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

Comfortability of quantum walks on embedded graphs on surfaces

Yusuke Higuchi1, Etsuo Segawa2
1Department of Mathematics, Gakushuin University,
Tokyo 171-8588, Japan
2Graduate School of Environment and Information Sciences, Yokohama National University,
Hodogaya, Yokohama 240-8501, Japan

Abstract. In this paper, a quantum walk model which reflects the underlying embedding on the surface is proposed. We obtain the scattering matrix of this quantum walk model characterized by the faces on the surface, and find a detection of the orientablility of the underlying embedding by the scattering information. The comfortability is the square norm of the stationary state restricted to the internal and reflected by the underlying embedding. We find that quantum walker feels more comfortable to a surface with small genus in some natural setting.

Key words and phrases. Discrete-time quantum walk, graph embedding on closed surfaces, stationary state, scattering matrix, comfortability

1 Introduction

How is the underlying geometric structure estimated by a reaction to an input? To tackle this inverse problem, we adopt a discrete-time quantum walk model, and try to extract a topological structure of graphs from the behavior of the quantum superposition. Among a lot of discrete-time quantum walk models [11], the Grover walk [19] may come to mind first. In fact, using the commutativity of the Grover matrix with any permutation matrix, the time evolution of the Grover walk is easily constructed [12], which exhibits interesting behaviors [9, 10] and also plays the key roles in the quantum search [17]. Moreover there are several important properties in the Grover walk connecting to not only a random walk[QCcoupling2, 2] but also the stationary Shrödinger equation [7].

Here, let us first apply the Grover walk as a discrete-time quantum walk model and focus on the embedding of graphs on surfaces as a geometric structure. For example, the complete graph with 44 vertices K4K_{4} has 1111 kinds of embedding way; see Figure 7. We expect that the quantum walk “feels” these underlying embeddings. Grover walk may decide whether an underlying graph admits a topological embedding on a given surface [3]. On the other hand, the time evolution of Grover walk depends only on the adjacent structure of the underlying graph because the time evolution operator UGroverU_{Grover} acts as

UGroverδe=(2deg(t(e))1)δe¯+2deg(t(e))ϵ:o(ϵ)=t(e),e¯δϵ,U_{\mathrm{Grover}}\delta_{e}=\left(\frac{2}{\mathrm{deg}(t(e))}-1\right)\delta_{\bar{e}}+\frac{2}{\mathrm{deg}(t(e))}\sum_{\epsilon:\;o(\epsilon)=t(e),\neq\bar{e}}\delta_{\epsilon},

for any standard base labeled by an arc ee of the underlying graph. Here o(e)o(e) and t(e)t(e) are origin and terminal vertices of the arc ee, and deg(x)\mathrm{deg}(x) is the degree of the vertex xx. Then the weights associated with transmitting and reflecting at vertex xx are 2/deg(x)2/\mathrm{deg}(x) and 2/deg(x)12/\deg(x)-1 in the Grover walk, which is independent of the underlying embedding. Therefore the Grover walk can not distinguish any two embedded surfaces of a graph. Thus for our object, we need to propose another kind of discrete-time quantum walk model for this problem.

To this end, let us introduce the idea of the drawing a graph without crossings of edges and extra faces on some closed surface, which is called the two-cell embedding [6, 14, 15]. More details are given in Section 3. See also Figures 5, 6 and 7. It is well known that the two-cell embedding on a closed surface of a graph G=(X,E)G=(X,E) is realized by the rotation system [6, 14, 15]. Here XX and EE are the set of vertices and the set ofunoriented edges. The rotation system is the triple of the symmetric digraph G=(X,A)G=(X,A), the rotation ρ:AA\rho:A\to A and the twist τ:A2\tau:A\to\mathbb{Z}_{2}, where AA is the set of the symmetric arcs induced by EE. Here the rotation ρ\rho is decomposed into cyclic permutations with respect to the incoming arcs of each vertex. Thus our target can be switched to how a quantum walk model is constructed for given rotation system (G,ρ,τ)(G,\rho,\tau).

In this paper, the abstract graph GG is deformed so that information of ρ,τ\rho,\tau are reflected and also the in- and out- degrees are equally 22. The regularity of the degree 22 derives from the implementation by some optical polarizers elements [13]. See Sections 4.1, 4.2 for more detailed construction and also Figure 8. Such a new graph G(ρ,τ)G(\rho,\tau) is obtained by replacing each vertex of the double covering graph [6, 14] induced by τ\tau with a directed cycle induced by ρ\rho. The replaced directed cycles are called islands, and the symmetric arcs between the islands, which reflects the original adjacent relation, are called the bridges. The tails, which are semi-infinite paths, are inserted into all the island arcs. Such an assignment of tails is called the hedgehog; see Figure 8 (d). Let us explain the time evolution below; see Sections 4.3 and 4.4 for more detail. The degree-22 regularity of the new graph makes it possible that the one-step time evolution on the whole space can be described by the local scattering at each vertex of the internal graph by the 2×22\times 2 unitary matrix

C=[abcd].C=\begin{bmatrix}a&b\\ c&d\end{bmatrix}.

Here a,b,c,da,b,c,d are the complex valued weights associated with moving of a quantum walk for one step “from an island to the same island”, “ a bridge to an island”, “an island to a bridge” and “a bridge to the inverse bridge”, respectively. This unitarity of the local coin matrix CC includes also the unitarty of the total time evolution operator. On the other hand, we set that the dynamics on the tail is free; see (4.16) for the detailed dynamics of the free. The initial state is set so that the internal graph receives the constant inflow at every time step. If a quantum walker goes out to the external, it never come back to the internal because of the free dynamics of the tails. Such a quantum walker can be regarded as the outflow. By the balance between the in- and out- flows, this quantum walk model converges to a fixed point in the long time limit [4, 5, 8].

Now, for a fixed abstract graph GoG_{o}, let us vary the embedding that underlies it. For every ρ\rho and τ\tau, set ψ\psi_{\infty} as the stationary state and ψ|Go\psi_{\infty}|_{G_{o}} as the restriction of ψ\psi_{\infty} to the internal graph GoG_{o}. In this paper, we focus on extracting embedding structures of a graph from the stationary state of this quantum walk model. To this end, we divide the stationary state into external and internal parts; we characterize the external part by the scattering matrix, which shows a reaction of the internal graph to an input at the tails, while the internal part by the comfortability, which is corresponding to the energy in the internal graph. We estimate how quantum walker

(i) gives a scattering on the underlying embedding by finding the expression of the scattering matrix

𝜷out=S𝜶in,\boldsymbol{\beta}_{out}=S\boldsymbol{\alpha}_{in},

where SS is independent of inflow 𝜶in\boldsymbol{\alpha}_{in} and outflow 𝜷out\boldsymbol{\beta}_{out} and a unitary operator [4, 5], and

(ii) feels comfortable to the embedding by defining

(Go,ρ,τ):=:=12||ψ|Go||2,\mathcal{E}(G_{o},\rho,\tau):=\mathcal{E}:=\frac{1}{2}||\;\psi_{\infty}|_{G_{o}}\;||^{2},

that is, the larger \mathcal{E} is, the more comfortable quantum walker feels to the underlying embedding.

Throughout this paper, in addition to the hedgehog tail condition, we subject the following assumption (2).

Assumption 1.
  1. (1)

    the assignment of tails is the hedgehog;

  2. (2)

    the (2,2)(2,2) element of CC (=d)(=d) is a real number.

Under such a construction of the quantum walk reflecting the underlying embedding on a closed surface, in this paper, now we are ready to state our main result on the scattering and comfortablity.

Theorem 1.1 (Scattering).

Assume dd\in\mathbb{R} and set ω=detC\omega=-\det C and the hedgehog assignment of tails. Let FF be the set of faces induced by the rotation system (G,ρ,τ)(G,\rho,\tau). The scattering matrix is decomposed into the following |F||F| unitary matrices as follows:

S=fFSf,S=\bigoplus_{f\in F}S_{f},

where

Sf=bcωPf(IfaωPf)1+dIf.S_{f}=bc\omega P_{f}\;(I_{f}-a\omega P_{f})^{-1}+dI_{f}.

Here the operators induced by each face ff, IfI_{f} is the identity operator and PfP_{f} are defined in (6.26).

Theorem 6.1 in Section 6 gives the scattering matrix in a more general setting. The operator PfP_{f} is a weighted permutation matrix induced by the closed walk along the boudary of the face ff, which is called the facial walk of ff. The rotation system (G,ρ,τ)(G,\rho,\tau) determines the set of faces FF and gives uniquely orientation of the closed walk along the boundary of the face fFf\in F. This closed walk corresponds to the facial walk, which is expressed by a sequence of arcs, and we simply write ff for this closed walk. We should remark f¯F\bar{f}\not\in F for fFf\in F, where f¯\bar{f} is the inverse direction of ff. The weight is determined by each edge type which is passed through the facial walk. Then using this property of the scattering matrix, the detection of the orientability by the scattering information is proposed in Theorem 6.2.

Theorem 1.2 (Comfortability).

Let G=(X,A)G=(X,A) and (G,ρ,τ)(G,\rho,\tau) be the abstract symmetric digraph and its rotation system. Let FF be the set of faces determined by the rotation system (G,ρ,τ)(G,\rho,\tau). Pick a tail at random from the hedgehog tails as an input. Under Assumption 1 with a>0a>0 and ω=1\omega=1, the average of the comfortability with respect to a randomly chosen input is expressed by

𝔼[]\displaystyle\mathbb{E}[\mathcal{E}] =1|A|2+|b|2|b|2fF|f|1+a|f|1a|f|1|A|a|b|2fF11a|f|eff¯(adistf(e,e¯)+adistf(e¯,e)).\displaystyle=\frac{1}{|A|}\frac{2+|b|^{2}}{|b|^{2}}\sum_{f\in F}|f|\frac{1+a^{|f|}}{1-a^{|f|}}-\frac{1}{|A|}\frac{a}{|b|^{2}}\sum_{f\in F}\frac{1}{1-a^{|f|}}\sum_{e\in f\cap\bar{f}}\left(\;a^{\mathrm{dist}_{f}(e,\bar{e})}+a^{\mathrm{dist}_{f}(\bar{e},e)}\;\right). (1.1)

Here ff¯f\cap\bar{f} is the set of self-intersection of the face fFf\in F; where if a facial walk ff passes through an arc ee and also its inverse e¯\bar{e}, then the face ff is said to have a self-intersection at the unoriented boundary edge |e||e|.

Thus, the larger the first term, the more comfortable the quantum walker feels, while the larger the second term, the more uncomfortable it feels. The first term is related to an integer partition of |A||A|, while the second term is related to the self-intersections of faces. See Figure 4 for the self-intersection. Figure 2 shows the ranking of the comfortability of the embeddings for the complete graph K4K_{4} with a=0.98a=0.98. We discuss the messages of combinatorial structures from Theorem 1.2 in Section 2.

This paper is organized as follows. Section 2 is devoted to how geometric and combinatorial information of the graph embeddings is extracted from Theorem 1.2 as its corollaries. We show the best and worst embedding of the complete graph KnK_{n} for quantum walker and also characterize the comfortability on the island by the Young diagram. In section 3, we give a short review on the graph embeddings on the orientable/non-orientable surfaces. In section 4, the time evolution of the quantum walk induced by the rotation system (G,ρ,τ)(G,\rho,\tau) is discussed. In section 5, we show the unitary equivalence of the time evolutions between the isomorphic embeddings. In section 6, we find that the scattering matrix is characterized by the resulting faces and described by direct sum of unitary and circurant matrices. Moreover a detection method of the orientability by using the scattering information is proved. In section 7, we give the proof of Theorem 1.2.

Refer to caption
Figure 1: The list of embeddings of K4K_{4}: The genus is described by gg and kk, for orientable and non-orientable surfaces, respectively. The boundary lenghts of faces of the resulting embedding are indicated by [λ1,λ2,,λκ][\lambda_{1},\lambda_{2},\dots,\lambda_{\kappa}]. For example, [6,3,3][6,3,3] indicates there are 11 hexagon and 22 triangles in the embedding.
Refer to caption
Figure 2: The ranking of the comfortability for the embeddings of K4K_{4} with a=0.98a=0.98: This ranking follows Corollary 2.1 in the setting of a1a\to 1.

2 Combinatorial observations (corollaries of Theorem 1.2)

In this section, we discuss how more detailed geometric information is extracted from Theorem 1.2.

2.1 Observation 1: In the limit of a1a\to 1

It is easy to observe that if a0a\to 0, then the comfortability converges to 33, which is completely independent of (G,ρ,τ)(G,\rho,\tau). It also confirms the consistency by considering that if a=0a=0, a walker is forced to trace a route “ tail\toisland\tobridge\toisland\totail” by the definition of this quantum walk.

On the other hand, if a1a\to 1, then the comfortability diverges. Then, taking a=1δa=1-\delta, we will find the appropriate scaling with respect to δ\delta, and its coefficient of the first order hoping that we will find some geometric information in the coefficient. Indeed, we obtain the following.

Corollary 2.1.

Let GG be a connected abstract graph with the vertex set XX and the edge set EE. Let (G,ρ,τ)(G,\rho,\tau) be the rotation system with the face set FF. Under Assumption 1 with a=1δa=1-\delta (δ1)(\delta\ll 1), the average of the comfortability with respect to the randomly chosen initial state is expressed by

limδ0𝔼[δ]δ2=|F||E|(11|F|fF|ff¯||f|).\lim_{\delta\downarrow 0}\mathbb{E}[\mathcal{E}_{\delta}]\;\delta^{2}=\frac{|F|}{|E|}\left(1-\frac{1}{|F|}\sum_{f\in F}\frac{|f\cap\bar{f}|}{|f|}\right). (2.2)

The small genus increases the number of faces, while large genus induces longer faces and, therefore, larger number of intersections. Thus for a1a\to 1, quantum walker feels more comfortable to the surface with a smaller genus. In particular, the single-face embedding on an orientable surface attains the above comfortablity 0. Figure 2 shows the ranking of the comfortability for K4K_{4} following Corollary 2.1.

Proof.

Inserting the following expansions by small δ1\delta\ll 1 into the first term of (1.1),

1+a|f|1a|f|=1δ1|f|{2δ+O(δ2)},2+|b|2|b|2=1δ(1+32δ+O(δ2)),\frac{1+a^{|f|}}{1-a^{|f|}}=\frac{1}{\delta}\frac{1}{|f|}\left\{2-\delta+O(\delta^{2})\right\},\;\frac{2+|b|^{2}}{|b|^{2}}=\frac{1}{\delta}\left(1+\frac{3}{2}\delta+O(\delta^{2})\right),

we have

the first term=1δ2(|F||E|+|F||E|δ+O(δ2)),\displaystyle\text{the first term}=\frac{1}{\delta^{2}}\left(\frac{|F|}{|E|}+\frac{|F|}{|E|}\delta+O(\delta^{2})\right),

On the other hand, inserting the following expansions by small δ1\delta\ll 1 into the second term of (1.1),

adistf(e,e¯)+adistf(e¯,e)=2|f|δ+O(δ2),\displaystyle a^{\mathrm{dist}_{f}(e,\bar{e})}+a^{\mathrm{dist}_{f}(\bar{e},e)}=2-|f|\delta+O(\delta^{2}),
11a|f|=1δ1|f|(1+|f|12δ+O(δ)+O(δ2)),\displaystyle\frac{1}{1-a^{|f|}}=\frac{1}{\delta}\frac{1}{|f|}\left(1+\frac{|f|-1}{2}\delta+O(\delta)+O(\delta^{2})\right),
a|b|2=12δ(112δ+(δ2))\displaystyle\frac{a}{|b|^{2}}=\frac{1}{2\delta}\left(1-\frac{1}{2}\delta+(\delta^{2})\right)

we have

the second term=1δ21|E|fF|ff¯||f|.\text{the second term}=\frac{1}{\delta^{2}}\frac{1}{|E|}\sum_{f\in F}\frac{|f\cap\bar{f}|}{|f|}.

Then we have the desired conclusion. ∎

Example: The best and worst embeddings of KnK_{n} for quantum walker.
For a fixed abstract graph, if the underlying embedding gives the most comfortability in all the possible embeddings of the graph, then it is called the best embedding of the graph, while if it gives the least comfortability, then it is called the worst embedding of the graph. Let us find the best and worst embeddings of the complete graph KnK_{n} by using the following famous graph theoretical facts.

Fact 1 (Minimal and maximal genera of KnK_{n}).
  1. (1)

    The (minimal) genus of KnK_{n}  (Rindel and Youngs (1968) [18]):

    orientable:\displaystyle\text{{\rm orientable}}:\; γ(Kn)=(n3)(n4)12,\displaystyle\gamma(K_{n})=\left\lceil\frac{(n-3)(n-4)}{12}\right\rceil,
    non-orientable:\displaystyle\text{{\rm non-orientable}}:\; γ~(Kn)={(n3)(n4)6n7,3n=7.\displaystyle\tilde{\gamma}(K_{n})=\begin{cases}\left\lceil\frac{(n-3)(n-4)}{6}\right\rceil&\text{: $n\neq 7$,}\\ 3&\text{: $n=7$.}\end{cases}
  2. (2)

    The maximal genus of KnK_{n} for the orientable surface  (Nordhaus and Stewart (1971) [16]):

    γM(Kn)=(n1)(n2)4.\gamma_{M}(K_{n})=\left\lfloor\frac{(n-1)(n-2)}{4}\right\rfloor.
  3. (3)

    The maximal genus of a connected graph G=(V,E)G=(V,E) for the non-orientable surface ([6, 14] and its reference therein):

    γ~M(G)=β(G),\tilde{\gamma}_{M}(G)=\beta(G), (2.3)

    where β(G)\beta(G) is the betti number of GG, that is, β(G)=|E||V|+1\beta(G)=|E|-|V|+1.

By combining Corollary 2.1 with Fact 1, the most comfortable underlying embedding of KnK_{n} for quantum walker can be chractorized as follows.

Corollary 2.2.

Under the setting of Corollary 2.1, the best and worst embeddings of the complete graph KnK_{n} for the comfortability of quantum walker must be on the closed surfaces with the minimal and maximal genus, respectively, whose orientability are divided into cases of mod(n,4)\mod(n,4) as follows.

  1. (1)

    n1,2mod4n\equiv 1,2\mod 4 case:

    • The best embedding: any non-orientable surface

    • The worst embedding: any orientable surface

  2. (2)

    n0,3mod4n\equiv 0,3\mod 4 and n3,4,7n\neq 3,4,7 case:

    • The best embedding: any orientable and non-orientable surfaces

    • The worst embedding: a non-orientable surface

  3. (3)

    n=3,4,7n=3,4,7 case:

    • The best embedding: any orientable surface

    • The worst embedding: a non-orientable surface

The best and worst surfaces of KnK_{n} for quantum walker is described in the following Table.

n1,2mod4n\equiv 1,2\mod 4 n0,3mod4n\equiv 0,3\mod 4, n3,4,7n\neq 3,4,7 n=3,4,7n=3,4,7
Best Non-ori Ori and Non-ori Ori
Worst Ori Non-ori Non-ori
Table 1: The best and worst surfaces of KnK_{n} for quantum walker: The genera for the best and worst surfaces are minimal and maximal genera, respectively. The best embedding is close to the triangulation, because 2|E|=3(|E||V|+2(n3)(n4)/6)2|E|=3(|E|-|V|+2-(n-3)(n-4)/6) holds. In particular, if n0,3mod4n\equiv 0,3\mod 4 with n8,11mod12n\not\equiv 8,11\mod 12, then the best embedding is the triangulation. Since every worst embedding is a single-face embedding, the comfortability of the worst embedding is 0 for n1,2mod4n\equiv 1,2\mod 4.
Proof.

Let us find that the best embedding and the worst embedding of KnK_{n} for quantum walker as follows.

  1. (1)

    Best: The expression (2.2) tells us that the large number of faces (the first term) and small number of self-intersection (the second term) makes quantum walker feel comfortable. Thus we expect that the small genus embedding will be the best embedding because the small genus accomplishes both of them.

    First let us estimate the number of faces for the minimum genus embedding. Combining Fact 1 with the Eular formula

    |F|=|E||V|+2{2γ(Kn): Orientable caseγ~(Kn): Non-orientable case,|F|=|E|-|V|+2-\begin{cases}2\gamma(K_{n})&\text{: Orientable case}\\ \tilde{\gamma}(K_{n})&\text{: Non-orientable case}\end{cases},

    we have

    {|Fori|=|Fnonori|:n0,3mod4 and n3,4,7,|Fori|<|Fnonori|:n1,2mod4,|Fori|>|Fnonori|:n=3,4,7.\left\{\,\begin{aligned} &|F^{ori}|=|F^{non-ori}|\quad:n\equiv 0,3\mod 4\text{ and }n\neq 3,4,7,\\ &|F^{ori}|<|F^{non-ori}|\quad:n\equiv 1,2\mod 4,\\ &|F^{ori}|>|F^{non-ori}|\quad:n=3,4,7.\end{aligned}\right. (2.4)

    Secondly, let us check whether the minimal genus embedding has a self-intersection. In the following, let us confirm that

    there are no intersections in the minimum genus embeddings on both “non-orientanble surfaces” and “orienable surfaces except n2,5mod12n\equiv 2,5\mod 12”.

    Put nn, m=n(n1)/2m=n(n-1)/2 and \ell as the numbers of vertices, edges and faces for the resulting embbeding of KnK_{n}, respectively. Assume that there is a face having a self-intersection in an embedding of KnK_{n}. Note that the boundary length of a face having a self-intersection must be at least 88. Then the following inequality holds.

    2m3×(1)+8×1.2m\geq 3\times(\ell-1)+8\times 1. (2.5)
    1. (a)

      Non-orientable case: Let kk be the genus of the underlying closed surface of the embedding having self-intersections. By the Euler formula and (2.5), we have

      k\displaystyle k =2nm+\displaystyle=2-n-m+\ell
      2n+m3+53\displaystyle\geq 2-n+\frac{m}{3}+\frac{5}{3}
      =16{(n3)(n4)+10}.\displaystyle=\frac{1}{6}\left\{(n-3)(n-4)+10\right\}. (2.6)

      This inequality is equivalent to

      k{γ~(Kn)+10/6n0,1,3,4mod6 and n7,(γ~(Kn)4/6)+10/6n2,5mod6,(γ~(Kn)1)+10/6n=7,k\geq\begin{cases}\tilde{\gamma}(K_{n})+10/6&\text{: $n\equiv 0,1,3,4\mod 6$ and $n\neq 7$,}\\ (\tilde{\gamma}(K_{n})-4/6)+10/6&\text{: $n\equiv 2,5\mod 6$,}\\ (\tilde{\gamma}(K_{n})-1)+10/6&\text{: $n=7$,}\end{cases}

      which implies

      k>γ~(Kn)k>\tilde{\gamma}(K_{n}) (2.7)

      and the embeddings on the non-orientable surfaces with the minimal genus has no self-intersections. Then the second term of (2.2) for the minimal genus embedding on the non-orientable surface is reduced to 0.

    2. (b)

      Orientable case: By the Euler formula for the orientable case, by replacing kk with 2g2g in (2.6),

      2g16{(n3)(n4)+10},2g\geq\frac{1}{6}\left\{(n-3)(n-4)+10\right\}, (2.8)

      which is equivalent to

      2g{2γ(Kn)+10/6n0,3,4,7mod12(2γ(Kn)1)+10/6n1,6,8,10mod122γ(Kn)n2,5mod12(2γ(Kn)4/6)+10/6n8,11mod12.2g\geq\begin{cases}2\gamma(K_{n})+10/6&\text{: $n\equiv 0,3,4,7\mod 12$}\\ (2\gamma(K_{n})-1)+10/6&\text{: $n\equiv 1,6,8,10\mod 12$}\\ 2\gamma(K_{n})&\text{: $n\equiv 2,5\mod 12$}\\ (2\gamma(K_{n})-4/6)+10/6&\text{: $n\equiv 8,11\mod 12$.}\end{cases}

      Then we have

      g>γ(Kn)g>\gamma(K_{n}) (2.9)

      except n2,5mod6n\equiv 2,5\mod 6. Thus it is ensured that there are no interactions in the minimum genus embedding except n2,5mod12n\equiv 2,5\mod 12. So the second term of (2.2) for the minimal genus embedding on the orientable surface is reduced to 0 except n2,5mod12n\equiv 2,5\mod 12.

    Thus combining (2.4) with (2.7) and (2.9), we obtain that the best embedding of the comfortability is the minimum genus embedding on

    {both orientable and non-orientable surfacesn0,3mod4 and n3,4,7non-orientable surfacen1,2mod4orientable surfacen=3,4,7.\begin{cases}\text{both orientable and non-orientable surfaces}&\text{: $n\equiv 0,3\mod 4$ and $n\neq 3,4,7$}\\ \text{non-orientable surface}&\text{: $n\equiv 1,2\mod 4$}\\ \text{orientable surface}&\text{: $n=3,4,7$.}\end{cases} (2.10)
  2. (2)

    Worst: The expression (2.2) tells us that the small number of faces (the first term) and large number of self-intersection (the second term) makes quantum walker feel uncomfortable. Thus we expect that the large genus embedding will be the worst embedding because the large genus accomplishes both of them.

    The number of faces of the maximal genus of orientable and non-orientable surfaces are

    |FMori|={1n1,2mod42n0,3mod4 and |FNnonori(Kn)|=1,|F_{M}^{ori}|=\begin{cases}1&\text{: $n\equiv 1,2\mod 4$}\\ 2&\text{: $n\equiv 0,3\mod 4$}\end{cases}\text{ and }|F_{N}^{non-ori}(K_{n})|=1,

    respectively.

    1. (a)

      n1,2mod4n\equiv 1,2\mod 4 case: If the underlying surface is orientable, then there is only one face and every boundary face has the self-intersection, which means that comfortability is reduced to 0 by (2.2). On the other hand, if the underlying surface is non-orientable, then there must exists at least one edge having no self-intersection in the face, which implies that the comfortability is non-zero by (2.2). Thus we have the worst embedding is the maximal genus embedding on the orientable surface.

    2. (b)

      n0,3mod4n\equiv 0,3\mod 4 case: If the underlying surface is orientable, then the number of faces is 22. Then the boundary edges of the two faces have no self-intersections. If a twist is inserted into one of the edges of the boundary edges so called the edge twisted surgery [6, 14], then the two faces are merged into a single face conserving the self intersected edges and the resulting surface becomes non-orientable. Then this single-face embedding on the non-orientable surface is worse than the embedding with double-face on the orientable surface.

    The worst embedding of the comfortability is the maximal genus embedding on

    {non-orientable surfacesn0,3mod4orientable surfacen1,2mod4\begin{cases}\text{non-orientable surfaces}&\text{: $n\equiv 0,3\mod 4$}\\ \text{orientable surface}&\text{: $n\equiv 1,2\mod 4$}\end{cases} (2.11)

Therefore (2.10) and (2.11) lead the desired conclusion. ∎

2.2 Observation 2: Comfortability on the island.

It is shown in the proof of Theorem 7.1 that the comfortability on the island |Ais\mathcal{E}|_{A_{is}} is proportional to the first term of (1.1). Then let us estimate the first term of (1.1) by using some combinatorial method.

Set

h(x)=x1+ax1ax.h(x)=x\frac{1+a^{x}}{1-a^{x}}.

Let F={f1,,fr}F=\{f_{1},\dots,f_{r}\}, with |f1||f2||fr||f_{1}|\geq|f_{2}|\geq\cdots\geq|f_{r}| be the set of the underlying faces, where r=|F|r=|F|. Then the first term of (1.1) in Theorem 1.2 can be reexpressed by

1|A|2+|b|2|b|2x{|f1|,,|fr|}h(x).\frac{1}{|A|}\frac{2+|b|^{2}}{|b|^{2}}\sum_{x\in\{|f_{1}|,\dots,|f_{r}|\}}h(x).

Note that |A|=|f1|++|fr||A|=|f_{1}|+\cdots+|f_{r}|. Then the boundary lenghts of FF give the integer partition, which is bijective to the Young diagram. Thus in the following, let us consider what is the integer partition λ|A|\lambda\vdash|A| makes the first term larger. Let us the important properties of h(x)h(x) for the above consideration be summarised below.

Properties of h(x)h(x):

  1. (1)

    For ,m\ell,m\in\mathbb{N},

    h(+m)<h()+h(m)(<h(+m)+h(0))h(\ell+m)<h(\ell)+h(m)\;(<h(\ell+m)+h(0)\;)
  2. (2)

    For j,mj\ell_{j},m_{j}\in\mathbb{N} (j{1,2}j\in\{1,2\}) with 1+m1=2+m2\ell_{1}+m_{1}=\ell_{2}+m_{2} and |1m1|<|2m2||\ell_{1}-m_{1}|<|\ell_{2}-m_{2}|, then

    h(1)+h(m1)<h(2)+h(m2).h(\ell_{1})+h(m_{1})<h(\ell_{2})+h(m_{2}).

For λ=(x1,x2,,xr)|A|\lambda=(x_{1},x_{2},\dots,x_{r})\vdash|A| with x1x2xrx_{1}\geq x_{2}\geq\cdots\geq x_{r}, we subject the condition xr3x_{r}\geq 3 because the length of face is larger than 33. For such a Young diagram λ=(x1,,xr)\lambda=(x_{1},\dots,x_{r}), let us put Q(λ)Q(\lambda) by

Q(λ)=x{x1,,xs}h(x).Q(\lambda)=\sum_{x\in\{x_{1},\dots,x_{s}\}}h(x).

We call it the island of the energy. Define the partially order λ1,λ2|A|\lambda_{1},\lambda_{2}\vdash|A| by λ1>λ2\lambda_{1}>\lambda_{2} if and only if Q(λ1)>Q(λ2)Q(\lambda_{1})>Q(\lambda_{2}). By property (1), large length of the Young diagram makes Q()Q(\cdot) large. By property (2), bias of sizes of rows also makes Q()Q(\cdot) large. Thus for any Young diagram λ|A|\lambda\vdash|A|, we have

[|A|]λ{[3,3,,3]|A|=0mod3,[4,3,,3]|A|=1mod3,[5,3,,3]|A|=2mod3.[\;|A|\;]\leq\lambda\leq\begin{cases}[3,3,\dots,3]&\text{: $|A|=0\mod{3}$},\\ [4,3,\dots,3]&\text{: $|A|=1\mod{3}$},\\ [5,3,\dots,3]&\text{: $|A|=2\mod{3}$}.\end{cases}

Then we immediately obtain the following corollary.

Corollary 2.3.

Let GG be a connected abstract graph. If there is a triangulation in the embeddings of GG, then that is one of the best embeddings for quantum walker.

Proof.

If there is a triangulation in the embeddings, then that is the most comfortable on the island. Moreover since there is no self-intersections in the triangulation, then that is the most comfortable embedding by Corollary 2.1. ∎

If n0,3,4,7mod12n\equiv 0,3,4,7\mod 12, the minimum genus embedding of KnK_{n} becomes the triangulation, because 3|F|=2|E|3|F|=2|E| holds by Fact 1. Then we can see that Corollary 2.3 is consistent with Corollary 2.2 for KnK_{n}.

Example: The Hasse diagram in Fig. 3 describes the case for |A|=12|A|=12 by only using the properties of h(x)h(x), (1) and (2).

Refer to caption
Figure 3: Hasse diagram with respect to the comfortability on the island.

This Hasse diagram shows the partial order of the comfortability restricted to the islands for K4K_{4}’s embeddings. In this Hasse diagraph, the order of the Young diagram [4,4,4][4,4,4] and [9,3][9,3] are not determined by only using the properties (1) and (2) of h(x)h(x). However it is possible to estimate the difference between the comfortabilities on the island to λ=[9,3]\lambda=[9,3] and μ=[4,4,4]\mu=[4,4,4] in the Hasse diagram of Fig 3, Q([9,3])Q([9,3]) and Q([4,4,4])Q([4,4,4]) as follows:

Q([9,3])<Q([5,4,3])<Q([4,4,3,1])<Q([4,4,4])+h(0).\displaystyle Q([9,3])<Q([5,4,3])<Q([4,4,3,1])<Q([4,4,4])+h(0).

Here all the inequalities derive from the property (1) of h(x)h(x). On the other hand,

Q([4,4,4])<Q([5,4,3])<Q([9,3])+h(0)\displaystyle Q([4,4,4])<Q([5,4,3])<Q([9,3])+h(0)

Here the first and second inequalities derive from the property (2) and (1), respectively. Then we have

|Q([9,3])Q([4,4,4])|<h(0)=2/|loga|.|Q([9,3])-Q([4,4,4])|<h(0)=2/|\log a|.

In the statement of Corollary 2.1, the comfortability of ``k=2,[8,4]"``k=2,\;[8,4]" is greater than that of ``g=1,[8,4]"``g=1,\;[8,4]" (see Figure 2), while these are the same by in the Hasse diagram of Fig. 3 on the island. The difference derives from the number of intersections; there are 22 points of the intersection in the enneagon for g=1g=1 while there is only 11 point of the intersection in the enneagon for k=2k=2. See Figure 4. The same reason can be applied to the case for ``k=2,[9,3]"``k=2,\;[9,3]" and ``g=1,[9,3]"``g=1,\;[9,3]". Therefore the non-orientability reduces the self-intersection and increases the comfortability.

Refer to caption
Figure 4: The self-intersection. The comparison between the embeddings on the torus and the Klein bottle whose both faces are 11 octagon and 11 square. The numbers of self-intersections in the octagon at embeddings on the torus and the Klein bottle are 22 and 11, respectively. Thus the embedding on the Klein bottle is better than the embedding on the torus in the setting of Corollary 2.1.

3 Quick review on the rotation system

In this section, we give a quick review on the rotation system following [6, 14], which will be important to construct our quantum walk model.

Abstract graph. Let G(X,A)G(X,A) be a symmetric digraph with the set of the vertices XX and the set of the symmetric arcs AA, that is, eAe\in A if and only if e¯A\bar{e}\in A, where e¯\bar{e} is the inverse arc of ee. In this paper, we assume there is no multiple arcs. Then an arc eAe\in A is sometimes represented by (o(e),t(e))(o(e),t(e)) to emphasize the origin and terminal vertices. The support edge of eAe\in A is denoted by |e|=|e¯||e|=|\bar{e}|. The set of edges (which are undirected) is defined by E={|e||eA}E=\{|e|\;|\;e\in A\}. For each arc eAe\in A, the terminus and origin vertices are denoted by t(e)Xt(e)\in X and o(e)Xo(e)\in X. The arc eAe\in A with o(e)=xo(e)=x, t(e)=yt(e)=y is also represented by a=(x,y)Aa=(x,y)\in A. Let AxAA_{x}\subset A be the set of arcs whose terminal vertices are xXx\in X, that is, Ax={eA|t(e)=x}A_{x}=\{e\in A\;|\;t(e)=x\}.

Rotation. A cyclic permutation on a countable set WW is a bijection map π:WW\pi:W\to W such that π(ω)ω\pi(\omega)\neq\omega for any ωW\omega\in W. On each vertex xXx\in X, a cyclic permutation on AxA_{x}, ρx:AxAx\rho_{x}:A_{x}\to A_{x}, is assigned. The extension of ρx\rho_{x} to the whole arc set AA is given by

ρ~x(e)={ρx(e)eAx,e: otherwise.\tilde{\rho}_{x}(e)=\begin{cases}\rho_{x}(e)&\text{: $e\in A_{x}$,}\\ e&\text{: otherwise.}\end{cases}

The rotation ρ\rho on the symmetric arc set AA is defined by

ρ=(xXρ~x).\rho=\left(\prod_{x\in X}\tilde{\rho}_{x}\right).

Twist. The designation for the twist of each edge is a map τ:A2\tau:A\to\mathbb{Z}_{2} such that τ(e)=τ(e¯)\tau(e)=\tau(\bar{e}) for any eAe\in A. If τ(e)=0\tau(e)=0, the edge eAe\in A is called type-0, otherwise, called type-11. The type-11 edge is regarded as twisted.

Rotation system. The triple (G,ρ,τ)(G,\rho,\tau) is called the rotation system of GG. GG alone is sometimes called an abstract graph. Let σ\sigma be the transposition such that σ(e)=e¯\sigma(e)=\bar{e} for any eAe\in A. A facial walk induced by the rotation system is the representative of the sequences of arcs identified by cyclic permutation and inverse

(e0,e1,,er1)(e_{0},e_{1},\dots,e_{r-1})

with some natural number r>2r>2 such that

ej+1={σ(ρ(ej))k=0jτ(|ej|)=0σ(ρ1(ej))k=0jτ(|ej|)=1e_{j+1}=\begin{cases}\sigma(\rho(e_{j}))&\text{: $\sum_{k=0}^{j}\tau(|e_{j}|)=0$}\\ \sigma(\rho^{-1}(e_{j}))&\text{: $\sum_{k=0}^{j}\tau(|e_{j}|)=1$}\end{cases} (3.12)

for j=0,1,,r1j=0,1,\dots,r-1 in modulus of rr. See Figure 5 for a rotation system of K4K_{4} and its facial walks. The set of facial walks is denoted by FF. Every fFf\in F gives the boundary of each face of the resulting two-cell embedding of the closed surface. Indeed, it is very starting point for our paper that the orientation system determines the embedding on a closed surface in the following meaning.

Theorem 3.1 ([6, 14, 15]).

Every rotation system on graph GG defines (up to equivalence of embeddings) a unique locally oriented graph embedding. Conversely, every locally oriented graph embedding defines a rotation system for GG.

The Eular’s formula gives the genus of the undelying embedding:

genus={g=12(|E||V||F|)+1: if the surface is orientable,k=|E||V||F|+2: if the surface is non-orientable.genus=\begin{cases}g=\frac{1}{2}(|E|-|V|-|F|)+1&\text{: if the surface is orientable,}\\ k=|E|-|V|-|F|+2&\text{: if the surface is non-orientable.}\end{cases}

The graph GG contains a cycle which has odd number of type-11 edges if and only if the underlying surface is non-orientable. The following method is useful to detect the orientability of the underlying surface[6, 14]. See also Figure 6:

Operation.

  1. i.

    Choose an arbitrary spanning tree TT of GG;

  2. ii.

    Choose a arbitrary vertex xx of TT; xx is called the root;

  3. iii.

    For any vertex which is adjacent to xx with the type-11 edge, say yy, the orientation is reversed from ρy\rho_{y} to ρy1\rho_{y}^{-1} and for any edge incident to yy in GG, say ee, the type is reversed from τ(e)\tau(e) to τ(e)+1\tau(e)+1;

  4. iv.

    Continue this process until all the types of edges in the tree are type-0;

  5. v.

    The surface is non-orientable if and only if there exists a type-11 edge in GTG\setminus T after the above updated orientation system.

The following remark will play an important role in the construction of our quantum walk model.

Remark 3.1.

For any closed walk (e0,e1,,es1)(e_{0},e_{1},\dots,e_{s-1}) on GG, the operation iii preserves the value of j=0s1τ(ej)\sum_{j=0}^{s-1}\tau(e_{j}), which ensures that the underlying embedding is preserved by this operation iii.

After the underlying closed surface is determined by the above procedures, for every vertex xx, we first arrange a vertex xx and its “half” edges connecting to xx clockwise according to the rotation ρx\rho_{x} on the surface. Secondly, if xx and yy are connected in GG, then connect the corresponding half-edges each other without any crossing of edges so that the type of edge is conserved. See Figure 7.

Refer to caption
Figure 5: A rotation system of K4K_{4} and facial closed walks: The rotation is assigned clockwise at each vertex, and the twist is assigned at the edge {0,1}\{0,1\}. There are 33 faces in this rotation system; 22 triangles and 11 hexagon.
Refer to caption
Figure 6: The detection of the orientability: In the rotation system, vertex 11 is selected as the root of the spanning tree, and twist the vertex 11 so that every type of edge in the tree become 0. There are 22 fundamental cycles containing a type-11 edge. Thus the closed surface must be non-orientable. By the Euler formula, the genus is k=2(|F||E|+|V|)=2(36+4)=1k=2-(|F|-|E|+|V|)=2-(3-6+4)=1. Then the surface is the projective plane.
Refer to caption
Figure 7: The drawing on the closed surface: Diagonally located places on the dotted boundary are identified with each other. Incident half edges of each vertex are arranged clockwise so that its rotation is conserved. Connect corresponding half-edges without any crossing and every type-0 edge passes through the dotted boundary evenly, while every type-11 does so oddly. In this case, there are no crossings across the dotted boundary in every the type-0, while the type-11 edge crosses the dotted boundary once.

4 Construction of quantum walk on the rotation system (G,ρ,τ)(G,\rho,\tau)

On the graph embedded on a closed surface, when a walker passes through a twisted edge, the rotation of the endpoint vertex is reversed. To reflect the parity of the “sheet”, that is, front/back, in the dynamics of our quantum walk, let us prepare the following notions.

4.1 Double covering

The rotation system (G,ρ1,τ)(G,\rho^{-1},\tau) is called the chiral rotation system of (G,ρ,τ)(G,\rho,\tau). To realize the parity and the rotation in the dynamics of the quantum walk model we first prepare the two rotation systems which are chiral to each other and secondly join them by rewiring the twisted edges to the corresponding chiral vertices so that the original and its chiral rotations are conserved.

More precisely, let us first set the double covering graph Gτ=(Xτ,Aτ)G^{\tau}=(X^{\tau},A^{\tau}) which is realized by the voltage assignment of τ:2A\tau:\mathbb{Z}_{2}\to A as follows. See also Figure 8 (b).

  1. (1)

    The vertex set : Xτ=X×2=X0X1X^{\tau}=X\times\mathbb{Z}_{2}=X_{0}\sqcup X_{1}, where Xj={(x.j)|xX}X_{j}=\{(x.j)\;|\;x\in X\} (j2)(j\in\mathbb{Z}_{2}).

  2. (2)

    The arc set : (x,j)Xτ(x,j)\in X^{\tau} and (y,k)Xτ(y,k)\in X^{\tau} are adjacent each other in GτG^{\tau} if and only if xx and yy are adjacent in GG and k=j+τ((x,y))k=j+\tau(\;(x,y)\;).

Secondly, to reproduce the rotation ρ\rho, the rotation ρ\rho is assigned to X0X_{0}, while the inverse rotation ρ1\rho^{-1} is assigned to X1X_{1}. Such a resulting rotation system is called the double covering of (G,ρ,τ)(G,\rho,\tau), which coincides with the rotation system (Gτ,ρρ1,id)(G^{\tau},\rho\oplus\rho^{-1},\mathrm{id}).

It is possible to draw the abstract graph GτG^{\tau} in the following way:

(*) X0X_{0} is placed in the “left”
while X1X_{1} is placed in the “right”.

We call the regions of the subgraphs in the left and right, the front and back sheets, respectively. Note that every vertex in the front (resp. back) sheet follows the rotation ρ\rho (resp. ρ1\rho^{-1}), respectively. Every arc crossing the boundary of the sheets ((x,j),(y,k))(\;(x,j),(y,k)\;) satisfies (x,y)A(x,y)\in A and kjk\neq j. See Figure 8 (b).

Lemma 4.1.

Under the drawing (*) of GτG^{\tau}, Operation iii to a vertex yXy\in X in (G,ρ,τ)(G,\rho,\tau) is equivalent to pulling (y,0)(y,0) and its chiral (y,1)(y,1) to their opposite sheets crossing the boundary and conserving the locations of the other vertices of GτG^{\tau}, and switching the labeling of each vertex.

Proof.

Note that the situation of a vertex yy with its connected edges Dy={e1,,er}D_{y}=\{e_{1},\dots,e_{r}\} in (G,ρ,τ)(G,\rho,\tau) is equivalent to the situation in the drawing (*) of GτG^{\tau} that all the edges of DyD_{y} and its chiral edges of DyD_{y^{\prime}} with τ(ej)=0\tau(e_{j})=0 are in the front sheet and back sheet, respectively, while all the edges with τ(ej)=1\tau(e_{j})=1 cross the boundary. The operation iii. to a vertex yy switches the rotation and types of all its incoming and outgoing arcs. Then in the situation of GτG^{\tau} with the drawing (*), the vertex yy and its chiral vertex yy^{\prime} are pulled to the opposite sheets conserving the locations of the other vertices. After this switching, the resulting drawing of GτG^{\tau}, the rotations of yy and yy^{\prime}, and also the types of all the connected arcs are switched. ∎

Proposition 4.1.

The rotation system (G,ρ,τ)(G,\rho,\tau) is non-orientable if and only if GτG^{\tau} is connected.

Proof.

Every orientable rotation system can be drawn without any twisted edges, (G,ρ,id)(G,\rho,\mathrm{id}), by Operation iiv. The rotation system (G,ρ,id)(G,\rho,\mathrm{id}) is orientable and the corresponding double covering graph is constructed by the two connected components, that is, disconnected because there are no type-11 edges. The operation reversing the rotation of a vertex and the types of its connected edges keeps the disconnectivity of the double covering graph by the previous lemma. ∎

4.2 Blow-up

The blow up graph of G=(V,A)G=(V,A) with the rotation ρ\rho is obtained by replacing each vertex of GG into the directed cycle following the rotation assigned to each vertex and conserving the adjacency relation of the original graph as follows. The vertex set of the blow up graph GBU(ρ)=(XBU(ρ),ABU(ρ))G^{BU}(\rho)=(X^{BU}(\rho),A^{BU}(\rho)) is denoted by

XBU(ρ)=A;X^{BU}(\rho)=A;

the arc set of the blow up graph is denoted by the disjoint union of

ABU(ρ)=AbrAis,A^{BU}(\rho)=A_{br}\sqcup A_{is},

which are called the bridge and the island, respectively. Here the “bridge” AbrA_{br} and the “island” AisA_{is} are both isomorphic to AA and denoted by

Ais={(e,ρ(e))|eA},Abr={(e,e¯)|eA}.A_{is}=\{(e,\rho(e))\;|\;e\in A\},\;A_{br}=\{(e,\bar{e})\;|\;e\in A\}.

See Figure 8 (c).

Let us give some remarks and introduce some new notations which will be important to describe the time evolution of our quantum walk model.

Remark 4.1.

It holds that for eAe\in A with t(e)=xXt(e)=x\in X in GG,

{(e,ρ(e))|t(e)=x in G}={(ρj(e),ρj+1(e))|j=0,1,,degG(x)1}Ais.\{(e^{\prime},\;\rho(e^{\prime}))\;|\;\text{$t(e^{\prime})=x$ in $G$}\}=\{(\;\rho^{j}(e),\;\rho^{j+1}(e)\;)\;|\;j=0,1,\dots,\deg_{G}(x)-1\}\subset A_{is}.

We call the above directed cycle induced by xXx\in X the island of xXx\in X which is denoted by Ais(x)A_{is}(x). On the other hand, for ϵE\epsilon\in E with the end vertices x,yXx,y\in X in GG, (e,e¯)(e,\bar{e}) and (e¯,e)(\bar{e},e), where |e|=ϵ|e|=\epsilon, are called the bridge between the islands xx and yy. Then let us define

ρ(𝒆is):=(ρ(e),ρ2(e))\rho(\boldsymbol{e}_{is}):=(\rho(e),\rho^{2}(e))

for an island arc 𝒆is=(e,ρ(e))\boldsymbol{e}_{is}=(e,\rho(e)), and

𝒆¯br:=(e¯,e),τ(𝒆br)=τ(|e|)\bar{\boldsymbol{e}}_{br}:=(\bar{e},e),\;\tau(\boldsymbol{e}_{br})=\tau(|e|)

for a bridge arc 𝒆br=(e,e¯)\boldsymbol{e}_{br}=(e,\bar{e}).

Remark 4.2.

The in-degree and out-degree for every vertex of the blow-up graph are equally 22.

For each vertex eXBUe\in X^{BU}, the incoming arcs to ee come from the vertex ρ1(e)\rho^{-1}(e) along the island of tG(e)t_{G}(e) and the vertex e¯\bar{e} along the bridge between the islands t(e)t(e) and o(e)o(e) while the outgoing arcs to ee go out to the vertex ρ(e)\rho(e) along the island of tG(e)t_{G}(e) and the vertex e¯\bar{e} along the bridge between the islands t(e)t(e) and o(e)o(e).

Then for any island arc 𝒆isAis\boldsymbol{e}_{is}\in A_{is}, there are exactly two bridge arcs 𝒆br\boldsymbol{e}_{br} and 𝒆brAbr\boldsymbol{e}_{br}^{\prime}\in A_{br} such that o(𝒆is)=t(𝒆br)o(\boldsymbol{e}_{is})=t(\boldsymbol{e}_{br}) and t(𝒆is)=t(𝒆br)t(\boldsymbol{e}_{is})=t(\boldsymbol{e}_{br}^{\prime}), respectively. Such bridge arcs for the island arc 𝒆is\boldsymbol{e}_{is} are denoted by

𝒆br:=br(𝒆is),𝒆br:=br(𝒆is),\boldsymbol{e}_{br}:=\mathrm{br}(\boldsymbol{e}_{is}),\;\boldsymbol{e}_{br}^{\prime}:=\mathrm{br}^{\sharp}(\boldsymbol{e}_{is}),

respectively. On the other hand, for any bridge arc 𝒆brAbr\boldsymbol{e}_{br}\in A_{br}, there is exactly two island arcs 𝒆is\boldsymbol{e}_{is} and 𝒆is\boldsymbol{e}_{is}^{\prime} such that o(𝒆br)=t(𝒆is)=o(𝒆is)o(\boldsymbol{e}_{br})=t(\boldsymbol{e}_{is})=o(\boldsymbol{e}_{is}^{\prime}). Such island arcs for the bridge arc 𝒆br\boldsymbol{e}_{br} are denoted by

𝒆is:=is(𝒆br),𝒆is:=is(𝒆br),\boldsymbol{e}_{is}:=\mathrm{is}(\boldsymbol{e}_{br}),\;\boldsymbol{e}_{is}^{\prime}:=\mathrm{is}^{\sharp}(\boldsymbol{e}_{br}),

respectively.

Refer to caption
Figure 8: The construction of G(ρ,τ)G(\rho,\tau): (a) the rotation system of K4K_{4} embedding in the projective plain (G,ρ,τ)(G,\rho,\tau); (b) the rotation system of the double covering graph GτG^{\tau}, (Gτ,ρρ1,id)(G^{\tau},\rho\oplus\rho^{-1},\mathrm{id}); (c) the blow-up graph of G(ρ,τ)G(\rho,\tau); (d) the blow-up graph with tails G~(ρ,τ)\tilde{G}(\rho,\tau) for δAis=Ais\delta A_{is}=A_{is} case (every tail is inserted on each island arc, and such a tail arrangement is called hedgehog.)

4.3 Quantum walk on the rotation system (G,ρ,τ)(G,\rho,\tau)

For given the rotation system (G,ρ,τ)(G,\rho,\tau), let us consider the blow up graph of GτG^{\tau} induced by the rotation ρρ1\rho\oplus\rho^{-1} by replacing GG discussed in the previous subsection 4.2 with GτG^{\tau} discussed also in subsection 4.1. This blow up graph is denoted by G(ρ,τ)=(X(ρ,τ),A(ρ,τ))G(\rho,\tau)=(X(\rho,\tau),A(\rho,\tau)). Note that G(ρ,τ)=GτBU(ρρ1)G(\rho,\tau)={G^{\tau}}^{BU}(\rho\oplus\rho^{-1}).

Now we are ready to describe the dynamics of this quantum walk model. The total state space of our quantum walk is given by the vector space

total=A(ρ,τ).\mathcal{H}_{total}=\mathbb{C}^{A(\rho,\tau)}.

In our quantum walk model, a local time evolution at each vertex of G(ρ,τ)G(\rho,\tau) is represented by a 22-dimensional unitary matrix, because the degree of the blow up graph is 22, and if a quantum walk passes through the twisted bridge, the phase of the quantum walker is converted by eiπ=1e^{i\pi}=-1.

In the following, we represent this dynamics of our quantum walk model in more precisely. Let W:A(ρ,τ)A(ρ,τ)W:\mathbb{C}^{A(\rho,\tau)}\to\mathbb{C}^{A(\rho,\tau)} be the unitary time evolution operator such that for each time step tt\in\mathbb{N}, the total state of the quantum walk at time tt, ψtA(ρ,τ)\psi_{t}\in\mathbb{C}^{A(\rho,\tau)}, is given by

ψt+1=Wψt for t0\psi_{t+1}=W\psi_{t}\text{ for $t\geq 0$}

with some initial state ψ0A(ρ,τ)\psi_{0}\in\mathbb{C}^{A(\rho,\tau)}. The unitary time evolution operator WW is defined as follows.

Definition 4.1.

Let us set a 2×22\times 2 unitary matrix by

C=[abcd].C=\begin{bmatrix}a&b\\ c&d\end{bmatrix}.

For any island arc 𝐞isAisA(ρ,τ)\boldsymbol{e}_{is}\in A_{is}\subset A(\rho,\tau), bridge arc 𝐞brAbrA(ρ,τ)\boldsymbol{e}_{br}\in A_{br}\subset A(\rho,\tau), the time evolution operator WW is defined by

(Wψ)(𝒆is)\displaystyle(W\psi)(\boldsymbol{e}_{is}) =aψ(ρ1(𝒆is))+bψ(br(𝒆is))\displaystyle=a\;\psi(\rho^{-1}(\boldsymbol{e}_{is}))+b\;\psi(\mathrm{br}(\boldsymbol{e}_{is})) (4.13)
(Wψ)(𝒆br)\displaystyle(W\psi)(\boldsymbol{e}_{br}) =(1)τ(𝒆br){cψ(is(𝒆br))+dψ(𝒆¯br)}\displaystyle=(-1)^{\tau(\boldsymbol{e}_{br})}\left\{c\;\psi(\mathrm{is}(\boldsymbol{e}_{br}))+d\;\psi(\bar{\boldsymbol{e}}_{br})\right\} (4.14)

for any ψA(ρ,τ)\psi\in\mathbb{C}^{A(\rho,\tau)}.

Remark 4.3.

For any bridge arc 𝐞brAbrA(ρ,τ)\boldsymbol{e}_{br}\in A_{br}\subset A(\rho,\tau), the local time evolution on the vertex o(𝐞br)X(ρ,τ)o(\boldsymbol{e}_{br})\in X(\rho,\tau) is described by

[ψt+1(is(𝒆br))ψt+1(𝒆br)]=[100(1)τ(𝒆br)]C[ψt(is(𝒆br))ψt(𝒆¯br)].\begin{bmatrix}\psi_{t+1}(\;\mathrm{is}^{\sharp}(\boldsymbol{e}_{br})\;)\\ \psi_{t+1}(\boldsymbol{e}_{br})\end{bmatrix}=\begin{bmatrix}1&0\\ 0&(-1)^{\tau(\boldsymbol{e}_{br})}\end{bmatrix}\;C\begin{bmatrix}\psi_{t}(\;\mathrm{is}(\boldsymbol{e}_{br})\;)\\ \psi_{t}(\bar{\boldsymbol{e}}_{br})\end{bmatrix}. (4.15)

4.4 Extension to an infinite system

For the blow-up graph G(ρ,τ)G(\rho,\tau), let us choose a subset of AisA_{is} by δAis\delta A_{is} as the boundary to the “outside”. We deform this blow-up graph to an infinite graph by the following procedure.

  1. (1)

    Every eA(ρ,τ)δAise\in A(\rho,\tau)\setminus\delta A_{is} is left as it is without any transformations.

  2. (2)

    Each eδAise\in\delta A_{is} is divided by two arcs by replacing ee with two new arcs ξout\xi_{out} and ξin\xi_{in} satisfying with o(e)=o(ξout)o(e)=o(\xi_{out}), t(ξout)=o(ξin)t(\xi_{out})=o(\xi_{in}) and t(ξin)=t(e)t(\xi_{in})=t(e).

  3. (3)

    Semi-infinite path e\mathbb{P}_{e} with the root vertex o(e)o(\mathbb{P}_{e}) is joined by identifying o(e)o(\mathbb{P}_{e}) with t(ξout)=o(ξin)t(\xi_{out})=o(\xi_{in}). The pair of (ξout,ξin)(\xi_{out},\xi_{in}) is called the quay of eδAise\in\delta A_{is}. The sets of such as ξout\xi_{out} and ξin\xi_{in} are denoted by δAqy\delta A_{qy}^{-} and δAqy+\delta A_{qy}^{+}, respectively.

The resulting graph is denoted by G~(ρ,τ)=(X~(ρ,τ),A~(ρ,τ)){\tilde{G}}(\rho,\tau)=(\tilde{X}(\rho,\tau),\tilde{A}(\rho,\tau)), which is an infinite graph. See Figure 8 (d). The set of boundary vertices of G~(ρ,τ)\tilde{G}(\rho,\tau) is defined by

δX=eδAiso(e)X~(ρ,τ).\delta X=\bigcup_{e\in\delta A_{is}}o(\mathbb{P}_{e})\subset{\tilde{X}}(\rho,\tau).

The set of arcs of tails is denoted by AtlA_{tl}. The subset of AtlA_{tl} called the pier is denoted by

δApr+={eAtl|t(e)δX},δApr={eAtl|o(e)δX}.\delta A_{pr}^{+}=\{e\in A_{tl}\;|\;t(e)\in\delta X\},\;\delta A_{pr}^{-}=\{e\in A_{tl}\;|\;o(e)\in\delta X\}.

Note that for each island having a boundary vertex, the local rotation in G~(ρ,τ)\tilde{G}(\rho,\tau) is “expanded” by the insertion of the tail. Let us use the same notation for the new rotation as ρ\rho to reduce the number of notations. Let A~isA~BU\tilde{A}_{is}\subset{\tilde{A}}^{BU} be the set of the islands of G~(ρ,τ)\tilde{G}(\rho,\tau) following the rotation ρ\rho. Let A~br\tilde{A}_{br} be the union of AbrA_{br} and δApr\delta A_{pr}. The designation of the twist to AtlA_{tl} is assigned by 0.

The total state space here is extended to

~total=A~(ρ,τ).\tilde{\mathcal{H}}_{total}=\mathbb{C}^{\tilde{A}(\rho,\tau)}.

The time evolution operator W~\tilde{W} on A~(ρ,τ)\mathbb{C}^{\tilde{A}(\rho,\tau)} is given as follows, which is essentially the same as Definition 4.1 except the boundaries and the tails but the setting of the initial state is crucial to obtain the stationary state:

Definition 4.2.

Let Ψt\Psi_{t} be the tt-th iteration of W~\tilde{W}, such that Ψt=W~Ψt1\Psi_{t}=\tilde{W}\Psi_{t-1} t=1,2,t=1,2,\dots with the initial state Ψ0\Psi_{0} defined in (4.17).

The time evolution W~\tilde{W}.

  1. (1)

    For 𝒆AtlδApr\boldsymbol{e}\in A_{tl}\setminus\delta A_{pr}^{-}: The arcs of a tail is put by e0,e1,e_{0},e_{1},\dots and e¯0,e¯1,\bar{e}_{0},\bar{e}_{1},\dots with t(e0)δXt(e_{0})\in\delta X t(ej+1)=o(ej)t(e_{j+1})=o(e_{j}) j=0,1,2,j=0,1,2,\dots. The time evolution on the tail is free, that is,

    Ψt+1(ej)=Ψt(ej+1),Ψt+1(e¯j+1)=Ψt(e¯j)\Psi_{t+1}(e_{j})=\Psi_{t}(e_{j+1}),\;\Psi_{t+1}(\bar{e}_{j+1})=\Psi_{t}(\bar{e}_{j}) (4.16)

    for any j=0,1,2,j=0,1,2,\dots.

  2. (2)

    For 𝒆A~isAbrδApr\boldsymbol{e}\in\tilde{A}_{is}\cup A_{br}\cup\delta A_{pr}^{-}: For any island arc 𝒆isA~is\boldsymbol{e}_{is}\in\tilde{A}_{is}, “bridge or pier” arc 𝒆brAbrδApr\boldsymbol{e}_{br}\in A_{br}\cup\delta A_{pr}^{-},

    Ψt+1(𝒆is)\displaystyle\Psi_{t+1}(\boldsymbol{e}_{is}) =aΨt(ρ1(𝒆is))+bΨt(br(𝒆is)),\displaystyle=a\;\Psi_{t}(\rho^{-1}(\boldsymbol{e}_{is}))+b\;\Psi_{t}(\mathrm{br}(\boldsymbol{e}_{is})),
    Ψt+1(𝒆br)\displaystyle\Psi_{t+1}(\boldsymbol{e}_{br}) =(1)τ(𝒆br){cΨt(is(𝒆br))+dΨt(𝒆¯br)}.\displaystyle=(-1)^{\tau(\boldsymbol{e}_{br})}\left\{c\;\Psi_{t}(\mathrm{is}(\boldsymbol{e}_{br}))+d\;\Psi_{t}(\bar{\boldsymbol{e}}_{br})\right\}.

    Here if 𝒆isδAqy±\boldsymbol{e}_{is}\in\delta A_{qy}^{\pm}, the bridge arc br(𝒆is)δApr\mathrm{br}(\boldsymbol{e}_{is})\in\delta A_{pr}^{-} is the pier from t(𝒆is)t(\boldsymbol{e}_{is}); if 𝒆brδApr\boldsymbol{e}_{br}\in\delta A_{pr}^{-}, the island arc is(𝒆br)δAqy\mathrm{is}(\boldsymbol{e}_{br})\in\delta A_{qy}^{-} is the quay of the boundary vertex o(𝒆br)o(\boldsymbol{e}_{br}).

The initial state Ψ0\Psi_{0}.
Let each tail is labeled by each element of δApr+\delta A_{pr}^{+}. Prepare the sequence of complex values (αe)eδApr+(\alpha_{e})_{e\in\delta A_{pr}^{+}}. Then we set the following uniformly bounded initial state on the tails:

Ψ0(e)={αeeA(e)dist(t(e),o(e))>dist(t(e),o(e)) (eδAis),0 otherwise.\Psi_{0}(e^{\prime})=\begin{cases}\alpha_{e}&\text{: $e^{\prime}\in A(\mathbb{P}_{e})$, ${\rm dist}(t(e^{\prime}),o(\mathbb{P}_{e}))>{\rm dist}(t(e^{\prime}),o(\mathbb{P}_{e}))$ ($e\in\delta A_{is}$),}\\ 0&\text{ otherwise.}\end{cases} (4.17)

Note that the initial state Ψ0\Psi_{0} is bounded but no longer square summable. Such a setting of initial state provides a constant inflow from the tails at very time step. On the other hand, the setting of the time evolution on tails implies an outflow from the interior. Under this setting, the convergence to a fixed point of this quantum walk is ensured in the following meaning:

Proposition 4.2 ([8]).

For any eA~(ρ,τ)e\in\tilde{A}(\rho,\tau),

limtΨt(e)=Ψ(e).\exists\lim_{t\to\infty}\Psi_{t}(e)=\Psi_{\infty}(e).

5 Unitary equivalence

Let (G,ρ,τ)(G,\rho,\tau) be a rotation system whose underlying abstract graph is G=(X,A)G=(X,A). Let us deform the rotation system as follows which is corresponding to Operation iii.

Operation (\star)
Choose a one vertex from GG, say xXx\in X. Then let us reverse its rotation and switch all the edge types of the incident edges of xx. Such a rotation system is denoted by (G,ρ(x),τ(x))(G,\rho^{(x)},\tau^{(x)}).

By Lemma 4.1, the resulting embeddings on a closed surface of the rotation systems of the double covering graph, (Gτ,ρρ1,id)(G^{\tau},\rho\oplus\rho^{-1},\mathrm{id}) and (Gτ(x),ρ(x)ρ(x)1,id)(G^{\tau^{(x)}},\rho^{(x)}\oplus{\rho^{(x)}}^{-1},\mathrm{id}), are isomorphic to each other. Under these isomorphic two embeddings, we take the blow up and choose the boundary island arcs δA\delta A from G(ρ,τ)G(\rho,\tau) and its isomorphic boundary δA\delta A^{\prime} from G(ρ(x),τ(x))G(\rho^{(x)},\tau^{(x)}). The resulting blow up graphs are denoted by G~(ρ,τ)\tilde{G}(\rho,\tau) and G~(ρ(x),τ(x))\tilde{G}(\rho^{(x)},\tau^{(x)}), respectively. Note that they are still isomorphic to each other: G~(ρ,τ)G~(ρ(x),τ(x))\tilde{G}(\rho,\tau)\cong\tilde{G}(\rho^{(x)},\tau^{(x)}) under the following bijection ϕ\phi for the incident arcs of islands xx and xx^{\prime}:

Bijection ϕ:G~(ρ,τ)G~(ρ,τ)\phi:\tilde{G}(\rho,\tau)\to\tilde{G}(\rho^{\prime},\tau^{\prime}) satisfies the following:

  1. (1)

    island :
    eis=(e,ρ~(e))Ais(x,k)A(ρ,τ)e_{is}=(e,\tilde{\rho}(e))\in A_{is}(x,k)\subset A(\rho,\tau)
    with o(e)=(y,k)o(e)=(y,k^{\prime}) and o(ρ~(e))=(z,k′′)o(\tilde{\rho}(e))=(z,k^{\prime\prime}) in GτG^{\tau}
    \leftrightarrow
    ϕ(eis)=(e,ρ~(e))Ais(x,k+1)A(ρ,τ)\phi(e_{is})=(e^{\prime},\tilde{\rho}(e^{\prime}))\in A_{is}(x,k+1)\subset A(\rho,\tau)
    with o(e)=(z,k′′+1)o(e^{\prime})=(z,k^{\prime\prime}+1) and t(ρ~(e))=(y,k+1)t(\tilde{\rho}(e^{\prime}))=(y,k^{\prime}+1) in GτG^{\tau}.

  2. (2)

    bridge :
    eAbrA(ρ,τ)e\in A_{br}\subset A(\rho,\tau)
    with o(e)=(x,)o(e)=(x,\ell) and t(e)=(y,m)t(e)=(y,m) in GτG^{\tau}
    \leftrightarrow
    ϕ(e)AbrA(ρ,τ)\phi(e)\in A^{\prime}_{br}\subset A(\rho^{\prime},\tau^{\prime})
    with o(e)=(x,+1)o(e)=(x,\ell+1) and t(e)=(y,m+1)t(e)=(y,m+1) in GτG^{\tau}

The other arcs are left nothing as it is.

In this section, we show that the time evolution operators W~\tilde{W} on G~(ρ,τ)\tilde{G}(\rho,\tau) and W~(x)\tilde{W}^{(x)} on G~(ρ(x),τ(x))\tilde{G}(\rho^{(x)},\tau^{(x)}) are unitarily equivalent. Indeed we have the following proposition.

Proposition 5.1.

Let the rotation system (G,ρ,τ)(G,\rho,\tau) and (G,ρ,τ)(G,\rho^{\prime},\tau^{\prime}) are isomorphic to each other, and let the boundaries of the induced blow-up graphs are also isomorphic to each other. Then the time evolution operators WW on G~(ρ,τ)\tilde{G}(\rho,\tau) and WW^{\prime} on G~(ρ,τ)\tilde{G}(\rho^{\prime},\tau^{\prime}) are unitarily equivalent, that is, there is an unitary map 𝒰\mathcal{U} such that

W=𝒰W𝒰.W^{\prime}=\mathcal{U}^{*}\;W\;\mathcal{U}.
Proof.

It is sufficient to consider the case for ρ=ρx\rho^{\prime}=\rho^{x} for arbitrary xXx\in X. By Lemma 4.1, “the reversing the rotation of xx” and “the reversing all the incident edges of xx” are reflected in the corresponding time evolution operator on G~(ρ,τ)\tilde{G}(\rho,\tau):

  1. (1)

    Switching the labels of all the arcs associated to (x,0)(x,0) and (x,1)(x,1) following the bijection ϕ\phi (\leftrightarrow Reversing the rotation of xx);

  2. (2)

    Changing the twist τ\tau to τ\tau^{\prime} by

    τ(e)={τ(e)+1e is incident to (x,0) or (x,1) in Gτ,τ(e): otherwise.\tau^{\prime}(e)=\begin{cases}\tau(e)+1&\text{: $e$ is incident to $(x,0)$ or $(x,1)$ in $G^{\tau}$,}\\ \tau(e)&\text{: otherwise.}\end{cases}

    (\leftrightarrow Reversing all the incident edges of xx)

Obviously, the operation (1) has the corresponding unitary map. This map is denoted by 𝒰ϕ\mathcal{U}_{\phi}, that is,

(𝒰ϕψ)(e)=ψ(ϕ1(e)).(\mathcal{U}_{\phi}\psi)(e)=\psi(\phi^{-1}(e)).

Then in the rest of the discussion, let us find the corresponding unitary map of the operation (2). To this end, for the blow-up graph G~(ρ,τ)\tilde{G}(\rho,\tau), set

𝒜:={θ:A(ρ,τ)|θ(e)=0 for all eAis,θ(e¯)=θ(e) for all eAbr}.\mathcal{A}:=\{\theta:A(\rho,\tau)\to\mathbb{C}\;|\;\theta(e)=0\text{ for all $e\in A_{is}$},\;\theta(\bar{e})=-\theta(e)\text{ for all $e\in A_{br}$}\}.

Note that if we set

θ(e)={πeAbr with τ(e)=10: otherwise\theta(e)=\begin{cases}\pi&\text{: $e\in A_{br}$ with $\tau(e)=1$}\\ 0&\text{: otherwise}\end{cases} (5.18)

then τ(e)=e𝐢θ(e)\tau(e)=e^{\boldsymbol{\rm i}\;\theta(e)} for any eAbre\in A_{br}, where 𝐢=1\boldsymbol{\rm i}=\sqrt{-1}. On the other hand if we set

θ(e)={πeAbr with τ(e)=10: otherwise\theta(e)=\begin{cases}\pi&\text{: $e\in A_{br}$ with $\tau^{\prime}(e)=1$}\\ 0&\text{: otherwise}\end{cases} (5.19)

then τ(e)=e𝐢θ(e)\tau^{\prime}(e)=e^{\boldsymbol{\rm i}\;\theta(e)} for any eAbre\in A_{br}. For θ𝒜\theta\in\mathcal{A}, and arbitrary path in G(ρ,τ)G(\rho,\tau), p=(a1,a2,,ak)p=(a_{1},a_{2},\dots,a_{k}) with t(aj)=o(aj+1)t(a_{j})=o(a_{j+1}) (j=1,,k1)(j=1,\dots,k-1), we define

pθ:=j=1kθ(ej).\int_{p}\theta:=\sum_{j=1}^{k}\theta(e_{j}).

A path p=(a1,,ar)p=(a_{1},\dots,a_{r}) with a1Abra_{1}\in A_{br}, arAisa_{r}\in A_{is} in G(ρ,τ)G(\rho,\tau) may be represented by

p=(e1,ξ1,ρ~1(ξ1),,ρ~k1(ξ1),e2,ξ2,ρ~1(ξ2),,ρ~k2(ξ2),e3,ξ3,ρ~1(ξ3),,ρ~k3(ξ3), ,er,ξr,ρ~1(ξr),,ρ~kr(ξr)),p=(\;e_{1},\xi_{1},\tilde{\rho}^{1}(\xi_{1}),\dots,\tilde{\rho}^{k_{1}}(\xi_{1}),\;e_{2},\;\xi_{2},\tilde{\rho}^{1}(\xi_{2}),\dots,\tilde{\rho}^{k_{2}}(\xi_{2}),\;e_{3},\;\xi_{3},\tilde{\rho}^{1}(\xi_{3}),\dots,\tilde{\rho}^{k_{3}}(\xi_{3}),\dots{\\ }\dots,e_{r},\xi_{r},\tilde{\rho}^{1}(\xi_{r}),\dots,\tilde{\rho}^{k_{r}}(\xi_{r})\;),

where ξj,ρ~s(ξj)Ais\xi_{j},\tilde{\rho}^{s}(\xi_{j})\in A_{is} and ejAbre_{j}\in A_{br}. Here if there is a back-tracking in pp, a corresponding subsequence of a walk on an island disappears. The inverse path of pp is defined by

p¯=(ρ~kr+1(ξr),ρ~kr+2(ξr),,ρ~dr1(ξr),e¯r, ,ρ~k3+1(ξ3),ρ~k3+2(ξ3),,ρ~d31(ξ3),e¯2,ρ~k1+1(ξ1),ρ~k1+2(ξ1),,ρ~d11(ξ1),e¯1),\bar{p}=(\;\tilde{\rho}^{k_{r}+1}(\xi_{r}),\tilde{\rho}^{k_{r}+2}(\xi_{r}),\dots,\tilde{\rho}^{d_{r}-1}(\xi_{r}),\;\bar{e}_{r},\dots{\\ }\dots,\tilde{\rho}^{k_{3}+1}(\xi_{3}),\tilde{\rho}^{k_{3}+2}(\xi_{3}),\dots,\tilde{\rho}^{d_{3}-1}(\xi_{3}),\;\bar{e}_{2},\tilde{\rho}^{k_{1}+1}(\xi_{1}),\tilde{\rho}^{k_{1}+2}(\xi_{1}),\dots,\tilde{\rho}^{d_{1}-1}(\xi_{1}),\;\bar{e}_{1}\;),

where ρ~:=ρρ1\tilde{\rho}:=\rho\oplus\rho^{-1}. The inverse path of pp for the other cases and pp with back-trackings case are also defined in the same way.

Lemma 5.1.

Set θ1,θ2𝒜\theta_{1},\theta_{2}\in\mathcal{A}. If cθ1=cθ2\int_{c}\theta_{1}=\int_{c}\theta_{2} for any close path cc in G(ρ,τ)G(\rho,\tau), then p(θ2θ1)=p(θ2θ1)\int_{p}(\theta_{2}-\theta_{1})=\int_{p^{\prime}}(\theta_{2}-\theta_{1}) for any path pp and pp^{\prime} with o(p)=o(p)o(p)=o(p^{\prime}) and t(p)=t(p)t(p)=t(p^{\prime}).

Proof.

It holds that p¯θ=pθ\int_{\bar{p}}\theta=-\int_{p}\theta for any path pp by the definition. Note that pp¯p\cup\bar{p^{\prime}} is a closed cycle. Then we have

0=pp¯(θ2θ1)=p(θ2θ1)+p¯(θ2θ1)=p(θ2θ1)p(θ2θ1),0=\int_{p\cup\bar{p}}(\theta_{2}-\theta_{1})=\int_{p}(\theta_{2}-\theta_{1})+\int_{\bar{p^{\prime}}}(\theta_{2}-\theta_{1})=\int_{p}(\theta_{2}-\theta_{1})-\int_{p^{\prime}}(\theta_{2}-\theta_{1}),

which is the conclusion. ∎

Lemma 5.1 tells us that the value p(θ2θ1)\int_{p}(\theta_{2}-\theta_{1}) is independent of routes from o(p)o(p) to t(p)t(p). Let us fix a vertex xx_{*}. Since p(θ2θ1)\int_{p}(\theta_{2}-\theta_{1}) for a path starting from xx_{*} is determined by t(p)=xt(p)=x, we set such a value by Δ(x)\Delta(x), which is well-defined.

Let ψt\psi_{t} (resp. ψt\psi_{t}^{\prime}) be the tt-th iteration of W~\tilde{W} (resp. W~\tilde{W^{\prime}}) with the twist τ\tau (resp. τ\tau^{\prime}) such that ψt+1=W~ψt\psi_{t+1}=\tilde{W}\psi_{t} (resp. ψt+1=W~ψt\psi^{\prime}_{t+1}=\tilde{W}^{\prime}\psi_{t}^{\prime}). By Remark 4.3, we have

[ψt+1(ξ)ψt+1(ϵ)]=[e𝐢θ1(ξ)00e𝐢θ1(ϵ)]C[ψt(ξ)ψt(ϵ¯)],\begin{bmatrix}\psi_{t+1}(\xi)\\ \psi_{t+1}(\epsilon)\end{bmatrix}=\begin{bmatrix}e^{\boldsymbol{\rm i}\;\theta_{1}(\xi)}&0\\ 0&e^{\boldsymbol{\rm i}\;\theta_{1}(\epsilon)}\end{bmatrix}\;C\begin{bmatrix}\psi_{t}(\xi^{\flat})\\ \psi_{t}(\bar{\epsilon})\end{bmatrix}, (5.20)

for any ϵAbr\epsilon\in A_{br}, where ξ=is(ϵ)\xi^{\flat}=\mathrm{is}(\epsilon) and ξ=is(ϵ)\xi=\mathrm{is}^{\sharp}(\epsilon). Here θ1\theta_{1} is defined in (5.18). Now let us set θ2\theta_{2} by (5.19). The operation ()(\star) keeps the parity of any cycles passing through the vertex xx, because τ(e)+τ(e)\tau(e)+\tau(e^{\prime}) with t(e)=x=o(e)t(e)=x=o(e) in GG is changed to

τ(e)+τ(e)=τ(e)+1+τ(e)+1=τ(e)+τ(e),\tau^{\prime}(e)+\tau^{\prime}(e^{\prime})=\tau(e)+1+\tau(e^{\prime})+1=\tau(e)+\tau(e^{\prime}),

which implies cθ1=cθ2\int_{c}\theta_{1}=\int_{c}\theta_{2} for any cycle cc in G~\tilde{G}. Let us consider a path pp with o(p)=xo(p)=x_{*} and t(p)=o(ξ)=o(ϵ)t(p)=o(\xi)=o(\epsilon) in G~\tilde{G}. Thus we have

Δ(t(ξ))\displaystyle\Delta(t(\xi)) =p+ξ(θ2θ1)=p(θ2θ1)+θ2(ξ)θ1(ξ)\displaystyle=\int_{p+\xi}(\theta_{2}-\theta_{1})=\int_{p}(\theta_{2}-\theta_{1})+\theta_{2}(\xi)-\theta_{1}(\xi)
=Δ(o(ξ))+θ2(ξ)θ1(ξ),\displaystyle=\Delta(o(\xi))+\theta_{2}(\xi)-\theta_{1}(\xi),
Δ(t(e))\displaystyle\Delta(t(e)) =p+e(θ2θ1)=p(θ2θ1)+θ2(e)θ1(e)\displaystyle=\int_{p+e}(\theta_{2}-\theta_{1})=\int_{p}(\theta_{2}-\theta_{1})+\theta_{2}(e)-\theta_{1}(e)
=Δ(o(ϵ))+θ2(e)θ1(e),\displaystyle=\Delta(o(\epsilon))+\theta_{2}(e)-\theta_{1}(e),

Here Δ(x)\Delta(x) with a fixed vertex xX~x_{*}\in\tilde{X} is well-defined by Lemma 5.1. Inserting them into (5.20), we have

[e𝐢Δ(t(ξ))ψt+1(ξ)e𝐢Δ(t(ϵ))ψt+1(ϵ)]\displaystyle\begin{bmatrix}e^{\boldsymbol{\rm i}\;\Delta(\;t(\xi)\;)}\;\psi_{t+1}(\xi)\\ e^{\boldsymbol{\rm i}\;\Delta(\;t(\epsilon)\;)}\;\psi_{t+1}(\epsilon)\end{bmatrix} =[e𝐢θ2(ξ)00e𝐢θ2(ϵ)][e𝐢Δ(o(ξ))00e𝐢Δ(o(ϵ))]C[ψt(ξ)ψt(ϵ¯)]\displaystyle=\begin{bmatrix}e^{\boldsymbol{\rm i}\;\theta_{2}(\xi)}&0\\ 0&e^{\boldsymbol{\rm i}\;\theta_{2}(\epsilon)}\end{bmatrix}\begin{bmatrix}e^{\boldsymbol{\rm i}\;\Delta(\;o(\xi)\;)}&0\\ 0&e^{\boldsymbol{\rm i}\;\Delta(\;o(\epsilon)\;)}\end{bmatrix}\;C\;\begin{bmatrix}\psi_{t}(\xi^{\flat})\\ \psi_{t}(\bar{\epsilon})\end{bmatrix}
=[e𝐢θ2(ξ)00e𝐢θ2(ϵ)]C[e𝐢Δ(t(ξ))ψt(ξ)e𝐢Δ(t(ϵ¯))ψt(ϵ¯)]\displaystyle=\begin{bmatrix}e^{\boldsymbol{\rm i}\;\theta_{2}(\xi)}&0\\ 0&e^{\boldsymbol{\rm i}\;\theta_{2}(\epsilon)}\end{bmatrix}\;C\;\begin{bmatrix}e^{\boldsymbol{\rm i}\;\Delta(\;t(\xi^{\flat})\;)}\;\psi_{t}(\xi^{\flat})\\ e^{\boldsymbol{\rm i}\;\Delta(\;t(\bar{\epsilon})\;)}\;\psi_{t}(\bar{\epsilon})\end{bmatrix}

The second equality derives from o(ξ)=o(ϵ))o(\xi)=o(\epsilon)) which gives the commutativity of the second diagonal matrix and CC in first equality. Therefore introducing the unitary map 𝒰Δ\mathcal{U}_{\Delta} by

(𝒰Δψ)(a)=e𝐢Δ(t(a))ψ(a),(\mathcal{U}_{\Delta}\psi)(a)=e^{\boldsymbol{\rm i}\;\Delta(t(a))}\;\psi(a),

we obtain

W=(𝒰Δ𝒰ϕ)W(𝒰Δ𝒰ϕ).W^{\prime}=(\mathcal{U}_{\Delta}\mathcal{U}_{\phi})^{*}\;W\;(\mathcal{U}_{\Delta}\mathcal{U}_{\phi}).

Remark 5.1.

For a vertex (x,j)X(ρ,τ)(x,j)\in X(\rho,\tau), define sheet(x,j):=j2\mathrm{sheet}(x,j):=j\in\mathbb{Z}_{2}. Then the unitary map 𝒰=𝒰Δ𝒰ϕ\mathcal{U}=\mathcal{U}_{\Delta}\;\mathcal{U}_{\phi} is described by

(𝒰ψ)(a)={ψ(ϕ1(a))sheet(t(a))+sheet(x)=sheet(ϕ(t(a)))+sheet(ϕ(x)),ψ(ϕ1(a))sheet(t(a))+sheet(x)sheet(ϕ(t(a)))+sheet(ϕ(x)).(\mathcal{U}\psi)(a)=\begin{cases}\psi(\phi^{-1}(a))&\text{: $\mathrm{sheet}(t(a))+\mathrm{sheet}(x_{*})=\mathrm{sheet}(\phi(t(a)))+\mathrm{sheet}(\phi(x_{*}))$,}\\ -\psi(\phi^{-1}(a))&\text{: $\mathrm{sheet}(t(a))+\mathrm{sheet}(x_{*})\neq\mathrm{sheet}(\phi(t(a)))+\mathrm{sheet}(\phi(x_{*}))$.}\end{cases}

6 Scattering

6.1 The scattering matrix represented by faces

For the blow-up graph G(ρ,τ)G(\rho,\tau), let δX={x1,,xκ}\delta X=\{x_{1},\dots,x_{\kappa}\} and δApr+={e1,,eκ}\delta A_{pr}^{+}=\{e_{1},\dots,e_{\kappa}\} such that t(ej)=xjt(e_{j})=x_{j}. The stationary state is denoted by ΨA~(ρ,τ)\Psi_{\infty}\in\mathbb{C}^{\tilde{A}(\rho,\tau)}. Let us represent the inflow from the outside by 𝜶inδX\boldsymbol{\alpha}_{in}\in\mathbb{C}^{\delta X} such that

𝜶in(x)=Ψ(e) with t(e)=x\boldsymbol{\alpha}_{in}(x)=\Psi_{\infty}(e)\text{ with $t(e)=x$}

for any xδXx\in\delta X, 2\ell\in\mathbb{Z}_{2}, while the outflow to the outside by 𝜷outδX\boldsymbol{\beta}_{out}\in\mathbb{C}^{\delta X} such that

𝜷out(x)=Ψ(e¯) with t(e)=x\boldsymbol{\beta}_{out}(x)=\Psi_{\infty}(\bar{e})\text{ with $t(e)=x$}

for any xδXx\in\delta X. The scattering matrix S:δXδXS:\mathbb{C}^{\delta X}\to\mathbb{C}^{\delta X} is defined by

𝜷out=S𝜶in,\boldsymbol{\beta}_{out}=S\boldsymbol{\alpha}_{in},

which is determined by the rotation system G(ρ,τ)G(\rho,\tau) and δX\delta X, and independent of 𝜶in\boldsymbol{\alpha}_{in}, 𝜷out\boldsymbol{\beta}_{out}. It is shown in [8] that such a matrix SS exists and is unitary. However its explicit expression is up to the individual setting.

We prepare important graph notions which express the scattering matrix. Let f=(e0,,eκ1)Ff=(e_{0},\dots,e_{\kappa-1})\in F be a facial closed walk of the rotation system (Gτ,ρρ1,id)(G^{\tau},\rho\oplus\rho^{-1},\mathrm{id}) with length κ=|f|\kappa=|f|. The extended facial walk of f~\tilde{f} in A(ρ,τ)A(\rho,\tau) is defined by just alternatively inserted corresponding island arcs into each arc of ff; (we use the same notation for the extended facial walk by ff):

f:=(ξ0,e0,,ξκ1,eκ1),f:=(\;\xi_{0},e_{0},\dots,\xi_{\kappa-1},e_{\kappa-1}\;), (6.21)

where ejAbrAe_{j}\in A_{br}\cong A and ξjAis\xi_{j}\in A_{is} with

ξj=is(ej),ej=br(ξj+1)\xi_{j}=\mathrm{is}(e_{j}),\;e_{j}=\mathrm{br}(\xi_{j+1}) (6.22)

for any j{0,1,,κ1}j\in\{0,1,\dots,\kappa-1\} in the modulus of κ\kappa. Here if ξjδAis\xi_{j}\in\delta A_{is}, ξj\xi_{j} represents the quay arcs (ξj+,ξj)(\xi_{j}^{+},\xi_{j}^{-}). Let δFF\delta F\subset F be the set of (extended) facial walk passing through a boundary vertex δX\delta X in G~\tilde{G}. For any facial walk ff, there exists a chiral facial walk ff^{*} defined by going around the opposite direction on the opposite sheet. We introduce the following useful lemma to consider the scattering.

Lemma 6.1.

Assume dd\in\mathbb{R}. Let ψ\psi_{\infty}\in\mathcal{H} be the stationary state. Set ω=det(C)\omega=-\det(C). For each facial walk fδFf\in\delta F of the orientation system (G,ρ,τ)(G,\rho,\tau) represented by a sequence of Ω\Omega,

f:=(ξ0,e0,ξ1,e1,,ξκ1,eκ1),f:=(\;\xi_{0},e_{0},\xi_{1},e_{1},\dots,\xi_{\kappa-1},e_{\kappa-1}\;),

we have

ψ(ξj+1)=(1)τ(ej)ωψ(ξj′′),(j=0,1,,κ1)\psi_{\infty}(\xi^{\prime}_{j+1})=(-1)^{\tau(e_{j})}\omega\;\psi_{\infty}(\xi_{j}^{\prime\prime}),\;(j=0,1,\dots,\kappa-1)

where if ξj+1δAis\xi_{j+1}\in\delta A_{is}, then ξj+1=ξj+1+\xi_{j+1}^{\prime}=\xi_{j+1}^{+}, otherwise ξj+1\xi_{j+1}; if ξjδAis\xi_{j}\in\delta A_{is}, then ξj′′=ξj\xi_{j}^{\prime\prime}=\xi_{j}^{-}, otherwise ξj′′=ξj\xi_{j}^{\prime\prime}=\xi_{j}.

Proof.

For any bridge arcs eAbre\in A_{br}, (of course) the vertices o(e)o(e) and t(e)t(e) are connected by the bridge arcs e1R:=ee^{R}_{1}:=e and e1L:=e¯e^{L}_{1}:=\bar{e}, moreover the island arcs e0R:=is(e)e_{0}^{R}:=\mathrm{is}(e) and e0L:=is(e)e_{0}^{L}:=\mathrm{is}^{\sharp}(e) are connected to o(e)o(e), while the island arcs e2R:=is(e¯)e_{2}^{R}:=\mathrm{is}^{\sharp}(\bar{e}) and e2L:=is(e¯)e_{2}^{L}:=\mathrm{is}(\bar{e}) are connected to o(t)o(t). Then the blow up graph has locally a path structure with 22 vertices and “66” symmetric arcs. Let us see this fact leads a transfer matrix discussed in the spectral analysis on the discrete-time quantum walks on the one-dimensional lattice and gives the conclusion. Let us set ψ:=ψ\psi:=\psi_{\infty}. The point to get the transfer matrix is to align the “subscriptions” for each vector: the local eigenequation is rewritten by

[a1c0][ψ(e0R)ψ(e0L)]=[0b1d][ψ(e1R)ψ(e1L)] and [1a0c][ψ(e2R)ψ(e2L)]=[b0d1][ψ(e1R)ψ(e1L)]\displaystyle\begin{bmatrix}-a&1\\ -c^{\prime}&0\end{bmatrix}\begin{bmatrix}\psi(e_{0}^{R})\\ \psi(e_{0}^{L})\end{bmatrix}=\begin{bmatrix}0&b\\ -1&d^{\prime}\end{bmatrix}\begin{bmatrix}\psi(e_{1}^{R})\\ \psi(e_{1}^{L})\end{bmatrix}\text{ and }\begin{bmatrix}1&-a\\ 0&-c^{\prime}\end{bmatrix}\begin{bmatrix}\psi(e_{2}^{R})\\ \psi(e_{2}^{L})\end{bmatrix}=\begin{bmatrix}b&0\\ d^{\prime}&-1\end{bmatrix}\begin{bmatrix}\psi(e_{1}^{R})\\ \psi(e_{1}^{L})\end{bmatrix} (6.23)

where c=(1)τ(e)cc^{\prime}=(-1)^{\tau(e)}c, d=(1)τ(e)dd^{\prime}=(-1)^{\tau(e)}d. Therefore we have

[ψ(e2R)ψ(e2L)]\displaystyle\begin{bmatrix}\psi(e_{2}^{R})\\ \psi(e_{2}^{L})\end{bmatrix} =[1a0c]1[b0d1][0b1d]1[a1c0][ψ(e0R)ψ(e0L)]\displaystyle=\begin{bmatrix}1&-a\\ 0&-c^{\prime}\end{bmatrix}^{-1}\begin{bmatrix}b&0\\ d^{\prime}&-1\end{bmatrix}\begin{bmatrix}0&b\\ -1&d^{\prime}\end{bmatrix}^{-1}\begin{bmatrix}-a&1\\ -c^{\prime}&0\end{bmatrix}\begin{bmatrix}\psi(e_{0}^{R})\\ \psi(e_{0}^{L})\end{bmatrix}
=1|c|2ω[ω001][1d¯2dd¯d+d¯1d2][ω001][ψ(e0R)ψ(e0L)]\displaystyle=\frac{1}{|c|^{2}\omega^{\prime}}\begin{bmatrix}\omega^{\prime}&0\\ 0&1\end{bmatrix}\begin{bmatrix}1-{\bar{d^{\prime}}}^{2}&d^{\prime}-\bar{d^{\prime}}\\ -d^{\prime}+\bar{d^{\prime}}&1-{d^{\prime}}^{2}\end{bmatrix}\begin{bmatrix}\omega^{\prime}&0\\ 0&1\end{bmatrix}\begin{bmatrix}\psi(e_{0}^{R})\\ \psi(e_{0}^{L})\end{bmatrix}

where ω=(1)τ(e)ω\omega^{\prime}=(-1)^{\tau(e)}\omega. Here the second equality is obtained by the properties of the unitarity matrix CC, for examples, a=ωd¯a=-\omega^{\prime}\bar{d^{\prime}}, b=ωc¯b=\omega^{\prime}\bar{c^{\prime}}, |d|2+|c|2=1|d^{\prime}|^{2}+|c^{\prime}|^{2}=1. Then dd\in\mathbb{R} if and only if ψ(e2R)=ωψ(e0R)\psi(e_{2}^{R})=\omega^{\prime}\psi(e_{0}^{R}) and ψ(e0L)=ωψ(e2L)\psi(e_{0}^{L})=\omega^{\prime}\psi(e_{2}^{L}), which is the desired conclusion. ∎

For a facial walk f=(ξ0,e0,,ξκ1,eκ1)δFf=(\xi_{0},e_{0},\dots,\xi_{\kappa-1},e_{\kappa-1})\in\delta F, set fδX={xk0,,xkq1}f\cap\delta X=\{x_{k_{0}},\dots,x_{k_{q-1}}\} and fδAis={ξk0,,ξkq}f\cap\delta A_{is}=\{\xi_{k_{0}},\dots,\xi_{k_{q}}\} with k0<k1<<kq1k_{0}<k_{1}<\cdots<k_{q-1} in the modulus of qq. Here 1qκ1\leq q\leq\kappa. Note that the qq tails interfere with the facial walk ff. The tail originating the boundary xkix_{k_{i}} is denoted by Taili\mathrm{Tail}_{i}. Put

f(,m):=xkxkmτ and distf(,m):=kkm,\bowtie_{f}(\ell,m):=\int_{x_{k_{\ell}}}^{x_{k_{m}}}\tau\text{ and }\mathrm{dist}_{f}(\ell,m):=k_{\ell}-k_{m},

which are the parity of number of type-11 edge between xmx_{m} and xx_{\ell} , and the distance between the boundary vertices kmk_{m} and kk_{\ell} along the facial walk ff. To simplify the notation, let us put fδX={0,1,,q1}f\cap\delta X=\{0,1,\dots,q-1\}. We set the identity matrix, the weighted cyclic permutation matrix on {0,,q1}\mathbb{C}^{\{0,\dots,q-1\}} by

(Ifh)(j)=h(j),\displaystyle(I_{f}h)(j)=h(j),
(Pf(ω)h)(j)=(1)f(j,j1)ωdistf(j,j1)h(j1)\displaystyle(P_{f}(\omega)h)(j)=(-1)^{\bowtie_{f}(j,j-1)}\omega^{\mathrm{dist}_{f}(j,j-1)}h(j-1) (6.24)

for any h{0,,q1}h\in\mathbb{C}^{\{0,\dots,q-1\}} and j{0,,q1}j\in\{0,\dots,q-1\} in the modulus of |f||f|, respectively. Now we are ready to give the theorem for the scattering:

Theorem 6.1.

Assume dd\in\mathbb{R} and set ω=detC\omega=-\det C. The scattering matrix is decomposed into the following |F||F| unitary matrices as follows:

S=fFSf,S=\bigoplus_{f\in F}S_{f},

where

Sf=bcPf(ω)(IfaPf(ω))1+dIf.S_{f}=bcP_{f}(\omega)\;(I_{f}-aP_{f}(\omega))^{-1}+dI_{f}.

Here the operators induced by each external facial closed walk IfI_{f} and Pf(ω)P_{f}(\omega) are defined in (6.1).

Remark 6.1.

Since SfS_{f} is the q×qq\times q unitary matrix and Pfq(ω)=IfP_{f}^{q}(\omega)=I_{f}, where q=|fδX|q=|f\cap\delta X|, then SfS_{f} is rewritten by

Sf=bc1aqω|f|Pf(ω)(If+(aPf(ω))+(aPf(ω))2++(aPf(ω))q1).S_{f}=\frac{bc}{1-a^{q}\omega^{|f|}}P_{f}(\omega)\left(I_{f}+(aP_{f}(\omega))+(aP_{f}(\omega))^{2}+\cdots+(aP_{f}(\omega))^{q-1}\right).

From this expression, if fδX={k0,,kq1}f\cap\delta X=\{k_{0},\dots,k_{q-1}\}, we have

(Sf)kj,ki={bcaji1ωdistf(kj,ki)1a|fδX|×(1)f(kj,ki)ij,bca|fδX|1ω|f|1a|fδX|+di=j.(S_{f})_{k_{j},k_{i}}=\begin{cases}bc\frac{a^{j-i-1}\omega^{{\rm dist}_{f}(k_{j},k_{i})}}{1-a^{|f\cap\delta X|}}\times(-1)^{\bowtie_{f}(k_{j},k_{i})}&\text{: $i\neq j$,}\\ \\ bc\frac{a^{|f\cap\delta X|-1}\omega^{|f|}}{1-a^{|f\cap\delta X|}}+d&\text{: $i=j$.}\end{cases}

Here “ji1j-i-1” corresponds to the number of boundary vertices where the facial walk ff passes through from kik_{i} to kjk_{j}.

Proof.

Pick up a facial walk f=(ξ0,e0,,ξκ1,eκ1)δFf=(\xi_{0},e_{0},\dots,\xi_{\kappa-1},e_{\kappa-1})\in\delta F, and set fδX={xj0,,xjq1}f\cap\delta X=\{x_{j_{0}},\dots,x_{j_{q-1}}\} and fδAis={ξk0,,ξkq}f\cap\delta A_{is}=\{\xi_{k_{0}},\dots,\xi_{k_{q}}\}. Here 1qκ1\leq q\leq\kappa. Note that the qq tails interfere with the facial walk ff. The tail originating the boundary xkix_{k_{i}} is denoted by Taili\mathrm{Tail}_{i}. Let us give the inflow 11 from the Taili\mathrm{Tail}_{i}, and consider the outflow to the tail Tailj\mathrm{Tail}_{j} (i,j{0,1,,κ1}i,j\in\{0,1,\dots,\kappa-1\}). Let us see that the (j,i)(j,i) element of the scattering matrix SS can be calculated according to the number of “nights” that a quantum walk stays at the face ff with the boundary vertices xkix_{k_{i}} and xkjx_{k_{j}} from the time the quantum walk enters at entrance ii to the time the quantum walk leaves at exit jj.

Consider the case for i=0i=0. Let tjinδApr+t^{in}_{j}\in\delta A_{pr}^{+} and tjoutδAprt^{out}_{j}\in\delta A_{pr}^{-} be the arcs of Tailj\mathrm{Tail}_{j} originating the vertex xkjδXx_{k_{j}}\in\delta X. Let “the day trip walk” from Tail0\mathrm{Tail}_{0} to Tailj\mathrm{Tail}_{j} be defined by

(t0in,𝒘0,tjout),(t_{0}^{in},\boldsymbol{w}_{0},t_{j}^{out}),

where

𝒘0:=(ξk0+,ek0,,ξk1,ξk1+,ek1,,ξkj1,ξkj1+,ekj1,,ξkj),\boldsymbol{w}_{0}:=(\xi_{k_{0}}^{+},e_{k_{0}},\dots,\xi_{k_{1}}^{-},\xi_{k_{1}}^{+},e_{k_{1}},\dots,\xi_{k_{j-1}}^{-},\xi_{k_{j-1}}^{+},e_{k_{j-1}},\dots,\xi_{k_{j}}^{-}),

and the “rr-night walk” from Tail0\mathrm{Tail}_{0} to Tailj\mathrm{Tail}_{j} be defined by

(t0in,𝒘0,ξkj+,ekj,,ξkj+1,ξkj+1+,ekj+1,,ξkj1,ξkj1+,ekj1,,ξkj,r times (nights)tjout)(t_{0}^{in},\boldsymbol{w}_{0},\overbrace{\xi_{k_{j}}^{+},e_{k_{j}},\dots,\xi_{k_{j+1}}^{-},\xi_{k_{j+1}}^{+},e_{k_{j+1}},\dots,\xi_{k_{j-1}}^{-},\xi_{k_{j-1}}^{+},e_{k_{j-1}},\dots,\xi_{k_{j}}^{-},}^{\text{$r$ times (nights)}}t_{j}^{out})

By Lemma 6.1, the weight associated with the moving along the facial walk from ξAis\xi_{\ell}\in A_{is} to ξ+1Ais\xi_{\ell+1}\in A_{is} must be ω×(1)τ(e)\omega\times(-1)^{\tau(e_{\ell})} in the stationary state. By the local time evolution denoted by the 2×22\times 2-unitary matrix

C=[abcd],C=\begin{bmatrix}a&b\\ c&d\end{bmatrix},

the weights associated with moving from the tail t0inδApr+t_{0}^{in}\in\delta A_{pr}^{+} to the quay ξk0+Ais\xi_{k_{0}}^{+}\in A_{is}, and from the quay ξkjAis\xi_{k_{j}}^{-}\in A_{is} to the tail tjoutδAprt_{j}^{out}\in\delta A_{pr}^{-} are cc and bb, respectively; the weight associated with moving from ξksAis\xi_{k_{s}}^{-}\in A_{is} to ξks+Ais\xi_{k_{s}}^{+}\in A_{is} is aa. Remark that the weight of the closed path starting from ξk0+\xi_{{k_{0}}}^{+} and returning back to ξk0+\xi_{k_{0}}^{+} along the boundary face ff is aqωκa^{q}\omega^{\kappa} since (1)fτ=1(-1)^{\int_{f}\tau}=1. Then set 𝒔(r)(,0){0,1,,q1}\boldsymbol{s}^{(r)}(\cdot,0)\in\mathbb{C}^{\{0,1,\dots,q-1\}} (r=0,1,r=0,1,\dots) as the weight of rr-night walk in the stationary state by 𝒔(r)(j,0)=(aqωκ)r×𝒔(0)(j,0)\boldsymbol{s}^{(r)}(j,0)=(a^{q}\omega^{\kappa})^{r}\times\boldsymbol{s}^{(0)}(j,0), (r=1,,κ1)(r=1,\dots,\kappa-1) with

𝒔(0)(j,0)={bcaj1ωkjk0(1)xk0xkjτj0bcaq1ωκj=0\boldsymbol{s}^{(0)}(j,0)=\begin{cases}b\;c\;a^{j-1}\;\omega^{k_{j}-k_{0}}\;(-1)^{\int_{x_{k_{0}}}^{x_{k_{j}}}\tau}&\text{: $j\neq 0$}\\ b\;c\;a^{q-1}\;\omega^{\kappa}\;&\text{: $j=0$}\end{cases}

The outflow from jj is obtained by the superposition of 𝒔(r)(j,0)\boldsymbol{s}^{(r)}(j,0)’s (r=0,1,2,)(r=0,1,2,\dots) because of the constant inflow 11 at every time step from the tail tk0t_{k_{0}}. Then we have

(S)kj,k0\displaystyle(S)_{k_{j},k_{0}} =r=0𝒔(r)(j,0)={bcaj1ωkjk01aqωκ×(1)xk0xkjτj0bcaq1ωκ1aqωκ+dj=0\displaystyle=\sum_{r=0}^{\infty}\boldsymbol{s}^{(r)}(j,0)=\begin{cases}bc\frac{a^{j-1}\omega^{k_{j}-k_{0}}}{1-a^{q}\omega^{\kappa}}\times(-1)^{\int_{x_{k_{0}}}^{x_{k_{j}}}\tau}&\text{: $j\neq 0$}\\ \\ bc\frac{a^{q-1}\omega^{\kappa}}{1-a^{q}\omega^{\kappa}}+d&\text{: $j=0$}\end{cases}
=bc1aqωκ(Pf(ω)+aPf(ω)2++aq2Pfq1(ω)+aq1ωκIf)j,0dδj,0\displaystyle=\frac{bc}{1-a^{q}\omega^{\kappa}}\left(P_{f}(\omega)+aP_{f}(\omega)^{2}+\cdots+a^{q-2}P_{f}^{q-1}(\omega)+a^{q-1}\omega^{\kappa}I_{f}\right)_{j,0}-d\delta_{j,0}
=(Sf)kj,k0.\displaystyle=(S_{f})_{k_{j},k_{0}}.

From the symmetricity of the rotation, we have

(S)kj,ki={bcaji1ωkjki1aqωκ×(1)xkixkjτijbcaq1ωκ1aqωκ+di=j,(S)_{k_{j},k_{i}}=\begin{cases}bc\frac{a^{j-i-1}\omega^{k_{j}-k_{i}}}{1-a^{q}\omega^{\kappa}}\times(-1)^{\int_{x_{k_{i}}}^{x_{k_{j}}}\tau}&\text{: $i\neq j$}\\ \\ bc\frac{a^{q-1}\omega^{\kappa}}{1-a^{q}\omega^{\kappa}}+d&\text{: $i=j$,}\end{cases} (6.25)

which leads the conclusion. ∎

Let (G,ρ1,τ1)(G,\rho_{1},\tau_{1}) and (G,ρ2,τ2)(G,\rho_{2},\tau_{2}) be the rotation systems which are isomorphic to each other. Let S1S_{1} and S2S_{2} be the scattering matrices of those resulting embbedings, respectively.

Proposition 6.1.

Let S1S_{1} and S2S_{2} be the above. Then we have

S2=DS1D,S_{2}=D^{*}S_{1}D,

where D=fFDfD=\oplus_{f\in F}D_{f} is the diagonal matrix such that

Df=diag[eiΔ(t(e0)),eiΔ(t(e1)),,eiΔ(t(eκ1))]D_{f}=\mathrm{diag}[e^{-i\Delta(t(e_{0}))},e^{-i\Delta(t(e_{1}))},\dots,e^{-i\Delta(t(e_{\kappa-1}))}]

for a face f=(e0,e1,,eκ1)f=(e_{0},e_{1},\dots,e_{\kappa-1}).

Proof.

Proposition 5.1 immediately leads to the conclusion. ∎

Let us consider the case for the following special assignment of the tail, which can be constructed independently of the embedding.

The hedgehog tail assignment: We call the hedgehog tail assignment if

δAis=Ais,\delta A_{is}=A_{is},

which is the setting of the tails so that a tail is inserted between each vertex in the islands.

In the hedgehog tail assignment, δF=F\delta F=F holds. Pick up a facial walk f=(ξ0,e0,,ξκ1,eκ1)δF=Ff=(\xi_{0},e_{0},\dots,\xi_{\kappa-1},e_{\kappa-1})\in\delta F=F. Since distf=1\mathrm{dist}_{f}=1, |δXf|=|f||\delta X\cap f|=|f|, Pf(ω)=ωPfP_{f}(\omega)=\omega P_{f}, where Pf:=Pf(1)P_{f}:=P_{f}(1) such that

(Pfh)(j)=(1)τ(ej)h(j)(P_{f}h)(j)=(-1)^{\tau(e_{j})}h(j) (6.26)

for any j=0,1,,|f|1j=0,1,\dots,|f|-1 and h{x0,,x|f|1}h\in\mathbb{C}^{\{x_{0},\dots,x_{|f|-1}\}}. Then Theorem 6.1 implies Theorem 1.1 which is the special case for the hedgehog.

6.2 Detection of the orientability

Let us estimate whether the underlying closed surface is orientable or not by observing the outflow of the internal graph to an inflow. The following theorem may be useful for such an estimation.

Theorem 6.2.

Under Assumption 1 with a>0a>0, the underlying surface is orientable if and only if for any x,yXτx,y\in X^{\tau} with xyx\neq y, the signature of the element of scattering matrix

sgn((S)ex,ey)\mathrm{sgn}((S)_{e_{x},e_{y}})

is invariant for any (ex,ey){e,eAτ;|(S)e,e0,t(e)=x,t(e)=y}(e_{x},e_{y})\in\{e,e^{\prime}\in A^{\tau};\;|\;(S)_{e,e^{\prime}}\neq 0,\;t(e)=x,t(e^{\prime})=y\}.

Proof.

Since a>0a>0,

(S)e,e/|Se,e|={(1)t(e)t(e)τe and e are included in the same face, 0: otherwise.(S)_{e,e^{\prime}}/|S_{e,e^{\prime}}|=\begin{cases}(-1)^{\int_{t(e)}^{t(e^{\prime})}\tau}&\text{: $e$ and $e^{\prime}$ are included in the same face, }\\ 0&\text{: otherwise.}\end{cases}

Here xyτ:=e′′pτ(e′′)\int_{x}^{y}\tau:=\sum_{e^{\prime\prime}\in p}\tau(e^{\prime\prime}) with a path pp satisfying o(p)=xo(p)=x and t(p)=yt(p)=y. Note that Lemma 4.1 implies that xyτ\int_{x}^{y}\tau is independent of the choice of path if and only if the underlying surface is orientable. ∎

This means that once a vertex is found where there exist different signatures among the outflows from it, then the underlying closed surface must be unorientable.

7 Comfortability (Proof of main theorem)

In this section, we subject to Assumption 1. Note that because of the hedgehog boundary condition, the set of tails has a bijection map to the set of bridges. Then for a bridge arc eAbre\in A_{br}, put t2(e):=t(is(e¯))δXt^{2}(e):=t(\mathrm{is}^{\sharp}(\bar{e}))\in\delta X. Let us set ηAbr\eta\in\mathbb{C}^{A_{br}} by

η(e):=1bcω(Q𝜶in)(t2(e)).\eta(e):=\frac{1}{bc\omega}(Q\boldsymbol{\alpha}_{in})(t^{2}(e)).

Here the matrix Q:=SdIQ:=S-dI represents the scattering after a quantum walker penetrates the interior at least once. By using the notation η\eta of such a scattering, the stationary state ψ\psi_{\infty} is expressed as follows.

Lemma 7.1.

Under Assumption 1, for any bridge arc eAbre\in A_{br},

ψ(e)=(1)τ(e)ω(η(e)+dη(e¯)).\psi_{\infty}(e)=(-1)^{\tau(e)}\;\omega\;(\;\eta(e)+d\eta(\bar{e})\;).

For a facial walk ff,

f:=(ξ0+,e0,ξ1,ξ1+,e1,ξ2,ξ2+,e2,,ξκ1,ξκ1,eκ1,ξ0),f:=(\;\xi_{0}^{+},e_{0},\;\;\xi_{1}^{-},\xi_{1}^{+},e_{1},\;\;\xi_{2}^{-},\xi_{2}^{+},e_{2},\;\dots\\ \dots,\;\xi_{\kappa-1}^{-},\xi_{\kappa-1}^{-},e_{\kappa-1},\;\;\xi_{0}^{-}\;),

the stationary state at the island arc is

ψ(ξm+)=(1)τ(em)bη(em),ψ(ξm+1)=ωbη(em),\psi_{\infty}(\xi_{m}^{+})=(-1)^{\tau(e_{m})}b\eta(e_{m}),\;\psi_{\infty}(\xi_{m+1}^{-})=\omega b\eta(e_{m}),

for any mκm\in\mathbb{Z}_{\kappa}.

Proof.

Let the outflow from the quay (ξm+1,ξm+1+)(\xi_{m+1}^{-},\xi_{m+1}^{+}) be βm+1\beta_{m+1}. Then βm+1=dαm+1+cψ(ξm+1)\beta_{m+1}=d\alpha_{m+1}+c\psi_{\infty}(\xi_{m+1}^{-}), which implies

ψ(ξm+1)=(1/c)(SfdIf)𝜶f(m+1)=ωbη(em).\psi_{\infty}(\xi_{m+1}^{-})=(1/c)(S_{f}-dI_{f})\boldsymbol{\alpha}_{f}(m+1)=\omega b\eta(e_{m}). (7.27)

by Theorem 1.1. Lemma 6.1 leads

ψ(ξm+)=(1)τ(em)ω1ψ(ξm+1)=(1)τ(em)bη(em)\psi_{\infty}(\xi_{m}^{+})=(-1)^{\tau(e_{m})}\omega^{-1}\psi_{\infty}(\xi_{m+1}^{-})=(-1)^{\tau(e_{m})}b\eta(e_{m}) (7.28)

By (6.23), under the assumption of dd\in\mathbb{R}, we have

[ψ(e1R)ψ(e1L)]\displaystyle\begin{bmatrix}\psi(e_{1}^{R})\\ \psi(e_{1}^{L})\end{bmatrix} =1b[db10][a1c0][ψ(e0R)ψ(e0L)]\displaystyle=\frac{1}{b}\begin{bmatrix}d^{\prime}&-b\\ 1&0\end{bmatrix}\begin{bmatrix}-a&1\\ -c^{\prime}&0\end{bmatrix}\begin{bmatrix}\psi(e_{0}^{R})\\ \psi(e_{0}^{L})\end{bmatrix}
=1b[1dd1][ω001][ψ(e0R)ψ(e0L),]\displaystyle=\frac{1}{b}\begin{bmatrix}1&d^{\prime}\\ d^{\prime}&1\end{bmatrix}\begin{bmatrix}\omega^{\prime}&0\\ 0&1\end{bmatrix}\begin{bmatrix}\psi(e_{0}^{R})\\ \psi(e_{0}^{L}),\end{bmatrix}

which implies

ψ(e1R)=ωb(ψ(e0R)+dψ(e2L)),ψ(e1L)=ωb(ψ(e0R)+dψ(e2L))\psi(e_{1}^{R})=\frac{\omega^{\prime}}{b}(\psi(e_{0}^{R})+d^{\prime}\psi(e_{2}^{L})),\;\psi(e_{1}^{L})=\frac{\omega^{\prime}}{b}(\psi(e_{0}^{R})+d^{\prime}\psi(e_{2}^{L})) (7.29)

by Lemma 6.1. Recalling that the arcs e1Re_{1}^{R}, e1Le_{1}^{L} are bridge arcs with e1L¯=e1R=e\bar{e_{1}^{L}}=e_{1}^{R}=e and e0R=is(e)e_{0}^{R}=\mathrm{is}(e), e0L=is(e)e_{0}^{L}=\mathrm{is}^{\sharp}(e), we obtain

ψ(e)=(1)τ(e)ω(η(e)+η(e¯))\psi_{\infty}(e)=(-1)^{\tau(e)}\omega(\eta(e)+\eta(\bar{e}))

by inserting (7.27) and (7.28) into (7.29). ∎

Let σ:AbrAbr\sigma:\mathbb{C}^{A_{br}}\to\mathbb{C}^{A_{br}} be the flip-flop matrix such that

(σψ)(a)=ψ(a¯)(\sigma\psi)(a)=\psi(\bar{a})

for any ψAbr\psi\in A_{br} and eAbre\in A_{br}, and set Q:AbrAbrQ:\mathbb{C}^{A_{br}}\to\mathbb{C}^{A_{br}} by

Q=SdI.Q=S-dI.

Here the scattering matrix SS is regarded as the operator on Abr\mathbb{C}^{A_{br}} with the bijection map AbrδXA_{br}\to\delta X by x=t(is(e¯))δXx=t(\mathrm{is}^{\sharp}(\bar{e}))\in\delta X for any eAbre\in A_{br}(: note that the boundary is the hedgehog). By using the expression of the stationary state ψ\psi_{\infty} in Lemma 7.1, the comfortability \mathcal{E} is described as follows.

Lemma 7.2.

Under Assumption 1, the comfortability with the inflow 𝛂in\boldsymbol{\alpha}_{in} is described by

=1|c|2Q𝜶2+12|bc|2(σ+dI)Q𝜶2.\mathcal{E}=\frac{1}{|c|^{2}}||Q\boldsymbol{\alpha}||^{2}+\frac{1}{2|bc|^{2}}||(\sigma+dI)Q\boldsymbol{\alpha}||^{2}.
Proof.

The comfortability \mathcal{E} is decomposed into

=island+bridge.\mathcal{E}=\mathcal{E}^{island}+\mathcal{E}^{bridge}.

Here island=(1/2)eAis|ψ(e)|2\mathcal{E}_{island}=(1/2)\sum_{e\in A_{is}}|\psi_{\infty}(e)|^{2} and bridge=(1/2)eAbr|ψ(e)|2\mathcal{E}^{bridge}=(1/2)\sum_{e\in A_{br}}|\psi_{\infty}(e)|^{2}.

  1. (1)

    Island: Since every island arc belongs to unique facial walk, island\mathcal{E}^{island} can be further decomposed into each facial walk by

    island=fFfisland,\mathcal{E}^{island}=\sum_{f\in F}\mathcal{E}_{f}^{island},

    where fisland=efAis|ψ(e)|2\mathcal{E}_{f}^{island}=\sum_{e\in f\cap A_{is}}|\psi_{\infty}(e)|^{2}. By Lemma 7.1, the comfortability on the island is deformed by

    fisland\displaystyle\mathcal{E}_{f}^{island} =12m=0|f|1{|(1)τ(em)bηm|2+|ωbηm|2}\displaystyle=\frac{1}{2}\sum_{m=0}^{|f|-1}\left\{\;|(-1)^{\tau(e_{m})}b\eta_{m}|^{2}+|\omega b\eta_{m}|^{2}\;\right\}
    =|b|2m=0|f|11|bcω|2|(Qf𝜶f)(m+1)|2\displaystyle=|b|^{2}\sum_{m=0}^{|f|-1}\frac{1}{|bc\omega|^{2}}|(Q_{f}\boldsymbol{\alpha}_{f})(m+1)|^{2}
    =1|c|2(SfdIf)𝜶f2.\displaystyle=\frac{1}{|c|^{2}}||\;(S_{f}-dI_{f})\boldsymbol{\alpha}_{f}\;||^{2}.

    Then we have

    island=fFfisland=1|c|2(SdI)𝜶in2.\mathcal{E}^{island}=\sum_{f\in F}\mathcal{E}_{f}^{island}=\frac{1}{|c|^{2}}||(S-dI)\boldsymbol{\alpha}_{in}||^{2}. (7.30)
  2. (2)

    Bridge: By Lemma 7.1, we have

    bridge\displaystyle\mathcal{E}^{bridge} =12eAbr|(1)τ(e)ω(η(e)+dη(e¯))|2\displaystyle=\frac{1}{2}\sum_{e\in A_{br}}|(-1)^{\tau(e)}\omega(\eta(e)+d\eta(\bar{e}))|^{2}
    =1+d22eAbr|η(e)|2+deAbrRe[eAbrη(e)η(e¯)¯]\displaystyle=\frac{1+d^{2}}{2}\sum_{e\in A_{br}}|\eta(e)|^{2}+d\cdot\sum_{e\in A_{br}}\mathrm{Re}\left[\sum_{e\in A_{br}}\eta(e)\overline{\eta(\bar{e})}\right]
    =12{(1+d2)η2+2dη,ση}\displaystyle=\frac{1}{2}\left\{(1+d^{2})\;||\eta||^{2}+2d\cdot\langle\eta,\sigma\eta\rangle\right\}
    =12(σ+dI)η2\displaystyle=\frac{1}{2}||(\sigma+dI)\eta||^{2}
    =12|bc|2(σ+dI)(SdI)𝜶in2.\displaystyle=\frac{1}{2|bc|^{2}}||(\sigma+dI)(S-dI)\boldsymbol{\alpha}_{in}||^{2}. (7.31)

Combining (7.30) with (7.31), we obtain the desired conclusion. ∎

Now let us set the inflow inserting 11 from a tail which is chosen uniformly at random, that is, an inflow 𝜶in\boldsymbol{\alpha}_{in} is selected randomly from

{δe|eAbr}.\{\delta_{e}\;|\;e\in A^{br}\}.

Each probability that 𝜶in=δe\boldsymbol{\alpha}_{in}=\delta_{e} (eAbre\in A_{br}) is 1/|Abr|1/|A_{br}|. We are interested in the average of the comfortability with respect to this randomly setting of the inflow, that is,

𝔼[]=1|Abr|eAbr(e),\mathbb{E}[\mathcal{E}]=\frac{1}{|A_{br}|}\sum_{e\in A_{br}}\mathcal{E}^{(e)},

where (e)\mathcal{E}^{(e)} is the comfortability with the inflow δe\delta_{e}. The average of the comfortability is expressed as follows.

Theorem 7.1.

Under the Assumption 1 and the uniformly at random inflow, the average of the comfortability is expressed by

𝔼[]\displaystyle\mathbb{E}[\mathcal{E}] =1|Abr|2+|b|2|b|2fF|f|1|a|2|f|| 1(aω)|f||2\displaystyle=\frac{1}{|A_{br}|}\frac{2+|b|^{2}}{|b|^{2}}\sum_{f\in F}|f|\frac{1-|a|^{2|f|}}{|\;1-(a\omega)^{|f|}\;|^{2}}
+1|Abr|d|b|2fF1| 1(aω)|f||2eff¯{(aω)distf(e,e¯)(1|a|2distf(e¯,e))\displaystyle\qquad+\frac{1}{|A_{br|}}\frac{d}{|b|^{2}}\sum_{f\in F}\frac{1}{|\;1-(a\omega)^{|f|}\;|^{2}}\sum_{e\in f\cap\bar{f}}\bigg{\{}\;(a\omega)^{\mathrm{dist}_{f}(e,\bar{e})}\;(1-|a|^{2\mathrm{dist}_{f}(\bar{e},e)})
+(aω¯)distf(e¯,e)(1|a|2distf(e,e¯))}.\displaystyle\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad\qquad+(\overline{a\omega})^{\mathrm{dist}_{f}(\bar{e},e)}\;(1-|a|^{2\mathrm{dist}_{f}(e,\bar{e})})\;\bigg{\}}.

Here |f||f| is the length of the facial walk ff in (G,ρ,τ)(G,\rho,\tau) and distf(e,e)\mathrm{dist}_{f}(e,e^{\prime}) is the distance from ee to ee^{\prime} along the facial walk ff in (G,ρ,τ)(G,\rho,\tau) for any e,efe,e^{\prime}\in f.

Remark 7.1.

Theorem 1.2 corresponds to the case for a>0a>0 and ω=1\omega=1 in Theorem 7.1.

Proof.

By Lemma 7.2, it holds that

𝔼[]\displaystyle\mathbb{E}[\mathcal{E}] =1|c|2𝔼[Q𝜶2]+1|bc|2𝔼[(σ+dI)Q𝜶2]\displaystyle=\frac{1}{|c|^{2}}\mathbb{E}[\;||Q\boldsymbol{\alpha}||^{2}\;]+\frac{1}{|bc|^{2}}\mathbb{E}[\;||(\sigma+dI)Q\boldsymbol{\alpha}||^{2}\;]
=1|c|21|Abr|tr(QQ)+12|bc|21|Abr|tr([(σ+dI)Q][(σ+dI)Q])\displaystyle=\frac{1}{|c|^{2}}\frac{1}{|A_{br}|}\mathrm{tr}(Q^{*}Q)+\frac{1}{2|bc|^{2}}\frac{1}{|A_{br}|}\mathrm{tr}(\;[(\sigma+dI)Q]^{*}[(\sigma+dI)Q]\;)
=1|Abr|2+|b|22|bc|2tr(QQ)+1|Abr|d|bc|2tr(QQσ)\displaystyle=\frac{1}{|A_{br}|}\frac{2+|b|^{2}}{2|bc|^{2}}\mathrm{tr}(QQ^{*})+\frac{1}{|A_{br}|}\frac{d}{|bc|^{2}}\mathrm{tr}(QQ^{*}\sigma)

Now to describe the above RHS more explicitly, let us compute QQQQ^{*} as follows. Note that Q=fFQfQ=\oplus_{f\in F}Q_{f}, and for each fFf\in F with f=(e0,e1,,e|f|1)f=(e_{0},e_{1},\dots,{e_{|f|-1}}), QfQ_{f} can be expanded by

Qf\displaystyle Q_{f} =bcωPf(IaωPf)1\displaystyle=bc\omega P_{f}(I-a\omega P_{f})^{-1}
=bcω1(aω)|f|(I+aωPf++(aωPf)|f|1).\displaystyle=\frac{bc\omega}{1-(a\omega)^{|f|}}(I+a\omega P_{f}+\cdots+(a\omega P_{f})^{|f|-1}).

since Pf|f|=IfP_{f}^{|f|}=I_{f}. Then we have

QfQf\displaystyle Q_{f}Q_{f}^{*} =|bc|2|1(aω)n|2k=0n1(,mn with m=k(aω)m(aω¯))Pfk.\displaystyle=\frac{|bc|^{2}}{|1-(a\omega)^{n}|^{2}}\sum_{k=0}^{n-1}\left(\sum_{{\scriptsize\begin{matrix}\ell,m\in\mathbb{Z}_{n}\\ \text{ with }m-\ell=k\end{matrix}}}(a\omega)^{m}(\overline{a\omega})^{\ell}\right)P_{f}^{k}. (7.33)

Here wk:=,mn,m=k(aω)m(aω¯)w_{k}:=\sum_{\ell,m\in\mathbb{Z}_{n},\;m-\ell=k}\;(a\omega)^{m}(\overline{a\omega})^{\ell} is reduced to

wk=1|b|2{(aω)k(1|a|2(|f|k))+(aω¯)|f|k(1|a|2k)}.w_{k}=\frac{1}{|b|^{2}}\left\{(a\omega)^{k}(1-|a|^{2(|f|-k)})+(\overline{a\omega})^{|f|-k}(1-|a|^{2k})\right\}.

Inserting the above expression for wkw_{k} into (7.33), we obtain

(QfQf)e,em=(1)emeτ|b|2|1(aω)|f||2×{(aω)m(1|a|2(|f|(m)))+(aω¯)|f|(m)(1|a|2(m))}.(Q_{f}Q_{f}^{*})_{e_{\ell},e_{m}}=(-1)^{\int_{e_{m}}^{e_{\ell}}\tau}\frac{|b|^{2}}{|1-(a\omega)^{|f|}|^{2}}\\ \times\left\{(a\omega)^{m-\ell}(1-|a|^{2(|f|-(m-\ell))})+(\overline{a\omega})^{|f|-(m-\ell)}(1-|a|^{2(m-\ell)})\right\}. (7.34)

By using (7.34), tr(QQ)\mathrm{tr}(QQ^{*}) and tr(QQσ)\mathrm{tr}(QQ^{*}\sigma) are described by

tr(QQ)\displaystyle\mathrm{tr}(QQ^{*}) =fFtr[|b|2|1(aω)|f||2(1|a|2|f|)Pf0]\displaystyle=\sum_{f\in F}\mathrm{tr}\left[\frac{|b|^{2}}{|1-(a\omega)^{|f|}|^{2}}(1-|a|^{2|f|})P_{f}^{0}\right]
=|b|2fF|f|1|a|2|f||1(aω)|f||2,\displaystyle=|b|^{2}\sum_{f\in F}|f|\frac{1-|a|^{2|f|}}{|1-(a\omega)^{|f|}|^{2}}, (7.35)

and

tr(QQσ)\displaystyle\mathrm{tr}(QQ^{*}\sigma) =ee(QQ)e,e(σ)e,e=e,e¯(QQ)e¯,e\displaystyle=\sum_{e}\sum_{e^{\prime}}(QQ^{*})_{e,e^{\prime}}(\sigma)_{e^{\prime},e}=\sum_{e,\bar{e}}(QQ^{*})_{\bar{e},e}
=fFeff¯(QQ)e¯,e\displaystyle=\sum_{f\in F}\sum_{e\in f\cap\bar{f}}(QQ^{*})_{\bar{e},e}
=fF|b|2|1(aω)|f||2\displaystyle=\sum_{f\in F}\frac{|b|^{2}}{|1-(a\omega)^{|f|}|^{2}}
eff¯{(aω)distf(e,e¯)(1|a|2distf(e¯,e))+(aω)distf(e¯,e)(1|a|2dist(e,e¯))}\displaystyle\qquad\sum_{e\in f\cap\bar{f}}\left\{(a\omega)^{\mathrm{dist}_{f}(e,\bar{e})}(1-|a|^{2\mathrm{dist}_{f}(\bar{e},e)})+(a\omega)^{\mathrm{dist}_{f}(\bar{e},e)(1-|a|^{2\mathrm{dist}(e,\bar{e})})}\right\} (7.36)

Here the last equality derives from distf(e¯,e)=|f|distf(e,e¯)\mathrm{dist}_{f}(\bar{e},e)=|f|-\mathrm{dist}_{f}(e,\bar{e}) and (1)ee¯τ=1(-1)^{\int_{e}^{\bar{e}}}\tau=1 since e¯F\bar{e}\in F and e,e¯e,\bar{e} are in the same sheet for any eff¯e\in f\cap\bar{f}. Combining (7) with (7.35) and (7.36), we obtain the desired conclusion. ∎


Acknowledgments We would like to thank Takumi Kakegawa, and professors Kenta Ozeki, Atsuhiro Nakamoto for the fruitful discussions and suggestions. Yu.H. acknowledges financial supports from the Grant-in-Aid of Scientific Research (C) Japan Society for the Promotion of Science (Grant No. 18K03401, No. 23K03203). E.S. acknowledges financial supports from the Grant-in-Aid of Scientific Research (C) Japan Society for the Promotion of Science (Grant No. 24K06863) and Research Origin for Dressed Photon.

References

  • [1] Ambainis, A., Bach, E., Nayak, A., Vishwanath, A. and Watrous, J., One-dimensional quantum walks, Proc. 33rd Annual ACM Symp. Theory of Computing, (2001) 37–49.
  • [2] Apers, S. and Sarlette, A., Quantum fast-forwarding; Markov chains and graph property testing, Quantum Information and Computation 19 (2019) 181–213.
  • [3] Colin de Verdière, Y., Sur un nouvel invariant des graphes et un critère de planarit, Journal of Combinatorial Theory B 50 (1990) 11–21.
  • [4] Feldman, E and Hillery, M., Quantum walks on graphs and quantum scattering theory, Contemporary Mathematics 381 (2005) 71–96.
  • [5] Feldman, E and Hillery, M., Modifying quantum walks: A scattering theory approach, Journal of Physics A: Mathematical and Theoretical 40 (2007) 11319.
  • [6] Gross, J. L. and Tucker, W. T., Topological Graph Theory, Dover Pulications, New York (2001).
  • [7] Higuchi, K., Feynman-type representation of the scattering matrix on the line via a discrete-time quantum walk, Journal of Physics A: Mathematical and Theoretical 54 (2021) 235203.
  • [8] Higuchi, Yu. and Segawa, E., Dynamical system induced by quantum walks, Journal of Physics A: Mathematical and Theoretical 52 (2019) 39520.
  • [9] Higuchi, Yu. and Segawa, E., Circuit equation of Grover walk, Annales Henri Poincaré25 (2024) 3739–3777.
  • [10] Ko, C, K., Konno, N., Yoo, H, J and Segawa, E., How does Grover walk recognize the shape of crystal lattice?, Quantum Information Processing 17 (2017) 167–185.
  • [11] Konno, N., Portugal, R., Sato, I. and Segawa, E., Partition-based discrete-time quantum walks, Quantum Information Processing 17 (2018) 1–35.
  • [12] Krovi, H. and Brun, T. A., Quantum walks on quotient graphs, Physical Review A 75 (2007) 062332.
  • [13] Mizutani, Y., Horikiri, T., Matsuoka, L., Higuchi, Yu. and Segawa E., Implementation of a discrete-time quantum walk with a circulant matrix on a graph by optical polarizing elements, Physical Review A 106 (2022) 022402.
  • [14] Mohar, B. and Thomassen, C., Graphs on Surfaces, Johns Hopkins University Press (2001).
  • [15] Nakamoto, A. and Ozeki, K., “Kyokumenjou no Gurahu Riron” (Graphs on Surfaces), Saiensu-Sha (2021) (Japanese book)
  • [16] Nordhaus, E. A. and Stewart, B. M., On the Maximum Genus of a Graph, Journal of Combinatorial Theory B 11 (1971) 258–267.
  • [17] Portugal, R., Quantum Walk and Search Algorithm, 2nd Ed., Springer Nature Switzerland (2018).
  • [18] Ringel, Y. and Youngs, J. W. T., Solution of the Heawood map-coloring problem, Procedings of National Academy of Sciences 60 (1968) 438–445.
  • [19] Watrous, J., Quantum simulations of classical random walks and undirected graph connectivity, Journal of Computer and System Sciences 62 (2007) 376–391.