Skip to main content
Log in

A Crossover Between Open Quantum Random Walks to Quantum Walks

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

Abstract

We propose an intermediate walk continuously connecting an open quantum random walk and a quantum walk with parameters \(M\in {{\mathbb {N}}}\) controlling a decoherence effect; if \(M=1\), the walk coincides with an open quantum random walk, while \(M=\infty \), the walk coincides with a quantum walk. We define a measure which recovers usual probability measures on \({{\mathbb {Z}}}\) for \(M=\infty \) and \(M=1\) and we observe intermediate behavior through numerical simulations for varied positive values M. In the case for \(M=2\), we analytically show that a typical behavior of quantum walks appears even in a small gap of the parameter from the open quantum random walk. More precisely, we observe both the ballistically moving towards left and right sides and localization of this walker simultaneously. The analysis is based on Kato’s perturbation theory for linear operator. We further analyze this limit theorem in more detail and show that the above three modes are described by Gaussian distributions.

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
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

References

  1. Ambainis, A.: Quantum walks and their algorithmic applications. Int. J. Quant. Inf. 1(2003), 507–518 (2003)

    Article  Google Scholar 

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

    Article  ADS  MathSciNet  Google Scholar 

  3. Attal, S., Petruccione, F., Sabot, C., Sinayskiy, I.: Open Quantum Random Walks. J. Stat. Phys. 147, 832–852 (2012)

    Article  ADS  MathSciNet  Google Scholar 

  4. Brun, T.A., Brun, Carteret, H.A., Ambainis, A.: Quantum random walks with decoherent coins. Phys. Rev. A 67, 032304 (2003)

    Article  ADS  MathSciNet  Google Scholar 

  5. Dhahri, A., Mukhamedov, F.: Open quantum random walks and quantum Markov chains. Funct. Anal. Appl. 53, 137–142 (2019)

    Article  MathSciNet  Google Scholar 

  6. Higuchi, Yu., Segawa, E.: Dynamical system induced by quantum walks. J. Phys. A 52, 39 (2019)

    Article  MathSciNet  Google Scholar 

  7. Kato, T.: A Short Introduction to Perturbation Theory for Linear Operators. Springer, New York (1982)

    Book  Google Scholar 

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

    Article  MathSciNet  Google Scholar 

  9. Konno, N., Komatsu, T., Sato, I.: Walk/Zeta Correspondence. J. Stat. Phys. 190, 36 (2023)

    Article  ADS  MathSciNet  Google Scholar 

  10. Konno, N.: Limit theorems and absorption problems for one-dimensional correlated random walks. Stoch. Models 25, 28–49 (2009)

    Article  MathSciNet  Google Scholar 

  11. Konno, N.: Quantum walks. In: Lecture Notes in Mathematics, Vol. 1954, pp. 309–452 Springer, Heidelberg (2008)

  12. Ko, C.K., Konno, N., Segawa, E., Yoo, H.: Limit theorems for open quantum random walks. J. Stat. Phys. 176, 710–735 (2019)

    Article  ADS  MathSciNet  Google Scholar 

  13. Portugal, R.: Quantum Walk and Search Algorithms, 2nd edn. Springer, Cham (2018)

    Book  Google Scholar 

  14. Segawa, E., Konno, N.: Limit theorems for quantum walks driven by many coins. Int. J. Quant. Inf. 6, 1231–1243 (2008)

    Article  Google Scholar 

  15. Sunada, T., Tate, T.: Asymptotic behavior of quantum walks on the line. J., Funct. Anal. 262, 2608–2645 (2012)

    Article  MathSciNet  Google Scholar 

  16. Suzuki, A.: Asymptotic velocity of a position-dependent quantum walk. Quant. Inf. Process. 15, 103–119 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  17. Whitfield, J.D., Rodríguez-Rosario, C.A., Aspuru-Guzik, A.: Quantum stochastic walks: a generalization of classical random walks and quantum walks. Phys. Rev. A 81, 022323 (2010)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

The authors thank to Jir̆í Marys̆ka, Stanisla Skoupý for fruitful discussion. The authors also thank to the referee for carefully reading and giving constructive suggestions to this paper. K.M. was supported by World Premier International Research Center Initiative (WPI), Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan and the grant-in-aid for young scientists No. JP17K14235, Japan Society for the Promotion of Science, Grant-in-Aid of Scientific Research (B) No. JP21H01001. E.S. acknowledges financial supports from the Grant-in-Aid of Scientific Research (C) No. JP19K03616, Japan Society for the Promotion of Science and Research Origin for Dressed Photon.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Etsuo Segawa.

Additional information

Communicated by Hal Tasaki.

