Skip to main content
UKPMC Funders Author Manuscripts logoLink to UKPMC Funders Author Manuscripts
. Author manuscript; available in PMC: 2017 Sep 22.
Published in final edited form as: Nature. 2017 Mar 22;543(7646):525–528. doi: 10.1038/nature21434

Simultaneous tracking of spin angle and amplitude beyond classical limits

Giorgio Colangelo 1,*, Ferran Martin Ciurana 1, Lorena C Bianchet 1, Robert J Sewell 1, Morgan W Mitchell 1,2,
PMCID: PMC5407441  EMSID: EMS71282  PMID: 28332519

Abstract

Measurement of spin precession is central to extreme sensing in physics,1,2 geophysics,3 chemistry,4 nanotechnology5 and neuroscience,6 and underlies powerful magnetic resonance spectroscopies.7 Because there is no spin-angle operator, any measurement of spin precession is necessarily indirect, e.g., inferred from spin projectors Fα at different times. Such projectors do not commute, and thus quantum measurement back-action (QMBA) necessarily enters the spin measurement record, introducing errors and limiting sensitivity. Here we show how to reduce this disturbance below δFαN, the classical limit for N spins, by directing the QMBA almost entirely into an unmeasured spin component. This generates a planar squeezed state8 which, because spins obey non-Heisenberg uncertainty relations,9,10 allows simultaneous precise knowledge of spin angle and amplitude. We use high-dynamic-range optical quantum non-demolition measurements1113 applied to a precessing magnetic spin ensemble, to demonstrate spin tracking with steady-state angular sensitivity 2.9 dB beyond the standard quantum limit, simultaneous with amplitude sensitivity 7.0 dB beyond Poisson statistics.14 This method for the first time surpasses classical limits in non-commuting observables, and enables orders-of-magnitude sensitivity boosts for state-of-the-art sensing1518 and spectroscopy.19,20


Spin-based magnetometers monitor precession of the collective spin F of a magnetically-sensitive atomic ensemble,1,3,21 while atomic clocks2 and other atomic sensors22 use pseudo-spin systems with equivalent quantum descriptions: all are described by the SU(2) Lie algebra. Many optical interferometers are also SU(2) systems.23 These SU(2) systems obey different uncertainty relations than do position/momentum or harmonic oscillator systems, with dramatic consequences for their quantum sensitivity limits. The classical quantity to be measured manifests as precession rate /dt about a known axis, which we take to be x. This signal is not directly observable, because there is no quantum mechanical operator for spin angle ψ. Rather, it must be estimated, e.g. from Fz = |Fρ| cos ψ, where |Fρ| is the spin amplitude in the yz plane, itself an observable to be measured. While some theoretical models assume Fρ to be precisely known a priori, this assumption cannot be applied to most systems of interest – relaxation necessarily adds noise not knowable a priori.24 We thus require a multi-component measurement: of amplitude and angle, or equivalently Fy and Fz, in both cases requiring tracking of non-commuting observables.

Similarly, magnetic resonance techniques20 employ simultaneous amplitude and angle tracking to correlate spin relaxation rate, which indicates the physical environment,19 with precession frequency, which indicates the chemical shift or, in imaging, the spin location.7 In these applications, joint angle-amplitude dynamics contain the important signal.

For simple harmonic oscillator systems, it is well known that QMBA couples angle and amplitude, or equivalently the quadratures X and P, as required to preserve the Heisenberg uncertainty relation δXδP ≥ 1/2 (we take ħ = 1 throughout). This limits angle tracking to the standard quantum limit25 (SQL), with uncertainty δψ = N−1/2, where here N is the mean number of excitations. In contrast, uncertainty principles do not prevent tracking spin systems beyond the SQL. As the spin components Fy and Fz precess about the x axis, they are governed by the Robertson (not Heisenberg) uncertainty relation9

δFyδFz12|[Fy,Fz]|=12|Fx|. (1)

In normal sensor operation, 〈Fx〉 is set to zero, to allow large polarization in the FyFz plane. Because it is a constant of the motion, 〈Fx〉 remains zero for all time, and Eq. (1) places no limit on how precisely Fy and Fz can be simultaneously known or tracked. Arithmetic uncertainty relations10 then set the relevant limit, var(Fy) + var(Fz) ~ N2/3, far below var(Fz) ~ N, the SQL. Because N is typically ~ 106 in cold atom systems and ~ 1012 in atomic vapors, this N1/3 advantage extends the quantum limits by orders of magnitude. Spin states with two sub-classical spin uncertainties have been studied theoretically as planar squeezed states.8

Our discussion thus far indicates only the absence of uncertainty-principle barriers to precision spin tracking. We now outline a proof, given in Methods, that continuous quantum non-demolition (QND) measurement achieves this goal. The state evolution is illustrated in Fig. 1 a) and summarized here: Fz is coupled to an optical “meter” variable Sz via the QND interaction Heff = gFzSz, where g is a coupling constant. The interaction with NL photons imprints a signal proportional to Fz on the meter, which when measured reduces var(Fz) by an amount Δm ~ g2NLvar2(Fz). This same interaction rotates F about Fz by a random angle θgSz , which increases var(Fy) by Δd ~ g2NLvar(Fx), much smaller than Δm, given that var(Fz) ≫ 1. Combining these effects, there is a net reduction of var(Fz) + var(Fy), the total variance in the plane of precession. Precessing and under continuous measurement, Fz and Fy alternate roles as the measured and disturbed variable, and each experiences a net uncertainty reduction. When NL reaches 1/(g2N), the measurement benefit Δm ~ N is of order the initial variance, while the in-plane back-action Δd ~ 1 is still negligible. Probing with this NL also induces a negligible loss of coherence, so that the sensitivity to both angular and radial perturbations improves beyond classical limits. It is important to note that the QMBA is not eliminated in this method, rather it is directed almost entirely to the Fx variable, which is never measured and acts as a depository for quantum uncertainty. A similar approach has been proposed for harmonic oscillators using auxiliary negative-mass oscillators to create uncertainty depositories.26,27

Fig. 1. Simultaneous, precise tracking of spin angle and amplitude.

Fig. 1

a) Bloch-sphere representation of the atomic state evolution. Ellipsoids show uncertainty volumes (not to scale) as the state evolves anti-clockwise from an initial, Fy-polarized state with isotropic uncertainty. An x-oriented magnetic field B drives a coherent spin precession in the FyFz plane. Quasi-continuous measurement of Fz produces a reduction in Fz and Fy variances, with a corresponding increase in var(Fx). b) Observed Faraday rotation angle φFz versus time. Each circle shows the rotation angle from one V-polarized pulse. A magnetic field of 37.6 mG produces the observed oscillation, while dephasing due to residual magnetic gradients and off-resonant scattering of probe photons cause the decay of coherence. Blue circles show a single, representative trace, overlaid on 453 repetitions of the experiment shown as orange dots. The time zero corresponds to the first probe pulse; the end of optical pumping is 58 µs earlier. c) Experimental geometry: 1.9 × 106 cold 87Rb atoms are confined in a weakly-focused single beam optical dipole trap (ODT). Transverse optical pumping is used to produce Fy polarisation. On-axis, 0.6 µs pulses with mean photon number 2.74 × 106 experience Faraday rotation by an angle φFz. A polarimeter consisting of waveplates, a polarising beamsplitter, high-quantum-efficiency photodiodes, and charge-sensitive amplifiers measures the output Stokes component Sy. A reference detector before the atoms measures input Stokes component S0 = |Sx|. The rotation angle is computed as φ = arcsin(Sy/Sx).

Realizing this in-principle advantage requires control of measurement dynamics28 and incoherent effects,29 as well as low-noise non-destructive detection with high dynamic range.30 We use an ensemble of N = 1.9 × 106 cold 87Rb atoms held in an optical dipole trap. The atoms are initially prepared in the Fy-polarized state by optical pumping and, due to an applied B-field in the x direction, precess coherently in the FyFz plane with Larmor period TL ≈ 38 μs. The “meter” variable is the polarisation of ~ 1 μs, off-resonance optical pulses, which experience Faraday rotation by an angle φ = gFz on the Poincare sphere as they propagate through the atomic cloud. We probe the atoms with V-polarized optical pulses, interspersed with Hpolarized compensation pulses to dynamically decouple the spin alignment,12,13 i.e., to produce the effective hamiltonian Heff = gFzSz without tensor light shifts. Earlier experiments have demonstrated sub-projection noise Faraday rotation measurements of either angular13 or amplitude14 variables. To measure both, we use high dynamic-range, shot-noise-limited optoelectronics30 and nonlinear signal reconstruction to achieve sub-projection-noise readout sensitivity for rotation up to φ ≈ 100 mrad. See Methods.

A representative sequence of measured Faraday rotation angles φ(tk) for QND measurements spread over 1 ms is shown in Fig. 1 b), and is well described by a free induction decay model that we use to estimate Fz and Fy at a time te

φ(t)=g[Fz(te)cosωLtrFy(te)sinωLtr]etr/T2+φ0 (2)

where trtte. The coupling constant g is found by an independent calibration, while the Larmor frequency ωL, the coherence time T2, and the offset φ0 are found by fitting to the measured φ(tk) over the the range te − Δttkte + Δt, where Δt = 270 μs (see Methods).

With these parameters fixed, we then use Eq. (2) to obtain a predictive estimate F1=(Fy(1),Fz(1)) at time te using the measurements {φ(tk)}te−Δttk<te from an interval Δt immediately before te; and to obtain a confirming estimate F2=(Fy(2),Fz(2)) using {φ(tk)}te<tktet from the interval Δt after te. Because the classical parameters g, ωL, T2 and φ0, are fixed beforehand, these are two linear, least-squares estimates of the vector F obtained from disjoint data sets. Estimating F for several values of te gives a predictive trajectory and a confirming one. We gather statistics over 453 repetitions of the experiment. Empirically, we find Δt = 270 μs minimizes the total conditional variance Tr(ΓF2|F1) (see Methods), reflecting a trade-off of photon shot noise versus scattering-induced decoherence and magnetic-field technical noise.

Fig. 2 a) shows the resulting mean predictive trajectory 〈F1〉, which spirals slowly toward the origin due to residual magnetic field gradient, and the discrepancy between the trajectories, F2F1. The scatter of this discrepancy rapidly decreases with increasing te , as more probe pulses become available for estimating F1, and reaches a steady state after about 250 μs of probing, at which point the number of pulses used for estimation is limited by Δt. With the optimum Δt = 270 μs, Np = 90 and the total number of photons used to estimate F is NL = Np . nL = 2.47 × 108.

Fig. 2. Experimental results.

Fig. 2

a) Measured trajectories in the FyFz phase space at different estimation times te. For each of the 453 traces shown in Fig. 1 b), the function of Eq. (2) is fit to the data to find predictive and confirming estimates F1, F2, respectively, for (Fy , Fz) at time te. Fits for F1 and F2 use disjoint sets of data covering the ranges te − Δttk < te and te < tkte + Δt, respectively. A single fit is a tightly-wound spiral shown as a thin blue line and the thick arrow shows the trajectory from t = 0 to t = te = 30 µs. For clarity, we show results for te values spaced by 40 µs, slightly more than one Larmor period. Each point shows 〈F1〉 + 100ℱ, where 〈F1〉 is the mean over the 453 repetitions, and F2Γ2,1Γ11F1 is the error of the best linear prediction (see SI). The factor 100 provides magnification for visualization purposes. Orange ellipses, with radial and azimuthal radii of 2σ, where σ=100CL, show the relevant classical limits: Poisson (radial, CL = N) and SQL (azimuthal, CL = 〈Fρ〉/2). b) Fits to estimate (Fy, Fz) for te = 400 µs and a measurement time Δt = 270 µs. Blue (red) shows fits based on prior (posterior) data. Shaded regions show fit residuals ×10. c) Evolution of tracking precision for different te. Blue circles and green squares show radial and azimuthal components of ΓF2|F1. Error bars show the ±1 standard error in the variance for 453 repetitions. Dashed blue and solid green curves show Poisson and SQL variances. These decrease during probing due to loss of coherence and loss of atoms. No readout noise has been subtracted.

To quantify the measurement uncertainty, we compute the vector conditional covariance ΓF2|F1=ΓF2ΓF2F1ΓF11ΓF1F2 where Γv matrix for vector v, and Γuv indicates the cross-covariance matrix for u and v. Defining the polar coordinate system (Fy, Fz) = ρ(−sin ψ, cos ψ), we identify the radial and azimuthal variances, var(Fρ) ≡ ρ̂TΓF2|F1ρ̂ and var(Fψ) ≡ ψ̂TΓF2|F1ψ̂, respectively, where ρ̂ ≡ (−sin ψ, cos ψ)T and ψ̂ ≡ (− cos ψ, − sin ψ)T are radial and azimuthal unit vectors.

As shown in Fig. 2 c), var(Fψ) drops below the SQL of 〈Fρ〉/2 after ≈ 150 μs of probing, and remains below it to the limit of the experiment. No read out noise has been subtracted. Considering the steady-state region te ≥ 270 μs, var(Fψ) is on average 2.9 dB below the SQL, and var(Fρ) is on average 7.0 dB below the Poissonian variance N, to give a precision surpassing classical limits in both dynamical variables. For any given value of te, var(Fρ) and var(Fψ) have standard errors of ≈ 0.3 dB, implying high statistical significance even without combining results for different te.

