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

Generalized eigenfunctions for quantum walks via path counting approach

Takashi Komatsu Department of Bioengineering, School of Engineering,The University of Tokyo, Bunkyo-ku, Tokyo, 113-8656, Japan komatsu@coi.t.u-tokyo.ac.jp Norio Konno Department of Applied Mathematics, Yokohama National University, Hodogaya, Yokohama, Kanagawa, 240-8501, Japan konno-norio-bt@ynu.ac.jp Hisashi Morioka Graduate School of Science and Engineering, Ehime University, Bunkyo-cho 3, Matsuyama, Ehime, 790-8577, Japan morioka@cs.ehime-u.ac.jp  and  Etsuo Segawa Graduate School of Environment Information Sciences, Yokohama National University, Hodogaya, Yokohama, Kanagawa, 240-8501, Japan segawa-etsuo-tb@ynu.ac.jp
Abstract.

We consider the time-independent scattering theory for time evolution operators of one-dimensional two-state quantum walks. The scattering matrix associated with the position-dependent quantum walk naturally appears in the asymptotic behavior at spatial infinity of generalized eigenfunctions. The asymptotic behavior of generalized eigenfunctions is a consequence of an explicit expression of the Green function associated with the free quantum walk. When the position-dependent quantum walk is a finite rank perturbation of the free quantum walk, we derive a kind of combinatorial constructions of the scattering matrix by counting paths of quantum walkers. We also mention some remarks on the tunneling effect.

Key words and phrases:
quantum walk, scattering matrix, tunneling effect
2010 Mathematics Subject Classification:
Primary 81U20, Secondary 47A40

1. Introduction

Discrete time quantum walks (QWs for short) have been investigated in both finite and infinite systems. For finite systems, the effectiveness and universality of quantum walks in the quantum search algorithms have been studied (see [2], [4], [20] and its references therein). For infinite systems, there are several mathematical works obtaining limiting behaviors different from those of classical random walks (see [12]). Recently, spectral and scattering theory for QWs have been intensively studied. For example, see [5], [6], [17], [15], [16], [18], [19], [21], [22], [23].

In this paper, we consider one-dimensional position-dependent QWs in view of the scattering theory. In our previous works [18] and [19], the method of our study was based on the scattering theory of quantum mechanics like Schrödinger equations. For general information of this research area, the monograph by Yafaev [24] is available and its reference is also worthwhile. In order to construct a generalized eigenfunction of the time evolution operator of QWs, we use the time-independent scattering theory of quantum mechanics. We also adopt another approach in the latter half of this paper. Namely, we introduce a combinatorial construction of the scattering matrix associated with perturbations of finite rank. A primitive form of this method can be seen in [7]. In our construction, we relate the generalized eigenfunction with the time evolution of the QW, and we can compute the scattering matrix by counting paths of quantum walkers. For Schrödinger operators with finite rank perturbations, the representation of the scattering matrix has been derived by [14].

Let us introduce our model of discrete time QWs. In the following, 𝐙{\bf Z}, 𝐑{\bf R} and 𝐂{\bf C} denote the set of integers, the set of real numbers and the set of complex numbers, respectively. We focus on two-state QWs on 𝐙{\bf Z}. Let ψ={ψ(x)}x𝐙\psi=\{\psi(x)\}_{x\in{\bf Z}} be a 𝐂2{\bf C}^{2}-valued sequence on 𝐙{\bf Z}. In the following, we denote the column vector ψ(x)\psi(x) by ψ(x)=[ψL(x),ψR(x)]𝖳\psi(x)=[\psi_{L}(x),\psi_{R}(x)]^{\mathsf{T}} for 𝐂{\bf C}-valued sequences ψL\psi_{L} and ψR\psi_{R} on 𝐙{\bf Z}. Here [a,b]𝖳[a,b]^{\mathsf{T}} means the transpose of the row vector [a,b][a,b]. First of all, we define the homogeneous QW. We define the operator U0U_{0} by

(U0ψ)(x)=P0ψ(x+1)+Q0ψ(x1),(U_{0}\psi)(x)=P_{0}\psi(x+1)+Q_{0}\psi(x-1),

where the matrices P0P_{0} and Q0Q_{0} are given by

P0=eiγ/2[pei(αγ/2)qei(βγ/2)00],\displaystyle P_{0}=e^{i\gamma/2}\left[\begin{array}[]{cc}pe^{i(\alpha-\gamma/2)}&qe^{i(\beta-\gamma/2)}\\ 0&0\end{array}\right],
Q0=eiγ/2[00qei(βγ/2)pei(αγ/2)],\displaystyle Q_{0}=e^{i\gamma/2}\left[\begin{array}[]{cc}0&0\\ -qe^{-i(\beta-\gamma/2)}&pe^{-i(\alpha-\gamma/2)}\end{array}\right],

for p(0,1]p\in(0,1], q[0,1)q\in[0,1), p2+q2=1p^{2}+q^{2}=1 and α,β,γ𝐓:=𝐑/2π𝐙\alpha,\beta,\gamma\in{\bf T}:={\bf R}/2\pi{\bf Z}. We have C0:=P0+Q0U(2)C_{0}:=P_{0}+Q_{0}\in\mathrm{U}(2). The operator U0U_{0} is unitary on 2(𝐙;𝐂2)\ell^{2}({\bf Z};{\bf C}^{2}). Note that we can write U0=SC0U_{0}=SC_{0} for the shift operator

(Sψ)(x)=[ψL(x+1)ψR(x1)].(S\psi)(x)=\left[\begin{array}[]{c}\psi_{L}(x+1)\\ \psi_{R}(x-1)\end{array}\right].

In Sections 3 and 4, we consider the free QW defined by U0=SU_{0}=S. This is a special case of homogeneous QWs such that C0C_{0} is the 2×22\times 2 identity matrix.

The position-dependent QW defined by the perturbed operator U=SCU=SC is given by the operator CC of multiplication by the matrix C(x)U(2)C(x)\in\mathrm{U}(2) for every x𝐙x\in{\bf Z}. Thus the operator UU is also unitary in 2(𝐙;𝐂2)\ell^{2}({\bf Z};{\bf C}^{2}). The time evolutions of the homogeneous QW and the position-dependent QW are given by

Ψ(t,)=Utψ,Ψ(0)(t,)=U0tψ,t𝐙,\Psi(t,\cdot)=U^{t}\psi,\quad\Psi^{(0)}(t,\cdot)=U_{0}^{t}\psi,\quad t\in{\bf Z},

for an initial state ψ\psi. Note that these time evolution preserve the norm of the initial state in 2(𝐙;𝐂2)\ell^{2}({\bf Z};{\bf C}^{2}).

In this paper, we assume that the following assumptions hold :

(A-1) There exist constants ϵ0,c0>0\epsilon_{0},c_{0}>0 such that

C(x)C0c0(1+|x|)1ϵ0,x𝐙,\|C(x)-C_{0}\|_{\infty}\leq c_{0}(1+|x|)^{-1-\epsilon_{0}},\quad x\in{\bf Z},

where \|\cdot\|_{\infty} is the maximum norm for 2 by 2 matrices.

(A-2) For all x𝐙x\in{\bf Z}, C(x)C(x) is not anti-diagonal.

In Section 4, the assumption (A-1) will be replaced by “UU0U-U_{0} is an operator of finite rank”.

In this paper, we consider the time-independent scattering theory for UU and U0U_{0}. In particular, we focus on the scattering matrix associated with the wave operators

W±=s-limt±UtU0tin2(𝐙;𝐂2).W_{\pm}={\mathop{{\rm s\text{-}lim}}_{t\to\pm\infty}}\,U^{-t}U_{0}^{t}\quad\text{in}\quad\ell^{2}({\bf Z};{\bf C}^{2}).

Existence and completeness of W±W_{\pm} for UU and U0U_{0} have already been proven by Suzuki [22]. Note that the assumption (A-1) is the well-known short-range condition which guarantees the existence and the asymptotic completeness of the wave operators.

Theorem 1.1 (Suzuki [22]).

The wave operators exist and are complete i.e. the ranges of W±W_{\pm} coincide with the absolutely continuous subspace ac(U)\mathcal{H}_{ac}(U) of 2(𝐙;𝐂2)\ell^{2}({\bf Z};{\bf C}^{2}) for UU. Precisely, for any ϕac(U)\phi\in\mathcal{H}_{ac}(U), there exist ψ±2(𝐙;𝐂2)\psi_{\pm}\in\ell^{2}({\bf Z};{\bf C}^{2}) such that U0tψ±Utϕ2(𝐙;𝐂2)0\|U_{0}^{t}\psi_{\pm}-U^{t}\phi\|_{\ell^{2}({\bf Z};{\bf C}^{2})}\to 0 as t±t\to\pm\infty. The wave operators are unitary on 2(𝐙;𝐂2)\ell^{2}({\bf Z};{\bf C}^{2}) and we have W±=W±1W_{\pm}^{*}=W_{\pm}^{-1}.

The assumption (A-2) guarantees the penetrability of barriers given by C(x)C(x). In fact, if C(x)C(x) is anti-diagonal at a point x=x0x=x_{0}, a quantum walker is reflected at x0x_{0}.

The scattering operator is defined by

Σ=W+W:ψψ+.\Sigma=W_{+}^{*}W_{-}:\psi_{-}\mapsto\psi_{+}.

Its Fourier transform Σ^\widehat{\Sigma} is decomposed by the scattering matrix :

Σ^=JγΣ^(θ)dθ,\widehat{\Sigma}=\int_{J_{\gamma}}\oplus\widehat{\Sigma}(\theta)d\theta,

where JγJ_{\gamma} is the interval defined in Lemma 2.1 (see also [21] or [18]). The scattering matrix (S-matrix for short) Σ^(θ)\widehat{\Sigma}(\theta) naturally appears in the generalized eigenfunction for UU. One of the purposes of this paper is to derive the asymptotic behavior as x±x\to\pm\infty of the generalized eigenfunction in (𝐙;𝐂2)\ell^{\infty}({\bf Z};{\bf C}^{2}). The S-matrix can be represented by the distorted Fourier transformation associated with UU and U0U_{0}. The distorted Fourier transformation is constructed as the spectral decomposition of unitary operators.

We also mention stationary measures of QWs as a related topic. Generalized eigenfunctions of UU give stationary measures for UU. Note that this is a special kind of stationary measures. For the topic of stationary measures, see e.g. Konno [12], Konno-Takei [13], Komatsu-Konno [10], Kawai et al. [9].

The plan of this paper is as follows. In Section 2, we consider the Green function of the homogeneous QW. Green functions have been derived in the momentum space by [18]. In this paper, we derive an explicit formula of Green functions on 𝐙{\bf Z}. In Section 3, we construct the generalized eigenfunction for UU in (𝐙;𝐂2)\ell^{\infty}({\bf Z};{\bf C}^{2}). The optimal functional space in which there exist generalized eigenfunctions is characterized by Agmon-Hörmander’s \mathcal{B}-\mathcal{B}^{*} spaces ([1]). By using the formula of the Green function, we obtain the asymptotic behavior at infinity of generalized eigenfunctions. In Section 4, instead of general theory of the time-independent scattering theory, we introduce a combinatorial construction of the S-matrix when UU0U-U_{0} is an operator of finite rank. The scattering matrix is computed by a finite rank submatrix induced from the operator UU. We also mention the resonant-tunneling effect (see also [17]) here. Finally, we prove that incident waves always pass through penetrable barriers. Some remarks on complex contour integrals are gathered in Appendix A.

1.1. Notation

The notation in this paper is as follows. 𝐓:=𝐑/2π𝐙{\bf T}:={\bf R}/2\pi{\bf Z} denotes the flat torus. Similarly, the complex flat torus is denoted by 𝐓𝐂=𝐂/2π𝐙={x+iy;x𝐓,y𝐑}{\bf T}_{{\bf C}}={\bf C}/2\pi{\bf Z}=\{x+iy\ ;\ x\in{\bf T},y\in{\bf R}\}. We often use the identifications 𝐓=[π,π){\bf T}=[-\pi,\pi) or [0,2π)[0,2\pi) under 2π2\pi-periodicity. 𝐓𝐂{\bf T}_{{\bf C}} appears in Appendix A. For a vector 𝐯𝐂2{\bf v}\in{\bf C}^{2}, |𝐯||{\bf v}| denotes the norm in 𝐂2{\bf C}^{2}. For two vectors 𝐯,𝐰𝐂2{\bf v},{\bf w}\in{\bf C}^{2}, we denote by 𝐯,𝐰\langle{\bf v},{\bf w}\rangle the inner product in 𝐂2{\bf C}^{2}. For a 𝐂2{\bf C}^{2}-valued sequence f={f(x)}x𝐙f=\{f(x)\}_{x\in{\bf Z}}, we define the mapping 𝒰\mathcal{U} by the Fourier transformation

f^(ξ):=(𝒰f)(ξ)=12πx𝐙eixξf(x),ξ𝐓.\widehat{f}(\xi):=(\mathcal{U}f)(\xi)=\frac{1}{\sqrt{2\pi}}\sum_{x\in{\bf Z}}e^{-ix\xi}f(x),\quad\xi\in{\bf T}.

The Fourier coefficients of g^(ξ)\widehat{g}(\xi) on 𝐓{\bf T} is given by

g(x)=(𝒰g^)(x)=12π𝐓eixξg^(ξ)𝑑ξ,x𝐙.g(x)=(\mathcal{U}^{*}\widehat{g})(x)=\frac{1}{\sqrt{2\pi}}\int_{{\bf T}}e^{ix\xi}\widehat{g}(\xi)d\xi,\quad x\in{\bf Z}.

Then 𝒰\mathcal{U} is a unitary transformation from 2(𝐙;𝐂2)\ell^{2}({\bf Z};{\bf C}^{2}) to L2(𝐓;𝐂2)L^{2}({\bf T};{\bf C}^{2}). Here 2(𝐙;𝐂2)\ell^{2}({\bf Z};{\bf C}^{2}) and L2(𝐓;𝐂2)L^{2}({\bf T};{\bf C}^{2}) are equipped with its inner products and norms

(f,g)2(𝐙;𝐂2)=x𝐙f(x),g(x),f2(𝐙;𝐂2)=(f,f)2(𝐙;𝐂2),(f,g)_{\ell^{2}({\bf Z};{\bf C}^{2})}=\sum_{x\in{\bf Z}}\langle f(x),g(x)\rangle,\quad\|f\|_{\ell^{2}({\bf Z};{\bf C}^{2})}=\sqrt{(f,f)_{\ell^{2}({\bf Z};{\bf C}^{2})}},
(f^,g^)L2(𝐓;𝐂2)=𝐓f^(ξ),g^(ξ)𝑑ξ,f^L2(𝐓;𝐂2)=(f^,f^)L2(𝐓;𝐂2).(\widehat{f},\widehat{g})_{L^{2}({\bf T};{\bf C}^{2})}=\int_{{\bf T}}\langle\widehat{f}(\xi),\widehat{g}(\xi)\rangle d\xi,\quad\|\widehat{f}\|_{L^{2}({\bf T};{\bf C}^{2})}=\sqrt{(\widehat{f},\widehat{f})_{L^{2}({\bf T};{\bf C}^{2})}}.

For Banach spaces XX and YY, 𝐁(X;Y){\bf B}(X;Y) denotes the space of bounded linear operators from XX to YY.

2. Green function

2.1. Resolvent for the homogeneous QW

The spectral theory for UU and U0U_{0} is studied in [21], [18], [19], [16]. Here let us list some basic results.

Let

U^0=𝒰U0𝒰.\widehat{U}_{0}=\mathcal{U}U_{0}\mathcal{U}^{*}.

It follows that U^0\widehat{U}_{0} is the operator of multiplication on 𝐓{\bf T} by the unitary matrix

U^0(ξ)=eiγ/2[pei(αγ/2)eiξqei(βγ/2)eiξqei(βγ/2)eiξpei(αγ/2)eiξ],ξ𝐓.\widehat{U}_{0}(\xi)=e^{i\gamma/2}\left[\begin{array}[]{cc}pe^{i(\alpha-\gamma/2)}e^{i\xi}&qe^{i(\beta-\gamma/2)}e^{i\xi}\\ -qe^{-i(\beta-\gamma/2)}e^{-i\xi}&pe^{-i(\alpha-\gamma/2)}e^{-i\xi}\end{array}\right],\quad\xi\in{\bf T}.

For any κ𝐂\kappa\in{\bf C}, we have

p(ξ,κ):=det(U^0(ξ)eiκ)= 2pei(κ+γ/2)(cos(ξ+αγ2)+1pcos(κγ2)).\displaystyle\begin{split}p(\xi,\kappa):=&\,\mathrm{det}(\widehat{U}_{0}(\xi)-e^{i\kappa})\\ =&\,2pe^{i(\kappa+\gamma/2)}\left(-\cos\left(\xi+\alpha-\frac{\gamma}{2}\right)+\frac{1}{p}\cos\left(\kappa-\frac{\gamma}{2}\right)\right).\end{split}

By using p(ξ,κ)p(\xi,\kappa), we can evaluate the continuous spectrum of U0U_{0}. Moreover, Weyl’s singular sequence method shows the essential spectrum of UU. For a unitary operator AA, let σ(A)\sigma(A) denote the spectrum of AA. We denote by σess(A)\sigma_{ess}(A), σd(A)\sigma_{d}(A) the essential spectrum and the discrete spectrum of AA. By using the spectral decomposition of AA, we can also define the absolutely continuous spectrum σac(A)\sigma_{ac}(A), the point spectrum σp(A)\sigma_{p}(A), and the singular continuous spectrum σsc(A)\sigma_{sc}(A). See [21], [18], or [19] for more details of definitions. Note that the absence of singular continuous spectrum was first proved by Asch et al. [3].

Lemma 2.1.

Following assertions hold.

  1. (1)

    Let Jγ=Jγ,1Jγ,2J_{\gamma}=J_{\gamma,1}\cup J_{\gamma,2} where

    Jγ,1=[arccosp+γ/2,πarccosp+γ/2],Jγ,2=[π+arccosp+γ/2,2πarccosp+γ/2].\displaystyle\begin{split}&J_{\gamma,1}=[\arccos p+\gamma/2,\pi-\arccos p+\gamma/2],\\ &J_{\gamma,2}=[\pi+\arccos p+\gamma/2,2\pi-\arccos p+\gamma/2].\end{split}

    Then we have σ(U0)=σac(U0)={eiθ;θJγ}\sigma(U_{0})=\sigma_{ac}(U_{0})=\{e^{i\theta}\ ;\ \theta\in J_{\gamma}\}.

  2. (2)

    We have σess(U)=σess(U0)={eiθ;θJγ}\sigma_{ess}(U)=\sigma_{ess}(U_{0})=\{e^{i\theta}\ ;\ \theta\in J_{\gamma}\} and σsc(U)=\sigma_{sc}(U)=\emptyset.

  3. (3)

    Let 𝒯={eiθ;θJγ,𝒯}\mathcal{T}=\{e^{i\theta}\ ;\ \theta\in J_{\gamma,\mathcal{T}}\} where

    Jγ,𝒯={arccosp+γ/2,πarccosp+γ/2,π+arccosp+γ/2, 2πarccosp+γ/2}.\displaystyle\begin{split}J_{\gamma,\mathcal{T}}=\left\{\begin{split}&\arccos p+\gamma/2,\ \pi-\arccos p+\gamma/2,\\ &\pi+\arccos p+\gamma/2,\ 2\pi-\arccos p+\gamma/2\end{split}\right\}.\end{split}

    Then there is no eigenvalue of UU in σess(U)𝒯\sigma_{ess}(U)\setminus\mathcal{T}.

  4. (4)

    For eiθ𝒯e^{i\theta}\in\mathcal{T}, the zeros of p(ξ,θ)p(\xi,\theta) are degenerate i.e. pξ(ξ,θ)=0\frac{\partial p}{\partial\xi}(\xi,\theta)=0 if p(ξ,θ)=0p(\xi,\theta)=0. For eiθσ(U0)𝒯e^{i\theta}\in\sigma(U_{0})\setminus\mathcal{T}, the zeros of p(ξ,θ)p(\xi,\theta) are non-degenerate i.e. pξ(ξ,θ)0\frac{\partial p}{\partial\xi}(\xi,\theta)\not=0 if p(ξ,θ)=0p(\xi,\theta)=0.

Let us turn to the resolvent R0(κ)=(U0eiκ)1R_{0}(\kappa)=(U_{0}-e^{i\kappa})^{-1} for κ𝐂𝐑\kappa\in{\bf C}\setminus{\bf R}. The operator R^0(κ)=𝒰R0(κ)𝒰\widehat{R}_{0}(\kappa)=\mathcal{U}R_{0}(\kappa)\mathcal{U}^{*} is the operator of multiplication on 𝐓{\bf T} by the matrix

(2.1) R^0(ξ,κ)=1p(ξ,κ)[pei(αγ)eiξeiκqeiβeiξqei(βγ)eiξpeiαeiξeiκ].\widehat{R}_{0}(\xi,\kappa)=\frac{1}{p(\xi,\kappa)}\left[\begin{array}[]{cc}pe^{-i(\alpha-\gamma)}e^{-i\xi}-e^{i\kappa}&-qe^{i\beta}e^{i\xi}\\ qe^{-i(\beta-\gamma)}e^{-i\xi}&pe^{i\alpha}e^{i\xi}-e^{i\kappa}\end{array}\right].

In [18], we have proven the limiting absorption principle of R0(κ)R_{0}(\kappa) as follows. Agmon-Hörmander’s \mathcal{B}-\mathcal{B}^{*} spaces ([1]) are often used in the context of the limiting absorption principle. The functional spaces (𝐙)\mathcal{B}({\bf Z}) and (𝐙)\mathcal{B}^{*}({\bf Z}) of 𝐂2{\bf C}^{2}-valued sequences on 𝐙{\bf Z} are defined by the norms

f(𝐙)=j=0rj1/2(rj1|x|<rj|f(x)|2)1/2,\|f\|_{\mathcal{B}({\bf Z})}=\sum_{j=0}^{\infty}r_{j}^{1/2}\left(\sum_{r_{j-1}\leq|x|<r_{j}}|f(x)|^{2}\right)^{1/2},

for r1=0r_{-1}=0, rj=2jr_{j}=2^{j}, j0j\geq 0, and

u(𝐙)2=supR>11R|x|<R|u(x)|2.\|u\|_{\mathcal{B}^{*}({\bf Z})}^{2}=\sup_{R>1}\frac{1}{R}\sum_{|x|<R}|u(x)|^{2}.

Note that the inclusion relation

(𝐙)2(𝐙;𝐂2)(𝐙)\mathcal{B}({\bf Z})\subset\ell^{2}({\bf Z};{\bf C}^{2})\subset\mathcal{B}^{*}({\bf Z})

holds.

Lemma 2.2.

Following assertions hold.

  1. (1)

    For θJγJγ,𝒯\theta\in J_{\gamma}\setminus J_{\gamma,\mathcal{T}} and f(𝐙)f\in\mathcal{B}({\bf Z}), there exist weak-* limits R0(θ±i0)f:=limϵ0R0(θilog(1ϵ))fR_{0}(\theta\pm i0)f:=\lim_{\epsilon\downarrow 0}R_{0}(\theta-i\log(1\mp\epsilon))f in the sense

    limϵ0(R0(θilog(1ϵ))f,g)=(R0(θ±i0)f,g),g(𝐙).\lim_{\epsilon\downarrow 0}(R_{0}(\theta-i\log(1\mp\epsilon))f,g)=(R_{0}(\theta\pm i0)f,g),\quad g\in\mathcal{B}({\bf Z}).
  2. (2)

    R0(θ±i0)fR_{0}(\theta\pm i0)f for f(𝐙)f\in\mathcal{B}({\bf Z}) satisfy R0(θ±i0)f(𝐙)cf(𝐙)\|R_{0}(\theta\pm i0)f\|_{\mathcal{B}^{*}({\bf Z})}\leq c\|f\|_{\mathcal{B}({\bf Z})} for a constant c>0c>0 which is independent of θ\theta if θ\theta varies over a compact interval in JγJγ,𝒯J_{\gamma}\setminus J_{\gamma,\mathcal{T}}.

  3. (3)

    The mappings JγJγ,𝒯θ(R0(θ±i0)f,g)J_{\gamma}\setminus J_{\gamma,\mathcal{T}}\ni\theta\mapsto(R_{0}(\theta\pm i0)f,g) are continuous for f,g(𝐙)f,g\in\mathcal{B}({\bf Z}).

2.2. Green function

In [18], we also derived the radiation condition which guarantees the uniqueness of the solution to the equation

(U0eiθ)u=fon𝐙,θJγJγ,𝒯,(U_{0}-e^{i\theta})u=f\quad\text{on}\quad{\bf Z},\quad\theta\in J_{\gamma}\setminus J_{\gamma,\mathcal{T}},

in (𝐙)\mathcal{B}^{*}({\bf Z}) for f(𝐙)f\in\mathcal{B}({\bf Z}). However, our result was proven in the momentum space i.e. ξ\xi-space. In the following, we consider this equation in the physical space i.e. xx-space.

For κ𝐂𝐑\kappa\in{\bf C}\setminus{\bf R}, we consider the equation

(2.2) (U0eiκ)u=fon𝐙,(U_{0}-e^{i\kappa})u=f\quad\text{on}\quad{\bf Z},

for f(𝐙)f\in\mathcal{B}({\bf Z}). Since eiκσ(U0)e^{i\kappa}\not\in\sigma(U_{0}), the resolvent R0(κ)R_{0}(\kappa) exists in 2(𝐙;𝐂2)\ell^{2}({\bf Z};{\bf C}^{2}). We seek the solution to the equation (2.2) of the form

(2.3) u(x)=y𝐙G0(xy,κ)f(y),\displaystyle u(x)=\sum_{y\in{\bf Z}}G_{0}(x-y,\kappa)f(y),

where the kernel G0(x,κ)G_{0}(x,\kappa) is the matrix

G0(x,κ)=[r11(x,κ)r12(x,κ)r21(x,κ)r22(x,κ)].G_{0}(x,\kappa)=\left[\begin{array}[]{cc}r_{11}(x,\kappa)&r_{12}(x,\kappa)\\ r_{21}(x,\kappa)&r_{22}(x,\kappa)\end{array}\right].

The following lemma is a consequence of the representation (2.3).

Lemma 2.3.

Let δ(L)=[δ,0]𝖳\delta^{(L)}=[\delta,0]^{\mathsf{T}} and δ(R)=[0,δ]𝖳\delta^{(R)}=[0,\delta]^{\mathsf{T}} for δ={δx0}x𝐙\delta=\{\delta_{x0}\}_{x\in{\bf Z}} where δx0\delta_{x0} is the Kronecker delta. The solutions u(L)u^{(L)} and u(R)u^{(R)} to the equations

(2.4) (U0eiκ)u(L)=δ(L),(U0eiκ)u(R)=δ(R),(U_{0}-e^{i\kappa})u^{(L)}=\delta^{(L)},\quad(U_{0}-e^{i\kappa})u^{(R)}=\delta^{(R)},

are given by

(2.5) u(L)(x)=[r11(x,κ)r21(x,κ)],u(R)(x)=[r12(x,κ)r22(x,κ)].u^{(L)}(x)=\left[\begin{array}[]{c}r_{11}(x,\kappa)\\ r_{21}(x,\kappa)\end{array}\right],\quad u^{(R)}(x)=\left[\begin{array}[]{c}r_{12}(x,\kappa)\\ r_{22}(x,\kappa)\end{array}\right].

Now let us compute an explicit expression of G0(x,κ)G_{0}(x,\kappa) by using Lemma 2.5. In view of (2.4), we have

(U^0(ξ)eiκ)u^(L)(ξ)=12π[10],(U^0(ξ)eiκ)u^(R)(ξ)=12π[01].(\widehat{U}_{0}(\xi)-e^{i\kappa})\widehat{u}^{(L)}(\xi)=\frac{1}{\sqrt{2\pi}}\left[\begin{array}[]{c}1\\ 0\end{array}\right],\quad(\widehat{U}_{0}(\xi)-e^{i\kappa})\widehat{u}^{(R)}(\xi)=\frac{1}{\sqrt{2\pi}}\left[\begin{array}[]{c}0\\ 1\end{array}\right].

Due to (2.1), we have

(2.6) r11(x,κ)=12π𝐓eixξ(pei(ξ+αγ)eiκ)p(ξ,κ)𝑑ξ,\displaystyle r_{11}(x,\kappa)=\frac{1}{2\pi}\int_{{\bf T}}\frac{e^{ix\xi}(pe^{-i(\xi+\alpha-\gamma)}-e^{i\kappa})}{p(\xi,\kappa)}d\xi,
(2.7) r21(x,κ)=12π𝐓qeixξei(ξ+βγ)p(ξ,κ)𝑑ξ,\displaystyle r_{21}(x,\kappa)=\frac{1}{2\pi}\int_{{\bf T}}\frac{qe^{ix\xi}e^{-i(\xi+\beta-\gamma)}}{p(\xi,\kappa)}d\xi,

and

(2.8) r12(x,κ)=12π𝐓qeixξei(ξ+β)p(ξ,κ)𝑑ξ,\displaystyle r_{12}(x,\kappa)=-\frac{1}{2\pi}\int_{{\bf T}}\frac{qe^{ix\xi}e^{i(\xi+\beta)}}{p(\xi,\kappa)}d\xi,
(2.9) r22(x,κ)=12π𝐓eixξ(pei(ξ+α)eiκ)p(ξ,κ)𝑑ξ.\displaystyle r_{22}(x,\kappa)=\frac{1}{2\pi}\int_{{\bf T}}\frac{e^{ix\xi}(pe^{i(\xi+\alpha)}-e^{i\kappa})}{p(\xi,\kappa)}d\xi.

We compute the integrals (2.6)-(2.9) by using the result in Appendix A. In fact, we have

(2.10) r11(x,κ)=12πeix(αγ/2)2pei(κ+γ/2)(peiγ/2I(x1,κ)eiκI(x,κ)),\displaystyle r_{11}(x,\kappa)=\frac{1}{2\pi}\frac{e^{-ix(\alpha-\gamma/2)}}{2pe^{i(\kappa+\gamma/2)}}\left(pe^{i\gamma/2}I(x-1,\kappa)-e^{i\kappa}I(x,\kappa)\right),
(2.11) r21(x,κ)=12πqei(βγ)2pei(κ+γ/2)I(x1,κ),\displaystyle r_{21}(x,\kappa)=\frac{1}{2\pi}\frac{qe^{-i(\beta-\gamma)}}{2pe^{i(\kappa+\gamma/2)}}I(x-1,\kappa),

and

(2.12) r12(x,κ)=12πqeiβei(x+1)(α+γ/2)2pei(κ+γ/2)I(x+1,κ),\displaystyle r_{12}(x,\kappa)=-\frac{1}{2\pi}\frac{qe^{i\beta}e^{i(x+1)(-\alpha+\gamma/2)}}{2pe^{i(\kappa+\gamma/2)}}I(x+1,\kappa),
(2.13) r22(x,κ)=12πeix(αγ/2)2pei(κ+γ/2)(peiγ/2I(x+1,κ)eiκI(x,κ)),\displaystyle r_{22}(x,\kappa)=\frac{1}{2\pi}\frac{e^{-ix(\alpha-\gamma/2)}}{2pe^{i(\kappa+\gamma/2)}}\left(pe^{i\gamma/2}I(x+1,\kappa)-e^{i\kappa}I(x,\kappa)\right),

where I(x,κ)I(x,\kappa) denotes the integral (A.1). In view of Lemma A.5, we can derive the formula of G0(x,κ)G_{0}(x,\kappa).

Lemma 2.4.

For κ𝐂𝐑\kappa\in{\bf C}\setminus{\bf R}, we have

(2.14) r11(x,κ)=±ieix(α+γ/2)(pe±i|x1|ζ(κ)ei(κγ/2)e±i|x|ζ(κ))2peiκsinζ(κ),\displaystyle r_{11}(x,\kappa)=\pm\frac{ie^{ix(-\alpha+\gamma/2)}(pe^{\pm i|x-1|\zeta(\kappa)}-e^{i(\kappa-\gamma/2)}e^{\pm i|x|\zeta(\kappa)})}{2pe^{i\kappa}\sin\zeta(\kappa)},
(2.15) r21(x,κ)=±iqei(αβ)eix(α+γ/2)e±i|x1|ζ(κ)2peiκsinζ(κ),\displaystyle r_{21}(x,\kappa)=\pm\frac{iqe^{i(\alpha-\beta)}e^{ix(-\alpha+\gamma/2)}e^{\pm i|x-1|\zeta(\kappa)}}{2pe^{i\kappa}\sin\zeta(\kappa)},

and

(2.16) r12(x,κ)=iqei(α+β)eix(α+γ/2)e±i|x+1|ζ(κ)2peiκsinζ(κ),\displaystyle r_{12}(x,\kappa)=\mp\frac{iqe^{i(-\alpha+\beta)}e^{ix(-\alpha+\gamma/2)}e^{\pm i|x+1|\zeta(\kappa)}}{2pe^{i\kappa}\sin\zeta(\kappa)},
(2.17) r22(x,κ)=±ieix(α+γ/2)(pe±i|x+1|ζ(κ)ei(κγ/2)e±i|x|ζ(κ))2peiκsinζ(κ),\displaystyle r_{22}(x,\kappa)=\pm\frac{ie^{ix(-\alpha+\gamma/2)}(pe^{\pm i|x+1|\zeta(\kappa)}-e^{i(\kappa-\gamma/2)}e^{\pm i|x|\zeta(\kappa)})}{2pe^{i\kappa}\sin\zeta(\kappa)},

for ±Imζ(κ)>0\pm\mathrm{Im}\,\zeta(\kappa)>0 respectively where ζ(κ)\zeta(\kappa) is defined by (A.2).

The formula of G0(x,κ)G_{0}(x,\kappa) is not so simple. In the following, we consider the case where U0U_{0} is the free QW for the sake of simplicity. In this case, we have J0,1=[0,π]J_{0,1}=[0,\pi], J0,2=[π,2π]J_{0,2}=[\pi,2\pi], J0,𝒯={0,π,2π}J_{0,\mathcal{T}}=\{0,\pi,2\pi\}. We can take ζ(κ)=κ\zeta(\kappa)=\kappa for 0<Reκ<π0<\mathrm{Re}\,\kappa<\pi and ζ(κ)=2πκ\zeta(\kappa)=2\pi-\kappa for π<Reκ<2π\pi<\mathrm{Re}\,\kappa<2\pi. Taking κ=κ±=θilog(1ϵ)\kappa=\kappa_{\pm}=\theta-i\log(1\mp\epsilon) for θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}} and ϵ>0\epsilon>0, we have

r11(x,κ±)=±i(e±iκ±|x1|e±iκ±(|x|±1))2eiκ±sinκ±,\displaystyle r_{11}(x,\kappa_{\pm})=\pm\frac{i(e^{\pm i\kappa_{\pm}|x-1|}-e^{\pm i\kappa_{\pm}(|x|\pm 1)})}{2e^{i\kappa_{\pm}}\sin\kappa_{\pm}},
r22(x,κ±)=±i(e±iκ±|x+1|e±iκ±(|x|±1))2eiκ±sinκ±,\displaystyle r_{22}(x,\kappa_{\pm})=\pm\frac{i(e^{\pm i\kappa_{\pm}|x+1|}-e^{\pm i\kappa_{\pm}(|x|\pm 1)})}{2e^{i\kappa_{\pm}}\sin\kappa_{\pm}},
r12(x,κ±)=r21(x,κ±)=0.\displaystyle r_{12}(x,\kappa_{\pm})=r_{21}(x,\kappa_{\pm})=0.

More precisely, we obtain the following explicit expression of G0(x,κ±)G_{0}(x,\kappa_{\pm}).

Lemma 2.5.

Suppose U0=SU_{0}=S. Let F(x)=1F(x)=1 for x0x\geq 0 and F(x)=0F(x)=0 for x1x\leq-1. The corresponding diagonal kernel G0(x,κ±)G_{0}(x,\kappa_{\pm}) is given by

(2.18) r11(x,κ+)=F(x1)eiκ+(x1),r22(x,κ+)=F(x1)eiκ+(x+1),\displaystyle r_{11}(x,\kappa_{+})=F(x-1)e^{i\kappa_{+}(x-1)},\quad r_{22}(x,\kappa_{+})=F(-x-1)e^{-i\kappa_{+}(x+1)},
(2.19) r11(x,κ)=F(x)eiκ+(x1),r22(x,κ)=F(x)eiκ(x+1),\displaystyle r_{11}(x,\kappa_{-})=-F(-x)e^{i\kappa_{+}(x-1)},\quad r_{22}(x,\kappa_{-})=-F(x)e^{-i\kappa_{-}(x+1)},

for θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}.

Remark. We can obtain Lemma 2.5 directly from

r11(x,κ)=12πππeixξeiξeiκ𝑑ξ,r22(x,κ)=12πππeixξeiξeiκ𝑑ξ,r_{11}(x,\kappa)=\frac{1}{2\pi}\int_{-\pi}^{\pi}\frac{e^{ix\xi}}{e^{i\xi}-e^{i\kappa}}d\xi,\quad r_{22}(x,\kappa)=\frac{1}{2\pi}\int_{-\pi}^{\pi}\frac{e^{ix\xi}}{e^{-i\xi}-e^{i\kappa}}d\xi,

by using the complex contour integral.

The formulas (2.18)-(2.19) hold for G0(x,θ±i0)G_{0}(x,\theta\pm i0), θ(J0,1J0,2)J0,𝒯\theta\in(J_{0,1}\cup J_{0,2})\setminus J_{0,\mathcal{T}}. We obtain the formulas

(2.20) (R0(θ+i0)f)(x)=eiθ(x1)yx1[eiθyfL(y)0]+eiθ(x+1)yx+1[0eiθyfR(y)],\displaystyle\begin{split}&(R_{0}(\theta+i0)f)(x)\\ &=e^{i\theta(x-1)}\sum_{y\leq x-1}\left[\begin{array}[]{c}e^{-i\theta y}f_{L}(y)\\ 0\end{array}\right]+e^{-i\theta(x+1)}\sum_{y\geq x+1}\left[\begin{array}[]{c}0\\ e^{i\theta y}f_{R}(y)\end{array}\right],\end{split}
(2.21) (R0(θi0)f)(x)=eiθ(x1)yx[eiθyfL(y)0]eiθ(x+1)yx[0eiθyfR(y)],\displaystyle\begin{split}&(R_{0}(\theta-i0)f)(x)\\ &=-e^{i\theta(x-1)}\sum_{y\geq x}\left[\begin{array}[]{c}e^{-i\theta y}f_{L}(y)\\ 0\end{array}\right]-e^{-i\theta(x+1)}\sum_{y\leq x}\left[\begin{array}[]{c}0\\ e^{i\theta y}f_{R}(y)\end{array}\right],\end{split}

for θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}} and f(𝐙)f\in\mathcal{B}({\bf Z}).

Remark. Maeda et al. [16] have given an essentially equivalent expression of the Green function G0(x,κ)G_{0}(x,\kappa) by using a computation of the transfer matrix associated with QWs (see Proposition 3.6 and Lemma 4.9 in [16]). Compared to our construction by the Fourier transforms, their argument gives more clearly a discrete and unitary analogue of the scattering theory for the Strum-Liouville differential equation. On the other hand, our method can be extended to multi-dimensional QWs.

3. Generalized eigenfunction

3.1. Spectral representation for Free QW

In the following, we assume U0=SU_{0}=S. In [18], we derived a characterization of the generalized eigenfunction

(3.1) (U0eiθ)u=0on𝐙,θJ0J0,𝒯,(U_{0}-e^{i\theta})u=0\quad\text{on}\quad{\bf Z},\quad\theta\in J_{0}\setminus J_{0,\mathcal{T}},

in (𝐙)\mathcal{B}^{*}({\bf Z}). In order to apply this theory, we recall the framework of the spectral representation for U0U_{0}.

Now we introduce the function θ(ξ)=arccos(cosξ)\theta(\xi)=\arccos(\cos\xi) for ξ𝐓\xi\in{\bf T}. We have J0,1={θ(ξ);ξ𝐓}J_{0,1}=\{\theta(\xi)\ ;\ \xi\in{\bf T}\} and J0,2={2πθ(ξ);ξ𝐓}J_{0,2}=\{2\pi-\theta(\xi)\ ;\ \xi\in{\bf T}\}. The matrix U^0(ξ)\widehat{U}_{0}(\xi) has the spectral decomposition

U^0(ξ)=eiθ(ξ)P^1(ξ)+eiθ(ξ)P^2(ξ),\widehat{U}_{0}(\xi)=e^{i\theta(\xi)}\widehat{P}_{1}(\xi)+e^{-i\theta(\xi)}\widehat{P}_{2}(\xi),

where

P^1(ξ)={[1000],0<ξ<π,[0001],π<ξ<2π,P^2(ξ)={[0001],0<ξ<π,[1000],π<ξ<2π.\displaystyle\widehat{P}_{1}(\xi)=\left\{\begin{split}\left[\begin{array}[]{cc}1&0\\ 0&0\end{array}\right],&\quad 0<\xi<\pi,\\ \left[\begin{array}[]{cc}0&0\\ 0&1\end{array}\right],&\quad\pi<\xi<2\pi,\end{split}\right.\quad\widehat{P}_{2}(\xi)=\left\{\begin{split}\left[\begin{array}[]{cc}0&0\\ 0&1\end{array}\right],&\quad 0<\xi<\pi,\\ \left[\begin{array}[]{cc}1&0\\ 0&0\end{array}\right],&\quad\pi<\xi<2\pi.\end{split}\right.

Note that e±iθ(ξ)=e±iξe^{\pm i\theta(\xi)}=e^{\pm i\xi} for 0ξπ0\leq\xi\leq\pi and e±iθ(ξ)=eiξe^{\pm i\theta(\xi)}=e^{\mp i\xi} for πξ<2π\pi\leq\xi<2\pi. Letting

M(θ)={ξ𝐓;p(ξ,θ)=0}={±θ(ξ)},\displaystyle M(\theta)=\{\xi\in{\bf T}\ ;\ p(\xi,\theta)=0\}=\{\pm\theta(\xi)\},

we introduce the operator 0(θ)=(0,1(θ),0,2(θ))\mathcal{F}_{0}(\theta)=(\mathcal{F}_{0,1}(\theta),\mathcal{F}_{0,2}(\theta)) where

0,j(θ)f={P^jf^|M(θ),θJ0,j,0,θJ0,j,f(𝐙).\displaystyle\mathcal{F}_{0,j}(\theta)f=\left\{\begin{split}\widehat{P}_{j}\widehat{f}\big{|}_{M(\theta)},&\quad\theta\in J_{0,j},\\ 0,&\quad\theta\not\in J_{0,j},\end{split}\right.\quad f\in\mathcal{B}({\bf Z}).

In order to characterize the range of 0(θ)\mathcal{F}_{0}(\theta), we introduce the vector space 𝐡(θ){\bf h}(\theta) as follows. Let 𝐡~(θ)\widetilde{{\bf h}}(\theta) be the space of 𝐂{\bf C}-valued functions on M(θ)M(\theta) with its inner product

(ϕ,ψ)𝐡~(θ)=ξ(θ)M(θ)ϕ(ξ(θ))ψ(ξ(θ))¯,ϕ,ψ𝐡~(θ).(\phi,\psi)_{\widetilde{{\bf h}}(\theta)}=\sum_{\xi(\theta)\in M(\theta)}\phi(\xi(\theta))\overline{\psi(\xi(\theta))},\quad\phi,\psi\in\widetilde{{\bf h}}(\theta).

The eigenvectors 𝐞1(ξ),𝐞2(ξ)𝐂2{\bf e}_{1}(\xi),{\bf e}_{2}(\xi)\in{\bf C}^{2} of U^0(ξ)\widehat{U}_{0}(\xi) with respect to eigenvalues eiθ(ξ)e^{i\theta(\xi)}, eiθ(ξ)e^{-i\theta(\xi)} are given by

𝐞1(ξ)={[10],0<ξ<π,[01],π<ξ<2π,,𝐞2(ξ)={[01],0<ξ<π,[10],π<ξ<2π.\displaystyle{\bf e}_{1}(\xi)=\left\{\begin{split}\left[\begin{array}[]{c}1\\ 0\end{array}\right],&\quad 0<\xi<\pi,\\ \left[\begin{array}[]{c}0\\ 1\end{array}\right],&\quad\pi<\xi<2\pi,\end{split}\right.,\quad{\bf e}_{2}(\xi)=\left\{\begin{split}\left[\begin{array}[]{c}0\\ 1\end{array}\right],&\quad 0<\xi<\pi,\\ \left[\begin{array}[]{c}1\\ 0\end{array}\right],&\quad\pi<\xi<2\pi.\end{split}\right.

We introduce the vector space

𝐡~(θ)𝐞j={{ϕ𝐞j|M(θ);ϕ𝐡~(θ)},θJ0,j,{0},θJ0,j,\displaystyle\widetilde{{\bf h}}(\theta){\bf e}_{j}=\left\{\begin{split}\{\phi{\bf e}_{j}\big{|}_{M(\theta)}\ ;\ \phi\in\widetilde{{\bf h}}(\theta)\}&,\quad\theta\in J_{0,j},\\ \{0\}&,\quad\theta\not\in J_{0,j},\end{split}\right.

and we define

𝐡(θ)=𝐡~(θ)𝐞1𝐡~(θ)𝐞2,θJ0J0,𝒯.{\bf h}(\theta)=\widetilde{{\bf h}}(\theta){\bf e}_{1}\oplus\widetilde{{\bf h}}(\theta){\bf e}_{2},\quad\theta\in J_{0}\setminus J_{0,\mathcal{T}}.

For ϕ𝐡(θ)\phi\in{\bf h}(\theta), we can take ϕ1,ϕ2𝐡~(θ)\phi_{1},\phi_{2}\in\widetilde{{\bf h}}(\theta) such that ϕ\phi is formally represented by ϕ=ϕ1𝐞1ϕ2𝐞2\phi=\phi_{1}{\bf e}_{1}\oplus\phi_{2}{\bf e}_{2} where ϕj𝐞j𝐡~(θ)𝐞j\phi_{j}{\bf e}_{j}\in\widetilde{{\bf h}}(\theta){\bf e}_{j}. Then we have 0(θ)𝐁((𝐙);𝐡(θ))\mathcal{F}_{0}(\theta)\in{\bf B}(\mathcal{B}({\bf Z});{\bf h}(\theta)) for θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}, and we can characterize the generalized eigenfunctions of U0U_{0} in (𝐙)\mathcal{B}^{*}({\bf Z}) by using the adjoint operator 0(θ)\mathcal{F}_{0}(\theta)^{*} as follows.

Lemma 3.1.

Let θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}. Then we have 0(θ)𝐡(θ)={u(𝐙);(U0eiθ)u=0}\mathcal{F}_{0}(\theta)^{*}{\bf h}(\theta)=\{u\in\mathcal{B}^{*}({\bf Z})\ ;\ (U_{0}-e^{i\theta})u=0\}.

Proof. See Theorem 3.11 in [18]. ∎

In fact, 0(θ)ϕ(𝐙)\mathcal{F}_{0}(\theta)^{*}\phi\in\mathcal{B}^{*}({\bf Z}) for the free QW is given by

(3.4) (0(θ)ϕ)(x)=12π[ϕj(θ)eiθxϕj(θ)eiθx],θJ0,jJ0,𝒯.\displaystyle(\mathcal{F}_{0}(\theta)^{*}\phi)(x)=\frac{1}{\sqrt{2\pi}}\left[\begin{array}[]{c}\phi_{j}(\theta)e^{i\theta x}\\ \phi_{j}(-\theta)e^{-i\theta x}\end{array}\right],\quad\theta\in J_{0,j}\setminus J_{0,\mathcal{T}}.

Here we adopt the identification

M(θ)±θ(ξ)={±θ,θJ0,1J0,𝒯,θ,θJ0,2J0,𝒯.\displaystyle M(\theta)\ni\pm\theta(\xi)=\left\{\begin{split}\pm\theta&,\quad\theta\in J_{0,1}\setminus J_{0,\mathcal{T}},\\ \mp\theta&,\quad\theta\in J_{0,2}\setminus J_{0,\mathcal{T}}.\end{split}\right.

The operator 0(θ)\mathcal{F}_{0}(\theta) appears in the asymptotic behavior of R0(θ±i0)R_{0}(\theta\pm i0) at x±x\to\pm\infty. Here we define the equivalence relation uvu\simeq v as x±x\to\pm\infty for u,v(𝐙)u,v\in\mathcal{B}^{*}({\bf Z}) by

uvu(x)v(x)=o(1)asx±.u\simeq v\Leftrightarrow u(x)-v(x)=o(1)\quad\text{as}\quad x\to\pm\infty.
Lemma 3.2.

Let f(𝐙)f\in\mathcal{B}({\bf Z}). We have

(3.5) (R0(θ+i0)f)(x)2πeiθe±iθx(0,1(θ)f)(±θ)x±,\displaystyle(R_{0}(\theta+i0)f)(x)\simeq\sqrt{2\pi}e^{-i\theta}e^{\pm i\theta x}(\mathcal{F}_{0,1}(\theta)f)(\pm\theta)\quad x\to\pm\infty,
(3.6) (R0(θi0)f)(x)2πeiθeiθx(0,1(θ)f)(θ),x±,\displaystyle(R_{0}(\theta-i0)f)(x)\simeq-\sqrt{2\pi}e^{-i\theta}e^{\mp i\theta x}(\mathcal{F}_{0,1}(\theta)f)(\mp\theta),\quad x\to\pm\infty,

for θJ0,1J0,𝒯\theta\in J_{0,1}\setminus J_{0,\mathcal{T}}, and

(3.7) (R0(θ+i0)f)(x)2πeiθe±iθx(0,2(θ)f)(±θ),x±,\displaystyle(R_{0}(\theta+i0)f)(x)\simeq\sqrt{2\pi}e^{-i\theta}e^{\pm i\theta x}(\mathcal{F}_{0,2}(\theta)f)(\pm\theta),\quad x\to\pm\infty,
(3.8) (R0(θi0)f)(x)2πeiθeiθx(0,2(θ)f)(θ),x±,\displaystyle(R_{0}(\theta-i0)f)(x)\simeq-\sqrt{2\pi}e^{-i\theta}e^{\mp i\theta x}(\mathcal{F}_{0,2}(\theta)f)(\mp\theta),\quad x\to\pm\infty,

for θJ0,2J0,𝒯\theta\in J_{0,2}\setminus J_{0,\mathcal{T}}.

Proof. This lemma is a direct consequence of (2.20)-(2.21). ∎

3.2. Position-dependent QW

For the position-dependent QW UU, we define R(κ)=(Ueiκ)1R(\kappa)=(U-e^{i\kappa})^{-1} for κ𝐂𝐑\kappa\in{\bf C}\setminus{\bf R}. The limiting absorption principle for R(κ)R(\kappa) holds as follows.

Lemma 3.3.

Following assertions hold.

  1. (1)

    For θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}} and f(𝐙)f\in\mathcal{B}({\bf Z}), there exist weak-* limits R(θ±i0)f:=limϵ0R(θilog(1ϵ))fR(\theta\pm i0)f:=\lim_{\epsilon\downarrow 0}R(\theta-i\log(1\mp\epsilon))f in the sense

    limϵ0(R(θilog(1ϵ))f,g)=(R(θ±i0)f,g),g(𝐙).\lim_{\epsilon\downarrow 0}(R(\theta-i\log(1\mp\epsilon))f,g)=(R(\theta\pm i0)f,g),\quad g\in\mathcal{B}({\bf Z}).
  2. (2)

    R(θ±i0)fR(\theta\pm i0)f for f(𝐙)f\in\mathcal{B}({\bf Z}) satisfy R(θ±i0)f(𝐙)cf(𝐙)\|R(\theta\pm i0)f\|_{\mathcal{B}^{*}({\bf Z})}\leq c\|f\|_{\mathcal{B}({\bf Z})} for a constant c>0c>0 which is independent of θ\theta if θ\theta varies over a compact interval in J0J0,𝒯J_{0}\setminus J_{0,\mathcal{T}}.

  3. (3)

    The mappings J0J0,𝒯θ(R(θ±i0)f,g)J_{0}\setminus J_{0,\mathcal{T}}\ni\theta\mapsto(R(\theta\pm i0)f,g) are continuous for f,g(𝐙)f,g\in\mathcal{B}({\bf Z}).

Proof. See Theorem 4.3 in [18]. ∎

In view of the well-known resolvent equation

R(κ)=R0(κ)(1VR(κ)),V=UU0,κ𝐂𝐑,R(\kappa)=R_{0}(\kappa)\left(1-VR(\kappa)\right),\quad V=U-U_{0},\quad\kappa\in{\bf C}\setminus{\bf R},

Lemmas 3.2 and 3.3 imply the following asymptotic behavior.

Lemma 3.4.

Let us define the distorted Fourier transformation ±(θ)\mathcal{F}_{\pm}(\theta) by

(3.9) ±(θ)=(±,1(θ),±,2(θ)),±,j(θ)=0,j(θ)(1VR(θ±i0)).\mathcal{F}_{\pm}(\theta)=(\mathcal{F}_{\pm,1}(\theta),\mathcal{F}_{\pm,2}(\theta)),\quad\mathcal{F}_{\pm,j}(\theta)=\mathcal{F}_{0,j}(\theta)(1-VR(\theta\pm i0)).

Taking a sequence f(𝐙)f\in\mathcal{B}({\bf Z}), we have

(3.10) (R(θ+i0)f)(x)2πeiθe±iθx(±,1(θ)f)(±θ),x±,\displaystyle(R(\theta+i0)f)(x)\simeq\sqrt{2\pi}e^{-i\theta}e^{\pm i\theta x}(\mathcal{F}_{\pm,1}(\theta)f)(\pm\theta),\quad x\to\pm\infty,
(3.11) (R(θi0)f)(x)2πeiθeiθx(±,1(θ)f)(θ),x±,\displaystyle(R(\theta-i0)f)(x)\simeq-\sqrt{2\pi}e^{-i\theta}e^{\mp i\theta x}(\mathcal{F}_{\pm,1}(\theta)f)(\mp\theta),\quad x\to\pm\infty,

for θJ0,1J0,𝒯\theta\in J_{0,1}\setminus J_{0,\mathcal{T}}, and

(3.12) (R(θ+i0)f)(x)2πeiθe±iθx(±,2(θ)f)(±θ),x±,\displaystyle(R(\theta+i0)f)(x)\simeq\sqrt{2\pi}e^{-i\theta}e^{\pm i\theta x}(\mathcal{F}_{\pm,2}(\theta)f)(\pm\theta),\quad x\to\pm\infty,
(3.13) (R(θi0)f)(x)2πeiθeiθx(±,2(θ)f)(θ),x±,\displaystyle(R(\theta-i0)f)(x)\simeq-\sqrt{2\pi}e^{-i\theta}e^{\mp i\theta x}(\mathcal{F}_{\pm,2}(\theta)f)(\mp\theta),\quad x\to\pm\infty,

for θJ0,2J0,𝒯\theta\in J_{0,2}\setminus J_{0,\mathcal{T}}.

Remark. When UU0U-U_{0} is a sufficiently small perturbation, the perturbed Green function exists. The perturbed Green function is the kernel function of R(κ)R(\kappa). For our purpose in this paper, it is sufficient to use the resolvent equation and the distorted Fourier transformation. The radiation condition for the equation (Ueiθ)u=f(U-e^{i\theta})u=f can be determined by the asymptotic behavior of G0(x,θ±i0)G_{0}(x,\theta\pm i0) as |x||x|\to\infty. We study the asymptotic behavior of the Green function and its application to the radiation condition for a multi-dimensional QW in the forthcoming paper [11].

The adjoint operator ±(θ)\mathcal{F}_{\pm}(\theta)^{*} of the distorted Fourier transformation characterizes the generalized eigenfunction of UU.

Lemma 3.5.

Let θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}. Then we have ±(θ)𝐡(θ)={u(𝐙);(Ueiθ)u=0}\mathcal{F}_{\pm}(\theta)^{*}{\bf h}(\theta)=\{u\in\mathcal{B}^{*}({\bf Z})\ ;\ (U-e^{i\theta})u=0\}.

Proof. See Theorem 5.6 in [18]. ∎

The generalized eigenfunction +(θ)ϕ\mathcal{F}_{+}(\theta)^{*}\phi derives the scattered wave as follows. In view of the equality

R(θ+i0)=eiθUR(θi0),R(\theta+i0)^{*}=-e^{i\theta}UR(\theta-i0),

we have

+(θ)ϕ=0(θ)ϕ+e2iθR(θi0)V0(θ)ϕ+eiθV0(θ)ϕ.\mathcal{F}_{+}(\theta)^{*}\phi=\mathcal{F}_{0}(\theta)^{*}\phi+e^{2i\theta}R(\theta-i0)V^{*}\mathcal{F}_{0}(\theta)^{*}\phi+e^{i\theta}V^{*}\mathcal{F}_{0}(\theta)^{*}\phi.

The third term on the right-hand side is negligible at infinity due to the assumption (A-1).

The S-matrix Σ^(θ)\widehat{\Sigma}(\theta) for θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}} is given by

(3.14) Σ^(θ)=12πeiθA(θ),A(θ)=(θ)V0(θ),\widehat{\Sigma}(\theta)=1-2\pi e^{i\theta}A(\theta),\quad A(\theta)=\mathcal{F}_{-}(\theta)V^{*}\mathcal{F}_{0}(\theta)^{*},

and unitary on 𝐡(θ){\bf h}(\theta). For the proof, see Theorem 5.3 in [18]. Due to the definition of (θ)\mathcal{F}_{-}(\theta) and the formula (3.4), we put

Σ^j(θ)={12πeiθ,j(θ)V0(θ),θJ0,j,1,θJ0,j,\displaystyle\widehat{\Sigma}_{j}(\theta)=\left\{\begin{split}1-2\pi e^{i\theta}\mathcal{F}_{-,j}(\theta)V^{*}\mathcal{F}_{0}(\theta)^{*}&,\quad\theta\in J_{0,j},\\ 1&,\quad\theta\not\in J_{0,j},\end{split}\right.

for j=1,2j=1,2. Then we have Σ^(θ)=(Σ^1(θ),Σ^2(θ))\widehat{\Sigma}(\theta)=(\widehat{\Sigma}_{1}(\theta),\widehat{\Sigma}_{2}(\theta)), and it follows that each Σ^j(θ)\widehat{\Sigma}_{j}(\theta) is a 2×22\times 2 unitary matrix for θJ0,jJ0,𝒯\theta\in J_{0,j}\setminus J_{0,\mathcal{T}}. In the following, we denote the S-matrix by

Σ^j(θ)=[τj(θ)ρ~j(θ)ρj(θ)τ~j(θ)]U(2),j=1,2.\widehat{\Sigma}_{j}(\theta)=\left[\begin{array}[]{cc}\tau_{j}(\theta)&\widetilde{\rho}_{j}(\theta)\\ \rho_{j}(\theta)&\widetilde{\tau}_{j}(\theta)\end{array}\right]\in\mathrm{U}(2),\quad j=1,2.

Now the asymptotic behavior of +(θ)ϕ\mathcal{F}_{+}(\theta)^{*}\phi comes from Lemma 3.4 as follows. The S-matrix naturally appears in the asymptotic behavior at infinity.

Lemma 3.6.

Let ϕ𝐡(θ)\phi\in{\bf h}(\theta). We have

(3.17) (+(θ)ϕ)(x)12π[ϕj(θ)eiθx(ρj(θ)ϕj(θ)+τ~j(θ)ϕj(θ))eiθx],x,\displaystyle(\mathcal{F}_{+}(\theta)^{*}\phi)(x)\simeq\frac{1}{\sqrt{2\pi}}\left[\begin{array}[]{c}\phi_{j}(\theta)e^{i\theta x}\\ (\rho_{j}(\theta)\phi_{j}(\theta)+\widetilde{\tau}_{j}(\theta)\phi_{j}(-\theta))e^{-i\theta x}\end{array}\right],\quad x\to\infty,
(3.20) (+(θ)ϕ)(x)12π[(τj(θ)ϕj(θ)+ρ~j(θ)ϕj(θ))eiθxϕj(θ)eiθx],x,\displaystyle(\mathcal{F}_{+}(\theta)^{*}\phi)(x)\simeq\frac{1}{\sqrt{2\pi}}\left[\begin{array}[]{c}(\tau_{j}(\theta)\phi_{j}(\theta)+\widetilde{\rho}_{j}(\theta)\phi_{j}(-\theta))e^{i\theta x}\\ \phi_{j}(-\theta)e^{-i\theta x}\end{array}\right],\quad x\to-\infty,

for θJ0,jJ0,𝒯\theta\in J_{0,j}\setminus J_{0,\mathcal{T}}.

In order to determine Σ^(θ)\widehat{\Sigma}(\theta), it is sufficient to consider the cases ϕj(θ)=2π\phi_{j}(\theta)=\sqrt{2\pi}, ϕj(θ)=0\phi_{j}(-\theta)=0, and ϕj(θ)=0\phi_{j}(\theta)=0, ϕj(θ)=2π\phi_{j}(-\theta)=\sqrt{2\pi}. Namely, we have

(3.23) (+(θ)ϕ)(x)[eiθxρj(θ)eiθx],x,\displaystyle(\mathcal{F}_{+}(\theta)^{*}\phi)(x)\simeq\left[\begin{array}[]{c}e^{i\theta x}\\ \rho_{j}(\theta)e^{-i\theta x}\end{array}\right],\quad x\to\infty,
(3.26) (+(θ)ϕ)(x)[τj(θ)eiθx0],x,\displaystyle(\mathcal{F}_{+}(\theta)^{*}\phi)(x)\simeq\left[\begin{array}[]{c}\tau_{j}(\theta)e^{i\theta x}\\ 0\end{array}\right],\quad x\to-\infty,

for ϕj(θ)=2π\phi_{j}(\theta)=\sqrt{2\pi}, ϕj(θ)=0\phi_{j}(-\theta)=0, and

(3.29) (+(θ)ϕ)(x)[0τ~j(θ)eiθx],x,\displaystyle(\mathcal{F}_{+}(\theta)^{*}\phi)(x)\simeq\left[\begin{array}[]{c}0\\ \widetilde{\tau}_{j}(\theta)e^{-i\theta x}\end{array}\right],\quad x\to\infty,
(3.32) (+(θ)ϕ)(x)[ρ~j(θ)eiθxeiθx],x,\displaystyle(\mathcal{F}_{+}(\theta)^{*}\phi)(x)\simeq\left[\begin{array}[]{c}\widetilde{\rho}_{j}(\theta)e^{i\theta x}\\ e^{-i\theta x}\end{array}\right],\quad x\to-\infty,

for ϕj(θ)=0\phi_{j}(\theta)=0, ϕj(θ)=2π\phi_{j}(-\theta)=\sqrt{2\pi}.

In the following, we drop the suffix jj in every component of Σ^j(θ)\widehat{\Sigma}_{j}(\theta) with respect to θJ0,jJ0,𝒯\theta\in J_{0,j}\setminus J_{0,\mathcal{T}} for the sake of simplicity of the notation. ϕL(θ)\phi_{L}(\theta) and ϕR(θ)\phi_{R}(\theta) denote ϕj(θ)\phi_{j}(\theta) and ϕj(θ)\phi_{j}(-\theta) for θJ0,jJ0,𝒯\theta\in J_{0,j}\setminus J_{0,\mathcal{T}}, respectively. The S-matrix Σ^(θ)\widehat{\Sigma}(\theta) for θJ0,jJ0,𝒯\theta\in J_{0,j}\setminus J_{0,\mathcal{T}} is simply denoted by

[τ(θ)ρ~(θ)ρ(θ)τ~(θ)]U(2).\left[\begin{array}[]{cc}\tau(\theta)&\widetilde{\rho}(\theta)\\ \rho(\theta)&\widetilde{\tau}(\theta)\end{array}\right]\in\mathrm{U}(2).

4. Combinatorial construction of scattering matrix

4.1. Finite rank perturbation

In the following, we assume that the position-dependent QW UU is a finite rank perturbation of the free QW U0=SU_{0}=S. We put Γ={0,1,,n}\Gamma=\{0,1,\ldots,n\} for a positive integer nn, and

(4.1) C(x)={[a(x)b(x)c(x)d(x)],xΓ,[1001],xΓ.\displaystyle C(x)=\left\{\begin{split}\left[\begin{array}[]{cc}a(x)&b(x)\\ c(x)&d(x)\end{array}\right],&\quad x\in\Gamma,\\ \left[\begin{array}[]{cc}1&0\\ 0&1\end{array}\right],&\quad x\not\in\Gamma.\end{split}\right.

Due to the assumption (A-2), we have a(x)0a(x)\not=0 (d(x)0\Leftrightarrow d(x)\not=0) for every xΓx\in\Gamma. Recall the rule of notation which has been introduced at the end of previous section. Since UU0U-U_{0} is an operator of finite rank, the asymptotic behaviors (3.23)-(3.26) and (3.29)-(3.32) become exactly

(4.2) (+(θ)ϕ)(x)={[eiθxρ(θ)eiθx],xn+1,[τ(θ)eiθx0],x1,\displaystyle(\mathcal{F}_{+}(\theta)^{*}\phi)(x)=\left\{\begin{split}\left[\begin{array}[]{c}e^{i\theta x}\\ \rho(\theta)e^{-i\theta x}\end{array}\right],&\quad x\geq n+1,\\ \left[\begin{array}[]{c}\tau(\theta)e^{i\theta x}\\ 0\end{array}\right],&\quad x\leq-1,\end{split}\right.

for ϕL(θ)=2π\phi_{L}(\theta)=\sqrt{2\pi}, ϕR(θ)=0\phi_{R}(\theta)=0, and

(4.3) (+(θ)ϕ)(x)={[0τ~(θ)eiθx],xn+1,[ρ~(θ)eiθxeiθx],x1,\displaystyle(\mathcal{F}_{+}(\theta)^{*}\phi)(x)=\left\{\begin{split}\left[\begin{array}[]{c}0\\ \widetilde{\tau}(\theta)e^{-i\theta x}\end{array}\right],&\quad x\geq n+1,\\ \left[\begin{array}[]{c}\widetilde{\rho}(\theta)e^{i\theta x}\\ e^{-i\theta x}\end{array}\right],&\quad x\leq-1,\end{split}\right.

for ϕL(θ)=0\phi_{L}(\theta)=0, ϕR(θ)=2π\phi_{R}(\theta)=\sqrt{2\pi}.

In order to introduce a combinatorial construction of the S-matrix, let us relate the generalized eigenfunction of UU and the dynamics of the QW. Let χ:(𝐙)2(Γ;𝐂2)\chi:\mathcal{B}^{*}({\bf Z})\to\ell^{2}(\Gamma;{\bf C}^{2}) be defined by (χu)(x)=u(x)(\chi u)(x)=u(x) for xΓx\in\Gamma and u(𝐙)u\in\mathcal{B}^{*}({\bf Z}). The operator χ:2(Γ;𝐂2)(𝐙)\chi^{*}:\ell^{2}(\Gamma;{\bf C}^{2})\to\mathcal{B}^{*}({\bf Z}) is defined by

(χψ)(x)={ψ(x),xΓ,0,xΓ,ψ2(Γ;𝐂2).\displaystyle(\chi^{*}\psi)(x)=\left\{\begin{split}\psi(x),&\quad x\in\Gamma,\\ 0,&\quad x\not\in\Gamma,\end{split}\right.\quad\psi\in\ell^{2}(\Gamma;{\bf C}^{2}).

Here the operator χ𝐁(2(Γ;𝐂2);2(𝐙;𝐂2))\chi^{*}\in{\bf B}(\ell^{2}(\Gamma;{\bf C}^{2});\ell^{2}({\bf Z};{\bf C}^{2})) is the adjoint operator of χ𝐁(2(𝐙;𝐂2);2(Γ;𝐂2))\chi\in{\bf B}(\ell^{2}({\bf Z};{\bf C}^{2});\ell^{2}(\Gamma;{\bf C}^{2})). Note that χχ\chi\chi^{*} is the identity on 2(Γ;𝐂2)\ell^{2}(\Gamma;{\bf C}^{2}), and χχ\chi^{*}\chi is the projection onto the subspace {u(𝐙);suppuΓ}\{u\in\mathcal{B}^{*}({\bf Z})\ ;\ \mathrm{supp}u\subset\Gamma\}. Now we define the 2(n+1)×2(n+1)2(n+1)\times 2(n+1) submatrix EnE_{n} of UU by En=χUχE_{n}=\chi U\chi^{*}. Precisely, identifying a vector ψ2(Γ;𝐂2)\psi\in\ell^{2}(\Gamma;{\bf C}^{2}) with [ψR(0),ψL(0),ψR(1),ψL(1),,ψR(n),ψL(n)]𝖳𝐂2(n+1)[\psi_{R}(0),\psi_{L}(0),\psi_{R}(1),\psi_{L}(1),\ldots,\psi_{R}(n),\psi_{L}(n)]^{\mathsf{T}}\in{\bf C}^{2(n+1)}, we have

En=[0P(1)Q(0)0P(2)Q(1)00P(n)Q(n1)0],E_{n}=\left[\begin{array}[]{ccccc}0&P(1)&&&\\ Q(0)&0&P(2)&&\\ &Q(1)&0&&\\ &&\ddots&\ddots&\\ &&&0&P(n)\\ &&&Q(n-1)&0\end{array}\right],

where

P(x)=[00b(x)a(x)],Q(x)=[d(x)c(x)00],P(x)=\left[\begin{array}[]{cc}0&0\\ b(x)&a(x)\end{array}\right],\quad Q(x)=\left[\begin{array}[]{cc}d(x)&c(x)\\ 0&0\end{array}\right],

for xΓx\in\Gamma.

Lemma 4.1.

Eigenvalues of EnE_{n} lie in {λ𝐂;|λ|<1}\{\lambda\in{\bf C}\ ;\ |\lambda|<1\}.

Proof. Let ψ2(Γ;𝐂2)\psi\in\ell^{2}(\Gamma;{\bf C}^{2}) be an eigenvector associated with the eigenvalue λ\lambda of EnE_{n}. Then we have

(4.4) |λ|2ψ2(Γ;𝐂2)2=Enψ2(Γ;𝐂2)2(Uχψ,Uχψ)2(𝐙;𝐂2)=ψ2(Γ;𝐂2)2.|\lambda|^{2}\|\psi\|^{2}_{\ell^{2}(\Gamma;{\bf C}^{2})}=\|E_{n}\psi\|^{2}_{\ell^{2}(\Gamma;{\bf C}^{2})}\leq(U\chi^{*}\psi,U\chi^{*}\psi)_{\ell^{2}({\bf Z};{\bf C}^{2})}=\|\psi\|^{2}_{\ell^{2}(\Gamma;{\bf C}^{2})}.

Now we suppose |λ|=1|\lambda|=1. Then the equality in (4.4) holds. Due to the definition of χ\chi and χ\chi^{*}, this implies χχUχψ=Uχψ\chi^{*}\chi U\chi^{*}\psi=U\chi^{*}\psi on 𝐙{\bf Z} so that Uχψ=λχψU\chi^{*}\psi=\lambda\chi^{*}\psi. This means that χψ\chi^{*}\psi is an eigenfunction of UU associated with the eigenvalue λ\lambda with a finite support. By the assumption of C(x)C(x), there is no eigenfunction of UU with a finite support. This is a contradiction. We obtain |λ|<1|\lambda|<1. ∎

Under the setting of this paper, we can prove the convergence of the series 1+eiθEn+e2iθEn2+1+e^{-i\theta}E_{n}+e^{-2i\theta}E_{n}^{2}+\cdots by using the Jordan canonical form of the matrix EnE_{n}. Let λ1,,λr\lambda_{1},\ldots,\lambda_{r}, 1r2(n+1)1\leq r\leq 2(n+1), be eigenvalues of EnE_{n} with algebraic multiplicities kjk_{j}, j=1,,rj=1,\ldots,r. There exists a 2(n+1)×2(n+1)2(n+1)\times 2(n+1) regular matrix GnG_{n} such that

Gn1EnGn=J(λ1,k1)J(λr,kr),G^{-1}_{n}E_{n}G_{n}=J(\lambda_{1},k_{1})\oplus\cdots\oplus J(\lambda_{r},k_{r}),

where J(λj,kj)J(\lambda_{j},k_{j}) is the Jordan block associated with the eigenvalue λj\lambda_{j}.

Lemma 4.2.

The Neumann series 1+eiθEn+e2iθEn2+1+e^{-i\theta}E_{n}+e^{-2i\theta}E_{n}^{2}+\cdots converges.

Proof. This lemma is a consequence of the Jordan canonical form of EnE_{n} and Lemma 4.1. ∎

Remark. When C(x)C(x) depends on the position xΓx\in\Gamma, it is difficult to know the eigenvalues of EnE_{n} in detail. We can show the following facts.

  • EnE_{n} has 0 as an eigenvalue. The dimension of the corresponding eigenspace is 22 or higher.

  • The dimension of the eigenspace associated with a non-zero eigenvalue of EnE_{n} is 11.

The following fact is a special case of Theorem 3.1 in [8]. We derive a shortcut of the proof under our settings.

Proposition 4.3.

Let Ψ0(𝐙)\Psi_{0}\in\mathcal{B}^{*}({\bf Z}) be given by

Ψ0(x)={αLeiθx[10],xn+1,αReiθx[01],x1,[00],otherwise,\displaystyle\Psi_{0}(x)=\left\{\begin{split}\alpha_{L}e^{i\theta x}\left[\begin{array}[]{c}1\\ 0\end{array}\right],&\quad x\geq n+1,\\ \alpha_{R}e^{-i\theta x}\left[\begin{array}[]{c}0\\ 1\end{array}\right],&\quad x\leq-1,\\ \left[\begin{array}[]{c}0\\ 0\end{array}\right],&\quad\text{otherwise},\end{split}\right.

for θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}, and for any constants αL,αR𝐂\alpha_{L},\alpha_{R}\in{\bf C}. We define Ψt\Psi_{t} for every positive integer tt by Ψt=UΨt1\Psi_{t}=U\Psi_{t-1}. Then there exists the limit

Ψ(x):=limteiθtΨt(x),x𝐙,\Psi_{\infty}(x):=\lim_{t\to\infty}e^{-i\theta t}\Psi_{t}(x),\quad x\in{\bf Z},

and Ψ(𝐙)\Psi_{\infty}\in\mathcal{B}^{*}({\bf Z}) satisfies

(4.5) UΨ=eiθΨon𝐙.U\Psi_{\infty}=e^{i\theta}\Psi_{\infty}\quad\text{on}\quad{\bf Z}.

In particular, we have Ψ=+(θ)ϕ\Psi_{\infty}=\mathcal{F}_{+}(\theta)^{*}\phi for ϕL(θ)=2παL\phi_{L}(\theta)=\sqrt{2\pi}\alpha_{L} and ϕR(θ)=2παR\phi_{R}(\theta)=\sqrt{2\pi}\alpha_{R}.

Proof. The initial state Ψ0\Psi_{0} represents the flow incoming from infinity. The incoming flow is split into two parts at x=0x=0 and nn by the time evolution. One is reflected and another one is transmitted. Once the flow comes out of Γ\Gamma, it goes to infinity. In view of this dynamics, we decompose Ψt\Psi_{t} as

Ψt=χχΨt+(1χχ)(ΨteitθΨ0)+eitθ(1χχ)Ψ0.\Psi_{t}=\chi^{*}\chi\Psi_{t}+(1-\chi^{*}\chi)(\Psi_{t}-e^{it\theta}\Psi_{0})+e^{it\theta}(1-\chi^{*}\chi)\Psi_{0}.

The first term on the right-hand side is the state in Γ\Gamma at time tt. The second term and the third term on the right-hand side are the outgoing flow and the incoming flow, respectively.

We put ϕt=χΨt\phi_{t}=\chi\Psi_{t}. Then we have

(4.6) ϕ0=0,ϕt+1=Enϕt+eiθtχUΨ0,t0.\phi_{0}=0,\quad\phi_{t+1}=E_{n}\phi_{t}+e^{i\theta t}\chi U\Psi_{0},\quad t\geq 0.

Here we have used the relation χU(1χχ)UtΨ0=eiθtχUΨ0\chi U(1-\chi^{*}\chi)U^{t}\Psi_{0}=e^{i\theta t}\chi U\Psi_{0}. This recurrence relation implies

ϕt=eiθ(t1)m=0t1eiθmEnmχUΨ0,t1.\phi_{t}=e^{i\theta(t-1)}\sum_{m=0}^{t-1}e^{-i\theta m}E_{n}^{m}\chi U\Psi_{0},\quad t\geq 1.

In view of Lemmas 4.1 and 4.2, the limit

ϕ:=limteiθtϕt=eiθ(1eiθEn)1χUΨ0\phi_{\infty}:=\lim_{t\to\infty}e^{-i\theta t}\phi_{t}=e^{-i\theta}(1-e^{-i\theta}E_{n})^{-1}\chi U\Psi_{0}

exists. This implies

(4.7) Enϕ+χUΨ0=eiθϕ.E_{n}\phi_{\infty}+\chi U\Psi_{0}=e^{i\theta}\phi_{\infty}.

Let us turn to the outgoing flow. We can see

(1χχ)(Ψ1eitθΨ0)=0,(1χχ)(ΨteitθΨ0)=(1χχ)m=1t1Umftm,t2,\displaystyle\begin{split}&(1-\chi^{*}\chi)(\Psi_{1}-e^{it\theta}\Psi_{0})=0,\\ &(1-\chi^{*}\chi)(\Psi_{t}-e^{it\theta}\Psi_{0})=(1-\chi^{*}\chi)\sum_{m=1}^{t-1}U^{m}f_{t-m},\quad t\geq 2,\end{split}

where the source ftmf_{t-m} is defined by ftm=(δ0+δn)χϕtmf_{t-m}=(\delta_{0}+\delta_{n})\chi^{*}\phi_{t-m} for δy={δxy}x𝐙\delta_{y}=\{\delta_{xy}\}_{x\in{\bf Z}}, y𝐙y\in{\bf Z}. For sufficiently large tt, the value of (1χχ)(ΨteitθΨ0)(1-\chi^{*}\chi)(\Psi_{t}-e^{it\theta}\Psi_{0}) at a point x𝐙Γx\in{\bf Z}\setminus\Gamma is given by UμxftμxU^{\mu_{x}}f_{t-\mu_{x}} where μx=x\mu_{x}=-x for x1x\leq-1 and μx=xn\mu_{x}=x-n for xn+1x\geq n+1. The limit Ψout:=limteitθ(1χχ)(ΨteitθΨ0)\Psi_{out}:=\lim_{t\to\infty}e^{-it\theta}(1-\chi^{*}\chi)(\Psi_{t}-e^{it\theta}\Psi_{0}) is determined by the limits

limteitθ(Uμxftμx)(x)=eiμxθ(Uf)(x±),x𝐙Γ,\lim_{t\to\infty}e^{-it\theta}(U^{\mu_{x}}f_{t-\mu_{x}})(x)=e^{-i\mu_{x}\theta}(Uf_{\infty})(x_{\pm}),\quad x\in{\bf Z}\setminus\Gamma,

where f=(δ0+δn)χϕf_{\infty}=(\delta_{0}+\delta_{n})\chi^{*}\phi_{\infty}, and x=1x_{-}=-1 for x1x\leq-1 or x+=n+1x_{+}=n+1 for xn+1x\geq n+1. Thus we obtain

(4.8) Ψout(x)=eiμxθ(Uf)(x±),x𝐙Γ.\Psi_{out}(x)=e^{-i\mu_{x}\theta}(Uf_{\infty})(x_{\pm}),\quad x\in{\bf Z}\setminus\Gamma.

Finally, we prove the equation (4.5). Note that

Ψ=χϕ+Ψout+Ψ0on𝐙.\Psi_{\infty}=\chi^{*}\phi_{\infty}+\Psi_{out}+\Psi_{0}\quad\text{on}\quad{\bf Z}.

From this decomposition, we have

(4.9) χUΨ=Enϕ+χUΨ0+χUΨout=eiθϕ,\chi U\Psi_{\infty}=E_{n}\phi_{\infty}+\chi U\Psi_{0}+\chi U\Psi_{out}=e^{i\theta}\phi_{\infty},

due to (4.7) and χUΨout=0\chi U\Psi_{out}=0. On the other hand, we have

(4.10) (1χχ)UΨ=(1χχ)U(χϕ+Ψ0)+UΨout=(1χχ)δx±Uf+eiθΨ0+eiθΨoutδx±Uf=eiθ(Ψout+Ψ0).\displaystyle\begin{split}(1-\chi^{*}\chi)U\Psi_{\infty}&=(1-\chi^{*}\chi)U(\chi^{*}\phi_{\infty}+\Psi_{0})+U\Psi_{out}\\ &=(1-\chi^{*}\chi)\delta_{x_{\pm}}Uf_{\infty}+e^{i\theta}\Psi_{0}+e^{i\theta}\Psi_{out}-\delta_{x_{\pm}}Uf_{\infty}\\ &=e^{i\theta}(\Psi_{out}+\Psi_{0}).\end{split}

Here we have used the relation UΨout=eiθ(Ψoutδx±Ψout)U\Psi_{out}=e^{i\theta}(\Psi_{out}-\delta_{x_{\pm}}\Psi_{out}). Plugging (4.9) and (4.10), we obtain the equation (4.5). ∎

Let us put αL=1\alpha_{L}=1 and αR=0\alpha_{R}=0. Then the generalized eigenfunction Ψ\Psi_{\infty} satisfies the behavior (4.2). The values Ψout(1)\Psi_{out}(-1) and Ψout(n+1)\Psi_{out}(n+1) determine τ(θ)\tau(\theta) and ρ(θ)\rho(\theta), respectively. In the following, Ak,lA_{k,l} denotes the (k,l)(k,l)-component of a matrix AA.

Theorem 4.4.

For θJ0,jJ0,𝒯\theta\in J_{0,j}\setminus J_{0,\mathcal{T}}, we have

τ(θ)=a(0)(Enn(1(eiθEn)2)1)2,2n+2,ρ(θ)=e2inθ(c(n)eiθ+d(n)e2iθ(En2(1(eiθEn)2)1)2n+1,2n+2),τ~(θ)=d(n)((Enn(1(eiθEn)2))1)2n+1,1,ρ~(θ)=a(0)e2iθ(En2(1(eiθEn)2)1)2,1+b(0)eiθ.\displaystyle\begin{split}&\tau(\theta)=a(0)(E_{n}^{n}(1-(e^{-i\theta}E_{n})^{2})^{-1})_{2,2n+2},\\ &\rho(\theta)=e^{2in\theta}\left(c(n)e^{i\theta}+d(n)e^{-2i\theta}(E_{n}^{2}(1-(e^{-i\theta}E_{n})^{2})^{-1})_{2n+1,2n+2}\right),\\ &\widetilde{\tau}(\theta)=d(n)((E_{n}^{n}(1-(e^{-i\theta}E_{n})^{2}))^{-1})_{2n+1,1},\\ &\widetilde{\rho}(\theta)=a(0)e^{-2i\theta}(E_{n}^{2}(1-(e^{-i\theta}E_{n})^{2})^{-1})_{2,1}+b(0)e^{i\theta}.\end{split}

Proof. Suppose αL=1\alpha_{L}=1 and αR=0\alpha_{R}=0. We have (Ψout)L(1)=τ(θ)eiθ(\Psi_{out})_{L}(-1)=\tau(\theta)e^{-i\theta} and (Ψout)R(n+1)=ρ(θ)eiθ(n+1)(\Psi_{out})_{R}(n+1)=\rho(\theta)e^{-i\theta(n+1)} in view of (4.2). We compute (Ψout)L(1)(\Psi_{out})_{L}(-1) and (Ψout)R(n+1)(\Psi_{out})_{R}(n+1) by using (4.8).

We have

Ψout(1)=eiθ[a(0)(ϕ)L(0)+b(0)(ϕ)R(0)0].\Psi_{out}(-1)=e^{-i\theta}\left[\begin{array}[]{cc}a(0)(\phi_{\infty})_{L}(0)+b(0)(\phi_{\infty})_{R}(0)\\ 0\end{array}\right].

In view of ϕ=eiθ(1+eiθEn+e2iθEn2+)φ\phi_{\infty}=e^{-i\theta}(1+e^{-i\theta}E_{n}+e^{-2i\theta}E_{n}^{2}+\cdots)\varphi where φ=χUΨ0\varphi=\chi U\Psi_{0}, we compute ϕ(0)\phi_{\infty}(0) as follows (see also Figure 1). Note that

φ=[φR(0)φL(0)φR(n)φL(n)]=[000eiθ(n+1)].\varphi=\left[\begin{array}[]{c}\varphi_{R}(0)\\ \varphi_{L}(0)\\ \vdots\\ \varphi_{R}(n)\\ \varphi_{L}(n)\end{array}\right]=\left[\begin{array}[]{c}0\\ 0\\ \vdots\\ 0\\ e^{i\theta(n+1)}\end{array}\right].

It follows (ϕ)R(0)=0(\phi_{\infty})_{R}(0)=0 from αR=0\alpha_{R}=0. On the other hand, we have

(ϕ)L(0)=eiθeiθ(n+1)(eiθnEnn+eiθ(n+2)Enn+2+)2,2n+2=(Enn(1(eiθEn)2)1)2,2n+2.\displaystyle\begin{split}(\phi_{\infty})_{L}(0)&=e^{-i\theta}e^{i\theta(n+1)}(e^{-i\theta n}E_{n}^{n}+e^{-i\theta(n+2)}E_{n}^{n+2}+\cdots)_{2,2n+2}\\ &=(E_{n}^{n}(1-(e^{-i\theta}E_{n})^{2})^{-1})_{2,2n+2}.\end{split}

Then we have (Ψout)L(1)=τ(θ)eiθ=eiθa(0)(Enn(1(eiθEn)2)1)2,2n+2(\Psi_{out})_{L}(-1)=\tau(\theta)e^{-i\theta}=e^{-i\theta}a(0)(E_{n}^{n}(1-(e^{-i\theta}E_{n})^{2})^{-1})_{2,2n+2}. This equality implies the formula of τ(θ)\tau(\theta).

Refer to caption
Figure 1. The dynamics of EnmφE_{n}^{m}\varphi for the case n=3n=3 implies the formulas of (ϕ)L(0)(\phi_{\infty})_{L}(0) and (ϕ)R(n)(\phi_{\infty})_{R}(n). Leftwards arrows and rightwards arrows on every point represent (Enmϕ)L(x)(E_{n}^{m}\phi)_{L}(x) and (Enmϕ)R(x)(E_{n}^{m}\phi)_{R}(x), respectively.

Next we compute ϕ(n)\phi_{\infty}(n). Note that

Ψout(n+1)=eiθ[0c(n)(ϕ)L(n)+d(n)(ϕ)R(n)].\Psi_{out}(n+1)=e^{-i\theta}\left[\begin{array}[]{c}0\\ c(n)(\phi_{\infty})_{L}(n)+d(n)(\phi_{\infty})_{R}(n)\end{array}\right].

We have (ϕ)L(n)=eiθ(n+1)(\phi_{\infty})_{L}(n)=e^{i\theta(n+1)} and

(ϕ)R(n)=eiθeiθ(n+1)(e2iθEn2+e4iθEn4+)1,2n+2=eiθ(n2)(En2(1(eiθEn)2)1)1,2n+2.\displaystyle\begin{split}(\phi_{\infty})_{R}(n)&=e^{-i\theta}e^{i\theta(n+1)}(e^{-2i\theta}E_{n}^{2}+e^{-4i\theta}E_{n}^{4}+\cdots)_{1,2n+2}\\ &=e^{i\theta(n-2)}(E_{n}^{2}(1-(e^{-i\theta}E_{n})^{2})^{-1})_{1,2n+2}.\end{split}

It follows from this equality that Ψout(n+1)=ρ(θ)eiθ(n+1)=eiθ(c(n)eiθ(n+1)+d(n)eiθ(n2)(En2(1(eiθEn)2)1)1,2n+2)\Psi_{out}(n+1)=\rho(\theta)e^{-i\theta(n+1)}=e^{-i\theta}(c(n)e^{i\theta(n+1)}+d(n)e^{i\theta(n-2)}(E_{n}^{2}(1-(e^{-i\theta}E_{n})^{2})^{-1})_{1,2n+2}). Thus we obtain the formula of ρ(θ)\rho(\theta).

Taking αL=0\alpha_{L}=0 and αR=1\alpha_{R}=1 and repeating the similar argument, we also have formulas of τ~(θ)\widetilde{\tau}(\theta) and ρ~(θ)\widetilde{\rho}(\theta). ∎

4.2. Resonant-tunneling effect for double barrier model

Let us consider the “double-barrier” model : C(x)C(x) is given by

C(x)={[ajbjcjdj],x=xj,[1001],xxj,\displaystyle C(x)=\left\{\begin{split}\left[\begin{array}[]{cc}a_{j}&b_{j}\\ c_{j}&d_{j}\end{array}\right]&,\quad x=x_{j},\\ \left[\begin{array}[]{cc}1&0\\ 0&1\end{array}\right]&,\quad x\not=x_{j},\end{split}\right.

for x0=0x_{0}=0 and x1=nx_{1}=n where aj0a_{j}\not=0 for j=0,1j=0,1. In this case, we can see the explicit formula of Σ^(θ)\widehat{\Sigma}(\theta) for θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}.

Proposition 4.5.

For θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}, we have

τ(θ)=a0a11b1c0e2iθn,ρ(θ)=c0Δ1+c1e2iθn1b1c0e2iθn,τ~(θ)=d0d11b1c0e2iθn,ρ~(θ)=b1Δ0e2iθn+b01b1c0e2iθn,\displaystyle\begin{split}&\tau(\theta)=\frac{a_{0}a_{1}}{1-b_{1}c_{0}e^{-2i\theta n}},\quad\rho(\theta)=\frac{c_{0}\Delta_{1}+c_{1}e^{2i\theta n}}{1-b_{1}c_{0}e^{-2i\theta n}},\\ &\widetilde{\tau}(\theta)=\frac{d_{0}d_{1}}{1-b_{1}c_{0}e^{-2i\theta n}},\quad\widetilde{\rho}(\theta)=\frac{b_{1}\Delta_{0}e^{-2i\theta n}+b_{0}}{1-b_{1}c_{0}e^{-2i\theta n}},\end{split}

for Δj=ajdjbjcj\Delta_{j}=a_{j}d_{j}-b_{j}c_{j}.

Note that we can compute the generalized eigenfunction which satisfies (4.2) or (4.3). Proposition 4.5 can be proven directly.

Now we consider the resonant-tunneling effect as an analogue of the one dimensional Schrödinger equation. Resonant-tunneling effect means that an incident wave passes through a barrier without loss of its energy. In view of the scattering theory, this phenomenon can be derived by ρ(θ)=0(|τ(θ)|=1)\rho(\theta)=0(\Leftrightarrow|\tau(\theta)|=1).

Lemma 4.6.

If ρ(θ)=0\rho(\theta)=0 for some θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}, there exist parameters αj,βj,γj𝐓\alpha_{j},\beta_{j},\gamma_{j}\in{\bf T}, j=0,1j=0,1, and p,q[0,1]p,q\in[0,1], p0p\not=0, p2+q2=1p^{2}+q^{2}=1, such that

(4.11) C(xj)=eiγj/2[pei(αjγj/2)qei(βjγj/2)qei(βjγj/2)pei(αjγj/2)],C(x_{j})=e^{i\gamma_{j}/2}\left[\begin{array}[]{cc}pe^{i(\alpha_{j}-\gamma_{j}/2)}&qe^{i(\beta_{j}-\gamma_{j}/2)}\\ -qe^{-i(\beta_{j}-\gamma_{j}/2)}&pe^{-i(\alpha_{j}-\gamma_{j}/2)}\end{array}\right],

for j=0,1j=0,1. ρ(θ)=0\rho(\theta)=0 if and only if ρ~(θ)=0\widetilde{\rho}(\theta)=0.

Proof. Suppose ρ(θ)=0\rho(\theta)=0. We have c0Δ1+c1e2iθn=0c_{0}\Delta_{1}+c_{1}e^{2i\theta n}=0. In view of |Δ1|=1|\Delta_{1}|=1, we have |c0|=|c1||c_{0}|=|c_{1}|. It follows from the unitarity of C(x)C(x) that |a0|=|a1||a_{0}|=|a_{1}|, |b0|=|b1||b_{0}|=|b_{1}|, and |d0|=|d1||d_{0}|=|d_{1}|. We obtain (4.11). Since C(x)C(x) is unitary, the equivalence of ρ~(θ)=0\widetilde{\rho}(\theta)=0 and ρ(θ)=0\rho(\theta)=0 is a consequence. ∎

The resonant-tunneling effect for the QW UU is derived as follows.

Theorem 4.7.

Suppose that the matrix C(xj)C(x_{j}) is given by (4.11) for j=0,1j=0,1. If q0q\not=0, we have ρ(θ)=0\rho(\theta)=0 for

θ=β1β0+γ0+π2n+πmn,\theta=\frac{\beta_{1}-\beta_{0}+\gamma_{0}+\pi}{2n}+\frac{\pi m}{n},

where m=0,1,,2n1m=0,1,\ldots,2n-1. If q=0q=0, we have ρ(θ)=0\rho(\theta)=0 for any θ\theta.

Proof. We prove for the case q0q\not=0. In view of the formula (4.11), ρ(θ)=0\rho(\theta)=0 implies

0=c0Δ1+c1e2iθn=qei(γ0β0)eiγ1qei(γ1β1)e2iθn.0=c_{0}\Delta_{1}+c_{1}e^{2i\theta n}=-qe^{i(\gamma_{0}-\beta_{0})}e^{i\gamma_{1}}-qe^{i(\gamma_{1}-\beta_{1})}e^{2i\theta n}.

Then we have e2iθn=ei(β1β0+γ0+π)e^{2i\theta n}=e^{i(\beta_{1}-\beta_{0}+\gamma_{0}+\pi)}. We have proven the theorem. ∎

As a direct consequence of Theorem 4.7, we see that an inverse scattering problem can be solved. Namely, we can determine x1=nx_{1}=n by the number of θ\theta such that ρ(θ)=0\rho(\theta)=0 for q0q\not=0.

Corollary 4.8.

Suppose that the matrix C(xj)C(x_{j}) is given by (4.11) for j=0,1j=0,1, and q0q\not=0. If there exist 2n2n number of θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}} such that ρ(θ)=0\rho(\theta)=0, we have x1=nx_{1}=n.

4.3. Incident wave always passes through a barrier

As has been seen in the previous subsection, ρ(θ)\rho(\theta) may vanish for some θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}. At the end of this paper, let us show that τ(θ)\tau(\theta) does not vanish for all θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}} under the assumption (A-2).

Theorem 4.9.

Suppose that C(x)C(x) is given by (4.1). The transmission coefficient τ(θ)\tau(\theta) does not vanish for all θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}.

Proof. Suppose τ(θ)=0\tau(\theta)=0 for some θJ0J0,𝒯\theta\in J_{0}\setminus J_{0,\mathcal{T}}. In view of (4.2), we have (+(θ)ϕ)(x)=0(\mathcal{F}_{+}(\theta)^{*}\phi)(x)=0 for x1x\leq-1. Since u:=+(θ)ϕu:=\mathcal{F}_{+}(\theta)^{*}\phi is a generalized eigenfunction of UU, we have

[eiθ0c(x)d(x)][uL(x)uR(x)][a(x+1)b(x+1)0eiθ][uL(x+1)uR(x+1)]=0.\left[\begin{array}[]{cc}e^{i\theta}&0\\ c(x)&d(x)\end{array}\right]\left[\begin{array}[]{c}u_{L}(x)\\ u_{R}(x)\end{array}\right]-\left[\begin{array}[]{cc}a(x+1)&b(x+1)\\ 0&e^{i\theta}\end{array}\right]\left[\begin{array}[]{c}u_{L}(x+1)\\ u_{R}(x+1)\end{array}\right]=0.

It follows

(4.12) [uL(x+1)uR(x+1)]=[a(x+1)b(x+1)0eiθ]1[eiθ0c(x)d(x)][uL(x)uR(x)],\left[\begin{array}[]{c}u_{L}(x+1)\\ u_{R}(x+1)\end{array}\right]=\left[\begin{array}[]{cc}a(x+1)&b(x+1)\\ 0&e^{i\theta}\end{array}\right]^{-1}\left[\begin{array}[]{cc}e^{i\theta}&0\\ c(x)&d(x)\end{array}\right]\left[\begin{array}[]{c}u_{L}(x)\\ u_{R}(x)\end{array}\right],

for any x𝐙x\in{\bf Z}. Note that the inverse matrix on the right-hand side exists due to the assumption (A-2). Since u(x)=0u(x)=0 for x1x\leq-1, we have u(x)=0u(x)=0 for any x𝐙x\in{\bf Z} by using the equation (4.12). This is a contradiction. ∎

If the assumption (A-2) does not hold, the incident wave is reflected at a point. This is a trivial case where there exist complete reflection phenomena of quantum walkers. For example, we take

C(x)={[0bc0],x=0,n,[1001],x0,n,\displaystyle C(x)=\left\{\begin{split}\left[\begin{array}[]{cc}0&b\\ c&0\end{array}\right]&,\quad x=0,n,\\ \left[\begin{array}[]{cc}1&0\\ 0&1\end{array}\right]&,\quad x\not=0,n,\end{split}\right.

where |b|=|c|=1|b|=|c|=1. Then a quantum walker is reflected at x=0x=0 and nn. Moreover, there exists an eigenfunction uu of UU such that suppu[0,n]\mathrm{supp}u\subset[0,n] (see [19]). The matrix C(x)C(x) gives a model of non-penetrable barrier at x=0x=0 and nn. The assumption (A-2) guarantees the penetrability condition of barriers given by C(x)C(x).

Acknowledgement. The authors greatly appreciate a valuable comment by Dr. Kenta Higuchi. In particular, his comment is helpful for the correction of Lemma 4.2. H. Morioka is supported by the JSPS Grant-in-aid for young scientists (B) No. 16K17630 and the JSPS Grant-in-aid for young scientists No. 20K14327. E. Segawa is supported by the JSPS Grant-in-Aid for Scientific Research (C) No. 19K03616 and Research Origin for Dressed Photon.

Appendix A Complex contour integration

Here we compute the integration

(A.1) I(x,κ)=ππeixηcosη+1pcos(κγ/2)𝑑η,x𝐙,κ𝐂𝐑,I(x,\kappa)=\int_{-\pi}^{\pi}\frac{e^{ix\eta}}{-\cos\eta+\frac{1}{p}\cos(\kappa-\gamma/2)}d\eta,\quad x\in{\bf Z},\quad\kappa\in{\bf C}\setminus{\bf R},

by using the residue theorem. In order to use the complex contour integral, we extend the integrand to the complex variable

F(z,x,κ)=eixzcosz+1pcos(κγ/2),z𝐓𝐂.F(z,x,\kappa)=\frac{e^{ixz}}{-\cos z+\frac{1}{p}\cos(\kappa-\gamma/2)},\quad z\in{\bf T}_{{\bf C}}.

The integrand F(z,x,κ)F(z,x,\kappa) has simple poles at z=±ζ(κ)z=\pm\zeta(\kappa) where

(A.2) ζ(κ)=arccos(1pcos(κγ2)).\zeta(\kappa)=\arccos\left(\frac{1}{p}\cos\left(\kappa-\frac{\gamma}{2}\right)\right).

Moreover, we have

Resz=±ζ(κ)F(z,x,κ)=±e±ixζ(κ)sinζ(κ).{\mathop{{\rm Res}}_{z=\pm\zeta(\kappa)}}F(z,x,\kappa)=\pm\frac{e^{\pm ix\zeta(\kappa)}}{\sin\zeta(\kappa)}.

We assume Imζ(κ)>0\mathrm{Im}\,\zeta(\kappa)>0. In order to compute I(x,κ)I(x,\kappa), we consider the complex contour integral

I𝒞(x,κ)=𝒞F(z,x,κ)𝑑z,I_{\mathcal{C}}(x,\kappa)=\int_{\mathcal{C}}F(z,x,\kappa)dz,

where 𝒞=j=03𝒞j\mathcal{C}=\sum_{j=0}^{3}\mathcal{C}_{j} with

𝒞0={w=t;t:ππ},𝒞1={w=π+it;t:0ρ},𝒞2={w=t+iρ;t:ππ},𝒞3={w=π+it;t:ρ0},\displaystyle\begin{split}&\mathcal{C}_{0}=\{w=t\ ;\ t:-\pi\to\pi\},\quad\mathcal{C}_{1}=\{w=\pi+it\ ;\ t:0\to\rho\},\\ &\mathcal{C}_{2}=\{w=t+i\rho\ ;\ t:\pi\to-\pi\},\quad\mathcal{C}_{3}=\{w=-\pi+it\ ;\ t:\rho\to 0\},\end{split}

for sufficiently large ρ>0\rho>0. Due to the residue theorem, we have

I𝒞(x,κ)=2πieixζ(κ)sinζ(κ).I_{\mathcal{C}}(x,\kappa)=\frac{2\pi ie^{ix\zeta(\kappa)}}{\sin\zeta(\kappa)}.

In view of the periodicity of the integrand, we can see

𝒞1+𝒞3F(z,x,κ)𝑑z=0.\int_{\mathcal{C}_{1}}+\int_{\mathcal{C}_{3}}F(z,x,\kappa)dz=0.

We also have

|𝒞2F(z,x,κ)𝑑z|ceρ(x+1),\left|\int_{\mathcal{C}_{2}}F(z,x,\kappa)dz\right|\leq ce^{-\rho(x+1)},

for a constant c>0c>0. This implies

𝒞2F(z,x,κ)𝑑z0,\int_{\mathcal{C}_{2}}F(z,x,\kappa)dz\to 0,

as ρ\rho\to\infty for x0x\geq 0. Now we obtain

(A.3) I(x,κ)=2πieixζ(κ)sinζ(κ),x0.I(x,\kappa)=\frac{2\pi ie^{ix\zeta(\kappa)}}{\sin\zeta(\kappa)},\quad x\geq 0.

Let us turn to the complex contour integral

I𝒞(x,κ)=2πieixζ(κ)sinζ(κ),I_{\mathcal{C}^{\prime}}(x,\kappa)=-\frac{2\pi ie^{-ix\zeta(\kappa)}}{\sin\zeta(\kappa)},

for 𝒞=j=03𝒞j\mathcal{C}^{\prime}=\sum_{j=0}^{3}\mathcal{C}^{\prime}_{j} with

𝒞0={w=t;t:ππ},𝒞1={w=πit;t:0ρ},𝒞2={w=tiρ;t:ππ},𝒞3={w=πit;t:ρ0},\displaystyle\begin{split}&\mathcal{C}^{\prime}_{0}=\{w=t\ ;\ t:\pi\to-\pi\},\quad\mathcal{C}^{\prime}_{1}=\{w=-\pi-it\ ;\ t:0\to\rho\},\\ &\mathcal{C}^{\prime}_{2}=\{w=t-i\rho\ ;\ t:-\pi\to\pi\},\quad\mathcal{C}^{\prime}_{3}=\{w=\pi-it\ ;\ t:\rho\to 0\},\end{split}

for sufficiently large ρ>0\rho>0. By the similar way for the case 𝒞\mathcal{C}, we can see

(A.4) I(x,κ)=2πieixζ(κ)sinζ(κ),x0.I(x,\kappa)=\frac{2\pi ie^{-ix\zeta(\kappa)}}{\sin\zeta(\kappa)},\quad x\leq 0.

Plugging (A.3) and (A.4), we obtain the following fact for Imζ(κ)>0\mathrm{Im}\,\zeta(\kappa)>0. For the case Imζ(κ)<0\mathrm{Im}\,\zeta(\kappa)<0, the proof is similar.

Lemma A.1.

For x𝐙x\in{\bf Z} and κ𝐂𝐑\kappa\in{\bf C}\setminus{\bf R}, we have

(A.5) I(x,κ)=±2πie±i|x|ζ(κ)sinζ(κ)for±Imζ(κ)>0.I(x,\kappa)=\pm\frac{2\pi ie^{\pm i|x|\zeta(\kappa)}}{\sin\zeta(\kappa)}\quad\text{for}\quad\pm\mathrm{Im}\,\zeta(\kappa)>0.

References

  • [1] S. Agmon and L. Hörmander, Asymptotic properties of solutions of differential equations with simple characteristics, J. Anal. Math., 30 (1976), 1-30.
  • [2] A. Ambainis, Quantum walks and their algorithmic applications, Int. J. Quantum Inf., 1 (2003), 507-518.
  • [3] J. Asch, O. Bourget and A. Joye, Spectral stability of unitary network models, Rev. Math. Phys., 27 (2015), 1530004.
  • [4] A. M. Childs, Universal computation by quantum walk, Phys. Rev. Lett. 102 (2009), 180501.
  • [5] E. Feldman and M. Hillery, Quantum walks on graphs and quantum scattering theory, Coding Theory and Quantum Computing, edited by D. Evans, J. Holt, C. Jones, K. Klintworth, B. Parshall, O. Pfister, and H. Ward, Contemporary Mathematics, 381 (2005), 71-96.
  • [6] E. Feldman and M. Hillery, Modifying quantum walks: A scattering theory approach, Journal of Physics A: Mathematical and Theoretical 40 (2007), 11343-11359.
  • [7] R. P. Feynman and A. R. Hibbs, “Quantum Mechanics and Path Integrals”, Dover Publications, Inc., Mineola, NY, emended edition, 2010.
  • [8] Yu. Higuchi and E. Segawa, Dynamical system induced by quantum walks, J. Phys. A: Math. Theor., 52 (2009), 395202.
  • [9] H. Kawai, T. Komatsu and N. Konno, Stationary measure for two-state space-inhomogenous quantum walk in one dimension, Yokohama Math. J., 64 (2018), 111-130.
  • [10] T. Komatsu and N. Konno, Stationary amplitudes of quantum walks on the higher-dimensional integer lattice, Quantum Inf. Process., 16 (2017), 291.
  • [11] T. Komatsu, N. Konno, H. Morioka and E. Segawa, Asymptotic properties of generalized eigenfunctions for multi-dimensional quantum walks, in preparation.
  • [12] N. Konno, Quantum Walks, Lecture Notes in Mathematics, 1954 (2008), 309-452, Springer-Verlag, Heidelberg.
  • [13] N. Konno and M. Takei, The non-uniform stationary measures for discrete-time quantum walks in one dimension, Quantum Inf. Comp., 15 (2015), 1060-1075.
  • [14] S. Kuroda, Finite-dimensional perturbation and a representation of scattering operator, Pacific J. Math., 13 (1963), 1305-1318.
  • [15] M. Maeda, H. Sasaki, E. Segawa, A. Suzuki and K. Suzuki, Scattering and inverse scattering for nonlinear quantum walks, Discrete and Continuous Dynamical Systems A, 38 (2018), 3687-3703.
  • [16] M. Maeda, H. Sasaki, E. Segawa, A. Suzuki and K. Suzuki, Dispersive estimates for quantum walks on 1D lattice, preprint. arXiv:1808.05714
  • [17] K. Matsue, L. Matsuoka, O. Ogurisu and E. Segawa, Resonant-tunneling in discrete-time quantum walk, Quantum Studies: Mathematics and Foundations 6 (2018), 35–44.
  • [18] H. Morioka, Generalized eigenfunctions and scattering matrices for position-dependent quantum walks, Rev. Math. Phys., 31 (2019), 1950019.
  • [19] H. Morioka and E. Segawa, Detection of edge defects by embedded eigenvalues of quantum walks, Quantum Inf. Process., 18 (2019), 283.
  • [20] R. Portugal, “Quantum Walk and Search Algorithms”, 2nd Ed., Springer Nature Switzerland, 2018.
  • [21] S. Richard, A. Suzuki and R. Tiedra de Aldecoa, Quantum walks with an anisotropic coin I: spectral theory, Lett. Math. Phys., 108 (2018), 331-357.
  • [22] A. Suzuki, Asymptotic velocity of a position-dependent quantum walk, Quantum Inf. Process, 15 (2016), 103-119.
  • [23] R. Tiedra de Aldecoa, Stationary scattering theory for unitary operators with an application to quantum walks, J. Funct. Anal., 279 (2020), 108704.
  • [24] D. Yafaev, “Mathematical Scattering Theory: General Theory”, Translations of Mathematical Monographs, Vol. 105, American Mathematical Society, Providence, RI, 2009.