Publisher's Note

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

Propositions A.1–A.3

Propositions A.1A.3

The matrix representation for \({\hat{W}}_{s,t}(0)=:{\hat{W}}_{s,t}\) is explicitly written as

$$\begin{aligned} {\hat{W}}_{s,t}= \frac{1}{2} \left[ \begin{array}{cccc|cccc} 1 &{} 0 &{} 0 &{} 1 &{} 0 &{} 1 &{} 0 &{} 0 \\ 1 &{} 0 &{} 0 &{} -1 &{} 0 &{} -1 &{} 0 &{} 0 \\ 1 &{} 0 &{} 0 &{} -1 &{} 0 &{} 1 &{} 0 &{} 0 \\ 1 &{} 0 &{} 0 &{} 1 &{} 0 &{} -1&{} 0 &{} 0 \\ \hline 0 &{} 0 &{} 1 &{} 0 &{} 1 &{} 0 &{} 0 &{} 1 \\ 0 &{} 0 &{} 1 &{} 0 &{} 1 &{} 0 &{} 0 &{} -1 \\ 0 &{} 0 &{} -1 &{} 0 &{} 1 &{} 0 &{} 0 &{} -1 \\ 0 &{} 0 &{} -1 &{} 0 &{} 1 &{} 0 &{} 0 &{} 1 \end{array} \right] . \end{aligned}$$

for \(s=-1\) and \(t=0\). Using the formula for \(\lambda _1(k)\) and \(\lambda _2(k)\), we know that eigenvalues of \({\hat{W}}_{s,t}\) are

$$\begin{aligned} \textrm{spec}({\hat{W}}_{s,t})=\left\{ 0,0,1,1,1,-\frac{1}{2}, \frac{-1\pm i\sqrt{7}}{4}\right\} , \end{aligned}$$
(51)

which shows that the matrix \({\hat{W}}_{s,t}\) possesses eigenvalues with multiplicity greater than 1. By the Cayley-Hamilton theorem, we know that \(F_{{\hat{W}}_{s,t}}({\hat{W}}_{s,t})= O\), where

$$\begin{aligned} F_{{\hat{W}}_{s,t}}(\lambda ) = \lambda ^2 (\lambda - 1)^3 (2\lambda ^2 + \lambda + 1)(2\lambda +1). \end{aligned}$$
(52)

Interestingly, we observe the following result.

Lemma A.1

The minimal polynomial \(f_{{\hat{W}}_{s,t}}(\lambda )\) of the matrix \({\hat{W}}_{s,t}\) is

$$\begin{aligned} f_{{\hat{W}}_{s,t}}(\lambda ) = \lambda (\lambda - 1) (2\lambda ^2 + \lambda + 1)(2\lambda +1). \end{aligned}$$

Proof

Direct calculations yield that

$$\begin{aligned} {\hat{W}}_{s,t} ({\hat{W}}_{s,t} - I) (2{\hat{W}}_{s,t}^2 + {\hat{W}}_{s,t} + I)(2{\hat{W}}_{s,t}+I) = O. \end{aligned}$$

\(\square \)

A direct consequence of this lemma is the following.

Proposition A.1

All eigenvalues of \({\hat{W}}_{s,t}\) are semisimple. In particular, the matrix \({\hat{W}}_{s,t}\) is diagonalizable.

Thanks to Proposition A.1, the eigenvalue \(\lambda = 1\) of \({\hat{W}}_{s,t}\) is semisimple and hence the associating eigenspace, which is a three-dimensional linear space, is generated by three linearly independent eigenvectors of \(\lambda = 1\). We have the following observation for eigenstructures at \(\lambda =1\) by a direct computation of the non-perturbed matrix \({\hat{W}}_{s,t}\).

Lemma A.2

The eigenspace of \({\hat{W}}_{s,t}\) associated with \(\lambda = 1\) is generated by the following vectors:

$$\begin{aligned}{}[0, 0, 0, 0, 1, 0, 0, 1]^\top ,\quad [1, 0, 0, 1, 0, 0, 0, 0]^\top , \quad [1, 0, 1, 0, 1, 1, 0, 0]^\top . \end{aligned}$$

One of their orthonormal choice is

$$\begin{aligned} \nonumber \phi _1&= \left[ 0, 0, 0, 0, \frac{1}{\sqrt{2}}, 0, 0, \frac{1}{\sqrt{2}}\right] ^\top ,\;\;\phi _2 = \left[ \frac{1}{\sqrt{2}}, 0, 0, \frac{1}{\sqrt{2}}, 0, 0, 0, 0\right] ^\top \\ \phi _3&= \left[ \frac{1}{2\sqrt{3}}, 0, \frac{1}{\sqrt{3}}, -\frac{1}{2\sqrt{3}}, \frac{1}{2\sqrt{3}}, \frac{1}{\sqrt{3}}, 0, -\frac{1}{2\sqrt{3}}\right] ^\top . \end{aligned}$$
(53)

Next consider the perturbation of these eigenstructures. Namely, consider eigenpairs of \({\hat{W}}_{s,t}(\delta )\) with sufficiently small \(\delta \). Thus to apply the reduction process [7], the next task is to obtain an explicit expression for the matrix \({\tilde{T}}^{(1)}\) because its eigenvalue \(\lambda _j^{(1)}\) gives the coefficient of the second order of the eigenvalue of \({\hat{W}}_{s,t}(\delta )\). To this end, we need to prepare an explicit expression for the eigenprojection \(\Pi \) of the eigenvalue 1 of \({\hat{W}}_{s,t}(0)\). Our strategy here is to compute eigenvalue \(\lambda _j^{(1)}\) of \({\tilde{T}}^{(1)}\). Since \(\Pi \) is an eigenprojection, it holds that \(\Pi ^2=\Pi \) and \(\Pi {\hat{W}}_{s,t}={\hat{W}}_{s,t}\Pi \). In addition, it can be shown that the property \(\Pi =\Pi ^*\) holds by [6] which treats more general setting of the truncated unitary matrix.

Proposition A.2

Let the centered generalized eigenspace of \({\hat{W}}_{s,t}\) be defined by

$$\begin{aligned} {{\mathcal {H}}}_c:=\{\psi \in \ell ^2(\{s,\dots ,t\};{{\mathbb {C}}}^4) \;|\; ({\hat{W}}_{s,t}-\lambda )^m\psi =0 \mathrm {\;for\; some\;} |\lambda |=1,\;m\ge 1\} \end{aligned}$$

Then we have the following properties of \({{\mathcal {H}}}_c\).

  1. 1.

    The centered generalized eigenspace of \({\hat{W}}_{s,t}:={\hat{W}}_{s,t}(0)\) is expressed by

    $$\begin{aligned} {{\mathcal {H}}}_c&= \bigoplus _{|\lambda |=1}\ker (\lambda -{\hat{W}}_{s,t}) \\&= \textrm{span}\{ \chi _{s,t}^*\psi \;|\; \textrm{supp}(\psi )\subset \{s,\dots ,t\},\;\psi \mathrm {\;is\;an\;eigenvector\;of\;} {\hat{W}}_{-\infty ,\infty }(0)={\hat{U}}(0) \} \end{aligned}$$
  2. 2.

    The complementary invariant subspace of \({{\mathcal {H}}}_c\) can be expressed by \({{\mathcal {H}}}_c^\perp \).

Proof

This is a direct consequence of Lemma 3.3, (3.12) and Lemma 3.4 in [6]. \(\square \)

By the first part of Proposition A.2 implies that \({{\mathcal {H}}}_c=\ker (1-{\hat{W}}_{s,t})\). Moreover the second part of Proposition A.2 implies that the eigenprojection \(\Pi \) of the eigenvalue 1 can be computed by using Lemma A.2 as follows.

$$\begin{aligned} \Pi&= \phi _1\phi _1^*+\phi _2\phi _2^*+\phi _3\phi _3^* \\&= \frac{1}{12}\begin{bmatrix} 7 &{}\quad 0 &{}\quad 2 &{}\quad 5 &{}\quad 1 &{}\quad 2 &{}\quad 0 &{}\quad -1\\ 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 \\ 2 &{}\quad 0 &{}\quad 4 &{}\quad -2 &{}\quad 2 &{}\quad 4 &{}\quad 0 &{}\quad -2\\ 5 &{}\quad 0 &{}\quad -2 &{}\quad 7 &{}\quad -1 &{}\quad -2 &{}\quad 0 &{}\quad 1\\ 1 &{}\quad 0 &{}\quad 2 &{}\quad -1 &{}\quad 7 &{}\quad 2 &{}\quad 0 &{}\quad 5\\ 2 &{}\quad 0 &{}\quad 4 &{}\quad -2 &{}\quad 2 &{}\quad 4 &{}\quad 0 &{}\quad -2\\ 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 \\ -1 &{}\quad 0 &{}\quad -2 &{}\quad 1 &{}\quad 5 &{}\quad -2 &{}\quad 0 &{}\quad 7\\ \end{bmatrix}, \end{aligned}$$

which is an orthogonal projection, in which case \(\Pi ^2 = \Pi \) and \(\Pi ^*= \Pi \) hold.

