This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

11affiliationtext: College of Science and Engineering Ritsumeikan University
1-1-1 Noji-higashi, Kusatsu, 525-8577, Japan
22affiliationtext: Graduate School of Engineering Science, Osaka University
1-3, Machikaneyama, Toyonaka, Osaka, 560-8531, Japan

Eigenvalue analysis of three-state quantum walks
with general coin matrices

Jirô Akahori Chusei Kiumi Corresponding Author: c.kiumi@osaka-u.ac.jp Norio Konno Takuya Watanabe
Abstract

Mathematical analysis on the existence of eigenvalues is vital, as it corresponds to the occurrence of localization, an exceptionally important property of quantum walks. Previous studies have demonstrated that eigenvalue analysis utilizing the transfer matrix proves beneficial for space inhomogeneous three-state quantum walks with a specific class of coin matrices, including Grover matrices. In this research, we turn our attention to the transfer matrix of three-state quantum walks with a general coin matrix. Building upon previous research methodologies, we dive deeper into investigating the properties of the transfer matrix and employ numerical analysis to derive eigenvalues for models that were previously unanalyzable.

1 Introduction

Quantum walks (QWs) are fundamental models for understanding quantum dynamics. Despite their apparent simplicity, quantum walks exhibit a wide range of intricate quantum features, making them essential tools in various domains, including the study of topological phases [1] and the formulation of models and designs for quantum algorithms [2, 3, 4]. Within the context of QWs, the combination of the coin operator and the shift operator determines how the system evolves. A vital part of studying QWs is looking at the spectral properties of the time evolution operator. This helps us understand the behaviors and features of the walk, thereby enriching our comprehension of the underlying dynamics.

In this paper, our focus is on three-state QWs on a one-dimensional integer lattice, which involves working within an infinite-dimensional Hilbert space. This model is intensively studied in various articles with different contexts [5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20]. It is well-known that the occurrence of localization, a characteristic property of QWs, is equivalent to the existence of eigenvalues of the time evolution operator. Moreover, a quantitative investigation of localization can be conducted by deriving the time-averaged limit distribution using the eigenvalues and eigenvectors [21]. The transfer matrix method stands as a powerful tool for eigenvalue analysis and has been applied to several types of QW models [22, 23, 24], and even for non-unitary QWs [25]. We specifically consider the two-phase model with a finite number of defects. Previous studies [24] have conducted eigenvalue analysis for three-state QWs with a specific class of coin matrices, including Grover matrices, through the transfer matrix method. In contrast, this work extends the result by demonstrating that the method remains applicable and effective even for cases involving a general coin matrix.

Calculating explicit eigenvalues analytically for a general coin matrix remains highly challenging. Therefore, we resort to numerical calculations to determine eigenvalues, utilizing the method of the transfer matrix. For our numerical analysis, we concentrate on one-defect and two-phase Fourier walks. The Fourier walk has been intensively studied by [26, 27, 28] for the space-homogeneous model, where the coin matrices are identical at any position. Notably, it has been proven that the Fourier walk does not exhibit localization for the homogeneous model. However, we numerically demonstrate that the Fourier walk does exhibit localization for both the one-defect and two-phase models.

The remainder of this paper is organized as follows. In Section 2, we define our three-state QWs with a self-loop on the integer lattice. Section 3 introduces the transfer matrix with a general coin matrix and establishes methods for the eigenvalue problem. Theorem 1 constitutes the main theorem, providing a necessary and sufficient condition for the eigenvalue problem. Section 4 is dedicated to the numerical analysis of the Fourier walk. We present figures to visually depict the results of the numerical analysis on the eigenvalue equation and its probability distribution, thereby illustrating the occurrence of localization and the existence of the eigenvalue.

2 Definition

We define a three-state QW on the integer lattice \mathbb{Z}. Let us define the Hilbert space \mathcal{H} as

=2(;3)={Ψ:3|xΨ(x)32<}.\mathcal{H}=\ell^{2}(\mathbb{Z};\mathbb{C}^{3})=\left\{\Psi:\mathbb{Z}\rightarrow\mathbb{C}^{3}\ \middle|\ \sum_{x\in\mathbb{Z}}\|\Psi(x)\|_{\mathbb{C}^{3}}^{2}<\infty\right\}.

Here, \mathbb{C} denotes the set of complex numbers. The quantum state Ψ\Psi\in\mathcal{H} is expressed as

Ψ(x)=[Ψ1(x)Ψ2(x)Ψ3(x)].\Psi(x)=\left[\begin{array}[]{ c }\Psi_{1}(x)\\ \Psi_{2}(x)\\ \Psi_{3}(x)\end{array}\right].

Let {Cx}x{±}\{C_{x}\}_{x\in\mathbb{Z}\cup\{\pm\infty\}} be a sequence of 3×33\times 3 unitary matrices, which is written as below:

Cx=[ax(1,1)ax(1,2)ax(1,3)ax(2,1)ax(2,2)ax(2,3)ax(3,1)ax(3,2)ax(3,3)].C_{x}=\left[\begin{array}[]{ c c c }a_{x}^{(1,1)}&a_{x}^{(1,2)}&a_{x}^{(1,3)}\\ a_{x}^{(2,1)}&a_{x}^{(2,2)}&a_{x}^{(2,3)}\\ a_{x}^{(3,1)}&a_{x}^{(3,2)}&a_{x}^{(3,3)}\end{array}\right].

Here, we assume that |ax(2,2)|1\left|a_{x}^{(2,2)}\right|\neq 1. The coin operator is then defined using the coin matrices as:

(CΨ)(x)=CxΨ(x).(C\Psi)(x)=C_{x}\Psi(x).

Next, we define a shift operator SS that moves Ψ1(x)\Psi_{1}(x) to the left and Ψ3(x)\Psi_{3}(x) to the right.

(SΨ)(x)=[Ψ1(x+1)Ψ2(x)Ψ3(x1)].\ (S\Psi)(x)=\left[\begin{array}[]{ c }\Psi_{1}(x+1)\\ \Psi_{2}(x)\\ \Psi_{3}(x-1)\end{array}\right].

Finally, the time evolution of the QW is determined by the following unitary operator

U=SC.U=SC.

We impose a condition essential for considering eigenvalue analysis with our method.

Cx={C,x[x+,),Cx,x(x,x+),C,x(,x],C_{x}=\begin{cases}C_{\infty},&x\in[x_{+},\infty),\\ C_{x},&x\in(x_{-},x_{+}),\\ C_{-\infty},&x\in(-\infty,x_{-}],\end{cases}

where x+>0,x<0x_{+}>0,x_{-}<0.  We say the model with this condition as two-phase QW with finite number of defects. As a remark, we can easily replace C±C_{\pm\infty} with periodic coin matrices by following the previous research [29]. However we opt not to explore this generalization as it leads to notably more complex notation. For an initial state Ψ0(Ψ02=1)\Psi_{0}\in\mathcal{H}\left(\|\Psi_{0}\|_{\mathcal{H}}^{2}=1\right), the finding probability of a walker in position xx at time t0t\in\mathbb{Z}_{\geq 0} is defined by

μt(Ψ0)(x)=(UtΨ0)(x)32,\mu_{t}^{(\Psi_{0})}(x)=\left\|\left(U^{t}\Psi_{0}\right)(x)\right\|_{\mathbb{C}^{3}}^{2},

where 0\mathbb{Z}_{\geq 0} is the set of non-negative integers. We say that the QW\mathrm{QW} exhibits localization if there exists a position x0x_{0}\in\mathbb{Z} and an initial state Ψ0\Psi_{0}\in\mathcal{H} satisfying limsuptμt(Ψ0)(x0)>0\lim\sup_{t\rightarrow\infty}\mu_{t}^{(\Psi_{0})}(x_{0})>0. It is known that the QW exhibits localization if and only if there exists an eigenvalue of UU, that is, there exists λ[0,2π)\lambda\in[0,2\pi) and Ψ{𝟎}\Psi\in\mathcal{H}\setminus\{\mathbf{0}\} such that

UΨ=eiλΨ.U\Psi=e^{i\lambda}\Psi.

Let σp(U)\sigma_{p}(U) hereafter denote the set of eigenvalues of UU.

3 Eigenvalue analysis with transfer matrix

The eigenvalue equation UΨ=eiλΨU\Psi=e^{i\lambda}\Psi can be rewritten as the following system of equations:

ei(λΔx)Ψ1(x1)=Ax(λ)Ψ1(x)+Bx(λ)Ψ3(x),\displaystyle e^{i(\lambda-\Delta_{x})}\Psi_{1}(x-1)=A_{x}(\lambda)\Psi_{1}(x)+B_{x}(\lambda)\Psi_{3}(x), (1)
ei(λΔx)Ψ3(x+1)=Cx(λ)Ψ1(x)+Dx(λ)Ψ3(x),\displaystyle e^{i(\lambda-\Delta_{x})}\Psi_{3}(x+1)=C_{x}(\lambda)\Psi_{1}(x)+D_{x}(\lambda)\Psi_{3}(x), (2)
Ψ2(x)=Ex(λ)Ψ1(x)+Fx(λ)Ψ3(x).\displaystyle\Psi_{2}(x)=E_{x}(\lambda)\Psi_{1}(x)+F_{x}(\lambda)\Psi_{3}(x). (3)

where

Ax(λ)=ax(1,1)+ax(1,2)ax(2,1)eiλax(2,2),Bx(λ)=ax(1,3)+ax(1,2)ax(2,3)eiλax(2,2),\displaystyle A_{x}(\lambda)=a_{x}^{(1,1)}+\frac{a_{x}^{(1,2)}a_{x}^{(2,1)}}{e^{i\lambda}-a_{x}^{(2,2)}},\ B_{x}(\lambda)=a_{x}^{(1,3)}+\frac{a_{x}^{(1,2)}a_{x}^{(2,3)}}{e^{i\lambda}-a_{x}^{(2,2)}},
Cx(λ)=ax(3,1)+ax(3,2)ax(2,1)eiλax(2,2),Dx(λ)=ax(3,3)+ax(3,2)ax(2,3)eiλax(2,2),\displaystyle C_{x}(\lambda)=a_{x}^{(3,1)}+\frac{a_{x}^{(3,2)}a_{x}^{(2,1)}}{e^{i\lambda}-a_{x}^{(2,2)}},\ D_{x}(\lambda)=a_{x}^{(3,3)}+\frac{a_{x}^{(3,2)}a_{x}^{(2,3)}}{e^{i\lambda}-a_{x}^{(2,2)}},
Ex(λ)=ax(2,1)eiλax(2,2),Fx(λ)=ax(2,3)eiλax(2,2).\displaystyle E_{x}(\lambda)=\frac{a_{x}^{(2,1)}}{e^{i\lambda}-a_{x}^{(2,2)}},\ F_{x}(\lambda)=\frac{a_{x}^{(2,3)}}{e^{i\lambda}-a_{x}^{(2,2)}}.

Here, we define transfer matrices Tx(λ)T_{x}(\lambda) by

Tx(λ)=1Ax(λ)[eiλBx(λ)Cx(λ)eiλ(Bx(λ)Cx(λ)Ax(λ)Dx(λ))],T_{x}(\lambda)=\frac{1}{A_{x}(\lambda)}\left[\begin{array}[]{ c c }e^{i\lambda}&-B_{x}(\lambda)\\ C_{x}(\lambda)&-e^{-i\lambda}(B_{x}(\lambda)C_{x}(\lambda)-A_{x}(\lambda)D_{x}(\lambda))\end{array}\right],

then equations (1), (2) can be expressed as

[Ψ1(x)Ψ3(x+1)]=Tx(λ)[Ψ1(x1)Ψ3(x)].\left[\begin{array}[]{ c }\Psi_{1}(x)\\ \Psi_{3}(x+1)\end{array}\right]=T_{x}(\lambda)\left[\begin{array}[]{ c }\Psi_{1}(x-1)\\ \Psi_{3}(x)\end{array}\right].

From the unitarity of the coin matrix, we can write

Cx=eiΔx[detCx(1,1)¯detCx(1,2)¯detCx(1,3)¯detCx(2,1)¯detCx(2,2)¯detCx(2,3)¯detCx(3,1)¯detCx(3,2)¯detCx(3,3)¯],C_{x}=e^{i\Delta_{x}}\left[\begin{array}[]{ c c c }\overline{\det C_{x}^{(1,1)}}&-\overline{\det C_{x}^{(1,2)}}&\overline{\det C_{x}^{(1,3)}}\\ -\overline{\det C_{x}^{(2,1)}}&\overline{\det C_{x}^{(2,2)}}&-\overline{\det C_{x}^{(2,3)}}\\ \overline{\det C_{x}^{(3,1)}}&-\overline{\det C_{x}^{(3,2)}}&\overline{\det C_{x}^{(3,3)}}\end{array}\right],

where eiΔx=detCxe^{i\Delta_{x}}=\det C_{x} and  Cx(i,j)C_{x}^{(i,j)} denote the 2×22\times 2 submatrix obtained by excluding the ii-th row and jj-th column of CxC_{x}. Using the above expression, we can simplify each component of the transfer matrix as:

Ax(λ)=ax(1,1)eiλeiΔxax(3,3)¯eiλax(2,2),Bx(λ)=ax(1,3)eiλ+eiΔxax(3,1)¯eiλax(2,2),\displaystyle A_{x}(\lambda)=\frac{a_{x}^{(1,1)}e^{i\lambda}-e^{i\Delta_{x}}\overline{a_{x}^{(3,3)}}}{e^{i\lambda}-a_{x}^{(2,2)}},\ B_{x}(\lambda)=\frac{a_{x}^{(1,3)}e^{i\lambda}+e^{i\Delta_{x}}\overline{a_{x}^{(3,1)}}}{e^{i\lambda}-a_{x}^{(2,2)}},
Cx(λ)=ax(3,1)eiλ+eiΔxax(1,3)¯eiλax(2,2),Dx(λ)=ax(3,3)eiλeiΔxax(1,1)¯eiλax(2,2)\displaystyle C_{x}(\lambda)=\frac{a_{x}^{(3,1)}e^{i\lambda}+e^{i\Delta_{x}}\overline{a_{x}^{(1,3)}}}{e^{i\lambda}-a_{x}^{(2,2)}},\ D_{x}(\lambda)=\frac{a_{x}^{(3,3)}e^{i\lambda}-e^{i\Delta_{x}}\overline{a_{x}^{(1,1)}}}{e^{i\lambda}-a_{x}^{(2,2)}}

and we can further calculate

Ax(λ)Dx(λ)Bx(λ)Cx(λ)=detCx(2,2)eiλeiΔxeiλax(2,2)=ei(λ+Δx)eiλax(2,2)¯eiλax(2,2).A_{x}(\lambda)D_{x}(\lambda)-B_{x}(\lambda)C_{x}(\lambda)=\frac{\det C_{x}^{(2,2)}e^{i\lambda}-e^{i\Delta_{x}}}{e^{i\lambda}-a_{x}^{(2,2)}}=-e^{i(\lambda+\Delta_{x})}\frac{e^{-i\lambda}-\overline{a_{x}^{(2,2)}}}{e^{i\lambda}-a_{x}^{(2,2)}}.

