Loading [MathJax]/jax/output/SVG/jax.js

Opinion fitness and convergence to consensus in homogeneous and heterogeneous populations

  • Received: 01 August 2020 Revised: 01 November 2020 Published: 07 February 2021
  • 91C20; 82B21; 60K35

  • In this work we study the formation of consensus in homogeneous and heterogeneous populations, and the effect of attractiveness or fitness of the opinions. We derive the corresponding kinetic equations, analyze the long time behavior of their solutions, and characterize the consensus opinion.

    Citation: Mayte Pérez-Llanos, Juan Pablo Pinasco, Nicolas Saintier. Opinion fitness and convergence to consensus in homogeneous and heterogeneous populations[J]. Networks and Heterogeneous Media, 2021, 16(2): 257-281. doi: 10.3934/nhm.2021006

    Related Papers:

    [1] Mayte Pérez-Llanos, Juan Pablo Pinasco, Nicolas Saintier . Opinion fitness and convergence to consensus in homogeneous and heterogeneous populations. Networks and Heterogeneous Media, 2021, 16(2): 257-281. doi: 10.3934/nhm.2021006
    [2] Sabrina Bonandin, Mattia Zanella . Effects of heterogeneous opinion interactions in many-agent systems for epidemic dynamics. Networks and Heterogeneous Media, 2024, 19(1): 235-261. doi: 10.3934/nhm.2024011
    [3] Sharayu Moharir, Ananya S. Omanwar, Neeraja Sahasrabudhe . Diffusion of binary opinions in a growing population with heterogeneous behaviour and external influence. Networks and Heterogeneous Media, 2023, 18(3): 1288-1312. doi: 10.3934/nhm.2023056
    [4] Clinton Innes, Razvan C. Fetecau, Ralf W. Wittenberg . Modelling heterogeneity and an open-mindedness social norm in opinion dynamics. Networks and Heterogeneous Media, 2017, 12(1): 59-92. doi: 10.3934/nhm.2017003
    [5] Yuntian Zhang, Xiaoliang Chen, Zexia Huang, Xianyong Li, Yajun Du . Managing consensus based on community classification in opinion dynamics. Networks and Heterogeneous Media, 2023, 18(2): 813-841. doi: 10.3934/nhm.2023035
    [6] Nastassia Pouradier Duteil . Mean-field limit of collective dynamics with time-varying weights. Networks and Heterogeneous Media, 2022, 17(2): 129-161. doi: 10.3934/nhm.2022001
    [7] Sergei Yu. Pilyugin, Maria S. Tarasova, Aleksandr S. Tarasov, Grigorii V. Monakov . A model of voting dynamics under bounded confidence with nonstandard norming. Networks and Heterogeneous Media, 2022, 17(6): 917-931. doi: 10.3934/nhm.2022032
    [8] Marina Dolfin, Mirosław Lachowicz . Modeling opinion dynamics: How the network enhances consensus. Networks and Heterogeneous Media, 2015, 10(4): 877-896. doi: 10.3934/nhm.2015.10.877
    [9] Sergei Yu. Pilyugin, M. C. Campi . Opinion formation in voting processes under bounded confidence. Networks and Heterogeneous Media, 2019, 14(3): 617-632. doi: 10.3934/nhm.2019024
    [10] Catherine Choquet, Ali Sili . Homogenization of a model of displacement with unbounded viscosity. Networks and Heterogeneous Media, 2009, 4(4): 649-666. doi: 10.3934/nhm.2009.4.649
  • In this work we study the formation of consensus in homogeneous and heterogeneous populations, and the effect of attractiveness or fitness of the opinions. We derive the corresponding kinetic equations, analyze the long time behavior of their solutions, and characterize the consensus opinion.



    In the last few years, several physicists and mathematicians devoted their attention to opinion dynamics. Different techniques were used, depending on the specific problem. For instance, computer simulations are preferred when agents contact through networks, although they appear also in mean field problems [5,15,32,33,37], Markov processes and other probabilistic tools for finitely many opinions [17,18,21], large systems of ordinary differential equations for active particles, and the associated Boltzmann type equations for the evolution of several observables [12,20,27,29,31,35,38], together with their Fokker-Planck limits. We can cite the books [8,9,19,28], and the surveys [1,24] for details.

    Despite these different tools, the microscopic interactions among agents are based on sociological theories like social impact and social pressure theory [3,13,23], where agents modify their opinions trying to fit in some social group, and the persuasive argument theory [11,25], where the new opinion appears after an interchange of arguments among agents.

    However, the space of opinions is typically assumed as homogeneous, without considering the advantages of holding some specific opinions. Recently, in [30] we have proposed a model where the opinions are weighted by a function λ(w) representing the attractiveness, advantages, or fitness of opinion w. If we call w, w the actual opinions of two interacting agents, the new opinions w, w are obtained from the following microscopic interaction rule

    w=w+γλ(w)(ww)w=w+γλ(w)(ww). (1)

    where 0<γ<1/2 is a fixed parameter related to the strength of the interaction. Notice that the new opinion w is obtained moving w toward w, thus implementing the tendency to compromise. Moreover the magnitude of the change is proportional to λ(w) i.e. to the attractiveness of opinion w.

    Let us mention briefly that in [6,7,22] there are two competing opinions, say ±L, and an attitude spectrum A={±1,±2,,±L} which represents the strength of the opinion of an individual or its degree of conviction. Now, the behavior of agents depends on their attitudes, and it is not a characteristic associated with an opinion. A similar mechanism can be found in [2,38], where the microscopic interaction rule is given by

    w=w+γP(w)(ww)w=w+γP(w)(ww), (2)

    where P is usually equal to zero at the boundary of the space of opinions, representing that extreme opinions are difficult to change. Despite the apparent similarity between (1) and (2), both dynamics are very different, see [30].

    In this work we study the long time behavior of a population interacting through rule (1), and we add more heterogeneity on the agents by introducing two parameters p and q modeling the power of persuasion and the stubbornness of each agent, as in [31].

    In Section §2 we introduce the model and the mean field first order equation, satisfied by the density of agents in the space of opinions. In Section §3 we focus on the effect of the fitness of opinions in an homogeneous population, and we derive the precise value of the consensus opinion m, namely

    m=Λ1(Λ(w)df0(w))

    where Λ is a primitive of λ, and f0 is the initial distribution of agents on the space of opinions. Let us mention that we need λ>c>0 for some constant c. This hypotheses cannot be relaxed, due to the singular behavior of the population located near the zeros of λ, and a symmetry breaking phenomena when λ(m)=0. In Section §4 we analyze the formation of consensus in a non-homogeneous population. We include in Section §5 some agent based simulations, showing a good agreement for a finite population of the results predicted by the mean field equation.

    Let us introduce our model of opinion formation. We consider a population composed of infinitely many agents. The opinion of an agent with respect to some statement is represented by a real number w[1,1] (meaning 1 being completely in disagreement with the statement and 1 in complete agreement).

    In addition, we take into account the ability (or difficulty) of an individual to persuade another agent, denoted by p[0,1]; if p=0 the agent has no persuasion at all, while p=1 corresponds to perfect orators. We also take into account the stubbornness of an agent denoted by q[0,1], where q=0 corresponds to a stubborn agent, who will never be affected by other's opinion, and q=1 entails a very volatile agent who always takes into account other agents' opinions.