On the other hand,

$$\begin{aligned} T^{(1)} = \begin{bmatrix} t^{(1)} &{} O \\ O &{} t^{(1)} \end{bmatrix},\quad t^{(1)}= -\frac{i}{2} \begin{bmatrix} 1 &{}\quad 0 &{}\quad 0 &{}\quad -1 \\ 1 &{}\quad 0 &{}\quad 0 &{}\quad 1 \\ 1 &{}\quad 0 &{}\quad 0 &{}\quad 1 \\ 1 &{}\quad 0 &{}\quad 0 &{}\quad -1 \\ \end{bmatrix} \end{aligned}$$

and hence

$$\begin{aligned} \Pi T^{(1)}\Pi&= \frac{-i}{6} \begin{bmatrix} 1 &{}\quad 0 &{}\quad 1 &{}\quad 0 &{}\quad 1 &{}\quad 1 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 \\ 1 &{}\quad 0 &{}\quad 0 &{}\quad 1 &{}\quad 1 &{}\quad 0 &{}\quad 0 &{}\quad 1 \\ 0 &{}\quad 0 &{}\quad 1 &{}\quad -1 &{}\quad 0 &{}\quad 1 &{}\quad 0 &{}\quad -1 \\ 1 &{}\quad 0 &{}\quad 1 &{}\quad 0 &{}\quad 1 &{}\quad 1 &{}\quad 0 &{}\quad 0 \\ 1 &{}\quad 0 &{}\quad 0 &{}\quad 1 &{}\quad 1 &{}\quad 0 &{}\quad 0 &{}\quad 1 \\ 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 1 &{}\quad -1 &{}\quad 0 &{}\quad 1 &{}\quad 0 &{}\quad -1 \\ \end{bmatrix} \end{aligned}$$

Direct calculations yield the following result.

Proposition A.3

The matrix \(\Pi T^{(1)}\Pi \) is a skew-Hermitian matrix. Eigenvalues of the matrix \(\Pi T^{(1)}\Pi \) on the eigenspace \(\textrm{R}(\Pi )\) are 0, \(i / \sqrt{3}\) and \(-i/\sqrt{3}\). In particular, all these eigenvalues are simple. The corresponding eigenprojections are \(\Pi ^{(1)}_1:=v_1v_1^{^*}\), \(\Pi ^{(1)}_2:=v_2v_2^*\) and \(\Pi ^{(1)}_3:=v_3v_3^*\), respectively. Here

$$\begin{aligned} v_1&=[1/2, 0,0,1/2,-1/2,0,0,-1/2]^\top ,\\ v_2&=\frac{1}{\sqrt{2}\;(3-\sqrt{3})}[2-\sqrt{3},0,1-\sqrt{3},1,2-\sqrt{3},1-\sqrt{3},0,1]^\top ,\\ v_3&=\frac{1}{\sqrt{2}\;(3+\sqrt{3})}[2+\sqrt{3},0,1+\sqrt{3},1,2+\sqrt{3},1+\sqrt{3},0,1]^\top . \end{aligned}$$

Remark A.1

  1. 1.

    \(\Pi =\Pi _1^{(1)}+\Pi _2^{(1)}+\Pi _3^{(1)}\) and \(\Pi _i^{(1)}\Pi _j^{(1)}=\delta _{ij}\Pi _i^{(1)}\) holds.

  2. 2.

    The eigenvalue on \(\textrm{R}(1-\Pi )\) is also 0. The eigenvector of eigenvalue \(\epsilon \) for the matrix \(\Pi (T^{(1)}-\epsilon I)\Pi \) corresponds to that of \(\textrm{R}(\Pi )\). Taking \(\epsilon \rightarrow 0\) to this eigenvector, we obtain the eigenvector \(v_1\). See [7] for more detail.

  3. 3.

    The coefficients of the second orders of the eigenvalues of \(\lambda _1(\delta )\) for \(j=0\) in (27) and \(\lambda _2(\delta )\) for \(j=1,2\) in (30) identify with the eingenvalues of \(\Pi T^{(1)}\Pi \) in \(R(\Pi )\). This derives from nothing but the contribution of the reduction process of the perturbation theory [7]. Moreover, it is possible to obtain the coefficients of more than the third orders by repeating the reduction process. For example, the coefficient of \(\lambda _2(\delta )\) for \(j=1\) can be computed by the same procedure by replacing \({\tilde{T}}^{(1)}\) with \({\tilde{T}}^{(2)}\), where

    $$\begin{aligned} {\tilde{T}}^{(2)}:=\Pi T^{(2)}\Pi -\Pi T^{(1)}\Pi T^{(1)} S- \Pi T^{(1)}S T^{(1)} \Pi -ST^{(1)}\Pi T^{(1)}\Pi . \end{aligned}$$

    Here S is the reduced resolvent of \({\hat{W}}_{s,t}\) at the eigenvalue 1.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Konno, N., Matsue, K. & Segawa, E. A Crossover Between Open Quantum Random Walks to Quantum Walks. J Stat Phys 190, 202 (2023). https://doi.org/10.1007/s10955-023-03211-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10955-023-03211-6

Keywords

Navigation