quant-ph0308131/aep.tex
1: %\documentclass[aps,preprint,superscriptaddress,showpacs]{revtex4}
2: 
3: \documentclass[aps,twocolumn,superscriptaddress,showpacs]{revtex4}
4: %\documentclass[aps,twocolumn,groupedaddress,showpacs]{revtex4}
5: \usepackage{epsf,latexsym}
6: \usepackage{times}
7: 
8: \newcommand{\bra}[1]{\langle #1 |}
9: \newcommand{\ket}[1]{| #1 \rangle}
10: \newcommand{\braket}[2]{\left \langle #1 | #2 \right\rangle}
11: 
12: \newcommand\R{{\mathrm {I\!R}}}
13: \newcommand\N{{\mathrm {I\!N}}}
14: \newcommand\h{{\cal H}}
15: 
16: \newcommand{\ra}{{\rightarrow}}
17: \newcommand{\be}{\begin{equation}}
18: \newcommand{\ee}{\end{equation}}
19: 
20: \def\CC{{\rm\kern.24em \vrule width.04em height1.46ex depth-.07ex
21:     \kern-.30em C}}
22: \def\P{{\rm I\kern-.25em P}}
23: 
24: \def\RR{{\rm
25:          \vrule width.04em height1.58ex depth-.0ex
26:          \kern-.04em R}}
27: 
28: \def\bbbone{{\mathchoice {\rm 1\mskip-4mu l} {\rm 1\mskip-4mu l}
29: {\rm 1\mskip-4.5mu l} {\rm 1\mskip-5mu l}}}
30: 
31: \def\bbbc{{\mathchoice {\setbox0=\hbox{$\displaystyle\rm C$}\hbox{\hbox
32: to0pt{\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
33: {\setbox0=\hbox{$\textstyle\rm C$}\hbox{\hbox
34: to0pt{\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
35: {\setbox0=\hbox{$\scriptstyle\rm C$}\hbox{\hbox
36: to0pt{\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}
37: {\setbox0=\hbox{$\scriptscriptstyle\rm C$}\hbox{\hbox
38: to0pt{\kern0.4\wd0\vrule height0.9\ht0\hss}\box0}}}}
39: 
40: \def\bbbz{{\mathchoice {\hbox{$\sf\textstyle Z\kern-0.4em Z$}}
41: {\hbox{$\sf\textstyle Z\kern-0.4em Z$}}
42: {\hbox{$\sf\scriptstyle Z\kern-0.3em Z$}}
43: {\hbox{$\sf\scriptscriptstyle Z\kern-0.2em Z$}}}}
44: 
45: \newcommand{\putfig}[2]{$$\leavevmode\hbox{\epsfxsize=#2 cm
46:    \epsffile{#1.eps}}$$}
47: \newcommand{\insertfig}[2]{\leavevmode \vcenter{\hbox{\epsfxsize=#2 cm
48:    \epsffile{#1.eps}}}}
49: 
50: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
51: 
52: \begin{document}
53: \title{Quantum entangling power of adiabatically connected hamiltonians }
54: \author{Alioscia Hamma$^{1,3,4}$, Paolo Zanardi$^{1,2}$}
55: \address{ Institute for Scientific Interchange (ISI), Villa
56: Gualino, Viale Settimio Severo 65, I-10133 Torino, Italy
57: \\
58: $^2$ Department of Mechanical Engineering, Massachusetts
59: Institute of Technology, Cambridge Massachusetts 02139
60: \\ 
61: $^3$ Dipartimento di Scienze Fisiche, Universit\`a
62: Federico II di Napoli, Via Cintia ed. G,\\
63: $^4$Istituto
64: Nazionale di Fisica Nucleare (INFN), sezione di Napoli
65: }
66: \begin{abstract}
67: The space of quantum Hamiltonians has a natural partition in
68: classes of operators that can be adiabatically deformed into each
69: other. We consider parametric families of Hamiltonians acting on a
70: bi-partite quantum state-space. When the different Hamiltonians in
71: the family fall in the same adiabatic class one can manipulate
72: entanglement by  moving through energy eigenstates corresponding
73: to different value of the control parameters. We introduce an
74: associated notion of adiabatic entangling power. This novel
75: measure is analyzed for general $d\times d$ quantum systems and
76: specific two-qubits examples are studied
77: \end{abstract}
78: \pacs{}
79: \maketitle
80: \section{ Introduction}
81: Adiabatic evolutions represent a very special class of quantum evolutions, nevertheless
82: they allow for a broad set of quantum state manipulations. In particular
83: a big deal of activity has been recently devoted to the study of adiabatic techniques
84: for Quantum Information Processing \cite{QIP}.
85: 
86: The notion of adiabatic quantum computing  emerged as an novel
87: intriguing paradigm for the development of  efficient quantum
88: algorithms \cite{aqc},\cite{berk},\cite{berk1}. In this approach
89: information e.g., the solution of an hard combinatorial problem,
90: is encoded in  the ground state of a properly designed many-qubits
91: Hamiltonian $H_f.$ This ground-state is  then generated by letting
92: the system evolve in adiabatic fashion from the ground state of a
93: simple initial Hamiltonian $H_0$ \cite{aqc}. In view of the
94: adiabatic theorem ( see e.g.,  \cite{mess}) the crucial property
95: which governs the scaling behaviour of the computational time is
96: the spectral gap i.e., energy difference between the ground and
97: the first excited state. The larger the gap the faster the
98: computation can be.
99: 
100: In adiabatic quantum computing  as defined in Ref. \cite{aqc} the parametric family of Hamiltonians has the simple form of a convex combination of
101: $H_0$ and $H_f;$ one can also consider more general  family of Hamiltonians and more complex paths in the control parameter space.