We have shown how quantum measurement back-action can be almost completely evaded in spin-based sensors and spectroscopies, allowing simultaneous tracking of spin angle and amplitude beyond classical limits, using the physics of planar squeezed states.8 Our method is very close to practical application in the highest-performance atomic sensors: Tracking of atomic spin precession by non-destructive optical measurement is already used in the highest-sensitivity magnetic field measurements1 and is also being developed for optical lattice clocks.15 The method is compatible with multi-pass17 and cavity build-up methods,18 that greatly reduce incoherent scattering, the limiting factor in our experiment. Together, these advances promise orders-of-magnitude sensitivity improvement in extreme sensing, in applications ranging from studies of macromolecular dynamics4 and geophysics,3 to non-invasive measurements of single-neurons6 and brain dynamics.16

Methods

A. Faraday rotation probing of atomic spins

The effective atom-light interaction is given by the hamiltonian

Heff=gSzFz (3)

which describes a quantum non-demolition measurement of the collective atomic spin Fz, where the operators Fαifα(i) (with α = x, y, z) describe the collective atomic spin, with fα(i) the spin orientation of individual atom spins. The optical polarization of the probe pulses is described by the Stokes operators Sk=12(aL,aR)σk(aL,aR)T, with Pauli matrices σk. The coupling constant g depends on the detuning from the resonance, the atomic structure and the geometry of the atomic ensemble and probe beam and is independently measured.3134,50

An input Sx-polarized optical pulse interacting with the atoms experiences a rotation by an angle φ = gFz because of the interaction given by eq. (3). The transformation produced by the measurement on Sy is

Sy=Sycosφ+Sxsinφ (4)

In our experiment we measure Sx at the input by picking off a fraction of the optical pulse and sending it to a reference detector, and Sy using a fast home-built balanced polarimeter.35 Both signals are recorded on a digital oscilloscope.

From the record of Sx and Sy , we calculate φ̂, the estimator for φ:

φ^=arcsin(SySx)=arcsin(sinφ+SySx1sin2φ)=φ+SySx+12(SySx)2tanφ+O(SySx)3. (5)

We note that due to shot noise Sy/Sx is normally distributed with zero mean and variance 1/(2Sx) ~ 5×10−7. The term containing tan φ thus describes a distortion of the signal at the ~ 10−6 level, which is negligible in the experiment.

B. Quantum limits for spin variances

Different classical limits provide benchmarks for the radial and azimuthal components of a spin precessing in the FyFz plane. In general, these benchmarks describe the minimal noise of quantum states describing uncorrelated particles. For our system of N spin-1 atoms, the lowest noise uncorrelated state is the coherent spin state defined as a pure product state in which each atom is fully polarized in the same direction. If this direction is ŷ cos θ sin θ, then the azimuthal component Fθ = − Fy sin θ + Fz cos θ has variance

var(Fψ)SQL=Fρ2. (6)

Any state that surpasses this limit implies entanglement among the atoms, and/or entanglement of the internal components of the individual atoms.36,37

For the radial component Fρ = Fy cos θ + Fz sin θ, the classical limit comes from the fact that accumulation of independent atoms into the ensemble is limited by Poisson statistics, var(N) = 〈N〉, so that for F = 1,

var(Fρ)Poisson=N. (7)

Noise below this level can be produced by a strong interaction among the atoms during accumulation,3841 or as here by precise non-destructive measurement.34,4247

C. Operator-level description of back-action evading measurement of two non-commuting spin observables

We consider a spin variable F, defined by commutation relations [Fx, Fy] = iFz and cyclic permutations, precessing about the Fx axis and subjected to brief, non-destructive measurements of the Fz variable. We assume the precession during the measurement is negligible. In the measurement, the spin is coupled to the polarization of a probe pulse, described by the Stokes operators S with [Sx, Sy] = iSz and cyclic permutations. The probe initial state is a coherent state polarized along Sx, so that |Sx|=NL/2,Sy=Sz=0,andvar(Sy)=var(Sz)=12|Sx|.. The system and meter are coupled by the quantum nondemolition hamiltonian

Heff=gSzFz (8)

which acts for unit time. The transformation produced is

Sx=SxcosgFzSysingFz (9)
Sy=SycosgFz+SxsingFz (10)
Sz=Sz (11)
Fx=FxcosgSzFysingSz (12)
Fy=FycosgSz+FxsingSz (13)
Fz=Fz (14)

Where primes indicate the output variables.

We assume a spin state in the Fy-Fz plane, i.e. with 〈Fx〉 = 0, and with zero initial cross-correlation, i.e. cov(Fx , Fy) = cov(Fx , Fz) = 0. Due to the zero mean of Sz , which is also independent of F, the transformation preserves these statistics in the primed variables, for example

cov(Fx,Fy)=cov(Fx,Fy)cos2gSzsin2gSz+[var(Fx)+var(Fy)]cosgSzsingSz=0 (15)

We can compute the statistics of the output variables using

cosgSz=112g2var(Sz)+O(g)4=114g2|Sx|+O(g)4 (16)

and similar expansions for 〈cos2 gSz〉 and 〈sin2 gSz〉. The mean of Fy changes due to the back-action as

Fy=FycosgSz+FxsingSz=FycosgSz=Fy14g2|Sx|Fy+O(g)4 (17)

while the means of Fx and Fz are unchanged.

The variance of Fx is coupled to the variance of Fy , due to the rotation about Fz by a random angle gSz :

var(Fx)=(FxcosgSzFysingSz)2Fy2singSz2=var(Fx)cos2gSz+Fy2sin2gSz=var(Fx)+g2|Sx|[14var(Fx)+12Fy2]+O(g)4 (18)

and similarly

var(Fy)=(FycosgSz+FxsingSz)2Fy2cosgSz2=Fy2cos2gSz+Fx2sin2gSzFy2cosgSz2=Fy2cos2gSzFy2cosgSz2+Fx2sin2gSz=var(Fy)+g2|Sx|[14var(Fy)+12var(Fx)]+O(g)4 (19)

after noting that, to order g3, 〈cos2 gSz〉 = 〈cos gSz2.

After the coupling, a projective measurement of Sy provides information about Fz , with readout variance

varRO(Fz)12g2|Sx|. (20)

The approximation comes from a linearization of Eq. (10), which as discussed in Sec. A introduces an error at the 10−6 level, negligible in this scenario.

The resulting Fz variance, including both the prior and posterior information, is then48,49

var(Fz)=1var1(Fz)+varRO1(Fz)=var(Fz)1+2g2|Sx|var(Fz) (21)

expanding in g this becomes

var(Fz)=var(Fz)2g2|Sx|var2(Fz)+O(g)4. (22)

Collecting Eqs. (18), (19) and (22), defining ΔFαFαFαandΔvar(Fα)var(Fα)var(Fα), and dropping terms of order O(g)4 we find

ΔFy=12g2|Sx|Fy (23)
Δvar(Fx)=g2|Sx|[14var(Fx)+12Fy2] (24)
Δvar(Fy)=g2|Sx|[14var(Fy)+12var(Fx)] (25)
Δvar(Fz)=2g2|Sx|var2(Fz) (26)

Considering an initial coherent spin state and choosing |〈Sx〉| = g−2N−1, where N is the number of spins, we note that Δvar(Fz) ~ N, implying a reduction in the uncertainty of Fz comparable to its initial uncertainty. Due to the Fy2 term, the increase in var(Fx) is ~ N, comparable to its initial value. The other changes are ~ 1, negligible relative to the initial values. In this way we see that uncertainty is moved from Fz to Fx with negligible effect on Fy.

Larmor precession then noiselessly rotates uncertainty from Fy into Fz , uncertainty that is moved into Fx by the next measurement. This procedure reduces the uncertainty of both Fy and Fz with negligible in uence from measurement back-action.

D. Implementation in an atomic ensemble

1. Experimental set up

The experimental set up is described in detail in references .31,50 The trap consists of a single beam laser at 1064 nm with 6.3 W of optical power, focused to a beam waist of 26 μm using an 80 mm lens. The trap is loaded with laser-cooled atoms from a magneto optical trap (MOT). After sub-doppler cooling in the final stage of the loading sequence, the trapped atoms have a temperature ~ 12 μK. The resulting atomic ensemble has an approximately Lorentzian distribution along the trap axis (which we label the z-axis) with a FWHM of w = 4 mm, and a gaussian distribution in the radial direction with of ω = 33 ± 3μm.

2. State preparation

The initial atomic state is prepared via optical pumping with circularly polarized light resonant with the F = 1 → F′ = 1 transition propagating along the y-axis. During the optical pumping stage the atoms are also illuminated with repumping light resonant with the F = 2 → F = 2′ transition using the six MOT beams, preventing accumulation of atoms in the F = 2 hyperfine level, and a small magnetic field is applied along the x-axis, with Bx = 37.6 mG, to coherently rotate the atomic spins in the yz plane. We use a stroboscopic pumping strategy, chopping the optical pumping light into a series of τpump = 1.5 μs duration pulses applied synchronously with the precessing atoms for total of 200 μs, to prepare the atoms in an Fy-polarized state with high efficiency (~ 98%), resulting in a input polarized atomic ensemble with 〈Fy〉 ≃ N (see Extended Data Fig. 1). The pulse duration τpumpTL is chosen to optimize the optical pumping efficiency.

3. Probing

We probe the atoms via off-resonant paramagnetic Faraday-rotation using τ = 0.6 μs duration pulses of linearly polarized light with a detuning of 700 MHz to the red of the 87Rb D2 line. The probe pulses are V-polarized, with on average NL = 2.74 × 106 photons, and sent through the atomic cloud at 3 μs intervals. Between the probe pulses, we send H-polarized compensation pulses with on average NL(H)=1.49×106 photons through the atomic cloud. As described in detail in references,48,51,52 the compensation pulses serve to cancel effects due to the tensor light shift, but do not otherwise contribute to the measurement. During the probing sequence, a magnetic field along the x direction drives a coherent rotation of the atoms in the yz plane with TL = 38 μs period. This ensures that the time taken to complete a single-pulse measurement is small compared to the Larmor precession period, i.e. τTL.

We correct for slow drifts in the polarimeter signal by subtracting a baseline φ0=1Nk=1Nφk(i) from each pulse, estimated by repeating the measurement without atoms in the trap.

4. Statistics of probing inhomogeneously-coupled atoms

We consider the statistics of Faraday rotation measurements on an ensemble of N atoms, described by individual spin operators fi. To define the SQL, we consider an ensemble in a coherent spin state, with the individual spins are independent and fully polarized in the FyFz plane. We take N to be Poisson-distributed. When the spatial structure of the probe beam is taken into account, the Faraday rotation is described by the input-output relation for the Stokes component Sy

Sy(out)=Sy(in)+Sx(in)i=1Ng(xi)fz(i) (27)

where g(xi) is the coupling strength for the ith atom, proportional to the intensity at the location xi of the atom. Sy(in) is has zero mean and variance |Sx(in)|/2. We consider first the case in which the spin is orthogonal to the measured Fz direction, i.e. a measurement of the azimuthal component. Here the uncertainty in g(xi) and in N make a negligible contribution, and the rotation angle φ=Sy(out)/Sx(in) has the statistics

φ=fzi=1Ng(xi)xiNfzμ1 (28)
var(φ)=var(φ0)+var(fz)i=1Ng2(xi)N,xivar(φ0)+Nvar(fz)μ2 (29)

where φ0 is the polarization angle of the input light, subject to shot-noise fluctuations and assumed independent of Fz , and the angle brackets indicate an average over the number and positions of the atoms.

Next we consider the case in which the spin is along the measured Fz direction, i.e., a measurement of the radial component. In this case, the uncertainty in fz is zero, and the variation in g and in N determines the measured variation

φ=Nfzμ1 (30)
var(φ)=var(φ0)+fz2var(i=1Ng(xi))var(φ0)+Nfz2v2 (31)

We note that v2 includes the variation of both the atom number and the coupling strength, and as such is lower-bounded by the Poisson statistics of N: v2 ≥ 〈g2(x)〉 = μ2.

For known 〈fz〉 and var(fz), measurements of 〈φ〉 and var(φ) versus N give the calibration factors μ1 and μ2 as described in Sections D5 and D6, respectively. To preserve the SQL var(Fz)=12|Fy| and similar, in the analysis leading to Fig. 2 we infer mean values as

F(a)=1μ1φ(a) (32)

and covariances, including cov(A, A) = var(A), as

cov(F(a),F(b))=1μ2cov(φ(a),φ(b)), (33)

where F(a,b) and φ(a,b) are corresponding spin and angle variables. We note that because the contribution of var(φ0) is not subtracted, this overestimates the spin variances.

5. Measurement of calibration factor μ1

We calibrate the measured rotation angle φ with a dispersive atom number measurements using absorption imaging, as shown in Extended Data Fig. 2. For the absorption imaging, atoms are transferred into the f = 2 hyperfine ground state by a 100 μs pulse of laser light tuned to the 5S1/2(f = 1) → 5P3/2(f′ = 2) transition. The dipole trap is switched off to avoid spatially dependent light shifts. An image is taken with a 100 µs pulse of circularly polarized light resonant to the 5S1/2(f = 2) → 5P3/2(f′ = 3) transition. We calculate the resonant interaction cross-section and take into account the finite observable optical depth. The statistical error in the absorption imaging is < 3%, including imaging noise and shot-to-shot trap loading variation.

6. Measurement of calibration factor µ2

To measure μ2 we prepare a Fy-polarized state by optical pumping, and then probe stroboscopically with Np = 36 pulses of NL = 3.15 × 107 photons each in the presence of a B-field of ≈ 71.5 mG along y , producing a Larmor precession of an angle π during the 10 μs pulse repetition period. In this way, the measured variable is always ±Fz , evading back-action effects.

If φn is the measured Faraday rotation angle for pulse n, and φ0(n) is the corresponding input angle, we can define the pulse-train-averaged rotation signal as

φ1Npn=1Np(1)n1φn (34)

with variance

var(φ)var(φ0)+μ2n=1Npvar(Fz,n) (35)

where φ0=1Npn=1Npφo(n), with zero mean and variance var(φ0) = (NpNL)−1, and Fz,n is the value of Fz at the time of the nth probe pulse.

During the measurement, off-resonant scattering of probe photons produces both a reduction in the number of probed atoms and introduces noise into F. We note that this is a single-atom process that preserves the independence of the atomic spins. We compute the resulting evolution of the state using the covariance matrix methods reported in,48 and specifically described for this case in Section D7, giving

var(gNpn=1NpFz,n)=μ212αN (36)

where 1/2 = var(fz) is the variance of the initial state, α = 0.86 describes the net noise reduction due to scattering.

Including the readout noise var(φ0) and a generic technical noise a2N2 in the preparation of the coherent spin state, we have the observable variance

var(φ)=var(φ0)+μ212αN+a2N2, (37)

in which the N scaling distinguishes the atomic quantum noise from other contributions. Experimental result shown in Extended Data Fig. 3 give μ2 = (1.5 ± 0.2) × 10−14.

7. Calculation of the noise contribution α

As reported in Colangelo et al.48 the full system is described by a state vector V={Fz,Sy(1),Sy(2),,Sy(N)} and covariance matrix Γ = 〈VV + (VV)T〉/2 − 〈V〉 ∧ 〈V〉, where Syn is the measured photon imbalance after the n-th pulse. The QND interaction leads to a transformation of the covariance matrix

Γ(n+1)=M(n)Γ(n)[M(n)]T (38)

where M is equal to the identity matrix apart from the elements M1,1 = −1 due to the precession by an angle π about the magnetic field, and Mn+1,1 = gSx, where Sx = NL/2 and NL is the number of photons per pulse and g is the coupling constant for uniform coupling.

Off-resonant scattering of photons introduces decoherence, noise and loss in the atomic state. During the spin-noise measurement, a fraction ξ = 1 − exp(−ηNL) = 0.01 of atoms scatter a photon during a single probe pulse, where η = 3 × 10−10 is the scattering rate per photon measured in an independent experiment, while a fraction χ = 1 − ξ remain in the coherent spin state. The scattered atoms are either lost from the F = 1 manifold, or return to F = 1 with probability p = 0.7 and random polarization. This has the effect of losing atomic polarization at each measurement. We calculate the effective measured polarization in terms of the initial atom number. We assume that the fraction p of scattered atoms the return to F = 1 have a random polarization and that the scattering rate η is independent of the atomic state.

After each pulse, the atomic part of the covariance matrix transforms according to

Γat(n+1)=χΓat(n)+23p(1χ)N(n)𝕀 (39)

where 𝕀 is the identity matrix. This follows from Eq.(A.6) of48 assuming Γ = NΓλ. We note that we have

N(n+1)=1(1χ)(1p)N(n)=(χ+pχp)N(n) (40)

which, assuming that N(0) = N, gives

N(n)=(χ+pχp)nN (41)

Including these terms, we get a linear transformation of the covariance matrix after the n-th pulse

Γ(n+1)=DM(n)Γ(n)[M(n)]TDT+N(n) (42)

where D is a zero matrix apart from the element D1,1=χ, and N(n) is the identity matrix apart from the element N(n)1,1=23p(1χ)(χ+pχp)nN.

We sum N individual polarimeter signals Sy(n) to find the net Stokes operator Syn=1N(1)n1Sy(n). This has a variance

var(Sy)=n=1Nvar(Sy(n))+2nmNcov(Sy(n),Sy(m))(1)nm=PΓ(N)P (43)

with the projector P = diag(0, 1, −1, 1, −1, … , −1). When evaluated analytically using χ = 0.99, this gives

var(Sy)=var(Sy,0)+βg212NNL2 (44)

where β ≈ 0.1081. Noting that var(φ)=var(Sy)/Sx2, where Sxn=1NSx(n) is the total input Stokes operator, dividing Eq. (44) by Sx2=NL2/4, and comparing against Eq. (37), we find that α = 8β ≈ 0.86.

E. Data analysis

1. Fitting procedure

As described in the main text, we follow a two-step fit procedure in our data analysis: we first fit Eq. 2 to the joint data set {φ(tk)} of the first and second measurements, to estimate the classical parameters g, ωL, T2 and φ0 near the measurement time te; then second, with the classical parameters fixed, we obtain a predictive estimate F1 using measurements {φ(tk)}te−Δttk<te from the interval Δt immediately before te; and a confirming estimate F2 using {φ(tk)}te<tktet from the interval Δt after te.

2. Measurement phase

We note that, as shown in Fig. 2, the observed squeezing is independent of the choice of te provided that sufficient measurements are available in the interval Δt prior to te, indicating that the observed squeezing is not dependent on the choice of a particular phase of the spin oscillation.

3. Fit gain

Since the classical parameters g, ωL, T2 and φ0 are fixed beforehand, the predictive and confirming fits to estimate Fy and Fz are linear, least-squares fits to disjoint data sets. We note that our condition for (conditional) spin squeezing is whether the conditional variance Tr(ΓF2|F1) is below classical limits - i.e., whether the estimate F1 can be used to precisely predict the estimate F2. This definition of squeezing is quite robust as regards the choice of estimators for Fy and Fz: they only need to have the right gain, i.e., the slope of the curve relating the mean estimate to the true value. The error propagation formula can then be used to find the variance of the true values in terms of the variance of the estimators.

We check that the least-squares fits give the correct gain by comparing the estimated F1,2 with the results of two independent fits using all free parameters in Eq. (2). Results, shown in Extended Data Fig. 4, show the gains are equal to within a ~ 10−3 fractional error, implying a similarly small ~ 10−3 fractional error in the inferred variances and conditional variances, with negligible effect on the squeezing results.

4. Weights

For the first fit to estimate the classical parameters, our data are weighted using an empirical function based on two observations: 1) the polarimeter signal shows increased technical noise in the optical variable at larger imbalance, i.e. when measuring a large instantaneous spin-projection along the z-axis; and 2) points closer in time to te should be given greater weight (minimizing errors introduced by small changes in ωL and T2 during the measurement). This motivates using the weight function

W(φ(tk))g(|tkte|)h(φ(tk)) (45)

where g(|tkte|)1+Aexp(w|tkteT2|) and h(φk) = 1 + r |φk|. This ensures that we accurately estimated the classical parameters g, ωL, T2 and φ0 at the measurement time te.

We numerically optimize W(φ(tk)) varying the parameters A, w and r and minimizing the resulting Tr(ΓF2|F1) from the predictive and confirming fits. We find an optimum with the parameters A = 15, w = 11 and r = 6, and note that the fit procedure gives similar results with variations of up to 30% in each of these parameters.

For the predictive and confirming fits, which are linear in Fy and Fz, all the points are weighted equally.

5. Optimal measurement length

The optimal measurement length Δt results from a trade off between the photon shot noise, the decoherences induced by the probing and the technical noise induced by the magnetic field. Longer measurements reduces the photon shot noise, while increasing the atomic decoherences and making the model eq. (2) less accurate. We empirically find the optimal Δt by minimizing the total variance Tr(ΓF2|F1) for measurements with different length, as shown in Extended Data Fig. 5.

6. Conditional Covariance

Estimating F for several values of te gives a predictive trajectory and a confirming one. Estimations are repeated on 453 repetitions of the experiment to gather statistics. Assuming gaussian statistics, to quantify the measurement uncertainty, we compute the conditional covariance matrix

ΓF2|F1=ΓF2ΓF2F1ΓF11ΓF1F2 (46)

which quantifies the error in the best linear prediction of F2 based on F1.53 Here Γv indicates the covariance matrix for vector v, and Γuv indicates the cross-covariance matrix for vectors u and v. The difference between the best linear prediction of F using F1 and the confirming estimate F2 is visualized using the vector 𝓕={y,z}=F˜2ΓF2F1ΓF11F˜1, where F˜i=FiFi. Standard errors in the estimated conditional covariance matrix are calculated from the statistics of {ℱ}.54

Data Availability

The datasets generated and analysed during the current study are available from the corresponding author on reasonable request. The data shown in Fig. 2 and all the data used to generate plots of Extended Data are included as Source Data.

Extended Data

Extended Data Fig. 1.

Extended Data Fig. 1

Optical pumping efficiency. We prepare an input atomic state with 〈Fy〉 ≃ N via stroboscopic optical pumping in the presence of a small magnetic field along the x-axis. Data is fit with an exponential growing curve ∼ a(1 − e−t/τ) (solid line) and we obtain a = 0.979 ± 0.004 and τ = 0.26 ± 0.02. Orange dashed line: Optical pumping efficiency of 98%. ±1 s.e.m. error bars are smaller than the points for most of the data.

Extended Data Fig. 2.

Extended Data Fig. 2

Calibration of average Faraday rotation signal. We calibrate the rotation angle φ against input atom number N, measured via absorption imaging. Solid line, the fit curve φ = a0 + µ1N, with we obtain µ1 = (7.07 ± 0.04) × 10−8 and a0 = (3.9 ± 0.3) × 10−3. Error bars indicate ±1 s.e.m.

Extended Data Fig. 3.

Extended Data Fig. 3

Calibration of quantum noise limited Faraday rotation probing of atomic spins. We plot the measured variance var(φ) as a function of the number of atoms N in an input coherent spins state with 〈F〉 = {0, N, 0}. Solid curve: a fit using the polynomial var(φ) = a0 + a1N + a2N2. The linear term a1 = αµ2N/2 corresponds to the atomic quantum noise from atoms in the input coherent spin state. We estimate a0 = (11.7 ± 0.7) × 10−10, a1 = (6.5 ± 0.8) × 10−15, and a2 = (2.8 ± 12) × 10−22, consistent with negligible technical noise in the atomic state preparation. Dashed line: var(φ) = a0 + a1N. Error bars indicate ±1 standard error in the variance for 206 repetitions.

Extended Data Fig. 4.

Extended Data Fig. 4

Fit gain. We compare the estimated Fz and Fy from a fit using Eq. (2); first, with the classical parameters g, ωL, T2 and φ0, fixed (labeled Fy,z(fix)) for measurements 1 and 2; and second, free to vary as independent parameters (labeled Fy,z(ind)). In blue (green) Fz (Fy) of the first measurement, in red (orange) Fz (Fy) of the second measurement. A linear fit γx + δ to points of plots a-d gives γa = 0.9981(8), γb = 1.0026(8), γc = 0.9923(4), γd = 1.0007(5) and δa = 0.003(1), δb = 0.0001(9), δc = 0.0004(3), δd = −0.0023(3), where the subscripts refer to the values shown in plots a-d. A grey line y = x is plotted on both the figures.

Extended Data Fig. 5.

Extended Data Fig. 5

Tracking precision as function of Δt. An optimum is found at Δt = 270 µs. Error bars indicate ±1 standard error in the variance for 453 repetitions.

Acknowledgements

We thank L. Bianchet, G. Vitagliano, M. D. Reid, P. D. Drummond, G. Tóth, N. Behbood, M. Napolitano, S. Palacios, D. T. Campbell and M. M. Fria. Work supported by MINECO/FEDER, MINECO projects MAQRO (Ref. FIS2015-68039-P), XPLICA (FIS2014-62181-EXP) and Severo Ochoa grant SEV-2015-0522, Catalan 2014-SGR-1295, by the European Union Project QUIC (grant agreement 641122), European Research Council project AQUMET (grant agreement 280169) and ERIDIAN (grant agreement 713682), and by Fundació Privada CELLEX.

Footnotes

Author Contributions

G.C. and F.M. performed the experiment, G.C analysed the results with the help of R.J.S., M.W.M. conceived the experiment and developed the theoretical model. All the authors designed the experiment, discussed the results and contributed to the manuscript.

Competing Interests

The authors declare no competing financial interests.

References

  • 1.Kominis I, Kornack T, Allred J, Romalis M. A subfemtotesla multichannel atomic magnetometer. Nature. 2003;422:596. doi: 10.1038/nature01484. [DOI] [PubMed] [Google Scholar]
  • 2.Ludlow AD, Boyd MM, Ye J, Peik E, Schmidt PO. Optical atomic clocks. Rev Mod Phys. 2015;87:637. [Google Scholar]
  • 3.Dang HB, Maloof AC, Romalis MV. Ultrahigh sensitivity magnetic field and magnetization measurements with an atomic magnetometer. Appl Phys Lett. 2010;97 151110. [Google Scholar]
  • 4.Gomes MD, Dao P, Jeong K, Slack CC, Vassiliou CC, Finbloom JA, Francis MB, Wemmer DE, Pines A. 129Xe NMR relaxation-based macromolecular sensing. J Am Chem Soc. 2016;138:9747. doi: 10.1021/jacs.6b02758. [DOI] [PubMed] [Google Scholar]
  • 5.Wolf T, Neumann P, Nakamura K, Sumiya H, Ohshima T, Isoya J, Wrachtrup J. Subpicotesla diamond magnetometry. Phys Rev X. 2015;5 041001. [Google Scholar]
  • 6.Jensen K, Budvytyte R, Thomas RA, Wang T, Fuchs AM, Balabas MV, Vasilakis G, Mosgaard LD, Stærkind HC, Müller JH, Heimburg T, et al. Non-invasive detection of animal nerve impulses with an atomic magnetometer operating near quantum limited sensitivity. Sci Rep. 2016;6 doi: 10.1038/srep29638. 29638. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Plewes DB, Kucharczyk W. Physics of MRI: A primer. J Magn Reson Imaging. 2012;35:1038. doi: 10.1002/jmri.23642. [DOI] [PubMed] [Google Scholar]
  • 8.He QY, Peng S-G, Drummond PD, Reid MD. Planar quantum squeezing and atom interferometry. Phys Rev A. 2011;84 022107. [Google Scholar]
  • 9.Robertson HP. The uncertainty principle. Phys Rev. 1929;34:163. [Google Scholar]
  • 10.Dammeier L, Schwonnek R, Werner RF. Uncertainty relations for angular momentum. New J Phys. 2015;17 doi: 10.1103/PhysRevLett.119.170404. 093046. [DOI] [PubMed] [Google Scholar]
  • 11.Grangier P, Levenson JA, Poizat J-P. Quantum non-demolition measurements in optics. Nature. 1998;396:537. [Google Scholar]
  • 12.Koschorreck M, Napolitano M, Dubost B, Mitchell MW. Quantum nondemolition measurement of large-spin ensembles by dynamical decoupling. Phys Rev Lett. 2010;105 doi: 10.1103/PhysRevLett.105.093602. 093602. [DOI] [PubMed] [Google Scholar]
  • 13.Sewell RJ, Napolitano M, Behbood N, Colangelo G, Mitchell MW. Certified quantum non-demolition measurement of a macroscopic material system. Nat Photon. 2013;7:517. [Google Scholar]
  • 14.Béguin J-B, Bookjans EM, Christensen SL, Sørensen HL, Müller JH, Polzik ES, Appel J. Generation and detection of a sub-poissonian atom number distribution in a one-dimensional optical lattice. Phys Rev Lett. 2014;113 doi: 10.1103/PhysRevLett.113.263603. 263603. [DOI] [PubMed] [Google Scholar]
  • 15.Lodewyck J, Westergaard PG, Lemonde P. Nondestructive measurement of the transition probability in a Sr optical lattice clock. Phys Rev A. 2009;79 061401. [Google Scholar]
  • 16.Sander TH, Preusser J, Mhaskar R, Kitching J, Trahms L, Knappe S. Magnetoencephalography with a chip-scale atomic magnetometer. Biomed Opt Express. 2012;3:981. doi: 10.1364/BOE.3.000981. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Sheng D, Li S, Dural N, Romalis MV. Subfemtotesla scalar atomic magnetometry using multipass cells. Phys Rev Lett. 2013;110 doi: 10.1103/PhysRevLett.110.160802. 160802. [DOI] [PubMed] [Google Scholar]
  • 18.Hosten O, Engelsen NJ, Krishnakumar R, Kasevich MA. Measurement noise 100 times lower than the quantum-projection limit using entangled atoms. Nature. 2016;529:505. doi: 10.1038/nature16176. [DOI] [PubMed] [Google Scholar]
  • 19.Hall LT, Cole JH, Hill CD, Hollenberg LCL. Sensing of fluctuating nanoscale magnetic fields using nitrogen-vacancy centers in diamond. Phys Rev Lett. 2009;103 doi: 10.1103/PhysRevLett.103.220802. 220802. [DOI] [PubMed] [Google Scholar]
  • 20.Bienfait A, Pla JJ, Kubo Y, Stern M, Zhou X, Lo CC, Weis CD, Schenkel T, Thewalt MLW, Vion D, Esteve D, et al. Reaching the quantum limit of sensitivity in electron spin resonance. Nat Nanotechnol. 2016;11:253. doi: 10.1038/nnano.2015.282. [DOI] [PubMed] [Google Scholar]
  • 21.Budker D, Romalis M. Optical magnetometry. Nature Phys. 2007;3:227. [Google Scholar]
  • 22.Miffre A, Jacquey M, Buchner M, Trenec G, Vigue J. Atom interferometry. Phys Scr. 2006;74:C15. [Google Scholar]
  • 23.Yurke B, McCall SL, Klauder JR. Su(2) and su(1,1) interferometers. Phys Rev A. 1986;33 doi: 10.1103/physreva.33.4033. 4033. [DOI] [PubMed] [Google Scholar]
  • 24.Colangelo G, Sewell RJ, Behbood N, Ciurana FM, Triginer G, Mitchell MW. Quantum atom–light interfaces in the gaussian description for spin-1 systems. New J Phys. 2013;15 103007. [Google Scholar]
  • 25.Braginsky VB, Vorontsov YI, Thorne KS. Quantum nondemolition measurements. Science. 1980;209:547. doi: 10.1126/science.209.4456.547. [DOI] [PubMed] [Google Scholar]
  • 26.Tsang M, Caves CM. Evading quantum mechanics: Engineering a classical subsystem within a quantum environment. Phys Rev X. 2012;2 031016. [Google Scholar]
  • 27.Polzik ES, Hammerer K. Trajectories without quantum uncertainties. Ann Phys. 2015;527:A15. [Google Scholar]
  • 28.Smith GA, Chaudhury S, Silberfarb A, Deutsch IH, Jessen PS. Continuous weak measurement and nonlinear dynamics in a cold spin ensemble. Phys Rev Lett. 2004;93 doi: 10.1103/PhysRevLett.93.163602. 163602. [DOI] [PubMed] [Google Scholar]
  • 29.Koschorreck M, Napolitano M, Dubost B, Mitchell MW. Sub-projection-noise sensitivity in broadband atomic magnetometry. Phys Rev Lett. 2010;104 doi: 10.1103/PhysRevLett.104.093602. 093602. [DOI] [PubMed] [Google Scholar]
  • 30.Ciurana FM, Colangelo G, Sewell RJ, Mitchell MW. Real-time shot-noise-limited differential photodetection for atomic quantum control. Opt Lett. 2016;41:2946. doi: 10.1364/OL.41.002946. [DOI] [PubMed] [Google Scholar]
  • 31.Kubasik M, Koschorreck M, Napolitano M, de Echaniz SR, Crepaz H, Eschner J, Polzik ES, Mitchell MW. Polarization-based light-atom quantum interface with an all-optical trap. Phys Rev A. 2009;79 043815. [Google Scholar]
  • 32.Deutsch IH, Jessen PS. Quantum control and measurement of atomic spins in polarization spectroscopy. Optics Communications. 2010;283:681. [Google Scholar]
  • 33.Kuzmich A, Mandel L, Bigelow NP. Generation of spin squeezing via continuous quantum nondemolition measurement. Phys Rev Lett. 2000;85:1594. doi: 10.1103/PhysRevLett.85.1594. [DOI] [PubMed] [Google Scholar]
  • 34.Appel J, Windpassinger PJ, Oblak D, Hoff UB, Kjær-gaard N, Polzik ES. Mesoscopic atomic entanglement for precision measurements beyond the standard quantum limit. Proc Nat Acad Sci. 2009;106 doi: 10.1073/pnas.0901550106. 10960. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Ciurana FM, Colangelo G, Sewell RJ, Mitchell MW. Real-time shot-noise-limited differential photodetection for atomic quantum control. Opt Lett. 2016;41:2946. doi: 10.1364/OL.41.002946. [DOI] [PubMed] [Google Scholar]
  • 36.Gühne O, Tóth G. Entanglement detection. Phys Rep. 2009;474:1. [Google Scholar]
  • 37.Sørensen AS, Mølmer K. Entanglement and extreme spin squeezing. Phys Rev Lett. 2001;86 doi: 10.1103/PhysRevLett.86.4431. 4431. [DOI] [PubMed] [Google Scholar]
  • 38.Béguin J-B, Bookjans EM, Christensen SL, Sørensen HL, Müller JH, Polzik ES, Appel J. Generation and detection of a sub-poissonian atom number distribution in a one-dimensional optical lattice. Phys Rev Lett. 2014;113 doi: 10.1103/PhysRevLett.113.263603. 263603. [DOI] [PubMed] [Google Scholar]
  • 39.Schlosser N, Reymond G, Protsenko I, Grangier P. Sub-poissonian loading of single atoms in a microscopic dipole trap. Nature. 2001;411:1024. doi: 10.1038/35082512. [DOI] [PubMed] [Google Scholar]
  • 40.Hofmann CS, Günter G, Schempp H, Robert-de Saint-Vincent M, Gärttner M, Evers J, Whitlock S, Weidemüller M. Sub-Poissonian statistics of Rydberg-interacting dark-state polaritons. Phys Rev Lett. 2013;110 doi: 10.1103/PhysRevLett.110.203601. 203601. [DOI] [PubMed] [Google Scholar]
  • 41.Gajdacz M, Hilliard AJ, Kristensen MA, Pedersen PL, Klempt C, Arlt JJ, Sherson JF. Preparation of ultracold atom clouds at the shot noise level. Phys Rev Lett. 2016;117 doi: 10.1103/PhysRevLett.117.073604. 073604. [DOI] [PubMed] [Google Scholar]
  • 42.Hosten O, Engelsen NJ, Krishnakumar R, Kasevich MA. Measurement noise 100 times lower than the quantum-projection limit using entangled atoms. Nature. 2016;529:505. doi: 10.1038/nature16176. [DOI] [PubMed] [Google Scholar]
  • 43.Stockton JK. Continuous quantum measurement of cold alkali-atom spins. Ph.D. thesis, California Institute of Technology; 2007. [Google Scholar]
  • 44.Takano T, Fuyama M, Namiki R, Takahashi Y. Spin squeezing of a cold atomic ensemble with the nuclear spin of one-half. Phys Rev Lett. 2009;102 doi: 10.1103/PhysRevLett.102.033601. 033601. [DOI] [PubMed] [Google Scholar]
  • 45.Schleier-Smith MH, Leroux ID, Vuletić V. States of an ensemble of two-level atoms with reduced quantum uncertainty. Phys Rev Lett. 2010;104 doi: 10.1103/PhysRevLett.104.073604. 073604. [DOI] [PubMed] [Google Scholar]
  • 46.Sewell RJ, Koschorreck M, Napolitano M, Dubost B, Behbood N, Mitchell MW. Magnetic sensitivity beyond the projection noise limit by spin squeezing. Phys Rev Lett. 2012;109 doi: 10.1103/PhysRevLett.109.253605. 253605. [DOI] [PubMed] [Google Scholar]
  • 47.Bohnet JG, Cox KC, Norcia MA, Weiner JM, Chen Z, Thompson JK. Reduced spin measurement back-action for a phase sensitivity ten times beyond the standard quantum limit. Nat Photon. 2014;8:731. [Google Scholar]
  • 48.Colangelo G, Sewell RJ, Behbood N, Ciurana FM, Triginer G, Mitchell MW. Quantum atom–light interfaces in the gaussian description for spin-1 systems. New J Phys. 2013;15 103007. [Google Scholar]
  • 49.Madsen LB, Mølmer K. Spin squeezing and precision probing with light and samples of atoms in the gaussian description. Phys Rev A. 2004;70 052324. [Google Scholar]
  • 50.Koschorreck M, Napolitano M, Dubost B, Mitchell MW. Sub-projection-noise sensitivity in broadband atomic magnetometry. Phys Rev Lett. 2010;104 doi: 10.1103/PhysRevLett.104.093602. 093602. [DOI] [PubMed] [Google Scholar]
  • 51.Koschorreck M, Napolitano M, Dubost B, Mitchell MW. Quantum nondemolition measurement of large-spin ensembles by dynamical decoupling. Phys Rev Lett. 2010;105 doi: 10.1103/PhysRevLett.105.093602. 093602. [DOI] [PubMed] [Google Scholar]
  • 52.Sewell RJ, Napolitano M, Behbood N, Colangelo G, Mitchell MW. Certified quantum non-demolition measurement of a macroscopic material system. Nat Photon. 2013;7:517. [Google Scholar]
  • 53.Behbood N, Martin Ciurana F, Colangelo G, Napolitano M, Tóth G, Sewell RJ, Mitchell MW. Generation of macroscopic singlet states in a cold atomic ensemble. Phys Rev Lett. 2014;113 doi: 10.1103/PhysRevLett.113.093601. 093601. [DOI] [PubMed] [Google Scholar]
  • 54.Kendall M, Stuart A. The advanced theory of statistics. Vol. 2. London: Griffin; 1979. [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

The datasets generated and analysed during the current study are available from the corresponding author on reasonable request. The data shown in Fig. 2 and all the data used to generate plots of Extended Data are included as Source Data.

RESOURCES