Skip to main content
Log in

A Differential Game Model of Opinion Dynamics: Accord and Discord as Nash Equilibria

  • Published:
Dynamic Games and Applications Aims and scope Submit manuscript

Abstract

This paper presents a noncooperative differential (dynamic) game model of opinion dynamics with open-loop information structure. In this game, the agents’ motives are shaped by their expectations of the nature of others’ opinions as well as how susceptible they are to get influenced by others, how stubborn they are, and how quick they are willing to change their opinions on a set of issues in a prescribed time interval. These motives are independently formed by all agents. The existence of a Nash equilibrium in the network means that a collective behavior emerges out of local interaction rules and these individual motives. We prove that a unique Nash equilibrium may exist in the game under quite different circumstances. It may exist not only if there is a harmony of perceptions among the agents of the network, but also when agents have different views about the correlation among issues. The first leads to an accord in the network usually expressed as a partial consensus, and the second to a discord in the form of oscillating opinions. In the case of an accord, the harmony in the network may be in the form of similarity in pairwise conceptions about the issues but may also be an agreement on the status of a “leader” in the network. A Nash equilibrium may fail to exist only if the network is in a state of discord.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  1. Acemoglu D, Ozdaglar A (2011) Opinion dynamics and learning in social networks. Dyn Games Appl 1(1):3–49

    Article  MathSciNet  Google Scholar 

  2. Albi G, Pareschi L, Toscani G, Zanella M (2017) Recent advances in opinion modeling: control and social influence. Act Part 1:49–98

    MathSciNet  Google Scholar 

  3. Altafini C (2013) Consensus problems on networks with antagonistic interactions. IEEE Trans Autom Control 58(4):935–946

    Article  MathSciNet  Google Scholar 

  4. Amelkin V, Bullo F, Singh AK (2017) Polar opinion dynamics in social networks. IEEE Trans Autom Control 62(11):5650–5665

    Article  MathSciNet  Google Scholar 

  5. Axelrod R (1997) The dissemination of culture: a model with local convergence and global polarization. J Conflict Resolut 41(2):203–226

    Article  Google Scholar 

  6. Başar T, Olsder GJ (1998) Dynamic noncooperative game theory. SIAM, Philadelphia, PA

    Book  Google Scholar 

  7. Bikhchandani S, Hirshleifer D, Welch I (1992) A theory of fads, fashion, custom, and cultural change as informational cascades. J Polit Econ 100(5):992–1026

    Article  Google Scholar 

  8. Bindel D, Kleinberg J, Oren S (2015) How bad is forming your own opinion? Games Econ Behav 92:248–265

    Article  MathSciNet  Google Scholar 

  9. Blondel VD, Hendrickx JM, Tsitsiklis JN (2010) Continuous-time average-preserving opinion dynamics with opinion-dependent communications. SIAM J Control Optim 48(8):5214–5240

    Article  MathSciNet  Google Scholar 

  10. Boccaletti S, Bianconi G, Criado R, Del Genio CI, Gómez-Gardenes J, Romance M, Sendina-Nadal I, Wang Z, Zanin M (2014) The structure and dynamics of multilayer networks. Phys Rep 544(1):1–122

    Article  MathSciNet  Google Scholar 

  11. Deffuant G, Neau D, Amblard F, Weisbuch G (2000) Mixing beliefs among interacting agents. Adv Complex Syst 3:87–98

    Article  Google Scholar 

  12. DeGroot MH (1974) Reaching a consensus. J Am Stat Assoc 69(345):118–121

    Article  Google Scholar 

  13. Engwerda J (2005) LQ dynamic optimization and differential games. Wiley, West Sussex

    Google Scholar 

  14. Etesami SR, Başar T (2015) Game-theoretic analysis of the Hegselmann-Krause model for opinion dynamics in finite dimensions. IEEE Trans Autom Control 60(7):1886–1897

    Article  MathSciNet  Google Scholar 

  15. Ferraioli D, Goldberg PW, Ventre C (2016) Decentralized dynamics for finite opinion games. Theoret Comput Sci 648:96–115

    Article  MathSciNet  Google Scholar 

  16. Friedkin NE, Johnsen EC (1990) Social influence and opinions. J Math Sociol 15(3–4):193–206

    Article  Google Scholar 

  17. Gabbay M (2007) The effects of nonlinear interactions and network structure in small group opinion dynamics. Physica A 378(1):118–126

    Article  Google Scholar 

  18. Ghaderi J, Srikant R (2014) Opinion dynamics in social networks with stubborn agents: Equilibrium and convergence rate. Automatica 50(12):3209–3215

    Article  MathSciNet  Google Scholar 

  19. Groeber P, Lorenz J, Schweitzer F (2014) Dissonance minimization as a microfoundation of social influence in models of opinion formation. J Math Sociol 38(3):147–174

    Article  MathSciNet  Google Scholar 

  20. Hegselmann R, Krause U (2002) Opinion dynamics and bounded confidence models, analysis, and simulation. J Art Soc Soc Simulat 5(3)

  21. Heider F (1946) Attitudes and cognitive organization 21(1):107–112

  22. Horn RA, Johnson CR (1991) Topics in matrix analysis. Cambridge University Press, Cambridge

    Book  Google Scholar 

  23. Horn RA, Johnson CR (2002) Matrix analysis. Cambridge University Press, Cambridge

    Google Scholar 

  24. Huckfeldt R (2007) Unanimity, discord, and the communication of public opinion. Am J Polit Sci 51(4):978–995

    Article  Google Scholar 

  25. Iñiguez G, Török J, Yasseri T, Kaski K, Kertész J (2014) Modeling social dynamics in a collaborative environment. EPJ Data Sci 3:7

    Article  Google Scholar 

  26. Jackson MO (2010) Social and economic networks. Princeton University Press, Princeton

    Book  Google Scholar 

  27. Kirk DE (1970) Optimal control theory: an introduction. Prentice-Hall, Englewood Cliffs, NJ

    Google Scholar 

  28. Krackhardt D (2009) A plunge into networks. Science 326(5949):47–48

    Article  Google Scholar 

  29. Lorenz J (2007) Continuous opinion dynamics under bounded confidence: a survey. Int J Mod Phys C 18(12):1819–1838

    Article  Google Scholar 

  30. Macy MW, Flache A (2009) Social dynamics from the bottom up: Agent-based models of social interaction. In: The Oxford handbook of analytical sociology. Oxford University Press, pp 245–268

  31. Mirtabatabaei A, Bullo F (2012) Opinion dynamics in heterogeneous networks: convergence conjectures and theorems. SIAM J Control Optim 50(5):2763–2785

    Article  MathSciNet  Google Scholar 

  32. Morarescu IC, Girard A (2011) Opinion dynamics with decaying confidence: application to community detection in graphs. IEEE Trans Autom Control 56(8):1862–1873

    Article  MathSciNet  Google Scholar 

  33. Niazi MUB (2017) A noncooperative dynamic game model of opinion dynamics in multilayer social networks. Dissertation, Bilkent University, Turkey

  34. Niazi MUB,Özgüler AB, Yildiz A (2016) Consensus as a Nash equilibrium of a dynamic game. In: 2016 12th international conference on signal-image technology and internet-based systems (SITIS), IEEE, pp 365–372

  35. Olfati-Saber R (2007) Consensus and cooperation in multi-agent networked systems. Proc IEEE 95(1):215–233

    Article  Google Scholar 

  36. Osborne MJ, Rubinstein A (1994) A course in game theory. MIT Press, Cambridge

    MATH  Google Scholar 

  37. Özgüler AB, Yıldız A (2014) Foraging swarms as Nash equilibria of dynamic games. IEEE Trans Cybern 44(6):979–987

    Article  Google Scholar 

  38. Parsegov SE, Proskurnikov AV, Tempo R, Friedkin NE (2017) Novel multidimensional models of opinion dynamics in social networks. IEEE Trans Autom Control 62(5):2270–2285

    Article  MathSciNet  Google Scholar 

  39. Proskurnikov AV, Matveev AS, Cao M (2016) Opinion dynamics in social networks with hostile camps: consensus versus polarization. IEEE Trans Autom Control 61(6):1524–1536

    Article  Google Scholar 

  40. Ren W, Cao Y (2011) Distributed coordination of multi-agent networks: emergent problems, models, and issues. Springer, New York

    Book  Google Scholar 

  41. Ren W, Beard RW, Atkins EM (2005) A survey of consensus problems in multi-agent coordination. In: Proceedings of the American control conference, IEEE, pp 1859–1864

  42. Shi G, Proutiere A, Johansson M, Baras JS, Johansson KH (2015) Emergent behaviors over signed random dynamical networks: state-flipping model. IEEE Trans Control Netw Syst 2(2):142–153

    Article  MathSciNet  Google Scholar 

  43. Smith DD (1968) Cognitive consistency and the perception of others’ opinions. Public Opin Q 32(1):1–15

    Article  Google Scholar 

  44. Yıldız A, Özgüler AB (2015) Partially informed agents can form a swarm in a Nash equilibrium. IEEE Trans Autom Control 60(11):3089–3094

    Article  MathSciNet  Google Scholar 

  45. Yıldız A, Özgüler AB (2016) Foraging motion of swarms with leaders as Nash equilibria. Automatica 73:163–168

    Article  MathSciNet  Google Scholar 

  46. Yildiz E, Acemoglu D, Ozdaglar A, Saberi A, Scaglione A (2011) Discrete opinion dynamics with stubborn agents. https://ssrn.com/abstract=1744113

Download references

Acknowledgements

This work is supported by the Science and Research Council of Turkey (TÜBİTAK) under the project EEEAG-114E270. We would like to thank Dr. Aykut Yıldız for his contributions at the earlier stages of this project. We would also like to thank the editor and the anonymous reviewer for their invaluable suggestions to improve the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Muhammad Umar B. Niazi.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

This work is supported by the Science and Research Council of Turkey (TÜBÍTAK) under the project EEEAG-114E270.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (pdf 214 KB)

Supplementary material 2 (pdf 145 KB)

Appendix

Appendix

1.1 The Game of Opinion Dynamics is of the Linear Quadratic Type

To show that (2) is of the linear quadratic type [6, Definition 6.5], we consider

$$\begin{aligned} {\mathbf {z}}_i=[\,\varDelta _{i1}'\;\dots \;\varDelta _{i(i-1)}'\;\;\varDelta _{ii}'\;\dots \;\varDelta _{in}']' \in \mathbb {R}^{nd}, \end{aligned}$$

where

$$\begin{aligned} \varDelta _{ij}:=\left\{ \begin{array}{ll} {\mathbf {x}}_i-{\mathbf {x}}_j, &{} \text {if }i\ne j; \\ {\mathbf {x}}_i-{\mathbf {b}}_i, &{} \text {if }i=j; \end{array}\right. \end{aligned}$$

for \(i,j=1,\dots ,n\). Then, we have \({\mathbf {z}}=[\,{\mathbf {z}}_1'\;\dots \;{\mathbf {z}}_n'\,]'\in \mathbb {R}^{n^2d}\) as the state vector, which satisfies

$$\begin{aligned} \dot{{\mathbf {z}}} = \sum _{i=1}^n B_i\mathbf {u_i}, \end{aligned}$$
(15)

