Photonics Research, Volume. 13, Issue 8, 2013(2025)

Pilot-reference-free continuous-variable quantum key distribution with efficient decoy-state analysis

Anran Jin1...2,†, Xingjian Zhang3,4,5,†, Liang Jiang5, Richard V. Penty2 and Pei Zeng6,* |Show fewer author(s)
Author Affiliations
  • 1Yau Mathematical Sciences Center, Tsinghua University, Beijing 100084, China
  • 2Electrical Engineering Division, Department of Engineering, University of Cambridge, Cambridge CB3 0FA, UK
  • 3QICI Quantum Information and Computation Initiative, School of Computing and Data Science, The University of Hong Kong, Hong Kong, China
  • 4Shenzhen Futian SUSTech Institute for Quantum Technology and Engineering, Shenzhen 518055, China
  • 5International Quantum Academy, Shenzhen 518048, China
  • 6Pritzker School of Molecular Engineering, The University of Chicago, Chicago, Illinois 60637, USA
  • show less

    Continuous-variable quantum key distribution (CV QKD) using optical coherent detectors is practically favorable due to its low implementation cost, flexibility of wavelength division multiplexing, and compatibility with standard coherent communication technologies. However, the security analysis and parameter estimation of CV QKD are complicated due to the infinite-dimensional latent Hilbert space. Also, the transmission of strong reference pulses undermines the security and complicates the experiments. In this work, we tackle these two problems by presenting a time-bin-encoding CV protocol with a simple phase-error-based security analysis valid under general coherent attacks. With the key encoded into the relative intensity between two optical modes, the need for global references is removed. Furthermore, phase randomization can be introduced to decouple the security analysis of different photon-number components. We can hence tag the photon number for each round, effectively estimate the associated privacy using a carefully designed coherent-detection method, and independently extract encryption keys from each component. Simulations manifest that the protocol using multi-photon components increases the key rate by two orders of magnitude compared to the one using only the single-photon component. Meanwhile, the protocol with four-intensity decoy analysis is sufficient to yield tight parameter estimation with a short-distance key-rate performance comparable to the best Bennett-Brassard-1984 implementation.

    1. INTRODUCTION

    Quantum key distribution (QKD) allows the generation of random secure keys between distant communication parties, of which the security is guaranteed by quantum physical laws. Apart from its theoretical advances, QKD is also one of the few quantum information processing technologies that can be robustly deployed in the field, where photonic systems are considered the most suitable carriers of QKD operation. In general, two types of QKD protocols exist based on the detection methods: discrete-variable (DV) QKD [1,2] uses the single-photon detector or photon-number-resolving detector to generate discrete detection information, while continuous-variable (CV) QKD [35] applies optical homodyne or heterodyne detection to generate continuous measurement information.

    CV QKD has its advantages over DV QKD in short distances, mainly attributing to the distinct features of the coherent detectors used. The homodyne and heterodyne detectors are compatible with the standard classical communication and can be operated in much milder conditions than single-photon detectors. The spatial-temporal filtering of the local oscillators (LOs) allows dense wavelength-division multiplexing with intense classical channels [68], and the high quantum efficiency and operation rate give CV QKD high key rates in metropolitan distances [911]. Moreover, the feasibility of on-chip implementations of the coherent detectors [12] promises large-scale integrated quantum networks. CV QKD is therefore considered highly practical and promising.

    However, there exist two major limitations to the reliability of CV QKD. First, the transmission of the strong local oscillators is usually necessary to set up the phase reference between the communication parties, yet this complicates the implementation in the multiplexing separation and the relative phase shift calibration with the signals [13]. The LO transmission also opens up security loopholes where the eavesdropper Eve can affect the estimation of the signal variance by manipulating the LO intensity [14,15], input time [16], and wavelength [17]. The “local” local oscillator scheme [13,18] is a valid solution, yet it still requires the transmission of pilot pulses and compensation in the classical post-processing layer, which increases the experimental complexity. Second, the security of CV QKD in the finite-data regime under coherent attacks is still incomplete. In fact, for the traditional entanglement-distillation approach [19], the finite-size coherent attack security is only tackled for Gaussian-modulated CV QKD [20], which is, however, impractical since continuous modulation is never possible in reality.

    Recently, several remarkable works on CV QKD have been proposed, aiming at closing its security loopholes. In Refs. [5,21], Koashi et al. proposed a DV-like security analysis for the binary phase shift keying CV QKD. Their analysis covers the finite size and coherent attack intrinsically since it follows the phase-error complementarity approach [22], yet their protocol still assumes the transmission of local oscillators. Qi [23] and Primaatmaja et al. [24] proposed CV QKD protocols with two-mode encoding, generating dual-rail qubits that do not require global references. Their security analyses, however, do not cover the finite-data regime and coherent attacks. In fact, Qi’s analysis requires repeated measurements and is only valid for individual attacks, and Primaatmaja et al.’s analysis is based on the Devetak-Winter formula [19] and is only valid for collective attacks. Hence, the gap in CV QKD between theory and practice is still a challenging problem to be tackled.

    In this work, we close this gap by proposing a new time-bin-encoding CV QKD protocol that enjoys both simple security proof and practical implementation. We remove the necessity of LO transmission by the two-mode encoding, hence closing the security loophole while simplifying the experimental setups. We follow the phase-error complementarity approach [22,25,26] so that the security naturally covers the general coherent-attack case. What is more, the intensity-based encoding allows phase randomization to be applied, where we can group the received signals based on the transmitted photon numbers and restore the tagging-based security analysis [27,28] and the decoy-state method [29,30]. Instead of generating secure key bits from all raw key bits, we can thus take advantage of the photon-number tags and distill key bits from the rounds with a low phase-error rate. Our tagging-based security analysis builds a direct connection between CV QKD and the normal Bennett-Brassard-1984 (BB84) protocol: we clearly show how the multi-photon components in the CV QKD protocol contribute to a higher key rate in short distances. Compared with a similar protocol in Ref. [24] with numerical optimization under collective attacks, our protocol generates higher key rates under coherent attacks with simpler parameter estimation using four decoy levels, while also demonstrating good finite-size performance with a data size of N=1010.

    We will start with the protocol description of the time-bin-encoding CV QKD in Section 2. We present its security analysis based on phase-error correction [22,25,26] in Section 3, identifying an equivalent protocol squashing the optical modes into qubits with identical key mapping statistics [5,27,31] in Section 3.A. We exploit the block-diagonal structures of both the source and the receiver in Section 3.B, thus invoking the photon-number tagging technique [27,28] standard in DV QKD in this CV protocol. In Section 3.C, we calculate the parameters in the key-rate formula with quantities on optical modes. The estimation of these quantities will be explained in Section 4 with homodyne tomography [32] and a decoy method [29,30]. We finally simulate the asymptotic and finite-size performances of the time-bin CV QKD under realistic fiber-channel setups in Section 5.

    2. PROTOCOL DESCRIPTION

    We present the proposed time-bin-encoding CV QKD protocol in Table 1 and depict its schematic diagram in Fig. 1. The two communication parties, Alice and Bob, employ the time-bin degree of freedom to encode keys. They use the Z-basis for key generation and X-basis for parameter estimation. At the moment, we do not present the details of the X-basis parameter settings. We will specify the choices of the random phase factors, φa1,φa2,φb1 and φb2, and the light intensity, μa{μ,ν1,ν2,0}, in Section 4.

    Phase-Randomized Time-Bin-Encoding CV QKD

    On Alice’s side (source):

    Z-basis:

    Randomly select a key bit ka{0,1}, a phase factor φa[0,2π), and a light intensity μa{μ,ν1,ν2,0}.

    Prepare a coherent state of |0A1|μaeiφaA2 for ka=0 or |μaeiφaA1|0A2 for ka=1.

    X-basis:

    Randomly select two phase factors φa1 and φa2 and a light intensity μa{μ,ν1,ν2,0}.

    Prepare a coherent state of |μa/2eiφa1|μa/2eiφa2.

    Alice sends the state through an authenticated channel to Bob.

    On Bob’s side (detection):

    Z-basis:

    Randomly select a phase factor φb[0,2π).

    Use homodyne detectors both with LO phases φb to measure the modes and obtain quadratures q1 and q2.

    Decode the key bit as 0 if |q1|<τ|q2|>τ, 1 if |q1|>τ|q2|<τ, and Ø otherwise.

    X-basis:

    Randomly select two phases φb1 and φb2 independently.

    Use homodyne detectors with LO phases φb1 and φb2 to measure the modes and obtain quadratures q1 and q2.

    Use φ1,φ2 and q1,q2 for phase-error estimation (see Section 4).

    Alice and Bob perform basis sifting, where they obtain raw keys in the rounds; they both choose Z-basis with light intensity μa=μ and kbØ.

    Based on parameter estimation, Alice and Bob perform information reconciliation and privacy amplification to obtain final keys.

    Schematic diagram of the experimental setup. The setups of Alice and Bob are shaded in green and blue, respectively. Alice prepares two-mode phase-randomized states according to the basis choice and raw key value in key generation rounds, as shown in the table. In this work, we consider a time-bin encoding, where one obtains two modes via time delay. The state modulation consists of intensity modulation (IM), phase modulation (PM), and necessary attenuation (ATTN). Upon receiving the state, Bob measures each mode with homodyne detectors. He uses a synchronized clock to distinguish adjacent modes and applies phase modulation (PM) to the local oscillator (LO).

    Figure 1.Schematic diagram of the experimental setup. The setups of Alice and Bob are shaded in green and blue, respectively. Alice prepares two-mode phase-randomized states according to the basis choice and raw key value in key generation rounds, as shown in the table. In this work, we consider a time-bin encoding, where one obtains two modes via time delay. The state modulation consists of intensity modulation (IM), phase modulation (PM), and necessary attenuation (ATTN). Upon receiving the state, Bob measures each mode with homodyne detectors. He uses a synchronized clock to distinguish adjacent modes and applies phase modulation (PM) to the local oscillator (LO).

    Here, we briefly explain the idea behind the protocol design. The source states of our scheme resemble the ones in the time-bin-encoding BB84 protocol with coherent states [1], where the light intensities of the consecutive pulses naturally encode the key-bit information. As the key information is encoded in the relative intensity between the two modes, Alice does not need to send a pilot phase reference as in common CV QKD. The source states also resemble those in the coherent-one-way (COW) protocol [33,34]. The difference is that we randomize the phase of the coherent state pulses and employ homodyne detection for the measurement devices.

    In our scheme, Bob decodes the key bit information, namely, the Z-basis information, by measuring the light intensity of the pulses using the homodyne detectors. For instance, when Bob observes q1 to be close to zero and q2 to be far away from zero, he may naturally guess that the original state sent by Alice corresponds to ka=0. However, unlike the key decoding with photon-number detectors, the result of measuring a coherent state’s quadrature is subjected to a Gaussian distribution rather than a fixed value. The inherent shot noise of the homodyne detection introduces an intrinsic error in distinguishing a vacuum state from a pulse with a non-zero intensity [35]. To suppress the bit error, we introduce a threshold value, τ, in key decoding. The pulse intensity will be considered non-zero only when the quadrature magnitude is larger than τ. So a bit 0 will be decoded if |q1|<τ and |q2|>τ and bit 1 if |q1|>τ and |q2|<τ. The choice of τ should be optimized with respect to the channel transmittance and pulse intensity. As we will show in Appendix A, although this key mapping scheme can be further optimized, the performance of the simple key mapping scheme is already near-optimal.

    The X-basis is designed to estimate the information leakage of different photon-number components of the Z-basis. Thanks to the phase randomization for both the sources and the detectors, the Z-basis states are block-diagonal on the total photon-number basis on the two optical modes after being emitted from the source and before being measured by the homodyne detectors. As we will clarify in Section 3, we can equivalently introduce total photon-number measurements at these two locations. As a result, Eve’s eavesdropping strategy is effectively “twirled” to a photonic channel that only acts on the states incoherently with respect to the total photon numbers. One can thus virtually tag the emitted and received pulses according to the photon-number space, allowing the Gottesman-Lütkenhaus-Lo-Preskill (GLLP) framework [27] for analyzing the key privacy contained in each photon-number subspace. In particular, dealing with photon-number spaces effectively brings our security analysis to the DV regime. In Section 3, we shall construct observables to estimate the m-photon component phase-error rates em,mX for privacy estimation. Intuitively, the phase-error rates em,mX provide upper bounds on the key information leakage to the eavesdropper, Eve.

    To estimate the m-photon component phase-error rates, em,mX, ideally, we need a source emitting the photon-number cat states, (|0|m±|m|0)/2, and photon-number-resolving measurements that distinguish the cat states. While this is not directly implementable, we can use only coherent states and homodyne measurements to establish unbiased estimators of em,mX, as shown in Table 2. On the source side, we employ a generalized decoy-state method to estimate the behaviors of photon-number-cat states using coherent states with various intensities [29,30], which shall be discussed in Section 4.B. On the detection side, ideally, we also want to measure the photon-number-cat states to obtain unbiased estimation of the phase-error rates, em,mX. While this is not directly measurable in practice, we employ the homodyne tomography technique and estimate the photon-number-cat state measurement via quadrature measurement results [3641], which shall be discussed in Section 4.A.

    State Preparation and Detection Settings in the Ideal Implementation and the Realistic Implementationa

    SourceDetection
    BasisIdealRealIdealReal
    Z|0|m|0|μeiφaΠ^(q1,q2), Eq. (9)Π^(q1,q2)
    |m|0|μeiφa|0
    X12(|0|m±|m|0)|μ2eiφa1|μ2eiφa212(|0|m±|m|0)(0|m|±m|0|)Q^φ1Q^φ2, Eq. (6),
    Estimation via Eq. (29)

    For brevity, we omit the subscripts of modes and express the detection with the measurement operators. In key generation rounds, Bob applies phase-randomized homodyne detection for key-decoding. The expression of measurement operator Π^(q1,q2) is given in Eq. (9), where q1 and q2 represent the quadratures of the two modes. The operator is block-diagonal on the total photon-number basis. For parameter estimation, ideally, Alice sends photon-number-cat states, and Bob performs a corresponding projective measurement. In the realistic setting, Alice can only prepare phase-randomized weak coherent states, and Bob can only perform phase-randomized homodyne measurements. The homodyne measurement operator, Q^φ1Q^φ2, is given in Eq. (6). Afterward, Bob estimates the photon-number-cat state measurement expectations via homodyne tomography methods, as shown in Eq. (29).

    We briefly remark on the performance of homodyne detection. In key decoding, one may consider the homodyne detection as ill-performed single-photon detectors that introduce an inevitable bit-error rate. On the other hand, homodyne detection allows for more efficient parameter estimation than single-photon detection. As we shall discuss later, the set of all quadrature operators spans the underlying mode, thus allowing one to express any linear operator in terms of the quadrature operators. Therefore, with proper transformation of the quadrature measurement results, homodyne detection allows one to obtain an unbiased estimation of linear operator expectations. This is the reason for accurately estimating phase-error rates with repeated homodyne measurements, including those of the multi-photon components. In comparison, as the single-photon detection is not information-complete, estimation of multi-photon observables requires more complex setups such as sequential beam splitting [42], and one can only obtain upper and lower bounds rather than an unbiased estimation.

    3. SECURITY ANALYSIS

    We analyze the security of our phase-randomized time-bin-encoding CV QKD protocol along the complementarity approach [22,26,27]. As outlined in Fig. 2, we shall set up a series of equivalent protocols of the realistic implementation that do not change the statistics of any observer, with which we define the phase-error observable and estimate the key privacy. In Section 3.A, we shall prove that raw key generation can be effectively regarded as qubit measurements on a pair of entangled qubits, which allows us to borrow the mature complementarity-based security analysis in the DV regime. In brief, on the source side, we transform the preparation of key states to an entanglement-based protocol [25,26], where a qubit measurement controls the key-encoding process, as shown in Fig. 2(b). On the detection side, we prove that the homodyne measurement can be squashed into an effective qubit measurement, as shown in Fig. 2(c). Moreover, in Section 3.B, we shall rigorously prove that phase randomization twirls the photonic modes into diagonal states on the Fock basis and explain how to apply the tagging idea of the GLLP framework [27,28]. We also show how to estimate the phase-error rates for different photon-number components from Fock-basis observables in Section 3.C. Later in Section 4, we show that the estimation can be realized in the realistic implementation with coherent states and homodyne detection.

    Equivalent quantum circuits in key generation rounds. Reductions in each step are plotted with red dashed boxes. (a) The realistic implementation. The operations on Alice’s and Bob’s sides are shaded in green and blue, respectively. Alice prepares weak coherent states on two modes, which depend on the basis choice and the raw key value. On Bob’s side, Bob measures the two modes with homodyne detectors (HDs) and obtains quadratures q1 and q2. Afterward, Bob performs classical post-processing (CP) on the data and obtains a raw key kb probabilistically, where the key decoding may fail due to the key mapping threshold, denoted as Ø. The blue rounded boxes represent phase randomization processes in state preparation or for the LOs in homodyne detection. (b) Equivalent entanglement-based state preparation. Key encoding can be interpreted as a qubit control operation on two modes where the control qubit measurement gives Alice’s raw key ka. The joint state on the two modes is diagonal on the Fock basis after phase randomization. One can insert a photon-number measurement, M^na, and read out the total photon number, m, without changing the state. (c) Equivalent key-decoding measurement. The joint state of the two modes becomes diagonal on the Fock basis due to detector phase randomization. In key decoding, the modes are first squashed into a qubit probabilistically, where the failure gives the abort signal Ø. Upon successful squashing into a qubit, the computational-basis measurement gives the raw key bit. (d) Due to detector phase randomization, one can insert a photon-number measurement, M^nb, and read out the total photon number, n, without changing the state. (e) Equivalent circuit for security analysis. After the above reductions, the key generation measurements can be equivalently defined on a pair of (sub-normalized) qubit states.

    Figure 2.Equivalent quantum circuits in key generation rounds. Reductions in each step are plotted with red dashed boxes. (a) The realistic implementation. The operations on Alice’s and Bob’s sides are shaded in green and blue, respectively. Alice prepares weak coherent states on two modes, which depend on the basis choice and the raw key value. On Bob’s side, Bob measures the two modes with homodyne detectors (HDs) and obtains quadratures q1 and q2. Afterward, Bob performs classical post-processing (CP) on the data and obtains a raw key kb probabilistically, where the key decoding may fail due to the key mapping threshold, denoted as Ø. The blue rounded boxes represent phase randomization processes in state preparation or for the LOs in homodyne detection. (b) Equivalent entanglement-based state preparation. Key encoding can be interpreted as a qubit control operation on two modes where the control qubit measurement gives Alice’s raw key ka. The joint state on the two modes is diagonal on the Fock basis after phase randomization. One can insert a photon-number measurement, M^na, and read out the total photon number, m, without changing the state. (c) Equivalent key-decoding measurement. The joint state of the two modes becomes diagonal on the Fock basis due to detector phase randomization. In key decoding, the modes are first squashed into a qubit probabilistically, where the failure gives the abort signal Ø. Upon successful squashing into a qubit, the computational-basis measurement gives the raw key bit. (d) Due to detector phase randomization, one can insert a photon-number measurement, M^nb, and read out the total photon number, n, without changing the state. (e) Equivalent circuit for security analysis. After the above reductions, the key generation measurements can be equivalently defined on a pair of (sub-normalized) qubit states.

    To focus on the essence of security analysis, we present the result in a single-round analysis in this section, where one can interpret it as the quantum Shannon limit under collective attacks. Namely, Eve applies the same attacking strategy to the quantum signal of Alice and Bob in each round. At the end of the protocol, Eve applies an optimal joint operation to guess the users’ final key. Note that a collective attack is stronger than an individual attack, where Eve must measure her side information before the classical post-processing in an individual attack. Nevertheless, the complementarity-based security analysis is inherently adapted to the most general case, namely, the coherent attack, where the statistics over the rounds may not be independent and identically distributed (i.i.d.) [43]. Unlike the collective attack, Eve may vary her attacking strategies in different rounds. Readers may refer to Ref. [44] for a detailed discussion on the various types of attacks. We will discuss the parameter estimation with non-i.i.d. finite statistics in Section 4.C.

    A. Entanglement-Based Squashing Protocol

    Here, we show the equivalence of the time-bin CV QKD protocol to a qubit-based entanglement distribution protocol, where the protocols generate the same transmitted quantum states and measurement statistics. The latter protocol enables us to simplify the security analysis and estimate the information leakage from phase-error rates.

    We first focus on the key-generation rounds in the protocol where both users choose the Z-basis, of which the whole procedure is depicted in Fig. 2(a). In the realistic implementation, Alice prepares phase-randomized coherent states, 02πdφa2π|Ψ(ka)φaA1A2Ψ(ka)φa|,where |Ψ(ka)φaA1A2={|0A1|μeiφaA2,if  ka=0,|μeiφaA1|0A2,if  ka=1.We denote the optical modes sent to Bob as A1 and A2, which are CV systems. Throughout this paper, we treat the phase of optical modes, φa, as fully randomized over [0,2π). Finite phase randomization, φa{2jπ/D}j[D], suffices for a practical implementation, where its difference from the full phase randomization is negligible when D is sufficiently large [45]. This is also the case in later discussions on the detector phase randomization. Alice’s key-state preparation can be effectively seen as an entanglement-based protocol [25,26]. Given the phase value, φa, Alice first prepares the following entangled state: |ΨφaAA1A2=12(|0A|Ψ(ka=0)φaA1A2+|1A|Ψ(ka=1)φaA1A2),where system A is a qubit system that superposes the two possible key states. The entangled state can be prepared by the quantum circuit in Fig. 2(b). Up to phase randomization, systems A and A1A2 are initialized in |+ and |0|μ, and a control-swap operation is then applied from the qubit system to the optical modes. Alice obtains raw key bit ka by measuring system A on the computational basis, and the optical modes are prepared into the corresponding key state, |Ψ(ka)φa. The complementary observable of Alice’s key-generation measurement can thus be defined over qubit system A, which measures the complementary basis of {|+,|}{(|0±|1)/2}.

    At the detection side in Fig. 2(a), Bob receives two optical modes B1 and B2, takes homodyne measurements, and maps the quadratures to a raw key or an abort signal. This process can be described by a trace-non-preserving completely positive map, FrandB1B2B(ρ^B1B2)=02πdφb2πR0dq1dq2K^(q1,q2,φb)ρ^B1B2K^(q1,q2,φb),where K^(q1,q2,φb)|0Bq1(φb),q2(φb)|B1,B2+|1Bq2(φb),q1(φb)|B1,B2,|q(φ) is the rotated position eigenstate of quadrature observable Q^φ=a^eiφ+a^eiφ,with a^ and a^ denoting the annihilation and creation operators, respectively, and R0R2 records the region that decodes the real-valued tuple, (q1,q2), as kb=0. Note that in our protocol, R0={|q1|<τ}×{|q2|>τ}, and the region decodes the quadratures to kb=1 under the mapping (q1,q2)(q2,q1), which we denote as R1={|q1|>τ}×{|q2|<τ}. The LOs of homodyne measurements are synchronically randomized, as denoted by φb in Eq. (4). As the key-decoding region does not cover the entire parameter space, FrandB1B2B is hence not trace-preserving, where Tr[FrandB1B2B(ρ^B1B2)] gives the probability of obtaining raw key bit kb{0,1}. Bob’s raw key can be equivalently seen as obtained by measuring the squashed sub-normalized qubit on the computational basis, and the probabilities are given by Pr(kb=0)=0|FrandB1B2B(ρ^B1B2)|0=02πdφb2πR0dq1dq2q1(φb),q2(φb)|ρ^B1B2|q1(φb),q2(φb),Pr(kb=1)=1|FrandB1B2B(ρ^B1B2)|1=02πdφb2πR1dq1dq2q1(φb),q2(φb)|ρ^B1B2|q1(φb),q2(φb).Similar to the treatment to A, we can define the complementary observable of Bob’s key generation measurement on qubit system B.

    B. Photon-Number Tagging of the Source and Receiver

    In the last section, we have shown that raw keys can be equivalently seen as generated from qubit measurements on A and B. Should Alice and Bob instead measure the qubit system on the complementary bases, the probability that they obtain different results, or the phase-error rate, eX, could be used to upper-bound the average privacy amplification cost per round as h(eX), where h(p)=plogp(1p)log(1p) is the binary entropy function. Nevertheless, the actual privacy leakage may be less than the direct calculation. Note that the above privacy leakage estimation is averaged over the overall quantum state transmitted from Alice to Bob. The contribution to the privacy leakage of different components in quantum signals can differ. For instance, Eve can apply the photon-number-splitting (PNS) attack in the rounds in which Alice transmits two photons and Bob receives only a single photon [46,47]; hence no privacy should be expected, rendering the phase-error probability to be 1/2 in these rounds. If Alice and Bob can distinguish such rounds from the others, they can simply discard them in privacy amplification. The GLLP framework makes the above statement rigorous [27,28]. Suppose Alice and Bob can categorize the transmitted quantum signals into different groups, or tags, and evaluate phase-error probabilities separately. The privacy amplification cost can be evaluated by iQih(eiX), where Qi is the probability that a signal in the ith group is transmitted and detected, namely, the gain, and eiX is the phase-error probability of the group. Due to the concavity of the entropy function, this estimation is no larger than h(iQieiX).

    In DV QKD, the tagging idea has been well practiced. In the coherent-state-based BB84 protocol, phase randomization on the source side diagonalizes the quantum signals on the Fock basis [29,30], and an ideal single-photon detector naturally distinguishes the single-photon components from other detected Fock components, allowing Alice and Bob to tag the quantum states with respect to the photon number [48]. Similarly, we now prove that the photon-number tag can also be applied to the phase-randomized CV QKD protocol in Table 1. On the source side, the phase randomization diagonalizes the state on the joint Fock basis, ρ^Z=02πdφa2π12|0A10||μeiφaA2μeiφa|+12|μeiφaA1μeiφa||0A20|=m=0Prμ(m)12(|0mA1A20m|+|m0A1A2m0|),where Prμ(m)=eμμm/m! is the Poisson distribution. Consequently, one can virtually insert a photon-number measurement after phase randomization to measure the total photon number on the two modes without changing the state, as shown in Fig. 2(b). On the detection side, when Bob takes the Z-basis measurement, the phase-randomized homodyne detector POVM elements can be expanded on the Fock basis [24], Π^(q1,q2)=02πdφb2π|q1(φb)B1q1(φb)||q2(φb)B2q2(φb)|=n=0k0=0nl0=0nψk0(q1)ψl0(q1)ψnk0(q2)ψnl0(q2)|k0,nk0B1B2l0,nl0|,where ψn(qj)=12nn!2πHn(qj/2)eqj2/4is the coordinate representation of Fock state |n, with Hn being the nth Hermite polynomial. Therefore, one can virtually insert another photon-number measurement after phase randomization before the squashing channel [Eq. (4)] on the detection side to measure the total photon number of the received state, as shown in Fig. 2(d).

    Based on the above results, we depict a virtual quantum circuit of the protocol when both Alice and Bob choose the Z-basis in Fig. 2(e). We denote the photon-number measurement results on the source side and the detection side as m and n, respectively. Alice and Bob can thus distill secrete keys separately based on the photon-number tag of (m,n).

    A lower bound on the key rate can then be given by [27,28] rm=0Qm,m[1h(em,mX)]fQZh(eZ),where Qm,m and em,mX denote the gain and the phase-error rate in the rounds where m photons are sent and m photons are accepted, QZ is the Z-basis gain, eZ is the bit-error rate, and f is the efficiency of information reconciliation (note not to confuse the gains with quadrature observables). In addition, since Bob’s key decoding succeeds probabilistically where he only accepts quadratures above the threshold, we use the term “accepting” to represent receiving a certain state and passing the post-selection. All the gains and error rates in the key-rate formula are restricted to the rounds with light intensity μa=μ. We discard the rounds where the total photon number decreases after state transmission, as the photons that are lost may come from Eve’s interception, with which Eve can apply a PNS attack. The corresponding phase-error probability is 1/2; hence these rounds do not contribute to key generation. In addition, as the transmission channel is naturally lossy in a usual setting, we do not account for the terms where the total photon number increases.

    Note that the key-rate formula in Eq. (11) assumes forward reconciliation, where Bob reconciles his raw keys to Alice’s, ka, and then the users perform privacy amplification. The rounds where Alice sends a non-vacuum state while Bob receives a vacuum state are hence insecure, since the information carriers are lost through the channel. Instead, if reverse reconciliation is used, where Alice reconciles her raw keys to Bob’s, the rounds where Bob receives a vacuum state become secure. One can interpret Bob’s raw keys in these rounds as generated from local random numbers, and no information is known a priori in transmission. This is a common practice in usual CV QKD and in accordance with the observation in Ref. [23]. The fact that the vacuum component can also contribute to the key-rate formula was first observed in Ref. [49]. We present as Theorem 1 the key-rate lower bound with reverse reconciliation as the main key-rate formula to be used throughout this paper.

    Theorem 1. For the time-bin CV QKD protocol in Table 1 with reverse reconciliation, in the asymptotic limit of an infinite data size, the distillable secure key rate r is lower bounded by rrev, rrrev=Q*,0+m=1Qm,m[1h(em,mX)]fQZh(eZ),where Qm,m and em,mX denote the gain and the phase-error rate in the rounds where m photons are sent and m photons are accepted, QZ is the Z-basis gain, eZ is the bit-error rate, and f is the efficiency of information reconciliation. Q*,0 represents the gain of the rounds where Bob accepts a vacuum state for whatever state is sent by Alice.

    C. Phase-Error Probability Calculation

    We now evaluate the key-rate formula in Eq. (12) with Fock-basis observables [5]. The bit-error rate eZ can be directly measured, as the Z-measurement statistics in the entanglement-based squashing model are the same as the realistic statistics. To evaluate the gains and phase-error probabilities, we first determine the state before the phase-error measurement under each photon-number tag. Define P^m as the projector onto the m-photon state on modes A1 and A2. When sending m photons, the source in Fig. 2(b) collapses to P^mA1A2ρ^AA1A2P^mA1A2=Prμ(m)|ΨmAA1A2Ψm|,where |ΨmAA1A2=12(|0A|0mA1A2+|1A|m0A1A2),Prμ(m)=eμμmm!.Upon transmitting the m-photon state, |ΨmAA1A2, the n-photon state is selected on the detection side after the squashing channel, FrandB1B2B{P^nB1B2NEA1A2B1B2[Prμ(m)|ΨmAA1A2Ψm|]P^nB1B2}=Qm,nρ^AB(m,n),where NEA1A2B1B2 represents Eve’s channel, and Qm,n denotes the probability of sending an m-photon state and accepting an n-photon state, namely, the gain for the states tagged by the photon-number tuple, (m,n). Note that the probability that Bob aborts the signal is reflected in Qm,n. The normalized state, ρ^AB(m,n), is a bipartite qubit, with which we evaluate the phase-error probability, em,nX=Tr[ρ^AB(m,n)(|+AB+|+|+AB+|)].With respect to the complementary-basis measurement result on the qubit A, + or , the state on modes A1 and A2 collapses to |Ψm±A1A2=12(|0m±|m0)A1A2with equal probabilities. For the state on Bob’s systems B1 and B2 under tag (m,n), ρ^B1B2(m,n), the statistics of the complementary measurement are given by B±|FrandB1B2B[ρ^B1B2(m,n)]|±B=Tr[ρ^B1B2(m,n)M^±]=Tr[ρ^B1B2(m,n)P^nM^±P^n],where M^±=12R0dq1dq202πdφb2π[|q1(φb),q2(φb)±|q2(φb),q1(φb)][q1(φb),q2(φb)|±q2(φb),q1(φb)|].In the last equation in Eq. (18), we utilize the fact that ρ^B1B2(m,n) acts on the n-photon space of system B1B2. Combining the above results, we can express the phase-error rate for each tag with observables on optical modes:

    Proposition 1. The phase-error rate em,nX of the rounds where m photons are sent and n photons are accepted can be calculated by Qm,nem,nXPrμ(m)=12Tr[NEA1A2B1B2(|Ψm+A1A2Ψm+|)P^nM^P^n+NEA1A2B1B2(|ΨmA1A2Ψm|)P^nM^+P^n],where Qm,n denotes the probability of sending an m-photon state and accepting an n-photon state, and Prμ(m) is the probability of the source emitting m photons. NE denotes Eve’s channel on the two optical modes and P^n denotes the projector onto the n-photon subspace. |Ψm± and M^± are defined in Eq. (17) and Eq. (19) respectively.

    We can write P^nM^+()P^n on the Fock basis using Eq. (9), expressing the phase-error rate as quantities on optical modes. Here, we list the final results for a protocol that utilizes up to the two-photon components. The detailed calculation is placed in Appendix C. For the single-photon component, Q1,1e1,1XPrμ(1)=c12{Tr[NEA1A2B1B2(|Ψ1+A1A2Ψ1+|)12(|01B1B2|10B1B2)(01|B1B210|B1B2)]+Tr[NEA1A2B1B2(|Ψ1A1A2Ψ1|)12(|01B1B2+|10B1B2)(01|B1B2+10|B1B2)]},where c1=R0dq1dq2[ψ02(q1)ψ12(q2)+ψ02(q2)ψ12(q1)],and gain Q1,1 is given by Q1,1Prμ(1)=c1Tr{NE[TrA(|Ψ1AA1A2Ψ1|)](|01B1B201|+|10B1B210|)}.Up to the less-than-unity factor c1 that arises from the data post-selection in key mapping, the formulae are the same as the complementary-basis result in the coherent-state-based BB84 protocol [50]. It can also be seen that the phase-error rate of the rounds where Alice transmits two photons and Bob accepts one photon involves the probability where Alice transmits (|02±|20)/2 and Bob receives (|01|10)/2. In a pure-loss channel, the superimposed two-photon state (|02±|20)/2 would lose coherence if one photon is lost during the channel, thus giving 50% phase-error rate. This observation validates the intuition of the PNS attack. For the two-photon subspace, based on Eqs. (9) and (20), we have Q2,2e2,2XPrμ(2)=12c2Tr[NEA1A2B1B2(|Ψ2+A1A2Ψ2+|)12(|02B1B2|20B1B2)(02|B1B220|B1B2)]+12c2+Tr[NEA1A2B1B2(|Ψ2A1A2Ψ2|)12(|02B1B2+|20B1B2)(02|B1B2+20|B1B2)]+c211Tr[NEA1A2B1B2(|Ψ2A1A2Ψ2|)|11B1B211|],where c2+=R0dq1dq2[ψ0(q1)ψ2(q2)+ψ2(q1)ψ0(q2)]2,c2=R0dq1dq2[ψ0(q1)ψ2(q2)ψ2(q1)ψ0(q2)]2,c211=R0dq1dq2[2ψ12(q1)ψ12(q2)],and the two-photon gain is given by Q2,2Prμ(2)=c2+Tr{NEA1A2B1B2[TrA(|Ψ2AA1A2Ψ2|)]12(|02B1B2+|20B1B2)(02|B1B2+20|B1B2)}+c2Tr{NEA1A2B1B2[TrA(|Ψ2AA1A2Ψ2|)]12(|02B1B2|20B1B2)(02|B1B220|B1B2)}+2c211Tr{NEA1A2B1B2[TrA(|Ψ2AA1A2Ψ2|)]|11B1B211|}.The probability of accepting a vacuum state when employing reverse reconciliation is given by Q*,0=c0Tr[P^0B1B2NEA1A2B1B2(ρ^Z)P^0B1B2],where c0=R02ψ02(q1)ψ02(q2)dq1dq2,and ρ^Z is the Z-basis state sent by the source given in Eq. (8); hence Q*,0 is given by the product of the probability of receiving a vacuum-state in the Z-basis rounds and a post-selection-related integration factor. Note that the former value is independent of the post-selection.

    4. PARAMETER ESTIMATION AND PRACTICAL PROTOCOL

    We briefly show how to estimate the parameters derived in Section 3.C with a practical setup. In the actual protocol, we do not have photon-number-resolving detectors, with which one can directly measure the above parameters. In addition, the phase-error probabilities and gains are defined by particular Fock-basis states, yet the actual photon source emits coherent states. Nevertheless, we can construct unbiased estimators with the available states and detection settings to evaluate these values. On the detection side, we apply the homodyne tomography technique to evaluate the photon-number observables [3641]. The homodyne tomography allows unbiased estimation of the expected value of a variety of observables, including the photon-number observables, of measuring an unknown quantum state. On the source side, we extend the decoy-state method [29,30,51] to evaluate the statistics defined by the non-classical Fock states with the use of the coherent states at hand. We will give a practical version of the protocol at the end of this section. A fully detailed discussion is placed in Ref. [52] on how the specific parameters related to key rate calculation can be practically estimated.

    A. Effective Photon-Number Resolving via Homodyne Tomography

    Since the eigenstates of the quadrature observables, |q(φ), form a complete basis on an optical mode, one can reconstruct a general observable on an optical mode with homodyne measurements. In our study, the parameters to be estimated involve photon-number measurements on two modes in the form of O^1O^2=|n1B1m1||n2B2m2|. Their measurements on an arbitrary state, ρ^, can be obtained from two independent homodyne measurements with randomized LO phases, O^1O^2=Tr[(O^1O^2)ρ^]0πdφ1πdq10πdφ2πdq2R[O^1](q1,φ1)R[O^2](q2,φ2)p(q1,q2|φ1,φ2),where p(q1,q2|φ1,φ2) is the joint probability of the quadrature measurements on the two modes conditioned on phases φ1 and φ2. The estimators, R[O^1](q1,φ1) and R[O^2](q2,φ2), link the quadrature measurement statistics with O^1O^2. For a homodyne detector with efficiency η, the estimator for observable |nn+d| is given by Rη[|nn+d|](q,φ)=eid(φ+π2)n!(n+d)!dk|k|exp(12η2ηk2ikq)kdLnd(k2),where Lnd is the generalized Laguerre polynomial. The estimator is shown to be bounded for detector efficiency η>1/2 [39,41], a mild requirement for current technologies [53,54]. Consequently, repeated measurements allow the users to obtain an unbiased estimation of the photon-number observables that converges in probability. Note that the detector imperfection does not need to be trusted. The homodyne tomography is valid as long as the detector is well-calibrated so that the quadrature measurement is genuine. In Ref. [52], we shall provide more details of the homodyne tomography techniques.

    B. Generalized Decoy-State Method

    To effectively realize the non-classical states on the source side, we extend the standard decoy-state method [29,30]. We take advantage of two-mode coherent states with simultaneous phase randomization on the two modes. We denote the state with phase difference φ as ρ^μφ=02πdθ2π|μ2eiθμ2eiθ||μ2ei(θ+φ)μ2ei(θ+φ)|=m=0k=0,l=0meμ(μ2)mei(lk)φk!l!(mk)!(ml)!|k,mkl,ml|,where we specify the light intensity with the subscript, μ. With proper linear combination of these states, we can effectively construct the photon-number-cat states that we are interested in. It is well-known that (|01±|10)/2 is the single-photon component of ρ^μ0(π), Prμ(1)|Ψ1+()Ψ1+()|=P^1ρ^μ0(π)P^1,where Prμ represents the Poisson distribution determined by light intensity μ, as given in Eq. (14). Thus, the estimation problem is transformed into the estimation of the single-photon yields of ρ^μ0 and ρ^μπ. For the multi-photon components (|0m±|m0)/2, a direct calculation shows Prμ(m)|Ψm+Ψm+|=P^m(ρ^μZ+2m2mk=0m1ρ^μ2πkm2m2mk=0m1ρ^μ2πkm+δ)P^m,Prμ(m)|ΨmΨm|=P^m(ρ^μZ2m2mk=0m1ρ^μ2πkm+2m2mk=0m1ρ^μ2πkm+δ)P^m,where δ=π for odd m and π/2 for even m, and ρ^μZ is the state emitted from the source in a key generation round. Consequently, the terms that define em,mX and Qm,m can be constructed from the statistics when emitting the states of ρ^μZ and ρ^μφ with φ{2πk/m,2πk/m+δ}k=0m1. Notably, the extended decoy method allows estimating the gains with the number of parameters increasing only linearly in the photon number. In later discussions, we shall utilize up to the two-photon components. Specifically, for m=2, Prμ(2)|Ψ2±Ψ2±|=P^2[ρ^μZ±(12ρ^μ0+12ρ^μπ)(12ρ^μπ2+12ρ^μ3π2)]P^2.

    One may notice in Eq. (35) that there are states outside the encoding subspace |02 and |20 being introduced, which gives Eve possibility to distinguish the X-basis states. In fact, the X basis is composed of the mixture of (|02±|20)/2 and |11. As a result, it is not possible for Eve to distinguish between the Z-basis states and the (|02±|20)/2 states of the X basis. It is possible for Eve to distinguish the |11 state, yet it does not yield knowledge on the encoded key information since it is orthogonal to the |02 and |20 space. Hence, the standard decoy argument still applies even if the parameter-estimation space consists of a direct sum of the key-encoding space and some orthogonal spaces.

    C. General Parameter Estimation under the Coherent Attack

    In this section, we discuss the security analysis and parameter estimation in the most general case. In the most general adversarial scenario, namely, the coherent attack, Eve can apply a joint quantum operation over the rounds for eavesdropping, which may correlate or even entangle the states transmitted to Bob. Eve collects all the side information leaked to her in the protocol and then guesses the legitimate users’ keys. Under such an attack, the measurement statistics obtained by Bob are generally correlated over the rounds [43].

    The complementarity-based security analysis remains valid with finite statistics under a coherent attack [22]. The information leakage is quantified via the number of phase errors, while the occurrence of a phase error in each round may be non-i.i.d. That is, one should interpret the gains and phase-error rates in Eq. (12) as frequencies in non-i.i.d. statistics. To avoid any misunderstanding, throughout this work, the term “frequency” is used in the context of probability theory and refers to the number of occurrences of an event. For instance, Q1,1 should be regarded as the frequency of the events that Alice sends a single-photon state and Bob accepts a single-photon state among key generation rounds in the virtual experiment. The remaining problem is to estimate these parameters via observed statistics.

    To tackle the non-i.i.d. parameter estimation problem, we can apply a martingale-based analysis. We shall present the details in Ref. [52]. Here, we explain its basic idea. As the starting point, in the ith round, the users can evaluate the probability of choosing some experimental setting and observing a particular event conditioned on the experimental history, including the events of sending an m-photon state and accepting an n-photon state and the occurrence of a phase error if they choose the key generation setting, and observing a particular homodyne detection result if they choose to perform the parameter estimation operations. The events’ correlations with the experimental history are inherently taken into account in the definitions of conditional probabilities. We can then set up martingales for a series of events, such as the occurrence of phase errors in each round of the virtual protocol, and link their frequencies with the associated conditional probabilities via concentration results like Azuma’s inequality [55]. Note that such concentration results work for general non-i.i.d. correlations. To tighten the estimation, we shall apply a new variant of Azuma’s inequality, Kato’s inequality [56]. Furthermore, the setting choices randomly chosen by Alice and Bob are independent of the experimental history and unknown to Eve. Therefore, conditioned on the experimental history, the probabilities of different possible events in a round are linked. For instance, the probability that the users take key generation measurements and a phase error occurs in a round is measurable via the probability that they instead take parameter estimation measurements and observe certain statistics. The relation is in the form of Eq. (20), while now the probabilities are interpreted as conditional ones that cover the correlations. The relations between conditional probabilities then link the martingales for the parameter estimation measurement with the ones for the gains and phase-error rates, completing the parameter estimation. In the end, the total number of keys that can be securely distilled from finite statistics under the coherent attack is given by a formula of the following form.

    Theorem 2 (Informal). For the CV QKD protocol with MZ,s rounds for key generation with the signal intensity s, given the failure probability in parameter estimation εpe, suppose the quantum bit error rate is eZ,s, the number of key generation rounds accepting a vacuum state is lower-bounded by M*,0Z,s_, the number of key generation rounds sending and accepting an m-photon state is lower-bounded by Mm,mZ,s_, and the m-photon phase-error rate is upper-bounded by em,mX¯. Then, given the failure probability in privacy amplification εpa, conditioned on the success of information reconciliation with efficiency f, except a total failure probability ε=εpe+εpa, the finite-size key length is lower-bounded by nM*,0Z,s_+m=1Mm,mZ,s_[1h(em,mX¯)]fMZ,sh(eZ,s)log1εpa,and the key rate is lower-bounded by rn/MZ,s.

    The term logεpa in the key-rate formula originates from the failure probability in privacy amplification [22,57]. The parameter estimation failure probability εpe comes from the use of concentration results to link frequencies and probabilities. In the asymptotic limit of infinite data size, εpe converges to zero, and the effect of εpa on the key rate becomes negligible; hence the key-rate formula degenerates to that in Eq. (12), where M*,0Z,s_MZ,sQ*,0, and Mm,mZ,s_MZ,sQm,m. In Ref. [52], we provide the details of non-i.i.d. parameter estimation and the formal description of the key-rate formula.

    D. Practical Protocol

    Combining the above ingredients, we provide a practical protocol that utilizes up to the two-photon components in Table 3. In parameter estimation, Bob applies homodyne tomography to estimate the statistics of measuring photon-number observables, including |00, (|01±|10)/2, (|02±|20)/2, and |11, on various states transmitted from the source, originally ρ^μaZ and ρ^μφa. Afterward, the users can obtain upper and lower bounds on the gains and phase-error rates by applying the extended decoy-state method.

    Practical Time-Bin CV QKD with Decoy States Using up to Two Photons

    On Alice’s side (source):

    Z-basis:

    Randomly select a key bit ka{0,1}, a phase factor φa[0,2π), and a light intensity μa{μ,ν1,ν2,0}.

    Prepare a coherent state of |0A1|μaeiφaA2 for ka=0 or |μaeiφaA1|0A2 for ka=1.

    X-basis:

    Randomly select a phase factor φa1[0,2π) and another phase factor with relative phase φa randomly in φa2φa1{0,π/2,π,3π/2}. Randomly select a light intensity μa{μ,ν1,ν2,0}.

    Prepare a coherent state of |μa/2eiφa1|μa/2eiφa2.

    Alice sends the state through an authenticated channel to Bob.

    On Bob’s side (detection):

    Z-basis:

    Randomly select a phase factor φb[0,2π).

    Use homodyne detectors with LO phases φb to measure the modes and obtain quadratures q1 and q2.

    Decode the key bit as 0 if |q1|<τ|q2|>τ, 1 if |q1|>τ|q2|<τ, and Ø otherwise.

    X-basis:

    Randomly select two phases φb1,φb2[0,π).

    Use homodyne detectors with LO phases φb1 and φb2 to measure the modes and obtain quadratures q1 and q2.

    Alice announces the light intensity in each round and relative phase between the two modes in X-basis states φa=φa2φa1.

    Alice and Bob perform basis sifting, where they obtain raw keys in the rounds; they both choose Z-basis with light intensity μa=μ and kbØ.

    Bob estimates the gains and phase-error rates from the statistics in the rounds where Alice sends the Z-basis states ρ^μaZ or X-basis states ρ^μaφa with φa{0,π/2,π,3π/2}.

    Alice and Bob perform reverse information reconciliation and privacy amplification to obtain final keys.

    In the end, we make some remarks on the protocol. Notice that in contrast to the conventional BB84-type protocols, our protocol also uses for parameter estimation the signals where Alice chooses the Z basis and Bob chooses the X basis. Alice’s announcement of the relative phase does not reveal key information since the key is encoded in the relative intensity between the two modes. We assume Alice and Bob apply continuous phase randomization, although it is only practical to use discrete random phases. The effect of the discretization requires further investigation. In addition, in the X basis, Alice only transmits coherent states with relative phases in {0,π/2,π,3π/2}. These relative phases are enough to estimate the phase-error rate of an up-to-two-photon protocol according to Eq. (35).

    5. PERFORMANCES AND COMPARISON

    (a) The solid lines illustrate the asymptotic key rates of protocols using maximal one, two, three, and four photons to generate keys. The dotted line is the linear key rate bound [58,59]. We plot the PLOB bound here. The channel and devices are assumed to be ideal with no excess noise and inefficiency. (b) The relative contribution of Qm,m, i.e., the gain of the rounds where m photons are sent and m photons are received. The m-photon contribution of the i-photon protocol is relative to the raw key rate. Each group of bars illustrates the contribution of vacuum, and one-, two-, three-, and four-photon components of the protocol at a certain distance.

    Figure 3.(a) The solid lines illustrate the asymptotic key rates of protocols using maximal one, two, three, and four photons to generate keys. The dotted line is the linear key rate bound [58,59]. We plot the PLOB bound here. The channel and devices are assumed to be ideal with no excess noise and inefficiency. (b) The relative contribution of Qm,m, i.e., the gain of the rounds where m photons are sent and m photons are received. The m-photon contribution of the i-photon protocol is relative to the raw key rate. Each group of bars illustrates the contribution of vacuum, and one-, two-, three-, and four-photon components of the protocol at a certain distance.

    It can be seen that the key rate improves as we make use of the multi-photon components. The improvement is most remarkable between the one- and two-photon protocols. This is reasonable since in the one-photon protocol, the multi-photon components are considered insecure, thus limiting the source intensity. The low source intensity would result in severe bit-error-rate and higher post-selection thresholds, which in turn suppress the key rate, while in the two-photon protocol where the two-photon components are considered secure, the limit on the source intensity can be lifted, and the bit error rate would drop, resulting in higher key rates. This is manifested in Fig. 3(b), where the single-photon protocol sees significant vacuum contribution, while the two-photon protocol, at short distances, does not. Since the vacuum component would yield a 50% bit error rate, we see the lower bit error rate of the two-photon protocol than the single-photon protocol as in Table 4.

    When we further make use of the three-photon components, the key rate as well as the source intensity still increase, yet less obviously. This is mainly because the fraction of the rounds where three photons are sent and three photons are received, decaying cubically with the channel transmittance, is not dominating, especially at longer distances. For example, we see in Fig. 3(b) that at 20 km, the contribution of the three-photon component is less than that of the single- and two-photon components, and at 40 km the three-photon component rarely has contribution to the key rate. This trend is justified further in the four-photon protocol, where in Fig. 3(b) we see the four-photon-component contribution is quite small for longer distances, and in turn the key rate of the four-photon protocol only improves marginally more than that of the three-photon protocol. Simulation shows that resorting to higher-than-four photon-number components has negligible increase to the key rate. Hence, if we consider the protocol with infinite photon-number components, the 0 km key rate is around 0.31 bit/channel, and the BB84 protocol with currently the best single-photon detector of 80% efficiency [60,61] has a 0 km key rate 0.29 bit/channel, based on the model from Ref. [48]. This key rate advantage lasts for about 1 km. Our key rate thus matches the best BB84 key rate with practically favorable devices.

    The practical performances of the two-photon time-bin CV QKD protocol are illustrated in Fig. 4. For a reasonable range of excess noise ξ from 103 to 102 with respect to channel output, the key rate decays mildly as shown in Fig. 4(a). Notice that the key rate is almost unaffected at 0 km since no noise photon is introduced to give phase error, and the bit error is almost unchanged for a negligible increase in the shot-noise variance. This demonstrates the robustness of the phase-error analysis to the excess noise.

    Practical performances of the two-photon protocol. (a) Asymptotic key rate against excess noise with respect to channel output, assuming infinite decoy levels. (b) Asymptotic key rate against misalignment, i.e., the phase-reference difference between the two optical modes generating the time-bin qubit, assuming infinite decoy levels. (c) Asymptotic key rate derived using decoy methods. The noiseless setup (blue curves) uses fixed decoy levels at 1.2×10−4, 1×10−4, and vacuum, and the optimized protocol parameters of the practical setup (red curves) are listed in Table 5. (d) Finite-size key rates with various data sizes derived using the decoy-state method under no excess noise and misalignment. The security parameter is at most 5×10−10, and the source Z/X basis setting is 1:2.

    Figure 4.Practical performances of the two-photon protocol. (a) Asymptotic key rate against excess noise with respect to channel output, assuming infinite decoy levels. (b) Asymptotic key rate against misalignment, i.e., the phase-reference difference between the two optical modes generating the time-bin qubit, assuming infinite decoy levels. (c) Asymptotic key rate derived using decoy methods. The noiseless setup (blue curves) uses fixed decoy levels at 1.2×104, 1×104, and vacuum, and the optimized protocol parameters of the practical setup (red curves) are listed in Table 5. (d) Finite-size key rates with various data sizes derived using the decoy-state method under no excess noise and misalignment. The security parameter is at most 5×1010, and the source Z/X basis setting is 1:2.

    Figure 4(b) illustrates the key rate against the mode reference misalignment, where the two optical modes generating the time-bin qubit differ by δ in the reference phases intrinsically. The misalignment in relative phases does not affect the Z basis as we encode the key bits into the relative intensities, and it only affects the X basis where the phase error is defined as the flips in relative phases. Our protocol thus has robustness against misalignment.

    Figure 4(c) illustrates the key rates of the decoy-state protocol in Section 4.B. We set one decoy level at vacuum, and heuristically optimize the two decoy intensities ν1 and ν2 and the signal intensity μ. The decoy estimations are done by linear programming with a cutoff photon number of 20 [62,63]. A detailed treatment of the decoy estimation of the two-photon protocol is placed in Ref. [52]. We see for both the noiseless setup, with no excess noise and misalignment, and the practical setup, with 103 excess noise and 5° misalignment, the four-level decoy estimation is almost exact. Especially on the two-photon components, the decoy estimation has only a 1% discrepancy for yield and no discrepancy for phase-error rate. This clearly surpassed the decoy performance of the protocol in Ref. [24], since our protocol uses simpler estimation of the phase error by identifying the principal components in key generation. The optimized parameters of the practical setup are listed in Table 5.

    Optimal Protocol Parameters in Generating the Key Rate Plot with 103 Excess Noise and 5° Misalignment, Using Four Decoy Levels, as in Fig. 4(c)a

    Distance (km)Signal Intensity μThreshold τDecoy Intensity ν1Decoy Intensity ν2
    01.4871.6411.737×1011.000×104
    51.1722.0493.406×1032.740×104
    100.9242.4572.993×1021.000×104
    150.9243.0681.861×1021.000×104
    200.7283.4761.355×1022.441×104
    250.7284.2911.355×1021.562×104

    The heuristically optimized signal intensity, post-selection threshold, and the decoy intensities are as given. There is one more decoy intensity set to be vacuum. The decoy estimation is done by linear programming with cutoff photon number 20.

    Table 6 demonstrates the performance comparison of our protocol with other discrete-modulated CV protocols. Compared with the protocol in Ref. [24] with similar Z-basis operations, our protocol generates a higher key rate and reaches longer distance with four-intensity-level decoy estimation. This manifests the effectiveness of our parameter estimation process since our extended decoy method imposes more accurate restriction to Eve’s attack identified by the phase-error rate. Compared with the binary-modulated single-mode CV protocol in Ref. [5], under 103 excess noise, our protocol generates about a quarter of their key rate yet reaches a longer distance without pilot-reference transmission. This could be attributed to the noise-robustness of the two-mode encoding.

    Comparison between Protocolsa

    Distance0 km10 km20 km30 km
    Ref. [24] (4-decoy and no noise)8.8×1029.8×103Below 1×104No key rate
    This work (4-decoy and no noise)1.2×1011.2×1027.3×1042.2×105
    Ref. [5] (103 excess noise)4.1×1016.2×1022.1×103No key rate
    This work (4-decoy and 103 excess noise)1.2×1011.2×1025.8×1041.2×105

    Our protocol gives higher key rate and longer distance than Ref. [24] under 4-intensity decoy estimation. Our protocol generates slightly lower key rate than the binary-modulated CV protocol in Ref. [5], yet reaches longer distance.

    Figure 4(d) illustrates the finite-size key rate when there is no channel excess noise or misalignment. The security parameter reflected by the failure probability [22] is set to be smaller than 5×1010 in this simulation. The ratio between the Z- and X-basis setting choices on the source side is fixed at 1:2. For this reason, the asymptotic key rate drops to a third of that, where the basis ratio is also optimized. Again, we use four decoy levels. We optimize their intensities and ratio settings as well as the basis setting from the detection side with respect to the distance. Given this huge parameter space, we apply efficient searching algorithms such as particle swarming with random seeding. The searching is rapid yet not guaranteed to reach the global maximum. This explains the non-smoothness in the key-rate plot. It can be seen that with a data size of 1020 rounds, the key rate approaches the asymptotic key rate. A high short-distance key rate can be reached with a reasonable data size of 1010 to 1011. Moreover, with a data size of 1012, the communication distance can almost reach the maximal distance of the asymptotic case. Our time-bin CV protocol thus enjoys good finite-size performance and is one of the very few CV schemes that admits robust finite-size analysis [5,20].

    6. CONCLUSION AND OUTLOOK

    In summary, we present the time-bin-encoding CV QKD protocol with a phase-error-based security analysis. Similar to the ideas in DV protocols [31] and other CV protocols [5], we introduce a squashing channel to “squash” the original privacy-estimation problem on two optical modes to a single qubit, enabling the definition of phase-error rate. Phase randomization on both the source and detector enables the introduction of the photon-number-tagging method, identifying the central components for key generation. Combined with the decoy-state estimation, the parameter estimation is made simple and efficient. We are thus able to obtain a finite-size analysis with decent performances under a practical setting. We expect our methods of constructing squashing models and applying phase randomization can be applied to many other CV protocols.

    In the protocol design, one of our major observations is that coherent detectors can be used to estimate the privacy of multi-photon signals. This is also pointed out in Ref. [24]. Such detectors may also be helpful to the DV protocols. In fact, we may consider a hybrid protocol: single-photon detectors for key generation and homodyne detectors for parameter estimation. The multi-photon components in this protocol can contribute to key generation compared with the single-photon BB84 protocol.

    To evaluate the finite-size performance of this CV protocol, we provide a detailed analysis based on martingale theory, which is valid against coherent attacks. Notably, combined with the photon-number tagging method, the phase-error approach greatly simplifies the finite-size analysis. From the numerical simulation under practical security parameters, it can be seen that our protocol enjoys a robust finite-size performance with a reasonable data size requirement for practical usage.

    It is tempting to further enhance the key rate and the maximal distance of this protocol. We may consider the high-dimensional time-bin encoding, which is relatively easy to implement experimentally [6466]. The high-dimensional complementarity security analysis [67] can be invoked, and the squashing channel should map the optical modes to a qudit. We can also apply the trusted-noise model to alleviate the effect of the detector noise [68,69]. The model requires the modification of the detector POVM, which is still block-diagonal on the Fock basis [24]. One may also consider using squeezed states as the light source to reduce the shot noise in one quadrature and use the other only for parameter estimation. This may tackle the large bit error rate due to the shot noise, the issue that renders the 0 km performance of our protocol not as good as the usual CV QKD scheme. We can also examine the variations of our protocol based on the combination with new DV QKD schemes such as the measurement-device-independent-type schemes [70,71] and their extensions, including the twin-field-type [7274] and the mode-pairing schemes [75,76].

    Acknowledgment

    Acknowledgment. We acknowledge insightful discussions with Hoi-Kwong Lo and Xiongfeng Ma. A.J. and R.V.P. acknowledge support from the UK EPSRC Quantum Communications Hub. A.J. acknowledges funding from Cambridge Trust. X.Z. acknowledges support from Shenzhen-Hong Kong Cooperation Zone for Technology and Innovation and Hong Kong Research Grant Council of the Research Impact Fund. P.Z. and L.J. acknowledge support from the ARO, ARO MURI, AFOSR MURI, NSF, NTT Research, Packard Foundation, and the Marshall and Arlene Bennett Family Research Program.

    APPENDIX A: REFINED KEY MAPPING SCHEME

    In the time-bin-encoded CV QKD protocol in Table 1, we consider a simple key mapping strategy with a threshold value τ illustrated in Fig. 5(a). It can be seen that the security analysis in Section 3 does not rely on the specific shapes of the key mapping regions R0 and R1, as long as they differ by a swap of the two optical modes. As a result, we can optimize R0 and R1 for a higher key-rate performance indicated by r in Eq. (12).

    The simple and the maximum-likelihood-based key mapping. The two axes represent the two homodyne measurement results (q1,q2). The yellow region is decoded as 0 and the blue decoded as 1. The gray region is discarded. (a) The simple key mapping scheme used throughout this paper, being rectangular. Point A has key mapping error 0.02% and point B 50%, yet point A is discarded while B is accepted. (b) The maximum-likelihood key mapping scheme, being approximately triangular. The post-selection is done for coherent light amplitude α=1, where only the region with a likelihood ratio greater than 10 is kept in this showcase. In this refined key mapping, point A is accepted and point B is discarded.

    Figure 5.The simple and the maximum-likelihood-based key mapping. The two axes represent the two homodyne measurement results (q1,q2). The yellow region is decoded as 0 and the blue decoded as 1. The gray region is discarded. (a) The simple key mapping scheme used throughout this paper, being rectangular. Point A has key mapping error 0.02% and point B 50%, yet point A is discarded while B is accepted. (b) The maximum-likelihood key mapping scheme, being approximately triangular. The post-selection is done for coherent light amplitude α=1, where only the region with a likelihood ratio greater than 10 is kept in this showcase. In this refined key mapping, point A is accepted and point B is discarded.

    To get a higher key rate r, we want to avoid the post-selection of the detected signal region (q1,q2) as long as the bit error rate eZ can be ensured low. To this end, we introduce a maximum-likelihood-based key mapping and analyze its performance. Corresponding to bit 0 or 1, Bob would receive coherent states |0|αeiφa or |αeiφa|0 for some randomized phase φa and α after attenuation in a pure-loss channel. Corresponding to bit values 0 and 1, the probability distributions f0 and f1 of the homodyne measurement results (q1,q2) are given by f0(q1,q2)=02πdφ4π2expq12(q22|α|cosφ)22,f1(q1,q2)=02πdφ4π2expq22(q12|α|cosφ)22,where φ is the difference between the source and detector phase randomization.

    Upon detecting a specific pair of quadratures (q1,q2), the maximum-likelihood key mapping scheme requires to decode bit 0 if f0(q1,q2)>f1(q1,q2) and bit 1 if f1(q1,q2)>f0(q1,q2). According to Eqs. (A1) and (A2), the maximum-likelihood key mapping is equivalent to the decoding of bit 0 if q12<q22 and bit 1 if q12>q22. This refined key mapping takes in detection results such as point A in Fig. 5(a) where q12 is significantly different from q22, thus reducing the post-selection loss of the gain. However, in the region where q12 and q22 are comparable, the key mapping error will be large, making a large contribution to the final bit error rate eZ. Conditioned on the detection outcome (q1,q2), the key mapping error e0Z(q1,q2) in R0 is related to the likelihood function by e0Z(q1,q2)=Pr[(q1,q2)R0|ka=1]=f1(q1,q2)Q(q1,q2),with the key mapping gain Q(q1,q2)=f0(q1,q2)+f1(q1,q2). We can similarly define e1Z(q1,q2) for the R1 region. The final bit error rate eZ is eZ=12R0dq1dq2e0Z(q1,q2)Q(q1,q2)+12R1dq1dq2e1Z(q1,q2)Q(q1,q2).

    For example, at point B in Fig. 5(a), the key mapping error is 50%. To discard the erroneous results, we can set a threshold t>1 and are required to decode bit 0 if f0(q1,q2)>tf1(q1,q2) and bit 1 if f1(q1,q2)>tf0(q1,q2). The region in between will be discarded. Figure 5(b) illustrates the numerically plotted key mapping regions given the light amplitude α=1 at t=10, where the gray region is discarded. It can be seen that the key mapping regions are nearly isosceles right triangles with non-zero intercepts ±τ with the quadrature axes. We thus present the approximately maximum-likelihood key mapping scheme as the following: kb={0if  (q2τ>±q1)(q2+τ<±q1),1if  (q1τ>±q2)(q1+τ<±q2),Øotherwise.

    As in Section 5, we regard the intercept τ as a protocol parameter. We optimize it and the source intensity μ for each distance to yield the optimal key rate. However, numerical optimization shows that this refined key mapping would only improve the key-rate performances marginally. It increases the key rate at 0 km of the ideal two-photon protocol from 0.1261 bit/channel to 0.1314 bit/channel and that of the ideal four-photon protocol from 0.3131 bit/channel to 0.3242 bit/channel, with almost no increase to the maximal transmission distance. This is due to the low gain of the newly accepted region such as point A in Fig. 5(a). Considering the experimental cost as well, we thus suggest that using the simple threshold key mapping as in Section 2 is good enough in practice.

    APPENDIX B: SIMULATION FORMULAE UNDER THERMAL-NOISE CHANNEL

    We present the simulation formulae of the asymptotic time-bin CV QKD under a thermal-noise channel with excess noise ξ from the output. We set =2 so that the vacuum variance is one. A thermal noise channel is characterized as a Gaussian completely positive map transforming the first and second moments (r¯,V), representing the mean vector and covariance matrix of the quadrature operators, of the input state as [35] r¯ηr¯,VηV+(1η)I+ξI,where η is the channel transmittance. Two thermal channels with transmittance η and η and excess noise ξ and ξ concatenate to another thermal channel with transmittance ηη and excess noise (ηξ+ξ) since r¯ηηr¯,Vη(ηV+(1η)I+ξI)+(1η)I+ξI=ηηV+(1ηη)I+(ηξ+ξ)I.

    On the bit-error side, the thermal noise can be seen as adding ξ to the unity variance of the coherent states. Hence, if Alice transmits a coherent state |μeiθ through a thermal-noise channel with transmittance η and excess noise ξ, and Bob applies homodyne detection with LO phase φ, the detection result q will follow a distribution Pr(q|μ,θφ)=12πexp{[q2ημcos(θφ)]22(1+ξ)}.Since both the signal states and the receiver LO are uniformly phase randomized, θφ is also uniformly randomized with [0,2π) in a cyclic manner. The bit error rate eμZ and the Z-basis gain QμZ can thus be calculated according to the post-selection threshold τ, uniformly randomizing over [0,2π).

    The calculation of the vacuum gain Q*,0, according to Eq. (27), requires the probability of sending the Z-basis state while receiving vacuum. This can be calculated via the Wigner function for the Gaussian state, and in specific, Tr[P^0B1B2NEA1A2B1B2(ρ^Z)P^0B1B2]=(22+ξ)2exp(2ημ2+ξ).

    The single- and two-photon gains, Q1,1 and Q2,2, and phase-error rates, e1,1X and e2,2X, are more complicated in calculation. In the infinite-decoy setup, we calculate the photon gains directly. We decompose the thermal noise ρ^th into Fock states, ρ^th=k=0k¯k(k¯+1)k+1|kk|,where k¯=ξ/2(1η) is the average photon number of the thermal noise. The optical mode from Alice can be seen as mixing with the thermal noise through an η-transmittance beam splitter. We calculate the effect of the thermal noise in an ensemble manner, that is, we calculate the case where the channel injects k and l noise photons to the two consecutive optical modes, respectively, and mix the results according to the noise-photon-number distribution in Eq. (B5). We set a cutoff photon number at Nc=3 since the thermal noise is relatively low. Simulation shows that higher cutoffs have negligible effects on the key rate. We also account for the effects of the misalignment angle δ, which introduces a sin2(mδ/2) error to the m-photon phase error rate. We ignore the correlation between the misalignment and thermal noise photon as a second-order small quantity.

    The calculations of the quantities of interest are listed below. The notation of (k,l) represents the conditional probability that the thermal sources emit k and l photons, respectively, to the two optical modes. System.Xml.XmlElementSystem.Xml.XmlElementSystem.Xml.XmlElementSystem.Xml.XmlElement

    In the finite-decoy setup, the estimations of photon gains are derived from the statistics of coherent-state gains. We need to calculate the probability of transmitting certain coherent states while receiving certain photon states. This can also be done by the Gaussian-state Wigner function. Let κ=2/(2+ξ). Denote the output of a thermal noise channel when transmitting the coherent state |α as ρα. Its Fock-basis matrix elements are 0|ρα|0=κexp(κ|α|2),1|ρα|1=κ(κ2|α|2+1κ)exp(κ|α|2),0|ρα|1=κ2α*exp(κ|α|2),2|ρα|2=κ[12κ4|α|4+2(κ2κ3)|α|2+(1κ2)]exp(κ|α|2),0|ρα|2=12κ3(α*)2exp(κ|α|2).The statistics required by the decoy method are all based on the gains of separable coherent states. For example, the probability of sending |α|β while receiving (|02+|20)/2 can be computed by 12(02|+20|)ραρβ(|02+|20)=12(0|ρα|02|ρβ|2+2|ρα|20|ρβ|0+0|ρα|22|ρβ|0+2|ρα|00|ρβ|2).

    APPENDIX C: PHASE-ERROR CALCULATION DETAILS

    We give the detailed derivation of the phase-error probability expressions Eqs. (21)–(27) for the zero-, one-, and two-photon components. According to Eq. (20), the calculation involves expanding the X-basis measurement operator M± based on Eq. (9). For the single-photon subspace, we have P^1[02πdφ2π|q1(φ)q1(φ)||q2(φ)q2(φ)|]P^1=ψ02(q1)ψ12(q2)|0101|+ψ12(q1)ψ02(q2)|1010|+ψ0(q1)ψ0(q2)ψ1(q1)ψ1(q2)|0110|+ψ0(q1)ψ0(q2)ψ1(q1)ψ1(q2)|1001|,P^1[02πdφ2π|q2(φ)q2(φ)||q1(φ)q1(φ)|]P^1=ψ12(q1)ψ02(q2)|0101|+ψ02(q1)ψ12(q2)|1010|+ψ0(q1)ψ0(q2)ψ1(q1)ψ1(q2)|0110|+ψ0(q1)ψ0(q2)ψ1(q1)ψ1(q2)|1001|,P^1[02πdφ2π|q1(φ)q2(φ)||q2(φ)q1(φ)|]P^1=ψ02(q1)ψ12(q2)|0110|+ψ12(q1)ψ02(q2)|1001|+ψ0(q1)ψ0(q2)ψ1(q1)ψ1(q2)|0101|+ψ0(q1)ψ0(q2)ψ1(q1)ψ1(q2)|1010|,P^1[02πdφ2π|q1(φ)q2(φ)||q2(φ)q1(φ)|]P^1=ψ12(q1)ψ02(q2)|0110|+ψ02(q1)ψ12(q2)|1001|+ψ0(q1)ψ0(q2)ψ1(q1)ψ1(q2)|0101|+ψ0(q1)ψ0(q2)ψ1(q1)ψ1(q2)|1010|.Hence the single-photon phase-error operator is given by P^1M^±P^1=R0dq1dq2[ψ0(q1)ψ1(q2)±ψ0(q2)ψ1(q1)]212(|01±|10)(01|±10|)=R0dq1dq2[ψ02(q1)ψ12(q2)+ψ12(q1)ψ02(q2)]12(|01±|10)(01|±10|).Note that the second equality is deduced as the cross terms are odd functions with respect to q1 and q2, and the key-mapping region is symmetrical. This gives Eq. (21). The gain Q1,1 involves measuring P^1(M^++M^)P^1, which is clearly in the form of Eq. (23).

    The two-photon case involves more terms, but we can make use of the symmetry of the key mapping region R0 to eliminate the odd terms. The calculation goes by P^2[02πdφ2π|q1(φ)q1(φ)||q2(φ)q2(φ)|]P^2=ψ02(q1)ψ22(q2)|0202|+ψ0(q1)ψ0(q2)ψ2(q1)ψ2(q2)|0220|+ψ0(q1)ψ0(q2)ψ2(q1)ψ2(q2)|2002|+ψ02(q2)ψ22(q1)|2020|+ψ12(q1)ψ12(q2)|1111|+oddterms.The X-basis measurement is thus given by P^2M^+P^2=R0dq1dq2[ψ0(q1)ψ2(q2)+ψ2(q1)ψ0(q2)]212(|02+|20)(02|+20|)+R0dq1dq22ψ12(q1)ψ12(q2)|1111|,P^2M^P^2=R0dq1dq2[ψ0(q1)ψ2(q2)ψ2(q1)ψ0(q2)]212(|02|20)(02|20|).This recovers Eq. (24), and adding the two equations together gives the expression for the gain Q2,2 as in Eq. (26). Expanding the vacuum subspace according to Eq. (9) gives the coefficients as in Eq. (27).

    [1] C. H. Bennett, G. Brassard. Quantum cryptography: public key distribution and coin tossing. Proceedings of the IEEE International Conference on Computers, Systems and Signal Processing, 175-179(1984).

    [4] J. Lin, T. Upadhyaya, N. Lütkenhaus. Asymptotic security analysis of discrete-modulated continuous-variable quantum key distribution. Phys. Rev. X, 9, 041064(2019).

    [13] D. B. Soh, C. Brif, P. J. Coles. Self-referenced continuous-variable quantum key distribution protocol. Phys. Rev. X, 5, 041010(2015).

    [18] B. Qi, P. Lougovski, R. Pooser. Generating the local oscillator “locally” in continuous-variable quantum key distribution based on coherent detection. Phys. Rev. X, 5, 041009(2015).

    [27] D. Gottesman, H.-K. Lo, N. Lütkenhaus. Security of quantum key distribution with imperfect devices. Quantum Info. Comput., 4, 325-360(2004).

    [28] X. Ma. Quantum cryptography: from theory to practice(2008).

    [32] G. M. D’Ariano, L. Maccone, M. F. Sacchi. Homodyne tomography and the reconstruction of quantum states of light. Quantum Information with Continuous Variables of Atoms and Light, 141-158(2007).

    [41] G. D’Ariano. Tomographic measurement of the density matrix of the radiation field. J. Eur. Opt. Soc. B, 7, 693(1995).

    [49] H.-K. Lo. Getting something out of nothing. Quantum Inf. Comput., 5, 413-418(2005).

    [55] K. Azuma. Weighted sums of certain dependent random variables. Tohoku Math. J. Second Ser., 19, 357-367(1967).

    [73] X. Ma, P. Zeng, H. Zhou. Phase-matching quantum key distribution. Phys. Rev. X, 8, 031043(2018).

    Tools

    Get Citation

    Copy Citation Text

    Anran Jin, Xingjian Zhang, Liang Jiang, Richard V. Penty, Pei Zeng, "Pilot-reference-free continuous-variable quantum key distribution with efficient decoy-state analysis," Photonics Res. 13, 2013 (2025)

    Download Citation

    EndNote(RIS)BibTexPlain Text
    Save article for my favorites
    Paper Information

    Category: Research Articles

    Received: Nov. 4, 2024

    Accepted: Mar. 13, 2025

    Published Online: Jul. 18, 2025

    The Author Email: Pei Zeng (peizeng@uchicago.edu)

    DOI:10.1364/PRJ.546543

    CSTR:32188.14.PRJ.546543

    Topics