102: For example in the so-called {\em geometric quantum computation} \cite{GQC}
103: one considers {\em loops} in the control space of a non-degenerate set of Hamiltonians
104: to the purpose of controlled   Berry phases generation \cite{berry}.
105: When even the non-degeneracy constraint is lifted one and high-dimensional eigenspaces
106: are allowed, one is led to consider non-abelian
107: holonomies which  mix non-trivially the ground-states of the system. This latter method,
108: which  provides a general approach to QIP as well, is termed {\em holonomic quantum computation}
109: \cite{HQC}
110: 
111: In this paper we shall investigate how one can adiabatically
112: generate quantum entanglement \cite{unan},\cite{pz}. The idea is a
113: simple one. One first prepares a bi-partite  quantum system in one
114: of its eigenstates e.g., the ground-state, and then drives the
115: control parameters of the system Hamiltonian along some path. If
116: this path is  adiabatic the system will stand at any any time in
117: the corresponding eigenstate. In general eigenstates associated to
118: different control parameters will have different entanglement,
119: therefore the described dynamical process will result in a
120: protocol for entanglement manipulation. We would like to
121: characterize a parametric family of Hamiltonians in terms of its
122: capability of entanglement generation according the above
123: protocol. In this paper we  will  focus on bi-partite e.g.,
124: two-qubits, quantum systems. The aim will be, given an Hamiltonian
125: family, to characterize its entangling capabilities by means of
126: adiabatic manipulations.
127: 
128: 
129: \section{Adiabatic connectibility}
130: 
131: Let us start by a few simple general considerations about adiabatically connectible Hamiltonians.
132: We would like to understand how the space of Hamiltonians over ${\cal H}\cong\CC^D,$
133: splits in classes of  elements that can be adiabatically deformed into each other.
134: 
135: {\em Definition} Two Hamiltonians $H_0$ and $H_1$ are adiabatically connectible
136: if its exists a continuous family of Hamiltonians $\{H_t\}_{t\in[0,1]}
137: $ such that i) $H(0)=H_0$ and $H(1)=H_1,$ ii)
138: the degeneracies of the spectra of the $H_t'$s do not depend on $t.$
139: 
140: The notion of adiabatic deformability of Hamiltonians is an important concept
141: in many-body and field theory quantum systems. Indeed when  two Hamiltonians
142: can be connected in this way they share several properties e.g., ground-state degeneracy,
143: quasi-particle quantum numbers,$\ldots,$
144: so that  in many respects they can be regarded  as belonging to the same kind of universality class
145: \cite{PWA}. On the other an obstruction to such a process will be typically
146: associated to a some sort of quantum phase transition. Unconnectible Hamiltonians
147: show  qualitative different features. Since we will study how entanglement changes
148: while remaining in the same adiabatic class our  analysis can be regarded
149: as complimentary to the one of entanglement behaviour in quantum phase transitions
150: \cite{QPT}.
151: 
152: In the simple finite-dimensional case we are interested in one can prove the following
153: 
154: {\em{ Proposition 1}}.-- Two Hamiltonians $H_0$ and $H_1$ over
155: ${\cal H}\cong\CC^D,$ are adiabatically connectible if and only if
156: they belong to the same connected component of the set of iso-degenerate Hamiltonians. 
157: 
158: {\em Proof.} 
159: Let $H_\alpha=\sum_{i=1}^R \epsilon_\alpha^i
160: \Pi_\alpha^i \,(\alpha=0,1)$ the spectral resolution of   $H_0$ and $H_1.$
161: %Isodegeneracy  of these latter two operators means that the sets
162: %${\cal D}_\alpha:=\{ \rm{tr} \Pi_\alpha^i \}_{i=1}^R,\,(\alpha=0,1)\subset \N$
163: %coincides.  
164: We now order their eigenvalues in ascending order i.e., $\epsilon_\alpha^1<...<\epsilon_\alpha^R.$ We define two { vectors} $D_\alpha\,(\alpha=0,1)$
165: in $\R^R$ as follows $D_\alpha:=(\rm{tr} \Pi_\alpha^1,\ldots,\Pi_\alpha^R),$ where the components are ordered according to the corresponding eigenvalues.
166: The Hamiltonians  $H_0$ and $H_1$ belong  to the same connected component of the set of iso-degenerate hamiltonians
167: iff  $D_0=D_1.$ Iso-degeneracy is given by the weaker condition that it exists a permutation $P$ of $R$ objects such
168: that $(D_1)_i=(D_0)_{P(1)},\,(i=1,\ldots,R).$
169: It is an elementary fact that, given the two systems of
170: ortho-projectors $\{\Pi_\alpha^i\}_{i=1}^R, \,(\alpha=0,1)$ such
171: that Tr$\Pi_1^i=\mbox{Tr} \Pi_2^i,\,(i=1,\ldots,R),$ it exist a
172: (non-unique ) unitary $W$ such that $W
173: \,\Pi_0^i\,W^\dagger=\Pi_1^i,\,(i=1,\ldots,R).$ 
174: Let us introduce $R$ real-valued functions $\epsilon^i\colon [0,\,1]\mapsto \R$ 
175:  such that $\epsilon^i(0)=\epsilon_0^i,$ and $\epsilon^i(1)=\epsilon_1^i\, (i=1,\ldots,R).$ 
176:  In view of the ordering assumption  we can choose them to satisfy the no-crossing constraints $\epsilon^{i+1}(t)>\epsilon^{i}(t)\,
177: (i=1,\ldots,R-1).$
178: Consider now the following family of Hamiltonians $H(t)=\sum_{i=1}^R \epsilon^i(t) U_t \Pi_0^i U_t^\dagger,$
179: where the continuous unitary family $\{U_t\}_{t=0}^1$ is such that $U_0=\openone$ and $U_1=U.$
180: Clearly $H(0)=H_0$ and $H(1)=H_1.$ Moreover, for the very way they have been constructed, all the $H(t)$
181: belong to the same connected component of the set of iso-degenerate Hamiltonians of $H_0$ and $H_1.$
182:  This shows that the latter condition is sufficient in order that $H_0$ and $H_1.$ are adiabatically connectible.