    Each agent is thus characterized by the three parameters (w,p,q). Moreover, we assume that the power of persuasion p and the stubbornness q remain fixed for each of the agents, who may modify their opinion w after binary encounters.

    Our model also takes into account the possibility that the fitness of an individual opinion may affect other agents. We introduce an influence function λ(w), with w[1,1], representing the influence exerted by an individual with opinion w. Usual examples of influence functions are

    ● Quadratic : λ(w)=|w|2;

    ● Linear : λ(w)=|w|;

    ● Uniform : λ(w)=1;

    ● Co-Linear : λ(w)=1|w|;

    ● Co-Quadratic : λ(w)=(1|w|)2.

    Individuals which held extreme opinions will have more influence under the linear and quadratic functions (extremist influence functions), while the co-linear and co-quadratic functions endow more influence to agents with moderate opinions, (centrist influence functions).

    We now describe the up-dating rules of the opinions. Consider two interacting agents with parameters (w,p,q) and (w,p,q) before the encounter. Denote by (w,p,q) and (w,p,q) the new values for the parameters after the interaction, respectively. As we mentioned before, the parameters p,p,q,q will remain unchanged: p=p, p=p, q=q and q=q. Regarding the up-dating of the opinion, we propose the following rule:

    w=w+γqpλ(w)(ww),w=w+γqpλ(w)(ww). (3)

    Observe that the term γλ(w)p(ww) reflects the tendency to compromise, which is proportional to the power of persuasion, p, and the influence exerted by the other agent opinion through λ(w), as well as her/his own stubbornness, q. Here, γ is a given real number in (0,1/2) modelling the strength of the tendency to compromise.

    First of all we establish some notations. Let K=[1,1]×[0,1]×[0,1] be the space of the triple ϖ=(w,p,q). We denote P(K) the convex cone of probability measures on K. We endow P(K) with the weak* convergence defined as fkf if ϕdfkϕdf for any ϕC(K). Since K is compact, P(K) is also compact. It is well-known that this topology can be metricized in various ways. We will use the Monge-Kantorovich W1 distance defined for any μ,νP(K) by

    W1(μ,ν)=supϕϕd(μν), (4)

    where the supremum is taken among the 1-Lipschitz functions ϕ:KR; namely, |ϕ(x)ϕ(y)||xy| for any x,yK. We refer to [39] for more details on this distance.

    Denote as fγt(ϖ) the distribution of agents on the triple ϖ at time t0 when agents interact following the rule (3). Indeed, fγt is a probability measure on K, denoted as well in the sequel as dfγt or fγt(ϖ)dϖ. Bear in mind that fγt may not necessarily be absolutely continuous with respect to Lebesgue measure. In fact, fγt could be a Dirac measure.

    In the case of binary interactions and assuming, as usually done, that the joint distibution fγt(ϖ,ϖ)dϖdϖ=fγt(ϖ)fγt(ϖ)dϖdϖ, it can be shown (see e.g. [28]) that fγt is the unique solution of a Boltzmann type equation that takes into account a mean value of all possible interactions. Namely, for any observable ϕC(K),

    ddtKϕ(ϖ)dfγt(ϖ)=K(ϕ(ϖ)ϕ(ϖ))dfγt(ϖ)dfγt(ϖ).

    A fixed point argument yields existence and uniqueness for this equation.

    Our next purpose is to describe the time evolution of this density for a given initial condition f0P(K). To this end, we perform a grazing limit in the Boltzmann type equation above, that is, as the parameter γ adjusting the strength of the interactions (3), goes to zero. Indeed we can then approximate

    ϕ(ϖ)ϕ(ϖ)(ww)wϕ(ϖ)+12(ww)2wwϕ(ϖ)=γqpλ(w)(ww)wϕ(ϖ)+12(γqpλ(w))2(ww)2wwϕ(ϖ)

    to obtain, after simplification,

    1γddtKϕ(ϖ)dfγt(ϖ)K(mγ(t)w)pλγqwϕ(ϖ)dfγt(ϖ)+γ2Kwwϕ(ϖ)Dγt(ϖ)dfγt(ϖ)

    with

    pλγ=Kpλ(w)dfγt(ϖ),mγ(t)=Kpλ(w)pλγwdfγt(ϖ),Dγt(ϖ)=q2(p2w2λ22wp2wλ2+w2p2λ2).

    Notice that pλγ is the mean value at time t of pλ(w), and mγ(t) the mean opinion weighted by the normalized actions of both the power of persuasion and the influence exerted by the agents.

    Rescaling time considering τ=γt, so that ddτ=1γddt, and letting fγτ:=fγt, mγ(τ):=mγ(t), we obtain the approximation

    ddτKϕ(ϖ)dfγτ(ϖ)K(mγ(τ)w)pλγqwϕ(ϖ)dfγτ(ϖ)+γ2Kwwϕ(ϖ)Dγτ(ϖ)dfγτ(ϖ). (5)

    This procedure can be justified showing that fγτ converges as γ0 to fτ, the unique solution of the first order transport equation

    ddτKϕ(ϖ)dfτ(ϖ)=K(m(τ)w)pλqwϕ(ϖ)dfτ(ϖ).

    This is the core of the following Theorem, whose proof (except for minor changes) can be found in [31].

    Theorem 2.1. There exists fC([0,+),P(K)) (where P(K) is endowed with the weak convergence) such that, as γ0, fγτ:=fγt with τ=γt converges to fτ in C([0,T],P(K)) for any T>0. Moreover f is the unique solution in C([0,+),P(K)) of

    Kϕdfτ=Kϕdf0+τ0K(m(s)w)pλqwϕ(ϖ)dfs(ϖ)ds, (6)

    for any τ0 and any ϕC1(K). Here,

    pλ=Kpλ(w)dfτ(ϖ) (7)

    is the mean value at time τ of pλ(w), and

    m(τ)=Kpλ(w)pλwdfτ(ϖ). (8)

    is the mean opinion weighted by the normalized actions of both the power of persuasion and the influence exerted by the agents.

    Remark 1. Notice that (6) is the weak formulation of the transport equation

    τf+w((m(τ)w)pλqf)=0.

    The rest of the paper is devoted to the study of the long-time behaviour of fτ. The next section deals with the simplest case of an homogeneous population i.e. where there are no parameter (p,q). The case of a hetereogeneous population with stubborn people is presented in section 4.

    In both cases we will prove that fτ converges as τ+ to an explicit measure f and provide an explicit estimation of the rate of convergence in term of the W1-distance defined in (4). We obtain in particular that the distribution of opinion converges to a Dirac mass located at an explicit limit opinion w. Since fγτfτ as γ0 we deduce that for a fixed γ1, the solution fγτ of the Boltzmann equation on the time scale τ will be very close to f for τ1. However for a fixed γ the approximation of the Boltzmann equation by the first order equation (6) is valid up to an error term. Indeed (5) shows that the Boltzmann equation will be well-approximated by the first order equation plus an aditional diffusion term γ2ww(Dγτ(ϖ)fγτ). We thus expect the solution fγτ of the Boltzmann equation to be very close to a smoothed version of f.

    This indeed happens as shown by the numerical experiments at the end of the paper. We also refer to [28] for more details on this issue.

    It is worth to emphasize here an important difference with respect to the model studied in [31], which corresponds to taking λ(w)=1 as the influence function. Indeed, when λ=1 then m(t)=1pKpwdgt, and the weight p is constant in time. However, for a non-constant λ, pλ is a priori non-constant in time. This forced us to develop new arguments in order to tackle this more delicate situation, when studying the asymptotic behaviour of m(t).

    From now on we denote ft:=fτ for ease of notation.

    Before going further we recall a useful trick concerning transport equations.

    If fP([1,1]) is a probablity measure on [1,1], and F:R[0,1] is its cumulative distribution function, (namely F(x)=f((,x]) - it is a non-decreasing and right-continuous function with left limit), one can consider the generalized inverse of F, defined as F1:[0,1][1,1]

    F1(ρ)=inf{x[1,1] such that F(x)ρ}. (9)

    Observe that F1 is also non-decreasing and left-continuous with right limit in (0,1]. Furthermore, the following inclusion holds

    [F1(0+),F1(1)]suppf. (10)

    In addition, for any x[1,1] and any ρ[0,1] we have the inequalities

     If F(x)>0 then F1(F(x))x while F(F1(ρ))ρ. (11)

    See the note of Embrechts and Hofert [16] for the above (and further) properties of F1. Moreover, it can be proved that

    10ϕ(F1(r))dr=11ϕ(w)df(w), (12)

    for any ϕ integrable.

    This change of variables will be a key point in the subsequent arguments. More precisely, consider fC([0,);P([1,1])) satisfying a transport equation of the form

    tft+x(V(t,x)ft)=0. (13)

    Denote as Ft the cumulative distribution function of ft, and Xt=F1t its generalized inverse. With the use of (12) we can rewrite the weak form of equation (13) in a much simpler form. This is the core of the next result:

    Proposition 1. Let v:[0,+)×[1,1]R be continuous and globally Lipschitz with respect to the second variable. Then, fC([0,+),P([1,1])) is a weak solution of (13) in the sense that for any ϕC1([1,1]) and any t>0,

    11ϕ(x)dft(x)=11ϕ(x)df0(x)+t011ϕ(x)v(s,x)dfs(x)ds, (14)

    if and only if for any r(0,1], Xt(r) is a solution of

    tXt(r)=v(t,Xt(r)). (15)

    Here, X0 is the generalized inverse of F0 (the cumulative distribution function of f0).

    The proof can be found essentially in Theorem 3.1 in [2]. See also [31], where it is rewritten under the point of view of the ordinary differential equation for the flux (15).

    In this section we evaluate the impact of the attractiveness of the opinion independently of any other consideration. Thus, an agent is now completely characterized by its opinion w[1,1], and when interacting with an agent of opinion w[1,1] the resulting opinion w is

    w=w+λ(w)(ww). (16)

    This synergy after encounters was indeed introduced in [30]. A simpler form that takes the model with homogeneity, allowed us to find the kinetic equations with the use of the empirical distribution for N agents, fN(w,t)=1NNi=1δwi(t), and taking limits as N.

    The resultant continuous distribution ftP([1,1]) verifies the mean field first order equation

    11ϕdft=11ϕdf0+t011(mtw)λϕ(w)dfs(w)ds. (17)

    where

    mt=11λ(w)λwdft(w),λ=11λ(w)dft(w),

    which is consistent with (6) taking p=q=1.

    Moreover, in [30] we identified a conserved quantity for the evolution equation (17), whenever λC([1,1]). We observed that the function 11Λ(w)dft(w), being Λ an antiderivative of λ, remains constant in time. Indeed, according to (17),

    ddt11Λ(w)dft=11(mtw)λΛ(w)dft(w)=11(mtw)λλ(w)dft(w).

    Recall that λmt=11wλ(w)dft(w), thus the right hand side vanishes and then

    11Λdft=11Λdf0for any t0.

    Sending t+ and assuming that consensus occurs in the sense that ftδm for some m, we deduce

    Λ(m)=Λdf0. (18)

    Hence, the candidate to be the value of consensus is

    m=Λ1(11Λ(w)df0(w)).

    In the next result we perform a rigorous proof of the dynamics contraction towards this value.

    Theorem 3.1. Assume that λ:[1,1]R is a continuous function such that, for some λ_>0,

    λ(w)λ_for any wconv(supp(f0)), (19)

    being conv(supp(f0)) the convex hull of supp(f0).

    Then, there exists m:=limt+mt[1,1] such that

    W1(ft,δm)|conv(supp(f0))|eλ_t. (20)

    Moreover, the limit opinion m is given by

    m=Λ1(Λ(w)df0(w)), (21)

    where Λ is an antiderivative of λ.

    Before the proof, some remarks are in order:

    Remark 2. Notice that we just require λ to be positive on the convex hull of the support of the initial distribution f0. This is due to the fact that the dynamic is contractive, in the sense that supp(ft)supp(f0), t0.

    Remark 3. In the case λ1 we have Λ(w)=w so that the consensus opinion m is simply the initial mean opinion, see [34].

    The proof of Theorem 3.1 goes as follows:

    Proof of Theorem 3.1. Let Xt be the generalized inverse of the cumulative distribution function corresponding to ft. According to Proposition 1 we can rewrite the equation satisfied by ft as

    tXt(r)=(mtXt(r))λ.

    Moreover,

    mt=11λ(w)λwdft(w)=10Xt(r)λ(Xt(r))λdr.

    Since Xt is non-decreasing,

    mtXt(1)10λ(Xt(r))λdr=Xt(1)11λ(w)λdft(w)=Xt(1),

    and similarly mtXt(0+). Thus,

    Xt(0+)mtXt(1).

    Note that λ0 implies that λ0. As a result tXt(1)0, thus Xt(1) is non-increasing. An identical argument shows that Xt(0+) is non-decreasing. Since [Xt(0+),Xt(1)] is the convex hull of supp(ft) this proves that

    conv(supp(ft))conv(supp(f0))t0.

    Indeed, λ(w)λ_>0 for any wsupp(ft). Consequently,

    t[(Xt(1)Xt(r))2]=2(Xt(1)Xt(r))2λ2λ_(Xt(1)Xt(r))2,

    and Gronwall's Lemma gives, for any r(0,1] and any t0,

    |Xt(1)Xt(r)||X0(1)X0(r)|eλ_t.

    In particular,

    |Xt(1)Xt(0+)||X0(1)X0(0+)|eλ_t,

    which reveals that the length of supp(ft) goes to 0.

    Recalling that Xt(0+)mtXt(1) for any t, Xt(0+) increases and Xt(1) decreases. Accordingly, supp(ft) shrinks to a limit point limt+mt:=m[Xt(0+),Xt(1)] for any t. In fact, for any r(0,1],

    |Xt(r)m||Xt(1)Xt(0+)||X0(1)X0(0+)|eλ_t.

    Inequality (20) follows now by observing that

    W1(ft,δm)11|wm|dft(w)=10|Xt(r)m|dr|X0(1)X0(0+)|eλ_t.

    It remains to show (21). Notice that Λ is C1, Λ is positive and increasing on conv(supp(f0)). Thus, Λ defines a bijection from the interval conv(supp(f0)) on its image

    A:=Λ(conv(supp(f0)))=[Λ(X0(0+)),Λ(X0(1))].

    Furthermore,

    Λ(X0(0+))11Λdf0(w)=10Λ(X0(r))drΛ(X0(1)),

    and consequently Λdf0A. Identity (18) entails that mconv(supp(f0)) and the proof of (21) concludes.

    We would like to close this section by emphasizing some relevant considerations of the result above.

    First of all, the hypothesis of strict positivity of the influence function on the convex hull of the initial density is absolutely necessary to obtain Theorem 3.1.

    Indeed in [30] we already noticed singular behaviour at the roots of λ in the simulations. Bear in mind that the value m specified in (21) is expected to hold in mean. That is the reason why, even starting from identical initial conditions, simulations could give different values for m, due to the random fluctuations from realization to realization.

    This fact becomes more evident the larger the order of the zero is. If λ has a high order zero at z, Λ is almost a constant function close to it, and the inverse function Λ is defined although it is very sensitive to small changes. If in addition we have that

    z=11Λ(w)df0(w),

    two phenomena occur. One is the slow formation of consensus due to a frozen dynamics due to the small values that λ takes. The second one is a symmetry rupture, and after a long time consensus is reached above or below z.

    Let us note that the first order equation is interpreted in a weak sense, so the differentiability of λ is not required. However, in the next section we need to impose the condition λW2,([1,1]), which seems to be a technical hypothesis. Indeed, we can consider λ(w)=|w|1/2 and given an uniform initial distribution of agents, the consensus is reached at m=0, see [30]

    It is worth to mention a related dynamics appearing in [2,38], where agents are influenced by their own opinion when interact,

    {w=w+γP(w)(ww)w=w+γP(w)(ww). (22)

    Typical examples of function P are non-increasing on |w|, representing the fact that the extremist people are more likely to remain in their believes. In fact, the roots of the function P represent the reticence of the individual that eventually adopted that opinion, to change it after subsequent encounters. Let us observe that in this model stubborn agents appear dynamically when they approach the zeros of P. The qualitative differences in the dynamic between (22) and our model (1) were shown numerically in [30]. Moreover, when consensus occurs in (22), its value is completely different from the value (21) obtained here. In the case of (22), equation (17) reads

    11ϕdft=11ϕdf0+t011P(w)(ww)ϕ(w)dfs(w)ds, (23)

    where w=11wdfs(w) is the mean opinion at time s. If we assume that P(w)P_>0, w[1,1], it follows that an antiderivative Π of 1/P (and not of λ as in our model) is conserved. Indeed

    ddtΠ(w)dft(w)=11P(w)(ww)1P(w)dfs(w)=11(ww)dfs(w)

    which is equal to 0. Consequently, if consensus occurs in the sense that ftδ˜m for some limit opinion ˜m, then ˜m must satisfy

    Π(˜m)=Πdf0

    so that ˜m=Π1(Πdf0). We present in section §5 below some numerical experiments to validate the formula for ˜m.

    This section is devoted to study the long-time behaviour of the unique solution to the transport equation (6), arising from an initial distribution f0 of the form

    ft=0=α0f00+(1α0)f10, (24)

    being α0(0,1] the proportion of stubborn agents in the population, and f00,f10P(K) the initial distributions of agents with parameters (w,p,q) in the stubborn and non-stubborn population, respectively. Observe that since the stubborn agents do not change their opinion in an interaction, ft will evolve as

    ft=α0f00+(1α0)f1t,t0.

    Before stating our main result, we give an informal deduction of the possible value to be the limiting opinion, presuming that the non-stubborn agents reach a consensus at m.

    In view of the transport term, whenever consensus is reached among the non-stubborn agents, it should take place at m:=limtmt, whenever this limit exists. To search for the candidate to m, we argue as follows. Accepting that consensus is reached at m, then

    f1tf10(p,q)dpdqδmt+,

    where f10(p,q)dpdq is the distribution of the non-stubborn population on the (p,q)-parameters. This density is constant in time since (p,q)-parameters are unaffected by the dynamics. As a result, we can pass to the limit as t+ in the definition of mt, namely

    pλmt=Kpλ(w)wdft(ϖ). (25)

    On the one hand,

    Kpλ(w)wdft(ϖ)=α01110pλ(w)wdf00(w,p)+(1α0)Kpλ(w)wdf1t(ϖ)α01110pλ(w)wdf00(w,p)+(1α0)λ(m)mKpdf10(ϖ).

    While on the other hand,

    pλ=α01110pλ(w)df00(w,p)+(1α0)Kpλ(w)df1t(ϖ)α01110pλ(w)df00(w,p)+(1α0)λ(m)Kpdf10(ϖ).

    Taking now limits in (25) we get

    m(α01110pλ(w)df00(w,p)+(1α0)λ(m)Kpdf10(ϖ))=α01110pλ(w)wdf00(w,p)+(1α0)λ(m)mKpdf10(ϖ).

    Since α0>0,

    m1110pλ(w)df00(w,p)=1110pλ(w)wdf00(w,p).

    Conclusion: If the non-stubborn population reaches consensus, then the consensus opinion m is specified by

    m:=1110pλ(w)wpλ0df00(w,p), (26)

    whenever the term

    pλ0:=1110pλ(w)df00(w,p), (27)

    that stands for the mean value of pλ within the stubborn population, does not vanish.

    This shows that, admitting long-time consensus among the non-stubborn population, its shared opinion m is the mean opinion value weighted by the normalized pλ (the power of conviction times the influence function). Observe that this mean value is taken just within the stubborn population. Thus, if a common limit opinion exists, it is determined by the stubborn agents.

    At this stage, it is crucial to underline a relevant fact about the candidate for consensus found above. If the influence function is constant, this model lays on our previous work studied in [31], and hence the value for m specified in (26) turns out to be

    m:=1110pwp0df00(w,p).

    Our main result shows that non-stubborn agents indeed reach consensus asymptotically at the value m determined by (26). We also provide an estimate on the rate of convergence towards the consensus, in terms of the W1-distance between f1t and its limit f10(p,q)dpdqδm.

    Before stating it recall that f00 and f1t denote the distribution of the stubborn and non-stubborn agents on (w,p,q), and we denote α0(0,1] the proportion of stubborn agents in the population. By f1t|(p,q)P([1,1]) we understand the distribution of opinions within the group of non-stubborn agents having parameters (p,q). Its existence is guaranteed by Jirina's Theorem. There are several classical references on this subject, for example [4,36].

    We are now ready to state our main result of this section

    Theorem 4.1. Assume f10P(K) is supported in {qε0} for some ε0>0 and that the map

    (p,q)[0,1]×[ε0,1]f10|(p,q)P([1,1]),

    is globally Lipschitz for the W1-distance: there exists L>0 such that for any (p,q),(p,q)[0,1]×[ε0,1],

    W1(f10|(p,q),f10|(p,q))L(|qq|+|pp|). (28)

    In addition, assume that λW2,([1,1]) verifies that there exists λ_>0 such that λ(w)λ_ for any w[1,1].

    Then, for any t0,

    W1(f1t,f10(p,q)dpdqδm)(4+κt)eε0α0pλ0t (29)

    where m is specified in (26) and stands for the mean opinion weighted by the normalized power of persuasion multiplied by the influence function within the group of stubborn agents. The mean value of pλ among the stubborn agents, pλ0 is defined in (27). Moreover

    κ=8λλ_(λ+λ)+4λ.

    The global idea of the proof of Theorem 4.1 follows the lines of the proof in [31], where the case λ1 is treated. Notice however that when λ1 then pλ=p is constant in time whereas for an arbitrary λ it is time-varying quantity. This introduces new difficulties in many of the steps into which the proof is divided.

    In the first step we consider f1t|(p,q)P([1,1]), the conditional distribution of opinion among the agents with parameter (p,q). This conditional distribution turns out to be the unique solution to the following transport equation. Furthermore, it is (p,q)-Lipschitz with respect to the Wasserstein distance. These facts are summarized below and the proof can be easily adapted from the case λ constant [31].

    Step 4.1. For any (p,q)supp(f0(p,q)dpdq), f1t|(p,q) is the unique solution to

    {tf1t|(p,q)+w((mtw)qpλf1t|(p,q))=0,f1t=0|(p,q)=f0|(p,q), (30)

    in C([0,+),P([1,1])).

    Moreover, the function (p,q)f1t|(p,q) is Lipschitz with respect to the Wasserstein distance W1. Namely, for any (p,q),(p,q)[0,1]×[ε0,1],

    W1(f1t|(p,q),f1t|(p,q))Ct(|qq|+|pp|).

    Furthermore, it fulfils

    Kϕdf1t=1010(11ϕdf1t|(p,q)(w))df10(p,q),ϕC(K). (31)

    In the next item, we take advantage of the tendency to compromise modeled by the interaction rules (3) to prove that non-stubborn agents with given (p,q) parameters tend to synchronize their opinions. Conditioning to values (p,q) we declare

    λ(p,q)=11λ(w)df1t|(p,q)(w), (32)

    and

    m(t,p,q)=11λ(w)λ(p,q)wdf1t|(p,q)(w), (33)

    the mean value of λ and the mean opinion among the agents with parameter (p,q)[0,1]×[ε0,1], respectively.

    Step 4.2. For any (p,q)[0,1]×[ε0,1] there holds

    W1(f1t|(p,q),δm(t,p,q))2eε0α0pλ0tt0, (34)

    being pλ0 defined in (27).

    Proof. Let Xt be the generalized inverse of the cumulative distribution function corresponding to ft|(p,q). According to Proposition 1 we can rewrite (30) as

    tXt(r)=(mtXt(r))qpλ.

    In particular

    t[(Xt(1)Xt(r))2]=2(Xt(1)Xt(r))2qpλ2(Xt(1)Xt(r))2ε0pλ.

    The nonnegativity of λ entails that

    pλ=pλ(w)dft=α0pλ0+(1α0)pλ(w)df1tα0pλ0,

    hence

    t[(Xt(1)Xt(r))2]2α0ε0(Xt(1)Xt(r))2pλ0.

    Gronwall's Lemma implies now that

    |Xt(1)Xt(r)||X0(1)X0(r)|eα0ε0pλ0t, (35)

    which sending r0+ gives

    |Xit(1)Xit(0+)||Xi0(1)Xi0(0+)|eα0ε0pλ0t. (36)

    This shows that the length of the support of ft|(p,q) goes to 0.

    Invoking (12), we express m(t,p,q) in terms of the generalized inverse,

    m(t,p,q)=11λ(w)λ(p,q)wdft|(p,q)(w)=10Xt(r)λ(Xt(r))λ(p,q)dr.

    Since Xt is non-decreasing,

    m(t,p,q)Xt(1)10λ(Xt(r))λ(p,q)dr=Xt(1)11λ(w)λ(p,q)dft|(p,q)(w)=Xt(1),

    and similarly m(t,p,q)Xt(0+). Thus,

    Xt(0+)mitXit(1). (37)

    Therefore, for any r(0,1],

    |Xt(r)m(t,p,q)||Xt(1)Xt(0+)||X0(1)X0(0+)|eα0ε0pλ0t.

    The proof now concludes by noticing that

    W1(ft|(p,q),δm(t,p,q))11|wm(t,p,q)|dft|(p,q)(w)=10|Xt(r)m(t,p,q)|dr|X0(1)X0(0+)|eα0ε0pλ0t.

    As a result of the previous step, it is desirable to study the asymptotic behavior of the function m(t,.) as t+.

    Step 4.3. For any t0 and any (p,q)[0,1]×[ε0,1] the function m(t,p,q) declared in (33) satisfies

    tm(t,p,q)=α0qpλ0[mm(t,p,q)]+(1α0)q11pλ(m(t,p,q))[m(t,p,q)m(t,p,q)]df10(p,q)+R(t,p,q) (38)

    where m and pλ0 are given in (26) and (27) respectively, and

    |R(t,p,q)|{8λλ_(λ+λ)+4λ}eε0α0pλ0t. (39)

    Proof. Using equation (30), the evolution in time of the conditioned mean λ(p,q) defined in (32) behaves as

    ddtλ(p,q)=ddt11λ(w)df1t|(p,q)(w)=qpλ11(mtw)λ(w)df1t|(p,q)(w)

    and

    ddt11wλ(w)df1t|(p,q)=qpλ11(mtw)(wλ(w)+λ(w))df1t|(p,q)(w).

    Thus,

    tm(t,p,q)=t{1λ(p,q)11wλ(w)df1t|(p,q)(w)}=qpλλ(p,q){11(mtw)λ(w)(wm(t,p,q))df1t|(p,q)(w)+11(mtw)λ(w)df1t|(p,q)(w)}.

    Summing up,

    tm(t,p,q)=qpλλ(p,q)11(mtw)λ(w)(wm(t,p,q))df1t|(p,q)(w)+qpλ(mtm(t,p,q)). (40)

    Estimate (34) allows to bound the intergral in the r.h.s. as follows. Let ϕ(w)=(mtw)λ(w)(wm(t,p,q)) so that

    11(mtw)λ(w)(wm(t,p,q))df1t|(p,q)(w)=11ϕ(w)ϕ(m(t,p,q))df1t|(p,q)(w)=11ϕ(w)(df1t|(p,q)(w)δm(t,p,q)).

    According to the definition of the W1-distance we obtain

    11(mtw)λ(w)(wm(t,p,q))df1t|(p,q)(w)Lip(ϕ)W1(f1t|(p,q),δm(t,p,q)).

    It is easily seen that Lip(ϕ)4(λ+λ). Accordingly to (34), we deduce

    11(mtw)λ(w)(wm(t,p,q))df1t|(p,q)(w)8(λ+λ)eε0α0pλ0t.

    Thus,

    tm(t,p,q)=qpλ(mtm(t,p,q))+˜R(t,p,q)

    with

    |˜R(t,p,q)|qpλλ(p,q)8(λ+λ)eε0α0pλ0t8λλ_(λ+λ)eε0α0pλ0t,

    given that the assumption λ(w)λ_>0 for any w[1,1] entails that λ(p,q)λ_>0 for any (p,q).

    We now focus on the second term in the right hand side of (40). First,

    pλmt=pwλ=α0pwλ0+(1α0)Kpwλ(w)df1t(ϖ),

    where the integral can be written using (31) as

    1010p(11wλ(w)df1t|(p,q)(w))df10(p,q)=1010pλ(p,q)m(t,p,q)df10(p,q).

    Recalling in addition that pwλ0=pλ0m with m given in (26), we obtain

    pλmt=α0pλ0m+(1α0)1010pλ(p,q)m(t,p,q)df10(p,q).

    On the other hand,

    pλ=α0pλ0+(1α0)Kpλ(w)df1t(ϖ)=α0pλ0+(1α0)1010p(11λ(w)df1t|(p,q)(w))df10(p,q)=α0pλ0+(1α0)1010pλ(p,q)df10(p,q).

    Merging the former identities together,

    tm(t,p,q)=α0qpλ0[mm(t,p,q)]+(1α0)q1010pλ(p,q)[m(t,p,q)m(t,p,q)]df10(p,q)+˜R(t,p,q).

    According to (34) we have for any (p,q) that

    |λ(p,q)λ(m(t,p,q))|=|(f1t|(p,q)δm(t,p,q),λ)|2eε0α0pλ0tLip(λ)2eε0α0pλ0tλ.

    As a result,

    tm(t,p,q)=α0qpλ0[mm(t,p,q)]+(1α0)qpλ(m(t,p,q))[m(t,p,q)m(t,p,q)]df10(p,q)+˜R(t,p,q)+ˆR(t,p,q)

    with

    ˆR(t,p,q)(1α0)q2eε0α0pλ0tλ×1010p|m(t,p,q)m(t,p,q)|df10(p,q)4eε0α0pλ0tλ

    where we used that m(t,p,q)[1,1] for any (t,p,q). We deduce (38) taking R(t,p,q)=˜R(t,p,q)+ˆR(t,p,q).

    To understand intuitively the infinite system of equations (38) it is useful to consider the elementary situation where only a finite number of values for (p,q), q0, are present in the population, namely (p1,q1),...,(pN,qN). Then, f1t takes the simpler form

    f1t=Ni=1αigit(w)dwδp=pi,q=qi,

    where git:=ft|(pi,qi) is the distribution of opinion in the (pi,qi)-population and αi(0,1] is the proportion of (pi,qi) agents in the non-stubborn population. Letting mit:=m(t,pi,qi) we can rewrite the system (38) as

    ddtmit=ABmit+Nj=1cjλ(mjt)(mjtmit)+Ri(t) (41)

    where Rit:=R(t,pi,qi), A=α0pwλ0, B=α0pλ0, cj=(1α0)αjpj. The right hand side of (41) is composed of three terms. The first one ABmit drives mit towards A/B=m, which is the desired asymptotic state. The sum includes a coupling between mit and all the mjt, j=1,,N, whose effect contributes to synchronize them as in, e.g. the Cucker-Smale model [14]. The error term decreases exponentially fast to 0, thus it lacks relevance in the asymptotic behaviour. We thus expect that limt+mit=m for any i=1,..,N. We will prove in the sequel that this intuition is indeed correct in general.

    The regularity of m(t,p,q) with respect to (p,q) plays an important role in the convergence of m(t,p,q) to m as t+.

    Step 4.4. For any t0, the function (p,q)supp(f10(p,q)dpdq)m(t,p,q) is Lipschitz.

    Proof. According to Step 4.1, (p,q)f1t|(p,q) is Lipschitz with respect to the Wasserstein distance W1: for any (p,q),(p,q)[0,1]×[ε0,1],

    W1(f1t|(p,q),f1t|(p,q))Ct(|qq|+|pp|).

    The functions λ(p,q) and wλ(p,q) are also Lipschitz in (p,q) for a given t, whenever we assume that λ is Lipschitz. Indeed for a given t0 and any (p,q),(p,q),

    |λ(p,q)λ(p,q)|=|(f1t|(p,q)f1t|(p,q),λ)|Lip(λ)W1(f1t|(p,q),f1t|(p,q))Ctλ(|pp|+|qq|)

    and in the same way

    |wλ(p,q)wλ(p,q)|CtLip(wλ(w))(|pp|+|qq|).

    Consequently, for a given t0, the functions λ(p,q) and wλ(p,q) are continuous in (p,q). Since λ(p,q)λ_>0 for any (p,q), it follows that m(t,p,q)=wλ(p,q)/λ(p,q) is continuous in (p,q) for any t0.

    Step 4.5. For any (p,q)supp(f10(p,q)dpdq) and any t0 it holds that

    |m(t,p,q)m|eε0α0pλ0t(|m(0,p,q)m|+κt)

    where

    κ=8λλ_(λ+λ)+4λ.

    Proof. Relation (38) implies that for any q[ε0,1] and t0,

    12t|m(t,p,q)m|2=tm(t,p,q)[m(t,p,q)m]=qα0pλ0[mm(t,p,q)]2+q(1α0)[m(t,p,q)m]×1010pλ(m(t,p,q))[m(t,p,q)m(t,p,q)]df10(p,q)+R(t,p,q)[m(t,p,q)m]. (42)

    Recall that m(t,.) is continuous by the previous Step and supp(f10(p,q)dpdq) is compact. Take some (p,q) such that

    maxsupp(f10(p,q)dpdq)|m(t,.)m|=|m(t,p,q)m|.

    In particular, we can write (42) at (p,q) as

    12t|m(t,.)m|2|(p,q)=qα0pλ0[mm(t,p,q)]2+q(1α0)[m(t,p,q)m]×1010pλ(m(t,p,q))[m(t,p,q)m]df10(p,q)q(1α0)[m(t,p,q)m]21010pλ(m(t,p,q))df10(p,q)+R(t,p,q)[m(t,p,q)m]=:I+II+III+IV.

    The choice of q assures that

    IIq(1α0)|m(t,p,q)m|21010pλ(m(t,p,q))df10(p,q)=III.

    The cancelation of these two terms and qε0 gives

    t|m(t,.)m|2|(p,q)2ε0α0pλ0|mm(t,p,q)|2+2R(t,p,q)|m(t,p,q)m|. (43)

    Let h(t;(p,q))=|m(t,p,q)m|2. Notice that th(t;(p,q)) is a C1 function, since m is C1 in time. Moreover, from (42) it follows that |th(t;(p,q))|C. Since h(t,.) is continuous for any t by the previous Step, we obtain that h is continuous in (t,p,q). According to the Measurable Selection Theorem (see e.g. §18.19 in [10]) we can choose (p,q) to be a measurable function of t.

    The Envelope Theorem (see Theorem 4.2 below) ensures that the function V(t) defined by

    V(t):=max(p,q)supp(f10)h(t;(p,q))

    is absolutely continuous with derivative

    V(t)=t(|m(t,p,q)m|2)a.e.

    Furthermore, in view of (43) and (39)

    V(t)2ε0α0pλ0V(t)+2κV(t)eε0α0pλ0t,

    where

    κ=8λλ_(λ+λ)+4λ.

    Dividing by 2V we obtain

    (V)(t)ε0α0pλ0V(t)+κeε0α0pλ0t

    which integrated gives

    V(t)eε0α0pλ0tV(0)+t0eε0α0pλ0(ts)κeε0α0pλ0sds=eε0α0pλ0t(V(0)+κt).

    This shows that

    |m(t,p,q)m|eε0α0pλ0t(|m(0,p,q)m|+κt),

    for any (p,q)supp(f10(p,q)dpdq), as desired.

    In the course of the proof of the previous step we used the following envelope Theorem, due to Milgrom and Segal in [26]:

    Theorem 4.2. Consider the function V(t):=maxxXh(x,t), t[0,1] being X a set. Suppose that h is absolutely continuous with respect to t, for any x. Moreover, admit that there exists bL1([0,1]) such that |th(x,t)|b(t) for any xX and almost any t[0,1]. Then V is absolutely continuous.

    Assuming further that h is differentiable in t, for any xX, and that for any t[0,1] the set X(t):=argmaxh(.,t) is non-empty. Then, for any selection of x(t)X(t) we have

    V(t)=V(0)+t0th(x(s),s)ds.

    We are now in position to finish the proof of Theorem 4.1.

    Step 4.6. There holds

    W1(f1t,f10(p,q)dpdqδm)(4+κt)eε0α0pλ0t,

    for any t0.

    Proof. Step 4.5 ensures that for any t0 and any (p,q)supp f10(p,q)dpdq,

    W1(δm(t,p,q),δm)=|m(t,p,q)m|(2+κt)eε0α0pλ0t,

    while according to (34), we have that

    W1(f1t|(p,q),δm)(4+κt)eε0α0pλ0t.

    In fact, we claim that

    W1(f1t,δmf10(p,q)dpdq)(4+κt)eε0α0pλ0t.

    Let ψ:KR be an arbitrary 1-Lipschitz function. Then,

    Kψ(df1tδmf10(p,q)dpdq)=1010(11ψ(w,p,q)(dft|(p,q)δm))df0(p,q).

    The inner integral is bounded above by W1(ft|(p,q),δm) since ψ(.,p,q) is 1-Lipschitz, which implies that

    Kψ(df1tδmf10(p,q)dpdq)(4+κt)eε0α0pλ0t1010df0(p,q)=(4+κt)eε0α0pλ0t.

    The claim follows by taking supremum among the functions ψ 1-Lipschitz.

    In this section we perform some agent based simulations of the results studied along this work. We assess qualitatively the effect of the influence function

    λ(w)=(w0.5)2+ε,ε=0.01,

    on the dynamics. We first consider a homogeneous population and ratify that, indeed, the conclusions of Theorem 3.1 are true. Then, to illustrate the conclusion of Theorem 4.1, we add to this scenario several stubborn agents positioned at two specific values for ω. We finally analyze the evolution of a completely heterogeneous population with different values of q.

    In all of the simulations presented here, we consider a population of N=1000 agents. The opinion of the non-stubborn agents are initially uniformly distributed in [1,1]. At every time slot, each one of the N agents interacts with a randomly selected agent and then, updates its opinion following the interaction rule (3) with γ=0.01.

    Since opinions are initially distributed uniformly in [1,1], i.e. f0=121[1,1], the theoretical value m of the consensus(given by (21) in Theorem 3.1), satisfies

    Λ(m)=1211Λ(w)dw,

    where Λ is an antiderivative of λ(w)=(w0.5)2+ε, ε=0.01. A numerical resolution of this nonlinear equation gives

    m0.35.

    In Figure 1 (left) we show the time evolution of the opinions of 10 agents of the population (blue curves). Consensus clearly occurs at the value m, indicated by the horizontal red dashed line.

    Figure 1.  Evolution of the opinion of 10 agents (blue) from a homogeneous population of N=1000 agents interacting according the interaction rule (16) studied in this paper (left) or the interaction rule (22) considered in [2,38] (right). The red dashed line indicates the theoretical limit opinion in both cases (m0.35 for interaction rule (16), (left), and ˜m0.41 for interaction rule (22), (right)) The early evolution is shown in inset.

    This is completely in contrast to consider that the influence is exerted, instead, by one's own opinion agent, see [2,38]. Indeed, under the interaction rules (22), we prove in subsection §3.3 that if consensus is reached then, the consensus opinion ˜m must satisfy

    Π(˜m)=11Π(w)df0(w),

    where Π is an antiderivative of 1/P. Taking as f0 the uniform distribution on [1,1] and P(w)=λ(w)=(w0.5)2+ε, we numerically obtain

    ˜m0.41.

    The result of the agent-based simulation shown in the right figure of Figure 1 confirms that consensus takes place at ˜m (indicated by the red dashed horizontal line).

    We can also observe from the simulation that the consensus is attained much faster for the interaction rule (16) than for (22). Recall that the interaction rule (22) takes into account the attractiveness of the opinion of the agent one is interacting with. Intuitively, this is due to the fact that agents with opinion w such that P(w)0 are almost stubborn agents: they change opinion very slowly. This is clear from the figures in Figure 2, where we plot for the two interaction rules (16) (left) and (22) (right) the logarithm of the length of the convex hull of suppft, the support of the distribution ft of opinion at time t. Namely, we depict

    ln(maxsuppftwminsuppftw).
    Figure 2.  Evolution of ln(maxwminw), where the max and min are taken on the support of the distribution of opinions, for interaction rule (16) studied in this paper (left), and interaction rule (22) considered in [2,38] (right).

    In the rest of this section we examine the impact of stubborn agents combined with the influence function.

    We consider a population of N=1000 agents with a proportion α0 of stubborn agents. To balance the effect between stubbornness and the attractiveness of the other's opinions, every agent has p=1. As before the influence function is λ(w)=(w0.5)2+ε, ε=0.01, and γ=0.01. Assume that half of the stubborn agents have opinion w=1/4 and the other half opinion w=3/4, so that

    f00=12δp=1δw=14+12δp=1δw=34.

    Notice that, in particular

    pλ0=pλ(w)df00(w,p)=11λ(w)df00(w)=12(λ(1/4)+λ(3/4))=116+ε.

    and

    pλ(w)w0=11λ(w)wdf0(w)=12(λ(1/4)14+λ(3/4)34)=12(116+ε).

    The theoretical limit opinion is then

    m=pλ(w)w0pλ0=12.

    To evaluate the impact driven just by the proportion α0 of stubborn agents on the dynamics, we suppose that all of the non-stubborn agents have q=1. We show in Figure 3 the time evolution of the opinion of 10 agents belonging to the non-stubborn population (blue curves). We consider a proportion of stubborn agents α0=2% (left) and α0=60% (right). The theoretical limit opinion m, is depicted on red dashed line.

    Figure 3.  Evolution of the opinion of 10 agents (blue) from a population of N=1000 agents with α0=2% (left) and α0=60% (right) stubborn agents. The red dashed line indicates the theoretical limit opinion m=1/2, and the blue dotted lines the opinion of the stubborn agents (half with opinion 1/4 and the other half with opinion 3/4). The early evolution is shown in inset.

    We observe a perfect compliance between the agent-based simulations and the theoretical prediction. Furthermore, in the simulations the consensus is clearly achieved in two steps. First non-stubborn agents quickly reach a consensus, and then all together move slowly towards the final limit opinion m.

    This fact is specially well observed on the left figure, where a smaller proportion of stubborn population is considered (α0=2%). At first the impact of the stubborn agents is almost negligible, so that the opinions of the non-stubborn population evolve first at a value close to the predicted consensus opinion in absence of stubborn agents, namely here 0.355. In other words, the influence function λ drives the dynamics at early stages, and then is wiped out by the stubborn agents. Comparing both figures, we also notice that a high proportion of stubborn agents accelerates the convergence towards the consensus, in accordance with estimation (29).

    Finally, we wish to appreciate the qualitative impact of the parameter q on the dynamics. We now consider that the values of q for the non-stubborn population are distributed uniformly in (0.2;1). The rest of the parameters are kept as before with α0=0.6. In Table 4 we show several snapshots at different times of the distribution of (w,q) in the non-stubborn population (w in the horizontal axis, q in the vertical axis). We can clearly appreciate that agents with a high q, i.e. the most volatile agents, are indeed changing opinion quicker than the rest.

    Figure 4.  Evolution of the distribution of (w,q) among the non-stubborn population during one simulation (w in the horizontal axis and q in the vertical axis) with λ(w)=(w0.5)2+ε, ε=0.01. From left to right and top to bottom, figures show the distribution of (w,q) at time 1,500,1000,1500,3000,5000,10000,15000,30000.

    The authors appreciated the fine comments of both referees which helped to enhance the previous version of the manuscript.



    [1] G. Aletti, A. K. Naimzada and G. Naldi., Mathematics and physics applications in sociodynamics simulation: The case of opinion formation and diffusion, Mathematical Modeling of Collective Behavior in Socio-Economic and Life Sciences, Birkhäuser Boston, (2010), 203–221. doi: 10.1007/978-0-8176-4946-3_8
    [2] First-order continuous models of opinion formation. SIAM J. Appl. Math. (2007) 67: 837-853.
    [3] Opinions and social pressure. Scientific American (1955) 193: 31-35.
    [4] R. B. Ash, Real Analysis and Probability, Probability and Mathematical Statistics, Academic Press, New York-London, 1972.
    [5] P. Balenzuela, J. P. Pinasco and V. Semeshenko, The undecided have the key: Interaction driven opinion dynamics in a three state model, PLoS ONE, 10 (2016), e0139572, 1–21.
    [6] Opinion strength influences the spatial dynamics of opinion formation. The Journal of Mathematical Sociology (2016) 40: 207-218.
    [7] B. O. Baumgaertner, P. A. Fetros, R. C. Tyson and S. M. Krone, Spatial Opinion Dynamics and the Efects of Two Types of Mixing, Phys Rev E., 98 (2018), 022310.
    [8] N. Bellomo, Modeling Complex Living Systems A Kinetic Theory and Stochastic Game Approach, Birkhauser, 2008.
    [9] N. Bellomo, G. Ajmone Marsan and A. Tosin, Complex Systems and Society. Modeling and Simulation, SpringerBriefs in Mathematics, 2013. doi: 10.1007/978-1-4614-7242-1
    [10] K. C. Border and C. D. Aliprantis, Infinite Dimensional Analysis - A Hitchhiker's Guide, 3rd Edition, Springer, 2006.
    [11] What a person thinks upon learning he has chosen differently from others: Nice evidence for the persuasive-arguments explanation of choice shifts. Journal of Experimental Social Psychology (1975) 11: 412-426.
    [12] A well-posedness theory in measures for some kinetic models of collective motion. Mathematical Models and Methods in Applied Sciences (2011) 21: 515-539.
    [13] R. B. Cialdini and M. R. Trost, Social influence: Social norms, conformity and compliance, The Handbook of Social Psychology, McGraw-Hill, (1998), 151–192.
    [14] Emergent behavior in flocks. IEEE Transactions on Automatic Control (2007) 52: 852-862.
    [15] Mixing beliefs among interacting agents. Advances in Complex Systems (2000) 3: 87-98.
    [16] A note on generalized inverse. Mathematical Methods of Operations Research (2013) 77: 423-432.
    [17] Social influence and opinions. Journal of Mathematical Sociology (1990) 15: 193-206.
    [18] A simple model of consensus formation. Okayama Economic Review (1999) 31: 95-100.
    [19] S. Galam, Sociophysics: A Physicist's Modeling of Psycho-Political Phenomena, , Springer Science & Business Media, 2012. doi: 10.1007/978-1-4614-2032-3
    [20] R. Hegselmann and U. Krause, Opinion dynamics and bounded confidence: Models, analysis and simulation, Journal of Artificial Societies and Social Simulation, 5 (2002).
    [21] Ergodic theorems for weakly interacting infinite systems and the voter model. The Annals of Probability (1975) 3: 643-663.
    [22] C. La Rocca, L. A. Braunstein and F. Vázquez, The influence of persuasion in opinion formation and polarization, Europhys. Letters, 106 (2014), 40004. doi: 10.1209/0295-5075/106/40004
    [23] The psychology of social impact. American Psychologist (1981) 36: 343-356.
    [24] Continuous opinion dynamics under bounded confidence: A survey. International Journal of Modern Physics C (2007) 18: 1819-1838.
    [25] M. Mäs and A. Flache, Differentiation without distancing. Explaining bi-polarization of opinions without negative influence, PloS one, 8 (2013), e74516.
    [26] Envelope theorems for arbitrary choice sets. Econometrica (2002) 70: 583-601.
    [27] Simulation of Sznajd sociophysics model with convincing single opinions. International Journal of Modern Physics C (2001) 12: 1091-1092.
    [28] (2014) Interacting Multiagent Systems: Kinetic Equations and Monte Carlo Methods. Oxford: Oxford University Press.
    [29] Measure-valued opinion dynamics. M3AS: Mathematical Models and Methods in Applied Sciences (2020) 30: 225-260.
    [30] M. Pérez-Llanos, J. P. Pinasco and N. Saintier, Opinion attractiveness and its effect in opinion formation models, Phys. A, 559 (2020), 125017, 9 pp. doi: 10.1016/j.physa.2020.125017
    [31] Opinion formation models with heterogeneous persuasion and zealotry. SIAM Journal on Mathematical Analysis (2018) 50: 4812-4837.
    [32] F. Vazquez, N. Saintier and J. P. Pinasco, The role of voting intention in public opinion polarization, Phys. Rev. E, 101 (2020), 012101, 13pp.
    [33] N. Saintier, J. P. Pinasco and F. Vazquez, A model for a phase transition between political mono-polarization and bi-polarization, Chaos: An Interdisciplinary Journal of Nonlinear Science, 30 (2020), 063146, 17 pp. doi: 10.1063/5.0004996
    [34] Modeling opinion dynamics: Theoretical analysis and continuous approximation. Chaos, Solitons & Fractals (2017) 98: 210-215.
    [35] F. Slanina and H. Lavicka, Analytical results for the Sznajd model of opinion formation, The European Physical Journal B, 35 (2003) 279–288. doi: 10.1140/epjb/e2003-00278-0
    [36] (1993) Probability Theory, An Analytic View.Cambridge University Press.
    [37] Opinion evolution in closed community. International Journal of Modern Physics - C (2000) 11: 1157-1165.
    [38] Kinetic models of opinion formation. Communications in Mathematical Sciences (2006) 4: 481-496.
    [39] C. Villani, Topics in optimal transportation, Grad.Studies in Math., American Mathematical Soc., (2003). doi: 10.1090/gsm/058
  • This article has been cited by:

    1. Natalia Lucía Kontorovsky, Carlo Giambiagi Ferrari, Juan Pablo Pinasco, Nicolas Saintier, Kinetic modeling of coupled epidemic and behavior dynamics: The social impact of public policies, 2022, 32, 0218-2025, 2037, 10.1142/S0218202522500488
    2. Juan Pablo Pinasco, Mauro Rodriguez Cartabia, Nicolas Saintier, Interacting particles systems with delay and random delay differential equations, 2022, 214, 0362546X, 112524, 10.1016/j.na.2021.112524
    3. Juan Pablo Pinasco, Mauro Rodriguez Cartabia, Nicolas Saintier, Evolutionary game theory in mixed strategies: From microscopic interactions to kinetic equations, 2021, 14, 1937-5077, 115, 10.3934/krm.2020051
    4. Lucía Pedraza, Juan Pablo Pinasco, Nicolas Saintier, Pablo Balenzuela, An analytical formulation for multidimensional continuous opinion models, 2021, 152, 09600779, 111368, 10.1016/j.chaos.2021.111368
    5. Azmy S Ackleh, Rainey Lyons, Nicolas Saintier, Finite difference schemes for a size structured coagulation-fragmentation model in the space of Radon measures, 2022, 0272-4979, 10.1093/imanum/drac071
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1894) PDF downloads(219) Cited by(5)

Figures and Tables

Figures(4)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog