Sie sind auf Seite 1von 25

Dear Author,

Here are the proofs of your article.

• You can submit your corrections online, via e-mail or by fax.

• For online submission please insert your corrections in the online correction form. Always
indicate the line number to which the correction refers.

• You can also insert your corrections in the proof PDF and email the annotated PDF.

• For fax submission, please ensure that your corrections are clearly legible. Use a fine black
pen and write the correction in the margin, not too close to the edge of the page.

• Remember to note the journal title, article number, and your name when sending your
response via e-mail or fax.

• Check the metadata sheet to make sure that the header information, especially author names
and the corresponding affiliations are correctly shown.

• Check the questions that may have arisen during copy editing and insert your answers/
corrections.

• Check that the text is complete and that all figures, tables and their legends are included. Also
check the accuracy of special characters, equations, and electronic supplementary material if
applicable. If necessary refer to the Edited manuscript.

• The publication of inaccurate data such as dosages and units can have serious consequences.
Please take particular care that all such details are correct.

• Please do not make changes that involve only matters of style. We have generally introduced
forms that follow the journal’s style.
Substantial changes in content, e.g., new results, corrected values, title and authorship are not
allowed without the approval of the responsible editor. In such a case, please contact the
Editorial Office and return his/her consent together with the proof.

• If we do not receive your corrections within 48 hours, we will send you a reminder.

• Your article will be published Online First approximately one week after receipt of your
corrected proofs. This is the official first publication citable with the DOI. Further changes
are, therefore, not possible.

• The printed version will follow in a forthcoming issue.

Please note
After online publication, subscribers (personal/institutional) to this journal will have access to the
complete article via the DOI using the URL: http://dx.doi.org/[DOI].
If you would like to know when your article has been published online, take advantage of our free
alert service. For registration and further information go to: http://www.link.springer.com.
Due to the electronic nature of the procedure, the manuscript and the original figures will only be
returned to you on special request. When you return your corrections, please inform us if you would
like to have these documents returned.
Metadata of the article that will be visualized in
OnlineFirst
ArticleTitle Estimation of entanglement in bipartite systems directly from tomograms
Article Sub-Title
Article CopyRight Springer Science+Business Media, LLC, part of Springer Nature
(This will be the copyright line in the final PDF)
Journal Name
Corresponding Author Family Name Sharmila
Particle
Given Name B.
Suffix
Division Department of Physics
Organization Indian Institute of Technology Madras
Address Chennai, 600036, India
Phone
Fax
Email sharmilab@physics.iitm.ac.in
URL
ORCID http://orcid.org/0000-0002-3947-3628

Author Family Name Lakshmibala


Particle
Given Name S.
Suffix
Division Department of Physics
Organization Indian Institute of Technology Madras
Address Chennai, 600036, India
Phone
Fax
Email slbala@physics.iitm.ac.in
URL
ORCID
Author Family Name Balakrishnan
Particle
Given Name V.
Suffix
Division Department of Physics
Organization Indian Institute of Technology Madras
Address Chennai, 600036, India
Phone
Fax
Email vbalki@physics.iitm.ac.in
URL
ORCID

Received 2 February 2019


Schedule Revised
Accepted 3 June 2019
Abstract We investigate the advantages of extracting the degree of entanglement in bipartite systems directly from
tomograms, as it is the latter that are readily obtained from experiments. This would provide a superior
alternative to the standard procedure of assessing the extent of entanglement between subsystems after
employing the machinery of state reconstruction from the tomogram. The latter is both cumbersome and
involves statistical methods, while a direct inference about entanglement from the tomogram circumvents
these limitations. In an earlier paper, we had identified a procedure to obtain a bipartite entanglement
indicator directly from tomograms. To assess the efficacy of this indicator, we now carry out a detailed
investigation using two nonlinear bipartite models by comparing this tomographic indicator with standard
markers of entanglement such as the subsystem linear entropy and the subsystem von Neumann entropy
and also with a commonly used indicator obtained from inverse participation ratios. The two-model
systems selected for this purpose are a multilevel atom interacting with a radiation field, and a double-well
Bose–Einstein condensate. The role played by the specific initial states of these two systems in the
performance of the tomographic indicator is also examined. Further, the efficiency of the tomographic
entanglement indicator during the dynamical evolution of the system is assessed from a time-series
analysis of the difference between this indicator and the subsystem von Neumann entropy.
Keywords (separated by '-') Entanglement indicator - Tomogram - Inverse participation ratios - Bipartite systems - Time-series analysis
Footnote Information
Quantum Information Processing _#####################_
https://doi.org/10.1007/s11128-019-2352-0

Estimation of entanglement in bipartite systems directly


Author Proof

of
from tomograms

B. Sharmila1 · S. Lakshmibala1 · V. Balakrishnan1

pro
Received: 2 February 2019 / Accepted: 3 June 2019
© Springer Science+Business Media, LLC, part of Springer Nature 2019

1 Abstract
2 We investigate the advantages of extracting the degree of entanglement in bipartite
3 systems directly from tomograms, as it is the latter that are readily obtained from exper-
4 iments. This would provide a superior alternative to the standard procedure of assessing
5 the extent of entanglement between subsystems after employing the machinery of state
6 reconstruction from the tomogram. The latter is both cumbersome and involves sta-
7
cted
tistical methods, while a direct inference about entanglement from the tomogram cir- 1
8 cumvents these limitations. In an earlier paper, we had identified a procedure to obtain
9 a bipartite entanglement indicator directly from tomograms. To assess the efficacy of
10 this indicator, we now carry out a detailed investigation using two nonlinear bipartite
11 models by comparing this tomographic indicator with standard markers of entangle-
12 ment such as the subsystem linear entropy and the subsystem von Neumann entropy
13 and also with a commonly used indicator obtained from inverse participation ratios.
14 The two-model systems selected for this purpose are a multilevel atom interacting with
15 a radiation field, and a double-well Bose–Einstein condensate. The role played by the
orre

16 specific initial states of these two systems in the performance of the tomographic indi-
17 cator is also examined. Further, the efficiency of the tomographic entanglement indica-
18 tor during the dynamical evolution of the system is assessed from a time-series analysis
19 of the difference between this indicator and the subsystem von Neumann entropy.

20 Keywords Entanglement indicator · Tomogram · Inverse participation ratios ·


21 Bipartite systems · Time-series analysis
unc

B B. Sharmila
sharmilab@physics.iitm.ac.in
S. Lakshmibala
slbala@physics.iitm.ac.in
V. Balakrishnan
vbalki@physics.iitm.ac.in

1 Department of Physics, Indian Institute of Technology Madras, Chennai 600036, India

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
_####_ Page 2 of 21 B. Sharmila et al.

22 1 Introduction

23 Quantum state reconstruction seeks to obtain the density matrix and the corresponding
Wigner function from tomograms. Experimental outcomes are in the form of tomo-
Author Proof

24

of
25 grams which are histograms of appropriate observables. However, even reconstruction
26 of qubit states, photon number states or unentangled states of bipartite systems from
27 relevant tomograms is typically a long and tedious task which employs statistical
28 tools that are inherently error-prone. The task becomes significantly more difficult
29 in the case of superpositions of states or entangled states of multipartite systems. It

pro
30 is therefore desirable, as far as possible, to extract information about the state (such
31 as indicators of its nonclassicality and entanglement) directly from the tomogram,
32 avoiding the reconstruction procedure. This has been demonstrated in bipartite qubit
33 systems by estimating state fidelity with respect to a specific target state directly from
34 the tomogram [1–5], and comparing the errors that arise with the corresponding errors
35 in procedures involving complete state reconstruction.
36 Obtaining information about a quantum state directly from tomograms for continu-
37 ous variable processes such as matter–radiation interactions qualifies to be a milestone
38 in state tomography. This is because reconstruction is far more cumbersome in the con-
39
cted
text of optical tomograms: the Hilbert space of the field is infinite-dimensional, and
40 homodyne measurements yield only limited information (namely, expectation values
41 of the density matrix in only a finite set of quadrature bases). Even in this case, how-
42 ever, a few results have been established. It has been shown that it is possible, solely
43 from tomograms, to assess qualitatively whether subsystems are entangled [6]. The
44 squeezing properties of the state in single and bipartite systems have been quantified
45 [7]. Further, specific nonclassical effects have been assessed from tomograms during
46 temporal evolution of single-mode and bipartite systems. For instance, qualitative sig-
47 natures of revivals (when a system returns to its initial state apart from an overall phase)
and fractional revivals (when the state of a system is a superposition of ‘copies’ of its
orre

48

49 initial state) have been identified in tomograms in the case of a single-mode radiation
50 field propagating in a Kerr medium [8]. Decoherence of entangled bipartite states has
51 also been investigated using tomograms [7]. Quadrature and higher-order squeezing
52 of a radiation field subject to cubic nonlinearities have been quantified, and an entan-
53 glement indicator which can be inferred from the tomogram for bipartite systems such
54 as a double-well Bose–Einstein condensate(BEC) has been proposed [9]. It is there-
55 fore important to appreciate that in multipartite systems, even if the entanglement
56 indicators obtained directly from tomograms do not qualify to be monotones, they
57 will substantially facilitate characterization of the properties of the state of the system,
unc

58 provided they capture qualitatively the salient features of entanglement dynamics.


