an important role in simple (i.e. without reinforcement) contagion. In complex contagion, a hub requires several other agents to change its opinion and enforce it to its neighbours. This may explain why the BA structure is more resistant to external influence than the random structure (ER model). Similarly, the difference in degree distribution explains the higher opinion fragmentation in the BA structure compared to the ER structure.

The results suggest that an external field can affect consensus formation. In the absence of external influence (α=0𝛼0\alpha=0italic_α = 0), consensus (global order) (m𝑚mitalic_m) and local order (nasubscript𝑛𝑎n_{a}italic_n start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT) can form, and both opinions are equally likely to reach a majority. The amount depends on the underlying network structure, but on average, no preference toward any specific opinion is observed, resulting in the absence of consensus. Introducing an external field pushing a particular opinion creates a more diverse set of possibilities where different regimes appear depending on Asuperscript𝐴\mathcal{E}^{A}caligraphic_E start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT and the weight α𝛼\alphaitalic_α. The external opinion tends to disrupt the equality between both opinions, going from being equally likely to achieve a majority, towards a situation where a favoured opinion is more likely than the other one to gain a majority. Without this external field, the difference between the average numbers and sizes of the resulting opinion clusters of both opinions is small, implying that both opinions can achieve a majority. In general, the symmetric model reliably results in lower levels of social stress because it restricts opinion change away from the media opinion. In contrast, the asymmetric model allows for more fluctuations.

5 Conclusions

We studied the concurrent impact of media influence (as an external field) and social interactions (peer pressure) on the opinion dynamics of a stochastic majority model. We introduced one model version in which the external field only influences those agents who disagree with the field, and another version in which all agents are subject to the external field. The external field led the population to different macroscopic regimes in which different opinions will likely become the majority. In the first model, the population could converge to one of the opinions or to a mixed state where both opinions coexist. In the mixed case, the population behaves qualitatively in the same way in the presence or absence of external influence. In the second model, either the external opinion dominated, or the population had mixed opinions. These macroscopic patterns depended on the intensity of the external influence and on how much weight of the external field compared to the weight of social contacts. If the external influence is not strong enough, the population can move towards the opposite opinion to the one being enforced. Furthermore, the external field is reinforced by social clustering, which helps align the population towards the external opinion, even in the case of low-intensity external influence. On the other hand, high-degree nodes help to prevent the population from moving towards the external field. By controlling whether agents feel the external field equally, we showed that agents being exposed to the external influence, independently of their opinions, led the population to states where one of the opinions (the one not being enforced) was never dominant. On the other hand, if only agents aligned with the external field were affected by this field, the space of possible macroscopic states was equal for both opinions. This can be relevant to business advertising, health, and other governmental intervention campaigns.

The impacts of the concurrent influence of peers and an external actor have been investigated in relatively simple social structures. More complex network topologies could reveal other underlying structures affecting the opinion dynamics in real populations. An example could be using stochastic block models to represent community structure or network models combining homophily with preferential attachment [44]. Additionally, agents could have heterogeneous weights representing differences in their susceptibility to external actors. A more in-depth study of the connection between complex opinion spread and topological characteristics should include using measures adapted to complex contagions [45]. The WS network is the sole topology where full consensus was reached. This result, of complete order, is understood as a finite-size effect for the classical voter model on WS networks [46]. Despite limitations, our study shows that introducing simultaneous external and peer influence in opinion dynamics can lead to a more diverse behaviour than a monotonous increase in consensus for increasing external influence. The emergence of different macroscopic population opinions can help our understanding of how groups with varying trust in the media and subject to certain levels of media presence, react.

References

  • [1] D. Stauffer, Introduction to statistical physics outside physics, Physica A: Statistical Mechanics and its Applications 336 (1-2) (2004) 1–5. doi:10.1016/j.physa.2004.01.004.
  • [2] C. Castellano, S. Fortunato, V. Loreto, Statistical physics of social dynamics, Reviews of Modern Physics 81 (2) (2009) 591–646. doi:10.1103/RevModPhys.81.591.
  • [3] P. Contucci, C. Vernia, On a statistical mechanics approach to some problems of the social sciences, Frontiers in Physics 8 (2020) 585383. doi:10.3389/fphy.2020.585383.
  • [4] J. Q. Stewart, The development of social physics, American Journal of Physics 18 (5) (1950) 239–253. doi:10.1119/1.1932559.
  • [5] P. Ball, Critical mass: How one thing leads to another, Arrow Books, London, 2005.
  • [6] H. Xia, H. Wang, Z. Xuan, Opinion dynamics: A multidisciplinary review and perspective on future research, International Journal of Knowledge and Systems Science 2 (4) (2011) 72–91. doi:10.4018/jkss.2011100106.
  • [7] F. Schweitzer, Sociophysics, Physics Today 71 (2) (2018) 40–46. doi:10.1063/PT.3.3845.
  • [8] C. Song, Z. Qu, N. Blumm, A.-L. Barabási, Limits of predictability in Human mobility, Science 327 (5968) (2010) 1018–1021. doi:10.1126/science.1177170.
  • [9] W. B. Arthur, Inductive reasoning and bounded rationality, The American Economic Review 84 (2) (1994) 406–411.
  • [10] M. Granovetter, Threshold models of collective behavior, American Journal of Sociology 83 (6) (1978) 1420–1443. doi:10.1086/226707.
  • [11] R. Axelrod, R. A. Hammond, The evolution of ethnocentric behavior, in: Proc. Midwest Political Science Convention, April 3-6, 2003, Chicago, IL, Chicago, IL, 2003.
  • [12] M. McPherson, L. Smith-Lovin, J. M. Cook, Birds of a feather: Homophily in social networks, Annual Review of Sociology 27 (2001) 415–444.
    URL http://www.jstor.org/stable/2678628
  • [13] T. C. Schelling, Dynamic models of segregation, The Journal of Mathematical Sociology 1 (2) (1971) 143–186. doi:10.1080/0022250X.1971.9989794.
  • [14] R. Axelrod, The dissemination of culture: A model with local convergence and global polarization, The Journal of Conflict Resolution 41 (2) (1997) 203–226.
  • [15] S. H. Strogatz, Exploring complex networks, Nature 410 (6825) (2001) 268–276. doi:10.1038/35065725.
  • [16] S. Murrar, M. Brauer, Entertainment-education effectively reduces prejudice, Group Processes & Intergroup Relations 21 (7) (2018) 1053–1077. doi:10.1177/1368430216682350.
  • [17] R. H. Thaler, C. R. Sunstein, Nudge: Improving decisions about health, wealth, and happiness, rev. and expanded ed Edition, Penguin Books, New York, 2009.
  • [18] B. H. Rahman, Conditional influence of media: Media credibility and opinion formation, Journal of Political Studies 21 (1) (2014) 299–314, copyright - Copyright Department of Political Science, University of the Punjab Summer 2014; Document feature - Tables; ; Last updated - 2015-05-11.
  • [19] V. Pansanella, A. Sîrbu, J. Kertesz, G. Rossetti, Mass media impact on opinion evolution in biased digital environments: A bounded confidence model, Scientific Reports 13 (2023) 14600.
  • [20] L. Helfmann, N. D. Conrad, P. Lorenz-Spreen, C. Schütte, Modelling opinion dynamics under the impact of influencer and media strategies, Scientific Reports 13 (2023) 19375.
  • [21] L. H. Hoffman, C. J. Glynn, M. E. Huge, R. B. Sietman, T. Thomson, The role of communication in public opinion processes: Understanding the impacts of intrapersonal, media, and social filters, International Journal of Public Opinion Research 19 (3) (2007) 287–312. doi:10.1093/ijpor/edm014.
  • [22] A. Sîrbu, V. Loreto, V. D. P. Servedio, F. Tria, Opinion dynamics: Models, extensions and external effects, in: V. Loreto, M. Haklay, A. Hotho, V. D. Servedio, G. Stumme, J. Theunis, F. Tria (Eds.), Participatory Sensing, Opinions and Collective Awareness, Springer International Publishing, 2017, pp. 363–401. doi:10.1007/978-3-319-25658-0\_17.
  • [23] P. Clifford, A. Sudbury, A model for spatial conflict, Biometrika 60 (3) (1973) 581–588. doi:10.1093/biomet/60.3.581.
  • [24] R. A. Holley, T. M. Liggett, Ergodic theorems for weakly interacting infinite systems and the voter model, The Annals of Probability 3 (4) (1975) 643–663, publisher: Institute of Mathematical Statistics.
  • [25] S. Redner, Reality-inspired voter models: A mini-review, Comptes Rendus Physique 20 (4) (2019) 275–292. doi:10.1016/j.crhy.2019.05.004.
  • [26] D. Centola, M. Macy, Complex contagions and the weakness of long ties, American Journal of Sociology 113 (3) (2007) 702–734. doi:10.1086/521848.
  • [27] K. Sznajd-Weron, J. Sznajd, Opinion evolution in closed community, International Journal of Modern Physics C 11 (06) (2000) 1157–1165. doi:10.1142/S0129183100000936.
  • [28] S. Galam, Minority opinion spreading in random geometry, The European Physical Journal B 25 (4) (2002) 403–406.
  • [29] M. J. De Oliveira, Isotropic majority-vote model on a square lattice, Journal of Statistical Physics 66 (1-2) (1992) 273–281. doi:10.1007/BF01060069.
  • [30] C. Castellano, M. A. Muñoz, R. Pastor-Satorras, Nonlinear q-voter model, Physical Review E 80 (4) (2009) 041129. doi:10.1103/PhysRevE.80.041129.
  • [31] A. R. Vieira, C. Anteneodo, Threshold q-voter model, Physical Review E 97 (5) (2018) 052106. doi:10.1103/PhysRevE.97.052106.
  • [32] L. Dall’Asta, C. Castellano, Effective surface-tension in the noise-reduced voter model, Europhysics Letters (EPL) 77 (6) (2007) 60005. doi:10.1209/0295-5075/77/60005.
  • [33] D. Volovik, S. Redner, Dynamics of confident voting, Journal of Statistical Mechanics: Theory and Experiment 2012 (04) (2012) P04003. doi:10.1088/1742-5468/2012/04/P04003.
  • [34] F. Zarei, Y. Gandica, L. E. C. Rocha, Bursts of communication increase opinion diversity in the temporal Deffuant model, Scientific Reports 14 (1) (2024) 2222. doi:10.1038/s41598-024-52458-w.
  • [35] N. Perra, L. E. C. Rocha, Modelling opinion dynamics in the age of algorithmic personalisation, Scientific Reports 9 (2019) 7261.
  • [36] N. Botte, J. Ryckebusch, L. E. C. Rocha, Clustering and stubbornness regulate the formation of echo chambers in personalised opinion dynamics, Physica A: Statistical Mechanics and its Applications 599 (2022) 127423. doi:10.1016/j.physa.2022.127423.
  • [37] J. R. Majmudar, S. M. Krone, B. O. Baumgaertner, R. C. Tyson, Voter models and external influence, The Journal of Mathematical Sociology 44 (1) (2020) 1–11. doi:10.1080/0022250X.2019.1625349.
  • [38] J. Civitarese, External fields, independence, and disorder in q-voter models, Physical Review E 103 (1) (2021) 012303. doi:10.1103/PhysRevE.103.012303.
  • [39] G. De Marzo, A. Zaccaria, C. Castellano, Emergence of polarization in a voter model with personalized information, Physical Review Research 2 (4) (2020) 043117. doi:10.1103/PhysRevResearch.2.043117.
  • [40] Azhari, R. Muslim, The external field effect on the opinion formation based on the majority rule and the q-voter models on the complete graph, International Journal of Modern Physics C (2022) 2350088doi:10.1142/S0129183123500882.
  • [41] N. Crokidakis, Effects of mass media on opinion spreading in the Sznajd sociophysics model, Physica A: Statistical Mechanics and its Applications 391 (4) (2012) 1729–1734. doi:10.1016/j.physa.2011.11.038.
  • [42] T. H. Cormen, C. E. Leiserson, R. L. Rivest, C. Stein, 20.2 Breadth-first search, in: Introduction to algorithms, fourth edition Edition, The MIT Press, Cambridge, Massachusett, 2022, p. 1291.
  • [43] V. Sood, T. Antal, S. Redner, Voter models on heterogeneous networks, Physical Review E 77 (4) (2008) 041121. doi:10.1103/PhysRevE.77.041121.
  • [44] F. Karimi, M. Génois, C. Wagner, P. Singer, M. Strohmaier, Homophily influences ranking of minorities in social networks, Scientific Reports 8 (1) (2018) 11077. doi:10.1038/s41598-018-29405-7.
  • [45] D. Guilbeault, D. Centola, Topological measures for identifying and predicting the spread of complex contagions, Nature Communications 12 (1) (2021) 4430. doi:10.1038/s41467-021-24704-6.
  • [46] C. Castellano, D. Vilone, A. Vespignani, Incomplete ordering of the voter model on small-world networks, Europhysics Letters (EPL) 63 (1) (2003) 153–158. doi:10.1209/epl/i2003-00490-0.