0712.2084/ck.tex
1: \documentclass[pra,twocolumn,superscriptaddress,floatfix,showpacs,10pt,letterpaper]{revtex4}
2: \usepackage{graphicx}
3: \usepackage{subfigure}
4: \usepackage{amssymb}
5: \usepackage{verbatim}
6: \usepackage{amsmath}
7: \usepackage{amscd}
8: \usepackage{amssymb}
9: 
10: \DeclareMathOperator{\tr}{tr} \DeclareMathOperator{\supp}{supp}
11: \DeclareMathOperator{\wt}{wt} \DeclareMathOperator{\aut}{aut}
12: 
13: \newcommand{\G}{{\cal G}}
14: \newcommand{\ket}[1]{|{#1}\rangle}
15: \newcommand{\bra}[1]{\langle{#1}|}
16: \newcommand{\X}{\sigma_x}
17: \newcommand{\Z}{\sigma_z}
18: \newcommand{\Y}{\sigma_y}
19: \newcommand{\Xs}[1]{\sigma_{x{#1}}}
20: \newcommand{\Ys}[1]{\sigma_{y{#1}}}
21: \newcommand{\Zs}[1]{\sigma_{z{#1}}}
22: \newcommand{\Xbar}{\overline{X}}
23: \newcommand{\Ybar}{\overline{Y}}
24: \newcommand{\Zbar}{\overline{Z}}
25: \newcommand{\A}{{\cal A}}
26: \newcommand{\ptj}[1]{p_{Tof}^{({#1})}}
27: \newcommand{\pgj}[1]{p_g^{({#1})}}
28: \newcommand{\psj}[1]{p_{stor}^{({#1})}}
29: \newcommand{\ppj}[1]{p_{prep}^{({#1})}}
30: \newcommand{\pmj}[1]{p_{meas}^{({#1})}}
31: \newcommand{\ttj}[1]{t_{Tof}^{({#1})}}
32: \newcommand{\tpj}[1]{t_{prep}^{({#1})}}
33: \newcommand{\tmj}[1]{t_{meas}^{({#1})}}
34: \newcommand{\Sum}{\sum}
35: \newcommand{\low}[1]{\raisebox{-0.3ex}{{#1}}}
36: \newcommand{\im}{{\rm i}}
37: \newcommand{\etal}{\mbox{{\em et al}}}
38: \newcommand{\abs}[1]{|#1|}
39: %\newcommand{\iden}{\bm{1}}% Change this!
40: %\newcommand{\iden}{\mbox$1 \hspace{-1.0mm}  {\bf l}$}}
41: \newcommand{\iden}{1 \hspace{-1.0mm}  {\bf l}}
42: 
43: \newcommand{\ncd}{\newcommand}
44: \ncd{\QC}{$\mbox{QC}_{\cal{C}}\;$}
45: \ncd{\QCpr}{${\mbox{QC}_{\cal{C}}}^\prime\;$}
46: \ncd{\QCns}{$\mbox{QC}_{\cal{C}}$}
47: \ncd{\QCprns}{${\mbox{QC}_{\cal{C}}}^\prime$}
48: \ncd{\cskN}{{|\phi_{\{\kappa\} } \rangle}_{{\cal{C}}_N}}
49: \ncd{\cskNpr}{{|\phi_{\{\kappa^\prime\} } \rangle}_{{\cal{C}}_N}}
50: \ncd{\cskNtil}{{|\phi_{\{\tilde{\kappa} \} } \rangle}_{{\cal{C}}_N}}
51: \ncd{\csk}{{|\phi_{\{\kappa\} } \rangle}_{\cal{C}}}
52: \ncd{\csktil}{{|\phi_{\{\tilde{\kappa} \} } \rangle}_{\cal{C}}}
53: \ncd{\cskf}{|\phi_{\{\kappa\} } \rangle_{\cal{C}}}
54: \ncd{\csktilf}{|\phi_{\{\tilde{\kappa} \} } \rangle_{\cal{C}}}
55: \ncd{\bracsk}{\mbox{}_{\cal{C}}\langle\phi_{\{\kappa\} }|}
56: \ncd{\bracsktil}{\mbox{}_{\cal{C}}\langle\phi_{\{\tilde{\kappa} \}
57: }|} \ncd{\nbracsk}{\mbox{}_{\cal{C}}\langle\phi_{\{\kappa\} }}
58: \ncd{\nbracsktil}{\mbox{}_{\cal{C}}\langle\phi_{\{\tilde{\kappa} \}
59: }} \ncd{\cs}{|\phi \rangle_{\cal{C}}\;} \ncd{\csns}{|\phi
60: \rangle_{\cal{C}}} \ncd{\nbgh}{\text{nbgh}} \ncd{\Sab}{S^{ab}}
61: \ncd{\Sba}{S^{ba}} \ncd{\ds}{\displaystyle} \ncd{\ovl}{\overline}
62: 
63: \newtheorem{conjecture}{Conjecture}
64: \newtheorem{fact}{Fact}
65: \newtheorem{definition}{Definition}
66: \newtheorem{example}{Example}
67: \newtheorem{lemma}{Lemma}
68: \newtheorem{theorem}{Theorem}
69: \newtheorem{corollary}{Corollary}
70: \newtheorem{remark}{Remark}
71: \newtheorem{proposition}{Proposition}
72: \newtheorem{procedure}{Procedure}
73: \newtheorem{scheme}{Scheme}
74: \newtheorem{errmod}{Error model}
75: 
76: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
77: 
78: \begin{document}
79: 
80: \title{Semi-Clifford operations, structure of $\mathcal{C}_k$ hierarchy, and \\ gate complexity for fault-tolerant quantum computation}
81: 
82: \author{Bei Zeng} \affiliation{Department of Physics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA}
83: 
84: \author{Xie Chen} \affiliation{Department of Physics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA}
85: 
86: \author{Isaac L. Chuang} \affiliation{Department of Physics, Massachusetts Institute of Technology, Cambridge, MA 02139, USA}
87: 
88: \date{\today}
89: 
90: \begin{abstract}
91: 
92: Teleportation is a crucial element in fault-tolerant quantum
93: computation and a complete understanding of its capacity is very
94: important for the practical implementation of optimal fault-tolerant
95: architectures. It is known that stabilizer codes support a natural
96: set of gates that can be more easily implemented by teleportation
97: than any other gates. These gates belong to the so called
98: $\mathcal{C}_k$ hierarchy introduced by Gottesman and Chuang (Nature
99: \textbf{402}, 390). Moreover, a subset of $\mathcal{C}_k$ gates,
100: called semi-Clifford operations, can be implemented by an even
101: simpler architecture than the traditional teleportation setup (Phys.
102: Rev. \textbf{A62}, 052316). However, the precise set of gates in
103: $\mathcal{C}_k$ remains unknown, even for a fixed number of qubits
104: $n$, which prevents us from knowing exactly what teleportation is
105: capable of. In this paper we study the structure of $\mathcal{C}_k$
106: in terms of semi-Clifford operations, which send by conjugation at least one maximal 
107: abelian subgroup of the $n$-qubit Pauli group into another one.
108: We show that for $n=1,2$, all
109: the $\mathcal{C}_k$ gates are semi-Clifford, which is also true for
110: $\{n=3,k=3\}$. However, this is no longer true for $\{n>2,k>3\}$. To
111: measure the capability of this teleportation primitive, we introduce
112: a quantity called `teleportation depth', which characterizes how
113: many teleportation steps are necessary, on average, to implement a
114: given gate. We calculate upper bounds for teleportation depth by
115: decomposing gates into both semi-Clifford $\mathcal{C}_k$ gates and
116: those $\mathcal{C}_k$ gates beyond semi-Clifford operations, and
117: compare their efficiency.
118: 
119: \end{abstract}
120: 
121: \pacs{03.67.Pp, 03.67.Lx} \maketitle %\narrowtext
122: 
123: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
124: 
125: \section{Introduction}
126: 
127: The discovery of quantum error-correcting codes and the theory of
128: fault-tolerant quantum computation have greatly improved the
129: long-term prospects for quantum computing technology
130: \cite{Nielsen,Preskill}. To implement fault-tolerant quantum
131: computation for a given quantum error-correcting code, protocols for
132: performing fault-tolerant operations are needed. The basic design
133: principle of a fault-tolerant operation protocol is that if only one
134: component in the procedure fails, then the failure causes at most
135: one error in each encoded block of qubits output from the procedure.
136: 
137: The most straightforward protocol is to use transversal gates
138: whenever possible. A transversal operation has the virtue that an
139: error occurring on the $k$th qubit in a block can only ever
140: propagate to the $k$th qubit of other blocks of the code, no matter
141: what other sequence of gates we perform before a complete
142: error-correction procedure \cite{Shor,Gottesman}. Unfortunately, it
143: is widely believed in the quantum information science community that
144: there does not exist a quantum error correcting code, upon which we
145: can perform universal quantum computations using just transversal
146: gates \cite{Gottesman}, and recently this belief is proved
147: \cite{ZCC}.
148: 
149: We therefore have to resort to other techniques, for instance
150: quantum teleportation \cite{Got} or state distillation
151: \cite{BravyiDistill}. The $\mathcal{C}_k$ hierarchy is introduced by
152: Gottesman and Chuang to implement fault-tolerant quantum computation
153: via teleportation \cite{Got}. The starting point is, if we can
154: perform the Pauli operations and measurements fault-tolerantly, we
155: can then perform all Clifford group operations fault-tolerantly by
156: teleportation. We can then use a similar technique to boot-strap the
157: way to universal fault-tolerant computation, using teleportation,
158: which gives a $\mathcal{C}_k$ hierarchy of quantum teleportation, as
159: defined below:
160: 
161: \begin{definition}
162: The sets $\mathcal{C}_k$ are defined in a recursive way as sets of
163: unitary operations $U$ that satisfy:
164: \begin{equation}
165: \mathcal{C}_{k+1}=\{U|U\mathcal{C}_1U^{\dagger}\subseteq\mathcal{C}_k\},
166: \end{equation}
167: where $\mathcal{C}_1$ is the Pauli group. We call a unitary
168: operation an $n$-qubit $\mathcal{C}_k$ gate if it belongs to the set
169: $\mathcal{C}_k$ and acts nontrivially on at most $n$ qubits.
170: \end{definition}
171: 
172: Note by definition $\mathcal{C}_2$ is the Clifford group, which
173: takes the Pauli group into itself. And
174: $\mathcal{C}_{k}\supset\mathcal{C}_{k-1}$, but $\mathcal{C}_k$ for
175: $k\geq 3$ is no longer a group.
176: 
177: \begin{figure}[htbp]
178: \includegraphics[width=3.00in]{2bit_teleportation.eps}\caption{Two-bit teleportation scheme. $<$ denotes an \textbf{EPR}
179: pair, $B$ represents Bell-basis measurement, $R_{xy}' = U R_{xy}
180: U^{\dagger}$, where $R_{xy}$ is a Pauli operator. The double wires
181: carry classical bits and single wire carries qubits. Any gate in the
182: $\mathcal{C}_k$ hierarchy can be implemented fault-tolerantly using
183: this teleporation scheme.}\label{2bit}
184: \end{figure}
185: 
186: \begin{figure}[htbp]
187: \includegraphics[width=3.00in]{1bit_teleportation.eps}
188: \caption{One-bit teleportation scheme. For $Z$-teleportation, $A=I$,
189: $B=H$, $D=Z$, and $E$ is a CNOT gate with the first qubit as its
190: target. For $X$-teleportation, $A=H$, $B=I$, $D=X$, and $E$ is a
191: CNOT gate with the first qubit as its control. All semi-Clifford
192: $\mathcal{C}_k$ gates can be implemented fault-tolerantly using this
193: scheme.}\label{1bit}
194: \end{figure}
195: 
196: 
197: All the gates in $\mathcal{C}_k$ can be performed with the two-bit
198: teleportation scheme (FIG. \ref{2bit}) in a fault-tolerant manner.
199: Because, as proved in \cite{Gottesman}, it is possible to
200: fault-tolerantly prepare the ancilla state $|\Psi_U^n\rangle$, apply the
201: classically controlled correction operation $R_{xy}^{'\dagger}$, and
202: measure in Bell basis on a stabilizer code. However the precise set
203: of gates which form $\mathcal{C}_k$ is unknown, even for a fixed
204: number of qubits. It is demonstrated in \cite{Xinlan} that a subset
205: of $\mathcal{C}_k$ gates could be implemented by a different
206: architecture than the standard teleportation, called one-bit
207: teleportation, as shown in FIG. \ref{1bit}. Those gates adopt the
208: form $L_1VL_2$, where $V$ is a diagonal gate in $\mathcal{C}_k$ and
209: $L_1,L_2$ are two Clifford operations. Gates of this form are
210: recently studied in literature and are called the semi-Clifford
211: operations \cite{Gross}. In the following we will denote the
212: $n$-qubit Pauli group as $\mathcal{P}_n$ and a semi-Clifford
213: operation is defined to be a gate which sends at least one maximal
214: abelian subgroup of $\mathcal{P}_n$ to another maximal abelian one
215: under conjugation.
216: 
217: Due to the fact that one-bit teleportation needs only half the
218: number of ancilla qubits per teleportation than the standard two-bit
219: teleportation, it is important to understand the difference of
220: capabilities between one and two-bit teleportation for the practical
221: implementations of fault-tolerant architecture. It is conjectured in
222: \cite{Xinlan} that those two capabilities coincide for
223: $\{n=2,k=3\}$, which means that all the $\mathcal{C}_3$ gates for
224: two qubits are semi-Clifford operations.
225: 
226: In this paper, we prove this conjecture for a more general situation
227: where $\{n=1,2,\forall k\}$, and $\{n=3,k=3\}$. We then disprove it
228: for parameters $\{n>2,k>3\}$ by explicit construction of
229: counterexamples. We leave open the question for the parameters
230: $\{n>2,k=3\}$, and a more general problem of fully characterizing
231: the structure of $\mathcal{C}_k$: we conjecture that all gates in
232: $\mathcal{C}_k$ are something we refer to as generalized
233: semi-Clifford operations, i.e. a natural generalization of the
234: concept of semi-Clifford operation to the case including classical
235: permutations. Our results about this semi-Clifford operations versus
236: $\mathcal{C}_k$ gates relation can be visualized in FIG.
237: \ref{Cksemi}.
238: 
239: \begin{figure}[htbp]
240: \includegraphics[width=3.00in]{gates.eps}
241: \caption{Semi-Clifford operations versus $\mathcal{C}_k$ gates. A:
242: all gates; B: generalized semi-Clifford gates; C: semi-Clifford
243: gates; D: $\mathcal{C}_k$ gates; E: $\mathcal{C}_3$ gates. C is
244: strictly contained in B and E is strictly contained in D. The two
245: question marks indicate two open problems we have: whether $D$ is a
246: subset of $B$; and whether $E$ is a subset of $C$.} \label{Cksemi}
247: \end{figure}
248: 
249: Just as in the usual circuit model, different gates are implemented
250: with different levels of complexity using this teleportation scheme.
251: It is then natural to ask the questions, how to characterize this
252: concept of gate complexity with concrete physical quantities, how
253: does this measure based on teleportation schemes compare with the
254: usual circuit depth, and what it implies for the practical
255: construction of quantum computation architecture. To answer these
256: questions, we introduce a quantity as a measure of gate complexity
257: for fault-tolerant quantum computation based on the $\mathcal{C}_k$
258: hierarchy, called the teleportation depth, which characterizes how
259: many teleportation steps are necessary, on average, to implement a
260: given gate. We demonstrate the effect of the existence of non
261: semi-Clifford operations in $\mathcal{C}_k$ on the estimation of the
262: upper bound for the teleportation depth, as well as some
263: quantitative difference between the capabilities of one and two-bit
264: teleportation.
265: 
266: The paper is organized as follows: Section II gives definition and
267: basic properties of semi-Clifford operations and generalized
268: semi-Clifford operations; in Section III we study the structure of
269: $\mathcal{C}_k$ hierarchy in terms of semi-Clifford and generalized
270: semi-Clifford operations; Section IV is devoted to the discussion of
271: teleportation depth and how it depends on the structure of
272: $\mathcal{C}_k$; and with Section V, we conclude our paper.
273: 
274: \section{Semi-Clifford operations and its generalization}
275: 
276: The concept of semi-Clifford operations was first introduced in
277: \cite{Gross}, to characterize the property of gates transforming
278: Pauli matrices acting on a single qubit. Here we generalize it to
279: the $n$-qubit case, through the following definition.
280: 
281: \begin{definition}
282: An $n$-qubit unitary operation is called semi-Clifford if it sends
283: by conjugation at least one maximal abelian subgroup of
284: $\mathcal{P}_n$ to another maximal abelian subgroup of
285: $\mathcal{P}_n$.
286: 
287: That is, if $U$ is an $n$-qubit semi-Clifford operation, then there
288: must exist at least one maximal abelian subgroup $G$ of
289: $\mathcal{P}_n$, such that $UGU^{\dagger}$ is another maximal
290: abelian subgroup of $\mathcal{P}_n$.
291: \end{definition}
292: 
293: The most basic property of a semi-Clifford operation is,
294: \begin{proposition}
295: If $R$ is a semi-Clifford operation, then there exist Clifford
296: operations $L_1,L_2$ such that $L_1RL_2$ is diagonal.
297: \end{proposition}
298: \textbf{Proof}: $Z_i$ represents the Pauli $Z$ operation on the
299: $i^{th}$ qubit. If  $R$ is an $n$-qubit semi-Clifford operation,
300: then there must exist $n$-qubit operations $L_1,L_2\in\mathcal{C}_2$
301: such that $RL_2Z_iL_2^{\dagger}R^{\dagger}=L_1^{\dagger}Z_iL_1$,
302: \cite{Gotthesis}, i.e.
303: $L_1RL_2Z_iL_2^{\dagger}R^{\dagger}L_1^{\dagger}=Z_i$ holds for any
304: $i=1...n$. Therefore, $(L_1RL_2)Z_i=Z_i(L_1RL_2)$, i.e. the
305: $n$-qubit gate $L_1RL_2$ is diagonal.$\square$
306: 
307: In other words, semi-Clifford operations are those gates
308: diagonalizable `up to Clifford multiplications'. Thus the structure
309: problem of the whole set of semi-Clifford operations is reduced to
310: that of the diagonal subset within it.
311: 
312: As we shall see later, the notion of semi-Clifford operations is
313: useful in characterizing some but not all gates in the
314: $\mathcal{C}_k$ hierarchy. More generally, we might also consider
315: those gates with properties of transforming the span, or in other
316: words the group algebra over the complex field, of a maximal abelian
317: subgroup of $\mathcal{P}_n$.
318: 
319: \begin{definition}
320: A generalized semi-Clifford operation on $n$ qubits is defined to
321: send by conjugation the span of at least one maximal abelian
322: subgroup of $\mathcal{P}_n$ to the span of another maximal abelian
323: subgroup of $\mathcal{P}_n$.
324: \end{definition}
325: 
326: Denote $\langle S_i\rangle$ the group generated by a set of operators $\{S_i\}$, and denote 
327: the span of the group $\langle S_i\rangle$ as
328: $\mathfrak{C}(\langle S_i\rangle)$. Then in a more mathmatical form we can write
329: the above definition as:
330: 
331: If $U$ is a generalized semi-Clifford operation on $n$ qubits, then
332: there must exist at least one maximal abelian subgroup $G=\langle g_i\rangle$ of
333: $\mathcal{P}_n$, such that for all $s\in \mathfrak{C}(\langle g_i\rangle)$, 
334: $UsU^{\dagger}\in\mathfrak{C}(U\langle g_i\rangle U^{\dagger})$, where $UGU^{\dagger}$ is another maximal abelian subgroup of $\mathcal{P}_n$.
335: 
336: Then the basic property of a generalized semi-Clifford operation is,
337: \begin{proposition}
338: If $R$ is a generalized semi-Clifford operation, then there exist
339: Clifford operations $L_1,L_2$, and a classical permutation operator
340: $P$ such that $PL_1RL_2$ is diagonal.
341: \end{proposition}
342: \textbf{Proof}: If  $R$ is a generalized semi-Clifford operation,
343: then there must exist $L_1,L_2\in\mathcal{C}_2$ such that
344: $RL_2\mathfrak{C}(\langle
345: Z_i\rangle_{i=1}^n)L_2^{\dagger}R^{\dagger}=L_1^{\dagger}\mathfrak{C}(\langle
346: Z_i\rangle_{i=1}^n)L_1$, i.e. $L_1RL_2\mathfrak{C}(\langle
347: Z_i\rangle_{i=1}^n)L_2^{\dagger}R^{\dagger}L_1^{\dagger}=\mathfrak{C}(\langle
348: Z_i\rangle_{i=1}^n)$. That is, $L_1RL_2$ maps all the diagonal
349: matrices to diagonal matrices, therefore $L_1RL_2$ must be a
350: monomial matrices, i.e. there exist a permutation matrix $P$ and a
351: diagonal matrix $V$, such that $L_1RL_2=P^{\dagger}V\Rightarrow
352: PL_1RL_2$ is diagonal.$\square$
353: 
354: Note for the single qubit case, i.e. $n=1$, the concepts of
355: semi-Clifford operation and generalized semi-Clifford operation
356: coincide.
357: 
358: \section{The structure of $\mathcal{C}_k$}
359: 
360: In this section we study the structure of gates in $\mathcal{C}_k$.
361: To begin with, we study some basic properties of $\mathcal{C}_k$
362: gates. Then we give our main results as structure theorems, which
363: state that all the $\mathcal{C}_k$ gates are semi-Clifford when
364: $\{n=1,2,\forall k\}$ and $\{n=3,k=3\}$, but for $\{n>2,k>3\}$ there
365: are examples of $\mathcal{C}_k$ gates which are non-semi-Clifford.
366: We then discuss the open question for the parameters $\{n>2,k=3\}$,
367: and based on the constructed counterexamples we conjecture that all
368: $\mathcal{C}_k$ gates are generalized semi-Clifford operations.
369: 
370: It should be noted that the set of $n$-qubit $\mathcal{C}_k$ gates
371: is always strictly contained in the set of $n$-qubit
372: $\mathcal{C}_{k+1}$ gates. In \cite{Got}, explicit examples are
373: given to support this statement. If we denote as $\Lambda_{n-1}(U)$
374: the $n$-qubit gate which applies $U$ to the $n$th qubit only if the
375: first $n-1$ qubits are all in the state $\ket{1}$, then
376: $\Lambda_{n-1}(\text{diag}(1,e^{2\pi/2^m}))$ is in
377: $\mathcal{C}_{m+n-1}\setminus\mathcal{C}_{m+n-2}$.
378: 
379: \subsection{Basic properties}
380: 
381: We first state an important property of gates in $\mathcal{C}_k$,
382: which reduce the problem of characterizing the structure of
383: $\mathcal{C}_k$ into a problem of characterizing a certain subset of
384: gates in $\mathcal{C}_k$.
385: 
386: \begin{proposition}
387: If $R\in \mathcal{C}_k$, then $L_1RL_2\in \mathcal{C}_k$, where
388: $L_1,L_2\in \mathcal{C}_2$, $k\geq 2$.
389: \end{proposition}
390: \textbf{Proof}: We prove this proposition by induction.
391: 
392: i) It is obviously true for $k=2$;
393: 
394: ii) Assume it is true for $k$;
395: 
396: iii) For $k+1$, $R\in \mathcal{C}_{k+1}$ implies $RAR^{\dagger}\in
397: \mathcal{C}_{k}$, where $A\in \mathcal{C}_{1}$. If we conjugate $A$
398: by $L_1RL_2$, we get
399: \begin{equation}
400: L_1RL_2A(L_1RL_2)^{\dagger}=L_1R(L_2AL_2^{\dagger})R^{\dagger}L_1^{\dagger}.
401: \end{equation}
402: 
403: Since $L_1,L_2\in \mathcal{C}_2$, $L_1^{\dagger},L_2^{\dagger}$ are
404: in $\mathcal{C}_2$ also. And because $L_2AL_2^{\dagger} \in
405: \mathcal{C}_1$, $R(L_2AL_2^{\dagger})R^{\dagger} \in \mathcal{C}_k$.
406: According to assumption ii),
407: $L_1R(L_2AL_2^{\dagger})R^{\dagger}L_1^{\dagger} \in \mathcal{C}_k$.
408: Finally as we can see from Eqn (2), $L_1RL_2 \in
409: \mathcal{C}_{k+1}$.$\square$
410: 
411: According to Proposition 3, in order to characterize the full
412: structure of $\mathcal{C}_k$, we only need to characterize the
413: structure of a subset of it which generates the whole set with
414: Clifford multiplications.
415: 
416: It is known that $\mathcal{C}_k$ is not a group for $k>2$ and its
417: structure is in general hard to characterize. However, if we denote
418: all the diagonal gates in $\mathcal{C}_k$ as $\mathcal{F}_k$, then
419: we have the following:
420: \begin{proposition}
421: $\mathcal{F}_{k}$ is a group.
422: \end{proposition}
423: 
424: If we can characterize the group structure of $\mathcal{F}_k$, then
425: the structure of the $\mathcal{C}_k$ subset $\{L_1F_kL_2\}$ is known
426: to us ($L_1, L_2 \in \mathcal{C}_2$, $F_k \in \mathcal{F}_k$).
427: According to Proposition 1, this is just the set of all
428: semi-Clifford operations in $\mathcal{C}_k$. In the next section, we
429: will repeatedly use this fact to gain knowledge about semi-Clifford
430: $\mathcal{C}_k$ gates from the group structure of $\mathcal{F}_k$
431: and for now we will give a brief proof of the above proposition.
432: 
433: \textbf{Proof}: We prove by induction.
434: 
435: i) It is of course true for $k=2$;
436: 
437: ii) Assume it is true for $k$, i.e. $\mathcal{F}_{k}$ is group;
438: 
439: iii) Then for $k+1$, note for any $F_{k+1} \in\mathcal{F}_{k+1}$,
440: $F_{k+1}MF_{k+1}^{\dagger}=F_{k}M=MF'_{k}$, for non-diagonal
441: $M\in\mathcal{C}_1$, where $F_{k}, F'_{k}\in \mathcal{F}_{k}$.
442: 
443: a) If $F_{k+1} \in\mathcal{F}_{k+1}$, then $F_{k+1}^{\dagger}
444: \in\mathcal{F}_{k+1}$, since
445: $F_{k+1}^{\dagger}MF_{k+1}=F_{k}^{\dagger}M=MF^{'\dagger}_{k}$,
446: which is in $\mathcal{F}_{k}$ by assumption ii).
447: 
448: b) If $F_{1k},F_{2k}\in\mathcal{F}_k$, then
449: $F_{1k}F_{2k}\in\mathcal{F}_k$, since
450: $F_{1k-1}F_{2k-1}\in\mathcal{F}_{k-1}$.$\square$
451: 
452: According to this proposition, all semi-Clifford $\mathcal{C}_k$
453: gates can be characterized by the group structure of diagonal
454: $\mathcal{C}_k$ gates.
455: 
456: \subsection{Structure theorems}
457: 
458: Our main results about the structure of $\mathcal{C}_k$ are the
459: following three theorems, which state that all the $\mathcal{C}_k$
460: gates are semi-Clifford when $\{n=1,2,\forall k\}$ and
461: $\{n=3,k=3\}$, but it is no longer true for $\{n>2,k>3\}$.
462: 
463: \begin{theorem}
464: All gates in $\mathcal{C}_k$ are semi-Clifford operations for
465: $(n=1,2,\forall k)$.
466: \end{theorem}
467: \textbf{Proof}: Here we prove the case of $n=2$. The proof of the
468: $n=1$ case is similar but can also be checked by direct calculation
469: and lead to a complete classification of all 1-qubit $\mathcal{C}_k$
470: gates according to the group structure of diagonal 1-qubit
471: $\mathcal{C}_k$ gates. We give details for the $n=1$ case in
472: appendix.
473: 
474: For $n=2$, we prove this theorem by induction:
475: 
476: i) It is obviously true for $k=1,2$;
477: 
478: ii) Assume it is true for $k$;
479: 
480: iii) For $k+1$:
481: 
482: a) We calculate the set $S_1=\{L_1V\}$ for all
483: $L_1\in\mathcal{C}_2$, where $V\in\mathcal{F}_k$. Note by assumption
484: ii), $S_1$ gives us all the elements in $\mathcal{C}_k$ up to
485: Clifford conjugation.
486: 
487: b) Note in general
488: $V=\text{diag}\{e^{i\alpha},e^{i\beta},e^{i\gamma},e^{i\delta}\}$
489: for some angles $\alpha,\beta,\gamma$ and $\delta$. By exhaustive
490: calculation with all $L_1\in\mathcal{C}_2$ we show that if there
491: exists an element $V_s\in S_1$ such that $V_s$ is trace zero and
492: Hermitian, then
493: $V=\text{diag}\{e^{-i\theta_1},e^{-i\theta_2},e^{i\theta_2},e^{i\theta_1}\}$
494: for some $\theta_1$ and $\theta_2$. Furthermore, we can again show
495: by  exhaustive calculation with all $L_1\in\mathcal{C}_2$ that the
496: only trace zero and Hermitian $V_s\in S_1$ is of the following form
497: up to Clifford conjugation:
498: \begin{equation}
499: V_s=\left(\begin{matrix}
500: 0 & 0 & 0 &e^{-i\theta_1} \\
501: 0 & 0 &e^{-i\theta_2} & 0\\
502: 0 & e^{i\theta_2} & 0 & 0\\
503: e^{i\theta_1} & 0 & 0 & 0
504: \end{matrix}\right).
505: \end{equation}
506: 
507: c) We calculate the set $S_2=\{L_1V_sL_1^{\dagger}\}$ for all
508: $L_1\in\mathcal{C}_2$, which by assumption ii) and fact b) gives all
509: the elements in $\mathcal{C}_k$ which are trace zero and Hermitian.
510: 
511: d) We show that for any two-qubit gate $U$ such that
512: $UV_sU^{\dagger}=Z_1$ and $\{U\mathcal{P}_2U^{\dagger}\}\subseteq
513: S_2$, there exist $L_1,L_2\in\mathcal{C}_2$ such that $L_1UL_2$ is
514: diagonal. This can be started from studying the eigenvectors of
515: $V_s$, which can be chosen of the form
516: 
517: \begin{equation}
518: U=\frac{1}{\sqrt{2}}\left(\begin{matrix}
519: 1 & 0 & 0 &1 \\
520: 0 & 1 &1 & 0\\
521: 0 & e^{i\theta_2} & -e^{i\theta_2} & 0\\
522: e^{i\theta_1} & 0 & 0 & -e^{i\theta_1}
523: \end{matrix}\right),
524: \end{equation}
525: and carefully considering the possible phase of each eigenvector and
526: the possible superposition of the eigenvectors due to the degeneracy
527: of the eigenvalues, similar to the process shown in Appendix
528: A.$\square$
529: 
530: \begin{theorem}
531: All gates in $\mathcal{C}_k$ are semi-Clifford operations for
532: $\{n=3,k=3\}$.
533: \end{theorem}
534: \textbf{Proof}: We prove this theorem exhaustively using the
535: following proposition:
536: \begin{proposition}
537: An $n$-qubit $\mathcal{C}_k$ gate $U$ is semi-Clifford if and only
538: if the group $\{U\mathcal{P}_nU^{\dagger}\}\cap\mathcal{P}_n$
539: contains a maximally abelian subgroup of $\mathcal{P}_n$.
540: \end{proposition}
541: \textbf{Proof}: Suppose $U=L_1VL_2$, then
542: $U\mathcal{P}_nU^{\dagger}=L_1VL_2\mathcal{P}_nL_1^{\dagger}V^{\dagger}L_1^{\dagger}=L_1V\mathcal{P}_nV^{\dagger}L_1^{\dagger}\supset\{L_1Z_iL_1^{\dagger}\}_{i=1}^n$.
543: 
544: On the contrary, if $\{U\mathcal{P}_nU^{\dagger}\}\cap\mathcal{P}_n$
545: contains a maximal abelian subgroup of $\mathcal{P}_n$, then there
546: must exist $L_1,L_2 \in \mathcal{C}_2$ such that
547: $UL_1^{\dagger}Z_iL_1U^{\dagger}=L_2Z_iL_2^{\dagger}$, i.e.
548: $L_2^{\dagger}UL_1^{\dagger}Z_iL_1U^{\dagger}L_2=Z_i$ holds for any
549: $i=1...n$. Therefore,
550: $(L_2^{\dagger}UL_1^{\dagger})Z_i=Z_i(L_2^{\dagger}UL_1^{\dagger})$,
551: $\Rightarrow L_2^{\dagger}UL_1^{\dagger}$ is diagonal. If we denote
552: this diagonal gate as $V$, $L_2^{\dagger}UL_1^{\dagger}=V\Rightarrow
553: U=L_1VL_2$.
554: 
555: Therefore, by exhaustive study with the subgroups of the three-qubit
556: Clifford group which are isomorphic to $\mathcal{P}_3$, we complete
557: the proof of this theorem. More detailed analysis about this is
558: given in Appendix B. The calculation is done using GAP
559: \cite{GAP}.$\square$
560: 
561: \begin{theorem}
562: Not all gates in $\mathcal{C}_k$ are semi-Clifford operations for
563: $(n>2, k>3)$.
564: \end{theorem}
565: \textbf{Proof}: Actually we only need to prove this theorem for
566: $n=3,k=4$ then it naturally holds for all the other parameters of
567: $\{n>2,k>3\}$. However we would like to explicitly construct
568: examples for all $\{n=3,k>4\}$. Define $W_{k}$ as in FIG. \ref{fig4}.
569: 
570: \begin{figure}[htbp]
571: \includegraphics[width=2.00in]{Wk.eps}
572: \caption{A non-Clifford-diagonalizable $\mathcal{C}_k$ gate $W_k$.
573: $V_{k}=diag(1,e^{i\pi/2^{k-1}})$.}\label{fig4}
574: \end{figure}
575: 
576: 
577: \begin{proposition}
578: The gate
579: \begin{equation}
580: W_k=T(c_1, c_2, t_3) \otimes V_{3,k}
581: \end{equation}
582: is a $\mathcal{C}_{k+1}$ operation but not a semi-Clifford
583: operation, where $T(c_1, c_2, t_3)$ is Toffoli gate with the 1st and
584: 2nd qubit as its control and the 3rd qubit as its target, $V_{3,k}$
585: is single qubit operator $diag(1,exp(i\pi/2^{k-1}))$ on the 3rd
586: qubit.
587: \end{proposition}
588: \textbf{Proof}:
589: 
590: To prove that $W_k$ is in $\mathcal{C}_{k+1}$,
591: 
592: i) When $k=2$, $V_{k} = diag\{1, i\} \in \mathcal{C}_2$. $W_2$ is of
593: the form $LR$, where $L$ is a Clifford operation and $R$ is the
594: Toffoli gate. According to Proposition 3, $W_2$ and the Toffoli gate
595: are both in $\mathcal{C}_3$.
596: 
597: ii) For $k > 2$, direct calculation shows that
598: $\{W_kZ_iW_k^{\dagger}\} \subset \mathcal{C}_2$, $i = 1, 2, 3$.
599: $W_kX_1W_k^{\dagger} \in \mathcal{C}_k$, $W_kX_2W_k^{\dagger} \in
600: \mathcal{C}_k$, $W_kX_3W_k^{\dagger} \in \mathcal{C}_{k-1}$. The
601: images of $X_i$'s under the conjugation of $W_k$ can all be written
602: in the form $W_kX_iW_k^{\dagger} = X_i F_{ki} = F'_{ki} X_i$, where
603: $F_{k1}$, $F'_{k1}$, $F_{k2}$, $F'_{k2}$ are diagonal gates in
604: $\mathcal{C}_k$ and $F_{k3}$, $F'_{k3}$ are diagonal single qubit
605: gates in $\mathcal{C}_{k-1}$ acting on the third qubit.
606: 
607: The image of the whole 3-qubit Pauli group $\{W_k\mathcal{P}_3
608: W_k^{\dagger}\}$ is generated by the six elements shown above. As
609: multiplication by Clifford gates preserves the $\mathcal{C}_k$
610: hierarchy, we only need to check the images of Pauli operations
611: which are composed of two or more $X_i$'s and see if their images
612: are still in $\mathcal{C}_k$.
613: 
614: This is obviously true considering the special form of
615: $\{W_kX_iW_k^{\dagger}\}$. Multiplication of any two of them is of
616: the form $W_kX_iX_jW_k^{\dagger} = X_i F_{ki} F'_{kj} X_j$. This is
617: in $\mathcal{C}_k$ as the diagonal $\mathcal{C}_k$ gates form a
618: group. Further more, multiplication of all of them takes the form
619: $W_kX_1X_2X_3W_k^{\dagger} = X_1 F_{ki} F'_{kj} X_2 F'_{k3} X_3$. As
620: $F'_{k3}$ is a single qubit operation on the third qubit,
621: $W_kX_1X_2X_3W_k^{\dagger} =X_1 F_{ki} F'_{kj} F'_{k3} X_2 X_3$.
622: This is again a $\mathcal{C}_k$ gate because of the group structure
623: of diagonal $\mathcal{C}_k$ gates.
624: 
625: Therefore, we have checked explicitly that $W_k \in
626: \mathcal{C}_{k+1}$.
627: 
628: To prove that $W_k$ is not semi-Clifford, we can exhaustively
629: calcultate $\{W_k\mathcal{P}_3W_k^{\dagger}\}$ and find its
630: intersection with $\mathcal{P}_3$. The fact that
631: $\{W_k\mathcal{P}_3W_k^{\dagger}\} \cap \mathcal{P}_3$ does not
632: contain a maximally abelian subgroup of $\mathcal{P}_3$ implies that
633: $W_k$ is not semi-Clifford, due to Proposition 5.
634: 
635: With this example we have directly proved Theorem 3. $\square$
636: 
637: 
638: \subsection{Open problems}
639: 
640: Let us try to understand more about the structure theorems we have
641: in the previous section.
642: 
643: First recall from \cite{Xinlan} that the controlled-Hadamard gate
644: $\Lambda_1(H)$, which is a $\mathcal{C}_3$ gate, is explicitly shown
645: to be semi-Clifford. We can also view this from the perspective of
646: Proposition 5, by noting that
647: $\Lambda_1(H)Z_1\Lambda_1(H)^{\dagger}=Z_1$,
648: $\Lambda_1(H)Y_2\Lambda_1(H)^{\dagger}=Z_1\otimes -Y_2$, which means
649: that the maximal abelian subgroup of the Pauli group generated by
650: $\langle Z_1,Y_2\rangle\times\langle\pm 1,\pm i\rangle$ is in the
651: image of $\Lambda_1(H)$. However, if we consider $W_3$ from the
652: perspective of Proposition 5, we get $W_3Z_1W_3^{\dagger}=Z_1$,
653: $W_3Z_2W_3^{\dagger}=Z_2$,
654: $W_3Z_3W_3^{\dagger}=\Lambda_1(Z_2)\otimes Z_3$. Note this do not
655: give us a maximal abelian subgroup of the Pauli group
656: $\langle Z_1,Z_2,Z_3\rangle\times\langle\pm 1,\pm i\rangle$, 
657: due to the effect of $\Lambda_1(Z_2)$ caused by
658: conjugating through the Toffoli gate. This intuitively explains why
659: Theorem 3 could be true, but no counterexample to Theorem 2 exists.
660: 
661: Note that $W_k$ is actually a generalized semi-Clifford operation,
662: which is apparent from its form. Also, the construction of the
663: series of gates $W_k$, as well as their extensions to $n>3$ qubits,
664: cannot give any non-semi-Clifford $\mathcal{C}_3$ gate. We then have
665: the following conjectures on the open problem of the structure of
666: $\mathcal{C}_k$ hierarchy in general.
667: 
668: \begin{conjecture}
669: All gates in $\mathcal{C}_3$ are semi-Clifford operations.
670: \end{conjecture}
671: 
672: \begin{conjecture}
673: All gates in $\mathcal{C}_k$ are generalized semi-Clifford operations.
674: \end{conjecture}
675: 
676: \section{The teleportation depth}
677: 
678: Teleportation, as a computational primitive, is a crucial element
679: providing universal quantum computation to fault-tolerant schemes
680: based on stabilizer codes. However, not all gates are of equal
681: complexity in this scheme. To actually incorporate this technique in
682: the construction of practical computational architecture, it is
683: useful to know which gates are easier to implement and which are
684: harder, so that we could achieve optimal efficiency in performing a
685: computational task. In the circuit model of quantum computation, we
686: face the same problem and in that case `circuit depth'
687: was introduced \cite{Yao} to characterize the number of simple one
688: and two-qubit gates needed to implement an operation. While this
689: provides a good measure of gate complexity, it does not take into
690: consideration of fault-tolerance. It is interesting to have measures
691: quantifying fault-tolerant gate complexity to be compared with
692: `circuit depth' to give us a better understanding of the
693: computational tasks at hand.
694: 
695: Based on the $\mathcal{C}_k$ hierarchy introduced in \cite{Got} and
696: the knowledge of its structure gained in previous section, we define
697: a measure of gate complexity for the teleportation protocol, called
698: the teleportation depth, which characterizes how many teleportation
699: steps are necessary, on average, to implement a given gate. Since
700: any teleportation unavoidably causes randomness, we need to figure
701: out a certain point to start with, i.e. we should assume in advance
702: that some kind of gates can be performed fault-tolerantly. We know
703: that a fault-tolerant protocol is usually associated with some
704: quantum error-correcting codes. Self-dual CSS codes, such as the
705: $7$-qubit Steane code, admit all gates in the Clifford group to be
706: transversal \cite{Gotthesis}. In such a situation, we only need to
707: teleport the gate outside the Clifford group, and in the following,
708: we will assume this as a starting point. The advantage of doing
709: this, in practice, is that due to Proposition 1, we have the freedom
710: of preparing the ancilla states up to some Clifford multiplications.
711: 
712: \subsection{Definition of the teleportation depth}
713: 
714: With the standard two-bit teleportation scheme (FIG. \ref{2bit}) in
715: mind, it is easy to see that all gates in the $\mathcal{C}_k$
716: hierarchy can be teleported fault-tolerantly as a whole in a
717: recursive manner. Suppose $U$ is an $n$-qubit $\mathcal{C}_k$ gate.
718: The ancilla state can be fault-tolerantly prepared and all the
719: elements in the teleportation circuit of $U$ are in $\mathcal{C}_2$
720: and can be performed fault-tolerantly, except the classically
721: controlled operation $U_1 = R_{xy}' = U R_{xy} U^{\dagger}$, where
722: $R_{xy}$ is an operator in $\mathcal{C}_1$ which depends on the
723: (random) Bell-basis measurement outcomes $xy$. However, as $U$ is in
724: $\mathcal{C}_k$, $U_1$ is in general a $\mathcal{C}_{k-1}$ operation
725: and can be implemented again by teleportation. In this way, after
726: each teleportation step, a $\mathcal{C}_k$ gate is mapped to another
727: gate one level lower. This recursive procedure terminates when $U_i$
728: is in $\mathcal{C}_2$.
729: 
730: Based on the above picture we give a more formal definition of
731: teleportation, which characterizes its randomness nature.
732: 
733: \begin{definition}
734: The teleportation map $f$ takes an $n$-qubit operator $A$ to a set
735: of operators via the following manner:
736: \begin{equation}
737: f: A\rightarrow \{AP_{j_1}A^{\dagger}\}_{j_1=1}^{4^{n+1}},
738: \end{equation}
739: where $P_i$ are elements of the $n$-qubit Pauli group
740: $\mathcal{P}_n$.
741: \end{definition}
742: 
743: Note
744: \begin{equation}
745: f\circ f: A\rightarrow
746: \{(AP_{j_1}A^{\dagger})P_{j_2}(AP_{j_1}A^{\dagger})^{\dagger}\}_{j_1,j_2=1}^{4^{n+1}},
747: \end{equation}
748: and
749: \begin{eqnarray}
750: f\circ f\circ f: A\rightarrow \nonumber\\
751: \{((AP_{j_1}A^{\dagger})P_{j_2}(AP_{j_1}A^{\dagger})^{\dagger})P_{j_3}((AP_{j_1}A^{\dagger})P_{j_2} \nonumber\\
752: (AP_{j_1}A^{\dagger})^{\dagger})^{\dagger}\}_{j_1,j_2,j_3=1}^{4^{n+1}},
753: \cdots
754: \end{eqnarray}
755: 
756: Each element of image of the map $f^{\circ m}$ on $A$ is associated
757: with a set
758: \begin{equation}
759: S=\{j_1,j_2\ldots,j_m\}.
760: \end{equation}
761: 
762: Denote $f^{\circ m}_S(A)$ as the element in image of the map
763: $f^{\circ m}$ on $A$ associated with the set $S$. Each element in
764: the image occurs with equal probability.
765: 
766: \begin{definition}
767: $f^{\circ m}_S(A)$ terminates if $f^{\circ m}_S(A)\in
768: \mathcal{C}_2$.
769: \end{definition}
770: 
771: If $f^{\circ m_1}_S(A)$ terminates, then $f^{\circ m_2}_{S'}(A)$
772: terminates for any $m_2\geq m_1$, and
773: $S'=\{j_1,j_2\ldots,j_{m_1},\ldots,j_{m_2}\}$. Therefore, for each
774: $f^{\circ m}_S(A)$ that terminates, there must exist a set $S_{min}$
775: with the minimal size such that $f^{\circ |S_{min}|}_{S_{min}}(A)$
776: terminates, where $S_{min}=\{j_1,j_2\ldots,j_{m'}\}$
777: ($m'=|S_{min}|$). In our following discussions, we will only
778: consider sets $S$ which are minimal in this sense.
779: 
780: This mapping procedure works directly on $\mathcal{C}_k$ gates. If
781: $W$ is an $n$-qubit $\mathcal{C}_k$ gate, then there is no need to
782: decompose it into consecutive application of several other gates and
783: we say we can `direct teleport' $W$. $W$ is in $\mathcal{C}_k$ iff
784: $\forall S$, $f^{\circ (k-2)}_S(A) \in \mathcal{C}_2$, and $\exists
785: S'$, s.t. $f^{\circ (k-3)}_{S'}(A) \notin \mathcal{C}_2$.
786: 
787: Among all $\mathcal{C}_k$ gates, the set of semi-Clifford operations
788: have the special property that they can be teleported with only half
789: the ancilla resources as in a standard teleportation scheme. This
790: `one-bit teleportation scheme' is illustrated in FIG. \ref{1bit}.
791: This scheme also complies with the mapping description given above.
792: Instead of Bell basis measurement, randomness in one-bit
793: teleportation scheme comes from single qubit measurement and $P_j$
794: belongs to a maximal abelian subgroup of the whole $n$-qubit Pauli
795: group in general.
796: 
797: To teleport an arbitrary $n$-qubit gate $A$, we can first decompose
798: $A$ into the $\mathcal{C}_k$ hierarchy, $A = A_1 A_2 \dots A_r$,
799: where $A_i \in \mathcal{C}_{k_i}$, because we only know how to
800: teleport $\mathcal{C}_k$ gates fault-tolerantly. We call this
801: procedure `decomposition of $A$  into $\mathcal{C}_{\infty}$'.
802: Suppose that to teleport each gate $A_i$, $m_i$ maps are needed on
803: average, with average taken over all possible set $S = \{j_1,
804: j_2\dots, j_m\}$. Then the teleportation depth of $A$ is defined as
805: follows.
806: 
807: \begin{definition}
808: The teleportation depth of a gate $A$, denoted as $T$, is the
809: minimal sum of all $m_i$--the average number of teleportation steps
810: needed to implement each component gate of $A$--where the minimum is
811: taken over all possible decompositions of $A$ into
812: $\mathcal{C}_{\infty}$.\label{depth}
813: \end{definition}
814: 
815: Due to Definition \ref{depth}, in order to calculate the teleportation
816: depth of a given gate $A$, one needs to find all possible
817: decompositions of $A$ into $\mathcal{C}_{\infty}$ gates and
818: calculate the corresponding depth, then minimize over all of them.
819: This is generally intractable, but one may expect to upper bound the
820: depth with some particular decomposition of $A$ into
821: $\mathcal{C}_{\infty}$ gates.
822: 
823: Let us first consider the case of an $n$-qubit $\mathcal{C}_k$ gate.
824: 
825: \begin{definition}
826: $T(n,k)$ is the teleportation depth of an $n$-qubit $\mathcal{C}_k$
827: gate.
828: \end{definition}
829: 
830: As such a gate can be teleported directly, $T(n,k)$ is upper bounded
831: by the average number of steps needed in this direct teleportation
832: scheme to terminate the teleportation procedure.
833: 
834: \begin{equation}
835: T(n,k) \leq \frac{1}{N} \sum_S |S|
836: \end{equation}where the summation is over all possible (minimal) sets $S$ and $N$
837: is the number of such sets.
838: 
839: However, when $k\rightarrow\infty$, it is not obvious that the above
840: summation will converge. We will show that this is true. Then for an
841: arbitratry gate $A$, by decomposing $A$ into a finite series of
842: $\mathcal{C}_k$ gates, we can see that the teleportation depth of
843: $A$ turns out to be finite. Then we do not actually require the
844: procedure to terminate within a finite number of steps.
845: 
846: Different teleportation schemes, for example one-bit and two-bit
847: teleportation, give different upper bounds on teleportation depth
848: for a certain circuit. While for some circuits one scheme is
849: obviously more efficient than others, the comparison among different
850: schemes in other case may not be so straightforward and may depend
851: sensitively on various parameters in the circuit. In the following
852: sections, we study such dependence and present surprising results
853: beyond our usual expectation with examples from important quantum
854: circuits.
855: 
856: \subsection{Teleportation depth of semi-Clifford $\mathcal{C}_k$ gates}
857: 
858: We first calculate explicitly an upper bound for the teleportation
859: depth of semi-Clifford $n$-qubit $\mathcal{C}_k$ gates. We know from
860: \cite{Xinlan} that this kind of gate can be teleported directly with
861: the architecture of one-bit teleporation and we denote the upper
862: bound calculated with this `\textbf{one-bit}' `\textbf{direct}'
863: teleportation procedure as $T_1(n,k)$. For a general $n$-qubit gate,
864: if it is possible to decompose it into a series of semi-Clifford
865: $\mathcal{C}_k$ operations, the upper bound of teleportation depth
866: obtained by teleporting each part separately using one-bit
867: teleportation scheme is in general denoted as $T_1$.
868: 
869: \begin{definition}
870: $T_1$ is the average total number of teleportation steps needed to
871: teleport separately each semi-Clifford $\mathcal{C}_k$ component of
872: a quantum circuit using the one-bit teleportation scheme, if such a
873: decomposition is possible.
874: 
875: More specifically, $T_1(n,k)$ is the average number of teleportation
876: steps needed to teleport an $n$-qubit semi-Clifford $\mathcal{C}_k$
877: gate directly (i.e. without decomposition) using the one-bit
878: teleportation scheme.
879: \end{definition}
880: 
881: Apparently we have $T(n,k) \leq T_1(n,k)$ in general.
882: 
883: The probability that the teleportation process terminates
884: immediately after one teleportation step equals the percentage
885: weight of a maximal abelian subgroup in the whole Pauli group, which
886: is $\frac{1}{2^n}$ for an $n$-qubit Pauli group. Now each
887: teleportation step may have two possible endings: i) with
888: probability $p=\frac{1}{2^n}$, $\{U\mathcal{P}_nU^{\dagger}\} \in
889: \mathcal{P}_n$ and the process terminates; ii) with probability
890: $1-p$, $\{U\mathcal{P}_nU^{\dagger}\}$ is a general $n$-qubit
891: $\mathcal{C}_{k-1}$ gate and the process goes on. The upper bound of
892: teleportation depth calculated with this process is then
893: 
894: \begin{eqnarray}
895: T_1(n,k) &=& p\sum\limits_{s=1}^{k-3} s(1-p)^{s-1}+(k-2)(1-p)^{k-3}\nonumber\\
896: &=&2^n\left(1-(1-\frac{1}{2^n})^{k-2}\right).\label{depthck}
897: \end{eqnarray}
898: 
899: It is clearly seen from Eq. (\ref{depthck}) that $T_1(n,k)$
900: converges to $2^n$ when $k\rightarrow\infty$, which means that
901: $T(n,k)$ is in general bounded. For instance, when $n=2$, Eq.
902: (\ref{depthck}) tells us $T(2,k) \leq T_1(2,k) = 4(1-(3/4)^{k-2})$.
903: The behavior of $T_1(2,k)$ is shown in FIG. \ref{depthtwoqubits}.
904: However, since $T_1(2,k) = 4(1-(3/4)^{k-2}) \leq 4(1-(1/2)^{k-2}) =
905: 2T_1(1,k)$, we find that teleporting two single-qubit semi-Clifford
906: $\mathcal{C}_k$ gates together using the one-bit teleportation
907: scheme needs fewer teleportation steps than to teleport each of them
908: separately.
909: 
910: \begin{figure}[htbp]
911: \includegraphics[width=3.00in]{T12k.eps}
912: \caption{The behavior of
913: $T_1(2,k)=4(1-(3/4)^{k-2})$}\label{depthtwoqubits}
914: \end{figure}
915: 
916: Since $1-\frac{1}{2^n}<1$, $T_1(n,k)$ quickly reaches $2^n$ as $k$
917: grows. Therefore, generally, the upper bound of the teleportation
918: depth of a $\mathcal{C}_k$ gate given by `direct teleportation' is
919: not determined by $k$, but by the number of qubits $n$ it actually
920: acts on. Moreover, since $T_1(n,\infty)=2^n$, i.e. the upper bound
921: of teleportation depth increases exponentially with $n$, in
922: generally, when $n,k$ are large, it is better to decompose an
923: $n$-qubit $\mathcal{C}_k$ gate into some one and two-qubits gates to
924: get a lower upper bound.
925: 
926: However, if $k\sim P(n)$, where $P(n)$ is a polynomial in $n$, then
927: $T_1(n,k)$ scales as $P(n)$.
928: 
929: Now we give two examples as applications of the above upper bounds,
930: through which we obtain some idea about the order of teleportation
931: depth in comparison with the usual circuit depth.
932: 
933: \subsubsection{Teleportation depth of the $n$-qubit QFT}
934: 
935: \begin{figure}[htbp]
936: \includegraphics[width=3.40in]{QFT.eps}
937: \caption{Circuit for $n$-qubit Quantum Fourier Transform}\label{QFT}
938: \end{figure}
939: 
940: The first example is the $n$-qubit Quantum Fourier Transform (QFT)
941: circuit, as shown in FIG. \ref{QFT}. $R_k$ denotes the unitary
942: transformation $R_k=\text{diag}(1,e^{2\pi i/2^k})$. The circuit
943: depth of $n$-qubit QFT goes as $n^2$ and we will soon find that the
944: teleportation depth of this circuit is of the same order.
945: 
946: Each block of gates within a single dashed box (Hadamard plus
947: controlled $z$-rotations on the $k^{th}$ qubit)is a semi-Clifford
948: $(n-k+1)$-qubit $\mathcal{C}_{n-k+2}$ gate, $k = 1, \dots, n-1$ and
949: can be teleported directly using the one-bit scheme. Therefore the
950: whole circuit can be teleported piece by piece by one-bit
951: teleportation. Note that
952: \begin{eqnarray}
953: T(n,k=n+1) &\leq& T_1(n,k=n+1)\\&=&
954: 2^n\left(1-(1-\frac{1}{2^n})^{n-1}\right)\\
955: &\sim& n-1
956: \end{eqnarray}
957: for large $n$. Actually,  numerical data shows that even when $n$ is
958: small, $T_1(n,k=n+1)\sim n-1$ is almost also true.
959: 
960: Therefore, the teleportation depth of the $n$-qubit QFT is
961: upper-bounded by
962: \begin{eqnarray}
963: \sum_{j=2}^{n} T(j,k=j+1)&\leq& \sum_{j=2}^{n} T_1(j,k=j+1)\\
964: & \leq & \sum_{j=1}^n (j-1)\nonumber\\
965: &=&\frac{1}{2}n(n-1)\sim\mathcal{O}(n^2).
966: \end{eqnarray}
967: 
968: Numerial calculation shows that $\sum_{j=2}^{n} T_1(j,k=j+1)$ is
969: almost $\frac{1}{2}n(n-1)-1$.
970: 
971: Note the probability for the teleportation process to terminate is
972: $1$ for teleporting an $n$-qubit $\mathcal{C}_{k = n+1}$ gate
973: $n-1=k-2$ times. This means that the upper bound we got for this
974: block teleportation scheme of QFT is just slightly lower than
975: naively assuming that we need $k-2$ teleportation steps to teleport
976: a $\mathcal{C}_k$ gate. The reason we do not benefit from the
977: avarage is that for QFT, $k$ is generally comparable with $n$.
978: 
979: \subsubsection{Uniformly Controlled rotation}
980: 
981: \begin{figure}[htbp]
982: \includegraphics[width=3.00in]{Uniformlycontrol.eps}
983: \caption{Definition of the $n-1$-fold uniformly controlled rotation
984: of a qubit about the axis $\vec{a}$}\label{fig3}
985: \end{figure}
986: 
987: Now we consider another example, the uniformly controlled rotations,
988: which are widely used in analyzing the circuit complexity of an
989: arbitrary $n$-qubit quantum gate \cite{Mott} \cite{Vivek}. This
990: circuit in general needs $2^{n+2}-4n-4$ CNOT gates and $2^{n+2}-5$
991: one-qubit elementary rotations to implement. For complexity analysis
992: of this circuit see for example \cite{Mott}.
993: 
994: The teleportation depth of this rotation is in general upper bounded
995: by $2^n$. However, if each $(n-1)$-qubit-controlled gate is in
996: $\mathcal{C}_k$, we might expect to do better. For instance, when
997: $k=cn$, for any positive constant $n$, the teleportation depth
998: scales as $cn$, i.e. linear in $n$. Moreover, if $k\sim P(n)$, where
999: $P(n)$ is a polynomial in $n$, then the teleportation depth scales
1000: as $P(n)$.
1001: 
1002: \subsection{Teleportation depth beyond semi-Clifford $\mathcal{C}_k$ gates}
1003: 
1004: Now recall our series of examples of non-semi-Clifford
1005: $\mathcal{C}_k$ gates given in FIG. \ref{fig4}. We know that if
1006: $V_k\in\mathcal{C}_k$, then $W_k\in\mathcal{C}_{k+1}$. And the group
1007: $W_k\mathcal{P}_3W_k^{\dagger}$ does not contain a maximally abelian
1008: subgroup of $\mathcal{P}_3$, i.e. $W_k\in\mathcal{C}_{k+1}$ is not
1009: directly one-bit teleportable.
1010: 
1011: Therefore, we know that there are some $W_k$ gates in the
1012: $\mathcal{C}_k$ hierarchy which can only be teleported directly by
1013: the standard two-bit teleportation scheme. Using this scheme, we can
1014: calculate another upper bound for teleportation depth, which we
1015: denote as $T_2(n,k)$.
1016: 
1017: \begin{definition}
1018: $T_2$ is the average total number of teleportation steps needed to
1019: teleport separately each $\mathcal{C}_k$ component of a quantum
1020: circuit using two-bit teleportation scheme, if such a decomposition
1021: is possible.
1022: 
1023: More specifically, $T_2(n,k)$ is the average number of teleportation
1024: steps needed to teleport an $n$-qubit $\mathcal{C}_k$ gate directly
1025: (i.e. without decomposition) using the two-bit teleportation scheme.
1026: \end{definition}
1027: 
1028: 
1029: For a general $n$-qubit $\mathcal{C}_k$ gate, $T_2(n,k)$ can be
1030: calculated by replacing $p$ with $\frac{1}{4^n}$ in Eq.
1031: (\ref{depthck})
1032: \begin{eqnarray}
1033: T_2(n,k) &=& p\sum\limits_{s=1}^{k-3} s(1-p)^{s-1}+(k-2)(1-p)^{k-3}\nonumber\\
1034: &=&4^n\left(1-(1-\frac{1}{4^n})^{k-2}\right)\label{depthck2}
1035: \end{eqnarray}which then converges to $4^n$ when $k\rightarrow\infty$.
1036: 
1037: One may guess that in general to teleport $W_k$ directly using the
1038: two-bit scheme will give a lower bound for teleportation depth than
1039: to teleport the Toffoli gate and $V_k=diag(1,e^{i\pi/2^{k-1}})$
1040: separately using the one-bit scheme. Surprisingly, this is not
1041: generally true.
1042: 
1043: When $V_k\in\mathcal{C}_3$, this is indeed true. Teleporting $W_k$
1044: directly gives a bound of $T_2(3,4)=1.875$, which is less than
1045: $T_1(3,4)=2$, i.e. the bound given by teleporting the Toffoli gate
1046: and $V_k$ separately with the one-bit scheme.
1047: 
1048: However, when $k\rightarrow\infty$, teleporting $W_k$ directly gives
1049: a bound of $T_2(3,4)=5.25$, which is greater than $T_1(3,4)=3$, i.e.
1050: the bound given by teleporting the Toffoli gate and $V_k$
1051: separately.
1052: 
1053: This means that there exists a critical value $k$ that determines
1054: which way is more efficient for teleporting $W_k$, directly or
1055: separately.
1056: 
1057: Note if $V_k\in\mathcal{C}_k$, we also have
1058: $W_k^{\dagger}\in\mathcal{C}_{k+1}$. Calculating the bounds of
1059: teleportation depth for $W_k^{\dagger}$ shows a similar behavior as
1060: that of $W_k$, however of a slightly different value. For instance,
1061: when $V_k\in\mathcal{C}_3$, teleporting $W_k^{\dagger}$ directly
1062: gives a bound of $1.5$, which is less than 2, the bound given by
1063: teleporting separately. However when $k\rightarrow\infty$,
1064: teleporting $W_k^{\dagger}$ directly gives a bound of $5.5$, but
1065: teleporting separately gives only a bound of $3$.
1066: 
1067: Up to now, our discussion is entirely based on the $\mathcal{C}_k$
1068: hierarchy. To summarize the capacity of $\mathcal{C}_k$ for
1069: fault-tolerant quantum computation and provide basis for comparison
1070: with non-$\mathcal{C}_k$ schemes discussed below, we introduce
1071: another notion of $T_k$.
1072: 
1073: \begin{definition}
1074: $T_k$ is the minimum number of total teleportation steps needed to
1075: teleport separately each $\mathcal{C}_k$ component of a quantum
1076: circuit using either one-bit or two-bit teleportation scheme.
1077: \end{definition}
1078: 
1079: $T_k$ is defined in a way that represents the maximum capacity of
1080: teleportation based on $\mathcal{C}_k$ hierarchy. In general $T_1
1081: \geq T_k$, $T_2 \geq T_k$. To understand exactly how they compare
1082: for a given circuit, a full characterization of the structure of
1083: $\mathcal{C}_k$ is necessary. Here based on the structure theorems
1084: given in Section III, we gave a simple example where $T_1$ or $T_2$
1085: could be strictly larger than $T_k$. The next question to ask is
1086: then whether we can go beyond $\mathcal{C}_k$ and this will be
1087: discussed in the following section.
1088: 
1089: \subsection{Teleportation beyond $\mathcal{C}_k$}
1090: 
1091: In the definition of teleportation depth, we require that $A$ be
1092: decomposed into a set of $\mathcal{C}_{\infty}$ gates. This is due
1093: to the fact that $\mathcal{C}_{\infty}$ are the only gates that we
1094: know so far how to perform fault-tolerantly by teleportation. In
1095: general, if we do not require the decomposition to be in
1096: $\mathcal{C}_{\infty}$, then we might get a better upper bound on
1097: teleportation depth than the one defined previously, i.e. there
1098: might exist upper bound $T^*$ of teleportation depth that is
1099: strictly less than $T_k$. We give two such examples below. We leave
1100: open the problem of how to implement teleportations fault-tolerantly
1101: for a general $n$-qubit gate.
1102: \\
1103: 
1104: \textbf{Example 1} For a general one-qubit gate $U$, we know that
1105: $U$ can be decomposed into three $\mathcal{C}_{\infty}$ gates, each
1106: of which has $T_1 < 2$. Hence through the decomposition we can bound
1107: its total teleportation depth by $6$.
1108: 
1109: However, to teleport $U$ directly without decomposition via two-bit
1110: teleportation gives a bound of $T_2 < 4^1 = 4$ less than $T_k$.
1111: \\
1112: 
1113: \textbf{Example 2} Consider a classical reversible circuit given in
1114: FIG. \ref{fig5}. We denote this series of three Toffoli gates as
1115: $R_{c3}$.
1116: \begin{figure}[htbp]
1117: \includegraphics[width=1.00in]{Threetoffoli.eps}
1118: \caption{The $R_{c3}$ gate--Three Toffoli gates in
1119: series}\label{fig5}
1120: \end{figure}
1121: 
1122: This gate $R_{c3}$ is not in $\mathcal{C}_k$ hierarchy as can be
1123: shown below:
1124: 
1125: Suppose that $R_{c3} \in \mathcal{C}_k$ is at certain level of the
1126: hierarchy, $R_{c3} X_1 R_{c3}^{\dagger}$ must be a gate in
1127: $\mathcal{C}_{k-1}$. Calculating explicitly as in FIG. \ref{fig6} we
1128: have
1129: 
1130: \begin{figure}[htbp]
1131: \includegraphics[width=3.50in]{TcXTc.eps}
1132: \caption{Conjugating $X_1$ by $R_{c3}$}\label{fig6}
1133: \end{figure}
1134: 
1135: The non-Clifford part of the right hand side of the equation is a
1136: series of two Toffoli gates, and we denote it as $R_{c2}$. Due to
1137: Proposition 3, $R_{c2}$ is also in $\mathcal{C}_{k-1}$.
1138: 
1139: As shown in FIG. \ref{fig7}, conjugating $X_1$ by $R_{c2}$ results in
1140: $L R_{c2}^{\dagger}$, where $L$ is a Clifford operation. However, by
1141: exchanging the second and third qubit in FIG. \ref{fig7}, we find
1142: that $R_{c2}^{\dagger} X_1 R_{c2} = L' R_{c2}$, i.e. conjugating
1143: $X_1$ by $R_{c2}^{\dagger}$ gives back $R_{c2}$. Therefore, $R_{c2}$
1144: cannot be in the $\mathcal{C}_k$ hierarchy and we can conclude that
1145: $R_{c3}$ is not a $\mathcal{C}_k$ gate either. $\square$
1146: 
1147: \begin{figure}[htbp]
1148: \includegraphics[width=3.00in]{Tc2XTc2.eps}
1149: \caption{Conjugating $X_1$ by $R_{c2}$}\label{fig7}
1150: \end{figure}
1151: 
1152: If we leave aside the problem of how to teleport gates beyond
1153: $\mathcal{C}_k$ fault-tolerantly, we can teleport $R_{c3}$ directly
1154: and obtain an upper bound of $2.75$, which is less than $T_k = 3$,
1155: the bound given by teleporting the three Toffoli gates separately.
1156: 
1157: 
1158: \section{Conclusion and Discussion}
1159: 
1160: In this paper we address the following questions: what is the
1161: capacity of the teleportation scheme in practical implementation of
1162: fault-tolerant quantum computation and what is the most efficient
1163: way to make use of the teleportation protocol. To answer these
1164: questions we first notice that  one-bit and two-bit teleportation
1165: schemes require different resources to implement and are of
1166: different capabilities. To understand what kind of gates can be
1167: teleported fault-tolerantly with these two schemes respectively, we
1168: study the structure of $\mathcal{C}_k$ hierarchy and its
1169: relationship with semi-Clifford operations.  We show for $n=1,2$,
1170: all the $\mathcal{C}_k$ gates are semi-Clifford operations, which is
1171: also true for $\{n=3,k=3\}$. However, this is no longer true for
1172: parameters $\{n>2,k>3\}$. Based on the counterexamples we
1173: constructed for $\{n=3,k \geq 3\}$, we conjecture that all
1174: $\mathcal{C}_3$ gates are semi-Clifford and all $\mathcal{C}_k$
1175: gates are generalized semi-Clifford.
1176: 
1177: Such an understanding of the $\mathcal{C}_k$ structure has great
1178: implications on the optimal design of fault-tolerant architectures.
1179: While all $\mathcal{C}_k$ gates can be teleported fault-tolerantly,
1180: the semi-Clifford subset of it requires less resources to implement
1181: than others. To quantify this notion of gate complexity in
1182: fault-tolerant quantum computation based on the $\mathcal{C}_k$
1183: hierarchy, we introduce a measure called the teleportation depth
1184: $T$, which characterizes how many teleportation steps are necessary,
1185: on average, to implement a given gate.  Using different
1186: teleportation schemes, we can give different upper bounds on $T$,
1187: for example $T_1$, $T_2$ and $T_k$. General assumption was that $T_1
1188: = T_2 = T_k = T$. However we showed in this work that, surprisingly
1189: for certain series of gates $T_1$ could be strictly greater than
1190: $T_k$ and $T_k$ could also be strictly greater than $T$.
1191: 
1192: The ultimate understanding of the structure of $\mathcal{C}_k$ will
1193: provide a clearer clue on how to teleport circuits most efficiently.
1194: To achieve this goal, some results from other branches of
1195: mathematics might be helpful. It is noted that the Barnes-Wall
1196: lattices, whose isometry group is a subgroup of index 2 in the real
1197: Clifford group, have been extensively studied and recently their
1198: involutions have been classified \cite{Griess}. It is our hope that
1199: the $\mathcal{C}_3$ structure might be further understood once we
1200: have a better understanding of the Clifford group.
1201: 
1202: For $n=1$, we fully characterize the structure of $\mathcal{C}_k$ by
1203: further study on the diagonal gates in $\mathcal{C}_k$, which form a
1204: group. It is interesting to note some evidence that $\mathcal{C}_k$
1205: gates might be the only non-Clifford gates which could be
1206: transversally implemented on a stabilizer code \cite{ZCC}. We also
1207: fully characterize the structure of $\mathcal{C}_3$ for $n=3$, but
1208: this seems not directly related to allowable transversal
1209: non-Clifford gates on stabilizer codes. It is shown that those
1210: transversal non-Clifford gates are allowed only if they are
1211: generalized semi-Clifford \cite{Xie}, therefore we might expect some
1212: generalized semi-Clifford $\mathcal{C}_k$ gates transversally
1213: implementable on some stabilizer codes. We believe such kind of
1214: exploration on the relationship between transversally implementable
1215: gates and teleportable gates will shed some light on further
1216: understanding of practical implementation of fault-tolerant
1217: architectures.
1218: 
1219: \section*{Acknowledgments}
1220: 
1221: We thank Daniel Gottesman, Debbie Leung, and Carlos Mochon for comments.
1222: 
1223: \section*{Appendix A: Single qubit $\mathcal{C}_k$ gates}
1224: 
1225: \subsection*{1. Single qubit gates with eigenvalues $\pm 1$}
1226: 
1227: In this section we discuss what kind of single qubit unitary gates
1228: could have eigenvalues $\pm 1$ apart from an overall phase factor,
1229: i.e. if $\lambda_+,\lambda_-$ denote the two eigenvalues of a single
1230: qubit unitary $U$, then what is the condition under which
1231: $\lambda_++\lambda_-=0$. This information is useful since only the
1232: unitary of this kind can be transformed into elements in Pauli group
1233: under conjugation, i.e. there exits a unitary operator $R$, such
1234: that $RAR^{\dag}=e^{i\theta}U$, where $A\in \mathcal{C}_{1}$. We'll
1235: see that those kind of unitary has very restricted form which is
1236: given by the following proposition.
1237: 
1238: \begin{proposition}:
1239: The single qubit unitary gates which have eigenvalues $\pm 1$ apart
1240: from an overall phase factor could only be of the following two
1241: forms:
1242: \begin{center}
1243: $\Gamma_1(\varphi)=\left[
1244: \begin{array}{cc}
1245: 0 & 1 \\
1246: e^{i\varphi} & 0
1247: \end{array}
1248: \right]$
1249: \end{center}
1250: or
1251: \begin{center}
1252: $\Gamma_2(\phi,\xi)=\left[
1253: \begin{array}{cc}
1254: \cos{\phi}& \sin{\phi}e^{i\xi} \\
1255: \sin{\phi}e^{-i\xi} & -\cos{\phi}
1256: \end{array}
1257: \right]$
1258: \end{center}
1259: \end{proposition}
1260: \textbf{Proof}: We begin to prove this proposition by writing down a
1261: general form of single qubit unitary gate as the following:
1262: \begin{equation}
1263: \Gamma=\left[
1264: \begin{array}{cc}
1265: \cos{\phi}e^{i\theta} & \sin{\phi}e^{i\xi} \\
1266: \sin{\phi}e^{-i\xi} & -\cos{\phi}e^{-i\theta}
1267: \end{array}
1268: \right]
1269: \end{equation}
1270: Direct calculation gives
1271: \begin{align}
1272: \lambda_{\pm}&=
1273: \frac{1}{2}\cos{\phi}e^{i\theta}-\frac{1}{2}\cos{\phi}e^{-i\theta}\nonumber\\
1274: &\pm\frac{1}{2} e^{-i\theta}(\cos{\phi}^2 e^{4i\theta}-2\cos{\phi}^2
1275: e^{2i\theta} +\cos{\phi}^2+4 e^{2i\theta})^{1/2}
1276: \end{align}
1277: Therefore $\lambda_++\lambda_-=0$ gives
1278: \begin{equation}
1279: \cos{\phi}\sin{\theta}=0
1280: \end{equation}
1281: If $\cos{\phi}=0$, the unitary must adopt the form of
1282: $\Gamma_1(\varphi)$; if $\sin{\theta}=0$, then apart from an overall
1283: phase, we can simply choose $\theta=0$ which leads to the form of
1284: $\Gamma_2(\phi,\xi)$.$\square$
1285: 
1286: Note $\Gamma_1$ could be viewed as a special situation of $\Gamma_2$
1287: for the case $\cos{\phi}=0$. However, we list $\Gamma_1$ separately
1288: for future convenience.
1289: 
1290: \subsection*{2. Gate series associated with $\Gamma_1(\varphi)$ and
1291: $\Gamma_2(\phi,\xi)$}
1292: 
1293: In this section we investigate the gate seises associated with
1294: $\Gamma_1(\varphi)$ and $\Gamma_2(\phi,\xi)$. It is obvious that if
1295: $\Gamma_1(\varphi), \Gamma_2(\phi,\xi) \in \mathcal{C}_{k}$, then
1296: the unitary $U(\varphi)$ whose columns are the eigenvectors of
1297: $\Gamma_1(\varphi)$ or $\Gamma_2(\phi,\xi)$ might be in
1298: $\mathcal{C}_{k+1}$, given that
1299: $U(\varphi)ZU(\varphi)^{\dag}=\Gamma_1(\varphi)$.
1300: 
1301: For $\Gamma_1(\varphi)$, the two normalized eigenvectors can be
1302: chosen as
1303: \begin{align}
1304: |\Gamma_1(\varphi)\rangle_+&=\frac{1}{\sqrt{2}}(|0\rangle+e^{i\varphi/2}|1\rangle)\nonumber\\
1305: |\Gamma_1(\varphi)\rangle_-&=\frac{1}{\sqrt{2}}(|0\rangle-e^{i\varphi/2}|1\rangle)
1306: \end{align}
1307: we now want a unitary whose columns is are eigenvectors of
1308: $\Gamma_1(\varphi)$ apart from an overall factor of each
1309: eigenvector, i.e.
1310: \begin{align}
1311: U(\varphi,\alpha)&=(e^{i\alpha}|\Gamma_1(\varphi)\rangle_+,
1312: |\Gamma_1(\varphi)\rangle_-)\nonumber\\
1313: &=\frac{1}{\sqrt{2}}\left[
1314: \begin{array}{cc}
1315: e^{i\alpha} & 1 \\
1316: e^{i\alpha}e^{i\varphi/2} & -e^{i\varphi/2}
1317: \end{array}
1318: \right].
1319: \end{align}
1320: If $U(\varphi,\alpha)\in \mathcal{C}_{k+1}$, then
1321: $U'=L_1U(\varphi,\alpha)L_2$ is also in $\mathcal{C}_{k+1}$. What is
1322: important for us is to find $U'$ which is either of the form
1323: $\Gamma_1$ or $\Gamma_2$, then from its eigenvectors we can generate
1324: gates in $\mathcal{C}_{k+1}$. It is noticed that if we choose
1325: $\alpha=0$, then
1326: \begin{align}
1327: U(\varphi,0)&=(|\Gamma_1(\varphi)\rangle_+,
1328: |\Gamma_1(\varphi)\rangle_-)\nonumber\\
1329: &=\frac{1}{\sqrt{2}}\left[
1330: \begin{array}{cc}
1331: 1 & 1 \\
1332: e^{i\varphi/2} & -e^{i\varphi/2}
1333: \end{array}
1334: \right],
1335: \end{align}
1336: and
1337: \begin{align}
1338: U(\varphi,0)HX=\left[
1339: \begin{array}{cc}
1340: 0 & 1 \\
1341: e^{i\varphi/2} & 0
1342: \end{array}
1343: \right]=\Gamma_1(\varphi/2).
1344: \end{align}
1345: Later we will show that for all the allowed value of $\alpha$, there
1346: exist $L_1,L_2 \in \mathcal{C}_2$, such that
1347: $L_1U(\varphi,0)L_2=U(\varphi,\alpha)$, so it is sufficient to
1348: consider the case of $\alpha=0$.
1349: 
1350: Therefore we get a set of unitary given by
1351: \begin{equation}
1352: V_k(\varphi)=\Gamma_1(\varphi/2^{k}),
1353: \end{equation}
1354: if $\Gamma_1(\varphi) \in \mathcal{C}_2$ then
1355: $\Gamma_1(\varphi/2^{k})$ could be in $\mathcal{C}_k$. We already
1356: know that $\Gamma_(\pi/2)$ is in $\mathcal{C}_2$, then we have
1357: \begin{equation}
1358: V_k=\Gamma_1(2\pi/2^{k})
1359: \end{equation}
1360: is in $\mathcal{C}_k$.
1361: 
1362: Note
1363: \begin{equation}
1364: S_kX=V_k,
1365: \end{equation}
1366: and we already know that $S_k \in \mathcal{C}_k$. Therefore by
1367: deriving $V_k$ we get nothing new due to proposition 1.
1368: 
1369: Now we come to the $\Gamma_2(\phi,\xi)$ case. Similarly, we begin
1370: from the two normalized eigenvectors of $\Gamma_2(\phi,\xi)$, which
1371: can be chosen as
1372: \begin{eqnarray}
1373: |\Gamma_2(\phi,\xi)\rangle_+&=&\frac{1}{\sqrt{2}}(\cos{\frac{\phi}{2}}|0\rangle+\sin{\frac{\phi}{2}}e^{-i\xi}|1\rangle)\nonumber\\
1374: |\Gamma_2(\phi,\xi)\rangle_-&=&\frac{1}{\sqrt{2}}(\sin{\frac{\phi}{2}}e^{i\xi}|0\rangle-\cos{\frac{\phi}{2}}|1\rangle)
1375: \end{eqnarray}
1376: we now construct a unitary whose columns are eigenvectors of
1377: $\Gamma_2(\varphi)$ apart from an overall factor of each
1378: eigenvector, i.e.
1379: \begin{align}
1380: U(\phi,\xi,\beta)&=(e^{i\beta}|\Gamma_2(\phi,\xi)\rangle_+,
1381: |\Gamma_2(\phi,\xi)\rangle_-)\nonumber\\
1382: &=\frac{1}{\sqrt{2}}\left[
1383: \begin{array}{cc}
1384: e^{i\beta}\cos{\frac{\phi}{2}} & \sin{\frac{\phi}{2}}e^{i\xi} \\
1385: e^{i\beta}\sin{\frac{\phi}{2}}e^{-i\xi} & -\cos{\frac{\phi}{2}}
1386: \end{array}
1387: \right].
1388: \end{align}
1389: If $U(\phi,\xi,\beta)\in \mathcal{C}_{k+1}$, then
1390: $U'=L_1U(\phi,\xi,\beta)L_2$ is also in $\mathcal{C}_{k+1}$. It is
1391: noticed that if we choose $\beta=0$, then
1392: \begin{align}
1393: U(\phi,\xi,0)&=(|\Gamma_2(\phi,\xi)\rangle_+,
1394: |\Gamma_2(\phi,\xi)\rangle_-)\nonumber\\
1395: &=\frac{1}{\sqrt{2}}\left[
1396: \begin{array}{cc}
1397: \cos{\frac{\phi}{2}} & \sin{\frac{\phi}{2}}e^{i\xi} \\
1398: \sin{\frac{\phi}{2}}e^{-i\xi} & -\cos{\frac{\phi}{2}}
1399: \end{array}
1400: \right].
1401: \end{align}
1402: Also later we will show that for all the allowed value of $\alpha$,
1403: there exist $L_1,L_2 \in \mathcal{C}_2$, such that
1404: $L_1U(\phi,\xi,0)L_2=U(\phi,\xi,\beta)$, so it is sufficient to
1405: consider the case of $\beta=0$.
1406: 
1407: Therefore we get a set of unitary given by
1408: \begin{equation}
1409: W_k(\phi,\xi)=\Gamma_2(\phi/2^{k-1},\xi),
1410: \end{equation}
1411: if $\Gamma_2(\phi,\xi) \in \mathcal{C}_2$ then
1412: $\Gamma_1(\phi/2^{k-1},\xi)$ could be in $\mathcal{C}_k$. We already
1413: know that only for $\Gamma_2(\pi/4,0)$ is in $\mathcal{C}_2$, then
1414: we have
1415: \begin{equation}
1416: W_k=\Gamma_2(\pi/2^{k},0)
1417: \end{equation}
1418: is in $\mathcal{C}_k$.
1419: 
1420: Note for other possible values of $\phi$ and $\xi$, it is
1421: straightforward to show that there exist $L_1,L_2 \in
1422: \mathcal{C}_2$, such that
1423: $L_1\Gamma_2(\pi/4,0)L_2=\Gamma_2(\phi,\xi)$, so it is sufficient to
1424: consider the case of $\phi=\pi/4$ and $\xi=0$.
1425: 
1426: Note
1427: \begin{equation}
1428: HPW_kPX \sim S_k,
1429: \end{equation}
1430: where $\sim$ means up to an overall phase, and we already know that
1431: $S_k \in \mathcal{C}_k$. Therefore again by deriving $W_k$ we get
1432: nothing new due to proposition 1.
1433: 
1434: \subsection*{3. Gates in $\mathcal{C}_k \setminus \mathcal{C}_{k-1}$ for single qubit}
1435: 
1436: We conclude this section by presenting the following proposition,
1437: which gives the structure of Gates in $\mathcal{C}_k \setminus
1438: \mathcal{C}_{k-1}$ for single qubit.
1439: 
1440: \begin{proposition}
1441: The set $\mathcal{C}_k \setminus \mathcal{C}_{k-1}$ for single qubit
1442: is given by
1443: \begin{equation}
1444: L_1S_kL_2\in \mathcal{C}_k
1445: \end{equation}
1446: where $L_1,L_2\in \mathcal{C}_2$, $k\geq 2$.
1447: \end{proposition}
1448: \textbf{Proof}: We almost reached the proof of this proposition by
1449: considering the results in subsections A and B. The only left we
1450: need to clarify is
1451: 
1452: 1. What happens when $\mathcal{C}_k$ is diagonal, which can not be
1453: directly obtained by considering the eigenvectors of $V_{k-1}$ and
1454: $W_{k-1}$. The answer is already known, since $S_k$ is the only
1455: diagonal gate in $\mathcal{C}_k \setminus \mathcal{C}_{k-1}$.
1456: 
1457: 2. The values of $\alpha$ and $\beta$. This can be answered by
1458: noting the fact the equations
1459: \begin{eqnarray}
1460: UZU^{\dag}&=&G_1\nonumber\\
1461: UXU^{\dag}&=&G_2
1462: \end{eqnarray}
1463: with $G_1,G_2$ known totally determines $U$ up to an overall phase.
1464: Let's start from
1465: \begin{equation}
1466: U(\varphi,\alpha)=\frac{1}{\sqrt{2}}\left[
1467: \begin{array}{cc}
1468: e^{i\alpha} & 1 \\
1469: e^{i\alpha}e^{i\varphi/2} & -e^{i\varphi/2}
1470: \end{array}
1471: \right].
1472: \end{equation}
1473: Note $U(\varphi,\alpha)ZU(\varphi,\alpha)^{\dag}\sim
1474: \Gamma_1(2\phi)$, and
1475: \begin{align}
1476: &U(\varphi,\alpha)XU(\varphi,\alpha)^{\dag}\nonumber\\
1477: =&\frac{1}{\sqrt{2}}\left[
1478: \begin{array}{cc}
1479: \cos{\alpha} & \sin{\alpha}e^{-i2\varphi}\\
1480: \sin{\alpha}e^{i2\varphi} & -\cos{\alpha}
1481: \end{array}
1482: \right].\square
1483: \end{align}
1484: 
1485: \section*{Appendix B: Detailed analysis about $\mathcal{C}_3$}
1486: 
1487: \subsection*{1. Notations}
1488: 
1489: Let's first define some notations.
1490: 
1491: Recall $\mathcal{P}_n$ is the Pauli group for $n$ qubit with order
1492: $4^{n+1}$. Now let $\widetilde{\mathcal{P}}_n$ be the quotient group
1493: $\mathcal{P}_n/Z(\mathcal{P}_n)$ with order $4^n$.
1494: 
1495: Let $\mathcal{C}_2(n)$ denote the Clifford group for $n$ qubit.
1496: Define the quotient group
1497: $\widetilde{\mathcal{C}}_2(n)=\mathcal{C}_2(n)/Z(\mathcal{C}_2(n))$.
1498: Since $\widetilde{\mathcal{P}}_n$ is a normal subgroup of
1499: $\widetilde{\mathcal{C}}_2(n)$, we could further define a quotient
1500: group
1501: $\widehat{\mathcal{C}}_2(n)=\widetilde{\mathcal{C}}_2(n)/\widetilde{\mathcal{P}}_n\cong
1502: Sp(2n,2)$. Note $Sp(2,2)\cong S_3$ and $Sp(4,2)\cong S_6$. Denote
1503: the set $\mathcal{K}(n)=\{A|A\in Sp(2n,2), A^2=1\}$, i.e.
1504: $\mathcal{K}(n)$ are the set of all involutions of the symplectic
1505: group $Sp(2n,2)$.
1506: 
1507: Denote the order of maximal Abelian subgroup of $\mathcal{K}(n)$ by
1508: $a(n)$. Hence $a(1)=2,a(2)=8,a(n)\leq 2^{\frac{n(n+1)}{2}}$
1509: \cite{Barry}.
1510: 
1511: Define the set $\mathcal{M}(n)=\{U|U\in
1512: \widetilde{\mathcal{C}}_2(n)\setminus\widetilde{\mathcal{P}}_n\cup\{I\}\}$.
1513: 
1514: Now recall the definition for $\mathcal{C}_k(n)$:
1515: \begin{equation}
1516: \mathcal{C}_k(n)=\{U|UP_nU^{\dag}\in \mathcal{C}_{k-1}(n)\}
1517: \end{equation}
1518: 
1519: For any $n$-qubit $U \in \mathcal{C}_k(n)$, the group $G_U(n)$ is
1520: defined by $G_U(n)=U\widetilde{\mathcal{P}}_nU^{\dag}$.
1521: 
1522: Define the set $\mathcal{R}_k(n)=\{U|U\in
1523: \widetilde{\mathcal{C}}_k(n),W^{\dag}=W,Tr(W)=0\}$.
1524: 
1525: And the set $\mathcal{F}_k(n)=\{U|U\in
1526: \widetilde{\mathcal{C}}_k(n)$,$U$ is diagonal$\}$.
1527: 
1528: Denote the group generated by $\{A_i\}_{i=1}^{n}$ by $\langle
1529: \{A_i\}_{i=1}^{n}\rangle$ for any set of operators $A_i$.
1530: 
1531: \subsection*{2. Some facts for calculating $\mathcal{C}_3$ structure}
1532: 
1533: We state some simple facts about $\mathcal{C}_3$ structure which we
1534: use to verify Theorem 3 numerically.
1535: 
1536: \begin{fact}
1537: We could always choose $G_U(n)\subset\mathcal{R}_{k-1}(n)$ for any
1538: $U$ in $\mathcal{C}_k(n)$.
1539: \end{fact}
1540: 
1541: Because we can always choose Hermitian and trace zero elements in
1542: ${\mathcal{P}}_n$ as the representative element for each element in
1543: $\widetilde{\mathcal{P}}_n$.
1544: 
1545: \begin{fact}
1546: If all $n-1$-qubit $\mathcal{C}_k$ gates are semi-Clifford, and if
1547: $G_U(n)\supset \langle \{B_i\}_{i=1}^{n}\rangle$, where $B_i\in
1548: \widetilde{\mathcal{P}}_n$ and $B_i\neq B_j, B_iB_j\neq B_k$ for
1549: $i\neq j\neq k$, then $G_U(n)\cap \widetilde{\mathcal{P}}_n \subset
1550: K_Z(n)$.
1551: \end{fact}
1552: 
1553: Because if $\langle \{B_i\}_{i=1}^{n}\rangle\neq K_Z(n)$, then
1554: $U(n)$ could be reduced to $U(1)\otimes U(n-1)$ via Clifford
1555: operation.
1556: 
1557: \begin{fact}
1558: If $A,B\in \mathcal{M}(n)\cap \mathcal{R}_2(n)$, and $A,B$
1559: correspond to the same element in $\widehat{\mathcal{C}}_2(n)$, then
1560: $AB\in \mathcal{P}_n$.
1561: \end{fact}
1562: 
1563: Because if $A,B$ correspond to the same element in
1564: $\widehat{\mathcal{C}}_2(n)$, then there exists $\alpha\in
1565: \widetilde{\mathcal{P}}_n$ such that $A=\alpha B$.
1566: 
1567: \begin{fact}
1568: For any $n$-qubit $\mathcal{C}_3$ gate $U$, if
1569: $G_U(n)\supseteq\langle \{Z_i\}_{i=1}^{m}\rangle$, where $m\leq n$,
1570: then the quotient group $G_U(n)/\langle \{Z_i\}_{i=1}^{m}\rangle\in
1571: \mathcal{K}(n)$ is Abelian.
1572: \end{fact}
1573: 
1574: For any $n$-qubit $\mathcal{C}_3$ gate $U$, if
1575: $G_U(n)\supseteq\langle \{Z_i\}_{i=1}^{m}\rangle$, where $m\leq n$,
1576: then the quotient group $G_U(n)/\langle \{Z_i\}_{i=1}^{m}\rangle\in
1577: \mathcal{K}(n)$ is Abelian. Because elements of $G_U(n)\in
1578: \widetilde{\mathcal{C}}_2(n)$ are either commute or anticommute, the
1579: corresponding elements in $\widehat{\mathcal{C}}_2(n)$ should
1580: commute.
1581: 
1582: \subsection*{3. $n=1$ case}
1583: 
1584: Since $Sp(2,2)\cong S_3$, $a(1)=2<4$. Hence $G_U(2)\cap
1585: \widetilde{\mathcal{P}}_2$ contains at least one element in
1586: $\widetilde{\mathcal{P}}_1$, i.e. $G_U(1)\cap
1587: \widetilde{\mathcal{P}}_1 \supseteq K_Z(1)$ holds for any single
1588: qubit $\mathcal{C}_3$ gate.
1589: 
1590: Furthermore, it is noted that any $U\in\mathcal{R}_k(1)$ can be
1591: parameterized by
1592: \begin{center}
1593: $U(\theta,\varphi)=\left[
1594: \begin{array}{cc}
1595: \cos{\theta} & \sin{\theta}e^{i\varphi} \\
1596: \sin{\theta}e^{-i\varphi} & -\cos{\theta}
1597: \end{array}
1598: \right]$,
1599: \end{center}
1600: and starting from elements in $\mathcal{R}_2(1)$ and calculate their
1601: eigenvectors, we understand that $\varphi$ can only be of the values
1602: $0,\frac{\pi}{2},\pi,\frac{3\pi}{2}$ for $\cos{\theta}\neq 0$. This
1603: directly leads to the fact that the conjecture is true for any $k$
1604: when $n=1$. See ck.pdf for the details of this.
1605: 
1606: \subsection*{4. $n=2$ case}
1607: 
1608: Since $Sp(4,2)\cong S_6$, $a(2)=8<16$. Hence $G_U(2)\cap
1609: \widetilde{\mathcal{P}}_2$ contains at least one element in
1610: $\widetilde{\mathcal{P}}_2$. However, this is not enough to claim
1611: $G_U(2)\cap \widetilde{\mathcal{P}}_2$ holds for any two-qubit $U$.
1612: We need to examine the structure of $G_U(2)\cap
1613: \widetilde{\mathcal{P}}_2$ in more detail.
1614: 
1615: Consider the maximal Abelian subgroup in $\mathcal{K}(2)$ of order
1616: $8$, and its corresponding elements in
1617: $\widetilde{\mathcal{C}}_2(n)$, direct calculation shows it does not
1618: contain a subgroup of structure $\widetilde{\mathcal{P}}_1\times
1619: Z_2$. Hence we need to further consider Abelian subgroup in
1620: $\mathcal{K}(2)$ of order $4$. Due to lemma 3, we result in
1621: $G_U(2)\cap \widetilde{\mathcal{P}}_2 \supseteq K_Z(2)$ holds for
1622: any two-qubit $\mathcal{C}_3$ gate.
1623: 
1624: Then using Lemma 1 and 2, we could calculate $\mathcal{C}_4(2)$
1625: numerically. The result then shows that all the $\mathcal{C}_3(2)$
1626: gates are semi-Clifford.
1627: 
1628: \subsection*{5. $n=3$ case}
1629: 
1630: Since a(3)=64, and direct calculation of this group shows that not
1631: all the elements could be in $\mathcal{R}_2(3)$, hence $G_U(3)\cap
1632: \widetilde{\mathcal{P}}_3$ contains at least one element in
1633: $\widetilde{\mathcal{P}}_3$. Again, this is not enough to claim
1634: $G_U(3)\cap \widetilde{\mathcal{P}}_3 \supseteq K_Z(3)$ holds for
1635: any three-qubit $U$. We need to examine the structure of $G_U(3)\cap
1636: \widetilde{\mathcal{P}}_3$ in more detail to dig out 2 more elements
1637: in $\widetilde{\mathcal{P}}_3$.
1638: 
1639: Using Facts 1, 2 and 3, we could calculate $\mathcal{C}_3(3)$
1640: numerically. The result shows that the conjecture is also true in
1641: this case. See next subsection for more about $\mathcal{C}_3(3)$.
1642: 
1643: \subsection*{6. Diagonal gates in $\mathcal{C}_3$}
1644: 
1645: Define a diagonal Matrix  $A$ by $A_{jk}=\delta_{jk}e^{i\theta_j}$,
1646: where $j=1,...,N,\ N=2^n$ , for $n$-qubit case.
1647: 
1648: We now prove the following
1649: 
1650: \begin{lemma}
1651: If  $A\in\mathcal{C}_3$, if we choose $A_{11}=1$, then
1652: $A_{jj}=e^{im_j\pi/4}$  for any $j\neq 1$ , where $m_j$  are some
1653: integers.
1654: \end{lemma}
1655: \textbf{Proof}: We first prove for $j=N$ . Note we choose $A_{11}=1$
1656: to get rid of the overall phase of  $A$. Denote  $A'=X^{\otimes
1657: n}AX^{\otimes n}A^{\dagger}$, and  $A''=X^{\otimes n}A'X^{\otimes
1658: n}A'^{\dagger}$. Note  $A'$, $A''$ are also diagonal. Since
1659: $A\in\mathcal{C}_3$,  $A''$ must be in Pauli apart from an overall
1660: phase. And we also have  $A''_{11}=e^{2i\theta_N}$,
1661: $A''_{NN}=e^{-2i\theta_N}$. Hence we must have
1662: $\frac{A''_{11}}{A''_{NN}}=e^{4i\theta_N}=\pm 1$ , i.e.
1663: $\theta=\frac{m_N\pi}{4}$  for some integer  $m_N$.
1664: 
1665: For $j\neq N$ , there always exists a Clifford group operation which
1666: keeps $\ket{j}$ invariant but maps
1667: $\ket{1}\leftrightarrow\ket{N+1-j}$. Hence the above procedure
1668: applies to any  $j\neq N$. $\square$
1669: 
1670: Note the similar idea applies to the diagonal $\mathcal{C}_k$ gates,
1671: i.e. if $A\in\mathcal{C}_k$ , if we choose $A_{11}=1$ ,
1672: $A_{jj}=e^{im_j\pi/2^{k-1}}$ for any $j\neq 1$, where $m_j$  are
1673: some integers.
1674: 
1675: Now we consider some concrete gates:
1676: 
1677: \begin{proposition}
1678: For  $n=3$,  the three qubit diagonal $\mathcal{C}_3$  gates are
1679: given by a group generated by  $\pi/8$ gate, control-phase gate and
1680: control-control-Z gate.
1681: \end{proposition}
1682: 
1683: \textbf{Proof}: The proof is directly given by numerical
1684: calculation, based on Lemma 1.$\square$
1685: 
1686: \begin{references}
1687: 
1688: \bibitem{Nielsen} M. A. Nielsen and I. L. Chuang, {\it Quantum Computation and Quantum Information}, Cambridge University Press, Cambridge, UK, (2000).
1689: 
1690: \bibitem{Preskill} J. Preskill, in {\it Introduction to Quantum Computation and Information}, eds. H. K. Lo, S. Popescu, and T. Spiller, pp. 213-269, World Scientific Publishing Company, (2001).
1691: 
1692: \bibitem{Shor} P. Shor, {\it 37th Symposium on Foundations of Computing}, IEEE Computer Society Press, pp. 56-65, (1996).
1693: 
1694: \bibitem{Gottesman} D. Gottesman, in {\it Encyclopedia of Mathematical Physics}, eds. J.-P. Francoise, G. L. Naber and S. T. Tsou, vol. 4, pp. 196-201, Oxford: Elsevier, (2006).
1695: 
1696: \bibitem{ZCC} B. Zeng, A. W. Cross, and I. L. Chuang, arXiv: 0706.1382.
1697: 
1698: \bibitem{Got} D. Gottesman, and I. L. Chuang, Nature \textbf{402}, 390, (1999).
1699: 
1700: \bibitem{BravyiDistill} Sergei Bravyi and Alexei Kitaev, Phys. Rev. \textbf{A71}, 022316, (2005).
1701: 
1702: \bibitem{Xinlan} X. Zhou, D. W. Leung, and I. L. Chuang, Phys. Rev. \textbf{A62}, 052316, (2000).
1703: 
1704: \bibitem{Gross} D. Gross, M. Van den Nest, arXiv: 0707.4000.
1705: 
1706: \bibitem{Gotthesis} D. Gottesman, Ph.D. thesis, arXiv:quant-ph/9705052.
1707: 
1708: \bibitem{GAP} GAP - Groups, Algorithms, Programming -
1709: a System for Computational Discrete Algebra,
1710: http://www.gap-system.org/.
1711: 
1712: \bibitem{Yao} A. Yao, {\it Annual Symposium on Foundations of Computer Science}, pg. 352, (1993).
1713: 
1714: \bibitem{Mott} M. Mottonen, J. J. Vartiainen, V. Bergholm, and M. M. Salomaa, Quant. Inf. Comp. \textbf{5}, 467 (2005).
1715: 
1716: \bibitem{Vivek} V. V. Shende, S. S. Bullock, I. L. Markov, IEEE Trans. on Computer-Aided Design, vol. 25, no. 6, pp.1000 - 1010, (2006).
1717: 
1718: \bibitem{Griess} R. L. Griess Jr, arXiv:math/0511084.
1719: 
1720: \bibitem{Xie} X. Chen, H. Chuang, A. W. Cross, B. Zeng, and I. L. Chuang, arXiv: 0801.2360.
1721: 
1722: \bibitem{Barry} M. J. J. Barry, J. Austral. Math. Soc. \textbf{A27}, 59, (1979).
1723: 
1724: \end{references}
1725: 
1726: \end{document}
1727: