Skip to main content
Log in

Discrete-time quantum walks in random artificial gauge fields

  • Regular Paper
  • Published:
Quantum Studies: Mathematics and Foundations Aims and scope Submit manuscript

Abstract

Discrete-time quantum walks (DTQWs) in random artificial electric and gravitational fields are studied analytically and numerically. The analytical computations are carried by a new method which allows a direct exact analytical determination of the equations of motion obeyed by the average density operator. It is proven that randomness induces decoherence and that the quantum walks behave asymptotically like classical random walks. Asymptotic diffusion coefficients are computed exactly. The continuous limit is also obtained and discussed.

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

Similar content being viewed by others

Notes

  1. Note however that it is possible to keep all time steps if one is only interested in the continuous limit of the probability of the original walk.

  2. A simple computation shows that the partly reduced density operator \({\hat{\rho }}^{pr}_j (p) = \sum _K {\hat{\rho }}_j(K, p)\) remains flat in p at all times, as is the initial condition.

References

  1. Feynman, R.P., Hibbs, A.R.: International Series in Pure and Applied Physics. Quantum mechanics and path integrals. McGraw-Hill, New York (1965)

    MATH  Google Scholar 

  2. Aharonov, Y., Davidovich, L., Zagury, N.: Quantum random walks. Phys. Rev. A 48, 1687 (1993)

    Article  Google Scholar 

  3. Meyer, D.A.: From quantum cellular automata to quantum lattice gases. J. Stat. Phys. 85, 551–574 (1996)

  4. Schmitz, H., Matjeschk, R., Schneider, Ch., Glueckert, J., Enderlein, M., Huber, T., Schaetz, T.: Quantum walk of a trapped ion in phase space. Phys. Rev. Lett. 103(090504), 090504 (2009)

    Article  Google Scholar 

  5. Zähringer, F., Kirchmair, G., Gerritsma, R., Solano, E., Blatt, R., Roos, C.F.: Realization of a quantum walk with one and two trapped ions. Phys. Rev. Lett. 104, 100503 (2010)

    Article  Google Scholar 

  6. Schreiber, A., Cassemiro, K.N., Gábris, A., Potoček, V., Mosley, P.J., Andersson, E., Jex, I., Silberhorn, Ch.: Photons walking the line. Phys. Rev. Lett. 104(050502), 050502 (2010)

    Article  Google Scholar 

  7. Karski, M., Förster, L., Cho, J.-M., Steffen, A., Alt, W., Meschede, D., Widera, A.: Quantum walk in position space with single optically trapped atoms. Science 325(5937), 174–177 (2009)

    Article  Google Scholar 

  8. Sansoni, L., Sciarrino, F., Vallone, G., Mataloni, P., Crespi, A., Ramponi, R., Osellame, R.: Two-particle bosonic–fermionic quantum walk via 3D integrated photonics. Phys. Rev. Lett. 108(010502), 010502 (2012)

    Article  Google Scholar 

  9. Sanders, B.C., Bartlett, S.D., Tregenna, B., Knight, P.L.: Two-particle bosonic–fermionic quantum walk via 3D integrated photonics. Phys. Rev. A 67, 042305 (2003)

    Article  Google Scholar 

  10. Perets, B., Lahini, Y., Pozzi, F., Sorel, M., Morandotti, R., Silberberg, Y.: Realization of quantum walks with negligible decoherence in waveguide lattices. Phys. Rev. Lett. 100, 170506 (2008)

    Article  Google Scholar 

  11. Giulini, D., Joos, E., Kiefer, C., Kupsch, J., Stamatescu, I.-O., Zeh, H.D.: Decoherence and the Appearance of a Classical World in Quantum Theory. Springer, Berlin (1996)

    Book  MATH  Google Scholar 

  12. Ambainis, A.: Quantum walk algorithm for element distinctness. SIAM J. Comput. 37, 210–239 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  13. Magniez, F.,  Roland, J.,  Nayak, A.,  Santha, M.: Search via quantum walk. In: SIAM Journal on Computing—Proceedings of the 39th Annual ACM Symposium on Theory of Computing. ACM, New York (2007)

  14. Aslangul, C.: Quantum dynamics of a particle with a spin-dependent velocity. J. Phys. A Math. Theor. 38, 1–16 (2005)

    MathSciNet  MATH  Google Scholar 

  15. Bose, S.: Quantum communication through an unmodulated spin chain. Phys. Rev. Lett. 91, 207901 (2003)

    Article  Google Scholar 

  16. Burgarth, D.: Quantum state transfer with spin chains. PhD thesis, University College London (2006)

  17. Bose, S.: Quantum communication through spin chain dynamics: an introductory overview. Contemp. Phys. 48(Issue 1), 13 – 30 (2007)

  18. Collini, E., Wong, C.Y., Wilk, K.E., Curmi, P.M.G., Brumer, P., Scholes, G.D.: Coherently wired light harvesting in photosynthetic marine algae at ambient temperature. Nature 463(7281), 644–647 (2010)

  19. Engel, G.S., Calhoun, T.R., Read, R.L., Ahn, T.K., Manal, T., Cheng, Y.C., Blankenship, R.E., Fleming, G.R.: Evidence for wavelike energy transfer through quantum coherence in photosynthetic systems. Nature 446(7137), 782–786 (2007)

  20. Di Molfetta, G., Debbasch, F.: Discrete-time quantum walks: continuous limit and symmetries. J. Math. Phys. 53, 123302 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  21. Di Molfetta, G., Debbasch, F., Brachet, M.: Quantum walks as massless dirac fermions in curved space. Phys. Rev. A 88 (2013)

  22. Di Molfetta, G., Debbasch, F., Brachet, M.: Quantum walks in artificial electric and gravitational fields. Phys. A 397 (2014)

  23. Freidberg, J.: Plasma Physics and Fusion Energy. Cambridge University Press, Cambridge (2008)

  24. Kolb, E.W., Turner, M.S.: The Early Universe. Frontiers in Physics. Addison-Wesley, Redwood City (1990)

    Google Scholar 

  25. Balescu, R.: Aspects of anomalous transport in plasmas. CRC Press (2005)

  26. Horton, W.: Turbulent Transport in Magnetized Plasmas. World Scientific, Singapore (2014)

  27. Debbasch, F.: What is a mean gravitational field? Eur. Phys. J. B 37(2), 257–270 (2004)

    Article  Google Scholar 

  28. Wiegand, A.,  Buchert, T.: Multi-scale cosmology and structure-emerging dark energy: a plausibility analysis. arXiv:1002.3912v1 (2010)

  29. Debbasch, F.: Mean field theory and geodesics in general relativity. Eur. Phys. J. B 43(1), 143–154 (2005)

    Article  MathSciNet  Google Scholar 

  30. Kendon, V.: Decoherence in quantum walks—a review. Math. Struct. Comput. Sci. 17(6), 1169–1220 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  31. Vieira, R., Amorim, E.P.M., Rigolin, G.: Entangling power of disordered quantum walks. Phys. Rev. A 89, 042307 (2014)

    Article  Google Scholar 

  32. Brun, T.A., Carteret, H.A., Ambainis, A.: Quantum to classical transition for random walks. Phys. Rev. Lett. 91, 130602 (2003)

    Article  Google Scholar 

  33. Ahlbrecht, A.,  Vogts, H., Werner, A.H., Werner, R.F.: Asymptotic evolution of quantum walks with random coin. J. Math. Phys. 52(4) (2011)

  34. Ahlbrecht, A., Cedzich, C., Matjeschk, R., Scholz, V.B., Werner, A.H., Werner, R.F.: Asymptotic behavior of quantum walks with spatio-temporal coin fluctuations. Quant. Inf. Process. 11(5), 1219–1249 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  35. Joye, A.: Random time-dependent quantum walks. Commun. Math. Phys. 307(1), 65–100 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  36. Lifshitz, E., Pitaevski, L.P.: Physical Kinetics. Pergamon Press, Oxford (1981)

  37. Kollar, B., Koniorczyk, M.: Entropy rate of message sources driven by quantum walks. Phys. Rev. A 89, 022338 (2014)

    Article  Google Scholar 

  38. Liu, C., Petulante, N.: On the von Neumann entropy of certain quantum walks subject to decoherence. Math. Struct. Comput. Sci. 20, 1099–1115 (2010)

  39. Chandrasekhar, C.M., Banerjee, S., Srikanth, R.: Relationship between quantum walks and relativistic quantum mechanics. Phys. Rev. A 81, 062340 (2010)

    Article  Google Scholar 

  40. Abal, G., Siri, R., Romanelli, A., Donangelo, R.: Quantum walk on the line: entanglement and nonlocal initial conditions. Phys. Rev. A 73, 042302 (2006)

  41. Chisaki, K., Konno, N., Segawa, E., Shikano, Y.: Crossovers induced by discrete-time quantum walks. Quant. Inf. Comput. 11(9–10), 741–760 (2011)

    MathSciNet  MATH  Google Scholar 

  42. Navarrete-Benlloch, C., Perez, A., Roldan, E.: Nonlinear optical galton board. Phys. Rev. A 75, 062333 (2010)

    Article  Google Scholar 

  43. Obuse, H., Kawakami, N.: Topological phases and delocalization of quantum walks in random environments. Phys. Rev. B 84(19), 195139 (2011)

    Article  Google Scholar 

  44. Shikano, Y.: From discrete time quantum walk to continuous time quantum walk in limit distribution. J. Comput. Theor. Nanosci. 10(7), 1558–1570 (2013)

    Article  Google Scholar 

  45. Shikano, Y., Wada, T., Horikawa, J.: Discrete-time quantum walk with feed-forward quantum coin. Sci. Rep. 4 (2014)

  46. Prokofev, N.V., Stamp, P.C.E.: Decoherence and quantum walks: anomalous diffusion and ballistic tails. Phys. Rev. A 74(2), 020102 (2006)

  47. Schreiber, A., Cassemiro, K.N., Potoček, V., Gábris, A., Jex, I., Silberhorn, Ch.: Decoherence and disorder in quantum walks: from ballistic spread to localization. Phys. Rev. Lett. 106(18), 180403 (2011)

    Article  Google Scholar 

  48. Inui, N., Konishi, Y., Konno, N.: Localization of two-dimensional quantum walks. Phys. Rev. A 69(5), 052323 (2004)

    Article  Google Scholar 

  49. Leggett, A.J., Chakravarty, S., Dorsey, A.T., Matthew P.A., Fisher, A.G., Zwerger, W.: Dynamics of the dissipative two-state system. Rev. Mod. Phys. 59, 1–85 (1987)

  50. Caldeira, A.O., Leggett, A.J.: Path integral approach to quantum Brownian motion. Phys. A Stat. Mech. Appl. 121(3), 587–616 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  51. Debbasch, F., Mallick, K., Rivet, J.P.: Relativistic Ornstein–Uhlenbeck process. J. Stat. Phys. 88, 945 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  52. Chevalier, C., Debbasch, F.: Relativistic diffusions: a unifying approach. J. Math. Phys. 49, 043303 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  53. Debbasch, F., Espaze, D., Foulonneau, V.: Can diffusions propagate? J. Stat. Phys. 149, 37–49 (2012)

  54. Di Molfetta, G.,  Debbasch, F.: Discrete-time quantum walks: continuous limit in 1 \(+\) 1 and 1 \(+\) 2 dimension. J. Comput. Theor. Nanosci. 10(7), 1621–1625 (2012)

  55. Arnault, P.,  Debbasch, F.: Landau levels for discrete-time quantum walks in artificial magnetic fields. arXiv:1412.4337 (2014)

  56. Wald, R.M.: General Relativity. The University of Chicago Press, Chicago (1984)

    Book  MATH  Google Scholar 

Download references

Acknowledgments

This work has been partially supported by the Spanish Ministerio de Educación e Innovación, MICIN-FEDER project FPA2014-54459-P, SEV-2014-0398 and “Generalitat Valenciana” Grant GVPROMETEOII2014-087.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to G. Di Molfetta.

Appendices

Appendix 1: Interpretation in terms of artificial gauge fields

It has been proven in [2022] that quantum walks in \((1 + 1)\) dimensional space-times can be viewed as modeling the transport of a Dirac fermion in artificial electric and gravitational fields generated by the time-dependance of the angles \(\theta \) and \(\xi \). We recall here some basic conclusions obtained in [2022] and also offer new developments useful in interpreting the results of the present article.

The DTQWs defined by (1) are part of a larger family whose dynamics reads:

$$\begin{aligned} \begin{bmatrix} \psi ^{\mathrm{L}}_{j+1, m }\\ \psi ^{\mathrm{R}}_{j+1, m} \end{bmatrix} \ = {\tilde{\mathcal B}}\left( \theta _{j, m} ,\xi _{j, m}, \zeta _{j, m}, \alpha _{j, m} \right) \begin{bmatrix} \psi ^{\mathrm{L}}_{j, m+1} \\ \psi ^{\mathrm{R}}_{j, m-1} \end{bmatrix}, \end{aligned}$$
(35)

where

$$\begin{aligned} {\tilde{\mathcal B}}(\theta ,\xi , \zeta , \alpha ) = \mathrm{e}^{i \alpha } \begin{bmatrix} \mathrm{e}^{i\xi } \cos \theta&\mathrm{e}^{i \zeta } \sin \theta \\ - \mathrm{e}^{i \zeta } \sin \theta&\mathrm{e}^{-i\xi } \cos \theta \end{bmatrix}. \end{aligned}$$
(36)

The walks in this larger family are characterized by three time- and space-dependent Euler angles \((\theta , \xi , \zeta )\) and by a global, also time- and space-dependent phase \(\alpha \). They have been shown to model the transport of Dirac fermions in artificial electric and relativistic gravitational fields generated by the time-dependence of the three Euler angles and of the global phase. In a \((1 + 1)\) dimensional space-time, an electric field derives from a 2-potential \(A_{j, m} = \left( V_{j, m}, {\mathcal A}_{, mj} \right) \) and a relativistic gravitational field is represented by 2D metrics \(G_{j, m}\). The walks considered in this article correspond to

$$\begin{aligned} \xi= & {} \frac{\pi }{2} + {\bar{\xi }}_j \nonumber \\ \theta= & {} \frac{\pi }{4} + {\bar{\theta }}_j \nonumber \\ \alpha= & {} \frac{\pi }{2} + {\bar{\alpha }} \nonumber \\ \zeta= & {} 0 \end{aligned}$$
(37)

where \({\bar{\xi }}_j\) and \({\bar{\theta }}_j \) are random variables which depend on the time j and \({\bar{\alpha }} = 3 \pi /2\). According to [22], these walks model the transport of a Dirac fermion in an electric field generated by the 2-potential

$$\begin{aligned} A_{j} = \left( V_{j}, {\mathcal A}_{j} \right) = \left( {\bar{\alpha }}_{j}, - {\bar{\xi }}_{j}\right) = \left( \pi /2, - {\bar{\xi }}_{j}\right) \end{aligned}$$
(38)

and in a gravitational field characterized by the metric

$$\begin{aligned} G_{ij} = \text{ diag } (1, - \cos ^{-2} (\theta _j)). \end{aligned}$$
(39)

Since relativistic gravitational fields are represented by space-time metrics [56], making the angle \(\theta \) a time-dependent random variable is equivalent to imposing a time-dependent random gravitational field. To better understand the electric aspects of the problem, let us recall that the DTQWs defined by (35) exhibit the following exact discrete gauge invariance [22]:

$$\begin{aligned}&\Psi '_{j, m} = \Psi _{j, m} \mathrm{e}^{- i \phi _{j, m}} \nonumber \\&\xi '_{j, m} = \xi _{j, m} + \delta _{j, m} \nonumber \\&\theta '_{j, m} = \theta _{j, m} \nonumber \\&{\alpha }'_{j, m} = \alpha _{j, m} + \frac{\sigma _{j, m}}{2} \nonumber \\&\zeta '_{j, m} = = \zeta _{j, m} - \delta _{j, m} \end{aligned}$$
(40)

where

$$\begin{aligned} \sigma _{j, m}= & {} \phi _{j, m+1} + \phi _{j, m-1} - 2 \phi _{j +1, m} \nonumber \\ \delta _{j, m}= & {} \frac{\phi _{j, m+1} - \phi _{j, m-1} }{2} \end{aligned}$$
(41)

and \(\phi \) is the arbitrary time- and space-dependent phase shift. Let us now define a new quantity \(E_{j, m}\) by

$$\begin{aligned} E_{j, m} = - \left( {\mathcal D}_s V\right) _{j, m} + \left( {\mathcal D}_t {\mathcal A}\right) _{j, m} \end{aligned}$$
(42)

where the actions of the operators \({\mathcal D}_s\) and \({\mathcal D}_s\) on an arbitrary time- and space-dependent quantity \(u_{j, m}\) are

$$\begin{aligned} \left( {\mathcal D}_s u\right) _{j, m} = \frac{u_{j, m+1} - u_{j, m-1}}{2} \end{aligned}$$
(43)

and

$$\begin{aligned} \left( {\mathcal D}_t u\right) _{j, m} = \frac{2 u_{j+1, m} - u_{j, m+1} - u_{j, m-1}}{2}. \end{aligned}$$
(44)

The operators \({\mathcal D}_s\) and \({\mathcal D}_t\) are discrete counterparts of space- and time-derivatives. It is straightforward to check that the quantity \(E_{j, m}\) is a gauge invariant, and coincides in the continuous limit with the standard electric field E(tx), defined by \(E(t, x) = - \partial _x V + \partial _t {\mathcal A}\). The quantity \(E_{j. m }\) is thus a bona fide electric field in discrete space-time. For the DTQWs considered in this article, this electric field depends only on the time j and is related to the angle \({\bar{\xi }}\) by \(E_{j} = - \left( {\bar{\xi }}_{j+1} - {\bar{\xi }}_j\right) \). Making this angle a time-dependent random variable is thus equivalent to imposing a random electric field.

Appendix 2: Asymptotic computation of the eigenvalues and eigenvectors of the averaged transport operators

Let us here compute the eigenvalues \(\lambda ^{\mathrm{e/g}}_r\) and eigenvectors \(w^{\mathrm{e/g}}_r\), \(r = 1, 2, 3, 4\) only for values of K much smaller than unity. We do not perform an expansion in p because the initial condition is uniform in p and the average evolution does not localize the density operator around \(p = 0\). Indeed, the initial condition is localized at \(x' = x\) , i.e., does not exhibit any spatial correlation and the dynamics does not create spatial correlations.

The second order expansions of the operators \({\bar{\mathcal R}}^\mathrm{e}\) and \({\bar{\mathcal R}}^\mathrm{g}\) in K read:

$$\begin{aligned} \displaystyle { {\bar{\mathcal R}}^\mathrm{e}_2(K, p, \sigma ) = \begin{bmatrix} 1 - 2K^2&\quad 2iK&\quad 0&\quad 0 \\ 0&\quad 0&\quad \text {sinc}(\sigma /2) \cos (2p)&\quad - i\, \text {sinc}(\sigma /2) \sin (2p) \\ 2i \text {sinc}(\sigma /2) K&\quad \text {sinc}(\sigma /2) \left( 1 - 2 K^2 \right)&\quad \frac{ (1 - \text {sinc}(\sigma ))}{2} \cos (2p)&\quad i \frac{( \text {sinc}(\sigma ) - 1)}{2} \sin (2p)\\ 0&\quad 0&\quad i \frac{(1 + \text {sinc}(\sigma )) }{2}\sin (2p)&\quad - \frac{( \text {sinc}(\sigma ) + 1)}{2} \cos (2p) \end{bmatrix},} \end{aligned}$$
(45)

and

$$\begin{aligned} {\bar{\mathcal R}}^\mathrm{g}_2(K, p, \sigma ) = \begin{bmatrix} 1 - 2K^2&\quad 2i K&\quad 0&\quad 0 \\ 0&\quad 0&\quad \text {sinc}(\sigma ) \cos (2p)&\quad - i\, \text {sinc}(\sigma ) \sin (2p) \\ 2 i \text {sinc}(\sigma ) K&\quad \text {sinc}(\sigma )&\quad 0&\quad 0\\ 0&\quad 0&\quad i \sin (2p)&\quad - \cos (2p) \end{bmatrix}. \end{aligned}$$
(46)

For \(K = 0\), these two matrices are both block diagonal and we write \({\bar{\mathcal R}}^{\mathrm{e/g}}_2(K = 0, p, \sigma ) = \text{ diag } (1, M^{\mathrm{e/g}}(p, \sigma ))\), where \(M^{\mathrm{e/g}}(p, \sigma )\) are \(3 \times 3\) matrices acting in the space spanned by \((u_2, u_3, u_4)\). The matrices \({\bar{\mathcal R}}^{\mathrm{e/g}}_2(K = 0, p, \sigma )\) share \(u_1\) as common eigenvector, which we identify as \(w^{\mathrm{e/g}}_1(K = 0, p, \sigma )\); the associated eigenvalue is \(\lambda _1^{\mathrm{e/g}}(K = 0, p, \sigma ) = 1\). The other eigenvectors and eigenvalues, at zeroth order in K, are those of \(M^{\mathrm{e/g}}(p, \sigma )\). These eigenvalues can be computed analytically by solving the third-order characteristic polynomials associated to these matrices. The explicit expressions of these eignevalues are quite involved and need not be replicated here. What is important is how the moduli of these eigenvalues compares to unity. Direct inspection reveals that the moduli of all three \(\lambda ^\mathrm{e}_r(0, p, \sigma )\), \(r = 2, 3, 4\) are strictly inferior to unity if \(\sigma \) is not vanishing. The same goes for all three eigenvalues in the gravitational case, except for one of them which reaches \(\pm 1\) independently of \(\sigma \) for \(p = \pm \pi \) and is also equal to \(+1\) for \(p = 0\); the eigenspaces corresponding to \(\lambda _4^\mathrm{g} (\pm \pi , \sigma )\) and \(\lambda _4^\mathrm{g} (0, \sigma )\) are identical and generated by \(u_4\), which we choose as \(w_4^\mathrm{g}(p = \pm \pi , \sigma ) = w_4^\mathrm{g}(0, \sigma )\). For other values of p, the eigenvalue \(\lambda _4^\mathrm{g}(p, \sigma )\) and the eigenvector \(w_4^\mathrm{g}(p, \sigma )\) are defined by continuity. All other eigenvectors need not be specified for what follows.

Let us now turn to nonvanishing values of K. The characteristic polynomials of \({\bar{\mathcal R}}^{\mathrm{e/g}}_2(K, p, \sigma )\) contain terms of order 2 and 4 in K; at lowest order in K, the corrections to the eigenvalues \(\lambda _j^{\mathrm{e/g}}(K = 0, p, \sigma )\) thus scale generically as \(K^2\). Let \(\lambda \) be the variable of the characteristic polynomials. At second order in K, the K-dependent correction to each of the zeroth order eigenvalues \(\lambda _r^{\mathrm{e/g}}(K = 0, p, \sigma )\) can be found by expanding the characteristic polynomial of \({\bar{\mathcal R}}^{\mathrm{e/g}}_2(K, p, \sigma )\) at first order in \((\lambda - \lambda _r^{\mathrm{e/g}}(K = 0, p, \sigma ))\) and by keeping only the terms scaling as \(K^2\). This gives rational expressions for the corrections to the eigenvalues; these rational expressions can be further simplified by a final expansion around \(K = 0\) if p is treated as a finite, non-infinitesimal quantity , i.e., \(\mid K \mid \ll \mid p \mid \). One then finds:

$$\begin{aligned} \lambda _1^{\mathrm{e/g}}(K, p, \sigma ) = 1 - \alpha ^{\mathrm{e/g}} (p, \sigma ) K^2 + O(K^4) \end{aligned}$$
(47)

with

$$\begin{aligned} \alpha ^{\mathrm{e}}(p, \sigma ) = 2\, \frac{3 + \left( \text {sinc}(\sigma ) \right) ^2 + 2 \left( \text {sinc}(\sigma /2) \right) ^2 \left( 1 + \text {sinc}(\sigma ) \right) + 4 \cos (2p) \left( \text {sinc}(\sigma ) + \left( \text {sinc}(\sigma /2) \right) ^2 \right) }{3 + \left( \text {sinc}(\sigma ) \right) ^2 - 2 \left( \text {sinc}(\sigma /2) \right) ^2 \left( 1 + \text {sinc}(\sigma ) \right) + 4 \cos (2p) \left( \text {sinc}(\sigma ) - \left( \text {sinc}(\sigma /2) \right) ^2 \right) } \end{aligned}$$
(48)

and

$$\begin{aligned} \alpha ^{\mathrm{g}}(p, \sigma ) = 2 \, \frac{1 + \left( \text {sinc}(\sigma )\right) ^2}{1 - \left( \text {sinc}(\sigma )\right) ^2}. \end{aligned}$$
(49)

Note that \(\alpha ^\mathrm{g}\) is actually independent of p. Note also that the condition \(\mid K \mid \ll \mid p \mid \) does not hinder asymptotic computations, at least on the infinite line. Indeed, as time increases, the density operator becomes more and more localized around \(K = 0\), but it does not localize in p-space.Footnote 2 If one works on the infinite line, both K and p are continuous variables and the localization of the density operator around \(K = 0\) implies that the size of the region in p-space where the condition \(\mid K \mid \ll \mid p \mid \) does not apply actually shrinks to zero with time. For dynamics taking place on a finite circle (finite value of M), computations are a little more involved but can nevertheless be carried out. We feel a detailed analysis of the problem for finite values of M does not bring any valuable insight on interesting physics or mathematics, and we thus restrict the analytical discussion of the asymptotic dynamics to DTQWs on the infinite line, where expressions (48) and (49) suffice.

A direct computation shows that the corrections to the eigenvectors are first order in K. By convention, we fix to unity the value of the first component of \(w_1^{\mathrm{e/g}}(K, p, \sigma )\) in the basis \((u_1, u_2, u_3, u_4)\). One thus gets, for example,

$$\begin{aligned} w_1^{\mathrm{g}}(K, p, \sigma ) = \ {\begin{matrix} 1,&\frac{2 i K \text {sinc}(\sigma )^2}{1-\text {sinc}(\sigma )^2},&\frac{2 i K \text {sinc}(\sigma )}{1-\text {sinc}(\sigma )^2},&\frac{- 2 K \text {sinc}(\sigma ) \tan (p)}{1-\text {sinc}(\sigma )^2} \end{matrix}}. \end{aligned}$$
(50)

The expression of \(w_1^\mathrm{e}\) is substantially more complicated and need not be reproduced here.

Appendix 3: Asymptotic expression of the density operator in Fourier space

Let us now use the above results to determine the time evolution of the average density operator in both cases under consideration. The first step is to express the initial condition, \({\hat{\rho }}_{j = 0} (K, p) = (u_1 - u_4)/2\) for all (Kp), as a linear combination of the eigenvectors \(w^{\mathrm{e/g}}_r(K, p, \sigma )\). We thus write, for \(a = 1, 2, 3, 4\)

$$\begin{aligned} u_ a= \sum _{r = 1}^4 u^{\mathrm{e/g}}_{ar} (K, p, \sigma ) w^{\mathrm{e/g}}_r(K, p, \sigma ) \end{aligned}$$
(51)

and, conversely,

$$\begin{aligned} w^{\mathrm{e/g}}_r(K, p, \sigma )= \sum _{a = 1}^4 w^{\mathrm{e/g}}_{ra} (K, p, \sigma ) u_a. \end{aligned}$$
(52)

By the above discussion of the eigenvalues and eigenvectors of \({\bar{\mathcal R}}^{\mathrm{e/g}}_2\), one has notably \(u^{\mathrm{e/g}}_{11} (K, p, \sigma ) = 1\), \(u^{\mathrm{e/g}}_{1r} (K, p, \sigma ) = O(K)\) for \(r = 2, 3, 4\), \(w^{\mathrm{e/g}}_{11} (K, p, \sigma ) = 1 + O(K)\).

One then writes, for all K and p:

$$\begin{aligned} {\hat{\rho }}_{j = 0} (K, p) = \frac{1}{2} \sum _{r = 1}^4 \left( u^{\mathrm{e/g}}_{1r} (K, p, \sigma ) - u^{\mathrm{e/g}}_{4r} (K, p, \sigma ) \right) w^{\mathrm{e/g}}_r(K, p, \sigma ). \end{aligned}$$
(53)

which leads to

$$\begin{aligned} {\hat{\rho }}_{j = J} (K, p) = \frac{1}{2} \sum _{r = 1}^4 \left( \lambda _r^{\mathrm{e/g}} (K, p, \sigma ) \right) ^J \left( u^{\mathrm{e/g}}_{1r} (K, p, \sigma ) - u^{\mathrm{e/g}}_{4r} (K, p, \sigma ) \right) w^{\mathrm{e/g}}_r(K, p, \sigma ) \end{aligned}$$
(54)

or, expressing the eigenvectors \(w^{\mathrm{e/g}}_r(K, p, \sigma )\) in terms of the original basis vectors \((u_1, u_2, u_3, u_4)\):

$$\begin{aligned} {\hat{\rho }}_{j = J} (K, p) = \frac{1}{2} \sum _{r = 1}^4 \sum _{a = 1}^4 \left( \lambda _r^{\mathrm{e/g}} (K, p, \sigma ) \right) ^J \left( u^{\mathrm{e/g}}_{1r} (K, p, \sigma ) - u^{\mathrm{e/g}}_{4r} (K, p, \sigma ) \right) w^{\mathrm{e/g}}_{ra} (K, p, \sigma ) u_a \end{aligned}$$
(55)

Now, for all r,

$$\begin{aligned} \lambda _r^{\mathrm{e/g}} (K, p, \sigma )/ \lambda _1^{\mathrm{e/g}} (K, p, \sigma ) = \lambda _r^{\mathrm{e/g}}(K =0, p, \sigma ) ( 1 + O(K^2)), \end{aligned}$$
(56)

since \(\lambda _r^{\mathrm{e/g}}(K =0, p, \sigma ) = 1\). It follows that for small enough K, the contributions to (55) proportional to \(( \lambda _r^{\mathrm{e/g}} (K, p, \sigma ))^J\) are much smaller than the contribution proportional to \((\lambda _1^{\mathrm{e/g}} (K, p, \sigma ))^J\) for all values of p and \(\sigma \) such that \(\mid \lambda _r^{\mathrm{e/g}}(K =0, p, \sigma )\mid < 1\). According to the above discussion, this is realized for all \(r \ne 1\) and for all values of p and \(\sigma \), except in case 2 (random gravitational field) for \(r = 4\), \(p = \pm \pi \) or \(p = 0\) and all values of \(\sigma \). What happens at \(p = \pm \pi \) has no incidence on the computation of the density operator in physical space. Indeed, for finite values of M, the maximum value \(p_{\text{ max }}\) of \(\mid p \mid \) is \(p_{\text{ max }} = (2M/(2M + 1)) \pi < \pi \). Thus \(\pm \pi \) is only reached in the limiting case of infinite M , i.e., for quantum walks in the infinite line. However, \(\pm p_{\text{ max }} = \pm \pi \) then only appear as upper and lower bounds for integrals over p, and the values taken by \({\hat{\rho }} (J, K, p)\) at points \(\pm \pi \) does not modify the values of the integrals. Moreover, all current computations are only valid for \(\mid p \mid \ll \mid K \mid \), and are thus a priori invalid for \(p = 0\). What happens around \(p = 0\) has however no relevance to asymptotic computations on the infinite line because, as time increases, the density operator becomes more and more localized around \(K = 0\) (see discussion below (23)). For large enough J and small enough K, the double sum in (55) thus simplifies into:

