1: \documentclass[amsmath,amssymb,amsfonts]{revtex4}
2: \usepackage{amsthm,bbm,graphicx,bm}
3:
4: \newcommand{\ket} [1] {\lvert #1 \rangle}
5: \newcommand{\bra} [1] {\langle #1 \rvert}
6: \newcommand{\braket}[2]{\langle #1\vert #2 \rangle}
7: \newcommand{\proj}[1]{\ket{#1}\!\bra{#1}}
8: \newcommand{\id}{\mathrm{Id}}
9:
10: \newtheorem{proposition}{Proposition}
11: \newtheorem{corollary}{Corollary}
12:
13: \begin{document}
14:
15: \title{Exploring scalar quantum walks on Cayley graphs}
16:
17: \author{Olga Lopez Acevedo}
18: \thanks{Present address: Department of Physics, Nanoscience Center,
19: FIN-40014 University of Jyv\"{a}skyl\"{a}, Finland}
20: \email{lopez@phys.jyu.fi}
21:
22: \author{J\'er\'emie Roland}
23: \email{jroland@ulb.ac.be}
24:
25: \author{Nicolas J. Cerf}
26: \email{ncerf@ulb.ac.be}
27:
28: \affiliation{Quantum Information and Communication, Ecole Polytechnique, CP 165/59,
29: Universit\'e Libre de Bruxelles, B-1050 Brussels, Belgium}
30:
31: \date{\today}
32:
33: \begin{abstract}
34: A quantum walk, \emph{i.e.}, the quantum evolution of a particle
35: on a graph, is termed \emph{scalar} if the internal space of
36: the moving particle (often called the coin) has a dimension one.
37: Here, we study the existence of scalar quantum walks
38: on Cayley graphs, which are built from the generators of a group.
39: After deriving a necessary condition on these generators for the existence
40: of a scalar quantum walk, we present a general method to express
41: the evolution operator of the walk, assuming homogeneity of the evolution.
42: We use this necessary condition and the subsequent constructive method
43: to investigate the existence of scalar quantum walks on Cayley graphs
44: of various groups presented with two or three generators.
45: In this restricted framework, we classify all groups --
46: in terms of relations between their generators -- that admit
47: scalar quantum walks, and we also derive the form of the most general
48: evolution operator.
49: Finally, we point out some interesting special cases, and extend our study
50: to a few examples of Cayley graphs built with more than three generators.
51: \end{abstract}
52:
53: \maketitle
54:
55: \section{Introduction}
56: Within the traditional (discrete-time) model of quantum computation,
57: there exist several definitions for a quantum walk on a given graph $G$ over a set of vertices $X$.
58: In general terms, a quantum walk simulates the evolution of a particle on the
59: graph with some unitary operator as an evolution operator. The most studied
60: model, the ``coined model'', defines the evolution operator as the product of two operators.
61: One is a block diagonal operator which acts non-trivially only on some internal space of the particle, termed the ``coin''. The other operator is a permutation which moves the particle along the edges of the graph. In that model, the most common definition, which has been used for
62: constructing existing quantum walk algorithms, is the one given by Aharonov \textit{et al.} in \cite{QW1}.
63: Within this definition, the condition for the existence of such operators is that the graph $G$
64: is an oriented $d$-regular graph and that it is possible to ``label each directed edge with a
65: number between one and $d$ such that for each possible label $a$, the directed edges labeled $a$ form a permutation''.
66:
67:
68: Kendon gave another definition \cite{Kendon03}, which
69: is valid for non oriented graphs where the coin is of dimension $|X|$, the
70: number of vertices. In fact both definitions can be seen as particular cases
71: of the definition given by Montanaro in \cite{Montanaro05}. Following
72: the latter work, a necessary and sufficient condition to define a
73: coined quantum walk is {\it reversibility}. When a graph is reversible,
74: it has a certain number of cycles which may
75: be used to define a coined quantum walk. The idea is that each cycle is
76: associated to a basis state of the internal space, using the fact that cycles that do not have a
77: vertex in common can be associated to the same basis state. The
78: definition of Aharonov \textit{et al.} requires the graph to be $d$-regular
79: (and therefore reversible), with exactly $d$ groups of disjoint cycles including all edges of the graph.
80: Kendon's definition also requires a reversible graph, and in this case the cycles used
81: are determined by each directed edge and the corresponding edge with opposite direction.
82: Therefore, reversibility and the number of cycles in the graph may be used to classify the different
83: possible definitions in a two operator (coined-shift) model of quantum walks.
84:
85:
86: If we do not restrict to the coined model (that is, we do not assume that the evolution operator
87: factorizes into two operators), no sufficient condition is known for
88: the existence of a unitary evolution operator on a given graph.
89: Presently, only some necessary conditions are known, one of them being reversibility.
90: In \cite{QW3}, quantum walks with one
91: operator and with internal space (coin) are explored. It is shown how the
92: existence of such walks is related to the properties of the graph and some
93: examples are given. If the model has an internal space of dimension one, the one-operator
94: model corresponds to the definition used in \cite{Severini02} and quantum walks
95: become in this case equivalent to quantum cellular automata as defined by Meyer in \cite{NG}.
96: Here, we will call this model (one operator and a one-dimensional internal space) a
97: \emph{scalar quantum walk}.
98:
99:
100: In \cite{Severini02}, Severini gave a necessary condition
101: for the existence of a scalar quantum walk on a graph, that he called
102: {\it strong quadrangularity}.
103: The strong quadrangularity implies another condition called
104: {\it quadrangularity}. Together with specularity, strong quadrangularity
105: becomes sufficient for the existence of a unitary evolution on the graph. In
106: general terms, a graph is called {\it specular} if any two vertices are connected by outgoing edges
107: to the same set of vertex or do not have any common first neighbor, the same
108: being required for the ingoing edges. Specularity is equivalent to the condition that the operator may be expressed as the product of
109: a permutation matrix and a block diagonal matrix, which means that the evolution operator takes
110: in this case the form of a coined evolution operator
111: on a smaller graph~\cite{Severini02-coined}.
112:
113: In this paper, we will address the general question of determining
114: on which Cayley graphs a homogeneous scalar quantum walk can be defined.
115: First, there is an obvious interest in this question from the point of view
116: of quantum algorithmics. Indeed, the answer will help defining gates over graphs, which could then be combined to construct quantum algorithms.
117: Moreover, algorithms based on scalar quantum walks in the circuit model
118: can also be compared in a more natural way to quantum walks
119: in continuous-time models of quantum computation,
120: such as adiabatic quantum computation.
121: Apart from these algorithmic applications, there are also true conceptual
122: differences between scalar and coined quantum walks, which make them both interesting topics of studies on their own.
123: In the classical case, note that random walks using
124: an internal state space exhibit a memory effect,
125: which makes their classical evolution non-Markovian \cite{CW} in contrast
126: to random walks with no internal space.
127: In the quantum case, the behavior of scalar quantum walks
128: is a more direct consequence of a purely quantum effect,
129: namely phase interference, since adding an
130: internal space necessarily prevents interferences between paths.
131: Finally, we observe that the efficiency of the algorithms using
132: quantum walks varies with
133: the dimension of the internal space without any obvious reason \cite{QW2},
134: so that it seems to be important to investigate both scalar and coined
135: quantum walks to solve an algorithmic problem on a given search space.
136:
137:
138: To identify the Cayley graphs that accept a scalar quantum walk, we will first
139: verify whether some necessary conditions are satisfied for various graphs.
140: First, we know that all Cayley graphs are reversible. The other known
141: conditions are quadrangularity and strong quadrangularity.
142: While the condition of strong quadrangularity depends on the number of subsets
143: of the vertex set of the graph, and thus requires an exponential time to be
144: checked, the quadrangular condition depends on the number of pairs of vertices
145: and may be checked in polynomial time. We will thus use quadrangularity
146: as a necessary condition. This paper is organized as follows. In Section~\ref{conditions}, we will derive the quadrangular condition
147: and adapt it to homogeneous quantum walks on Cayley graphs.
148: In Section~\ref{operator}, we will present a general method
149: allowing us to determine whether a homogeneous scalar quantum walk exists
150: on a given Cayley graph and, if so, to explicitly derive its evolution operator.
151: In Section~\ref{applic}, we will apply the above necessary condition
152: and constructive method in order to find new scalar quantum walks.
153: We will consider a variety of Cayley graphs, built from two, three,
154: and occasionally more than three generators.
155:
156:
157:
158:
159: \section{Conditions on the existence of scalar quantum walks on Cayley graphs\label{conditions}}
160: \subsection{Quadrangularity for general graphs}
161: Let $G(X,E)$ be a graph with vertex set $X$ and edge set $E$. A scalar quantum
162: walk is a model of the evolution of a particle on the graph $G$. The state of the
163: particle is described by a unit vector in the Hilbert space $H=\ell^2(X)$ and we will
164: use the set of states $\{ |x \rangle \}_{x \in X}$ as a basis for $H$.
165: The evolution operator $W$ is a
166: unitary operator on $H$. The evolution of $|\psi_t\rangle$, the state
167: describing the position of the particle at time $t$, is given by the following equation:
168: \begin{equation}
169: \ket{\psi_t} = W^t \ket{\psi_0},
170: \end{equation}
171: where $|\psi_0\rangle$ is a given initial state.
172: We denote as $W_{x,y}$ the matrix elements $\bra x W \ket y$.
173: A necessary condition for the existence of a scalar quantum walk on a graph
174: is called quadrangularity.
175: It was first presented in \cite{Severini02}, and can also be
176: obtained as a consequence of the analysis presented in \cite{QW3}.
177: Let us recall it here.
178:
179: \begin{proposition}\label{quadrangularity-general}
180: Given a graph $G(X,E)$, if a scalar quantum walk on this graph exists then for
181: each pair of different edges of the form $(x,z)$ and $(y,z)$ there exists
182: another pair of different edges of the form $(x,z')$ and $(y,z')$.
183: \end{proposition}
184:
185: \begin{proof}
186: The equation $W^\dagger W=\id_H$, where $W^\dagger$ is the conjugate transpose
187: of $W$ and $\id_H$ is the identity on $H$, is equivalent to the set of equations:
188: \begin{equation}\label{unitary1}
189: \begin{cases} \sum_{z \in X} \overline W_{z,x} W_{z,y}=0 & \forall\ x,y \in X, x \not = y\\
190: \sum_{z \in X} \overline W_{z,x} W_{z,x}=1 & \forall\ x \in X\\
191: \end{cases}
192: \end{equation}
193: And the matrix equation $W W^\dagger=\id_H$ is equivalent to
194: \begin{equation}\label{unitary2}
195: \begin{cases} \sum_{z \in X} W_{x,z} \overline W_{y,z}=0 & \forall\ x,y \in X, x \not = y\\
196: \sum_{z \in X} W_{x,z} \overline W_{x,z}=1 & \forall\ x \in X\\
197: \end{cases}
198: \end{equation}
199: where $\overline \alpha$ denotes the complex conjugate of $\alpha$.
200: Suppose that the sum in the first type of equation contains only one term for some couple $(x,y)$.
201: Then, the only possibility for the condition to be satisfied is to have one of the two coefficients equal to
202: zero. This is equivalent to modifying the graph, so that in this case, there is no scalar quantum
203: walk for the original graph $G$.
204: \end{proof}
205:
206: This proposition can also be reformulated as follows:\\
207: {\it Given a graph $G(X,E)$, if a scalar quantum
208: walk on this graph exists then all consecutive edges in the graph with different
209: orientations belong to at least one closed path of length four with
210: alternating orientations}.
211:
212: \subsection{Quadrangularity for Cayley graphs}
213: A Cayley graph is a graph constructed using the relations of a group.
214: Given a group $\Gamma$ and a set $\Delta$ of elements generating the group (all elements
215: in $\Gamma$ may be obtained by multiplication of elements and inverse elements of
216: the generating set), the Cayley graph $C_{\Delta}(\Gamma)=(X,E)$ is defined by
217: \begin{align}
218: X&=\Gamma\\
219: E&=\{(x,x \delta):x \in \Gamma,\delta \in \Delta \}
220: \end{align}
221:
222: By this definition it will be natural to suppose that the coefficients of the
223: evolution operator depends only on the generator used and not on the starting
224: vertex, which can be interpreted as a homogeneity condition
225: \begin{equation}\label{homogeneity}
226: W_{x \delta,x}=W_{\delta}.
227: \end{equation}
228: In this case, the evolution operator may be written
229: \begin{equation}
230: W=\sum_{\delta\in\Delta} W_\delta U_\delta,
231: \end{equation}
232: where $U_\delta=\sum_x \ket{x\delta}\bra{x}$ is the unitary operator that
233: right-multiplies each group element by $\delta$.
234: The unitarity equations (\ref{unitary1}-\ref{unitary2}) becomes
235: \begin{equation}\label{unitaryCayley}
236: \sum_{\delta_1 \delta_2^{-1}=u} {\overline{W}}_{\delta_1}
237: W_{\delta_2} = \delta_{u=e}\qquad \forall\ \delta_1,\delta_2 \in \Delta,
238: \end{equation}
239: where $\delta_{u=e}=1$ if $u$ equals the identity element $e$ in the group $\Gamma$, and $0$ otherwise.
240:
241: For quantum walks over a Cayley graph,
242: Proposition~\ref{quadrangularity-general} implies a necessary condition over the elements of
243: the generating set $\Delta$:
244: \begin{proposition}\label{quadrangularity-cayley}
245: Given a group $\Gamma$ and a set $\Delta$ of elements generating $\Gamma$,
246: a necessary condition for the existence of a scalar quantum walk of the form
247: $W=\sum_{\delta\in\Delta} W_\delta U_\delta$ on the Cayley graph
248: $C_{\Delta}(\Gamma)$
249: is that for all couples $(\delta_1,\delta_2)$ of
250: different elements of $\Delta$, there exists at least one different couple
251: $(\delta_3,\delta_4)$ such that $\delta_1 \delta_2^{-1}=\delta_3
252: \delta_4^{-1}$.
253: \end{proposition}
254: \begin{proof}
255: From Equation (\ref{unitaryCayley}), we see that when the necessary condition is
256: not satisfied, at least one of the coefficients $W_i$ has to be zero. This is
257: equivalent to modifying the graph (to removing the corresponding edges labeled
258: with~$i$).
259: \end{proof}
260:
261: In principle, it is impossible to solve the general problem of determining
262: and classifying all groups obtained by the procedure of taking the quotient
263: of a free group. This result is a consequence of the existence
264: of finitely presented groups for which the word problem is unsolvable
265: (see Theorem $7.2$ in \cite{LS77}).
266: For this reason, it is probably hopeless to try solving
267: the problem of determining and classifying all groups
268: that verify the quadrangularity condition.
269:
270: Here, we will instead consider groups with a small number of generators
271: and build all possible minimal sets of relations verifying the quadrangularity condition. For the resulting
272: groups, we will completely solve the system of equations~(\ref{unitaryCayley})
273: arising from the unitarity condition, and thus derive
274: the most general evolution operator for a scalar quantum walk
275: on the associated Cayley graphs. The general method to solve this problem is exposed in the next Section.
276:
277:
278: \section{Solving the unitarity condition for the evolution operator\label{operator}}
279:
280: Let us suppose we want to build a scalar quantum walk on a Cayley graph verifying
281: the necessary condition from Proposition~\ref{quadrangularity-cayley}.
282: When the homogeneity condition (\ref{homogeneity}) is imposed,
283: the number of variables describing the evolution operator decreases considerably since it is reduced
284: to $d=|\Delta|$, the number of elements in the generating set $\Delta$.
285: The number of equations in the system (\ref{unitaryCayley}) is also reduced,
286: it is upper bounded by the number of different pairs of generators, that is, $\frac{d(d-1)}{2}$.
287: Each equation in this system may be transformed into a bilinear equation
288: \begin{eqnarray}
289: \bm{v}^\dagger P \bm{v} = 0,\label{bilinear}\\
290: \bm{v}^\dagger \bm{v} = 1,\label{normalization}
291: \end{eqnarray}
292: where $P$ is a $d\times d$ matrix with entries in $\{0,1\}$
293: depending on the graph $C_\Delta(\Gamma)$ and
294: $\bm{v}\in\mathbbm{C}^d$ is a column vector having as components the elements $W_i$ of
295: the evolution operator.
296: More precisely, if we define as
297: \begin{equation}
298: \bm{e}_1=\left( \begin{matrix} 1\\ 0 \\ \vdots \\ 0\end{matrix} \right),
299: \bm{e}_2=\left( \begin{matrix} 0\\ 1 \\ \vdots \\ 0\end{matrix} \right),
300: \cdots,
301: \bm{e}_d=\left( \begin{matrix} 0\\ 0 \\ \vdots \\ 1\end{matrix} \right)
302: \end{equation}
303: the vectors in the standard basis of $\mathbbm{C}^d$, we have
304: $\bm{v}=\sum_{i=1}^d W_i \bm{e}_i$.
305: Let $\{\lambda_1,\dots, \lambda_d\}$ be the $d$ eigenvalues of $P$
306: (with algebraic multiplicities). If the geometric multiplicities of these eigenvalues
307: equal their algebraic multiplicities, we may build an orthonormal basis
308: $\{\bm{v}_1, \dots, \bm{v}_d\}$ out of eigenvectors of $P$.
309: Developing $\bm{v}$ in this basis, we may write
310: $\bm{v} = \sum_{j=1}^d \alpha_j \bm{v}_j$
311: so that Equation (\ref{bilinear}) becomes
312: \begin{equation} \label{linear1}
313: \sum_{j=1}^d \lambda_j | \alpha_j |^2=0,
314: \end{equation}
315: whereas the normalization equation (\ref{normalization}) is simply
316: \begin{equation} \label{linear2}
317: \sum_{j=1}^n | \alpha_j |^2 =1.
318: \end{equation}
319: The relation between the new coefficients $\alpha_j$ and the evolution operator
320: coefficients $W_i$ is then given by
321: \begin{equation}
322: W_i = \sum_j \alpha_j\ \bm{e_i}^\dagger\bm{v_j}.
323: \end{equation}
324:
325: In the next section, we will use this approach to build scalar quantum walks on groups with
326: few generators, or prove that no such walk exists.
327:
328:
329: \section{Application}
330: \label{applic}
331:
332: Let us consider groups denoted as $\Gamma=\langle \Delta|R\rangle$,
333: using the standard free group representation where
334: $\Delta$ is a set of generators for the group $\Gamma$
335: and $R$ is a set of relations among these generators,
336: and determine whether the associated Cayley graph admits a scalar
337: quantum walk.
338:
339: \subsection{Groups finitely presented with two generators\label{two-generators}}
340: The simplest case is a group finitely generated with two generators.
341: We denote the set of generators as $\Delta=\{x,y\}$. There are only two possible couples $(x,y)$
342: and $(y,x)$ and therefore the only relation which makes the necessary condition satisfied is $x y^{-1}=y x^{-1}$. Let us consider the corresponding group
343: $\Gamma=\langle x,y|x y^{-1}=y x^{-1}\rangle$, and check whether there is a scalar
344: quantum walk on the associated Cayley graph $C_\Delta(\Gamma)$.
345: The unitarity equations (\ref{unitaryCayley}) become
346: \begin{equation}
347: \begin{cases} \overline W_x W_y+ \overline W_y W_x=0\\
348: \overline W_x W_x+ \overline W_y W_y=1,
349: \end{cases}
350: \end{equation}
351: which can be written as bilinear equations (\ref{bilinear}-\ref{normalization})
352: with $\bm{v}=\left(
353: \begin{matrix} W_x\\W_y \end{matrix} \right)$
354: and $P=\left(
355: \begin{matrix} 0 & 1\\1 & 0 \end{matrix} \right)$.
356: Developing $\bm{v}$ in the basis formed by the eigenvectors of $P$, that is
357: $\bm{v}_1=\frac{1}{\sqrt{2}}\left(
358: \begin{matrix} 1\\1 \end{matrix} \right)$ and
359: $\bm{v}_2=\frac{1}{\sqrt{2}}\left(
360: \begin{matrix} 1\\-1 \end{matrix} \right)$,
361: with respective eigenvalues $\lambda_1=1$ and $\lambda_2=-1$, we see that
362: Equations~(\ref{linear1}-\ref{linear2}) become
363: \begin{equation}
364: \begin{cases}
365: |\alpha_1|^2 -|\alpha_2|^2=0\\
366: |\alpha_1|^2 + |\alpha_2|^2=1,
367: \end{cases}
368: \end{equation}
369: so that the solution is up to a global phase
370: \begin{equation}\label{sol2gen}
371: \begin{cases}
372: W_x= \cos \phi\\
373: W_y= i \sin \phi
374: \end{cases}
375: \end{equation}
376: for $\phi \in [ 0,2\pi]$.
377: Thus, there exists a scalar quantum
378: walk on the Cayley graph $C_\Delta(\Gamma)$ for the group considered here.
379:
380:
381: Interestingly, this is also true for all groups finitely presented with
382: two generators satisfying $x y^{-1}=y x^{-1}$, since adding other types of relation will not modify the system of equations~(\ref{unitaryCayley}).
383: In particular, we can add some relations to the previous infinite group in order to make
384: it finite. Hence, we show that there is a scalar quantum walk for the
385: dihedral group $D_{n}=\langle x,y|x^{n}=e,y^2=e,(xy)^2=e\rangle$,
386: with generating set $\Delta=\{x,y\}$. The evolution operator coefficients are also given by Eqs.~(\ref{sol2gen}), and the
387: corresponding Cayley graph is both strongly quadrangular and specular
388: (it was first presented in \cite{Severini02}).
389:
390: Finally, let us note that the hypercube of dimension 2 can be obtained as
391: the Cayley graph of the $n=2$ case of $D_{n}$, that is,
392: $D_2=\langle x,y|x^2=e,y^2=e,xy=yx\rangle$, with $\Delta=\{x,y\}$.
393: We conclude that the hypercube of dimension 2
394: therefore also accepts a scalar quantum walk.
395:
396: \subsection{Groups finitely presented with three generators\label{three-generators}}
397: In the case of three generators there are six possible couples, and then a
398: large number of groups, in principle $6^5$.
399: We will use the following representation
400: in order to simplify the counting of the
401: possibles graphs verifying the necessary condition. We associate each pair of
402: generators $(\delta_i,\delta_j)$ to the vertex element $M_{i,j}$ of a $3
403: \times 3$ grid $M$ (see Fig.~\ref{reduction}). A relation $\delta_i\delta_j^{-1}=\delta_k\delta_l^{-1}$
404: may then be represented as an edge between vertices $(\delta_i,\delta_j)$
405: and $(\delta_k,\delta_l)$.
406:
407: \begin{figure}[htb]
408: \begin{center}
409: \includegraphics[height=3cm]{reduction.eps}
410: \end{center}
411: \caption{Representation of the relations between generators as a grid.
412: A relation $\delta_i\delta_j^{-1}=\delta_k\delta_l^{-1}$ will be represented as
413: an edge between vertices $(\delta_i,\delta_j)$ and $(\delta_k,\delta_l)$.\label{reduction}}
414: \end{figure}
415:
416: From the necessary condition, we know that the grids corresponding to a valid graph are those where all vertices are connected with
417: at least another vertex. Note that there is no element in the diagonal, since
418: there are no couples of the form $(\delta_i,\delta_i)$. Moreover, the different edges are not independent, since $\delta_i\delta_j^{-1}=\delta_k\delta_l^{-1}$ is equivalent to
419: $\delta_j\delta_i^{-1}=\delta_l\delta_k^{-1}$
420: (this implies a symmetry with respect to the diagonal axis), while by transitivity the relations
421: $\delta_i\delta_j^{-1}=\delta_k\delta_l^{-1}$ and $\delta_k\delta_l^{-1}=\delta_m\delta_n^{-1}$
422: imply $\delta_i\delta_j^{-1}=\delta_m\delta_n^{-1}$
423: We also see that not all edges are
424: interesting since a vertical or an horizontal edge implies a relation of
425: the form $\delta_i=\delta_j$ and thus a trivial reduction of the number of generators.
426:
427: \begin{figure}[htb]
428: \begin{center}
429: \includegraphics[height=2.5cm]{without_reduction.eps}
430: \end{center}
431: \caption{Representation of the three sets of relations implying no reduction.
432: Valid sets of relations are those for which any vertex is connected to at least one
433: other vertex. Other sets of relations either reduce to one of the three sets represented here
434: (possibly up to a permutation of the generators) or imply a reduction of the number of generators.
435: \label{groups-without-reduction}}
436: \end{figure}
437:
438: By a straightforward (but tedious) inspection of all grids,
439: it can be shown that among the $6^5$ groups finitely presented
440: with three generators, there are only three non-equivalent groups
441: (without trivial reduction) that satisfy the
442: necessary condition (see Fig.~\ref{groups-without-reduction}).
443: There remains to check whether there actually exists a scalar quantum walk
444: on the Cayley graph built from these three groups:
445: \begin{itemize}
446: \item [{\it i})]
447: $\Gamma=\langle x,y,z|xy^{-1}=yz^{-1},xy^{-1}=zx^{-1} \rangle$ and $\Delta=\{x,y,z\}$\\
448: Note that the group $\Gamma$ may equivalently be written as
449: $\langle x,y,z|z=xy^{-1}x,(xy^{-1})^3=e \rangle$.
450: The unitarity condition equations (\ref{unitary1}-\ref{unitary2}) become
451: \begin{equation}
452: \begin{cases} \overline W_x W_y+ \overline W_z W_x +\overline W_y W_z =0\\
453: \overline W_x W_x+ \overline W_y W_y + \overline W_z W_z =1,
454: \end{cases}
455: \end{equation}
456: which can be written as a bilinear hermitian form with $\bm{v}=\left(
457: \begin{matrix} W_x\\W_y\\W_z \end{matrix} \right)$
458: and $P=\left(
459: \begin{matrix} 0 & 1 & 0\\ 0 & 0 & 1\\1 & 0 & 0 \end{matrix} \right)$.
460: Using the expansion of $\bm{v}$ in terms of the eigenvectors of $P$, the system reduces to
461: \begin{equation}
462: \begin{cases}
463: |\alpha_1|^2 + e^{i \frac{2\pi}{3}}|\alpha_2|^2 + e^{-i \frac{2\pi}{3}}|\alpha_3|^2=0\\
464: |\alpha_1|^2 + |\alpha_2|^2 + |\alpha_3|^2=1,
465: \end{cases}
466: \end{equation}
467: so that the solution is (up to a global phase)
468: \begin{equation}\label{sol3gen}
469: \begin{cases}
470: W_x= \frac{1}{3}(1+ e^{i \phi_1} + e^{i \phi_2})\\
471: W_y= \frac{1}{3}(1+ e^{i \frac{2\pi}{3}}e^{i \phi_1} + e^{-i \frac{2\pi}{3}}e^{i \phi_2})\\
472: W_z= \frac{1}{3}(1+ e^{-i \frac{2\pi}{3}}e^{i \phi_1} + e^{i \frac{2\pi}{3}}e^{i \phi_2}).
473: \end{cases}
474: \end{equation}
475: \item [{\it ii})]
476: $\Gamma=\langle x,y,z| xy^{-1}=zx^{-1},yz^{-1}=zy^{-1} \rangle$ and $\Delta=\{x,y,z\}$\\
477: Using a similar method, we find the following general solution:
478: \begin{equation}\label{sol3gen-2}
479: \begin{cases}
480: W_x= \cos\phi \\
481: W_y= \frac{1+i(-1)^q}{2} \sin\phi \\
482: W_z= -\frac{1-i(-1)^q}{2} \sin\phi,
483: \end{cases}
484: \end{equation}
485: for $\phi \in [ 0,2 \pi ]$ and $q$ any integer.
486: \item [{\it iii})]
487: $\Gamma=\langle x,y,z| xy^{-1}=yx^{-1},xz^{-1}=zx^{-1}, yz^{-1}=zy^{-1}
488: \rangle$ and $\Delta=\{x,y,z\}$\\
489: We obtain that there is no solution
490: (the proof is also based on the same method as above).
491: % and we can demonstrate it using the
492: % same method with the three vectors $v_1=\left(
493: % \begin{matrix}W_x\\W_y\end{matrix} \right)$, $v_2=\left(
494: % \begin{matrix}W_x\\W_z\end{matrix} \right) $ and $v_3=\left( \begin{matrix}W_y\\W_z\end{matrix} \right) $.
495: \end{itemize}
496:
497: \begin{figure}[htb]
498: \begin{center}
499: \includegraphics[height=2.5cm]{with_reduction.eps}
500: \end{center}
501: \caption{Representation of the two sets of relations implying
502: reductions of the number of generators. In the first case, we obtain $x=y$, while in the second,
503: we have $x=y=z$. Note that other sets of relations may reduce to these cases, possibly up to
504: some permutation of the generators.\label{groups-with-reduction}}
505: \end{figure}
506:
507: If we add more relations of the type $\delta_i\delta_j^{-1}=\delta_k\delta_l^{-1}$
508: to these groups, reductions appear and we end up with one of the following groups
509: with less than three generators (see Fig.~\ref{groups-with-reduction}):
510: \begin{itemize}
511: \item [{\it iv})]
512: $\Gamma=\langle x,y | x y^{-1}=y x^{-1} \rangle$ and $\Delta=\{x,y\}$\\
513: This group with two generators has been presented in Section~\ref{two-generators}
514: and admits a scalar quantum walk.
515: \item [{\it v})]
516: $\Gamma=\langle x \rangle$ and $\Delta = \{ x \}$\\
517: The free group with one generator admits a trivial scalar quantum walk,
518: which acts as a shift operator
519: on an infinite line. Adding a relation $x^n=e$, we obtain the cyclic group
520: $C_n=\langle x| x^n=e\rangle$, and the associated Cayley graph becomes a cycle,
521: which thus also accepts such a trivial scalar quantum walk.
522: \end{itemize}
523:
524:
525: Of course, there exist other relations beyond these of the type
526: $\delta_i\delta_j^{-1}=\delta_k\delta_l^{-1}$, so that we may build further groups by adding
527: other relations. Nonetheless, as long as these additional relations do not imply relations
528: of the previous type nor trivial reductions, they will have no influence on the
529: possible existence of a scalar quantum walk on the associated Cayley graphs.
530:
531:
532: For instance, adding relations $x^2=e$ and $y^2=e$
533: to the above group ({\it i}),
534: we obtain the symmetric group $S_3=\langle x,y|x^2=e,y^2=e,(xy)^3=e\rangle$ with two generators. This is
535: the group of permutations of three elements, $x$ corresponding
536: to a permutation of the first two elements, $y$ to a permutation of the last two elements, and $xyx$ to a permutation of the first and third
537: elements. Taking $\Delta=\{x,y,xyx\}$, the corresponding Cayley graph therefore admits a scalar quantum walk, with evolution operator coefficients given by Eq.~(\ref{sol3gen}).
538:
539: Similarly, if we add the relation $y^2=x$ to the group ({\it ii}),
540: the set of relations imply $y^3=z$ and $y^4=e$, so that we obtain the cyclic graph
541: $C_4=\langle y|y^4=e\rangle$. Since we use as generators all the group elements
542: except the identity $e$, that is, $\Delta=\{y,y^2,y^3\}$, the Cayley graph reduces to the complete
543: graph over four vertices, which thus admits a scalar quantum walk with evolution operator coefficients given by Eq.~(\ref{sol3gen-2}).
544:
545: Finally, if we impose the additional relations $x^2=y^2=z^2=e$ to the group ({\it iii}) without reduction,
546: the associated Cayley graph becomes the hypercube in dimension $3$. It follows that even though
547: there exists a scalar quantum walk on the hypercube in dimension $2$, as we have seen above, there is no such walk in dimension $3$.
548:
549:
550: \subsection{Other examples}
551: \begin{itemize}
552: \item [{\it vi})]
553: Cayley graphs of the cyclic group $C_n=\langle a| a^n=e\rangle$.\\
554: We have already found a way to define a scalar quantum walk on a cycle, which is the simplest Cayley graph of the cyclic group $C_n$. Here, we show how to build
555: homogeneous scalar quantum walk on more general Cayley graphs
556: of the cyclic group (see also \cite{Severini03}).
557: The elements of $C_n$ may be written as $\{a^i|i=0,\dots,n\}$.
558: The necessary condition implies that if $a^i$ and $a^j$ are taken as generators
559: to build the Cayley graph, we must also take some elements $a^k$ and $a^l$
560: such that $(i-j)-(k-l) \mod n=0$.
561:
562: In particular, if $n$ is a multiple of $d$, this condition is satisfied if the generator set,
563: with $d$ elements, is taken as $\Delta = \{a^{j \frac{n}{d}+l}|j=0,\dots,d-1 \}$,
564: where $l$ is some integer.
565: The unitarity condition equations (\ref{unitary1}-\ref{unitary2})
566: become
567: \begin{equation}
568: \sum_{j=0}^{d-1}\overline W_j W_{(j+i)\!\!\!\mod n} = \delta_{i,0}
569: \end{equation}
570: for $i = 0,\dots,d-1$ and the solution is
571: \begin{equation}
572: W_j=\frac{1}{d} \sum_{k=0}^{d-1}e^{i \theta_k} e^{i \frac{2 \pi k
573: j}{d}}.
574: \end{equation}
575: Hence, there are $d-1$ real parameters that define the evolution operator
576: for this family of Cayley graphs of the cyclic groups. This is not exhaustive for cyclic groups and
577: there are also scalar quantum walks on Cayley graphs
578: outside this family, we have already seen one example, the complete graph
579: over $4$ vertices, and another example will be given below, in the context of Johnson graphs.
580:
581:
582: The $d=n$ special case
583: corresponds to the complete graph with self-loops (the self-loops come from the fact that
584: we also take as generator the identity $e$, which maps any group element to itself).
585: This graph trivially accepts scalar quantum walks since any unitary matrix without zero entries may be taken as a diffusion operator (note that here we give the general form of \emph{homogeneous}
586: scalar quantum walks). In particular, taking $\theta_0=\pi$ and $\theta_k=0$ for all $k\neq 0$
587: yields the usual diffusion operator from Grover's algorithm, with $W_0=1-2/n$ and $W_k=-2/n$ for all $k\neq 0$.
588: The $d=1$ special case reduces to the cycle, that we have already seen
589: above in the context of the free group with one generator.
590:
591: \item [{\it vii})]
592: The Johnson graphs $J(n,k)$.\\
593: The Johnson graph $J(n,k)$ is defined as the graph having as vertices the $C^k_n$ subsets
594: of size $k$ of a set of $n$ elements, and such that two subsets are linked by an edge
595: if they have $k-1$ elements in common. Let us note that $J(n,n-k)=J(n,k)$.
596:
597: All Johnson graphs verify the quadrangularity condition. Numerically, we have
598: obtained that the Johnson graphs $J(n,k)$ for small values of $n$ and $k$ also verify the strong
599: quadrangular condition. Whether they admit a scalar quantum walk in general is an open question.
600: However, our method may help to solve this problem on a case by case basis.
601: Indeed, while until now we have built Cayley graphs from groups, we may take the opposite point of view and try to build a group that
602: accepts a given graph as Cayley graph.
603: For instance, the graph $J(4,2)$ can be described as a Cayley graph of the cyclic group
604: of order six $C_6=\langle\delta_1|\delta_1^6=e\rangle$. Defining $\delta_2=\delta_1^2$,
605: the Cayley graph, represented in Fig.~\ref{johnson}, is then constructed with the generating set $\Delta=\{\delta_1,\delta_1^{-1},\delta_2,\delta_2^{-1}\}$.
606: The equations are thus
607: \begin{equation}
608: \begin{cases}
609: \overline W_{\delta_2} W_{\delta_1^{-1}}+\overline W_{\delta_2^{-1}}
610: W_{\delta_1}+\overline W_{\delta_1} W_{\delta_2^{-1}}+\overline
611: W_{\delta_1^{-1}} W_{\delta_2}=0\\
612: \overline W_{\delta_2} W_{\delta_1}+\overline W_{\delta_1^{-1}} W_{\delta_2^{-1}}=0\\
613: \overline W_{\delta_1} W_{\delta_1^{-1}}+\overline W_{\delta_2^{-1}} W_{\delta_2}=0\\
614: \overline W_{\delta_1} W_{\delta_1}+\overline W_{\delta_1^{-1}}
615: W_{\delta_1^{-1}}+\overline W_{\delta_2} W_{\delta_2}+\overline
616: W_{\delta_2^{-1}} W_{\delta_2^{-1}}=1.
617: \end{cases}
618: \end{equation}
619: Using the previous method, we show that the general solution is
620: \begin{equation}
621: \begin{cases}
622: W_{\delta_1}=\frac{1}{2}e^{i \phi}\\
623: W_{\delta_1^{-1}}=\frac{1}{2}\\
624: W_{\delta_2}=\frac{(-1)^q}{2}\\
625: W_{\delta_2^{-1}}=\frac{(-1)^{q+1}}{2}e^{i \phi}
626: \end{cases}
627: \end{equation}
628: for $\phi \in [ 0,2 \pi ]$ and $q$ any integer.
629:
630:
631: \begin{figure}[htb]
632: \begin{center}
633: \includegraphics[height=3.5cm]{johnson.eps}
634: \end{center}
635: \caption{Johnson graph $J(4,2)$, which may be seen as the Cayley graph of the cyclic group
636: $C_6=\langle\delta_1|\delta_1^6=e\rangle$ with generating set $\Delta=\{\delta_1,\delta_1^{-1},\delta_2,\delta_2^{-1}\}$, where $\delta_2=\delta_1^2$.
637: Note that we have only represented the oriented edges corresponding to $\delta_1$ and
638: $\delta_2$, and not their inverses, corresponding to $\delta_1^{-1}$ and $\delta_2^{-1}$.
639: \label{johnson}}
640: \end{figure}
641:
642:
643: Going beyond $J(4,2)$, we have also tested
644: numerically whether Johnson graphs verify the strong quadrangularity
645: condition. We have obtained that, at least within a
646: computationally achievable range of values $n$ and $k$, they
647: always verify the strong quadrangularity condition,
648: except for the pathological case $J(3,1)$ (and consequently $J(3,2)$)
649: which corresponds to the complete graph over three vertices
650: (without self-loops) and does obviously not admit a scalar quantum walk.
651: Consequently, it may be the case that,
652: in addition to the special case $J(4,2)$,
653: most Johnson graphs with large $n$ and $k$ admit a scalar quantum walk.
654: \end{itemize}
655:
656: \section{Conclusion}
657: We have obtained a simple necessary condition for the existence of a scalar
658: quantum walk on Cayley graphs, as well as a general method
659: to construct its evolution operator (or to conclude that no such walk
660: exists) when the necessary condition is fulfilled.
661: Even if the homogeneity of the evolution
662: operator is required, scalar quantum walks often exist on Cayley graphs,
663: and we have presented a series of examples for finitely presented groups (see Table~\ref{table}).
664:
665: %It may be interesting to determine how the number
666: %of possible sets of relations varies with the increment of the number of
667: %generators.
668:
669: \begin{table}[htb]
670: \begin{center}
671: % use packages: array
672: \begin{tabular}{|c|c|c|c|}
673: \hline
674: Generators & Relations & Existence of & Groups or graphs \\
675: $\Delta$ & $R$ & homogeneous SQW & \\
676: \hline
677: $x$ & $\emptyset$ & yes & Cycle graphs \\
678: $x,y$ & $xy^{-1}=yx^{-1}$ & yes & Dihedral groups $D_n$, \\
679: & & & Hypercube in dim. $2$ \\
680: $x,y,z$ & $xy^{-1}=yz^{-1}=zx^{-1}$ & yes & Symmetric group $S_3$\\
681: $x,y,z$ & $xy^{-1}=zx^{-1},yz^{-1}=zy^{-1}$ & yes & Complete graph over $4$ vertices\\
682: $x,y,z$ & $xy^{-1}=yx^{-1},xz^{-1}=zx^{-1}, yz^{-1}=zy^{-1}$ & no &
683: Hypercube in dim. $3$\\
684: \hline
685: $\ a^{j \frac{n}{d}+k}\ (j=0,\dots,d-1)\ $ & $a^n=e$ & yes & Complete graph with self-loops, \\
686: & & & Cycle graphs \\
687: $a,a^{-1},a^2,a^{-2}$ & $a^6=e$ & yes & Johnson graph $J(4,2)$ \\
688: \hline
689: \end{tabular}
690: \end{center}
691: \caption{Summary of our results.
692: The first two columns define the groups by giving their free
693: group representation $\Gamma=\langle \Delta | R \rangle$,
694: where $\Delta$ is the set of generators and $R$ is the set of relations between these
695: generators. The third column specifies whether there exists a homogeneous scalar quantum walk (SQW) on the associated Cayley graph. The last column lists groups or graphs also concerned by this result.\label{table}}
696: \end{table}
697:
698: We have unfortunately not been able to answer completely
699: the open question of deriving a general necessary and sufficient
700: condition for the existence of a quantum operator on a graph.
701: We have shown, however, that assuming homogeneity significantly
702: simplifies the problem and allows to solve it explicitly
703: for a large class of Cayley graphs. For non Cayley graphs,
704: there is no general procedure to assign equal
705: coefficients to different edges in a non-arbitrary way. However,
706: non-homogeneous quantum walks may also be used to construct quantum
707: algorithms, so that exploring
708: which general graphs admit scalar quantum walks
709: certainly deserves further investigation.
710:
711: \begin{acknowledgments}
712: O.~L.~A. thanks D. Leemans for helpful discussions
713: about the presentation of groups. O.~L.~A. and J.~R. acknowledge financial
714: support from the Belgian National Foundation for Scientific Research (FNRS).
715: This work has also been supported
716: by the IUAP programme of the Belgian government under grant V-18
717: and by the European Commission under the Integrated Project Qubit
718: Applications (QAP) funded by the IST directorate as Contract Number 015848.
719: \end{acknowledgments}
720:
721: \begin{thebibliography}{99}
722: \bibitem{QW1} D. Aharonov, A. Ambainis, J. Kempe and U. Vazirani. Quantum walks
723: on graphs. In {\it Proc. 33th STOC}, New York, NY,
724: 2001. ACM. quant-ph/0012090
725: \bibitem{Kendon03} V. Kendon. Quantum walks on general graphs. quant-ph/0306140 (2003).
726: \bibitem{Montanaro05} A. Montanaro. Quantum walks on directed
727: graphs. quant-ph/0504116 (2005), to appear in Quantum Information and Communication .
728: \bibitem{QW3} O. Lopez Acevedo and T. Gobron. Quantum walks on Cayley graphs,
729: {\it J. Phys. A: Math. Gen} \textbf{39}, 585--599 (2006).
730: \bibitem{Severini02} S. Severini. On the digraph of a unitary matrix.
731: {\it SIAM Journal on Matrix Analysis and Applications} \textbf{25}, 1, 295--300 (2003).
732: \bibitem{NG} D. A. Meyer, On the absence of homogeneous scalar unitary
733: cellular automata. {\it Phys. Lett. A} \textbf{223}, 5, 337--340 (1996).
734: \bibitem{Severini02-coined} S. Severini. The underlying digraph of a coined quantum random walk.
735: quant-ph/0210055 (2002).
736: \bibitem{CW} A. Chen and E. Renshaw. The general correlated random walk. {\it
737: J. Appl. Prob} \textbf{31}, 869--884 (1994).
738: \bibitem{QW2} A. Ambainis, J. Kempe and A. Rivosh. Coins make quantum walks
739: faster. In {\it Proc. SODA}, 2005. quant-ph/0402107
740: \bibitem{LS77} R. C. Lyndon and P. E. Schupp. {\it Combinatorial group
741: theory}. Classics in Mathematics. Springer-Verlag, Berlin, 2001. Reprint of the 1977 edition.
742: \bibitem{Severini03} S. Severini. Graphs of unitary
743: matrices. math.CO/0303084 (2003).
744: \end{thebibliography}
745: \end{document}
746:
747: