Paper The following article is Open access

Revisiting 129Xe electric dipole moment measurements applying a new global phase fitting approach

, , , , , , , and

Published 24 June 2021 © 2021 The Author(s). Published by IOP Publishing Ltd on behalf of the Institute of Physics and Deutsche Physikalische Gesellschaft
, , Citation Tianhao Liu et al 2021 New J. Phys. 23 063076 DOI 10.1088/1367-2630/ac09ca

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

1367-2630/23/6/063076

Abstract

By measuring the spin precession frequencies of polarized 129Xe and 3He, a new upper limit on the 129Xe atomic electric dipole moment (EDM) ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$ was reported in Sachdev et al (2019 Phys. Rev. Lett. 123, 143003). Here, we propose a new evaluation method based on global phase fitting (GPF) for analyzing the continuous phase development of the 3He–129Xe comagnetometer signal. The Cramer–Rao lower bound on the 129Xe EDM for the GPF method is theoretically derived and shows the potential benefit of our new approach. The robustness of the GPF method is verified with Monte-Carlo studies. By optimizing the analysis parameters and adding data that could not be analyzed with the former method, we obtain a result of ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)=\left[1.1{\pm}3.6\enspace \left(\mathrm{s}\mathrm{t}\mathrm{a}\mathrm{t}\right){\pm}2.0\enspace \left(\mathrm{s}\mathrm{y}\mathrm{s}\mathrm{t}\right)\right]{\times}1{0}^{-28}\enspace \text{e}\enspace \mathrm{c}\mathrm{m}$ in an unblinded analysis. For the systematic uncertainty analyses, we adopted all methods from the aforementioned PRL publication except the comagnetometer phase drift, which can be omitted using the GPF method. The updated null result can be interpreted as a new upper limit of $\vert {d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)\vert {< }8.3\enspace {\times}1{0}^{-28}\enspace \text{e}\enspace \mathrm{c}\mathrm{m}$ at the 95% C.L.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

A quantum field theory that models the formation of the imbalance of matter over antimatter in our universe must fulfill the Sakharov conditions [1, 2]. One of those conditions is the $\mathcal{CP}$ violation ($\mathcal{C}$ is charge conjugation and $\mathcal{P}$ is parity reversal). The best-tested standard model (SM) of particle physics provides two sources of $\mathcal{CP}$ violation, the phase of the Cabibbo–Kobayashi–Maskawa matrix and the term $\bar{\theta }$ in the QCD Lagrangian [3]. However, the $\mathcal{CP}$ violation within the SM is too small to produce the observed rate of the matter to antimatter asymmetry, motivating searches for physics beyond-the-SM (BSM). BSM theories generally include additional sources of $\mathcal{CP}$ violation [3, 4], producing a larger permanent electric dipole moment (EDM) of fundamental or composite particles [2, 5]. So far, all measurement results of EDMs in more than ten diverse systems, with the first published in 1957 [6] 4 , are consistent with zero. These null results are interpreted as upper limits on EDMs and place constraints on various sources of $\mathcal{CP}$ violation and masses of BSM particles, thus directing the search of BSM scenarios [5, 8].

Long spin-coherence time and obtainable high polarization leading to high signal-to-noise ratios (SNR) make several diamagnetic systems such as the 199Hg and 129Xe atom promising candidates for EDM experiments. Over the last 40 years, significant progress was made in the determination of upper limits for EDMs of diamagnetic systems (see figure 1). At present, the 199Hg atomic EDM measurement is the most sensitive, and its upper limit sets constraints to multiple sources of $\mathcal{CP}$ violation [9]. Considering various potential contributions to an atomic EDM, an improved limit on other systems, like the 129Xe EDM ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$, will tighten these constraints. The theoretical results for 129Xe EDM are more accurate and reliable than those obtained for 199Hg EDM, therefore 129Xe has the potential to probe new physics in the future, even though the experimental accuracy may not reach the actual 199Hg sensitivity [10].

Figure 1.

Figure 1. Selected upper bounds of EDM of diamagnetic systems performed since 1980 at the 95% C.L. For all systems, the current upper bound has decreased more than an order of magnitude compared to their first published result (TIF [11], 225Ra [12], Neutron [13], 129Xe [1417], 199Hg [9]).

Standard image High-resolution image

Recently, new upper bounds on the 129Xe EDM using 3He comagnetometry and SQUID detection have been reported by a joint collaboration between the University of Michigan, the Technical University of Munich and the Physikalisch-Technische Bundesanstalt (PTB) [17] as well as another independent group with almost identical sensitivities [16], which are about five times smaller than the previous limit set in 2001 [15]. One of the challenges in both experiments is the comagnetometer frequency drift, which is several magnitudes larger than the expected frequency shift due to a potential 129Xe EDM [18]. One approach to correct for the impact of the comagnetometer drift on the measured ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$ is using a deterministic physical model to fit the comagnetometer frequency drift [16, 19]. However, the physical origin of the comagnetometer frequency instability is subject of a controversial debate [20, 21], which was inspired by another theoretical model and motivated the performance of recent experiments to substantiate the former criticism [22, 23]. Instead, in reference [17] a phenomenological method was used, which does not need any physical model on the comagnetometer frequency drift, but a distinct pattern of electric fields with switching polarity. We will refer to that as the pattern combination (PC) method from here on.

Here, we propose a new analysis based on a global phase fitting (GPF) method, where the EDM value is estimated by a single fit to the comagnetometer phase development within one complete measurement. Besides an experimentally deduced EDM function as used in reference [16], allowing to analyze any electric field pattern, our GPF method uses a polynomial function to account for the comagnetometer frequency drift. Section 2 gives a short description of the basic principle of measuring the 129Xe EDM ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$ using comagnetometry. In addition, the PC method is introduced for comparison with the GPF method. The GPF method is elucidated in detail in Section 3, including the derivation of the Cramer–Rao lower bound (CRLB). The CRLB of the variance on the EDM value estimation using the GPF method is a factor of four smaller than that of the PC method. In Section 4 we validate the GPF method with Monte-Carlo simulations and compare the results of the PC and GPF method using the experimental data obtained for reference [17]. Eventually we recalculate the systematic uncertainties based on reference [17] and derive a new upper limit for the permanent 129Xe EDM.

2.  3He–129Xe-comagnetometry

2.1. Basic principle

For 129Xe atoms stored in a cell permeated by a uniform magnetic field $ \overrightarrow {B}$ and an electric field $ \overrightarrow {E}$, that is parallel to $ \overrightarrow {B}$, their spin precesses at an angular frequency

Equation (1)

where FXe = 1/2 is the total angular momentum number and γXe is the gyromagnetic ratio of 129Xe. The magnetic field B in equation (1) becomes an interference term for directly calculating ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$ from ωXe. To overcome the experimental difficulties on controlling and measuring $ \overrightarrow {B}$, comagnetometry was introduced with two collated species measured at the same time [15, 18, 24]. 3He is an ideal candidate for comagnetometry due to its potentially high SNR and its negligible EDM compared to ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$ [25]. The weighted frequency difference between 129Xe atoms and 3He atoms is defined as

Equation (2)

and commonly named the comagnetometer frequency. Here ωHe = |γHe B| is the spin precession frequency of 3He atoms with γHe being its gyromagnetic ratio. Therefore, ωco can be written as

Equation (3)

showing that ωco is independent of the magnitude of the background magnetic field but depends on its orientation $\hat{B}$ relative to the applied electric field. The current measurement sensitivity of ωco is in the nHz range for a single measurement, while the comagnetometer frequency drift is at the μHz level, which causes a non-negligible systematic error [22, 23, 26]. Multiple physical models to describe the comagnetometer drift were proposed. The dominant terms thereby vary in different models. Furthermore, several parameters, such as the longitudinal relaxation time T1 of the spins, used in these models are unknown or difficult to measure, making the frequency drift correction with a deterministic model inaccurate. By using a phenomenological model such as proposed here and in reference [17], these difficulties can be omitted 5 .

2.2. Parameters of two measurement campaigns

The data used in our analysis were collected in the joint collaboration at the Berlin magnetically shielded room (BMSR-2) facility at PTB Berlin. Table 1 summarizes the typical values of the main experimental parameters of the two measurement campaigns carried out in 2017 and 2018, respectively. More details on the setup and process are given in reference [27]. The spin precession signal of the transverse magnetization of 3He and 129Xe was recorded by a dc-SQUID system comprising two channels (Z1, Z2) with Z2 sensitive to the same B-field direction as Z1 but 12 cm further away from the 129Xe and 3He sample. The high voltage and leakage current between the two electrodes of the cell were monitored. In the 2017(2018) campaign, a background magnetic field B0 of 2.6 μT (3 μT) was applied to shift ωXe and ωHe to 30 Hz (36 Hz) and 90 Hz (98 Hz), respectively, which are well above the vibrational interference signals (see figure 2). In order to decrease potential crosstalk (bleed-through) from vibrational noise onto the 129Xe and 3He frequencies, a software SQUID axial gradiometer (Z1−Z2) was used. The power of vibrational noise was decreased by a factor of about 70, at a cost of 2% lower signal amplitude as compared to the Z1 magnetometer signal.

Table 1. Starting amplitude A0, transverse relaxation time T2, background magnetic field B0, electric field E0 and segment length ts of both measurement campaigns.

 20172018
A0,Xe/pT30–4070–80
A0,He/pT4–520–25
${T}_{2}^{\mathrm{X}\mathrm{e}}$/s6000–70008000–9000
${T}_{2}^{\mathrm{H}\mathrm{e}}$/s7000–80008000–9000
B0/μT2.63
E0/(kV cm−1)2.754.13
ts/s400 or 800100–800
Figure 2.

Figure 2. (Left) The SQUID gradiometer Z1−Z2 signal (grey curve) and the modulated high voltage signal (blue line) of one run from the 2018 campaign. (Right) The amplitude spectral density of data lasting 100 s from the starting of the first sub-run for the two magnetometer channels (Z1 and Z2) and the software gradiometer (Z1−Z2). The white noise level of the gradiometer is ${\rho }_{\omega }\approx 7.2\enspace \mathrm{f}\mathrm{T}\enspace {\sqrt{\mathrm{H}\mathrm{z}}}^{-1/2}$. The variance of the white noise is ${\sigma }_{\omega }^{2}={f}_{\mathrm{s}}{\rho }_{\omega }^{2}/2={\left(154\enspace \mathrm{f}\mathrm{T}\right)}^{2}$, with the sampling frequency fs = 915.5245 Hz.

Standard image High-resolution image

The left panel of figure 2 shows the raw SQUID gradiometer signal in pT (grey) of one run from the 2018 campaign lasting 35 000 s exemplarily. This run comprises two so called sub-runs with 36 segments each. A segment is defined as the time of constant electric field. For the two sub-runs shown in figure 2, the segments last 300 s and 600 s, respectively. The first sub-run ranging from 50 s to 12 400 s is used as an example in the data analysis section.

2.3. PC method

As mentioned above, one approach to mitigate the effect of the comagnetometer frequency drift is repetitively reversing the direction of the electric field $ \overrightarrow {E}$. This allows to separate the impact of ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$ on ωco from other interference terms. The E modulation method has been applied in diverse EDM experiments with varied modulation patterns [12, 28]. For the PC method, the common E pattern for one sub-run consisted of 36 segments with an equal time interval ts, and the sign of $ \overrightarrow {E}$ changed according to the following sequence ±[0 + - - + - + + - - + + - + - - + 0, 0 - + + - + - - + + - - + - + + - 0].

The PC method determines the EDM value from averaging the comagnetometer frequencies ωco from 2n ($n\in \mathbb{N}$) consecutive segments omitting those with zero voltages. This pattern is constructed to cancel the effect of the comagnetometer frequency drift up to n − 1 order when parametrized in polynomials. The effect of the higher order (above n − 1) drift dependency imposing a false EDM on each sub-run is deduced by applying polynomial fits to all ωco within the sub-runs, leading to a correction for the EDM and an additional systematic uncertainty (for more details see reference [27]).

3. Global phase fitting method

The general data-processing procedure for the GPF method is illustrated in figure 3. For this method, the raw SQUID data of a sub-run is cut into successive blocks of equal length. Each block data is fitted to deduce precession phases of both species 3He and 129Xe (see section 3.1) and the continuous comagnetometer phase is derived for each block (see section 3.2). For data blinding an additional phase, bound to the measured high-voltage signal, is added to the comagnetometer phase at this point (see section 3.3). The EDM value is acquired by fitting the blinded comagnetometer phases using a polynomial function together with a function comprising the phase evolution introduced by a hypothetical 129Xe EDM. The unblinded EDM result is obtained by reanalyzing the raw comagnetometer phases, as illustrated in figure 3.

Figure 3.

Figure 3. Schematic process of the GPF method.

Standard image High-resolution image

3.1. The phase of each block

The block length tb is a free parameter with a suitable range from 1 s to 20 s, to exclude the amplitude decay and frequency drift and to ensure a good performance of the fit for our data [17]. The SQUID data in each block are fitted to the function

Equation (4)

where aXe/He/i , bXe/He/i , ωXe/He, c, and d are the fit parameters and ωi=1,2,3,4 = 2π × 50i s−1 represent the power frequency and its harmonics. The constant and linear terms c and d ⋅ t describe the background magnetic field and its small drift as seen by the SQUID. The variable projection (VP) method is applied [29], where the nonlinear parameters ωXe/He are estimated separately from the linear parameters aXe/He/i , bXe/He/i , c, and d. To minimize the correlation between the fit terms in equation (4), the time of each block is assigned to be symmetrical around zero from −tb/2 to tb/2.

Figure 4 shows the raw SQUID data of a 5 s block from the start of the exemplary sub-run and the residual of the fit to this data. The residual is dominated by the mechanical vibration in the frequency range of 4–25 Hz as shown in the right plot of figure 2. We can assume approximate orthogonality between the precession signal and the vibrational noise of our setup. Therefore, the error on the fit parameter values caused by the latter one is negligible compared to that caused by the white noise, although its integrated power is much larger than the white noise power. This was validated with Monte-Carlo simulations using the recorded vibrational noise (see appendix A.1) as well as a synthetic noise with a sinusoidal function (see reference [30]). The phase of each species for the block k in the range of [−π, π) can be obtained by

Equation (5)

where Arg is the function to get the principle argument of a complex number, i is the imaginary unit and m = Xe or He. Note that due to the time centering, the estimated phase ϕk is referred to the middle time of each block tk = (k − 1/2)tb. The time interval of the block k is defined as (tk−1, tk ). The parameter uncertainties $\delta {a}_{\text{m}}^{k}$ and $\delta {b}_{\text{m}}^{k}$ are estimated from the covariance matrix of the fit

Equation (6)

where $ \overrightarrow {r}$ is the residual, ν is the number of degrees of freedom and J is the Jacobian matrix. The standard deviation of the derived phase $\delta {\phi }_{\mathrm{X}\mathrm{e}/\mathrm{H}\mathrm{e}}^{k}$ is

Equation (7)

Equation (6) assumes that the residual $ \overrightarrow {r}$ stems from the wideband white noise, which is a conservative approach for our case since the main signal in the residuals was the narrowband vibrational noise, leading to an overestimation of the uncertainty $\delta {\phi }_{\text{m}}^{k}$. However, the ratio between $\delta {\phi }_{\text{m}}^{k}$ for different blocks reflects the decaying SNR. Therefore, these estimated uncertainties are used as weights in the subsequent GPF routine.

Figure 4.

Figure 4. The gradiometer signal (grey) for one block with tb = 5 s and the residual of the fit (red).

Standard image High-resolution image

3.2. The accumulated comagnetometer phase

The accumulated phase ${{\Phi}}_{\text{m}}^{k}$ in a block k of the continuously precessing spins is the sum of the wrapped phase ${\phi }_{\text{m}}^{k}$ and a multiple of 2π

Equation (8)

where the integer ${n}_{\text{m}}^{k}$ is determined as

Equation (9)

rounded to the lower integer and ${n}_{\text{m}}^{1}=0$. Here, the frequencies ${\omega }_{\text{m}}^{k-1}$ are obtained by the fit of the block k using equation (4). If ${{\Phi}}_{\text{m}}^{k}-\left({{\Phi}}_{\text{m}}^{k-1}+{\omega }_{\text{m}}^{k-1}{t}_{\mathrm{b}}\right)$ is either > π or < −π, ${n}_{\text{m}}^{k}$ is incremented or decremented by one, respectively, to ensure a continuous phase evaluation. The standard deviation of the accumulated phase $\delta {{\Phi}}_{\text{m}}^{k}$ is equal to $\delta {\phi }_{\text{m}}^{k}$ as equation (8) does not introduce any additional uncertainty. According to equation (2), the evolved comagnetometer phase ${{\Phi}}_{\mathrm{c}\mathrm{o}}^{k}$ for each block k is determined by

Equation (10)

3.3. The fitted EDM value

By integrating equation (3), the accumulated phase due to a hypothetical 129Xe EDM dh at the block k is

Equation (11)

where ${ \overrightarrow {E}}_{i}$ is the average electric field within the block i. By replacing dh with a computer-generated pseudo-random EDM value dbias, the bias phase ${{\Phi}}_{\text{bias}}^{k}$ is calculated and then used to blind the comagnetometer phase ${{\Phi}}_{\mathrm{c}\mathrm{o},\mathrm{b}}^{k}={{\Phi}}_{\mathrm{c}\mathrm{o}}^{k}+{{\Phi}}_{\text{bias}}^{k}$ in order to avoid operator induced bias during process optimization.

The measured phase ${{\Phi}}_{\mathrm{c}\mathrm{o}}^{k}$ originates not only from the potential 129Xe EDM, but also from other sources such as chemical shift [22, 27]. These contributions are phenomenologically parametrized by a polynomial of gth order. Hence, the comagnetometer phase is fitted with the function

Equation (12)

where a, p0, p1, p2, ..., pg are the global fit parameters, and the atomic EDM of 129Xe is ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)=a\cdot {d}_{\mathrm{h}}$. Here the time series tk are normalized to the interval [0, 1] and shifted Legendre polynomials ${\tilde {P}}_{n}\left({t}_{k}\right)$ are applied to decrease the correlation between polynomial coefficients [31]. The fit was performed by using the iterative weighted least squares estimation method 6 with the built-in function nlinfit in MATLAB. Thereby the inverse values of the phase variances ${\left(\delta {{\Phi}}_{\mathrm{c}\mathrm{o}}^{k}\right)}^{2}$ are used as weights. Figure 5 shows the comagnetometer phase ${{\Phi}}_{\mathrm{c}\mathrm{o}}^{k}$, the fit phase ${{\Phi}}_{\text{fit}}^{k}$, and the EDM function ${{\Phi}}_{\text{EDM}}^{k}$ constructed from the measured E-field pattern of the exemplary sub-run.

Figure 5.

Figure 5. The comagnetometer phase ${{\Phi}}_{\mathrm{c}\mathrm{o}}^{k}$ and the data of the fit ${{\Phi}}_{\text{fit}}^{k}$ (top) as well as the EDM function ${{\Phi}}_{\text{EDM}}^{k}$ constructed from the measured electric field (bottom) with dh = 1 × 10−27 e cm for the exemplary sub-run.

Standard image High-resolution image

To determine the order needed for the polynomial function in equation (12), we apply an F-test where the significance of adding q terms to the fitting function with g terms was evaluated by the integral probability

Equation (13)

where PF is the probability density function of the F-distribution and N is the number of data points [33]. The upper bound of the integral is

Equation (14)

The order of the fit was defined sufficient when Pg,g+1 as well as Pg,g+2 are both smaller than a chosen threshold of Pmin.

The correlation matrix for the exemplary sub-run (see figure 2) using 7th order (fulfilling the criteria for Pmin = 0.6) is given in table 2. The correlated uncertainties of the parameters are determined as the square root of the diagonal of the covariance matrix, which inherently includes the uncertainty of the correlations between a and all polynomial parameters. The influences of these correlations to the estimation of a are small due to the orthogonality between the constructed function ${{\Phi}}_{\text{EDM}}^{k}$ and the polynomial function of the order up to n − 2 where 2n is the number of nonzero high-voltage segments. For the exemplary run, the correlations between the EDM parameter a and the polynomial coefficients are significantly smaller than 1, but nonzero, since the polynomials of higher than 3rd order are not orthogonal to ${{\Phi}}_{\text{EDM}}^{k}$ (with 25 = 32 high-voltage segments). The derived uncertainty is in good agreement with the result using the log profile likelihood method (see reference [30]). We also applied the linear regression method with the model in equation (12) and obtained consistent results.

Table 2. Correlation matrix of the first sub-run for the fit with a 7th order polynomial and the block length tb = 5 s.

  a p0 p1 p2 p3 p4 p5 p6 p7
a 1.0 0.0 0.0 0.0 0.0 0.2 0.0 0.2 0.2
p0 0.0 1.00.50.20.10.00.10.10.0
p1 0.0 0.51.00.50.20.10.10.10.0
p2 0.0 0.20.51.00.50.20.10.00.0
p3 0.0 0.10.20.51.00.50.20.10.0
p4 0.2 0.00.10.20.51.00.50.20.1
p5 0.0 0.10.10.10.20.51.00.50.1
p6 0.2 0.10.10.00.10.20.51.00.5
p7 0.2 0.00.00.00.00.10.10.51.0

3.4. The modified Allan deviation

The modified Allan deviation (MAD) is an established tool to evaluate the low-frequency drift of a time series of phases Φ, which is defined as

Equation (15)

where the integration time τ is n times the block length tb, and the total measurement time T is subdivided into P time intervals of equal length τ, such that T [34]. The residual phase ${{\Phi}}_{\mathrm{c}\mathrm{o}}^{k}-{{\Phi}}_{\text{fit}}^{k}$ to the sub-run B881 is shown in figure 6 (left) with the adjacent histogram showing a Gaussian-like distribution. The residual generally increases with time due to the signal decay. The MAD of this sub-run is plotted on the right in figure 6. σf of ${{\Phi}}_{\mathrm{c}\mathrm{o}}^{k}$ reaches the minimum at the integration time τ of 550 s and then increases due to the comagnetometer frequency drift. For the residual phase of this sub-run, the MAD decreases with increasing integration time according to σfτ−3/2 (dashed line in figure 6) over the full range, down to 0.4 nHz. This behaviour that is also observed in all other sub-runs is an indicator that the comagnetometer phase ${{\Phi}}_{\mathrm{c}\mathrm{o}}^{k}$ is adequately described by the fit model of equation (12), since the residual is white phase noise dominated.

Figure 6.

Figure 6. The result of the sub-run B881. (Left) Residual phase for the fit with a 7th order polynomial and the histogram of it. (Right) The modified Allan deviation and its error bar of the accumulated comagnetometer phase and the residual phases. To fulfil the MAD statistics criteria [34], only data are shown for integration time τ < 4000 s.

Standard image High-resolution image

3.5. The theoretical statistical uncertainty bound

The theoretical limit of the 129Xe EDM uncertainty can be derived as the CRLB, which also provides insights into optimizing experimental parameters. For the sake of simplicity, only a single species spin-precession signal is considered and its amplitude is assumed to be a constant over the whole sub-run. For the GPF method, ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$ is estimated with two steps: the VP fitting to obtain the phase of each block and GPF of the sub-runs. Therefore, the overall CRLB is the combination of the results of these two fits.

For the phase ϕ of a sinusoid embedded with white Gaussian noise (WGN) observed over one block with time being symmetrically around 0 s, the CRLB is

Equation (16)

where ${\sigma }_{\mathrm{w}}^{2}$ is the variance of the WGN, A the amplitude and N the number of data points in one block [35]. The CRLB for the parameters in the GPF model equation (12) is the reciprocal of the Fisher information matrix

Equation (17)

where M is the number of segments in one sub-run, and J is the number of blocks in one segment. For the sake of simplicity here, the standard polynomial is used in equation (12). Assuming $\sum _{k=1}^{JM}{{\Phi}}_{\text{EDM}}^{k}{t}_{k}^{i}=0$ for i going from 0 to g and the phase uncertainty δϕ is a constant, the considered CRLB can be simplified to the so called ideal or uncorrelated CRLB as

Equation (18)

By substituting equations (11), (16) and (17) into equation (18), and exploiting the periodic property of the constructed EDM function (see figure 5), for our case, $\sum _{k=1}^{JM}{\left({{\Phi}}_{\text{EDM}}^{k}\right)}^{2}=M\sum _{k=1}^{J}{\left({{\Phi}}_{\text{EDM}}^{k}\right)}^{2}$, the overall CRLB for ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$ becomes

Equation (19)

Equation (20)

where T = MJNΔt is the total measurement time and Δt = 1/fs is the sampling interval. Note that the number of segments M should be large enough to ensure the orthogonality between ${{\Phi}}_{\text{EDM}}^{k}$ and the polynomial functions. In case of an exponentially decaying amplitude A of the precession signal, the CRLB has to be calculated with equation (17). For the PC method, the CRLB on the 129Xe EDM for M segments is derived in reference [27] as

Equation (21)

The PC method applies linear fits to the comagnetometer phases within each segment to derive the comagnetometer frequency, which requires the addition of an interception term as a starting phase, increasing the variance by a factor of four compared to a linear fit without interception term. In the GPF method the accumulated comagnetometer phases within one sub-run are analyzed in a single fit, therefore the uncertainty does not increase as the single interception term is orthogonal to the EDM function (see equation (17)), showing up as the difference of equation (20) to equation (21). Furthermore, the PC method requires the unweighted average of at least four segment frequencies, which increases its statistical uncertainty even further.

4. Results

A Monte-Carlo study was conducted to confirm that the GPF method can reach the higher sensitivity as shown by the CRLB compared to the PC method. Later, the GPF method was used to obtain the 129Xe EDM from the data set as taken in reference [17] using the same magnetometer channel and block length for analysis. As there were data sets in the 2017 and 2018 campaigns which were not usable with the PC method but could be analyzed with the GPF method, we combined all data and optimized the analysis parameters to obtain the minimum uncertainty from the data. Ultimately, an improved upper limit of the 129Xe EDM was derived using the unblinded data.

4.1. Monte-Carlo tests

The accumulated phase of each spin species for the sampling point j was generated as

Equation (22)

Equation (23)

where the drift of the background field B(t) was parametrized with a 4th order polynomial. ${f}_{\text{lin}}^{\mathrm{X}\mathrm{e}/\mathrm{H}\mathrm{e}}$ represent the frequency shifts caused by the chemical shift and Earth's rotation. uXe/He are the drift amplitudes of the respective precession frequencies. The frequency drift was modeled as exponentially decaying functions with the characteristic time of T1 [22, 23, 26]. Thereby it was assumed that T1 is larger than T2 with range as listed in table 3. fEDM is the frequency shift due to the coupling of a synthetic EDM dsyn with the electric field according to equation (3). Substituting equations (22) and (23) into equation (10) results in the synthetic comagnetometer phase, whose time dependence is designed to mimic the measured data (for details see appendix A.2). The exponentially decaying spin precession signals of 129Xe and 3He atoms can be described by

Equation (24)

with tj = jΔt, which is the time for the sampling point j.

Table 3. The range of the parameter values used for generating synthetic spin precession data for Monte-Carlo simulations.

Para.Range
uHe 3.5–4.5 μHz
uXe 9–11 μHz
${f}_{\text{lin}}^{\mathrm{H}\mathrm{e}}$ 4–10 μHz
${T}_{1}^{\mathrm{H}\mathrm{e}}$ 9000–14 000 s
${T}_{1}^{\mathrm{X}\mathrm{e}}$ 9000–14 000 s
${f}_{\text{lin}}^{\mathrm{X}\mathrm{e}}$ 4–10 μHz

The parameters used to generate synthetic data were taken from 18 sub-runs of high sensitivity from the 2018 campaign. The starting amplitude of 129Xe and 3He were set to ${A}_{0}^{\mathrm{X}\mathrm{e}}=70$ pT, ${A}_{0}^{\mathrm{H}\mathrm{e}}=25$ pT and ${T}_{2}^{\mathrm{X}\mathrm{e}/\mathrm{H}\mathrm{e}}=8000$ s. The electric field contained 36 segments of 200 s up to 800 s length, as used in the measurement campaign. The values of other parameters in equations (22) and (23) were random and uniformly distributed in the ranges listed in table 3. Three different kinds of noise were separately added into the synthetic data of equation (24), including two WGN with σ = 154 fT, the standard deviation of the white noise in the real data, and a 5 times lower value of σ = 30.8 fT as well as real SQUID gradiometer noise derived as shown in appendix A.1. The overall EDM values obtained with the GPF method from the 18 synthetic sub-runs for four synthetic values dsyn = (1, 2, 5 and 10) × 10−28 e cm are plotted in figure 7. The averaged overall EDM uncertainty for WGN data with σ = 154 fT is 1.74 × 10−28 e cm, which is roughly a factor of 5 larger than that obtained from the data with σ = 30.8 fT and a factor of 1.1 higher than the calculated CRLB for these 18 sub-runs, which is 1.59 × 10−28 e cm. This mainly results from the correlation between the EDM and the parameters of the polynomials in the phase fit. The uncertainty for the real noise is 1.85 × 10−28 e cm, being similar to that for the white noise with σ = 154 fT, confirming the negligible influence of vibrational noise crosstalk. Most of the 1σ confidence intervals of the derived EDM cover the added EDM values dsyn, showing that the GPF method is capable of obtaining dsyn ⩾ 1 × 10−28 e cm in the presence of the chosen noise.

Figure 7.

Figure 7. The derived EDM values using the synthetic data sets. The horizontal coordinates for the real noise (in red) and WGN with σ = 30.6 fT (in purple) are shifted with + −2 × 10−29 e cm, respectively. The green shade illustrates the 1σ confidence interval with the added EDM as the center value and the uncertainty derived from the CRLB for σ = 154 fT.

Standard image High-resolution image

4.2. Overall results

4.2.1. Statistical uncertainty

Applying the GPF method to the same data set of 41 runs (80 sub-runs) as analyzed by the PC method [17] and using the same channel and analysis parameters, the statistical uncertainty is decreased by a factor of 2.1 from 6.6 × 10−28 e cm to 3.1 × 10−28 e cm.

Due to fewer constraints in the GPF method, runs with the number of segments M ≠ 4n with $n\in \mathbb{N}$ or having SQUID jumps could be included in the data analysis, leading to a total of 45 runs (87 sub-runs). Furthermore, the segments with a zero voltage were included in the analysis. For the analysis, the block length tb = 5 s was chosen and the threshold for the F-test set to Pmin = 0.6 (see appendix B) and the minimum order of the polynomial was set to 4 in order to adequately describe the comagnetometer phase drift. The average polynomial order used for all sub-runs was 6.4 and the maximum order 13.

The overall averaged result using the full data set is ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)=1.1{\pm}3.1{\times}1{0}^{-28}\enspace \text{e}\enspace \mathrm{c}\mathrm{m}$ with χ2/dof = 115.5/86. The detailed results and parameters for each sub-run are given in reference [30]. According to the PDG guidelines [36] we accounted for these random variations by scaling the statistical uncertainty with the factor $\sqrt{{\chi }^{2}/\mathrm{dof}}=1.16$ leading to 3.6 × 10−28 e cm. To cross-check the derived confidence interval of the weighted mean on ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$, we applied the bootstrap method with 5000 data samples [37] to the 87 EDM values and obtained a similar estimate of the statistical uncertainty of 3.14 × 10−28 e cm. Figure 8 shows the derived EDM results per sub-run. Sorting all EDM measurements into groups based on the experimental parameters, such as the cell geometry, B0 field direction, number and duration of segments and the gas pressure, shows no correlation between the deduced EDM value and these parameters, as can be seen in figure 9. Furthermore, no correlation between the chosen polynomial order and the derived sub-run EDM values was seen.

Figure 8.

Figure 8. EDM results of the 2017 and 2018 campaigns derived with the GPF method by sub-runs. The thin orange bar is the confidence interval of 1σ around the weighted mean. The reason for a lower uncertainty in the last 20 sub-runs is a change in the experimental parameters as explained in detail in reference [17].

Standard image High-resolution image
Figure 9.

Figure 9. The EDM results for grouping the data set by number of segments M, segment duration ts, gas pressure p, ${T}_{2}^{\text{Xe}}$, ${T}_{2}^{\text{He}}$, and the statistical uncertainty threshold of 1.4 × 10−27 e cm. The dashed line is at ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)=1.1{\times}1{0}^{-28}\enspace \text{e}\enspace \mathrm{c}\mathrm{m}$ and the grey region indicates the confidence interval of 1σ with the unscaled statistical uncertainty. For clarity of the figure, only a few parameters are plotted here.

Standard image High-resolution image

4.2.2. Systematic uncertainty

The systematic uncertainties of the two experiment campaigns were extensively studied in reference [17]. We applied the same analysis to the full data set used here and the derived systematic uncertainties are summarized in table 4. The correction to the comagnetometer frequency drift of order higher than one (due to the unweighted average of four segments in the PC method), as it has been done in reference [17], becomes obsolete for the GPF method since the model of equation (12) considers the higher order drifts implicitly. Furthermore, when applying the GPF method one has to check the additional impact of the charging current acting as leakage currents since it uses data during the high-voltage ramps. This effect is calculated in appendix C and turned out to be negligible.

Table 4. The systematic uncertainties determined as done in reference [17] based on the data set used for the GPF method.

 2017 (e cm)2018(e cm)
Leakage current (incl. impact of ICharging during ramping)1.20 × 10−28 4.40 × 10−31
Magnetization due to charging current1.70 × 10−29 1.20 × 10−29
Cell motion (rotation)4.20 × 10−29 4.0 × 10−29
Cell motion (translation)2.60 × 10−28 1.90 × 10−28
|E|2 effect1.20 × 10−29 2.20 × 10−30
|E| uncertainty9.90 × 10−29 5.70 × 10−30
Geometric phase⩽2 × 10−31 ⩽1 × 10−29
Total systematic uncertainty3.07 × 10−28 1.95 × 10−28
Scaled statistical uncertainty15.57 × 10−28 3.67 × 10−28

We further looked for the potential systematic effect caused by the comagnetometer drift and the vibrational noise with Monte-Carlo simulations and did not find observable systematic errors, see appendix A.

The overall systematic uncertainty is the weighted average of the systematic uncertainties of the two measurement campaigns 2017 and 2018 using the reciprocal of its statistical variance as weights, yielding 2.0 × 10−28 e cm. The final result, separating the statistical and systematic uncertainties, is

Equation (25)

from which we can set an upper limit $\vert {d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)\vert {< }8.3{\times}1{0}^{-28}\enspace \text{e}\enspace \mathrm{c}\mathrm{m}$ at the 95% C.L. This reanalysis leads to a limit that is a factor of 1.7 smaller compared to the previous result [17] and a factor of 8.0 compared to the result in 2001 [15].

5. Summary and outlook

We proposed a GPF method to analyze spin precession data. Applying the GPF method to the data set used in reference [17] yields a consistent result for ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$ but a two times smaller statistical uncertainty compared to the PC method, as predicted by the theoretical CRLB analysis. Using additional data which had to be discarded for the PC method due to incomplete electric field patterns and optimizing the analysis parameters, the upper limit of the 129Xe EDM improves by a factor of 1.7 to $\vert {d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)\vert {< }8.3{\times}1{0}^{-28}\enspace \text{e}\enspace \mathrm{c}\mathrm{m}$ at the 95% C.L. This enables 129Xe to be used as a comagnetometer in future neutron EDM experiments [38] with a systematic error contribution down to $\vert {d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)\vert {\times}{\gamma }_{\mathrm{n}}/{\gamma }_{\mathrm{X}\mathrm{e}}=2.1{\times}1{0}^{-27}\enspace \text{e}\enspace \mathrm{c}\mathrm{m}$. Our GPF method relieves the demands on the physical model describing the comagnetometer frequency drift and could be generally used in similar spin precession experiments, such as the Lorentz-invariance test. By optimizing the experimental parameters for the GPF method (see appendix D), the upper limit for ${d}_{\text{A}}\left({}^{129}\mathrm{X}\mathrm{e}\right)$ could be reduced even further, as planned for an upcoming EDM campaign.

Authors contributions

All the authors collaborated in the same way in the ideas, development and writing of the manuscript.

Acknowledgments

The authors gratefully thank the joint experimental efforts of the HeXe collaboration for the 129Xe EDM measurement campaigns in 2017 and 2018. We acknowledge the support of the Core Facility 'Metrology of Ultra-Low Magnetic Fields' at Physikalisch-Technische Bundesanstalt which receives funding from the Deutsche Forschungsgemeinschaft (DFG KO 5321/3-1 and TR 408/11-1). This work was supported by Deutsche Forschungsgemeinschaft grants TR408/12 and FA1456/1-1. T Liu acknowledges the support from the Funds for International Cooperation and Exchange of the National Natural Science Foundation of China (Grant No. 51861135308). We gratefully thank Dr. U Steinhoff and Dr. G Wübbeler for fruitful discussion.

Data availability statement

The data that support the findings of this study are available upon reasonable request from the authors.

Appendix A.: Experiment dependent factors

Here we used a Monte-Carlo simulation to investigate the influence of two systematic effects on the derived EDM result, namely the vibrational noise and the comagnetometer phase drift.

A.1. Vibrational noise

The effect of the real measurement noise (e.g. vibrational noise) on the estimated phase is quantitatively analyzed. Here, a synthetic precession signal was generated, using a single sinusoidal function with a constant amplitude A = 30 pT and a length of 10 000 s yielding 2000 blocks. Furthermore, white noise with σ = 154 fT (the standard deviation of the white noise in real gradiometer data) generated with MATLAB, or real noise (from the exemplary sub-run with a total noise power of (0.4 pT)2 and the precession signals filtered out) were added separately to the simulated precession signal. The error of the fitted phase for block i is defined as epsiloni = ϕfit,i ϕreal,i . Here ϕreal,i was known and ϕfit,i was obtained from the fit to block i.

The histograms of epsiloni for these two synthetic data sets are plotted in figure A1. The phase error for the white noise data is in good agreement with the normal distribution with σ = 1.11 × 10−4 rad, which is close to 1.08 × 10−4 rad, the CRLB on the phase estimator in equation (16). The phase error for the real noise data also satisfies the Gaussian distribution with σ = 1.24 × 10−4 rad, which is similar to the result of the white noise data. This implies that the vibrational noise did not cause evident additional phase deviation, although the standard deviation of the vibrational noise is around 7 times bigger than the white noise. As evident in figure A1, the vibrational noise does not cause an observable systematic error on the derived phase. Modelling the vibrational noise with a constant sinusoidal signal provides another way to study the phase error caused by the crosstalk effect. By doing so, we found that the error due to the modelled vibrational signal is more than ten-fold smaller than the uncertainty caused by the white noise. Detailed analysis results can be found in reference [30].

Figure A1.

Figure A1. Histograms of the phase error from the synthetic data with real gradiometer noise (light blue) and white noise (yellow). The data lasts for 10 000 s and consists of 2000 blocks. The mean value for the real noise data is (−2.47 ± 5.46) μrad and the standard deviation is (1.24 ± 0.04) × 10−4 rad.

Standard image High-resolution image

A.2. Comagnetometer phase drift

The analyzed comagnetometer phase drift ${{\Phi}}_{\mathrm{c}\mathrm{o}}^{k}$ for 9 sub-runs with high sensitivities on 129Xe EDM are plotted in the left panel of figure A2. Even though the subtraction of chemical shift and Earth's rotation is unnecessary for the EDM analysis using the GPF method, we subtracted both effects from the measured phase using the deterministic equations [18] to show only the comagnetometer phase drift. The synthetic comagnetometer phase drifts generated with two exponential functions for 18 random sub-runs are plotted in the right panel of figure A2, showing a similar behaviour as the experimentally obtained ones. The parameter ranges are listed in table 3.

Figure A2.

Figure A2. (Left) Measured comagnetometer phase drift of 9 sub-runs reduced by the linear deterministic term stemming from the Earth's rotation and chemical shift. The green curve is the result of the exemplary sub-run used in section 3. The outliers are caused by SQUID jumps in the time domain data leading to a large uncertainty so that the impact of it on EDM estimation is negligible. (Right) Synthetic data of 18 sub-runs.

Standard image High-resolution image

To investigate the potential systematic effect caused by the comagnetometer phase drift, we changed the ranges of the drift amplitudes uXe and uHe as given in table 3 by factors of 0.05 to 20 in the synthetic phase data. White phase noise was added into the phase data and the noise deviation σ increased with time in an exponential way with T2 = 8000 s. Figure A3 shows the derived EDM values as a function of the scale ratio of the drift amplitudes for two starting noise uncertainties σ0 = 100 μrad, similar to the real data, and σ0 = 1 μrad. No significant dependence between the obtained EDM value and the drift amplitudes could be observed. For the low noise case, the maximum deviation of the central EDM value among 7 results is less than 1 × 10−29 e cm. So, we did not assign a model dependent uncertainty for the comagnetometer drift when applying the GPF method, also supported by the analysis shown in figure 7 and the F-test results as shown in appendix B.

Figure A3.

Figure A3. Derived EDM value as a function of the scale ratio of the drift amplitude to the observed drift in two campaigns. The dotted line shows the added EDM value 1 × 10−27 e cm. Each result is an average of 40 sub-runs lasting 12 800 s and with 32 high-voltage segments.

Standard image High-resolution image

Appendix B.: The F-test threshold

The F-test threshold Pmin affects the polynomial order used in the GPF method, as listed in table B1. The EDM values for various Pmin are overlapped within the 1σ statistical uncertainty and are all consistent with zero. Additionally, the upper limit of the 129Xe EDM is almost insensitive to the threshold. We have chosen 0.6 as the F-test threshold as was done also in reference [17]. The highest upper 129Xe EDM bound shows that this choice even leads to the most conservative result.

Table B1. The EDM results and upper limit at 95% C.L. with various F-test threshold Pmin.

  EDMUncertainty  Upper limit
Pmin Average order(10−28 e cm)(10−28 e cm)Reduced χ2 P-value(10−28 e cm)
0.48.20.083.221.320.038.2
0.57.3−0.363.201.240.068.0
0.66.41.063.081.340.028.3
0.76.0−0.073.061.260.057.8
0.85.5−0.873.051.310.038.1

Appendix C.: Systematic effect of charging currents

As the GPF used the full data set, including data during high voltage ramps, the charging current can have an impact on the result in two ways. First the charging current could magnetize parts of the experimental equipment, and change the magnetic field seen by the spins, generating a false EDM. This effect has been carefully analyzed in reference [17] and was adapted for the data set used for GPF (see charging current in table 4). Secondly, the charging currents just as the leakage currents will generate magnetic fields correlated with the electric field direction, which will change the measured spin precession frequencies.

We used synthetic leakage current data over a whole sub-run to study the impact of charging currents. Both the amplitudes of leakage currents and charging currents were constants and their directions were changed according to the electric field pattern. The maximum conversion factor from leakage current to frequency shift was measured to be Ccurr = 1.7(0.01) nHz nA−1 for the 2017(2018) campaign. Figure C1 compares the accumulated phase over a sub-run caused by a leakage current with and without considering the charging currents during ramping. The sharp jumps (red dashed line) are caused by charging currents during the high voltage ramp, whose direction is opposite to the leakage current of the following segment. Applying the GPF method to this accumulated phase, we directly obtained the systematic error caused by leakage currents, which are 1.0 × 10−28 e cm and 1.2 × 10−28 e cm, for Icharge = 10 nA and no charging currents (Icharge = 0 nA), respectively. In the course of deriving the EDM error, one has to transfer the frequency shift to the EDM value using the electric field strength, which is set to 2.75 kV cm−1, the number of the 2017 campaign as listed in table 1. This result shows that charging currents during ramping even reduces the systematic error as compared to the effect of the leakage currents alone. However, this reduction is not significant due to the weak correlation between the accumulated phase caused by pulse-like charging currents and the EDM-induced phase shown in figure 5. Applying the same analysis to the sub-run with the maximum systematic error caused by leakage currents in each campaign, the systematic error reduced from 1.20 × 10−28 e cm to 1.19 × 10−28 e cm for the 2017 campaign (A921 sub-run with Ileakage = 100 pA, Icharge = 11 nA, U0 = 6 kV, ramp rate is 2 kV s−1, ts = 800 s, and M = 18), and from 4.5 × 10−31 e cm to 4.4 × 10−31 e cm for the 2018 campaign (B703 sub-run with Ileakage = 73 pA, Icharge = 1 nA, U0 = 7 kV, ramp rate is 1 kV s−1, ts = 400 s, and M = 36). These results show that the impact of the charging currents acting as the leakage currents is negligible.

Figure C1.

Figure C1. Simulated accumulated phase caused by leakage currents over one sub-run. The used parameters are: M = 36, ts = 400 s, Ileakage = 0.1 nA, Ccurr = 1.7 nHz nA−1, U0 = 6 kV and the voltage ramp rate is 1 kV s−1. The two curves were obtained for different charging currents as given in the legend.

Standard image High-resolution image

Appendix D.: Design of experimental parameters

The number of segments M in one sub-run has a significant impact on the estimation uncertainty derived by the GPF method. According to the ideal CRLB, a smaller number of segments results in a lower uncertainty, shown as the red line in figure D1. To search for the optimum segment number, we used the synthetic comagnetometer phase data with added WGN. The phase uncertainty increases with time and starts with 0.1 mrad. The time constants T2 for 129Xe atoms and 3He atoms are a random number ranged from 8000 s to 9000 s. The total measurement time length is fixed to 38 400 s, while M is varied from 2 to 64. The averaged EDM value over 100 runs for each M are plotted as blue crosses. The fit uncertainty is larger than the ideal CRLB due to the correlation between the EDM function and the phase drift. The gap is reduced with the increase of M, since the orthogonality condition is satisfied better. A relatively flat optimum is found around M = 16. Note that this optimum value also depends on the total measurement time. A sub-run with longer measurement time calls for a higher number of segments, hence the optimum number for T = 6400 s and T = 64 000 s is 8 and 64, respectively. In order to reduce the impact of the comagnetometer frequency drift, we applied a high-voltage pattern comprising up to 36 segments. According to the CRLB, the larger the number of high-voltage segments, the higher the uncertainty of the EDM result. An accurate model of this drift can ease the requirement on the number of the high-voltage segments in one run, thus significantly increasing the measurement sensitivity.

Figure D1.

Figure D1. The averaged EDM value over 100 random simulated runs. The reduced chi-square value was normalized to 1 for each result. The solid lines are guides for the eye.

Standard image High-resolution image

Footnotes

  • The first published result on a permanent EDM of neutron was done by Purcell and Ramsey in 1950 using the data of neutron scattering in lead [7].

  • A physical description providing the correct parameters of the drift may increase the sensitivity of the EDM by adapting experimental parameters in future measurement campaigns, see appendix D.

  • Since the phase error satisfies Gaussian distribution, the weighted least square estimator is identical to the maximum likelihood estimator [32].

Please wait… references are loading.
10.1088/1367-2630/ac09ca