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

Critical Exponents of Master-Node Network Model

Antonio Mihara
Physics Department, Federal University of São Paulo, Diadema–SP, Brazil
Anderson A.  Ferreira
Physics Department, Federal University of São Paulo, Diadema–SP, Brazil
André C. R. Martins
School of Arts, Science and Humanity, University of São Paulo, São Paulo-SP, Brazil
Fernando F. Ferreira
[email protected]
Physics Department of FFCLRP, University of São Paulo, Ribeirão Preto-SP, Brazil
Abstract

The dynamics of competing opinions in social network play an important role in society, with many applications in diverse social contexts as consensus, elections, morality and so on. Here we study a model of interacting agents connected in networks to analyze their decision stochastic process.  We consider a first-neighbor interaction between agents in a one-dimensional network with a shape of ring topology. Moreover, some agents are also connected to a hub, or master node, that has preferential choice or bias. Such connections are quenched. As the main results we observed a continuous non-equilibrium phase transition to an absorbing state as a function of control parameters. By using the finite size scaling method, we analyzed the static and dynamic critical exponents to show that this model probably cannot match any universality class already known.

1 Introduction

Absorbing state phase transitions with quenched disorder have attracted considerable attention [1]. Such phase transitions can be found in many areas, including physics, chemistry,social systems and biology, with examples in quantum systems [2], catalysis [3], surface and interface growth [4],opinion dynamics [5], ecology [6] and spreading of epidemics [7]. Such nonequilibrium phase transitions phenomena has been witnessing considerable theoretical progress, but many issues remain to be addressed in the literature and surprises may arise.

The critical behavior of absorbing phase transitions can be categorized into a finite number of universality classes. The most important universality class of absorbing-state transitions is the called directed percolation (DP)[8], which is a fundamental class of nonequilibrium phase transitions, playing a similar role as the Ising universality class in equilibrium statistical mechanics. Just to give an example of a particular DP model, consider the spreading disease process. One may consider a network of (infected or healthy) agents spreading the disease through the neighborhood. Such a system may exhibit an emerging phase transition between endemic or pandemic regime.

In the last years fews experimental achievements of the DP class have been observed [9]. However, the main difficulty in verifying the critical exponents of this class is related to inherent impurities present in real experiments. Because such disorders effectively change the intensity of local interactions. And, according to the Harris criterion [10], the introduction of a spatial quenching in the DP class is relevant for dimensions 3\leq 3. That is, the spatial correlation length, as well as, the correlation functions scale logarithmic , modifying the values of the critical exponents, as observed by numerical simulation in two dimensions ( see [11] and references therein) and confirmed by Hoobergys et al. [12] for the contact process model with quenched dilution in one and two dimensions.

Recently it was introduced a model of dynamics of opinions that assumes a network topology in the form of a ring with some nodes connected to a node outside the ring called a master node that has a fixed opinion on a subject. The connections between the ring nodes and the master node are quenched. Additionally, the dynamics has two absorbing states (0(0 or 1)1), where 11 is the opinion of the master node and 0 is the selfish opinion placed as a natural bias of agents in the social network. This model calls attention to presenting a second-order phase transition [13].

In this article, we will explore this model by making a more rigorous analysis of the critical exponents, both static and dynamic.As the main result, we find examples of critical exponents that depend on the model parameters and are different from the exponents of the voter model or directed percolation classes.

This paper is organized as follows. In section II, we describe the model. Then, we present the results of the study based on finite size scaling for static and dynamic critical exponents. Section IV shows the discussion and finally in section V the conclusions.

2 The Master–Network Model

The model consists of a network in the shape of a ring formed by nodes with periodic boundary conditions and a master node as illustrates in Figure 1. Initially, each node ii represents a particle or agent which may assume two different states si=0,1s_{i}=0,1 with probability wsw_{s}. The state si=0s_{i}=0 represents, for instance, some opinion or behavior with negative meaning like immoral, selfish, or corrupt behaviors, which harm the collective welfare, while si=1s_{i}=1 represents the opposite.

Each node interacts with its first nearest neighbors and may interact with a master node with probability ρ\rho (quenched disorder). This connection is drawn at the beginning and is fixed. In other words, ρ\rho is the fraction of the master node’s followers. The interaction strength between the master and each connected node is denoted by rr, which represents the master node’s influence power.

The general configuration of the system is given by (si,Γi)(s_{i},\Gamma_{i}), with i=1,,Li=1,\dots,L representing the individual nodes and Γi=1(0)\Gamma_{i}=1\,(0) if there is a connection between it and the master node (or not). In turn, the probability of the node/citizen being influenced by the master node is equal to rr.

Refer to caption
Figure 1: Model representation. Small spheres represent interacting nodes (citizens). Each node is in the state s=0s=0 or 1. The large sphere represents the master node and its interaction strength with other particles (the citizens) is fixed (denoted by rr) and the links are in quenched disorder, with density denoted by ρ\rho.

We are interested here in consensus-seeking agents. That fact limits the choices of possible interaction parameters. However, it is interesting to start considering the general case.

Assuming the central agent is not influenced by its neighbors, we can write the interaction matrix for all other agents, on the absence of the external field as summarized in Table 1.

Table 1: Transition probabilities in the general cases for Γi=0\Gamma_{i}=0.
sisis_{i}\rightarrow s_{i}^{\prime}
si1s_{i-1} si+1s_{i+1} 000\rightarrow 0 010\rightarrow 1 101\rightarrow 0 111\rightarrow 1
0 0 p1p_{1} 1p11-p_{1} p2p_{2} 1p21-p_{2}
1 1 p3p_{3} 1p31-p_{3} p4p_{4} 1p41-p_{4}
0 1 1w01-w_{0} w0w_{0} w1w_{1} 1w11-w_{1}
1 0 1w21-w_{2} w2w_{2} w3w_{3} 1w31-w_{3}

The two last lines of the matrix in Table 1 correspond to a case where the agent is surrounded by one neighbor with sj=0s_{j}=0 and another with sj=1s_{j}=1. In most applications, those two cases are symmetrical and identical, but it is worth noticing that we could introduce a chirality effect. For this paper, we assume there is no chirality and, therefore, w0=w2w_{0}=w_{2} and w1=w3w_{1}=w_{3}. That is, the two last lines in the matrix are identical.

It is also interesting to map how traditional models translate in Table 1 parameters. Normal majority models [14, 15, 16, 17, 18, 19] where both states are identical would have p1=1p4=1p_{1}=1-p_{4}=1 and, since both states are symmetrical, p2=1p3p_{2}=1-p_{3}. In the simplest case, where majority always convince and it is the only way to change the opinion of an agent, we would have p1=p2=1p_{1}=p_{2}=1. We also have that w0=w1=0w_{0}=w_{1}=0, reflecting the fact there is no update when no majority is observed. In the case all agents were contrarians [14, 15, 16, 17, 18, 19], we would have p1=p20.5p_{1}=p_{2}\leq 0.5, meaning each agent would tend to disagree with the majority. And we can get Galam-like models [20, 21, 22], where there is a status quo state (here, si=0s_{i}=0) that tends to be preserved in discussions (with the difference only the central agent can change its opinion), if we make p1=p4=0p_{1}=p_{4}=0, p2=1p3p_{2}=1-p_{3}, w0=0w_{0}=0, and p2>w10p_{2}>w_{1}\neq 0. Traditionally, we would also have p2=1p_{2}=1 in this case, but that is not necessary. Another interesting scenario corresponds to w1=1w0w_{1}=1-w_{0}, p1=p2p_{1}=p_{2}, and p3=p4p_{3}=p_{4}. In this situation, the central agent has no memory of its previous state, as its new state depends only on the neighbors and not on its initial preference.

The introduction of a central agent connected to some agents allows us to have agents influenced by their neighbors using distinct rules. That difference can, in principle, be used to describe many different scenarios. In the original paper and in much of the analysis that follows, this framework was used to explore a scenario where a central agent tries to impose the choice s=1s=1. While the terminology in the previous paper [13] identifies that option as a moral rule, the same model can also correspond to other scenarios, such as compliance to law, obedience to a norm, or even the adoption of a specific product.

We suppose that the interaction between node ii and its neighbors i1i-1 and i+1i+1 depends on the value of Γi\Gamma_{i}. For Γi=0\Gamma_{i}=0, sis_{i} will align with the majority in the neighborhood, a situation which leads to consensus. If there are differences in opinions between neighbors (frustration), the state si=0(1)si=1(0)s_{i}=0\,(1)\rightarrow s_{i}^{\prime}=1\,(0) with probability w0(w1)w_{0}\,(w_{1}) as we can see in Table 2.

Table 2: Transition probabilities for Γi=0\Gamma_{i}=0.
sisis_{i}\rightarrow s_{i}^{\prime}
si1s_{i-1} si+1s_{i+1} 000\rightarrow 0 010\rightarrow 1 101\rightarrow 0 111\rightarrow 1
0 0 p1p_{1} 1p11-p_{1} p2p_{2} 1p21-p_{2}
1 1 p3p_{3} 1p31-p_{3} p4p_{4} 1p41-p_{4}
0 1 1w01-w_{0} w0w_{0} w1w_{1} 1w11-w_{1}
1 0

For the case Γi=1\Gamma_{i}=1, the transition probabilities will be different. If both of its first nearest neighbors are si1=si+1=0s_{i-1}=s_{i+1}=0 then the particle will change from state si=0(1)s_{i}=0\,(1) to si=1s_{i}^{\prime}=1, with probability q(r)q\,(r). If both of its first nearest neighbors are si1=si+1=1s_{i-1}=s_{i+1}=1, then the particle will change from state si=0(1)s_{i}=0\,(1) to si=1s_{i}^{\prime}=1 with probability 11. Finally, if there is frustration between neighbors, the state si=0(1)si=1(0)s_{i}=0\,(1)\rightarrow s_{i}^{\prime}=1\,(0) with probability r0(r1)r_{0}\,(r_{1}), respectively. This case is summarized in Table 3.

Table 3: Transition probabilities for Γi=1\Gamma_{i}=1.
sisis_{i}\rightarrow s_{i}^{\prime}
si1s_{i-1} si+1s_{i+1} 000\rightarrow 0 010\rightarrow 1 101\rightarrow 0 111\rightarrow 1
0 0 1q11-q_{1} q1q_{1} 1r1-r rr
1 1 q2q_{2} 1q21-q_{2} q3q_{3} 1q31-q_{3}
0 1 1r01-r_{0} r0r_{0} r1r_{1} 1r11-r_{1}
1 0

It is worth mentioning that, in principle, the central agent can also work as a marker for the heterogeneity of the agents. Heterogeneity, of course, does not need this framework, but it is nice to notice that it can be included here. For example, if we have normal interactions when Γi=0\Gamma_{i}=0, we can easily introduce inflexibles [23, 24, 25, 26, 22] if we make, when Γi=1\Gamma_{i}=1, q1=q3=r0=r1=0q_{1}=q_{3}=r_{0}=r_{1}=0 and q2=r=1q_{2}=r=1.

Aiming to analyse this model, we choose some constraints to the parameters r0,r1,w0r_{0},r_{1},w_{0} and w1w_{1} in terms of rr, ρ\rho and Δ=w0w1\Delta=w_{0}-w_{1}.

The parameter Δ\Delta measures the intrinsic bias (or tendency) of an agent be in the state s=0(1)s=0(1) when Δ>0(<0)\Delta>0(<0) in the absence of any interaction or influence. If Δ\Delta is positive, the system has a tendency to be opposite to the master node. This is the most interesting situation. So, we study the regime in which 0<Δ<10<\Delta<1.

The probabilities r0r_{0} and r1r_{1} are parameterized so that when the master’s influence is null (r=0r=0), we have r0=w0r_{0}=w_{0} and r1=w1r_{1}=w_{1}. Otherwise, when r=1r=1, we should have necessarily r0=1r_{0}=1 and r1=0r_{1}=0. The simplest way is through a linear relation:

r0=r+(1r)(1Δ2),r_{0}=r+(1-r)(\frac{1-\Delta}{2})\,, (1)
r1=(1r)(1+Δ2).r_{1}=(1-r)(\frac{1+\Delta}{2})\,. (2)

Finally, the parameterized version of the model has only three free parameters:

0r1,;    0Δ1,;and     0ρ1.0\leq r\leq 1,;\;\;\;\;0\leq\Delta\leq 1,;\;\;\;\mbox{and}\;\;\;\;\;0\leq\rho\leq 1. (3)

3 Results

The most important characteristic of the critical phenomena is the set of critical exponents related with the singular behaviour of the thermodynamics functions near the critical point [27]. In equilibrium statistical mechanics the principal scheme of classification of the continuous phase transitions in two dimesions started to be structured after the seminal work of ref. [28]. However in non-equilibrium phase transitions the situation is more complex, due to the extra dimension time and other degrees of freedom present in dynamics, other critical exponents (dynamical exponents) are necessary to classify a non-equilibrium phase transition [29, 30, 31, 1]

3.1 Static Exponents

Monte Carlo simulations were used for obtaining the critical exponents for the order parameter II, the average density of immoral citizens. By setting the values of Δ\Delta and ρ\rho, one can expect a power–law behavior I(r)|rrc|βI(r)\sim|r-r_{c}|^{\beta}, near the critical point rcr_{c}. For large LL and |rrc|0|r-r_{c}|\rightarrow 0, II can be written in the following form:

ILβ/νf((rrc).L1/ν),I\sim L^{-\beta/\nu_{\bot}}\,f\left((r-r_{c}).L^{1/\nu_{\bot}}\right)\,, (4)

with the scaling function f(x)|x|βf(x)\sim|x|^{\beta}, for large |x||x|.

We performed numerical simulations for Δ=0.5\Delta=0.5 with two different values of ρ\rho (ρ=0.7\rho=0.7 and ρ=0.9\rho=0.9) and obtained estimates for β\beta and ν\nu_{\bot} for both cases. In Fig.2 we present (on the top panel) the results obtained with Δ=0.5\Delta=0.5 and ρ=0.7\rho=0.7 for different network sizes; and the corresponding finite–size scaling (FSS) analysis on the right panel.

Refer to caption
Refer to caption
Figure 2: Top: I×rI\,\times\,r, for parameters Δ=0.5,ρ=0.7\Delta=0.5,\,\rho=0.7, for different network sizes.Bottom: finite-size scaling. The dashed red line shows the behavior

From FSS we obtained the critical point rc=0.1931(9)r_{c}=0.1931(9), and the exponents a1/ν=0.54(9)a\equiv 1/\nu_{\bot}=0.54(9) and bβ/ν=0.160(12)b\equiv\beta/\nu_{\bot}=0.160(12). From both results one can obtain the following estimates: β=0.30(5)\beta=0.30(5) and ν=1.9(3)\nu_{\bot}=1.9(3).

However on the bottom panel of Fig.2 we observe that the scaling function ff behaves as f(x)|x|0.288(4)f(x)\sim|x|^{0.288(4)}, for large |x||x| (red dashed line), then we can obtain a more precise estimate: β=0.288(4)\beta=0.288(4).

For the parameters Δ=0.5\Delta=0.5 and ρ=0.9\rho=0.9, we obtained rc=0.1426(9)r_{c}=0.1426(9), a1/ν=0.530(8)a\equiv 1/\nu_{\bot}=0.530(8) and bβ/ν=0.1393(18)b\equiv\beta/\nu_{\bot}=0.1393(18). From these exponents we obtained the following estimates: β=0.263(5)\beta=0.263(5) and ν=1.887(28)\nu_{\bot}=1.887(28).

3.2 Dynamic Exponents

Simulations were performed with initial configuration very near the absorbing state: all sites with si=1s_{i}=1, except eleven sites in the “center” of the network (the central site, i=L/2i=L/2, and five sites both sides) with si=0s_{i}=0. Three time–dependent quantities were measured at the critical point:

  • N(t)=N(t)= the mean number of immoral citizens (sites with s=s=0).

  • P(t)=P(t)= the mean survival probability of immoral citizens.

  • R2(t)=R^{2}(t)= the mean quadratic distance of immoral citizens from the center (i=L/2i=L/2) of network.

For the asymptotic evolution we observed power–law behavior: NtηN\sim t^{\eta}, P(t)tδP(t)\sim t^{-\delta} and R2(t)tzR^{2}(t)\sim t^{z}. In Fig. 3 we present our results for Δ=0.5,ρ=0.7\Delta=0.5,\rho=0.7 and some values of rr. We obtained the exponents: η=0.3272(9)\eta=0.3272(9), δ=0.0684(9)\delta=0.0684(9) and z=0.930(3)z=0.930(3), for r=0.1923r=0.1923.

We also studied the case with Δ=0.5,ρ=0.9\Delta=0.5,\rho=0.9. Now for r=0.1434r=0.1434, the exponents obtained: η=0.4228(16)\eta=0.4228(16), δ=0.0798(7)\delta=0.0798(7) and z=1.164(7)z=1.164(7).

Finally one can expect the following scaling/hyperscaling relations [29]

βν=2δz,2η+4δ=dz.\frac{\beta}{\nu_{\bot}}=2\,\frac{\delta}{z}\,,\quad 2\eta+4\delta=dz\,. (5)

In table 4 we summarize the results above, and in table 5 we observe that both scaling relation above are satisfied, within the error estimates.

Table 4: Static and dynamic exponents obtained for Δ=0.5\Delta=0.5, for different values of ρ\rho.
ρ\rho β\beta ν\nu_{\bot} η\eta δ\delta zz
0.7 0.30(5) 1.9(3) 0.3272(9) 0.0684(9) 0.930(3)
0.9 0.263(5) 1.887(28) 0.4228(16) 0.0798(7) 1.164(7)
Table 5: Scaling relations for the exponents.
ρ\rho β/ν\beta/\nu_{\bot} = 2δ/z2\delta/z 2η+4δ2\eta+4\delta = zz
0.7 0.160(12) 0.1471(20) 0.928(4) 0.930(3)
0.9 0.1393(18) 0.1371(14) 1.165(4) 1.164(7)

4 Discussions

An important issue in obtaining dynamic exponents concerns the time behavior of the quantities N,P,R2N,P,R^{2}: at the critical point, such quantities depend on tt as a power law or as a power of lnt\ln t? In Figure 3 one can observe that NN behaves as a power law of tt, since the critical curve (corresponding to rrcr\approx r_{c}) is very close to straight line on a log-log plot.

Refer to caption
Figure 3: N(t)N(t) for Δ=0.5\Delta=0.5, ρ=0.7\rho=0.7 and some values of rr.

Quenched models are not always relevant compared to pure (non-quenched) models, i.e, they do not bring fundamental novelties already present in pure models. According to Harris [10] a quenched perturbation is relevant when the relation νd<2\nu_{\perp}d<2 is satisfied. In this model, the non-quenched versions (ρ=0\rho=0 or ρ=1\rho=1) have no transition. Therefore, we have a relevant quenched model that does not satisfy the Harris criterion.

One aspect to be questioned is whether the critical exponents of the model under study belong to some already known universality class. In a first analysis we could try identifying these critical exponents with some class already well known in the literature of one-dimensional stochastic processes with absorbing states, such as the DP class [8], the parity conserving (PC) universality class [32], the universality class of pair contact process with diffusion [33],[34], the Manna universality class in sandpile models [35] and  the voter universality class [36] since those models are the promising ca

For comparison , we will highlight the exponent set for the case where the density of connected followers is ρ=0.9\rho=0.9. At this choice, the static critical exponents are β=0.263(3)\beta=0.263(3) and ν=1.887(2)\nu_{\perp}=1.887(2), while in the DP class at d=1d=1 we have β=0.276486(8)\beta=0.276486(8) and ν=1.096854(4)\nu_{\perp}=1.096854(4). Likewise, there is a discrepancy in the dynamics critical exponents.

Another possibility would be to compare these results with the quenched DP university class. However, as shown analytically by Hoobergys et al. [37] through the quenched renormalization group, an introduction of an impurity (quenched perturbation) in dimension d=1d=1 in the DP class establishes that the exponent ν=2\nu_{\perp}=2 and the exponent β=352\beta=\frac{3-\sqrt{5}}{2}. Furthermore, as argued by these same authors, the temporal correlation functions scale logarithmically instead of a power law, as observed in our numerical analysis for our model in these parameterizations.

Alternatively, a priori, one can map a nonequilibrium stochastic system in dd-dimensions to a quantum Hamiltonian [38] which in turn can be related to a (d+1)(d+1)-dimensional equilibrium system [39]. So it seems plausible that such a coincidence is not accidental. We could not prove the connection between our (1dd) model and the 2dXYh4dXYh_{4} model, but we speculate that such a connection probably exists and we put the exponents (β,ν\beta,\nu) and those of the 2dXYh4dXYh_{4} model [40] on the same plot (Fig.4

As it was predicted in ref.[41]  and presented in Fig.5 of ref.[40], the exponents of 2dXYh4dXYh_{4} model depends linearly on 1/h41/h_{4}. In Fig.4  we plot the exponents β,ν\beta,\nu of 2dXYh4dXYh_{4} model corresponding to  to different values of 1/h41/h_{4} (blue and black triangles), such results were obtained through several Monte Carlo/finite size scaling studies, see ref.[40] and references therein.The yellow and green bands are to guide the eye and to reinforce that β\beta and ν\nu are proportional to 1/h41/h_{4}. By assuming that a mapping between both models is possible, we also plot the results (red symbols) of our 1dd model corresponding to some hypothetical values of h4h_{4}.

Refer to caption
Figure 4: Comparison between our exponents (red points) and those of the 2dXYh42dXYh_{4} model (blue and black points).

5 Conclusions

The model described and studied here introduces a central agent that can influence a proportion of agents and, in that sense, it is similar to other models where the effects of media and well-connected agents [42], [43] or norms [44]  were studied. However, the central agent here can not only influence the opinions of those connected to it. It can also change how connected agents are influenced by their neighbors. In that sense, it can act as a catalyst. And, as we have seen, we can describe numerous distinct models of the literature using the framework presented here. Therefore, the model can also be seen as an effort to generalize and compare the existing literature [45, 46, 47].

But the model is not interesting only for opinion dynamics. Given its dynamic properties, it seems to be novel and relevant for the non equilibrium statistical mechanics literature, as far as we know.

To determine whether the model is something new in relation to many others that are already well studied, we attempted to identify an eventual universality class in which it could be classified. Surprisingly, it doesn’t map onto existing classes. Furthermore, critical exponents have a dependency on model parameters. This feature makes the analysis quite laborious to be conducted numerically. The ideal approach is to seek analytical approaches that allow a broader understanding of the model.

Finally, it is worth remembering that the solutions presented were obtained for a very particular choice of constraints between the parameters described in the  tables 2 and 3 . This parameterization was just to simplify the analysis. Certainly, other cases may also have even more intriguing properties.

References

  • [1] Haye Hinrichsen. Non-equilibrium critical phenomena and phase transitions into absorbing states. Advances in physics, 49(7):815–958, 2000.
  • [2] Ricardo Gutiérrez, Cristiano Simonelli, Matteo Archimi, Francesco Castellucci, Ennio Arimondo, Donatella Ciampini, Matteo Marcuzzi, Igor Lesanovsky, and Oliver Morsch. Experimental signatures of an absorbing-state phase transition in an open driven many-body quantum system. Physical Review A, 96(4):041602, 2017.
  • [3] Robert M Ziff, Erdagon Gulari, and Yoav Barshad. Kinetic phase transitions in an irreversible surface-reaction model. Physical review letters, 56(24):2553, 1986.
  • [4] Lei-Han Tang and Heiko Leschhorn. Pinning by directed percolation. Physical Review A, 45(12):R8309, 1992.
  • [5] Franco Bagnoli, Fabio Franci, and Raúl Rechtman. Opinion formation and phase transitions in a probabilistic cellular automaton with two absorbing states. In International Conference on Cellular Automata, pages 249–258. Springer, 2002.
  • [6] Claudio Borile, Amos Maritan, and Miguel A Munoz. The effect of quenched disorder in neutral theories. Journal of Statistical Mechanics: Theory and Experiment, 2013(04):P04032, 2013.
  • [7] Romualdo Pastor-Satorras, Claudio Castellano, Piet Van Mieghem, and Alessandro Vespignani. Epidemic processes in complex networks. Reviews of modern physics, 87(3):925, 2015.
  • [8] Peter Grassberger and Alberto de la Torre. Reggeon field theory (schlögl’s first model) on a lattice: Monte carlo calculations of critical behaviour. Annals of Physics, 122(2):373–396, 1979.
  • [9] Haye Hinrichsen. On possible experimental realizations of directed percolation. Brazilian Journal of Physics, 30:69–82, 2000.
  • [10] André J Noest. New universality for spatially disordered cellular automata and directed percolation. Physical review letters, 57(1):90, 1986.
  • [11] Alexander HO Wada and Mário J De Oliveira. Critical properties of the contact process with quenched dilution. Journal of Statistical Mechanics: Theory and Experiment, 2017(4):043209, 2017.
  • [12] Jef Hooyberghs, Ferenc Iglói, and Carlo Vanderzande. Absorbing state phase transitions with quenched disorder. Physical Review E, 69(6):066140, 2004.
  • [13] Anderson A Ferreira, Leandro A Ferreira, Antonio Mihara, and Fernando F Ferreira. A stochastic quenched disorder model for interaction of network-master node systems. Physica A: Statistical Mechanics and its Applications, 556:124781, 2020.
  • [14] Serge Galam. Contrarian deterministic effects on opinion dynamics: the hung elections scenario. Physica A: Statistical Mechanics and its Applications, 333:453–460, 2004.
  • [15] Marta S de la Lama, Juan M López, and Horacio S Wio. Spontaneous emergence of contrarian-like behaviour in an opinion spreading model. EPL (Europhysics Letters), 72(5):851, 2005.
  • [16] André CR Martins and Cleber D Kuba. The importance of disagreeing: Contrarians and extremism in the coda model. Advances in Complex Systems, 13(05):621–634, 2010.
  • [17] Hyunsuk Hong and Steven H Strogatz. Kuramoto model of coupled oscillators with positive and negative coupling parameters: An example of conformist and contrarian oscillators. Physical Review Letters, 106(5):054102, 2011.
  • [18] Joao Paulo Gambaro and Nuno Crokidakis. The influence of contrarians in the dynamics of opinion formation. Physica A: Statistical Mechanics and its Applications, 486:465–472, 2017.
  • [19] Yucheng Dong, Min Zhan, Gang Kou, Zhaogang Ding, and Haiming Liang. A survey on the fusion process in opinion dynamics. Information Fusion, 43:57–65, 2018.
  • [20] Serge Galam. Majority rule, hierarchical structures, and democratic totalitarianism: A statistical approach. Journal of Mathematical Psychology, 30(4):426–434, 1986.
  • [21] Serge Galam. Social paradoxes of majority rule voting and renormalization group. Journal of Statistical Physics, 61(3):943–951, 1990.
  • [22] Serge Galam and Taksu Cheon. Tipping points in opinion dynamics: a universal formula in five dimensions. Frontiers in physics, 8:446, 2020.
  • [23] Serge Galam. Heterogeneous beliefs, segregation, and extremism in the making of public opinions. Physical Review E, 71(4):046123, 2005.
  • [24] André CR Martins and Serge Galam. Building up of individual inflexibility in opinion dynamics. Physical Review E, 87(4):042807, 2013.
  • [25] Nuno Crokidakis and Paulo Murilo Castro de Oliveira. Inflexibility and independence: Phase transitions in the majority-rule model. Physical Review E, 92(6):062122, 2015.
  • [26] Mauro Mobilia. Nonlinear q-voter model with inflexible zealots. Physical Review E, 92(1):012803, 2015.
  • [27] H Eugene Stanley. Scaling, universality, and renormalization: Three pillars of modern critical phenomena. Reviews of modern physics, 71(2):S358, 1999.
  • [28] Alexander A Belavin, Alexander M Polyakov, and Al-exander B Zamolodchikov. Infinite conformal symmetry in two-dimensional quantum field theory. Nuclear Physics B, 241(2):333–380, 1984.
  • [29] Joaquín Marro and Ronald Dickman. Nonequilibrium phase transitions in lattice models. Cambridge University Press, 2005.
  • [30] Géza Ódor. Universality classes in nonequilibrium lattice systems. Reviews of modern physics, 76(3):663, 2004.
  • [31] Sven Lübeck. Universal scaling behavior of non-equilibrium phase transitions. International Journal of Modern Physics B, 18(31n32):3977–4118, 2004.
  • [32] H. Takayasu and A. Yu. Tretyakov. Extinction, survival, and dynamical phase transition of branching annihilating random walk. Phys. Rev. Lett., 68:3060–3063, May 1992.
  • [33] Martin J Howard and Uwe C Täuber. Real versus imaginary noise in diffusion-limited reactions. Journal of Physics A: Mathematical and General, 30(22):7721–7731, nov 1997.
  • [34] Géza Ódor. Critical behavior of the one-dimensional diffusive pair contact process. Physical review. E, Statistical, nonlinear, and soft matter physics, 67:016111, 01 2003.
  • [35] S S Manna. Two-state model of self-organized criticality. Journal of Physics A: Mathematical and General, 24(7):L363–L369, apr 1991.
  • [36] Ivan Dornic, Hugues Chaté, Jérôme Chave, and Haye Hinrichsen. Critical coarsening without surface tension: The universality class of the voter model. Phys. Rev. Lett., 87:045701, Jul 2001.
  • [37] Jef Hooyberghs, Ferenc Iglói, and Carlo Vanderzande. Strong disorder fixed point in absorbing-state phase transitions. Physical review letters, 90(10):100601, 2003.
  • [38] Gunter M Schütz. Reaction-diffusion processes of hard-core particles. Journal of statistical physics, 79(1):243–264, 1995.
  • [39] John A. Hertz. Quantum critical phenomena. Phys. Rev. B, 14:1165–1184, Aug 1976.
  • [40] Andrea Taroni, Steven T Bramwell, and Peter CW Holdsworth. Universal window for two-dimensional critical exponents. Journal of Physics: Condensed Matter, 20(27):275233, 2008.
  • [41] Jorge V José, Leo P Kadanoff, Scott Kirkpatrick, and David R Nel-son. Renormalization, vortices, and symmetry-breaking perturbations in the two-dimensional planar model. Physical Review B, 16(3):1217, 1977.
  • [42] Gary Mckeown and Noel Sheehy. Mass media and polarization processes in the bounded confidence model of opinion dynamics. JASSS, 9(1):11, 2006.
  • [43] André C. R. Martins. Mobility and social network effects on extremist opinions. Phys. Rev. E, 78:036104, 2008.
  • [44] Corinna Elsenbroich and Nigel Gilbert. Socially Situated Social Norms, pages 115–132. Springer Netherlands, Dordrecht, 2014.
  • [45] André C. R. Martins. Bayesian updating as basis for opinion dynamics models. AIP Conf. Proc., 1490:212–221, 2012.
  • [46] Lucas Böttcher, Jan Nagler, and Hans J. Herrmann. Critical behaviors in contagion dynamics. Physical Review Letters, 118:088301, 2017.
  • [47] Serge Galam and Taksu Cheon. Tipping points in opinion dynamics: A universal formula in five dimensions. Frontiers in Physics, 8:446, 2020.