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

inverse spectral problem for the Schrödinger operator on the square lattice

Dongjie Wu1 Chuan-Fu Yang2  and  Natalia Pavlovna Bondarenko3
11footnotetext: Department of Applied Mathematics, School of Mathematics and Statistics, Nanjing University of Science and Technology, Nanjing, 210094, Jiangsu, China, Email: [email protected]22footnotetext: Department of Applied Mathematics, School of Mathematics and Statistics, Nanjing University of Science and Technology, Nanjing, 210094, Jiangsu, China, Email: [email protected]33footnotetext: S.M. Nikolskii Mathematical Institute, Peoples’ Friendship University of Russia (RUDN University), 6 Miklukho-Maklaya Street, Moscow, 117198, Russian Federation, Email: [email protected]

Abstract: We consider an inverse spectral problem on a quantum graph associated with the square lattice. Assuming that the potentials on the edges are compactly supported and symmetric, we show that the Dirichlet-to-Neumann map for a boundary value problem on a finite part of the graph uniquely determines the potentials. We obtain a reconstruction procedure, which is based on the reduction of the differential Schrödinger operator to a discrete one. As a corollary of the main results, it is proved that the S-matrix for all energies in any given open set in the continuous spectrum uniquely specifies the potentials on the square lattice.

Keywords: inverse spectral problem, Schrödinger operator, square lattice, Dirichlet-to-Neumann map, inverse scattering

1. introduction

Recently, there have been a lot of studies on quantum graphs, which are one-dimensional Schrödinger (Sturm-Liouville) operators d2dz2+qe(z)-\frac{d^{2}}{dz^{2}}+q_{e}(z) acting on the edges of a metric graph, while some matching conditions are imposed at the graph vertices. Such operators are used for modeling various processes on graph-like structures in physics, mechanics, chemistry, and other applications. Expositions of spectral theory results for quantum graphs can be found, e.g., in the monographs [13, 8, 9, 12] and references therein.

This paper is mostly focused on inverse spectral and scattering problems. Such problems consist in the reconstruction of unknown operator characteristics from spectral information. Till now, inverse problems have been studied for several types of quantum graphs. Reconstruction of differential operators on compact graphs has been investigated in [6, 7, 10, 16, 23, 24, 28, 29, 30] and other studies. Inverse spectral-scattering problems on non-compact graphs with finite and infinite edges were solved, e.g., in [11, 17, 27].

In recent years, Schrödinger operators on infinite periodic graphs have attracted considerable attention of scholars in connection with applications in material studies and nanotechnology (see [22, 20] and references therein). Spectral properties of such operators were studied in [22, 20, 21, 25, 26] and other papers. We also mention that in [14] some spectral theory issues were considered for infinite quantum graphs of general structure (not necessarily periodic). However, to the best of the authors’ knowledge, there were no studies on inverse problems for differential Schrödinger operators on periodic graphs except for the two preprints [4, 5] of Ando et al. Let us discuss their results in more detail.

The two preprints [4, 5] present the same results on the inverse scattering for the Schrödinger operator on the hexagonal lattice with finitely supported potential. The authors of [4, 5] have shown that, in this case, the scattering matrix (S-matrix) uniquely determines the Dirichlet-to-Neumann (D-N) map for a boundary value problem on a finite part of the graph. Furthermore, the differential (continuous) operator was reduced to a difference (discrete) one and the potentials on the graph edges were reconstructed from the D-N map. Thus, it has been shown that the S-matrix uniquely specifies a finitely supported potential on the hexagonal lattice. The preprints [4, 5] continue the previous studies of their authors [1, 2, 3, 18, 19], in which analogous ideas and methods were developed for the inverse scattering on discrete periodic graphs. The difference between [4] and [5] is that, in [5], the preliminary steps up to determining the D-N map by the S-matrix are implemented for the hexagonal lattice, while in [4] they are provided for a general periodic lattice. This opens a perspective of studying inverse spectral problems for different types of periodic graphs.

This paper is concerned with a family of one-dimensional Schrödinger operators d2dz2+qe(z)-\frac{d^{2}}{dz^{2}}+q_{e}(z) defined on the edges of the square lattice as in Figure 1, assuming the Kirchhoff conditions at the vertices (the details are given in Section 2). Here, zz varies over the interval (0,1)(0,1) and eEe\in E, where EE is the set of all edges of the square lattice.

Let us impose the following assumptions on the potentials.

(Q-1) qe(z)q_{e}(z) is real-valued, and qeL2(0,1)q_{e}\in L^{2}(0,1).

(Q-2) qe(z)=0q_{e}(z)=0 on (0,1)(0,1) except for a finite number of edges.

(Q-3) qe(z)=qe(1z)q_{e}(z)=q_{e}(1-z) for z(0,1)z\in(0,1).

Refer to caption
Figure 1.

Since the support of the potential qq is finite, we can choose a sufficiently large square domain DN:={n1+in2:0n1,n2N}D_{N}:=\{n_{1}+\mathrm{i}n_{2}\colon 0\leq n_{1},n_{2}\leq N\} such that suppq\mbox{supp}\,q is located inside of DND_{N} and consider the edge Dirichlet-to-Neumann map ΛE\Lambda_{E} associated with this domain (see Section 2 for details). This paper is devoted to the following inverse spectral problem.

Inverse Problem 1.1.

Given the edge D-N map ΛE\Lambda_{E}, find the potential qq.

The main results of this paper are the uniqueness theorem (Theorem 2.1) for Inverse Problem 1.1 and the reconstruction procedure in Section 5. For solving Inverse Problem 1.1, we use the relation between the edge D-N map and the vertex D-N map, which is defined in Section 3. In other words, we reduce the continuous problem to the discrete one. Consequently, we develop a constructive algorithm for the recovery of the potentials qeq_{e} from the vertex D-N map. Note that, although we follow the general idea of Ando et al, our reconstruction procedure is different from [4, 5]. First, our algorithm is based on specific properties of the square lattice. Second, the authors of [4, 5] apply the strategy of recovering the potentials along any zigzag line, which perfectly works for discrete operators with constant coefficients [3, 19] but causes difficulties for the reduced vertex Laplacian whose coefficients depend on the unknown potentials (see Remark 5.1 for details). In this paper, we step-by-step recover qeq_{e} together with the Laplacian coefficients required at the next steps. For reconstruction of the potential on each fixed edge, we use the classical results of the inverse spectral theory for the Schrödinger operators on finite intervals (see, e.g., [15]). Finally, our results imply the uniqueness for solution of the inverse scattering problem by the S-matrix on the square lattice.

It is worth mentioning that Inverse Problem 1.1 can be treated as the inverse spectral problem on a finite metric graph by the Weyl matrix associated to the boundary vertices. However, our results are novel in this direction. It is well-known the the Weyl matrix uniquely specifies the Schrödinger operator on a tree-graph (see [7, 28]). But, for graphs with cycles, this is not the case in general. Even for the simplest graphs with loops additional data are required (see [24, 30]). Our paper provides a new class of finite quantum graphs whose potentials are uniquely determined by the Weyl matrix. Here, the symmetry of the potentials (Q-3) is crucial.

The paper is organized as follows. In Section 2, we give the definitions related to the edge Schrödinger operator and the edge D-N map. In Section 3, we define the reduced vertex Laplacian and the vertex D-N map, and study the relation between the edge D-N map and the vertex D-N map. In Section 4, some auxiliary solutions of the vertex Schrödinger equation are obtained. In Section 5, we reconstruct the potentials from the D-N map. In Section 6, the inverse scattering by the S-matrix is discussed.

2. Edge Laplacian and Dirichlet-to-Neumann map

Let us define an infinite square lattice with the vertex set

V:={n1+in2:n1,n2}V:=\{n_{1}+\mathrm{i}n_{2}\colon n_{1},n_{2}\in\mathbb{Z}\}

and the edge set

E:={(v,v+1),(v,v+i):vV}.E:=\{(v,v+1),(v,v+\mathrm{i})\colon v\in V\}.

In other words, we consider the vertices as points on the complex plane \mathbb{C} and suppose that two vertices are joined by an edge if the distance between them equals 11. For two vertices w,vVw,v\in V, the notation wvw\sim v means that there exists an edge eEe\in E such that v,wv,w are end points of ee.

Let each edge ee be endowed with arclength metric and identified with the interval (0,1):e={(1z)e(0)+ze(1):0z1}(0,1):e=\{(1-z)e(0)+ze(1)\colon 0\leqslant z\leqslant 1\}, where e(0),e(1)Ve(0),e(1)\in V. Put

Ev:=Ev(0)Ev(1),Ev(i):={eE:e(i)=v},i=0,1.E_{v}:=E_{v}(0)\cup E_{v}(1),\quad E_{v}(i):=\{e\in E\colon e(i)=v\},\ i=0,1.

Let E\mathcal{E}\subseteq E. Then, we call f={fe}ef=\{f_{e}\}_{e\in\mathcal{E}} a function on \mathcal{E} if each fef_{e} (ee\in\mathcal{E}) is a function on [0,1][0,1]. We will write that f={fe}e𝒜(E)f=\{f_{e}\}_{e\in\mathcal{E}}\in\mathcal{A}(E) if fe𝒜[0,1]f_{e}\in\mathcal{A}[0,1] for all ee\in\mathcal{E}, where 𝒜\mathcal{A} is any functional class, e.g., 𝒜=C,L2\mathcal{A}=C,L^{2}, etc. For vVv\in V, eEve\in E_{v}\cap\mathcal{E}, and f={fe}ef=\{f_{e}\}_{e\in\mathcal{E}}, introduce the notations

f(v,e)={fe(0),e(0)=v,fe(1),e(1)=v,,ef(v)={fe(0),e(0)=v,fe(1),e(1)=v.f(v,e)=\begin{cases}f_{e}(0),&e(0)=v,\\ f_{e}(1),&e(1)=v,\end{cases},\qquad\partial_{e}f(v)=\begin{cases}-f_{e}^{\prime}(0),&e(0)=v,\\ f_{e}^{\prime}(1),&e(1)=v.\end{cases}

In other words, f(v,e)f(v,e) is the value of the function ff and ef(v)\partial_{e}f(v), of its derivative in the vertex vv along the edge ee.

Let vVv\in V be a vertex such that EvE_{v}\subseteq\mathcal{E}. Then, we say that f={fe}ef=\{f_{e}\}_{e\in\mathcal{E}} satisfies the Kirchhoff conditions at vv if

(K-1) ff is continuous at vv, that is, f(v,e1)=f(v,e2)f(v,e_{1})=f(v,e_{2}) for any e1,e2Eve_{1},e_{2}\in E_{v}. In this case, we denote f(v)=f(v,e)f(v)=f(v,e), eEve\in E_{v}.

(K-2) fC1(Ev)f\in C^{1}(E_{v}) and eEvef(v)=0\sum_{e\in E_{v}}\partial_{e}f(v)=0.

We will also use the notation f(v)=f(v,e)f(v)=f(v,e) if ee is the only edge in \mathcal{E} incident to the vertex vv, that is, Ev={e}\mathcal{E}\cap E_{v}=\{e\}.

Let us define the Schrödinger operator on a finite subgraph of the square lattice. Suppose that the potential q={qe}eEq=\{q_{e}\}_{e\in E} satisfies the assumptions (Q-1), (Q-2), and (Q-3). Let Ω̊\mathring{\Omega} be the vertex set of some connected finite subgraph of the lattice (V,E)(V,E). Denote

Ω:={vΩ̊:wΩ̊:wv},Ω:=Ω̊Ω,\displaystyle\partial\Omega:=\{v\notin\mathring{\Omega}\colon\exists w\in\mathring{\Omega}\colon w\sim v\},\quad\Omega:=\mathring{\Omega}\cup\partial\Omega, (1)
EΩ:={e=(v,w)E:v,wΩand(vΩ̊orwΩ̊)},\displaystyle E_{\Omega}:=\{e=(v,w)\in E\colon v,w\in\Omega\>\text{and}\>(v\in\mathring{\Omega}\>\text{or}\>w\in\mathring{\Omega})\},
E̊Ω:={e=(v,w)EΩ:v,wΩ̊},EΩ:=EΩE̊Ω.\displaystyle\mathring{E}_{\Omega}:=\{e=(v,w)\in E_{\Omega}\colon v,w\in\mathring{\Omega}\},\quad\partial E_{\Omega}:=E_{\Omega}\setminus\mathring{E}_{\Omega}.

Thus, EΩE_{\Omega} is the edge set of the subgraph with the vertices Ω\Omega. The notations Ω̊\mathring{\Omega}, Ω\partial\Omega, E̊Ω\mathring{E}_{\Omega}, and EΩ\partial E_{\Omega} are used for the interior vertices, the boundary vertices, the interior edges, and the boundary edges, respectively, of the subgraph (Ω,EΩ)(\Omega,E_{\Omega}).

Denote by ΔE\Delta_{E} and call the edge Laplacian the second derivative operation along each edge:

ΔEue(z)=d2dz2ue(z),eE.\Delta_{E}u_{e}(z)=\frac{d^{2}}{dz^{2}}u_{e}(z),\quad e\in E.

For the subgraph (Ω,EΩ)(\Omega,E_{\Omega}), consider the Dirichlet boundary value problem for the edge Laplacian

{(ΔE+qeλ)ue=0,inE̊Ω,u=f,onΩ,\begin{cases}(-\Delta_{E}+q_{e}-\lambda)u_{e}=0,\ in\ \mathring{E}_{\Omega},\\ u=f,\ on\ \partial\Omega,\end{cases} (2)

where a function u={ue}eEΩH2(EΩ)u=\{u_{e}\}_{e\in E_{\Omega}}\in H^{2}(E_{\Omega}) is assumed to fulfill the Kirchhoff conditions at every interior vertex vΩ̊v\in\mathring{\Omega} and f={f(v)}vΩf=\{f(v)\}_{v\in\partial\Omega} is a function given on the set of the boundary vertices.

Denote by (ΔE,Ω+qE,Ω)(-\Delta_{E,\Omega}+q_{E,\Omega}) the operator acting by the rule (ΔE+qe)ue(-\Delta_{E}+q_{e})u_{e}, whose domain D(ΔE,Ω+qE,Ω)D(-\Delta_{E,\Omega}+q_{E,\Omega}) is the set of all the functions u={ue}eEΩH2(EΩ)u=\{u_{e}\}_{e\in E_{\Omega}}\in H^{2}(E_{\Omega}) satisfying the Dirichlet condition u(v)=0u(v)=0 at any boundary vertex vΩv\in\partial\Omega and the Kirchhoff conditions at any interior vertex vΩ̊v\in\mathring{\Omega}. By the standard argument, the operator (ΔE,Ω+qE,Ω)(-\Delta_{E,\Omega}+q_{E,\Omega}) is self-adjoint and its spectrum σ(ΔE,Ω+qE,Ω)\sigma(-\Delta_{E,\Omega}+q_{E,\Omega}) is a countable set of real eigenvalues. Below, we assume that

λσ(ΔE,Ω+qE,Ω).\lambda\not\in\sigma(-\Delta_{E,\Omega}+q_{E,\Omega}). (3)

The edge Dirichlet-to-Neumann (D-N) map ΛE,Ω(λ)\Lambda_{E,\Omega}(\lambda) is defined as follows:

ΛE,Ω(λ)f(v)=eue(v),vΩ,\Lambda_{E,\Omega}(\lambda)f(v)=\partial_{e}u_{e}(v),\quad v\in\partial\Omega, (4)

where uu is the solution of the boundary value problem (2) for the boundary data ff and ee is the edge of EΩE_{\Omega} having vv as its end point. Below we assume that the vertex set Ω\Omega is fixed and use the short notation ΛE:=ΛE,Ω\Lambda_{E}:=\Lambda_{E,\Omega}. Note that f={f(v)}vΩf=\{f(v)\}_{v\in\partial\Omega} can be treated as a vector of M\mathbb{C}^{M}, M:=|Ω|M:=|\partial\Omega|, and ΛE(λ)\Lambda_{E}(\lambda), as the (M×M)(M\times M) matrix function such that ΛE(λ)f=g\Lambda_{E}(\lambda)f=g, g={ddzue(v)}vΩMg=\left\{\frac{d}{dz}u_{e}(v)\right\}_{v\in\partial\Omega}\in\mathbb{C}^{M}. The matrix function ΛE(λ)\Lambda_{E}(\lambda) is analytic in λ\lambda satisfying (3).

For simplicity, we consider the domain

Ω̊=DN:={n1+in2:0n1,n2N},\mathring{\Omega}=D_{N}:=\{n_{1}+\mathrm{i}n_{2}\colon 0\leq n_{1},n_{2}\leq N\},

where a natural number NN is chosen so large that suppqE̊Ω\mbox{supp}\,q\subseteq\mathring{E}_{\Omega}. In other words, qe=0q_{e}=0 on all the edges except eE̊Ωe\in\mathring{E}_{\Omega}. In Section 6, we show that the D-N map is uniquely specified by the S-matrix and vice versa, so the form of the domain is actually unimportant.

Consider Inverse Problem 1.1 for the domain Ω\Omega. In fact, we have to find the potential on E̊Ω\mathring{E}_{\Omega}, since on all the other edges qe=0q_{e}=0. Our main result is the following uniqueness theorem.

Theorem 2.1.

Suppose that the potential q={qe}eEq=\{q_{e}\}_{e\in E} on the square lattice (V,E)(V,E) satisfies (Q-1), (Q-2), and (Q-3) and ΛE(λ)\Lambda_{E}(\lambda) is the edge Dirichlet-to-Neumann map for the region Ω\Omega such that Ω̊=DN\mathring{\Omega}=D_{N} and suppqE̊Ω\mbox{supp}\,q\subseteq\mathring{E}_{\Omega}. Then ΛE(λ)\Lambda_{E}(\lambda) uniquely specifies the potential qq.

3. Reduced vertex Laplacian

The proof of Theorem 2.1 is based on the reduction of the edge Laplacian to the vertex one.

Consider the Schrödinger equation

(d2dz2+qe(z)λ)ϕ=0,z(0,1),\left(-\frac{d^{2}}{dz^{2}}+q_{e}(z)-\lambda\right)\phi=0,\quad z\in(0,1), (5)

on each fixed edge eEe\in E. Let ϕe0(z,λ),ϕe1(z,λ)\phi_{e0}(z,\lambda),\phi_{e1}(z,\lambda) be the solutions of (5) with the initial data ϕe0(0,λ)=0,ϕe0(0,λ)=1\phi_{e0}(0,\lambda)=0,\,\phi_{e0}^{\prime}(0,\lambda)=1 and ϕe1(1,λ)=0,ϕe1(1,λ)=1\phi_{e1}(1,\lambda)=0,\,\phi_{e1}^{\prime}(1,\lambda)=-1, respectively.

Denote by HeH_{e} the Schrödinger operator (d2dz2+qe)\left(-\frac{d^{2}}{dz^{2}}+q_{e}\right) with the domain D(He)D(H_{e}) which consists of functions uH2[0,1]u\in H^{2}[0,1] satisfying the Dirichlet conditions u(0)=u(1)=0u(0)=u(1)=0. It is well-known that the operator HeH_{e} is self-adjoint and its spectrum is a countable set of real eigenvalues. In the following, we assume that λσ(He)\lambda\not\in\sigma(H_{e}), eEe\in E. This guarantees that ϕe0(1,λ)0\phi_{e0}(1,\lambda)\neq 0 and ϕe1(0,λ)0\phi_{e1}(0,\lambda)\neq 0.

If w,vVw,v\in V are two end points of an edge eEe\in E, then we denote

ψwv(z,λ)={ϕe0(z,λ),e(0)=v,ϕe1(1z,λ),e(1)=v.\psi_{wv}(z,\lambda)=\begin{cases}\phi_{e0}(z,\lambda),&e(0)=v,\\ \phi_{e1}(1-z,\lambda),&e(1)=v.\end{cases}

Note that, by the assumption (Q-3), we have ϕe0(z,λ)=ϕe1(1z,λ)\phi_{e0}(z,\lambda)=\phi_{e1}(1-z,\lambda), hence ψwv(1,λ)=ψvw(1,λ)\psi_{wv}(1,\lambda)=\psi_{vw}(1,\lambda). In particular, if qe=0q_{e}=0, then ψwv(z,λ)=sinλzλ\psi_{wv}(z,\lambda)=\frac{\sin\sqrt{\lambda}z}{\sqrt{\lambda}}.

We define the reduced vertex Laplacian ΔV,λ\Delta_{V,\lambda} on VV by

(ΔV,λu)(v)=14wv1ψwv(1,λ)u(w),vV,(\Delta_{V,\lambda}u)(v)=\frac{1}{4}\sum_{w\sim v}\frac{1}{\psi_{wv}(1,\lambda)}u(w),\quad v\in V, (6)

for uLloc2(V)u\in L^{2}_{loc}(V). We also define the scalar multiplication operator:

(QV,λu)(v)=qv,λu(v),qv,λ:=14wvψwv(1,λ)ψwv(1,λ).(Q_{V,\lambda}u)(v)=q_{v,\lambda}u(v),\quad q_{v,\lambda}:=\frac{1}{4}\sum_{w\sim v}\frac{\psi_{wv}^{\prime}(1,\lambda)}{\psi_{wv}(1,\lambda)}. (7)
Lemma 3.1.

Let vv be a fixed vertex. If a function u={ue}eEvH2(Ev)u=\{u_{e}\}_{e\in E_{v}}\in H^{2}(E_{v}) satisfies equation (5) for all eEve\in E_{v} and the Kirchhoff conditions in vv, then the relation (ΔV,λ+QV,λ)u=0(-\Delta_{V,\lambda}+Q_{V,\lambda})u=0 holds at vv.

Proof.

Consider the solutions Se(z,λ)S_{e}(z,\lambda) and Ce(z,λ)C_{e}(z,\lambda) of equation of (5) satisfying the initial conditions Se(0,λ)=Ce(0,λ)=0S_{e}(0,\lambda)=C^{\prime}_{e}(0,\lambda)=0 and Se(0,λ)=Ce(0,λ)=1S^{\prime}_{e}(0,\lambda)=C_{e}(0,\lambda)=1. Then, any solution ueu_{e} of (5) can be written as

ue(z)=aeSe(z,λ)+beCe(z,λ)u_{e}(z)=a_{e}S_{e}(z,\lambda)+b_{e}C_{e}(z,\lambda) (8)

where aea_{e} and beb_{e} are constants.

Let vVv\in V be fixed. For convenience, suppose that e(0)=ve(0)=v for all eEve\in E_{v}. Then (8) together with (K-1) imply (see Figure 2):

Refer to caption
Figure 2.
b(v,v+1)=b(v,v1)=b(v,v+i)=b(v,vi)=u(v),b_{(v,v+1)}=b_{(v,v-1)}=b_{(v,v+i)}=b_{(v,v-i)}=u(v),

and (K-2) implies

a(v,v+1)+a(v,v1)+a(v,v+i)+a(v,vi)=0.a_{(v,v+1)}+a_{(v,v-1)}+a_{(v,v+i)}+a_{(v,v-i)}=0.

By (8), we have

ae=ue(1)Se(1,λ)Ce(1,λ)Se(1,λ)u(v).a_{e}=\frac{u_{e}(1)}{S_{e}(1,\lambda)}-\frac{C_{e}(1,\lambda)}{S_{e}(1,\lambda)}u(v).

Thus,

k=1,1,i,ia(v,v+k)=k=1,1,i,iu(v,v+k)(1)S(v,v+k)(1,λ)C(v,v+k)(1,λ)Sv,v+k(1,λ)ue(0)=0.\sum_{k=1,-1,i,-i}a_{(v,v+k)}=\sum_{k=1,-1,i,-i}\frac{u_{(v,v+k)}(1)}{S_{(v,v+k)}(1,\lambda)}-\frac{C_{(v,v+k)}(1,\lambda)}{S_{v,v+k}(1,\lambda)}u_{e}(0)=0. (9)

Note that Se(1,λ)=ψwv(1,λ)S_{e}(1,\lambda)=\psi_{wv}(1,\lambda) and, under the assumption (Q-3), Ce(1,λ)=Se(1,λ)C_{e}(1,\lambda)=S^{\prime}_{e}(1,\lambda). Consequently, we arrive at the relation (ΔV,λ+QV,λ)u(v)=0(-\Delta_{V,\lambda}+Q_{V,\lambda})u(v)=0. ∎

Consider the vertex set Ω\Omega defined by (1) for Ω̊=DN\mathring{\Omega}=D_{N} and the interior boundary value problem

{(ΔV,λ+QV,λ)u=0inΩ̊,u=fonΩ.\begin{cases}(-\Delta_{V,\lambda}+Q_{V,\lambda})u=0\ in\ \mathring{\Omega},\\ u=f\ on\ \partial\Omega.\end{cases} (10)

By virtue of Lemma 3.1, if a function u={ue}eEΩu=\{u_{e}\}_{e\in E_{\Omega}} solves the edge boundary value problem (2), then its values {u(v)}vΩ\{u(v)\}_{v\in\Omega} in the vertices satisfy (10).

Define the vertex degree in the subgraph (Ω,EΩ)(\Omega,E_{\Omega}) as follows:

degΩ(v)={#{wΩ:wv},vΩ̊,#{wΩ̊:wv},vΩ.\mbox{deg}_{\Omega}(v)=\begin{cases}\#\{w\in\Omega\colon w\sim v\},&v\in\mathring{\Omega},\\ \#\{w\in\mathring{\Omega}\colon w\sim v\},&v\in\partial\Omega.\end{cases}

Then, the vertex D-N map is defined by

ΛV,Ω(λ)f(v)=1degΩ(v)wv,wΩ̊u(w),vΩ.\Lambda_{V,\Omega}(\lambda)f(v)=-\frac{1}{\mbox{deg}_{\Omega}(v)}\sum_{w\sim v,w\in\mathring{\Omega}}u(w),\quad v\in\partial\Omega.

Below we use a shorter notation ΛV:=ΛV,Ω\Lambda_{V}:=\Lambda_{V,\Omega}. Note that, for the region Ω=DN\Omega=D_{N}, the degrees degΩ(v)\mbox{deg}_{\Omega}(v) of the boundary vertices vΩv\in\partial\Omega equal to 11. Hence

ΛV(λ)f(v)=u(w),vΩ,wΩ̊,wv.\Lambda_{V}(\lambda)f(v)=-u(w),\quad v\in\partial\Omega,\quad w\in\mathring{\Omega},\quad w\sim v. (11)

The edge and the vertex D-N maps are closely related to each other, which is shown in the following lemma.

Lemma 3.2.

The following equality holds:

ΛV(λ)=cosλI+sinλλΛE(λ).\Lambda_{V}(\lambda)=-\cos\sqrt{\lambda}I+\frac{\sin\sqrt{\lambda}}{\sqrt{\lambda}}\Lambda_{E}(\lambda). (12)

where II is the unit operator in M\mathbb{C}^{M} and λ\lambda satisfies (3).

Proof.

Let e(1)=v,vΩe(1)=v,\ v\in\partial\Omega, then

ΛV(λ)f(v)=ue(0),ΛE(λ)f(v)=ue(1).\Lambda_{V}(\lambda)f(v)=-u_{e}(0),\quad\Lambda_{E}(\lambda)f(v)=u^{\prime}_{e}(1). (13)

By (8), we have

ue(1)=aeSe(1,λ)+ue(0)Ce(1,λ)=f(v),u_{e}(1)=a_{e}S_{e}(1,\lambda)+u_{e}(0)C_{e}(1,\lambda)=f(v), (14)

and

ue(1)=aeSe(1,λ)+ue(0)Ce(1,λ).u^{\prime}_{e}(1)=a_{e}S^{\prime}_{e}(1,\lambda)+u_{e}(0)C^{\prime}_{e}(1,\lambda). (15)

Then

ae=1Se(1,λ)(f(v)ue(0)Ce(1,λ)).a_{e}=\frac{1}{S_{e}(1,\lambda)}(f(v)-u_{e}(0)C_{e}(1,\lambda)). (16)

Substituting (16) into (15) and using the relation

Se(1,λ)Ce(1,λ)Ce(1,λ)Se(1,λ)=1S_{e}(1,\lambda)C^{\prime}_{e}(1,\lambda)-C_{e}(1,\lambda)S^{\prime}_{e}(1,\lambda)=-1

we obtain

ue(1)=f(v)Se(1,λ)Se(1,λ)+(Ce(1,λ)Ce(1,λ)Se(1,λ)Se(1,λ))ue(0)=f(v)Se(1,λ)Se(1,λ)1Se(1,λ)ue(0).\begin{split}u^{\prime}_{e}(1)&=f(v)\frac{S^{\prime}_{e}(1,\lambda)}{S_{e}(1,\lambda)}+(C^{\prime}_{e}(1,\lambda)-\frac{C_{e}(1,\lambda)S^{\prime}_{e}(1,\lambda)}{S_{e}(1,\lambda)})u_{e}(0)\\ &=f(v)\frac{S^{\prime}_{e}(1,\lambda)}{S_{e}(1,\lambda)}-\frac{1}{S_{e}(1,\lambda)}u_{e}(0).\end{split} (17)

By (13) and (17), we get

ΛV(λ)f(v)=f(v)Se(1,λ)+(ΛE(λ)f(v))Se(1,λ).\Lambda_{V}(\lambda)f(v)=-f(v)S^{\prime}_{e}(1,\lambda)+(\Lambda_{E}(\lambda)f(v))S_{e}(1,\lambda). (18)

Note that the potentials on the edge ee connected to the vertex vΩv\in\partial\Omega are zero, so we have

Se(1,λ)=sinλλ,Se(1,λ)=cosλ.S_{e}(1,\lambda)=\frac{\sin\sqrt{\lambda}}{\sqrt{\lambda}},\quad S^{\prime}_{e}(1,\lambda)=\cos\sqrt{\lambda}.

Using these relations together with (18), we arrive at (12).

Now, let e(0)=v,vΩe(0)=v,\ v\in\partial\Omega, then

ΛV(λ)f(v)=ue(1),ΛE(λ)f(v)=ue(0).\Lambda_{V}(\lambda)f(v)=-u_{e}(1),\quad\Lambda_{E}(\lambda)f(v)=-u^{\prime}_{e}(0). (19)

By the boundary condition, we get ue(0)=f(v)u_{e}(0)=f(v). By (8), we have

ue(1)=aeSe(1,λ)+f(v)Ce(1,λ),u_{e}(1)=a_{e}S_{e}(1,\lambda)+f(v)C_{e}(1,\lambda), (20)

and

ue(0)=ae.u^{\prime}_{e}(0)=a_{e}. (21)

Substitute (21) into (20), we obtain

ue(1)=ue(0)Se(1,λ)+f(v)Ce(1,λ).u_{e}(1)=u^{\prime}_{e}(0)S_{e}(1,\lambda)+f(v)C_{e}(1,\lambda). (22)

By (19) and (22), we have

ΛV(λ)f(v)=f(v)Ce(1,λ)+(ΛE(λ)f(v))Se(1,λ).\Lambda_{V}(\lambda)f(v)=-f(v)C_{e}(1,\lambda)+(\Lambda_{E}(\lambda)f(v))S_{e}(1,\lambda).

Note that

Se(1,λ)=sinλλ,Ce(1,λ)=cosλ,S_{e}(1,\lambda)=\frac{\sin\sqrt{\lambda}}{\sqrt{\lambda}},\quad C_{e}(1,\lambda)=\cos\sqrt{\lambda},

so we directly derive (12). ∎

Therefore, Inverse Problem 1.1 of recovering the potential from the edge D-N map can be easily reduced to the following inverse problem.

Inverse Problem 3.1.

Given the vertex D-N map ΛV\Lambda_{V}, find the potential qq.

4. Special solutions of the vertex Schrödinger equation

We are now in a position to construct the solution of Inverse Problem 3.1. For this purpose, we first obtain some special solutions of the vertex Schrödinger equation

(ΔV,λ+QV,λ)u=0.(-\Delta_{V,\lambda}+Q_{V,\lambda})u=0.

The boundary Ω\partial\Omega of the domain Ω̊=DN\mathring{\Omega}=D_{N} consists of the four parts, see Figure 3:

Refer to caption
Figure 3. N=4
(Ω)T={(N+1)i+m:0mN}={tm}m=0N,\displaystyle(\partial\Omega)_{T}=\{(N+1)\mathrm{i}+m\colon 0\leqslant m\leqslant N\}=\{t_{m}\}_{m=0}^{N},
(Ω)B={i+m:0mN}={bm}m=0N,\displaystyle(\partial\Omega)_{B}=\{-\mathrm{i}+m\colon 0\leqslant m\leqslant N\}=\{b_{m}\}_{m=0}^{N},
(Ω)L={1+mi:0mN}={lm}m=0N,\displaystyle(\partial\Omega)_{L}=\{-1+m\mathrm{i}\colon 0\leqslant m\leqslant N\}=\{l_{m}\}_{m=0}^{N},
(Ω)R={N+1+mi:0mN}={rm}m=0N.\displaystyle(\partial\Omega)_{R}=\{N+1+m\mathrm{i}\colon 0\leqslant m\leqslant N\}=\{r_{m}\}_{m=0}^{N}.

Recall that ΛV=ΛV,Ω\Lambda_{V}=\Lambda_{V,\Omega} is the vertex D-N map defined by (11). The key to the inverse procedure is the following partial data problem.

Lemma 4.1.

(1) Given a partial Dirichlet data ff on Ω\(Ω)R\partial\Omega\backslash(\partial\Omega)_{R}, and a partial Neumann data gg on (Ω)L(\partial\Omega)_{L}, there is a unique solution uu in Ω\Omega of the boundary value problem

{(ΔV,λ+QV,λ)u=0inΩ̊,u=fonΩ\(Ω)R,vΩu=gon(Ω)L.\begin{cases}(-\Delta_{V,\lambda}+Q_{V,\lambda})u=0\ in\ \mathring{\Omega},\\ u=f\ on\ \partial\Omega\backslash(\partial\Omega)_{R},\\ \partial_{v}^{\Omega}u=g\ on\ (\partial\Omega)_{L}.\end{cases} (23)

where

vΩu=1degΩ(v)wv,wΩ̊u(w).\partial_{v}^{\Omega}u=-\frac{1}{\mbox{deg}_{\Omega}(v)}\sum_{w\sim v,w\in\mathring{\Omega}}u(w).

(2) Given the D-N map ΛV\Lambda_{V}, a partial Dirichlet data f1f_{1} on Ω\(Ω)R\partial\Omega\backslash(\partial\Omega)_{R} and a partial Neumann data gg on (Ω)L(\partial\Omega)_{L}, there exists a unique ff on Ω\partial\Omega such that f=f1f=f_{1} on Ω\(Ω)R\partial\Omega\backslash(\partial\Omega)_{R} and ΛVf=g\Lambda_{V}f=g on (Ω)L(\partial\Omega)_{L}.

Proof.

(1) The values of u(mi),m=0,,Nu(m\mathrm{i}),\ m=0,\cdots,N are computed from the values of gg. Using equation (23), the Dirichlet data ff and the known values of uu, one can then compute u(1+mi),m=0,,Nu(1+m\mathrm{i}),\ m=0,\cdots,N. Next, we obtain the values of u(2+mi),m=0,,Nu(2+m\mathrm{i}),\ m=0,\cdots,N. Repeating this procedure, we get u(v)u(v) for all vΩv\in\Omega.

(2) For subsets A,BΩA,B\subset\partial\Omega, we denote the associated submatrix of ΛV\Lambda_{V} by ΛV(A;B)\Lambda_{V}(A;B). Suppose f2=0f_{2}=0 on Ω\(Ω)R\partial\Omega\backslash(\partial\Omega)_{R} and ΛVf2=0\Lambda_{V}f_{2}=0 on (Ω)L(\partial\Omega)_{L}. By (1) and the boundary conditions, the solution uu vanishes in Ω\Omega identically. Hence f2=0f_{2}=0 on (Ω)R(\partial\Omega)_{R}. This implies the submatrix

ΛV((Ω)L;(Ω)R):(Ω)R(Ω)L\Lambda_{V}((\partial\Omega)_{L};(\partial\Omega)_{R}):(\partial\Omega)_{R}\longrightarrow(\partial\Omega)_{L}

is a bijection. We seek ff in the form

(ΛVf)|(Ω)L=ΛV((Ω)L;(Ω)R)f3+ΛV((Ω)L;Ω\(Ω)R)f1=g,(\Lambda_{V}f)\big{|}_{(\partial\Omega)_{L}}=\Lambda_{V}((\partial\Omega)_{L};(\partial\Omega)_{R})f_{3}+\Lambda_{V}((\partial\Omega)_{L};\partial\Omega\backslash(\partial\Omega)_{R})f_{1}=g,

where

f3=(ΛV((Ω)L;(Ω)R))1(gΛV((Ω)L;Ω\(Ω)R)f1).f_{3}=\left(\Lambda_{V}((\partial\Omega)_{L};(\partial\Omega)_{R})\right)^{-1}\left(g-\Lambda_{V}((\partial\Omega)_{L};\partial\Omega\backslash(\partial\Omega)_{R})f_{1}\right).

This proves the lemma. ∎

Now, for N+1k2NN+1\leqslant k\leqslant 2N, let us consider the diagonal line

Ak={x1+ix2:x1+x2=k}.A_{k}=\{x_{1}+\mathrm{i}x_{2}\colon x_{1}+x_{2}=k\}. (24)

The vertices on AkΩA_{k}\cap\Omega are written as

αk,l=αk,0+l(1i),l=0,1,2,,2N+2k,\alpha_{k,l}=\alpha_{k,0}+l(1-\mathrm{i}),\quad l=0,1,2,\dots,2N+2-k, (25)

where αk,0=tk(N+1)=k(N+1)+i(N+1)\alpha_{k,0}=t_{k-(N+1)}=k-(N+1)+\mathrm{i}(N+1).

Lemma 4.2.

Let AkΩ={αk,0,αk,m},k=N+1,,2NA_{k}\cap\partial\Omega=\{\alpha_{k,0},\alpha_{k,m}\},k=N+1,\cdots,2N. Then, there exists a unique solution uu in Ω\Omega of the boundary value problem

(ΔV,λ+QV,λ)u=0inΩ̊,(-\Delta_{V,\lambda}+Q_{V,\lambda})u=0\ in\ \mathring{\Omega}, (26)

with partial Dirichlet data ff such that

{f(αk,0)=1,f(z)=0,forzΩ\((Ω)Rαk,0)\begin{cases}f(\alpha_{k,0})=1,\\ f(z)=0,\ for\ z\in\partial\Omega\backslash\left((\partial\Omega)_{R}\cup\alpha_{k,0}\right)\end{cases} (27)

and partial Neumann data g=0g=0 on (Ω)L(\partial\Omega)_{L}. It satisfies

u(x1+ix2)=0ifx1+x2<k.u(x_{1}+\mathrm{i}x_{2})=0\ if\ x_{1}+x_{2}<k. (28)
Proof.

The uniqueness and the existence of uu on Ω\Omega follow from Lemma 4.1. By the condition on f,gf,g, one can compute u(x1+ix2)u(x_{1}+\mathrm{i}x_{2}) successively to obtain (28). ∎

An important feature of the solution uu in Lemma 4.2 is that uu vanishes below the line AkA_{k}. For such solutions, we obtain the following property.

Let uu be a solution of the equation

(ΔV,λ+QV,λ)u=0inΩ̊,(-\Delta_{V,\lambda}+Q_{V,\lambda})u=0\ in\ \mathring{\Omega}, (29)

which vanishes in the region x1+x2<kx_{1}+x_{2}<k. Let a,b,b,c,dVa,b,b^{\prime},c,d\in V and e,eEe,e^{\prime}\in E be as in Figure 4.

Refer to caption
Figure 4.

Then, evaluating the equation (29) at v=av=a and using (6), (7), we obtain

1ψba(1,λ)u(b)+1ψba(1,λ)u(b)=0.\frac{1}{\psi_{ba}(1,\lambda)}u(b)+\frac{1}{\psi_{b^{\prime}a}(1,\lambda)}u(b^{\prime})=0. (30)

The special solutions of Lemma 4.2 and their property (30) play a crucial role in the proof of the main theorem and in the reconstruction procedure in the next section.

5. Reconstruction procedure

The goal of this section is to prove Theorem 2.1 on the uniqueness of the inverse spectral problem solution. The proof consists in a reconstruction procedure, which is based on the reduction to the discrete inverse problem by the vertex D-N map (i.e. Inverse Problem 3.1), on specific properties of the square lattice, and on applying the classical results for the recovery of the Schrödinger potential on each edge.

Before proceeding to the reconstruction, we provide two well-known results for inverse spectral problems on a finite interval. Let e=(w,v)e=(w,v) be a fixed edge. Clearly, the eigenvalues of the Dirichlet problem

(d2dz2+qe(z)λ)u=0,u(0)=u(1)=0\left(-\frac{d^{2}}{dz^{2}}+q_{e}(z)-\lambda\right)u=0,\quad u(0)=u(1)=0

coincide with the zeros of the characteristic function ψwv(1,λ)\psi_{wv}(1,\lambda).

Lemma 5.1.

([15, Theorem 1.4.3]) If the potential qeq_{e} is symmetric, then the Dirichlet eigenvalues for the operator d2dz2+qe(z)-\frac{d^{2}}{dz^{2}}+q_{e}(z) on (0,1)(0,1) uniquely determine the potential qeq_{e}. In other words, qeq_{e} is uniquely specified by the function ψwv(1,λ)\psi_{wv}(1,\lambda).

Lemma 5.2.

([15, Theorem 1.4.7]) The Weyl function ψwv(1,λ)ψwv(1,λ)\frac{\psi^{\prime}_{wv}(1,\lambda)}{\psi_{wv}(1,\lambda)} uniquely specifies the potential qeq_{e}.

The both inverse problems of Lemmas 5.1 and 5.2 can be solved constructively by the well-known methods, e.g., the Gelfand-Levitan method and the method of spectral mappings (see [15]).

Now, let us prove Theorem 2.1 by providing a reconstruction procedure for solving Inverse Problem 1.1.

Reconstruction procedure. Suppose that the potential qq on the square lattice fulfills the conditions (Q-1), (Q-2), and (Q-3). Let Ω\Omega be the square region defined in Section 2 and let the corresponding edge D-N map ΛE\Lambda_{E} be given.

Step 1. We obtain the vertex D-N map ΛV\Lambda_{V} by the formula (12).

Step 2. For k=2N,2N1,,N+1k=2N,2N-1,\dots,N+1, implement the following steps 3–5.

Step 3. For the value of kk fixed at step 2, draw the line AkA_{k} defined by (24) and take the boundary data ff having the properties in Lemma 4.2. Under the assumption (Q-2), we have qe(z)=0q_{e}(z)=0 on all the boundary edges EΩ\partial E_{\Omega}. So, we know the functions ψwv(1,λ)=sinλλ\psi_{wv}(1,\lambda)=\frac{\sin\sqrt{\lambda}}{\sqrt{\lambda}} for (w,v)EΩ(w,v)\in\partial E_{\Omega}. By Lemma 4.1 (2), we can find uu on (Ω)R(\partial\Omega)_{R} with the vertex D-N map ΛV\Lambda_{V}. Thus, we know uu on Ω\partial\Omega.

Step 4. If k=2Nk=2N, using (11) and the value of u(αk,1+i)u(\alpha_{k,1}+\mathrm{i}), we find

u(αk,1)=ΛVu(αk,1+i),u(\alpha_{k,1})=-\Lambda_{V}u(\alpha_{k,1}+\mathrm{i}),

By (30), we have

u(αk,0)ψαk,0i,αk,0(1,λ)+u(αk,1)ψαk,11,αk,1(1,λ)=0,\frac{u(\alpha_{k,0})}{\psi_{\alpha_{k,0}-\mathrm{i},\alpha_{k,0}}(1,\lambda)}+\frac{u(\alpha_{k,1})}{\psi_{\alpha_{k,1}-1,\alpha_{k,1}}(1,\lambda)}=0, (31)

and

u(αk,1)ψαk,1i,αk,1(1,λ)+u(αk,2)ψαk,21,αk,2(1,λ)=0.\frac{u(\alpha_{k,1})}{\psi_{\alpha_{k,1}-\mathrm{i},\alpha_{k,1}}(1,\lambda)}+\frac{u(\alpha_{k,2})}{\psi_{\alpha_{k,2}-1,\alpha_{k,2}}(1,\lambda)}=0. (32)

By the known values u(αk,0)u(\alpha_{k,0}), u(αk,1)u(\alpha_{k,1}), u(αk,2)u(\alpha_{k,2}), ψαk,1i,αk,1(1,λ)\psi_{\alpha_{k,1}-\mathrm{i},\alpha_{k,1}}(1,\lambda), ψαk,21,αk,2(1,λ)\psi_{\alpha_{k,2}-1,\alpha_{k,2}}(1,\lambda), we can find

ψαk,11,αk,1(1,λ)=u(αk,1)u(αk,0)ψαk,0i,αk,0(1,λ),\psi_{\alpha_{k,1}-1,\alpha_{k,1}}(1,\lambda)=-\frac{u(\alpha_{k,1})}{u(\alpha_{k,0})}\psi_{\alpha_{k,0}-\mathrm{i},\alpha_{k,0}}(1,\lambda), (33)

and

ψαk,1i,αk,1(1,λ)=u(αk,1)u(αk,2)ψαk,21,αk,2(1,λ).\psi_{\alpha_{k,1}-\mathrm{i},\alpha_{k,1}}(1,\lambda)=-\frac{u(\alpha_{k,1})}{u(\alpha_{k,2})}\psi_{\alpha_{k,2}-1,\alpha_{k,2}}(1,\lambda). (34)

Note that the zeros of ψαk,11,αk,1(1,λ)\psi_{\alpha_{k,1}-1,\alpha_{k,1}}(1,\lambda) and ψαk,1i,αk,1(1,λ)\psi_{\alpha_{k,1}-\mathrm{i},\alpha_{k,1}}(1,\lambda) are the Dirichlet eigenvalues for the operator d2dz2+qe(z)-\frac{d^{2}}{dz^{2}}+q_{e}(z) on (0,1)(0,1). Since the potential is symmetric, by Lemma 5.1, these eigenvalues uniquely determine the potentials qe(z)q_{e}(z) on (αk,11,αk,1)(\alpha_{k,1}-1,\alpha_{k,1}) and (αk,1i,αk,1)(\alpha_{k,1}-\mathrm{i},\alpha_{k,1}).

Step 5. If k2N1k\leq 2N-1, then implement steps 5.1–5.6.

Step 5.1 Using (11) and the values of u(αk,1+i)u(\alpha_{k,1}+\mathrm{i}), u(αk,2N+1k+1)u(\alpha_{k,2N+1-k}+1), we find

u(αk,1)=ΛVu(αk,1+i),u(αk,2N+1k)=ΛVu(αk,2N+1k+1).u(\alpha_{k,1})=-\Lambda_{V}u(\alpha_{k,1}+\mathrm{i}),\quad u(\alpha_{k,2N+1-k})=-\Lambda_{V}u(\alpha_{k,2N+1-k}+1).

By (30), we get

ψαk,11,αk,1(1,λ)=u(αk,1)u(αk,0)ψαk,0i,αk,0(1,λ),\psi_{\alpha_{k,1}-1,\alpha_{k,1}}(1,\lambda)=-\frac{u(\alpha_{k,1})}{u(\alpha_{k,0})}\psi_{\alpha_{k,0}-\mathrm{i},\alpha_{k,0}}(1,\lambda), (35)

and

ψαk,2N+1ki,αk,2N+1k(1,λ)=u(αk,2N+1k)u(αk,2N+2k)ψαk,2N+2k1,αk,2N+2k(1,λ).\psi_{\alpha_{k,2N+1-k}-\mathrm{i},\alpha_{k,2N+1-k}}(1,\lambda)=-\frac{u(\alpha_{k,2N+1-k})}{u(\alpha_{k,2N+2-k})}\psi_{\alpha_{k,2N+2-k}-1,\alpha_{k,2N+2-k}}(1,\lambda). (36)

Then, by Lemma 5.1, the zeros of ψαk,11,αk,1(1,λ)\psi_{\alpha_{k,1}-1,\alpha_{k,1}}(1,\lambda) and ψαk,2N+1ki,αk,2N+1k(1,λ)\psi_{\alpha_{k,2N+1-k}-\mathrm{i},\alpha_{k,2N+1-k}}(1,\lambda) uniquely determine the potentials on (αk,11,αk,1)(\alpha_{k,1}-1,\alpha_{k,1}) and (αk,2N+1ki,αk,2N+1k)(\alpha_{k,2N+1-k}-\mathrm{i},\alpha_{k,2N+1-k}).

Step 5.2. Evaluating equation (29) at αk,1\alpha_{k,1}, we obtain

14(u(αk,1+i)ψαk,1,αk,1+i(1,λ)+u(αk,1+1)ψαk,1,αk,1+1(1,λ))+qv,λ(αk,1)u(αk,1)=0.-\frac{1}{4}\left(\frac{u(\alpha_{k,1}+\mathrm{i})}{\psi_{\alpha_{k,1},\alpha_{k,1}+\mathrm{i}}(1,\lambda)}+\frac{u(\alpha_{k,1}+1)}{\psi_{\alpha_{k,1},\alpha_{k,1}+1}(1,\lambda)}\right)+q_{v,\lambda}(\alpha_{k,1})u(\alpha_{k,1})=0.

Then, we can get the value of qv,λ(αk,1)q_{v,\lambda}(\alpha_{k,1}). By (7), we have

qv,λ(αk,1)\displaystyle q_{v,\lambda}(\alpha_{k,1}) =14(ψαk,11,αk,1(1,λ)ψαk,11,αk,1(1,λ)+ψαk,1,αk,1+1(1,λ)ψαk,1,αk,1+1(1,λ)\displaystyle=\frac{1}{4}\Biggl{(}\frac{\psi^{\prime}_{\alpha_{k,1}-1,\alpha_{k,1}}(1,\lambda)}{\psi_{\alpha_{k,1}-1,\alpha_{k,1}}(1,\lambda)}+\frac{\psi^{\prime}_{\alpha_{k,1},\alpha_{k,1}+1}(1,\lambda)}{\psi_{\alpha_{k,1},\alpha_{k,1}+1}(1,\lambda)}
+ψαk,1i,αk,1(1,λ)ψαk,1i,αk,1(1,λ)+ψαk,1,αk,1+i(1,λ)ψαk,1,αk,1+i(1,λ))\displaystyle+\frac{\psi^{\prime}_{\alpha_{k,1}-\mathrm{i},\alpha_{k,1}}(1,\lambda)}{\psi_{\alpha_{k,1}-\mathrm{i},\alpha_{k,1}}(1,\lambda)}+\frac{\psi^{\prime}_{\alpha_{k,1},\alpha_{k,1}+\mathrm{i}}(1,\lambda)}{\psi_{\alpha_{k,1},\alpha_{k,1}+\mathrm{i}}(1,\lambda)}\Biggr{)}

where the values of ψαk,11,αk,1(1,λ)\psi_{\alpha_{k,1}-1,\alpha_{k,1}}(1,\lambda), ψαk,1,αk,1+i(1,λ)\psi_{\alpha_{k,1},\alpha_{k,1}+\mathrm{i}}(1,\lambda) and ψαk,1,αk,1+1(1,λ)\psi_{\alpha_{k,1},\alpha_{k,1}+1}(1,\lambda) are known.

Step 5.3. Obtain the value

ψαk,1i,αk,1(1,λ)ψαk,1i,αk,1(1,λ)=4qv,λ(αk,1)ψαk,11,αk,1(1,λ)ψαk,11,αk,1(1,λ)\displaystyle\frac{\psi^{\prime}_{\alpha_{k,1}-\mathrm{i},\alpha_{k,1}}(1,\lambda)}{\psi_{\alpha_{k,1}-\mathrm{i},\alpha_{k,1}}(1,\lambda)}=4q_{v,\lambda}(\alpha_{k,1})-\frac{\psi^{\prime}_{\alpha_{k,1}-1,\alpha_{k,1}}(1,\lambda)}{\psi_{\alpha_{k,1}-1,\alpha_{k,1}}(1,\lambda)}
ψαk,1,αk,1+1(1,λ)ψαk,1,αk,1+1(1,λ)ψαk,1,αk,1+i(1,λ)ψαk,1,αk,1+i(1,λ)\displaystyle-\frac{\psi^{\prime}_{\alpha_{k,1},\alpha_{k,1}+1}(1,\lambda)}{\psi_{\alpha_{k,1},\alpha_{k,1}+1}(1,\lambda)}-\frac{\psi^{\prime}_{\alpha_{k,1},\alpha_{k,1}+\mathrm{i}}(1,\lambda)}{\psi_{\alpha_{k,1},\alpha_{k,1}+\mathrm{i}}(1,\lambda)} (37)

which is the Weyl function associated with the potential on (αk,1i,αk,1)(\alpha_{k,1}-\mathrm{i},\alpha_{k,1}).

Step 5.4. Similarly to steps 5.2–5.3, evaluating equation (29) at αk,2N+2k\alpha_{k,2N+2-k}, we get the Weyl function

ψαk,2N+1k1,αk,2N+1k(1,λ)ψαk,2N+1k1,αk,2N+1k(1,λ)=4qv,λ(αk,2N+1k)ψαk,2N+1ki,αk,2N+1k(1,λ)ψαk,2N+1ki,αk,2N+1k(1,λ)\displaystyle\frac{\psi^{\prime}_{\alpha_{k,2N+1-k}-1,\alpha_{k,2N+1-k}}(1,\lambda)}{\psi_{\alpha_{k,2N+1-k}-1,\alpha_{k,2N+1-k}}(1,\lambda)}=4q_{v,\lambda}(\alpha_{k,2N+1-k})-\frac{\psi^{\prime}_{\alpha_{k,2N+1-k}-\mathrm{i},\alpha_{k,2N+1-k}}(1,\lambda)}{\psi_{\alpha_{k,2N+1-k}-\mathrm{i},\alpha_{k,2N+1-k}}(1,\lambda)}
ψαk,2N+1k,αk,2N+1k+1(1,λ)ψαk,2N+1k,αk,2N+1k+1(1,λ)ψαk,2N+1k,αk,2N+1k+i(1,λ)ψαk,2N+1k,αk,2N+1k+i(1,λ)\displaystyle-\frac{\psi^{\prime}_{\alpha_{k,2N+1-k},\alpha_{k,2N+1-k}+1}(1,\lambda)}{\psi_{\alpha_{k,2N+1-k},\alpha_{k,2N+1-k}+1}(1,\lambda)}-\frac{\psi^{\prime}_{\alpha_{k,2N+1-k},\alpha_{k,2N+1-k}+\mathrm{i}}(1,\lambda)}{\psi_{\alpha_{k,2N+1-k},\alpha_{k,2N+1-k}+\mathrm{i}}(1,\lambda)} (38)

associated to the potential on (αk,2N+1k1,αk,2N+1k)(\alpha_{k,2N+1-k}-1,\alpha_{k,2N+1-k}).

Step 5.5. Repeating the procedure analogous to steps 5.2–5.4, we can obtain the values of ψαk,l1,αk,l(1,λ)ψαk,l1,αk,l(1,λ),l=2,,2N+1k\frac{\psi^{\prime}_{\alpha_{k,l}-1,\alpha_{k,l}}(1,\lambda)}{\psi_{\alpha_{k,l}-1,\alpha_{k,l}}(1,\lambda)},\ l=2,\cdots,2N+1-k and ψαk,li,αk,l(1,λ)ψαk,li,αk,l(1,λ),l=1,,2Nk\frac{\psi^{\prime}_{\alpha_{k,l}-\mathrm{i},\alpha_{k,l}}(1,\lambda)}{\psi_{\alpha_{k,l}-\mathrm{i},\alpha_{k,l}}(1,\lambda)},\ l=1,\cdots,2N-k.

Step 5.6. The potentials on (αk,l1,αk,l),l=2,,2N+1k(\alpha_{k,l}-1,\alpha_{k,l}),\ l=2,\cdots,2N+1-k and (αk,li,αk,l),l=1,,2Nk(\alpha_{k,l}-\mathrm{i},\alpha_{k,l}),\ l=1,\cdots,2N-k are uniquely determined by the Weyl functions by Lemma 5.2.

Thus, we have constructed all the potentials on the upper triangular region of the square domain.

Step 6. Rotate the whole system by the angle π\pi and take a square domain congruent to the previous one. Repeat the steps 2–5 to determine the potentials on all the remaining edges of EΩE_{\Omega}.

This reconstruction procedure determines the potential uniquely on each edge, so it implies the proof of Theorem 2.1.

Here we give a simple example of the reconstruction for N=3N=3.

As in Figure 5, by step 4, we get u(α6,1)u(\alpha_{6,1}) by the D-N map. By (33) and (34), we obtain

ψα6,11,α6,1(1,λ)=u(α6,1)u(α6,0)ψα6,0i,α6,0(1,λ),\psi_{\alpha_{6,1}-1,\alpha_{6,1}}(1,\lambda)=-\frac{u(\alpha_{6,1})}{u(\alpha_{6,0})}\psi_{\alpha_{6,0}-\mathrm{i},\alpha_{6,0}}(1,\lambda),
ψα6,1i,α6,1(1,λ)=u(α6,1)u(α6,2)ψα6,21,α6,2(1,λ).\psi_{\alpha_{6,1}-\mathrm{i},\alpha_{6,1}}(1,\lambda)=-\frac{u(\alpha_{6,1})}{u(\alpha_{6,2})}\psi_{\alpha_{6,2}-1,\alpha_{6,2}}(1,\lambda).

Then, the potentials on e1,e2e_{1},e_{2} are determined by Lemma 5.1.

Refer to caption
Figure 5. N=3,k=6

As in Figure 6, by (35),(36), we obtain

ψα5,11,α5,1(1,λ)=u(α5,1)u(α5,0)ψα5,0i,α5,0(1,λ),\psi_{\alpha_{5,1}-1,\alpha_{5,1}}(1,\lambda)=-\frac{u(\alpha_{5,1})}{u(\alpha_{5,0})}\psi_{\alpha_{5,0}-\mathrm{i},\alpha_{5,0}}(1,\lambda),
ψα5,2i,α5,2(1,λ)=u(α5,2)u(α5,3)ψα5,31,α5,3(1,λ).\psi_{\alpha_{5,2}-\mathrm{i},\alpha_{5,2}}(1,\lambda)=-\frac{u(\alpha_{5,2})}{u(\alpha_{5,3})}\psi_{\alpha_{5,3}-1,\alpha_{5,3}}(1,\lambda).

Then, the potentials on e3,e6e_{3},e_{6} are determined. By (37) and (38), we have

ψα5,1i,α5,1(1,λ)ψα5,1i,α5,1(1,λ)=4qv,λ(α5,1)ψα5,11,α5,1(1,λ)ψα5,11,α5,1(1,λ)\displaystyle\frac{\psi^{\prime}_{\alpha_{5,1}-\mathrm{i},\alpha_{5,1}}(1,\lambda)}{\psi_{\alpha_{5,1}-\mathrm{i},\alpha_{5,1}}(1,\lambda)}=4q_{v,\lambda}(\alpha_{5,1})-\frac{\psi^{\prime}_{\alpha_{5,1}-1,\alpha_{5,1}}(1,\lambda)}{\psi_{\alpha_{5,1}-1,\alpha_{5,1}}(1,\lambda)}
ψα5,1,α5,1+1(1,λ)ψα5,1,α5,1+1(1,λ)ψα5,1,α5,1+i(1,λ)ψα5,1,α5,1+i(1,λ),\displaystyle-\frac{\psi^{\prime}_{\alpha_{5,1},\alpha_{5,1}+1}(1,\lambda)}{\psi_{\alpha_{5,1},\alpha_{5,1}+1}(1,\lambda)}-\frac{\psi^{\prime}_{\alpha_{5,1},\alpha_{5,1}+\mathrm{i}}(1,\lambda)}{\psi_{\alpha_{5,1},\alpha_{5,1}+\mathrm{i}}(1,\lambda)},

and

ψα5,21,α5,2(1,λ)ψα5,21,α5,2(1,λ)=4qv,λ(α5,2)ψα5,2i,α5,2(1,λ)ψα5,2i,α5,2(1,λ)\displaystyle\frac{\psi^{\prime}_{\alpha_{5,2}-1,\alpha_{5,2}}(1,\lambda)}{\psi_{\alpha_{5,2}-1,\alpha_{5,2}}(1,\lambda)}=4q_{v,\lambda}(\alpha_{5,2})-\frac{\psi^{\prime}_{\alpha_{5,2}-\mathrm{i},\alpha_{5,2}}(1,\lambda)}{\psi_{\alpha_{5,2}-\mathrm{i},\alpha_{5,2}}(1,\lambda)}
ψα5,2,α5,2+1(1,λ)ψα5,2,α5,2+1(1,λ)ψα5,2,α5,2+i(1,λ)ψα5,2,α5,2+i(1,λ).\displaystyle-\frac{\psi^{\prime}_{\alpha_{5,2},\alpha_{5,2}+1}(1,\lambda)}{\psi_{\alpha_{5,2},\alpha_{5,2}+1}(1,\lambda)}-\frac{\psi^{\prime}_{\alpha_{5,2},\alpha_{5,2}+\mathrm{i}}(1,\lambda)}{\psi_{\alpha_{5,2},\alpha_{5,2}+\mathrm{i}}(1,\lambda)}.

Then, we can determine the potentials on e4,e5e_{4},e_{5} by Lemma 5.2.

Refer to caption
Figure 6. N=3,k=5

As in Figure 7, by (35),(36) in step 5, we get

ψα4,11,α4,1(1,λ)=u(α4,1)u(α4,0)ψα4,0i,α4,0(1,λ),\psi_{\alpha_{4,1}-1,\alpha_{4,1}}(1,\lambda)=-\frac{u(\alpha_{4,1})}{u(\alpha_{4,0})}\psi_{\alpha_{4,0}-\mathrm{i},\alpha_{4,0}}(1,\lambda),
ψα4,3i,α4,3(1,λ)=u(α4,3)u(α4,4)ψα4,41,α4,4(1,λ).\psi_{\alpha_{4,3}-\mathrm{i},\alpha_{4,3}}(1,\lambda)=-\frac{u(\alpha_{4,3})}{u(\alpha_{4,4})}\psi_{\alpha_{4,4}-1,\alpha_{4,4}}(1,\lambda).

Then, by using Lemma 5.1, the potentials on e7,e12e_{7},\ e_{12} are specified.

By (37),(38) and the argument in step 5, we get the values of ψα4,1i,α4,1(1,λ)ψα4,1i,α4,1(1,λ)\frac{\psi^{\prime}_{\alpha_{4,1}-\mathrm{i},\alpha_{4,1}}(1,\lambda)}{\psi_{\alpha_{4,1}-\mathrm{i},\alpha_{4,1}}(1,\lambda)}, ψα4,31,α4,3(1,λ)ψα4,31,α4,3(1,λ)\frac{\psi^{\prime}_{\alpha_{4,3}-1,\alpha_{4,3}}(1,\lambda)}{\psi_{\alpha_{4,3}-1,\alpha_{4,3}}(1,\lambda)}, ψα4,2i,α4,2(1,λ)ψα4,2i,α4,2(1,λ)\frac{\psi^{\prime}_{\alpha_{4,2}-\mathrm{i},\alpha_{4,2}}(1,\lambda)}{\psi_{\alpha_{4,2}-\mathrm{i},\alpha_{4,2}}(1,\lambda)} and ψα4,21,α4,2(1,λ)ψα4,21,α4,2(1,λ)\frac{\psi^{\prime}_{\alpha_{4,2}-1,\alpha_{4,2}}(1,\lambda)}{\psi_{\alpha_{4,2}-1,\alpha_{4,2}}(1,\lambda)}. Then, the potentials on e8,e9,e10,e11e_{8},e_{9},e_{10},e_{11} are determined by Lemma 5.2.

Refer to caption
Figure 7. N=3,k=4

By step 6, rotate the whole system by the angle π\pi and the potentials on the rest of edges are determined.

Remark 5.1.

Note that our reconstruction procedure is different from the one provided in [4, 5]. First, our algorithm depends on specific properties of the square lattice, while the procedure in [4, 5] was developed for the hexagonal lattice. Second, the reconstruction strategies differ. The authors of [4, 5] fix an arbitrary line AkA_{k} and compute the solution uu of the boundary value problem generated by the reduced vertex Laplacian (ΔV,λ+QV,λ)(-\Delta_{V,\lambda}+Q_{V,\lambda}) according to the analog of Lemma 4.2 for the hexagonal lattice. Then, they consider a zigzag line and find the ratios of the functions ψwv(1,λ)\psi_{wv}(1,\lambda) for adjacent edges of that line. This helps to recover the potentials qeq_{e}. This method perfectly works for the discrete Schrödinger operators with constant coefficients (see [3, 19]). But the analogous treatment of the reduced vertex Laplacian causes difficulties, since the both ΔV,λ\Delta_{V,\lambda} and QV,λQ_{V,\lambda} depend on some potentials qeq_{e}. Therefore, the solution uu cannot be found until the potentials are known. In our procedure, this problem does not arise, because we step-by-step reconstruct the potentials qeq_{e} together with the reduced vertex Laplacian coefficients, which are needed at the next steps. A similar approach can be applied to the hexagonal lattice to improve the algorithm of [4, 5].

6. Inverse scattering problem by the S-matrix

This section is concerned with the inverse scattering problem for the Schrödinger operator HEH_{E} by the S-matrix. It has been shown in [4] that the edge D-N map uniquely specifies the S-matrix. Consequently, the uniqueness theorem by the D-N map (Theorem 2.1) implies the uniqueness of solution for the inverse scattering problem. In this section, we provide the definition of the S-matrix S(λ)S(\lambda), following the papers [2, 4, 5] and show that it uniquely determines the potential on the square lattice.

In this section, it will be convenient for us to associate the vertex set VV of the square lattice with 2\mathbb{Z}^{2}. In other words, every vertex vVv\in V is a pair of integers n=(n1,n2)n=(n_{1},n_{2}). Define the vertex Laplacian

(ΔVf)(v)=14wvf(w),vV,\left(\Delta_{V}f\right)(v)=\frac{1}{4}\sum_{w\sim v}f(w),\quad v\in V,

which is self-adjoint on L2(V)L^{2}(V) equipped with the inner product

(f,g)=4n2f(n)g(n)¯.(f,g)=4\sum_{n\in\mathbb{Z}^{2}}f(n)\cdot\overline{g(n)}.

Put 𝕋2:=2(2π)2\mathbb{T}^{2}:=\mathbb{R}^{2}\setminus(2\pi\mathbb{Z})^{2} and define the discrete Fourier transform UV:L2(2;2)L2(𝕋2;2)U_{V}\colon L^{2}(\mathbb{Z}^{2};\mathbb{C}^{2})\longrightarrow L^{2}(\mathbb{T}^{2};\mathbb{C}^{2}) by

(UVf)(x)=1πn2einxf(n),x=(x1,x2)𝕋2.(U_{V}f)(x)=\frac{1}{\pi}\sum_{n\in\mathbb{Z}^{2}}e^{\mathrm{i}n\cdot x}f(n),\quad x=(x_{1},x_{2})\in\mathbb{T}^{2}. (39)

The adjoint operator UV:L2(𝕋2;2)L2(2;2)U_{V}^{*}\colon L^{2}(\mathbb{T}^{2};\mathbb{C}^{2})\longrightarrow L^{2}(\mathbb{Z}^{2};\mathbb{C}^{2}) is given by

(UV)g(n)=14π𝕋2einxg(x)𝑑x,n2.(U_{V}^{*})g(n)=\frac{1}{4\pi}\int_{\mathbb{T}^{2}}e^{-\mathrm{i}n\cdot x}g(x)dx,\quad n\in\mathbb{Z}^{2}. (40)

Then, on L2(𝕋2;2)L^{2}(\mathbb{T}^{2};\mathbb{C}^{2}), UV(ΔV)UVU_{V}(-\Delta_{V})U_{V}^{*} is the operator of multiplication by

H0(x)=12(cosx1+cosx2).H_{0}(x)=-\frac{1}{2}(\cos x_{1}+\cos x_{2}). (41)

For qe=0q_{e}=0, we denote ΔV,λ-\Delta_{V,\lambda} by ΔV,λ(0)-\Delta_{V,\lambda}^{(0)}. By Lemma 3.1 of [2], we define the characteristic surface of ΔV,λ(0)-\Delta_{V,\lambda}^{(0)} by

Mλ={x𝕋2:H0(x)+cosλ=0},M_{\lambda}=\{x\in\mathbb{T}^{2}\colon H_{0}(x)+\cos\sqrt{\lambda}=0\},

which is smooth if cosλ0,±1,λ\cos\sqrt{\lambda}\neq 0,\pm 1,\lambda\in\mathbb{R}. We put

T(0)={λ:cosλ=0,±1},T^{(0)}=\{\lambda\colon\cos\sqrt{\lambda}=0,\pm 1\},
T=T(0)(eEσ(He)),T=T^{(0)}\cup\left(\cup_{e\in E}\sigma\left(H_{e}\right)\right), (42)

where the operator HeH_{e} was defined for equation (5) in Section 3.

The resolvent re(λ)=(Heλ)1r_{e}(\lambda)=\left(H_{e}-\lambda\right)^{-1} is written as

(re(λ)f)(z)=0zϕe1(z,λ)ϕe0(t,λ)ϕe0(1,λ)f(t)𝑑t+z1ϕe0(z,λ)ϕe1(t,λ)ϕe1(0,λ)f(t)𝑑t.(r_{e}(\lambda)f)(z)=\int_{0}^{z}\frac{\phi_{e1}(z,\lambda)\phi_{e0}(t,\lambda)}{\phi_{e0}(1,\lambda)}f(t)dt+\int_{z}^{1}\frac{\phi_{e0}(z,\lambda)\phi_{e1}(t,\lambda)}{\phi_{e1}(0,\lambda)}f(t)dt.

We put

Φe0(λ)f=ddz(re(λ)f)|z=0=01ϕe1(t,λ)ϕe1(0,λ)f(t)𝑑t,\Phi_{e0}(\lambda)f=\frac{d}{dz}(r_{e}(\lambda)f)\bigg{|}_{z=0}=\int_{0}^{1}\frac{\phi_{e1}(t,\lambda)}{\phi_{e1}(0,\lambda)}f(t)dt,
Φe1(λ)f=ddz(re(λ)f)|z=1=01ϕe0(t,λ)ϕe0(1,λ)f(t)𝑑t.\Phi_{e1}(\lambda)f=-\frac{d}{dz}(r_{e}(\lambda)f)\bigg{|}_{z=1}=\int_{0}^{1}\frac{\phi_{e0}(t,\lambda)}{\phi_{e0}(1,\lambda)}f(t)dt.

Their adjoints acting from \mathbb{C} to L2(0,1)L^{2}(0,1) are defined for cc\in\mathbb{C} by

Φe1(λ)c=cϕe0(z,λ¯)ϕe0(1,λ¯),Φe0(λ)c=cϕe1(z,λ¯)ϕe1(0,λ¯).\Phi_{e1}(\lambda)^{*}c=c\frac{\phi_{e0}(z,\bar{\lambda})}{\phi_{e0}(1,\bar{\lambda})},\quad\Phi_{e0}(\lambda)^{*}c=c\frac{\phi_{e1}(z,\bar{\lambda})}{\phi_{e1}(0,\bar{\lambda})}.

Define the operator TV(λ):Lloc2(E)Lloc2(V)T_{V}(\lambda):L^{2}_{loc}(E)\rightarrow L^{2}_{loc}(V) by

(TV(λ)fe)(v)=14(eEv(1)Φe1(λ)fe+eEv(0)Φe0(λ)fe),vV.(T_{V}(\lambda)f_{e})(v)=\frac{1}{4}\left(\sum_{e\in E_{v}(1)}\Phi_{e1}(\lambda)f_{e}+\sum_{e\in E_{v}(0)}\Phi_{e0}(\lambda)f_{e}\right),\quad v\in V. (43)

The adjoint operator TV(λ):Lloc2(V)Lloc2(E)T_{V}(\lambda)^{*}:L^{2}_{loc}(V)\rightarrow L^{2}_{loc}(E) has the form

(TV(λ)u)e(z)=Φe1(λ)u(e(1))+Φe0(λ)u(e(0)).(T_{V}(\lambda)^{*}u)_{e}(z)=\Phi_{e1}(\lambda)^{*}u(e(1))+\Phi_{e0}(\lambda)^{*}u(e(0)).

For qe=0q_{e}=0, we denote TV(λ)T_{V}(\lambda) by TV(0)(λ)T_{V}^{(0)}(\lambda). Then,

(TV(0)(λ)f)(v)=14λsinλ(eEv(1)01sinλzλf(z)dz+eEv(0)01sinλ(1z)λf(z)dz)\begin{split}(T_{V}^{(0)}(\lambda)f)(v)=&\frac{1}{4}\frac{\sqrt{\lambda}}{\sin\sqrt{\lambda}}\bigg{(}\sum_{e\in E_{v}(1)}\int_{0}^{1}\frac{\sin\sqrt{\lambda}z}{\sqrt{\lambda}}f(z)dz+\\ &\sum_{e\in E_{v}(0)}\int_{0}^{1}\frac{\sin\sqrt{\lambda}(1-z)}{\sqrt{\lambda}}f(z)dz\bigg{)}\end{split}

and

(TV(0)(λ¯)u)e(z)=λsinλ(sinλzλu(e(1))+sinλ(1z)λu(e(0))).(T_{V}^{(0)}(\bar{\lambda})^{*}u)_{e}(z)=\frac{\sqrt{\lambda}}{\sin\sqrt{\lambda}}\left(\frac{\sin\sqrt{\lambda}z}{\sqrt{\lambda}}u(e(1))+\frac{\sin\sqrt{\lambda}(1-z)}{\sqrt{\lambda}}u(e(0))\right).

Under the assumptions (Q-1), (Q-2), (Q-3), we define the Schrödinger operator

HE={d2dz2+qe(z):eE}H_{E}=\left\{-\frac{d^{2}}{dz^{2}}+q_{e}(z)\colon e\in E\right\}

with domain D(HE)D(H_{E}) consisting of functions u={ue}eEu=\{u_{e}\}_{e\in E} such that ueH2(0,1)u_{e}\in H^{2}(0,1) satisfy the Kirchhoff condition (K-1) and (K-2) in all the vertices vVv\in V and eEd2dz2ue+qeueL2(0,1)2<\sum_{e\in E}\|-\frac{d^{2}}{dz^{2}}u_{e}+q_{e}u_{e}\|_{L^{2}(0,1)}^{2}<\infty. Due to [4], the operator HEH_{E} is self-adjoint with the essential spectrum σe(HE)=[0,)\sigma_{e}(H_{E})=[0,\infty). Furthermore, σe(HE)T\sigma_{e}(H_{E})\setminus T is absolutely continuous, where TRT\subset R is defined by (42). Then, one can introduce the S-matrix S(λ)S(\lambda) for λ(0,)T\lambda\in(0,\infty)\setminus T.

Define the spaces B(E)B^{*}(E) and B0(E)B_{0}^{*}(E) by

fB(E)\displaystyle f\in B^{*}(E)\quad fB(E)2=supR>11R|c(e)|<RfeL2(0,1)2<,\displaystyle\Leftrightarrow\quad\|f\|_{B^{*}(E)}^{2}=\sup_{R>1}\frac{1}{R}\sum_{|c(e)|<R}\|f_{e}\|_{L^{2}(0,1)}^{2}<\infty,
fB0(E)\displaystyle f\in B_{0}^{*}(E)\quad limR1R|c(e)|<RfeL2(0,1)2=0,\displaystyle\Leftrightarrow\quad\lim_{R\rightarrow\infty}\frac{1}{R}\sum_{|c(e)|<R}\|f_{e}\|_{L^{2}(0,1)}^{2}=0,

where c(e)=12|e(0)+e(1)|c(e)=\frac{1}{2}|e(0)+e(1)|. For f,gB(E)f,g\in B^{*}(E), we use the notation fgf\simeq g in the following sense:

fgfgB0(E).f\simeq g\Leftrightarrow f-g\in B_{0}^{*}(E).

Now, we give the definition of the S-matrix S(λ)S(\lambda). The following lemma is a special case of Theorem 5.8 from [4].

Lemma 6.1 ([4]).

Let λ(0,)T\lambda\in(0,\infty)\setminus T. Then, for any incoming data ϕinL2(Mλ)\phi^{in}\in L^{2}(M_{\lambda}), there exist a unique solution uB(E)u\in B^{*}(E) of the equation

(HEλ)u=0,(H_{E}-\lambda)u=0,

satisfying the Kirchoff conditions, and an outgoing data ϕoutL2(Mλ)\phi^{out}\in L^{2}(M_{\lambda}) satisfying

u\displaystyle u\simeq TV(0)(λ)UV1λ(x)+cosλ+i0σ(λ)ϕin\displaystyle-T_{V}^{(0)}(\lambda)^{*}U_{V}^{*}\frac{1}{\lambda(x)+\cos\sqrt{\lambda}+\mathrm{i}0\sigma(\lambda)}\phi^{in}
+TV(0)(λ)UV1λ(x)+cosλi0σ(λ)ϕout,\displaystyle+T_{V}^{(0)}(\lambda)^{*}U_{V}^{*}\frac{1}{\lambda(x)+\cos\sqrt{\lambda}-\mathrm{i}0\sigma(\lambda)}\phi^{out}, (44)

where x=(x1,x2)x=(x_{1},x_{2}), λ(x)=12(cosx1+cosx2)\lambda(x)=-\frac{1}{2}(\cos x_{1}+\cos x_{2}) is the eigenvalue of H0(x)H_{0}(x) and σ(λ)=1\sigma(\lambda)=1 if sinλ>0\sin\sqrt{\lambda}>0, σ(λ)=1\sigma(\lambda)=-1 if sinλ<0\sin\sqrt{\lambda}<0.

The mapping S(λ):ϕinϕoutS(\lambda)\colon\phi^{in}\rightarrow\phi^{out} is called the S-matrix.

The results of Section 6 in [4] imply the following lemma for the square lattice.

Lemma 6.2 ([4]).

For the edge Schrödinger operator on the square lattice, the S-matrix S(λ)S(\lambda) and the edge D-N map ΛE\Lambda_{E} uniquely determine each other.

Therefore, Theorem 2.1, Lemma 6.2, and the reconstruction procedure in Section 5 imply the following corollary.

Corollary 6.3.

Assume (Q-1), (Q-2) and (Q-3). Then, given any open interval I(0,)TI\subset(0,\infty)\setminus T, and the S-matrix S(λ)S(\lambda) for all λI\lambda\in I, one can uniquely reconstruct the potential qe(z)q_{e}(z) for all eEe\in E.

Indeed, under the assumptions (Q-1), (Q-2), and (Q-3), S(λ)S(\lambda) is meromorphic in the half-plane {λ:Reλ>0}\{\lambda\colon\mbox{Re}\,\lambda>0\} with possible branch points at TT. Therefore, given S(λ)S(\lambda) for λI\lambda\in I, one can find S(λ)S(\lambda) for all λ(0,)T\lambda\in(0,\infty)\setminus T by analytic continuation. Analogously, one can obtain the following result.

Corollary 6.4.

Assume (Q-1) and (Q-3), and given a real q0(z)L2(0,1)q_{0}(z)\in L^{2}(0,1) satisfying q0(z)=q0(1z)q_{0}(z)=q_{0}(1-z) and qe(z)=q0(z)q_{e}(z)=q_{0}(z) on (0,1)(0,1) except for a finite number of edges eEe\in E. Given an open interval Iσe(HE)TI\subset\sigma_{e}(H_{E})\setminus T and the S-matrix S(λ)S(\lambda) for all λI\lambda\in I, one can uniquely reconstruct the potential qe(z)q_{e}(z) for all edges eEe\in E.

The reconstruction procedure for Corollary 6.4 requires no essential changes. Instead of sinλzλ\frac{\sin\sqrt{\lambda}z}{\sqrt{\lambda}} and sinλ(1z)λ\frac{\sin\sqrt{\lambda}(1-z)}{\sqrt{\lambda}}, we have only to use the corresponding solutions to the Schrödinger equation (d2dz2+q0(z)λ)φ=0(-\frac{d^{2}}{dz^{2}}+q_{0}(z)-\lambda)\varphi=0.

Corollaries 6.3 and 6.4 are analogous to the results of [4, 5] for the hexagonal lattice.

References

  • [1] K. Ando, Inverse scattering theory for discrete Schrödinger operators on the hexagonal lattice, Ann. Henri Poincaré 14 (2013), 347-383.
  • [2] K. Ando, H. Isozaki and H. Morioka, Spectral properties of Schrödinger operators on perturbed lattices, Ann. Henri Poincaré 17 (2016), 2103-2171.
  • [3] K. Ando, H. Isozaki and H. Morioka, Inverse scattering for Schrödinger operators on perturbed lattices, Ann. Henri Poincaré 19 (2018), 3397-3455.
  • [4] K. Ando, H. Isozaki, E. Korotyaev and H. Morioka, Inverse scattering on the quantum graph — Edge model for graphene, arXiv:1911.05233.
  • [5] K. Ando, H. Isozaki, E. Korotyaev and H. Morioka, Inverse scattering on the quantum graph for graphene, arXiv:2102.05217.
  • [6] S. A. Avdonin and V. V. Kravchenko, Method for solving inverse spectral problems on quantum star graphs, Journal of Inverse and Ill-Posed Problems 31 (2023), no. 1, 31-42.
  • [7] M. I. Belishev, Boundary spectral inverse problem on a class of graphs (trees) by the BC method, Inverse Problems 20 (2004), 647-672.
  • [8] G. Berkolaiko, R. Carlson, S. Fulling and P. Kuchment, Quantum Graphs and Their Applications, Contemp. Math. 415, Amer. Math. Soc., Providence, RI (2006).
  • [9] G. Berkolaiko and P. Kuchment, Introduction to Quantum Graphs, Mathematical Surveys and Monographs 186, AMS (2013).
  • [10] N. P. Bondarenko, Spectral data characterization for the Sturm-Liouville operator on the star-shaped graph, Anal. Math. Phys. 10 (2020), 83.
  • [11] S. Buterin and G. Freiling, Inverse spectral-scattering problem for the Sturm-Liouville operator on a noncompact star-type graph. Tamkang J. Math., 44 (2013), 327–349.
  • [12] F. Chung, Spectral Graph Theory, AMS. Providence, Rhodse Island (1997).
  • [13] D. Cvetkovic, M. Doob and H. Saks, Spectra of graphs, Theory and applications, 3rd edition, Johann Ambrosius Barth, Heidelberg (1995).
  • [14] P. Exner, A. Kostenko, M. Malamud and H. Neidhardt, Spectral theory for infinite quantum graph, Ann. Henri Poincaré 19 (2018), 3457-3510.
  • [15] G. Freiling, V. A. Yurko. Inverse Sturm-Liouville problems and their applications. Nova Science pub(2001).
  • [16] B. Gutkin and U. Smilansky, Can one hear the shape of a graph? J. Phys. A 34 (2001), 6061-6068.
  • [17] M. Ignatyev, Inverse scattering problem for Sturm–Liouville operator on non-compact A-graph. Uniqueness result, Tamkang J. Math., 46 (2015), 401-422.
  • [18] H. Isozaki and E. Korotyaev, Inverse problems, trace formulae for discrete Schrödinger operators, Ann. Henri Poincaré, 13 (2012), 751-788.
  • [19] H. Isozaki and H. Morioka, Inverse scattering at a fixed energy for discrete Schrödinger operators on the square lattice, Ann. l’Inst. Fourier 65 (2015), 1153-1200.
  • [20] E. Korotyaev and I. Lobanov, Schrödinger operators on zigzag nanotubes, Annales henri poincaré. Basel: Birkhäuser-Verlag, 8 (2007), 1151-1176.
  • [21] E. Korotyaev and N. Saburova, Scattering on periodic metric graphs, Rev. Math. Phy., 32 (2020), 2050024.
  • [22] P. Kuchment and O. Post, On the spectra of carbon nano-structures, Commun. Math. Phy., 275 (2007), 805-826.
  • [23] P. Kurasov and M. Nowaczyk, Inverse spectral problem for quantum graphs, J. Phys. A: Math. Gen. 38 (2005), No. 22, 4901-4915.
  • [24] P. Kurasov, Inverse problems for Aharonov-Bohm rings, Math. Proc. Cambridge Phil. Soc. 148 (2010), no. 2, 331-362.
  • [25] Y. C. Luo, E. O. Jatulan, and C. K. Law, Dispersion relations of periodic quantum graphs associated with Archimedean tilings (I), J. Phy. A: Math. Theor. 52 (2019), 165201.
  • [26] Y. C. Luo, E. O. Jatulan, and C. K. Law, Dispersion relations of periodic quantum graphs associated with Archimedean tilings (II), J. Phy. A: Math. Theor. 52 (2019), 445204.
  • [27] K. Mochizuki, I. Y. Trooshin, On the scattering on a loop-shaped graph, Progress Math. 301 (2012), 227-245.
  • [28] V. Yurko, Inverse spectral problems for Sturm-Liouville operators on graphs, Inverse Problems 21 (2005), 1075-1086.
  • [29] V. Yurko, Inverse problems for Sturm-Liouville operators on bush-type graphs, Inverse Problems 25 (2009), no. 10, 105008.
  • [30] V. Yurko, Inverse spectral problems for differential operators on spatial networks, Russ. Math. Surveys 71 (2016) No. 3, 539-584.