13
LETTERS Experimental demonstration of a BDCZ quantum repeater node Zhen-Sheng Yuan 1,2 *, Yu-Ao Chen 1,2 *, Bo Zhao 1 , Shuai Chen 1 , Jo ¨rg Schmiedmayer 3 & Jian-Wei Pan 1,2 Quantum communication is a method that offers efficient and secure ways for the exchange of information in a network. Large-scale quantum communication 1–4 (of the order of 100 km) has been achieved; however, serious problems occur beyond this distance scale, mainly due to inevitable photon loss in the trans- mission channel. Quantum communication eventually fails 5 when the probability of a dark count in the photon detectors becomes comparable to the probability that a photon is correctly detected. To overcome this problem, Briegel, Du ¨ r, Cirac and Zoller (BDCZ) introduced the concept of quantum repeaters 6 , combining entan- glement swapping 7 and quantum memory to efficiently extend the achievable distances. Although entanglement swapping has been experimentally demonstrated 8 , the implementation of BDCZ quantum repeaters has proved challenging owing to the difficulty of integrating a quantum memory. Here we realize entanglement swapping with storage and retrieval of light, a building block of the BDCZ quantum repeater. We follow a scheme 9,10 that incorporates the strategy of BDCZ with atomic quantum memories 11 . Two atomic ensembles, each originally entangled with a single emitted photon, are projected into an entangled state by performing a joint Bell state measurement on the two single photons after they have passed through a 300-m fibre-based communication channel. The entanglement is stored in the atomic ensembles and later verified by converting the atomic excitations into photons. Our method is intrinsically phase insensitive and establishes the essential ele- ment needed to realize quantum repeaters with stationary atomic qubits as quantum memories and flying photonic qubits as quantum messengers. Although the BDCZ protocol 6 attracted much interest as a solution to extending the communication length, the absence of quantum memory has hindered the implementation of quantum repeaters. In 2001, Duan, Lukin, Cirac and Zoller (DLCZ) proposed an alterna- tive quantum repeater scheme 11 where linear optics and atomic ensembles are used to incorporate entanglement connection and quantum memory into a single unit. Motivated by the DLCZ pro- tocol, number-state entanglement between two atomic ensembles has been observed 12,13 . Most recently, a functional quantum node 14 based on asynchronous preparation of number-state entanglement for two pairs of atomic ensembles—the basic element of the DLCZ protocol—has also been demonstrated. However, two serious drawbacks make the DLCZ-type functional quantum nodes 11,14 unlikely to be a realistic solution for long-distance quantum communication 9,10,15 . First, the required long-term sub- wavelength stability of the path difference between two arms of a large-scale single-photon interferometer spanning the whole commu- nication distance is very difficult to achieve 9,10 , even with the latest and most sophisticated technology for coherent optical phase transfer 16 . Second, the swapping of number-state entanglement using a single- photon interferometer leads to the growth of a vacuum component in the generated state, and to the rapid growth of errors due to multiple emissions from individual ensembles 15 . A novel solution 9,10 is to combine the atomic quantum memory in DLCZ and the strategy of BDCZ. In this scheme, two-photon inter- ference is used to generate long-distance entanglement, so the stability requirement for the path differences is determined by the coherence length of the photons and is consequently seven orders of magnitude looser 17 than in the DLCZ protocol. Moreover, the vacuum component can be suppressed and is no longer a dominant term after a few entanglement connections 9,10 . Following this scheme, we demonstrate here the implementation of a quantum repeater node, involving entanglement swapping with the function of storage and retrieval of light. A high precision of local operations has been achieved that surpasses the theoretical threshold 6 required for the realization of robust quantum repeaters for long-distance quantum communication. In our experiment, to demonstrate entanglement swapping with storage and retrieval of light, we follow three steps: implementing two atom–photon entanglement sources, sending the flying qubits (the photons) to an intermediate station for a Bell state measurement (BSM), and verifying the entanglement between the stationary qubits (the two remote atomic ensembles). Unlike previous atom–photon entanglement sources realized with trapped ions 18 , single atoms in a cavity 19 , or two spatially separated atomic ensembles 20 , we use here two collective excitations in different spatial modes of a single atomic ensemble to implement the atom– photon entanglement 21 . In contrast to the method in which two separated spatial regions in one atomic cloud are covered by their own ‘write’ and ‘read’ beams 14 , here the two excitation modes share the same write and read beams, which offers high-quality entangle- ment and long-term stability. The basic principle is shown in Fig. 1 (see Methods). Alice and Bob each have a cold atomic ensemble consisting of about 10 8 atoms of 87 Rb loaded by magneto-optical traps (MOTs). At each site, atoms are first prepared in the initial state a ji, followed by a weak write pulse. Two anti-Stokes fields AS L and AS R induced by the write beam via spontaneous Raman scattering are collected at 63u relative to the propagating direction of the write beam. This defines two spatial modes of excitation in the atomic ensembles (L and R), which con- stitute our memory qubit. The two anti-Stokes fields in modes L and R are adjusted to have equal excitation probability and orthogonal polarizations. The two fields are then overlapped at a polarizing beam splitter PBS2 (Fig. 1) and coupled into a single-mode fibre. Neglecting the vacuum state and higher order excitations, the entangled state between the atomic 1 Physikalisches Institut, Ruprecht-Karls-Universita¨t Heidelberg, Philosophenweg 12, 69120 Heidelberg, Germany. 2 Hefei National Laboratory for Physical Sciences at Microscale and Department of Modern Physics, University of Science and Technology of China, Hefei, Anhui 230026, China. 3 Atominstitut der O ¨ sterreichischen Universita ¨ten, TU-Wien, A-1020 Vienna, Austria. *These authors contributed equally to this work. Vol 454 | 28 August 2008 | doi:10.1038/nature07241 1098 ©2008 Macmillan Publishers Limited. All rights reserved

Experimental demonstration of a BDCZ quantum repeater node

  • Upload
    tuwien

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

LETTERS

Experimental demonstration of a BDCZ quantumrepeater nodeZhen-Sheng Yuan1,2*, Yu-Ao Chen1,2*, Bo Zhao1, Shuai Chen1, Jorg Schmiedmayer3 & Jian-Wei Pan1,2

Quantum communication is a method that offers efficient andsecure ways for the exchange of information in a network.Large-scale quantum communication1–4 (of the order of 100 km)has been achieved; however, serious problems occur beyond thisdistance scale, mainly due to inevitable photon loss in the trans-mission channel. Quantum communication eventually fails5 whenthe probability of a dark count in the photon detectors becomescomparable to the probability that a photon is correctly detected.To overcome this problem, Briegel, Dur, Cirac and Zoller (BDCZ)introduced the concept of quantum repeaters6, combining entan-glement swapping7 and quantum memory to efficiently extend theachievable distances. Although entanglement swapping has beenexperimentally demonstrated8, the implementation of BDCZquantum repeaters has proved challenging owing to the difficultyof integrating a quantum memory. Here we realize entanglementswapping with storage and retrieval of light, a building block of theBDCZ quantum repeater. We follow a scheme9,10 that incorporatesthe strategy of BDCZ with atomic quantum memories11. Twoatomic ensembles, each originally entangled with a single emittedphoton, are projected into an entangled state by performing a jointBell state measurement on the two single photons after they havepassed through a 300-m fibre-based communication channel. Theentanglement is stored in the atomic ensembles and later verifiedby converting the atomic excitations into photons. Our method isintrinsically phase insensitive and establishes the essential ele-ment needed to realize quantum repeaters with stationary atomicqubits as quantum memories and flying photonic qubits asquantum messengers.

Although the BDCZ protocol6 attracted much interest as a solutionto extending the communication length, the absence of quantummemory has hindered the implementation of quantum repeaters.In 2001, Duan, Lukin, Cirac and Zoller (DLCZ) proposed an alterna-tive quantum repeater scheme11 where linear optics and atomicensembles are used to incorporate entanglement connection andquantum memory into a single unit. Motivated by the DLCZ pro-tocol, number-state entanglement between two atomic ensembleshas been observed12,13. Most recently, a functional quantum node14

based on asynchronous preparation of number-state entanglementfor two pairs of atomic ensembles—the basic element of the DLCZprotocol—has also been demonstrated.

However, two serious drawbacks make the DLCZ-type functionalquantum nodes11,14 unlikely to be a realistic solution for long-distancequantum communication9,10,15. First, the required long-term sub-wavelength stability of the path difference between two arms of alarge-scale single-photon interferometer spanning the whole commu-nication distance is very difficult to achieve9,10, even with the latest andmost sophisticated technology for coherent optical phase transfer16.

Second, the swapping of number-state entanglement using a single-photon interferometer leads to the growth of a vacuum component inthe generated state, and to the rapid growth of errors due to multipleemissions from individual ensembles15.

A novel solution9,10 is to combine the atomic quantum memory inDLCZ and the strategy of BDCZ. In this scheme, two-photon inter-ference is used to generate long-distance entanglement, so thestability requirement for the path differences is determined by thecoherence length of the photons and is consequently seven orders ofmagnitude looser17 than in the DLCZ protocol. Moreover, thevacuum component can be suppressed and is no longer a dominantterm after a few entanglement connections9,10. Following this scheme,we demonstrate here the implementation of a quantum repeaternode, involving entanglement swapping with the function of storageand retrieval of light. A high precision of local operations has beenachieved that surpasses the theoretical threshold6 required for therealization of robust quantum repeaters for long-distance quantumcommunication.

In our experiment, to demonstrate entanglement swapping withstorage and retrieval of light, we follow three steps: implementing twoatom–photon entanglement sources, sending the flying qubits (thephotons) to an intermediate station for a Bell state measurement(BSM), and verifying the entanglement between the stationary qubits(the two remote atomic ensembles).

Unlike previous atom–photon entanglement sources realized withtrapped ions18, single atoms in a cavity19, or two spatially separatedatomic ensembles20, we use here two collective excitations in differentspatial modes of a single atomic ensemble to implement the atom–photon entanglement21. In contrast to the method in which twoseparated spatial regions in one atomic cloud are covered by theirown ‘write’ and ‘read’ beams14, here the two excitation modes sharethe same write and read beams, which offers high-quality entangle-ment and long-term stability.

The basic principle is shown in Fig. 1 (see Methods). Alice and Bobeach have a cold atomic ensemble consisting of about 108 atoms of87Rb loaded by magneto-optical traps (MOTs). At each site, atomsare first prepared in the initial state aj i, followed by a weak writepulse. Two anti-Stokes fields ASL and ASR induced by the write beamvia spontaneous Raman scattering are collected at 63u relative to thepropagating direction of the write beam. This defines two spatialmodes of excitation in the atomic ensembles (L and R), which con-stitute our memory qubit.

The two anti-Stokes fields in modes L and R are adjusted to haveequal excitation probability and orthogonal polarizations. The twofields are then overlapped at a polarizing beam splitter PBS2 (Fig. 1)and coupled into a single-mode fibre. Neglecting the vacuum stateand higher order excitations, the entangled state between the atomic

1Physikalisches Institut, Ruprecht-Karls-Universitat Heidelberg, Philosophenweg 12, 69120 Heidelberg, Germany. 2Hefei National Laboratory for Physical Sciences at Microscale andDepartment of Modern Physics, University of Science and Technology of China, Hefei, Anhui 230026, China. 3Atominstitut der Osterreichischen Universitaten, TU-Wien, A-1020Vienna, Austria.*These authors contributed equally to this work.

Vol 454 | 28 August 2008 | doi:10.1038/nature07241

1098

©2008 Macmillan Publishers Limited. All rights reserved

and photonic qubits can be described effectively as

Yj iatph~1ffiffiffi2p Hj i Rj izeiw1 Vj i Lj i

ð1Þ

where Hj i= Vj i denotes horizontal/vertical polarization of the singleanti-Stokes photon, Lj i= Rj i denotes a single collective excitation inensemble L/R, and w1 is the propagating phase difference between thetwo anti-Stokes fields before they overlap at PBS2. The atom–photonentangled state (equation (1)) is equivalent to the maximally polar-ization-entangled state generated by spontaneous parametric down-conversion22.

In this way, one can implement two separate and remote atom–photon entanglement sources at Alice’s site (I) and Bob’s site (II),respectively. To make the higher order excitations negligible, a lowexcitation probability (xm < 0.01) is chosen for the collective modesm (m 5 L, R). Owing to the imperfect coupling of light modes, thetransmission loss, and the inefficiency of single-photon detectors, theoverall detection efficiency of an emerging anti-Stokes photon (gAS)is around 25%. To check the quality of atom–photon entanglement, aread pulse (see Methods) is applied after a controllable time delay dts

to convert the atomic collective excitation back into a Stokes field.Ideally, the retrieve efficiency of the Stokes fields should reach unity.However, various imperfections, such as low optical depth of theatomic ensembles and mode mismatching between the write andread pulses, lead to a 35% retrieve efficiency. Together with thenon-ideal collection and detection efficiency (,40%) of single-photon detectors, the overall detection efficiency of the Stokesphoton is around 15%. After combining the two retrieved Stokesfields on PBS1 (see Fig. 1), the anti-Stokes and Stokes fields are inthe following maximally polarization-entangled state

Yj iAS,S~1ffiffiffi2p Hj iAS Hj iSzei(w1zw2) Vj iAS Vj iS

ð2Þ

where w2 represents the propagating phase difference between twoStokes fields before they overlap at PBS1. In our experiment, the totalphase w1 1 w2 is actively stabilized via a built-in Mach–Zehnder inter-ferometer and fixed to zero (see Supplementary Information). With a

time delay dts 5 0.5ms, the measured polarization correlations of theStokes and anti-Stokes photons show a strong violation of a CHSH-type Bell’s inequality, with a visibility of 92%. Further measurementshows our atom–photon entanglement still survives up to a storagetime of dts 5 20ms (see Supplementary Information).

We now describe the entanglement generation between atomicensembles I and II via entanglement swapping. As shown in Fig. 1,photon 2 from Alice and photon 3 from Bob are both sent through a3-m optical fibre to an intermediate station for a joint BSM. In theexperiment, we chose to analyse the projection onto the Bell stateWzj i2,3~ 1

ffiffiffi2p

Hj i2 Hj i3z Vj i2 Vj i3

, which is achieved byoverlapping photons 2 and 3 onto a polarizing beam splitter(PBSm) and performing a proper polarization decomposition inthe output modes and a subsequent coincidence detection23.Conditioned on detecting a Wzj i2,3 state at the intermediate station,the two remote atomic ensembles are projected onto an identicalentangled state7,8: wz

I,II

~ 1 ffiffiffi

2p

Lj iI Lj iIIz Rj iI Rj iII

.It is noteworthy that double excitations in either atomic ensemble I

or II will cause false events in the BSM9,10, which reduce the successprobability of entanglement swapping by a factor of 2.Experimentally, the false events can be eliminated at the stage ofentanglement verification by the fourfold coincidence measurementof photons 1, 2, 3 and 4. Note that the detection time of photons 1 and4 is later than that of photons 2 and 3 by an interval dts, the storagetime in quantum memories. More importantly, such false events donot affect the applications of our experimental method in quantumrepeaters, since the generation of entanglement will be deterministicafter a second step of connecting two such nodes, where doubleexcitations are excluded automatically9,10.

The entanglement established between atomic ensembles I and IIcan be verified by converting the atomic spins into an entangledphoton pair 1 and 4, which are in the state Wzj i1,4. Here we measurethe S parameter in a CHSH-type Bell’s inequality,

S~ E(h1,h4)E(h1,h=4)E(h

=1,h4)E(h

=1,h

=4)

ð3Þ

where E(h1, h4) is the correlation function, and h1 and h=1 (h4 and h

=4)

âS âS

âS

âAS

âAS

âAS

Alice

87Rb 87Rb

87Rb

SMF1

PBS1

Write andreadprocess

W

∆ ∆Ω

SLM4 M3

SMF2M2L2

PBS2

ASL

ASR

HWP1HWP2

W1

M1

R1

SR

L1

BSM

MOT control5 ms

Trig. on

CleaningWrite

N

∆T

δtw δtsRead pulse

Cleaning andwrite pulses

6 m or 300 m fibre

| +⟩2,3

|e⟩ |e⟩

|a⟩ |a⟩|b⟩ |b⟩

Φ

PBSm1

1

2

2

I II

3 4

Bob

a

b

Figure 1 | The experimental scheme for entanglement swapping.a, Photons 2 and 3 overlap at BSM (Bell state measurement) and areprojected to the state of Wzj i2,3 through which the entanglement isgenerated between the two atomic ensembles I and II confined by magneto-optical traps (MOTs) in different glass cells separated by ,60 cm. Here 1, 2, 3and 4 indicate four photons emerging from the anti-Stokes modes (aaAS) andStokes modes (aaS), and Wzj i2,3 is one of the four Bell states. Inset,atom–photon entanglement. Shown are energy levels aj i, bj i, ej i 5 5S1=2,F~2

, 5S1=2,F~1

, 5P1=2,F~2

and theconfiguration of light beams. PBS, polarizing beam splitter; HWP, half-wave

plate; L, lens; M, mirror; SMF, single-mode fibre; W, write beam; R, readbeam; S, Stokes field; AS, anti-Stokes field; V, Rabi frequency of light fields.b, The time sequence of the experimental procedure at each site. For 6-m(300-m) fibre connection, there are 250 (200) experiment cycles in 5 ms andDT is 16 ms (20 ms) for one cycle, which contains N 5 10 (N 5 8) writesequences. The interval between two neighbouring write pulses is dtw 5 1ms(1.5 ms) and dts is the storage time. Whenever there is a desired coincidenceevent between photons 2 and 3, the following write sequence is stopped by afeedback circuit and the retrieve process can be started (at the time pointlabelled ‘Trig. on’).

NATURE | Vol 454 | 28 August 2008 LETTERS

1099

©2008 Macmillan Publishers Limited. All rights reserved

are the measured polarization bases of photon 1 (4). In the measure-ment, the polarization settings are (0u, 22.5u), (0u, 222.5u), (45u,22.5u) and (45u, 222.5u), respectively.

At a storage time dts 5 500 ns, the measured correlation functions(shown in Fig. 2) result in S 5 2.26 6 0.07, which violates Bell’sinequality by 3 standard deviations. To observe the lifetime of theentanglement between two remote memory qubits, we measure theinterference visibility of photons 1 and 4 as a function of the storagetime by choosing the polarization basis of z= (shown in Fig. 3,with zj i~ 1

ffiffiffi2p

Hj iz Vj ið Þ and j i~ 1 ffiffiffi

2p

Hj i Vj ið Þ). Upto a storage time of 4.5 ms, the visibility is still higher than the thresh-old of 1

ffiffiffi2p

, sufficient for the violation of Bell’s inequality. From thevisibilities of the atom–photon and atom–atom entanglements, theprecision of local operations at the BSM station is estimated to bebetter than 97% (see Supplementary Information). We emphasizethat this precision achieved here surpasses the threshold of 95% forlocal operations of independent photons necessary for future entan-glement purification and connections6, and therefore fits the require-ment for a scalable quantum network.

To demonstrate the robustness of our protocol in the generation ofquantum entanglement between two atomic ensembles over largedistances, we changed the length of the two connecting fibres from3 m to 150 m. The anti-Stokes photon is delayed 730 ns and theconnection length between Alice and Bob is 300 m. The entanglementswapping can be quantified by the fidelity of the measured state of theatomic ensembles. To determine the fidelity, we write the densitymatrix of wz

I,II

in terms of the Pauli matrices:

wz

wz

I,II~

1

4Izssx ssxssy ssyzssz ssz

ð4Þ

Here sx~ zj i zh j j i h j, sy~ Pj i Ph j Qj i Qh j and sz~

Hj i Hh j Vj i Vh j, with Qj i~ 1 ffiffiffi

2p

Hj izi Vj ið Þ and Pj i~1 ffiffiffi

2p

Hj ii Vj ið Þ. After a storage time of 1,230 ns (with a 730-

ns delay being taken into account), the two retrieved photons 1 and 4are sent to their own polarization analyser. Three series of polariza-tion settings are used and the measured local observables are shown

in Fig. 4. The fidelity of final state rexp on wz

is given by

F~Tr rexp wz

I,IIwz

~0:83+0:02, with 2.5 standard devia-

tions beyond the threshold of 0.78 to violate the CHSH-type Bell’sinequality for Werner states, demonstrating the success of entangle-ment swapping. This fidelity is comparable to the average valueachieved in the DLCZ-type functional quantum node14.

In summary, we have successfully demonstrated high-precisionentanglement swapping with storage and retrieval of light, a buildingblock for quantum repeaters. The extension of our work to longerchains will involve many quantum repeater nodes. To achieve thisambitious goal, several quantities—such as the lifetime and retrieveefficiency of the quantum memory, and the fidelity and generation rateof the entanglement state—still need to be improved significantly. Wesuggest three ways forward. First, better compensation of the residualmagnetic field and trapping the atoms in ‘clock states’24 with a blue-detuned optical trap25 should improve the lifetime to ,1 s. Second, ahigh optical density of the atomic cloud, achieved with the help of trapsor by coupling the atoms into an optical cavity26, should increase theretrieve efficiency close to unity. These improvements of the quantummemory would greatly enhance the fidelity and generation rate of theentanglement. Last, by local generation of entangled pairs of atomicexcitations together with the present technique of entanglement swap-ping, the entanglement distribution rate can be greatly improved27.Not only does our work enable immediate experimental investigationsof various quantum information protocols, but—with the above-mentioned future improvements—entanglement swapping withstorage and retrieval of light would also open the way to long-distancequantum communication.

METHODS SUMMARY

As shown in Fig. 1, Alice and Bob each have a cold atomic ensemble consisting of

about 108 atoms of 87Rb with temperature ,100mK. After 20 ms of loading

atoms into their MOTs separated by ,60 cm, we switch off the laser beams

and magnetic fields of the MOTs and start a 5-ms-long experiment cycle. At

each site, atoms are first prepared in the initial state aj i, followed by a (50 ns long,

0.6

(0°, 22.5°) (0°, –22.5°) (45°, 22.5°) (45°, –22.5°)

Cor

rela

tion

func

tion,

E(q

1, q

4) 0.4

0.2

0.0

–0.2

–0.4

–0.6

Figure 2 | Correlation functions of a CHSH-type Bell’s inequality with astorage time dts 5 500 ns. The x axis of the bar graph is labelled with (h1,h4), where h1 and h4 are the measured polarization bases of photons 1 and 4,respectively. Error bars represent statistical errors, which are 61 s.d.

0.9

0.8

0.7

0.6

0.50 2 4 6 8

Storage time, δts (µs)

Vis

ibili

ty

Figure 3 | Visibility of the atom–atom entanglement as a function of thestorage time with 6-m fibre connection. Filled black squares, visibility;dotted line, threshold for the violation of the CHSH-type Bell’s inequality.Error bars represent statistical errors, which are 61 s.d.

0.5

+ + + –

Polarization setting

– + – – HH HV VH VV

Frac

tion

0.4

0.3

0.2

0.1

0.0

a b c

Figure 4 | Polarization analysis of photons 1 and 4 when the connectionchannel is a 300-m fibre. The bars indicate the observed fractions for thecounts of specific polarization observables over the total counts of all thepolarization observables. For instance, the bar at 11 shows the ratiobetween the counts of the observable zzj i zzh j and the total counts ofobservables zzj i zzh j, zj i zh j, zj i zh j and j i h j.The polarization bases are chosen as a, 1/2; b, H/V; c, Q=P.

LETTERS NATURE | Vol 454 | 28 August 2008

1100

©2008 Macmillan Publishers Limited. All rights reserved

,1mW) weak write pulse, which has a beam waist of 240mm and is 10 MHz red-detuned from the aj i? ej i transition. Two anti-Stokes fields ASL and ASR

induced by the write beam via spontaneous Raman scattering are collected at

63u relative to the propagating direction of the write beam (70mm waist,

ej i? bj i). The excitation probability (xm) of the collective modes m (m 5 L,

R) is low (xm=1); thus the state of the atom–photon field can be expressed as11

Yj im< 0AS0bj imzffiffiffiffiffiffixm

p1AS1bj imzO(xm)

and iASibj im denotes the i-fold excitation of the anti-Stokes field and the collect-

ive spin in the atomic ensemble. The read beam is counter-propagating and

mode-matched with the write beam with a pulse length of 50 ns, a power of60 mW and a frequency close to resonance of the bj i? ej i transition.

Received 9 March; accepted 3 July 2008.

1. Peng, C.-Z. et al. Experimental long-distance decoy-state quantum keydistribution based on polarization encoding. Phys. Rev. Lett. 98, 010505 (2007).

2. Rosenberg, D. et al. Long-distance decoy-state quantum key distribution in opticalfiber. Phys. Rev. Lett. 98, 010503 (2007).

3. Schmitt-Manderbach, T. et al. Experimental demonstration of free-space decoy-state quantum key distribution over 144 km. Phys. Rev. Lett. 98, 010504 (2007).

4. Ursin, R. et al. Entanglement-based quantum communication over 144 km. NaturePhys. 3, 481–486 (2007).

5. Gisin, N., Ribordy, G., Tittel, W. & Zbinden, H. Quantum cryptography. Rev. Mod.Phys. 74, 145–195 (2002).

6. Briegel, H.-J., Dur, W., Cirac, J. I. & Zoller, P. Quantum repeaters: The role ofimperfect local operations in quantum communication. Phys. Rev. Lett. 81,5932–5935 (1998).

7. Zukowski, M., Zeilinger, A., Horne, M. A. & Ekert, A. K. ‘‘Event-ready-detectors’’Bell experiment via entanglement swapping. Phys. Rev. Lett. 71, 4287–4290 (1993).

8. Pan, J.-W., Bouwmeester, D., Weinfurter, H. & Zeilinger, A. Experimentalentanglement swapping: Entangling photons that never interacted. Phys. Rev. Lett.80, 3891–3894 (1998).

9. Zhao, B., Chen, Z.-B., Chen, Y.-A., Schmiedmayer, J. & Pan, J.-W. Robust creation ofentanglement between remote memory qubits. Phys. Rev. Lett. 98, 240502 (2007).

10. Chen, Z.-B., Zhao, B., Chen, Y.-A., Schmiedmayer, J. & Pan, J.-W. Fault-tolerantquantum repeater with atomic ensembles and linear optics. Phys. Rev. A 76,022329 (2007).

11. Duan, L.-M., Lukin, M. D., Cirac, J. I. & Zoller, P. Long-distance quantumcommunication with atomic ensembles and linear optics. Nature 414, 413–418(2001).

12. Matsukevich, D. N. & Kuzmich, A. Quantum state transfer between matter andlight. Science 306, 663–666 (2004).

13. Chou, C.-W. et al. Measurement-induced entanglement for excitation stored inremote atomic ensembles. Nature 438, 828–832 (2005).

14. Chou, C.-W. et al. Functional quantum nodes for entanglement distribution overscalable quantum networks. Science 316, 1316–1318 (2007).

15. Jiang, L., Taylor, J. M. & Lukin, M. D. Fast and robust approach to long-distancequantum communication with atomic ensembles. Phys. Rev. A 76, 012301 (2007).

16. Foreman, S. M. et al. Coherent optical phase transfer over a 32-km fiber with 1 sinstability at 10217. Phys. Rev. Lett. 99, 153601 (2007).

17. Yuan, Z.-S. et al. Synchronized independent narrow-band single photons andefficient generation of photonic entanglement. Phys. Rev. Lett. 98, 180503 (2007).

18. Moehring, D. L. et al. Entanglement of single-atom quantum bits at a distance.Nature 449, 68–72 (2007).

19. Wilk, T., Webster, S. C., Kuhn, A. & Rempe, G. Single-atom single-photonquantum interface. Science 317, 488–490 (2007).

20. Chen, Y.-A. et al. Memory-built-in quantum teleportation with photonic andatomic qubits. Nature Phys. 4, 103–107 (2008).

21. Chen, S. et al. A stable atom-photon entanglement source for quantum repeaters.Phys. Rev. Lett. 99, 180505 (2007).

22. Kwiat, P. G. et al. New high-intensity source of polarization-entangled photonpairs. Phys. Rev. Lett. 75, 4337–4341 (1995).

23. Pan, J.-W. & Zeilinger, A. Greenberger-Horne-Zeilinger-state analyzer. Phys.Rev. A 57, 2208–2211 (1998).

24. Harber, D. M., Lewandowski, H. J., McGuirk, J. M. & Cornell, E. A. Effect of coldcollisions on spin coherence and resonance shifts in a magnetically trappedultracold gas. Phys. Rev. A 66, 053616 (2002).

25. Kuga, T. et al. Novel optical trap of atoms with a doughnut beam. Phys. Rev. Lett.78, 4713–4716 (1997).

26. Simon, J., Tanji, H., Thompson, J. K. & Vuletic, V. Interfacing collective atomicexcitations and single photons. Phys. Rev. Lett. 98, 183601 (2007).

27. Sangouard, N. et al. Robust and efficient quantum repeaters with atomicensembles and linear optics. Phys. Rev. A 77, 062301 (2008).

Supplementary Information is linked to the online version of the paper atwww.nature.com/nature.

Acknowledgements We thank W. Dur for discussions. This work was supported bythe Deutsche Forschungsgemeinschaft, the Alexander von Humboldt Foundation,and the European Commission through the Marie Curie Excellence Grant and theERC Grant. This work was also supported by the National Fundamental ResearchProgram (grant 2006CB921900), the CAS and the NNSFC.

Author Information Reprints and permissions information is available atwww.nature.com/reprints. Correspondence and requests for materials should beaddressed to Y.-A.C. ([email protected]) or J.-W.P.([email protected]).

NATURE | Vol 454 | 28 August 2008 LETTERS

1101

©2008 Macmillan Publishers Limited. All rights reserved

Entanglement Swapping with Storage and Retrieval of Light

Zhen-Sheng Yuan2,1, Yu-Ao Chen1,2, Bo Zhao1, Shuai Chen1, Jörg Schmiedmayer3 and

Jian-Wei Pan1,2

1Physikalisches Institut, Ruprecht-Karls-Universität Heidelberg, Philosophenweg 12,

69120 Heidelberg, Germany

2Hefei National Laboratory for Physical Sciences at Microscale and Department of

Modern Physics, University of Science and Technology of China Hefei, Anhui 230026,

China

3Atominstitut der österreichischen Universitäten, TU-Wien, A-1020 Vienna, Austria

SUPPLEMENTARY INFORMATION

doi: 10.1038/nature07241

www.nature.com/nature 1

1. Atom-photon entanglement source

As discussed in the present paper, generating atom-photon entanglement is the first step

to demonstrate a quantum repeater. We will provide here the information on the quality

of the atomic-ensemble based quantum memory and of the atom-photon entanglement

source.

Shown in Fig. 1 of the present paper, the collective excitations in either of the two

spatial modes (labelled L for the left one and R for the right) are used as stationary

qubits that are entangled with photonic qubits. That is, the entangling components are

the polarization of the photon and the excitation of the spatial mode. We can investigate

the two modes individually addressed by a specific polarization of the emitted single

photons from either of the two spatial modes. For example, if vertical/horizontal (V/H)

polarized light are chosen at the outputs of PBS1 and PBS2, the property of the L/R

mode is being analyzed and can be operated separately.

0 5 10 15 20 250

10

20

30

40

g(2) AS

,S

Storage time (μs)

Supplementary Figure 1: The decay of cross correlation (2),AS Sg with the storage time.

The detection probability of anti-Stokes photons is fixed at 2×10-3.

doi: 10.1038/nature07241 SUPPLEMENTARY INFORMATION

www.nature.com/nature 2

The cross correlation function (2),AS Sg can be used to characterize the property of a

quantum memory. Theoretically, (2), 1 1/AS Sg χ= + , where χ is the excitation

probability of an atomic ensemble. Neglecting the imperfections of the optics, the

interference visibility V between anti-Stokes and Stokes photons can be obtained from

an optimistic estimation as28

(2),

(2),

11

AS S

AS S

gV

g−

=+

. (S1)

The cross correlation (2),AS Sg decays along the storage time, which also causes decay of

the visibility and therefore the atom-photon entanglement. Shown in Supplementary Figure 1, (2)

,AS Sg for one spatial mode becomes lower than 6 when the storage time

reaches 24 μs, indicating the visibility is lower than 71%.

0 5 10 15 20 251.8

1.9

2.0

2.1

2.2

2.3

2.4

2.5

2.6

2.7

Bel

l par

amet

er S

Storage time (μs)

Supplementary Figure 2: Decay of the S parameter in the Bell’s inequality with the

storage time at excitation rate of 2×10-3.. The solid dots are the measured data and the circles are calculated from the correlation function (2)

,AS Sg . The dashed line shows the

classical bound S=2.

doi: 10.1038/nature07241 SUPPLEMENTARY INFORMATION

www.nature.com/nature 3

The atom-photon entanglement is written as Eqn. (1) and can be converted to the

entanglement between the Stokes and anti-Stokes photons, Eqn. (2). It is verified by the

measurement of the Bell's inequality,

/ / / /, ( , ) ( , ) ( , ) ( , )AS S AS S AS S AS S AS SS E E E Eθ θ θ θ θ θ θ θ= − − − (S2)

where ( , )AS SE θ θ is the correlation function, in which ASθ and /ASθ ( Sθ and /

Sθ ) are

the measured polarization bases of the anti-Stokes (Stokes) photon. During the

measurement, the polarization settings are (0 ,22.5 ) , (0 , 22.5 )− , (0 , 22.5 ) , and

(45 , 22.5 )− , respectively. The parameter S can be approximately estimated by

2 2S V= . The violation of 2S ≤ shows quantum entanglement between the two

photons.

Shown in Supplementary Figure 2, the S parameter is 2.60±0.03 (resulting in a

visibility of 92%), at storage time of 500 ns and goes to lower than 2 at 24 μs. The small

difference between the measured S and the estimated one should arise from the

imperfections of the optics.

2. Phase stabilization method

As discussed in the present paper, the phase in the DLCZ-type functional quantum

node14 should be stabilized in a single-photon Mach-Zehnder interferometer whose arms

span the whole communication distance, where a sub-wavelength stability of the

transmission channel is needed. When the distance between the communication sites

becomes tens of kilometers, the path length fluctuations is extremely large and

impossible to stabilize using the passive method. Even for an active stabilization

method it is currently experimentally challenging. While in our experiment, since the

two-photon Hong-Ou-Mandel-type interferometer is exploited in the entanglement

connection, the requirement is to overlap the wave packets of two “messenger” photons

at the BSM. Therefore, the stability requirement for our interferometer is on the order of

doi: 10.1038/nature07241 SUPPLEMENTARY INFORMATION

www.nature.com/nature 4

coherence length of the photons, which is several-meter-long. In our experiment, a

Mach-Zehnder interferometer is used at each site to generate the atom-photon

entanglement. Here, the only task we need to do is to actively stabilize the phase of the

local interferometer. This is a standard technique and can be easily performed with a

high precision. We describe this technique as below.

Different from the technique in a previous atom-photon entanglement source20,

the current method for phase stabilization is simpler and more stable. Generally, the

phase in Eqn. (3) consists of four terms arising from the write beam, the read beam, the

anti-Stokes and the Stokes modes. However, since the two spatial modes share the same

write and read beams, they always observe a same phase of the write and of the read. So,

only the anti-Stokes and Stokes modes contribute to the phase 1 2φ φ+ , with 1φ ( 2φ )

arising from the path difference between LAS ( LS ) and RAS ( RS ). The present setup

provides high-quality entanglement and long-term stability by simply locking the phase

of 1 2φ φ+ . This is an obvious advantage compared with the atom-photon entanglement

implemented with two atomic ensembles12,20.

Supplementary Figure 3: Phase stabilization method. The prism P1 is mounted on a

piezo and the phase difference between the two arms L and R can be controlled by

driving this piezo.

To stabilize the phase 1 2φ φ+ in equation (3) actively, a Mach-Zehnder

interferometer are used as shown in Supplementary Figure 3. A locking beam with the

frequency of read and polarized at 45 is switched on during the 20 ms MOT loading

stage and sent into the interferometer at PBS2 by the weak reflection of a glass slice.

doi: 10.1038/nature07241 SUPPLEMENTARY INFORMATION

www.nature.com/nature 5

This locking beam is overlapped with the two spatial modes L and R. As the anti-Stokes

and Stokes light are perpendicularly polarized, the output of the locking beam is from

another port of PBS1. After the locking beam goes through a polarizer at 45 , the

interference signal can be detected by a photodiode and used to lock the phase 1 2φ φ+ .

During the experimental stage, this locking beam is switched off by an acousto-optical

modulator to prevent the collective excitation being destroyed.

Each arm of the Mach-Zehnder interferometer is about 1 meter long in the present

setup. With this active locking method, the visibility of the atom-photon entanglement

was kept at 92% during the experiment. In contrast, the visibility became nearly random

within a few minutes without the locking.

3. The estimation of the precision of local operations

To calculate the precision of local operations in the entanglement swapping, we assume

the generated state of the two remote atomic ensembles is a Werner state, since all the

mixed state can be transformed to a Werner state by bilateral unitary transformation.

In our experiment, the overall visibilities of the two atom-photon entanglement

sources were obtained as VI=92% and VII=90% from measured CHSH parameter via

2 2S V= . The relation between the initial visibilities VI , VII and the final visibility V

( 2.26 / (2 2) 80%= ) can be written as V=p12p2(4η2-1)VIVII/3 according to Eqn. (5) in

Ref. [6], where p1 and p2 are the reliabilities of one- and two-qubit operations, and η is

the quality of a single-qubit measurement. With the visibilities above, one get

p12p2(4η2-1)/3 =96.6%. Assume η=99.5% since the extinction ratio of the polarization

analyzer is better than 200:1, p12p2=97.9%, so p2>97.9%, which indicates the precision

of local operations. Assume the same precision can be achieved in the further

connections and purifications. Actually, this assumption is reasonable since the

precision of local operations depends on the quality of optical devices like beam

doi: 10.1038/nature07241 SUPPLEMENTARY INFORMATION

www.nature.com/nature 6

splitters, polarizers and half-wave plates, whose extinction ratio can be very high.

Plugging η=99.5% into Eqn. (6) in Ref. [6], we obtained a threshold of

p2-th=95.2% above which future entanglement purification works. When we used the

measured precision p2=97.9%, a fidelity region of 57.7%=Fmin<F< Fmax=94.2% was

obtained, which means any initial fidelity between this region can be purified up to

94.2% with the help of auxiliary entanglement pairs. With the current precision of local

operations, if we do another connection between two entanglement pairs with both

fidelities of 94.2%, the connected longer pair has a fidelity of 86.7%, which is in the

region of [Fmin, Fmax]. If we have seven nodes with all fidelities of 94.2%, after six

connections the finally connected entanglement pair will have a fidelity of 59.8%, still

in the region of [Fmin, Fmax]. This estimation shows that the quantum network is scalable

with the current precision of local operations, given a sufficiently long lifetime of the

generated entanglement pairs.

4. Towards long-distance quantum communication

Following the BDCZ scheme6,9, connecting two of the present building blocks provides

a chance to generate entanglement deterministically. Double excitations are “washed”

out automatically with this connection by a carefully designed Bell-state measurement.

Considering double emissions, the entangled state generated in the two atomic

ensembles is described by

† † † † † † † †

, ,vac

2 2I II I II I I II IIL L R R L R L R

I II I II

S S S S S S S Sψ

⎛ ⎞+ += +⎜ ⎟⎜ ⎟⎝ ⎠

(S3)

where the spatial modes and atomic ensembles are distinguished by subscript (L,R) and

(I,II), respectively. The first part is the maximally entangled state needed for further

doi: 10.1038/nature07241 SUPPLEMENTARY INFORMATION

www.nature.com/nature 7

operation, while the second part is the unwanted two-excitation state coming from

second-order excitations.

It is obvious the spurious contributions of two-excitation terms prevent further

entanglement manipulation and must be eliminated by some means. However we find

that it is not necessary to worry about these terms, because they can be automatically

washed out if the BSM in further entanglement swapping is carefully designed.

Let us consider four communication sites I, II, III and IV and assume we have created complex entangled states

,I IIψ and

,III IVψ between sites (I,II) and (I,IV)

respectively. Note the atomic ensembles II and III are at the same location. The memory

qubits II and III are illuminated simultaneously by read laser pulses. The retrieved

anti-Stokes photons are subject to BSM. Note that the arrangement of the PBSs in this BSM is designed to identify Bell state φ± at + / − polarization basis (in contrast

to the BSM used in the paper at H / V polarization basis), so that the two-photon

states converted from the unwanted two-excitation terms are directed into the same

output and thus won't induce a coincidence count on the detectors. Conditioned on the

coincidence events in this BSM, the memory qubits are projected into a maximally

entangled state.

† † † †, ,

1 ( ) vac2 I IV I IVL L R RI IV I IV

S S S Sψ = + . (S4)

Considering the imperfections in experiments, we find that the coincidence counts

prepare the memory qubits into a mixed entangled state of the form , 2 2 1 1 0 0I IV p p pρ ρ ρ ρ= + + , where the coefficients p2, p1 and p0 are determined by

the retrieve efficiency and detection efficiency. Here 2 ,I IVρ ψ ψ= is a maximally

entangled state, ρ1 is a maximally mixed state where only one of the four spatial modes

has one excitation and ρ0 is the vacuum state that all the spatial modes are in the ground

states. It is easy to see that ρI,IV is in fact an effectively maximally entangled states,

doi: 10.1038/nature07241 SUPPLEMENTARY INFORMATION

www.nature.com/nature 8

which can be projected automatically to a maximally entangled state in the

entanglement based quantum cryptography schemes. When we implement quantum

cryptography via the Ekert protocol29, only the first term ρ2 can contribute to a

coincidence count between the detectors at the two sites and will be registered after

classical communication. The maximally mixed state term ρ1 and the vacuum term ρ0

have no contribution to the experimental results, and thus ρI,IV is equivalent to the Bell state

,I IVψ .

The effectively entangled state can be connected to longer communication

distance via further entanglement swapping. A detailed calculation shows that, during

the entanglement connection process, the coefficients p2, p1 and p0 are stable to the first

order thanks to the two-photon detection6. The time overhead scales quadratically with

the communication distance. The entangled state can also be actively purified by use of

linear optics quantum purification30.

Supplementary Notes

28. de Riedmatten, H. et al. Direct measurement of decoherence for entanglement

between a photon and stored atomic excitation. Phys. Rev. Lett. 97, 113603 (2006).

29. Ekert, Artur K. Quantum cryptography based on Bell’s theorem. Phys. Rev. Lett. 67,

661–663 (1991).

30. Pan, J.-W., Gasparoni, S., Ursin, R., Weihs, G. & Zeilinger, A. Experimental

entanglement purification of arbitrary unknown states. Nature 423, 417 (2003).

doi: 10.1038/nature07241 SUPPLEMENTARY INFORMATION

www.nature.com/nature 9