59 The last of the foregoing is a significant step in exploiting tomograms, as entangle-
60 ment is an essential resource in quantum information processing. In this context, it is
61 important to consider interesting phenomena such as sudden death and birth of entan-
62 glement [10], and its collapse to a constant nonzero value over a significant interval of
63 time [11], as have been found in model systems. Since we are directly concerned with
64 quantum entanglement indicators here, we recall some standard measures of entan-
65 glement between the two subsystems A and B of a bipartite system. Two important
66 measures of entanglement are the subsystem von Neumann entropy (SVNE), given by

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Estimation of entanglement in bipartite systems directly… Page 3 of 21 _####_

67 −Tr (ρi log ρi ) where ρi (i = A, B) is the reduced density matrix of the subsystem,
68 and the subsystem linear entropy (SLE), given by 1 − Tr (ρi2 ). It is evident that the
69 SVNE and the SLE involve both off-diagonal and diagonal elements of the density
matrix in any given basis. In contrast, the tomogram only provides information about
Author Proof

70

of
71 the diagonal elements, although in several complete bases. Even though the tomo-
72 graphic indicator mentioned earlier mimics quite closely the qualitative features of
73 the SVNE and SLE [9], it does not qualify to be a measure. It is therefore necessary
74 to carry out a detailed comparative study between the tomographic indicator, on the
75 one hand, and the SLE and SVNE, on the other hand, in order to assess its efficacy

pro
76 and limitations vis-à-vis standard entanglement measures.
77 Another interesting entanglement measure with which the tomographic indicator
78 is to be compared is the inverse participation ratio (IPR). In a bipartite system, this
79 ratio is a measure of the spread of the wave function of the system over subsystem
80 basis states. The ratio itself is defined in terms of the fourth power of the system wave
81 function. A procedure to extract the extent of entanglement in spin systems from the
82 inverse participation ratio has been outlined in [12], and a measure of entanglement
83 has been defined. Extensive examination of the efficacy of this measure in interacting
84 spin systems has been carried out (see, for instance, [13–21]). The limitations of this
85
cted
measure in capturing the salient features of the SVNE in entangled qubit systems have
86 been pointed out in [15]. We identify the analog ξipr of this measure in the case of a
87 continuous variable system, and assess the performance of the tomographic indicator
88 relative to that of ξipr .
89 The models we consider for our purposes describe two experimentally viable bipar-
90 tite systems, namely, the double-well BEC with nonlinear interactions between the
91 condensate atoms [22], and a multilevel nonlinear atomic medium interacting with a
92 radiation field [23]. Extensive literature on entanglement dynamics exists in the case
93 of these systems. For instance, in a binary Bose–Einstein condensate, the dynamics
of the variance of appropriate observables was found to mimic entanglement dynam-
orre

94

95 ics [24]. Further, the experimental realization of atomic homodyne measurements


96 [25] has enabled quantum state tomography in BECs, while optical homodyne mea-
97 surements are an integral part of field tomography [26]. We have also obtained a
98 long data set of the difference between the SVNE and the tomographic indicator,
99 and carried out a time-series analysis of this difference for several initial states and
100 for different amounts of nonlinearity. The ergodicity properties of this difference
101 may be expected to shed light on the dynamical behavior of the tomographic indi-
102 cator.
103 The plan of the rest of this paper is as follows. In Sect. 2 , we describe how the
unc

104 tomographic indicator is obtained, and relate it to the participation ratio. In Sect. 3,
105 we introduce the two bipartite models mentioned above and compare various indi-
106 cators during dynamical evolution. Section 4 is devoted to the time-series analysis
107 referred to above. State reconstruction from the tomogram for a zero-photon state
108 has been demonstrated in “Appendix 1”. This helps appreciate the importance of
109 extracting nonclassical effects directly from tomograms. The crucial steps in the pro-
110 cedure for obtaining the time-evolved state of the double-well BEC are outlined in
111 “Appendix 2”.

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
_####_ Page 4 of 21 B. Sharmila et al.

112 2 Bipartite entanglement indicator from the tomogram

113 We begin with a quick summary of the salient features of a tomogram, followed by
a description of the procedure for extracting an indicator of bipartite entanglement
Author Proof

114

of
115 solely from the tomogram.

116 2.1 Salient features of a tomogram

117 The starting point is a quorum of observables (which can, in principle, be measured

pro
118 through appropriate experiments) whose statistics gives us tomographically complete
119 information about the state. In the case of a single-mode radiation field, for instance,
120 a quorum is constituted by the set of rotated quadrature operators [27,28]

 √
Xθ = a † eiθ + a e−iθ / 2

121 (1)

122 where 0 ≤ θ < π , and a and a † are photon annihilation and creation operators
satisfying [a, a † ] = 1. The expectation value of the density matrix ρ can be computed
123

124
cted
in each complete basis set {|X θ , θ } for a given θ . The tomogram [27,29] w(X θ , θ ) =
125 X θ , θ |ρ|X θ , θ  is usually represented as a three-dimensional plot of w versus X θ (on
126 the x-axis) and θ (on the y-axis).
127 Of immediate relevance to us is the optical tomogram [28] corresponding to a
128 bipartite system with subsystems A and B with rotated quadrature operators

 √
Xθ A = a e−iθ A + a † eiθ A / 2

129 (2)
orre

130 and  √
Xθ B = be−iθ B + b† eiθ B / 2.

131 (3)

132 Here a, a † (resp., b, b† ) are the annihilation and creation operators for subsystem A
133 (resp., B), and 0 ≤ θ A , θ B < π . The bipartite tomogram is given by

134 w(X θ A , θ A ; X θ B , θ B )
135
136
= X θ A , θ A ; X θ B , θ B |ρ AB |X θ A , θ A ; X θ B , θ B , (4)
unc

137 where ρ AB denotes the bipartite density matrix. This is just a straightforward extension
138 of the single-mode tomogram. Here, Xθi |X θi , θi  = X θi |X θi , θi  (i = A, B), and
139 |X θ A , θ A ; X θ B , θ B  stands for |X θ A , θ A  ⊗ |X θ B , θ B . The normalization condition is

 ∞  ∞
140 dX θ A dX θ B w(X θ A , θ A ; X θ B , θ B ) = 1 (5)
−∞ −∞

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Estimation of entanglement in bipartite systems directly… Page 5 of 21 _####_

141 for every value of θ A and θ B . The reduced tomogram for subsystem A is
 ∞
142 w A (X θ A , θ A ) = dX θ B w(X θ A , θ A ; X θ B , θ B )
Author Proof

−∞

of
143
144
= X θ A , θ A |ρ A |X θ A , θ A , (6)

145 where ρ A = TrB (ρ AB ) is the reduced density matrix of the subsystem A. A similar
146 definition holds for the subsystem B.
147 The bipartite tomographic entropy and the subsystem tomographic entropy are

pro
148 important concepts that we require for our tomographic entanglement indicator. The
149 former is given by
 ∞  ∞
150 S(θ A , θ B ) = − dX θ A dX θ B w(X θ A , θ A ; X θ B , θ B )
−∞ −∞
151
152
× log [w(X θ A , θ A ; X θ B , θ B )]. (7)

153 The subsystem tomographic entropy is


cted  ∞
154 S(θi ) = − dX θi wi (X θi , θi )
−∞
155
156
× log [wi (X θi , θi )] (i = A, B). (8)

157 As shown in [9], the mutual information is given by

158 S(θ A : θ B ) = S(θ A , θ B ) − S(θ A ) − S(θ B ). (9)

159 The tomographic entanglement indicator, given by


orre

160 ξtei = S(θ A : θ B ), (10)

161 is obtained by averaging the mutual information over an ideally very large number
162 of values of θ A and θ B over the interval [0, π ). In practice (as shown explicitly [9]
163 in the case of the double-well BEC system), however, even as few as 25 values of
164 S(θ A : θ B ) for θ A and θ B chosen at equally-spaced intervals in the range [0, π ) suffice
165 to yield a ξtei that compares well with standard entanglement measures such as the
166 SVNE. For unentangled bipartite pure states, S(θ A : θ B ) vanishes as a consequence
unc

167 of the factorization property w(X θ A , θ A ; X θ B , θ B ) = w A (X θ A , θ A )w B (X θ B , θ B ) for


168 every θ A , θ B .

169 2.2 Participation ratio and entanglement indicator

170 The generalized eigenstates of conjugate pairs of quadrature operators constitute a


171 pair of mutually unbiased bases [30], as


 √
X θ , θ |X θ+π/2 , θ + π/2 = 1/ 2π  > 0. (11)
 
172

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
_####_ Page 6 of 21 B. Sharmila et al.

173 The specific averaging procedure mentioned above obviously involves calculating
174 S(θ A : θ B ) in several sets of mutually unbiased bases. Parallels can be drawn between
175 ξipr and ξtei , as a similar averaging procedure (in this case, over inverse participation
ratios) is followed in calculating ξipr [12]. This is seen by writing the inverse par-
Author Proof

176

of
177 ticipation ratio η AB corresponding to a given bipartite pure state |ψ AB  in terms of
178 basis states of relevance to the problem at hand, namely, the rotated quadrature basis
179 {|X θ A , θ A ; X θ B , θ B }, for specific values of θ A and θ B . We have
 ∞  ∞  4
η AB = dX θ A dX θ B X θ A , θ A ; X θ B , θ B |ψ AB  . (12)

pro
180
−∞ −∞

181 For ease of notation, the dependence of η AB on θ A and θ B has been omitted in the
182 left-hand side of the forgoing expression. η AB is readily expressed in terms of the
183 tomogram as
 ∞  ∞
184 η AB = dX θ A dX θ B [w(X θ A , θ A ; X θ B , θ B )]2 . (13)
−∞ −∞

185
cted
The inverse participation ratio for each subsystem is given by
 ∞
186 ηi = dX θi [wi (X θi , θi )]2 (i = A, B). (14)
−∞

187 Note that ηi depends on θi . The numerical results reported in the next section seem to
188 suggest that 1 − (η A + η B − η AB ) which we denote by ξipr captures the qualitative
189 features of SVNE during temporal evolution. The symbol  represents average over
190 θ A and θ B . While in principle, all θ A and θ B are to be used in the averaging procedure,
in practice a sufficiently large number of θ A and θ B have been numerically found to
orre

191

192 be adequate. The choice of the notation ξipr is inspired merely by the fact that it is
193 obtained from inverse participation ratios (IPRs), though it is itself not an IPR.
194 It is appropriate to compare ξtei and ξipr , because both these quantities can be
195 obtained directly from the tomogram, and the similarity in form of the two relations
196 (1 − ξipr ) = η A + η B − η AB  and ξtei = S(θ A , θ B ) − S(θ A ) − S(θ B ) is manifest.
197 (Here ·· denotes the relevant average in each case.) In obtaining ξtei , it is evident that
198 (in practice) a good approximation would be to average only over the dominant values
199 of S(θ A : θ B ). We have compared the results obtained numerically by averaging over
200 the complete set of values of S(θ A : θ B ) and those obtained by averaging only over
unc

201 the dominant values (values that exceed the mean value by one standard deviation).
202 For this purpose, we have numerically generated several histograms of S(θ A : θ B ) for
203 both the atom–field interaction and the double-well BEC models. For a given initial
204 state, the system is unitarily evolved in time, and each histogram corresponds to a
205 specific instant of time. A typical histogram is presented in Fig. 1. The average of
206 S(θ A : θ B ) over all the values in the histogram in Fig. 1 has been compared with
207 that obtained by averaging only over the contribution from the shaded portion. The
208 qualitative features were found to be essentially the same in both cases in all the
209 histograms considered. The results in the following sections are therefore based on

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Estimation of entanglement in bipartite systems directly… Page 7 of 21 _####_

Fig. 1 A histogram of 8
S(θ A : θ B ) from 100 7
combinations of θ A and θ B . The 6 σ
shaded area marks the values of

Frequency
5
S(θ A : θ B ) that exceed the mean
Author Proof

4
value ∼ 0.07 by one standard

of
deviation 3
2
1
0
0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18
S(θA:θB)

pro
210 the latter averaging procedure, since it is clearly computationally less intensive. We
211 denote the entanglement indicator thus obtained by ξtei′ , and in subsequent sections

212

compare ξtei with the SLE and ξipr .

213 3 Entanglement indicators in generic bipartite models

214
cted ′ in two generic bipartite models describing, respec-
We investigate the properties of ξtei
215 tively, a multilevel atom interacting with a radiation field [23] and the double-well BEC
216 [22]. Both systems are inherently nonlinear and provide an ideal platform to exam-
217 ine nonclassical effects as the corresponding state evolves in time. In this paper, we
218 consider initial states of the total bipartite system that are pure states subsequently
219 governed by unitary evolution. This ensures that the SVNE (and similarly, the SLE)
220 remain independent of the subsystem label (A or B) for all t.

221 3.1 Atom–field interaction model


orre

222 We consider a nonlinear multilevel atomic medium coupled with strength g to a radi-
223 ation field of frequency ω F . The effective Hamiltonian (setting  = 1) is [23]

224 HAF = ω F a † a + ω A b† b + γ b† 2 b2 + g(a † b + ab† ). (15)

225 a, a † are photon annihilation and creation operators. The multilevel atom is modeled
226 as an oscillator with harmonic frequency ω A and ladder operators b, b† . The nonlinear
atomic medium is effectively described by the Kerr-like term in HAF with strength
unc

227

228 γ . A variety of initial states of the total system has been judiciously selected in order
229 to explore the range of possible nonclassical effects during time evolution. The unen-
230 tangled initial states considered correspond to the atom in its ground state |0 and the
231 field in either a coherent state (CS), or an m-photon added coherent state (m-PACS).
232 In the photon number basis, the CS |α (α ∈ C) is of course given by


2 /2  αk
233 |α = e−|α| √ |k. (16)
k=0
k!

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
_####_ Page 8 of 21 B. Sharmila et al.

234 The normalized m-PACS |α, m (where m is a positive integer), which possesses a
235 precisely quantified departure from perfect coherence, is given by

a † m |α a † m |α
Author Proof

236 |α, m = = , (17)

of
α|a m a † m |α

m! L m (−|α|2 )

237 where L m is the Laguerre polynomial of order m. We also consider two entangled
238 initial states, namely, the binomial state |ψbin  and the two-mode squeezed state |ζ .
239 These states are defined as follows. The total number operator

pro
240 Ntot = a † a + b† b (18)

241 commutes with HAF , and |ψbin  is an eigenstate of this operator with eigenvalue N (a
242 nonnegative integer). In explicit form, it is given by

N
  N 1/2
243 |ψbin  = 2−N /2 n |N − n; n, (19)
n=0
cted
244 where |N − n; n ≡ |N − n ⊗ |n, the product state corresponding to the field and the
245 atom in the respective number states |N − n and |n. The two-mode squeezed state
246 is given by
∗ † †
247 |ζ  = eζ ab−ζ a b |0; 0, (20)
248 where ζ ∈ C and |0; 0 is the product state corresponding to N = 0, n = 0.
249 Corresponding to these initial states, we have numerically generated tomograms at
250 approximately 2000 instants, separated by a time step 0.2 π/g. From these, we have
251
′ and the differences
obtained ξtei
orre

′ ′
252 d1 (t) = |SVNE − ξtei |, d2 (t) = |SLE − ξtei | (21)

253 as the system evolves. These differences are plotted against the scaled time gt/π in
254 Fig. 2a for an initial two-mode squeezed state, and in Fig. 2b for a factored product of
255 a CS and atomic ground state |0. From these plots, it is evident that ξtei′ is in much

256 better agreement with the SLE than with the SVNE over the time interval considered,
2 257 independent of the parameter values and the nature of the initial state. We therefore
258 choose the SLE as the reference entanglement indicator. Next, we compare d2 (t) with
unc

259 the difference


260 (t) = |SVNE − SLE|. (22)
261 We have verified that, In all the cases considered, (t) > d2 (t) (see, for instance, Fig.
262 3). In what follows, we therefore focus only on d2 (t) and the difference

263 d3 (t) = |SLE − ξipr |. (23)

264 This comparison brings out interesting features of both the indicators. When the
265 strength of the nonlinearity is low relative to that of the coupling (e.g., γ /g = 0.01),

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Estimation of entanglement in bipartite systems directly… Page 9 of 21 _####_

(a) (b)

Entropy difference

Entropy difference
0.08 0.8

0.06 0.6

0.04 0.4
Author Proof

of
0.02 0.2

0 0
0 10 20 30 0 100 200 300
gt/π gt/π

Fig. 2 d1 (t) (black dashed curve) and d2 (t) (pink curve) versus scaled time gt/π with ω F = 1, ω A =

pro
1, γ = 1. a g = 0.2, initial squeezed state |ζ , ζ = 0.1 b g = 100, initial state |α ⊗ |0, |α|2 = 1

Fig. 3 d2 (t) (blue dotted curve) 0.6

Entropy difference
and (t) (pink curve) versus
time, with ω F = 1, ω A = 1, 0.4
γ = 1, g = 100, initial state
|α ⊗ |0, |α|2 = 1
0.2

0
0 100 200 300
gt/π
cted
266 it is known [31] that full and fractional revivals occur, and entanglement measures
267 may be expected to display signatures of these revival phenomena. Figure 4a shows
268 that, at the revival time gTrev /π = 400, ξtei ′ agrees with the SLE much more closely

269 than ξipr does. Further, over the entire time interval (0, Trev ), d2 (t) is significantly
270 smaller than d3 (t). This feature holds even for larger values of the ratio γ /g, as shown
′ is therefore to be favored over ξ
in Fig. 4b. ξtei
271 ipr as an entanglement indicator. The
272 time evolution of the difference d2 (t) is drastically different from that of d3 (t) for
initial field states that depart from ideal coherence. In this case, over the entire time
orre

273

274 considered, ξipr performs significantly better than ξtei ′ for small values of γ /g, as in

275 Fig. 4c. As the value of γ /g is increased, the two indicators have essentially the same
276 behavior, as shown in Fig. 4d.
277 We turn now to entangled initial states. In the case of the two-mode squeezed state
|ζ , we see from Fig. 5a, b that ξtei ′ fares much better than ξ
278 ipr over the entire time
279 interval considered, for small values of ζ . With an increase in the value of ζ , both the
280 indicators show comparable departures from SLE.
281 In the case of an initial binomial state, on the other hand, ξipr fares significantly
282
′ . This can be understood by examining the Hamming distance between
better than ξtei
unc

283 the basis states constituting the binomial state. We define this distance in the context of
284 continuous variables by extrapolating the idea of Hamming distance for qudits given
285 below. The Hamming distance between two qubits |u and |v (u, v = 0, 1) is 1 if u =
286 v, and 0 if u = v. It is evident that this can be extended to qudits in a straightforward
287 fashion. The generalized Hamming distance (see, for instance, [12]), i.e., the distance
288 between two bipartite qudit states |u 1 ; u 2  (where u 1 , u 2 = 0, 1, . . . , d − 1) and
289 |v1 ; v2  (where v1 , v2 = 0, 1, . . . , d − 1), is 0 if u 1 = v1 and u 2 = v2 . The Hamming
290 distance is 1 if u 1 = v1 and u 2 = v2 , or vice versa. The distance is 2 (the maximum
291 possible value in bipartite systems) if u 1 = v1 and u 2 = v2 . Further, the efficacy of

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
_####_ Page 10 of 21 B. Sharmila et al.

(a) (b)

Entropy difference
0.4 0.4

Entropy difference
0.2 0.2
Author Proof

of
0 0
0 100 200 300 0 100 200 300
gt/π g t/ π
(c) (d)
Entropy difference

Entropy difference
0.4 0.4

pro
0.2 0.2

0 0
0 100 200 300 0 100 200 300
g t/ π g t/ π

Fig. 4 d2 (t) (blue dotted curve) and d3 (t) (brown curve) versus time, with ω F = 1, ω A = 1, γ = 1. a
g = 100, initial state |α ⊗ |0. b g = 0.2, initial state |α ⊗ |0. c g = 100, initial state |α, 5 ⊗ |0. d
g = 0.2, initial state |α, 5 ⊗ |0. |α|2 = 1 in all cases
cted
(a) (b)
Entropy difference
Entropy difference

0.4 0.4

0.2 0.2

0
0
0 10 20 30 0 10 20 30 40
gt/π gt/π
orre

Fig. 5 d2 (t) (blue dotted curve) and d3 (t) (brown curve) versus time, with ω F = 1, ω A = 1, γ = 1, g =
0.2, for the initial two-mode squeezed state |ζ . a ζ = 0.1. b ζ = 0.7

292 ξipr as an entanglement indicator increases with an increase in the Hamming distance.
293 This indicator is therefore especially useful for states which are Hamming-uncorrelated
294 (i.e., separated by a Hamming distance equal to 2) [12,13].
295 In the case of interest to us, both the subsystems are infinite-dimensional. We would
296 like to examine whether the efficacy of ξipr is correlated with the Hamming distance
297 in this case as well. For this purpose, we extend the notion of the Hamming dis-
unc

298 tance between two unentangled basis states in a straightforward manner: The distance
299 between |m; n and | p; q (where m, n, p, q = 0, 1, 2, . . . ad inf.) is equal to its max-
300 imum value of 2 if m = p and n = q; 1 if m = p, n = q or vice versa; and 0 if
301 m = p, n = q. Note that the binomial state |ψbin  can be expanded as a superposition
302 of states which are Hamming-uncorrelated [Eq. (19)]. Figure 6 shows that in this case,
303 too, ξipr is a significantly better entanglement indicator than ξtei′ .

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Estimation of entanglement in bipartite systems directly… Page 11 of 21 _####_

Fig. 6 d2 (t) (blue dotted curve) 0.6

Entropy difference
and d3 (t) (brown curve) versus
time, with ω F = 1, ω A = 0.4
1, γ = 1, g = 0.2. Initial state
|ψbin  with N = 10
Author Proof

0.2

of
0
0 100 200 300
g t/ π

pro
304 3.2 The double-well BEC model

305 The effective Hamiltonian for the system (setting  = 1) is given by [22]

306 HBEC = ω0 Ntot + ω1 (a † a − b† b) + U Ntot


2
− λ(a † b + ab† ). (24)

307 Ntot = a † a + b† b as before, but (a, a † ) and (b, b† ) are now the boson annihilation
308 and creation operators of the atoms in wells A and B respectively. U is the strength
309
cted
of the nonlinearity (both in the individual modes as well as in their interaction), λ is
310 the linear interaction strength, and ω0 , ω1 are constants. As in the previous instance,
311 we select a representative variety of initial states: (i) the unentangled direct product
312 |αa , m 1  ⊗ |αb , m 2  of boson-added coherent states of atoms in the wells A and B,
313 respectively, where αa , αb ∈ C; (ii) the binomial state |ψbin  [Eq. (19)], and (iii) the
314 two-mode squeezed vacuum state |ζ  [Eq. (20)], with the understanding that the basis
315 states are now product states of the species in the two wells.
316 In each of these cases, we must first obtain the state of the system at any time t ≥ 0
317 as it evolves under the Hamiltonian HBEC . It turns out that, in the case of an initial
state of type (i) above, the state of the system can be calculated explicitly as a function
orre

318

319 of t, as outlined in “Appendix 2”. In the cases (ii) and (iii), the state vector at time
320 t is computed numerically. Using the results obtained for the state of the system at
321 time t, we have generated, for each of the initial states listed above, tomograms at
322 approximately 1000 instants, separated by a time step 0.001 π/U . We have verified
323
′ agrees better with the SLE
that, as in the case of the atom–field interaction model, ξtei
324

than with the SVNE, and that the difference between ξtei and the SLE is smaller than
325 that between the SVNE and the SLE. In what follows, therefore, we have chosen the
326 SLE as the reference entanglement measure and compared d2 (t) with d3 (t).
327 The effective frequency parameter for the linear part of the Hamiltonian HBEC is
unc

328 given (see “Appendix 2”) by λ1 = (ω12 + λ2 )1/2 , while the strength of the nonlinearity
329 is parametrized by U . Hence, the relevant ratio for the characterization of the dynamics
330 is U /λ1 . Figure 7 depicts a representative example of the temporal behavior of d2 (t)
331 and d3 (t). The effect of increasing U /λ1 can be seen by comparing Fig. 7a, b, while
332 that of a departure from coherence of the initial state can be seen by comparing (a) and
333 (c). We have also carried out analogous studies in the case of entangled initial states
334 |ψbin  and |ζ . These results are not presented here owing to limitations of space, but
335 the general trends in the behavior of the entanglement indicators in these cases are
336 consistent with, and corroborate, those found in the atom–field interaction model.

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
_####_ Page 12 of 21 B. Sharmila et al.

(a) (b) (c)


Entropy difference

Entropy difference

Entropy difference
0.6 0.4 0.4

0.4
0.2 0.2
Author Proof

0.2

of
0 0 0
0 0.25 0.5 0.75 1 0 0.25 0.5 0.75 0 0.25 0.5 0.75 1
t/π t/π t/π

Fig. 7 d2 (t) (blue dotted curve) and d3 (t) (brown curve) versus time, with ω0 = 1, U = 1. a ω1 = 1, λ = 1,
initial state |α ⊗ |α, b ω1 = 0.1, λ = 0.1, initial state |α ⊗ |α, c ω1 = 1, λ = 1, initial state |α, 1 ⊗ |α.

pro
In all cases, |α|2 = 1

337 3.2.1 Decoherence effects

338 We now proceed to obtain the entanglement indicator from the tomogram in the pres-
339 ence of decoherence in the system. The latter is the most important issue from the
340 experimental point of view, and therefore needs to be addressed. For the purpose of
341 illustration, we
√ take the initial state of the double-well BEC to be of the form |αa ⊗|αb ,
342 where αa = 0.001 and αb = 1: since αa is sufficiently small in magnitude, |αa  can
343
cted
be approximated by a superposition of a zero-boson and one-boson states, whereas
344 |αb  is an infinite superposition of the boson number states. The states get entangled
345 during unitary evolution. We consider the entangled state of the system at any instant,
346 say U t = π/2, and apply damping to subsystem A alone for a time interval τ . This
347 toy example suffices to examine to what extent an efficient entanglement indicator can
348 be extracted from the tomogram, when the system decoheres. We have set U , ω0 , ω1
349 and λ equal to 1, for numerical computation. Decoherence takes place via amplitude
350 damping and phase damping [32].
351 The master equation for amplitude decay is given by
orre


352 = −Ŵ(a † aρ − 2aρa † + ρa † a). (25)

353 Here, Ŵ is the rate of loss and τ the time parameter is reckoned from the instant
354 U t = π/2. The solution to this master equation is [32]


355 ρ(τ ) = ρn,m,n ′ ,m ′ (τ )|n; mn ′ ; m|, (26)
n,m,n ′ ,m ′ =0
unc

356 where


n + r n′ + r

−Ŵτ (n+n ′ )

357 ρn,m,n ′ ,m ′ (τ ) = e
r r
r =0
−2Ŵτ r
358 (1 − e ) ρn+r ,m,n ′ +r ,m ′ (τ = 0). (27)

359 Note that ρ(τ = 0) is chosen to be |ψ00 (π/2U )ψ00 (π/2U )|. (The expression for
360 |ψ00 (t) as a function of time t is given in “Appendix 2”).

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Estimation of entanglement in bipartite systems directly… Page 13 of 21 _####_

′ , SVNE and ξ
Fig. 8 ξtei a qmi 1.6
versus τ for Ŵ = 0.001
1.2

Entropy
0.8
0.4
Author Proof

of
0

0π 250π 500π 750π 1000π


Ut

pro
′ , SVNE and ξ
Fig. 9 ξtei 1.6
a qmi
versus τ for Ŵ p = 0.001 1.2

Entropy
0.8
0.4
0

0π 250π 500π 750π 1000π


cted Ut

361 As the bipartite state is a mixed state owing to decoherence, the SVNE correspond-
362 ing to subsystems A and B are not equal to each other and are now denoted by SVNEa
363 and SVNEb , respectively. The entropy of the full system is denoted by SVNEab . The
364 quantum mutual information ξqmi = SVNEa + SVNEb − SVNEab is of immediate
′ , SVNE and ξ
interest. At each instant τ , ξtei
365 a qmi have been calculated. Since the sub-
366 system entropies are not equal, we expect that (if the tomogram captures decoherence
′ must match ξ
effects well) ξtei
367 qmi during dynamics. This is shown in Fig. 8, where
368

ξtei , ξqmi and the SVNE of interest, namely, SVNEa are compared. We point out that
369 in the absence of decoherence for a bipartite pure state, as examined earlier, ξqmi = 2
orre

370 SVNEa = 2 SVNEb .


371 Phase damping in A is modeled by [32]


372 = −Ŵ p ((a † a)2 ρ − 2a † aρa † a + ρ(a † a)2 ), (28)

373 where Ŵ p is the rate of decoherence. The solution is of the form given in Eq. (26),
374 with
unc

′ 2
375 ρn,m,n ′ ,m ′ (τ ) = e−Ŵ p τ (n−n ) ρn,m,n ′ ,m ′ (τ = 0).

′ mimics ξ
As before, ξtei
376 qmi reasonably well (Fig. 9).

377 4 Time-series analysis of d1 (t)

378 Finally, we turn to an assessment of the long-time behavior of the tomographic entan-
379 glement indicator, by means of a detailed time–series analysis.

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
_####_ Page 14 of 21 B. Sharmila et al.

380 As we have shown in the foregoing, the deviation of ξtei ′ from the SVNE is much

381 more pronounced than its deviation from the SLE. It is therefore appropriate to investi-
382
′ and the SVNE changes with time. With this
gate how an initial difference between ξtei
in mind, a time series of d1 (t) has been obtained for each of the models at hand, and
Author Proof

383

of
384 used to compute local Lyapunov exponents along the lines customary [33–35] in the
385 study of dynamical systems. This involves reconstruction of the effective phase space,
386 estimation of the minimum embedding dimension demb of this space, and calculation
387 of the exponents themselves. The procedure used is outlined below.
388 The time series had 20000 data points. The effective phase space was reconstructed

pro
389 using the TISEAN package [36]. 100 different initial values d1 (0) were randomly
390 chosen in this phase space. The maximum local Lyapunov exponent corresponding to
391 each d1 (0) was computed over the same time interval L. (The term ‘local’ refers to the
392 fact that L is much smaller than the time interval over which the maximum Lyapunov
393 exponent ∞ is obtained in the standard method). The average value  L of these 100
394 maximum local Lyapunov exponents was obtained following the prescription in [35].
395 This procedure was repeated for as many as 14 different values of L. Further, in each
396 case, it was verified that, with an increase in L,  L tends to ∞ + (m/L q ), where
397 m and q are constants [35]. Note that two neighboring initial values of the dynamical
398
cted
variable of interest [in our case d1 (t)] diverges exponentially with the exponent  L
399 in L steps. We also present the power spectra of the time series for completeness. We
400 would expect a broadband spectra for chaotic data. On the other hand, a ‘spiky’ power
401 spectrum (i.e., a cluster of clearly defined sharp peaks in the power spectrum) points
402 to a possible quasi-periodic behavior. The results are presented below.

403 4.1 Atom–field interaction model

404 The difference d1 (t) has been obtained at each instant with time step δt = 0.1 for
orre

405 20000 time steps, and the effective phase space has been reconstructed. We see that
406 for both the initial states |α⊗|0 and |α, 5⊗|0 with |α 2 | = 1 and weak nonlinearity
407 (γ /g = 0.01),  L is positive, and both  L and ∞ are larger for the second initial
408 state (compare Fig. 10a, d). ∞ increases with an increase in |α|2 for the initial state
409 |α ⊗ |0 (compare Fig. 10a, b). In contrast, for strong nonlinearity (e.g., as in Fig.
410 10c, γ /g = 5),  L is negative.
411 For completeness, we present the power spectrum S( f ) of the time series as a
412 function of the frequency f in units of g, for each of the cases corresponding to
413 Fig. 10a–d. The nearly quasi-harmonic power spectrum for weak nonlinearity (Fig.
10a) changes into a broadband spectrum with increasing |α|2 (Fig. 10b), while it
unc

414

415 loses its quasi-harmonicity without becoming a broadband spectrum with increasing
416 nonlinearity (Fig. 10c). The lack of coherence in the initial state makes the power
417 spectrum broadbanded (Fig. 10d). When a power spectrum exhibits regular or quasi-
418 regular spikes at different frequencies, it signals quasi-periodicity in the processes
419 generating the time-series. A broadband spectrum, on the other hand, is an indicator
420 of non-periodic (including chaotic) time evolution. The time-series under discussion is
421
′ and the SVNE. When the power
that of the difference d1 (t) between our indicator ξtei
422 spectrum of this time–series shows quasi-periodic behavior (Fig. 11a, c) in the sense

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Estimation of entanglement in bipartite systems directly… Page 15 of 21 _####_

(a) (b)
0.6 2.2
Λ∞=0.032, m=3.4, q=1.2 Λ∞=1.2, m=2.8, q=0.77
2
0.4
1.8
ΛL

ΛL
1.6
Author Proof

0.2

of
1.4
0 1.2
0 125 250 375 500 0 125 250 375 500
L L
(c) (d)
-0.03 1.4
Λ∞=-0.059, m=0.78, q=1.2 Λ∞=1.0, m=4.6, q=0.96
-0.04 1.3

pro
ΛL

ΛL
-0.05 1.2

-0.06 1.1

-0.07 1
125 250 375 500 125 250 375 500
L L

Fig. 10  L obtained from the time series of d1 (t) (blue crosses) and the fit ∞ + (m/L q ) (red curve)
versus L, with ω F , ω A and γ equal to 1. Initial state |α ⊗ |0: a g = 100, |α|2 = 1. b g = 100, |α|2 = 5.
c g = 0.2, |α|2 = 1. d Initial state |α, 5 ⊗ |0, g = 100, |α|2 = 1
cted
(a) (b)
1 100
0.01 1
log S(f)

0.0001 0.01
log S(f)

1e-06 0.0001
1e-08 1e-06
1e-10 1e-08
1e-12 1e-10
1e-14 1e-12
0 0.001 0.002 0.003 0.004 0.005 0 0.001 0.002 0.003 0.004 0.005
f/g f/g
(c) (d)
100
1
orre

0.01 1
log S(f)
log S(f)

0.0001 0.01
1e-06 0.0001
1e-08 1e-06
1e-10 1e-08
1e-12 1e-10
0 0.001 0.002 0.003 0.004 0.005 1e-12
0 0.001 0.002 0.003 0.004 0.005
f/g
f/g

Fig. 11 Power spectrum S( f ) on a logarithmic scale (red curve) versus f /g for the same respective sets of
parameters and initial states as in Fig. 10a, d
unc

423 above, we see that the local Lyapunov exponent is either negative, or small positive.
424 The two entanglement indicators (namely, ξtei ′ and the SVNE) therefore do not diverge

425 from each other in any significant way. We regard this as a broad corroboration of the
426
′ as an entanglement indicator. In contrast, when the power spectrum is
validity of ξtei
427 markedly broadband (Fig. 11b, d), the local Lyapunov exponent is distinctly positive,
428 showing that ξtei′ diverges significantly from the SVNE. Hence we conclude that ξ ′
tei
429 is not a reliable or satisfactory indicator of entanglement in these cases.

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
_####_ Page 16 of 21 B. Sharmila et al.

430 4.2 The double-well BEC model

431 As in the foregoing, we generate the time series of d1 (t) by calculating this difference
for 20,000 time steps, in this case with δt = 0.01. As seen in Fig. 12a–c, in this
Author Proof

432

of
433 instance  L is positive regardless of the degree of coherence of the initial states of the
434 subsystems, for a wide range of values of the ratio U /λ1 (recall that λ1 = (λ2 +ω12 )1/2 ).
435 With an increase in U /λ1 , ∞ increases (Fig. 12a, b). In contrast to the atom–field
436 interaction model, a departure of the initial state from perfect coherence causes ∞
437 to decrease (Fig. 12a, c).

pro
(a) 0.8 (b) 0.9
Λ∞=0.44, m=6.6, q=1.1 Λ∞=0.54, m=14, q=1.4
0.7 0.8
ΛL

0.6 ΛL 0.7

0.5 0.6

0.4 0.5
125 250 375 500 125 250 375 500
L L
cted (c) 0.4
Λ∞=0.15, m=3.7, q=1.0

0.3
ΛL

0.2

0.1
125 250 375 500
L

Fig. 12  L obtained from the time series of d1 (t) (blue crosses) and the fit ∞ + (m/L q ) (red curve)
versus L, with ω0 = ω1 = 1. Initial state |α, 1 ⊗ |α and a hopping frequency λ = 5, U = 0.5. b λ = 1,
orre

U = 1. c Initial state |α, 5 ⊗ |α, 5, λ = 1, U = 1. |α| = 1 in all three cases

(a) (b)
1 1
0.01 0.01
0.0001 0.0001
log S(f)
log S(f)

1e-06 1e-06
1e-08 1e-08
1e-10 1e-10
1e-12 1e-12
1e-14 1e-14
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
f/Uab f/Uab
unc

(c) 100
1
0.01
log S(f)

0.0001
1e-06
1e-08
1e-10
1e-12
1e-14
1e-16
0 0.1 0.2 0.3 0.4 0.5
f/Uab

Fig. 13 Power spectrum S( f ) on a logarithmic scale (red curve) versus f /U for the same respective sets
of parameters and initial states as in Fig. 12a–c

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Estimation of entanglement in bipartite systems directly… Page 17 of 21 _####_

438 The power spectra corresponding to the three cases in Fig. 12 are shown in Fig.
439 13. When the linear part of HBEC is dominant (λ dominates over U , Fig. 13a), S( f )
440 reflects a degree of quasiperiodicity in the time series. When U becomes comparable
to λ, however, the nonlinearity in the Hamiltonian takes over, and S( f ) exhibits a
Author Proof

441

of
442 broadband spectrum (Fig. 13b, c).

443 5 Concluding remarks

pro
444 Entanglement monotones such as the standard measures of entanglement can be con-
445 structed only if the off-diagonal elements of the density operator are known. For a
446 single-mode system, the optical tomogram is related to the density matrix ρ by


447 w(X θ , θ ) = X θ , θ |nρnm m|X θ , θ , (29)
n,m=0

448 where ρnm = n|ρ|m and {|n}, {|m} constitute the Fock basis. To get ρnm from the
tomogram w(X θ , θ ), we need to invert Eq. 29. But this is essentially the procedure
449

450
cted
for a full state reconstruction which is error-prone. As has been pointed out in [37],
451 even in linear inversion procedures to get ρ from histograms obtained experimentally,
452 small errors in the experimental data get magnified substantially. Improved proce-
453 dures for minimizing errors are of great interest and possibilities for improvement
454 have been explored both experimentally and theoretically (see, for instance, [38,39]).
455 Experimental challenges are primarily associated with gaining sufficient phase infor-
456 mation from more than one copy of the system. Studies in this regard currently focus
457 on reconstructing single-mode density matrices minimizing the errors. Adapting and
458 extending these ideas to entangled states, which is of direct relevance to us, would be
orre

459 a significant step toward extracting entanglement monotones.


460 We have demonstrated that, even without information on the off-diagonal elements
461 of the density matrix, substantial reproduction of the qualitative aspects of entangle-
462 ment dynamics can be achieved using the tomograms alone. The performance of the
463 entanglement indicator thus obtained in quantifying the extent of entanglement has
464 been assessed using two-model bipartite systems with inherent nonlinearities. It has
465 been shown that the indicator fares significantly better for generic initial states of
466 the system even during temporal evolution, compared to better-known entanglement
467 indicators such as one based on inverse participation ratios. In order to quantify the
unc

468 reliability of the indicator over long intervals of time, the difference between the SVNE
469 and our tomographic indicator has been examined using a time-series analysis. The
470 manner in which this difference is sensitive to the nonlinearity of the system, the nature
471 of the interaction, and the precise initial state is revealed by the time-series analysis.
472 The importance and relevance of this investigation lies in the fact that detailed state
473 reconstruction from the tomogram is completely avoided in identifying an efficient
474 entanglement indicator for generic bipartite systems involving continuous variables.

475 Acknowledgements One of the authors (SL) thanks M. Santhanam of IISER Pune for discussions pertaining
476 to inverse participation ratios.

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
_####_ Page 18 of 21 B. Sharmila et al.

477 Appendix 1: Reconstructing a single-mode vacuum state from the


478 tomogram

For our present purpose, we start with the tomogram for a zero-photon state obtained
Author Proof

479

of
480 numerically from the corresponding density matrix. The aim is to demonstrate how
481 to reconstruct the state from this tomogram. Since we have the exact density matrix
482 (equivalently, state) to compare it with, we can assess the efficiency of the reconstruc-
483 tion program. The procedure we employ is based on iterative maximum likelihood
484 estimates. Here, starting with a trial density matrix, an appropriate transformation

pro
485 outlined in [40] is applied iteratively on the density matrix to achieve a reconstructed
486 state with the maximum possible likelihood of being the true state (here, the vacuum
487 state). It is evident that the choice of the trial state is crucial for the efficiency of this
488 procedure. Figure 14a–c shows the different reconstructed states arising from three
489 different trial density matrices. Figure 14d is the photon number distribution of the
490 true state. It is evident that Fig. 14c is closer to the true state compared to Fig. 14a,
491 b. However, the probability of the zero-photon state is only ∼ 0.85 compared to 1 in
492 Fig. 14d.
493 More sophisticated techniques will lead to a reconstructed state closer to the zero-
494
cted
photon state. However, these are computationally intensive even in this case and
495 significantly more time-consuming for bipartite and multipartite states. It is there-
496 fore useful and efficient to extract as much information about a state directly from
497 tomograms as possible.

(a) (b)
0.7 0.8
0.6 0.7
Probability

Probability
orre

0.5 0.6
0.4 0.5
0.3 0.4
0.3
0.2 0.2
0.1 0.1
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
|n〉 |n〉
(c) (d)
0.9 1
0.8
Probability

0.7
Probability

0.8
0.6
0.5 0.6
0.4
unc

0.3 0.4
0.2 0.2
0.1
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
|n〉 |n〉

Fig. 14 Photon number distribution of the reconstructed state corresponding to trial density matrix a
0.1 9p=0 | p p|, b 0.2 4p=0 | p p| and c 0.9|00| + 0.1|11|. d Photon number distribution of the

true state |00|

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Estimation of entanglement in bipartite systems directly… Page 19 of 21 _####_

498 Appendix 2: Time evolution in the double-well BEC model

499 The double-well BEC effective Hamiltonian is given by Eq. (24). We require the
state |ψ(t) corresponding to an initial state that is a direct product of normalized
Author Proof

500

of
501 boson-added coherent states of the atoms in the wells A and B, namely,

502 |ψm 1 ,m 2 (0) = |αa , m 1  ⊗ |αb , m 2 , (30)

pro
503 where αa , αb ∈ C and m 1 , m 2 are nonnegative integers. The dependence of the state
504 on αa and αb has been suppressed on the left-hand side for notational simplicity. The
505 m-PACS |α, m [defined in Eq. (17)] reduces to the standard oscillator coherent state
506 |α for m = 0.
507 In the special case m 1 = 0, m 2 = 0, the state at any time t corresponding to the
508 initial state ψ0,0 (0) has been shown in Ref. [22] to be given by

∞ p q
1 2 +|α |2 )  β1 (t) β2 (t)
509 |ψ00 (t) = e− 2 (|αa | b

cted p,q=0
p!q!

510
511
e−it( p+q)[ω0 +U ( p+q)] | p ⊗ |q, (31)

512 where
β1 (t) = αa cos (λ1 t) + (i/λ1 )(λαb − ω1 αa ) sin (λ1 t),
513 (32)
β2 (t) = αb cos (λ1 t) + (i/λ1 )(λαa + ω1 αb ) sin (λ1 t),

514 and λ1 = (λ2 + ω12 )1/2 . It can then be shown [9] that the state vector at time t is given
by
orre

515

516 |ψm 1 ,m 2 (t) = Mm 1 ,m 2 (t)|ψ00 (t), (33)

517 where the operator Mm 1 ,m 2 (t) is as follows. Let k, l, p, q denote nonnegative integers,
518 and let
519 s = k + l + p + q, p = (k + m 2 − l), q = (l + m 1 − k). (34)

520 Further, let −1/2


κ = m 1 !m 2 !L m 1 (−|αa |2 )L m 2 (−|αb |2 )

521 (35)
unc

522 and Ŵ = cos−1 (ω1 /λ1 ). Then


⎨m1  p 
m2  q
(−1)k− p mk1 ml 2 pp qq e2i(l−k)λ1 t
      
523 Mm 1 ,m 2 (t) = κ

k=0 l=0 p=0 q=0
s  2(m 1 +m 2 )−s 
× cos 21 Ŵ sin 21 Ŵ (a † ) p+q−q (b† )q+ p− p

524

525 ×e−iω0 t(m 1 +m 2 )+iλ1 t(m 1 −m 2 ) e−iU t(m 1 +m 2 )(2Ntot +m 1 +m 2 ) . (36)

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
_####_ Page 20 of 21 B. Sharmila et al.

526 References
527 1. Ekert, A.K., Alves, C.M., Oi, D.K.L., Horodecki, M., Horodecki, P., Kwek, L.C.: Direct estimations
528 of linear and nonlinear functionals of a quantum state. Phys. Rev. Lett. 88, 217901 (2002). https://doi.
Author Proof

529 org/10.1103/PhysRevLett.88.217901

of
530 2. Horodecki, P., Ekert, A.: Method for direct detection of quantum entanglement. Phys. Rev. Lett. 89,
531 127902 (2002). https://doi.org/10.1103/PhysRevLett.89.127902
532 3. Bovino, F.A., Castagnoli, G., Ekert, A., Horodecki, P., Alves, C.M., Sergienko, A.V.: Direct measure-
533 ment of nonlinear properties of bipartite quantum states. Phys. Rev. Lett. 95, 240407 (2005). https://
534 doi.org/10.1103/PhysRevLett.95.240407
535 4. Blume-Kohout, R., Yin, J.O.S., van Enk, S.J.: Entanglement verification with finite data. Phys. Rev.

pro
536 Lett. 105, 170501 (2010). https://doi.org/10.1103/PhysRevLett.105.170501
537 5. Li, X., Shang, J., Ng, H.K., Englert, B.G.: Optimal error intervals for properties of the quantum state.
538 Phys. Rev. A 94, 062112 (2016). https://doi.org/10.1103/PhysRevA.94.062112
539 6. Rohith, M., Sudheesh, C.: Signatures of entanglement in an optical tomogram. J. Opt. Soc. Am.
540 B 33, 126 (2016). https://doi.org/10.1364/JOSAB.33.000126. URL http://josab.osa.org/abstract.cfm?
541 URI=josab-33-2-126
542 7. Laha, P., Lakshmibala, S., Balakrishnan, V.: Estimation of nonclassical properties of multiphoton coher-
543 ent states from optical tomograms. J. Mod. Opt. 65, 1466 (2018). https://doi.org/10.1080/09500340.
544 2018.1454527
545 8. Rohith, M., Sudheesh, C.: Visualizing revivals and fractional revivals in a Kerr medium using an optical
546 tomogram. Phys. Rev. A 92, 053828 (2015). https://doi.org/10.1103/PhysRevA.92.053828
547
548
cted
9. Sharmila, B., Saumitran, K., Lakshmibala, S., Balakrishnan, V.: Signatures of nonclassical effects in
optical tomograms. J. Phys. B: At. Mol. Opt. 50, 045501 (2017). https://doi.org/10.1088/1361-6455/
549 aa51a4
550 10. Yu, T., Eberly, J.: Sudden death of entanglement. Science 323, 598 (2009)
551 11. Laha, P., Sudarsan, B., Lakshmibala, S., Balakrishnan, V.: Entanglement dynamics in a model tripartite
552 quantum system. Int. J. Theor. Phys. 55, 4044 (2016)
553 12. Viola, L., Brown, W.G.: Generalized entanglement as a framework for complex quantum systems:
554 purity versus delocalization measures. J. Phys. A: Math. Theor. 40, 8109 (2007). URL http://stacks.
555 iop.org/1751-8121/40/i=28/a=S17
556 13. Brown, W.G., Santos, L.F., Starling, D.J., Viola, L.: Quantum chaos, delocalization, and entanglement
557 in disordered Heisenberg models. Phys. Rev. E 77, 021106 (2008). https://doi.org/10.1103/PhysRevE.
558 77.021106
orre

559 14. Buccheri, F., De Luca, A., Scardicchio, A.: Structure of typical states of a disordered Richardson model
560 and many-body localization. Phys. Rev. B 84, 094203 (2011). https://doi.org/10.1103/PhysRevB.84.
561 094203
562 15. Giraud, O., Martin, J., Georgeot, B.: Entropy of entanglement and multifractal exponents for random
563 states. Phys. Rev. A 79, 032308 (2009). https://doi.org/10.1103/PhysRevA.79.032308
564 16. Giraud, O., Martin, J., Georgeot, B.: Entanglement of localized states. Phys. Rev. A 76, 042333 (2007).
565 https://doi.org/10.1103/PhysRevA.76.042333
566 17. Beugeling, W., Andreanov, A., Haque, M.: Global characteristics of all eigenstates of local many-body
567 Hamiltonians: participation ratio and entanglement entropy. J. Stat. Mech. 2015, P02002 (2015). URL
568 http://stacks.iop.org/1742-5468/2015/i=2/a=P02002
569 18. Dukesz, F., Zilbergerts, M., Santos, L.F.: Interplay between interaction and (un)correlated disorder
570 in one-dimensional many-particle systems: delocalization and global entanglement. New J. Phys. 11,
unc

571 043026 (2009). URL http://stacks.iop.org/1367-2630/11/i=4/a=043026


572 19. Karthik, J., Sharma, A., Lakshminarayan, A.: Entanglement, avoided crossings, and quantum chaos in
573 an Ising model with a tilted magnetic field. Phys. Rev. A 75, 022304 (2007). https://doi.org/10.1103/
574 PhysRevA.75.022304
575 20. Lakshminarayan, A., Srivastava, S.C.L., Ketzmerick, R., Bäcker, A., Tomsovic, S.: Entanglement and
576 localization transitions in eigenstates of interacting chaotic systems. Phys. Rev. E 94, 010205 (2016).
577 https://doi.org/10.1103/PhysRevE.94.010205
578 21. Kannawadi, A., Sharma, A., Lakshminarayan, A.: Persistent entanglement in a class of eigenstates of
579 quantum Heisenberg spin glasses. EPL 115, 57005 (2016). URL http://stacks.iop.org/0295-5075/115/
580 i=5/a=57005

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Estimation of entanglement in bipartite systems directly… Page 21 of 21 _####_

581 22. Sanz, L., Moussa, M., Furuya, K.: Generation of generalized coherent states with two coupled Bose–
582 Einstein condensates. Ann. Phys. (N.Y.) 321, 1206 (2006). https://doi.org/10.1016/j.aop.2005.09.003.
583 URL http://www.sciencedirect.com/science/article/pii/S0003491605001569
584 23. Agarwal, G.S., Puri, R.R.: Collapse and revival phenomenon in the evolution of a resonant field in a
Author Proof

585 Kerr-like medium. Phys. Rev. A 39, 2969 (1989). https://doi.org/10.1103/PhysRevA.39.2969

of
586 24. Mistakidis, S., Katsimiga, G., Kevrekidis, P., Schmelcher, P.: Correlation effects in the quench-induced
587 phase separation dynamics of a two species ultracold quantum gas. New J. Phys. 20, 043052 (2018)
588 25. Gross, C., Strobel, H., Nicklas, E., Zibold, T., Bar-Gill, N., Kurizki, G., Oberthaler, M.: Atomic
589 homodyne detection of continuous-variable entangled twin-atom states. Nature 480, 219 (2011)
590 26. Smithey, D.T., Beck, M., Raymer, M.G., Faridani, A.: Measurement of the Wigner distribution and
591 the density matrix of a light mode using optical homodyne tomography: application to squeezed states

pro
592 and the vacuum. Phys. Rev. Lett. 70, 1244 (1993). https://doi.org/10.1103/PhysRevLett.70.1244
593 27. Vogel, K., Risken, H.: Determination of quasiprobability distributions in terms of probability dis-
594 tributions for the rotated quadrature phase. Phys. Rev. A 40, 2847 (1989). https://doi.org/10.1103/
595 PhysRevA.40.2847
596 28. Ibort, A., Man’ko, V.I., Marmo, G., Simoni, A., Ventriglia, F.: An introduction to the tomographic
597 picture of quantum mechanics. Phys. Script. 79, 065013 (2009). URL http://stacks.iop.org/1402-4896/
598 79/i=6/a=065013
599 29. Lvovsky, A.I., Raymer, M.G.: Continuous-variable optical quantum-state tomography. Rev. Mod. Phys.
600 81, 299 (2009). https://doi.org/10.1103/RevModPhys.81.299
601 30. Weigert, S., Wilkinson, M.: Mutually unbiased bases for continuous variables. Phys. Rev. A 78, 020303
602 (2008). https://doi.org/10.1103/PhysRevA.78.020303
31. Sudheesh, C., Lakshmibala, S., Balakrishnan, V.: Manifestations of wave packet revivals in the moments
603
604
cted
of observables. Phys. Lett. A 329, 14 (2004). https://doi.org/10.1016/j.physleta.2004.06.085. URL
605 http://www.sciencedirect.com/science/article/pii/S0375960104009132
606 32. Biswas, A., Agarwal, G.S.: Nonclassicality and decoherence of photon-subtracted squeezed states.
607 Phys. Rev. A 75, 032104 (2007). https://doi.org/10.1103/PhysRevA.75.032104
608 33. Abarbanel, H.D.I.: Analysis of Observed Chaotic Data. Springer, New York (1996)
609 34. Grassberger, P., Procaccia, I.: Characterization of strange attractors. Phys. Rev. Lett. 50, 346 (1983).
610 https://doi.org/10.1103/PhysRevLett.50.346
611 35. Abarbanel, H.D.I., Brown, R., Kennel, M.B.: Local Lyapunov exponents computed from observed
612 data. J. Nonlinear Sci. 2, 343 (1992). https://doi.org/10.1007/BF01208929
613 36. Hegger, R., Kantz, H., Schreiber, T.: Practical implementation of nonlinear time series methods: the
614 TISEAN package. Chaos 9, 413 (1999)
37. Bartkiewicz, K., Černoch, A., Lemr, K., Miranowicz, A.: Priority choice experimental two-qubit tomog-
orre

615
616 raphy: measuring one by one all elements of density matrices. Sci. Rep. 6, 19610 (2016)
617 38. Opatrný, T., Welsch, D.G.: Density-matrix reconstruction by unbalanced homodyning. Phys. Rev. A
618 55, 1462 (1997). https://doi.org/10.1103/PhysRevA.55.1462
619 39. Brida, G., Genovese, M., Gramegna, M., Traina, P., Predazzi, E., Olivares, S., Paris, M.: Toward a full
620 reconstruction of density matrix by on/off measurements. Int. J. Quant. Inf. 7, 27 (2009)
621 40. Lvovsky, A.: Iterative maximum-likelihood reconstruction in quantum homodyne tomography. J. Opt.
622 B: Quant. Semiclass. Opt. 6(6), S556 (2004)

623 Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps
624 and institutional affiliations.
unc

123
Journal: 11128 Article No.: 2352 TYPESET DISK LE CP Disp.:2019/6/8 Pages: 21 Layout: Small-Ex
Journal: 11128
Article: 2352

Author Query Form


Author Proof

of
Please ensure you fill out your response to the queries raised below
and return this form along with your corrections

pro
Dear Author
During the process of typesetting your article, the following queries have arisen. Please
check your typeset proof carefully against the queries listed below and mark the
necessary changes either directly on the proof/online grid or in the ‘Author’s response’
area provided below

Query Details required Author’s response


cted
1. Kindly check and confirm whether
the corresponding author is correctly
identified.
2. As per the information provided by
the publisher, Figs. 2–7, 10, 12, 13
will be black and white in print; hence,
please confirm whether we can add
“colour figure online” to the caption.
orre
unc

Das könnte Ihnen auch gefallen