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

Bounds to precision for quantum interferometry with Gaussian states and operations

Open Access Open Access

Abstract

We address high-precision measurements by active and passive interferometric schemes based on Gaussian states and operations. In particular, we look for the best states to be injected into their ports according to the quantum Cramér–Rao bound, i.e., maximizing the quantum Fisher information over all the involved parameters, given a constraint on the overall mean number of photons entering into the interferometer. We found that for passive interferometers involving only beam splitters, the optimal input leading to Heisenberg scaling is a pair of identical squeezed-coherent states with at most one-third of the total energy employed in squeezing. For active interferometers involving optical amplifiers, input coherent signals are enough to achieve Heisenberg scaling, given an optimal value of the amplification gain. For passive schemes our results clarify the role of squeezing in improving both the reference phase and the signal phase of an interferometer.

© 2015 Optical Society of America

1. INTRODUCTION

Optical interferometry is among the most precise measurement techniques available in physics. Therefore, it got involved in the most challenging task of modern cosmology, the direct detection of gravitational waves (GWs) [1]. In an interferometer the space-time ripples induced by a GW change the optical path of the light beams traveling along the setup, thus leading to a measurable fluctuation of the phase shift. The perturbations induced by a GW are extremely small and interferometers involved in GW detection must resolve tiny phase shifts with high precision. In the last decades, the technological developments of laser physics and optical components permitted mechanical and thermal noise, i.e., the noise of classical origin, to be largely suppressed. Yet, interferometric precision is limited by the quantum nature of the radiation field used to probe the presence of GWs.

The general interferometric problem is that of maximizing the precision in the estimation of phase-shift fluctuations at a fixed energy of the involved light beams. Actually, fluctuations associated with the very quantum nature of light pose bounds to the precision. On the other hand, the nonclassical features of the probe beams may also be exploited to improve these bounds compared to the corresponding classical ones, thus leaving room for quantum enhanced interferometry. Much work has been devoted to investigating the ultimate bounds to the precision in estimating the phase shift in an interferometer and to find the states that achieve these limits [218]. Overall, two main lines of research have been pursued: the one devoted to the use of finite superpositions of number states, e.g., the so-called NOON states [19]; the other involving Gaussian states and operations, e.g., squeezing [20]. While NOON states may lead to optimal or near-optimal performances at a fixed number of particles, they are not considered of practical interest because they are very sensitive to detection noise and optical losses and their generation is still experimentally challenging.

For these reasons we focus on interferometric schemes involving Gaussian states and operations, including active and passive elements, and analyze in detail the ultimate quantum bounds to precision. Here the number of particles is not fixed and optimization is performed at a fixed energy of the beams, i.e., fixing the average number of photons. This is relevant, since the energy is directly linked to the radiation pressure, which should be under control in high precision optical interferometers.

An interferometer is a device where the phase shift occurring to one or more modes of the radiation field are mapped onto the variations of some measurable quantity at the output. In other words, interferometers aim at monitoring the stability of a given configuration against perturbations rather than at estimating the value of an unknown phase shift. The general interferometric scheme we consider is depicted in Fig. 1. It may be described as a unitary operation Uint(Θ) acting on two input modes k=a,b, [k,k]=1, excited in the two-mode initial state |ψa|ψb, followed by a phase shift Uph(ϕ)=U(ϕ)I, U(ϕ)=exp(iϕaa) imposed onto one of them. Finally, a suitable joint measurement, described by a two-mode positive operator-valued measure (POVM), {Π}, is performed on the evolved modes. The vector Θ contains the information about the parameters characterizing the interferometer. Overall, the state before the measurement stage is

|Ψ(Θ,ϕ)=Uph(ϕ)Uint(Θ)|ψa|ψb.

 figure: Fig. 1.

Fig. 1. Schematic diagram of a two-mode interferometric setup. The two modes, prepared in the factorized state |ψa and |ψb, are coupled by the unitary operator Uint(Θ). Then, one of the modes is phase-shifted by an amount ϕ and, finally, the two-mode state undergoes a measurement described by the POVM {Π}.

Download Full Size | PDF

It depends on the phase shift ϕ and on the specific features of the interferometer through the set of parameters contained in the vector Θ.

The sensitivity of an interferometer in revealing fluctuations of the phase shift ϕ is governed by the quantum Fisher information (QFI) of the family of states |Ψ(Θ,ϕ). In recent years many efforts have been made to obtain analytical expressions of the QFI for single- and multimode states in the framework of single- and multi-parameter estimation [2124] and interferometry [25]. However, given the QFI of a family of states, it is not always straightforward to find the optimal state maximizing it. In our interferometric scenario, upon evaluating analytically the QFI we will be able to optimize the interferometer over all the involved parameters and to find the optimal input states. As we will see, we found that squeezing is indeed a resource and then evaluating exactly how much squeezing is needed to achieve optimal performances in the different configurations. Our analysis extends and generalizes previous results [26], where a squeezed vacuum has been shown to be the best state for the second port of a passive interferometer, when the first port is fed by a laser source, i.e., a coherent state.

In this paper, we investigate the ultimate quantum limit to the sensitivity of interferometers based on Gaussian states and operations. In particular, we focus on two classes of interferometers, sometimes referred to as SU(2) and SU(1,1) interferometers, respectively [3]. In the first case the interaction operator Uint(Θ) belongs to the SU(2) group and describes the action of passive devices, e.g., beam splitters. The energy is thus conserved during the evolution. Conversely, SU(1,1) interferometers also involve active devices such as optical amplifiers, which add some energy to the system besides making the two modes interfere. As we will see, our analysis indicates that for passive interferometers the optimal input leading to Heisenberg scaling is a pair of identical squeezed-coherent states with at most one-third of the total energy employed in squeezing, whereas for active schemes involving optical amplifiers input coherent signals are enough to achieve Heisenberg scaling, given an optimal value of the amplification gain.

The paper is structured as follows. In the next section we briefly review the quantum Cramér–Rao bound governing the ultimate precision achievable in quantum interferometry. In Section 3 we focus on passive interferometers, whereas in Section 4 we devote attention to active ones. Section 5 closes the paper by discussing results, with some concluding remarks.

2. QUANTUM CRAMÉR–RAO BOUND

The quantum Cramér–Rao bound [2730]

δϕ21MH(ϕ),
establishes that apart from the statistical scaling (M is the number of measurements), the variance of any unbiased estimator of the phase shift, i.e., the minimum detectable fluctuation in the estimation of the parameter ϕ, is bound by the inverse of the QFI H(ϕ)[31,32]. The quantum Cramér–Rao bound is a convenient tool used to evaluate the performance of an interferometer since it applies to all quantum measurements on the two modes, and to all procedures for estimating the phase shifts from the measurement results. In general, precision depends on the value of the phase shift itself, since the QFI is given by
H(ϕ)=Tr[RϕLϕ2],
where Rϕ is the two-mode state at the output of the interferometer and Lϕ, the so-called symmetric logarithmic derivative, is the operator satisfying the Lyapunov-like equation,
ϕRϕ=12(LϕRϕ+RϕLϕ).

In the present case, where the state of the two modes is pure, i.e., Rϕ=|Ψ(Θ,ϕ)Ψ(Θ,ϕ)|, the solution of the above equation is available explicitly as

Lϕ=2(|ϕΨ(Θ,ϕ)Ψ(Θ,ϕ)|+|Ψ(Θ,ϕ)ϕΨ(Θ,ϕ)|),
and the QFI turns out to be independent of the value of the phase shift. In particular, the QFI is proportional to the fluctuations of the phase-shift generator G=aaI, namely, H=4(G2G2), where the expectation value is taken over the output state |Ψ(Θ,ϕ).

3. PASSIVE INTERFEROMETERS

In a typical passive interferometer, such as the Mach–Zehnder or the Michelson interferometer, the energy of the input |ψa|ψb is conserved during the evolution and the interaction operator Uint(Θ) corresponds to the action of a balanced beam splitter (BS), namely, UBSUint(Θ)=exp[π4(abab)]. Unbalanced devices may be also considered, which however lead to inferior performances [33,34]. The precision of this kind of interferometer depends on the choice of |ψa|ψb and on the detection stage. We take the most general pure (and factorized) Gaussian signals as input |ψa|ψb=|α,ξ|γ,ζ, where |z,s=D(z)S(s)|0 is a squeezed-coherent state D(z)=exp(zkz*k) and S(s)=exp{12[s(k)2s*k2]} are the displacement and squeezing operators of modes k=a,b, respectively. Without loss of generality, we can assume α,ξ,γR and ζ=reiθC, and for this choice the QFI HH(α,γ,ξ,r,θ) is given by

H=14{4(α+γ)2[e2ξ+cosh2r]+cosh4ξ+cosh4r+2cosθsinh2r[2(α+γ)2+sinh2ξ]+cosh2(rξ)+cosh2(ξ+r)4}.

