Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Generation of a non-zero discord bipartite state with classical second-order interference

Open Access Open Access

Abstract

We report an investigation on quantum discord in classical second-order interference. In particular, we theoretically show that a bipartite state with D = 0.311 of discord can be generated via classical second-order interference. We also experimentally verify the theory by obtaining D = 0.197 ± 0.060 of non-zero discord state. Together with the fact that the nonclassicalities originated from physical constraints and information theoretic perspectives are not equivalent, this result provides an insight to understand the nature of quantum discord.

© 2017 Optical Society of America

1. Introduction

Quantum correlation plays essential roles in many quantum information processes as well as fundamental experiments. Quantum entanglement has been spotlighted as a representative example of quantum correlation [1]. However, it has been known that entanglement is not the only kind of quantum correlation [2–4]. In order to capture all the nonclassical correlation, Ollivier and Zurek introduced a measure, quantum discord, from the information-theoretic perspectives [5].

Quantum discord covers a broader concept of quantum correlation than entanglement, and thus it has entanglement as a subset. In other words, all the entangled states have non-zero discord, however, there exist non-zero discord states which are separable [6]. Moreover, almost all quantum states have non-zero discord, and thus, generating non-zero discord state is not difficult [7].

Quantum discord has been actively studied due to the extensive coverage beyond entanglement. It has been applied in various research fields such as quantum communication [8,9], quantum computation [10, 11], quantum metrology [12], thermodynamics of information [13, 14], and dynamics of open system [15,16]. There are also several studies on the generation of quantum discord in continuous variable regime [17–19] and discrete variable regime [11,20,21].

In the sense that interference is a fundamental phenomenon which may induce some correlations, it would be of interest to investigate quantum discord with the context of interference. Indeed, studying quantum discord in the context of interference may provide distinct quantumness criteria of physical constraints and information theoretic perspectives [22]. There are a few studies that relate interference to discord [23–25]. In these works, the authors showed that the interference between Gaussian states can be revealed with the aid of an ancillary Gaussian discordant state [23], and the discord can be a resource of estimating parameters in metrology [24]. The relation between discord and interference for mixed states was also investigated [25]. However, there is no study that directly shows the role of interference for generating discord.

In this paper, we study quantum correlation of bipartite systems in the context of interference. In particular, we consider both entanglement and discord with classical second-order interference. The theoretical and experimental results show that the second-order interference can have an important role for generating non-zero discord bipartite states. Since interference is a fundamental phenomenon both in classical and quantum physics, linking quantum correlation with interference will provide an insight into quantum correlation.

2. Theory

2.1. Quantum discord: the definition

We briefly review the concept of quantum discord. In the classical information theory, given two random variables, A and B, we can define the mutual information in two alternative ways:

I(A:B)=H(A)+H(B)H(A,B),
J(A:B)=H(A)H(A|B).
where H(A) = −Σa PA=a log PA=a is the Shannon entropy, H(A, B) = −Σa,b PA=a,B=b log PA=a,B=b is the joint entropy of A and B, and H(A|B) = Σb PB=b H(A|B=b) is the conditional entropy of A given B with the probability distribution P and the outcome a(b) for A(B). The marginal probability distributions, PA and PB, are obtained from the joint probability distribution, PA,B, such that PA = Σb PA,B=b and PB = Σa PA=a,B. By applying the Bayes rule, PA|B=b = PA,B=b/PB=b, we can easily show that H(A|B) = H(A, B) − H(B), and thus find out that the two definitions of mutual information are equivalent.

Considering the quantum version of the mutual information, we replace the classical probability distributions with the density matrices of two parties, ρA, ρB, and ρAB, and the Shannon entropy with the von Neumann entropy, S(ρA) = −TrA ρA log ρA, where ρA = TrB ρAB and ρB = TrA ρAB are the reduced density matrices of the joint density matrix, ρAB. For instance, I(ρAB) = S(ρA) + S(ρB) − S(ρAB). The two definitions of mutual information are no longer same in quantum version due to the conditional entropy in J(A : B). The quantum conditional entropy, S(ρA|B), is not directly formulated because the state of A can be affected by the measurement on B. Therefore, S(ρA|B) has to be defined according to the measurement on B such that S(ρA|B)=S(ρA|{ΠJB})=jPjS(ρA|ΠjB) where ρA|ΠjB=ΠjBρABΠjB/TrA,BΠjBρAB, Pj=TrA,BΠjBρAB, and {ΠJB} is a set of measurement on B. Since J(ρAB) is measurement-dependent in quantum physics, we express it as J(ρAB){ΠjB}=S(ρA)S(ρA|B)=S(ρA)S(ρA|{ΠjB}).

Quantum discord is then given by the difference between two expressions of the mutual information in quantum version:

D(ρAB){ΠjB}=I(ρAB)J(ρAB){ΠjB}
As I(ρAB) and J(ρAB){ΠjB} are total and classical correlations [26], respectively, we finally obtain the measurement-independent quantum discord by maximizing J(ρAB){ΠjB} over all possible measurement set of {ΠjB}:
D(ρAB)=I(ρAB)max{ΠjB}J(ρAB){ΠjB}.
Note that we present the above expression with projective measurements since a projector provides the optimal measurement for the conditional entropy of two qubit states [27]. Moreover, it is worth noting that quantum discord for bipartite pure states is equivalent to the measure for entanglement [10,26].

