Skip to main content
Log in

Open Quantum Random Walks: Ergodicity, Hitting Times, Gambler’s Ruin and Potential Theory

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

In this work we study certain aspects of open quantum random walks (OQRWs), a class of quantum channels described by Attal et al. (J Stat Phys 147: 832–852, 2012). As a first objective we consider processes which are nonhomogeneous in time, i.e., at each time step, a possibly distinct evolution kernel. Inspired by a spectral technique described by Saloff-Coste and Zúñiga (Stoch Proc Appl 117: 961–979, 2007), we define a notion of ergodicity for finite nonhomogeneous quantum Markov chains and describe a criterion for ergodicity of such objects in terms of singular values. As a second objective, and based on a quantum trajectory approach, we study a notion of hitting time for OQRWs and we see that many constructions are variations of well-known classical probability results, with the density matrix degree of freedom on each site giving rise to systems which are seen to be nonclassical. In this way we are able to examine open quantum versions of the gambler’s ruin, birth-and-death chain and a basic theorem on potential theory.

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.

Similar content being viewed by others

References

  1. Accardi, L., Koroliuk, D.: Quantum Markov chains: the recurrence problem. Quantum Prob. Relat. Top. VII, 6373 (1991)

    MathSciNet  MATH  Google Scholar 

  2. Accardi, L., Koroliuk, D.: Stopping times for quantum Markov chains. J. Theor. Probab. 5(3), 521–535 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  3. Alicki, R., Fannes, M.: Quantum Dynamical Systems. Oxford University Press, Oxford (2000)

    MATH  Google Scholar 

  4. Attal, S., Petruccione, F., Sabot, C., Sinayskiy, I.: Open quantum random walks. J. Stat. Phys. 147, 832–852 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  5. Attal, S., Guillotin-Plantard, N., Sabot, C.: Central limit theorems for open quantum random walks and quantum measurement records. Ann. Henri Poincaré 16, 15–43 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  6. Baraviera, A., Lardizabal, C.F., Lopes, A.O., Terra Cunha, M.: Quantum stochastic processes, quantum iterated function systems and entropy. São Paulo Journ. Math. Sci. 5(1), 51–84 (2011)

    MathSciNet  MATH  Google Scholar 

  7. Benatti, F.: Dynamics, Information and Complexity in Quantum Systems. Springer, Berlin (2009)

    MATH  Google Scholar 

  8. Biane, P., Bouten, L., Cipriani, F., Konno, N., Privault, N., Xu, Q.: Quantum Potential Theory. Lecture Notes in Mathematics, vol. 1954. Springer, Berlin (2008)

  9. Bourgain, J., Grünbaum, F.A., Velázquez, L., Wilkening, J.: Quantum recurrence of a subspace and operator-valued Schur functions. Commun. Math. Phys. 329, 1031–1067 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Brémaud, P.: Markov Chains: Gibbs Fields, Monte Carlo Simulation and Queues. Texts in Applied Mathematics 31. Springer, Berlin (1999)

  11. Burgarth, D., Chiribella, G., Giovannetti, V., Perinotti, P., Yuasa, K.: Ergodic and mixing quantum channels in finite dimensions. New J. Phys. 15, 073045 (2013)

    Article  ADS  MathSciNet  Google Scholar 

  12. Burgarth, D., Giovannetti, V.: The generalized Lyapunov theorem and its application to quantum channels. New J. Phys. 9, 150 (2007)

    Article  ADS  MathSciNet  Google Scholar 

  13. Carbone, R., Pautrat, Y.: Homogeneous open quantum random walks on a lattice. Preprint arXiv:1408.1113v2

  14. Carbone, R., Pautrat, Y.: Open quantum random walks: reducibility, period, ergodic properties. Preprint arXiv:1405.2214v3

  15. de Oliveira, C.R.: Intermediate Spectral Theory and Quantum Dynamics. Birkhäuser Verlag, Basel (2009)

    Book  MATH  Google Scholar 

  16. Fagnola, F., Rebolledo, R.: Transience and recurrence of quantum Markov semigroups. Probab. Theory Relat. Fields 126, 289306 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  17. Grimmett, G.R., Stirzaker, D.R.: Probability and Random Processes, 3rd edn. Oxford University Press, Oxford (2001)

    MATH  Google Scholar 

  18. Grünbaum, F.A., Velázquez, L., Werner, A.H., Werner, R.F.: Recurrence for discrete time unitary evolutions. Commun. Math. Phys. 320, 543569 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  19. Gudder, S.: Quantum Markov chains. J. Math. Phys. 49, 072105 (2008)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Gudder, S.: Transition effect matrices and quantum Markov chains. Found. Phys. 39, 573–592 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  21. Horn, R.A., Johnson, C.R.: Matrix Analysis. Cambridge University Press, Cambridge (1985)

    Book  MATH  Google Scholar 

  22. Horn, R.A., Johnson, C.R.: Topics in Matrix Analysis. Cambridge University Press, Cambridge (1991)

    Book  MATH  Google Scholar 

  23. Konno, N., Yoo, H.J.: Limit theorems for open quantum random walks. J. Stat. Phys. 150, 299–319 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  24. Kümmerer, B., Maassen, H.: A pathwise ergodic theorem for quantum trajectories. J. Phys. A Math. Gen. 37, 11889–11896 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. Lardizabal, C.F.: A quantization procedure based on completely positive maps and Markov operators. Quantum Inf. Process. 12, 1033–1051 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  26. Lardizabal, C.F., Souza, R.R.: On a class of quantum channels, open random walks and recurrence. J. Stat. Phys. 159, 772–796 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Liu, C., Petulante, N.: On limiting distributions of quantum Markov chains. Int. J. Math. Math. Sci., Vol 2011 (2011). ID 740816

  28. Liu, C., Petulante, N.: Asymptotic evolution of quantum walks on the N-cycle subject to decoherence on both the coin and position degrees of freedom. Phys. Rev. E 81, 031113 (2010)

    Article  ADS  MathSciNet  Google Scholar 

  29. Lozinski, A., Życzkowski, K., Słomczyński, W.: Quantum iterated function systems. Phys. Rev. E 68, 04610 (2003)

    Article  MathSciNet  Google Scholar 

  30. Lytvynov, E., Mei, L.: On the correlation measure of a family of commuting Hermitian operators with applications to particle densities of the quasi-free representations of the CAR and CCR. J. Funct. Anal. 245, 6288 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  31. Mukhamedov, F.: Weak ergodicity of nonhomogeneous Markov chains on noncommutative \(L^1\)-spaces. Banach J. Math. Anal. 7(2), 53–73 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  32. Nielsen, M.A., Chuang, I.L.: Quantum Computation and Quantum Information. Cambridge University Press, Cambridge (2000)

    MATH  Google Scholar 

  33. Norris, J.R.: Markov Chains. Cambridge University Press, Cambridge (1997)

    Book  MATH  Google Scholar 

  34. Novotný, J., Alber, G., Jex, I.: Asymptotic evolution of random unitary operations. Cent. Eur. J. Phys. 8(6), 1001–1014 (2010)

    Google Scholar 

  35. Novotný, J., Alber, G., Jex, I.: Asymptotic properties of quantum Markov chains. J. Phys. A Math. Theor. 45, 485301 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  36. Paulsen, V.: Completely Bounded Maps and Operator Algebras. Cambridge University Press, Cambridge (2003)

    Book  MATH  Google Scholar 

  37. Pellegrini, C.: Continuous time open quantum random walks and non-Markovian Lindblad master equations. J. Stat. Phys. 154, 838–865 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. Petz, D.: Quantum Information Theory and Quantum Statistics. Springer, Berlin (2008)

    MATH  Google Scholar 

  39. Platis, A., Limnios, N., Le Du, M.: Hitting time in a finite non-homogeneous Markov chain with applications. Appl. Stoch. Models Data Anal. 14, 241–253 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  40. Portugal, R.: Quantum Walks and Search Algorithms. Springer, Berlin (2013)

    Book  MATH  Google Scholar 

  41. Saloff-Coste, L., Zúñiga, J.: Convergence of some time inhomogeneous Markov chains via spectral techniques. Stoch. Proc. Appl. 117, 961–979 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  42. Szehr, O., Wolf, M.M.: Perturbation bounds for quantum Markov processes and their fixed points. J. Math. Phys. 54, 032203 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  43. Temme, K., Kastoryano, M.J., Ruskai, M.B., Wolf, M.M., Verstraete, F.: The \(\chi ^2\)-divergence and mixing times of quantum Markov processes. J. Math. Phys. 51, 12201 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  44. Venegas-Andraca, S.E.: Quantum walks: a comprehensive review. Quantum Inf. Process. 11, 1015–1106 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  45. Watrous, J.: Theory of Quantum Information Lecture Notes from Fall 2011. Institute for Quantum Computing, University of Waterloo

  46. Wolf, M.M.: Quantum Channels & Operations Guided Tour Lecture Notes (unpublished). 5 July 2012

Download references

Acknowledgments

The authors are grateful to an anonymous referee for several useful suggestions that led to marked improvements of the paper. C.F.L. is partially supported by a CAPES/PROAP grant to the Graduate Program in Mathematics—PPGMat/UFRGS. R.R.S. is partially supported by FAPERGS (proc. 002063-2551/13-0). The authors would like to thank C. Liu, T. Machida, S. E. Venegas-Andraca, N. Petulante, F. Petruccione and F. A. Grünbaum for stimulating discussions on this line of research.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Carlos F. Lardizabal.

Appendix: Proofs

Appendix: Proofs

Proof of Proposition 4.2

We have for \(B=(B_{ij})\) QTM, for all j,

$$\begin{aligned}&\sum _{i=1}^n \left\| B_{j}^{i*}B_{j}^i-\frac{I}{n}\right\| _2^2=\sum _{i=1}^n \left\langle B_{j}^{i*}B_{j}^i,B_{j}^{i*}B_{j}^i\right\rangle _2-2\sum _{i=1}^n\left\langle B_{j}^{i*}B_{j}^i,\frac{I}{n}\right\rangle _2+n\left\langle \frac{I}{n},\frac{I}{n}\right\rangle _2 \nonumber \\&\quad =\sum _{i=1}^ntr\big (\big (B_{j}^{i*}B_{j}^i\big )^2\big )-2\left\langle I,\frac{I}{n}\right\rangle _2+\frac{1}{n}\langle I,I\rangle _2=\sum _{i=1}^ntr\big (\big (B_{j}^{i*}B_{j}^i\big )^2\big )-\frac{1}{n}\langle I,I\rangle _2 \nonumber \\&\quad =\sum _{i=1}^n tr((B_{j}^{i*}B_{j}^i)^2) -\frac{k}{n} \end{aligned}$$
(10.1)

With (10.1) in mind, we perform another calculation. Consider an orthonormal basis of eigenstates for \(\Phi ^*\Phi \) given by \(\mathcal {F}=\{\eta _i\}_{i=1}^{N_\Phi }\), such that

$$\begin{aligned} \eta _1=\frac{1}{\sqrt{kn}}\left( I\otimes |1\rangle \langle 1|+\cdots +I\otimes |n\rangle \langle n|\right) , \;\;\;I=I_k\in M_k(\mathbb {C}), \end{aligned}$$
(10.2)

where \(I=I_k\) is the order k identity matrix. Note that \(\Vert \eta _1\Vert =1\). Recall that by Remark 2.1, whatever is the initial state \(\rho \) on \(\mathcal {H}\otimes \mathcal {K}\), the density \(\Phi (\rho )\) is of the form \(\sum _i\rho _i\otimes |i\rangle \langle i|\). This imposes a restriction on the kind of eigenstates present in \(\mathcal {F}\). Define

$$\begin{aligned} \rho _1\,:\,=\begin{bmatrix} I&0&\dots&0\end{bmatrix}^T=\sum _{i=1}^{N_\Phi } d_i\eta _i,\, \, \, d_i\in \mathbb {C}, \end{aligned}$$
(10.3)

with \(N_\Phi \) as in Remark 4.1. We have

$$\begin{aligned} \langle \Phi ^*\Phi \rho _1,\rho _1\rangle _2=\langle \Phi ^*\Phi \sum _{i} d_i\eta _i,\sum _{j} d_j\eta _j \rangle =\sum _{i,j} d_i\overline{d_j}\langle \Phi ^*\Phi \eta _i,\eta _j\rangle =\sum _{i=1}^{N_\Phi } |d_i|^2\sigma _i^2 \end{aligned}$$
(10.4)

Also,

$$\begin{aligned} \Phi (\rho _1)=\begin{bmatrix} B_{1}^1IB_{1}^{1*}&B_{1}^2IB_{1}^{2*}&\cdots&B_{1}^nIB_{1}^{n*}\end{bmatrix}^T, \end{aligned}$$
(10.5)

so

$$\begin{aligned} \langle \Phi (\rho _1),\Phi (\rho _1)\rangle _2= & {} \left\langle \begin{bmatrix} B_{1}^1IB_{1}^{1*} \\ B_{1}^2IB_{1}^{2*} \\ \vdots \\ B_{1}^nIB_{1}^{n*}\end{bmatrix},\begin{bmatrix} B_{1}^1IB_{1}^{1*} \\ B_{1}^2IB_{1}^{2*} \\ \vdots \\ B_{1}^nIB_{1}^{n*}\end{bmatrix}\right\rangle _2=tr((B_{1}^{1*}B_{1}^1)^2)\nonumber \\&+tr((B_{1}^{2*}B_{1}^2)^2)+\cdots +tr((B_{1}^{n*}B_{1}^n)^2). \end{aligned}$$
(10.6)

Therefore on one hand we have

$$\begin{aligned} \sum _{i=1}^n tr((B_{1}^{i*}B_{1}^i)^2)=\sum _{i=1}^{N_\Phi } |d_i|^2\sigma _i^2, \end{aligned}$$
(10.7)

and on the other we obtained, by (10.1),

$$\begin{aligned} \sum _{i=1}^n \left\| B_{1}^{i*}B_{1}^i-\frac{I}{n}\right\| _2^2=\sum _{i=1}^n tr((B_{1}^{i*}B_{1}^i)^2) -\frac{k}{n} \end{aligned}$$
(10.8)

Finally, note that \(d_1=\langle \rho _1,\eta _1\rangle =\frac{1}{\sqrt{kn}}tr(I)=\frac{k}{\sqrt{kn}}\) and so \(|d_1|^2=\frac{k}{n}\). We can repeat an analogous reasoning where we define \(\rho _2\), \(\rho _3,\dots \) in a similar way as \(\rho _1\) in (10.3). \(\square \)

Proof of Theorem 4.7

Assume that \(\sigma =\max _{1,\dots , q}\sigma _2(Q_j)<1\). Let \(\{B_i\}_{i=1}^\infty \) be a sequence of OQRWs such that

$$\begin{aligned} N_l=\#\{i\in \{1,\dots ,l\}:B_i\in \mathcal {Q}\} \end{aligned}$$
(10.9)

tends to infinity with l. By (4.9) we have that for every j,

$$\begin{aligned} \left( \sum _{i=1}^n \left\| \mathcal {B}_{0,l}(i,j)\mathcal {B}_{0,l}(i,j)^*-\frac{I}{n}\right\| _2^2\right) ^{1/2}\le \sigma ^{N_l}C(j,n) \end{aligned}$$
(10.10)

which tends to zero as \(l\rightarrow \infty \). Conversely, assume that the pair \((\mathcal {Q},\rho _\pi )\) is ergodic. Then equation (4.6) holds for any sequence \((B_i)_1^\infty \) of QTMs with invariant measure \(\rho _\pi \) such that \(B_i\in \mathcal {Q}\) for infinitely many i’s. That is, the columns of the iterated product are becoming equal to I / n. By contradiction, assume that one of the \(Q_i\), say \(Q_1\) satisfies \(\sigma _2(Q_1)=1\) and consider the following sequence of QTMs: \(B_{2i+1}=Q_1\), \(B_{2i}=Q_1^*\), \(i=1,2,\dots \). Now we consider \(Q_1Q_1^*\), for which \(\sigma _2(Q_1)=1\) is an eigenvalue with algebraic and geometric multiplicity at least 2, i.e., \(\mu _\Phi (1)=\gamma _\Phi (1)\ge 2\) (see Prop. 4.5).

Now let \(\Psi _r=(Q_1Q_1^*)^r\). It is clear that each \(\Psi _r\) is a quantum channel with real spectrum in [0, 1], by the remarks preceding the statement of this theorem. By standard arguments such as the one seen in Novotný et al. [34, 35] (via Jordan blocks), the asymptotic behavior of \(\Psi _r\) is determined by the peripheral spectrum, as contributions of eigenspaces associated to eigenvalues with norm less than 1 tend to disappear. Since 1 is the only eigenvalue in the unit circle associated to \(\Psi _r\), for all r, it is clear that the limit of \(\Psi _r\) as \(r\rightarrow \infty \) exists. The QTM \(B=\lim _{r\rightarrow \infty }\Psi _r=\lim _{r\rightarrow \infty }(Q_1Q_1^*)^r\) is such that there must be two matrices B(ij) and B(lm) which are distinct. Indeed, let \(\rho _0\) be an eigenstate of B associated to eigenvalue 1. Suppose that \(\rho _0\) is not the maximally mixed column. This assumption is possible since we have that the geometric multiplicity of 1 for \(Q_1Q_1^*\) is at least 2. In particular, by writing \(\rho _0=\sum _i \eta _i\otimes |i\rangle \langle i|\) we may assume there are kl such that \(\eta _k\ne \eta _l\). Now the fact that \(B(\rho _0)=\rho _0\), corresponds to the system of equations

$$\begin{aligned} B(1,1)\eta _1 B(1,1)^{*}+\cdots +B(1,n)\eta _n B(1,n)^{*}=\eta _1 \end{aligned}$$
(10.11)
$$\begin{aligned} B(2,1)\eta _1 B(2,1)^{*}+\cdots +B(2,n)\eta _n B(2,n)^{*}=\eta _2 \end{aligned}$$
(10.12)
$$\begin{aligned} \vdots \end{aligned}$$
$$\begin{aligned} B(n,1)\eta _1 B(n,1)^{*}+\cdots +B(n,n)\eta _n B(n,n)^{*}=\eta _n \end{aligned}$$
(10.13)

If, on the contrary, all B(ij) are equal, then by considering the k-th and l-th equation we conclude that \(\eta _k=\eta _l\), which is absurd. Therefore the QTM \(B=\lim _{r\rightarrow \infty }\Psi _r\) is such that there must be two matrices B(ij) and B(lm) which are distinct.

Moreover, there must be a row in B with two different entries, that is, two matrices \(B(i,r)\ne B(i,s)\) for some irs. In fact, suppose \(i_1\) and \(i_2\) are rows where we found two distinct elements of B, say, \(B(i_1,j_1)\) and \(B(i_2,j_2)\). If all entries of row \(i_1\) are equal then these must be equal to the maximally mixed column. The same conclusion holds for row \(i_2\). But then we would have \(B(i_1,j_1)=B(i_2,j_2)\), which is absurd. We conclude there must be a row in B with two different entries. Hence we are able to obtain xyz such that

$$\begin{aligned} \lim _{r\rightarrow \infty } {\mathcal {B}}_{0,2r}(x,y)- {\mathcal {B}}_{0,2r}(x,z)\ne 0, \end{aligned}$$
(10.14)

as required. \(\square \)

Proof of Theorem 8.3

  1. (a)

    Gambler’s ruin. Let \(A=\{0\}\) so that \(h_i^A(\rho _i)\) is the ruin probability starting from i. Using Theorem 6.3 we have the system of equations

    $$\begin{aligned} \left\{ \begin{array}{ll} h_0^A(\rho _i)=1 &{} \\ h_i^A(\rho _i)=p(\rho _i)h_{i+1}^A\left( \frac{B_i^{i+1}\rho _i B_i^{i+1*}}{tr(B_i^{i+1}\rho _i B_i^{i+1*})}\right) +q(\rho _i)h_{i-1}^A\left( \frac{B_i^{i-1}\rho _i B_i^{i-1*}}{tr(B_i^{i-1}\rho _i B_i^{i-1*})}\right) &{} i=1,2,\dots \end{array} \right. \end{aligned}$$
    (10.15)

    For instance if \(i=2\), using expression (5.9) and recalling that \(B_i^{i+1}=R\), \(B_i^{i-1}=L\),

    $$\begin{aligned} h_2^A(\rho )= & {} tr(B_2^3\rho B_2^{3*})\sum _{D\in \pi (3;A)} \frac{tr(DB_2^3\rho B_2^{3*}D^*)}{tr(B_2^3\rho B_2^{3*})}\nonumber \\&+\,tr(B_2^1\rho B_2^{1*})\sum _{C\in \pi (1;A)} \frac{tr(CB_2^1\rho B_2^{1*}C^*)}{tr(B_2^1\rho B_2^{1*})} \end{aligned}$$
    (10.16)

    We will also write \(q(\rho )=tr(L\rho L^*)\) and \(p(\rho )=tr(R\rho R^*)\). Let \(\Phi _L(X)=LXL^*\) and \(\Phi _R(X)=RXR^*\). Then if \(LR=RL\) we have \(\Phi _R\Phi _L=\Phi _L\Phi _R\) so let \(\{\eta _i\}\) be a basis for \(M_2(\mathbb {C})\) consisting of eigenstates for \(\Phi _L\) and \(\Phi _R\) and write \(\Phi _L(\eta _i)=\lambda _i\eta _i\), \(\Phi _R(\eta _i)=\mu _i\eta _i\). Then we have

    $$\begin{aligned} h_2^A(\eta _i)= & {} tr(R\eta _i R^*)\sum _{D\in \pi (3;A)} \frac{tr(DB_2^3\eta _i B_2^{3*}D^*)}{tr(B_2^3\eta _i B_2^{3*})}\\&+\,tr(L\eta _i L^*)\sum _{C\in \pi (1;A)} \frac{tr(CB_2^1\eta _i B_2^{1*}C^*)}{tr(B_2^1\eta _i B_2^{1*})} \end{aligned}$$
    $$\begin{aligned} tr(R\eta _i R^*)\sum _{D\in \pi (3;A)} \frac{tr(D\mu _i\eta _i D^*)}{\mu _i tr(\eta _i)}+tr(L\eta _i L^*)\sum _{C\in \pi (1;A)} \frac{tr(C\lambda _i\eta _i C^*)}{\lambda _i tr(\eta _i)} \end{aligned}$$
    $$\begin{aligned} =p(\eta _i)h_{3}^A(\eta _i)+q(\eta _i)h_{1}^A(\eta _i) \end{aligned}$$
    (10.17)

    In a simpler notation, let \(\eta \) be such that \(\Phi _L(\eta )=\lambda \eta \) and \(\Phi _R(\eta )=\mu \eta \), \(\lambda ,\eta \in \mathbb {C}\). Then for such choice the above equation becomes

    $$\begin{aligned} h_i^A(\eta )=\mu h_{i+1}^A(\eta )+\lambda h_{i-1}^A(\eta ). \end{aligned}$$
    (10.18)

    If \(\mu \ne \lambda \) the recurrence relation has, for fixed \(\eta \), a general solution

    $$\begin{aligned} h_i^A(\eta )=B+C\left( \frac{\lambda }{\mu }\right) ^i,\,\,\,\lambda =\lambda (\eta ),\, \mu =\mu (\eta ) \end{aligned}$$
    (10.19)

    Now, if \(\mu <\lambda \) then the restriction \(0\le h_i\le 1\) forces \(C=0\) so \(h_i=1\) for all i. That is, the ruin of the quantum gambler is certain. On the other hand, if \(\mu >\lambda \) then we get the family of solutions

    $$\begin{aligned} h_{i}^{A}({\eta })=\left( \frac{\lambda }{\mu }\right) ^{i} +A\left( 1-\left( \frac{\lambda }{\mu }\right) ^{i}\right) \end{aligned}$$
    (10.20)

    For a nonnegative solution we must have \(B\ge 0\) and so the minimal solution is

    $$\begin{aligned} h_{i}^{A}(\eta )=\left( \frac{\lambda }{\mu }\right) ^{i} \end{aligned}$$
    (10.21)

    Finally if \(\mu =\lambda \), the recurrence relation has a general solution \(h_i^{A}=B+Ci\) and again the restriction \(0\le h_i^A(\eta )\le 1\) forces \(C=0\), so \(h_i^{A}(\eta )=1\) for all i.

  2. (b)

    Birth-and-death chain. We write \(p_i(\rho )=tr(R_i{\rho } R_i^{*})\), \(q_i(\rho )=tr(L_i\rho L_{i}^{*})\). We still assume commutativity of the translation operators (moves to the left or right). Let \(\Phi _{L_{i}}(\eta _{j})=\lambda _{i;j}\eta _j\), \(\Phi _{R_{i}}(\eta _{j})=\mu _{i;j}\eta _j\). Then for any eigenstate \(\eta _j\) we have

    $$\begin{aligned} h_i^{A}(\eta _j)=p_i(\eta _j)h_{i+1}^A\left( \frac{R_i\eta _j R_i^{*}}{tr(R_i\eta _j R_i^{*})}\right) +q_i(\eta _j)h_{i-1}^A\left( \frac{L_i\eta _j L_i^{*}}{tr(L_i\eta _j L_i^*)}\right) , \end{aligned}$$
    (10.22)

    thus implying, due to the eigenvalue conditions \(p_i(\eta _j)=\mu _{ij}\), \(q_i(\eta _j)=\lambda _{ij}\) and using the same calculation used in the gambler’s ruin,

    $$\begin{aligned} h_i^A(\eta _j)=\mu _{i;j}h_{i+1}^A(\eta _j)+\lambda _{i;j} h_{i-1}^A(\eta _j). \end{aligned}$$
    (10.23)

    Let \(u_i=h_{i-1}^A-h_i^A\), then

    $$\begin{aligned} \mu _{i;j}p_iu_{i+1}=\mu _{i;j}p_i(h_i-h_{i+1})=\mu _{i;j}(p_ih_i-p_ih_{i+1})=\mu _{i;j}(p_ih_i+q_ih_{i-1}-h_i), \end{aligned}$$
    (10.24)

    the last equality due to (10.22), by isolating \(p_i\) (we omit the arguments for simplicity). Then

    $$\begin{aligned}&\mu _{i;j}p_iu_{i+1}=\mu _{i;j}(h_i(p_i-1)+q_ih_{i-1})=\mu _{i;j}(-h_iq_i+q_ih_{i-1})\nonumber \\&\quad =\mu _{i;j}q_i(h_{i-1}-h_i)=\mu _{i;j}q_iu_i \end{aligned}$$
    (10.25)

    Hence \(\mu _{i;j}p_iu_{i+1}=\mu _{i;j}q_iu_i\) and so \(p_iu_{i+1}=q_iu_i\). Therefore,

    $$\begin{aligned} u_{i+1}(\eta _j)=\left( \frac{q_i(\eta _j)}{p_i(\eta _j)}\right) u_i(\eta _j)=\left( \frac{q_i(\eta _j)q_{i-1}(\eta _j)\cdots q_1(\eta _j)}{p_i(\eta _j)p_{i-1}(\eta _j)\cdots p_1(\eta _j)}\right) u_1(\eta _j)=\gamma _i(\eta _j)u_1(\eta _j) \end{aligned}$$
    (10.26)

    where

    $$\begin{aligned} \gamma _i(\eta _j):=\frac{q_i(\eta _j)q_{i-1}(\eta _j)\cdots q_1(\eta _j)}{p_i(\eta _j)p_{i-1}(\eta _j)\cdots p_1(\eta _j)}=\frac{\lambda _{i;j}\lambda _{i-1;j}\cdots \lambda _{1;j}}{\mu _{i;j}\mu _{i-1;j}\cdots \mu _{1;j}}, \end{aligned}$$
    (10.27)

    the last equality due to the eigenvalue conditions. Then note that \(u_1+\cdots +u_i=h_0^A-h_i^A\) and recalling that \(h_0^A=1\) we can write

    $$\begin{aligned} h_i^A=h_0^A-u_1-u_2-\cdots - u_i=1-\gamma _0u_1-\gamma _1u_1-\cdots -\gamma _{i-1}u_1 \end{aligned}$$
    $$\begin{aligned} =1-B(\gamma _0+\cdots +\gamma _{i-1}),\,\,\,A=u_1,\,\,\,\gamma _0=1 \end{aligned}$$
    (10.28)

    Here B is yet to be determined. If \(\sum _{i=0}^\infty \gamma _i(\eta _j)=\infty \) the restriction \(0\le h_i^A\le 1\) forces \(B=0\) and \(h_i^A=1\) for all i. If \(\sum _{i=0}^\infty \gamma _i(\eta _j)< \infty \) then we may take \(B>0\) such that

    $$\begin{aligned} 1-B(\gamma _0+\cdots +\gamma _{i-1})\ge 0,\,\,\,\forall i \end{aligned}$$
    (10.29)

    So the minimal nonnegative solution occurs when \(B=B(j)=(\sum _{k=0}^\infty \gamma _k(\eta _j))^{-1}\) and then for all j,

    $$\begin{aligned} h_i^A(\eta _j)=\frac{\sum _{k=i}^\infty \gamma _k(\eta _j)}{\sum _{k=0}^\infty \gamma _k(\eta _j)},\,\,\,\forall i=1,2,\dots \end{aligned}$$
    (10.30)

\(\square \)

Proof of Theorem 8.4

The proof is inspired by a classical description [33]. Item a) was proved before the statement of this theorem. b) Consider the expected cost up to time n:

$$\begin{aligned}&\phi _i(\rho ;n):=\sum _{X_0=i,X_1,X_2,\dots \in (-a,a), X_r\in \{-a,a\},r\le n}\left[ c(X_0)+c(X_1)+c(X_2)+\cdots +c(X_{r-1})\right. \nonumber \\&\qquad \qquad \qquad \left. +f(X_r)\right] P_i(X_1,\dots ,X_r;\rho ),\,\,\,n\ge 1 \end{aligned}$$
(10.31)

and we define \(\phi _i(\rho ;0)=0\). Then it should be clear that \(\phi _i(\rho ;n)\uparrow \phi _i(\rho )\) as \(n\rightarrow \infty \). Similarly,

$$\begin{aligned} \phi _i(\rho ;n+1)=c(i)+\sum _{j\in I}p_{ij}\phi _j\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})};n\right) ,\,\,\,n\ge 0 \end{aligned}$$
(10.32)

Now suppose \(\psi \ge 0\) satisfies

$$\begin{aligned} \psi _i(\rho )\ge c(i)+\sum _j p_{ij}(\rho )\psi _j\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) , \end{aligned}$$
(10.33)

in D and \(\psi _i(\rho )\ge f(i)\) in \(\partial D\). Correspondingly, we call (10.33) the open quantum version of \(\psi \ge c+P\psi \). Note that for all \(\rho \), \(\psi _i(\rho )\ge 0=\phi _i(\rho ;0)\) so

$$\begin{aligned} \psi _i(\rho )\ge & {} c(i)+\sum _j p_{ij}(\rho )\psi _j\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) \ge c(i)+\sum _j p_{ij}(\rho )\phi _j\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})};0\right) \nonumber \\= & {} \phi _i(\rho ;1) \end{aligned}$$
(10.34)

in D. By induction we conclude \(\psi _i(\rho )\ge \phi _i(\rho ;n)\) and hence \(\psi _i(\rho )\ge \phi _i(\rho )\) for all \(\rho \), all i. For simplicity we write \(\psi \ge \phi \).

c) Now we assume \(P_i(T<\infty )=1\) for all i, i.e., we assume the probability of ever reaching \(\partial D\) equals 1. Note that a choice of initial density \(\rho \) is implicit here. We would like to show that the open quantum version of the problem \(\phi =c+P\phi \) in D, \(\phi =f\) in \(\partial D\) has at most one bounded solution. Let \(\psi _i(\rho )\) be another solution. For \(i\in D\) we have, writing \(p_{ij}=p_{ij}(\rho )=tr(B_i^j\rho B_i^{j*})\) for simplicity,

$$\begin{aligned} \psi _i(\rho )= & {} c(i)+\sum _{j\in I}p_{ij}\psi _j\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) =c(i)+\sum _{j\in \partial D}p_{ij}f(j)\nonumber \\&+\sum _{j\in D}p_{ij}\psi _j\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) \end{aligned}$$
(10.35)

By performing a repeated substitution we get

$$\begin{aligned}&\psi _i(\rho )= c(i)+\sum _{j\in \partial D}p_{ij}(\rho )f(j)+\sum _{j\in D}p_{ij}(\rho )\psi _j\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) \\&\quad =c(i)+\sum _{j\in \partial D}p_{ij}(\rho )f(j)+\sum _{j\in D}p_{ij}(\rho )\left[ c(j)+\sum _{j_1\in \partial D}p_{jj_1}\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) f(j_1)\right. \nonumber \\&\qquad \left. +\sum _{j_1\in D}p_{jj_1}\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) \psi _{j_1}\left( \frac{B_j^{j_1}\frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})} B_j^{j_1*}}{tr(B_j^{j_1}\frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})} B_j^{j_1*})}\right) \right] \nonumber \\&\quad =c(i)+\sum _{j\in \partial D}p_{ij}(\rho )f(j)+\sum _{j\in D}p_{ij}(\rho )\left[ c(j)+\sum _{j_1\in \partial D}p_{jj_1}\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) f(j_1)\right. \nonumber \\&\qquad \left. +\sum _{j_1\in D}p_{jj_1}\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) \psi _{j_1}\left( \frac{B_j^{j_1}B_i^j\rho B_i^{j*} B_j^{j_1*}}{tr(B_j^{j_1}B_i^j\rho B_i^{j*} B_j^{j_1*})}\right) \right] \nonumber \\&\quad =c(i)+\sum _{j\in \partial D}p_{ij}(\rho )f(j)+\sum _{j\in D}p_{ij}(\rho )c(j)+\sum _{j\in D}\sum _{j_1\in \partial D}p_{ij}(\rho )p_{jj_1}\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) f(j_1)\nonumber \\ \end{aligned}$$
(10.36)
$$\begin{aligned}&\qquad +\sum _{j\in D}\sum _{j_1\in D}p_{ij}(\rho )p_{jj_1}\left( \frac{B_i^j\rho B_i^{j*}}{tr(B_i^j\rho B_i^{j*})}\right) \psi _{j_1}\left( \frac{B_j^{j_1}B_i^j\rho B_i^{j*} B_j^{j_1*}}{tr(B_j^{j_1}B_i^j\rho B_i^{j*} B_j^{j_1*})}\right) =\cdots \nonumber \\&\qquad \cdots =c(i)+\sum _{j\in \partial D} p_{ij}(\rho )f(j)+\sum _{j\in D} p_{ij}(\rho )c(j) \nonumber \\&\qquad +\cdots +\sum _{j_1\in D}\cdots \sum _{j_{n-1}\in D} p_{ij}p_{jj_1}\cdots p_{j_{n-2}j_{n-1}}c(j_{n-1})\nonumber \\&\qquad +\sum _{j_1\in D}\cdots \sum _{j_{n-1}\in D}\sum _{j_n\in \partial D} p_{ij}p_{jj_1}\cdots p_{j_{n-1}j_n}f(j_n) \nonumber \\ \end{aligned}$$
(10.37)
$$\begin{aligned}&\qquad +\sum _{j_1\in D}\cdots \sum _{j_n\in D} p_{ij_1}p_{jj_1}\cdots p_{j_{n-1}j_n}\psi _{j_n}\nonumber \end{aligned}$$
(10.38)

Note that in the last equality we have omitted the dependence of the \(p_{ij}\) on the density matrices. A convenient cancellation of terms occur in each of them. For instance, in the last term, \(p_{ij_1}\cdots p_{j_{n-1}j_n}\) simplifies to

$$\begin{aligned} p_{ij_1}\cdots p_{j_{n-1}j_n}= & {} tr(B_i^{j_1}\rho B_i^{j_1*})tr\left( \frac{B_{j_1}^{j_2}B_i^{j_1}\rho B_i^{j_1*}B_{j_1}^{j_2*}}{tr(B_i^{j_1}\rho B_i^{j_1*})}\right) \cdots \nonumber \\= & {} tr(B_{j_{n-1}}^{j_n}\cdots B_i^{j_1}\rho B_i^{j_1*}\cdots B_{j_{n-1}}^{j_n*}) \end{aligned}$$
(10.39)

Now suppose \(P_i(T<\infty )=1\) for all i and that \(|\psi _i|\le M\) then as \(n\rightarrow \infty \),

$$\begin{aligned} \left| \sum _{j_1\in D}\cdots \sum _{j_n\in D} p_{ij_1}\cdots p_{j_{n-1}j_n}\psi _{j_n}\right| \le MP_i(T\ge n)\rightarrow 0, \end{aligned}$$
(10.40)

which means the last term in (10.39) vanishes when \(n\rightarrow \infty \). Recalling definition (10.31) we obtain in the case of equality in Eqs. (10.36)–(10.39) that \(\psi _i=\lim _{n\rightarrow \infty }\phi _i(\rho ;n)=\phi _i\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lardizabal, C.F., Souza, R.R. Open Quantum Random Walks: Ergodicity, Hitting Times, Gambler’s Ruin and Potential Theory. J Stat Phys 164, 1122–1156 (2016). https://doi.org/10.1007/s10955-016-1578-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-016-1578-9

Keywords

Navigation