183: 
184: Iso-degeneracy of $H_0$ and $H_1$ is also an obvious necessary condition for adiabatic  connectibility
185: because otherwise level crossing would necessarily occur. But level crossing would necessarily occur even if $D_0\neq D_1$ because, for some $t\in[0,\,1],$
186: and $1\le i\le R,$ it would be $\epsilon^{i+1}=\epsilon^{i}.$ This proves the necessity part of the Proposition. 
187: 
188: $\hfill\Box$
189: 
190: The role of the functions $\epsilon^i(t)$ in the Proof above is to map the spectrum of $H_0$ onto the one of $H_1$
191: whereas all the information about the eigenvectors is contained in the family of unitaries $U_t.$
192: By setting all the connecting functions $\epsilon^i_t/\epsilon_0^i$ to one, one gets a final Hamiltonian $\tilde{H}_1$
193: {\em iso-spectral} to $H_ 0$ having the same eigenvectors of $H_1.$
194: This latter remark is important for the following in that
195: it allows one to restrict to iso-spectral Hamiltonian families.
196: The actual spectrum structure e.g., the energy gaps, just imposes
197: an upper bound  over the speed at which the adiabatic deformation process can be  carried on.
198: Moreover in order to have a  one-to-one  correspondence between eigenvalues and eigenstates we shall
199: assume that our Hamiltonians are   {\em non-degenerate} i.e., $d_i=1,\,(i=1,\ldots,R).$
200: Notice that in Hamiltonian space the condition of non-degeneracy is a {\em generic} one.
201: 
202: The simplest case one can consider  is of course provided by two-level Hamiltonians
203: with eigenvalues $\epsilon_1$ and $\epsilon_2.$
204: Using the standard pauli matrices, one can write $H =\epsilon_S \openone +\epsilon_A \vec{n}\cdot\vec{\sigma}\,
205: (\epsilon_S:= (\epsilon_1+ \epsilon_2)/2,\,\epsilon_A:=( \epsilon_1- \epsilon_2)/2$
206: %Where the components of $\vec{n},$ up to normalization, are given by Tr$(\sigma_i H),\,(i=1,2,3).$
207: Here we have just two possibility 1) $\epsilon_A=0$ the
208: Hamiltonian is a rescaled identity and we have just one degree of
209: freedom ii) $\epsilon_A\neq 0;$ all possible operators of this
210: kind are then parameterized by a triple $(\epsilon_S, \epsilon_A,
211: \vec{n})$ where $\epsilon_A\in \RR,\, \epsilon_A\in\RR-\{0\}$ and
212: $\vec{n}\in S^2\cong SU(2)/U(1).$
213: For each of the two iso-degeneracy classes above there is just one connected component
214: i.e., any  the non (totally) degenerate Hamiltonian is adiabatically connectible 
215: any other  non (totally) degenerate Hamiltonian. Notice that this latter statement holds
216: for any dimension of ${\cal H}.$
217: %
218: %As a representative in this class one can take
219: %$H_0=\openone+\sigma_z.$ 
220: %In higher dimensions  the class of
221: %Hamiltonians connected to $H_0$ is then parameterized by the $R$
222: %distinct real numbers $\{\epsilon^i)\}_{i=1}^R$ and by an element of quotient
223: %manifold  $ U(D)/\prod_{i=1}^R U(d_i).$
224: 
225: \section{ Adiabatic entangling power.}
226: 
227: We move now to introduce our definition of adiabatic entangling power. Let ${\cal H}\cong
228: \CC^d\otimes\CC^d$ be a bi-partite quantum state space. We
229: consider a  family $\cal F$ of non-degenerate Hamiltonians
230: over $\cal H,$
231: %\begin{equation}
232: ${\cal F}_H:=\{H(\lambda)\,/\, \lambda \in{\cal M}\}
233: $
234: %\end{equation}
235: where $\cal M$ is a $n$-dimensional compact and connected manifold.
236: The points of $\cal M$ are to be seen as dynamically controllable parameters.
237: Let $E\colon {\cal H}\rightarrow \RR^+_0$ be a measure of bi-partite pure state
238: entanglement over $E$ e.g., von Neumann entropy of the reduced density matrix.
239: If $H(\lambda)=\sum_{i=1}^{d^2} \varepsilon_i |\Psi_i(\lambda)\rangle\langle \Psi_i(\lambda)|$
240: is the spectral resolution of an element of $\cal F$ we define
241: the {\em adiabatic entangling power} of $\cal F$ by
242: \begin{equation}
243: e({\cal F}_H):= \max_{i} \,\sup_{\lambda,\lambda^\prime}
244: |E(|\Psi_i(\lambda)\rangle)-E(|\Psi_i(\lambda^\prime)\rangle)|
245: \label{ep}
246: \end{equation}
247: $(i=1,\ldots,d^2,\,\lambda,\lambda^\prime\in{\cal M})$
248: %\begin{itemize}
249: 
250: %\item[0)]
251: We will assume  that it exists $H_{\lambda_0}\in{\cal F}_H$ such that the
252: associated eigenvectors are all{ \em product states}.
253: %, then the definition (\ref{ep})
254: %reduces to $e({\cal F}):= \max_{i} \,\sup_{\lambda}
255: %|E(|\Psi_i(\lambda)\rangle)|$
256: %\end{itemize}
257: Let us stress once again that the physical idea underneath these definitions is quite simple:
258: one starts from  the (unentangled) eigenvectors of $H_{\lambda_0}$
259: then by adiabatically driving the control parameters $\lambda$
260: the states $|\Psi_i(\lambda)\rangle$ can be reached.
261: If $\lambda^*$ denotes the point of where the maximum (\ref{ep})
262: is achieved ($\cal M$ is compact) any adiabatic path connecting $\lambda_0$
263: to $\lambda^*$ realizes an {\em optimal} entanglement generation procedure
264: within the family ${\cal F}_H.$
265: 
266: An explicit evaluation of (\ref{ep}) is, for a general $\cal F,$
267: quite a difficult task. In the light of the observations after
268: Proposition 1,  we can, without loss of generality,   consider
269: only the case in which $\cal F$ is an {\em iso-spectral} family of
270: {non-degenerate } Hamiltonians. Let ${\cal F}_U\subset {\cal
271: U}(\CC^d\otimes\CC^d)$ be a set (compact and connected) of unitary transformations
272: containing the identity. The isospectral family is
273: \begin{equation}
274: {\cal F}_H:=\{ U\,H_0 U^\dagger\,/\,
275: \,U\in{\cal F}_U\}
276: \label{family}
277: \end{equation}
278: where $H_0=\sum_{i=1}^{d^2}
279: \varepsilon_i|\Psi_i\rangle\langle\Psi_ i|,\; i\neq j\Rightarrow
280: \varepsilon_i\neq\varepsilon_j,$ and  the $|\Psi_i\rangle$'s are
281: an orthonormal basis of product states. Moreover one can also
282: restrict herself  to {\em ground-state} entanglement i.e., to
283: consider the entanglement contents of just the eigenvector
284: $|\Psi_0\rangle$ corresponding to the minimum energy eigenvalue.
285: If this is the case one can forget about the maximization over the
286: eigenvalue index $i$ in Eq. (\ref{ep}). The ground state of
287: $H(\lambda)$ ($H_0$) will be denoted as $|\Psi_0(\lambda)\rangle$
288: ($|\Psi_0\rangle$). For an iso-spectral family as in Eq.
289: (\ref{family}) we will use the notation $e({\cal F}_U).$
290: 
291: The adiabatic entangling power (\ref{ep}) induces, for the  class of Hamiltonian families (\ref{family}) the following 
292: real-valued function over the subsets ${\cal F}_U$ of ${\cal U}(\CC^d\otimes\CC^d).$
293: \begin{equation}
294: e({\cal F}_U)=  \max_i \sup_{ U\in{\cal F}_U}
295: E[U |\Psi_i\rangle].
296: \label{ep1}
297: \end{equation}
298: It is important to stress that this expression has the physical meaning of entanglement achievable  
299: by adiabatically manipulating the parametersi, living in a manifold ,say, $\cal M$, on which the $U$'s in ${\cal F}_U$ depend.
300: Indeed, for an iso-spectral Hamiltonian family (\ref{family}) the adiabatic evolution operator corresponding
301: to the path $\gamma\colon [0,\,T]\mapsto \cal M$ is given by the product of three different kinds of contributions
302: %\begin{equation}
303: $U_{ad} (\gamma)=U(\gamma(T))\, e^{-i H_0 T}\,U_B(\gamma).
304: $
305: %\label{3}
306: %\end{equation}
307: The first term $U(\gamma(T))$ is simply the unitary corresponding
308: to the end-point of the path $\gamma.$ Due to the adiabatic
309: theorem an initial eigenstate $|\Psi_i\rangle$ is indeed mapped,
310: up to a phase, onto the final eigenstate
311: $U(\gamma(T))|\Psi_i\rangle.$ The  second factor in $U_{ad}$
312: is clearly just the dynamical phase associated with $H_0$ whereas
313: the third is an operator taking into account the geometric
314: contribution to the phase accumulated by the eigenvectors
315: %\begin{equation}
316: $U_B(\gamma)= \sum_{i=1}^{d^2}e^{i\phi_g(\gamma)}|\Psi_i\rangle\langle \Psi_i|
317: $
318: %\end{equation}
319: in which $\phi_g(\gamma)=i\int_\gamma \langle\Psi_i(\lambda)|d|\Psi_i(\lambda)\rangle$
320: are the Berry's phases associated to $\gamma.$
321: Notice in passing that when $\gamma$ is a loop i.e., $\gamma(0)=\gamma(T)=\lambda_0$
322: then $U(\gamma(T))=\openone.$
323: As far as the adiabatic entangling power (\ref{ep}) is concerned
324: the phases can be obviously neglected. 
325: 
326: The adiabatic entangling power is invariant under left (and not
327: right in general) multiplication by bi-local unitary operators
328: i.e., $e({\cal F}_U)= e((U_1\otimes U_2){\cal F}_U),\,\forall
329: U_1,U_2\in{\cal U}(d).$ This implies that, as far adiabatic
330: entangling capabilities are concerned, a unitary family ${\cal
331: F}_U$ can be always considered closed under the
332: left-multiplication by local unitary operators.
333: 
334: 
335: We want now to establish a connection between the adiabatic entangling power (\ref{ep1})
336: and a variation of entangling power $e_p^{(av)}$ of -bi-partite unitaries introduced
337: in Ref. \cite{dur}[For a different definition, based on {\em average} entanglement production,
338: see also \cite{ZZF}]. In this paper we define $e_p(U)$ as the {\em maximum} entanglement obtainable
339: by the action of $U$ over all possible product states i.e., 
340: $e_p(U)=\sup_{\psi_1, \psi_2} E[U|\psi_1\rangle\otimes|\psi_1\rangle].$ 
341: 
342: Since the $|\Psi_i\rangle'$s are  by hypothesis  product states one clearly has
343: $E[U |\Psi_i\rangle]\le \sup_{\psi_1,\psi_2} E[U |\psi_1\rangle\otimes|\psi_1\rangle],$
344: Therefore  one obtains  the upper bound
345: \begin{equation}
346: e({\cal F}_U)\le \sup_{U\in{\cal F}_U} e_p( U)
347: \label{bound}
348: \end{equation}
349: In some circumstances one can get the equality.