2.2. Quantum discord and classical second-order interference

Let us consider a typical two-photon interference with a lossless symmetric beamsplitter (BS) that has two input modes a, b and output modes c, d. The BS transformation between the input and output modes can be written with creation operators, a12(c+id), and b12(d+ic). With this BS transformation, we can think of various kinds of two-photon interference by changing the input states.

First, let us investigate the case that two identical single photons are interfering at the BS, which corresponds to a typical Hong-Ou-Mandel interference [28]. In this case, the coincidences between the outputs, c and d, become null, showing a Hong-Ou-Mandel (HOM) dip with an unit visibility, V = 1. Note that the visibility of a HOM dip is defined as the relative depth of the dip with respect to the non-interfering cases.

If we change the polarization states of the incoming photons orthogonal to each other, the HOM dip disappears. However, in this setup, we can get a maximally entangled two-qubit state in polarization mode by post-selecting the case only when there is one photon at each output mode [29]. It is remarkable that the second-order interference is responsible for generating the entangled two-qubit state although there is no HOM dip with orthogonally polarized input photons. If the arrival time difference of the input photons is larger than the coherence time of the single-photon states, the output two-qubit state becomes completely mixed state [30]. Note that a HOM dip with V > 0.5 cannot be explained with classical theory, and thus it is considered as quantum second-order interference [31]. Therefore, this result shows that entanglement can be generated by quantum interference. Since all the entangled states have non-zero discord, we can conclude that quantum interference can generate quantum discord as well.

Now, let us turn our interest to classical second-order interference. For example, we can think of the case that two identical but phase randomized laser pulses are incoming to the input modes. In this case, one can get the HOM dip with limited visibility of V≤0.5. Note that the first-order interference would be washed out due to the randomized phase between the two inputs. This result can be completely explained by the classical theory of superposition of electromagnetic waves, and thus it is considered as classical interference [31–33]. We can get a two-qubit state in polarization mode by setting the polarization states of the inputs orthogonal to each other and post-selecting the case when there is only one photon at each output, as we did for the single-photon inputs.

We can classify the classical two-photon interference according to the existence of the mutual coherence between the laser pulses on the two input modes, a and b. Assuming that the polarization states of the input pulses are horizontal and vertical respectively, the two-qubit states of the mutually coherent inputs ρcoh and the mutually incoherent inputs ρincoh after the BS are given as

ρcoh(ϕ)=14(1ieiϕieiϕie2iϕieiϕ11ieiϕieiϕ11ieiϕie2iϕieiϕieiϕ1)
ρincoh=14(1000011001100001)
where ϕ is the phase difference between the input modes. The basis set of the density matrices is {|HH〉, |HV〉, |VH〉, |VV〉}, where H and V are horizontal and vertical polarization states, respectively. Note that 〈ρcoh (ϕ)〉ϕ = ρincoh, where 〈Xϕ represents the average of X over many events of randomly varying ϕ. The detailed derivation can be found in Appendix A.

From a given two-qubit density matrix, we can estimate the amount of entanglement and quantum discord [34,35]. Note that the amount of entanglement can be quantified by concurrence. Regardless of the phase difference ϕ, the concurrence and quantum discord of ρcoh (ϕ) are C(ρcoh(ϕ)) = D(ρcoh(ϕ)) = 0, showing that ρcoh (ϕ) does not have any quantum correlation. On the other hand, the concurrence and discord of ρincoh are C(ρincoh) = 0 and D(ρincoh)=(34)log2(34)0.311, respectively. This result shows that classical second-order interference can generate a certain type of quantum correlation that can be captured by quantum discord. It is interesting to note that the existence of mutual coherence between the two input modes degrades the quantum discord.

Let us consider the physics behind these results. Taking into account the theorem that non-classical inputs are indispensable for generating entanglement with a BS [36,37], it is obvious that the concurrence is zero in both cases. Quantum discord is not simply explained, but can be understood by investigating the reduced single-qubit states. Once we trace out one of the two qubits of Eq.(5) and (6), we obtain the reduced single-qubit density matrices for each output of the BS, c and d. The reduced single-qubit density matrices are given as

ρcoh|c(ϕ)=12(1ieiϕieiϕ1),
ρcoh|d(ϕ)=12(1ieiϕieiϕ1),
ρincoh|c=ρincoh|d=12(1001),
where ρcoh|c(d)(ϕ) and ρincoh|c(d) denote the reduced single-qubit density matrices at output c(d) for mutually coherent and incoherent inputs, respectively. Note that ρcoh|c (0) = |R〉〈R| and ρcoh|d (0) = |L〉〈L| where |R12(|H+i|V) and |L12(|H+i|V) By changing ϕ, one can get single-qubit states that reside on the equator of a Bloch sphere.

From the investigation of the reduced single-qubit states, the role of the mutual coherence between the input modes becomes clear. When the two inputs are mutually coherent, both output single-qubit states become pure and orthogonal to each other. As a result, the overall two-qubit state can be represented as a product of two single-qubit states, ρcoh = ρcoh|c ρcoh|d, and thus no correlation exists in the bipartite system. When input pulses are incoherently mixed at a BS, on the other hand, both of the reduced single-qubit states become completely mixed. In this case, the two-qubit state becomes a classical mixture of anti-correlated qubits on the equator of a Bloch sphere. Therefore, we can infer that the non-zero quantum discord of the mutually incoherent inputs case arises from the statistical mixing process. It is remarkable that the same two-qubit state can be obtained by randomly rotating the polarization of two beams by the same amount.

3. Experiment

3.1. Mutually coherent inputs

In order to verify the theory, we construct an experimental setup as shown in Fig. 1(a) for the mutually coherent inputs case. A Ti:Sapphire mode-locked femtosecond laser (fs-laser) generates laser pulses whose temporal width is less than 200 fs and the repetition rate is about 80 MHz. The center wavelength of the pulses is tuned to 785 nm. The filter set consists of two parts: One is neutral density filters that attenuate the intensity of the pulses and the other is an interference filter whose full width at half maximum (FWHM) of the transmission profile is about 3 nm.

 figure: Fig. 1

Fig. 1 Experimental setup where the phase of the two input pulses are mutually (a) coherent and (b) incoherent. fs-laser : femtosecond laser, filter set : filter set of neutral density filters and an interference filter, BS : 50:50 beamsplitter, AOM : acousto-optic modulator, FC : fiber coupler, PC : polarization controller, FBS : fiber beamsplitter, C : collimation lens, Q : quarter-wave plate, H : half-wave plate, P : polarizer, SPD : single photon detector, CCU : coincidence counting unit.

Download Full Size | PDF

To minimize the contribution of multi-photon states (more than two photons) to the coincidence detection, we attenuate the laser pulses so that the average photon number per pulse is about µ ≈ 0.1. With this condition, the multi-photon states contribution is less than 10 % of the total coincidence detection, see Appendix B for details.

After passing through the filters, the laser pulse enters a displaced Sagnac interferometer. Note that the second-order interference happens when two split laser pulses come out from the Sagnac interferometer. The Sagnac interferometer is employed to stabilize the mutual phase difference between the inputs. Although entanglement and discord are not affected by the phase difference, the stabilization of the phase difference during quantum state tomography is essential. One can achieve high-level of stabilization with a displaced Sagnac interferometer configuration [38,39].

In the Sagnac interferometer, two half-wave plates (H1 and H2) are employed to make the polarization states of the input pulses orthogonal, |H〉 and |V〉. Note that H1 and H2 can be horizontally tilted to adjust the phase difference between the two inputs. During the experiment, the mutual phase difference is set at ϕ = 0. At the two outputs of the BS, the two-qubit density matrix is reconstructed by means of quantum state tomography with maximum likelihood estimation using quarter-wave plates (Q), half-wave plates (H), and polarizers (P) [40, 41]. A home-made coincidence counting unit (CCU) based on an FPGA is used to register the coincidence counts between the SPDs [42]. The coincidence counts are accumulated for 10 seconds for each projection state.

Figure 2(a) shows the theoretically (upper) and experimentally (lower) reconstructed two-qubit density matrices. The fidelity between these two states, F = 0.992 0.001, shows that the experimentally reconstructed state is very close to the theoretical one. The concurrence and discord of the experimental two-qubit state is calculated as C(ρcoh) = 0.005 ±0.002 and D(ρcoh) = 0.002 ± 0.001, which verifies that the state has neither entanglement nor quantum discord.

 figure: Fig. 2

Fig. 2 Two-qubit density matrices where the phase of the pulses incident on the two input modes of a BS are mutually (a) coherent or (b) incoherent. ρcoh(ρcohexp) is the theoretical (experimental) density matrix of the coherent case. ρincoh(ρcohexp) is the theoretical (experimental) density matrix of the incoherent case. Re[·] and Im[·] denote the real and imaginary parts of the density matrix, respectively.

Download Full Size | PDF

3.2. Mutually incoherent inputs

The experimental setup for the mutually incoherent inputs is depicted in Fig. 1(b). The light source and the tomographic setup (Q, H, P) are identical to those of the mutually coherent inputs case. Two asynchronous acousto-optic modulators (AOM1,2) are employed to wash out the first-order coherence between the two inputs [32, 33]. To minimize the frequency shift by the AOMs, the operating RF frequencies of both AOM drivers are identically set to 40MHz. We block out all but the first-order diffracted pulses. H1 and H2 are employed to make the polarization states orthogonal to each other (|H〉 and |V〉). The pulses incident on fiber couplers (FC) interfere at a fiber beamsplitter (FBS). The polarization controllers (PC) at the input and output ports of the FBS compensate the polarization drift during fiber transmission. The output polarization states are investigated with the following tomographic setup.

