This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Effects of Systematic Error on Quantum-Enhanced Atom Interferometry

Joshua Goldsmith Joshua.Goldsmith@anu.edu.au Department of Quantum Science and Technology, Research School of Physics, The Australian National University, Canberra, Australia    Joseph Hope Department of Quantum Science and Technology, Research School of Physics, The Australian National University, Canberra, Australia    Simon Haine Department of Quantum Science and Technology, Research School of Physics, The Australian National University, Canberra, Australia
Abstract

We develop a framework for describing the effects of systematic state preparation error in quantum-enhanced atom interferometry on sensing performance. We do this in the context of both spin-squeezed and non-Gaussian states for the two-axis-twisting (TAT), one-axis-twisting (OAT), and twist-and-turn (TNT) state preparation schemes, and derive general conditions for robustness and susceptibility of quantum states to state preparation error. In the spin-squeezing regime, we find that OAT is more susceptible to state preparation error than TAT due to its parameter-dependent phase space rotation. In the non-Gaussian regime, we find that OAT is robust to state preparation errors, which can be explained by a small ratio of off-diagonal to diagonal elements in its Fisher-covariance matrix. In contrast, TNT does not exhibit this robustness. We find that the single parameter unbiased estimators that are habitually used in quantum-enhanced atom interferometry are not always optimal, and that there may be occasions where biased estimators, or two-parameter unbiased estimators, lead to lower net error.

preprint: APS/123-QED

I Introduction

Atom interferometers are measurement instruments relying on the interference of matter-waves [1, 2], and are currently used for precision sensing of a wide variety of quantities, including magnetic fields [3, 4], accelerations [5], rotations [6], gravitational fields [7, 8, 9], and gravity gradients [10]. Unlike classical devices, which may drift over time and thus require regular calibration, the response of an atom interferometer can, in principle, be locked to fundamental constants of nature [11, 12, 9], making it inherently robust against long-term drift. This key advantage of atom interferometry enables highly sensitive precision measurements, such as those of the fine-structure constant [13] and Newton’s gravitational constant [14].

There is considerable recent interest in using quantum entanglement to increase the sensitivity of atom interferometry through reducing the atomic shot-noise [15]. Holding everything else equal, reducing the atomic shot-noise would improve the per-shot sensitivity, increasing the rate at which a given sensitivity is reached when integrating down. This would ultimately improve the measurement bandwidth, allowing for more precise measurements of slowly changing signals, or allow for higher sensitivities when the experiment is limited by drift in other quantities [16, 17]. Alternatively, incorporating quantum-entanglement while holding precision fixed would allow for smaller interrogation times, leading to significantly smaller devices [17], and devices less prone to external perturbations [18]. There have been several proof-of-principle demonstrations of entanglement-enhanced atom interferometry, with entanglement generated via either atom-atom [19, 20, 21, 22, 23, 24, 25, 26, 27, 28, 29] or atom-light [30, 31, 32, 33, 34, 35] interactions. Several alternate schemes have also been theoretically proposed, but are yet to be experimentally demonstrated [36, 37, 38, 39, 40, 41, 42, 43, 44, 45].

While there have been many studies on how quantum-enhancement affects the precision of atom interferometers, there has yet to be an investigation on how it affects their accuracy. In particular, the key advantage of atom interferometry is that the response only depends on fundamental constants, which are well-known and static. Once atom-atom, or atom-light interactions are introduced into the scheme, the response now depends on less easy to characterise and control parameters. For example, the one-axis twisting scheme [46] can be implemented via atom-atom interactions [47, 39]. Deviations in any number of parameters, such as the atomic scattering length or even the trapping configuration, will affect the initial state and ultimately the response of the device. Similarly, atom-light interactions could be affected by parameters such as laser power, or even the size of the beam focus, both of which are susceptible to drift [48, 49]. These considerations raise the question: Given that the stability of calibration is the key advantage of atom interferometry, do the added complexities of entanglement-generation schemes outweigh the precision gains they offer?

In this paper, we investigate the impact of systematic errors in entangled state preparation on the theoretical precision and accuracy of atom interferometers. Through investigating the one-axis-twisting (OAT) and two-axis-twisting (TAT) processes, we consider systematic state preparation errors in spin-squeezed states, and demonstrate how in some circumstances, biased estimators may meaningfully increase parameter resolution. We examine the degree of state preparation error in OAT that would degrade quantum enhancement by various factors for metrologically useful atom ensemble sizes, providing a schema for calculating the effect of entangled state preparation error for an entangled atom interferometry sequence. We then consider errors in state preparation in the non-Gaussian regime through the OAT and twist-and-turn (TNT) state preparation schemes. In both the spin-squeezed and non-Gaussian regimes, we identify general classes of quantum states that are resistant to state preparation error. We present a means of overcoming the biasing effects of state preparation error using dual parameter maximum likelihood estimation.

II Quantum-enhanced atom interferometry and parameter estimation

An atom interferometry sequence can be divided into four stages: state preparation, parameter encoding, measurement, and parameter estimation, as depicted in Figure 1.

Refer to caption
Figure 1: A quantum-enhanced atom interferometry sequence. In (a), an initial coherent-spin state, entangled in (b) according to U^𝚲\hat{U}_{\mathbf{\Lambda}}, with the parameter φ\varphi encoded onto the state in (c). In (d) a projective measurement in the J^z\hat{J}_{z} basis is made, with an estimate φ\varphi^{*} occurring based on the measurement results according to an estimator φ~\tilde{\varphi}.

Quantum enhancement occurs in the state preparation stage, where correlations are induced in an initially unentangled atomic ensemble to surpass the shot-noise limit. The preparation of a metrologically useful entangled state, and the subsequent encoding of a parameter φ\varphi of interest, can be represented as

|ψφ,𝚲=U^φU^𝚲|ψ0,\ket{\psi_{\varphi,\mathbf{\Lambda}}}=\hat{U}_{\varphi}\hat{U}_{\mathbf{\Lambda}}\ket{\psi_{0}}, (1)

where 𝚲\mathbf{\Lambda} represents a vector of state preparation parameters. In this paper, we only consider state preparation schemes with one degree of freedom, which we denote as λ\lambda. All subsequent state preparation parameters are either constant or a function of λ\lambda. As an example, the OAT spin-squeezing scheme relies on a nonlinear interaction between the atoms to generate entanglement. In this case, the strength of that interaction would be the λ\lambda state preparation degree of freedom in 𝚲\mathbf{\Lambda}, with the subsequent phase space rotation a λ\lambda-dependent state preparation parameter. After the parameter φ\varphi is encoded onto the system, a measurement P^\hat{P} is made in some basis, generating a set of measurement results 𝐗={X1,,Xn}\mathbf{X}=\{X_{1},...,X_{n}\}, which are used to estimate φ\varphi via a function φ~\tilde{\varphi}, termed an estimator, producing an estimate φ=φ~\varphi^{*}=\langle\tilde{\varphi}\rangle. For a given estimator, the uncertainty in the estimate is given by the mean-square error:

MSE =(φ~φ)2\displaystyle=\langle\left(\tilde{\varphi}-\varphi\right)^{2}\rangle
=σQ2+σB2,\displaystyle=\sigma_{Q}^{2}+\sigma_{B}^{2}, (2)

where

σQ2\displaystyle\sigma_{Q}^{2} =φ~2φ~2\displaystyle=\langle\tilde{\varphi}^{2}\rangle-\langle\tilde{\varphi}\rangle^{2}
=φ~2(φ)2\displaystyle=\langle\tilde{\varphi}^{2}\rangle-\left(\varphi^{*}\right)^{2} (3)

is the contribution from sampling error and statical fluctuations resulting from the quantum projection noise in the measurement P^\hat{P}, and

σB\displaystyle\sigma_{B} =(φ~φ)\displaystyle=\langle\left(\tilde{\varphi}-\varphi\right)\rangle
=(φφ)\displaystyle=\left(\varphi^{*}-\varphi\right) (4)

is the systematic error, or bias. The MSE is the total mean-squared distance between our estimate of the parameter and its true value, and is an overall metric for error accounting for both bias in a parameter estimate, and fluctuations around that parameter estimate.

For an unbiased estimator (σB=0\sigma_{B}=0), the optimal precision of a readout distribution is given by the Cramer-Rao bound (CRB)

MSE=σQ2=Q2m1mFc,\text{MSE}=\sigma_{Q}^{2}=\frac{Q^{2}}{m}\geq\dfrac{1}{mF_{c}}, (5)

where mm is the number of repetitions of the experiment, QQ is the single-shot quantum projection noise, and FcF_{c} is the classical Fisher information, given by

Fc=j(φPφ(j))2Pφ(j)F_{c}=\sum_{j}\frac{(\partial_{\varphi}P_{\varphi}(j))^{2}}{P_{\varphi}(j)} (6)

for measurement values jj and a probability distribution P(j)P(j) [50]. The estimator φ~\tilde{\varphi} that is guaranteed to saturate the CRB for all distributions is the maximum likelihood estimator, but other classes of estimators may also fulfil this constraint in particular scenarios. The fundamental limit to single parameter quantum sensing is achieved through saturating the Quantum-Cramér-Rao bound (QCRB): Fc=FQF_{c}=F_{Q}, where FQF_{Q} is the quantum Fisher information [51, 15].

The minimally noisy unbiased estimator for a given state preparation sequence is a function of the degree of entanglement in the state, and hence on the state preparation degree of freedom. The estimator parameterised by the state preparation degree of freedom λp\lambda_{p} is denoted φ~(𝐗|λp\tilde{\varphi}(\mathbf{X}|{\lambda_{p}}) for a set of measurement data 𝐗\mathbf{X}.

II.1 Effects of state preparation error

State preparation error occurs when we believe that we have prepared the state

|ψφ,𝚲=U^φU^𝚲|ψ0,\ket{\psi_{\varphi,\mathbf{\Lambda^{\prime}}}}=\hat{U}_{\varphi}\hat{U}_{\mathbf{\Lambda^{\prime}}}\ket{\psi_{0}}\,, (7)

but have actually prepared the state in Equation 1, with 𝚲𝚲\mathbf{\Lambda}\neq\mathbf{\Lambda^{\prime}}. We use the notation that the dashed superscript denotes that a variable is assumed, while the absence of a dash implies the actual physical value. Given a state preparation error, the λp\lambda_{p} value used in our estimator will be λp=λ\lambda_{p}=\lambda^{\prime} instead of λp=λ\lambda_{p}=\lambda, as this will correspond to the estimator we think will be unbiased. As our estimator is then parameterised by the wrong value, it will output an expected value φ~(𝐗|λ)=φ\langle\tilde{\varphi}(\mathbf{X}|\lambda^{\prime})\rangle=\varphi^{*}, where the true phase value φφ\varphi\neq\varphi^{*} in most instances. In addition to the introduction of bias, the biased estimator will lead to different single shot quantum noise QQ relative to an unbiased estimator. Furthermore, assuming the wrong state preparation degree of freedom may bias other state preparation parameters dependent on λ\lambda, thus changing the measurement data as well as the estimator. These distinct measurement data may cause the state preparation error to further change QQ relative to an unbiased quantum-enhanced atom interferometry sequence. Figure 2 demonstrates how each of the actual, assumed, and estimator state preparation values interact to produce a value for the error in the parameter of interest.

Refer to caption
Figure 2: A depiction of how the set of state preparation parameters 𝚲\mathbf{\Lambda}, and the actual, assumed, and estimator state preparation degrees of freedom (λ,λ,λp\lambda,\lambda^{\prime},\lambda_{p}) interact to give error in state preparation. The real 𝚲\mathbf{\Lambda} values totally characterise the measurement data 𝐗\mathbf{X}, and the value λp\lambda_{p} is determined by λ\lambda^{\prime}, which in turn is used to process the measured data to obtain a parameter value with a particular mean squared error. Assuming a particular state preparation degree of freedom may result in later parameters in the state preparation sequence being suboptimal, hence the dashed line.

In many cases, such as when using spin-squeezed states, bias induced by state preparation error will be linearly proportional to the parameter φ\varphi for φ1\varphi\ll 1, and the total MSE will be characterised by the parameters BB and QQ:

MSE=(Bφ)2+Q2m,\text{MSE}=(B\varphi)^{2}+\frac{Q^{2}}{m}\,, (8)

where BB is defined by σB=Bφ\sigma_{B}=B\varphi. As the bias decreases when operating close to the bias-free operating point of φ=0\varphi=0, it is desirable to operate as close to this point as possible. Adaptive schemes will be ultimately limited by the resolution of the measurement. As such, we can obtain a self-consistent lower bound on the MSE by setting φ=Q/m\varphi=Q/\sqrt{m} in Equation 8, which yields MSEE2/mN\text{MSE}\geq E^{2}/mN, where

E\displaystyle E =NQ2(1+B2)\displaystyle=\sqrt{NQ^{2}(1+B^{2})}\, (9)

is the total error of the measurement, normalised by the precision attainable in a single bias-free, uncorrelated atomic ensemble. Alongside BB and QQ, E>1E>1 is a useful indicative quantity as it implies that the presence of state preparation error has worsened the precision of atom interferometry relative to using an unentangled state - one would have been better off dispensing with entangled state preparation entirely.

III Systematic state preparation error in spin-squeezed states

We now turn our attention to the specific example of spin-squeezed states. In this paper, we confine the states explored to two-mode bosonic states. We describe our quantum state using second quantised notation, and define a set of pseudo-spin operators

J^k=12(a^1a^2)σk(a^1a^2)T\hat{J}_{k}=\frac{1}{2}(\hat{a}^{\dagger}_{1}\hat{a}^{\dagger}_{2})\sigma_{k}(\hat{a}_{1}\hat{a}_{2})^{T} (10)

acting on our state, where σk\sigma_{k} corresponds to the kth{k}^{\text{th}} Pauli matrix, and a^1\hat{a}_{1} and a^2\hat{a}_{2} represent the first and second annihilation operators for a Fock space respectively [52]. The J^z\hat{J}_{z} operator corresponds to half the number difference of the two modes the atoms can populate, and is usually our physically accessible measurement. Certain classes of two-mode bosonic states can yield subshot noise sensitivity based on moments of a readout distribution. An important example of this is spin-squeezed states, where subshot noise phase sensitivity is engineered through reducing variance in the measurement axis at the expense of variance in a perpendicular axis according to a Heisenberg uncertainty relation. Given a set of measurement results, the phase estimation strategy that achieves the degree of quantum enhancement ξ\xi is the method of moments (MOM) strategy, where J^z\langle\hat{J}_{z}\rangle is assigned to a phase estimate φ\varphi [53]. Assuming the measurement axis is J^z\hat{J}_{z}, the degree of spin-squeezing is expressed in the Wineland spin-squeezing parameter [54]

ξ=NVar(J^z))|J^x|,\xi=\dfrac{\sqrt{N\text{Var}(\hat{J}_{z}))}}{|\langle\hat{J}_{x}\rangle|}, (11)

where ξ<1\xi<1 implies the state is spin-squeezed along the J^z\hat{J}_{z} axis, and surpasses the shot noise limit by a factor 1ξ\frac{1}{\xi}. States with NFQ<1\dfrac{\sqrt{N}}{\sqrt{F_{Q}}}<1 but ξ>1\xi>1 are known as metrologically useful non-Gaussian quantum states [24], and cannot achieve subshot noise sensitivity with a MOM estimator. We will develop separate frameworks for analysing state preparation error in spin-squeezed and non-Gaussian states.

Refer to caption
Figure 3: In (a), the average JzJ_{z} value as a function of the parameter φ\varphi and the estimator φ~\tilde{\varphi} for biased and unbiased estimators parameterised by λ=0.01\lambda^{\prime}=0.01 and λ=0.02\lambda=0.02 respectively for a two-axis twisted spin-squeezed state of N=100N=100 (these two values correspond to ξ=0.38\xi=0.38 and ξ=0.2\xi=0.2 respectively). Given a measured J^z\langle\hat{J}_{z}\rangle, using the incorrect estimator will lead to a biased φ\varphi value, as demonstrated by the red and green arrows. In (b), the absolute value of the bias using the estimator φ~(λp=λ)\tilde{\varphi}(\lambda_{p}=\lambda^{\prime}) for varying φ\varphi. Where there is no error, as in the estimator parameterised by λ\lambda, there is no bias. In (c). The variance of the biased and unbiased estimators. Variance is minimised at the φ=0\varphi=0 operating point, and differs for the biased and unbiased estimators.

We now examine how bias can arise, and variance can change under state preparation error for a method of moments estimator and a spin-squeezed state. This estimator involves a particular mapping between J^z\langle\hat{J}_{z}\rangle and the parameter of interest, with J^z\langle\hat{J}_{z}\rangle being the average of the measured results 𝐗\mathbf{X}. In the case of spin-squeezed atom interferometry sequences, the unitary U^𝚲\hat{U}_{\mathbf{\Lambda}} from Equation 1 can be broken up as

U^𝚲=U^θU^λ,\hat{U}_{\mathbf{\Lambda}}=\hat{U}_{\theta}\hat{U}_{\lambda}, (12)

where λ\lambda parametrises the strength of spin-squeezing independent of readout, and where θ\theta parameterises a constant or λ\lambda-dependent J^x\hat{J}_{x} phase space rotation designed to minimise variance in the measurement axis J^z\hat{J}_{z}. In order to determine the appropriate MOM estimator, we first determine the phase response of the device. Working in the Heisenberg picture, J^z\hat{J}_{z} evolves as

J^z=sin(φ)J^x0+cos(φ)(J^z0cos(θ)+J^y0sin(θ)),\hat{J}_{z}=-\sin(\varphi)\hat{J}_{x_{0}}+\cos(\varphi)\left(\hat{J}_{z_{0}}\cos(\theta)+\hat{J}_{y_{0}}\sin(\theta)\right), (13)

where J^i0\hat{J}_{i_{0}} refers to the operator post the application of U^λ\hat{U}_{\lambda} but before the application of U^φU^θ\hat{U}_{\varphi}\hat{U}_{\theta}. Given that |ψ0\ket{\psi_{0}} is a J^x\hat{J}_{x} eigenstate, we approximate that J^z0\langle\hat{J}_{z_{0}}\rangle,J^y0=0\langle\hat{J}_{y_{0}}\rangle=0 and thus we can write J^z\langle\hat{J}_{z}\rangle as

J^z=J^x0sin(φ).\langle\hat{J}_{z}\rangle=-\langle\hat{J}_{x_{0}}\rangle\sin(\varphi)\,. (14)

Informed by this, our estimate of φ\varphi is then

φ\displaystyle\varphi^{*} =sin1(J^zJ^x0),\displaystyle=\sin^{-1}\left(-\frac{\langle\hat{J}_{z}\rangle}{\langle\hat{J}_{x_{0}}\rangle^{\prime}}\right)\,, (15)

where J^x0J^x0\langle\hat{J}_{x_{0}}\rangle^{\prime}\equiv\langle\hat{J}_{x_{0}}\rangle evaluated at λp=λ\lambda_{p}=\lambda^{\prime}, and we restrict the φ\varphi domain to π2φπ2-\frac{\pi}{2}\leq\varphi\leq\frac{\pi}{2}. Expanding Equation 15 to linear order around J^z=0\langle\hat{J}_{z}\rangle=0, we can write

σB=φ(1J^x0J^x0).\sigma_{B}=\varphi\left(1-\dfrac{\langle\hat{J}_{x_{0}}\rangle^{\prime}}{\langle\hat{J}_{x_{0}}\rangle}\right). (16)

Similarly, we can calculate σQ\sigma_{Q} via the error propagation equation

σQ2\displaystyle\sigma_{Q}^{2} =1mVar(J^z)(φJ^z)2,\displaystyle=\frac{1}{m}\frac{\mathrm{Var}(\hat{J}_{z})}{\left(\partial_{\varphi}\langle\hat{J}_{z}\rangle^{\prime}\right)^{2}}, (17)

where the numerator is based on the measured variance in J^z\langle\hat{J}_{z}\rangle and the denominator is evaluated at the assumed state preparation value λ\lambda^{\prime}. Using Equation 13, this becomes

σQ2\displaystyle\sigma_{Q}^{2} =1m|J^x0|2(cos2(θ)Var(J^z0)+sin2(θ)Var(J^y0)\displaystyle=\dfrac{1}{m|\langle\hat{J}_{x_{0}}\rangle^{\prime}|^{2}}\left(\cos^{2}(\theta)\text{Var}(\hat{J}_{z_{0}})+\sin^{2}(\theta)\text{Var}(\hat{J}_{y_{0}})\right.
+12sin(2θ)Covar¯(J^z0,J^y0)),\displaystyle\quad+\left.\frac{1}{2}\sin(2\theta)\overline{\text{Covar}}(\hat{J}_{z_{0}},\hat{J}_{y_{0}})\right), (18)

where 18 is derived in Appendix VI A, and Covar¯(a,b)=Covar(a,b)+Covar(b,a)\bar{\text{Covar}}(a,b)=\text{Covar}(a,b)+\text{Covar}(b,a). Figure 3 (a) shows J^z\langle\hat{J}_{z}\rangle for two different values of λ\lambda, corresponding to ξ=0.20\xi=0.20 and ξ=0.38\xi=0.38, leading to different phase responses. Assuming one value of λ\lambda, while in reality, the other is true, leads to a miss-estimate of φ\varphi. Processing J^z\langle\hat{J}_{z}\rangle with the wrong estimator leads to a biased estimate of φ\varphi, with the red and green arrows in (a) demonstrating a biased estimate given a J^z\langle\hat{J}_{z}\rangle measurement at a particular operating point. In (b) and (c) we see σB\sigma_{B} and σQ\sigma_{Q} respectively. In this case, miss-characterising the state leads to a bias that is approximately linear in φ\varphi. The variance of the estimated parameter around the mean also changes relative to the unbiased estimator, that is, Q(λ)Q(λ)Q(\lambda)\neq Q(\lambda^{\prime}), in Figure 3 (c). In this particular example, the noise in the measurement data J^z\langle\hat{J}_{z}\rangle is not changed through state preparation error; the change in QQ is due purely to the use of a different estimator affecting the denominator of Equation 17. We will also see cases in Section III.1 where the data 𝐗\mathbf{X} are affected by a poorly characterised λ\lambda.