Hence, we can simplify the expression of the transfer matrix as

Tx(λ)=1ax(1,1)eiλeiΔxax(3,3)¯[eiλ(eiλax(2,2))ax(1,3)eiλeiΔxax(3,1)¯ax(3,1)eiλ+eiΔxax(1,3)¯eiΔx(eiλax(2,2)¯)].T_{x}(\lambda)=\frac{1}{a_{x}^{(1,1)}e^{i\lambda}-e^{i\Delta_{x}}\overline{a_{x}^{(3,3)}}}\begin{bmatrix}e^{i\lambda}\left(e^{i\lambda}-a_{x}^{(2,2)}\right)&-a_{x}^{(1,3)}e^{i\lambda}-e^{i\Delta_{x}}\overline{a_{x}^{(3,1)}}\\ a_{x}^{(3,1)}e^{i\lambda}+e^{i\Delta_{x}}\overline{a_{x}^{(1,3)}}&-e^{i\Delta_{x}}\left(e^{-i\lambda}-\overline{a_{x}^{(2,2)}}\right)\end{bmatrix}.

Note that, when Ax(λ)=0A_{x}(\lambda)=0, we cannot construct a transfer matrix. Therefore, we must treat this case separately. Let us define

Λ0={eiλx{±},Ax(λ)=0}\Lambda_{0}=\left\{e^{i\lambda}\in\mathbb{C}\mid\exists x\in\mathbb{Z}\cup\{\pm\infty\},\ A_{x}(\lambda)=0\right\}

and we consider the case eiλΛ0e^{i\lambda}\notin\Lambda_{0} first. For simplification, we may write Tx(λ)T_{x}(\lambda) as TxT_{x} henceforward. Let eiλΛ0e^{i\lambda}\notin\Lambda_{0} and φ2\varphi\in\mathbb{C}^{2}, we define Ψ~φ:2\tilde{\Psi}_{\varphi}:\mathbb{Z}\rightarrow\mathbb{C}^{2} as follows:

Ψ~φ(x)={Txx+T+φ,x+x,Tx1T0φ,0<x<x+,φ,x=0,Tx1T11φ,x<x<0,TxxTφ,xx,\tilde{\Psi}_{\varphi}(x)=\begin{cases}T_{\infty}^{x-x_{+}}T_{+}\varphi,&x_{+}\leq x,\\ T_{x-1}\cdots T_{0}\varphi,&0<x<x_{+},\\ \varphi,&x=0,\\ T_{x}^{-1}\cdots T_{-1}^{-1}\varphi,&x_{-}<x<0,\\ T_{-\infty}^{x-x_{-}}T_{-}\varphi,&x\leq x_{-},\end{cases} (4)

where T+=Tx+1T0,T=Tx1T11T_{+}=T_{x_{+}-1}\cdots T_{0},T_{-}=T_{x_{-}}^{-1}\cdots T_{-1}^{-1} and T±=Tx±T_{\pm\infty}=T_{x_{\pm}}. Let VλV_{\lambda} be a set of generalized eigenvectors and WλW_{\lambda} be a set of reduced vectors Ψ~φ\tilde{\Psi}_{\varphi} defined by (4):

Vλ={Ψ:3UΨ=eiλΨ},\displaystyle V_{\lambda}=\left\{\Psi:\mathbb{Z}\rightarrow\mathbb{C}^{3}\mid U\Psi=e^{i\lambda}\Psi\right\},
Wλ={Ψ~φφ2}.\displaystyle W_{\lambda}=\left\{\tilde{\Psi}_{\varphi}\mid\varphi\in\mathbb{C}^{2}\right\}.

For eiλΛ0e^{i\lambda}\notin\Lambda_{0}, we define a bijective map ι:VλWλ\iota:V_{\lambda}\rightarrow W_{\lambda} by

(ιΨ)(x)=[Ψ1(x1)Ψ3(x)].(\iota\Psi)(x)=\left[\begin{array}[]{ c }\Psi_{1}(x-1)\\ \Psi_{3}(x)\end{array}\right].

Here, the inverse of ι\iota is given as

(ι1Ψ~)(x)=[Ψ~1(x+1)Ex(λ)Ψ~1(x+1)+Fx(λ)Ψ~2(x)Ψ~2(x)]\left(\iota^{-1}\tilde{\Psi}\right)(x)=\left[\begin{array}[]{ c }\tilde{\Psi}_{1}(x+1)\\ E_{x}(\lambda)\tilde{\Psi}_{1}(x+1)+F_{x}(\lambda)\tilde{\Psi}_{2}(x)\\ \tilde{\Psi}_{2}(x)\end{array}\right]

for Ψ~Wλ,Ψ~(x)=[Ψ~1(x)Ψ~2(x)]t\tilde{\Psi}\in W_{\lambda},\ \tilde{\Psi}(x)=\left[\begin{array}[]{ c c }\tilde{\Psi}_{1}(x)&\tilde{\Psi}_{2}(x)\end{array}\right]^{t}. Thus, from the definition, ΨVλ\Psi\in V_{\lambda} if and only if there exists Ψ~Wλ\tilde{\Psi}\in W_{\lambda} such that Ψ=ι1Ψ~\Psi=\iota^{-1}\tilde{\Psi}. We can also see that ι1Ψ~{𝟎}\iota^{-1}\tilde{\Psi}\in\mathcal{H}\setminus\{\mathbf{0}\} is equivalent to Ψ~2(;2){𝟎}\tilde{\Psi}\in\ell^{2}\left(\mathbb{Z};\mathbb{C}^{2}\right)\setminus\{\mathbf{0}\}.

Corollary 3.1.

Let λ[0,2π)\lambda\in[0,2\pi) satisfying Ax(λ)0A_{x}(\lambda)\neq 0 for all x,eiλx\in\mathbb{Z},\ e^{i\lambda} is the eigenvalue of UU, i.e., eiλσp(U)e^{i\lambda}\in\sigma_{p}(U), if and only if there exists Ψ~Wλ{𝟎}\tilde{\Psi}\in W_{\lambda}\setminus\{\mathbf{0}\} such that Ψ~2(;2)\tilde{\Psi}\in\ell^{2}\left(\mathbb{Z};\mathbb{C}^{2}\right), and the associated eigenvector of eiλe^{i\lambda} becomes ι1Ψ~\iota^{-1}\tilde{\Psi}.

The subsequent property is pivotal in formulating the method for the eigenvalue problem.

Proposition 3.2.

For all λ[π,π)\lambda\in[-\pi,\pi) and x{±}x\in\mathbb{Z}\cup\{\pm\infty\}, we have

|detTx(λ)|=1.\left|\det T_{x}(\lambda)\right|=1.
Proof.

By direct calculation, we have

|ax(1,1)eiλeiΔxax(3,3)¯|2=|ax(3,3)eiλeiΔxax(1,1)¯|2\displaystyle\left|a_{x}^{(1,1)}e^{i\lambda}-e^{i\Delta_{x}}\overline{a_{x}^{(3,3)}}\right|^{2}=\left|a_{x}^{(3,3)}e^{i\lambda}-e^{i\Delta_{x}}\overline{a_{x}^{(1,1)}}\right|^{2}
=|ax(1,1)|2+|ax(3,3)|22(ax(1,1)ax(3,3)ei(λΔx)).\displaystyle=\left|a_{x}^{(1,1)}\right|^{2}+\left|a_{x}^{(3,3)}\right|^{2}-2\Re\left(a_{x}^{(1,1)}a_{x}^{(3,3)}e^{i(\lambda-\Delta_{x})}\right).

Therefore,

|Ax(λ)Dx(λ)|2=|ax(1,1)eiλeiΔxax(3,3)¯|2|ax(3,3)eiλeiΔxax(1,1)¯|2=1\left|\frac{A_{x}(\lambda)}{D_{x}(\lambda)}\right|^{2}=\frac{\left|a_{x}^{(1,1)}e^{i\lambda}-e^{i\Delta_{x}}\overline{a_{x}^{(3,3)}}\right|^{2}}{\left|a_{x}^{(3,3)}e^{i\lambda}-e^{i\Delta_{x}}\overline{a_{x}^{(1,1)}}\right|^{2}}=1

which implies

|detTx(λ)|=|Ax(λ)Dx(λ)|=1.|\det T_{x}(\lambda)|=\left|\frac{A_{x}(\lambda)}{D_{x}(\lambda)}\right|=1.

Thus, we can convince that the following two eigenvalues  ζx+\zeta_{x}^{+}, ζx\zeta_{x}^{-} of the transfer matrix satisfies |ζx+||ζx|=1.|\zeta_{x}^{+}||\zeta_{x}^{-}|=1.

ζx±=tr(Tx)±tr(Tx)24detTx2.\zeta_{x}^{\pm}=\frac{\mathrm{tr}(T_{x})\pm\sqrt{\mathrm{tr}(T_{x})^{2}-4\det T_{x}}}{2}.

In this paper, we define the square root of the complex number a=|a|eiθ,θ[0,2π)a=|a|e^{i\theta},\ \theta\in[0,2\pi) as a=|a|eiθ2\sqrt{a}=\sqrt{|a|}e^{i\frac{\theta}{2}}. Consequently, we designate ζx<\zeta_{x}^{<} as one of ζx±\zeta_{x}^{\pm} such that |ζx<|1\left|\zeta_{x}^{<}\right|\leq 1, and ζx>\zeta_{x}^{>} such that |ζx>|1.\left|\zeta_{x}^{>}\right|\geq 1. Additionally, we introduce |vx>\ket{v_{x}^{>}} and |vx<\ket{v_{x}^{<}} as normalized eigenvector of TxT_{x} corresponding to ζx>\zeta_{x}^{>} and ζx<\zeta_{x}^{<}, respectively.

Proposition 3.3.

The condition |ζ±+|=|ζ±|=1\left|\zeta_{\pm\infty}^{+}\right|=\left|\zeta_{\pm\infty}^{-}\right|=1 holds if and only if:

detT±¯trT±,detT±¯(trT±)240.\sqrt{\overline{\det T_{\pm\infty}}}\operatorname{tr}T_{\pm\infty}\in\mathbb{R},\ \overline{\det T_{\pm\infty}}\left(\operatorname{tr}T_{\pm\infty}\right)^{2}-4\leq 0.
Proof.

We aim to determine the conditions for

|ζ±+|=|ζ±|=1.\ \left|\zeta_{\pm\infty}^{+}\right|=\left|\zeta_{\pm\infty}^{-}\right|=1.

This is equivalent to

|tr(T±)+tr(T±)24detT±|2=|tr(T±)tr(T±)24detT±|2=4.\left|\mathrm{tr}(T_{\pm\infty})+\sqrt{\mathrm{tr}(T_{\pm\infty})^{2}-4\det T_{\pm\infty}}\right|^{2}=\left|\mathrm{tr}(T_{\pm\infty})-\sqrt{\mathrm{tr}(T_{\pm\infty})^{2}-4\det T_{\pm\infty}}\right|^{2}=4. (5)

For this to hold, it is necessary that

|tr(T±)+tr(T±)24detT±|2=|tr(T±)tr(T±)24detT±|2\displaystyle\left|\mathrm{tr}(T_{\pm\infty})+\sqrt{\mathrm{tr}(T_{\pm\infty})^{2}-4\det T_{\pm\infty}}\right|^{2}=\left|\mathrm{tr}(T_{\pm\infty})-\sqrt{\mathrm{tr}(T_{\pm\infty})^{2}-4\det T_{\pm\infty}}\right|^{2}
\displaystyle\iff (tr(T±)¯tr(T±)24detT±)=0.\displaystyle\Re\left(\overline{\mathrm{tr}(T_{\pm\infty})}\sqrt{\mathrm{tr}(T_{\pm\infty})^{2}-4\det T_{\pm\infty}}\right)=0.

Here, \iff denotes “if and only if”. By using (5), we deduce

|tr(T±)+tr(T±)24detT±|2=4\displaystyle\left|\mathrm{tr}(T_{\pm\infty})+\sqrt{\mathrm{tr}(T_{\pm\infty})^{2}-4\det T_{\pm\infty}}\right|^{2}=4
\displaystyle\iff |tr(T±)24detT±|2=4|tr(T±)|2\displaystyle\left|\sqrt{\mathrm{tr}(T_{\pm\infty})^{2}-4\det T_{\pm\infty}}\right|^{2}=4-|\mathrm{tr}(T_{\pm\infty})|^{2}
\displaystyle\iff |tr(T±)24detT±|=4|tr(T±)|2.\displaystyle\left|\mathrm{tr}(T_{\pm\infty})^{2}-4\det T_{\pm\infty}\right|=4-|\mathrm{tr}(T_{\pm\infty})|^{2}.

Squaring both side, we obtain the equivalent statement:

(tr(T±)detT±¯tr(T±)2detT±¯)2=0and|tr(T±)|240\displaystyle\left(\mathrm{tr}(T_{\pm\infty})\sqrt{\overline{\det T_{\pm\infty}}}-\mathrm{tr}(T_{\pm\infty})^{2}\overline{\det T_{\pm\infty}}\right)^{2}=0\ \text{and}\ |\mathrm{tr}(T_{\pm\infty})|^{2}-4\leq 0 (6)
\displaystyle\iff tr(T±)detT±¯tr(T±)2detT±¯40.\displaystyle\mathrm{tr}(T_{\pm\infty})\sqrt{\overline{\det T_{\pm\infty}}}\in\mathbb{R}\land\mathrm{tr}(T_{\pm\infty})^{2}\overline{\det T_{\pm\infty}}-4\leq 0. (7)

It can also be verified that if (7) holds, so does (5). This completes the proof. ∎

Let

Λ={eiλ||ζ±>||ζ±<|}.\Lambda=\left\{e^{i\lambda}\in\mathbb{C}\ \middle|\ \left|\zeta_{\pm\infty}^{>}\right|\neq\left|\zeta_{\pm\infty}^{<}\right|\right\}.

With this, we present the following theorem:

Theorem 1.

eiλΛ0e^{i\lambda}\notin\Lambda_{0} is the eigenvalue of UU, i.e., eiλσp(U)e^{i\lambda}\in\sigma_{p}(U), if and only if the following two conditions are satisfied:

1.eiλΛ,\displaystyle 1.\ e^{i\lambda}\in\Lambda,
2.ker((Tζ<)T+)ker((Tζ>)T){𝟎}.\displaystyle 2.\ \ker\left(\left(T_{\infty}-\zeta_{\infty}^{<}\right)T_{+}\right)\cap\ker\left(\left(T_{-\infty}-\zeta_{-\infty}^{>}\right)T_{-}\right)\neq\{\mathbf{0}\}.
Proof.

We know from Corollary 3.1 that eiλσp(U)e^{i\lambda}\in\sigma_{p}(U) if and only if there exists Ψ~Wλ{𝟎}\tilde{\Psi}\in W_{\lambda}\setminus\{\mathbf{0}\} such that xΨ~(x)22<\sum_{x\in\mathbb{Z}}\|\tilde{\Psi}(x)\|_{\mathbb{C}^{2}}^{2}<\infty.  If eiλΛe^{i\lambda}\notin\Lambda,  both |ζ±<||\zeta_{\pm\infty}^{<}| and |ζ±>||\zeta_{\pm\infty}^{>}| become 1 by definition. Since  Ψ~(x)\tilde{\Psi}(x) is given as (4), xΨ~(x)22=\sum_{x\in\mathbb{Z}}\|\tilde{\Psi}(x)\|_{\mathbb{C}^{2}}^{2}=\infty for all Ψ~Wλ{𝟎}\tilde{\Psi}\in W_{\lambda}\setminus\{\mathbf{0}\}. Therefore, eiλΛe^{i\lambda}\in\Lambda is a necessary condition for eiλσp(U)e^{i\lambda}\in\sigma_{p}(U). Next we assume that  eiλΛe^{i\lambda}\in\Lambda, then |ζ±>|>1|\zeta_{\pm\infty}^{>}|>1 and |ζ±<|<1|\zeta_{\pm\infty}^{<}|<1 hold. Since Ψ~Wλ{𝟎}\tilde{\Psi}\in W_{\lambda}\setminus\{\mathbf{0}\} is expressed by φ2{𝟎}\varphi\in\mathbb{C}^{2}\setminus\{\mathbf{0}\} and transfer matrices, there exists Ψ~Wλ{𝟎}\tilde{\Psi}\in W_{\lambda}\setminus\{\mathbf{0}\} such that xΨ~(x)22<\sum_{x\in\mathbb{Z}}\|\tilde{\Psi}(x)\|_{\mathbb{C}^{2}}^{2}<\infty if and only if there exists φ2{𝟎}\varphi\in\mathbb{C}^{2}\setminus\{\mathbf{0}\} such that T+φker(Tζ<),Tφker(Tζ>)T_{+}\varphi\in\ker\left(T_{\infty}-\zeta_{\infty}^{<}\right),\ T_{-}\varphi\in\ker\left(T_{-\infty}-\zeta_{-\infty}^{>}\right), that is, φker((Tζ<)T+)ker((Tζ>)T).\varphi\in\ker\left(\left(T_{\infty}-\zeta_{\infty}^{<}\right)T_{+}\right)\cap\ker\left(\left(T_{-\infty}-\zeta_{-\infty}^{>}\right)T_{-}\right). The conclusions drawn from the above discussions validate the statement. ∎

Let T+T1|v>T_{+}T_{-}^{-1}\ket{v_{-\infty}^{>}} be represented as [u1u2]\begin{bmatrix}u_{1}&u_{2}\end{bmatrix}, and let |v<\ket{v_{\infty}^{<}} be given by [v1v2]\begin{bmatrix}v_{1}&v_{2}\end{bmatrix}. Define

χ(λ)=u1v2u2v1.\chi(\lambda)=u_{1}v_{2}-u_{2}v_{1}.

Then, we can rewrite Theorem 1 with equation χ(λ)\chi(\lambda).

Corollary 3.4.

eiλΛ0e^{i\lambda}\notin\Lambda_{0} is the eigenvalue of UU, i.e., eiλσp(U)e^{i\lambda}\in\sigma_{p}(U), if and only if the following conditions are met:

eiλΛ,χ(λ)=0.e^{i\lambda}\in\Lambda,\ \chi(\lambda)=0.

From Theorem 1, for eiλσp(U)Λ0e^{i\lambda}\in\sigma_{p}(U)\setminus\Lambda_{0}, the eigenvector Ψker(Ueiλ){𝟎}\Psi\in\ker(U-e^{i\lambda})\setminus\{\mathbf{0}\} is given by Ψ=ι1Ψ~\Psi=\iota^{-1}\tilde{\Psi}, where Ψ~Wλ{𝟎}\tilde{\Psi}\in W_{\lambda}\setminus\{\mathbf{0}\} becomes

Ψ~(x)={(ζ<)xx+T+φ,x+x,Tx1T0φ,0<x<x+,φ,x=0,Tx1T11φ,x<x<0,(ζ>)xxTφ,xx,\displaystyle\tilde{\Psi}(x)=\begin{cases}(\zeta_{\infty}^{<})^{x-x_{+}}T_{+}\varphi,&x_{+}\leq x,\\ T_{x-1}\cdots T_{0}\varphi,&0<x<x_{+},\\ \varphi,&x=0,\\ T_{x}^{-1}\cdots T_{-1}^{-1}\varphi,&x_{-}<x<0,\\ (\zeta_{-\infty}^{>})^{x-x_{-}}T_{-}\varphi,&x\leq x_{-},\end{cases}

with φker((Tζ<)T+)ker((Tζ>)T){𝟎}.\varphi\in\ker\left(\left(T_{\infty}-\zeta_{\infty}^{<}\right)T_{+}\right)\cap\ker\left(\left(T_{-\infty}-\zeta_{-\infty}^{>}\right)T_{-}\right)\setminus\{\mathbf{0}\}. Furthermore, it is straightforward to verify that dimker(Ueiλ)=1\dim\ker(U-e^{i\lambda})=1.

If Ax=0A_{x}=0 for some xx\in\mathbb{Z}, we cannot construct a transfer matrix at xx. Therefore, we have to deal with this case separately.

Lemma 3.5.

For all λ(0,2π]\lambda\in(0,2\pi], we have

Ax(λ)=0Dx(λ)=0A_{x}(\lambda)=0\iff D_{x}(\lambda)=0
Proof.

We assume that Ax(λ)=0A_{x}(\lambda)=0, which means

ei(λΔx)=ax(3,3)¯ax(1,1) and |ax(3,3)|=|ax(1,1)|.e^{i(\lambda-\Delta_{x})}=\frac{\overline{a_{x}^{(3,3)}}}{a_{x}^{(1,1)}}\text{ and }\left|a_{x}^{(3,3)}\right|=\left|a_{x}^{(1,1)}\right|.

Then, we can easily calculate that

Dx(λ)=|ax(3,3)|2|ax(1,1)|2ax(3,3)¯ax(1,1)ax(2,2)eiΔx=0.D_{x}(\lambda)=\frac{\left|a_{x}^{(3,3)}\right|^{2}-\left|a_{x}^{(1,1)}\right|^{2}}{\overline{a_{x}^{(3,3)}}-a_{x}^{(1,1)}a_{x}^{(2,2)}e^{-i\Delta_{x}}}=0.

Similarly Dx(λ)=0D_{x}(\lambda)=0 implies

ei(λΔx)=ax(1,1)¯ax(3,3) and |ax(1,1)|=|ax(3,3)|e^{i(\lambda-\Delta_{x})}=\frac{\overline{a_{x}^{(1,1)}}}{a_{x}^{(3,3)}}\text{ and }\left|a_{x}^{(1,1)}\right|=\left|a_{x}^{(3,3)}\right|

and Ax(λ)=0A_{x}(\lambda)=0.

Proposition 3.6.

Consider Ψ:3\Psi:\mathbb{Z}\rightarrow\mathbb{C}^{3}. The equation UΨ=eiλΨU\Psi=e^{i\lambda}\Psi is satisfied if and only if:

[Ψ1(x)Ψ3(x+1)]=Tx[Ψ1(x1)Ψ3(x)],\left[\begin{array}[]{ c }\Psi_{1}(x)\\ \Psi_{3}(x+1)\end{array}\right]=T_{x}\left[\begin{array}[]{ c }\Psi_{1}(x-1)\\ \Psi_{3}(x)\end{array}\right],

for xx with Ax(λ)0A_{x}(\lambda)\neq 0 and

Ψ1(x1)=ax(3,3)ax(3,2)¯ax(1,1)¯ax(2,1)Ψ3(x),Ψ1(x)=ax(1,1)¯ax(1,2)ax(3,3)ax(2,3)¯Ψ3(x+1),\Psi_{1}(x-1)=\frac{a_{x}^{(3,3)}\overline{a_{x}^{(3,2)}}}{\overline{a_{x}^{(1,1)}}a_{x}^{(2,1)}}\Psi_{3}(x),\ \Psi_{1}(x)=\frac{\overline{a_{x}^{(1,1)}}a_{x}^{(1,2)}}{a_{x}^{(3,3)}\overline{a_{x}^{(2,3)}}}\Psi_{3}(x+1), (8)

for xx with Ax(λ)=0A_{x}(\lambda)=0.

Proof.

Assume Ax(λ)=0.A_{x}(\lambda)=0. By our previous lemma, this implies Dx(λ)=0.D_{x}(\lambda)=0. Given this, we have ei(λΔx)=ax(3,3)¯ax(1,1)=ax(1,1)¯ax(3,3)e^{i(\lambda-\Delta_{x})}=\frac{\overline{a_{x}^{(3,3)}}}{a_{x}^{(1,1)}}=\frac{\overline{a_{x}^{(1,1)}}}{a_{x}^{(3,3)}} and

Ψ1(x1)=eiλBx(λ)Ψ3(x),\displaystyle\Psi_{1}(x-1)=e^{-i\lambda}B_{x}(\lambda)\Psi_{3}(x),
Ψ1(x)=eiλCx1(λ)Ψ3(x+1),\displaystyle\Psi_{1}(x)=e^{i\lambda}C_{x}^{-1}(\lambda)\Psi_{3}(x+1),

where

Bx(λ)=a13a33¯+a11a31¯a33¯a11a22eiΔ,Cx(λ)=a31a33¯+a11a13¯a33¯a11a22eiΔ.B_{x}(\lambda)=\frac{a_{13}\overline{a_{33}}+a_{11}\overline{a_{31}}}{\overline{a_{33}}-a_{11}a_{22}e^{-i\Delta}},\ C_{x}(\lambda)=\frac{a_{31}\overline{a_{33}}+a_{11}\overline{a_{13}}}{\overline{a_{33}}-a_{11}a_{22}e^{-i\Delta}}.

This is equivalent to

Ψ1(x1)=ax(3,3)ax(3,2)¯ax(1,1)¯ax(2,1)Ψ3(x),Ψ1(x)=ax(1,1)¯ax(1,2)ax(3,3)ax(2,3)¯Ψ3(x+1).\Psi_{1}(x-1)=\frac{a_{x}^{(3,3)}\overline{a_{x}^{(3,2)}}}{\overline{a_{x}^{(1,1)}}a_{x}^{(2,1)}}\Psi_{3}(x),\ \Psi_{1}(x)=\frac{\overline{a_{x}^{(1,1)}}a_{x}^{(1,2)}}{a_{x}^{(3,3)}\overline{a_{x}^{(2,3)}}}\Psi_{3}(x+1).

As a consequence of the last proposition, we obtain the following prposition.

Proposition 3.7.

eiλ=a±(3,3)¯a±(1,1)eiΔ±e^{i\lambda}=\frac{\overline{a_{\pm\infty}^{(3,3)}}}{a_{\pm\infty}^{(1,1)}}e^{i\Delta_{\pm\infty}} is eigenvalue of UU if and only if

(a±(3,3))2a±(3,2)¯a±(2,3)¯=(a±(1,1)¯)2a±(1,2)a±(2,1).\left(a_{\pm\infty}^{(3,3)}\right)^{2}\overline{a_{\pm\infty}^{(3,2)}}\overline{a_{\pm\infty}^{(2,3)}}=\left(\overline{a_{\pm\infty}^{(1,1)}}\right)^{2}a_{\pm\infty}^{(1,2)}a_{\pm\infty}^{(2,1)}.

We get the following corollary, which was also proved in [30].

Corollary 3.8.

If the model is homogeneous, i.e., the coin matrix is

Cx=[a11a12a13a21a22a23a31a32a33]C_{x}=\left[\begin{array}[]{ c c c }a_{11}&a_{12}&a_{13}\\ a_{21}&a_{22}&a_{23}\\ a_{31}&a_{32}&a_{33}\end{array}\right]

for xx\in\mathbb{Z}. Then σp(U)={a33¯a11eiΔ}\sigma_{p}(U)=\left\{\frac{\overline{a_{33}}}{a_{11}}e^{i\Delta}\right\} if and only if

a332a32¯a23¯=a11¯2a12a21.a_{33}^{2}\overline{a_{32}}\overline{a_{23}}=\overline{a_{11}}^{2}a_{12}a_{21}.
Proof.

