Skip to main content
Log in

A Quantum-Like View to a Generalized Two Players Game

  • Published:
International Journal of Theoretical Physics Aims and scope Submit manuscript

Abstract

This paper consider the possibility of using some quantum tools in decision making strategies. In particular, we consider here a dynamical open quantum system helping two players, \(\mathcal {G}_{1}\) and \(\mathcal {G}_{2}\), to take their decisions in a specific context. We see that, within our approach, the final choices of the players do not depend in general on their initial mental states, but they are driven essentially by the environment which interacts with them. The model proposed here also considers interactions of different nature between the two players, and it is simple enough to allow for an analytical solution of the equations of motion.

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.

Institutional subscriptions

Similar content being viewed by others

Roman Rietsche, Christian Dremel, … Jan-Marco Leimeister

Notes

  1. Notice that in any choice X 1 Y 2 the indices 1 and 2 refer to the players, while X and Y refer to the possible choices of the players, 0 or 1.

  2. This is what in the literature is called loss-aversion

  3. In principle we should use a discrete variable to label each element of the reservoirs. However, since integrals are quite often easier to be computed than series, as usually done in the literature we consider this label to be real.

  4. Please consider that h is not really the full hamiltonian, see (2.5). In fact, \(\hat {n}_{j}(t)\) stays really constant in time if \([h,\hat {n}_{j}]=0\), which is surely true if λ 1=λ 2=0 and if μ e x =μ c o o p =0, see (2.5).

References

  1. Baaquie, B.E.: Quantum Finance. Cambridge University Press, Cambridge (2004)

    Book  MATH  Google Scholar 

  2. Khrennikov, A.Y.: Information Dynamics in Cognitive, Psychological, Social and Anomalous Phenomena. Kluwer, Dordrecht (2004)

    Book  MATH  Google Scholar 

  3. Khrennikov, A.: Ubiquitous Quantum Structure: from Psychology to Finances. Springer, Berlin (2010)

    Book  Google Scholar 

  4. Bagarello, F.: Quantum Dynamics for Classical Systems: with Applications of the Number Operator. Wiley, New York (2012)

    Book  Google Scholar 

  5. Haven, E., Khrennikov, A.: Quantum Social Science. Cambridge University Press, New York (2013)

    Book  Google Scholar 

  6. Busemeyer, J.R., Bruza, P.D.: Quantum Models of Cognition and Decision. Cambridge University Press, Cambridge (2012)

    Book  Google Scholar 

  7. Asano, M., Ohya, M., Tanaka, Y., Basieva, I., Khrennikov, A.: Quantum-like model of brain’s functioning: decision making from decoherence. J. Theor. Biol. 281, 56–64 (2011)

    Article  MathSciNet  Google Scholar 

  8. Asano, M., Ohya, M., Tanaka, Y., Basieva, I., Khrennikov, A.: Quantum-like dynamics of decision-making. Phys. A 391, 2083–2099 (2012)

    Article  Google Scholar 

  9. Manousakis, E.: Quantum formalism to describe binocular rivalry. Biosystems 98(2), 5766 (2009)

    Article  Google Scholar 

  10. Martínez-Martínez, I.: A connection between quantum decision theory and quantum games: the Hamiltonian of strategic interaction. J. Math. Psychology 58, 3344 (2014)

  11. Busemeyer, J.R., Pothos, E.M.: A quantum probability explanation for violations of ’rational’ decision theory. Proc. R. Soc. B. doi:10.1098/rspb.2009.0121

  12. Agrawal, P.M., Sharda, R.: OR Forum - Quantum mechanics and human decision making. Oper. Res. 61(1), 1–16 (2013)

    Article  MATH  MathSciNet  Google Scholar 

  13. Conte, E., Todarello, O., Federici, A., Vitiello, F., Lopane, M., Khrennikov, A., Zbilut, J.P.: Some remarks on an experiment suggesting quantum-like behavior of cognitive entities and formulation of an abstract quantum mechanical formalism to describe cognitive entity and its dynamics. Chaos Solitons Fractals 31, 1076–1088 (2006)

    Article  ADS  Google Scholar 

  14. Busemeyer, J.R., Wang, Z., Townsend, J.T.: Quantum dynamics of human decision-making. J. Math. Psyc. 50, 220–241 (2006)

    Article  MATH  MathSciNet  Google Scholar 

  15. Pothos, E.M., Perry, G., Corr, P.J., Matthew, M.R., Busemeyer, J.R.: Understanding cooperation in the Prisoners Dilemma game. Personal. Individ. Differ. 51, 210215 (2011)

  16. Manousakis, E.: Founding quantum theory on the basis of consciousness. Found. of Phys. 36(6), 795–838 (2006)

    Article  MathSciNet  ADS  Google Scholar 

  17. Donald, M.J.: Quantum theory and the brain. Proc. Roy. Soc. Lond. A 427, 43–93 (1990)

    Article  MATH  MathSciNet  ADS  Google Scholar 

  18. Khrennikov, A.: On the physical basis of the theory of mental waves. Neuroquantology Suppl. 1 8(4), S71–S80 (2010)

    MathSciNet  Google Scholar 

  19. Roman, P.: Advanced Quantum Mechanics. Addison–Wesley, New York (1965)

    Google Scholar 

  20. Merzbacher, E.: Quantum Mechanics. Wiley, New York (1970)

    Google Scholar 

Download references

Acknowledgments

The author acknowledges partial support from Palermo University and from G.N.F.M. The author also thanks the referees for their suggestions, useful to produce an improved version of the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to F. Bagarello.

Appendix: Few Results on the Number Representation

Appendix: Few Results on the Number Representation

To keep the paper self-contained, we discuss here few important facts in quantum mechanics and in the so–called number representation. More details can be found, for instance, in [19, 20], as well as in[4].

Let \(\mathcal {H}\) be an Hilbert space, and \(B(\mathcal {H})\) the set of all the (bounded) operators on \(\mathcal {H}\). Let \(\mathcal {S}\) be our physical system, and \(\mathfrak {A}\) the set of all the operators useful for a complete description of \(\mathcal {S}\), which includes the observables of \(\mathcal {S}\). For simplicity, it is convenient (but not really necessary) to assume that \(\mathfrak {A}\) coincides with \(B(\mathcal {H})\) itself. The description of the time evolution of \(\mathcal {S}\) is related to a self–adjoint operator H = H which is called the Hamiltonian of \(\mathcal {S}\), and which in standard quantum mechanics represents the energy of \(\mathcal {S}\). In this paper we have adopted the so–called Heisenberg representation, in which the time evolution of an observable \(X\in \mathfrak {A}\) is given by

$$ X(t)=\exp(iHt)X\exp(-iHt), $$
(A.1)

or, equivalently, by the solution of the differential equation

$$ \frac{dX(t)}{dt}=i\exp(iHt)[H,X]\exp(-iHt)=i[H,X(t)], $$
(A.2)

where [A,B]:=A BB A is the commutator between A and B. The time evolution defined in this way is a one–parameter group of automorphisms of \(\mathfrak {A}\).

An operator \(Z\in \mathfrak {A}\) is a constant of motion if it commutes with H. Indeed, in this case, equation (A.2) implies that \(\dot Z(t)=0\), so that Z(t)=Z for all t.

In some previous applications, [4], a special role was played by the so–called canonical commutation relations. Here, these are replaced by the so–called canonical anti–commutation relations (CAR): we say that a set of operators \(\{a_{\ell },~a_{\ell }^{\dagger }, \ell =1,2,\ldots ,L\}\) satisfy the CAR if the conditions

$$ \{a_{\ell},a_{n}^{\dagger}\}=\delta_{\ell n}\mathbb{1},\quad \{a_{\ell},a_{n}\}=\{a_{\ell}^{\dagger},a_{n}^{\dagger}\}=0 $$
(A.3)

hold true for all ,n = 1,2,…,L. Here, \(\mathbb {1}\) is the identity operator and {x,y}:=x y+y x is the anticommutator of x and y. These operators, which are widely analyzed in any textbook about quantum mechanics (see, for instance, [19, 20]) are those which are used to describe L different modes of fermions. From these operators we can construct \(\hat {n}_{\ell }=a_{\ell }^{\dagger } a_{\ell }\) and \(\hat {N}={\sum }_{\ell =1}^{L} \hat {n}_{\ell }\), which are both self–adjoint. In particular, \(\hat {n}_{\ell }\) is the number operator for the –th mode, while \(\hat {N}\) is the number operator of \(\mathcal {S}\). Compared with bosonic operators, the operators introduced here satisfy a very important feature: if we try to square them (or to rise to higher powers), we simply get zero: for instance, from (A.3), we have \(a_{\ell }^{2}=0\). This is related to the fact that fermions satisfy the Fermi exclusion principle [19].

The Hilbert space of our system is constructed as follows: we introduce the vacuum of the theory, that is a vector φ 0 which is annihilated by all the operators a : a φ 0 =0 for all = 1,2,…,L. Such a non zero vector surely exists. Then we act on φ 0 with the operators \(a_{\ell }^{\dagger }\) (but not with higher powers, since these powers are simply zero!):

$$ \varphi_{n_{1},n_{2},\ldots,n_{L}}:={(a_{1}^{\dagger})}^{n_{1}}{(a_{2}^{\dagger})}^{n_{2}}\cdots {(a_{L}^{\dagger})}^{n_{L}}\varphi_{\mathbf{0}}, $$
(A.4)

n =0,1 for all . These vectors form an orthonormal set and are eigenstates of both \(\hat {n}_{\ell }\) and \(\hat {N}\): \(\hat {n}_{\ell }\varphi _{n_{1},n_{2},\ldots ,n_{L}}=n_{\ell }\varphi _{n_{1},n_{2},\ldots ,n_{L}}\) and \(\hat {N}\varphi _{n_{1},n_{2},\ldots ,n_{L}}=N\varphi _{n_{1},n_{2},\ldots ,n_{L}},\) where \(N={\sum }_{\ell =1}^{L}n_{\ell }\). Moreover, using the CAR, we deduce that

$$\hat{n}_{\ell}\left(a_{\ell}\varphi_{n_{1},n_{2},\ldots,n_{L}}\right)=(n_{\ell}-1)(a_{\ell}\varphi_{n_{1},n_{2},\ldots,n_{L}}) $$

and

$$\hat{n}_{\ell}\left(a_{\ell}^{\dagger}\varphi_{n_{1},n_{2},\ldots,n_{L}}\right)=(n_{\ell}+1)(a_{l}^{\dagger}\varphi_{n_{1},n_{2},\ldots,n_{L}}), $$

for all . Then a and \(a_{\ell }^{\dagger }\) are called the annihilation and the creation operators. Notice that, in some sense, \(a_{\ell }^{\dagger }\) is also an annihilation operator since, acting on a state with n =1, we destroy that state.

The Hilbert space \(\mathcal {H}\) is obtained by taking the linear span of all these vectors. Of course, \(\mathcal {H}\) has a finite dimension. In particular, for just one mode of fermions, \(\dim (\mathcal {H})=2\). This also implies that, contrarily to what happens for bosons, all the fermionic operators are bounded.

The vector \(\varphi _{n_{1},n_{2},\ldots ,n_{L}}\) in (A.4) defines a vector (or number) state over the algebra \(\mathfrak {A}\) as

$$ \omega_{n_{1},n_{2},\ldots,n_{L}}(X)= \langle\varphi_{n_{1},n_{2},\ldots,n_{L}},X\varphi_{n_{1},n_{2},\ldots,n_{L}}\rangle, $$
(A.5)

where 〈 , 〉 is the scalar product in \(\mathcal {H}\). As we have discussed in [4], these states are useful to project from quantum to classical dynamics and to fix the initial conditions of the considered system.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bagarello, F. A Quantum-Like View to a Generalized Two Players Game. Int J Theor Phys 54, 3612–3627 (2015). https://doi.org/10.1007/s10773-015-2599-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10773-015-2599-x

Keywords

Navigation