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

Experimentally obtaining maximal coherence via assisted distillation process

Open Access Open Access

Abstract

Quantum coherence, which quantifies the superposition properties of a quantum state, plays an indispensable role in quantum resource theory. A recent theoretical work [Phys. Rev. Lett. 116, 070402 (2016) [CrossRef]  ] studied the manipulation of quantum coherence in bipartite or multipartite systems under the protocol local incoherent operation and classical communication (LQICC). Here we present what we believe is the first experimental realization of obtaining maximal coherence in the assisted distillation protocol based on a linear optical system. The results of our work show that the optimal distillable coherence rate can be reached even in one-copy scenario when the overall bipartite qubit state is pure. Moreover, the experiments for mixed states showed that distillable coherence can be increased with less demand than entanglement distillation. Our work might be helpful in remote quantum information processing and quantum control.

© 2017 Optical Society of America

1. INTRODUCTION

Quantum coherence, which exhibits the fundamental signature of superposition in quantum mechanics, has been exploited in many fields of quantum physics, such as biological systems [13], transport theory [4,5], thermodynamics [612], nanoscale physics [13], and other scientific work associated with quantum theory [1421]. Recently, a rigorous and unambiguous framework for quantifying quantum coherence has also been put forward, which has enhanced the exploitation of its operational significance in the context of quantum resource theory [2225].

Quantifying quantum coherence starts from the definition of incoherent states (free states) and incoherent operations (free operations) [22,23,25,26]. A quantum state ρ is incoherent if it is diagonal in a given reference basis {|i}, i.e., ρ=ipi|ii|, with some probability {pi} [23,27]. Incoherent operators are required to fulfill K^nIK^nI for all n represented by the set of Kraus operators {K^n}, where I is the set of incoherent states. Moreover, in a d-dimensional Hilbert space H, the maximally coherent state is |ϕd=1/di|i, and |ϕ|ϕ2 denotes the unit coherence resource state [22].

As a quantification and measure of quantum superposition in a fixed basis {|i}, various coherence measures have been proposed [22,24,27,28]. In this paper, we choose relative entropy of coherence [22] as the measure of this property. Relative entropy of coherence of a quantum state ρ is given by Cr(ρ)=S(Δ(ρ))S(ρ), where Δ(ρ)=Σi|ii|ρ|ii| is the dephasing in the reference basis. In many recent works, this kind of measure has been endowed with special significance as well as operational meaning [2224]. Winter and Yang [23] showed that asymptotically the standard distillable coherence of a general quantum state is given by the relative entropy of quantum coherence. Likewise, Yuan et al. [24] has shown that the intrinsic randomness as a measure of coherence is just equal to the coherence of formation.

Recently, the manipulation and conversion of quantum coherence in bipartite or multipartite systems has been a hot topic [2931]. Ma et al. [30] studied the interconversion of coherence and correlation between two parties under incoherent operations. The manipulation of coherence in a bipartite system was discussed by Chitambar et al. [31] and Streltsov et al. [32] when performing assisted distillation of quantum coherence. The task is considered with protocols where Alice can perform arbitrary operations and Bob is restricted to only incoherent operations while classical communication between Alice and Bob is allowed, which is referred as local quantum-incoherent operations and classical communication (LQICC). With this LQICC protocol, coherence can not be generated on Bob’s subsystem without classical communication from Alice, thus the theoretical framework forms a new kind of resource theory [26] in which the distillable coherence on one subsystem is demanded and manipulated by the set of free operations LQICC. Also, the key conclusion is that for pure bipartite states ρAB, the maximal distillable coherence after assisted distillation on Bob’s subsystem is equal to the von Neumann entropy of the reduced density matrix of Bob after it is dephased in the incoherent basis, i.e., S(Δ(ρB)).

Quantum coherence is a highly demanded resource in many physical systems, especially in multipartite systems; however, to date there is no relevant experimental implementation of assisted distillation of quantum coherence. To fill this gap, we experimentally implement a class of such LQICC protocols based on a linear optic system in our laboratory, for extracting the maximal distillable coherence on one subsystem in our single-shot photonic scheme. Our results show that even in the presence of experimental imperfections and limitations, a maximal increase in distillable coherence can be observed even in the single-copy scenario for pure states. Furthermore, we also reported the experimental study with a specific class of mixed states, which showed that distillable coherence can be increased with less demand than entanglement distillation.

2. THEORY

A. Resource Theory of Assisted Distillation of Quantum Coherence

In the context of quantum resource theory, the optimal rate of generating resources (distillation rate) from a quantum state is always bounded as the manipulation and transformation are under certain restrictions [26]. For instance, distillable coherence of a general quantum state ρ under incoherent operations is equal to the relative entropy of coherence [23].

The framework considered in the present task involves a bipartite system (denoted as Alice and Bob) in which coherence on Bob’s side is viewed as a resource, and the set of free operations is restricted to LQICC. Under this constraint, a resource can never be generated from any quantum-incoherent state [31,32] (free states), which has the form of ΣipiσiA|ii|B, where {|iB} is the incoherent basis with respect to Bob. Our goal is to maximize distillable coherence on Bob’s side with the assistance of Alice under such constraint. To quantify the optimal rate of preparing |ϕ that can be possibly obtained, distillable coherence of collaboration [31,32] is defined as

CdA|B(ρ)=sup{R:limn(infΛΛ(ρn)ϕRn)=0},
where the infimum is taken over all LQICC operations Λ and x returns the maximum integer no larger than x.

Similarly, the bound of distillable coherence of collaboration CdA|B(ρ) is given by the quantum-incoherent relative entropy [31,32], i.e.,

CdA|B(ρAB)CrA|B(ρAB),
which can be evaluated directly as CrA|B(ρAB)=S(ΔB(ρAB))S(ρAB), where S(·) denote the von Neumann entropy and ΔB is the dephasing in the reference incoherent basis with respect to Bob.

However, for a general bipartite resource state ρAB, whether the inequality in Eq. (2) can be an equality is still unknown. As shown in the relevant theoretical work, at least for pure states, the equal sign is affirmative [31,32].

B. Linking One-Copy Scenario to Asymptotic Settings

The task elaborated in the context of resource theory challenges one greatly even for a pure resource state as the optimal rate may only be obtained by working in the asymptotic settings, which involves complicated collective measurements on many copies of the shared resource states and hinders its operational significance as well as its physical implementation. Thus, linking the one-copy scenario to asymptotic settings is essential for experimentally obtaining the maximal increase in coherence.

Suppose that Alice and Bob share a pure resource state |ΨAB, through local measurement MA on Alice together with broadcast of the measurement outcomes, any possible pure decomposition of ρB can be realized, i.e., ρB=Σipi|ΨiΨi| for any set of {pi} and pure state {|Ψi}. It is obvious that Alice can help Bob get an average distillable coherence of CdA(ρB)=ΣipiCr(|Ψi) in a one-copy scenario (Note that Cd(ρ)=Cr(ρ)), which is beyond the original distillable coherence Cd(ρB) on Bob’s side, as the relative entropy of coherence never increases under the mixing of quantum states [22].

To quantify the best performance that Alice can achieve to help Bob get the maximal distillable coherence when she is restricted to just local measurement on her system and classical broadcast processing, we use the coherence of assistance (COA) [31] defined as

Ca(ρ)=maxipiS(Δ(Ψi)).
As shown in [31], for pure states, the coherence of collaboration CdA|B(|ΨAB) is equal to the regularized COA defined as Ca(ρ)=limn1nCa(ρn)=S(Δ(ρ)) due to a higher distillation rate can be achieved through joint measurement performed on many copies of |ΨAB. The situation becomes more interesting as COA has additivity for a qubit state ρ, i.e., Ca(ρ)=Ca(ρ) [31]. Thus, obtaining the maximal increase in distillable coherence on Bob’s side involves just the same local measurement applied for every copy of a purified resource state combined with classical communication. Hence, optimal LQICC only needs to be performed on every copy of the resource states instead of complex collective operations on many copies. This makes it possible to observe the maximal increase in distillable coherence in single-shot photonic experiments. Actually, the optimal LQICC to reach coherence of assistance is constructed in the Methods section below.

3. EXPERIMENTAL SETUP

The whole experimental setup is shown in Fig. 1 (details are provided in the Methods section below). It involves two modules–a state preparation module and an LQICC module. In the state preparation module, the pure entangled state |ΨAB=cos2θ|HH+sin2θ|VV can be generated with an arbitrary θ ranging from 0° to 45°. The optical arrangement in Bob’s path (see the state preparation module in Fig. 1) is used to generate Werner states as ρAB=p|ΨΨ|+(1p)I4 with an arbitrary p. The LQICC module includes an assisted distillation part and a tomography [33] part. In the assisted distillation part, we implement the optimal local measurement on every copy of Alice’s state. On Bob’s side, however, the operation is restricted to incoherent operations. Finally, tomography is used to evaluate the final distillable coherence of Bob’s state. For pure states, the distillable coherence of collaboration CdA|B(|ΨAB) can be evaluated directly by final distillable coherence of Bob. For Werner states, we replaced CdA|B(ρAB) with Cd(ρ1B) because we didn’t know whether the distillable coherence of collaboration could be reached in single-shot experiments.

 figure: Fig. 1.

Fig. 1. Experimental setup has two modules: state preparation and LQICC. In the state preparation module, two classes of pure states |ΨAB=cos2θ|HH+sin2θ|VV and |ΨAB=12(cos2θ|HH+cos2θ|HV+sin2θ|VHsin2θ|VV), and arbitrary Werner states in the form ρAB=p|ΨΨ|+(1p)I4 can be generated. After preparation of the desired two-photon resource states, the two photons are distributed to Alice and Bob. In the LQICC module, optimal assisted operations are then performed by Alice on her photon and the measurement results are sent to Bob via a classical communication channel. According to the classical message from Alice, desired corresponding incoherent operations are performed on Bob’s photon. After the assisted distillation protocol, quantum state tomography is used to characterize the final state of Bob’s photon to identify the final distillable coherence. Key to components: HWP, half-wave plate; QWP, quarter-wave plate; BS, beam splitter; IF, interference filter; SPD, single photon detector; and PBS, polarizing beam splitter.

Download Full Size | PDF

4. EXPERIMENTAL RESULTS

In the experimental test, we divided the procedure into two parts, respectively, testing pure and mixed states in an assisted distillation process.

In the first part of the experiments, we generated two classes of pure states.

The first class of pure states was prepared as |ΨAB=cos2θ|HH+sin2θ|VV, the rotation angle θ of HWP1 was set from 0° to 45°. Without assisted distillation, the state of Bob’s photon was ρB=cos22θ|HH|+sin22θ|VV|, whose distillable coherence was theoretically evaluated as Cd(ρB)=0. In the experiment, tomography was used for further calculation of distillable coherence on the state of Bob’s photon [see red upward-pointing triangles in Fig. 2(a)]. In the assisted distillation protocol, Alice’s optimal operation was to perform von Neumann measurements in the basis |η±A, which only need to be mutually unbiased with |H and |V for the first class of pure states (details in the Methods section below). Here our experiment chose |η±A=|y±A=12(|H±i|V). When Alice measured |y+, the state of Bob’s photon collapsed to cos2θ|Hisin2θ|V with a distillable coherence of S(Δ(ρB))=cos22θlog2cos22θsin22θlog2sin22θ. When |y was measured by Alice, the state of Bob’s photon collapsed to cos2θ|H+isin2θ|V with the same amount of distillable coherence S(Δ(ρB)). After receiving Alice’s measurement result (classical information) via classical communication channel, we just need to divide Bob’s photons into two ensembles for future applications according to Alice’s two measurement results. Since the photon in either ensemble had a distillable coherence of S(Δ(ρB)), the final distillable coherence on Bob’s side after assisted distillation is S(Δ(ρB)) per photon. The distillable coherence of collaboration was evaluated using tomography as the final distillable coherence on Bob’s side and denoted as CdA|B(|ΨAB) according to the two measurement results [see pink downward-pointing triangles in Fig. 2(a)]. The increase in distillable coherence can be obtained from δCd(ρB)=CdA|B(|ΨAB)Cd(ρB) and equals S(ρB) theoretically.

 figure: Fig. 2.

Fig. 2. Experimental results for the pure states. As shown in Fig. 1, Cd(ρB) (red upward-pointing triangles) represent the distillable coherence of Bob without distillation, CdA|B(|ΨAB) (pink downward-pointing triangles) represent the distillable coherence of Bob after assisted distillation, δCd(ρB) (blue squares) represent an increase in coherence, and black dashed-dotted lines represent the theoretical curve. In Fig. 1, the resource state has the form of |ΨAB=cos2θ|HH+sin2θ|VV and in Fig. 3 below, |ΨAB=12(cos2θ|HH+cos2θ|HV+sin2θ|VHsin2θ|VV).

Download Full Size | PDF

The second class of pure states for assisted distillation was prepared as |ΨAB=12(cos2θ|HH+cos2θ|HV+sin2θ|VHsin2θ|VV). Similarly, without assisted distillation, the state of Bob’s photon is ρB=cos22θ|x+x+|+sin22θ|xx| with |x±=12(|H±|V), which had a distillable coherence of Cd(ρB)=12[(1+cos4θ)log2(1+cos4θ)+(1cos4θ)log2(1cos4θ)] [black dash-dotted line in Fig. 2(b)]. The experimental results [red upward-pointing triangles in Fig. 2(b)] were also obtained from quantum state tomography and agreed well with the theoretical predictions. In the assisted protocol, Alice’s optimal operation was to perform von Neumann measurements mutually unbiased with cos2θ|H+sin2θ|V and cos2θ|Hsin2θ|V for the second class of pure states, and we also chose |η±A=|y±A. The state of Bob’s photon collapsed to 12(cos2θisin2θ)|H+12(cos2θ±isin2θ)|V according to Alice’s two measurement results |y± and the distillable coherence of collaboration CdA|B(|ΨAB)=1 theoretically. We used quantum state tomography to obtain experimental results of CdA|B(|ΨAB) (pink downward-pointing triangles) and δCd(ρB) (blue squares).

In the second part of the experiments, the task of assisted distillation of coherence for mixed states is much more complicated because it is still an open problem that needs answers to two questions: Is it possible to achieve the upper bound of the inequality in Eq. (2) and what are the optimal LQICC operations. For theoretical and experimental simplicity, the optimal LQICC operation was considered for assisted distillation in one-copy instead of n-copy scenarios. In this part, Werner states were prepared as ρAB=p|ΨΨ|+(1p)I4 with p ranging from 0.05 to 0.95 by the adjustment of apertures. We denoted ρ0B and ρ1B as the reduced density operator for Bob before and after the assistance of Alice. Without Alice’s assistance, Bob’s state was a maximally mixed state, i.e., ρ0B=I2, whose distillable coherence was Cd(ρ0B)=0 (see the upward-pointing triangles).

First we found out optimal von Neumann measurements in the single-shot experiment of assisted coherence distillation, which turned out to be a maximally coherent basis |η±=12(|H±eiφ|V) with arbitrary φ for Werner states. For simplicity we chose |y±. After measurements along |y± were performed on Alice’s photons, Bob’s state collapsed to ρ1B=p|yy|+(1p)I2 according to Alice’s two measurement results, respectively. Then Bob’s distillable coherence after Alice’s assistance theoretically was Cd(ρ1B)=12[(1+p)log2(1+p)+(1p)log2(1p)] (black dash-dotted line in Fig. 3). The experimental results for Cd(ρ1B), denoted as pink downward-pointing triangles, are shown in Fig. 3, which agreed well with the theoretical curve. The increase in distillable coherence δCd(ρB) can be easily obtained from δCd(ρB)=Cd(ρ1B)Cd(ρ0B) (relevant experimental results denoted as blue squares in Fig. 3). We also calculated the upper bound QI relative entropy in Eq. (2) as CrA|B(ρAB)=14[(1p)log2(1p)+(1+3p)log2(1+3p)2(1+p)log2(1+p)] and the theoretical curve is shown in Fig. 3 as a brown dash-dotted line.

 figure: Fig. 3.

Fig. 3. Experimental results for Werner states [the colored symbols have the same meaning as shown in Fig. 2, except the brown dash-dotted line represents the theoretical calculated upper bound in inequality in Eq. (2)]. The resource state has the form ρAB=p|ΨΨ|+(1p)I4.

Download Full Size | PDF

As shown in Figs. 2(a) and 2(b), for the two classes of pure states, both experimental results of CdA|B(|ΨAB) and δCd(ρB) are slightly smaller than the theoretical prediction around θ=22.5° due to difficulty in preparing highly entangled states.

For Werner state, experimental results in Fig. 3 show that the increased coherence on Bob’s side in single-shot experiments agreed well with the numerical simulation and was close to the upper bound CrA|B(ρAB). However, as it is obvious that CdA|B(ρAB)Cd(ρB), whether the upper bound can be reached in the asymptotic limit is still unknown. What’s interesting is that the distillable coherence on Bob’s side after assisted distillation is more than zero as long as p>0, i.e., CdA|B(ρAB)Cd(ρB)>0 if and only if p>0, which is much less demanding than the entanglement requirement with p>1/3 [34]. Thus the distillable coherence of collaboration CdA|B(ρAB) is nonzero if and only if the state ρAB is not quantum-incoherent in the reference basis. Hence, the coherence on Bob’s side can still be increased with Alice’s assistance even if the shared resource state ρAB is not entangled.

5. CONCLUSION

We have reported what we believe is the first experimental study of assisted distillation of quantum coherence for both pure and mixed two-qubit states. In experiments for pure states, optimal assisted coherence distillation is achieved and the increased distillable coherence on Bob’s subsystem is equal to the von Neumann entropy of the reduced density matrix of Bob after it is dephased in the incoherent basis, i.e., S(Δ(ρB)). For mixed states, optimal assisted coherence distillation is considered in a one-copy scenario and the maximal distillable coherence on Bob’s side after assistance is close to the upper bound CrA|B(ρAB).

The task of assisted distillation of quantum coherence can be applied in many scenarios. Quantum coherence is a significant property of superposition that is highly demanded in quantum information processing and quantum computation, and one needs to distill as many copies of a pure maximally coherent state as possible for better performance in these tasks. However, when considering a system where coherence is demanded, and is remote or inaccessible, the theoretical framework gives an optimal solution to this problem. As the results of our first experiment showed, one can extract maximal coherence from a bipartite resource state in an assisted distillation process based on linear optical systems. Thus, the procedure can be physically implemented in most related tasks of the quantum information process with sufficiently high fidelity. Furthermore, the test for Werner states linked the quantum coherence to quantum correlation, which is beyond entanglement, i.e., for some special form of mixed states, one can always extract distillable coherence on one subsystem as long as correlation exists. This result sheds light on the strong relation between quantum coherence and nonclassical correlation.

6. METHODS

A. Selection of Measurement Basis

In this section, we will show how to select the optimal measurement basis when the shared state is a pure two-qubit state ρAB.

According to theoretical analysis [31], if we expand the purification |ΨAB in the incoherent basis,

|ΨAB=k=11ck|ΨkA|kB,
then orthogonal states |η± always exist that are mutually unbiased with respect to the two states |ΨkA. When Alice performs a von Neumann measurement in the |η± basis, Bob will be in one of two post-measurement states |ϕ±B associated with the outcome of the measurement on Alice; in both cases, the state has coherence Cr(|ϕ±B)=S(Δ(ρB)) and with Ca(ρ)=S(Δ(ρ)), we can find [31]
Ca(ρ)=Ca(ρ)=S(Δ(ρ)).
For pure states, the coherence of assistance is equal to the coherence of collaboration in the asymptotic setting, CdA|B(|ΨAB)=Ca(ρB). Moreover, when Bob’s system is a qubit and the overall state is pure, the coherence of assistance and the coherence of collaboration are equivalent even in the single-copy case, Ca(ρ)=Ca(ρ).

Hence we can use the theoretical analysis to maximize the distillable coherence on Bob’s system. In the case where the shared state is a pure two-qubit state, when we perform von Neumann measurement on Alice in the mutually unbiased basis, we can obtain the maximal increase in distillable coherence in Bob’s system. In our experiments, when we expand the two pure states in incoherent basis, we obtain the sets of states |ΨkA are |H, |V for state |ΨAB=cos2θ|HH+sin2θ|VV and cos2θ|H±sin2θ|V for state |ΨAB=12(cos2θ|HH+cos2θ|HV+sin2θ|VHsin2θ|VV). Because there obviously exists a shared mutually unbiased basis |y± with respect to both cases, we can apply von Neumann measurement on Alice along the |y± basis in the first part of our experiments.

In the experiments for Werner states, we considered the task in a one-copy scenario because the complex collective measurement on many copies of a state is hard to implement because of experimental limitations. Because the two-qubit Werner states have a form of ρAB=p|ΨΨ|+(1p)I4 with high symmetry, when Alice measures her system in the basis |η±=12(|H±eiφ|V) with arbitrary φ, Bob’s state will collapse to ρ1B=p|ηη|+(1p)I2, and we chose the same |y± as that for pure states.

B. Detail Experimental Apparatus

We now move to the detailed experimental arrangement in our laboratory based on the polarization-entangled photon pairs.

In state preparation module in Fig. 1, two type I phase-matched β-barium borate (BBO) crystals, whose optic axes are normal to each other [35], are pumped by the continuous Ar+ laser at 351.1 nm with power up to 60 mW for the generation of photon pairs with a central wavelength at λ=702.2nm via spontaneous parametric down conversion process (SPDC). HWP1 working at λ=351.1nm is set in front of the BBO crystals to control the quantum state of the generated photon pairs encoded in polarization. Half-wave plates at both ends of the two single-mode fibers (SMF) are used to control polarization. In one branch, QWP3 is tilted to compensate the phase of the two-photon state for the generation of required states. This setup is capable of preparing two classes of pure states |ΨAB=cos2θ|HH+sin2θ|VV and |ΨAB=12(cos2θ|HH+cos2θ|HV+sin2θ|VHsin2θ|VV), where θ is the rotation angle of HWP1.

As for the generation of Werner states, two 50/50 beam splitters (BS) are inserted into one branch. In the transmission path, the two-photon state is prepared as the singlet state |Ψ=12(|HV|VH) when the rotation angle of HWP1 is set as θ=22.5°. In the reflected path, three 446λ quartz crystals and a half-wave plate with 22.5° are used to dephase the two-photon state into a completely mixed state I4. The ratio of the two states mixed at the output port of the second BS can be changed by the two adjustable apertures for the generation of arbitrary Werner states with ρAB=p|ΨΨ|+(1p)I4. Out of the state preparation module, the two photons are distributed to Alice and Bob, as shown in Fig. 1.

In the LQICC module, on Alice’s side, QWP1, HWP3, PBS, and two single photon detectors are used to perform arbitrary assisted projective measurements and the measurement results are sent to Bob via a classical communication channel. In the next stage, QWP2, HWP4, PBS, and two single photon detectors are used to perform tomography on Bob’s photon. Actually, we used a beam displacer (BD) together with a 45° holophote, which can act as a PBS and has a rather high extinction ratio.

Funding

National Natural Science Foundation of China (NSFC) (11574291); National Key R&D Program (2016YFA0301700).

Acknowledgment

We thank Jayne Thompson, Mile Gu, and Vlatko Vedral for fruitful discussions.

REFERENCES

1. S. F. Huelga and M. B. Plenio, “Vibrations, quanta and biology,” Contemp. Phys. 54, 181–207 (2013). [CrossRef]  

2. S. Lloyd, “Quantum coherence in biological systems,” J. Phys. 302, 012037 (2011). [CrossRef]  

3. M. B. Plenio and S. F. Huelga, “Dephasing-assisted transport: quantum networks and biomolecules,” New J. Phys. 10, 113019 (2008). [CrossRef]  

4. M. Herranen, K. Kainulainen, and P. M. Rahkila, “Kinetic transport theory with quantum coherence,” Nucl. Phys. A 820, 203c–206c (2009). [CrossRef]  

5. P. Rebentrost, M. Mohseni, and A. Aspuru-Guzik, “Role of quantum coherence and environmental fluctuations in chromophoric energy transport,” J. Phys. Chem. B 113, 9942–9947 (2009). [CrossRef]  

6. J. Åberg, “Catalytic coherence,” Phys. Rev. Lett. 113, 150402 (2014). [CrossRef]  

7. M. Lostaglio, K. Korzekwa, D. Jennings, and T. Rudolph, “Quantum coherence, time-translation symmetry, and thermodynamics,” Phys. Rev. X 5, 021001 (2015). [CrossRef]  

8. K. Korzekwa, M. Lostaglio, J. Oppenheim, and D. Jennings, “The extraction of work from quantum coherence,” New J. Phys. 18, 023045 (2016). [CrossRef]  

9. P. Ćwikliński, M. Studziński, M. Horodecki, and J. Oppenheim, “Limitations on the evolution of quantum coherences: towards fully quantum second laws of thermodynamics,” Phys. Rev. Lett. 115, 210403 (2015). [CrossRef]  

10. V. Narasimhachar and G. Gour, “Low-temperature thermodynamics with quantum coherence,” Nat. Commun. 6, 7689 (2015). [CrossRef]  

11. C. A. Rodríguez-Rosario, T. Frauenheim, and A. Aspuru-Guzik, “Thermodynamics of quantum coherence,” arXiv:1308.1245 (2013).

12. B. Gardas and S. Deffner, “Thermodynamic universality of quantum Carnot engines,” Phys. Rev. E 92, 042126 (2015). [CrossRef]  

13. O. Karlström, H. Linke, G. Karlström, and A. Wacker, “Increasing thermoelectric performance using coherent transport,” Phys. Rev. B 84, 113415 (2011). [CrossRef]  

14. J.-J. Chen, J. Cui, and H. Fan, “Coherence susceptibility as a probe of quantum phase transitions,” arXiv:1509.03576 (2015).

15. S. Cheng and M. J. W. Hall, “Complementarity relations for quantum coherence,” Phys. Rev. A 92, 042101 (2015). [CrossRef]  

16. X. Hu and H. Fan, “Coherence extraction from measurement-induced disturbance,” arXiv:1508.01978 (2015).

17. I. Marvian, R. W. Spekkens, and P. Zanardi, “Quantum speed limits, coherence, and asymmetry,” Phys. Rev. A 93, 052331 (2016). [CrossRef]  

18. Z. Bai and S. Du, “Maximally coherent states,” arXiv:1503.07103 (2015).

19. X. Liu, Z. Tian, J. Wang, and J. Jing, “Protecting quantum coherence of two-level atoms from vacuum fluctuations of electromagnetic field,” Ann. Phys. 366, 102–112 (2016). [CrossRef]  

20. S. Du, Z. Bai, and Y. Guo, “Conditions for coherence transformations under incoherent operations,” Phys. Rev. A 91, 052120 (2015). [CrossRef]  

21. M. N. Bera, T. Qureshi, M. A. Siddiqui, and A. K. Pati, “Duality of quantum coherence and path distinguishability,” Phys. Rev. A 92, 012118 (2015). [CrossRef]  

22. T. Baumgratz, M. Cramer, and M. B. Plenio, “Quantifying coherence,” Phys. Rev. Lett. 113, 140401 (2014). [CrossRef]  

23. A. Winter and D. Yang, “Operational resource theory of coherence,” Phys. Rev. Lett. 116, 120404 (2016). [CrossRef]  

24. X. Yuan, H. Zhou, Z. Cao, and X. Ma, “Intrinsic randomness as a measure of quantum coherence,” Phys. Rev. A 92, 022124 (2015). [CrossRef]  

25. A. Streltsov, G. Adesso, and M. B. Plenio, “Quantum coherence as a resource,” arXiv preprint arXiv:1609.02439 (2016).

26. F. G. S. L. Brandão and G. Gour, “Reversible framework for quantum resource theories,” Phys. Rev. Lett. 115, 070503 (2015). [CrossRef]  

27. A. Streltsov, U. Singh, H. S. Dhar, M. N. Bera, and G. Adesso, “Measuring quantum coherence with entanglement,” Phys. Rev. Lett. 115, 020403 (2015). [CrossRef]  

28. D. Girolami, “Observable measure of quantum coherence in finite dimensional systems,” Phys. Rev. Lett. 113, 170401 (2014). [CrossRef]  

29. S. Du, Z. Bai, and X. Qi, “Coherence measures and optimal conversion for coherent states,” Quantum Inf. Comput. 15, 1307–1316 (2015).

30. J. Ma, B. Yadin, D. Girolami, V. Vedral, and M. Gu, “Converting coherence to quantum correlations,” Phys. Rev. Lett. 116, 160407 (2016). [CrossRef]  

31. E. Chitambar, A. Streltsov, S. Rana, M. N. Bera, G. Adesso, and M. Lewenstein, “Assisted distillation of quantum coherence,” Phys. Rev. Lett. 116, 070402 (2016). [CrossRef]  

32. A. Streltsov, S. Rana, M. N. Bera, and M. Lewenstein, “Towards resource theory of coherence in distributed scenarios,” Phys. Rev. X 7, 011024 (2017). [CrossRef]  

33. B. Qi, Z. Hou, L. Li, D. Dong, G. Xiang, and G. Guo, “Quantum state tomography via linear regression estimation,” Sci. Rep. 3, 3496 (2013).

34. A. Miranowicz, “Violation of Bell inequality and entanglement of decaying Werner states,” Phys. Lett. A 327, 272–283 (2004). [CrossRef]  

35. P. G. Kwiat, E. Waks, A. G. White, I. Appelbaum, and P. H. Eberhard, “Ultrabright source of polarization-entangled photons,” Phys. Rev. A 60, R773(R) (1999). [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 (3)

Fig. 1.
Fig. 1. Experimental setup has two modules: state preparation and LQICC. In the state preparation module, two classes of pure states | Ψ A B = cos 2 θ | H H + sin 2 θ | V V and | Ψ A B = 1 2 ( cos 2 θ | H H + cos 2 θ | H V + sin 2 θ | V H sin 2 θ | V V ) , and arbitrary Werner states in the form ρ A B = p | Ψ Ψ | + ( 1 p ) I 4 can be generated. After preparation of the desired two-photon resource states, the two photons are distributed to Alice and Bob. In the LQICC module, optimal assisted operations are then performed by Alice on her photon and the measurement results are sent to Bob via a classical communication channel. According to the classical message from Alice, desired corresponding incoherent operations are performed on Bob’s photon. After the assisted distillation protocol, quantum state tomography is used to characterize the final state of Bob’s photon to identify the final distillable coherence. Key to components: HWP, half-wave plate; QWP, quarter-wave plate; BS, beam splitter; IF, interference filter; SPD, single photon detector; and PBS, polarizing beam splitter.
Fig. 2.
Fig. 2. Experimental results for the pure states. As shown in Fig. 1, C d ( ρ B ) (red upward-pointing triangles) represent the distillable coherence of Bob without distillation, C d A | B ( | Ψ A B ) (pink downward-pointing triangles) represent the distillable coherence of Bob after assisted distillation, δ C d ( ρ B ) (blue squares) represent an increase in coherence, and black dashed-dotted lines represent the theoretical curve. In Fig. 1, the resource state has the form of | Ψ A B = cos 2 θ | H H + sin 2 θ | V V and in Fig. 3 below, | Ψ A B = 1 2 ( cos 2 θ | H H + cos 2 θ | H V + sin 2 θ | V H sin 2 θ | V V ) .
Fig. 3.
Fig. 3. Experimental results for Werner states [the colored symbols have the same meaning as shown in Fig. 2, except the brown dash-dotted line represents the theoretical calculated upper bound in inequality in Eq. (2)]. The resource state has the form ρ A B = p | Ψ Ψ | + ( 1 p ) I 4 .

Equations (5)

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

C d A | B ( ρ ) = sup { R : lim n ( inf Λ Λ ( ρ n ) ϕ R n ) = 0 } ,
C d A | B ( ρ A B ) C r A | B ( ρ A B ) ,
C a ( ρ ) = max i p i S ( Δ ( Ψ i ) ) .
| Ψ A B = k = 1 1 c k | Ψ k A | k B ,
C a ( ρ ) = C a ( ρ ) = S ( Δ ( ρ ) ) .
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.