1: %% ****** Start of file template.aps ****** %
2: %%
3: %%
4: %% This file is part of the APS files in the REVTeX 4 distribution.
5: %% Version 4.0 of REVTeX, August 2001
6: %%
7: %%
8: %% Copyright (c) 2001 The American Physical Society.
9: %%
10: %% See the REVTeX 4 README file for restrictions and more information.
11: %%
12: %
13: % This is a template for producing manuscripts for use with REVTEX 4.0
14: % Copy this file to another name and then work on that file.
15: % That way, you always have this original template file to use.
16: %
17: % Group addresses by affiliation; use superscriptaddress for long
18: % author lists, or if there are many overlapping affiliations.
19: % For Phys. Rev. appearance, change preprint to twocolumn.
20: % Choose pra, prb, prc, prd, pre, prl, prstab, or rmp for journal
21: % Add 'draft' option to mark overfull boxes with black boxes
22: % Add 'showpacs' option to make PACS codes appear
23: % Add 'showkeys' option to make keywords appear
24: %
25: %
26: \documentclass[aps,pra,twocolumn,groupedaddress,amsmath,amssymb,showpacs]{revtex4}
27: %\documentclass[aps,pra,preprint,groupedaddress,amsmath,amssymb,showpacs]{revtex4}
28: %
29: %
30: \usepackage{graphicx}
31: \usepackage{amsmath}
32: \usepackage{dcolumn}% Align table columns on decimal point
33: \usepackage{bm}% bold math
34: \bibliographystyle{apsrev}
35: %
36: \newcommand{\tr}{\mathrm{Tr}}
37: \newcommand{\inp}{\mathrm{in}}
38: \newcommand{\out}{\mathrm{out}}
39:
40: \begin{document}
41: %
42: \title{Entangled mixed-state generation by twin-photon scattering}
43: \author{G. Puentes$^1$}
44: \author{A. Aiello$^1$}
45: \author{D. Voigt$^{1,2}$}
46: \author{J.P. Woerdman$^1$}
47: \affiliation{1. Huygens Laboratory, Leiden University, P.O. Box
48: 9504, 2300 RA Leiden, The
49: Netherlands. \\
50: 2. Cosine Research bv, Niels Bohrweg 11, 2333 CA Leiden, The
51: Netherlands.}
52: %
53: \begin{abstract}
54: We report experimental results on mixed-state generation by
55: multiple scattering of polarization-entangled photon pairs created
56: from parametric down-conversion. By using a large variety of
57: scattering optical systems we have experimentally obtained
58: entangled mixed states that lie upon and below the Werner curve in
59: the linear entropy-tangle plane. We have also introduced a simple
60: phenomenological model built on the analogy between classical
61: polarization optics and quantum maps. Theoretical predictions from
62: such model are in full agreement with our experimental findings.
63: %
64: \end{abstract}
65:
66: \pacs{03.67.Mn, 42.50.Ct, 42.50.Dv, 42.65.Lm}
67:
68:
69:
70: \maketitle
71: %42.50.Ct Quantum description of interaction of light and
72: %matter; related experiments.
73:
74: %03.67.Mn Entanglement production, characterization, and
75: %manipulation
76:
77: %42.50.Dv Nonclassical states of the electromagnetic field,
78: %including entangled photon states; quantum state engineering and
79: %measurements
80:
81: %42.65.Lm Parametric down conversion and production of entangled
82: %photons
83: %
84: %03.65.Db Functional analytical methods in quantum mechanics
85: %
86: %03.65.Ca Formalism in quantum mechanics
87: %
88: %03.70.+k Theory of quantized fields
89: %
90: %41.20.Jb Electromagnetic wave propagation; radiowave propagation
91: %
92: %41.85.-p Beam optics
93: %
94: %03.65.Ud Entanglement and quantum nonlocality (e.g. EPR
95: %paradox, Bell's inequalities, GHZ states, etc.) (for entanglement
96: %production in quantum information, see 03.67.Mn; for entanglement
97: %in Bose-Einstein condensates, see 03.75.Gg)
98: %
99: %03.65.Nk Scattering theory
100: %
101: %03.67.Mn Entanglement production, characterization, and
102: %manipulation (see also 03.65.Ud Entanglement and quantum
103: %nonlocality; for entanglement in Bose-Einstein condensates, see
104: %03.75.Gg)
105: %
106: %42.25.Ja Polarization
107: %
108: %42.50.Dv Nonclassical states of the electromagnetic field,
109: %including entangled photon states; quantum state engineering and
110: %measurements (see also 03.65.Ud Entanglement and quantum
111: %nonlocality, e.g. EPR paradox, Bell's inequalities, GHZ states,
112: %etc.)
113: %
114: %
115: \section{Introduction}
116: %
117: %
118: The study of spatial, temporal and polarization correlations of
119: light scattered by inhomogeneous and turbid media has a history of
120: more than a century \cite{rayleigh}. Due to the high complexity of
121: scattering media only single-scattering properties are known at a
122: microscopic level \cite{hulst}. Conversely, for
123: multiple-scattering processes the emphasis is mainly on
124: macroscopic theoretical descriptions of the correlation phenomena
125: \cite{rossum}. In most examples of the latter
126: \cite{albada1,chabanov,rikken,freund} the intensity correlations
127: of the interference pattern generated by multiple-scattered light
128: are explained in terms of \emph{classical} wave-coherence. On the
129: other hand, the recent availability of reliable single-photon
130: sources has triggered the interest in \emph{quantum} correlations
131: of multiple-scattered light \cite{lohdal}. Generally speaking,
132: quantum correlations of scattered photons depend on the quantum
133: state of the light illuminating the sample. In Ref. \cite{lohdal},
134: \emph{spatial} quantum correlations of scattered light were
135: analyzed for Fock, coherent and thermal input states.
136: %
137:
138: In this paper we present the first experimental results on quantum
139: polarization correlations of scattered photon pairs. Specifically,
140: we study the entanglement content in relation to the polarization
141: purity of multiple-scattered twin-photons, initially generated in
142: a polarization-entangled state by spontaneous parametric
143: down-conversion (SPDC). The initial entanglement of the input
144: photon pairs will in general be degraded by multiple scattering.
145: This can be understood by noting that the scattering process
146: distributes the initial correlations of the twin-photons over the
147: many spatial modes excited along the propagation in the medium. In
148: the case of spatially inhomogeneous media the polarization degrees
149: of freedom are coupled to the spatial degrees of freedom
150: generating polarization dependent speckle patterns. If the spatial
151: correlations of such patterns are averaged out by multi-mode
152: detection, the polarization state of the scattered photon(s) is
153: reduced to a mixture, and the resulting polarization-entanglement
154: of the photon pairs is degraded with respect to the initial one. A
155: related theoretical background was elaborated in
156: \cite{aiello1,vanvelsen}.
157: %
158:
159: This paper is structured as follows: In section II we report our
160: experiments on light scattering with entangled photons. First, we
161: present our experimental set-up and briefly describe the many
162: different optical systems that we used as scatterers. Next, we
163: show our experimental results. The notions of generalized Werner
164: and sub-Werner states are introduced to illustrate these results.
165: %
166: %In addition, we also discuss how we determined experimental
167: %errors.
168: %
169: In section III we introduce a simple phenomenological model for
170: photon scattering that fully reproduces our experimental findings.
171: Finally, in section IV we draw our conclusions.
172: %
173: \section{Experiments on light scattering with entangled photons}
174: %
175: \subsection{Experimental set-up}
176: %
177: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
178: %
179: Our experimental set-up is shown in Fig.~\ref{fig:1}.~A
180: Krypton-ion laser at 413.1~nm pumps a 1~mm thick
181: $\beta-\mathrm{Ba} \mathrm{B}_2 \mathrm{O}_4$ (\textsf{BBO})
182: crystal, where polarization-entangled photon pairs at wavelength
183: 826.2~nm are created by SPDC in degenerate type II phase-matching
184: configuration \cite{sergienko}. Single-mode fibers (\textsf{SMF})
185: are used as spatial filters to assure that each photon of the
186: initial SPDC pair travels in a single transverse mode. Spurious
187: birefringence along the fibers is compensated by suitably oriented
188: polarization controllers (\textsf{PC}). The total retardation
189: introduced by the fibers and walk-off effects at the \textsf{BBO}
190: crystal are compensated by compensating crystals (\textsf{CC}:
191: 0.5~mm thick \textsf{BBO} crystals) and half-wave plates
192: ($\lambda/2$), in both signal and idler paths. In this way the
193: initial two-photon state is prepared in the polarization singlet
194: state $|\psi_s \rangle=(|HV\rangle-|VH\rangle)/\sqrt{2}$, where
195: $H$ and $V$ are labels for horizontal and vertical polarizations
196: of the two photons, respectively.
197: %
198: %
199: \begin{figure}[!h]
200: \includegraphics[angle=0,width=8truecm]{fig1}
201: \caption{\label{fig:1}(Color online) Experimental scheme: After
202: singlet preparation, the idler photon propagates through the
203: scattering system $\mathcal{T}_A$. The polarization state of the
204: scattered photon-pairs is then reconstructed via a quantum
205: tomographic procedure (see text for details).}
206: \end{figure}
207: %
208: The experimentally prepared initial singlet state
209: $\rho_s^\mathrm{exp}$ has a fidelity \cite{fidelity} with the
210: theoretical singlet state $\rho_s = | \psi_s \rangle \langle
211: \psi_s |$ of $F(\rho_s,\rho_s^\mathrm{exp}) \sim 98\%$.
212: %
213: %
214: In the second part of the experimental set-up the idler photon
215: passes though the scattering device $\mathcal{T}_A$ before being
216: collimated by a photographic objective (\textsf{PO}) with focal
217: distance $f = 5~\mathrm{cm}$.
218: %
219: %
220: The third and last part of the experimental set-up, consists of
221: two tomographic analyzers (one per photon), each made of a
222: quarter-wave plate ($\lambda/4$) followed by a linear polarizer
223: (\textsf{P}). Such analyzers permit a full tomographic
224: reconstruction, via a maximum-likelihood technique \cite{james},
225: of the two-photon state. Additionally, interference filters
226: (\textsf{IF}) put in front of each detector ($\Delta \lambda = 5$~
227: nm) provide for bandwidth selection. Detectors
228: $\mathrm{\textsf{D}}_\mathrm{\textsf{A}}$ and
229: $\mathrm{\textsf{D}}_\mathrm{\textsf{B}}$ are ``bucket''
230: detectors, that is they do not distinguish which spatial mode a
231: photon comes from, thus each photon is detected in a
232: \emph{mode-insensitive} way.
233: %
234: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
235: %
236: \subsection{Scattering devices}
237: %
238: All the scattering optical systems that we used were located in
239: the path of only \emph{one} of the photons of the entangled-pair
240: (the idler one), as shown in Fig.~\ref{fig:1}. For this reason, we
241: refer to such systems as \emph{local} scatterers. Such scatterers
242: can be grouped in three general categories according to the
243: optical properties of the media they are made of \cite{puentes}:
244: %
245: \begin{description}
246: \item[Type I] Purely depolarizing media, or diffusers. Such media do not affect directly
247: the polarization state of the impinging light but change the
248: spatial distribution of the impinging electromagnetic field.
249: %
250: \item[Type II] Birefringent media, or retarders. These media introduce
251: a polarization-dependent \emph{delay} between different components
252: of the electromagnetic field.
253: %
254: \item[Type III] Dichroic media, or diattenuators. Such media introduce
255: polarization-dependent \emph{losses} for the different components of the
256: electromagnetic field.
257: \end{description}
258: %
259: Type I scattering systems produce an isotropic spread in the
260: momentum of the impinging photons. Examples of such scattering
261: devices are: spherical-particle suspensions (such as milk or
262: polymer micro-spheres), polymer and glass multi-mode fibers and
263: surface diffusers.
264: %
265: Type II scattering systems are made of birefringent
266: media, which introduce an optical axis that breaks polarization-isotropy.
267: Birefringence can be classified as ``material birefringence''
268: when it is an intrinsic property of the bulk medium (for example a
269: birefringent wave-plate), and as ``topological birefringence''
270: when it is induced by a special geometry of the system that
271: generates polarization-anisotropy, an example of a system with
272: topological birefringence is an array of cylindrical particles.
273: %
274: Finally, type III scattering systems are made of dichroic media
275: that produce polarization-dependent photon absorbtion. Examples of
276: such devices are commonly used polarizers.
277: %
278: A systematic characterization of all
279: the scattering devices that we used was given in Ref.
280: \cite{puentes}.
281: %
282: \subsection{ Experimental results \\
283: in the tangle versus linear entropy plane }
284: %
285: The degree of entanglement and the degree of mixedness of the
286: scattered photon pairs can be quantified by the tangle ($T$),
287: namely, the concurrence squared \cite{wooters}, and the linear
288: entropy ($S_L$) \cite{vedral}. These quantities were calculated
289: from the $4 \times 4$ polarization two-photon density matrix
290: $\rho$, by using $T(\rho)$=(max$\{0,\sqrt{\lambda_{1}}-
291: \sqrt{\lambda_{2}}-\sqrt{\lambda_{3}}-\sqrt{\lambda_{4} }\})^2$,
292: where $\lambda_{1} \geq \lambda_{2} \geq\lambda_{3}
293: \geq\lambda_{4}\geq 0 $ are the eigenvalues of $\rho(\sigma_{2}
294: \otimes \sigma_{2})\rho^{*}(\sigma_{2} \otimes \sigma_{2})$, where
295: $\sigma_2 = \left[ \begin{array}{cc}
296: 0 & -i \\
297: i & 0
298: \end{array} \right]$, and
299: $S_{L}(\rho)=\frac{4}{3}[1-\mathrm{Tr}(\rho^2)]$.
300: %
301: % The density
302: %matrix $\rho$ was reconstructed via maximum likelihood tomography
303: %\cite{james} from coincidence measurements.
304: %
305: \begin{figure}[ht]
306: \includegraphics[angle=0,width=7truecm]{fig2}
307: \caption{\label{fig:2} Experimental data in the linear
308: entropy-tangle ($S_{L}$-$T$) plane. The grey area corresponds to
309: unphysical density matrices. Dashed upper curve: Maximally
310: entangled mixed states (MEMS), continuous lower curve: Werner
311: states. (a)~Polarization-isotropic scatterers (type I).
312: (b)~Birefringent scatterers (type II).}
313: \end{figure}
314: %
315: Figures~\ref{fig:2}~(a) and (b) show experimental data reported on
316: the linear entropy-tangle plane. The position of each
317: experimental point in such plane has been calculated from a
318: tomographically reconstructed \cite{james} two-photon density
319: matrix $\rho^\mathrm{exp}$. The uniform grey area corresponds to
320: non-physical states \cite{mems}. The dashed curve that bounds the
321: physically admissible region from above is generated by the
322: so-called maximally entangled mixed states (MEMS)
323: \cite{mems2,mems3}. The lower continuous curve is produced by the
324: Werner states \cite{werner} of the form: $\rho_{W}=p \rho_s
325: +\frac{1-p}{4}I_4$, ($0\leq p \leq 1 $), where $I_4$ is the $4
326: \times 4$ identity matrix.~Figure~\ref{fig:2}~(a) shows
327: experimental data generated by isotropic scatterers (type I).
328: Specifically, our type I scatterers consisted of the following
329: categories. (i) Suspensions of milk and micro-spheres in distilled
330: water, where the sample dilution was varied to obtain different
331: points; (ii) Multi-mode glass and polymer fibers, where the tuning
332: parameter exploited to obtain different points was the length of
333: the fiber (cut-back method); (iii) Surface diffusers, where the
334: full width scattering angle was used as tuning parameter. It
335: should be noted that suspensions of milk and micro-spheres are
336: \emph{dynamic} media, where Brownian motion of the micro-particles
337: induces temporal fluctuations within the detection integration
338: time \cite{puentes}.
339:
340: In Fig.~2~(a), the experimental point at the top-left corner
341: (nearby $T=1$, $S_L=0$), is generated by the un-scattered initial
342: singlet state. The net effect of scattering systems with
343: increasing thickness is to shift the initial datum toward the
344: bottom-right corner ($T=0$, $S_L=1$), that corresponds to a fully
345: mixed state.
346:
347: Figure~\ref{fig:2}~(b) displays experimental data generated by
348: birefringent scattering systems (type II). As an example of a
349: system with ``material birefringence'' we used a pair of wedge
350: depolarizers in cascade \cite{kliger}. Different experimental
351: points where obtained by varying the relative angle between the
352: optical axis of the two wedges \cite{puentes2}.
353: %
354: The systems with ``topological birefringence'' we considered
355: consisted of two different devices:
356: %
357: (i) The first one was a bundle of parallel optical fibers
358: \cite{vdmark}. Translational invariance along the fibers axes
359: restricts the direction of the wave-vectors of the scattered
360: photons in a plane orthogonal to the common axis of the fibers.
361: %
362: (ii) The second device was a stack of parallel microscope slides
363: (with uncontrolled air layers in between). This optical system is
364: depolarizing because it amplifies any initial spread in the
365: wave-vector of the impinging photon. This photon enters via a
366: single-mode-fiber (numerical aperture $\mathrm{NA} = 0.12$), from
367: one side of the stack and travels in a plane parallel to the
368: slides.
369: %
370:
371: In Fig.~\ref{fig:3}, experimental data generated by dichroic
372: scattering systems (type III) are shown. We used:
373: (i) Surface diffusers
374: followed by a stacks of microscope slides at the Brewster angle
375: and (ii) Commercially available polaroid sheets with
376: manually-added surface roughness on its front surface to provide
377: for wave-vector spread.
378: %
379: All data thus obtained fall below the Werner curve, generating
380: what we called sub-Werner states, namely states with a lower value
381: of tangle ($T$) than a Werner state, for a given value of the
382: linear entropy ($S_L$).
383: %
384: \begin{figure}[!h]
385: \includegraphics[angle=0,width=6.5truecm]{fig3}
386: \caption{\label{fig:3} Experimental data generated by dichroic
387: scattering systems (type III).}
388: \end{figure}
389: %
390:
391: In summary, Figs.~\ref{fig:2} (a)-(b) show that all data generated
392: by type I and II scattering systems fall on the Werner curve,
393: within the experimental error; while data generated by scattering
394: samples type III, which are presented in Fig.~\ref{fig:3}, lay
395: below the Werner curve. In Section III we shall present a simple
396: theoretical interpretation for such results.
397: %
398: \subsection{Error estimate}
399: %
400: In order to estimate the errors in our measured data, we
401: numerically generated 16 Monte Carlo sets $N_i$ ($i=1,\ldots,16$)
402: of $10^{3}$ simulated photon counts, corresponding to each of the
403: $16$ actual coincidence count measurements $\{n_{i}^\mathrm{exp}
404: \}$ ($i=1,\ldots,16$) required by tomographic analysis to
405: reconstruct a single two-photon density matrix.
406: %
407: Each set $N_i$ had a Gaussian distribution centered around the
408: mean value $\mu_{i} = n_{i}^\mathrm{exp}$, with standard deviation
409: $\sigma_{i}=\sqrt{n_{i}^\mathrm{exp}}$. The sets $N_i$ where
410: created by using the ``NormalDistribution'' built-in function of
411: the program Mathematica 5.2.
412: % \cite{Wolfram}.
413: Once we generated the $16$ Monte Carlo sets $N_i$, we
414: reconstructed the corresponding $10^3$ density matrices using a
415: maximum likelihood estimation protocol, to assure that they could
416: represent physical states. Finally, from this ensemble of matrices
417: we calculated the average tangle $T^\mathrm{av}$ and linear
418: entropy $S_{L}^\mathrm{av}$. The error bars were estimated as the
419: absolute distance between the mean quantities ($\mathrm{av}$) and
420: the measured ones ($\mathrm{exp}$):
421: $\sigma_{T}=|T^\mathrm{exp}-T^\mathrm{av}|$,
422: $\sigma_{S_{L}}=|S_{L}^\mathrm{exp}-S_{L}^\mathrm{av}|$. It should
423: be noted that this procedure produces an overestimation of the
424: experimental errors. In the cases where part of the overestimated
425: error bars fell into the unphysical region, the length of such
426: bars was limited to the border of the physically allowed density
427: matrices.
428: %
429: \subsection{Generalized Werner states}
430: %
431: Close inspection of the reconstructed density matrices generated
432: by type II scattering systems revealed that in some cases the
433: measured states represented a generalized form of Werner states.
434: These are equivalent to the original Werner states $\rho_W$ with
435: respect to their values of $T$ and $S_L$, but the form of their
436: density matrices is different. Werner states $\rho_{W}$ of two
437: qubits were originally defined \cite{werner} as such states which
438: are $U \otimes U$ invariant: $\rho_{W}=U \otimes U \rho_{W}
439: U^{\dagger}\otimes U^{\dagger}$. Here $U \otimes U$ is any
440: symmetric separable unitary transformation acting on the two
441: qubits. The generalized Werner states $\rho_{GW}$ we
442: experimentally generated, can be obtained from $\rho_{W}$ by
443: applying a \emph{local} unitary operation $V$ acting upon only one
444: of the two qubits:
445: $\rho_{GW}=V\otimes I \rho_{W} V^{\dagger}\otimes I$, where $I=\left[\begin{array}{cc} 1 & 0\\
446: 0 & 1 \\
447: \end{array}\right]$, and
448: %
449: \begin{eqnarray}\label{first}
450: \label{eq:1}
451: V(\alpha,\beta,\gamma)=\left[\begin{array}{cc} e^{-\frac{i}{2}(\alpha+\beta)}\cos{\gamma/2} & -e^{-\frac{i}{2}(\alpha-\beta)}\sin{\gamma/2}\\
452: e^{\frac{i}{2}(\alpha-\beta)}\sin{\gamma/2} &
453: e^{\frac{i}{2}(\alpha+\beta)}\cos{\gamma/2}\\
454: \end{array}\right],
455: \end{eqnarray}
456: %
457: where $\alpha, \beta, \gamma$ can be identified with the three
458: Euler angles characterizing an ordinary rotation in $\mathbb{R}^3$
459: \cite{nielsen}. These generalized Werner states have the same
460: values of $T$ and $S_L$ as the original $\rho_{W}$ (since a local
461: unitary transformation does not affect neither the degree of
462: entanglement nor the degree of purity) but are no longer invariant
463: under unitary transformations of the form $U \otimes U$. By using
464: Eq. (\ref{first}), we calculated the average maximal fidelity of
465: the measured states $\rho^\mathrm{exp}_{GW}$ with the target
466: generalized Werner states
467: $\rho^\mathrm{th}_{GW}(p,\alpha,\beta,\gamma)$. We found
468: $\bar{F}(\rho^\mathrm{exp}_{GW},\rho^\mathrm{th}_{GW})\approx 96
469: \%$, revealing that our data are well fitted by this
470: four-parameter class of generalized Werner states.
471: %
472: %
473: \section{The phenomenological model}
474: %
475: \begin{figure}[h!]
476: \includegraphics[angle=0,width=9.5truecm]{fig4}
477: \caption{\label{fig:4} Numerical simulation for our
478: phenomenological model. Fig. 4 (a) isotropic and birefringent
479: scattering, Fig. 4 (b) dichroic scattering.}
480: \end{figure}
481:
482: In Ref. \cite{aiello3}, a theoretical study of the analogies
483: between classical linear optics and quantum maps was given. Within
484: this theoretical framework it is possible to build a simple
485: phenomenological model capable of explaining all our experimental
486: results. To this end let us consider the experimental set-up
487: represented in Fig. \ref{fig:1}. The linear optical scattering
488: element $\mathcal{T}_A$ inserted across path $A$ can be
489: \emph{classically} represented by some Mueller matrix
490: $\mathcal{M}$ \cite{hulst} which describes its
491: polarization-dependent interaction with a classical beam of
492: light. However, $\mathcal{T}_A$ can also be represented by a
493: linear, completely positive, local \emph{quantum} map
494: $\mathcal{E}: \, \rho \rightarrow \mathcal{E}[\rho]$, which
495: describes the interaction of the scattering element with a
496: two-photon light beam encoding a pair of polarization qubits.
497: These qubits are, in turn, represented by a $4 \times 4$ density
498: matrix $\rho$. Since $\mathcal{T}_A$ interacts with only one of
499: the two photons, the map $\mathcal{E}$ is said to be \emph{local}
500: and it can be written as $\mathcal{E} = \mathcal{E}_A \otimes
501: \mathcal{I}$, where $\mathcal{E}_A$ is the single-qubit (or
502: single-photon) quantum map representing $\mathcal{T}_A$, and
503: $\mathcal{I}$ is the single-qubit identity map.
504: %
505:
506: It can be shown that the classical Mueller matrix $\mathcal{M}$
507: and the single-qubit quantum map $\mathcal{E}_A$ are univocally
508: related. Specifically, if with $\mathcal{M}$ we denote the
509: complex-valued Mueller matrix written in the standard basis, then
510: the following decomposition holds:
511: %
512: \begin{equation}\label{Mueller1}
513: \mathcal{M}=\sum_{\mu = 0}^3 \lambda_{\mu} T_{\mu} \otimes
514: T_{\mu}^*,
515: \end{equation}
516: %
517: where $\{ T_\mu \}$ is a set of four $2 \times 2$ Jones matrices
518: \cite{hulst}, each representing a non-depolarizing linear optical
519: element in classical polarization optics, and $\{ \lambda_\mu\}$ are the
520: four non-negative eigenvalues of the ``dynamical'' matrix $H$
521: associated to $\mathcal{M}$. Given Eq. (\ref{Mueller1}), it is
522: possible to show that the two-qubit quantum map $\mathcal{E}$ can
523: be written as
524: %
525: %
526: \begin{eqnarray}\label{deco}
527: %
528: \rho_\mathcal{E} = \mathcal{E}[\rho] \propto \sum_{\mu = 0}^3
529: \lambda_\mu \,
530: T_\mu \otimes I \, \rho \, T_\mu^\dagger
531: \otimes I,
532: \end{eqnarray}
533: %
534: where the proportionality symbol ``$\propto$'' on the right hand
535: side of Eq.~(\ref{deco}) accounts for a possible renormalization
536: to ensure $\tr (\rho_\mathcal{E})=1$. Such renormalization becomes
537: necessary when $\mathcal{T}_A$ presents polarization-dependent
538: losses (i.e., dichroism). We anticipate that when such
539: re-normalization is necessary the map is considered non-trace
540: preserving. We shall briefly discuss this issue in the conclusion.
541: %
542:
543: With these ingredients, a phenomenological polarization-scattering
544: model can be built as follows. First we use the polar
545: decomposition \cite{Lu96} to write an arbitrary Mueller matrix
546: $\mathcal{M}=\mathcal{M}_{\Delta}\mathcal{M}_{B}\mathcal{M}_{D}$,
547: where $\mathcal{M}_{\Delta}$, $\mathcal{M}_{B}$ and
548: $\mathcal{M}_{D}$ represent a purely depolarizing element, a
549: birefringent (or retarder) element, and a dichroic (or
550: diattenuator) element, respectively. Specific analytical
551: expressions for $\mathcal{M}_{\Delta}$, $\mathcal{M}_{B}$ and
552: $\mathcal{M}_{D}$ can be found in the literature \cite{kliger}.
553: Second, we use Eq.~(\ref{Mueller1}) to find the quantum maps
554: corresponding to $\mathcal{M}_{\Delta}$, $\mathcal{M}_{B}$ and
555: $\mathcal{M}_{D}$ and, by using such maps, we calculate the
556: scattered two-photon state $\rho_{\mathcal{E}}$. In our
557: experimental realizations we used isotropic scatterers
558: $\mathcal{M}_{IS}=\mathcal{M}_{\Delta}$ with isotropic
559: depolarization factor $0 \leq \Delta < 1$, birefringent scattering
560: media $\mathcal{M}_{BS}$, described in terms of the product of a
561: purely birefringent medium $\mathcal{M}_{B}$ and an isotropic
562: depolarizer $\mathcal{M}_{\Delta}$, i.e.
563: $\mathcal{M}_{BS}=\mathcal{M}_{B}\mathcal{M}_{\Delta}$, and
564: finally, dichroic scattering media
565: $\mathcal{M}_{DS}=\mathcal{M}_{D}\mathcal{M}_{\Delta}$, which are
566: in turn described by a product of a purely dichroic medium
567: $\mathcal{M}_{D}$ and a purely depolarizing medium
568: $\mathcal{M}_{\Delta}$. It should be noted that these product
569: decompositions are not unique. Other decompositions with different
570: orders are possible but the elements of each matrix might change,
571: since the matrices $\mathcal{M}_{\Delta}$, $\mathcal{M}_{B}$ and
572: $\mathcal{M}_{D}$ do not commute.
573: %
574:
575: Filling in the above expressions with random numbers selected from
576: suitably chosen ranges, we simulated all scattering processes
577: occurring in our experiments. Fig.~\ref{fig:4} shows a numerical
578: simulation of the scattered states in the tangle vs. linear
579: entropy plane, obtained with the singlet two-photon state as input
580: state. Fig. \ref{fig:4}~(a) corresponds to isotropic and
581: birefringent scatterers, and Fig.~\ref{fig:4}~(b) to dichroic
582: scatterers. The qualitative agreement between this model and the
583: experimental results shown in Fig.~\ref{fig:2} and
584: Fig.~\ref{fig:3} is manifest.
585:
586: %
587: \section{Conclusions}
588: %
589: In summary, we have presented experimental results on entanglement
590: properties of scattered photon-pairs for three varieties of
591: optical scattering systems. In this way we were able to generate
592: two distinct types of two-photon mixed states; namely Werner-like
593: and sub-Werner-like states. Moreover, we have introduced a simple
594: phenomenological model based onto the analogy between classical
595: polarization optics and quantum mechanics of qubits, that fully
596: reproduces our experimental findings. In the case of sub-Werner
597: states, the phenomenological model represents a non-trace
598: preserving quantum map. Such description might be considered
599: controversial since a non-trace preserving local map can in
600: principle lead to violation of causality when it describes the
601: evolution of a composite system made of two spatially separate
602: subsystems \cite{Aiello062}. However, we argue that our measured
603: states do not violate the signaling condition as they are
604: post-selected by the coincidence measurement, a procedure which
605: involves classical communication between the two detectors.
606: Finally, we expect it to be possible to create states \emph{above}
607: the Werner curve (in particular MEMS) \cite{mems2,mems3}, by
608: post-selective detection when acting on a single photon
609: \cite{Aiello062}. Work along this line
610: is in progress in our group.\\
611:
612: \begin{acknowledgments}
613: This project is part of the program of FOM and is also supported
614: by the EU under the IST-ATESIT contract.
615:
616: We gratefully acknowledge M. B. van der Mark for making available
617: the bundle of parallel fibers \cite{vdmark}.
618: \end{acknowledgments}
619:
620: \begin{thebibliography}{a16}
621: \expandafter\ifx\csname
622: natexlab\endcsname\relax\def\natexlab#1{#1}\fi
623: \expandafter\ifx\csname bibnamefont\endcsname\relax
624: \def\bibnamefont#1{#1}\fi
625: \expandafter\ifx\csname bibfnamefont\endcsname\relax
626: \def\bibfnamefont#1{#1}\fi
627: \expandafter\ifx\csname citenamefont\endcsname\relax
628: \def\citenamefont#1{#1}\fi
629: \expandafter\ifx\csname url\endcsname\relax
630: \def\url#1{\texttt{#1}}\fi
631: \expandafter\ifx\csname
632: urlprefix\endcsname\relax\def\urlprefix{URL }\fi
633: \providecommand{\bibinfo}[2]{#2}
634: \providecommand{\eprint}[2][]{\url{#2}}
635: %
636: \bibitem{rayleigh}Lord Rayleigh, Philos. Mag. \textbf{47}, 375 (1899); Proc. Roy. Soc. London,
637: Ser. A \textbf{79}, 399 (1907).
638: %
639: \bibitem{hulst}H. C. van de Hulst, \emph{Light Scattering by Small
640: Particles} (Dover, New York, 1981).
641: %
642: \bibitem{rossum}M. C. W. van Rossum and T. M. Nieuwenhuizen, Rev.
643: Mod. Phys. \textbf{71}, 313 (1999).
644: %
645: \bibitem{albada1} M. P. Van Albada, and A. Lagendijk, Phys. Rev.
646: Lett. \textbf{55}, 2692 (1985).
647: %
648: \bibitem{chabanov} A. A Chabanov, M. Stoytchev, and A. Z. Genack,
649: Nature (London) \textbf{404}, 850 (2000).
650: %
651: %\bibitem{abrahams}E. Abrahams, P. W. Anderson, D. C. Licciardello,
652: %and T. V. Ramakrishnana, Phys. Rev. Lett \textbf{42}, 673 (1979).
653: %
654: \bibitem{rikken} G. L. J. A. Rikken and B. A. van Tiggelen, Nature (London) \textbf{381}, 54 (1996).
655: %
656: %\bibitem{albada2}M. P. can Albada, J. F. de Boer, and A.
657: %Lagendijk, Phys. Rev. Lett. \textbf{64}, 2787 (1990).
658: %
659: \bibitem{freund}F. Scheffold and G. Maret, Phys. Rev. Lett. \textbf{81}, 5800 (1998).
660: %
661: %\bibitem{mac}F. C. MacKintosh, J. X. Zhu, D. J. Pine, and D. A. Weitz, Phys. Rev. B \textbf{40}, 9342 (1989).
662: %
663: \bibitem{lohdal} P. Lodahl, A. P. Mosk, and A. Lagendijk, Phys.
664: Rev. Lett \textbf{95}, 173901 (2005).
665: %
666: \bibitem{aiello1}A. Aiello and J. P. Woerdman, Phys. Rev. A \textbf{70}, 023808 (2004).
667: %
668: \bibitem{vanvelsen}J. L. van Velsen and C. W. J. Beenakker, Phys. Rev. A \textbf{70}, 032325 (2004).
669: %
670: \bibitem{sergienko}P. G. Kwiat, K. Mattle, H. Weinfurter, A. Zeilinger, A. V. Sergienko, and Y. Shih,
671: Phys. Rev. Lett. \textbf{75}, 4337 (1995).
672: %
673: %\bibitem{detectors}The "bucket" detector (SPCM-AQR-1X, PerkinElmer)
674: %consists of an avalanche photo diode (APD) with a circular
675: %effective area diameter $d=150~\mu$m, placed at the focal plane of
676: %a lens of $NA=0.33$. Thus, the maximum number of modes $N$
677: %detected by this configuration, for the given down-conversion
678: %wavelength $\lambda=826.2~$nm, can be estimated as
679: %$N=(\frac{d}{\lambda}NA)^2 \approx 3600$.
680: %
681: \bibitem{fidelity}R. Jozsa, J. Mod. Opt. \textbf{41}, 2315 (1994).
682: The fidelity between the target state $\rho^\mathrm{tar}$ and the
683: measured state $\rho^\mathrm{exp}$ was calculated as
684: $F(\rho^\mathrm{tar}, \rho^\mathrm{exp}) =|\tr\left(
685: \sqrt{\sqrt{\rho^\mathrm{tar}} \rho^\mathrm{exp}
686: \sqrt{\rho^\mathrm{tar}}}\right)|^2$.
687: %
688: \bibitem{james} D. F. V. James, P. G. Kwiat, W. J. Munro, and A. G. White,
689: Phys. Rev. A \textbf{64}, 052312 (2001).
690: %
691: %
692: \bibitem{puentes} G. Puentes, D. Voigt, A. Aiello, and J. P. Woerdman, Opt. Lett. \textbf{30}, 3216 (2005).
693: %
694: \bibitem{wooters} W. K. Wootters, Phys. Rev. Lett. \textbf{80}, 2245 (1998).
695: %
696: \bibitem{vedral} S. Bose, and V. Vedral, Phys. Rev. A \textbf{61}, 040101(R) (2000).
697: %
698: %\bibitem{buzek}M. Ziman and V. Bu\v{z}ek, Phys. Rev. A
699: %\textbf{72}, 052325 (2005).
700: %
701: %\bibitem{white} A. G. White, D. F. V. James, P. H. Eberhard, and
702: %P. G. Kwiat, Phys. Rev. Lett. \textbf{83}, 3103 (1999).
703: %
704: %\bibitem{peters} N. A. Peters, J. B. Altepeter, D. Branning, E. R.
705: %Jeffrey, T. C. Wei, and P. G. Kwiat, Phys. Rev. Lett. \textbf{92},
706: %133601 (2004).
707: %
708: \bibitem{mems} W. J. Munro, D. F. V. James, A. G. White, and P. G. Kwiat, Phys. Rev. A \textbf{64}, 030302(R)
709: (2001).
710: %
711: %
712: \bibitem{mems2} N. A. Peters, J. B. Altepeter, D. A.
713: Branning, E. R. Jeffrey, T.-C. Wei, and P. G. Kwiat, Phys. Rev.
714: Lett. \textbf{92}, 133601 (2004).
715: %
716: \bibitem{mems3}M. Barbieri, F. De Martini, G. Di Nepi, and P.
717: Mataloni, Phys. Rev. Lett. \textbf{92}, 177901 (2004).
718: %
719: \bibitem{werner} R. F. Werner, Phys. Rev. A \textbf{40}, 4277 (1989).
720: %
721: \bibitem{kliger}D. S. Kliger, J. W. Lewis, and C. E. Randall, \emph{Polarized Light in Optics and Spectroscopy}
722: (Academic Press, Inc., 1990).
723: %
724: \bibitem{puentes2} G. Puentes, D. Voigt, A. Aiello, and J. P. Woerdman, Opt.
725: Lett. \textbf{31}, 2057 (2006).
726: %
727: \bibitem{vdmark} M. B. van der Mark, PhD thesis, Univ. Amsterdam
728: (1990).
729: %
730: \bibitem{nielsen}M. A. Nielsen and I. L. Chuang, \emph{Quantum
731: Computation and Quantum Information} (Cambridge U. Press, 2000),
732: Chap. 12.
733: %
734: %\bibitem{aiello2} A. Aiello, and J. P. Woerdman, e-print math-ph/0412061, (2004).
735: %
736: \bibitem{sakurai} J. J. Sakurai, \emph{Modern Quantum Mechanics}
737: (Addison Wesley, 1994).
738: %
739: \bibitem[{\citenamefont{Lu and Chipman}(1996)}]{Lu96}
740: \bibinfo{author}{\bibfnamefont{S.-Y.} \bibnamefont{Lu}} \bibnamefont{and}
741: \bibinfo{author}{\bibfnamefont{R.~A.} \bibnamefont{Chipman}},
742: \bibinfo{journal}{J. Opt. Soc. Am. A} \textbf{\bibinfo{volume}{13}},
743: \bibinfo{pages}{1106} (\bibinfo{year}{1996}).
744: %
745: \bibitem{aiello3} A. Aiello, G. Puentes, and J. P. Woerdman, quant-ph/0611179, submitted to Phys. Rev. A (2006).
746: %
747: \bibitem[{Aie({\natexlab{b}})}]{Aiello062}
748: \bibinfo{note}{A. Aiello, G. Puentes, D. Voigt, and J. P. Woerdman,
749: quanth-ph/0603182 (2006).}
750: %
751: \end{thebibliography}
752: %
753: \end{document}
754: