Effect of social behaviors in the opinion dynamics -voter model
Abstract
Order-disorder phase transition is one of the interesting features in the opinion dynamics model of sociophysics. This phenomenon is caused by noise parameters, which can be treated as social responses such as anticonformity and independence (nonconformity) in a social context. This paper explores the effect of nonconformity behavior on changes in the macroscopic state of the system. A probability represents anticonformity and independence, and the model is defined on the complete graph and the two-dimensional (2D) square lattice. We also introduce a skepticism parameter represented by , which describes a voter’s propensity to behave nonconformist. Our results analytically and numerically show that the model undergoes a continuous and discontinuous phase for all values of non-zero and at specific -sized agent. No phase transition is observed in the model with independent voters on the 2D square lattice, and the model with anticonformist voters undergoes a continuous phase transition. We also analyze the effect of the probability and skepticism level on the outcome of the system’s opinion.
I Introduction
Over the past few decades, scholars have tried applying statistical physics’s general concepts and rules to interdisciplinary fields such as socio-political, economics, biology, medicine, computer science, and many more. Particularly in its application in understanding socio-political phenomena, scientists have proposed several models of opinion dynamics that explain the rules of interaction between individuals that influence each other to make an agreement or vice versa. The basis for developing an opinion dynamics model does not have standard rules; it can be in direct observation of real social phenomena or can be the form of reasonable imagination. To read further how the views of scientists regarding the application of statistical physics to social phenomena can refer to Refs. [1, 2, 3, 4].
In addition to their analysis of real socio-political phenomena, scientists have tried to understand further the physical phenomena that arise in the dynamics of opinion, such as the order-disorder phase transition in the Ising model. The order-disorder phase transition in the opinion dynamics model is often called the transition of the consensus-status quo at a critical point of a social parameter. The order-disorder phase transition is one of the interesting physical phenomena to study in opinion dynamics models. This phenomenon can explain the macroscopic changes in a system due to a complex interaction at the microscopic level. This phenomenon is usually caused by the presence of small changes in the fluctuation parameters of the system, such as the ferromagnetic-antiferromagnetic phenomenon due to temperature changes in the magnetic spin system, even in the opinion dynamics models of sociophysics [5, 6, 7, 8, 9, 10, 11, 12, 13]. In addition to phase changes due to small changes in noise parameters in the system; scientists have also observed scaling behavior and universality in opinion dynamics. Scaling can be interpreted as the collapse of data on a particular parameter, which we call parameter scaling. Likewise, the model’s universality is defined by several critical parameters that make the macroscopic parameters of the model collapse. The universality class for the mean-field model corresponds to the model defined on the complete graph (fully-connected network) [9, 10, 12, 13], as well as the opinion dynamics model which is defined on the two and three-dimensional [14], according to the universality class of the two and three-dimensional Ising models [15, 16].
In the context of a social system, several social parameters such as conformity, anticonformity, and independence are tried to be applied to observe social phenomena that arise and are interpreted based on the principles of physics. In the social literature, these three social parameters are generally grouped into conformity and nonconformity [17, 18, 19], extensively studied in the opinion dynamics model in recent years [9, 10, 12, 13, 20, 21, 22]. As mentioned, these social parameters make the dynamic models richer in statistical physics features produced and more related to social phenomena, as mentioned in the previous study [23], how the external field (related to the independence behavior) and the behavior of voters’ scepticism affect the opinion outcome.
This paper studies the opinion dynamics model discussed in the previous work [23], by adding a social anticonformist behavior represented by a probability . Several studies consider a similar scenario as in Ref. [23], namely external field as the influence of the mass media, which can affect the macroscopic state of the system, such as changes in consensus behavior [24, 25, 26, 27, 28]. The concept of this external field is an analogy of the effect of the external field on the Ising model, where it can affect the critical point that makes the system undergoes an order-disorder phase transition [9, 10, 12, 13, 29, 30, 14]. As mentioned by Civitarese [23], the influence of this external field uniformly affects all agents, meaning that all agents will change their opinion when exposed to the effects of the external field, for example, changing from “up” to “down” opinion (state) or vice versa.
We follow the scenario conducted by Civitarese [23]. However, as mentioned, this paper considers another social response called anticonformity and defines the model in a complete graph and 2D square lattices to analyze the order-disorder phase transition. This parameter is not discussed in the previous study. We also analyze the effect of the probability anticonformity and the skepticism level on the outcome of the system opinion. In addition, we analyze other statistical properties such as the exit probability, scaling parameters, and model universality class, which are lacking in Ref. [23]. Finally, the structure of this paper is as follows: Section I: Introduction, Section II: Model and method, Section III: Result and discussion, and Section IV: Final remark.
II Model
The nonlinear -voter model was introduced by Castellano et al. [31], is an extended model from the original voter [32]. The nonlinear -voter model based on the nonlinear form on the flip function where is the fraction of disagreeing neighbors or the opposite of the voter’s opinion, and is the probability that the agent changes the opinion whenever the neighbor is not in a homogeneous state. We can easily check that either for (any value of ) or , the function reduced to linear form, , as the original model. However, in fact, the nonlinear condition of the model can be obtained for and . In this paper, we extend the model by introducing “external fields” and and analyzing their effect on the macroscopic properties of the system.
The parameter in the flip function acts like noise and destroys the ordered state of the model, and it can be considered in different ways, for example, by relating them to social behaviors. In the social context, we borrow psychological terms called independence or anticonformity (contrarian), represented by the probability of the external field . Anticonformity is a social behavior that attempts to contradict the opinion of a majority group. Independence can be defined as a social behavior resistant to influence, acting according to one’s own will [17, 18]. Both of these behaviors belong to nonconformity behavior. Previous studies showed that the nonconformity behavior relates to the individualism index [33]. Although this nonconformity behavior is not directly related to the individualism index, in countries with a high individualism distance index, these two behaviors are commonly found [19, 34]. Further individualism index can be seen in [35].
As we stated previously, in this model, the independence and anticonformity behaviors are presented by the probability and the parameter can be considered as a skepticism level, how likely are the agents to behave anticonformist or independent. In other words, skepticism means that voters find it difficult to believe the persuasion of others, with representing the easiest the voters to behave nonconformist.
In the case of the two behaviors above, the model can be described as follows: (1) We generate the initial random state, where the density of fraction opinion (state) up is . (2) In that random state, we randomly chose a group of voters who formed a committee (-sized agent) and another voter who will follow the -sized agent. We generate a random number , and if is less than the probability independence , the voter acts independently. With probability , the voter behaves skeptically; with probability , the voters change their opinion. Therefore, the probability that the voters change their opinion from up to down or down to up is . Otherwise, the voter follows the standard rule of the -voter model, i.e., with probability , the voter adopts the -sized agent whenever the -sized agent is in a homogeneous state (conformity behavior). The illustration of the model is illustrated in (1).
independence | (1) | ||||||
conformity |
where stands for the state up (down) of the -sized agent, while stands for the state up (down) of the voter. The conformity model in 1 is the case for .
In the model with anticonformist voters, the voters change their opinion oppositely to the -sized agent whenever the voter and the -sized agent are homogeneous. In this case, the probability of a voter changing its state from up to down (down to up) is . Otherwise, the voter follows the standard rule of the -voter model. The illustration of the model is exhibited in (2).
anticonformity | (2) | ||||
conformity |
In the 2D square lattices, the voter has four neighbor agents (), similar to the Ising model in the 2D square lattices lattice [36].
The nonlinear -voter model with external field effect is considered in the previous work [23] and is similar to the independence type in this paper. This paper explores the model analytically and numerically for two types of social interaction, namely, independence and anticonformity, represented by probability . We study the model on the complete graph (mean-field approximation) and 2D lattices, especially to analyze the critical point and the critical exponent of the model using finite-size analysis, which is defined as and [37], where and are the fourth order Binder cumulant and susceptibility parameters, respectively [38]. The parameter is the dimensionless scaling function that controls the finite-size scaling effect, where and . In addition, the order parameter (magnetization) can be computed using
(3) |
since each agent has two possible opinions , represented by Ising spin and down .
In this model, we also analyze the effect of the skepticism on the outcome of probability anticonformity , which is not discussed in the previous study [23].
III Result and Discussion
III.1 Mean-field approximation
The mean-field approximation postulates that the system is homogeneous and isotropic. In other words, all fluctuation in the system is neglected, implying that local and global concentration can be considered the same [39, 40]. In addition, the system does not depend on the dimensions of space. At this point, we can describe the system’s state using one parameter of the system, such as the fraction opinion in this model. Regarding graph structure, the graph’s complexity is reduced in the mean-field regime, allowing us to analyze the system analytically more easily. Even though the complexity is reduced, we can observe statistical physics features such as continuous and discontinuous phase transitions, scaling, universality, relaxation time, and exit probability, which are parameters that have been studied frequently in recent years [12, 13, 41, 42, 43, 28].
Model with independent voters- An independent agent acts independently to change their opinion in the population; the agents do not follow the neighboring agents (-sized agent). We define a probability , a parameter that controls the independent behavior of the agents, and another parameter , which represents skepticism of the agents. In this model, with probability , agents act independently of other agents whenever agents influence the voter-agent, while with probability , they act skeptically to the agents’ opinions. With probability the voter-agent change its opinion . Therefore, the total probability that makes the agent-voter change its opinion is . To describe the dynamics of the model, we follow the procedure in Ref. [23], in defining the probability of fraction opinion increases and decreases and decreases for in finite system as:
(4) | ||||
(5) |
and for :
(6) | ||||
(7) |
Equations (4)-(7) describe the increasing (decreasing) fraction opinion density by or remain constant during the dynamics process with probability for a specific range of fraction opinion density . To test the accuracy of the analytical result with the M-C result, we can expand Eqs. (4)-(7) for any . However, the mean-field results are more suitable for large . Therefore, we are more interested in examining Eqs. (4)-(7) as:
(8) | |||
(9) |
where . It should be emphasized that because we are analyzing the model with the mean-field approach, it is more suitable when considering the system for a large population such in Eqs. (8)-(9) instead of Eqs.(4)-(7).
Model with anticonformist voters- In this model, with probability , the agent-voter act as anticonformist, while probability acts skeptically to the opinion’ -sized agent. When the -sized agent shares the same opinion, the voter adopts the opposite opinion to the opinion’ -sized agent . Similar to the model with independent voters, the probability of the fraction opinion up increases and decreases (in the finite system for ), respectively are:
(10) | ||||
(11) |
and for :
(12) | ||||
(13) |
Again for comparison to the numerical simulation, defined on the complete graph, we consider Eqs. (10)-(13) for a large population size . Therefore, Eqs. (10)-(13) reduced to more simple forms as:
(14) | |||
(15) |
One can see in Eqs. (8)-(9) and Eq. (14)-(15), the model reduces to the original -voter model for [31]. For the stationary condition of the fraction opinion, will be discussed in the next subsection.
III.2 Time evolution and steady state
The recursive formula of the fraction opinion at time can be driven from the discrete-time master equation, which can be written as:
(16) |
where is a transition probability of fraction opinion from state to at a single time . By combining the equation of the probability of finding fraction opinion at time , to Eq. (16), the recursive formula of the fraction opinion at a time step can be written as:
(17) |
The functions and are the probability density of fraction opinion density increases and decreases which defined in Eqs. (8)-(9) and Eq. (14)-(15). Note that the time step since the total time for step is . Therefore we can rewrite Eq. (17) in continue version by taking limit , or as:
(18) |
Equation Eq. (17) is also called the rate equation of fraction opinion , which describes the [44, 45].
The macroscopic behavior of the system can be analyzed easily by considering the equilibrium condition of Eq. (18) = 0, implies . Therefore, from Eqs. (8)-(9) and (18) the probability of independence at a stationary state is:
(19) |
Eq. (19) describes the relation of the probability and skepticism for stationary state for specific ranges of and and to be the same for . The transition of skepticism decreases non-smoothly at low and a large -sized agent and decreases smoothly for small and all values of -sized agent, with increases as and increase [23].
We can also observe the order-disorder phase transition of the model from the equilibrium condition of Eqs. (8)-(9). Thus, the independence probability at the equilibrium state can be written as:
(20) |
Note that Eqs. (19) and (20) are slightly different, where Eq. (19) describes the order parameter for a specific range of , while Eq. (20) describes the order parameter for all value of . The probability in Eq. (20) is the critical independence that makes the system undergoes an order-disorder phase transition. In Ref. [29], authors have shown that for , the model undergoes a discontinuous phase transition for and continuous phase transition for . In this model, we also show that model undergoes a discontinuous phase transition for and a continuous phase transition for for all values of . Although responsible for the occurrence of the phase transition, the parameter has no impact on changing the phase-transition kind; namely cannot change from continuous to discontinuous phase transitions, and vice versa, but only shifts the critical point of the system. Panel (a) in Fig. 1 shows the plot of Eq. (20) for several values of -sized agent and . One can see the model undergoes a continuous phase transition for and discontinuous for . Based on phenomena, in the social context, one can say that at , the system is in the consensus (there is an agreement), the complete consensus occurred when there is no independent agent , and the consensus decreases as increases until at the critical point , and at the system is in a status-quo or stalemate situation. As we can see that the skepticism also controls the critical point since for , the model is reduced to the model introduced in the previous study [29].

Similar to the model with independent voters, we are interested in analyzing the effect of the probability on the transition of the skepticism level . Therefore by considering the steady state, , the anticonformity probability at equilibrium state for specific range of can be written as:
(21) |
Based on Eq. (21), the stationary concentration is affected by the probability , skepticism level , -sized agent, and the initial concentration . Similar to the for , the transition of skepticism is smooth both for and for all range (see Fig. 2), which indicates the existence of the continuous phase transition in this model. The effect of probability to the skepticism differs from the model with independent voters, showing the non-smoothness (discontinuous) transition of for at low [23].

In the same way as the model with independent voters, by considering the stationary condition of Eq. (18), the anticonformity probability at the equilibrium state, which makes the system undergoes an order-disorder phase transition when can be written as:
(22) |
Plot of Eq. (22) for several values of -sized agents and is exhibited in Fig. 1, panel (b). One can see that the model undergoes a continuous phase transition for all values -sized agents. As shown in Eq. (22), besides depending on the -sized agent, the critical point also depends on the skepticism , i.e., it decreases as increases. For a specific case , the model is reduced to the one introduced in Ref. [29].
In general, Eqs. (20) and (22) can be written as (after re-scaling ), where for all values of . For example, for the model with independent voters, for , we have , where the critical point is , that is when . The is typical of the critical exponent of the mean-field Ising model for the magnetization that makes the best data collapse near the critical point [46]. We also find the same for the model with anticonformist voters based on Eq. (22). Other critical exponents corresponding to susceptibility and Binder cumulant will be obtained using M-C simulation in section the next section.
The time evolution of the fraction opinion can be analyzed from Eq. (17) or Eq. (18) for a large population version. One can see that, obtaining the exact solution of Eq. (18) for in time step , namely , for any values of -sized agent and will much need work. However, we can solve (17) numerically using for example, Runge-Kutta 4th order (R-K) [47] . The comparison of the M-C simulation (data point) to the R-K method (dashed line) for the model with independent voters for typical values of and is exhibited in Fig. 3. One can see that the M-C result agrees very well with the R-K method. Based on Fig. 3, for and the fraction opinion evolves to two stable states, while for and the fraction opinion evolves to three stable states, indicating the model undergoes a continuous phase transition for and discontinuous phase transition for as discussed on previous section [see Fig. 1 panel (a)]. Fig. 4 shows the time evolution of the fraction opinion for the model with anticonformist voter. In this model, the fraction opinion evolves to two stable states for all , indicating the model undergoes a continuous phase transition as discussed in the previous section [see Fig. (1) panel (b)]. The case of this stability can be analyzed from the effective potential and the probability of the density function of the system, which is discussed in the section.


III.3 Scaling behavior
Statistically, we can formulate the scaling parameter of the model that makes data collapse. We can follow the procedure introduced in Ref. [48] to find the scaling parameter for several macroscopic parameters of the model, such as the order parameter , susceptibility , and Binder cumulant . Firstly, we prepare the system’s initial state in a disordered state, i.e., the fraction opinion . We randomly select a group of a -sized agent and another agent (voter) interacting with each other based on the model. In this case, there are three rules followed by the voter. Rule (1): With probability , the voter acts as a nonconformist (anticonformist or independent), and with probability , the voters act skeptical. Rule (2): With probability , the voter acts as a conformist. For the model with anticonformist voters, the voters will change their opinion whenever they have the same opinion as the -sized agent (homogeneous state). For the model with independent voters, with probability , the voter changes their opinion. Thus the total probability of the voters changing their opinion is for the model with anticonformist voters and for the model with independent voters. Rule (3) with probability and , the voters do not change their opinion, respectively. Therefore, rule (3) does not change the system’s state. Then, based on these rules, the scaling parameter depends on the probability of nonconformist and skepticism .
During the dynamic process, the voters follow one of the three rules above and make the sequence states where the state depends on the initial condition and the probability rules (1) and (2) with ratio , where probability rule (1) and generally depends on the model. We can say that if we have two systems with the same ratio , they are statistically the same. To compare the two systems, we need to obtain the average time during the dynamic process as a function of probability rules (1) and (2). The average time of two systems is inversely proportional to the total probability that makes the voters change their opinion [48]. If two systems have the same ratio probability and state at an instant time , then the macroscopic parameters of the systems will be the same, namely and . Therefore, by plotting
(23) |
The data will be collapsed. Theoretically, these scaling parameters work for any system which statistically follows the rules (1), (2), and (3). This paper will test the model on the complete graph and the 2D square lattice, which will be discussed in the next section.
III.4 Landau potential
The Landau theory of free energy can be used to analyze an equilibrium phase transition of a thermodynamics system [49, 50]. In his hypothesis, Landau argued that the free energy could be expanded in power series near the critical point in terms of the order parameter. Landau potential also can be applied to analyze a nonequilibrium system such as the Langevin equation for two absorbing states using mean-field approximation [51, 52]. Generally, the Landau potential is not only described by thermodynamic parameters such as temperature, pressure, and volume but can also depend on the order parameters of the system as in Eq. (3). Here, we use the Landau potential to analyze the order-disorder phase transition of the model. Based on the Landau theory, a potential can be written as , where is defined as in Eq. (3), and in general as a function of the macroscopic parameters of the system such as the probability , and the skepticism . The potential is symmetric under inversion ; therefore, the odd terms will vanish.
To analyze the order-disorder phase transition of the model, we only need to consider at least two even terms of the potential as:
(24) |
where is defined in Eq. (3), and depend of the model. The maximum-minimum conditions that describe the order-disorder phase transitions can be analyzed from the extremum conditions of Eq. (24). In addition, the continuous and discontinuous phase conditions can be analyzed by selecting the appropriate values of and . For , both continuous and discontinuous phase transition conditions occur. At the critical point, the continuous and discontinuous phase occurs for and , respectively.
To obtain the effective potential of the system, we will use the probability of the fraction opinion increases and decreases as shown in Eqs. (8) and (9) for the model with independent voters and Eqs. (14) and (15) for the model with anticonformist voters to define an effective force of the model. In general, the effective potential of the system can be obtained as , where is called an effective force that drives the spins to flip [53].
The general solution of the effective potential for the model with independent voters is:
(25) |
The plot of Eq. (III.4) for typical values of and probability is exhibited in Fig. 5. As seen for and , the model undergoes a continuous phase transition characterized by the bistable potential for (dashed line), and the potential is monostable for . However, three stable states for and indicate that the model undergoes a discontinuous phase transition. These potentials confirm the time evolution of the fraction opinion in Fig. 3; that is, the fraction opinion evolves to two stable states for and and three stable states for and .

For the model with independent voters, the general solution of potential is:
(26) |
The plot of Eq. (III.4) can be seen in Fig. 6. For all values of , the potential has two stable states for and monostable for , indicating the model undergoes a continuous phase transition. This potential also confirms the time evolution of the fraction opinion in the previous section (see Fig. 4).
The effective potential in Eqs. (III.4) and (III.4) maximum at , that is , with two bistable states, minimum at , that is , with one stable state, and the transition maximum-minimum at , that is . Therefore we can obtain the critical point of the system as follows:
(27) |
for the model with independent voters, and
(28) |
for the mode with anticonformist voters.
The potential in Eqs. (III.4) and (III.4) are functions of the fraction opinion . However, we can easily convert as a function of the order parameter by substituting into equations (III.4) and (III.4) to get a similar form as the Landau potential in Eq. (24). Therefore, for the model with the anticonformist agents, we obtain
(29) | ||||
(30) |
We have the same result as Eq. (27) for the critical point by setting . The parameter is always positive for (typically a continuous phase transition) and negative for (typically a discontinuous phase transition) for all values of .
(31) | ||||
at and for all and . | (32) |
As seen in Eq. (31), we also find the same for as in Eq. (28). The parameter is always positive for all values of and , indicating the typical of the continuous phase transition for all and .

III.5 Fokker-Planck description
In the previous section, we analyzed the model’s order-disorder phase transition by considering the equilibrium condition of and the effective potential . Here, we analyze the order-disorder phase transition of the model by analyzing the stationary of the probability density function, which can be obtained from the Fokker-Planck (F-P) equation below [54]:
(33) |
where and are the diffusion-like and drift-like coefficients which depend on the model. As we can see, Eq. (33) is just a one-dimensional of the F-P equation since the model is defined on the complete graph. The diffusion-like and drift-like coefficients for the model with independent voters, respectively, are:
(34) | ||||
(35) |
and for the model with anticonformist voters, the diffusion-like and drift-like coefficients, respectively, are:
(36) | ||||
(37) |
We only need the stationary condition of Eq. (33) to analyze the order-disorder phase transition. By a little mathematical modification, the general solution of Eq. (33) for the stationary condition can be written as:
(38) |
where is a constant that satisfied the normalized condition of . One can see that obtaining the exact solution of Eq. (38) for any values of the -sized agent, probability , and are difficult to carry out. However, we can solve numerically Eq. (38) for typical values of , and . The plot of Eq. (38) for the model with independent voters is exhibited in Fig. 7. One can see the typical of the discontinuous phase transition, namely has three maxima corresponding to three stable states on the effective potential in Eq. (III.4). For low , has two maxima, and the maxima appearing in is getting higher as increases. For high , the probability density function has only one maximum at , indicating the system is stable at .

The plot of Eq. (38) for the model with anticonformist voters is exhibited in Fig. 8. In this model, the probability density function has two maxima at low corresponding to two stable states in the effective potential in Eq. (III.4). The ‘peaks’ of are getting lower as increases and only one maximum for at . This corresponds to the occurrence of a continuous phase transition.

III.6 Exit probability
Exit probability is of the parameters that compute in social dynamics. Exit probability can be defined as the probability that the system is in a homogeneous state with one of two homogeneous states. In this paper, we consider the model to reach a homogeneous state with , namely the probability of all agents having the same state up in the end. The differential equation of the exit probability can be derived from the F-P equation as [55]:
(39) |
with boundary conditions and , since the probability of the system in opinion up is zero when and one when .
Theoretically, we can solve Eq. (39) exactly but obtaining in any values of and can need much work. However, in this paper, we solve Eq. (39) numerically using finite-different method [47]. As an illustration of the exact solution of Eq. (39), we can solve it easily for a simple case, for example, and without anticonformist voter (), we have . One can check that the exit probability will coincide in for all values of , indicating the existence of the separator point , which separates two ordered states and , with the same probability as shown in panel (a) of Figs. 9. It is easy to understand that for the case without the nonconformist voter (), all agents reach a consensus state depending on the state of the initial number of fractions opinion . If the fraction opinion , then the system will reach a full consensus state with all agents having the same state or opinion up (down) [28].

The plot of Eq. (39) for typical values of and for the model with independent voters is shown in Figs. 9 [panel (b)]. We also obtain that the exit probability of the model coincides at , indicating the separation of two consensus states with the same probability at . One can see that the effect of the probability does not affect the ‘symmetry form’ of the exit probability to the fraction opinion because the probability ‘acts’ homogeneous in reaching the order-disorder phase transition of the model. Therefore the probability of the system reaching two consensus states are the same. The same result is obtained for the model with anticonformist voters (not shown). It can also be seen in Fig. 1 that the fraction opinion is symmetrical to either the independence or anticonformity probability .
We can also obtain the scaling parameters between the exit probability versus fraction opinion using the standard finite-size scaling , and [56]. As seen in the main panels of Fig. 9, the data are collapsed for and . These and are universal; we obtain the same value for all values of N. These scaling parameters are the same as the majority and the -voter models with mass media effect, defined on the complete graph [28, 27].
III.7 Scaling and critical exponents
III.7.1 The model on the complete graph
Model with independent voters. We perform numerical simulation for the model with several population sizes , a typical value of -sized agent, and scepticism level . We variate population size to analyze the order parameter , susceptibility , and Binder cumulant and obtain the critical exponents using the finite-size relations to define the universality class of the model. The result for is exhibited in Fig. 10. The model undergoes a continuous phase transition with the critical point obtained from the crossing lines of Binder cumulant versus probability , which occurred at [see panel (c)]. This result agrees to the analytical treatment in Eq. (20), which is , where . The inset graphs show the scaling plot near the critical point , and all the data are collapsed for and . These scaling parameters indicate the model’s universality class is the same as the Sznajd model [9, 12], the majority rule [13], the kinetic exchange model [57, 45], as well as the mean-field Ising universality class. These scaling parameters work for all values of scepticism level and -sized agent .

Model with anticonformist voters. The simulation result of this model is exhibited in Fig. 11. One can see that the critical point that makes the model undergo a continuous phase transition is [see panel (c)]. Using the finite-size scaling relations, we also obtain the same and as the model with independent voters. These critical exponents indicating the model is identical to the model with independent voters and belongs to the mean-field Ising universality class.

Fig. 12 shows the comparison between numerical simulation versus analytical treatment in Eq. (22) for , several values of skepticism level and population size . It can be seen that the numerical simulation is in well-agreement to the analytical treatment in Eq. (20) after re-scaling [see the biggest panel]. Three small panels show the scaling plot of the order parameter , susceptibility , and Binder cumulant based on Eq. (23) and show the data collapse for all .

III.7.2 The model on the 2D square lattice
Model with independent voters. In this model, we consider the voter has four neighbors and follows the microscopic rules as the model, namely the voter adopts the opinion of the four neighbors whenever they have the same opinion. Otherwise, the voter acts independently. We consider several population sizes , where is the width size lattice with periodic boundary conditions. The result for order parameter [panel (a)], susceptibility [panel (b)], and Binder cumulant is exhibited in Fig. (13). One can see no crossing lines between Binder cumulant versus probability indicating the order-disorder phase transition is absent in this model.

Model with anticonformist voters. The simulation result of this model is exhibited in Fig. 14. It can be seen that the critical point that makes the model undergo a continuous phase transition is . Using the finite-size scaling analysis, the critical exponents that make the best collapse of all data are (see the inset graphs). As we can see, the critical exponents of the model are not the same as the critical exponents of the 2D Ising model. Therefore the model is not identical to the 2D Ising model. Although the model is defined on the 2D square lattice, it does not necessarily has the same universality class as the 2D Ising model.

As mentioned in the previous section, Eq. (23) works for all models that statistically follow the rules. Here, we perform numerical simulation for the model with the anticonformist agent for several values of skepticism level as shown in Fig. 15. The left panels are the normal plot, and the right panels are the scaling plot of the data. As we can see in the right panels, all data collapse and works for all values of .

The effect of the skepticism decreases the critical point as increases through the relation , which is obtained from Eq. (20) for . In general, the relation of the critical point to the skepticism parameter can be written as:
(40) |
where is a constant that depends on the -sized agent, namely, for the model with independent voters, we have and for the model with anticonformist voters, we have .
Fig. 16 shows the behavior of the critical point over -sized agent for several values of skepticism level based on Eqs. (27) and (28) for the model with independent and anticonformist voter, respectively. For the model with independent voters, the point decreases as increases [panel (a)], while for the model with anticonformity [panel (b)], the critical point increases as increases. Both models are the same for ; there is no order-disorder phase transition.

IV Final Remark
This paper considers the nonlinear -voter model defined on a complete graph and a 2D square lattice. We consider two social behaviors, namely anticonformity and independence, and analyze the effect of both social behaviors on the macroscopic phenomena of the model. We also introduce a parameter , which represents the skepticism level of the voter to be either independent or anticonformist. The social behaviors are represented by a probability that acts like a noise that makes the model undergoes an order-disorder phase transition.
Based on our results for the model on the complete graph in both analytical and numerical simulation, the model with anticonformist voters undergoes a continuous phase transition for all values of and skepticism level , while the model with independent voters undergoes a continuous phase transition for and a discontinuous phase transition for for all values of . The order-disorder phase transition is also analyzed from the effective potential and the stationary probability density function of the fraction opinion and shows the same results. In addition, we observe that the outcome of the fraction opinion increases as the probability of anticonformist and skepticism level increase. We can see that the skepticism transition is smooth for all ranges of anticonformist probability , in contrast to what has been reported for the model with independence in the previous study [23].
The models are defined on the complete graph for the models with independent voters and anticonformist voters; based on the critical exponents obtained using finite-size scaling analysis, they are identical and have the same universality class as the Ising mean-field model. For the model in the 2D square lattice, we observe that the order-disorder phase transition is absent for the model with independent voters and undergoes a continuous phase transition for the model with anticonformist voters with the critical point at . We obtain the critical exponents that make the best collapse of all data, namely and .
Our results show that regardless of the microscopic interaction form of opinion dynamics models, if they are defined in a fully connected network, they are identical to the mean-field Ising model. This case is not always true for opinion dynamic models on a 2D lattice; they are not necessarily identical to the 2D Ising models.
For the model on the complete graph, we analyze the exit probability, the probability of the model is in the consensus state. The separator point of the model is , a point that separates two equilibrium states at and . This point separates two consensus states, the consensus state with and , with the same probability. We also obtain the scaling parameter and based on the standard finite-size scaling that makes the best collapse of all data. On the other hand, when we examine two models with interaction microscopically different, if they are statistically the same, then the scaling behavior of these two models is identical. This scaling behavior has been proven in four scenarios/models in two different networks.
Based on the social point of view, one can say that the existence of this nonconformist behavior (represented by a probability ) can disrupt social cohesion. However, the impact is greater on the systems with independent voters than those with anticonformist voters. When this disruptive behavior (nonconformity) is absent, a unanimous agreement (complete consensus) within a population will always be reached. This unanimous agreement will continue to be distorted as this disruptive behavior strengthens. However, this agreement continues to exist. In other words, majority and minority opinions side by side up to a critical point . One could say nonconformist critical behavior. When , the system will always be in status quo.
The skepticism adopted by independent and anticonformist agents has the same impact on the critical point . When this skepticism increases, the tendency of the system to undergo the status quo will be even greater, both in models with independent voters and anticonformist voters. As mentioned previously, the difference is that independent agents have a stronger impact on bringing the system into a status quo than anticonformist agents. In other words, these results corroborate the fact that the existence of contrary socio-political entities is more stable than independent-disorganized socio-political entities. These results can provide an overview for policymakers to provide their policies more carefully.
Author contribution
R. Muslim: Main contributor, Conceptualization, Methodology, Investigation, Formal analysis, Software, Data Curation, Visualization, Writing – original draft H. Lugo H.: Software, Formal analysis, Visualization, Validation.
Acknowlegment
R. Muslim thanks the Quantum Matter Theory Group of the BRIN Research Center for Quantum Physics for its mini HPC so that the numerical data in this paper can be obtained.
Declaration of Interests
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
References
- Galam [2012] S. Galam, Sociophysics: A Physicist’s Modeling of Psycho-political Phenomena (Springer-Verlag, New York, 2012).
- Castellano et al. [2009a] C. Castellano, S. Fortunato, and V. Loreto, Statistical physics of social dynamics, Rev. Mod. Phys. 81, 591 (2009a).
- Sen and Chakrabarti [2014] P. Sen and B. K. Chakrabarti, Sociophysics: an introduction (Oxford, Oxford University Press, 2014).
- Schweitzer [2018] F. Schweitzer, Sociophysics, Physics today 71, 40 (2018).
- Hołyst et al. [2000] J. A. Hołyst, K. Kacperski, and F. Schweitzer, Phase transitions in social impact models of opinion formation, Physica A: Statistical Mechanics and its Applications 285, 199 (2000).
- Li et al. [2012] L. Li, A. Scaglione, A. Swami, and Q. Zhao, Phase transition in opinion diffusion in social networks, in 2012 IEEE international conference on Acoustics, speech and signal processing (ICASSP) (IEEE, 2012) pp. 3073–3076.
- Mukherjee and Chatterjee [2016] S. Mukherjee and A. Chatterjee, Disorder-induced phase transition in an opinion dynamics model: Results in two and three dimensions, Physical Review E 94, 062317 (2016).
- Velásquez-Rojas and Vazquez [2017] F. Velásquez-Rojas and F. Vazquez, Interacting opinion and disease dynamics in multiplex networks: discontinuous phase transition and nonmonotonic consensus times, Physical Review E 95, 052315 (2017).
- Muslim et al. [2020] R. Muslim, R. Anugraha, S. Sholihun, and M. F. Rosyid, Phase transition of the sznajd model with anticonformity for two different agent configurations, Int. J. Mod. Phys. C 31, 2050052 (2020).
- Muslim et al. [2021] R. Muslim, R. Anugraha, S. Sholihun, and M. F. Rosyid, Phase transition and universality of the three-one spin interaction based on the majority-rule model, International Journal of Modern Physics C 32, 2150115 (2021).
- Schawe and Hernández [2022] H. Schawe and L. Hernández, Higher order interactions destroy phase transitions in deffuant opinion dynamics model, Communications Physics 5, 32 (2022).
- Muslim et al. [2022a] R. Muslim, M. J. Kholili, and A. R. Nugraha, Opinion dynamics involving contrarian and independence behaviors based on the sznajd model with two-two and three-one agent interactions, Physica D: Nonlinear Phenomena 439, 133379 (2022a).
