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

Rigorous characterization method for photon-number statistics

Open Access Open Access

Abstract

Characterization of photon statistics of a light source is one of the most basic tools in quantum optics. Existing methods rely on an implicit and unverifiable assumption that the source never emits too many photons to stay within the measuring range of the detectors. As a result, they fail to fulfill the demand arising from emerging applications of quantum information such as quantum cryptography. Here, we propose a characterization method using a conventional Hanbury-Brown-Twiss setup to produce rigorous bounds on emission probabilities of low photon numbers from an unknown source. As an application, we show that our characterization method can be used for a practical light source in a quantum key distribution protocol to forsake the commonly used a priori assumption without significant change in efficiency. Our versatile and flexible formula for rigorous bounds will make an essential contribution to the optics toolbox in the era of quantum information.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Autocorrelation measurement of a light source has been known to be a convenient method for investigating the characteristics of the source. It dates back to Hanbury-Brown and Twiss (HBT) who measured [1, 2] the correlation of photocurrents from two photodetectors shone by a common light source to determine its second-order intensity correlation function. The meaning of the correlation functions in quantum optics was clarified by Glauber [3, 4], who showed that the second-order correlation function G(2)(r1,r2;τ)=I(r1,t)I(r2,t+τ) can also be determined from a HBT setup with two photon detectors by measuring the rate of coincidence counts. It was successfully used in the direct observation of the non-classical property of light [5]. In the case of a pulsed light, a proper integration around τ = 0 and a normalization lead to a normalized factorial moment g(2)(0)=n(n1)/n2 of the photon number n in the pulse [6]. Since g(2)(0)=0 is achieved only by an ideal single photon source (SPS), the measurement of g(2)(0) by the HBT setup has been widely adopted for the characterization of experimentally developed SPSs [7, 8]. Extension of the method to higher-order moments g(r)(0) is also straightforward by increasing the number of detectors to r [9, 10].

Despite the apparent success for the non-classical light sources, the above characterization method is not particularly suited to the emergent applications in quantum information. The most severe drawback is the fact that the HBT setup with conventional on-off (threshold) photon detectors provides the accurate value of the factorial moment only in the limit of low detection efficiencies. Here, on-off detectors such as avalanche photodiodes in Geiger mode or superconducting nano-wire detectors respond only to the absence or presence of photons. As a result, the value of g(r)(0) from an actual experiment is only approximate, and how it is deviated from the true value is unknown. This is problematic for the applications involving the security [11, 12], for which rigorous security statements are mandatory. For the computational tasks using photons such as boson sampling [13], reliability of the outcome will eventually be ascribed to that of the constituent components including light sources. Another drawback is that we encounter the moments {g(r)(0)}r much less frequently in the quantum information theory than the photon-number probability distribution {pn}n. For example, the requirement for the light source used in the decoy-state quantum key distribution (QKD) [14–16] is given by a set of inequalities in terms of the photon-number probability distribution {pn}n [17, 18]. Hence, as a characterization method in the era of quantum information, it is vital that it provides rigorous statements over {pn}n, and preferably it is tight.

Main difficulty of the problem lies in the fact that a practical characterization scheme must use a finite size of data, such as expectation values of a finite number of observables, while the input photon number is not bounded. Some of the previous methods [19, 20] using multiple on-off detectors assumed a cutoff on the input photon number and estimated likely values of {pn} rather than their bounds. Other methods to produce rigorous statements were limited to specific properties of light such as the non-classical property [21–26] and the non-Gaussian property [27, 28]. Very recently, Kovalenko et al. [29] have discussed the problem in a more general framework, but the obtained state-independent bounds on the systematic error for {pn} were rather large and will not be suited for applications such as QKD. Here we propose a characterization method using an HBT setup to produce rigorous bounds on emission probabilities of low photon numbers and show that the derived bounds are tight enough to improve the security of a quantum key distribution protocol without significant reduction in the key rate.

 figure: Fig. 1

Fig. 1 An implementation of the D = 4 case of our characterization method. The overall detection efficiencies are given by η1=R1T2η1(det), η2=R1R2η2(det), η3=T1R3η3(det), and η4=T1T3η4(det), where Tk (k=1,,3) and Rk (k=1,,3) are transmittance and reflectance, respectively.

Download Full Size | PDF

2. Characterization method

2.1. Formalism for constructing the rigorous bounds

Our proposed method uses a generalized HBT setup with D photon detectors, where the light pulses to be characterized are divided into D parts by beam splitters. We assume that each detector is an on-off detector with quantum efficiency ηi(det), which is modeled as reporting a presence of nonzero photons after a linear absorber with transmission ηi(det). Figure 1 shows an example for D = 4. The configuration of the beam splitters is not relevant. The only relevant system parameters are the overall efficiencies, {ηi}i=1,,D, including both the branching and the detector efficiencies. We denote their average as η:=D1iηi.

 figure: Fig. 2

Fig. 2 Comparison between the true values and the bounds. Each graph depicts the true value and the bounds from the characterization method in the case of D=2,3, and 4 with η=ηi=0.025 (regardless of D) when applied to an ideal Poissonian source with pn=eμμn/n! where μ is the mean photon number. (a), p0U and p0L, (b), p1U and p1L, (c), p2U, p2L, and p2U, (d), p3U, p3L, and p3U.

Download Full Size | PDF

We define an averaged r-fold coincidence probability cobs,r to represent the observed data in the following way. We represent a subset of the D detectors by the corresponding index set WZD, where ZD:={1,2,,D}. We denote by |W| the cardinality of set W. We first define a full coincidence event at a detector set W as the case where all the |W| detectors report detection, regardless of presence or absence of detections at the remaining detectors ZDW. We denote the corresponding probability by cobs,W. The averaged r-fold coincidence probability cobs,r is then defined as the arithmetic mean of cobs,W over all the (Dr) combination of r detectors, namely, cobs,r:=(Dr)1WIrcobs,W with Ir:={WZD | |W|=r}.

Let us first consider the case where the measured pulse contains exactly n photons. We can express probabilities for various coincidence patterns in terms of the overall efficiencies {ηi}i=1,,D and combine those to construct a closed-form expression for the averaged r-fold coincidence probability cobs,r. As shown in Appendix A, the probability for the n-photon input (pn = 1) is then given by

cobs,r=cn,r:=j=0r(1)jωr,jWIj(1iWηi)n,
where we defined ωr,j:=(Djrj)/(Dr). Note that cn,r=0 if r>n. For the case of a general input pulse with distribution {pn}n, the averaged coincidences should satisfy
cobs,r=n=0pncn,r (r=1,,D).

Our goal is to find rigorous bounds on {pn}n under the constraint of Eq. (2).

The above expressions should not be confused with the standard model [30] of a photon number resolving detector based on multiple on-off detectors. For a setup with D on-off detectors, the latter considers the probability c'obs,k of reporting detection of k photons, which occurs when k on-off detectors click and the other (Dk) detectors do not click. Events with distinct values of k are exclusive and k=0Dc'obs,k=1 holds. In contrast, our definition of r-fold coincidence involves events with k>r as well as k=r, and events with distinct values of r are not exclusive.

To describe the formula for the bounds, it is convenient to introduce vectors of order D+1 as follows. Let cobs:=(1,cobs,1,,cobs,D), cn:=(1,cn,1,,cn,D), and c:=(1,1,,1). Consider a basis {cj}jS specified by the index set S={0,1,,D}, and let {dj(S)}jS be its reciprocal basis, namely, cjdi(S)=δij for i,jS. Similarly, define di(S) for S:={0,1,,D1,}. We can obtain {di(S)}i and {di(S)}i as the rows of inverses of matrices C:=(c0T,c1T,,cDT) and C:=(c0T,c1T,,cD1T,cT), respectively. We postulate that the variations among the efficiencies {ηi}i are moderate, and they satisfy

iWηi<iWηi  if |W|<|W|
for any W,WZD. Now our main result is stated in the form of the following theorem.

Theorem 1 Under the condition (3), the constraint (2) implies

pnpnU:={cobsdn(S) (Dn:even)(4)cobsdn(S) (Dn:odd)(5)
pnpnL:={cobsdn(S) (Dn:even)(6)cobsdn(S) (Dn:odd)(7)
for n=0,,D1 and
n=DpnpDU:=cobsdD(S)
n=DpnpDL:=cobsd(S).

An analytical proof of Theorem 1 is given in Appendix B. Calculation of the bounds pnL and pnU in Eqs. (4)-(9) is straightforward because C and C are triangular matrices. Each bound is given as a linear function of {cobs,r}r with their coefficients written in terms of {ηi}i. It is flexible and versatile since the formula for an arbitrary number D of detectors is given in a closed-form expressionand the detection efficiencies do not need to be uniform. The explicit forms of the derived formulas are given in Appendix C. In Fig. 2, we plotted the expected numerical values of these bounds along with the true values when we characterize an ideal Poissonian source with pn=eμμn/n!. We see that the bounds on p0 and p1 are fairly good for μ0.5 with a D = 3 setup, and so are those on p2 and p3 with a D = 4 setup.

From Fig. 2, we notice that a bound pnL,U often improves only slightly as an increment of D. The corresponding formulas like Eq. (32) to Eq. (38) and Eq. (40) to Eq. (48) in Appendix C show that the dominant O(1) terms in η do not change. In general, we can prove that pnU improves only by O(η) from a (D1)-detector setup to a D-detector setup when Dn is odd, and the same goes for pnL when Dn is even (see Appendix D). In other words, each bound pnU,L shows a major improvement for every other increment of the number D of detectors used in the setup. The origin of this unexpected behavior may be understood from the threshold or saturation behavior of the detectors, which an adversary may exploit to fool us to believe in a wrong distribution. Given a distribution {pn}n and the corresponding {cobs,r}r, one may change {pn}n only slightly by replacing a O(ηD) potion of the pulses with ones with an extremely large photon number to flood the D detectors. Through this small modification, the adversary can increase cobs,D to any value without changing the rest of {cobs,r}r. Existence of such an attack forces us to trust the observed value of cobs,D only in one direction, which results in improving only half of the bounds. The above argument tells us that, in the derivation of the rigorous bound, it is essential to take the saturation behavior of the detectors into consideration.

2.2. Optimality of the proposed bounds

The presented formulation of the bounds has several merits. First, one can make sure the optimality (nonexistence of a better bound) of the obtained bounds easily. In short, the optimality follows if all the calculated bounds are nonnegative (see Appendix E). Second, it is optimal for a broad class of light with a small mean photon number n. We consider the uniform case in the limit of η0. For weak light sources where pn rapidly decreases with n, the most severe condition for the optimality of the S-related bounds is expected to be pD1L=cobsdD1(S)0. Indeed, if a light source satisfies

pn>1D(Dn)pD1
for n=D3,D5,, the condition cobsdD1(S)0 is sufficient for the optimality. The last condition holds true for η0 if the light source satisfies
n<g(D1)(0)g(D)(0).

Equations (10) and (11) form a sufficient condition for the optimality of the S-related bounds. We can obtain the conditions for S-related bounds by replacing D with D1. The proof of these statements is given in Appendix F. Examples for sources satisfying these conditions are a coherent light source with n<1 and a thermal light source with n<1/D (see Appendix F).

2.3. Parameter ambiguity and statistical errors

The closed-form expression allows us to adapt our method easily to the cases where there are ambiguities in the values of the system parameters {ηi}i and {cobs,r}r, simply by calculating the worst-case values and by introducing confidence levels if necessary. It also helps us to estimate how the degrees of such ambiguities will affect the tightness of the bounds. As an example, we consider the D = 4 setup with a uniform efficiency η1. We assume that the actual distribution {pn}n is similar to that of a weak coherent light source or a single photon source, namely, pnpn+1 for n1. It implies p01c˜obs,1 and pnc˜obs,n for n1, where c˜obs,r:=cobs,r/cr,r. Since cr,r=r!ηr, we have Δc˜obs,r/c˜obs,r=Δcobs,r/cobs,rr(Δη/η). By applying it to the dominant terms in the equations for the D = 4 case in Appendix C, we obtain

Δp0U,Lp0p1p0(|Δcobs,1|cobs,1+|Δη|η),
ΔpnU,Lpn|Δcobs,n|cobs,n+n|Δη|η(n1).

These show that the relative errors in cobs,r and η affect the bounds only proportionally, unless p0=0. While the above relations are derived from the explicit form for D = 4, it is not difficult to show that they hold for arbitrary values of D.

 figure: Fig. 3

Fig. 3 Asymptotic secure key rates per pulse for the decoy-state BB84 protocol with various assumptions on the light source. The protocol uses pulses with three different intensities, termed signal, decoy, and vacuum. We assume that the vacuum pulses contain no photons. (a), An ideal Poissonian source with known distributions, pn=eμμn/n! for the signal pulses and pn=eμμn/n! for the decoy pulses. (b)-(d), Based on the characterization method with D=4,3,2 detectors applied to the same source. For the characterization, we assume that all the overall detection efficiencies {ηi}i have the same value η. We generated simulated values of {cobs,r} assuming η=η0:=0.1/D. For application of Theorem 1, we used all the values of η in the region [0.99η0,1.01η0], and adopted the worst-case value of R as the key rate. We assume that statistical errors in estimating cobs,r are negligible. For the protocol, we assume q=0.8, q=0.1, Y0=108 and that the channel causes a constant error of 1% regardless of the transmission. The detection rate Q and the bit error rate E are modeled as Q/q=1exp (μτ)+Y0 and QE/q=0.01(1exp (μτ))+0.5Y0, where τ is the channel transmission. Q and E are defined similarly. The mean photon numbers are optimized for each value of τ under the condition μ=μ/10.

Download Full Size | PDF

3. The decoy-state BB84 protocol with a characterized source

As an application of the proposed characterization method, we consider the characterization of the laser light source used in the decoy-state BB84 protocol [14–16], which is one of the most frequently demonstrated QKD protocols [31]. The security of the protocol has usually been proved under the a priori Poissonian assumption, namely, pn=exp (μ)μnn! with mean photon number μ. Since the Poissonian assumption involves an infinite number of conditions, it is impossible to verify it experimentally. The proposals [17, 32, 33] for the use of light sources other than lasers also rely on an infinite set of conditions. It is thus important to replace such a priori assumptions with experimentally verifiable ones [34–36]. What we seek here is to use experimentally available bounds from Theorem 1.

We focus on the decoy-state BB84 protocol using pulses with three different intensities, termed signal, decoy, and vacuum. Let {pn}n and {pn}n be the photon-number distributions of the signal and the decoy pulses, respectively. We assume that the vacuum pulses contain no photons. The crux of the decoy-state protocol is to estimate the amount of valid signals, namely, the conditional detection probability Y1 given the sender emits exactly one photon (Y1 is often called the one-photon yield). Comparison between the observed probability of the detected signal pulses, Q, and that of the detected decoy pulses, Q, establishes a lower bound Y1L on Y1. Combined with an upper bound e1U on the error probability in the single photon emission events, an asymptotic secure key rate is given by [11]

R=(qp1L+qp1L)Y1L(1H(e1U))(Q+Q)H((QE+QE)/(Q+Q)).
where q is the probability of choosing the signal pulse and E is the observed error probability for the signal pulses, with q and E similarly defined for the decoy pulses.

Under the a priori Poissonian assumption of pn=exp (μ)μnn! and pn=exp (μ)μnn! with 1>μ>μ, the bounds [37] are then given by

Y1L=p2Q/qp2Q/q(p0p2p0p2)Y0p1p2p1p2
e1U=(QE/qp0e0Y0)/(p1Y1L),
where the zero-photon yield Y0 is determined from the observed probability of the detected vacuum pulses, and e0=1/2. For a general light source which is not necessarily Poissonian, Wang et al. [18] have shown that Eqs. (15) and (16) are still valid by replacing each term by the worst-case values, provided that pnL/pnUp2L/p2Up1L/p1U holds for all n2. We can easily extend them to bounds valid when pnL/pnUp2L/p2Up1L/p1U holds only for 2nD1, which can be verified by the proposed characterization method. The new bounds are
Y1L=p2LQ/qp2UQ/q(p0Up2Lp0Lp2U)Y0p1Up2Lp1Lp2Up2LpDUp1Up2Lp1Lp2U
and
e1U=(QE/qp0Le0Y0)/(p1LY1L).

The detailed derivation is given in Appendix G.

Figure 3 shows a comparison of the asymptotic secure key rate with the a priori assumption of ideal Poissonian and those with our characterization method. To calculate these curves, we used Eqs. (15) and (16) for curve (a) and Eqs. (17) and (18) for curves (b) and (c) in Fig. 3. For curve (d) with D = 2, Eq. (17) does not hold, and we used a trivial bound shown in Appendix G. We assumed that the overall quantum efficiencies of the detectors are uniform and known with an accuracy of 1 percent. From Fig. 3, we find that the secure key rate improves as D increases, and that for D = 4 it is comparable to that with the a priori assumption of ideal Poissonian. As mentioned in Sec. 1, previous schemes for characterizing the photon number statistics of a light source either require additional assumptions, concern properties not directly relevant to QKD, or produce too loose bounds. Hence the computed key rates (b)–(d) are the first instances with their security guaranteed with no a priori assumptions on the photon number statistics of the light source.

In order to test the robustness of our proposed scheme, we also computed the same key rates when the true distribution of the light source is slightly deviated from Poissonian. The result is given in Appendix H, which shows that the key rates do not change significantly.

4. Discussion and conclusion

We have presented a characterization method for photon-number distribution of a light source and shown the explicit formula for rigorous bounds on probabilities for small photon numbers. We believe that our characterization method makes a significant contribution to the quantum optics toolbox, and is especially useful in quantum information, such as in applications involving security and in computational tasks whose outcome cannot be verified efficiently.

For the measurement of the photon-number distribution of a light source, other methods with more compact setups [38–40] are known, but so far they only give us the most probable values of the probabilities and it is by no means obvious how one may derive rigorous bounds. Homodyne detection [39] and photon-number-resolving detectors [40–42] have an advantage of responding to arrival of two or more photons. On the other hand, these devices tend to require many system parameters to model them, which may make it difficult to produce a rigorous bound. They cannot avoid saturation behavior either, namely, a photon-number-resolving detector can resolve photons only up to a certain photon number, and homodyne tomography usually introduces a threshold manually to avoid artifacts. On these grounds, we believe that the HBT setup has an advantage of being made up of components, each of which is represented by a simple model. Of course, this inevitably raises a question of how closely an actual detector behaves to the on-off detector model. An obvious deviation is the dark counting, which may be treated as an estimation problem of genuine coincidence rates from the actually observed coincidence rates including the contribution of the dark counts. It is also important to verify [43] the validity of the on-off detector model in future experimental researches.

Appendix A: coincidence probabilities for n photon input

For simplicity, we consider the case when the input optical pulse to our HBT setup is treated as a single optical mode labeled as “in”, and each of the pulses incident on the D on-off detectors is also a single optical mode. Then, our setup is modeled by a quantum channel χ from the “in” mode to D distinct optical modes, each of which is followed by a perfect on-off detector. The channel χ is a 1-to-D linear loss channel with transmissions {ηi}i=1,,D. A perfect on-off detector is modeled by positive-operator-valued measure (POVM) {πnc,πc} with πnc:=|00| for no-click and πc:=1πnc for click, where |0 represents the vacuum (zero-photon) state of the incident mode.

Let us denote by Ci the event where the i-th detector clicks, and by C¯i its complement (no-click). In terms of this notation, the full coincidence probability cobs,W at a detector set W is given by

cobs,W=Prob{iWCi}.

Using the inclusion-exclusion principle, it is rewritten in terms of no-click events as

cobs,W=1Prob{iWC¯i}=1W:WW,|W|1(1)|W|+1Prob{iWC¯i}.

For an input state ρin, we have

Prob{iWC¯i}=Tr(χ(ρin)iW|00|i).

The righthand side is the probability of observing no photons in the |W| output modes of χ. Hence, when the input state has exactly n photons (ρin=|nn|), it is given by

Prob{iWC¯i}=(1iWηi)n.

Note that this relation is valid even when the input pulse and the incident pulses to the detectors span over multiple modes as long as the efficiencies are invariant over those modes. Substituting Eq. (22) to Eq. (20), we have, for the n-photon input,

cobs,W=W:WW(1)|W|(1iWηi)n.

To calculate cobs,r, we need to take a sum of cobs,W over WIr. In this sum, a set W appears with multiplicity given by

|{WIr|WW}|=(D|W|r|W|),
which is the number of combinations to choose (r|W|) elements from ZDW. Hence, we have
(Dr)cobs,r=W:WIrcobs,W=W:|W|r(D|W|r|W|)(1)|W|(1iWηi)n,
which leads to Eq. (1).

Alternatively, it is also possible to derive the expression of cn,r in Eq. (1) by considering the case where the input distribution is Poissonian with mean μ, namely, ρin=eμn(μn/n!)|nn|. Then, the D output modes of the lossy channel χ are independent and the i-th mode is Poissonian with mean ηiμ. As a result, Eq. (19) can be directly evaluated as

cobs,W=iW(1eηiμ)=W:WW(1)|W|exp (μiWηi),
leading to
(Dr)cobs,r=W:WIrcobs,W=W:|W|r(D|W|r|W|)(1)|W|exp (μiWηi).

Since cobs,r=eμn(μn/n!)cn,r for this input, we have

nμnn!cn,r=j=0r(1)jωr,jWIjexp (μμiWηi).

Taking the n-th derivative and setting μ = 0 reproduces the definition of cn,r in Eq. (1).

Appendix B: proof of theorem 1

We first prove the inequalities (4), (7), and (8) involving S. The crux of the proof is to show that cmdn(S) has a constant sign for m>D. To do so, we focus on the property of cmdn(S) as a function of m for a fixed value of n. Let us introduce a smooth function f(m) over as

f(m):=z0+j=1D(1)jzjWIjeαWm,
with αW:=ln(1iWηi)>0. Assume that the constants {zj}j are given by zj=ωjdn(S), where ωj:=(ω0,j,ω1,j,,ωD,j) with ωr,j:=0 for r<j. From Eq. (1), we see that f(m)=cmdn(S) for nonnegative integer m. By definition of dn(S), f(n)=1 and f(m)=0 for mS{n}.

Let us show that the numbers of zeros of f(m) and its derivative f(m) are D and D1, respectively. The prerequisite (3) ensures that there exists {βj}j=1D1 such that maxWIjαW<βj<minWIj+1αW. Let us temporarily assume that all {zj}j are nonzero and have the same sign. Using the notation Dγ:=eγm(ddm)eγm, we have

Dβ1D0f(m)=j=1DWIj(1)j(β1αW)αWzjeαWm.

We see that the coefficients of eαWm for j = 1 and 2 have the same sign. In a similar vein, all the coefficients in DβD1Dβ1D0f(m) have the same sign, implying that this function has no zeros. Since multiplication of e±γm does not change zeros, Rolle’s theorem assures that f(m) should have no more than D zeros. If some of {zj}j were zero or had the opposite sign, we could remove the zeros by a fewer number of operating Dγ, contradicting the fact that f(m) has at least D zeros. Hence, we conclude that f(m) has exactly D zeros at mS{n}, and f(m) has exactly D1 zeros.

Since Rolle’s theorem also implies that the D1 zeros of f(m) must be strictly between neighboring zeros of f(m), it follows that f(m)0 for mS{n}. Hence, f(m) changes its sign across every point in S{n}. When Dn is even, f(n)=1 then implies f(m)>0 for m>D. Then we have cobsdn(S)=mpmf(m)pn, proving Eq. (4). The case with Dn being odd similarly leads to Eq. (7).

For the special case of n=D, the largest zero of f(m) lies in (D2,D1). Since f(D1)=0 and f(D)=1, we have f(m)>0 for mD1, and hence f(m)1 for mD. This leads to Eq. (8).

The proof of Eqs. (5) and (6) proceeds in a similar way. By setting zj=ωjdn(S) in Eq. (29) for a fixed value of n{0,,D1}, we see that f(m)=cmdn(S) for nonnegative integer m. Since ω0=(1,1,łdots,1)=c, we have z0=cdn(S)=0. If {zj}j=1,,D are nonzero and have the same sign, it follows that DβD1Dβ1f(m) has no zeros and f(m) has no more than D1 zeros. If some of {zj}j=1,,D are zero or have the opposite sign, f(m) has fewer zeros. Since f(m)=0 for m{0,,D1}/{n}, the number of zeros of f(m) is D1. It also means that the number of zeros of Dβ1f(m) is D2, and Dβ1f(m)0 at zeros of f(m). This property implies that f(m)=Dβ1f(m)β1f(m) is non-zero at zeros of f(m). When D1n is even, f(m) is positive for m>D1, leading to Eq. (5). When D1n is odd, f(m) is negative for m>D1, leading to Eq. (6).

Finally, the proof of Eq. (9) proceeds exactly as that of Eq. (8) in the main text by setting zj=ωjd(S), except that f(D)=1 is replaced by limmf(m)=1. Combined with f(D1)=0, we have f(m)>0 for mD1 and hence f(m)<1 for mD. This leads to Eq. (9).

Appendix C: explicit formulas for the bounds

We present explicit formulas for the bounds calculated from Theorem 1 with D=2,3, and 4. We define c˜obs,r:=cobs,r/cr,r, which reduces to c˜obs,r=cobs,r/(r!ηr) for the uniform cases of η=η1=η2=η3=η4. They are of O(1) in the limit of η0. To simplify the notations, we also define sj:=WIjiWηi/(Dj) (j=2,,D), and ξi,j:=si/(sjηij)1 (i,j=2,,D). The formula for the uniform case is simply given by setting ξi,j=0 for all i,j.

(D=2)

p0L=1c˜obs,1+2(1+ξ2,1)η(1η)c˜obs,2,
p0U=1c˜obs,1+[1(1ξ2,1)η]c˜obs,2,
p1L=c˜obs,1[2(1ξ2,1)η]c˜obs,2,
p1U=c˜obs,12(1+ξ2,1)ηc˜obs,2,
p2L=2!(1+ξ2,1)η2c˜obs,2,
p2U=c˜obs,2.

(D = 3)

p0L=1c˜obs,1+[1(12ξ2,1)η]c˜obs,2[1(33ξ3,2/2)η+(29ξ3,2/2+2ξ3,1)η2]c˜obs,3,
p0U=1c˜obs,1+[1(12ξ2,1)η]c˜obs,23(1+ξ3,2)η(13η+2(1+ξ2,1)η2)c˜obs,3,
p1L=c˜obs,1[2(12ξ2,1)η]c˜obs,2+3(1+ξ3,2)η(23η)c˜obs,3,
p1U=c˜obs,1[2(12ξ2,1)η]c˜obs,2+[3(63ξ3,2)η+(29/2ξ3,2+2ξ3,1)η2]c˜obs,3,
p2L=c˜obs,23[1(1ξ3,2/2)η]c˜obs,3,
p2U=c˜obs,23(1+ξ3,2)ηc˜obs,3,
p3L=3!(1+ξ3,1)η3c˜obs,3,
p3U=c˜obs,3.

(D=4)

p0L=1c˜obs,1+[1(13ξ2,1)η]c˜obs,2[1(33ξ3,2)η+(212ξ3,2+6ξ3,1)η2]c˜obs,3+4(1+ξ4,3)η[16η+(11+3ξ3,1)η26(1+ξ3,1)η3]c˜obs,4,
p0U=1c˜obs,1+[1(13ξ2,1)η]c˜obs,2[1(33ξ3,2)η+(212ξ3,2+6ξ3,1)η2]c˜obs,3+[1(62ξ4,3)η+(11+6ξ2,18ξ3,2/312ξ4,3+11ξ4,2/3)η2(6+24ξ2,132ξ3,2/316ξ4,3+44ξ4,2/36ξ4,1)η3]c˜obs,4,
p1L=c˜obs,1[2(13ξ2,1)η]c˜obs,2+[3(66ξ3,2)η+(212ξ3,2+6ξ3,1)η2]c˜obs,3[4(186ξ4,3)η+(22+12ξ2,116ξ3,2/324ξ4,3+22ξ4,2/3)η2(6+24ξ2,132ξ3,2/316ξ4,3+44ξ4,2/36ξ4,1)η3]c˜obs,4,
p1U=c˜obs,1[2(13ξ2,1)η]c˜obs,2+[3(66ξ3,2)η+(212ξ3,2+6ξ3,1)η2]c˜obs,34(1+ξ4,3)η[312η+(11+3ξ3,1)η2]c˜obs,4,
p2L=c˜obs,23[1(1ξ3,2)η]c˜obs,3+12(1+ξ4,3)η(12η)c˜obs,4,
p2U=c˜obs,23[1(1ξ3,2)η]c˜obs,3+[6(186ξ4,3)η+(11+6ξ2,18ξ3,2/312ξ4,3+11ξ4,2/3)η2]c˜obs,4,
p3L=c˜obs,3[42(3ξ4,3)η]c˜obs,4,
p3U=c˜obs,34(1+ξ4,3)ηc˜obs,4,
p4L=4!(1+ξ4,1)η4c˜obs,4,
p4U=c˜obs,4.

Appendix D: relation between bounds for a D-detector setup and a (D-1)-detector setup

We discuss the relation between the S-related bounds for a D-detector setup and the S-related bounds for a (D1)-detector setup. We assume that the efficiencies are uniform (ηi=η). We also assume that the factorial moment of order D,

(n)D:=nn(n1)(nD+1)pn,
is finite and cobs,r=O(ηr). We denote pn(S,D):=cobsdn(S) for a D-detector setup and pn(S,D1):=cobsdn(S) for a (D1)-detector setup. From the biorthogonality relations, we have
n=0D1cn,rpn(S,D)+p(S,D)=cobs,r
n=0D1cn,rpn(S,D1)=cobs,r
for r=1,,D1. They also formally hold true for r = 0 if we define cn,0:=1 and cobs,0:=1. From Eq. (56) with r=D, we have p(S,D)=cobs,D. By subtracting Eq. (56) from Eq. (57), we obtain
n=0D1ηrcn,r(pn(S,D1)pn(S,D))=ηrcobs,D
for r=0,,D1. Notice that the right-hand side is at most O(η). Since {ηrcn,r}n,r converge to a triangular matrix with nonzero diagonal elements for η0, we see that
pn(S,D1)pn(S,D)=O(η),
implying that the S-related bounds for (D1)-detector setup improves only by O(η) through adding another detector.

Appendix E: conditions for the optimality of the proposed bounds

The present formula allows us to check the optimality of the obtained bounds. From the first row of the equality CC1cobsT=cobsT, we have n=0Dcobsdn(S)=1. Hence, if all the S-related bounds are nonnegative, namely, if

cobsdn(S)0 (0nD),
we find that {pn}n defined as pn=cobsdn(S) for 0nD and pn = 0 for nD+1 is a probability distribution, and it fulfills Eq. (2) as is seen from the remaining rows of CC1cobsT=cobsT. The inequalities (4),(7) and (8) can thus be simultaneously saturated and no tighter bounds exist. Similarly, if
cobsdn(S)0 (0nD1)
and
cobsd(S)0,
we can construct a probability distribution defined by pn=cobsdn(S) for 0nD1, pn=cobsd(S) for n=n0, and pn = 0 otherwise. It saturates the inequalities (5),(6) and (9), while it fulfills Eq. (2) in the limit of n0, implying the optimality of Eqs. (5),(6) and (9).

Appendix F: types of light sources which can be optimally characterized

We will prove that the conditions of Eqs. (10) and (11) are sufficient for the optimality of the S-related bounds. We consider the limit of η0, and assume that the efficiencies are uniform (ηi=η). Then it is not difficult to show cn,r/ηr  n!/(nr)! for η0, since cn,r is equal to the probability for at least r out of n photons to survive to reach r detectors. We also exclude sources which are singular. More precisely, we assume that (n)D in Eq. (55) is finite and cobs,r=O(ηr).

Since cm,rηr converges to m!/(mr)!, which is a polynomial of m of order r, so does cmdn(S) to a polynomial of order at most D, which we denote by h(m). The function h(m) satisfies h(m)=0 for 0mD except for m=n and h(n)=1. The former condition means that h(m)m(m1)(mD)/(mn). The latter condition means that the normalization factor is (1)Dn/(n!(Dn)!). We thus obtain

cmdn(S)m(m1)(mD)mn(1)Dnn!(Dn)!.

Since cobsdn(S)=m0cmdn(S)pm, we have

cobsdn(S)pnmD+1m(m1)(mD)mnχnpm
for n=D1,D3,, where χn:=1/(n!(Dn)!). Comparison of its right-hand side with the one with n=D1, we see that
cobsdn(S)pnχnχD1(pD1cobsdD1(S))=pn1D(Dn)(pD1cobsdD1(S))
holds for η0. It means that if Eq. (10) holds, positivity of cobsdn(S) follows that of cobsdD1(S).

For n=D1, Eq. (63) can be rewritten as cobsdD1(S)((n)D1(n)D/(D1)!. Since g(r)(0) is (n)r/nr , we see that if Eq. (11) holds, cobsdn(S)>0 in the limit of η0. Due to (59), the whole argument is applicable to the S-related bounds just by replacing D by D1.

Next, we consider how these conditions are applied to typical light sources. For a coherent light source, we have pn=exp (n)nn/n! and g(r)(0)=1, and Eq. (11) becomes n<1 . For a thermal light source with pn=1n+1(nn+1)n , we have g(r)(0)=r! and Eq. (11) becomes n<1/D. In these two cases, we can also check that Eq. (10) holds if Eq. (11) is satisfied. On the other hand, the condition (10) is not fulfilled by a pseudo single-photon source with an exceptionally small value of p0. While the optimality may not be achieved then, practically we will seldom encounter such a source due to coupling losses in experiments.

Appendix G: derivation of Y 1L and e 1U

We justify the lower bound Y1L given in Eq. (17) and the upper bound e1U in Eq. (18). The conditions Q=qnpnYn, Q=qnpnYn and 0Yn1 leads to

Qqn=0D1pnUYn+pDU
and
Qqn=0D1pnLYn.

Combined with pnL/pnUp2L/p2U for 2nD1, we have

p2LQqp2UQqn=0,1(p2LpnUp2UpnL)Yn+p2LpDU.

Since p2L/p2Up1L/p1U, we obtain Eq. (17). For the derivation of e1U, we use QE=qnpnYnen and Yn0, leading to

QEqp0LY0e0+p1LY1e1,
and obtain Eq. (18).

For D = 2, we cannot use the bound in Eq. (17). Instead, we obtain a bound Y1L directly from Eq. (66) as

Y1L=(Qqp0UY0pDU)/p1U.

The bound in Eq. (18) is still valid for D = 2.

 figure: Fig. 4

Fig. 4 Asymptotic secure key rates per pulse for the decoy-state BB84 protocol with a light source deviated from Poissonian. The photon number distribution for the signal pulse is a uniform mixture of Poissonian with mean μ˜ over the range μ˜[0.7μ,1.3μ]. The distribution for the decoy pulse is the same except μ replaced by μ=μ/10. All the other parameters are the same as those in Fig. 3. (a’), The key rate when the distributions of the signal and the decoy pulse are known. (b’)-(d’), Based on the characterization method with D=4,3,2 detectors applied to the same source. The black curve (a) which is slightly above (a’) is the same as the curve (a) in Fig. 3, shown for helping the comparison.

Download Full Size | PDF

Apeendix H: effect of an imperfection in the light source

We consider the key rate of the decoy-state protocol when we use a light source with its photon number distribution deviated from Poissonian. We take a fluctuating laser source as an example, and assume that the squared amplitude of the coherent state is uniformly distributed over the range [0.7μ,1.3μ]. The g(2)(0) value of this source is 1.03. Assuming such sources for signal and decoy pulses, we have done otherwise the same calculation as that in Fig. 3. The result is shown in Fig. 4. Comparison between the two figures shows that our method is not specialized for the Poissonian distribution but works for different distributions.

Funding

Cross-ministerial Strategic Innovation Promotion Program (SIP) and ImPACT Program (Council for Science, Technology and Innovation [CSTI]); Photon Frontier Network Program (Ministry of Education, Culture, Sports, Science and Technology); Core Research for Evolutional Science and Technology (CREST) (JPMJCR1671); Japan Society for the Promotion of Science (KAKENHI JP18K13469).

Acknowledgments

We thank Marco Lucamarini for valuable discussions.

References

1. R. Hanbury Brown and R. Q. Twiss, “A test of a new type of stellar interferometer on sirius,” Nature 178, 1046–1048 (1956). [CrossRef]  

2. R. Hanbury Brown and R. Q. Twiss, “Interferometry of the intensity fluctuations in light - i. basic theory: the correlation between photons in coherent beams of radiation,” Proc. Royal Soc. Lond. A: Math. Phys. Eng. Sci. 242, 300–324 (1957). [CrossRef]  

3. R. J. Glauber, “The quantum theory of optical coherence,” Phys. Rev. 130, 2529–2539 (1963). [CrossRef]  

4. R. J. Glauber, “Coherent and incoherent states of the radiation field,” Phys. Rev. 131, 2766–2788 (1963). [CrossRef]  

5. H. J. Kimble, M. Dagenais, and L. Mandel, “Photon antibunching in resonance fluorescence,” Phys. Rev. Lett. 39, 691–695 (1977). [CrossRef]  

6. M. Koashi, K. Kono, T. Hirano, and M. Matsuoka, “Photon antibunching in pulsed squeezed light generated via parametric amplification,” Phys. Rev. Lett. 71, 1164–1167 (1993). [CrossRef]   [PubMed]  

7. B. Lounis and M. Orrit, “Single-photon sources,” Rep. Prog. Phys. 68, 1129 (2005). [CrossRef]  

8. S. Buckley, K. Rivoire, and J. Vuckovic, “Engineered quantum dot single-photon sources,” Rep. Prog. Phys. 75, 126503 (2012). [CrossRef]  

9. M. J. Stevens, S. Glancy, S. W. Nam, and R. P. Mirin, “Third-order antibunching from an imperfect single-photon source,” Opt. Express 22, 3244–3260 (2014). [CrossRef]   [PubMed]  

10. A. Rundquist, M. Bajcsy, A. Majumdar, T. Sarmiento, K. Fischer, K. G. Lagoudakis, S. Buckley, A. Y. Piggott, and J. Vučković, “Nonclassical higher-order photon correlations with a quantum dot strongly coupled to a photonic-crystal nanocavity,” Phys. Rev. A 90, 023846 (2014). [CrossRef]  

11. D. Gottesman, H.-K. Lo, N. Lütkenhaus, and J. Preskill, “Security of quantum key distribution with imperfect device,” Quant. Inf. Comp. 4, 325 (2004).

12. V. Dunjko, E. Kashefi, and A. Leverrier, “Blind quantum computing with weak coherent pulses,” Phys. Rev. Lett. 108, 200502 (2012). [CrossRef]   [PubMed]  

13. S. Aaronson and A. Arkhipov, “The computational complexity of linear optics,” Theory Comput. 9, 143–252 (2013). [CrossRef]  

14. W.-Y. Hwang, “Quantum key distribution with high loss: Toward global secure communication,” Phys. Rev. Lett. 91, 057901 (2003). [CrossRef]   [PubMed]  

15. X.-B. Wang, “Beating the photon-number-splitting attack in practical quantum cryptography,” Phys. Rev. Lett. 94, 230503 (2005). [CrossRef]   [PubMed]  

16. H.-K. Lo, X. Ma, and K. Chen, “Decoy state quantum key distribution,” Phys. Rev. Lett. 94, 230504 (2005). [CrossRef]   [PubMed]  

17. Y. Adachi, T. Yamamoto, M. Koashi, and N. Imoto, “Simple and efficient quantum key distribution with parametric down-conversion,” Phys. Rev. Lett. 99, 180503 (2007). [CrossRef]   [PubMed]  

18. X.-B. Wang, L. Yang, C.-Z. Peng, and J.-W. Pan, “Decoy-state quantum key distribution with both source errors and statistical fluctuations,” New J. Phys. 11, 075006 (2009). [CrossRef]  

19. J. Řeháček, Z. Hradil, O. Haderka, J. Peřina, and M. Hamar, “Multiple-photon resolving fiber-loop detector,” Phys. Rev. A 67, 061801 (2003). [CrossRef]  

20. G. Harder, C. Silberhorn, J. Rehacek, Z. Hradil, L. Motka, B. Stoklasa, and L. L. Sánchez-Soto, “Time-multiplexed measurements of nonclassical light at telecom wavelengths,” Phys. Rev. A 90, 042105 (2014). [CrossRef]  

21. J. Sperling, W. Vogel, and G. S. Agarwal, “Correlation measurements with on-off detectors,” Phys. Rev. A 88, 043821 (2013). [CrossRef]  

22. R. Filip and L. Lachman, “Hierarchy of feasible nonclassicality criteria for sources of photons,” Phys. Rev. A 88, 043827 (2013). [CrossRef]  

23. J. Sperling, M. Bohmann, W. Vogel, G. Harder, B. Brecht, V. Ansari, and C. Silberhorn, “Uncovering quantum correlations with time-multiplexed click detection,” Phys. Rev. Lett. 115, 023601 (2015). [CrossRef]   [PubMed]  

24. L. Lachman, L. Slodička, and R. Filip, “Nonclassical light from a large number of independent single-photon emitters,” Sci. Reports 6, 19760 (2016). [CrossRef]  

25. P. Obšil, L. Lachman, T. Pham, A. Lešundák, V. Hucl, M. Čížek, J. Hrabina, O. Číp, L. Slodička, and R. Filip, “Nonclassical light from large ensembles of trapped ions,” Phys. Rev. Lett. 120, 253602 (2018). [CrossRef]  

26. L. Qi, M. Manceau, A. Cavanna, F. Gumpert, L. Carbone, M. de Vittorio, A. Bramati, E. Giacobino, L. Lachman, R. Filip, and M. Chekhova, “Multiphoton nonclassical light from clusters of single-photon emitters,” New J. Phys. 20, 073013 (2018). [CrossRef]  

27. L. Lachman and R. Filip, “Quantum non-Gaussianity from a large ensemble of single photon emitters,” Opt. Express 24, 27352–27359 (2016). [CrossRef]   [PubMed]  

28. I. Straka, L. Lachman, J. Hloušek, M. Miková, M. Mičuda, M. Ježek, and R. Filip, “Quantum non-Gaussian multiphoton light,” npj Quantum Inf. 4, 4 (2018). [CrossRef]  

29. O. P. Kovalenko, J. Sperling, W. Vogel, and A. A. Semenov, “Geometrical picture of photocounting measurements,” Phys. Rev. A 97, 023845 (2018). [CrossRef]  

30. J. Sperling, W. Vogel, and G. S. Agarwal, “True photocounting statistics of multiple on-off detectors,” Phys. Rev. A 85, 023820 (2012). [CrossRef]  

31. V. Scarani, H. Bechmann-Pasquinucci, N. J. Cerf, M. Dušek, N. Lütkenhaus, and M. Peev, “The security of practical quantum key distribution,” Rev. Mod. Phys. 81, 1301–1350 (2009). [CrossRef]  

32. T. Horikiri and T. Kobayashi, “Decoy state quantum key distribution with a photon number resolved heralded single photon source,” Phys. Rev. A 73, 032331 (2006). [CrossRef]  

33. Q. Wang, X.-B. Wang, and G.-C. Guo, “Practical decoy-state method in quantum key distribution with a heralded single-photon source,” Phys. Rev. A 75, 012312 (2007). [CrossRef]  

34. Y. Zhao, B. Qi, and H.-K. Lo, “Quantum key distribution with an unknown and untrusted source,” Phys. Rev. A 77, 052327 (2008). [CrossRef]  

35. M. Lucamarini, J. F. Dynes, I. Choi, M. B. Ward, B. Frohlich, Z. L. Yuan, and A. J. Shields, “Practical security of a quantum key distribution transmitter,” in Invited Talk 9648–41, SPIE Security + Defence, Toulouse, 2015 (unpublished), (2015).

36. J. F. Dynes, M. Lucamarini, K. A. Patel, A. W. Sharpe, M. B. Ward, Z. L. Yuan, and A. J. Shields, “Testing the photon-number statistics of a quantum key distribution light source,” Opt. Express 26, 22733–22749 (2018). [CrossRef]   [PubMed]  

37. X. Ma, B. Qi, Y. Zhao, and H.-K. Lo, “Practical decoy state for quantum key distribution,” Phys. Rev. A 72, 012326 (2005). [CrossRef]  

38. G. Zambra, A. Andreoni, M. Bondani, M. Gramegna, M. Genovese, G. Brida, A. Rossi, and M. G. A. Paris, “Experimental reconstruction of photon statistics without photon counting,” Phys. Rev. Lett. 95, 063602 (2005). [CrossRef]   [PubMed]  

39. A. I. Lvovsky, H. Hansen, T. Aichele, O. Benson, J. Mlynek, and S. Schiller, “Quantum state reconstruction of the single-photon fock state,” Phys. Rev. Lett. 87, 050402 (2001). [CrossRef]   [PubMed]  

40. E. Waks, E. Diamanti, B. C. Sanders, S. D. Bartlett, and Y. Yamamoto, “Direct observation of nonclassical photon statistics in parametric down-conversion,” Phys. Rev. Lett. 92, 113602 (2004). [CrossRef]   [PubMed]  

41. D. Rosenberg, A. E. Lita, A. J. Miller, and S. W. Nam, “Noise-free high-efficiency photon-number-resolving detectors,” Phys. Rev. A 71, 061803 (2005). [CrossRef]  

42. M. Fujiwara and M. Sasaki, “Direct measurement of photon number statistics at telecom wavelengths using a charge integration photon detector,” Appl. Opt. 46, 3069–3074 (2007). [CrossRef]   [PubMed]  

43. M. Bohmann, R. Kruse, J. Sperling, C. Silberhorn, and W. Vogel, “Direct calibration of click-counting detectors,” Phys. Rev. A 95, 033806 (2017). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (4)

Fig. 1
Fig. 1 An implementation of the D = 4 case of our characterization method. The overall detection efficiencies are given by η 1 = R 1 T 2 η 1 ( det ), η 2 = R 1 R 2 η 2 ( det ), η 3 = T 1 R 3 η 3 ( det ), and η 4 = T 1 T 3 η 4 ( det ), where T k   ( k = 1 , , 3 ) and R k   ( k = 1 , , 3 ) are transmittance and reflectance, respectively.
Fig. 2
Fig. 2 Comparison between the true values and the bounds. Each graph depicts the true value and the bounds from the characterization method in the case of D = 2 , 3 , and 4 with η = η i = 0.025 (regardless of D) when applied to an ideal Poissonian source with p n = e μ μ n / n ! where μ is the mean photon number. (a), p 0 U and p 0 L, (b), p 1 U and p 1 L, (c), p 2 U, p 2 L, and p 2 U, (d), p 3 U, p 3 L, and p 3 U.
Fig. 3
Fig. 3 Asymptotic secure key rates per pulse for the decoy-state BB84 protocol with various assumptions on the light source. The protocol uses pulses with three different intensities, termed signal, decoy, and vacuum. We assume that the vacuum pulses contain no photons. (a), An ideal Poissonian source with known distributions, p n = e μ μ n / n ! for the signal pulses and p n = e μ μ n / n ! for the decoy pulses. (b)-(d), Based on the characterization method with D = 4 , 3 , 2 detectors applied to the same source. For the characterization, we assume that all the overall detection efficiencies { η i } i have the same value η. We generated simulated values of { c obs , r } assuming η = η 0 : = 0.1 / D. For application of Theorem 1, we used all the values of η in the region [ 0.99 η 0 , 1.01 η 0 ], and adopted the worst-case value of R as the key rate. We assume that statistical errors in estimating c obs , r are negligible. For the protocol, we assume q = 0.8, q = 0.1, Y 0 = 10 8 and that the channel causes a constant error of 1% regardless of the transmission. The detection rate Q and the bit error rate E are modeled as Q / q = 1 exp   ( μ τ ) + Y 0 and Q E / q = 0.01 ( 1 exp   ( μ τ ) ) + 0.5 Y 0, where τ is the channel transmission. Q and E are defined similarly. The mean photon numbers are optimized for each value of τ under the condition μ = μ / 10.
Fig. 4
Fig. 4 Asymptotic secure key rates per pulse for the decoy-state BB84 protocol with a light source deviated from Poissonian. The photon number distribution for the signal pulse is a uniform mixture of Poissonian with mean μ ˜ over the range μ ˜ [ 0.7 μ , 1.3 μ ]. The distribution for the decoy pulse is the same except μ replaced by μ = μ / 10. All the other parameters are the same as those in Fig. 3. (a’), The key rate when the distributions of the signal and the decoy pulse are known. (b’)-(d’), Based on the characterization method with D = 4 , 3 , 2 detectors applied to the same source. The black curve (a) which is slightly above (a’) is the same as the curve (a) in Fig. 3, shown for helping the comparison.

Equations (68)

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

c obs , r = c n , r : = j = 0 r ( 1 ) j ω r , j W I j ( 1 i W η i ) n ,
c obs , r = n = 0 p n c n , r   ( r = 1 , , D ) .
i W η i < i W η i   if  | W | < | W |
p n p n U : = { c obs d n ( S )   ( D n : even ) ( 4 ) c obs d n ( S )   ( D n : odd ) ( 5 )
p n p n L : = { c obs d n ( S )   ( D n : even ) ( 6 ) c obs d n ( S )   ( D n : odd ) ( 7 )
n = D p n p D U : = c obs d D ( S )
n = D p n p D L : = c obs d ( S ) .
p n > 1 D ( D n ) p D 1
n < g ( D 1 ) ( 0 ) g ( D ) ( 0 ) .
Δ p 0 U , L p 0 p 1 p 0 ( | Δ c obs , 1 | c obs , 1 + | Δ η | η ) ,
Δ p n U , L p n | Δ c obs , n | c obs , n + n | Δ η | η ( n 1 ) .
R = ( q p 1 L + q p 1 L ) Y 1 L ( 1 H ( e 1 U ) ) ( Q + Q ) H ( ( Q E + Q E ) / ( Q + Q ) ) .
Y 1 L = p 2 Q / q p 2 Q / q ( p 0 p 2 p 0 p 2 ) Y 0 p 1 p 2 p 1 p 2
e 1 U = ( Q E / q p 0 e 0 Y 0 ) / ( p 1 Y 1 L ) ,
Y 1 L = p 2 L Q / q p 2 U Q / q ( p 0 U p 2 L p 0 L p 2 U ) Y 0 p 1 U p 2 L p 1 L p 2 U p 2 L p D U p 1 U p 2 L p 1 L p 2 U
e 1 U = ( Q E / q p 0 L e 0 Y 0 ) / ( p 1 L Y 1 L ) .
c obs , W = Prob { i W C i } .
c obs , W = 1 Prob { i W C ¯ i } = 1 W : W W , | W | 1 ( 1 ) | W | + 1 Prob { i W C ¯ i } .
Prob { i W C ¯ i } = Tr ( χ ( ρ in ) i W | 0 0 | i ) .
Prob { i W C ¯ i } = ( 1 i W η i ) n .
c obs , W = W : W W ( 1 ) | W | ( 1 i W η i ) n .
| { W I r | W W } | = ( D | W | r | W | ) ,
( D r ) c obs , r = W : W I r c obs , W = W : | W | r ( D | W | r | W | ) ( 1 ) | W | ( 1 i W η i ) n ,
c obs , W = i W ( 1 e η i μ ) = W : W W ( 1 ) | W | exp  ( μ i W η i ) ,
( D r ) c obs , r = W : W I r c obs , W = W : | W | r ( D | W | r | W | ) ( 1 ) | W | exp  ( μ i W η i ) .
n μ n n ! c n , r = j = 0 r ( 1 ) j ω r , j W I j exp  ( μ μ i W η i ) .
f ( m ) : = z 0 + j = 1 D ( 1 ) j z j W I j e α W m ,
D β 1 D 0 f ( m ) = j = 1 D W I j ( 1 ) j ( β 1 α W ) α W z j e α W m .
p 0 L = 1 c ˜ obs , 1 + 2 ( 1 + ξ 2 , 1 ) η ( 1 η ) c ˜ obs , 2 ,
p 0 U = 1 c ˜ obs , 1 + [ 1 ( 1 ξ 2 , 1 ) η ] c ˜ obs , 2 ,
p 1 L = c ˜ obs , 1 [ 2 ( 1 ξ 2 , 1 ) η ] c ˜ obs , 2 ,
p 1 U = c ˜ obs , 1 2 ( 1 + ξ 2 , 1 ) η c ˜ obs , 2 ,
p 2 L = 2 ! ( 1 + ξ 2 , 1 ) η 2 c ˜ obs , 2 ,
p 2 U = c ˜ obs , 2 .
p 0 L = 1 c ˜ obs , 1 + [ 1 ( 1 2 ξ 2 , 1 ) η ] c ˜ obs , 2 [ 1 ( 3 3 ξ 3 , 2 / 2 ) η + ( 2 9 ξ 3 , 2 / 2 + 2 ξ 3 , 1 ) η 2 ] c ˜ obs , 3 ,
p 0 U = 1 c ˜ obs , 1 + [ 1 ( 1 2 ξ 2 , 1 ) η ] c ˜ obs , 2 3 ( 1 + ξ 3 , 2 ) η ( 1 3 η + 2 ( 1 + ξ 2 , 1 ) η 2 ) c ˜ obs , 3 ,
p 1 L = c ˜ obs , 1 [ 2 ( 1 2 ξ 2 , 1 ) η ] c ˜ obs , 2 + 3 ( 1 + ξ 3 , 2 ) η ( 2 3 η ) c ˜ obs , 3 ,
p 1 U = c ˜ obs , 1 [ 2 ( 1 2 ξ 2 , 1 ) η ] c ˜ obs , 2 + [ 3 ( 6 3 ξ 3 , 2 ) η + ( 2 9 / 2 ξ 3 , 2 + 2 ξ 3 , 1 ) η 2 ] c ˜ obs , 3 ,
p 2 L = c ˜ obs , 2 3 [ 1 ( 1 ξ 3 , 2 / 2 ) η ] c ˜ obs , 3 ,
p 2 U = c ˜ obs , 2 3 ( 1 + ξ 3 , 2 ) η c ˜ obs , 3 ,
p 3 L = 3 ! ( 1 + ξ 3 , 1 ) η 3 c ˜ obs , 3 ,
p 3 U = c ˜ obs , 3 .
p 0 L = 1 c ˜ obs , 1 + [ 1 ( 1 3 ξ 2 , 1 ) η ] c ˜ obs , 2 [ 1 ( 3 3 ξ 3 , 2 ) η + ( 2 12 ξ 3 , 2 + 6 ξ 3 , 1 ) η 2 ] c ˜ obs , 3 + 4 ( 1 + ξ 4 , 3 ) η [ 1 6 η + ( 11 + 3 ξ 3 , 1 ) η 2 6 ( 1 + ξ 3 , 1 ) η 3 ] c ˜ obs , 4 ,
p 0 U = 1 c ˜ obs , 1 + [ 1 ( 1 3 ξ 2 , 1 ) η ] c ˜ obs , 2 [ 1 ( 3 3 ξ 3 , 2 ) η + ( 2 12 ξ 3 , 2 + 6 ξ 3 , 1 ) η 2 ] c ˜ obs , 3 + [ 1 ( 6 2 ξ 4 , 3 ) η + ( 11 + 6 ξ 2 , 1 8 ξ 3 , 2 / 3 12 ξ 4 , 3 + 11 ξ 4 , 2 / 3 ) η 2 ( 6 + 24 ξ 2 , 1 32 ξ 3 , 2 / 3 16 ξ 4 , 3 + 44 ξ 4 , 2 / 3 6 ξ 4 , 1 ) η 3 ] c ˜ obs , 4 ,
p 1 L = c ˜ obs , 1 [ 2 ( 1 3 ξ 2 , 1 ) η ] c ˜ obs , 2 + [ 3 ( 6 6 ξ 3 , 2 ) η + ( 2 12 ξ 3 , 2 + 6 ξ 3 , 1 ) η 2 ] c ˜ obs , 3 [ 4 ( 18 6 ξ 4 , 3 ) η + ( 22 + 12 ξ 2 , 1 16 ξ 3 , 2 / 3 24 ξ 4 , 3 + 22 ξ 4 , 2 / 3 ) η 2 ( 6 + 24 ξ 2 , 1 32 ξ 3 , 2 / 3 16 ξ 4 , 3 + 44 ξ 4 , 2 / 3 6 ξ 4 , 1 ) η 3 ] c ˜ obs , 4 ,
p 1 U = c ˜ obs , 1 [ 2 ( 1 3 ξ 2 , 1 ) η ] c ˜ obs , 2 + [ 3 ( 6 6 ξ 3 , 2 ) η + ( 2 12 ξ 3 , 2 + 6 ξ 3 , 1 ) η 2 ] c ˜ obs , 3 4 ( 1 + ξ 4 , 3 ) η [ 3 12 η + ( 11 + 3 ξ 3 , 1 ) η 2 ] c ˜ obs , 4 ,
p 2 L = c ˜ obs , 2 3 [ 1 ( 1 ξ 3 , 2 ) η ] c ˜ obs , 3 + 12 ( 1 + ξ 4 , 3 ) η ( 1 2 η ) c ˜ obs , 4 ,
p 2 U = c ˜ obs , 2 3 [ 1 ( 1 ξ 3 , 2 ) η ] c ˜ obs , 3 + [ 6 ( 18 6 ξ 4 , 3 ) η + ( 11 + 6 ξ 2 , 1 8 ξ 3 , 2 / 3 12 ξ 4 , 3 + 11 ξ 4 , 2 / 3 ) η 2 ] c ˜ obs , 4 ,
p 3 L = c ˜ obs , 3 [ 4 2 ( 3 ξ 4 , 3 ) η ] c ˜ obs , 4 ,
p 3 U = c ˜ obs , 3 4 ( 1 + ξ 4 , 3 ) η c ˜ obs , 4 ,
p 4 L = 4 ! ( 1 + ξ 4 , 1 ) η 4 c ˜ obs , 4 ,
p 4 U = c ˜ obs , 4 .
( n ) D : = n n ( n 1 ) ( n D + 1 ) p n ,
n = 0 D 1 c n , r p n ( S , D ) + p ( S , D ) = c obs , r
n = 0 D 1 c n , r p n ( S , D 1 ) = c obs , r
n = 0 D 1 η r c n , r ( p n ( S , D 1 ) p n ( S , D ) ) = η r c obs , D
p n ( S , D 1 ) p n ( S , D ) = O ( η ) ,
c obs d n ( S ) 0   ( 0 n D ) ,
c obs d n ( S ) 0   ( 0 n D 1 )
c obs d ( S ) 0 ,
c m d n ( S ) m ( m 1 ) ( m D ) m n ( 1 ) D n n ! ( D n ) ! .
c obs d n ( S ) p n m D + 1 m ( m 1 ) ( m D ) m n χ n p m
c obs d n ( S ) p n χ n χ D 1 ( p D 1 c obs d D 1 ( S ) ) = p n 1 D ( D n ) ( p D 1 c obs d D 1 ( S ) )
Q q n = 0 D 1 p n U Y n + p D U
Q q n = 0 D 1 p n L Y n .
p 2 L Q q p 2 U Q q n = 0 , 1 ( p 2 L p n U p 2 U p n L ) Y n + p 2 L p D U .
Q E q p 0 L Y 0 e 0 + p 1 L Y 1 e 1 ,
Y 1 L = ( Q q p 0 U Y 0 p D U ) / p 1 U .
Select as filters


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