The condition (8) holds for all xx\in\mathbb{Z}, that is,

a33a11¯a32¯a21=a11¯a33a12a23¯\frac{a_{33}}{\overline{a_{11}}}\frac{\overline{a_{32}}}{a_{21}}=\frac{\overline{a_{11}}}{a_{33}}\frac{a_{12}}{\overline{a_{23}}}

if and only if a332a32¯a23¯=a11¯2a12a21a_{33}^{2}\overline{a_{32}}\overline{a_{23}}=\overline{a_{11}}^{2}a_{12}a_{21}. Also, Theorem 1 shows that homogeneous model does not have eigenvalue for eiλΛ0e^{i\lambda}\notin\Lambda_{0}. ∎

4 Numerical analysis on Fourier walk

In this section, we deal with the Fourier walk studied in [26, 27, 28]. Let us consider the following Fourier matrix as a coin matrix.

F=13[1111ωω21ω2ω]F=\frac{1}{\sqrt{3}}\begin{bmatrix}1&1&1\\ 1&\omega&\omega^{2}\\ 1&\omega^{2}&\omega\end{bmatrix}

where ω=ei2π3\omega=e^{i\frac{2\pi}{3}}.

4.1 One-defect model

Consider the coin matrix defined as:

Cx={eiθFx=0,Fx0.C_{x}=\begin{cases}e^{i\theta}F&x=0,\\ F&x\neq 0.\end{cases}

with some θ(0,2π]\theta\in(0,2\pi].

Proposition 4.1.

For a one-defect Fourier walk,

Λ0σp(U)={a0(3,3)¯a0(1,1)eiΔ0,a(3,3)¯a(1,1)eiΔ}σp(U)=.\Lambda_{0}\cap\sigma_{p}(U)=\left\{\frac{\overline{a_{0}^{(3,3)}}}{a_{0}^{(1,1)}}e^{i\Delta_{0}},\frac{\overline{a^{(3,3)}}}{a^{(1,1)}}e^{i\Delta}\right\}\cap\sigma_{p}(U)=\emptyset.
Proof.

From Proposition 3.7, we know that eiλ=a(3,3)¯a(1,1)eiΔe^{i\lambda}=\frac{\overline{a^{(3,3)}}}{a^{(1,1)}}e^{i\Delta} is not an eigenvalue of UU since

(a(3,3))2a(3,2)¯a(2,3)¯=ω33133=a(1,1)¯2a(1,2)a(2,1).\displaystyle\left(a^{(3,3)}\right)^{2}\overline{a^{(3,2)}}\overline{a^{(2,3)}}=\frac{\omega}{3\sqrt{3}}\neq\frac{1}{3\sqrt{3}}=\ \overline{a^{(1,1)}}^{2}a^{(1,2)}a^{(2,1)}.

Also, we prove that a0(3,3)¯a0(1,1)eiΔ0\frac{\overline{a_{0}^{(3,3)}}}{a_{0}^{(1,1)}}e^{i\Delta_{0}} is not an eigenvalue of UU by contradiction. Assuming that eiλ=a0(3,3)¯a0(1,1)eiΔ0σp(U)e^{i\lambda}=\frac{\overline{a_{0}^{(3,3)}}}{a_{0}^{(1,1)}}e^{i\Delta_{0}}\in\sigma_{p}(U) then

φ1:=(Ψ1(1)Ψ3(0))=k1(ω21),φ2:=(Ψ1(0)Ψ3(1))=k2(ω1)\varphi_{1}:=\begin{pmatrix}\Psi_{1}(-1)\\ \Psi_{3}(0)\end{pmatrix}=k_{1}\begin{pmatrix}\omega^{2}\\ 1\end{pmatrix},\varphi_{2}:=\ \begin{pmatrix}\Psi_{1}(0)\\ \Psi_{3}(1)\end{pmatrix}=k_{2}\begin{pmatrix}\omega\\ 1\end{pmatrix} (9)

and

φ1ker(Tζ<)andφ2ker(Tζ>)\varphi_{1}\in\ker\left(T-\zeta^{<}\right)\ \text{and}\ \varphi_{2}\in\ker\left(T-\zeta^{>}\right)

must be satisfied. However, by direct calculation, we obtain the contradiction. ∎

From this proposition, we understand that we only need to consider eiλΛ0e^{i\lambda}\notin\Lambda_{0}. Figure 1 displays the plot of χ(λ)\chi(\lambda) for different values of θ\theta where λΛ\lambda\in\Lambda. Meanwhile, Figure 2 illustrates the probability distribution corresponding to Figure 1 for the one-defect Fourier walk.

Refer to caption
(a) θ=π12\theta=\frac{\pi}{12}
Refer to caption
(b) θ=3π12\theta=\frac{3\pi}{12}
Refer to caption
(c) θ=7π12\theta=\frac{7\pi}{12}
Refer to caption
(d) θ=11π12\theta=\frac{11\pi}{12}
Figure 1: The plot of χ(λ)\chi(\lambda) for different values of θ\theta in the one-defect Fourier walk. Only values of χ(λ)\chi(\lambda) for λΛ\lambda\in\Lambda are plotted. The red line indicate λΛ0\lambda\in\Lambda_{0}. Numerically, we can confirm that (a) has three, (b) has four, and (c) and (d) each have six eigenvalues of UU causing localization.
Refer to caption
(a) θ=π12\theta=\frac{\pi}{12}
Refer to caption
(b) θ=3π12\theta=\frac{3\pi}{12}
Refer to caption
(c) θ=7π12\theta=\frac{7\pi}{12}
Refer to caption
(d) θ=11π12\theta=\frac{11\pi}{12}
Figure 2: Probability distribution μt(Ψ0)(x)\mu_{t}^{(\Psi_{0})}(x) for the one-defect Fourier walk with different values of θ\theta at time 100100 with initial state Ψ0(0)=[13,i3,13]t\Psi_{0}(0)=[\frac{1}{\sqrt{3}},\frac{i}{\sqrt{3}},\frac{1}{\sqrt{3}}]^{t} and Ψ0(x)=𝟎\Psi_{0}(x)=\mathbf{0} for x0x\neq 0.

4.2 Two-phase model

Consider the coin matrix defined as:

Cx={Fx<0,eiθFx0.C_{x}=\begin{cases}F&x<0,\\ e^{i\theta}F&x\geq 0.\end{cases}

with some θ(0,2π]\theta\in(0,2\pi]. The following proposition can be proved similarly to Proposition 4.1.

Proposition 4.2.

For a two-phase Fourier walk,

Λ0σp(U)={a±(3,3)¯a±(1,1)eiΔ±}σp(U)=.\Lambda_{0}\cap\sigma_{p}(U)=\left\{\frac{\overline{a_{\pm\infty}^{(3,3)}}}{a_{\pm\infty}^{(1,1)}}e^{i\Delta_{\pm\infty}}\right\}\cap\sigma_{p}(U)=\emptyset.

From this proposition, we understand that we only have to consider eiλΛ0e^{i\lambda}\notin\Lambda_{0}. Figure 3 displays the plot of χ(λ)\chi(\lambda) for different values of θ\theta where λΛ\lambda\in\Lambda. Meanwhile, Figure 4 illustrates the probability distribution corresponding to Figure 3 for the two-phase Fourier walk.

Refer to caption
(a) θ=π12\theta=\frac{\pi}{12}
Refer to caption
(b) θ=3π12\theta=\frac{3\pi}{12}
Refer to caption
(c) θ=7π12\theta=\frac{7\pi}{12}
Refer to caption
(d) θ=11π12\theta=\frac{11\pi}{12}
Figure 3: The plot of χ(λ)\chi(\lambda) for different values of θ\theta in two-phase Fourier walk. Only values of χ(λ)\chi(\lambda) for λΛ\lambda\in\Lambda are plotted. The red line indicate λΛ0\lambda\in\Lambda_{0}. Numerically, we can confirm that (a) has no eigenvalue. On the other hand, (b) has one (c) has two and (d) have three eigenvalues of UU causing localization.
Refer to caption
(a) θ=π12\theta=\frac{\pi}{12}
Refer to caption
(b) θ=3π12\theta=\frac{3\pi}{12}
Refer to caption
(c) θ=7π12\theta=\frac{7\pi}{12}
Refer to caption
(d) θ=11π12\theta=\frac{11\pi}{12}
Figure 4: Probability distribution μt(Ψ0)(x)\mu_{t}^{(\Psi_{0})}(x) for the two-phase Fourier walk with different values of θ\theta at time 100100 with initial state Ψ0(0)=[13,i3,13]t\Psi_{0}(0)=[\frac{1}{\sqrt{3}},\frac{i}{\sqrt{3}},\frac{1}{\sqrt{3}}]^{t} and Ψ0(x)=𝟎\Psi_{0}(x)=\mathbf{0} for x0x\neq 0.

Acknowledgment

The author C. Kiumi was supported by JSPS KAKENHI Grant Number JP22KJ1408.

References

  • [1] T. Kitagawa, M. S. Rudner, E. Berg, E. Demler, Exploring topological phases with quantum walks, Phys. Rev. A 82 (3) (2010) 033429.
  • [2] S. Apers, A. Scarlet, Quantum fast-forwarding: Markov chains and graph property testing, Quantum Inf. Comput. 19 (3&4) (2019) 181–213.
  • [3] A. Gilyén, Y. Su, G. H. Low, N. Wiebe, Quantum singular value transformation and beyond: exponential improvements for quantum matrix arithmetics, in: Proceedings of the 51st Annual ACM SIGACT Symposium on Theory of Computing, ACM, 2019.
  • [4] S. Apers, A. Gilyén, S. Jeffery, A unified framework of quantum walk search, in: 38th International Symposium on Theoretical Aspects of Computer Science (STACS 2021), Vol. 187 of Leibniz International Proceedings in Informatics (LIPIcs), Schloss Dagstuhl – Leibniz-Zentrum für Informatik, 2021, pp. 6:1–6:13.
  • [5] N. Inui, N. Konno, Localization of multi-state quantum walk in one dimension, Phys. A: Stat. Mech. Appl. 353 (2005) 133–144.
  • [6] N. Inui, N. Konno, E. Segawa, One-dimensional three-state quantum walk, Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 72 (5 Pt 2) (2005) 056112.
  • [7] S. Falkner, S. Boettcher, Weak limit of the three-state quantum walk on the line, Phys. Rev. A 90 (1) (2014) 012307.
  • [8] M. Štefaňák, I. Bezděková, I. Jex, Limit distributions of three-state quantum walks: The role of coin eigenstates, Phys. Rev. A 90 (1) (2014) 012342.
  • [9] D. Li, M. Mc Gettrick, W. Zhang, K. Zhang, One-dimensional lazy quantum walks and occupancy rate, Chin. Phys. B 24 (5) (2015) 050305.
  • [10] T. Machida, Limit theorems of a 3-state quantum walk and its application for discrete uniform measures, Quantum Inf. Comput. (2015) 406–418.
  • [11] C. Wang, W. Wang, d. X. Lu, Limit theorem for a Time-Inhomogeneous Three-State quantum walk on the line, J. Comput. Theor. Nanosci. 12 (12) (2015) 5164–5170.
  • [12] Y.-Z. Xu, G.-D. Guo, S. Lin, One-Dimensional Three-State quantum walk with Single-Point phase defects, Int. J. Theor. Phys. 55 (9) (2016) 4060–4074.
  • [13] H. Kawai, T. Komatsu, N. Konno, Stationary measures of three-state quantum walks on the one-dimensional lattice, Yokohama Math. J. 63 (2017) 59–74.
  • [14] J. Rajendran, C. Benjamin, Playing a true parrondo’s game with a three-state coin on a quantum walk, EPL 122 (4) (2018) 40004.
  • [15] T. Endo, T. Komatsu, N. Konno, T. Terada, Stationary measure for three-state quantum walk, Quantum Inf. Comput. 19 (11&12) (2019) 901–912.
  • [16] A. Saha, S. B. Mandal, D. Saha, A. Chakrabarti, One-Dimensional lazy quantum walk in ternary system, IEEE Trans. Quant. Eng. 2 (2021) 1–12.
  • [17] C. Wang, X. Lu, W. Wang, The stationary measure of a space-inhomogeneous three-state quantum walk on the line, Quantum Inf. Process. 14 (3) (2015) 867–880.
  • [18] P. R. N. Falcão, A. R. C. Buarque, W. S. Dias, G. M. A. Almeida, M. L. Lyra, Universal dynamical scaling laws in three-state quantum walks, Phys. Rev. E 104 (2021) 054106.
  • [19] C. Kiumi, A new type of quantum walks based on decomposing quantum states, Quantum Inf. Comput. 21 (7&8) (2021) 541–556.
  • [20] T. Yamagami, E. Segawa, T. Mihana, A. Röhm, R. Horisaki, M. Naruse, Bandit algorithm driven by a classical random walk and a quantum walk, Entropy 25 (6) (2023).
  • [21] C. Kiumi, K. Saito, Strongly trapped space-inhomogeneous quantum walks in one dimension, Quantum Inf. Process. 21 (9) (2022).
  • [22] S. Endo, T. Endo, T. Komatsu, N. Konno, Eigenvalues of Two-State quantum walks induced by the hadamard walk, Entropy 22 (1) (2020).
  • [23] C. Kiumi, K. Saito, Eigenvalues of two-phase quantum walks with one defect in one dimension, Quantum Inf. Process. 20 (5) (2021).
  • [24] C. Kiumi, Localization of space-inhomogeneous three-state quantum walks, J. Phys. A: Math 55 (22) (2022) 225205.
  • [25] C. Kiumi, K. Saito, Spectral analysis of non-unitary two-phase quantum walks in one dimension (2022). arXiv:2205.11046.
  • [26] K. Saito, Periodicity for the fourier quantum walk on regular graphs, Quantum Inf. and Comput. 19 (1&2) (2019) 23–34.
  • [27] M. Asano, T. Komatsu, N. Konno, A. Narimatsu, The fourier and grover walks on the two-dimensional lattice and torus, Yokohama Math. J. 65 (2019) 13–32.
  • [28] A. Narimatsu, Localization does not occur for the fourier walk on the multi-dimensional lattice, Quantum Inf. Comput. 21 (5&6) (2021) 387–394.
  • [29] C. Kiumi, Localization in quantum walks with periodically arranged coin matrices, Int. J. Quantum Inf. 20 (05) (2022).
  • [30] C. K. Ko, E. Segawa, H. J. Yoo, One-dimensional three-state quantum walks: Weak limits and localization, Infin. Dimens. Anal. Quantum Probab. Relat. Top. 19 (04) (2016) 1650025.