- Muslim et al. [2022b] R. Muslim, S. A. Wella, and A. R. Nugraha, Phase transition in the majority rule model with the nonconformist agents, Physica A: Statistical Mechanics and its Applications 608, 128307 (2022b).
- Calvelli et al. [2019] M. Calvelli, N. Crokidakis, and T. J. Penna, Phase transitions and universality in the sznajd model with anticonformity, Physica A: Statistical Mechanics and its Applications 513, 518 (2019).
- Stanley [1971] H. E. Stanley, Phase transitions and critical phenomena, Vol. 7 (Clarendon Press, Oxford, 1971).
- Campostrini et al. [2002] M. Campostrini, M. Hasenbusch, A. Pelissetto, P. Rossi, and E. Vicari, Critical exponents and equation of state of the three-dimensional heisenberg universality class, Physical Review B 65, 144520 (2002).
- Willis [1963] R. H. Willis, Two dimensions of conformity-nonconformity, Sociometry , 499 (1963).
- Willis [1965] R. H. Willis, Conformity, independence, and anticonformity, Hum. Relat. 18, 373 (1965).
- Hofstede et al. [2010] G. Hofstede, G. J. Hofstede, and M. Minkov, Cultures and organizations: Software of the mind (New York, McGraw Hill, 2010).
- Grabisch and Rusinowska [2020] M. Grabisch and A. Rusinowska, A survey on nonstrategic models of opinion dynamics, Games 11, 65 (2020).
- Baron [2021] J. W. Baron, Consensus, polarization, and coexistence in a continuous opinion dynamics model with quenched disorder, Physical Review E 104, 044309 (2021).
- Lipiecki and Sznajd-Weron [2022] A. Lipiecki and K. Sznajd-Weron, Polarization in the three-state q-voter model with anticonformity and bounded confidence, Chaos, Solitons & Fractals 165, 112809 (2022).
- Civitarese [2021] J. Civitarese, External fields, independence, and disorder in q-voter models, Physical Review E 103, 012303 (2021).
- Candia and Mazzitello [2008] J. Candia and K. I. Mazzitello, Mass media influence spreading in social networks with community structure, Journal of Statistical Mechanics: Theory and Experiment 2008, P07007 (2008).
- Martins et al. [2010] T. V. Martins, M. Pineda, and R. Toral, Mass media and repulsive interactions in continuous-opinion dynamics, EPL (Europhysics Letters) 91, 48003 (2010).
- Crokidakis [2012] N. Crokidakis, Effects of mass media on opinion spreading in the sznajd sociophysics model, Physica A: Statistical Mechanics and its Applications 391, 1729 (2012).
- Azhari and Muslim [2022] A. Azhari and R. Muslim, The external field effect on the opinion formation based on the majority rule and the -voter models on the complete graph, International Journal of Modern Physics C (2022).
- Muslim et al. [2023] R. Muslim, R. A. NQZ, and M. A. Khalif, Mass media and its impact on opinion dynamics of the nonlinear -voter model, arXiv preprint arXiv:2305.06053 (2023).
- Nyczka et al. [2012a] P. Nyczka, K. Sznajd-Weron, and J. Cisło, Phase transitions in the q-voter model with two types of stochastic driving, Physical Review E 86, 011105 (2012a).
- Vieira and Crokidakis [2016] A. R. Vieira and N. Crokidakis, Phase transitions in the majority-vote model with two types of noises, Physica A: Statistical Mechanics and its Applications 450, 30 (2016).
- Castellano et al. [2009b] C. Castellano, M. A. Muñoz, and R. Pastor-Satorras, Nonlinear q-voter model, Physical Review E 80, 041129 (2009b).
- Clifford and Sudbury [1973] P. Clifford and A. Sudbury, A model for spatial conflict, Biometrika 60, 581 (1973).
- Hofstede [2001] G. Hofstede, Culture’s consequences: Comparing values, behaviors, institutions and organizations across nations (California, Sage publications, 2001).
- Solomon et al. [2010] M. R. Solomon, G. Bamossy, S. Askegaard, and M. K. Hogg, Consumer behavior: a European perspective 4th edn (England, Pearson Education, 2010).
- [35] See https://www.hofstede-insights.com/country-comparison for further details about IDV of various countries in the world.
- McCoy and Wu [1973] B. M. McCoy and T. T. Wu, The two-dimensional Ising model (Harvard University Press, 1973).
- Cardy [1996] J. Cardy, Scaling and renormalization in statistical physics, Vol. 5 (Cambridge, Cambridge University Press, 1996).
- Binder et al. [2012] K. Binder, D. M. Ceperley, J.-P. Hansen, M. Kalos, D. Landau, D. Levesque, H. Mueller-Krumbhaar, D. Stauffer, and J.-J. Weis, Monte Carlo methods in statistical physics, Vol. 7 (Springer Science & Business Media, 2012).
- Binney et al. [1992] J. J. Binney, N. J. Dowrick, A. J. Fisher, and M. E. Newman, The theory of critical phenomena: an introduction to the renormalization group (Oxford University Press, 1992).
- Amit and Martin-Mayor [2005] D. J. Amit and V. Martin-Mayor, Field theory, the renormalization group, and critical phenomena: graphs to computers (World Scientific Publishing Company, 2005).
- Mili et al. [2022] I. Mili, H. Latelli, Z. Charifi, H. Baaziz, and T. Ghellab, A simple formula for calculating the carrier relaxation time, Computational Materials Science 213, 111678 (2022).
- Rocha et al. [2021] G. S. Rocha, G. S. Denicol, and J. Noronha, Novel relaxation time approximation to the relativistic boltzmann equation, Physical review letters 127, 042301 (2021).
- Li and Li [2020] J. Li and S. Li, Some state-specific exit probabilities in a markov-modulated risk model, Mathematical Problems in Engineering 2020, 1 (2020).
- Hinrichsen [2000] H. Hinrichsen, Non-equilibrium critical phenomena and phase transitions into absorbing states, Advances in physics 49, 815 (2000).
- Krapivsky et al. [2010] P. L. Krapivsky, S. Redner, and E. Ben-Naim, A kinetic view of statistical physics (Cambridge University Press, 2010).
- Gitterman [2013] M. Gitterman, Phase transitions: Modern applications (World Scientific Publishing Company, 2013).
- Pinder [2018] G. F. Pinder, Numerical methods for solving partial differential equations: a comprehensive introduction for scientists and engineers (John Wiley & Sons, 2018).
- Sznajd-Weron et al. [2011] K. Sznajd-Weron, M. Tabiszewski, and A. M. Timpanaro, Phase transition in the sznajd model with independence, Europhys. Lett. 96, 48002 (2011).
- Landau [1937] L. D. Landau, On the theory of phase transitions, Zh. Eksp. Teor. Fiz. 7, 19 (1937).
- Plischke and Bergersen [1994] M. Plischke and B. Bergersen, Equilibrium statistical physics (World scientific, 1994).
- Al Hammal et al. [2005] O. Al Hammal, H. Chaté, I. Dornic, and M. A. Munoz, Langevin description of critical phenomena with two symmetric absorbing states, Physical review letters 94, 230601 (2005).
- Vazquez and López [2008] F. Vazquez and C. López, Systems with two symmetric absorbing states: relating the microscopic dynamics with the macroscopic behavior, Physical Review E 78, 061127 (2008).
- Nyczka et al. [2012b] P. Nyczka, J. Cisło, and K. Sznajd-Weron, Opinion dynamics as a movement in a bistable potential, Physica A 391, 317 (2012b).
- Frank [2005] T. D. Frank, Nonlinear Fokker-Planck equations: fundamentals and applications (Berlin, Springer, 2005).
- Gardiner et al. [1985] C. W. Gardiner et al., Handbook of stochastic methods, Vol. 3 (springer Berlin, 1985).
- Sousa et al. [2008] A. Sousa, T. Yu-Song, and M. Ausloos, Effects of agents’ mobility on opinion spreading in sznajd model, The European Physical Journal B 66, 115 (2008).
- Crokidakis [2014] N. Crokidakis, Phase transition in kinetic exchange opinion models with independence, Physics Letters A 378, 1683 (2014).