Before we investigate the two-qubit state, we observe the classical second-order interference. To this end, we temporally set the polarization states of the inputs identical and remove the tomographic setup. By scanning one of the optical paths with a translation stage, we measure the coincidences between the two outputs of the BS, see Fig. 3. The 49.0 ± 1.03% visibility of classical HOM dip shows that the spatial and polarization modes of the pulses are matched well. Note that the classical limit of the HOM dip visibility is 50% [31]. The FWHM of the HOM dip, which is measured as 190 µm, also corresponds well to the coherence length elongated by the interference filter.

 figure: Fig. 3

Fig. 3 Classical Hong-Ou-Mandel dip. The means and standard deviations of the coincidence count rate data are represented with blue circles and error bars, respectively. The red solid line denotes the Gaussian fitting for the data. The visibility of the dip, V, is calculated from the fitted Gaussian function.

Download Full Size | PDF

In order to get a two-qubit state in polarization mode, we locate the stage at the zero position, make the input polarization states orthogonal, |H〉 and |V〉, and process the quantum state tomography. The theoretical and experimental density matrices are shown in Fig. 2(b). The concurrence and discord of this experimental two-qubit state are calculated as C(ρincoh) = 0 and D(ρincoh) = 0.197 ± 0.060, which reveal that the two-qubit state is separable, however has non-zero discord. It is remarkable that this result clearly shows that the criteria of quantumness originated from physical constraints and information theoretic perspectives do not coincide [22].

For the purpose of showing the role of the second-order interference, we experimentally investigate the purity and discord of the two-qubit states at various optical path length differences, see Fig. 4. As the optical path length difference between the two inputs increases, the purity and discord of the two-qubit state decrease. Note that the width of the decreasing discord with respect to the optical path length difference 125 µm is comparable with the coherence length measured by the HOM dip in Fig. 3. When the optical path length difference becomes larger than the coherence length, the two-qubit state asymptotically becomes a completely mixed state which is a zero-discord state.

 figure: Fig. 4

Fig. 4 Purity (red and right vertical axis) and discord (blue and left vertical axis) versus the optical path length difference for mutually incoherent case. The circles and error bars denote the experimental average values and standard deviations, respectively. The solid lines are the Gaussian fittings for the data.

Download Full Size | PDF

We summarize the theoretical and experimental results in Table 1. It shows that the experiment verifies the theory well, except a discrepancy on the discord of ρincoh. The discrepancy might come from experimental imperfections such as the multi-photon contribution of weak laser pulses. Also, the large standard deviations suggest that estimating quantum discord is inherently a noisy procedure, i.e., sensitive to small perturbation [43,44].

Tables Icon

Table 1. Characteristics of the theoretical and experimental two-qubit density matrices. ρmixed is the two-qubit state when the optical path length difference between two mutually incoherent optical pulses are larger than the coherence length of the optical pulses.

4. Conclusion

In conclusion, we theoretically and experimentally study entanglement and discord in classical second-order interference. When the laser pulses of the two input modes are mutually coherent, both entanglement and quantum discord of the post-selected output state become zero. On the contrary, when the pulses are mutually incoherent, the output state becomes separable, yet non-zero discord state. It shows that classical interference with post-selection can provide a certain type of quantum correlation, quantum discord. This result stimulates us to understand more about quantum correlation based on interference that is a fundamental phenomenon in both classical and quantum physics.

5. Appendix A: Two qubit states in classical second-order interference

We develop the theory describing post-selected two-qubit state in the classical second-order interference. We assume that the light incident on a lossless symmetric beamsplitter (BS) is from a monochromatic single-mode laser and the photon number of the output modes of the BS can be resolved. We post-select only the case where two photons are involved and one photon emerges from each output mode. The unitary transformation of the BS can be written as

(ab)12(1ii1)(cd)
where a and b denote the creation operators at the input modes, c and d are those for the output modes. Note that the polarization states of the two input modes a and b are always orthogonal to each other. Therefore, we can simplify the situation as aaH and bbV, where the subscripts H and V denote horizontal and vertical polarization states, respectively.

5.1. Mutually coherent inputs

Since the photon number statistics of a laser follows Poissonian, we can write the laser pulse at each input mode of the BS as

|α=e|α|2/2m=0αmm!(aH)m|0
|β=e|β|2/2n=0βnn!(eiϕbV)n|0
where ϕ is the phase difference between the two input modes. Assuming that the average photon numbers of the two input modes are the same, we can express α and β as µ1/2 where µ is the average photon number. The overall input state of the BS can be represented as
|ψcohin=eμm,n=0μ(m+n)/2m!n!(aH)m(eiϕbV)n|0|0.

One can obtain the full expression of the output state with the BS transformation, Eq.(10), and the input state, Eq.(13) as follows.

|ψcohout=eμm,n=0(μ/2)(m+n)/2m!n!(cH+idH)m(eiϕ(icV+dV))n|0|0.
Since we are interested in such output state that each output mode has only one photon, the post-selected output state is given as
|ψcoh=12(icHdH+eiϕcHdVeiϕcVdH+iei2ϕcVdV)|0|0.
Note that this state is identical to the coherent output state, Eq. (5), i.e., ρcoh (ϕ) = |ψcohψ|.

5.2. Mutually incoherent inputs

When two input modes of the BS are mutually incoherent to each other, we can deal with the input state separately according to the incident photon numbers at each input mode, i and j. Because we are interested in such output state that each output mode has only one photon, there are three cases for the input state contributing to the post-selected output state, ρincoh. The output state will be the statistical mixture of the output states of these three input states. In order to find the output density matrix, we calculate each case as follows.

Case 1) |ψ(1,1)in=aHbV|0|0. This input state corresponds to that each input mode has one photon, respectively. After the BS transformation, the output state is presented as

|ψ(1,1)=12(icHcV+cHdVcVdH+idHdV)|0|0.
After the post-selection of one photon at each output mode, we can obtain the output state in the density matrix form as
ρincoh(1,1)=12(0000011001100000),
which corresponds to a Bell state |ψ〉.

Case 2) |ψ(2,0)in=12(aH)2|0|0. This input state corresponds to that there are two photons and no photon at input a and b, respectively. After the BS transformation, the output state is given as

|ψ(2,0)=122((cH)2+2icHdH(dH)2)|0|0.
Therefore, the post-selected output state can be presented in the density matrix form as
ρincoh(2,0)=(1000000000000000).
Case 3) |ψ(0,2)in=12(bV)2|0|0 This input state corresponds to that there are no photon and two photons at input a and b, respectively. After the BS transformation, the output state is given as
|ψ(0,2)=122((cV)2+2icVdV+(dV)2)|0|0.
Therefore, the post-selected output state can be presented in the density matrix form as
ρincoh(0,2)=(0000000000000001).
The overall two-qubit state ρincoh of Eq.(6) is given as the statistical mixture of ρincoh(1,1), ρincoh(2,0), and ρincoh(0,2). Taking account of the photon number distribution of a laser pulse, we can get the overall two-qubit state as
ρincoh=p(1,1)ρincoh(1,1)+p(2,0)ρincoh(2,0)+p(0,2)ρincoh(0,2)Tr[p(1,1)ρincoh(1,1)+p(2,0)ρincoh(2,0)+p(0,2)ρincoh(0,2)]
where p(1,1) = P(μ, 1)2/2, p(2, 0) = P(μ, 2)P(μ, 0)/2, and p(0, 2) = P(μ, 0)P(μ, 2)/2 are the conditional probabilities that correspond to ρincoh(1,1), ρincoh(2,0) and ρincoh(0,2), respectively. Since the P(µ, n) = eµ µn/n! is the Poissonian distribution of the photon numbers where µ and n is the average photon number and the number of photons, the conditional probability p(i, j) means that the probability of occurring coincidence event when i and j photons are incident on a and b inputs, respectively.

6. Appendix B : Multi-photon states contribution to coincidence detection

Although we assume ideal single-photon detectors that resolve the number of photons in theory, practical single-photon detectors cannot resolve photon numbers, and thus the multi-photon states can affect the coincidence detections. The multi-photon effect can be estimated by investigating the ratio of the coincidence detections between two-photon states and multi-photon states.

To this end, let us assume that the efficiencies of single photon detectors are not dependent on the number of photons incident on the detectors. Then, we can define an error function with respect to the average photon number µ, which estimates the proportion of erroneous coincidence detections to the total coincidence detections:

E(μ)=i+j3p(μ,i,j)i+j2p(μ,i,j)
where p(µ, i, j) = P(µ, i)P(µ, j)Q(i + j) is the conditional probability, P(µ, n) = eµ µn/n! is the Poisson distribution of the photon numbers, and Q(n)=1(12)n1 is the probability function of the number of photons involved in the coincidence detection.

Figure 5 shows E(µ) as a function of the average photon number µ. Note that we limit the number of photons involved (i+ j) to 100. As the average photon number increases, the proportion of erroneous coincidence detection due to the multi-photon states increases. For µ = 0.1, which is investigated for our experiment, E(µ) = 0.096, and therefore we can conclude that most of the coincidence detections are due to the two-photon states.

 figure: Fig. 5

Fig. 5 The proportion of the erroneous coincidence detections due to multi-photon states.

Download Full Size | PDF

Funding

ICT R&D programs of MSIP/IITP (B0101-16-1355, R0101-15-0060); National Research Foundation of Korea (2015K2A1B8047228, 2016R1A2A1A05005202); KIST Research Programs (2E27231, 2V05340, 2E26680-16-P024).

References and links

1. R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, “Quantum entanglement,” Rev. Mod. Phys. 81, 865 (2009). [CrossRef]  

2. C. H. Bennett, D. P. DiVincenzo, C. A. Fuchs, T. Mor, E. Rains, P. W. Shor, J. A. Smolin, and W. K. Wootters, “Quantum nonlocality without entanglement,” Phys. Rev. A 59, 1070 (1999). [CrossRef]  

3. M. Horodecki, P. Horodecki, R. Horodecki, J. Oppenheim, A. Sen, U. Sen, and B. Synak-Radtke, “Local versus nonlocal information in quantum-information theory: Formalism and phenomena,” Phys. Rev. A 71, 062307 (2005). [CrossRef]  

4. J. Niset and N. J. Cerf, “Multipartite nonlocality without entanglement in many dimensions,” Phys. Rev. A 74, 052103 (2006). [CrossRef]  

5. H. Ollivier and W.H. Zurek, “Quantum discord: A measure of the quantumness of correlations,” Phys. Rev. Lett. 88, 017901 (2001). [CrossRef]  

6. K. Modi, A. Brodutch, H. Cable, T. Paterek, and V. Vedral, “The classical-quantum boundary for correlations: Discord and related measures,” Rev. Mod. Phys. 84, 1655 (2012). [CrossRef]  

7. A. Ferraro, L. Aolita, D. Cavalcanti, F. M. Cucchietti, and A. Acín, “Almost all quantum states have nonclassical correlations,” Phys. Rev. A 81, 052318 (2010). [CrossRef]  

8. B. Dakic, Y.O. Lipp, X.S. Ma, M. Ringbauer, S. Kropatschek, S. Barz, T. Paterek, V. Vedral, A. Zeilinger, C. Brukner, and P. Walther, “Quantum discord as resource for remote state preparation,” Nat. Phys. 8, 666–670 (2012). [CrossRef]  

9. S. Pirandola, “Quantum discord as a resource for quantum cryptography,” Sci. Rep. 4, 6956 (2014). [CrossRef]   [PubMed]  

10. A. Datta, A. Shaji, and C. M. Caves, “Quantum discord and the power of one qubit,” Phys. Rev. Lett. 100, 050502 (2008). [CrossRef]   [PubMed]  

11. B. P. Lanyon, M. Barbieri, M. P. Almeida, and A. G. White, “Experimental quantum computing without entanglement,” Phys. Rev. Lett. 101, 200501 (2008). [CrossRef]   [PubMed]  

12. K. Modi, H. Cable, M. Williamson, and V. Vedral, “Quantum correlations in mixed-state metrology,” Phys. Rev. X 1, 021022 (2011).

13. W. H. Zurek, “Quantum discord and Maxwell’s demons,” Phys. Rev. A 67, 012320 (2003). [CrossRef]  

14. A. Brodutch and D. R. Terno, “Quantum discord, local operations, and Maxwell’s demons,” Phys. Rev. A 81, 062103 (2010). [CrossRef]  

15. A. Shabani and D. A. Lidar, “Vanishing quantum discord is necessary and sufficient for completely positive maps,” Phys. Rev. Lett. 102, 100402 (2009). [CrossRef]   [PubMed]  

16. L. Mazzola, J. Piilo, and S. Maniscalco, “Sudden transition between classical and quantum decoherence,” Phys. Rev. Lett. 104, 200401 (2010). [CrossRef]   [PubMed]  

17. R. Blandino, M. G. Genoni, J. Etesse, M. Barbieri, M. G. A. Paris, P. Grangier, and R. Tualle-Brouri, “Homodyne estimation of Gaussian quantum discord,” Phys. Rev. Lett. 109, 180402 (2012). [CrossRef]   [PubMed]  

18. S. Hosseini, S. Rahimi-Keshari, J. Y. Haw, S. M. Assad, H. M. Chrzanowski, J. Janousek, T. Symul, T. C. Ralph, and P. K. Lam, “Experimental verification of quantum discord in continuous-variable states,” J. Phys. B: At. Mol. Opt. Phys. 47, 025503 (2014). [CrossRef]  

19. V. Chille, N. Quinn, C. Peuntinger, C. Croal, L. Mista Jr., C. Marquardt, G. Leuches, and N. Korolkova, “Quantum nature of Gaussian discord: Experimental evidence and role of system-environment correlations,” Phys. Rev. A 91, 050301 (2015). [CrossRef]  

20. J.-S. Xu, X.-Y. Xu, C.-F. Li, C.-J. Zhang, X.-B. Zou, and G.-C. Guo, “Experimental investigation of classical and quantum correlations under decoherence,” Nat. Comm. 1, 7 (2010). [CrossRef]  

21. M. P. Almeida, M. Gu, A. Fedrizzi, M. A. Broome, T. C. Ralph, and A. G. White, “Entanglement-free certification of entangling gates,” Phys. Rev. A 89, 042323 (2014). [CrossRef]  

22. A. Ferraro and M. G. A. Paris, “Nonclassicality criteria from phase-space representations and information-theoretical constraints are maximally inequivalent,” Phys. Rev. Lett. 108, 260403 (2012). [CrossRef]   [PubMed]  

23. A. Meda, S. Olivares, I. P. Degiovanni, G. Brida, M. Genovese, and M. G. A. Paris, “Revealing interference by continuous variable discordant states,” Opt. Lett. 38, 3099–3102 (2013). [CrossRef]   [PubMed]  

24. D. Girolami, A. M. Souza, V. Giovannetti, T. Tufarelli, J. G. Filgueiras, R. S. Sarthour, D. O. Soares-Pinto, I. S. Oliveira, and G. Adesso, “Quantum Discord Determines the Interferometric Power of Quantum States,” Phys. Rev. Lett. 112, 210401 (2014). [CrossRef]  

25. L. Zhang, A. K. Pati, and J. Wu, “Interference visibility, entanglement, and quantum correlation,” Phys. Rev. A 92, 022316 (2015). [CrossRef]  

26. L. Henderson and V. Vedral, “Classical, quantum and total correlations,” J. Phys. A: Math. Gen. 34, 6899 (2001). [CrossRef]  

27. S. Hamieh, R. Kobes, and H. Zaraket, “Positive-operator-valued measure optimization of classical correlations,” Phys. Rev. A 70, 052325 (2004). [CrossRef]  

28. C. K. Hong, Z. Y. Ou, and L. Mandel, “Measurement of subpicosecond time intervals between two photons by interference,” Phys. Rev. Lett. 59, 2044 (1987). [CrossRef]   [PubMed]  

29. Y. H. Shih and C. O. Alley, “New type of Einstein-Podolsky-Rosen-Bohm experiment using pairs of light quanta produced by optical parametric down conversion,” Phys. Rev. Lett 61, 2921 (1998). [CrossRef]  

30. H.-T. Lim, Y.-S. Kim, Y.-S. Ra, J. Bae, and Y.-H Kim, “Experimental realization of an approximate partial transpose for photonic two-qubit systems,” Phys. Rev. Lett. 107, 160401 (2011). [CrossRef]   [PubMed]  

31. J. G. Rarity, P. R. Tapster, and R. Loudon, “Non-classical interference between independent sources,” J. Opt. B: Quantum Semiclass. Opt. 7, S171 (2005). [CrossRef]  

32. Y.-S. Kim, O. Slattery, P. S. Kuo, and X. Tang, “Conditions for two-photon interference with coherent pulses,” Phys. Rev. A 87, 063843 (2013). [CrossRef]  

33. Y.-S. Kim, O. Slattery, P. S. Kuo, and X. Tang, “Two-photon interference with continuous-wave multi-mode coherent light,” Opt. Express 22, 3611 (2014). [CrossRef]   [PubMed]  

34. W. K. Wootters, “Entanglement of formation of an arbitrary state of two qubits,” Phys. Rev. Lett. 80, 2245 (1998). [CrossRef]  

35. J. Yune, K.-H. Hong, H.-T. Lim, J.-C. Lee, O. Kwon, S.-W. Han, Y.-S. Kim, S. Moon, and Y.-H. Kim, “Quantum discord protection from amplitude damping decoherence,” Opt. Express 23, 26012 (2015). [CrossRef]   [PubMed]  

36. M. S. Kim, W. Son, V. Bužek, and P. L. Knight, “Entanglement by a beam splitter: Nonclassicality as a prerequisite for entanglement,” Phys. Rev. A 65, 032323 (2002). [CrossRef]  

37. X.-B. Wang, “Theorem for the beam-splitter entangler,” Phys. Rev. A 66, 024303 (2002). [CrossRef]  

38. J.-C. Lee, Y.-C. Jeong, Y.-S. Kim, and Y.-H. Kim, “Experimental demonstration of decoherence suppression via quantum measurement reversal,” Opt. Express 19, 16309 (2011). [CrossRef]   [PubMed]  

39. Y.-S. Kim, J-C. Lee, O. Kwon, and Y.-H. Kim, “Protecting entanglement from decoherence using weak measurement and quantum measurement reversal,” Nat. Phys. 8, 117 (2012). [CrossRef]  

40. K. Banaszek, G. M. D’Ariano, M. G. A. Paris, and M. F. Sacchi, “Maximum-likelihood estimation of the density matrix,” Phys. Rev. A 61, 010304 (1999). [CrossRef]  

41. D. F. V. James, P. G. Kwiat, W. J. Munro, and A. G. White, “Measurement of qubits,” Phys. Rev. A 64, 052312 (2001). [CrossRef]  

42. B. K. Park, Y.-S. Kim, O. Kwon, S.-W. Han, and S. Moon, “High-performance reconfigurable coincidence counting unit based on a field programmable gate array,” Appl. Opt. 54, 4727 (2015). [CrossRef]   [PubMed]  

43. C. Benedetti, A. P. Shurupov, M. G. A. Paris, G. Brida, and M. Genovese, “Experimental estimation of quantum discord for a polarization qubit and the use of fidelity to assess quantum correlations,” Phys. Rev. A 87, 052136 (2013). [CrossRef]  

44. A. Mandarino, M. Bina, C. Porto, S. Cialdi, S. Olivares, and M. G. A. Paris, “Assessing the significance of fidelity as a figure of merit in quantum state reconstruction of discrete and continuous-variable systems,” Phys. Rev. A 93, 062118 (2016). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 Experimental setup where the phase of the two input pulses are mutually (a) coherent and (b) incoherent. fs-laser : femtosecond laser, filter set : filter set of neutral density filters and an interference filter, BS : 50:50 beamsplitter, AOM : acousto-optic modulator, FC : fiber coupler, PC : polarization controller, FBS : fiber beamsplitter, C : collimation lens, Q : quarter-wave plate, H : half-wave plate, P : polarizer, SPD : single photon detector, CCU : coincidence counting unit.
Fig. 2
Fig. 2 Two-qubit density matrices where the phase of the pulses incident on the two input modes of a BS are mutually (a) coherent or (b) incoherent. ρ c o h ( ρ c o h exp ) is the theoretical (experimental) density matrix of the coherent case. ρ i n c o h ( ρ c o h exp ) is the theoretical (experimental) density matrix of the incoherent case. Re[·] and Im[·] denote the real and imaginary parts of the density matrix, respectively.
Fig. 3
Fig. 3 Classical Hong-Ou-Mandel dip. The means and standard deviations of the coincidence count rate data are represented with blue circles and error bars, respectively. The red solid line denotes the Gaussian fitting for the data. The visibility of the dip, V, is calculated from the fitted Gaussian function.
Fig. 4
Fig. 4 Purity (red and right vertical axis) and discord (blue and left vertical axis) versus the optical path length difference for mutually incoherent case. The circles and error bars denote the experimental average values and standard deviations, respectively. The solid lines are the Gaussian fittings for the data.
Fig. 5
Fig. 5 The proportion of the erroneous coincidence detections due to multi-photon states.

Tables (1)

Tables Icon

Table 1 Characteristics of the theoretical and experimental two-qubit density matrices. ρmixed is the two-qubit state when the optical path length difference between two mutually incoherent optical pulses are larger than the coherence length of the optical pulses.

Equations (23)

Equations on this page are rendered with MathJax. Learn more.

I ( A : B ) = H ( A ) + H ( B ) H ( A , B ) ,
J ( A : B ) = H ( A ) H ( A | B ) .
D ( ρ A B ) { Π j B } = I ( ρ A B ) J ( ρ A B ) { Π j B }
D ( ρ A B ) = I ( ρ A B ) max { Π j B } J ( ρ A B ) { Π j B } .
ρ c o h ( ϕ ) = 1 4 ( 1 i e i ϕ i e i ϕ i e 2 i ϕ i e i ϕ 1 1 i e i ϕ i e i ϕ 1 1 i e i ϕ i e 2 i ϕ i e i ϕ i e i ϕ 1 )
ρ i n c o h = 1 4 ( 1 0 0 0 0 1 1 0 0 1 1 0 0 0 0 1 )
ρ c o h | c ( ϕ ) = 1 2 ( 1 i e i ϕ i e i ϕ 1 ) ,
ρ c o h | d ( ϕ ) = 1 2 ( 1 i e i ϕ i e i ϕ 1 ) ,
ρ i n c o h | c = ρ i n c o h | d = 1 2 ( 1 0 0 1 ) ,
( a b ) 1 2 ( 1 i i 1 ) ( c d )
| α = e | α | 2 / 2 m = 0 α m m ! ( a H ) m | 0
| β = e | β | 2 / 2 n = 0 β n n ! ( e i ϕ b V ) n | 0
| ψ c o h i n = e μ m , n = 0 μ ( m + n ) / 2 m ! n ! ( a H ) m ( e i ϕ b V ) n | 0 | 0 .
| ψ c o h o u t = e μ m , n = 0 ( μ / 2 ) ( m + n ) / 2 m ! n ! ( c H + i d H ) m ( e i ϕ ( i c V + d V ) ) n | 0 | 0 .
| ψ c o h = 1 2 ( i c H d H + e i ϕ c H d V e i ϕ c V d H + i e i 2 ϕ c V d V ) | 0 | 0 .
| ψ ( 1 , 1 ) = 1 2 ( i c H c V + c H d V c V d H + i d H d V ) | 0 | 0 .
ρ i n c o h ( 1 , 1 ) = 1 2 ( 0 0 0 0 0 1 1 0 0 1 1 0 0 0 0 0 ) ,
| ψ ( 2 , 0 ) = 1 2 2 ( ( c H ) 2 + 2 i c H d H ( d H ) 2 ) | 0 | 0 .
ρ i n c o h ( 2 , 0 ) = ( 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 ) .
| ψ ( 0 , 2 ) = 1 2 2 ( ( c V ) 2 + 2 i c V d V + ( d V ) 2 ) | 0 | 0 .
ρ i n c o h ( 0 , 2 ) = ( 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 ) .
ρ i n c o h = p ( 1 , 1 ) ρ i n c o h ( 1 , 1 ) + p ( 2 , 0 ) ρ i n c o h ( 2 , 0 ) + p ( 0 , 2 ) ρ i n c o h ( 0 , 2 ) Tr [ p ( 1 , 1 ) ρ i n c o h ( 1 , 1 ) + p ( 2 , 0 ) ρ i n c o h ( 2 , 0 ) + p ( 0 , 2 ) ρ i n c o h ( 0 , 2 ) ]
E ( μ ) = i + j 3 p ( μ , i , j ) i + j 2 p ( μ , i , j )
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.