350: 
351: {\em Proposition 2}.-- Suppose  that the unitary family ${\cal
352: F}_U$ is such that for all $U_1, U_2\in{ U}({d})$ one has ${\cal
353: F}_U (U_1\otimes U_2)\subset {\cal F}_U$ i.e., the  family is
354: closed also under right multiplication of bi-local operators.  It
355: follows that that the adiabatic entangling power coincides with
356: the supremum over ${\cal F}_U$ of the entangling power $e_p(U).$
357: 
358: {\em Proof.}
359: It is straightforward
360: \begin{eqnarray}
361: e({\cal F})&=&\max_i \sup_{U\in{\cal F}_U, U_1, U_2}E[U  (U_1\otimes U_2)|\Psi_i\rangle] \nonumber \\
362: &=&\sup_{U\in{\cal F}_U, \psi_1, \psi_2}  E[U |\psi_1\rangle\otimes|\psi_2\rangle]
363: \ge \sup_{U\in{\cal F}_U}e_p(U).
364: \end{eqnarray}
365: Therefore using Eq. (\ref{bound}) one obtains
366: $e({\cal F})=\sup_{U\in{\cal F}_U} e_p(U_\lambda).$
367: Notice also that for such a family the maximization over the eigenvalue index $i$ in Eq. (\ref{ep}) is irrelevant.
368: $\hfill\Box$
369: 
370: \section{Examples}
371: 
372: We will now illustrate the use of the general notions introduced so far
373: by considering in a detailed fashion some concrete Hamiltonian
374: families acting on a two-qubits space. Before doing that let us
375: remind a few basic facts about two-qubits entanglement in pure
376: states. We  denote the standard  product basis by
377: $|\Psi_i\rangle,\,(i=1,\ldots,4) $ and consider a generic
378: two-qubits state $\ket{\Phi}=U \ket{\Psi}=\sum_{i=1}^{4} a_i
379: \ket{\Psi_i}$. The eigenvalues of the associated reduced density
380: matrix
381:  are given by $\lambda=(1+\sqrt{1-4C^2})/2$ and $1-\lambda$, where
382: $C^2=\left| a_1 a_2- a_3 a_4\right| ^{2}$ and $2\,C$ is the so called
383: 'concurrence'. The entanglement measure is given by $E=-\left[
384: \lambda \log_2 \lambda +\left( 1-\lambda \right) \log_2 \left(
385: 1-\lambda \right) \right] $. Since $\frac{dE}{d\lambda}<0$,
386: finding the maximum possible entanglement for the output state
387: $\ket{\Phi}$ means minimizing $\lambda$, or, which is the same,
388: maximizing $C^2$. The state $\ket{\Phi}$ is maximally entangled
389: for $\lambda=\frac{1}{2}$, or $C^2=\frac{1}{4}$.
390: 
391: {\em Example 0.}-- It is useful to start with an example of a
392: two-qubits  Hamiltonian family with {\em zero} adiabatic
393: entangling power. Let $H(\lambda)=\sum_{\alpha=x,y,z}
394: \lambda_\alpha \sigma_\alpha\otimes\sigma_\alpha,$ where the
395: $\lambda'$s are such that the corresponding hamiltonian is always
396: not degenerate One has that
397: $[H(\lambda),\,H(\lambda^\prime)]=0,\,(\forall\lambda,\lambda^\prime),$
398: then all the elements of the family can be simultaneously
399: diagonalized. The joint eigenvectors are clearly given by the
400: Bell's basis
401: $|\Phi^\pm\rangle:=1/\sqrt{2}(|00\rangle\pm|11\rangle),\,|\Psi^\pm\rangle:=1/\sqrt{2}(|10\rangle\pm|01\rangle).$
402: Entanglement in the eigenstates is therefore maximal an cannot be
403: changed by varying the control parameters $\lambda.$ Analogously
404: one can easily build examples of Hamiltonian families having joint
405: constant eigenvectors given by products.
406: 
407: 
408: \begin{figure}
409: \putfig{aep_fig1}{6} \caption{(Color online) Entanglement generated by the
410: hamiltonian of the example 1 for the input state $\ket{01}$ as a
411: function of the parameters $\mu,\mu_z$.} \label{fig1}
412: \end{figure}
413: 
414: {\em Example 1.}--
415: The  non-degenerate Hamiltonian we  consider is the following
416: \begin{equation}
417: H_0= \lambda_1 \sigma_z\otimes \openone + \lambda_2  \openone
418: \otimes\sigma_z,\quad (\lambda_1\neq\lambda_2)
419: \end{equation}
420: The eigenvectors are given by the standard product basis.
421: %($|\lambda_1|\neq |\lambda_2|$)
422: %Eigenvalues and the associated eigenvectors of $H_0$ are given by
423: %$\{ -1-(\lambda_1-\lambda_2),\, -1+(\lambda_1-\lambda_2),\, 1-(\lambda_1+\lambda_2),\,
424: %1+(\lambda_1+\lambda_2)\}$ and $\{ |01\rangle,\,|10\rangle,\,|00\rangle,\,|11\rangle\}$
425: We introduce  the family of unitaries $U(\mu,\mu_z)=\exp[i K(\mu,\mu_z)]$
426: where
427: \begin{equation}
428: K(\mu,\mu_z):=\mu \sigma^+\otimes\sigma^-+\bar{\mu} \sigma^-\otimes\sigma^+
429: +\mu_z(\sigma_z\otimes \openone - \openone \otimes\sigma_z)
430: \label{trans}
431: \end{equation}
432: and the associated iso-spectral family
433: $H(\mu,\mu_z):=U(\mu,\mu_z) H_0 U(\mu,\mu_z)^\dagger.$
434: The Hilbert space is given by $\h={\rm span}\left\{\ket{00}
435: ,\ket{01},\ket{10} ,\ket{11} \right\}$
436:  and we can split it in the two subspaces $\h_0={\rm span}\left\{\ket{00}
437: ,\ket{11}\right\}$
438:  and $\h_1={\rm span}\left\{ \ket{01} ,\ket{10}\right\}$, where obviously
439: $\h=\h_0\oplus \h_1$.
440: 
441: The evolution operator $U$ is the identity on $\h_0$ while it is a
442: straightforward exercise to verify that on $\h_1$ it yields:
443: $U\ket{01} \equiv \ket{\xi}\equiv a \ket{01} +b\ket{10}$ and $U\ket{10}
444: \equiv \ket{\zeta} \equiv-
445: \overline{b}\ket{01} +\overline{a}\ket{10}$, where $a=\cos \theta
446: +\frac{2i\sin \theta }{\theta }\mu _{z}$,
447:  $b=\frac{4i\sin \theta }{\theta }\overline{\mu}$ and
448:   $\vec \theta \equiv 2(\mu +\overline{\mu},i(\mu -\overline{\mu }),\mu _{z})$.
449: %$\theta=2\sqrt{4|\mu|^2+\mu_z^2}$
450:  For the
451: generic state $\ket{\Psi}= \alpha\ket{01} +\beta\ket{10}
452: +\gamma\ket{00}+ \delta\ket{11}$ one has 
453: $C^2=\left| xy-\gamma\delta\right|^{2},$ where $x=\alpha
454: a-\beta\overline{b}$ and $y=\alpha b+\beta\overline{a}$.
455: 
456: For  $\ket{01}$ the evolved state is $\left| \xi \right\rangle
457: =a\ket{01} +b\ket{10} $ and its reduced density
458: matrix is obviously $\rho =\rm{diag}(\left| a\right| ^{2},\left|b\right|^{2})$
459: whose eigenvalues are $\left|a\right|^{2}$ and $1-\left| a\right| ^{2}$.
460: The condition to obtain a maximally entangled state is hence $\left|a\right|^{2}=\frac{1}{2}$, that is,
461: $
462: \sin ^{2}\theta =\frac{1}{2}\left[ 1+\left( \frac{\mu _{z}}{2\left| \mu
463: \right| }\right) ^{2}\right]   \label{eq1}$. This equation admits (at
464: least) one solution iff
465: $\left|\mu _{z}\right| \leq 2\left| \mu\right| .$ Thus a maximally
466: entangled state can be reached starting from either  $\ket{01} ,\ket{10}.$
467: In Fig.  \ref{fig1} is showed  the reachable entanglement from the input
468: state $|01\rangle$  
469: as a function of the parameters $\mu ,\mu _{z}.$ We see how moving in the parameter space to 
470:  higher values of $\mu_z$ spoils the reachibility of a maximally entangled state.
471: % Notice first that if we choose the state $\ket{01}$, the function to maximize
472: %becomes $\left| a\right| ^{2}\left|b\right|^{2}$. Also notice %that
473: % $U\ket{01}$ is
474: %always more entangled than $U\left( \alpha \ket{01} +\gamma
475: %\ket{00} \right) $. 
476: 
477: {\em Example 2}.--
478: Let us examine now the following unitary family:
479: $U=\exp(i \sum_{j=1}^{3} \lambda_j \sigma_j \otimes \sigma_j)
480: $
481: In the so-called magic basis $[
482: \ket{\Psi_1}=({\ket{00}+\ket{11}})/\sqrt{2},\ket{\Psi_2}=-i(\ket{00}-\ket{11})/\sqrt{2}
483: ,\ket{\Psi_3}=
484: (\ket{01}-\ket{10})/\sqrt{2},\ket{\Psi_4}=-i(\ket{01}+\ket{10})/\sqrt{2}
485: ]$
486:  (as well in the Bell basis) these unitaries are diagonal and read
487: $U=\sum_{k=1}^{4}e^{i h_k} \ket{\Psi_k} \bra{\Psi_k}$
488:  where $\{h_1=\lambda_1 -\lambda_2 +\lambda_3, h_2=\lambda_1 +\lambda_2
489: -\lambda_3, h_3=-\lambda_1 +\lambda_2 +\lambda_3,
490:  h_4=-\lambda_1 -\lambda_2 -\lambda_3\}.$ So in this basis the input state
491: is $\ket{\Psi}=\sum_{k}w_k\ket{\Psi_k}$ and the
492:   output state is $\ket{\Phi}=\sum_{k}w_k e^{-i h_k}\ket{\Psi_k}.$  The concurrence is given
493:    by $C^2=\sum_{k,l} (w_k e^{-i h_k})^2 (w^*_l e^{i h_l})^2$. Following
494: Ref \cite{kc}, we find that the
495:     maximum reachable concurrence is $C=\max_{k,l} |\sin(h_k-h_l)|$ and the
496: product input state which gives the
497:     best entangling capability as a  function of the parameters $\lambda_k$
498: is then $\frac{1}{\sqrt{2}} (\ket{\Psi_k} +i\ket{\Psi_l}).$
499:      So for instance a maximally entangled state can be reached from the
500: input state $\frac{1}{\sqrt{2}} (\ket{\Psi_1} +i\ket{\Psi_2})=|00\rangle$ for
501:       parameters such that $\lambda_3 - \lambda_2 = \pi/4$ (see Fig. \ref{fig2}).
502: 
503: Before passing to the conclusions we would like to show that the
504: first two-qubit Hamiltonian family associated with the unitaries
505: (\ref{trans}) can be used to generate a non trivial entangling
506: gate in an adiabatic fashion. 
507: 
508: {\em{Proposition 3}}
509: An adiabatic  loop in the parameter space $(\mu,\mu_z)\,(|\mu|^2+\mu_z^2=\mbox{const})$
510: gives rise to the diagonal  unitary mapping $|\alpha\beta\rangle\rightarrow
511: \exp(i\phi_{\alpha\beta})|\alpha\beta\rangle$
512: where, if $\gamma$ denotes the geometric contribution , $E_{\alpha\beta}$ the eigenvelues
513: and $T$ is the operation time, one has
514: $\phi_{01}= E_{01} \,T +\gamma,\,\phi_{10}= E_{01} \,T -\gamma,\,\phi_{00}= E_{00}\,T,\,
515: \phi_{11}= E_{11}\,T$
516: For $\phi_{01}+\phi_{10}-(\phi_{00}+\phi_{11})= -4\,T \neq 0 \mbox{mod}\, 2\pi$
517: the obtained transformation is equivalent to a controlled-phase-shift.
518: %\vskip 0.5truecm
519: 
520: {\em{Proof.}}
521: Indeed it is easy to check that
522: i) by the adiabatic theorem the evolution has to be diagonal in the
523: product basis
524: ii) the geometric contribution of the states $|\alpha\alpha\rangle\,(\alpha=0,1)$
525: is zero ($\Leftarrow U(\mu,\mu_z)|\alpha\alpha\rangle=|\alpha\alpha\rangle$)
526: iii) In the one-qubit subspace spanned by $|{\bf{0}}\rangle:= |01\rangle$ and $|{\bf{1}}\rangle:=|10\rangle$
527: the  unitaries $e^{iK}$ with the $K$ defined in (\ref{trans}) look like
528: $U(\mu,\mu_z)=exp  [i(\mu \tilde{\sigma}^+ +\bar{\mu}\tilde{\sigma}^- +\mu_z\tilde{\sigma}_z)].$
529: This latter equation can be of course written as ${\bf{B}}\cdot \dot{\bf{\sigma}}$
530: where a fictious magnetic field ${\bf{B}}$ has been introduced.
531: One can then use the standard Berry-phase argument for a spin $1/2$ particle in
532: an adiabatically  changing magnetic field to claim that under a ${\bf{B}}$ going along an adiabatic loop,
533:  one has $|{\bf{0}}\rangle\mapsto e^{i\gamma} |{\bf{0}}\rangle$
534: and $|{\bf{1}}\rangle\mapsto e^{-i\gamma} |{\bf{1}}\rangle$. Here $\gamma$ denotes the standard geometric phase
535: i.e., proportional to the solid angle swept by ${\bf{B}}.$
536: The final equivalence claim stems from a known result in literature \cite{calarco}.
537: $\hfill\Box$
538: \begin{figure}
539: \putfig{aep_fig2}{6}
540: \caption{(Color online) Entanglement for the unitary $U=e^{i \sum_{i=1}^{3} \lambda_i
541: \sigma^i \otimes \sigma^i}$, with the input state
542:  $\frac{1}{\sqrt{2}} (\ket{\Psi_1} +i\ket{\Psi_2})$ as a  function of
543: $\lambda_2,\lambda_3$, with $\lambda_1=1$.}
544: \label{fig2}
545: \end{figure}
546: 
547: Of course the general fact that entangling gates  can be obtained  via adiabatic  manipulations is not new see e.g.,\cite{HQC} \cite{ekert}.
548: The point of Prop 4. is  to show explicitly how the   particular  two-qubit  Hamiltonian family associated to the untarries
549: (\ref{trans}) can be exploited for enacting controlled phase via a simple  adiabatic protocol.
550: 
551: 
552: 
553: 
554: 
555: 
556: \section{ Conclusions.}
557: 
558:  In this paper we analyzed the entanglement
559: generation capabilities of  a parametric family of adiabatically
560: connected non-degenerate Hamiltonians. One prepares the system in 
561: a separable eigenstate of of a distinguished Hamiltonian $H_0$ in the 
562: family and then the space of parameters is adiabatically explored. 
563: The system remains then in an energy eigenstate and the (bi-partite) entanglement
564: contained in such an eiegenstate can be maximized over the
565: manifold of control parameters. 
566: We introduced an  associated measure $e$  of adiabatic entangling
567: power and discussed its properties and relations with a previously introduced measure
568: for the case of iso-spectral families of  Hamiltonians.
569: We illustrated the general ideas
570: by studying explicitly the adiabatic entangling power of concrete
571: two-qubits Hamiltonian families. We also showed how to generate a
572: non-trivial two-qubits entangling gate by means of adiabatic
573: loops.
574: 
575: 
576: 
577: We thank M. C. Abbati, A. Mani\`a, L. Faoro and an anonymous referee for useful comments.
578: P.Z. gratefully acknowledges financial support by
579: Cambridge-MIT Institute Limited and by the European Union project  TOPQIP
580: (Contract IST-2001-39215)
581: 
582: 
583: 
584: %\end{document}
585: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
586: \begin{thebibliography}{}
587: 
588: \bibitem{QIP}   D.P.~DiVincenzo and C.~Bennett, Nature {\bf 404}, 247 (2000)
589: \bibitem{aqc} E. Farhi et al, Science {\bf 292}, 472 (2001)
590: \bibitem{berk} Wim van Dam, M. Mosca, U. Vazirani, quant-ph/0206003
591: \bibitem{berk1} D. Aharonov, A. Ts-Shma,  quant-ph/0301023
592: \bibitem{mess} A. Messiah, {\em Quantum mechanics}, John Wiley and Sons (1958)
593: \bibitem{GQC} A. Jones  {\it et al.} , Nature  {\bf 403}, 869 (2000);
594: G. Falci  {\it et al.} , Nature  {\bf 407}, 355 (2000).
595: \bibitem{berry}M.V. Berry, {\it Proc. R. Soc. Lond. A} {\bf 392}, 45 (1984)
596: \bibitem{HQC}P. Zanardi and M. Rasetti,
597: {Phys. Lett. A} {\bf 264}, 94 (1999);
598: J. Pachos, P. Zanardi and M. Rasetti,
599: {\it Phys. Rev.A} {\bf 61}, 010305(R) (2000);
600: L.-M. Duan,J. I.  Cirac and P. Zoller,
601: {\it Science} {\bf 292}, 1695 (2001)
602: \bibitem{unan}
603:  Unanyan et al, Phys. Rev. Lett. {\bf{87}}, 137902 (2001);
604: S. Guérin et al,  Phys. Rev. A {\bf 66}, 032311 (2002); 
605: R. G. Unanyan et al, Phys. Rev. A {\bf 66}, 042101 (2002)
606:  R. G. Unanyan, M. Fleischhauer, quant-ph/0208144;
607: \bibitem{pz} U. Dorner et al, Phys. Rev. Lett {\bf{91}}, 073601 (2003)
608: \bibitem{PWA} P.W. Anderson,
609: Concepts in Solids: Lectures on the Theory of Solids (World Scientific Lecture Notes in Physics , Vol 58)
610: \bibitem{QPT}   A. Osterloh et al, Nature (London) {\bf 416}, 608 (2002);
611: T. J. Osborne, M. A. Nielsen, Phys. Rev. A {\bf 66}, 032110 (2002)
612: G. Vidal et al,  Phys. Rev. Lett. {\bf 90}, 227902 (2003)
613: \bibitem{loss}Here we are assuming, without loss of generality, that Ker$ H_0=\{0\}$
614: \bibitem{dur}W. D\"ur {\em et al}, Phys. Rev. Lett {\bf{87}}, 137901 (2000)
615: \bibitem{ZZF} P. Zanardi {\em et al},  Phys. Rev A,  {\bf 62}  030301(R) (2000)
616: \bibitem{kc} B. Krauss, I. Cirac,  Phys. Rev. A {\bf 63}, 062309 (2001)
617: \bibitem{ekert} Ekert et al, J. Mod. Opt. {\bf{47}}, 2501 (2000))
618: \bibitem{calarco} See appendix B.1 of T. Calarco, I. Cirac, P. Zoller, Phys. Rev. A {\bf 63},
619:  62304 (2001) 
620: \end{thebibliography}
621: 
622: \end{document}
623: The condition iso-degeneracy  is clearly necessary,
624: because otherwise one would have (at least) one level crossing
625: that in turn would prevent the applicability of the adiabatic
626: theorem. 
627: To prove the converse let us observe that the set   ${\cal I}_D(d_1,\ldots,d_R)$ of 
628: $D$-dimensional Hamiltonians with $R$ distinct eigenvalues $\{\epsilon_i\}_{i=1}^R$  
629: each of them with degeneracy $d_i$ is parametrized
630: by the product of the quotient   $ U(D)/\prod_{i=1}^R U(d_i)$
631: with set $\{x\in \RR^R\,/\, x_i\neq x_j, \forall i,j,\,(i\neq j) \}$
632: Suppose now that $H_0$ and $H_1$ are iso-degenerate, on
633: can then write $H_\alpha=\sum_{i=1}^R \epsilon_\alpha^i
634: \Pi_\alpha^i \,(\alpha=0,1)$ where tr$ \Pi_\alpha^i=d_i$ is the
635: degeneracy of the $i$-th eigenvalue (of {\em both} $H_0$ and
636: $H_1$). It is an elementary fact that, given the two systems of
637: ortho-projectors $\{\Pi_\alpha^i\}_{i=1}^R, \,(\alpha=0,1)$ such
638: that Tr$\Pi_1^i=\mbox{Tr} \Pi_2^i,\,(i=1,\ldots,R),$ it exist a
639: (non-unique ) unitary $U$ such that $U
640: \,\Pi_0^i\,U^\dagger=\Pi_1^i,\,(i=1,\ldots,R);$ it follows that
641: $H_1= U \,H_0 \,U^\dagger.$ Therefore if one consider any
642: one-parameter family of unitaries $\{U_t\}_{t\in[0,1]}$ with
643: $U_0=\openone$ and $U_1=U,$ then the  Hamiltonian family $H_t:=
644: W_t U_t \,H_0 \,U^\dagger_t W_t$ where $W_t= \sum_{i=1}^R
645: \sqrt{\chi_i(t)} U_t \,\Pi_0^i\,U_t^\dagger,\,
646: \chi_i(0)=1,\,\chi_i(1)=\epsilon_1^i/\epsilon_0^i$ \cite{loss} adiabatically
647: connects $H_0$ and $H_1$ if
648: $i\neq j\Rightarrow \chi_i(t)\epsilon_0^i\neq \chi_j(t)\epsilon_0^j \forall t\in [0,1].$
649: This latter condition amounts to say that $H_0$ and $H_1$ are in
650: the same connected component of ${\cal I}_D(d_1,\ldots,d_R)$
651: 
652: