quant-ph0401127/qnn.tex
1: \documentclass[fleqn]{article}
2: 
3: \usepackage[sumlimits]{amsmath}
4: \usepackage{bbold}                    % this must precede "amsfonts" package
5: \usepackage{amsfonts}
6: \usepackage{amssymb}
7: \usepackage{amsopn}
8: \usepackage{amsthm}                   % AMS {theorem}
9: \usepackage{mathrsfs}                 % Ralph Smith Formal Script fonts
10: \usepackage{titlesec}
11: \usepackage{graphicx}
12: \usepackage{multicol}
13: \usepackage{cite}
14: 
15: %
16: % Margins & fonts _____________________________________________________________
17: %
18: \setlength{\textwidth}{7.0in}
19: \setlength{\textheight}{9.6in}
20: \setlength{\topmargin}{0.2cm}           % originally {2cm}, but arXiv shifts everything vertically
21: \setlength{\leftmargin}{0cm}
22: \setlength{\voffset}{-3cm}
23: \setlength{\hoffset}{-2.9cm}          % instead of {-3.2cm} to squeeze arXiv's margin note
24: \setlength{\mathindent}{\parindent}
25: \setlength{\headheight}{0pt}
26: 
27: \def\rmdefault{ppl}                   % Palatino as default text font
28: \def\sfdefault{phv} %{panr}
29: \titleformat{\section}[hang]
30:  {\large\sffamily\bfseries}
31:  {\thesection}{.5em}{}
32: \titleformat{\subsection}[hang]
33:  {\normalfont\sffamily\bfseries}
34:  {\thesubsection}{.5em}{}
35: \titleformat{\subsubsection}[hang]
36:  {\normalfont\sffamily}
37:  {\thesubsubsection}{.5em}{}
38: \titleformat{\paragraph}[runin]
39:  {\normalfont\sffamily\bfseries}
40:  {}{.5em}{}
41: 
42: %
43: % Common environments _________________________________________________________
44: %
45: \renewenvironment{figure}%
46:  {\par\medskip\refstepcounter{figure}\footnotesize\sffamily\selectfont}%
47:  {\par\medskip}
48: \renewenvironment{table}%
49:  {\par\medskip\refstepcounter{cnt_table}\footnotesize\sffamily\selectfont}%
50:  {\par\medskip}
51: 
52: 
53: %
54: % Operators, abbreviations and symbols ________________________________________
55: %
56: \def\dif{\mathrm{d}}                  % full differential
57: \def\Dif{\mathrm{D}}                  % covariant differential
58: \def\imu{\mathrm{i}}                  % imaginary unit
59: \DeclareMathOperator{\re}{re}         % real part
60: \DeclareMathOperator{\im}{im}         % imaginary part
61: \DeclareMathOperator{\id}{id}         % identity
62: \DeclareMathOperator{\tr}{tr}         % trace
63: \DeclareMathOperator{\rank}{rank}     % rank
64: \DeclareMathOperator{\Div}{div}       % divergence
65: \DeclareMathOperator{\diag}{diag}     % elements of a diagonal
66: \DeclareMathOperator{\dom}{dom}       % domain
67: \DeclareMathOperator{\img}{img}       % image
68: \DeclareMathOperator{\vol}{vol}       % volume
69: \DeclareMathOperator{\Ln}{Ln}         % principal branch of the natural logarithm
70: \DeclareMathOperator{\fix}{fix}       % fixed subset
71: \DeclareMathOperator{\vspan}{span}    % (sub)space spanned by certain vectors
72: \DeclareMathOperator{\supp}{supp}     % support
73: \DeclareMathOperator{\sign}{sgn}      % sign
74: \DeclareMathOperator{\prob}{prob}     % probability
75: \DeclareMathOperator{\prol}{pr}       % field prolongation
76: \DeclareMathOperator{\spct}{sp}       % spectra
77: \DeclareMathOperator{\lin}{lin}       % linear part of an operator/transformation etc.
78: \DeclareMathOperator{\Arctan}{Arctan} % principal branch of arctan
79: \DeclareMathOperator{\Arccos}{Arccos} % principal branch of arccos
80: \DeclareMathOperator{\ord}{ord}       % order
81: \DeclareMathOperator{\argc}{arg}      % argument of complex number
82: \DeclareMathOperator{\Argc}{Arg}      % argument of complex number (principal branch)
83: \DeclareMathOperator{\Int}{int}       % interior
84: \newcommand{\Est}{\mathsf{E}}         % estimate/expectation value
85: \renewcommand{\mod}{\,\mathrm{mod}\,} % standard spaces around `mod' are too wide
86: \newcommand{\e}[1]{\mathrm{e}^{#1}}   % short form of exponent function
87: \newcommand{\df}[1]{{\rm\textbf{#1}}} % term being defined
88: \newcommand{\term}[1]{\textit{#1}}    % textbook term recalled
89: \newcommand{\rhs}{\textsc{rhs}}       % "right-hand side"
90: \newcommand{\lhs}{\textsc{lhs}}       % "left-hand side"
91: %newcommand{\ode}[1]{\textsc{ode}#1}  % "ordinary differential equation[s]"
92: %newcommand{\pde}[1]{\textsc{pde}#1}  % "partial differential equation[s]"
93: \newcommand{\wrt}{with respect to}    % "with respect to"
94: \newcommand{\dof}[1]{\textsc{dof}#1}  % "degree[s] of freedom"
95: \newcommand{\const}{\mathrm{const}}   % a constant
96: \newcommand{\ie}{\textit{i.e.}}       % "id est"
97: \newcommand{\eg}{\textit{e.g.}}       % "exempli gratia"
98: \newcommand{\cf}{\textit{cf.}}        % "confer" (compare)
99: \newcommand{\etc}{\textit{etc.}}      % "et cetera"
100: \newcommand{\etal}{\textit{et al.}}   % "et al"
101: \newcommand{\acr}[1]{{\text\rm\small{#1}}} %acronyms
102: \newcommand{\bra}[1]{{\langle{#1}|}}  % bra
103: \newcommand{\ket}[1]{{|{#1}\rangle}}  % ket
104: \newcommand{\braket}[2]{{\langle{#1}|{#2}\rangle}}
105: \newcommand{\ketbra}[2]{{|{#1}\rangle\langle{#2}|}}
106: %newcommand{\Bra}[1]{{\left\langle{#1}\right|}}
107: %newcommand{\Ket}[1]{{\left|{#1}\right\rangle}}
108: %newcommand{\Braket}[2]{{\left\langle{#1}\big|{#2}\right\rangle}}
109: \newcommand{\avr}[1]{\langle{#1}\rangle}
110: \def\half{\frac{1}{2}}                % 1/2
111: \def\thalf{\tfrac{1}{2}}              % 1/2 (tiny)
112: \def\iff{\Leftrightarrow}             % "if and only if" (<=>)
113: \def\implies{\Rightarrow}             % implication (=>)
114: \def\implied{\Leftarrow}              % implication (<=)
115: \def\idmatrix{\text{\textbb 1}}       % identity matrix (from "bbold" package)
116: \def\nullmatrix{\text{\textbb 0}}     % null matrix (from "bbold" package)
117: \def\till{..}                         % range 1..n
118: \def\Till{,\!...,}                    % enumeration scope a,...,z
119: \def\fK{\mathbb{K}}                   % field K (any)
120: \def\fN{\mathbb{N}}                   % natural numbers N
121: \def\fZ{\mathbb{Z}}                   % integer numbers Z
122: \def\fQ{\mathbb{Q}}                   % field of rational numbers C
123: \def\fR{\mathbb{R}}                   % field of real numbers R
124: \def\fC{\mathbb{C}}                   % field of complex numbers C
125: \def\fH{\mathbb{H}}                   % quaternion numbers H
126: \def\Lie{\mathscr{L}}                 % Lie derivative
127: \def\intprod%                         % interior product
128:  {{\unitlength 1pt\begin{picture}(9,8)(-7,0)\line(0,1){6}\line(-1,0){6}\end{picture}}}
129: \newcommand\roplus%                   % right semi-simple sum
130:  {{\unitlength 1pt\begin{picture}(11,8)(0,0)\linethickness{.3pt}%
131:   \put(5.0,-.3){\line(0,1){5.4}}\put(2,0){$\ni$}%
132:  \end{picture}}}
133: \newcommand\loplus%                   % left semi-simple sum
134:  {{\unitlength 1pt\begin{picture}(11,8)(0,0)\linethickness{.3pt}%
135:   \put(5.3,-.3){\line(0,1){5.4}}\put(2,0){$\in$}%
136:  \end{picture}}}
137: \def\T#1{\mathrm{T}_{{#1}}}           % tangent space
138: \def\cT#1{\mathrm{T}^*_{{#1}}}        % cotangent space
139: \def\grG{G}                           % group G (arbitrary)
140: \def\grGL{\mathrm{GL}}                % classical groups: GL
141: \def\grL{\mathrm{L}}                  % - L (Lorentz)
142: \def\grO{\mathrm{O}}                  % - O
143: \def\grU{\mathrm{U}}                  % - U
144: \def\grSO{\mathrm{SO}}                % - SO
145: \def\grSp{\mathrm{Sp}}                % - Sp (symplectic)
146: \def\grSpin{\mathrm{Spin}}            % - Spin
147: \def\grSU{\mathrm{SU}}                % - SU
148: \def\grSL{\mathrm{SL}}                % - SL
149: \def\grS{\mathrm{S}}                  % - symmetric (permutations)
150: \def\grDiff{\mathrm{Diff}}            % other groups: diffeomorphisms
151: \def\grHol{\mathrm{Hol}}              % - holomorphic functions
152: \def\grHom{\mathrm{Hom}}              % - homomorphisms
153: \def\grAut{\mathrm{Aut}}              % - automorphisms
154: \def\grEnd{\mathrm{End}}              % - endomorphisms
155: \def\grGr{\mathrm{Gr}}                % - Grassmanian
156: \def\grSym{\mathrm{S}}                % - symmetric of permutations
157: \def\grAff{\mathrm{Aff}}              % - affine transformations
158: \def\alGL{\mathfrak{gl}}              % classical algebras: gl
159: \def\alL{\mathfrak{l}}                % - l
160: \def\alO{\mathfrak{o}}                % - o
161: \def\alU{\mathfrak{u}}                % - u
162: \def\alSO{\mathfrak{so}}              % - so
163: \def\alSp{\mathfrak{sp}}              % - sp
164: \def\alSpin{\mathfrak{spin}}          % - spin
165: \def\alSU{\mathfrak{su}}              % - su
166: \def\alSL{\mathfrak{sl}}              % - sl
167: \def\alAff{\mathfrak{aff}}            % - affine
168: 
169: %
170: % Formating macros ____________________________________________________________
171: %
172: 
173: \renewcommand{\thetable}{\arabic{cnt_table}}
174: \newcommand{\figref}[1]{Fig.~\ref{#1}}
175: \newcommand{\tabref}[1]{Table~\ref{#1}}
176: \newcommand{\scref} [1]{Sect.~\ref{#1}}
177: \renewcommand{\caption}[1] {\par{\footnotesize\sffamily Figure \thefigure{}: #1}}
178: \newcommand{\fignumber}    {\par\nopagebreak{\textbf{Figure \thefigure{}.}}}
179: \newcommand{\figcaption}[1]{\par\nopagebreak{\textbf{Figure \thefigure{}} #1}}
180: \newcommand{\tabcaption}[1]{{\textbf{Table \thetable{}}: #1}\par\nopagebreak}
181: 
182: \newcommand\publ[1]  {{#1}}
183: \newcommand\eprint[1]{{\tt #1}}
184: \def\refno#1         {\noindent{\footnotesize #1}\par\noindent\rule[5mm]{\textwidth}{0.1mm}\par}
185: \newenvironment{articlehead}%
186:  {\begin{quote}\setlength{\parskip}{0pt}\setlength{\parindent}{0pt}}%
187:  {\end{quote}}
188: \def\title#1         {{\sffamily\Large\selectfont\textbf{#1}}\\\par\medskip}
189: \renewcommand{\author}[1]{{\textsc{#1}}}
190: \def\authore#1#2     {\textsc{#1}\footnote{e-mail: {#2}}\null}
191: \newcommand\authora[2]{\textsc{#1}$^{#2}$}
192: \def\authorae#1#2#3  {\textsc{#1}$^{#2,}$\footnote{e-mail: {#3}}\null}
193: \def\address#1       {\par{\it #1}}
194: \renewenvironment{abstract}%
195:  {\par\medskip\small\begin{center}\textbf{Abstract}\end{center}\par~~~}%
196:  {\par}
197: \def\keywords#1      {\par{\bf Keywords:~}{#1}\par}
198: \def\pacs#1          {\par{\bf PACS:~}{#1}\par}
199: \def\subjclass#1     {\par{\bf Subj. class.:~}{#1}\par}
200: 
201: \newcommand\TRASH[1]{}                % text to be ignored
202: 
203: \def\Gamma{\varGamma}
204: \def\Theta{\varTheta}
205: \def\Lambda{\varLambda}
206: \def\Xi{\varXi}
207: \def\Pi{\varPi}
208: \def\Sigma{\varSigma}
209: \def\Upsilon{\varUpsilon}
210: \def\Phi{\varPhi}
211: \def\Psi{\varPsi}
212: \def\Omega{\varOmega}
213: \def\rho{\varrho}
214: 
215: % RevTex compatibility
216: \def\leq{\leqslant}
217: \def\alt{\leqslant}
218: \def\le {\leqslant}
219: \def\geq{\geqslant}
220: \def\agt{\geqslant}
221: \def\ge {\geqslant}
222: 
223: \def\SD   {\textrm{\small SD}}
224: \def\QM   {\textrm{\small QM}}
225: \def\QC   {\textrm{\small QC}}
226: \def\HAD  {\textrm{\small H}}
227: \def\NOT  {\textrm{\small NOT}}
228: \def\CNOT {\textrm{\small CNOT}}
229: \def\POVM {\textrm{\small POVM}}
230: \def\DOF  {\textrm{\small DOF}}
231: %def\XOR  {\textrm{\small XOR}}
232: 
233: \DeclareMathOperator{\cov}{cov}
234: %DeclareMathOperator{\Int}{int}
235: \DeclareMathOperator{\Ext}{ext}
236: \DeclareMathOperator{\erf}{erf}
237: \newcommand\scprod[2]{\langle{#1},{#2}\rangle}
238: 
239: \def\spH{\mathscr{H}}
240: \def\spP{\varOmega}
241: \def\fT#1{\mathbb{T}^{#1}}
242: \def\fCP#1{\mathbb{C}P^{#1}}
243: \def\grG{\mathscr{G}}
244: \def\stV{\mathscr{V}}
245: \def\stU{\mathscr{U}}
246: \def\stE{\mathscr{E}}
247: 
248: \def\qbs{\hat{\sigma}}          % Clifford basis (1+Pauli)
249: \def\ebs{\hat{e}}               % event basis
250: \def\vbs{\hat{v}}               % vertex basis
251: \def\vbsl#1{\vbs_{\mathrm{#1}}} % labeled vertex
252: \def\Xl#1{X^{\mathrm{#1}}}      % labeled spike
253: 
254: \def\tGate{\tau_\mathrm{gate}}
255: \def\tAvr {\tau_\mathrm{avr}}
256: \def\tSig {\tau_\mathrm{sig}}
257: 
258: % synapse-type symbols
259: \newcommand\synInh{\unitlength 0.5pt%
260:  \begin{picture}(18,14)(0,1)%
261:   \put(0,5){\line(1,0){11}}\put(14,5){\circle*{6}}%
262:  \end{picture}}
263: \newcommand\synExc{\unitlength 0.5pt%
264:  \begin{picture}(18,14)(0,1)%
265:   \put(0,5){\line(1,0){11}}\thicklines\put(5,5){\vector(-1,0){0}}%
266:  \end{picture}}
267: \newcommand\synGat{\unitlength 0.5pt%
268:  \begin{picture}(18,14)(0,1)%
269:   \put(0,5){\line(1,0){14}}\put(14,-1){\line(0,1){12}}%
270:  \end{picture}}
271: \newcommand\synMod{\unitlength 0.5pt%
272:  \begin{picture}(18,14)(0,1)%
273:   \put(0,5){\line(1,0){11}}\put(14,5){\circle{6}}%
274:  \end{picture}}
275: 
276: \begin{document}
277: %refno{\today}
278: \begin{articlehead}
279: \begin{raggedright}
280: \title{Quantum computing in neural networks}
281: \author{P. Gralewicz}
282: \address{Department of Theoretical Physics, University of \L\'{o}d\'{z},
283:   Pomorska 149/153, 90-236, \L\'od\'z, Poland}
284: \end{raggedright}
285: 
286: \begin{abstract}
287: According to the  statistical interpretation of quantum theory, quantum
288: computers form a distinguished class of probabilistic machines (PMs) by encoding
289: $n$ qubits in $2n$ pbits (random binary variables). This raises the possibility
290: of a large-scale quantum computing using PMs, especially with neural networks
291: which have the innate capability for probabilistic information processing.
292: Restricting ourselves to a particular model, we construct and numerically
293: examine the performance of neural circuits implementing universal quantum gates.
294: A discussion on the physiological plausibility of proposed coding scheme is also
295: provided.
296: \end{abstract}
297: \end{articlehead}
298: \begin{multicols}{2}
299: 
300: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
301: \section{Introduction}
302:   Neural networks are naturally evolved systems for information processing.
303: Despite decades of experimental and theoretical research, there is no agreement
304: upon the information encoding employed by these circuits -- the
305: problem of what exactly is being communicated \textit{via} seemingly chaotic
306: spike trains is still largely open \cite{RWRB1997}. Advancement in understanding
307: of this neural language is obstructed by variety of cell types, working
308: conditions and molecular factors to be taken into account \cite{ChS1992}.
309: Generally accepted schemes, the \term{rate code} and the \term{phase code}, may
310: turn out to be only the first two in a sequence of progressively more intricate
311: codes, where higher order correlations within cellular complexes are utilized.
312: 
313:   Quantum information science, on the other hand, had matured over the last two
314: decades making significant contributions to both information theory and quantum
315: mechanics (\QM). The latter, having historical roots in particle physics, is
316: still often identified with the micro-world. Yet, there is nothing in the
317: mathematical foundations of {\QM} which could justify that point of view. In
318: fact, apart from that microscopic realizations, quantum theory has found
319: many avatars, from mechanical \cite{A1984}, linguistic \cite{AC2003}, purely
320: geometric \cite{B2004}, to statistical \cite{Bal70,P1995,PeTe98,CFS2002}. In
321: this article, that last, widely accepted interpretation, is being used to study
322: the feasibility of a hypothesis that spike trains may actually encode for
323: quantum states.
324: 
325:   Such hypothesis appears particularly attractive in that the Nature is notorious
326: in repeating itself at various scales, and if quantum computing (\QC) proves to
327: be practical, it would be rather surprising if one could not find it implemented
328: at a higher level. From this point of view neural networks are the obvious
329: candidates for such implementations. By examining two neural circuits,
330: designed to perform quantum operations ($1$-qubit rotations, and $2$-qubit
331: {\CNOT} gate), we demonstrate the feasibility of our hypothesis within the
332: limits of a simple model. Although quantum registers are realized efficiently
333: with just two neurons per qubit, the major costs are in the processing of
334: information carried by the spike trains. The simulations provided are intended
335: to emphasize the amount of these resources as well as the functionality required
336: for implementation.
337: 
338:   We begin with a short review of the formalism which allows for the
339: identification of pairs of spiking neurons with qubits. In Section \ref{sc:ann}
340: a reduced model of neural network is described, which in Sec. \ref{sc:q-gates}
341: is further used as a basis for construction of quantum gates. The results of
342: simulations, in terms of achieved fidelity and coherence, are promising enough
343: to look toward more realistic implementations. We touch briefly on these issues
344: in the last section.
345: 
346: 
347: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
348: \section{Manipulation of quantum states embedded in probabilistic space}
349: \label{sc:povm}
350: The operational approach to quantum mechanics, through the formalism of positive
351: operator-valued measures ({\POVM}s), allows one to
352: express the states of a quantum system defined in a finite-dimensional
353: Hilbert space $\spH$, in terms of probability distributions. If the dimension
354: is $d:=\dim\spH$, then a generic density matrix $\hat{\rho}$ representing the
355: state has $d^2-1$ degrees of freedom ({\DOF}s). A distribution obtained through
356: particular {\POVM} has length $d^2$, and -- due to normalization constraint --
357: the same number of {\DOF}s as the density matrix \cite{PeTe98}. For $n$-qubit
358: states, this distribution can be associated with joint probability of $2n$
359: binary random variables\footnote{
360:   This is in close analogy to complex numbers which extend the reals, and at the
361: same time are embeddable in a real vector space of doubled dimension equipped
362: with complex structure.}.
363: 
364: %-------------------------------------------------------------------------------
365: Let $\hat{\rho}$ be a generic density matrix of a $1$-qubit state, which
366: using summation convention, we write as
367: \[
368:   \hat{\rho}
369:    := \rho^\mu\,\qbs_\mu
370:     = \begin{pmatrix}
371:         \thalf + \rho^3 & \rho^1 + \imu \rho^2 \\
372:         \rho^1 - \imu \rho^2 & \thalf - \rho^3
373:       \end{pmatrix},
374: \]
375: where $\qbs_0=\idmatrix$, $\qbs_{1,2,3}$ are the Pauli matrices,
376: $\rho^0\equiv \thalf$, and $\rho^1,\rho^2,\rho^3\in[-\thalf,\thalf]$ are the
377: three real coordinates of a Bloch vector.
378: Let $\{\hat{A}^{z}\}$ be a normalized $4$-element positive operator-valued
379: measure %(\POVM)
380: \begin{equation}
381:  \label{eq:Anorm1}
382:   \sum_{z} \hat{A}^{z} = \idmatrix,
383:   \qquad z=0\Till 3.
384: \end{equation}
385: Typically, one associates such a {\POVM} with the Pauli basis, that is
386: \[
387:   \hat{A}^{z}
388:    := A^z\null_\mu \qbs^\mu
389:     = \tfrac{1}{2}\idmatrix + A^z\null_i \qbs^i,\qquad
390:   i=1\Till 3.
391: \]
392: where $A^z\null_0\equiv \tfrac{1}{2}$, and $\qbs^\mu:=\qbs_\mu^\dagger=\qbs_\mu$,
393: is the basis dual with respect to the scalar product
394: $
395:   \scprod{\qbs^\mu}{\qbs_\nu}
396:    :=\thalf\tr[\qbs^\mu\qbs_\nu]
397:     =\delta^\mu\null_\nu
398: $.
399: Although not a strict necessity, it is reasonable to assume the same {\POVM} for
400: all $n$ qubits within a register, and consequently take the entire {\POVM} as a
401: $n$-fold tensor product
402: \[
403:   \hat{A}^{z_1...z_n}
404:     := \hat{A}^{z_1}\otimes\cdots\otimes\hat{A}^{z_n}.
405: \]
406: This leads to the following distribution
407: \[
408:  \hspace{-5mm}
409:  \begin{aligned}
410:   \null&p^{z_1...z_n}
411:    := \scprod{\hat{A}^{z_1...z_n}}{\hat{\rho}}
412:     = 2^{-2n} + A^{z_1}\null_{i_1}\cdots A^{z_n}\null_{i_n}\rho^{i_1...i_n},\\
413:   \null&\qquad \sum_{z_1...z_n} p^{z_1...z_n} = 1.
414:  \end{aligned}
415: \]
416: Introducing the event basis $\{\ebs_{z}\}$, the transformation can concisely be
417: written as
418: \begin{equation}
419:  \label{eq:rho_p}
420:   \hat{p} = A \hat{\rho},
421: \end{equation}
422: where
423: \[
424:  \hspace{-5mm}
425:  \begin{aligned}
426:   \hat{p} &= p^{z_1...z_n}\,\ebs_{z_1}\otimes\cdots\otimes\ebs_{z_1},\\
427:   A &= A^{z_1}\null_{\mu_1}\cdots A^{z_n}\null_{\mu_n}\,
428:       \ebs_{z_1}\otimes\cdots\otimes\ebs_{z_1}\otimes
429:       \qbs^{\mu_1}\otimes\cdots\otimes\qbs^{\mu_n}.
430:  \end{aligned}
431: \]
432: Conversely, if $\{\hat{A}^z\}$, are linearly independent, then one can invert
433: the relation \eqref{eq:rho_p} and take the distribution $\hat{p}$ as an
434: equivalent representation of the quantum state $\hat{\rho}=A^{-1}\hat{p}$.
435: 
436: A unitary transformation $U\in\grU(2^n)$ of the state is a linear
437: operator\footnote{
438:   Note, that the embedding allows to consider a wider range of isometries to
439:   be implemented, not only the ones corresponding to unitary operations.
440:   For instance the $1$-qubit antipode (unfortunately also called the
441:   quantum universal-NOT) can only be approximated in unitary QM
442:   \cite{BHW1999,MBSS2002}. In probabilistic approach one can realize it exactly.
443: } $L\in 1\oplus\grSO(2^{2n}-1)$
444: \[
445:   \hat{\rho}
446:    \mapsto U^\dagger\!\hat{\rho}\,U
447:    = L\hat{\rho},
448: \]
449: with elements
450: \begin{equation}
451:  \label{eq:L}
452:  \begin{aligned}
453:   \null&L^{\mu_1...\mu_n}\null_{\nu_1...\nu_n}\\
454:   \null&\quad
455:    = \scprod{\qbs^{\mu_1}\otimes\cdots\otimes\qbs^{\mu_n}}
456:             {U^\dagger\qbs_{\nu_1}\otimes\cdots\otimes\qbs_{\nu_n}U}.
457:  \end{aligned}
458: \end{equation}
459: After transformation of the basis $A^{-1}:\{\qbs_\mu\}\to\{\ebs_{z}\}$ one has
460: the same operation acting on probability distribution
461: \begin{equation}
462:  \label{eq:Lp}
463:   \hat{p} \mapsto (A L A^{-1}) \hat{p}.
464: \end{equation}
465: There is, however, an important difference between the linear dynamics
466: of Eq. \eqref{eq:Lp} and Markovian transitions usually considered in association
467: with stochastic evolution: Denote by $\Omega^{2n}$ the space of joint
468: probability distributions of $2n$ pbits. Since the operator $A L A^{-1}$ is by
469: definition invertible, it follows that, in general, it is not a positive one,
470: hence only a subset of $\Omega^{2n}$ will be mapped back into itself. We denote
471: this subset -- the (closure of) positive domain of quantum operators -- by
472: \[
473:   \overline{\Omega^{2n}_+}:=\{\hat{p}\in\Omega^{2n}\mid A L A^{-1}\hat{p}\in\Omega^{2n}\}.
474: \]
475: This is simply the image of all quantum states under the {\POVM} $A$. The
476: boundary $\Omega^{2n}_0 := \partial\overline{\Omega^{2n}_+}$, which is the image
477: of the Bloch sphere in $\Omega^{2n}$ contains pure states, while its interior
478: $\Omega^{2n}_+ := \overline{\Omega^{2n}_+}\setminus\Omega^{2n}_0$
479: is the subset of mixed states. All remaining distributions
480: $\Omega^{2n}_- := \Omega^{2n}\setminus \overline{\Omega^{2n}_+}$
481: are mapped by $A^{-1}$ to the exterior of the Bloch sphere. Therefore, the
482: {\POVM} partitions the set of possible distributions into three disjoint
483: subsets:
484: 
485: \par\medskip\begin{tabular}{ll}
486:  $\Omega^{2n}_0$                     & -- pure quantum states \\
487:  $\Omega^{2n}_+ = \Int\Omega^{2n}_0$ & -- mixed/decohered states\\
488:  $\Omega^{2n}_- = \Ext\Omega^{2n}_0$ & -- overcohered states
489: \end{tabular}\par\medskip
490: 
491: To explain the term \term{overcohered} used above, let us take a closer look
492: at the limitations imposed by the {\POVM} on distributions in
493: $\overline{\Omega^{2n}_+}$. Positivity of $A$ implies, that the probabilities
494: are bound by
495: \[
496:   p^{z_1...z_n}\leq 2\avr{p}=2(A^z\null_0\rho^0)^{n}=2^{1-2n}.
497: \]
498: Furthermore, if, as we assume, $A$ is non-degenerate, then for any quantum state
499: only one of the elements $\{p^{z_1...z_n}\}$ can either vanish, or reach the
500: maximal value $2\avr{p}$. This means, that there is a non-zero lower bound
501: on the entropy of distributions in $\overline{\Omega^{2n}_+}$, and hence no
502: distribution with certain outcome can represent a quantum state. Moreover, all
503: single-pbit marginals are non-vanishing.
504: 
505:   A quantitative characterization of the \term{coherence} can be given by the
506: radius of the state's Bloch vector. The metric
507: $g:\Omega^{2n}\times\Omega^{2n}\to\fR$ induced by the
508: {\POVM} on the distribution space permits to obtain this radius directly for
509: an arbitrary $\hat{p}\in\Omega^{2n}$. Let
510: $\hat{p}_1,\hat{p}_2\in\overline{\Omega^{2n}_+}$, and $\hat{\rho}_1$,
511: $\hat{\rho}_2$ be corresponding quantum states. Then $g$ is given by
512: \[
513:   g(\hat{p}_1,\hat{p}_2)
514:    := \tr[\hat{\rho}_1^\dagger\hat{\rho}_2]
515:     = (A^{-1}\hat{p}_1)^\dagger\cdot(A^{-1}\hat{p}_2).
516: \]
517: Because this is a bilinear map with coefficients independent of
518: $\hat{p}_1,\hat{p}_2$, one is free to extend its domain onto the entire space
519: $\Omega^{2n}$. Since $A^0\null_\mu=\thalf$, the radius is
520: \[
521:   r^2 := \|\vec{\rho}\,\|^2 = 2^{-n} g(\hat{p},\hat{p})-2^{2n}.
522: \]
523: In particular, for a pure state
524: \[
525:   r_\mathrm{pure}^2 = 2^{-n}(1-2^{-n}),
526: \]
527: and the Bloch radius of any mixed state is always bound by $r<r_\mathrm{pure}$.
528: The ratio
529: \begin{equation}
530:  \label{eq:defR}
531:   R := \frac{r}{r_\mathrm{pure}},
532: \end{equation}
533: can be adopted for a measure of coherence -- ranging from $R=0$ for
534: maximally decohered state $\hat{\rho}=2^{-n}\idmatrix$, through $R=1$ for any
535: pure $\hat{\rho}=\hat{\rho}^2$, and beyond $R>1$ for all overcohered ones.
536: 
537:   In order to quantify the performance of circuits considered latter in this
538: article, we will also employ another, independent measure
539: by which one can estimate the angular disparity between expected and obtained
540: states. The \term{fidelity} or normalized overlap between
541: $\hat{p},\hat{q}\in\Omega^{2n}$ is defined here as
542: \[
543:   F := \frac{g(\hat{p},\hat{q})}{\sqrt{g(\hat{p},\hat{p})}\sqrt{g(\hat{q},\hat{q})}}.
544: \]
545: 
546:   We choose fidelity as a commonly adopted measure, for the purpose of
547: comparison, notwithstanding direct estimate of the unitary error between
548: the desired pure state $\hat{p}\in\Omega^{2n}_0$ and obtained distribution
549: $\hat{q}\in\Omega^{2n}$, which is readily computable:
550: \[
551:  \alpha = \arccos\left(\frac{g(\hat{p},\hat{q})-2^{-n}}{r(\hat{q})\sqrt{2^{n}-1}}\right).
552: \]
553: The two quantities $\alpha$ and $F$, are nevertheless dependent.
554: 
555: 
556: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
557: \section{A toy-model neural-network}
558: \label{sc:ann}
559:   The information in neural networks is carried by spike trains, which after
560: appropriate discretization can be transformed to binary strings. The model
561: network described in this section is a much simplified version of what usually
562: is considered realistic -- the purpose of such reduction is to retain only the
563: essential features.
564:   Consistently with discretization of transmitted signals, the model operates
565: in explicitly step-wise manner, instead of continuous-time evolution. Likewise,
566: the delays effected along the inter-neuron paths are also taken to be
567: integers.
568: %\footnote{
569: %  This latter feature can, to some extent be justified by the physiology of
570: % neural connections: The propagation speed of action potentials along
571: % myelin-sheathed and unsheathed parts of the axon differ by at least one order
572: % of magnitude. The introduced delays are therefore dominated by the number of
573: % unsheathed segments which is roughly an integer.
574: %}.
575: 
576:   Let $\grG=(\stV,\stE)$, be a multiply-connected digraph, where
577: $\stV=\{\vbs_i\}$ is the vertex basis of neurons, (we shall also write
578: $\stV_{N}$ to explicit the number $N$ of vertices involved) and
579: $\stE=\stV\otimes\stV\otimes\fN$ is the basis of edges, that is the
580: possible synaptic connections. The actual couplings between $i^\mathrm{th}$ and
581: $j^\mathrm{th}$ neuron are set by the weights $W^i\null_{js}$ where $s\in\fN$
582: enumerates the delays introduced along multiple edges. For each vertex we define
583: two variables: the binary \term{output state} $X^i\in\{0,1\}$, and the
584: \term{residual potential} $u^i\in\fR$.
585: 
586:   We adopt the discrete \term{integrate-and-fire} scheme for the dynamics of
587: this network. In each time step the potential is first updated by accumulating
588: the incoming signals
589: \[
590:   u^{i,t-1}\mapsto u^{it}_\star := u^{i,t-1} + W^i\null_{js}X^{j,t-s}
591: \]
592: where the summation runs over connected vertices ($j$) and edge delays
593: ($s\geq 1$).
594: %This `excited' value is a source of spike generation --
595: %The probability of spike generation depends on this `excited' value
596: Subsequent spike generation ($X^{it}=1$) occurs with probability
597: $P(u^{it}_\star)$, where $P:\fR\to[0,1]$, is a `noisy'
598: activation function with firing threshold fixed at $u_\mathrm{thr}=\thalf$.
599: Its actual form used in simulations is given by
600: \[
601:   P(u_\star) := \half\left(1+\erf\frac{u_\star - u_\mathrm{thr}}{\sigma}\right),
602: \]
603: where $\sigma\geq 0$ is a global control parameter characterizing the noise
604: standard deviation ({\SD}). In particular, in the limit $\sigma\to 0$ the spikes
605: are produced deterministically, as $P$ becomes a step-function. The excited
606: state $u^{it}_\star$ is eventually reduced by release of a spike (refractory
607: potential), and further quenched with a bound, nonlinear map $S$
608: \[
609:   u^{it}_\star \mapsto u^{it} = S(u^{it}_\star- X^{it}).
610: \]
611: We assume $S$ to have an attractive fixed point at the origin (the resting
612: potential),
613: $
614:   \forall_u:\;\lim_{t\to\infty} S^t(u)=0
615: $,
616: to be linear in its neighborhood
617: $
618:   S'(0) = 1
619: $,
620: and having finite, but non-zero asymptotes
621: $
622:   |S(\pm\infty)|<\infty
623: $.
624: The motivation for introduction of this mapping is twofold:
625:   First, the physiological mechanisms of signal transmission imply existence of
626: \term{saturations} in both positive and negative direction. The cell can be
627: depolarized or hyperpolarized through synaptic channels only to certain extent,
628: and adding more excitatory or inhibitory connections will not have a significant
629: effect.
630:   Second, the reason to have $u=0$ for an attractive fixed point, is to
631: imitate the `leaky' integration scenario, by which in the absence of input
632: the potential returns back to its resting point.
633:   In the simulations this function was taken to be a simple, skew-symmetric
634: mapping
635: \[
636:   S(u) := \gamma\tanh\frac{u}{\gamma},\qquad \gamma \geq 0.
637: \]
638: Here, the asymptotes are $S(\pm\infty)=\pm\gamma$, therefore we call $\gamma$
639: the `saturation parameter'. If we assume, that the neuron is left without input and
640: some residual $u$, so that no spikes are generated, then the potential $u$ will
641: decay sub-exponentially in time, as
642: \[
643:   u \mapsto S^t(u) \underset{t\gg 1}{\to} \frac{\gamma u}{\sqrt{\gamma^2 + \tfrac{2}{3} t u^2}},
644: \]
645: where $S^t$ means $t$-fold composition.
646: In the limit $\gamma\to 0$, the residual potential is reset to zero after each
647: cycle, and this situation can be associated with time steps longer than the
648: total refractory time ($\sim 20\,\mathrm{ms}$), within which the cell relaxes to
649: its resting point. If $\gamma>0$, then the probability of a consecutive spike is
650: modified by the residual potential: The cell is within the \emph{relative}
651: refractory period, when the the potassium channels are still open, but the
652: sodium gates are already reverted to their normal state. This mode corresponds
653: to time steps of order $\sim 5\,\mathrm{ms}$. Shorter times are generally
654: unrealistic due to high suppression of spike generation during the
655: \emph{absolute} refractory period, when the sodium channels are closed.
656: 
657: %Fixing the time step to the duration of absolute refractory period
658: %$\sim 1\,\mathrm{ms}$, the neurons are still operating within the relative
659: %refractory period, before the resting potential is reached. This modifies
660: 
661:   The choice of a specific value of $\gamma$ is therefore indirectly related to
662: the time scale, and consequently to the discretization window of action
663: potentials. If this window is too short, the discretization becomes
664: ambiguous and the model breaks down -- this is another reason not to
665: consider high saturation values.
666: 
667: %Another comes from the mechanisms of spike
668: %generation, by which multiple spikes within the refractory period are
669: %suppressed.
670: 
671:   The qualitative behavior of the above model is best understood by analyzing
672: single neuron at the limits of the two control parameters $\sigma$, and $\gamma$.
673: Assume the cell is fed with a stimulus at a constant frequency
674: $\nu_\mathrm{in}\in[0,1]$, and consider at first the noiseless regime $\sigma=0$.
675: If $\gamma=0$, then the only memory of past
676: input values is stored in delayed connections. The cell fires only if the value
677: of the convolution $W_{js}X^{j,t-s}$ exceeds the threshold $u_\mathrm{thr}$.
678: Such neurons acts like a high-pass filter and its firing rate is
679: $\nu_\mathrm{out}=P(\nu_\mathrm{in}\sum_{s,j} W_{js})$. By increasing the noise
680: {\SD} $\sigma$, the shape of this filtering function changes along with the
681: spiking probability $P$, nevertheless it never becomes close to an ideal
682: multiplier -- the response is always nonlinear.
683: 
684:   If $\gamma\to\infty$ the cell accumulates and `remembers' the residual value
685: of convolution left over after subtraction of generated spikes. This makes it
686: into a perfect multiplier with spike rate
687: $\nu_\mathrm{out}=\nu_\mathrm{in}\sum_{s,j} W_{js}$. Raising $\sigma$ above zero
688: does not change this average response, but the determinism initially apparent in
689: the spike patterns is gradually being washed away.
690: 
691:   In between of these two regimes, lies a surprisingly complex area of
692: fractal-spaced frequency thresholds and output patterns, particularly
693: conspicuous at $\sigma=0$ and $\nu_\mathrm{in}=1$. Presence of these features,
694: found in many nonlinear deterministic systems do not critically depend on the
695: specific shape of the function $S$.
696: %The system itself can exhibit a very rich dynamics and is worth a separate study.
697: 
698: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
699: \section{Implementation of universal quantum gates}
700: \label{sc:q-gates}
701:   According to the discussion provided in section \ref{sc:povm}, one needs $2n$
702: random binary variables to implement an $n$-qubit register. In our model
703: of the neural network, these variables are identified with discretized spikes
704: registered at $2n$ network sites. The question we set up to address in this
705: section is, whether there are circuits which can implement state-independent
706: rotations of the joint probability distributions, that is -- quantum gates.
707: 
708:   The set of gates universal for quantum computation \cite{BBCVMSSSW1995}
709: includes the whole algebra of $1$-qubit rotations, and an arbitrary $2$-qubit
710: entangling gate, typically chosen to be the {\CNOT} (controlled-{\NOT}).
711: Although probabilistic encoding of qubits is efficient ({\ie} linear in $n$),
712: manipulation of their $2^{2n}$ degrees of freedom ({\DOF}s), by definition
713: requires exponential amount of resources. From this perspective the construction
714: of circuits described below should appear at least conceptually straightforward:
715: The space of binary functions over the vertices $\stV_{2n}$ is
716: $\stV^{2n}_*=\fZ_{2n}$. We first embed an element $X^t=\{X^{it}\}\in\stV^{2n}_*$
717: into $\Omega^{2n}$, then apply the gate $G := A L A^{-1}$, and finally project
718: the result back onto $\stV^{2n}_*$. The entire quantum gate transforming one
719: set of spike trains $X^t$ to another $Y^t\in\stV^{2n}_*$, is then a composition
720: \[
721:   \Pi\circ G\circ\Pi^{-1}:\; X^t \mapsto Y^t,
722: \]
723: where $\Pi^{-1}:\stV^{2n}_*\to\Omega^{2n}$, and $\Pi\circ\Pi^{-1}=\id_{\stV^{2n}_*}$.
724: The main problem in this approach is to construct a reliable projection
725: $\Pi$, since any information loss during that operation will affect the
726: quality of entire gate.
727: 
728:   Concrete realization, requires also to decide upon particular {\POVM} being
729: used. It is possible to choose this transformation in such a way, that some of
730: the gates will be significantly simplified, for instance acquiring convenient
731: form of permutations. Our choice is dictated by the optimization of the {\CNOT}
732: gate, discussed latter in this section. This {\POVM} is given by\footnote{
733:   From the point of view of state estimation, the optimal POVM is a conformal
734:  transformation, which maps the Bloch sphere into a sphere inscribed in the
735:  standard simplex of $\fR^{2^{2n}}$. Thanks to the many symmetries of such
736:  geometric configuration, some of the rotations are expressible as permutations
737:  of the simplex' vertices and can be implemented with high efficiency. In the
738:  case of $A$ given by Eq. \eqref{eq:A}, the permutation $(00,10)(01,11)$
739:  corresponds to the $1$-qubit NOT gate. With a different POVM %associated with the real Fourier transform,
740:  one can bring the Hadamard's gate H to a permutation, therefore if an algorithm
741:  relies on frequent applications of this operation, that could be a preferred
742:  choice.
743: }
744: \begin{equation}
745:  \label{eq:A}
746:   A^z\null_\mu := \frac{1}{2}\begin{pmatrix}
747:       1 &         - \tfrac{1}{\sqrt{3}} &         - \tfrac{1}{\sqrt{3}} &         - \tfrac{1}{\sqrt{3}} \\
748:       1 &\phantom{-}\tfrac{1}{\sqrt{3}} &\phantom{-}\tfrac{1}{\sqrt{3}} &         - \tfrac{1}{\sqrt{3}} \\
749:       1 &         - \tfrac{1}{\sqrt{3}} &\phantom{-}\tfrac{1}{\sqrt{3}} &\phantom{-}\tfrac{1}{\sqrt{3}} \\
750:       1 &\phantom{-}\tfrac{1}{\sqrt{3}} &         - \tfrac{1}{\sqrt{3}} &\phantom{-}\tfrac{1}{\sqrt{3}}
751:     \end{pmatrix}.
752: \end{equation}
753: 
754: %-------------------------------------------------------------------------------
755: \subsection{Single-qubit gates}
756: 
757: The neural circuit implementing arbitrary $1$-qubit gate is presented in
758: \figref{fig:G1D}. The projection $\Pi$ which transforms the `sparse' code
759: $\{\Xl{00'},\Xl{01'},\Xl{10'},\Xl{11'}\}\in\Omega^2$ onto a `dense' one
760: $\{\Xl{A'},\Xl{B'}\}\in\stV^2_*$, is a linear mapping implemented with
761: weights
762: \[
763:   W_\Pi
764:     = \begin{pmatrix}
765:         0 & 0 & 1 & 1 \\
766:         0 & 1 & 0 & 1
767:       \end{pmatrix}.
768: \]
769: But its inverse, $\Pi^{-1}$ is nonlinear and we realize this function, in a
770: two-step linear-feedback operation. The first step requires, apart from the
771: input signals, an additional supply of constant `current' of units from the
772: vertex $\vbsl{1}$. The effect of such a coupling to unity on a cell is to alter
773: its firing threshold. The weights of this part, effecting a linear injection
774: from $\{\Xl{A},\Xl{B},1\}$ to $\{\Xl{00},\Xl{01},\Xl{10},\Xl{11}\}$ are
775: \[
776:   W_{\Pi^{-1}} = \begin{pmatrix}
777:              - 1 &         - 1 &\phantom{-}1 \\
778:              - 1 &\phantom{-}1 &\phantom{-}0 \\
779:     \phantom{-}1 &         - 1 &\phantom{-}0 \\
780:     \phantom{-}1 &\phantom{-}1 &         - 1
781:     \end{pmatrix}.
782: \]
783: While the composition $W_\Pi W_{\Pi^{-1}}=\id_{\stV^2_*}$ as required, the
784: reciprocal is not an identity and needs a rectifying feedback sent from the
785: `winning' neuron to its neighbors, in order to bring their residual potentials
786: back to zero. Because of the one-step delay, this signal has to be adjusted
787: to match the attenuation already done by the function $S$. Hence the weight
788: matrix of this rectification step is determined by
789: \begin{equation}
790:  \label{eq:Wrec}
791:  \begin{aligned}
792:   {[}W_\mathrm{rec}{]}^i\null_j
793:    &= -S\big([W_{\Pi^{-1}}W_\Pi]^i\null_j - \delta^i\null_j\big)\\
794:    &= -S(-1)[\idmatrix_{\Omega^2} - W_{\Pi^{-1}}W_\Pi]^i\null_j.
795:  \end{aligned}
796: \end{equation}
797: Note, that for vanishing saturation parameter, this correction also disappears
798: due to $S\equiv 0$.
799: 
800: \smallskip
801:   In the absence of noise ($\sigma=0$), the conversion $\Pi^{-1}$ between
802: dense and sparse coding is completely error-free. As $\sigma$ increases, the
803: imperfections start to appear in the form of either multiple, or `void' spiking
804: in the first hidden layer. Although we found the circuit to behave stably in
805: these conditions, an improvement, in terms of both fidelity and coherence, can
806: be achieved by adding a second, normalizing feedback (not shown explicitly in
807: \figref{fig:G1D}).
808: 
809: %\input{fig-G1D}
810: \begin{figure}
811: \includegraphics{fig-G1D.eps}
812: \figcaption{Schematic for the 1-qubit gate. The hidden nodes are drawn in
813:   inverted lexicographical order to avoid excessive entanglement of the graph.
814:   Explicit connections of the normalization (nor) feedback are omitted, to
815:   improve legibility. Inhibitory connections are marked with (\synInh),
816:   excitatory with (\synExc), and those capped with (\synGat) depend on the
817:   applied gate $G$. The double connections (\synMod)
818:   consist of inhibition followed by delayed and attenuated excitation. The
819:   transient time is $\tGate = 4 + \tAvr$. Including the normalization feedback,
820:   and doubled gate connections ($\tAvr=2$), the circuit comprises $10$ nodes,
821:   and $62$ edges.}
822: \label{fig:G1D}
823: \end{figure}
824: 
825:   Normalizing feedbacks are commonly proposed for explanation of the observed
826: behavior in cortical neurons \cite{CHM1997,TLDRN1998}. The main difference
827: between these and our proposal is that while the former are multiplicative, this
828: one acts additively. Its role is to adjust the residual potentials for the
829: difference
830: \[
831:   1 - \sum_i X^i,\qquad i=00,01,10,11.
832: \]
833: Because we do not know which of the four neurons spiked mistakenly, the
834: normalizing signal is sent evenly to all of them. Its strength is determined by
835: the average excess of a signal encountered on a double-spike event:
836: \[
837:   \tfrac{1}{4}\sum_j [W_\mathrm{rec}]^i\null_j = -\tfrac{1}{4}S(-1).
838: \]
839: The deficit, which happens upon lack of a single spike has the same magnitude
840: but opposite sign. Therefore, the weight matrix of this normalization reads
841: \[
842:   W_\mathrm{nor}
843:     = -S(-1)\begin{pmatrix}
844:         -\tfrac{1}{4} & -\tfrac{1}{4} & -\tfrac{1}{4} & -\tfrac{1}{4} & 1 \\
845:         -\tfrac{1}{4} & -\tfrac{1}{4} & -\tfrac{1}{4} & -\tfrac{1}{4} & 1 \\
846:         -\tfrac{1}{4} & -\tfrac{1}{4} & -\tfrac{1}{4} & -\tfrac{1}{4} & 1 \\
847:         -\tfrac{1}{4} & -\tfrac{1}{4} & -\tfrac{1}{4} & -\tfrac{1}{4} & 1
848:       \end{pmatrix},
849: \]
850: where the last column refers to the unit vertex $\vbsl{1}$.
851: 
852: \smallskip
853:   Making the embedding $\Pi^{-1}$ robust is crucial for achieving correct
854: projection $\Pi$. Multiple, or void events linearly projected \textit{via}
855: $W_\Pi$ typically lead to a significant loss of coherence (although with less
856: impact on fidelity). In order to compensate for these non-exclusive events, the
857: spiking neuron sends a composite inhibitory signal down the hierarchy.
858: This is implemented with double connections: the first transmits
859: inhibitory signal at certain level $\eta$, and is followed by a delayed,
860: excitatory one aimed at bringing the residual potential of target neuron back
861: to its prior value. The full accomplishment of this goal is impossible with
862: nonlinear function $S$ -- the value can only be fully restored in the linear
863: limit $\gamma\to\infty$. Given the inhibitory coupling $\eta$, our best estimate
864: of the following excitation strength is $\eta'=\avr{u} - S(\avr{u}-\eta)$, where
865: $\avr{u}$ is the average residual potential. Because $\avr{u}\approx 0$, we set
866: $\eta'=-S(-\eta)$. The optimal value of $\eta$ was found numerically, by
867: minimizing the variation of fidelity over a range of gates acting on test
868: states (see Results).
869: 
870: \smallskip
871:   Application of a gate $G$ requires no additional node of the network,
872: only manipulation of the weights between the embedding and projecting parts.
873: In simplest case these are directly set to
874: \[
875:   W_G = A L A^{-1} = G.
876: \]
877: We have found however, that within some limits, the mechanism of \term{synaptic
878: averaging} may provide improvement of the performance. In real networks, a
879: single synapse contributes only a tiny fraction of the total input signal
880: \cite{ChS1992}. Multiple connections of similar lengths lead to signal
881: accumulation, different delays -- to temporal averaging. In our toy-model, the
882: first case is replaced by single, but strong connections, while for
883: implementation of the latter we directly use several edges having different
884: lengths with proportionally attenuated couplings. In the case of a single-qubit
885: gate, of the several configurations tested, the best results were obtained with
886: just two-step average $\tAvr=2$, hence, the connections were fixed at
887: $W_{G,1}=W_{G,2}=\thalf G$.
888: 
889: \paragraph{Results.}
890:   In order to reduce statistical uncertainties, all gates were tested on a fixed
891: set of $14$ pure states approximately evenly distributed on the Bloch sphere:
892: \[
893:  \hspace{-1mm}
894:  \begin{aligned}
895:   \Big\{&\ket{0}, \ket{1},
896:         \tfrac{1}{\sqrt{2}}\big(\ket{0}+\e{\imu k\pi/2}\ket{1}\big),\\
897:   \null&\frac{\sqrt{\sqrt{3}\pm 1}\ket{0}+\e{\imu(2k+1)\pi/4}\sqrt{\sqrt{3}\mp 1}\ket{1}}
898:              {\sqrt{2\sqrt{3}}}\Big\},
899:  \end{aligned}
900:  \quad
901:  k \in \fZ_4.
902: \]
903: While the input nodes were fed with spike trains $\{\Xl{A},\Xl{B}\}$ of joint
904: probability distributions $\hat{p}_\mathrm{in}\in\Omega^2_0$ corresponding to
905: the above states, the output $\{\Xl{A'},\Xl{B'}\}$ was tested for its coherence
906: $R$, and fidelity with the desired distribution $G\hat{p}_\mathrm{in}$.
907:   The initial test runs were made for several gates including Hadamard's, {\NOT},
908: the antipode (non-unitary), and two rotations:
909: $U_\theta = \exp(\imu\theta\hat{\sigma}_2/2)$, and
910: $U_\phi   = \exp(\imu\phi\hat{\sigma}_3/2)$. For the representative, the phase
911: gate $U_\phi$ was selected -- the effects of other operations were
912: quantitatively similar, or better. Its representation $G_\phi=A L_\phi A^{-1}$,
913: acting in $\Omega^2$ reads
914: \[
915:  \hspace{-7mm}
916:   G_\phi
917:     = \half\begin{pmatrix}
918:         1 + \cos\phi & 1 - \cos\phi &-\sin\phi     & \sin\phi     \\
919:         1 - \cos\phi & 1 + \cos\phi & \sin\phi     &-\sin\phi     \\
920:         \sin\phi     &-\sin\phi     & 1 + \cos\phi & 1 - \cos\phi \\
921:        -\sin\phi     & \sin\phi     & 1 - \cos\phi & 1 + \cos\phi
922:       \end{pmatrix}.
923: \]
924: The results presented in \figref{fig:FC-1D} are averages over $36$
925: rotation angles evenly spaced across the entire interval $[0,2\pi)$. The best
926: performance was observed for $\phi=0$ (identity) and $\phi=\pi$, while the worst
927: cases were encountered around $\phi\approx \pm \pi/2$ (but not exactly at these
928: angles). For each setting $(\sigma,\gamma)$, the inhibition level $\eta$ was
929: adjusted to minimize the variance of fidelity across the test states and
930: rotation angles ({\cf} \figref{fig:FC-1D}-insets). Note that while this
931: optimization was mainly coincident with maximization of the fidelity itself,
932: the trend in coherence was typically opposite. Had we chosen to optimize for
933: purity of states ($R\to 1$), the figures would look different.
934: 
935:   The prominent feature of \figref{fig:FC-1D}a, is the overcoherence of output
936: states in the limit $\gamma\to 0$. This means these distributions are too sharp
937: to represent quantum states, and any subsequent application of another gate
938: would certainly lead to a loss of accuracy. Interestingly, the
939: average fidelity remains at relatively high level. This suggests a possibility
940: of correcting the distributions by rescaling about the average. On the other
941: hand, the fidelity {\SD} is significant for small saturations, and becomes
942: comparable with statistical uncertainties only above $\gamma\gtrsim 1$.
943: 
944:   The conclusion drawn from \figref{fig:FC-1D}b is clear: the circuit considered
945: here is designed to work in deterministic regime $\sigma\to 0$. This makes an
946: interesting contrast between stochastic nature of quantum states and the
947: determinism of gates acting on them. As we are going to show, this dichotomy is
948: not limited to the $1$-qubit gate, but persists also in the case of entangling
949: operation {\CNOT}.
950: 
951:   Finally, we have sought for an estimate of the time needed to complete the
952: quantum rotations with this gate. Apart from the spatial resources, measured in
953: terms of cells and connections being used, time is an important factor
954: contributing to the overall cost of the realization. To assess this property, we
955: have run the circuit while varying the \term{signal length} $\tSig$: After an
956: initial transient of $\tGate = 4 + \tAvr$, the network was ran for $\tSig\geq 1$
957: successive steps, after which the cells were re-set to their initial state
958: ($u^i=0$, $X^i=0$), ensuring that all memory traces stored in residual
959: potentials were erased. This procedure was repeated until satisfactory
960: statistics ($N\tSig\approx 10^4$) was gathered.
961: 
962: %\input{fig-FC-1D.tex}
963: \begin{figure}
964: \includegraphics{fig-FC-1D.eps}
965: \figcaption{Performance of the 1-qubit phase gate in function of \textbf{a)}:
966:   saturation level of the residual potential, \textbf{b)}: noise SD of the
967:   activation function $P$, \textbf{c)}: length of the input signal (note the
968:   scale difference between graphs). Synaptic averaging was fixed at $\tAvr=2$.
969:   Each point is a mean over $36$ rotation angles within the whole interval
970:   $[0,2\pi)$, and $14$ pure states on which the phase gate was tested. With
971:   statistics of $10^4$ steps per setting, the associated uncertainties are
972:   negligible -- shaded regions ($F$) and broken lines ($R$) represent the
973:   standard deviations across the states and rotation angles.
974:   \textbf{Insets}: The inhibition levels used during simulations, optimized for
975:   minimization of the fidelity variance.}
976: \label{fig:FC-1D}
977: \end{figure}
978: 
979:   The results provided in \figref{fig:FC-1D}c evidently show that the real
980: temporal cost is not only the delay $\tGate$, but a significant number of
981: further steps are needed to `tune' this gate to a signal. After approximately
982: $\tSig\approx 30$ events the output quality no longer improves, and
983: consequently one can identify $\tSig$ with the statistics needed for maximal
984: efficiency. Since the latter is a function of saturation $\gamma$ and noise
985: $\sigma$, one expects $\tSig$ to raise monotonically with $\gamma$ and decrease
986: as $\sigma$ increases. In particular, the ideal case $\gamma\to\infty$,
987: $\sigma\to 0$ would also require infinite statistics to achieve the best
988: performance. One therefore finds yet another reason for the low saturation
989: values: The finiteness of signals encoded in spike trains, limits the
990: attainable efficiency of transformations, and high saturation values cannot
991: provide improvement beyond these limitations.
992: 
993: %-------------------------------------------------------------------------------
994: \subsection{The CNOT gate}
995: Unlike the single-qubit gates which can, by means of a special choice of the
996: {\POVM}, be transformed to a permutation, the {\CNOT} operation does not admit
997: such representation\footnote{
998:   Would it be possible, then either the gate could have no entangling
999:   capability, or the marginal probabilities of the two qubits be not conserved.
1000: }.
1001: With $A$ given by \eqref{eq:A}, its operator $G_\mathrm{CNOT}=A L_\mathrm{CNOT} A^{-1}$ has the following
1002: structure
1003: \[
1004:  \hspace{-9mm}
1005:   G_\mathrm{CNOT}
1006:     = \begin{pmatrix}
1007:         H_1 - J_1            & H_1^\mathrm{T} + J_2 & H_2                  & H_2^\mathrm{T}       \\
1008:         H_1^\mathrm{T} + J_2 & H_1 - J_1            & H_2^\mathrm{T}       & H_2                  \\
1009:        -H_2                  &-H_2^\mathrm{T}       & H_1 - J_2            & H_1^\mathrm{T} + J_1 \\
1010:        -H_2^\mathrm{T}       &-H_2                  & H_1^\mathrm{T} + J_1 & H_1 - J_2
1011:       \end{pmatrix},
1012: \]
1013: where
1014: \begin{align*}
1015:   H_1 &= \frac{1}{4}\begin{pmatrix}
1016:           1 & \tfrac{1}{\sqrt{3}} & 1 & \tfrac{1}{\sqrt{3}}\\
1017:          -\tfrac{1}{\sqrt{3}} & 1 &-\tfrac{1}{\sqrt{3}} & 1\\
1018:           1 & \tfrac{1}{\sqrt{3}} & 1 & \tfrac{1}{\sqrt{3}}\\
1019:          -\tfrac{1}{\sqrt{3}} & 1 &-\tfrac{1}{\sqrt{3}} & 1
1020:         \end{pmatrix},\\
1021:   H_2 &= \frac{1}{4}\begin{pmatrix}
1022:           \tfrac{1}{\sqrt{3}} &-1 &-\tfrac{1}{\sqrt{3}} & 1\\
1023:           1 & \tfrac{1}{\sqrt{3}} &-1 &-\tfrac{1}{\sqrt{3}}\\
1024:          -\tfrac{1}{\sqrt{3}} & 1 & \tfrac{1}{\sqrt{3}} &-1\\
1025:          -1 &-\tfrac{1}{\sqrt{3}} & 1 & \tfrac{1}{\sqrt{3}}
1026:         \end{pmatrix},\\
1027:   J_1 &= \frac{1}{\sqrt{3}}\begin{pmatrix}
1028:           1 & 0 & 0 & 0 \\
1029:           0 & 0 & 0 &-1 \\
1030:           0 & 0 & 1 & 0 \\
1031:           0 &-1 & 0 & 0
1032:         \end{pmatrix},\\
1033:   J_2 &= \frac{1}{\sqrt{3}}\begin{pmatrix}
1034:           0 & 0 & 1 & 0 \\
1035:           0 &-1 & 0 & 0 \\
1036:           1 & 0 & 0 & 0 \\
1037:           0 & 0 & 0 &-1
1038:         \end{pmatrix}.
1039: \end{align*}
1040: As already mentioned, the {\POVM} \eqref{eq:A} has been
1041: chosen to optimize the {\CNOT} gate. Indeed, the linear projection
1042: $\Pi:\Omega^4\to\stV^4_*$,
1043: \[
1044:   W_\Pi =
1045:     \begin{pmatrix}
1046:       0 & 0 & 0 & 0 & 0 & \ldots & 1 & 1 \\
1047:       0 & 0 & 0 & 0 & 1 & \ldots & 1 & 1 \\
1048:       0 & 0 & 1 & 1 & 0 & \ldots & 1 & 1 \\
1049:       0 & 1 & 0 & 1 & 0 & \ldots & 0 & 1
1050:     \end{pmatrix},
1051: \]
1052: when combined with $G_\mathrm{CNOT}$, %where $L$ is the {\CNOT} operator,
1053: shows the two pbits $\Xl{A}$ and $\Xl{D}$ are invariant under $G_\mathrm{CNOT}$.
1054: These can be directly copied to the output $\Xl{A'}$, $\Xl{D'}$, as shown in
1055: \figref{fig:CNOT}. For that same reason, while implementing the embedding part
1056: $\Pi^{-1}$, we pair $\{\Xl{A},\Xl{D}\},\{\Xl{B},\Xl{C}\}$, rather than using the
1057: natural order $\{\Xl{A},\Xl{B}\},\{\Xl{C},\Xl{D}\}$. The section $\Pi^{-1}$ is
1058: a two-stage procedure: First, with the same construction as in $1$-qubit case
1059: we separately embed the two marginals  $\{\Xl{A},\Xl{D}\}$ and
1060: $\{\Xl{B},\Xl{C}\}$:
1061: \[
1062:   \Pi^{-1}_\mathrm{I}:\stV^4_*\to\Omega^2\times\Omega^2.
1063: \]
1064: Next, we combine these into a single map
1065: \[
1066:   \Pi^{-1}_\mathrm{II}:\Omega^2\times\Omega^2\to\Omega^4.
1067: \]
1068: Since this is done with a linear mapping, there is again a rectifying feedback
1069: $W_\mathrm{rec,II}$ obtainable from Eq. \eqref{eq:Wrec} applied to
1070: $\Pi_\mathrm{II}$ on $\Omega^4$. In contrast to $\Pi^{-1}_\mathrm{I}$, the
1071: second-stage embedding $\Pi^{-1}_\mathrm{II}$ turns out to be unstable against
1072: noise, and the normalization feedback is now a necessity. Because of
1073: \[
1074:   \forall_i\; 2^{-4}\sum_j [W_\mathrm{rec,II}]^i\null_j =-\tfrac{9}{16} S(-1),
1075: \]
1076: the normalization weights are set to
1077: \[
1078:   W_\mathrm{nor,II} = -9S(-1)(-D_4\oplus 1),
1079: \]
1080: where $[D_4]^i\null_j\equiv 2^{-4}$ is the diffusion operator, and `$1$' refers
1081: to the unit vertex $\vbsl{1}$.
1082: 
1083: %\input{fig-GCNOT.tex}
1084: \begin{figure}
1085: \hspace{-3mm}\includegraphics{fig-GCNOT.eps}
1086: \figcaption{Schematic of the CNOT gate. The circuit has $38$ nodes, and $1309$
1087:   edges if the synaptic averaging is set to $\tAvr=4$. The transient time is 
1088:   $\tGate = 5 + \tAvr$.}
1089: \label{fig:CNOT}
1090: \end{figure}
1091: 
1092:   Thanks to the invariance of two pbits $\Xl{A'}=\Xl{A}$, $\Xl{D'}=\Xl{D}$, the
1093: hierarchical projection $\Pi$ have been significantly simplified (in comparison
1094: to what is needed for general $2$-qubit gate). The four partial projections
1095: from $G_\mathrm{CNOT}$, shown on the right hand side of \figref{fig:CNOT}, are
1096: modulated directly by the marginal $\Pi^{-1}_\mathrm{I}(\Xl{A},\Xl{D})$. The
1097: mechanism of this modulation is the same as explained before (inhibition
1098: followed by attenuated excitation) and the same parameter $\eta$ is set common
1099: on those connections.
1100: 
1101: %\smallskip
1102:   Finally, the gate edges were multiplied, in order to use the synaptic
1103: averaging mechanism. We found no dramatic improvement while varying the
1104: averaging length $\tAvr$, at $\gamma\gtrsim 1$, nevertheless the
1105: performance was significantly better for small values of the saturation
1106: parameter. At $\gamma=1$ the optimal length was $\tAvr=4$.
1107: 
1108: \paragraph{Results.}
1109:   The performance was assessed upon a testing set of $28$ pure states, which
1110: included both separable and entangled ones:
1111: \[
1112:  \hspace{-9mm}
1113:  \begin{aligned}
1114:   \big\{
1115:    &\ket{00},\ket{01},\ket{10},\ket{11},\\
1116:    &\tfrac{1}{\sqrt{2}}\big(\ket{00}+\e{\imu k\pi/2}\ket{01}\big),\,&
1117:    &\tfrac{1}{\sqrt{2}}\big(\ket{00}+\e{\imu k\pi/2}\ket{10}\big),\\
1118:    &\tfrac{1}{\sqrt{2}}\big(\ket{00}+\e{\imu k\pi/2}\ket{11}\big),\,&
1119:    &\tfrac{1}{\sqrt{2}}\big(\ket{01}+\e{\imu k\pi/2}\ket{10}\big),\\
1120:    &\tfrac{1}{\sqrt{2}}\big(\ket{01}+\e{\imu k\pi/2}\ket{11}\big),\,&
1121:    &\tfrac{1}{\sqrt{2}}\big(\ket{10}+\e{\imu k\pi/2}\ket{11}\big)
1122:   \big\},
1123:  \end{aligned}
1124:  \quad
1125:  k\in\fZ_4.
1126: \]
1127: Interestingly, although some of these states are `preferred' in terms of
1128: achieved fidelity $F$, there was no correlation between this measure and the
1129: entanglement property. This observation should not be surprising, because the
1130: mapping of $2$-qubit states into joint probability distributions does not make
1131: entangled states distinct. It follows that even imperfect gate implementation
1132: should not distinguish these states from separable ones.
1133: 
1134: The results of simulations are presented in \figref{fig:FC-CNOT}. Like before,
1135: for each setting of the control parameters $(\gamma,\sigma)$, the inhibition
1136: level $\eta$ was adjusted to minimize the variance of fidelity across the
1137: test states.
1138: 
1139: Qualitatively, the figures \ref{fig:FC-CNOT} are largely similar to what had
1140: been obtained for $1$-qubit gates (\figref{fig:FC-1D}), the major difference
1141: is in the range of achieved fidelities and output coherences. Standard deviations
1142: of coherence $R$ increased evenly by approximately a factor of $2$, while the
1143: fidelity {\SD} multiplied by about $4-5$. The most dramatic changes are
1144: observed in \figref{fig:FC-CNOT}a: Whereas at low saturation values
1145: ($\gamma<1$) the $1$-qubit gate worked relatively well, in the case of
1146: {\CNOT} a huge overcoherence takes place along with significant fidelity
1147: loss.
1148: 
1149:   For a reasonable performance at $\sigma=0$ and $\tSig \gtrsim 30$ one
1150: needs $\gamma \gtrsim 1$. In this regime one finds $F\gtrsim 0.97\,(-0.03,+0.02)$,
1151: corresponding to the unitary error $\alpha \lesssim 14^\circ\,(-5,+4)$; with
1152: noise at $\sigma=0.3$ and $\gamma=1$ the fidelity drops down to
1153: $F=0.77\,(-0.11,+0.15)$, or $\alpha=42^\circ\,(-22,+12)$, what is hardly
1154: acceptable for a large-scale quantum computation. While comparing these
1155: values with the best to-date experimental achievements ($F \sim 0.7-0.8$ with
1156: trapped ions \cite{SKHRGLDBREB2003}, $F \sim 0.6-0.8$ with Josephson junctions
1157: \cite{YPANT2003}, $F \sim 0.85$ in optical setup \cite{BPWR2003}) one has to
1158: take into account the many simplifications of our toy-model. More realistic
1159: simulations, or ultimately -- realizations, may not necessarily prove as good as
1160: this one, and would probably require additional resources to implement some of
1161: the error correcting schemes, nevertheless the principle of quantum computing
1162: with neural networks has been demonstrated.
1163: 
1164: %\input{fig-FC-CNOT.tex}
1165: \begin{figure}
1166: \includegraphics{fig-FC-CNOT.eps}
1167: \figcaption{Performance of the CNOT gate for synaptic averaging fixed at
1168:   $\tAvr=4$. Note the differences in ranges, while comparing with
1169:   Fig.~2. The statistics for each setting is $10^4$ steps, and each point is an
1170:   average over $28$ test states.}
1171: \label{fig:FC-CNOT}
1172: \end{figure}
1173: 
1174: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1175: \section{Discussion}
1176: \label{sc:done}
1177: 
1178:   We have studied the potential of an artificial neural network to operate on
1179: correlated spike trains assuming the latter to encode quantum states. The model
1180: neurons are reduced here down to the essential ingredients of computational
1181: capability. Few comments concerning the simplifications made are in order at
1182: this point:
1183: 
1184:   First, we have completely neglected the synaptic noise, by assuming the
1185: signals to be relayed undisturbed between cells. The justification is that
1186: here the few edges of each node represent averages over $10^3-10^4$ real
1187: synaptic connections therefore the impact of faulty transmission through a
1188: single synapse is greatly reduced. But inclusion of this likely source of errors
1189: may still be a significant factor reducing the overall performance.
1190: 
1191:   Second, the time duration of processed signals are assumed to be much shorter
1192: than the synaptic plasticity scale. Adaptation is an inherent element of
1193: information processing in the brain, but it conflicts with the objective of
1194: reliable signal transformations in that there is a trade-off between computing
1195: efficiency and adaptive capability. The resolution is provided by separation in
1196: time scales between the two processes -- transformations act over short signals,
1197: typically in response to rapidly varying external stimuli. This is consistent
1198: with the optimal signal length which was found here for both $1$- and $2$-qubit
1199: gates to be of order $\sim 30$ steps. Assuming the time step is set to
1200: $\sim 5\,\mathrm{ms}$ leads to a realistic signal duration of
1201: $\sim 150\,\mathrm{ms}$.
1202: 
1203:   Third, the detrimental effect of cellular noise on the performance of quantum
1204: gates clearly shows the deterministic regime to be preferable at least for the
1205: coding scheme considered here. On one hand, a sharp firing threshold needed for
1206: the neurons to act as `counters' which discretize linearly accumulated input
1207: signals, corroborates with the theoretical analysis of optimality in terms of
1208: information encoding \cite{BRP2003}.
1209:   But on the other, the noise itself which blurs this
1210: threshold has been shown to be a viable resource acting through the mechanisms
1211: of stochastic resonance \cite{GHJM1998}. This suggests to consider alternative
1212: quantum coding schemes, which would make use of the inherent uncertainty in spike
1213: generation, provided the relevant conditions are stable enough ({\eg}, noise
1214: variance at a constant, moderate level).
1215: 
1216:   It is worthwhile to note at this point, that the quantum states are not
1217: absolute entities, and the same set of spike trains may be `quantized' in many
1218: different ways depending on the assumed definition of a state. Accordingly, the
1219: quantum transformations as well as their implementations will differ. We have
1220: discussed here only two coding schemes (referred to as the `dense' and `sparse'
1221: spatial code), but it appears plausible, that the real networks may actually
1222: alternate (or combine) many different encodings, depending on the nature of
1223: the input signal and the functional properties of the circuit. An evident
1224: possibility is the \emph{sparse temporal code} based on probability waves,
1225: particularly attractive for at least two reasons: First, the brain waves
1226: provide the frequency basis necessary for phase discrimination, and there is an
1227: experimental indication for independence between rate and phase variables
1228: \cite{HBK2003}. The question is not whether the spiking probability oscillation
1229: does have a role, but rather what is the relevant number of modes involved
1230: in computation (if more than two then one should consider qudits instead of just
1231: qubits).
1232: Second, while the `dense' code requires two random binary variables per qubit,
1233: by trading spatial for temporal resources, probability waves allow to encode
1234: one-qubit per neuron.
1235:   The drawback is that the mechanisms of short term synaptic plasticity
1236: \cite{MLFS1997,BP1998,FD2002} makes the neural circuits operating on this form
1237: of a code susceptible to unwanted modifications \cite{FLBR2001}. From this
1238: perspective, the use of sparse spatial coding \cite{F1987,BNMLL1990,OF1996,ASGA1996,BS1997,KBTGA2003},
1239: appears to be advantageous, since such spike trains have by definition no
1240: temporal correlations, and hence the circuits operating in this fashion are
1241: expected to be more stable.
1242: 
1243: \smallskip
1244:   In summary, we have demonstrated the principle of employing quantum
1245: coding in artificial neural networks, by providing examples of circuits which
1246: realize quantum gates. There is a room for improvement and further investigation
1247: with more realism put into the model, alternative circuits, and algorithm
1248: implementations. Exploring the possible ways in which neural networks can handle
1249: quantum codes, can certainly benefit both the quantum mechanics and neuroscience.
1250: On one hand, applications of {\QM} to neural systems broaden the range of
1251: possibilities to be considered when seeking to understand the language of
1252: spikes, on the other -- macroscopic realizations can provide clues about the
1253: microscopic phenomena upon which {\QM} originated.
1254: 
1255: 
1256: %\clearpage
1257: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1258: \def\bbaut#1#2{{#1}~{#2},}
1259: \def\bbttl#1{\textsl{``#1''},}
1260: \def\bbpub#1{\textit{#1}}
1261: \begin{thebibliography}{AAA}{\footnotesize
1262: \setlength{\itemsep}{0pt}\setlength{\parskip}{0pt}
1263: 
1264: \bibitem{RWRB1997}
1265:   \bbaut{F.}{Rieke}
1266:   \bbaut{D.}{Warland}
1267:   \bbaut{R.R.}{de Ruyter van Steveninck}
1268:   \bbaut{W.}{Bialek}
1269:   \bbttl{Spikes: Exploring the Neural Code}
1270:   The MIT Press, Cambridge, Massachusetts 1997.
1271: 
1272: \bibitem{ChS1992}
1273:   \bbaut{P.S.}{Churchland}
1274:   \bbaut{T.J.}{Sejnowski}
1275:   \bbttl{The Computational Brain}
1276:   The MIT Press, Cambridge, Massachusetts 1992.
1277: 
1278: \bibitem{A1984}
1279:   \bbaut{D.}{Aerts}
1280:   \bbttl{A possible explanation for the probabilities of quantum mechanics}
1281:   \bbpub{J. Math. Phys.} \textbf{27} (1986), 202--210.
1282: 
1283: \bibitem{AC2003}
1284:   \bbaut{D.}{Aerts}
1285:   \bbaut{M.}{Czachor}
1286:   \bbttl{Bag-of-words problem and semantic analysis in Fock space}
1287:   \eprint{quant-ph/0309022}
1288: 
1289: \bibitem{B2004}
1290:   \bbaut{D.C.}{Brody}
1291:   \bbttl{Shapes of quantum states}
1292:   \bbpub{J. Phys. A: Math. Gen.} \textbf{37} (2004), 251–-257.
1293: 
1294: \bibitem{Bal70}
1295:   \bbaut{L.E.}{Ballentine}
1296:   \bbttl{The statistical interpretation of quantum mechanics}
1297:   \bbpub{Rev. Mod. Phys.} \textbf{42}:2 (1970), 358--381.
1298: 
1299: \bibitem{P1995}
1300:   \bbaut{A.}{Peres}
1301:   \bbttl{Quantum Theory: Concepts and Methods}
1302:   Kluwer Academic Publishers, Dordrecht 1995.
1303: 
1304: \bibitem{PeTe98}
1305:   \bbaut{A.}{Peres} \bbaut{D.R.}{Terno}
1306:   \bbttl{Convex probability domain of generalized quantum measurements}
1307:   \bbpub{J. Phys. A} \textbf{31} (1998), L671--L675.
1308: 
1309: \bibitem{CFS2002}
1310:   \bbaut{C.M.}{Caves}
1311:   \bbaut{Ch.A.}{Fuchs}
1312:   \bbaut{R.}{Schack}
1313:   \bbttl{Quantum probabilities as Bayesian probabilities}
1314:   \bbpub{Phys. Rev. A} \textbf{65} (2002), 022305.
1315: 
1316: \bibitem{BHW1999}
1317:   \bbaut{V.}{Bu\v{z}ek}
1318:   \bbaut{M.}{Hillery}
1319:   \bbaut{R.F.}{Werner}
1320:   \bbttl{Optimal manipulations with qubits: Universal-NOT gate}
1321:   \bbpub{Phys. Rev. A} \textbf{60} (1999), R2626–-9.
1322: 
1323: \bibitem{MBSS2002}
1324:   \bbaut{F.}{De Martini}
1325:   \bbaut{V.}{Bu\v{z}ek}
1326:   \bbaut{F.}{Sciarrino}
1327:   \bbaut{C.}{Sias}
1328:   \bbttl{Experimental realization of the quantum universal NOT gate}
1329:   \bbpub{Nature} \textbf{419} (2002), 815--818.
1330: 
1331: \bibitem{BBCVMSSSW1995}
1332:   \bbaut{A.}{Barenco} \bbaut{C.H.}{Bennett} \bbaut{R.}{Cleve}
1333:   \bbaut{D.P.}{DiVincenzo} \bbaut{N.}{Margolus} \bbaut{P.}{Shor}
1334:   \bbaut{T.}{Sleator} \bbaut{J.A.}{Smolin} \bbaut{H.}{Weinfurter}
1335:   \bbttl{Elementary gates for quantum computation}
1336:   \bbpub{Phys. Rev. A} \textbf{52} (1995), 3457--3467.
1337:   % quant-ph/9503016
1338: 
1339: % Quantum Circuits with Mixed States
1340: % quant-ph/9806029
1341: 
1342: \bibitem{SKHRGLDBREB2003}
1343:   \bbaut{F.}{Schmidt-Kaler} {\etal}
1344:   \bbttl{Realization of the Cirac-Zoller controlled-NOT quantum gate}
1345:   \bbpub{Nature} \textbf{422} (2003), 408--411.
1346: 
1347: \bibitem{YPANT2003}
1348:   \bbaut{T.}{Yamamoto}
1349:   \bbaut{Yu.A.}{Pashkin}
1350:   \bbaut{O.}{Astafiev}
1351:   \bbaut{Y.}{Nakamura}
1352:   \bbaut{J.S.}{Tsai}
1353:   \bbttl{Demonstration of conditional gate operation using superconducting charge qubits}
1354:   \bbpub{Nature} \textbf{425} (2003), 941--946.
1355: 
1356: \bibitem{BPWR2003}
1357:   \bbaut{J.L.}{O'Brien} \bbaut{G.J.}{Pryde} \bbaut{A.G.}{White}
1358:   \bbaut{T.C.}{Ralph}
1359:   \bbttl{Demonstration of an all-optical quantum controlled-NOT gate}
1360:   \bbpub{Nature} \textbf{426} (2003), 264--267.
1361: 
1362: % Linear optical controlled-NOT gate in the coincidence basis
1363: % T. C. Ralph, N. K. Langford, T. B. Bell, and A. G. White
1364: % Phys. Rev. A 65, 062324 (2002)
1365: % quant-ph/0112088
1366: 
1367: \bibitem{CHM1997}
1368:   \bbaut{M.}{Carandini}
1369:   \bbaut{D.}{Heeger}
1370:   \bbaut{J.A.}{Movshon}
1371:   \bbttl{Linearity and normalization in simple cells of the macaque primary visual cortex}
1372:   \bbpub{J. Neurosc.} \textbf{17} (1997), 8621--8644.
1373: 
1374: \bibitem{TLDRN1998}
1375:   \bbaut{G.G.}{Turrigiano}
1376:   \bbaut{K.R.}{Leslie}
1377:   \bbaut{N.S.}{Desai}
1378:   \bbaut{L.C.}{Rutheford}
1379:   \bbaut{S.B.}{Nelson}
1380:   \bbttl{Activity-dependent scaling of quantal amplitude in neocortical neurons}
1381:   \bbpub{Nature} \textbf{391} (1998), 892--896.
1382: 
1383: \bibitem{MLFS1997}
1384:   \bbaut{H.}{Markram}
1385:   \bbaut{J.}{Lubke}
1386:   \bbaut{M.}{Frotscher}
1387:   \bbaut{B.}{Sakmann}
1388:   \bbttl{Regulation of synaptic efficacy by coincidence of postsynaptic SPs and EPSPs}
1389:   \bbpub{Science} \textbf{275} (1997), 213--215.
1390: 
1391: \bibitem{BP1998}
1392:   \bbaut{G.Q.}{Bi}
1393:   \bbaut{M.M.}{Poo}
1394:   \bbttl{Synaptic modifications in cultured hippocampal neurons: dependence on spike timing, synaptic strength, and postsynaptic cell type}
1395:   \bbpub{J. Neurosci.} \textbf{18} (1998), 10464--10472.
1396: 
1397: \bibitem{FD2002}
1398:   \bbaut{R.C.}{Froemke}
1399:   \bbaut{Y.}{Dan}
1400:   \bbttl{Spike-timing-dependent synaptic modification induced by natural spike trains}
1401:   \bbpub{Nature} \textbf{416} (2002), 433--438.
1402: 
1403: \bibitem{BRP2003}
1404:   \bbaut{M.}{Bethge} \bbaut{D.}{Rotermund} \bbaut{K.}{Pawelzik}
1405:   \bbttl{Optimal neural rate coding leads to bimodal firing rate distributions}
1406:   \bbpub{Network: Comput. Neural Syst.} \textbf{14} (2003), 303-–319.
1407: 
1408: \bibitem{GHJM1998}
1409:   \bbaut{L.}{Gammaitoni} \bbaut{P.}{H\"{a}nggi} \bbaut{P.}{Jung}
1410:   \bbaut{F.}{Marchesoni}
1411:   \bbttl{Stochastic resonance}
1412:   \bbpub{Rev. Mod. Phys.} \textbf{70} (1998), 223--287.
1413: 
1414: %\bibitem{WPPDM1994}
1415: %  \bbttl{Stochastic resonance on a circle}
1416: %  \bbaut{K.}{Wiesenfeld}
1417: %  \bbaut{D.}{Pierson}
1418: %  \bbaut{E.}{Pantazelou}
1419: %  \bbaut{C.}{Dames}
1420: %  \bbaut{F.}{Moss}
1421: %  \bbpub{Phys. Rev. Lett.} \textbf{72} (1994), 2125–-2129.
1422: 
1423: %\bibitem{OXDP2002}
1424: %  \bbaut{M.W.}{Oram}
1425: %  \bbaut{D.}{Xiao}
1426: %  \bbaut{B.}{Dritschel}
1427: %  \bbaut{K.R.}{Payne}
1428: %  \bbttl{The temporal resolution of neural codes: does response latency have a unique role?}
1429: %  \bbpub{Phil. Trans. R. Soc. Lond. B} \textbf{357} (2002), 987--1001.
1430: 
1431: \bibitem{HBK2003}
1432:   \bbaut{J.}{Huxter}
1433:   \bbaut{N.}{Burgess}
1434:   \bbaut{J.}{O’Keefe}
1435:   \bbttl{Independent rate and temporal coding in hippocampal pyramidal cells}
1436:   \bbpub{Nature} \textbf{425} (2003), 828--832.
1437: 
1438: \bibitem{F1987}
1439:   \bbaut{D.J.}{Field}
1440:   \bbttl{Relations between the statistics of natural images and the response properties of cortical cells}
1441:   \bbpub{J. Opt. Soc. Am. A} \textbf{4} (1987), 2379--2394.
1442: 
1443: \bibitem{BNMLL1990}
1444:   \bbaut{C.A.}{Barnes}
1445:   \bbaut{B.L.}{McNaughton}
1446:   \bbaut{S.J.}{Mizumori}
1447:   \bbaut{B.W.}{Leonard}
1448:   \bbaut{L.H.}{Lin}
1449:   \bbttl{Comparison of spatial and temporal characteristics of neuronal activity in sequential stages of hippocampal processing}
1450:   \bbpub{Prog. Brain Res.} \textbf{83} (1990), 287--300.
1451: 
1452: \bibitem{OF1996}
1453:   \bbaut{B.A.}{Olshausen}
1454:   \bbaut{D.J.}{Field}
1455:   \bbttl{Emergence of simple-cell receptive field properties by learning a sparse code for natural images}
1456:   \bbpub{Nature} \textbf{381} (1996), 607--609.
1457: 
1458: \bibitem{ASGA1996}
1459:   \bbaut{A.}{Arieli}
1460:   \bbaut{A.}{Sterkin}
1461:   \bbaut{A.}{Grinvald}
1462:   \bbaut{A.}{Aertsen}
1463:   \bbttl{Dynamics of ongoing activity: Explanation of the large variability in evoked cortical responses}
1464:   \bbpub{Science} \textbf{273} (1996), 1868-–1871.
1465: 
1466: \bibitem{BS1997}
1467:   \bbaut{A.J.}{Bell}
1468:   \bbaut{T.J.}{Sejnowski}
1469:   \bbttl{The ``independent components'' of natural scenes are edge filters}
1470:   \bbpub{Vision Res.} \textbf{37} (1997), 3327--3338.
1471: 
1472: \bibitem{KBTGA2003}
1473:   \bbaut{T.}{Kenet}
1474:   \bbaut{D.}{Bibitchkov}
1475:   \bbaut{M.}{Tsodyks}
1476:   \bbaut{A.}{Grinvald}
1477:   \bbaut{A.}{Arieli}
1478:   \bbttl{Spontaneously emerging cortical representations of visual attributes}
1479:   \bbpub{Nature} \textbf{425} (2003), 954--956.
1480: 
1481: \bibitem{FLBR2001}
1482:   \bbaut{A.L.}{Fairhall}
1483:   \bbaut{G.D.}{Lewen}
1484:   \bbaut{W.}{Bialek}
1485:   \bbaut{R.R.}{de Ruyter van Steveninck}
1486:   \bbttl{Efficiency and ambiguity in an adaptive neural code}
1487:   \bbpub{Nature} \textbf{412} (2001), 787--792
1488: 
1489: % Andreas V.M. Herz
1490: % Global analysis of recurrent neural networks
1491: % (tmp) Springer
1492: 
1493: % E.A. Stern, A.E. Kincaid, C.J. Wilson
1494: % Spontaneous subthreshold membrane potential fluctuations and action potential variability of rat corticostriatal and striatal neurons in vivo
1495: % J. Neurophysiol. 77:4 (1997), 1697--715
1496: 
1497: }\end{thebibliography}
1498: \clearpage
1499: \end{multicols}
1500: \end{document}
1501: 
1502: 
1503: