Optimal phase estimation with photon-number difference measurement using twin-Fock states of light
Xu J H1, Wang J Z1, Chen A X1, Li Y2, Jin G R1, †
Key Laboratory of Optical Field Manipulation of Zhejiang Province and Physics Department of Zhejiang Sci-Tech University, Hangzhou 310018, China
Beijing Computational Science Research Center, Beijing 100193, China

 

† Corresponding author. E-mail: grjin@zstu.edu.cn

Project supported by the National Natural Science Foundation of China (Grant Nos. 91636108, 11775190, and 11774024), Science Foundation of Zhejiang Sci-Tech University, China (Grant No. 18062145-Y), Open Foundation of Key Laboratory of Optical Field Manipulation of Zhejiang Province, China (Grant No. ZJOFM-2019-002), and Science Challenge Project, China (Grant No. TZ2018003).

Abstract

Quantum phase measurement with multiphoton twin-Fock states has been shown to be optimal for detecting equal numbers of photons at the output ports of a Mach–Zehnder interferometer (i.e., the so-called single-fringe detection), since the phase sensitivity can saturate the quantum Cramér–Rao lower bound at certain values of phase shift. Here we report a further step to achieve a global phase estimation at the Heisenberg limit by detecting the particle-number difference (i.e., the measurement). We show the role of experimental imperfections on the ultimate estimation precision with the six-photon twin-Fock state of light. Our results show that both the precision and the sensing region of the measurement are better than those of the single-fringe detection, due to combined contributions of the measurement outcomes. We numerically simulate the phase estimation protocol using an asymptotically unbiased maximum likelihood estimator.

1. Introduction

Optimal measurement scheme with a proper choice of data processing is important to realize high-precision phase measurement.[13] For the commonly used particle-number difference measurement (or the measurement) over quasi-classical coherent states, the phase sensitivity is constrained by the shot-noise limit (SNL) , where N is the number of particles of the input state. Various schemes have been proposed and demonstrated to improve the phase estimation precision beyond the SNL by using quantum entangled states,[4,5] e.g., the so-called NOON states . This is a maximally entangled state with all the particles being either in the mode a or in the mode b, leading to the Heisenberg limited phase sensitivity δ θ = 1/N.[36] However, the NOON states are difficult to prepare and are easily subject to the loss-induced decoherence.[7,8]

Compared with various kinds of entangled states, the twin-Fock states of photons (or atoms) |n, na,b are believed to be more robust to the decoherence and hence result in a better phase estimation precision under typical experimental noise.[9] With the twin-Fock states as the input of the Mach–Zehnder interferometer (MZI), the interferometric signal is vanishing if one performs the measurement at the output ports.[10] Therefore, various measurement schemes such as the parity detection,[1117] the measurement,[1820] and the so-called single-fringe detection[21] have been proposed to realize high-precision phase measurement beyond the SNL. These phase estimation protocols are based on the inversion estimators with their uncertainties predicted by the error-propagation formula.

According to the general theory of quantum metrology, the ultimate estimation precision is determined by the Cramér–Rao lower bound (CRB)[2225] , where is the classical Fisher information (CFI), dependent on the measurement probabilities. Furthermore, the CRB can be saturated asymptotically by maximum likelihood estimation (MLE) or Bayesian estimation.[26,27] In this paper, we investigate high-precision phase measurement with the twin-Fock states of light. We show that the measurement is optimal to achieve a global phase estimation at the Heisenberg limit. More specially, we find that the CRB saturates the quantum Cramér–Rao lower bound (QCRB) for any value of the phase shift, i.e., , where N = 2n is the number of particles and FQ = N(N + 2)/2 denotes the quantum Fisher information (QFI). This result remains valid for the measurement. To demonstrate it clearly, we use the six-photon twin-Fock state as the input and consider experimental imperfections in the measurement probabilities, similar to a recent experimental work.[21] Our results indicate that both the precision and the sensing region of the measurement are better than those of the single-fringe measurement. Finally, we numerically simulate the phase estimation protocol using the asymptotically unbiased maximum likelihood estimator.

2. Fisher information of particle-number difference measurement

For the input twin-Fock states |n, na,b, the probabilities for detecting particle-number difference (i.e., the so-called measurement) are given by

conditioned on a phase shift θ. For brevity, we have introduced Schwinger’s representation of the angular momentum , where and denote two bosonic modes and is the Pauli matrix. The unitary operator represents a sequence of actions of the 50:50 beamsplitter at the output port, the phase accumulation at the two paths, and the 50:50 beamsplitter at the input port.[26,28] In Eq. (1), the Wigner’s d-matrix elements are defined as usual (see e.g., Ref. [29]), . Hereinafter, |j,μ⟩ ≡ |n, Nna,b denote the eigenstates of , with μ = nj ∈ [−j, +j] and j = N/2, representing n particles in the a-mode and Nn particles in the b-mode. As a result, the input twin-Fock states can be expressed as |j, μ⟩ with μ = 0 and hence the output state . In Fig. 1, we show the phase-dependence of pμ(θ) for the six-particle twin-Fock state (i.e., N = 6), which satisfies pμ (−θ) = pμ(θ) and pμ(θ) = pμ(θ). As depicted by the red dashed lines, one can easily see that the probabilities p0(θ) and p± 1(θ) exhibit 6-fold oscillatory patterns.

Fig. 1. The conditional probabilities for detecting particle-number difference: pμ(θ) for (a) μ = 0, (b) μ = ± 1, (c) μ = ± 2, and (d) μ = ± 3. The red dashed lines show the ideal results; the blue solid lines give the occurrence probabilities with the experimental imperfections, obtained from Eq. (6). Numerical simulations give the statistical average of the occurrence frequency (the solid circles) and its standard derivation (the bars) of each outcome, simulated by M = 20 replicas of random numbers at each θ ∈ (−π, π). Hereinafter, denotes the statistical average.

The amount of phase information that is encoded by an arbitrary pure state is given by the quantum Fisher information , where the expectation value ⟨···⟩in is taken with respect to the input state. For the N-photon twin-Fock states, it is easy to obtain the QFI and hence the QCRB . To saturate it, we consider the measurement and calculate the CFI[2225]

which determines the CRB . The occurrence probabilities {pμ (θ)} have been defined by Eq. (1), satisfying
Similar to Refs. [30] and [31], one can note that the first-order derivations of the d-matrix elements are real, which simplify the CFI as
where we have used the completeness condition ∑μ|j,μ⟩ ⟨j,μ | = 1. For the 6-particle twin-Fock state, one can obtain the ideal result of the CFI (see the upper horizontal line of Fig. 2), independent of θ. The above results indicate that the measurement is optimal for the N-particle twin-Fock states due to δ θCRB = δθQCRBO(1/N), which holds for an arbitrary phase shift θ ∈ (−π, π). Such a Heisenberg limited phase sensitivity can be saturated by the maximum likelihood estimation (see below).

Fig. 2. The CFI for the measurement (thick solid line), the single-fringe detection (thin solid line), and the multi-fringe detection (red dashed line), obtained respectively, from Eqs. (2), (5), and (7), with . The dashed horizontal lines are the classical limit of the QFI FQ = N and the ideal result of the QFI with the twin-Fock input state FQ = N(N + 2)/2 for N = 6. The shaded area is a region where the CFI of the measurement outperforms that of the single-fringe detection.

For the twin-Fock input states, the output signal of the measurement is always vanishing,[32] implying that the inversion estimator based on the average signal cannot infer the true value of the phase shift. To overcome it, other measurement schemes have been proposed, e.g., the parity measurement[1117] and the measurement.[1820] With the error propagation formula, it has been shown that the phase sensitivity of the two measurements can only saturate the QCRB at a certain value of the phase shift (e.g., θ = 0).[21] We show that the CFI of the measurement is the same with that of the measurement, leading to a global phase estimation (see Appendix A).

Recently, Xiang et al.[21] have demonstrated a very simple and efficient single-fringe detection. Their measurement scheme has been proposed originally by Sun et al.[33] and is equivalently to a projection measurement over the output state, with the signal , corresponding to Eq. (1) for μ = 0. Note that p0(θ) describes the occurrence probability for detecting equal numbers of particles at the two output ports n1 = n2 = n, i.e., μ = (n1n2)/2 = 0, denoted hereafter by the outcome “0”. All the other detection events can be treated as a single outcome “∅”, with the occurrence probability .

Obviously, the single-fringe detection is indeed a binary-outcome measurement scheme,[3436] where the two outcomes “0” and “∅” can be obtained by coarse-graining over the original (N + 1) measurement outcomes of the measurement. For any binary-outcome measurement, it has been shown that the inversion estimator based on the average signal always saturates the CRB ,[3436] with the CFI given by

where . Using such a simple measurement scheme, Xiang et al.[21] have shown that the achievable phase sensitivity can saturate the QCRB at the optimal working points (e.g., θ = 0) due to . However, the role of the experimental imperfections reduces this ideal result, also introducing a phase-dependent Fisher information.[21]

3. Experimental imperfections

The experimental imperfections can be taken into account artificially.[37] Similar to Ref. [38], we adopt a simple method by introducing the peak heights and the visibility in the probabilities {pμ(θ)}, i.e., replacing the occurrence probabilities by

where we have introduced the θ-independent parameters A0 = 0.9293, B0 = 0.0245, B1 = 0.0087, B2 = 0.0068, and B3 = 0.0076 (see Appendix B). For brevity, we still assume , which requires B−μ = B. The above treatments ensure the normalization condition . In Fig. 1, we plot the occurrence probabilities as a function of θ (the solid blue lines), which also exhibits a multi-fold oscillatory pattern, but with the visibility less than that of the ideal results pμ(θ). For instance, as shown in Fig. 1(a), we find that the visibility of is about 94%, consistent with the experimental result of Ref. [21].

To simulate the measurement, we first generate random numbers according to at a given phase shift (see Appendix B). The occurrence numbers of all the outcomes and hence the occurrence frequencies can be obtained from the random numbers, which comprise a single phase measurement.[27] Next, we repeat the above process to obtain the average occurrence frequencies (see the circles of Fig. 1) for any value of θ ∈ (−π, π), which can be fitted by . In a real experiment, the phase-dependent occurrence probabilities can be obtained from the interferometer calibration.

Using the occurrence probabilities , we obtain the CFI of the measurement and that of the single-fringe detection , as depicted by Fig. 2. One can see that both and cannot reach the ideal result of the QFI FQ = 24. For the single-fringe detection (see the thin solid line), the phase-dependence of shows a quite good agreement with that of Ref. [21]. Maximum of the CFI is found at the optimal working point θmax = 0.29 (16.5°), with . While for the measurement (the solid thick line), appears at θmax = 0.35 (20°).

The shaded area in Fig. 2 is a region where the amount of phase information extracted by the measurement is larger than that of the single-fringe detection. Obviously, this is straightforward since the former one contains (N + 1) measurement contributions and hence . However, to find out which measurement outcomes dominate in the CFI is still nontrivial. We now consider the CFI of a multi-outcome measurement, where besides the outcome μ = 0, additional contributions from the outcomes μ = ± 2 are taken into account,

Here, the factor 2 in the second term comes from the symmetry property p−2(θ) = p2(θ). Note that all the other outcomes with μ = ± 1, ± 3 have been treated as a single outcome “∅”, with the occurrence probability p(θ) = 1 −p0(θ) − 2p2(θ).

Compared with the single-fringe case, is indeed the CFI of a multi-fringe detection.[36] Using the occurrence probabilities , we calculate the phase-dependence of , as illustrated by the red dashed line of Fig. 2. One can see that the maximum of also appears at θmax = 0.35 (20°), with . Obviously, both and are obtained by {coarse-graining} over the original (N + 1) measurement outcomes, leading to . Indeed, the coarse-graining method over a general multi-outcome measurement (i.e., combining several different outcomes into a single measurement outcome) will reduce the CFI; see Appendix C. Our results in Fig. 2 show that the outcomes μ = 0, ± 2 are of importance to achieve the best sensitivity of the measurement .

4. Maximum likelihood estimation

To saturate the CRB, we consider the well-known phase-estimation protocol based on the MLE, which is unbiased as the number of independent measurements (see e.g. Ref. [22]). Numerically, the MLE θest can be solved by maximizing the likelihood function (i.e., a multinomial distribution)

where denotes the occurrence number of each outcome μ at a given true value of phase shift θ0, and is a fit of the averaged occurrence frequency of each outcome (see Appendix B). Hereinafter, we directly use to speed up numerical simulations.

In Fig. 3, we plot the scaled phase distribution as a function of θ for θ0 = π/4. From the inset, one can see that there are four peaks of over the entire phase interval θ ∈(−π, π), and only one of them can be used to predict θ0. This leads to the so-called phase ambiguity.[3941] To avoid it, we simply introduce a prior knowledge about θ0 using the phase distribution , where P(θ) = 1 for θ ∈ (0,π/2), and 0 outside. For large enough , one can find that the phase distribution can be well approximated by a Gaussian with respect to the phase shift θ,

where σ is 68.3% confidence interval of the Gaussian around the MLE,[38] and determined by

Fig. 3. Maximum likelihood estimator. For a given true value of phase shift θ0 = π/4, the phase distribution as a function of θ (the red dashed line), which approaches to a Gaussian (solid blue line) and hence can be used to determine the MLE θest and its variance σ (i.e, 68.3% confidence interval). The inset: four peaks of over the entire phase interval θ ∈ (−π, π). The occurrence number of each outcome is obtained from random numbers.

In Fig. 4, we plot the statistical average of the phase uncertainty (see the circles) and its standard derivation (the bars) for each true value of phase shift θ0 ∈ (0,π/2). From the inset, one can note that the standard derivation of the MLE ≫ (θ0 − ⟨θests), indicating that the MLE is unbiased.[27] As a result, the phase uncertainty per measurement shows a good agreement with the CRB , see the blue solid line. Remarkably, the CRB of the measurement can beat the shot-noise limit within a phase interval, e.g., 0 < θ0 < arctan 2 (∼ 63°); while for the single-fringe detection, it is the interval (∼ 39°). This means that the quantum-enhancement sensing region of the measurement is about {twice} wider than that of the single-fringe detection (see the vertical grid lines of Fig. 4). The enlarged sensing region is important in quantum multiparticle microscopy that uses nonclassical states of light for illumination.[42]

Fig. 4. For each given true value of the phase shift θ0 ∈ (0, π/2), statistical average of (the solid circles) and its standard derivation (the bars), simulated by M = 20 replicas of random numbers. Blue solid line is the CRB of the measurement; and black dashed line is the CRB of the single-fringe measurement. The inset: statistical average of (θ0θest) as a function of θ0 and the standard derivation. Horizontal grid lines are the shot-noise limit and the ideal result of the QCRB for N = 6. The vertical grid lines give and arctan 2.

Finally, it should be mentioned that compared with other entangled states, the twin-Fock states are more robust to decoherence and give better phase estimation precision under typical experimental noise.[9] To date, small N twin-Fock states of light have been prepared using spontaneous parametric down-conversion. Nevertheless, it is still a significant challenge to generate high photon-number states with a deterministic way. Fortunately, the deterministic preparations of atomic twin-Fock states have been proposed and demonstrated using spin-1 Bose–Einstein condensates.[20,43,44] In particular, Luo et al.[44] have shown that the atomic twin-Fock states can be deterministically generated by controlled quantum phase transition. These results give a renewed interest to realize phase measurements near the Heisenberg limit with large N twin-Fock states.[45]

5. Conclusion

In summary, we have investigated an optimal phase estimation protocol with the twin-Fock states as the input of a two-mode Mach–Zendner interferometer. Our analytical results indicate that the CFI of the measurement is the same as that of the measurement and both of them saturate the QFI for any true value of phase shift, i.e., , which leads to a global phase estimation . To demonstrated it clearly, we consider the six-particle twin-Fock state as the input and numerically simulate the (N + 1) occurrence probabilities of the measurement, where the experimental imperfections are taken into account phenomenologically. Our results show that the maximum value of the CFI is about 20.1, approaching to the ideal result of the QFI FQ = 24. Furthermore, the quantum-enhancement sensing region of the measurement is about twice wider than that of the single-fringe detection, which is important to improve the global imaging quality of a quantum multiphoton microscopy.[42] Finally, we numerically simulate the phase estimation protocol using the asymptotically unbiased maximum likelihood estimator.

Reference
[1] Ma J Wang X Sun C P Nori F 2011 Phys. Rep. 509 89
[2] Pezzé L Smerzi A Oberthaler M K Schmied R Treutlein P 2018 Rev. Mod. Phys. 90 035005
[3] Dowling J P 2008 Contemp. Phys. 49 125
[4] Giovannetti V Lloyd S Maccone L 2004 Science 306 1330
[5] Durkin G A Dowling J P 2007 Phys. Rev. Lett. 99 070801
[6] Gerry C C Campos R A 2003 Phys. Rev. 68 025602
[7] Dorner U Demkowicz-Dobrzanski R Smith B J Lundeen J S Wasilewski W Banaszek K Walmsley I A 2009 Phys. Rev. Lett. 102 040403
[8] Zhang Y M Li X W Yang W Jin G R 2013 Phys. Rev. 88 043832
[9] Zhang Z Duan L M 2014 New J. Phys. 16 103037
[10] Holland M J Burnett K 1993 Phys. Rev. Lett. 71 1355
[11] Bollinger J J Itano W M Wineland D J Heinzen D J 1996 Phys. Rev. 54 R4649
[12] Gerry C C 2000 Phys. Rev. 61 043811
[13] Gerry C C Benmoussa A Campos R A 2002 Phys. Rev. 66 013804
[14] Gerry C C Mimih J 2010 Contemp. Phys. 51 497
[15] Anisimov P M Raterman G M Chiruvelli A Plick W N Huver S D Lee H Dowling J P 2010 Phys. Rev. Lett. 104 103602
[16] Chiruvelli A Lee H 2011 J. Mod. Opt. 58 945
[17] Seshadreesan K P Kim S Dowling J P Lee H 2013 Phys. Rev. 87 043833
[18] Kim T Pfister O Holland M J Noh J Hall J L 1998 Phys. Rev. 57 4004
[19] Zhang Y M Li X W Jin G R 2013 Chin. Phys. 22 114206
[20] Lücke B Scherer M Kruse J Pezzé L Deuretzbacher F Hyllus P Topic O Peise J Ertmer W Arlt J Santos L Smerzi A Klempt C 2011 Science 334 773
[21] Xiang G Y Hofmann H F Pryde G J 2013 Sci. Rep. 3 2684
[22] Helstrom C W 1976 Quantum Detection and Estimation Theory New York Academic Press
[23] Braunstein S L Caves C M 1994 Phys. Rev. Lett. 72 3439
[24] Braunstein S L Caves C M Milburn G J 1996 Ann. Phys. 247 135
[25] Paris M G A 2009 In. J. Quantum Inform. 7 125
[26] Yurke B McCall S L Klauder J R 1986 Phys. Rev. 33 4033
[27] Pezzé L Smerzi A Khoury G Hodelin J F Bouwmeester D 2007 Phys. Rev. Lett. 99 223602
[28] Gerry C C Knight P L 2005 Introductory Quantum Optics Cambridge Cambridge University Press
[29] Feng X M Wang P Yang W Jin G R 2015 Phys. Rev. 92 043307
[30] Liu P Wang P Yang W Jin G R Sun C P 2017 Phys. Rev. 95 023824
[31] Liu P Jin G R 2017 J. Phys. A: Math. Theor. 50 405303
[32] Gerry C C Campos R A Benmoussa A 2004 Phys. Rev. Lett. 92 209301
[33] Sun F W Liu B H Gong Y X Huang Y F Ou Z Y Guo G C 2008 Europhys. Lett. 82 24001
[34] Feng X M Jin G R Yang W 2014 Phys. Rev. 90 013807
[35] Ghirardi L Siloi I Bordone P Troiani F Paris M G A 2018 Phys. Rev. 97 012120
[36] Wang J Z Yang Z Q Chen A X Yang W Jin G R 2019 Opt. Express 27 10343
[37] Xiang G Y Higgins B L Berry D W Wiseman H M Pryde G J 2010 Nat. Photon. 5 43
[38] Jin G R Yang W Sun C P 2017 Phys. Rev. 95 013835
[39] Pezzé L Smerzi A 2007 Europhys. Lett. 78 30004
[40] Higgins B L Berry D W Bartlett S D Wiseman H M Pryde G J 2007 Nature 450 393
[41] Berry D W Higgins B L Bartlett S D Mitchell M W Pryde G J Wiseman H M 2009 Phys. Rev. 80 052114
[42] Israel Y Rosen S Silberberg Y 2014 Phys. Rev. Lett. 112 103604
[43] Zhang Z Duan L M 2013 Phys. Rev. Lett. 111 180401
[44] Luo X Y Zou Y Q Wu L N Liu Q Han M F Tey M K You L 2017 Science 355 620
[45] Zou Y Q Wu L N Liu Q Luo X Y Guo S F Cao J H Tey M K You L 2018 Pro. Natl. Acad. Sci. USA 115 6381