In order to characterize the interferometer and maximize the QFI with respect to the input state it is useful to introduce the following parameterization: the total number of photons in the initial state,

Ntot=α2+γ2+sinh2r+sinh2ξ,
the signal fraction,
Δ=α2/(α2+γ2),
and the total and partial squeezing fractions,
βtot=(sinh2r+sinh2ξ)/Ntotβ=sinh2ξ/Ntot,
respectively. Note that 0Δ1, 0βtot1, and 0ββtot. Thereafter, given Ntot and βtot, the maximum QFI is achieved for Δ=1/2, β=βtot/2, and θ=0, namely, the inputs have the same coherent amplitude α=γ, and the squeezing parameter ξ=r, and explicitly reads
Hmax(Ntot,βtot)=2Ntot[2+Ntotβtot(2βtot)+2(1βtot)Ntotβtot(2+Ntotβtot)],
where Ntot=2(α2+sinh2r) and Ntotβtot=2sinh2r. We may perform a further numerical maximization of Eq. (3) with respect to βtot and the results are reported in the upper panel of Fig. 2, where one sees that the optimal value of βtot increases with Ntot. In the large energy regime Ntot1, we have βtot2/3, corresponding to
Hmax89Ntot2[2+(1+3Ntot1)1/2]+4Ntot83Ntot2.

 figure: Fig. 2.

Fig. 2. Top: optimal value of the total squeezing fraction βtot, maximizing the QFI for a passive interferometer, as a function of Ntot. Bottom: corresponding phase sensitivity δϕ as a function of Ntot (blue solid line). The shot-noise limit Ntot1/2 (red dashed line) and the Heisenberg limit Ntot1 (green dotted line) are also reported.

Download Full Size | PDF

Notice that Heisenberg scaling HNtot2 is achieved with an improved proportionality constant (compared to known results, see, e.g., [2]) due to the fact that we have optimized the input signals over all the state parameters.

It is worth noting that, for the optimal input states leading to Eq. (3), we have

UBS|α,r|α,r=|2α,r|0,r,
that is, the state undergoing the phase shift is still factorized and, in particular, the single mode acquiring the phase is the displaced-squeezed state |2α,r. This result may be better appreciated upon recalling previous findings for the single-mode case [35]. In that case, an optimal estimation of a phase shift imposed onto a Gaussian state is achieved when all the energy is used to squeeze the vacuum, i.e., the displacement is useless. The corresponding QFI reads Hsm=8Ntot(1+Ntot) [35]. One could have naïvely expected that, for an interferometric setup involving two modes, energy has be used to generate squeezing as well, leading to βtot=1 and β=1/2, where the last condition ensures that UBS|0,r|0,r=|0,r|0,r [34]. Actually, this is not the case since it would lead to Hmax=4Ntot(1+12Ntot), which follows from Hsm halving Ntot (just one mode undergoes the phase shift). The need of a displacement is due to the realistic description of the phase reference (the other mode of the interferometer), which is neglected in the single-mode analysis, where the existence of a stable classical reference is somehow assumed. It also shows that the quantum enhancement coming from the use of squeezing results from the improvement of both the reference phase and the signal phase within the interferometric setup.

The lower panel of Fig. 2 shows the behavior of the phase sensitivity δϕ=[Hmax]1/2 as a function of Ntot. We clearly observe two regimes: for Ntot1 sensitivity is shot-noise limited, namely, δϕNtot1/2, while the Heisenberg limit δϕNtot1 is reached for Ntot1. Overall, our result shows that in order to achieve optimal estimation of a phase shift using a passive interferometer and Gaussian states, we should excite the two input modes with identical signals and distribute the coherent and the squeezing photons according to the value of βtot maximizing Eq. (3). On the one hand, this choice allows one to obtain a precise phase reference [36,37], i.e., the squeezed vacuum after the interaction. On the other hand, it shows that a coherent amplitude is always needed.

The optimization over all the state parameters allows one to improve the proportionality constant in the Heisenberg scaling. In order to see this explicitly, let us consider the case where the input signals are given by |α,0|0,ζ, with ζ=reiθ, which is a coherent state and a squeezed vacuum state (we still assume αR). This is a configuration often considered for GW detectors [38] and has been recently addressed in [26] (see also [2,39]). In Ref. [26] the coherent amplitude is fixed and the optimality of a squeezed vacuum for the second port has been shown. Here, we optimize the QFI:

H(gw)(Ntot,β)=2Ntot[1+(12+Ntot)β+(1β)Ntotβ(1+Ntotβ)],
also over the squeezing fraction β=sinh2r/Ntot at fixed Ntot=α2+sinh2r. We found the maximum for β=1, corresponding to
Hmax(gw)=2Ntot(32+Ntot)Ntot12Ntot2.

4. ACTIVE INTERFEROMETERS

As a paradigmatic example of interferometers involving amplification we consider the so-called coherent-light-boosted interferometer (CLBI) [40], in which Uint(Θ)=exp(ξabξ*ab), i.e., Θ=ξ=reiθ (see Fig. 1). The interaction imposes phase-sensitive amplification (i.e. two-mode squeezing) and introduces quantum correlations between the two modes, which suggests that quantum enhancement may already be achieved starting from classical input signals. In order to check this conjecture we consider as input just a couple of coherent states |ψa|ψb=|α|γ, with α,γR. This also makes the interferometer feasible with current technology, since it minimizes the impact of phase-matching imperfections at the input of the amplifier [4143]. The QFI H(CLBI)H(CLBI)(α,γ,r,θ) for the CLBI is given by

H(CLBI)=(α2+γ2+12)(1+cosh4r)1+2αγcosθsinh4r+2(α2γ2)cosh2r.

In order to simplify the notation, we observe that the maximum over θ is obtained for θ=π, and introduce the following parameters characterizing the interferometer: the signal fraction Δ=α2/(α2+γ2), the total number of photons after the interaction, namely,

Ntot=(α2+γ2+1)cosh(2r)+2αγcos(θ)sinh(2r)1,
and the squeezing fraction β=2sinh2r/Ntot. The analytical expression of the corresponding H(CLBI)(Ntot,Δ,β) is clumsy and is not explicitly reported here. If we fix the value of Ntot and focus on Δ>1/2, then it is possible to find numerically the values Δmax and βmax maximizing H(CLBI). These are reported in the upper panel of Fig. 3, where we can see that they are both an increasing function of Ntot, with Δmax1/2 and βmax2/3 for Ntot1.

 figure: Fig. 3.

Fig. 3. Top: optimal value of the signal fraction Δmax (red dashed line) and of the squeezing fraction βmax (blue solid line), maximizing the QFI for the CLBI, as functions of Ntot. Bottom: corresponding sensitivity δϕ as a function of Ntot. The shot-noise limit Ntot1/2 (red dashed line) and the Heisenberg limit Ntot1 (green dotted line) are also reported.

Download Full Size | PDF

In Fig. 3 we show the behavior of the phase sensitivity δϕ=[Hmax(CLBI)]1/2 as a function of Ntot. Two regimes may be easily identified: for Ntot1 we find shot-noise-limited precision δϕNtot1/2, while for Ntot1 we obtain Hmax(CLBI)43Ntot(Ntot+2) and thus the Heisenberg scaling δϕNtot1. The comparison between the lower panels of Figs. 2 and 3 reveals that for Ntot1 passive interferometers offer better performance compared to the active ones, at least for the class of setup considered here. On the other hand, active interferometers are worth being investigated because there is evidence that the amplification mechanism is suitable to fight the possible photon loss occurring in the interferometer [44].

5. DISCUSSION AND CONCLUSIONS

One may wonder whether and how the ultimate bound to precision may be achieved in practice, i.e., whether there exists a feasible measurement in the final stage of the interferometer whose Fisher information is equal to the QFI or approaches it in some limit. In principle, an observation of this kind is provided by the spectral measure of the symmetric logarithmic derivative Lϕ. However, even for pure states where the solution is available explicitly it is usually very challenging to implement this kind of measurement. As a consequence, the performance of feasible measurements, e.g., photodetection after a further mixing or amplification stage, have been explored. Also, in these realistic scenarios, squeezing has been shown to represent a resource in achieving the Heisenberg limit [45], or at least to beat the shot-noise limit in some regimes [46]. It should also be mentioned that the use of Bayesian analysis [47,48] or maximum-likelihood estimators [49,50] allows one to achieve the Cramér–Rao bound (either classical or quantum) for any value of the phase shift itself, not only for a specific setting.

Our results show that precision may be indeed improved by optimizing over all the degrees of freedom and indicates which parameters are relevant in the different scenarios. Our results also provide a set of general benchmarks for Gaussian interferometry, which are independent of the measurement scheme used at the output and may be compared to performances of realistic interferometric schemes.

In conclusion, we have analyzed two-mode interferometers based on Gaussian states and operations and found the best input states, optimizing the precision given a constraint on the total mean number of photons entering the setup. For passive interferometers, e.g., Mach–Zehnder-like devices, the optimal input is made of a pair of identical squeezed-coherent states with at most one-third of the energy used in squeezing. Our results clarify the role of squeezing, which improves both the reference phase and the signal phase in the interferometer. For active interferometers, i.e., setups involving optical amplifiers, we focused on the CLBI. In this case coherent signals at the input are enough to achieve Heisenberg scaling, provided that the amplification gain is tuned to an optimal value. Our results go beyond the traditional analysis of laser interferometry and may be relevant for the development of novel quantum-enhanced interferometric schemes.

FUNDING INFORMATION

Ministry of Education, Universities and Research (MIUR) FIRB project “LiCHIS” (RBFR10YQ3H); European Union (EU) Collaborative Project QuProCS (641277); University of Milan (UniMI) (H2020 Transition Grant 15-6-3008000-625).

REFERENCES AND NOTES

1. R. X. Adhikari, “Gravitational radiation detection with laser interferometry,” Rev. Mod. Phys. 86, 121–151 (2014). [CrossRef]  

2. C. M. Caves, “Quantum-mechanical noise in an interferometer,” Phys. Rev. D 23, 1693–1708 (1981). [CrossRef]  

3. B. Yurke, S. L. McCall, and J. R. Klauder, “SU(2) and SU(1, 1) interferometers,” Phys Rev. A 33, 4033–4054 (1986). [CrossRef]  

4. M. J. Holland and K. Burnett, “Interferometric detection of optical phase shifts at the Heisenberg limit,” Phys. Rev. Lett. 71, 1355–1358 (1993). [CrossRef]  

5. M. G. A. Paris, “Small amount of squeezing in high-sensitive realistic interferometry,” Phys. Lett. A 201, 132–138 (1995). [CrossRef]  

6. T. Kim, O. Pfister, M. J. Holland, J. Noh, and J. L. Hall, “Influence of decorrelation on Heisenberg-limited interferometry with quantum correlated photons,” Phys. Rev. A 57, 4004–4013 (1998). [CrossRef]  

7. C. C. Gerry, “Heisenberg-limit interferometry with four-wave mixers operating in a nonlinear regime,” Phys. Rev. A 61, 043811 (2000). [CrossRef]  

8. R. A. Campos, C. C. Gerry, and A. Benmoussa, “Optical interferometry at the Heisenberg limit with twin Fock states and parity measurements,” Phys. Rev. A 68, 023810 (2003). [CrossRef]  

9. F. W. Sun, B. H. Liu, Y. X. Gong, Y. F. Huang, Z. Y. Ou, and G. C. Guo, “Experimental demonstration of phase measurement precision beating standard quantum limit by projection measurement,” EPL 82, 24001 (2008). [CrossRef]  

10. L. Pezzé and A. Smerzi, “Mach-Zehnder interferometry at the Heisenberg limit with coherent and squeezed-vacuum light,” Phys. Rev. Lett. 100, 073601 (2008). [CrossRef]  

11. M. Kacprowicz, R. Demkowicz-Dobrzanski, W. Wasilewski, K. Banaszek, and I. A. Walmsley, “Experimental quantum-enhanced estimation of a lossy phase shift,” Nat. Photonics 4, 357–360 (2010). [CrossRef]  

12. H. Cable and G. A. Durkin, “Parameter estimation with entangled photons produced by parametric down-conversion,” Phys. Rev. Lett. 105, 013603 (2010). [CrossRef]  

13. S. Knysh, V. N. Smelyanskiy, and G. A. Durkin, “Scaling laws for precision in quantum interferometry and the bifurcation landscape of the optimal state,” Phys. Rev. A 83, 021804(R) (2011). [CrossRef]  

14. M. G. Genoni, S. Olivares, and M. G. A. Paris, “Optical phase estimation in the presence of phase diffusion,” Phys. Rev. Lett. 106, 153603 (2011). [CrossRef]  

15. S. Pirandola, “Quantum reading of a classical digital memory,” Phys. Rev. Lett. 106, 090504 (2011). [CrossRef]  

16. A. Bisio, M. Dall’Arno, and G. M. D’Ariano, “Tradeoff between energy and error in the discrimination of quantum-optical devices,” Phys. Rev. A 84, 012310 (2011). [CrossRef]  

17. S. Pirandola, C. Lupo, V. Giovannetti, S. Mancini, and S. L. Braunstein, “Quantum reading capacity,” New J. Phys. 13, 113012 (2011). [CrossRef]  

18. L. Pezzè, P. Hyllus, and A. Smerzi, “Phase-sensitivity bounds for two-mode interferometers,” Phys. Rev. A 91, 032103 (2015). [CrossRef]  

19. J. P. Dowling, “Quantum optical metrology at the lowdown on high-N00N states,” Contemp. Phys. 49, 125–143 (2008). [CrossRef]  

20. S. Olivares, “Quantum optics in the phase space,” EPJ-ST 203, 3–24 (2012). [CrossRef]  

21. A. Monras, “Phase space formalism for quantum estimation of Gaussian states,” arXiv:1303.3682 (2013).

22. O. Pinel, P. Jian, N. Treps, C. Fabre, and D. Braun, “Quantum parameter estimation using general single-mode Gaussian states,” Phys. Rev. A 88, 040102(R) (2013). [CrossRef]  

23. Z. Jiang, “Quantum Fisher information for states in exponential form,” Phys. Rev. A 89, 032128 (2014). [CrossRef]  

24. D. Šafránek, A. R. Lee, and I. Fuentes, “Quantum parameter estimation using multi-mode Gaussian states,” arXiv:1502.07924 (2015).

25. G. Adesso, “Gaussian interferometric power,” Phys. Rev. A 90, 022321 (2014). [CrossRef]  

26. M. D. Lang and C. M. Caves, “Optimal quantum-enhanced interferometry using a laser power source,” Phys. Rev. Lett. 111, 173601 (2013). [CrossRef]  

27. C. W. Helstrom, Quantum Detection and Estimation Theory (Academic, 1976).

28. D. C. Brody and L. P. Hughston, “Statistical geometry in quantum mechanics,” Proc. R. Soc. London A 454, 2445–2475 (1998). [CrossRef]  

29. D. C. Brody and L. P. Hughston, “Geometrization of statistical mechanics,” Proc. R. Soc. London A 455, 1683–1715 (1999). [CrossRef]  

30. M. G. A. Paris, “Quantum estimation for quantum technology,” Int. J. Quantum Inform. 7, 125–137 (2009). [CrossRef]  

31. S. L. Braunstein and C. M. Caves, “Statistical distance and the geometry of quantum states,” Phys. Rev. Lett. 72, 3439–3443 (1994). [CrossRef]  

32. S. L. Braunstein, C. M. Caves, and G. J. Milburn, “Generalized uncertainty relations: Theory, examples, and Lorentz invariance,” Ann. Phys. 247, 135–173 (1996). [CrossRef]  

33. S. Olivares and M. G. A. Paris, “Optimized interferometry with Gaussian states,” Opt. Spectrosc. 103, 231–236 (2007). [CrossRef]  

34. S. Olivares and M. G. A. Paris, “Fidelity matters: The birth of entanglement in the mixing of Gaussian states,” Phys. Rev. Lett 107, 170505 (2011). [CrossRef]  

35. A. Monras, “Optimal phase measurements with pure Gaussian states,” Phys. Rev. A 73, 033821 (2006). [CrossRef]  

36. M. Jarzyna and R. Demkowicz-Dobrzański, “Quantum interferometry with and without an external phase reference,” Phys. Rev. A 85, 011801(R) (2012).

37. Since the squeezing of the states is the same, after the evolution we still have factorized states, otherwise correlations will arise reducing the QFI, since now the single mode involved in the phase shift is no longer a pure state. It is worth noting that in the single-mode case one has Hsm>Hmax, but it is not possible to fix a phase reference.

38. J. Abadie, B. P. Abbott, R. Abbott, R. Adhikari, S. B. Anderson, K. Arai, M. C. Araya, J. Betzwieser, G. Billingsley, E. Black, J. K. Blackburn, R. Bork, A. F. Brooks, C. Cepeda, T. Chalermsongsak, A. Corsi, D. C. Coyne, R. Dannenberg, B. Daudert, V. Dergachev, R. DeSalvo, J. C. Driggers, P. Ehrens, R. Engel, T. Etzel, N. Fotopoulos, E. K. Gustafson, C. Hanna, J. Harms, J. Heefner, A. W. Heptonstall, K. A. Hodge, A. Ivanov, M. Jacobson, P. Kalmus, W. Kells, P. J. King, V. Kondrashov, W. Z. Korth, D. Kozak, A. Lazzarini, P. E. Lindquist, M. Mageswaran, K. Mailand, E. Maros, J. N. Marx, G. McIntyre, S. Meshkov, T. Nash, G. H. Ogin, C. Osthelder, P. Ajith, P. Patel, M. Pedraza, M. Phelps, L. R. Price, S. Privitera, N. A. Robertson, V. Sannibale, L. Santamaría, A. C. Searle, F. Seifert, A. S. Sengupta, A. Singer, L. Singer, M. R. Smith, A. Stochino, R. Taylor, C. I. Torrie, S. Vass, A. E. Villar, L. Wallace, R. L. Ward, A. J. Weinstein, S. E. Whitcomb, P. A. Willems, H. Yamamoto, D. Yeaton-Massey, L. Zhang, and J. Zweizig (the LIGO Scientific Collaboration), “A gravitational wave observatory operating beyond the quantum shot-noise limit,” Nat. Phys. 7, 962–965 (2011). [CrossRef]  

39. K. McKenzie, D. Shaddock, D. McClelland, B. Buchler, and P. Lam, “Experimental demonstration of a squeezing-enhanced power-recycled Michelson interferometer for gravitational wave detection,” Phys. Rev. Lett. 88, 231102 (2002). [CrossRef]  

40. N. P. William, J. P. Dowling, and G. S. Agarwal, “Coherent-light-boosted, sub-shot noise, quantum interferometry,” New J. Phys. 12, 083014 (2010). [CrossRef]  

41. A. Ferraro, M. G. A. Paris, A. Allevi, A. Andreoni, M. Bondani, and E. Puddu, “Three-mode entanglement by interlinked nonlinear interactions in optical χ(2) media,” J. Opt. Soc. Am. B 21, 1241 (2004). [CrossRef]  

42. A. Allevi, A. Andreoni, M. Bondani, E. Puddu, A. Ferraro, and M. G. A. Paris, “Properties of two interlinked χ(2) interactions in non-collinear phase-matching,” Opt. Lett. 29, 180 (2004). [CrossRef]  

43. A. Ferraro and M. G. A. Paris, “Multimode entanglement and telecloning in a noisy environment,” Phys. Rev. A 72, 032312 (2005). [CrossRef]  

44. C. Sparaciari, S. Olivares, and M. G. A. Paris, “Bounds to precision in realistic Gaussian interferometry,” in preparation.

45. M. D. Lang and C. M. Caves, “Optimal quantum-enhanced interferometry,” Phys. Rev. A 90, 025802 (2014). [CrossRef]  

46. R. Demkowicz-Dobrzanski, K. Banaszek, and R. Schnabel, “Fundamental quantum interferometry bound for the squeezed-light-enhanced gravitational wave detector GEO 600,” Phys. Rev. A 88, 041802(R) (2013). [CrossRef]  

47. L. Pezzé, A. Smerzi, G. Khoury, J. F. Hodelin, and D. Bouwmeester, “Phase detection at the quantum limit with multiphoton Mach–Zehnder interferometry,” Phys. Rev. Lett. 99, 223602 (2007). [CrossRef]  

48. S. Olivares and M. G. A. Paris, “Bayesian estimation in homodyne interferometry,” J. Phys. B 42, 055506 (2009). [CrossRef]  

49. A. S. Lane, S. L. Braunstein, and C. M. Caves, “Maximum-likelihood statistics of multiple quantum phase measurements,” Phys. Rev. A 47, 1667–1696 (1993). [CrossRef]  

50. Z. Hradil, R. Myška, T. Opatrný, and J. Bajer, “Entropy of phase measurement: Quantum phase via quadrature measurement,” Phys. Rev. A 53, 3738–3742 (1996). [CrossRef]  

Cited By

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

Alert me when this article is cited.


Figures (3)

Fig. 1.
Fig. 1. Schematic diagram of a two-mode interferometric setup. The two modes, prepared in the factorized state | ψ a and | ψ b , are coupled by the unitary operator U int ( Θ ) . Then, one of the modes is phase-shifted by an amount ϕ and, finally, the two-mode state undergoes a measurement described by the POVM { Π } .
Fig. 2.
Fig. 2. Top: optimal value of the total squeezing fraction β tot , maximizing the QFI for a passive interferometer, as a function of N tot . Bottom: corresponding phase sensitivity δ ϕ as a function of N tot (blue solid line). The shot-noise limit N tot 1 / 2 (red dashed line) and the Heisenberg limit N tot 1 (green dotted line) are also reported.
Fig. 3.
Fig. 3. Top: optimal value of the signal fraction Δ max (red dashed line) and of the squeezing fraction β max (blue solid line), maximizing the QFI for the CLBI, as functions of N tot . Bottom: corresponding sensitivity δ ϕ as a function of N tot . The shot-noise limit N tot 1 / 2 (red dashed line) and the Heisenberg limit N tot 1 (green dotted line) are also reported.

Equations (16)

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

| Ψ ( Θ , ϕ ) = U ph ( ϕ ) U int ( Θ ) | ψ a | ψ b .
δ ϕ 2 1 M H ( ϕ ) ,
H ( ϕ ) = Tr [ R ϕ L ϕ 2 ] ,
ϕ R ϕ = 1 2 ( L ϕ R ϕ + R ϕ L ϕ ) .
L ϕ = 2 ( | ϕ Ψ ( Θ , ϕ ) Ψ ( Θ , ϕ ) | + | Ψ ( Θ , ϕ ) ϕ Ψ ( Θ , ϕ ) | ) ,
H = 1 4 { 4 ( α + γ ) 2 [ e 2 ξ + cosh 2 r ] + cosh 4 ξ + cosh 4 r + 2 cos θ sinh 2 r [ 2 ( α + γ ) 2 + sinh 2 ξ ] + cosh 2 ( r ξ ) + cosh 2 ( ξ + r ) 4 } .
N tot = α 2 + γ 2 + sinh 2 r + sinh 2 ξ ,
Δ = α 2 / ( α 2 + γ 2 ) ,
β tot = ( sinh 2 r + sinh 2 ξ ) / N tot β = sinh 2 ξ / N tot ,
H max ( N tot , β tot ) = 2 N tot [ 2 + N tot β tot ( 2 β tot ) + 2 ( 1 β tot ) N tot β tot ( 2 + N tot β tot ) ] ,
H max 8 9 N tot 2 [ 2 + ( 1 + 3 N tot 1 ) 1 / 2 ] + 4 N tot 8 3 N tot 2 .
U BS | α , r | α , r = | 2 α , r | 0 , r ,
H ( gw ) ( N tot , β ) = 2 N tot [ 1 + ( 1 2 + N tot ) β + ( 1 β ) N tot β ( 1 + N tot β ) ] ,
H max ( gw ) = 2 N tot ( 3 2 + N tot ) N tot 1 2 N tot 2 .
H ( CLBI ) = ( α 2 + γ 2 + 1 2 ) ( 1 + cosh 4 r ) 1 + 2 α γ cos θ sinh 4 r + 2 ( α 2 γ 2 ) cosh 2 r .
N tot = ( α 2 + γ 2 + 1 ) cosh ( 2 r ) + 2 α γ cos ( θ ) sinh ( 2 r ) 1 ,
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.