where \(B_i\in \mathbb {R}^{n^2d\times d}\) is given as

$$\begin{aligned} B_i=\left[ \begin{array}{ccccccc} -({\mathbf {e}}_i\otimes I_d)'&\cdots&-({\mathbf {e}}_i\otimes I_d)'&(\mathbf {1}_n\otimes I_d)'&-({\mathbf {e}}_i\otimes I_d)'&\cdots&-({\mathbf {e}}_i\otimes I_d)' \end{array}\right] ', \end{aligned}$$
(16)

where the matrix block \(\mathbf {1}_n\otimes I_d\) is the ith block and \({\mathbf {e}}_i=[\,0\dots 0\;\;1\;\;0\dots 0\,]'\in \mathbb {R}^n\) is the standard ith basis vector of \(\mathbb {R}^n\).

The cost functional (1) can now be written as

$$\begin{aligned} \mathscr {J}_i({\mathbf {z}},{\mathbf {u}}_i) = \frac{1}{2}\int _0^\tau \left( {\mathbf {z}}'L_i{\mathbf {z}} + {\mathbf {u}}_i'{\mathbf {u}}_i\right) dt, \end{aligned}$$
(17)

where \(L_i=H_i\otimes K_i \in \mathbb {R}^{n^2d\times n^2d}\) with \(H_i\in \mathbb {R}^{n\times n}\) whose jkth entry, for \(j,k=1,\dots ,n\), is defined as

$$\begin{aligned}{}[H_i]_{jk} = \left\{ \begin{array}{ll} 1, &{} \text {if } j=k=i; \\ 0, &{} \text {otherwise}; \end{array}\right. \end{aligned}$$

and \(K_i=\text {diag}[\,W_{i1},\dots ,W_{in}\,]\in \mathbb {R}^{nd\times nd}\). From the above formulation, it can be seen that the game described by (15) and (17) is a linear quadratic differential game. By Assumption 1, \(L_{i}\) is positive semi-definite, which gives that (17), equivalently (1), is a strictly convex function of \({\mathbf {u}}_i(t)\). Hence, the Nash equilibrium, if it exists, is unique [6, Proof of Theorem 6.12]. Thus, we can skip a more familiar approach of coupled Riccati differential equations, which involve manipulations with large sparse matrices like \(L_i\) above.

1.2 Proof of Theorem 1 Completed

The necessary conditions ([6, Theorem 6.11], [27, Chapter 5]) for the existence of a minimum of the cost function (1) are

$$\begin{aligned} {\mathbf {u}}_i^*(t)= & {} \arg \min _{{\mathbf {u}}_i\in {\mathsf {S}}_i} \mathcal{{H}}_i(\mathbf {p}_i,{\mathbf {x}},{\mathbf {b}}_i,{\mathbf {u}}_i), \nonumber \\ \dot{\mathbf {x}}_i(t)= & {} \frac{\partial \mathcal{{H}}_i}{\partial \mathbf {p}_i}, \ \ \dot{\mathbf {p}}_i(t) = -\frac{\partial \mathcal{{H}}_i}{\partial {\mathbf {x}}_i}, \nonumber \\ {\mathbf {x}}_i(0)= & {} {\mathbf {b}}_i, \ \ \mathbf {p}_i(\tau ) = 0, \ \ i\in {\mathsf {N}}, \end{aligned}$$
(18)

where \(\mathbf {p}_i:[0,\tau ] \rightarrow {\mathbb {R}}^{nd}\) is a costate function and \(\mathcal{{H}}_i\) is the Hamiltonian given by

$$\begin{aligned} \mathcal{{H}}_i= & {} \frac{1}{2}\left\{ \left[ \sum _{j\in {\mathsf {N}}_i} \varDelta _{ij}' W_{ij} \varDelta _{ij} \right] + \varDelta _{ii}' W_{ii} \varDelta _{ii} + {\mathbf {u}}_i' {\mathbf {u}}_i \right\} +\,\mathbf {p}_i' {\mathbf {u}}_i. \end{aligned}$$

Let \(\mathbf {p} = [\mathbf {p}_{1}'\;\cdots \;\mathbf {p}_{n}']'\) and note that the combined state and costate equations are, by (18),

$$\begin{aligned} \left[ \begin{array}{c} \dot{\mathbf {x}} \\ \dot{\mathbf {p}} \end{array} \right] = \left[ \begin{array}{cc} 0 &{} -I \\ -Q &{} 0 \end{array} \right] \left[ \begin{array}{c} {\mathbf {x}} \\ \mathbf {p} \end{array} \right] + \left[ \begin{array}{cc} 0 &{} 0 \\ W &{} 0 \end{array} \right] \left[ \begin{array}{c} {\mathbf {b}} \\ \mathbf {p}(0) \end{array} \right] , \end{aligned}$$

which has the solution of the form

$$\begin{aligned} \left[ \begin{array}{c} {\mathbf {x}} \\ \mathbf {p} \end{array} \right] = \left( \varPhi (t) + \varPsi (t)B \right) \left[ \begin{array}{c} {\mathbf {b}} \\ \mathbf {p}(0) \end{array} \right] , \end{aligned}$$
(19)

where

$$\begin{aligned} \varPhi (t)= & {} e^{At} = \left[ \begin{array}{cc} \phi _{11}(t) &{} \phi _{12}(t) \\ \phi _{21}(t) &{} \phi _{22}(t) \end{array} \right] , \\ \varPsi (t)= & {} \int _0^t e^{A(t-\mu )} d\mu = \left[ \begin{array}{cc} \psi _{11}(t) &{} \psi _{12}(t) \\ \psi _{21}(t) &{} \psi _{22}(t) \end{array} \right] , \\ A= & {} \left[ \begin{array}{cc} 0 &{} -I \\ -Q &{} 0 \end{array} \right] , \qquad B = \left[ \begin{array}{cc} 0 &{} 0 \\ W &{} 0 \end{array} \right] . \end{aligned}$$

Note that

$$\begin{aligned} \begin{array}{lll} \varPhi (t) &{}=&{} {\mathscr {L}}^{-1}\{(sI - A)^{-1}\} \\ &{}=&{} {\mathscr {L}}^{-1} \bigg \{ { \left[ \begin{array}{cc} s(s^2I - Q)^{-1} &{} -(s^2I - Q)^{-1} \\ -Q(s^2I - Q)^{-1} &{} s(s^2I - Q)^{-1} \end{array} \right] } \bigg \}. \end{array} \end{aligned}$$
(20)

The state transition matrix \(\varPhi (t)\) and the matrix \(\varPsi (t)\) are calculated using the formal power series in \(s^{-1}\) of each block in (20) and, with (3) in view, are given by

$$\begin{aligned} \varPhi (t) = \left[ \begin{array}{cc} \phi _{11}(t) &{} \phi _{12}(t) \\ \phi _{21}(t) &{} \phi _{22}(t) \end{array} \right] = \left[ \begin{array}{cc} f_{Q}(t) &{} -g_{Q}(t) \\ -Qg_{Q}(t) &{} f_{Q}(t) \end{array} \right] , \\ \varPsi (t) = \left[ \begin{array}{cc} \psi _{11}(t) &{} \psi _{12}(t) \\ \psi _{21}(t) &{} \psi _{22}(t) \end{array} \right] = \left[ \begin{array}{cc} g_{Q}(t) &{} -h_{Q}(t) \\ -Qh_{Q}(t) &{} g_{Q}(t) \end{array} \right] . \end{aligned}$$

From (19), we have

$$\begin{aligned} {\mathbf {x}}(t)=[\phi _{11}(t)+\psi _{12}(t)W] {\mathbf {b}}+\phi _{12}(t)\mathbf {p}(0). \end{aligned}$$

Also,

$$\begin{aligned} \mathbf {p}(t)=~[\phi _{21}(t)+\psi _{22}(t)W]{\mathbf {b}}+\phi _{22}(t)\mathbf {p}(0). \end{aligned}$$

Evaluating at \(t=\tau \) and employing the boundary condition \(\mathbf {p}(\tau ) = 0\), we obtain

$$\begin{aligned} \phi _{22}(\tau ) \mathbf {p}(0)= -[\phi _{21}(\tau ) + \psi _{22}(\tau )W] {\mathbf {b}}, \end{aligned}$$

or substituting the expressions in terms of \(f_{Q}(t), g_{Q}(t), h_{Q}(t)\) above,

$$\begin{aligned} f_{Q}(\tau )\mathbf {p}(0)= g_{Q}(\tau )(Q-W) {\mathbf {b}} \end{aligned}$$
(21)

and

$$\begin{aligned} {\mathbf {x}}(t)={\mathbf {b}}+h_{Q}(t)(Q-W){\mathbf {b}}-g_{Q}(t)\mathbf {p}(0). \end{aligned}$$
(22)

If \(f_{Q}(\tau )\) is nonsingular, then a unique solution \(\mathbf {p}(0)\) to (21) exists for all \({\mathbf {b}}\). If \(f_{Q}(\tau )\) is singular, then there is a negative eigenvalue \(-r^2\) of Q. Let \(\mathbf {v}\) be a left eigenvector associated with \(-r^2\) and note that

$$\begin{aligned} \begin{array}{l} \mathbf {v}' Q= -r^2 \mathbf {v}', \mathbf {v}' f_{Q}(\tau )= \cos (r\tau ) \mathbf {v}', \mathbf {v}' g_{Q}(\tau )= r^{-1}\,\sin (r\tau ) \mathbf {v}'. \end{array} \end{aligned}$$

It follows, by multiplying (21) by \(\mathbf {v}'\) on the left, that \(\cos (r\tau )\mathbf {v}'\mathbf {p}(0)=-r^{-1}\sin (r\tau )\mathbf {v}'(r^2 I+W){\mathbf {b}}\). If \(r\tau \) is an odd multiple of \(\pi /2\), then the left-hand side is zero and the right-hand side is nonzero for at least \({\mathbf {b}}=\mathbf {v}\) as \(r^2 I+W\) is symmetric, positive definite. It follows that \(f_{Q}(\tau )\) must be nonsingular for a solution to exist for all \({\mathbf {b}}\).

Summarizing, in order to be able to solve (21) uniquely for \(\mathbf {p}(0)\) for any initial state \({\mathbf {b}}\) it is necessary and sufficient that \(f_{Q}(\tau )\) is invertible. This establishes the necessity and sufficiency of the condition (C1) of Theorem 1(i) for the existence of a unique Nash solution. If \(f_{Q}(\tau )\) is invertible, we then obtain \(\mathbf {p}(0)=f_{Q}(\tau )^{-1}g_{Q}(\tau )(Q-W){\mathbf {b}}\). Substituting in \({\mathbf {x}}(t)=[f_{Q}(t)-h_{Q}(t)W] {\mathbf {b}}-g_{Q}(t)\mathbf {p}(0)\), we obtain

$$\begin{aligned} {\mathbf {x}}(t) = [f_{Q}(t)-h_{Q}(t)W] {\mathbf {b}}-g_{Q}(t)f_{Q}(\tau )^{-1}g_{Q}(\tau )(Q-W){\mathbf {b}}, \end{aligned}$$

which gives (7) upon employing \(f_{Q}(t)=I+Qh_{Q}(t)\) and noting again that functions of Q commute. Taking the derivative with respect to t and using \(\frac{d}{dt}h_{Q}(t)=g_{Q}(t)\), \(\frac{d}{dt}g_{Q}(t)=f_{Q}(t)\), we also obtain (6). This completes the proof of Theorem 1. \(\square \)

Remark 4

(Non-unique equilibria) Here, we briefly characterize non-unique Nash profiles that may arise. Suppose \(f_{Q}(\tau )\) is singular, so that (C2) of Theorem 1 holds. Then, with \(J_{Q}=P^{-1}QP\) denoting the Jordan form of Q, Eq. (21) is equivalent to

$$\begin{aligned} \left[ \begin{array}{cc}f_{J_{r}}(\tau )&{}0\\ 0&{}f_{J}(\tau )\end{array}\right] \left[ \begin{array}{c}\mathbf {p}_{1}\\ \mathbf {p}_{2}\end{array}\right] = \left[ \begin{array}{c}G_{1}(\tau )\\ G_{2}(\tau )\end{array}\right] (Q-W){\mathbf {b}}, \end{aligned}$$

where \(\text {diag}[J_{r}, J] = J_{Q}\), \([G_{1}(\tau )'\;G_{2}(\tau )']' = P^{-1}g_{Q}(\tau )\), \([\mathbf {p}_{1}'\;\mathbf {p}_{2}']' = \mathbf {p}(0)\) with \(J_{r}\) denoting the Jordan blocks associated with the negative eigenvalue \(-r^2\) of Q; and J, associated with the other eigenvalues. It follows that, if \(\tau \) is an odd multiple of \(\pi /(2r)\), then \(f_{J_{r}}(\tau )\) is singular and \(f_{J}(\tau )\) is nonsingular. In this case, \(\mathbf {p}(0)\) and \({\mathbf {b}}\) satisfy (21) if and only if \(\mathbf {p}_{1}\) is in the null space of \(f_{J_{r}}(\tau )\), \(\mathbf {p}_{2}=[f_{J}(\tau )]^{-1}G_{2}(Q-W){\mathbf {b}}\), and \(G_{1}(Q-W){\mathbf {b}}=0\). (It is easy to see that \(\mathbf {p}_{1}\) is in the null space of \(f_{J_{r}}(\tau )\) if and only if \(\mathbf {v}'f_{Q}(\tau )\mathbf {p}(0)=0\) for all left eigenvectors of Q associated with \(-r^2\).) Since the null space of \(f_{J_{r}}(\tau )\) for \(\tau \) that are odd multiples of \(\pi /(2r)\) is nontrivial, \(\mathbf {p}_{1}\ne 0\) results in infinitely many different \({\mathbf {x}}(t)\) in (22). Among \({\mathbf {b}}\) such that \(G_{1}(Q-W){\mathbf {b}}=0\), one can distinguish between \({\mathbf {b}}=\tilde{\mathbf {b}}\) that are in the null space of \(Q-W\) and those with nonzero components \(\hat{\mathbf {b}}\) in the range space of \((Q-W)'\), see Remark 1(c). In the first case, the choice \(\mathbf {p}_{2}=0\) is necessary and for \(\mathbf {p}_{1}=0\), (22) gives the constant profile \({\mathbf {x}}(t)={\mathbf {b}}\), while any \({\mathbf {p}}_{1}\ne 0\) in the null space of \(f_{J_{r}}(\tau )\) gives \({\mathbf {x}}(t)={\mathbf {b}}-g_{Q}(t)\mathbf {p}(0)\) as another profile that is not constant since \(g_{Q}(t)\) is nonsingular for almost all t, by Remark 1(d). In the second, \((Q-W){\mathbf {b}}\) \(=(Q-W){\hat{\mathbf {b}}}\) gives \(\mathbf {p}_{2}=[f_{J}(\tau )]^{-1}G_{2}(Q-W) {\hat{\mathbf {b}}}\) and there are again infinitely many profiles (22) that result from different choices of \(\mathbf {p}_{1}\) in the null space of \(f_{J_{r}}(\tau )\). This characterizes all non-unique solutions to (21). We emphasize that constant profiles that are in the null space of \(Q-W\) exist whether \(f_{Q}(\tau )\) is singular or not, but they lead to unique opinion profiles only if \(f_{Q}(\tau )\) is nonsingular. \(\triangle \)

1.3 Proof of Theorem 2(iii)

By Section 2.1, \(H_{p}=H_{+}-H_{-}\) for a square root \(H=H_{+}+H_{-}\) of Q so that

$$\begin{aligned} \begin{array}{lll} &{}&{}\displaystyle \lim _{\tau \rightarrow \infty } \cosh [H(\tau -t)]\cosh (H\tau )^{-1}\\ &{}&{}\quad = \displaystyle \lim _{\tau \rightarrow \infty } \left\{ \begin{array}{l} \left\{ \exp [(H_{+}+H_{-})(\tau -t)]+\exp [-(H_{+}+H_{-})(\tau -t) \right\} \\ \times \left\{ \exp [(H_{+}+H_{-})\tau ]+\exp [-(H_{+}+H_{-})\tau ] \right\} ^{-1} \end{array} \right\} \\ &{}&{}\quad = \displaystyle \exp (-H_{+}t)\exp (H_{-}t)\lim _{\tau \rightarrow \infty } \left\{ \begin{array}{l} \left\{ \exp [2H_{-}(\tau -t)]+\exp [-2H_{+}(\tau -t)]\right\} \\ \times [\exp (2H_{-}\tau )+\exp (-2H_{+}\tau )]^{-1} \end{array} \right\} \\ &{}&{}\quad = \exp (-H_{p}t), \end{array} \end{aligned}$$

where the last equality follows by

$$\begin{aligned} \lim _{\tau \rightarrow \infty }\{\exp [2H_{-}(\tau -t)]+\exp [-2H_{+}(\tau -t)]\}=I \end{aligned}$$

and by

$$\begin{aligned} \lim _{\tau \rightarrow \infty }[\exp (2H_{-}\tau )+\exp (-2H_{+}\tau ]]=I. \end{aligned}$$

Therefore, applying the limit to (8), the expression for \({\mathbf {x}}^{\text {inf}}(t\)) in (9) is obtained. This completes the proof of Theorem 2. \(\square \)

Remark 5

(Connection with the infinite horizon game) If Q is free of any real negative eigenvalue, then it has real square roots H, all of which have eigenvalues with nonzero real parts including a square root \(H_{p}\) with eigenvalues of all positive real parts. Consider the state and costate vectors given by

$$\begin{aligned} \begin{array}{l} {\mathbf {x}}(t) = [Q^{-1}W +\exp (-H_{p}t)\,(I-Q^{-1}W)] {\mathbf {b}},\\ \mathbf {p}(t) = H_{p}\exp (-H_{p}t)\,(I-Q^{-1}W) {\mathbf {b}}. \end{array} \end{aligned}$$
(23)

It is straightforward to verify that state–costate equation above is satisfied for any initial state \({\mathbf {x}}(0)={\mathbf {b}}\). Moreover, the final condition \(\lim _{t\rightarrow \infty }\mathbf {p}(t)=\mathbf {0}\) also holds, due to the fact that \(H_{p}\) has all its eigenvalues with positive real parts. It follows that \({\mathbf {x}}(t)={\mathbf {x}}^{\text {inf}}(t)\) satisfies the necessary conditions for the infinite horizon game \(\tau \rightarrow \infty \). In order for these to constitute the unique solution for that game, we may revert to counterparts for n players of Theorems 7.16 and 7.10 in [13] and show that the corresponding matrix “M” has a strongly stable solution. The fact that this can be done in a number of special cases suggests that the limiting Nash solution of Theorem 2(iii) is also the unique solution of the infinite horizon game.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Niazi, M.U.B., Özgüler, A.B. A Differential Game Model of Opinion Dynamics: Accord and Discord as Nash Equilibria. Dyn Games Appl 11, 137–160 (2021). https://doi.org/10.1007/s13235-020-00350-3

Download citation

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s13235-020-00350-3

Keywords

Navigation