$$\begin{aligned} {\hat{\rho }}^{\mathrm{e/g}}_{j = J}(K, p) = \frac{1}{2} \sum _{a = 1}^4 \left( \lambda _1^{\mathrm{e/g}} (K, p, \sigma ) \right) ^J \left( u^{\mathrm{e/g}}_{11} (K, p, \sigma ) - u^{\mathrm{g/e}}_{41} (K, p, \sigma ) \right) w^{\mathrm{e/g}}_{1a} (K, p, \sigma ) u_a \end{aligned}$$
(57)

Now, \(u_{11}^{\mathrm{e/g}} (K, p, \sigma ) = 1 + O(K)\), \(u_{41}^{\mathrm{e/g}} (K, p, \sigma ) = O(K)\), \(w_{11}^{\mathrm{e/g}} (K, p, \sigma ) = 1 + O(K)\) and \(w_{1b}^{\mathrm{e/g}} (K, p, \sigma ) = O(K)\) for \(b = 2, 3, 4\). As far as orders of magnitude are concerned, Eq. (57) gives:

$$\begin{aligned} {\hat{\rho }}^{\mathrm{e/g}}_{j = J} (K, p) = \frac{1}{2} \left( 1 - \alpha ^{\mathrm{e/g}} (p, \sigma ) K^2 \right) ^N \left( 1 + O(K)\right) u_1 + \sum _{b = 2}^4 O(K) u_b. \end{aligned}$$
(58)

At lowest order in K, \(( 1 - \alpha ^{\mathrm{e/g}} (K, p, \sigma ) K^2 )^J = 1 - \alpha ^{\mathrm{e/g}} (K, p, \sigma ) J K^2\). We will now restrict the discussion to scales K and times J obeying \(JK^2 \gg K\) i.e., \(J K \gg 1\). Note that the maximum spatial spread of \({\bar{\rho }}\) at time J is \(L_{\text {max}}(J) = 2J\), so that the minimum value of K for which \({\hat{\rho }}\) takes non-negligible values at time J is of order \(K_{\text {min}}(N) = 1/J\). The condition \(J K \gg 1\) thus restricts the discussion to length scales much smaller than \(L_{\text {max}}(N)\). In particular, consider the time-dependent scale \(K_J= K_*/\sqrt{J}\), where \(K_*\) is an arbitrary time-independent wave-vector. The wave-vector \(K_J\) obeys \(JK_J^2 = K_*^2 \gg K_J\) for sufficiently large J. Thus, the possible diffusive behavior of the averaged transport is encompassed by the present discussion.

With the above assumption, Eq. (57) implies the following approximate, but very simple expression for the long time (large J) density operator in Fourier space:

$$\begin{aligned} {\hat{\rho }}^{\mathrm{e/g}}_{j = J} (K, p) = \frac{1}{2} \left( 1 - \alpha ^{\mathrm{e/g}} (p, \sigma ) K^2 \right) ^J u_1. \end{aligned}$$
(59)

In particular, for \(K_J= K_*/\sqrt{J}\) (where \(K_*\) is an arbitrary but J-independent wave number) and large enough J,

$$\begin{aligned} {\hat{\rho }}^{\mathrm{e/g}}_{j = J} (K_J, p) = \frac{1}{2} \left( 1 - \alpha ^{\mathrm{e/g}} (p, \sigma ) \frac{K_*^2}{J} \right) ^J u_1 \sim \frac{1}{2} \exp \left( - \alpha ^{\mathrm{e/g}} (p, \sigma ) K_*^2\right) u_1 = \frac{1}{2} \exp \left( - \alpha ^{\mathrm{e/g}} (p, \sigma ) J K_J^2\right) u_1. \end{aligned}$$
(60)

This is an approximate expression for the asymptotic density operator presented in the main body of this article.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Di Molfetta, G., Debbasch, F. Discrete-time quantum walks in random artificial gauge fields. Quantum Stud.: Math. Found. 3, 293–311 (2016). https://doi.org/10.1007/s40509-016-0078-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s40509-016-0078-6

Keywords

Navigation