III.1 Spin-squeezed OAT vs TAT

We consider two spin-squeezing methods to study spin-squeezed states under state preparation error, namely two-axis twisting (TAT) [46, 55] and one-axis twisting (OAT) [56, 57, 39, 32, 20]. TAT has yet to be experimentally realised, but provides the “simplest” example of an entangled state preparation sequence, given that its optimal J^x\hat{J}_{x} rotation parameter is independent of λ\lambda. It has been shown that TAT dynamics can be produced in cavity-mediated atom-atom interactions [58], and can produce a state similar to that prepared by quantum-nondemolition measurement (QND) squeezing [48, 32, 59, 34, 49, 35]. For an initial J^x\hat{J}_{x} eigenstate, the TAT Hamiltonian and state preparation sequence are

H^TAT\displaystyle\hat{H}_{\text{TAT}} =χ(J^zJ^y+J^yJ^z),\displaystyle=-\hbar\chi(\hat{J}_{z}\hat{J}_{y}+\hat{J}_{y}\hat{J}_{z}), (19)
U^TAT\displaystyle\hat{U}_{\text{TAT}} =exp((iπJ^x2))exp((iλ(J^zJ^y+J^yJ^z))),\displaystyle=\exp{\left(\frac{-i\pi\hat{J}_{x}}{2}\right)}\exp{\left(i\lambda(\hat{J}_{z}\hat{J}_{y}+\hat{J}_{y}\hat{J}_{z})\right)}, (20)

where λ=χt\lambda=\chi t. OAT has also been experimentally demonstrated [19, 20, 31]. The OAT Hamiltonian is given by

H^OAT=χJ^z2,\hat{H}_{\text{OAT}}=-\hbar\chi\hat{J}_{z}^{2}, (21)

and the OAT state preparation sequence is given by

U^OAT=exp((iθ(λ)J^x))exp((iλJ^z2)),\hat{U}_{\text{{OAT}}}=\exp{\left(-i\theta(\lambda^{\prime})\hat{J}_{x}\right)}\exp{\left(i\lambda\hat{J}_{z}^{2}\right)}, (22)

where the optimal J^x\hat{J}_{x} rotation θ\theta is dependent on λ\lambda, and thus a suboptimal rotation will be applied if λλ\lambda^{\prime}\neq\lambda.

We simulate the TAT and OAT interferometry sequence according to Equations 22 and 20 for N=100N=100 atoms in an initial J^x\hat{J}_{x} eigenstate given a sample number m=104m=10^{4}, a phase φ=Qm\varphi=\frac{Q}{\sqrt{m}} encoded anti-clockwise around the J^y\hat{J}_{y} axis, and a range of actual and assumed λ\lambda and λ\lambda^{\prime} values. For each λ\lambda, λ\lambda^{\prime} pair, we calculate BB, QQ, and EE according to Equations 16, 18, and 9 respectively. In the case of both OAT and TAT, we find that λλ\lambda^{\prime}\neq\lambda leads to a biased estimate of the encoded phase as φ~\tilde{\varphi} is parameterised by λp=λ\lambda_{p}=\lambda^{\prime} instead of λp=λ\lambda_{p}=\lambda. In both cases, λ<λ\lambda^{\prime}<\lambda leads to negative bias and λ>λ\lambda^{\prime}>\lambda leads to positive bias, as can be seen in Figure 4 (b) and (f). The magnitude and sign of the bias depend on the discrepancy between the assumed and actual J^x0\langle\hat{J}_{x_{0}}\rangle values, as per Equation 16, with the dependency of J^x0\langle\hat{J}_{x_{0}}\rangle for both TAT and OAT displayed in Figure 4 (a) and (e) respectively. For both OAT and TAT in the spin-squeezing regime, the J^x0\langle\hat{J}_{x_{0}}\rangle value begins to decrease at an increasing rate as λ\lambda increases. This results in the slope of the graph of BB increasing for both OAT and TAT with respect to the assumed state preparation parameter.

The plots of EE for OAT and TAT are qualitatively distinct. In the case of TAT in Figure 4 (d), we find that EE increases uniformly as the assumed state preparation value λ\lambda^{\prime} increases, with EE for λ<λ\lambda^{\prime}<\lambda lower than for λ>λ\lambda^{\prime}>\lambda. This has a surprising implication: that the “wrong” estimator may be capable of improving the precision of an atom interferometry experiment. Whilst EE is a lower bound on the achievable error, given the small contribution of the φ\varphi term in Equation 18 and the neighbouring plot of QQ in (c), we expect that EE will be close to the actual error achievable for some mm and φ\varphi values. We will investigate this in the next subsection. The decrease in QQ comes because the variance of the parameter estimate is reduced as a result of underestimating J^x0\langle\hat{J}_{x_{0}}\rangle according to Equation 18. The increased rate of QQ increase for λ>λ\lambda^{\prime}>\lambda owes to the increased rate of decrease in J^x0\langle\hat{J}_{x_{0}}\rangle in the same expression. For OAT, we find that EE is optimal when λ=λ\lambda^{\prime}=\lambda. This is because state preparation error in OAT leads to the “wrong” U^θ\hat{U}_{\theta} being applied, meaning that the axis of maximal spin-squeezing is not along the measurement axis J^z\hat{J}_{z} - the data 𝐗\mathbf{X} produced are suboptimal. As a result, greater single shot variance ensues, as can be seen in (g). We note that QQ climbs much more steeply for λ<λ\lambda^{\prime}<\lambda than for λ>λ\lambda^{\prime}>\lambda. This owes to the fact that the optimal θ\theta angle changes much more quickly for low λ\lambda values than for higher λ\lambda values.

The qualitative distinction between EE for OAT and TAT can be understood by considering the Q-sphere phase space representations of TAT and OAT states before parameter encoding in Figure 4 (i), (j), and (k). In (i), we see a TAT quantum state for λ=0.02\lambda=0.02 for all λ\lambda^{\prime} values. As the J^x\hat{J}_{x} rotation angle does not depend on λ\lambda, the axis of maximal spin squeezing is always aligned to the J^z\hat{J}_{z} axis. In (j), we see an OAT state with λ=0.034\lambda=0.034 prepared with an assumed value λ=0.014\lambda^{\prime}=0.014, which, in contrast to (i), results in the axis of maximal spin squeezing not aligning to the readout axis J^z\hat{J}_{z} as θ\theta is dependent on λ\lambda. The corresponding OAT state for λ=λ=0.034\lambda=\lambda^{\prime}=0.034 is displayed in (k), and is aligned to the J^z\hat{J}_{z} readout axis.

The quantum state that is most robust to state preparation error in Figure 4 is the TAT state for small λ\lambda (λ=0.005\lambda=0.005). This is not only because θ\theta is independent of λ\lambda, but also because its J^x0\langle\hat{J}_{x_{0}}\rangle moment changes negligibly with λ\lambda around this region. By Equations 16 and 18, this means bias will be small, and QQ will nearly be equal to Q(Δλ=0)Q(\Delta\lambda=0). In general, thsee equations suggest spin-squeezed quantum states around which λJ^x0=0\partial_{\lambda}\langle\hat{J}_{x_{0}}\rangle=0 (and where Δλ0\Delta\lambda\neq 0 doesn’t influence later state preparation parameters) will be robust to state preparation error.

Refer to caption
Figure 4: In (a) and (e), plots of J^x0\langle\hat{J}_{x_{0}}\rangle as a function of the state preparation parameter λ\lambda for OAT and TAT, respectively. In (b) and (f), the bias coefficient BB for TAT and OAT for an assumed state preparation parameter λ\lambda^{\prime} for a set of states parameterised by varying λ\lambda. In (c), (d) and (g), (h), the single shot variance QQ, and the lower bound on error EE for OAT and TAT respectively given the same λ\lambda and λ\lambda^{\prime} values as in (b) and (f). λλ\lambda\neq\lambda^{\prime} leads to a biased estimator φ~(λp=λ)\tilde{\varphi}(\lambda_{p}=\lambda^{\prime}), leading to a biased parameter estimate, with the magnitude of the bias dependent on the discrepancy between J^x0\langle\hat{J}_{x_{0}}\rangle and Jx0^\langle\hat{J_{x_{0}}}\rangle^{\prime} for both OAT and TAT, as per Equation 16. Biased estimators lead to superior QQ and EE in (c) and (d) for TAT for λ<λ\lambda^{\prime}<\lambda, whilst the unbiased estimator is the optimal estimator for OAT in (g) and (h). This distinction is due to the state preparation parameter dependence of the J^x\hat{J}_{x} rotation for OAT, but not for TAT. In (i) we see a λ=0.02\lambda=0.02 TAT state aligned to the measurement axis J^z\hat{J}_{z} for all λ\lambda^{\prime} values, whereas in (j) and (k), we see differing alignment to the J^z\hat{J}_{z} axis for λ=0.034,λ=0.014\lambda=0.034,\lambda^{\prime}=0.014, and λ=λ=0.034\lambda=\lambda^{\prime}=0.034.

III.2 Bias-variance tradeoff

In Figure 4 (d), we see for λ<λ\lambda^{\prime}<\lambda that EE decreases below the error for a TAT interferometry sequence free from state preparation error (for Δλ=0\Delta\lambda=0, EE is saturated). Using an estimator φ~(λp=λ)\tilde{\varphi}(\lambda_{p}=\lambda^{\prime}) rather than φ~(λp=λ)\tilde{\varphi}(\lambda_{p}=\lambda) led to a decreasing lower bound on the state preparation error. Although this did not demonstrate that a biased estimator would necessarily reduce MSE, it indicated that such a scenario may arise for TAT, since we expect EE to be close to actual MSE. To investigate how using a biased estimator may decrease MSE, we fix the parameter of interest at φ=0.001\varphi=0.001, and plot the ratio of the lowest mean squared error MSEopt\text{MSE}_{\text{opt}} to MSE(λp=λ)(\lambda_{p}=\lambda) given a set of estimators φ~(λp)\tilde{\varphi}(\lambda_{p}) with λp\lambda_{p} spanning from 0 to λ\lambda, over a range of state preparation parameters and sample sizes.

Refer to caption
Figure 5: (a): the ratio of the optimal MSE for a biased estimator divided by the MSE of the unbiased estimator, for a range of λ\lambda and shot numbers mm for two-axis-twisting. The λp\lambda_{p} values parameterising the biased estimator φ~(λp)\tilde{\varphi}(\lambda_{p}) are 0λp<λ0\leq\lambda_{p}<\lambda. Biased estimators with lower noise are able to achieve increased precision over unbiased estimators, with this advantage dissipating for larger mm due to bias staying constant and variance decreasing as 1m\frac{1}{\sqrt{m}}. (b): the λp\lambda_{p} value that optimises the MSE for each state preparation value with the λ\lambda values displayed as yy-axis asymptotes. For larger λ\lambda, bias more heavily counteracts the decreased variance from the biased estimator, meaning that the optimal λp\lambda_{p} gets closer to λ\lambda for smaller mm.

In Figure 5 (a), we see that for every λ\lambda, the unbiased estimator never produces the least error - the optimal ratio between the best biased MSE and the unbiased MSE is always less than 1. We thus see that atom interferometry with TAT states would be amenable to the bias variance trade-off, where a biased estimator decreases overall error relative to an unbiased estimator because of its lower variance.

The bias variance trade-off becomes more useful for higher λ\lambda values. This owes to the fact that J^x0\langle\hat{J}_{x_{0}}\rangle is smaller for higher λ\lambda as shown in Figure 4 (a), meaning the denominator in Equation 18 can be artificially increased by picking an estimator that “overestimates” J^x0\langle\hat{J}_{x_{0}}\rangle. We note that the larger mm is, the lower the achievable bias variance trade-off. This is because as mm increases, the non-zero bias term from the biased estimator stays constant, whilst σQ\sigma_{Q} decreases as 1m\frac{1}{\sqrt{m}} - the decrease in the variance possible from choosing a biased estimator is counteracted by the bias. In Figure 5 (b), we see the λp\lambda_{p} value parameterising the MSE-optimising estimator for a range of λ\lambda and mm values. The mm value at which the optimal λpλ\lambda_{p}\approx\lambda decreases for more highly squeezed states. This is because more highly squeezed states are more susceptible to the biasing effects of state preparation error. As per Figure 4 (b), this bias is reduced by having λp\lambda_{p} closer to λ\lambda, where we interpret λ\lambda^{\prime} as being λp\lambda_{p} in this subsection. The above results suggest that biased estimators may lead to lower measurement error if TAT can be realised and manipulated in an atom interferometry experiment. For other state preparation schemes where the measurement results 𝐗\mathbf{X} are not affected by λ\lambda^{\prime}, it is likely that similar bias variance trade-offs exist.

III.3 Varying atomic ensemble size

We now investigate the effect of state preparation error on atomic ensembles large enough to produce useful measurements (N>104N>10^{4}). Numerically simulating an interferometry sequence with this number of atoms is intractable, due to the size of the Hilbert space. To probe the sensitivity of larger number of atoms under systematic state preparation error, we use the analytic expressions for the moments of OAT states in Appendix VI.2, combined with Equations 16 and 18. We assume that an experimenter seeks to produce maximally spin-squeezed states. The maximally spin-squeezed OAT state is reached at a λ\lambda value of [46]

λ(N)=241/621/3N2/3,\lambda(N)=\frac{24^{1/6}}{2^{1/3}N^{2/3}}, (23)

where the above expression is derived in the limit of small λ\lambda which is valid so long as N>>1N>>1. We define Δλcrit\Delta\lambda_{\text{crit}} as the value of |λλ||\lambda^{\prime}-\lambda| such that EE passes a certain threshold (in the case of our results, Δλ<0\Delta\lambda<0 when these thresholds are reached). This places a bound on the maximum sensitivity achievable with a NN-atom OAT ensemble under state preparation error at the λ\lambda^{\prime} where this threshold is passed for all φ\varphi operating points.

We first consider a set of thresholds relative to the quantum noise of an unbiased state preparation scheme - E(Δλ=0)E(\Delta\lambda=0). We also consider a set of thresholds relative to the amount of error achievable with an unentangled state, denoted EunE_{\text{un}}.

Refer to caption
Figure 6: (a): Amount of spin-squeezing as a function of the state preparation parameter λ\lambda for varying one-axis-twisted atomic ensemble sizes NN. (b) and (c): |λλ|λ\frac{|\lambda^{\prime}-\lambda|}{\lambda} values such that error is at least 282-8, times E(Δλ=0)E(\Delta\lambda=0), and 0.3310.33-1 times the shot-noise limit EunE_{\text{un}} respectively. In (b), the larger the ensemble size, the less state preparation error required to reduce the efficacy of spin-squeezing by a constant factor relative to the sensitivity of the state produced. This is due to greater spin-squeezing for larger atomic number, with more spin-squeezed states being more sensitive to suboptimal JxJ_{x} rotation angles. However, relative to an unentangled state in (c), we see that the robustness of OAT states for increasing atomic ensemble size is nearly constant.

We plot Δλcritλ\frac{\Delta\lambda_{\text{crit}}}{\lambda} for a range of thresholds as a function of NN. In Figure 6 (b), we see Δλcritλ\frac{\Delta\lambda_{\text{crit}}}{\lambda} for E(Δλ)/E(Δλ=0)=2E(\Delta\lambda)/E(\Delta\lambda=0)=2, 44, 66, and 88. For each of these thresholds, less percentage error in λ\lambda is required to cause an equivalent error factor as NN increases. This is because as NN increases, so does the maximal spin-squeezing one can generate, as shown in Figure 6 (a). More squeezed states are more susceptible to suboptimal J^x\hat{J}_{x} rotation angles, which means a lower percentage of systematic error is required to decrease the precision by a particular factor. In (c), we see that roughly equivalent magnitudes of state preparation error at varying NN values lead to the same precision decrease relative to constant fractions of shot noise. This suggests that squeezing under state preparation does not get much worse for large NN relative to the precision achievable with unentangled ensembles of atoms.

The values in Figure 6 (b) and (c) suggest that entangled state preparation error is a significant hindrance to precision in quantum-enhanced atom interferometry. Given the multitude of ways of realising OAT, we will not attempt to analyse state preparation error in the context of a specific experimental configuration. However, our results provide a useful starting point for an experimenter wishing to do so. More generally, they offer a schematic for how to calculate error in parameter resolutions under a particular state-generation Hamiltonian for atomic ensembles large enough to produce metrologically useful measurements.

IV Systematic state preparation error in non-Gaussian states

To systematically surpass the shot-noise limit with a non-Gaussian state, one cannot use a method of moments estimator. An estimator that will saturate the precision limit for a particular basis is the maximum likelihood estimator. The data 𝐗\mathbf{X} operated on by this estimator are the entire distribution of measurement results, rather than a moment of the results. The φ\varphi^{*} value is determined by calculating the φ\varphi value that maximises an objective function for the set of experimental results, known as a likelihood function, denoted \mathcal{L}. Maximum likelihood estimation in quantum-enhanced atom interferometry is performed according to the equation

argmaxφλp(𝐗|φ)=argmaxφi=1mlog(|Xi|ψφ,λp|2),\text{argmax}_{\varphi}\mathcal{L}_{\lambda_{p}}(\mathbf{X}|\varphi)=\text{argmax}_{\varphi}\sum^{m}_{i=1}\text{log}(|\innerproduct{X_{i}}{\psi_{\varphi,\lambda_{p}}}|^{2}), (24)

where the data are produced by the actual state preparation parameter vector 𝚲\mathbf{\Lambda}, and where |ψ\ket{\psi} is parametrised by both the estimator state preparation parameter λp\lambda_{p} and the parameter φ\varphi that we wish to optimise over. Under state preparation error where λp=λ\lambda_{p}=\lambda^{\prime} for λλ\lambda^{\prime}\neq\lambda, a biased likelihood function is optimised, a scenario termed maximum likelihood estimation under model misspecification. Using a biased likelihood function for parameter estimation will mostly lead to biased estimates of the parameter of interest.

For an unbiased estimator in the optimal basis, the single shot variance QQ aligns with the Quantum Cramér-Rao Bound (QCRB). However, under model misspecification, the single-shot variance of the estimate will not necessarily agree with the QCRB and is given by

Q2=Var(Zλp(X,φ))𝐄[φZλp(X,φ)]2,Q^{2}=\dfrac{\text{Var}(Z_{\lambda_{p}}(X,\varphi^{*}))}{\mathbf{E}[\partial_{\varphi}Z_{\lambda_{p}}(X,\varphi^{*})]^{2}}, (25)

where Zλp(X,φ)=φlog(|X|ψφ,λp|2)Z_{\lambda_{p}}(X,\varphi)=\frac{\partial}{\partial\varphi}\log(|\innerproduct{X}{\psi_{\varphi,\lambda_{p}}}|^{2}) and where XX denotes the random variable corresponding to measurement outputs. By expanding the unbiased likelihood function to second order in the state preparation degree of freedom λ\lambda, and the parameter of interest φ,\varphi, we find in Appendix VI.3 that the bias for a small error in λ\lambda is

σB=FφλFφφ(λλ),\sigma_{B}=-\dfrac{F_{\varphi\lambda}}{F_{\varphi\varphi}}(\lambda^{\prime}-\lambda), (26)

where FφλF_{\varphi\lambda} and FφφF_{\varphi\varphi} are the off-diagonal and second diagonal terms of the Fisher covariance matrix evaluated at the actual λ\lambda and φ\varphi values, the entries of a Fisher information matrix given by

Fij=miPmjPmPm.F_{ij}=\sum_{m}\dfrac{\partial_{i}P_{m}\partial_{j}P_{m}}{P_{m}}. (27)

This suggests that the ratio FφλFφφ\frac{F_{\varphi\lambda}}{F_{\varphi\varphi}} encapsulates the local susceptibility of a quantum state to state preparation error in a particular measurement basis. If this ratio is near 0 for a continuous range of λ\lambda values, the state should be immune from bias under state preparation error. Furthermore, assuming that all other state preparation parameters are independent of λ\lambda, the misspecified model will remain constant under state preparation error, leading to QQ also remaining constant.

IV.1 Non-Gaussian OAT and TNT

We investigate OAT and TNT non-Gaussian states under state preparation error using Monte Carlo techniques for small state preparation errors, and compare them to the Taylor expansion-based estimate. The TNT Hamiltonian is defined by

H^TNT=(χJ^z2+ΩJ^x)\hat{H}_{\text{TNT}}=\hbar(\chi\hat{J}_{z}^{2}+\Omega\hat{J}_{x}) (28)

where Ω=χN2\Omega=\frac{\chi N}{2} [26, 60, 61]. Non-Gaussian OAT states have yet to be produced, whilst non-Gaussian TNT states have been produced but have yet to yield subshot noise precision in an atom interferometry experiment [24]. The QFI for a non-Gaussian OAT is maximised without a J^x\hat{J}_{x} rotation, and thus the state preparation sequence is

U^OAT=eiλJ^z2.\hat{U}_{\text{OAT}}=e^{i\lambda\hat{J}_{z}^{2}}. (29)

In the case of TNT, the QFI can be increased by applying an appropriate J^x\hat{J}_{x} rotation just before parameter encoding. Whilst the optimal rotation is dependent on the state preparation parameter λ\lambda, highly metrologically useful states can be obtained with θ=1\theta=-1, with the upside of a system that is simpler to experimentally implement and analyse. This results in a TNT state preparation sequence of

U^TNT=eiJ^xei(λJ^z2+δJ^x),\hat{U}_{\text{TNT}}=e^{i\hat{J}_{x}}e^{-i(\lambda\hat{J}_{z}^{2}+\delta\hat{J}_{x})}, (30)

where λ=χt\lambda=\chi t and δ=Ωt\delta=\Omega t, with the optimal value of δ\delta being δ=λN2\delta=\frac{\lambda N}{2}. For both OAT and TNT, we seek to perform an optimal readout, i.e. the readout that saturates the QCRB. The optimal readout basis for OAT and TNT non-Gaussian states is J^x\hat{J}_{x} [62, 61, 63]. A readout in this basis can be realised through a π/2\pi/2 rotation clockwise about the J^y\hat{J}_{y}-axis, followed by a standard number-difference measurement. Thus, we attempt to find the value φ\varphi that maximises the value of the log-likelihood function

λp(𝐗|φ)=i=1i=mlog(|Xi|ei(π2φJ^y)U^OAT/TAT|ψ0|2),\mathcal{L}_{\lambda_{p}}(\mathbf{X}|\varphi)=\sum^{i=m}_{i=1}\text{log}\left(|\bra{X_{i}}e^{i(\frac{\pi}{2}-\varphi\hat{J}_{y})}\hat{U}_{\text{OAT/TAT}}\ket{\psi_{0}}|^{2}\right), (31)

for OAT and TNT respectively.

We simulate the generation, readout, and estimation of a parameter encoded onto the quantum states through Monte-Carlo sampling. We perform the simulation for states in both the spin-squeezed and non-Gaussian regimes, with the non-Gaussian regime starting for OAT at λ0.1\lambda\approx 0.1, and for TNT at λ0.045\lambda\approx 0.045. For OAT, we consider λ=λ±0.0025\lambda^{\prime}=\lambda\pm 0.0025 for 4545 λ\lambda values between 0.010.01 and 0.170.17, and for TNT, we consider λ=λ±0.0015\lambda^{\prime}=\lambda\pm 0.0015 for 4545 λ\lambda values between 0.010.01 and 0.110.11. For each λ,λ\lambda,\lambda^{\prime} pair, we perform a maximum likelihood estimate parameterised by λp=λ\lambda_{p}=\lambda^{\prime} given 10510^{5} measurements of the prepared quantum states, and repeat these 55 and 2525 times for each λ,λ\lambda,\lambda^{\prime} pair for OAT and TNT respectively. We calculate the bias, and the ratio of the quantum noise QQ to the QCRB, comparing the biased estimate to that derived from Equation 26. Given that bias is not proportional to the parameter of interest for an MLE estimator, a similarly useful lower bound to EE for spin-squeezed states cannot be devised, and the parameter of interest is fixed at φ=0.02\varphi=0.02 for the simulations.

In the case of OAT in Figure 7 (a), we see that the covariance matrix ratio in Equation 26 reliably predicts the bias in both the Gaussian and non-Gaussian regimes, where the start of the non-Gaussian regime is denoted by the dotted red line. In (b), we see that under small state preparation error in the non-Gaussian regime, OAT is essentially immune to state preparation error - Q/QCRBQ/\text{QCRB} remains constant. This is because the likelihood function being optimised changes negligibly from the unbiased likelihood function despite λpλ\lambda_{p}\neq\lambda, which can be inferred from the small FφλFφφ\frac{F_{\varphi\lambda}}{F_{\varphi\varphi}}. In the case of TNT, we see that the approximation correctly predicts bias in the Gaussian regime, but is inaccurate past early in the non-Gaussian regime. This can be attributed to higher-order derivatives being required to characterise the likelihood function even at this small amount of state preparation error - the metric of Equation 26 is of limited use in the absence of likelihood functions that change sufficiently slowly with respect to the state preparation parameter. We see that in contrast to OAT, QQ in the case of non-Gaussian TNT states strays significantly from the QCRB even for small state preparation error. Currently, the only non-Gaussian atomic states that have been generated with a significant (N>103N>10^{3}) atomic ensemble are non-Gaussian TNT state. Our results indicate that state preparation error is likely to degrade the overall precision of metrology with non-Gaussian TNT states.

Refer to caption
Figure 7: In (a) and (c) the bias for OAT and TNT states in their respective Gaussian and non-Gaussian regimes for state preparation parameter λ\lambda and assumed state preparation parameter λ\lambda^{\prime} under misspecified maximum likelihood estimation, compared with the bias predicted by Equation 26. In (b) and (d), the ratio between single-shot error QQ and the Quantum Cramér-Rao bound for OAT and TNT respectively under the same λ,λ\lambda,\lambda^{\prime} values as in (a) and (c). OAT in its non-Gaussian regime is robust to state preparation error as can be seen in both (b) and (d), and the metric Equation 26 is able to predict the bias in both Gaussian and non-Gaussian regimes. TNT in its non-Gaussian regime is not robust to state preparation error, as it is too sensitive to its state preparation parameter in the non-Gaussian for Equation 26 to be accurate for even the small Δλ\Delta\lambda chosen.

IV.2 Comparing OAT estimation strategies

OAT’s robustness to state preparation error under a maximum likelihood estimator raises an interesting proposition - that under a drifting state preparation parameter, such as in the case of phase diffusion, one would be better off dispensing with a method of moments of estimator in favour of a maximum likelihood estimator, even in the spin-squeezing regime. To investigate this, we compare the σB\sigma_{B} value for a MOM estimator and an MLE estimator for constant Δλ\Delta\lambda in the absence of any J^x\hat{J}_{x} rotation. We calculate σB\sigma_{B} for the MLE estimator through the first order Taylor expansion in Equation 26, as this approximation is reliable in the case of OAT for larger Δλ\Delta\lambda, and calculate σB\sigma_{B} for the spin-squeezed state with Equation 16. For the OAT states where an MLE estimator is used, the measurement occurs in the J^x\hat{J}_{x} basis as opposed to the J^z\hat{J}_{z} basis.

Refer to caption
Figure 8: In (a) and (b) the bias for OAT spin-squeezed states as a function of λ\lambda for varying Δλ\Delta\lambda for MOM and MLE estimators given JzJ_{z} and JxJ_{x} bases respectively. For λ>0.015\lambda>0.015, bias can be significantly reduced through measuring in the J^x\hat{J}_{x} basis and processing the measurement data with an MLE estimator, rather than measuring in J^z\hat{J}_{z} and using an MOM estimtor.

In Figure 8 (a) and (b), we see the σB\sigma_{B} value for an OAT state for varying λ\lambda and constant Δλ\Delta\lambda given a fixed φ\varphi value of 0.020.02, for an MOM and MLE estimator respectively. Whilst σB\sigma_{B} is smaller in the case of the MOM estimator for λ<0.015\lambda<0.015, the MOM estimator dramatically reduces the total bias for larger λ\lambda values. This suggests that under considerable phase diffusion for larger λ\lambda values, it is wise to use the MLE estimator instead of the canonical MOM estimator.

IV.3 Mitigating bias in quantum-enhanced atom interferometry

In both spin-squeezed and non-Gaussian states, we see that systematic state preparation error mostly leads to bias in parameter estimation. Whilst we show that there exist classes of spin-squeezed and non-Gaussian quantum states that are robust to bias from state preparation error, it may not always be possible to manufacture such a state. We present a solution that keeps the estimate of the encoded parameter φ\varphi unbiased - we discard our assumed value of the state preparation degree of freedom λ\lambda^{\prime}, and estimate both λ\lambda, and φ\varphi according to two-parameter maximum likelihood estimation.

For a two-parameter maximum likelihood estimate of a state preparation parameter λ\lambda and a relative phase φ\varphi, the sensitivity of an estimate of φ\varphi is given by one on the square root of the upper-diagonal term of the Fisher information matrix,

Q=FλλFφφFλλFφλ2.Q=\sqrt{\dfrac{F_{\lambda\lambda}}{F_{\varphi\varphi}F_{\lambda\lambda}-F_{\varphi\lambda}^{2}}}. (32)

In a non-diagonal Fisher covariance matrix, Qtwo-param>Qone-paramQ_{\text{two-param}}>Q_{\text{one-param}} - estimating both parameters at once leads to more error in φ\varphi, than using λp=λ\lambda_{p}=\lambda, and estimating φ\varphi. However, under state preparation error that induces bias, overall error may be lower under two parameter estimation for sufficient sample size.

We demonstrate that this two parameter estimation strategy can result in lower error for a sample size of m=250m=250 in both the spin-squeezed and non-Gaussian regimes in Figure 9. We plot NMSE\sqrt{N\text{MSE}} for TAT with from λ=0.005\lambda=0.005 to λ=0.015\lambda=0.015 for λ=λ+0.005\lambda^{\prime}=\lambda+0.005, and TNT for λ=0.05\lambda=0.05 to λ=0.08\lambda=0.08 and λ=λ+0.01\lambda^{\prime}=\lambda+0.01, alongside the two-parameter CRB. The TAT state is prepared according to Equation 20, and the TNT state is prepared according to Equation 30. In Figure 9 (a) and (b), we see precision for TNT and TAT under systematic state preparation error, and under two parameter maximum likelihood estimation, respectively. In both instances, we see that the two parameter estimate leads to increased precision. This indicates that using two parameter MLE where a state preparation parameter is not perfectly known is a viable method for precise quantum-enhanced atom interferometry, as well as a means of eliminating measurement bias.

Refer to caption
Figure 9: In (a), the normalised MSE for λ=λ+0.01\lambda^{\prime}=\lambda+0.01 for a TNT state, and in (b) the normalised MSE for a TAT state for λ=λ+0.005\lambda^{\prime}=\lambda+0.005, each for a sample size of m=250m=250 and φ=0.02\varphi=0.02. The normalised two-parameter Cramér-Rao bound is plotted alongside the MSE. In both cases, we see that the two parameter method is superior to the MSE under a biased single parameter-estimator.

V Conclusion

We have shown that using entangled atom states in atom interferometry may undermine its most attractive feature - the ability to make calibration-free measurements. We derive analytic expressions for both the bias and mean-squared error under state preparation in the spin-squeezed regime, and a metric for local susceptibility to state preparation error under a maximum likelihood estimator that can be used in the non-Gaussian regime. This allows the exploration of the susceptibility of quantum states produced by important Hamilitonians such as OAT, TAT, and TNT, to state preparation error and to devise mathematical conditions on quantum states that are robust to state preparation error. We demonstrate how precision can counterintuitively be increased under a biased estimator in the absence of additional interferometer operations dependent on the state preparation parameter in the case of TAT. Finally, we present a means of retaining the accuracy of atom interferometry under state preparation error using two-parameter maximum likelihood estimation, which may come at the cost of precision for small sample sizes.

Acknowledgements.
We acknowledge the Ngunnawal people as the traditional custodians of the land upon which this research was conducted, and recognise that sovereignty was never ceded. We would like to acknowledge fruitful discussions had with Karandeep Gill, Zain Mehdi and John Close. SAH acknowledges support through an Australian Research Council Future Fellowship Grant No. FT210100809.

References

  • Keith et al. [1991] D. W. Keith, C. R. Ekstrom, Q. A. Turchette, and D. E. Pritchard, An interferometer for atoms, Phys. Rev. Lett. 66, 2693 (1991).
  • Kasevich and Chu [1991] M. Kasevich and S. Chu, Atomic interferometry using stimulated Raman transitions, Phys. Rev. Lett. 67, 181 (1991).
  • Braje et al. [2014] D. Braje, S. DeSavage, C. Adler, J. Davis, and F. Narducci, A frequency selective atom interferometer magnetometer, Journal of Modern Optics 61, 61–71 (2014).
  • Vengalattore et al. [2007] M. Vengalattore, J. M. Higbie, S. R. Leslie, J. Guzman, L. E. Sadler, and D. M. Stamper-Kurn, High-resolution magnetometry with a spinor Bose-Einstein condensate, Phys. Rev. Lett. 98, 200801 (2007).
  • Canuel et al. [2006] B. Canuel, F. Leduc, D. Holleville, A. Gauguet, J. Fils, A. Virdis, A. Clairon, N. Dimarcq, C. J. Bordé, A. Landragin, and P. Bouyer, Six-axis inertial sensor using cold-atom interferometry, Phys. Rev. Lett. 97, 010402 (2006).
  • Gustavson et al. [1997] T. L. Gustavson, P. Bouyer, and M. A. Kasevich, Precision rotation measurements with an atom interferometer gyroscope, Phys. Rev. Lett. 78, 2046 (1997).
  • Peters et al. [2001] A. Peters, K. Y. Chung, and S. Chu, High-precision gravity measurements using atom interferometry, Metrologia 38, 25 (2001).
  • Hu et al. [2013] Z.-K. Hu, B.-L. Sun, X.-C. Duan, M.-K. Zhou, L.-L. Chen, S. Zhan, Q.-Z. Zhang, and J. Luo, Demonstration of an ultrahigh-sensitivity atom-interferometry absolute gravimeter, Phys. Rev. A 88, 043610 (2013).
  • Canuel et al. [2018] B. Canuel, A. Bertoldi, L. Amand, E. Pozzo di Borgo, T. Chantrait, C. Danquigny, M. Dovale Alvarez, B. Fang, A. Freise, R. Geiger, J. Gillot, S. Henry, J. Hinderer, D. Holleville, J. Junca, G. Lefèvre, M. Merzougui, N. Mielec, T. Monfret, S. Pelisson, M. Prevedelli, S. Reynaud, I. Riou, Y. Rogister, S. Rosat, E. Cormier, A. Landragin, W. Chaibi, S. Gaffet, and P. Bouyer, Exploring gravity with the MIGA large scale atom interferometer, Scientific Reports 8, 14064 (2018).
  • Snadden et al. [1998] M. J. Snadden, J. M. McGuirk, P. Bouyer, K. G. Haritos, and M. A. Kasevich, Measurement of the earth’s gravity gradient with an atom interferometer-based gravity gradiometer, Phys. Rev. Lett. 81, 971 (1998).
  • Geiger et al. [2011] R. Geiger, V. Ménoret, G. Stern, N. Zahzam, P. Cheinet, B. Battelier, A. Villing, F. Moron, M. Lours, Y. Bidel, A. Bresson, A. Landragin, and P. Bouyer, Detecting inertial effects with airborne matter-wave interferometry, Nature Communications 2 (2011).
  • Geiger et al. [2020] R. Geiger, A. Landragin, S. Merlet, and F. P. D. Santos, High-accuracy inertial measurements with cold-atom sensors, AVS Quantum Science 2, 024702 (2020).
  • Parker et al. [2018] R. H. Parker, C. Yu, W. Zhong, B. Estey, and H. Müller, Measurement of the fine-structure constant as a test of the standard model, Science 360, 191 (2018).
  • Rosi et al. [2014] G. Rosi, F. Sorrentino, L. Cacciapuoti, M. Prevedelli, and G. M. Tino, Precision measurement of the Newtonian gravitational constant using cold atoms, Nature 510, 518 (2014).
  • Pezze et al. [2018] L. Pezze, A. Smerzi, M. K. Oberthaler, R. Schmied, and P. Treutlein, Quantum metrology with nonclassical states of atomic ensembles, Rev. Mod. Phys. 90, 035005 (2018).
  • Szigeti et al. [2020] S. S. Szigeti, S. P. Nolan, J. D. Close, and S. A. Haine, High-precision quantum-enhanced gravimetry with a Bose-Einstein condensate, Phys. Rev. Lett. 125, 100402 (2020).
  • Szigeti et al. [2021] S. S. Szigeti, O. Hosten, and S. A. Haine, Improving cold-atom sensors with quantum entanglement: Prospects and challenges, Applied Physics Letters 118, 140501 (2021).
  • Wang et al. [2023] X. Wang, A. Kealy, C. Gilliam, S. Haine, J. Close, B. Moran, K. Talbot, S. Williams, K. Hardman, C. Freier, P. Wigley, A. White, S. Szigeti, and S. Legge, Improving measurement performance via fusion of classical and quantum accelerometers, The Journal of Navigation 76, 91 (2023).
  • Gross et al. [2010] C. Gross, T. Zibold, E. Nicklas, J. Esteve, and M. K. Oberthaler, Nonlinear atom interferometer surpasses classical precision limit, Nature 464, 1165 (2010).
  • Riedel et al. [2010] M. F. Riedel, P. Böhi, Y. Li, T. W. Hänsch, A. Sinatra, and P. Treutlein, Atom-chip-based generation of entanglement for quantum metrology, Nature 464, 1170 (2010).
  • Lücke et al. [2011] B. Lücke, M. Scherer, J. Kruse, L. Pezze, F. Deuretzbacher, P. Hyllus, O. Topic, J. Peise, W. Ertmer, J. Arlt, L. Santos, A. Smerzi, and C. Klempt, Twin matter waves for interferometry beyond the classical limit, Science 334, 773 (2011).
  • Hamley et al. [2012] C. D. Hamley, C. S. Gerving, T. M. Hoang, E. M. Bookjans, and M. S. Chapman, Spin-nematic squeezed vacuum in a quantum gas, Nat Phys 8, 305 (2012).
  • Berrada et al. [2013] T. Berrada, S. van Frank, R. Bücker, T. Schumm, J. F. Schaff, and J. Schmiedmayer, Integrated Mach–Zehnder interferometer for Bose–Einstein condensates, Nature Communications 4, 2077 EP (2013).
  • Strobel et al. [2014] H. Strobel, W. Muessel, D. Linnemann, T. Zibold, D. B. Hume, L. Pezzè, A. Smerzi, and M. K. Oberthaler, Fisher information and entanglement of non-Gaussian spin states, Science 345, 424 (2014).
  • Muessel et al. [2014] W. Muessel, H. Strobel, D. Linnemann, D. B. Hume, and M. K. Oberthaler, Scalable spin squeezing for quantum-enhanced magnetometry with Bose-Einstein condensates, Phys. Rev. Lett. 113, 103004 (2014).
  • Muessel et al. [2015] W. Muessel, H. Strobel, D. Linnemann, T. Zibold, B. Juliá-Díaz, and M. K. Oberthaler, Twist-and-turn spin squeezing in Bose-Einstein condensates, Phys. Rev. A 92, 023603 (2015).
  • Kruse et al. [2016] I. Kruse, K. Lange, J. Peise, B. Lücke, L. Pezzè, J. Arlt, W. Ertmer, C. Lisdat, L. Santos, A. Smerzi, and C. Klempt, Improvement of an atomic clock using squeezed vacuum, Phys. Rev. Lett. 117, 143004 (2016).
  • Zou et al. [2018] Y.-Q. Zou, L.-N. Wu, Q. Liu, X.-Y. Luo, S.-F. Guo, J.-H. Cao, M. K. Tey, and L. You, Beating the classical precision limit with spin-1 Dicke states of more than 10,000 atoms, Proceedings of the National Academy of Sciences 115, 6381 (2018).
  • Laudat et al. [2018] T. Laudat, V. Dugrain, T. Mazzoni, M.-Z. Huang, C. L. G. Alzar, A. Sinatra, P. Rosenbusch, and J. Reichel, Spontaneous spin squeezing in a Rubidium BEC, New Journal of Physics 20, 073018 (2018).
  • Appel et al. [2009] J. Appel, P. J. Windpassinger, D. Oblak, U. B. Hoff, N. Kjaergaard, and E. S. Polzik, Mesoscopic atomic entanglement for precision measurements beyond the standard quantum limit, Proceedings of the National Academy of Sciences 106, 10960 (2009).
  • Leroux et al. [2010] I. D. Leroux, M. H. Schleier-Smith, and V. Vuletic, Implementation of cavity squeezing of a collective atomic spin, Phys. Rev. Lett. 104, 073602 (2010).
  • Schleier-Smith et al. [2010] M. H. Schleier-Smith, I. D. Leroux, and V. Vuletic, Squeezing the collective spin of a dilute atomic ensemble by cavity feedback, Phys. Rev. A 81, 021804 (2010).
  • Sewell et al. [2012] R. J. Sewell, M. Koschorreck, M. Napolitano, B. Dubost, N. Behbood, and M. W. Mitchell, Magnetic sensitivity beyond the projection noise limit by spin squeezing, Phys. Rev. Lett. 109, 253605 (2012).
  • Hosten et al. [2016] O. Hosten, N. J. Engelsen, R. Krishnakumar, and M. A. Kasevich, Measurement noise 100 times lower than the quantum-projection limit using entangled atoms, Nature 529, 505 EP (2016).
  • Greve et al. [2022] G. P. Greve, C. Luo, B. Wu, and J. K. Thompson, Entanglement-enhanced matter-wave interferometry in a high-finesse cavity, Nature 610, 472 (2022).
  • Hammerer et al. [2010] K. Hammerer, A. S. Sørensen, and E. S. Polzik, Quantum interface between light and atomic ensembles, Rev. Mod. Phys. 82, 1041 (2010).
  • Haine and Ferris [2011] S. A. Haine and A. J. Ferris, Surpassing the standard quantum limit in an atom interferometer with four-mode entanglement produced from four-wave mixing, Phys. Rev. A 84, 043624 (2011).
  • Haine [2013] S. A. Haine, Information-recycling beam splitters for quantum enhanced atom interferometry, Phys. Rev. Lett. 110, 053002 (2013).
  • Haine et al. [2014] S. A. Haine, J. Lau, R. P. Anderson, and M. T. Johnsson, Self-induced spatial dynamics to enhance spin squeezing via one-axis twisting in a two-component Bose-Einstein condensate, Phys. Rev. A 90, 023613 (2014).
  • Haine and Lau [2016] S. A. Haine and W. Y. S. Lau, Generation of atom-light entanglement in an optical cavity for quantum enhanced atom interferometry, Phys. Rev. A 93, 023607 (2016).
  • Szigeti et al. [2017] S. S. Szigeti, R. J. Lewis-Swan, and S. A. Haine, Pumped-up SU(1,1) interferometry, Phys. Rev. Lett. 118, 150401 (2017).
  • Lewis-Swan et al. [2018] R. J. Lewis-Swan, M. A. Norcia, J. R. K. Cline, J. K. Thompson, and A. M. Rey, Robust spin squeezing via photon-mediated interactions on an optical clock transition, Phys. Rev. Lett. 121, 070403 (2018).
  • Haine and Hope [2020] S. A. Haine and J. J. Hope, Machine-designed sensor to make optimal use of entanglement-generating dynamics for quantum sensing, Phys. Rev. Lett. 124, 060402 (2020).
  • Fuderer et al. [2023] L. A. Fuderer, J. J. Hope, and S. A. Haine, Hybrid method of generating spin-squeezed states for quantum-enhanced atom interferometry, Phys. Rev. A 108, 043722 (2023).
  • Barberena et al. [2024] D. Barberena, A. Chu, J. K. Thompson, and A. M. Rey, Trade-offs between unitary and measurement induced spin squeezing in cavity QED, Phys. Rev. Res. 6, L032037 (2024).
  • Kitagawa and Ueda [1993] M. Kitagawa and M. Ueda, Squeezed spin states, Phys. Rev. A 47, 5138 (1993).
  • Haine and Johnsson [2009] S. A. Haine and M. T. Johnsson, Dynamic scheme for generating number squeezing in Bose-Einstein condensates through nonlinear interactions, Phys. Rev. A 80, 023611 (2009).
  • Kuzmich et al. [2000] A. Kuzmich, L. Mandel, and N. P. Bigelow, Generation of spin squeezing via continuous quantum nondemolition measurement, Phys. Rev. Lett. 85, 1594 (2000).
  • Kritsotakis et al. [2021] M. Kritsotakis, J. A. Dunningham, and S. A. Haine, Spin squeezing of a Bose-Einstein condensate via a quantum nondemolition measurement for quantum-enhanced atom interferometry, Phys. Rev. A 103, 023318 (2021).
  • Fisher [1922] R. A. Fisher, On the mathematical foundations of theoretical statistics, Philosophical Transactions of the Royal Society of London. Series A, Containing Papers of a Mathematical or Physical Character 222, 309 (1922).
  • Braunstein and Caves [1994] S. L. Braunstein and C. M. Caves, Statistical distance and the geometry of quantum states, Phys. Rev. Lett. 72, 3439 (1994).
  • Yurke et al. [1986] B. Yurke, S. L. McCall, and J. R. Klauder, SU(2) and SU(1,1) interferometers, Phys. Rev. A 33, 4033 (1986).
  • Pezzè et al. [2018] L. Pezzè, A. Smerzi, M. Oberthaler, R. Schmied, and P. Treutlein, Quantum metrology with nonclassical states of atomic ensembles, Reviews of Modern Physics 90 (2018).
  • Wineland et al. [1992] D. J. Wineland, J. J. Bollinger, W. M. Itano, F. L. Moore, and D. J. Heinzen, Spin squeezing and reduced quantum noise in spectroscopy, Phys. Rev. A 46, R6797 (1992).
  • Ma and Wang [2009] J. Ma and X. Wang, Fisher information and spin squeezing in the Lipkin-Meshkov-Glick model, Phys. Rev. A 80, 012318 (2009).
  • Estève et al. [2008] J. Estève, C. Gross, A. Weller, S. Giovanazzi, and M. K. Oberthaler, Squeezing and entanglement in a Bose-Einstein condensate, Nature 455, 1216 (2008).
  • Sørensen and Mølmer [2001] A. S. Sørensen and K. Mølmer, Entanglement and extreme spin squeezing, Phys. Rev. Lett. 86, 4431 (2001).
  • Luo et al. [2024] C. Luo, H. Zhang, A. Chu, C. Maruko, A. Rey, and J. Thopmpson, Hamiltonian engineering of collective XYZ spin models in an optical cavity, arXiv:2402.19429  (2024).
  • Haine and Szigeti [2015] S. A. Haine and S. S. Szigeti, Quantum metrology with mixed states: When recovering lost information is better than never losing it, Phys. Rev. A 92, 032317 (2015).
  • Law et al. [2001] C. K. Law, H. T. Ng, and P. T. Leung, Coherent control of spin squeezing, Phys. Rev. A 63, 055601 (2001).
  • Mirkhalaf et al. [2018] S. S. Mirkhalaf, S. P. Nolan, and S. A. Haine, Robustifying twist-and-turn entanglement with interaction-based readout, Phys. Rev. A 97, 053618 (2018).
  • Nolan and Haine [2017] S. P. Nolan and S. A. Haine, Quantum Fisher information as a predictor of decoherence in the preparation of spin-cat states for quantum metrology, Phys. Rev. A 95, 043642 (2017).
  • Haine [2018] S. A. Haine, Using interaction-based readouts to approach the ultimate limit of detection-noise robustness for quantum-enhanced metrology in collective spin systems, Phys. Rev. A 98, 030303 (2018).

VI Appendixes

In part A, we derive the expression for the quantum noise in a parameter of interest in a spin-squeezed state under state preparation error. In part B, we discuss how Kitagawa and Ueda’s analytic expressions for OAT moments are used in Section 3 C. In part C, we derive an analytic expression for the bias of a maximum likelihood estimate under model misspecification. In part D, we briefly consider random, as opposed to systematic state preparation error, for TAT and OAT spin-squeezed states.

VI.1 Deriving Equation 18

We wish to calculate

σQ2=Var(J^z)m|φJz^|2,\sigma_{Q}^{2}=\dfrac{\text{Var}(\hat{J}_{z})}{m|\partial_{\varphi}\langle\hat{J_{z}}\rangle|^{2}}, (33)

in terms of the J^x\hat{J}_{x} rotation θ\theta, and the encoded parameter φ\varphi. We take Equation 13, and compute the variance on the Heisenberg evolved operators as

Var(J^z)=cos2(θ)cos2(φ)Var(J^z0)+sin2(φ)Var(J^x0)+cos2(φ)sin2(θ)Var(J^y0)12sin(θ)sin(2φ)Covar¯(J^x0,J^y0)12sin(2φ)cos(θ)Covar¯(J^z0,J^x0)+12sin(2θ)cos2(φ)Covar¯(J^z0,J^y0),\text{Var}(\hat{J}_{z})=\cos^{2}(\theta)\cos^{2}(\varphi)\text{Var}(\hat{J}_{z_{0}})+\sin^{2}(\varphi)\text{Var}(\hat{J}_{x_{0}})+\cos^{2}(\varphi)\sin^{2}(\theta)\text{Var}(\hat{J}_{y_{0}})-\frac{1}{2}\sin(\theta)\sin(2\varphi)\overline{\text{Covar}}(\hat{J}_{x_{0}},\hat{J}_{y_{0}})\\ -\frac{1}{2}\sin(2\varphi)\cos(\theta)\overline{\text{Covar}}(\hat{J}_{z_{0}},\hat{J}_{x_{0}})+\frac{1}{2}\sin(2\theta)\cos^{2}(\varphi)\overline{\text{Covar}}(\hat{J}_{z_{0}},\hat{J}_{y_{0}}), (34)

where Covar¯(a,b)=Covar(a,b)+Covar(b,a)\overline{\text{Covar}}(a,b)=\text{Covar}(a,b)+\text{Covar}(b,a). Given that the initial spin-squeezed state is roughly symmetric about the JxJ_{x}-axis, the covariance terms in x,zx,z and x,yx,y will be approximately zero. This, in combination with taking the expected value of Equation 13 allows us to write the standard error in φ\varphi based on Equation 33 as

σQ2=1m|cos(φ)J^x0+sin(φ)(J^z0cos(θ)+J^y0sin(θ))|2(cos2(θ)cos2(φ)Var(J^z0)+sin2(φ)Var(J^x0)+cos2(φ)(sin2(θ)Var(J^y0)+12sin(2θ)Covar¯(J^z0,J^y0))).\sigma_{Q}^{2}=\dfrac{1}{m|\cos(\varphi)\langle\hat{J}_{x_{0}}\rangle+\sin(\varphi)(\langle\hat{J}_{z_{0}}\rangle\cos(\theta)+\langle\hat{J}_{y_{0}}\rangle\sin(\theta))|^{2}}\Bigg{(}\cos^{2}(\theta)\cos^{2}(\varphi)\text{Var}(\hat{J}_{z_{0}})+\sin^{2}(\varphi)\text{Var}(\hat{J}_{x_{0}})\\ +\cos^{2}(\varphi)\left(\sin^{2}(\theta)\text{Var}(\hat{J}_{y_{0}})+\frac{1}{2}\sin(2\theta)\overline{\text{Covar}}(\hat{J}_{z_{0}},\hat{J}_{y_{0}})\right)\Bigg{)}. (35)

By picking our initial state to be a CSS to be a J^x\hat{J}_{x} eigenstate as before, we can set J^y0=0\langle\hat{J}_{y_{0}}\rangle=0, J^z0=0\langle\hat{J}_{z_{0}}\rangle=0. Linearising around φ=0\varphi=0, we obtain

σQ2=1m|J^x0|2(cos2(θ)Var(J^z0)+sin2(θ)Var(J^y0)+12sin(2θ)Covar¯(J^z0,J^y0)).\begin{split}\sigma_{Q}^{2}&=\dfrac{1}{m|\langle\hat{J}_{x_{0}}\rangle|^{2}}\left(\cos^{2}(\theta)\text{Var}(\hat{J}_{z_{0}})+\sin^{2}(\theta)\text{Var}(\hat{J}_{y_{0}})\right.\\ &\left.+\frac{1}{2}\sin(2\theta)\overline{\text{Covar}}(\hat{J}_{z_{0}},\hat{J}_{y_{0}})\right).\end{split} (36)

as desired.

VI.2 Analytic expressions for OAT moments

We use analytic expressions to evaluate the spin-moments of OAT developed by Kitagawa and Ueda [46]. Using S¯i\bar{S}_{i} instead of J^i\hat{J}_{i} to denote a collective spin operator in a direction ii, for S=N/2,S=N/2, they found that

S¯x\displaystyle\left\langle\bar{S}_{x}\right\rangle =Scos2S1μ2,S¯y=0,S¯z=0,\displaystyle=S\cos^{2S-1}\frac{\mu}{2},\quad\left\langle\bar{S}_{y}\right\rangle=0,\quad\left\langle\bar{S}_{z}\right\rangle=0,
ΔS¯x2\displaystyle\left\langle\Delta\bar{S}_{x}^{2}\right\rangle =S2[2S(1cos2(2S1)μ2)(S12)A],\displaystyle=\frac{S}{2}\left[2S\left(1-\cos^{2(2S-1)}\frac{\mu}{2}\right)-\left(S-\frac{1}{2}\right)A\right],
ΔS¯y,z2=S2{1+12(S12)×[A±A2+B2cos(2ν+2δ)]}.\displaystyle\begin{split}\left\langle\Delta\bar{S}_{y,z}^{2}\right\rangle&=\frac{S}{2}\Bigg{\{}1+\frac{1}{2}\left(S-\frac{1}{2}\right)\\ &\quad\times\left[A\pm\sqrt{A^{2}+B^{2}}\cos(2\nu+2\delta)\right]\Bigg{\}}.\end{split}

where μ=2λ\mu=2\lambda, A=1cos2S2μ,B=4sinμ2cos2S2μ2A=1-\cos^{2S-2}\mu,B=4\sin\frac{\mu}{2}\cos^{2S-2}\frac{\mu}{2}, and δ=12arctanBA\delta=\frac{1}{2}\arctan\frac{B}{A}. ν\nu is the actual rotation angle about the xx axis, δ\delta is the assumed angle that is applied to the OAT state in the simulation. The assumed angle ν\nu applied in the simulation is πδ\pi-\delta^{\prime}, where δ\delta is based on the assumed state preparation parameter λ\lambda.

VI.3 Derivation of analytic expressions for MLE bias

Consider a likelihood function

λp(𝐗(𝚲)|φ)=klog(P(Xk|φ,λp)),\mathcal{L}_{\lambda_{p}}(\mathbf{X(\mathbf{\Lambda})}|\varphi)=\sum_{k}\log(P(X_{k}|\varphi,\lambda_{p})), (37)

where 𝐗\mathbf{X} represents a set of measurements dependent on the actual state of state parameters 𝚲\mathbf{\Lambda}, PP represents a probability distribution parametrised by the estimator state preparation parameter λp\lambda_{p}, and φ\varphi is the parameter of interest. We can expand \mathcal{L} to second order around its maximum value φ0\varphi_{0} as

\displaystyle\mathcal{L} =(λp,φ0)+λp(λpλ0)+φ(φφ0)\displaystyle=\mathcal{L}(\lambda_{p},\varphi_{0})+\dfrac{\partial\mathcal{L}}{\partial\lambda_{p}}(\lambda_{p}-\lambda_{0})+\dfrac{\partial\mathcal{L}}{\partial\varphi}(\varphi-\varphi_{0})
+12(2λp2(λpλ0)2+2φ2(φφ0)2\displaystyle\quad+\dfrac{1}{2}\left(\dfrac{\partial^{2}\mathcal{L}}{\partial\lambda_{p}^{2}}(\lambda_{p}-\lambda_{0})^{2}+\dfrac{\partial^{2}\mathcal{L}}{\partial\varphi^{2}}(\varphi-\varphi_{0})^{2}\right.
+22λpφ(λpλ0)(φφ0)).\displaystyle\qquad\left.+2\dfrac{\partial^{2}\mathcal{L}}{\partial\lambda_{p}\partial\varphi}(\lambda_{p}-\lambda_{0})(\varphi-\varphi_{0})\right).

The derivative around the maximum with respect to φ\varphi will be 0, so we can write

φ=2φ2(φφ0)+2λpφ(λpλ0)=0\begin{split}\dfrac{\partial\mathcal{L}}{\partial\varphi}=\dfrac{\partial^{2}\mathcal{L}}{\partial\varphi^{2}}(\varphi-\varphi_{0})+\dfrac{\partial^{2}\mathcal{L}}{\partial\lambda_{p}\partial\varphi}(\lambda_{p}-\lambda_{0})=0\end{split} (38)

from which we can calculate the bias as

φφ0=2λpφ2φ2(λpλ0).\varphi-\varphi_{0}=-\dfrac{\frac{\partial^{2}\mathcal{L}}{\partial\lambda_{p}\partial\varphi}}{\frac{\partial^{2}\mathcal{L}}{\partial\varphi^{2}}}(\lambda_{p}-\lambda_{0}). (39)

The numerator and denominator turn out to be the off-diagonal and diagonal terms of the Fisher covariance matrix respectively, thus we can write the above equation as

φφ0=FλφFφφ(λpλ0),\varphi-\varphi_{0}=-\dfrac{F_{\lambda\varphi}}{F_{\varphi\varphi}}(\lambda_{p}-\lambda_{0}), (40)

so we can interpret the ratio of the off-diagonal and φ\varphi elements of the Fisher covariance as a measure of the susceptibility of a quantum state in a particular basis to state preparation error.

Given that λp=λ\lambda_{p}=\lambda^{\prime}, the assumed state preparation degree of freedom, and λ0=λ\lambda_{0}=\lambda, the actual state preparation degree of freedom, we can write that

φφ=FλφFφφ(λλ).\varphi^{*}-\varphi=\dfrac{F_{\lambda\varphi}}{F_{\varphi\varphi}}(\lambda^{\prime}-\lambda). (41)

VI.4 Random state preparation error in spin-squeezed states

Let us consider random state preparation error with multiple state preparation parameters 𝚲\mathbf{\Lambda}, but only one degree of freedom λ\lambda that fluctuates randomly with a known standard deviation Δλ\Delta\lambda. This Δλ\Delta\lambda is defined separately from the Δλ\Delta\lambda in the main text. The randomness in λ\lambda results in the transition from a pure to a mixed state. After parameter encoding, the state before readout can be described as

ρ^𝚲,φ\displaystyle\hat{\rho}_{\mathbf{\Lambda},\varphi} =U^φ𝑑λ12πΔλ\displaystyle=\hat{U}_{\varphi}\int d\lambda\,\dfrac{1}{\sqrt{2\pi\Delta\lambda}} (42)
×exp((λλ0)22Δλ)|ψ𝚲ψ𝚲|U^φ,\displaystyle\quad\times\exp\left(-\frac{(\lambda-\lambda_{0})^{2}}{2\Delta\lambda}\right)\ket{\psi_{\mathbf{\Lambda}}}\bra{\psi_{\mathbf{\Lambda}}}\hat{U}^{\dagger}_{\varphi},

where |ψ𝚲ψ𝚲|\ket{\psi_{\mathbf{\Lambda}}}\bra{\psi_{\mathbf{\Lambda}}} represents the density matrix of a pure quantum state entangled according to 𝚲\mathbf{\Lambda}. Assuming that the experimenter has complete knowledge of the mixed state prepared, there will be no systematic error in the parameter estimate, i.e. B=0B=0.

We assume that our state preparation parameter vector is 𝚲={λ,θ}\mathbf{\Lambda}=\{\lambda,\theta\}, where λ\lambda is again the only degree of freedom. For a spin-squeezed state, the phase can be estimated based off J^z\langle\hat{J}_{z}\rangle as in Figure 3 under an estimator accounting for the state preparation parameter fluctuations, and the variance can be calculated as

Δφ=ΔJzm|φ𝐓𝐫]ρ^λJz^]|.\Delta\varphi=\dfrac{\Delta J_{z}}{\sqrt{m}|\partial_{\varphi}\mathbf{Tr}]\hat{\rho}_{\lambda}\hat{J_{z}}]|}. (43)

We consider the sensitivity ratio for varying levels of mixed state preparation error for TAT and OAT. In the case of OAT, we apply the optimal J^x\hat{J}_{x} rotation for the mean state preparation value in the distribution, λ0\lambda_{0}. We find that increased uncertainty in the λ\lambda value leads to greater state preparation error for both OAT and TAT under a method of moments estimator. In Figure 10 (a) for TAT and (b) for OAT, as Δλ\Delta\lambda increases, the variance in J^x\hat{J}_{x} increases. This is because the state’s total variance takes the possible range of JzJ_{z} and JxJ_{x} distributions given by the randomly distributed λ\lambda. Given that the state is squeezed along the JzJ_{z} axis, the variance of the mixed state for all fixed values of λ0\lambda_{0} will be greater in J^x\hat{J}_{x} than in J^z\hat{J}_{z}, and monotonically increasing in both, leading to an increase in error according to Equation 18.

In the cases of both OAT and TAT, we see that a maximum likelihood estimation strategy improves on a method of moments strategy for Δλ0\Delta\lambda\neq 0. This is because more highly mixed states tend away from being Gaussian, where the quantum state can be characterised by a single moment, and thus the CRB can be saturated by a MOM estimator. When using a maximum likelihood estimation strategy, we see a distinction between OAT and TAT. In the case of TAT, increased uncertainly in λ\lambda does not necessarily decrease precision. This is because as λ\lambda increases, less spin-squeezed state are mixed in with more highly spin-squeezed states. The J^z\hat{J}_{z} distribution becomes weighted with the more highly distinguishable distributions of spin-squeezed states, resulting in a distribution that is overall more distinguishable, thus having higher Fisher information than mixed states with lower Δλ\Delta\lambda values. In the case of OAT on the other hand, we find increased state preparation uncertainty leads to uniformly increasing QQ. This is because the phase space rotation after squeezing varies for the state preparation parameter, unlike for TAT. The varied state preparation parameter results in a mixture of states that are not all squeezed along the same axis, as the optimal J^x\hat{J}_{x} rotation can not be applied for all states - far fewer highly distinguishable distributions become part of the new probability distribution.

Refer to caption
Figure 10: In (a) and (b), the scaled quantum noise for TAT and OAT states respectively for varying magnitudes of state preparation parameter Δλ\Delta\lambda noise. The solid lines indicate the normalised φ\varphi noise for a method of moments estimator, while the dotted lines in the same colour indicate the phase sensitivity of the state according the CRB, i.e. the sensitivity attainable with a maximum likelihood estimate. In both (a) and (b), we see that MSE increases monotonically for increasing state preparation parameter fluctuations. This is due to the increase of variance in the pseudo-spin operators as a result of a mixture of a wide variety of states. In (a) for TAT, we see that a maximum likelihood estimate diverges from the MOM estimate due to our mixed state no longer being Gaussian, and we see higher Δλ\Delta\lambda leading to lower noise in some cases as higher CFI states are mixed in. In (b) for OAT, the maximum likelihood estimate improves on the MOM estimate, but QQ is monotonically increasing, as all but one state in the OAT fixture are rotated by a suboptimal angle, meaning comparatively lower CFI states.