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

No Infinite Tail Beats Optimal Spatial Search

Weichen Xie xiew@clarkson.edu Department of Mathematics, Clarkson University, Potsdam, New York, USA 13699-5815    Christino Tamon tino@clarkson.edu Department of Computer Science, Clarkson University, Potsdam, New York, USA 13699-5815
Abstract

Farhi and Gutmann (Physical Review A, 57(4):2403, 1998) proved that a continuous-time analogue of Grover search (also called spatial search) is optimal on the complete graphs. We extend this result by showing that spatial search remains optimal in a complete graph even in the presence of an infinitely long path (or tail). If we view the latter as an external quantum system that has a limited but nontrivial interaction with our finite quantum system, this suggests that spatial search is robust against a coherent infinite one-dimensional probe. Moreover, we show that the search algorithm is oblivious in that it does not need to know whether the tail is present or not, and if so, where it is attached to.

I Introduction

The celebrated quantum search algorithm of Grover [12] provides a provable quadratic speedup over any classical algorithm. Shortly thereafter, Farhi and Gutmann proposed an analogue analog of Grover’s algorithm [8]. They defined the Grover search algorithm via a continuous-time quantum walk on a complete graph where the oracle or target vertex is marked by a suitably weighted self-loop. Remarkably, the Farhi-Gutmann algorithm achieved perfect fidelity on complete graphs of any size. In contrast, this property does not hold for Grover search (viewed as a discrete-time quantum walk) as it is inherently a bounded-error probabilistic algorithm.

This continuous-time search problem was later generalized by Childs and Goldstone [6] to arbitrary finite graphs where it is known as the spatial search problem. A collection of different families of finite graphs had been studied in this context; for example, see [6, 13, 21, 9, 4, 19]. But, to date, spatial search has not been studied on infinite graphs as it seems that the quantum walk will escape or diffuse to infinity before having a chance to localize on the marked vertex (or oracle). The goal of this work is to disabuse ourselves of this highly plausible intuition.

We consider infinite graphs which are obtained by attaching an infinite path (or tail) to a finite graph. This family of graphs with tails was explored by Golinskii [11]. In this work, we view the finite graph as our operational quantum system for performing quantum search and the tail as an external (possibly, adversarial) infinite-dimensional quantum system which interacts with our finite sytem in a coherent manner.

The main question we explore in this work is: can spatial search still be performed optimally in the presence of an infinite-dimensional probe? We provide a positive answer to this question for complete graphs. This extends the result of Farhi-Gutmann [8] to the infinite setting. Moreover, the quantum search algorithm is oblivious as it does not need to know whether the infinite-dimensional probe is present (or not) and where it is attached to (if present). Since we give our adversary the benefit of an infinite-dimensional quantum system, this serves only to strengthen the result.

Our technique relies on the theory of Jacobi operators (see [11, 7]). The main idea is to decompose the adjacency operator AA of our infinite graph using two pairwise orthogonal invariant subspaces (see Golinskii [11], Theorem 1.2) where the first one is finite-dimensional while the second one is infinite-dimensional. The next crucial observation is that spatial search takes place in the infinite-dimensional invariant subspace of AA. Moreover, the action of AA in this infinite-dimensional invariant subspace is given by a finite rank Jacobi matrix JJ. Finally, we show that the initial and target states of the spatial search are nearly confined in a two-dimensional subspace spanned by two bound states of JJ. This yields the claimed spatial search result. To the best of our knowledge, this is the first result which explores optimal spatial search on infinite graphs.

As outlined above, our argument is the standard argument for showing optimal spatial search in the finite setting (see [3, 4, 5]). Namely, we show that the initial and target states are spanned by two distinct eigenvectors of the perturbed adjacency matrix. A contribution of this work is to show that this argument holds in the infinite setting via bound states of the reduced adjacency operator. We believe that this argument might be useful in other settings.

Following a brief discussion of basic notation and terminology in Section II, we prove our main results in Sections III and IV. Then, we justify the optimality of our search algorithm in Section V. Finally, we conclude with some open questions in Section VI.

II Preliminaries

We introduce the basic notation and terminology that we will use throughout. The set of all positive integers is denoted +\mathbb{Z}^{+} and the set of all complex numbers with unit modulus is denoted U(1)U(1). For vectors x,yx,y where x=ζyx=\zeta y, for some ζU(1)\zeta\in U(1), we write xyx\equiv y. We adopt standard asymptotic notation: o(fn)o(f_{n}) denotes any function gng_{n} so that gn/fn0g_{n}/f_{n}\to 0, 𝒪(fn)\mathcal{O}(f_{n}) denotes functions gng_{n} for which gn/fng_{n}/f_{n} is bounded from above by a constant, and Ω(fn)\Omega(f_{n}) denotes functions gng_{n} where gn/fng_{n}/f_{n} is bounded from below by a constant, where in each case nn\to\infty; see [20]. In our case, the asymptotic parameter nn corresponds to the size of a finite graph.

Graphs and operators.

We study undirected and connected graphs G=(V,E)G=(V,E) with vertex set VV and edge set EE, respectively. The adjacency matrix AA of GG is a symmetric matrix whose (i,j)(i,j) entry is 11 if (i,j)E(i,j)\in E and 0 otherwise. For a vertex uu, let N(u)={vV:(u,v)E}N(u)=\{v\in V:(u,v)\in E\} denote the set of neighbors of uu. The degree of vertex uu, denoted deg(u)\deg(u), is the cardinality of N(u)N(u). The complete graph (or clique) on nn vertices is denoted KnK_{n}. A rooted graph (G,r)(G,r) is a graph GG with a distinguished vertex rr which we call the root. See [10] for further background on algebraic graph theory.

We allow countably infinite graphs, in which case, V=+V=\mathbb{Z}^{+} (see [16]). For example, the infinite path PP_{\infty} has edges which are consecutive positive integers; its adjacency matrix is known as the free Jacobi matrix (see [11]). Related to these graphs, we associate a complex separable Hilbert space =2(V)\mathcal{H}=\ell^{2}(V) equipped with the inner product x,y=uVxu¯yu\langle x,y\rangle=\sum_{u\in V}\overline{x_{u}}y_{u}, for vectors x,yx,y\in\mathcal{H}. For 2(V)\ell^{2}(V), a standard basis is {eu:uV}\{e_{u}:u\in V\}, where eue_{u} is the unit vector corresponding to vertex uu. An infinite graph 𝒢\mathcal{G} is locally finite if deg(u)<\deg{(u)}<\infty for all uVu\in V. For such a graph 𝒢\mathcal{G}, the adjacency operator AA is a linear operator that maps the standard basis vector eve_{v} to the vector associated with the neighboring vertices N(v)N(v); that is, Aev=uN(v)au,veuAe_{v}=\sum_{u\in N(v)}a_{u,v}e_{u}, or simply au,v=eu,Aeva_{u,v}=\langle e_{u},Ae_{v}\rangle. Let deg(𝒢)=sup{deg(u):uV}\deg{(\mathcal{G})}=\sup\{\deg{(u)}:u\in V\}. If deg(𝒢)<\deg(\mathcal{G})<\infty, then the adjacency operator AA is a bounded self-adjoint operator (see [15]).

The spectrum of a linear operator AA is the set σ(A)\sigma(A) of all complex numbers λ\lambda where λIA\lambda I-A is not invertible. For a bounded and self-adjoint operator AA, its spectrum can be classified further into the point spectrum σp(A)\sigma_{p}(A) and the continuous spectrum σc(A)\sigma_{c}(A). The point spectrum consists of all eigenvalues λ\lambda\in\mathbb{C} of AA such that Ax=λxAx=\lambda x for some nonzero x2(V)x\in\ell^{2}(V). On the contrary, the values in σc(A)\sigma_{c}(A) are not eigenvalues of AA and have no corresponding eigenvectors in 2(V)\ell^{2}(V).

The spectral theorem (see [1, 17]) states that for a bounded self-adjoint operator AA on a complex Hilbert space \mathcal{H}, there exists a unique resolution of the identity EE on the Borel subsets of σ(A)\sigma(A) so that A=σ(A)λ𝑑E(λ)A=\int_{\sigma(A)}\lambda\ dE(\lambda). Moreover, if ff is a bounded Borel function on σ(A)\sigma(A), then f(A)=σ(A)f(λ)𝑑E(λ)f(A)=\int_{\sigma(A)}f(\lambda)\ dE(\lambda). We also use a decomposition induced by invariant subspaces (see [1], Theorem 3, section 40) which states that if WkW_{k} (k=1,2,,m)(k=1,2,\cdots,m) are pairwise orthogonal invariant subspaces of AA, that is, =k=1mWk\mathcal{H}=\bigoplus_{k=1}^{m}W_{k} and AWkWkAW_{k}\subset W_{k}, for each kk, then A=k=1mAkPkA=\sum_{k=1}^{m}A_{k}P_{k}, where PkP_{k} is the projection on WkW_{k} and AkA_{k} is the restriction of AA to WkW_{k}.

Spatial Search.

A continuous-time quantum walk on an infinite graph 𝒢\mathcal{G} with bounded self-adjoint adjacency operator AA is given by the unitary operator eitAe^{-itA} (acting on the Hilbert space 2(V)\ell^{2}(V)). Our focus is on infinite graphs obtained from a finite connected rooted graph (G,r)(G,r) by attaching an infinite path PP_{\infty} at the root vertex rr; denote the resulting infinite graph as 𝒢=G(r,P)\mathcal{G}=G(r,P_{\infty}). As we use {1,2,,n}\{1,2,\ldots,n\} to label the vertices of a finite rooted graph of order nn, we take the liberty to designate the last vertex nn as the root (without loss of generality). These are the graphs with tails studied by Golinskii [11].

We say the infinite graph 𝒢=Gn(P)\mathcal{G}=G_{n}(P_{\infty}) has optimal spatial search (adopting [4]) if there is a real γ>0\gamma>0 so that for each vertex ww of GnG_{n}, a continuous-time quantum walk on 𝒢\mathcal{G} with a self-loop on ww of weight γ\gamma will unitarily map the principal eigenvector z1z_{1} of GnG_{n} to the unit vector ewe_{w} with constant fidelity in time t=𝒪(1/ϵ1)t=\mathcal{O}(1/\epsilon_{1}), where ϵ1=|ew,z1|\epsilon_{1}=|\langle e_{w},z_{1}\rangle|. That is,

|ew,eit(A+γPw)z1|=Ω(1),\displaystyle\left|\langle e_{w},e^{-it(A+\gamma P_{w})}z_{1}\rangle\right|=\Omega(1),

where AA is the adjacency operator of 𝒢\mathcal{G} and PwP_{w} is the projection onto the subspace spanned by ewe_{w}. It is customary to assume ϵ1=o(1)\epsilon_{1}=o(1) as otherwise we already have a constant overlap between the target state ewe_{w} and the initial state z1z_{1}.

III Infinite lollipop is optimal

For n2n\geq 2, consider the infinite lollipop graph n=Kn(P)\mathcal{L}_{n}=K_{n}(P_{\infty}). The vertices of the infinite path PP_{\infty} are labelled with n+1,n+2,n+1,n+2,\ldots. To this lollipop graph n\mathcal{L}_{n}, we place a self-loop of weight γ\gamma at vertex 11 (the oracle or target vertex) and denote the resulting infinite graph as n(γ)\mathcal{L}_{n}(\gamma). Please see Figure 1.

Refer to caption
Figure 1: Optimal spatial search on the infinite lollipop with the oracle hiding at vertex 11.

The adjacency operator HH of n(γ)\mathcal{L}_{n}(\gamma) is obviously a bounded self-adjoint linear operator on l2(+)l^{2}(\mathbb{Z}^{+}). It can be written as an infinite dimensional matrix under the standard basis {ej}j=1\{e_{j}\}_{j=1}^{\infty} as

H=(γ1110111011 01 101).\displaystyle H=\begin{pmatrix}\gamma&1&\cdots&1&&&&\\ 1&0&\cdots&1&&&&\\ \vdots&\vdots&\ddots&\vdots&&&&\\ 1&1&\cdots&0&1&&&&\\ &&&1&\ 0&1&&\\ &&&&\ 1&0&1&\\ &&&&&\ddots&\ddots&\ddots\end{pmatrix}.

Our goal is to prove the following infinite analogue of the Farhi-Gutmann result.

Theorem 1.

For γ=n+𝒪(1)\gamma=n+\mathcal{O}(1), we have

|e1,eitHz1|=Ω(1),\displaystyle|\langle e_{1},e^{-itH}z_{1}\rangle|=\Omega(1),

for t=π/2nt=\pi/2\sqrt{n}, where z1=n1/2j=1nejz_{1}=n^{-1/2}\sum_{j=1}^{n}e_{j}.

By a theorem of Golinskii ([11], Theorem 1.2), using a change of basis, HH can be written in a block diagonal form. This new basis {e~k}k=1\{\tilde{e}_{k}\}_{k=1}^{\infty} is defined as follows. Let

e~k=ek,k=n,n+1,\displaystyle\tilde{e}_{k}=e_{k},\ \ k=n,n+1,\cdots (1)

and

e~n1=1n1j=1n1ej.\displaystyle\tilde{e}_{n-1}=\frac{1}{\sqrt{n-1}}\sum_{j=1}^{n-1}e_{j}. (2)

The next orthogonal basis vector is then given by

e~n2\displaystyle\tilde{e}_{n-2}^{\prime} =He~n1e~n1,He~n1e~n1e~n,He~n1e~n.\displaystyle=H\tilde{e}_{n-1}-\langle\tilde{e}_{n-1},H\tilde{e}_{n-1}\rangle\tilde{e}_{n-1}-\langle\tilde{e}_{n},H\tilde{e}_{n-1}\rangle\tilde{e}_{n}.

which, after normalization, yields

e~n2=1(n2)(n1)((n1)e1e~n1).\displaystyle\tilde{e}_{n-2}=\frac{1}{\sqrt{(n-2)(n-1)}}((n-1)e_{1}-\tilde{e}_{n-1}). (3)

We take the remaining basis vectors to be the non-principal columns of the Fourier matrix of order n2n-2. In particular, the basis vector e~k\tilde{e}_{k} is defined as

e~k=1n2j=1n2e2πi(j1)k(n2)ej(k=1,,n3).\displaystyle\tilde{e}_{k}=\frac{1}{\sqrt{n-2}}\sum_{j=1}^{n-2}e^{\frac{2\pi i(j-1)k}{(n-2)}}e_{j}\ \ \ (k=1,\ldots,n-3).

Since He~n2=(γn2n11)e~n2+γn2n1e~n1H\tilde{e}_{n-2}=\left(\gamma\frac{n-2}{n-1}-1\right)\tilde{e}_{n-2}+\gamma\frac{\sqrt{n-2}}{n-1}\tilde{e}_{n-1}, the subspace 𝒮=span{e~n2,e~n1,}\mathcal{S}=\operatorname{span}\{\tilde{e}_{n-2},\tilde{e}_{n-1},\cdots\} is HH-invariant. Under the new basis, the adjacency operator HH becomes

H=(In2OOH^)\displaystyle H=\begin{pmatrix}-I_{n-2}&O\\ O&\widehat{H}\end{pmatrix}

As will be clear soon, it suffices for us to restrict our focus on the operator H^\widehat{H}, which is the operator HH restricted to the subspace 𝒮=span{e~n2,e~n1,}\mathcal{S}=\operatorname{span}\{\tilde{e}_{n-2},\tilde{e}_{n-1},\cdots\}. This is because the initial state z1z_{1} and the target state e1e_{1} of our spatial search problem both have non-negligible overlap with 𝒮\mathcal{S}.

It follows from the preceding analysis that the operator HH under the basis {e~n2,e~n1,}\{\tilde{e}_{n-2},\tilde{e}_{n-1},\ldots\} is given by

H^=e~n2e~n1e~ne~n+1(γn2n11γn2n1γn2n1n2+γn1n1n10110 1).\displaystyle\widehat{H}=\begin{array}[]{rl}\begin{matrix}{\scriptstyle\tilde{e}_{n-2}}\\ {\scriptstyle\tilde{e}_{n-1}}\\ {\scriptstyle\tilde{e}_{n}}\\ {\scriptstyle\tilde{e}_{n+1}}\\ \vdots\end{matrix}\!\!\!\!\!&\begin{pmatrix}\gamma\frac{n-2}{n-1}\!-\!1&\gamma\frac{\sqrt{n-2}}{n-1}&&&&&\\ \gamma\frac{\sqrt{n-2}}{n-1}&n\!\!-\!\!2\!+\!\frac{\gamma}{n-1}&\sqrt{n\!\!-\!\!1}&&&&\\ &\sqrt{n\!\!-\!\!1}&0&1&&&\\ &&1&0&\ \ 1&&\\ &&&\ddots&\ddots&\ddots\end{pmatrix}.\end{array} (5)

This symmetric tridiagonal matrix is an eventually-free or finite rank Jacobi matrix (see [11]) whose full spectrum can be computed via the so-called Jost solution (see [7], Appendix). The Jost solution is a vector y(x)=(y1(x),y2(x),)Ty(x)=(y_{1}(x),y_{2}(x),\cdots)^{T} that satisfies the eigen-equation of H^\widehat{H} with eigenvalue of the form x+1xx+\frac{1}{x}, namely,

H^y(x)=(x+1x)y(x),\displaystyle\widehat{H}y(x)=\left(x+\frac{1}{x}\right)y(x), (6)

and also the degree condition

limkxkyk(x)=1.\displaystyle\lim_{k\to\infty}x^{-k}y_{k}(x)=1.

Given the special form of H^\widehat{H}, we can set

yk(x)=xk,k=3,4,,\displaystyle y_{k}(x)=x^{k},\ \ k=3,4,\cdots, (7)

and use (6), to get the Jost polynomial

y0(x)\displaystyle y_{0}(x) =1γn1n2[(2n)x4+[(n3)γ+42n)]x3\displaystyle=\frac{1}{\gamma}\sqrt{\frac{n-1}{n-2}}\cdot\left[(2-n)x^{4}+[(n-3)\gamma+4-2n)]x^{3}\right.
+[(n3)γ+52n]x2+[3nγ]x+1].\displaystyle+\left.[(n-3)\gamma+5-2n]x^{2}+[3-n-\gamma]\,x+1\right]. (8)

In order to compute the eigenvalues of H^\widehat{H}, we need the following spectral theorem for finite rank Jacobi operators. We will only need information about the point spectrum as will be clear soon.

Theorem 2.

([11], p8) Let JJ be an eventually-free Jacobi matrix and y0(x)y_{0}(x) be its Jost function. Then all roots of y0(x)y_{0}(x) in the complex unit disk are real and simple, y0(0)0y_{0}(0)\not=0. A real number λj\lambda_{j} is an eigenvalue of JJ if and only if

λj=xj+1xj,xj(1,1),y0(xj)=0.\displaystyle\lambda_{j}=x_{j}+\frac{1}{x_{j}},\quad x_{j}\in(-1,1),\quad y_{0}(x_{j})=0.

By choosing γ=n+𝒪(1)\gamma=n+\mathcal{O}(1), the roots of y0(x)y_{0}(x) in the complex unit disk can be approximated consecutively. First note that there are four real roots for y0(x)y_{0}(x) and two of them lie in the unit disk as indicated by the following table:

xx -\infty 1-1 1n{\scriptstyle\frac{1}{n}} 11 ++\infty
γn2n1y0{\gamma}{\scriptstyle\sqrt{\frac{n-2}{n-1}}}\cdot y_{0} -\infty γ+1>0\gamma\!+\!1\!>\!0 1n+𝒪(1n2)<0{\scriptstyle-\frac{1}{n}+\mathcal{O}(\frac{1}{n^{\scriptscriptstyle 2}})}\!<\!0 2nγ+𝒪(n)>02n\gamma\!+\!{\scriptstyle\mathcal{O}(n)}\!>\!0 -\infty
Table 1: For γ=n+𝒪(1)\gamma=n+\mathcal{O}(1) and sufficiently large nn, there are 44 sign changes of y0(x)y_{0}(x).

Denote the two roots within the unit disk as x±=1n+δ±x_{\pm}=\frac{1}{n}+\delta_{\pm}. Hence,

0\displaystyle 0 =γn2n1y0(x±)=1n+n2δ±2+o(1n)+o(n2δ±2).\displaystyle={\gamma}\sqrt{\frac{n-2}{n-1}}\cdot y_{0}(x_{\pm})=-\frac{1}{n}+n^{2}\delta_{\pm}^{2}+o\left(\frac{1}{n}\right)+o(n^{2}\delta_{\pm}^{2}).

In order for the left-hand side to attain 0 exactly, for all nn, at least the two highest order terms on the right-hand side should cancel perfectly; that is, o(1n)=1n+n2δ±2o(\frac{1}{n})=-\frac{1}{n}+n^{2}\delta_{\pm}^{2} which implies

x±=1n±1n3/2+o(1n3/2).\displaystyle x_{\pm}=\frac{1}{n}\pm\frac{1}{n^{3/2}}+o\left(\frac{1}{n^{3/2}}\right). (9)

Therefore, the two distinct eigenvalues of H^\widehat{H} are given by

λ±=n±n+𝒪(1).\displaystyle\lambda_{\pm}=n\pm\sqrt{n}+\mathcal{O}(1). (10)

The corresponding eigenvectors (or bound states) are given by y±=(y1(x),y2(x),)Ty_{\pm}=(y_{1}(x_{\mp}),y_{2}(x_{\mp}),\ldots)^{T} where the entries are defined by the Jost polynomials

y1(x)\displaystyle y_{1}(x_{\mp}) =±1γn1n21n3/2+𝒪(1n3),\displaystyle=\pm\frac{1}{\gamma}\sqrt{\frac{n-1}{n-2}}\cdot\frac{1}{n^{3/2}}+\mathcal{O}\left(\frac{1}{n^{3}}\right), (11)
y2(x)\displaystyle y_{2}(x_{\mp}) =1n11n21n12n5/2+𝒪(1n7/2),\displaystyle=\frac{1}{\sqrt{n-1}}\cdot\frac{1}{n^{2}}\mp\frac{1}{\sqrt{n-1}}\frac{2}{n^{5/2}}+\mathcal{O}\left(\frac{1}{n^{7/2}}\right), (12)
yk(x)\displaystyle y_{k}(x_{\mp}) =1nkkn(2k+1)/2+𝒪(1nk+1),k=3,4,.\displaystyle=\frac{1}{n^{k}}\mp\frac{k}{n^{(2k+1)/2}}+\mathcal{O}\left(\frac{1}{n^{k+1}}\right),\quad k=3,4,\cdots. (13)

Now, we are ready to prove Theorem 1.

First, note that both e1=1n1e~n2+n2n1e~n1e_{1}=\sqrt{\frac{1}{n-1}}\tilde{e}_{n-2}+\sqrt{\frac{n-2}{n-1}}\tilde{e}_{n-1} and z1=e~n1z_{1}=\tilde{e}_{n-1} are completely in the invariant subspace 𝒮\mathcal{S}. Thus, we can restrict our unitary evolution to 𝒮\mathcal{S}. Moreover, as e1e_{1} overlaps almost completely with e~n1\tilde{e}_{n-1}, it suffices to consider e~n1\tilde{e}_{n-1} as the target state. Hence, the fidelity can be further approximated as

|e1,eitHz1|n2n1|e¯1,eitH^e¯2|+𝒪(1n1/2),\displaystyle\left|\langle e_{1},e^{-itH}z_{1}\rangle\right|\geq\sqrt{\frac{n-2}{n-1}}\left|\langle\overline{e}_{1},e^{-it\widehat{H}}\overline{e}_{2}\rangle\right|+\mathcal{O}\left(\frac{1}{n^{1/2}}\right),

where e¯k\overline{e}_{k} is a unit vector which is the kk-th basis vector for the invariant subspace 𝒮=span{e~n2,e~n1,}\mathcal{S}=\operatorname{span}\{\tilde{e}_{n-2},\tilde{e}_{n-1},\cdots\}.

Notice that

limne¯k,y+y+,e¯ky+,y++e¯k,yy,e¯ky,y=1,k=1,2,\displaystyle\mathop{\mathrm{\lim}}\limits_{n\rightarrow\infty}{\frac{\langle\overline{e}_{k},y_{+}\rangle\langle y_{+},\overline{e}_{k}\rangle}{\langle y_{+},y_{+}\rangle}+\frac{\langle\overline{e}_{k},y_{-}\rangle\langle y_{-},\overline{e}_{k}\rangle}{\langle y_{-},y_{-}\rangle}}=1,\ \ k=1,2,

that is, both the initial and the target states lie in the two-dimensional invariant subspace spanned by the eigenvectors y+y_{+} and yy_{-}. Thus, it again suffices to consider the fidelity in this subspace.

Straightforward calculation shows that when γ=n+𝒪(1)\gamma=n+\mathcal{O}(1), the fidelity satisfies |e1,eitHz1|=1+o(1)|\langle e_{1},e^{-itH}z_{1}\rangle|=1+o(1) for time t=π2nt=\frac{\pi}{2\sqrt{n}}. If we normalize the adjacency matrix of KnK_{n}, we obtain 𝒪(n)\mathcal{O}(\sqrt{n}) time (matching Grover search).

IV Oracle at the edge of infinity

In this section, we show that even when the oracle is placed at the attachment vertex of the infinite path, spatial search remains optimal on n\mathcal{L}_{n}. See Figure 2.

Together with Section III, this will show that the search algorithm is oblivious as it does not need to know if the external probe (tail) is present or not. This is because the two cases surprisingly require the same asymptotic time for optimal spatial search. The claim follows from a similar analysis as before.

The adjacency operator of n(γ)\mathcal{L}_{n}(\gamma) is given as

H=(01110111γ11 01 101).\displaystyle H=\begin{pmatrix}0&1&\cdots&1&&&&\\ 1&0&\cdots&1&&&&\\ \vdots&\vdots&\ddots&\vdots&&&&\\ 1&1&\cdots&\gamma&1&&&&\\ &&&1&\ 0&1&&\\ &&&&\ 1&0&1&\\ &&&&&\ddots&\ddots&\ddots\end{pmatrix}.

First, we restrict our focus to the invariant subspace 𝒮=span{e~n1,e~n,}\mathcal{S}=\operatorname{span}\{\tilde{e}_{n-1},\tilde{e}_{n},\cdots\} since en=e~n𝒮e_{n}=\tilde{e}_{n}\in\mathcal{S} and z1=n1ne~n1+1ne~n𝒮z_{1}=\sqrt{\frac{n-1}{n}}\tilde{e}_{n-1}+\sqrt{\frac{1}{n}}\tilde{e}_{n}\in\mathcal{S}, where e~n1=1n1j=1n1ej\tilde{e}_{n-1}=\frac{1}{\sqrt{n-1}}\sum_{j=1}^{n-1}e_{j} and e~k=ek\tilde{e}_{k}=e_{k}, knk\geq n. Under the new basis that spans 𝒮\mathcal{S}, the operator HH is a rank-22 Jacobi matrix

H^=e~n1e~ne~n+1(n2n1n1γ110 1).\displaystyle\widehat{H}=\begin{array}[]{rl}\begin{matrix}{\scriptstyle\tilde{e}_{n-1}}\\ {\scriptstyle\tilde{e}_{n}}\\ {\scriptstyle\tilde{e}_{n+1}}\\ \vdots\end{matrix}\!\!\!\!\!&\begin{pmatrix}n\!\!-\!\!2\!&\sqrt{n\!\!-\!\!1}&&&&\\ \sqrt{n\!\!-\!\!1}&\gamma&1&&&\\ &1&0&\ \ 1&&\\ &&\ddots&\ddots&\ddots\end{pmatrix}.\end{array} (15)

Hence, the following Jost polynomials are obtained:

y0(x)\displaystyle y_{0}(x) =1n1[γx3+(n2)(γ1)x2\displaystyle=\frac{1}{\sqrt{n-1}}\cdot[-\gamma x^{3}+(n-2)(\gamma-1)x^{2}
+(2nγ)x+1],\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ +(2-n-\gamma)x+1], (16)
y1(x)\displaystyle y_{1}(x) =1n1(xγx2),\displaystyle=\frac{1}{\sqrt{n-1}}\cdot(x-\gamma x^{2}), (17)
yk(x)\displaystyle y_{k}(x) =xk,k=2,3,.\displaystyle=x^{k},\ \ k=2,3,\cdots. (18)

By choosing γ=n+𝒪(1)\gamma=n+\mathcal{O}(1), we can compute the two distinct roots for y0y_{0} that is within the interval [1,1][-1,1]

x±=1n±1n3/2+o(1n3/2),\displaystyle x_{\pm}=\frac{1}{n}\pm\frac{1}{n^{3/2}}+o\left(\frac{1}{n^{3/2}}\right), (19)

which yields the corresponding eigenvalues for H^\widehat{H}

λ±=n±n+𝒪(1).\displaystyle\lambda_{\pm}=n\pm\sqrt{n}+\mathcal{O}(1). (20)

As shown previously, the corresponding eigenvectors are defined by the values of Jost polynomials:

y1(x)\displaystyle y_{1}(x_{\mp}) =±1n11n3/2+𝒪(1n5/2)\displaystyle=\pm\frac{1}{\sqrt{n-1}}\cdot\frac{1}{n^{3/2}}+\mathcal{O}\left(\frac{1}{n^{5/2}}\right) (21)
y2(x)\displaystyle y_{2}(x_{\mp}) =1n22n5/2+𝒪(1n3)\displaystyle=\frac{1}{n^{2}}\mp\frac{2}{n^{5/2}}+\mathcal{O}\left(\frac{1}{n^{3}}\right) (22)
yk(x)\displaystyle y_{k}(x_{\mp}) =1nkkn(2k+1)/2+𝒪(1nk+1),k=3,4,.\displaystyle=\frac{1}{n^{k}}\mp\frac{k}{n^{(2k+1)/2}}+\mathcal{O}\left(\frac{1}{n^{k+1}}\right),\ \ k=3,4,\cdots. (23)

Following the same argument, we can see that for time t=π2nt=\frac{\pi}{2\sqrt{n}}, the fidelity satisfies |en,eitHz1|=1+o(1)|\langle e_{n},e^{-itH}z_{1}\rangle|=1+o(1).

Refer to caption
Figure 2: Optimal spatial search still occurs on the infinite lollipop even if the oracle sits at gateway vertex nn.

V Optimality

We show that the time bound obtained in Theorem 1 (and in Section IV) is optimal. To this end, we generalize an argument of Farhi and Gutmann [8] to a class of infinite graphs with tails.

For a finite graph GnG_{n} on nn vertices, take the cone G^n=K1+Gn\widehat{G}_{n}=K_{1}+G_{n} which is obtained by adding a new vertex (called the conical vertex) and connecting it to all vertices of GnG_{n}. Then, attach a tail to the conical vertex and denote this infinite graph as G^n(P)\widehat{G}_{n}(P_{\infty}). Notice that we recover the infinite lollipop when GnG_{n} is a clique.

Theorem 3.

Let GnG_{n} be a (n,d)(n,d)-regular graph, where d=ω(n)d=\omega(\sqrt{n}) for some δ>0\delta>0, and let 0\mathcal{H}_{0} be the adjacency operator of G^n(P)\widehat{G}_{n}(P_{\infty}). Let z12z_{1}\in\ell^{2} being natural embedding of the principal eigenvector of GnG_{n}. Suppose there is a time t0t_{0} so that for some γ\gamma\in\mathbb{R} and ζU(1)\zeta\in U(1), and for a vertex ww of GnG_{n}, we have

eit0(0+γPw)z1ζew2=o(1).\left\lVert e^{-it_{0}(\mathcal{H}_{0}+\gamma P_{w})}z_{1}-\zeta e_{w}\right\rVert^{2}=o(1).

Then, γt0=Ω(1/ϵ1)\gamma t_{0}=\Omega(1/\epsilon_{1}), where ϵ1=|ew,z1|\epsilon_{1}=|\langle e_{w},z_{1}\rangle|, provided 1/γϵ1o(1)1/\gamma\epsilon_{1}\in o(1).

The largest eigenvalue in σp(0)\sigma_{p}(\mathcal{H}_{0}) has a unique bound eigenstate β1\beta_{1} which satisfies β1z1=o(1)\left\lVert\beta_{1}-z_{1}\right\rVert=o(1). This can be shown using similar techniques as in previous sections.

We call a time-dependent state ψ(t)2(+)\psi(t)\in\ell^{2}(\mathbb{Z}^{+}) exponentially decaying if there is a positive integer MM so that for all mMm\geq M we have |em,ψ(t)|=𝒪(κm)|\langle e_{m},\psi(t)\rangle|=\mathcal{O}(\kappa^{m}), for some κ<1\kappa<1, for all tt.

Proof.

(of Theorem 3) Let w=0+γPw\mathcal{H}_{w}=\mathcal{H}_{0}+\gamma P_{w}. It suffices to show the claim under the assumption eitwβ1ζew2=o(1)\left\lVert e^{-it\mathcal{H}_{w}}\beta_{1}-\zeta e_{w}\right\rVert^{2}=o(1), since β1z1=o(1)\left\lVert\beta_{1}-z_{1}\right\rVert=o(1) and by using the triangle inequality for squared norm (see [18], eq. 18.5).

Following [8], we compare two Schrödinger evolutions given by

ψw(t)\displaystyle\psi_{w}^{\prime}(t) =iwψw(t),\displaystyle=-i\mathcal{H}_{w}\psi_{w}(t), ψw(0)=β1,\displaystyle\psi_{w}(0)=\beta_{1},
ψ0(t)\displaystyle\psi_{0}^{\prime}(t) =i0ψ0(t),\displaystyle=-i\mathcal{H}_{0}\psi_{0}(t), ψ0(0)=β1.\displaystyle\psi_{0}(0)=\beta_{1}.

Note ψ0(t)β1\psi_{0}(t)\equiv\beta_{1}, for all tt, as β1\beta_{1} is an eigenstate of 0\mathcal{H}_{0}.

For simplicity, we assume that spatial search achieves perfect fidelity, namely, ψw(t0)ew\psi_{w}(t_{0})\equiv e_{w}. The general case is handled using triangle inequality for squared norm.

The key quantity is M(t):=ψw(t)ψ0(t)2M(t):=\left\lVert\psi_{w}(t)-\psi_{0}(t)\right\rVert^{2}. First, notice that

M(t0)\displaystyle M(t_{0}) =2(1𝚁𝚎ew,β1)2(1ϵ1).\displaystyle=2(1-\mathtt{Re}\langle e_{w},\beta_{1}\rangle)\geq 2(1-\epsilon_{1}). (24)

Furthermore, we have M(t)=2𝚁𝚎ψw(t),ψ0(t)M^{\prime}(t)=-2\mathtt{Re}\langle\psi_{w}(t),\psi_{0}(t)\rangle^{\prime}. Given that the inner product is an infinite series, the existence of its derivative requires uniform convergence. As ψ0(t)β1\psi_{0}(t)\equiv\beta_{1} is exponentially decaying (by virtue of being a Jost solution), we have |(ψw(t))m¯(β1)m||(β1)m||\overline{(\psi_{w}(t))_{m}}(\beta_{1})_{m}|\leq|(\beta_{1})_{m}| and

m=1|(β1)m|CM+mMκm=CM+κM1κ<\displaystyle\sum_{m=1}^{\infty}|(\beta_{1})_{m}|\leq C_{M}+\sum_{m\geq M}\kappa^{m}=C_{M}+\frac{\kappa^{M}}{1-\kappa}<\infty

for a constant CMC_{M}. Thus, ψw(t),β1\langle\psi_{w}(t),\beta_{1}\rangle is uniformly convergent (see Titchmarsh [20], 1.11).

Since w\mathcal{H}_{w} is a finite-rank Jacobi matrix and ψw(t)=iwψw(t)\psi^{\prime}_{w}(t)=-i\mathcal{H}_{w}\psi_{w}(t), we see that ψw(t),ψ0(t)\langle\psi^{\prime}_{w}(t),\psi_{0}(t)\rangle is also uniformly convergent. This allows us to take the derivative of ψw(t),ψ0(t)\langle\psi_{w}(t),\psi_{0}(t)\rangle by termwise differentiation (see Titchmarsh [20], 1.72), i.e.,

ψw(t),ψ0(t)\displaystyle\langle\psi_{w}(t),\psi_{0}(t)\rangle^{\prime} =ψw(t),ψ0(t)+ψw(t),ψ0(t)\displaystyle\!=\!\langle\psi_{w}(t),\psi^{\prime}_{0}(t)\rangle\!+\!\langle\psi^{\prime}_{w}(t),\psi_{0}(t)\rangle
=iψw(t),0ψ0(t)+iwψw(t),ψ0(t).\displaystyle\!=\!-i\langle\psi_{w}(t),\mathcal{H}_{0}\psi_{0}(t)\rangle\!+\!i\langle\mathcal{H}_{w}\psi_{w}(t),\psi_{0}(t)\rangle.

So, we obtain

M(t)=2γ𝙸𝚖Pwψw(t),ψ0(t)2γPwψ0(t).\displaystyle M^{\prime}(t)=2\gamma\mathtt{Im}\langle P_{w}\psi_{w}(t),\psi_{0}(t)\rangle\leq 2\gamma\left\lVert P_{w}\psi_{0}(t)\right\rVert.

Thus, M(t)2γϵ1M^{\prime}(t)\leq 2\gamma\epsilon_{1}, which further implies

M(t0)=0t0M(t)𝑑t2γϵ1t0.\displaystyle M(t_{0})=\int_{0}^{t_{0}}M^{\prime}(t)dt\leq 2\gamma\epsilon_{1}t_{0}.

By combining the lower and upper bounds on M(t0)M(t_{0}), we get γt0(1ϵ1)/ϵ1=Ω(1/ϵ1)\gamma t_{0}\geq(1-\epsilon_{1})/\epsilon_{1}=\Omega(1/\epsilon_{1}) as ϵ1=o(1)\epsilon_{1}=o(1). ∎

Theorem 3 justifies the optimal time 1/ϵ11/\epsilon_{1} used to define spatial search. For the infinite lollipop, we have ϵ1=1/n\epsilon_{1}=1/\sqrt{n}, and thus, γt0=Ω(n)\gamma t_{0}=\Omega(\sqrt{n}). Since the search algorithm uses γ=n+𝒪(1)\gamma=n+\mathcal{O}(1), we get t0=Ω(1/n)t_{0}=\Omega(1/\sqrt{n}), which matches the bound achieved by Theorem 1. Hence, the algorithm is optimal.

VI Conclusion

In this work, we proved optimal spatial search occurs on cliques even in the presence of an infinite path. This generalized a known result of Farhi and Gutmann [8] to the infinite setting. We view this as a first step in showing that optimal spatial search is robust against an adversary modeled as an infinite-dimensional external quantum probe. Interesting directions for future work include extending the result to multiple tails or to tails induced by more general Jacobi matrices and strengthening the lower bound to other families of infinite graphs.

Our work was motivated by a question posed in [2]. We mention that discrete-time quantum walk on graphs with tails was studied in [14].

Acknowledgments

Work started during the workshop “Graph Theory, Algebraic Combinatorics, and Mathematical Physics” at Centre de Recherches Mathématiques (CRM), Université de Montréal. C.T. would like to thank CRM for its hospitality and support during his sabbatical visit. W.X. was supported by NSF grant DMS-2212755. We thank Pierre-Antoine Bernard and Luc Vinet for discussions.

References

  • [1] N.I. Akhiezer and I.M. Glazman. Theory of linear operators in Hilbert space. Courier Corporation, 2013.
  • [2] P.-A. Bernard, C. Tamon, L. Vinet, and W. Xie. Quantum state transfer in graphs with tails, 2022. arxiv:quant-ph/2211.14704.
  • [3] S. Chakraborty, L. Novo, A. Ambainis, and Y. Omar. Spatial search by quantum walk is optimal for almost all graphs. Physical Review Letters, 116(3):100501, 2016.
  • [4] S. Chakraborty, L. Novo, and J. Roland. Optimality of spatial search via continuous-time quantum walks. Physical Review A, 102(3):032214, 2020.
  • [5] A. Chan, C. Godsil, C. Tamon, and W. Xie. Of shadows and gaps in spatial search. Quantum Information and Computation, 22(13& 14):1110–1131, 2022.
  • [6] A. Childs and J. Goldstone. Spatial search by quantum walk. Physical Review A, 70:022314, 2004.
  • [7] D. Damanik and B. Simon. Jost functions and Jost solutions for Jacobi matrices, II. Decay and Analyticity. International Mathematics Research Notices, 2006:19396, 2006.
  • [8] E. Farhi and S. Gutmann. Analog analogue of a digital quantum computation. Physical Review A, 57(4):2403, 1998.
  • [9] A. Glos and T. Wong. Optimal quantum walk search on Kronecker graphs with dominant or fixed regular initiators. Physical Review A, 98:062334, 2018.
  • [10] C. Godsil and G. Royle. Algebraic Graph Theory. Springer, 2001.
  • [11] L. Golinskii. Spectra of infinite graphs with tails. Linear and Multilinear Algebra, 64(11):2270–2296, 2016.
  • [12] L. Grover. Quantum mechanics help in searching for a needle in a haystack. Physical Review Letters, 79:325, 1997.
  • [13] J. Janmark, D. Meyer, and T. Wong. Global symmetry is not necessary for fast quantum search. Physical Review Letters, 112:210502, 2014.
  • [14] N. Konno, E. Segawa, and M. Štefaňák. Relation between quantum walks with tails and quantum walks with sinks on finite graphs, 2021. arxiv:math-ph/2105.03111.
  • [15] B. Mohar. The spectrum of an infinite graph. Linear algebra and its applications, 48:245–256, 1982.
  • [16] B. Mohar and W. Woess. A survey on spectra of infinite graphs. Bulletin of the London Mathematical Society, 21(3):209–234, 1989.
  • [17] W. Rudin. Functional Analysis. McGraw-Hill, 2nd edition, 1991.
  • [18] S. Sternberg. A mathematical companion to quantum mechanics. Dover, 2019.
  • [19] H. Tanaka, M. Sabri, and R. Portugal. Spatial search on Johnson graphs by continuous-time quantum walk. Quantum Information Processing, 21(74), 2022.
  • [20] E.C. Titchmarsh. The Theory of Functions. Oxford University Press, 2nd edition, 1939.
  • [21] T. Wong. Quantum walk search on Johnson graphs. Journal of Physics A: Mathematical and Theoretical, 49(19):195303, 2016.