<<

Quantum dynamics of collective spin states in a thermal

Roy Shaham,1, 2, ∗ Or Katz,1, 2 and Ofer Firstenberg1 1Department of Physics of Complex Systems, Weizmann Institute of Science, Rehovot 76100, Israel 2Rafael Ltd, IL-31021 Haifa, Israel Ensembles of alkali-metal or noble-gas atoms at room temperature and above are widely applied in quantum optics and metrology owing to their long-lived spins. Their collective spin states maintain nonclassical nonlocal correlations, despite the atomic thermal motion in the bulk and at the boundaries. Here we present a stochastic, fully quantum description of the effect of atomic diffusion in these systems. We employ the Bloch-Heisenberg- Langevin formalism to account for the quantum noise originating from diffusion and from various boundary conditions corresponding to typical wall coatings, thus modeling the dynamics of nonclassical spin states with spatial interatomic correlations. As examples, we apply the model to calculate spin noise spectroscopy, temporal relaxation of squeezed spin states, and the coherent coupling between two spin species in a hybrid system.

I. INTRODUCTION bulk and at the system’s boundaries, still lack a proper fully quantum description. Gaseous spin ensembles operating at room temperature and above have attracted much interest for decades. At ambi- ent conditions, alkali-metal vapors and odd of noble In this paper, we describe the effect of spatial diffusion exhibit long spin-coherence times, ranging from mil- on the quantum state of warm spin gases. Using the Bloch- liseconds to hours [1–6]. These spin ensembles, consisting Heisenberg-Langevin formalism, we identify the dissipation of a macroscopic number of atoms, are beneficial for preci- and noise associated with atomic thermal motion and with the sion sensing, searches of new physics, and demonstration of scattering off the cell boundaries. Existing significant work macroscopic quantum effects [7–13]. In particular, manipula- in this field relies primarily on mean-field models, which ad- tions of collective spin states allow for demonstrations of ba- dress both wall coupling [40] and diffusion in unconfined sys- sic quantum phenomena, including entanglement, squeezing, tems [41]. The latter work derives the correlation function of and teleportation [14–17] as well as storage and generation of diffusion-induced quantum noise from the correlation func- photons [18–21]. It is the collectively enhanced coupling and tion of mass diffusion in unconfined systems. Here we derive the relatively low noise offered by these spin ensembles that the quantum noise straight out of Brownian motion consider- make them particularly suitable for metrology and quantum ations and provide a solution for confined geometries. Our information applications. model generalizes the mean-field results and enables the de- Thermal atomic motion is an intrinsic property of the scription of interatomic correlations and collective quantum dynamics in gaseous systems. Gas-phase atoms, in low- states of the ensemble. We apply the model to highly polar- pressure room-temperature systems, move at hundreds of me- ized spin vapor and analyze the effect of diffusion in various ters per second in ballistic trajectories, crossing the cell at sub- conditions, including spin noise (SN) spectroscopy [12, 42– millisecond timescales and interacting with its boundaries. To 46], spin squeezing [14, 32], and coupling of alkali-metal to suppress wall collisions, buffer gas is often introduced, which noble-gas spins in the strong-coupling regime [33, 34]. renders the atomic motion diffusive via velocity-changing col- lisions [22]. At the theory level, the effect of diffusion on the mean spin has been extensively addressed, essentially by de- The paper is arranged as follows. We derive in Sec. II scribing the evolution of an impure (mixed) spin state in the the Bloch-Heisenberg-Langevin model for the evolution of cell using a mean-field approximation [23–28]. This common

arXiv:2006.04243v2 [quant-ph] 2 Aug 2020 the collective spin operator due to atomic Brownian motion formalism treats the spatial dynamics of an average atom in and cell boundaries. We focus on highly polarized ensembles any given position using a spatially dependent density matrix. in Sec. III and provide the model solutions. In Sec. IV, we It accurately captures the single-atom dynamics but neglects present several applications of our model. We discuss how it is both interatomic correlations and thermal fluctuations associ- employed to describe the temporal evolution, to calculate ex- ated with the spin motion and collisions. perimental results, to provide insight, and to optimize setups Nonclassical phenomena involving collective spin states, for specific tasks. Limits of our model and differences from such as transfer of quantum correlations between nonoverlap- existing models, as well as future prospects, are discussed in ping light beams by atomic motion [29–31], call for a quan- Sec. V. We provide appendices that elaborate on the quantum tum description of the thermal motion. For spin-exchange noise produced by thermal motion (Appendix A), a simpli- collisions, which are an outcome of thermal motion, such a fied model for analyzing the scattering off the cell walls (Ap- quantum description has received much recent attention [32– pendix B), means of solving the Bloch-Heisenberg-Langevin 39]. However, the more direct consequences of thermal mo- equation (Appendix C), and the Faraday rotation scheme used tion, namely, the stochasticity of the spatial dynamics in the herein (Appendix D). 2

(a) (b) A. Diffusion in the bulk

We consider the limit of gas-phase atoms experiencing fre- (c) quent, spin-preserving, velocity-changing collisions, such as those characterizing a dilute alkali-metal vapor in an inert buffer gas. In this so-called Fickian diffusion regime, the atomic motion is diffusive, and the local density evolution can (d) be described by the stochastic differential equation [47] 2 √ ∂na/∂t = D∇ na + ∇(η na), (2) where D is the diffusion coefficient, and η is a white Gaus- sian stochastic process the components of which satisfy 0 0 0 0 hηi(r, t)ηj(r , t )ic = 2Dδijδ(r − r )δ(t − t ) for i, j = Figure 1. (a) Atomic spins in the gas phase, comprising a collective x, y, z. We use h·ic to represent ensemble average over the quantum spin ˆs(r, t) and undergoing thermal motion. (b) In the dif- classical atomic trajectories, differing from the quantum ex- fusive regime, the spins spatially redistribute via frequent velocity- pectation value h·i. The first term in Eq. (2) leads to delo- changing collisions. (c) Collisions (local interaction) with the walls calization of the atomic position via deterministic diffusion, of the gas cell may fully or partially depolarize the spin state. (d) while the second term introduces fluctuations that localize the Diffusion and wall collisions lead to a multimode evolution, here ex- atoms to discrete positions. Equation (2), derived by Dean emplified for a spin excitation aˆ(r, t) ∝ ˆs (r, t) − iˆs (r, t) with x y for Brownian motion in the absence of long-range interactions an initial Gaussian-like spatial distribution haˆ(r, t)i and for destruc- [47], is valid under the coarse-grain approximation, when the tive wall collisions. In addition to the mode-specific decay Γn, each spatial mode accumulates mode-specific quantum noise Wˆ n(t). temporal and spatial resolutions are coarser than the mean- free time and path between collisions. Substituting ∂na/∂t into Eq. (1), we obtain the Bloch- Heisenberg-Langevin dynamical equation for the collective II. MODEL spin ∂ˆs/∂t = i[H, ˆs] + D∇2ˆs + fˆ. (3) Consider a warm ensemble of Na atomic spins confined in a cell, as illustrated in Fig. 1a. Let r (t) be the classical lo- Here H is the spin Hamiltonian in the absence of atomic mo- a ˆ cation of the ath atom at time t and define the single-body tion, originating from the ∂sa/∂t term in Eq. (1). The quan- tum noise operator fˆ = fˆ(r, t) is associated with the local density function at some location r as na(r) = δ[r − ra(t)]. We denote the spin operator of the ath atom by ˆs and de- fluctuations of the atomic positions. It can be formally written a ˆ √ P fine the space-dependent collective spin operator as ˆs(r, t) = as fµ = ∇(ˆsµη/ n), where µ = x, y, z, and n = a na PNa is the atomic density. The noise term has an important role a=1 ˆsana(r). While formally ˆs(r, t) is sparse and spiked, practical experiments address only its coarse-grained proper- in preserving the mean spin moments of the ensemble. The fˆ = ties, e.g. ,due to finite spatial scale of the employed optical commutation relation of different instances of the noise µ ˆ ˆ0 ˆ 0 0 or magnetic fields. The time evolution of the collective spin fµ(r, t) and fν = fν (r , t ) satisfies operator is given by ˆ ˆ0 0 0 0 h[fµ, fν ]ic = 2iξµν D(∇∇ )ˆsξδ(r − r )δ(t − t ), (4)

where ξµν is the Levi-Civita antisymmetric tensor. These Na ∂ˆs X ∂ˆsa ∂na commutation relations ensure the conservation of spin com- = na + ˆsa . (1) 0 0 ∂t ∂t ∂t mutation relations [ˆsµ(r, t),ˆsν (r , t)] = iξµνˆsξδ(r − r ) on a=1 the operator level, compensating for the diffusion-induced de- cay in the bulk due to the D∇2 term. We provide the full Here the first term accounts for the internal degrees of free- derivation of fˆ and its properties in Appendix A. dom, including the local Hamiltonian evolution of the spins The spin noise process is temporally white and spatially and spin-spin interactions, while the second term accounts for colored, with higher noise content for shorter wavelengths. the external degrees of freedom, namely, for motional effects. The increase of noise at a fine-grain scale counteracts the The focus of this paper is on the second term, considered in diffusion term, which decreases the spin variations faster at the diffusion regime as illustrated in Fig. 1b. We consider the smaller length scales; this is a manifestation of the fluctuation- first term only for its contribution to the boundary conditions, dissipation theorem. Finally, as expected, ensemble averaging via the effect of wall collisions as illustrated in Fig. 1c. In the over the noise realizations leaves only the diffusion term in the following, we first derive the equations governing the quan- mean-field Bloch equation for the spin ∂hsi/∂t = D∇2hsi, tum operator ˆs(r, t) in the bulk and subsequently introduce where hsi = hˆs(r, t)i is the spin expectation value at a course- the effect of the boundaries. grained position r. 3

B. Boundary conditions In the limit of perfect spin-preserving coating, the bound- ary condition becomes a no-flux (Neumann) condition satis- We now turn to derive the contribution of wall collision fying (nˆ · ∇)ˆs = 0, and depolarization is minimized. This to the quantum dynamics of the collective spin. When the limit is realized for N  R/λ, where R is the dimension atoms diffuse to the boundaries of the cell, their spin inter- of the cell [58] [7]. In the opposite limit of strongly depo- acts with the surface of the walls. This interaction plays an larizing walls, i.e., N . 1, the (Dirichlet) boundary condi- −1/N important role in determining the depolarization and deco- tion is ˆs = wˆ /(1 − e ) [59], rendering the scattered spin herence times of the total spin [7, 22] and may also induce state random. For any other value of N (partially depolariz- frequency shifts [24, 48–51]. Bare glass strongly depolarizes ing walls), the boundary condition in Eq. (7) is identified as a alkali-metal atoms, and magnetic impurities in the glass affect stochastic Robin boundary condition [60]. the nuclear spin of noble-gas atoms. To attenuate the depolar- The two mechanisms discussed in this section — the bulk ization at the walls, cells can be coated with spin-preserving diffusion and the wall coupling — are independent physical coatings such as paraffin [4, 52, 53] or octadecyltrichlorosi- processes. This is evident by the different parameters charac- lane (OTS) [40] for alkali-metal vapor and Surfasil or SolGel terizing them — D and N — which are dictated by different [54–57] for spin-polarized xenon. The coupling between the physical scales, such as buffer gas pressure and the quality of spins and the cell walls constitutes the formal boundary con- the wall coating. These processes are different in nature; while ditions of Eq. (3). wall coupling leads to spin depolarization and thermalization, In the mean-field picture, the wall coupling can be de- diffusion leads to spin redistribution while conserving the total scribed as a local scatterer for the spin-density matrix ρ. In spin. They introduce independent fluctuations and dissipation, this picture, assisted by kinetic gas theory, the boundary con- and they affect the spins at different spatial domains (the bulk ditions can be written as [23] and the boundary). That being said, both processes are nec- (1 + 2 λnˆ · ∇)ρ = (1 − 2 λnˆ · ∇)Sρ, (5) essary to describe the complete spin dynamics in a confined 3 3 volume, simultaneously satisfying Eqs. (3) and (7). where S is the wall scattering matrix. Here λ denotes the mean free path of the atoms, related to the diffusion coefficient via D = λv/¯ 3, where v¯ is the mean thermal velocity. III. POLARIZED ENSEMBLES We adopt a similar perspective in order to derive the cou- pling of the collective spin ˆs with the walls in the Bloch- When discussing nonclassical spin states for typical ap- Heisenberg-Langevin formalism. In this formalism, the scat- plications, it is beneficial to consider the prevailing limit of tering off the walls introduces not only decay, but also fluc- highly polarized ensembles. Let us assume that most of tuations. In the Markovian limit, when each scattering event the spins point downwards (−zˆ). In this limit, we follow is short, its operation on a single spin becomes a stochastic the Holstein-Primakoff transformation [61, 62] and approx- density matrix imate the longitudinal spin component by its mean value

−1/N ˆsz(r, t) = sz (with sz = −n/2 for spin 1/2). The ladder Sˆsa = e ˆsa + wˆ a. (6) operator ˆs− = ˆsx − iˆsy, which flips a single spin downwards at position r, can be represented by the annihilation operator Here N denotes the average number of wall collisions a spin p withstands before depolarizing [40]. The accompanied quan- aˆ = ˆs−/ 2|sz|. This operator satisfies the bosonic commu- tation relations [ˆa(r, t), aˆ†(r0, t)] = δ(r − r0). Under these tum noise process is wˆ a; it ensures the conservation of spin commutation relations at the boundary. transformations, Eqs. (3) and (7) become Using the stochastic scattering matrix, we generalize the mean-field boundary condition [Eq. (5)] for collective spin op- 2 ˆ erators as ∂a/∂tˆ = i[H, aˆ] + D∇ aˆ + f, (8) −1/N 2 −1/N −1/N 2 −1/N (1 − e )ˆa = − 3 λ(1 + e )nˆ · ∇aˆ +w, ˆ (9) (1 − e )ˆs + 3 λ(1 + e )(nˆ · ∇)ˆs = wˆ . (7) ˆ ˆ ˆ p Here wˆ (r, t) = P wˆ n , for positions r on the cell bound- where both f = (fx − ify)/ 2|sz| and wˆ = (w ˆx + a a a p ary, is the collective wall-coupling noise process affecting the iwˆy)/ 2|sz| are now vacuum noise processes (see Appen- local spin on the wall. wˆ is zero on average and its statisti- dices A and B; note that fˆ is spatially colored). Here, Eq. (8) cal properties, together with the derivation of Eq. (6), are dis- describes the spin dynamics in the bulk, while Eq. (9) holds at cussed in Appendix B. The first term in Eq. (7) describes the the boundary. fractional depolarization by the walls, and the second term We solve Eqs. (8) and (9) by decomposing the operators describes the difference between the spin flux entering and into a superposition of nonlocal diffusion modes aˆ(r, t) = P exiting the wall. If the wall coupling also includes a coher- n aˆn(t)un(r). We first identify the mode functions un(r) ent frequency-shift component, it can be appropriately added by solving the homogeneous Helmholtz equation (D∇2 + to these terms. The term on the right-hand side describes the Γn)un(r) = 0, where the eigenvalues −Γn are fixed by the associated white fluctuations. Robin boundary condition [Eq. (9) without the noise term]. 4

R ∗ 3 The operator aˆn(t) = V aˆ(r, t)un(r)d r, where V is the depend on the beam size, cell dimensions, and diffusion char- cell volume, annihilates a collective transverse spin excita- acteristics. On the mean-field level, this effect has been de- 2 tion with a nonlocal distribution |un(r)| and a relaxation rate scribed by motion of atoms in and out of the beam [28, 65]. Γn. These operators satisfy the bosonic commutation relation Here we calculate the SNS directly out of the quantum noise † ˆ [ˆan, aˆm] = δnm. The noise terms f and wˆ are decomposed us- induced by the thermal motion as derived above. ing the same mode-function basis. This leads to mode-specific For concreteness, we consider two cylindrical cells of ra- noise terms Wˆ n(t), operating as independent sources. dius R = 1 cm and length L = 3 cm. One cell con- Assuming, for the sake of example, a magnetic (Zeeman) tains 100 Torr of buffer gas, providing for λ = 0.5 µm and 2 Hamiltonian H = ω0ˆsz, where ω0 is the Larmor precession D = 1 cm /s, and no spin-preserving coating N . 1 (e.g., as frequency around a zˆ magnetic field, the time evolution of the in Ref. [32]). The other cell has a high-quality paraffin coat- mode operators is given by ing, allowing for N = 106 wall collisions before depolariza- tion [4], and only dilute buffer gas originating from outgassing −(iω0+Γn)t 3 2 aˆn(t) =a ˆn(0)e + Wˆ n(t). (10) of the coating, such that λ = 1 mm and D = 3 · 10 cm /s rendering the atomic motion in the Fickian regime (λ  R,L) The multimode decomposition and evolution are illustrated in [66, 67]. A probe beam with waist radius w0 measures the Fig. 1d, showing the first angular-symmetric mode distribu- alkali-metal spin xˆ component, oriented along the cylinder tions un(r) of a cylindrical cell. In Table I, we provide explicit axis as presented in Fig. 2a. The cell is placed inside a mag- solutions of the mode bases and associated decay rates for any netic field B = 2πf0/ga · zˆ pointing along the spin polariza- given boundary properties in either rectangular, cylindrical, or tion, where ga is the alkali-metal gyromagnetic ratio. spherical cells. The solution procedure and the correspond- In Appendix D, we review the measurement details and cal- ing decomposition of the noise terms are demonstrated for culate the spin noise spectral density Sxx(f) for both cells an exemplary one-dimensional (1D) geometry in Appendix C. (G) 2 ˜ We note that, asymptotically, the decay of high-order modes X |In | 2P Γn S (f) = , (11) n  1 xx 2 2 2 ( ) is independent of cell geometry and is approximately 4 Γn + 4π (f − f0) −1/3 2 −1/3 n given by Γn ∼ D(πnV ) , where πnV approximates the mode’s wave number. where f is the frequency in which the SNS is examined, Γn (G) is again the decay rate of the nth diffusion mode, In is the overlap of the Gaussian probe beam with that mode, and P˜ IV. APPLICATIONS depends on the spin statistics and on the polarization, such that P˜ = 1 for highly polarized ensembles and for spin 1/2. The outlined Bloch-Heisenberg-Langevin formalism ap- The calculated spectra are shown in Figs. 2b and 2c for plies to various experimental configurations and applications. w0 = 1 mm. The cusp-like spectra originate from a sum of It should be particularly useful when two constituents of the Lorentzians, the relative weights of which correspond to the (G) 2 same system have different spatial characteristics, leading to overlap of the probe beam with each given mode |In | . In different spatial modes. That occurs, for example, when cou- the past, this cusp was identified as a universal phenomena pling spins to optical fields [Fig. 1d] or when mixing atomic [65], while here we recreate this result using the eigenmodes species with different wall couplings. In this section, we con- and accounting for the boundary. With spin-preserving coat- sider three such relevant, real-life cases. ing, the uniform mode n = 0 decays slower, and its contri- bution to the noise spectrum is much more pronounced, while the higher-order modes decay faster due to lack of buffer gas. A. Spin noise spectroscopy The dominance of the central narrow feature thus depends on the overlap of the probe with the least-decaying mode (G) 2 SNS allows one to extract physical data out of the noise |I0 | . To quantify it, we define the unitless noise content R f1/2 ζ = Sxx(f)df as the fraction of the noise residing properties of the spin system. It is used for magnetometry with −f1/2 atomic ensembles in or out of equilibrium [12, 43, 44, 46, 63], within the full width at half maximum of the spectrum. Figure for low-field NMR [64], for fundamental noise studies aimed 2d shows ζ for different beam sizes w0/R. Evidently, the spin at increasing metrological sensitivity [9, 12], and more [42]. resonance is more significant in the buffer gas cell, unless the SNS is also used to quantify interatomic correlations in probe beam covers the entire cell. This should be an important squeezed states, when it is performed with precision surpass- consideration in the design of such experiments. ing the standard quantum limit [14, 32, 39, 43]. Spin noise in an alkali-metal vapor is affected by various dephasing mechanisms. Here we describe the effect of diffu- B. Squeezed-state lifetime sion, given a spatially fixed light beam employed to probe the spins. Since this probe beam may overlap with several spa- When the spin noise is measured with a sensitivity below tial modes of diffusion, the measured noise spectrum would the standard quantum limit, the spin ensemble is projected 5

Figure 2. Effect of thermal motion in spin noise spectroscopy. (a) The spins are initially polarized along a magnetic field Bzˆ, and the spin projection hˆsxi is measured by Faraday rotation using a nondiverging Gaussian probe beam. The calculation assumes a probe waist radius of w0 = 1 mm and a cylindrical cell with radius R = 1 cm and length L = 3 cm. (b) Spin noise spectrum for an uncoated cell with a dense buffer gas (D = 1 cm2/s, N ≤ 1), calculated using Eq. (11). Many spatial (diffusion) modes contribute to the noise, and thus the total signal is a weighted sum of varying Lorentzians, producing a cusp profile. (c) Spin noise spectrum for a cell coated with high-quality paraffin coating (D = 3 · 103 cm2/s, N = 106). The cusp is wider (since D is larger), except for an additional sharp feature associated with the uniform spatial mode, the slow decay of which is governed by wall collisions. The dotted lines in (b) and (c) are a simple Lorentzian with 2 2 width Γw = π D/w0, provided as reference for a single-mode approximation. (d) Noise content in the vicinity of the resonance (as defined in the main text) for the same two cells and varying probe waists. The narrower the probe beam, the larger its overlap with the high-order, rapidly decaying, diffusion modes, thus leading to weaker signal in the central resonance feature. This effect becomes more pronounced for lower buffer gas pressures. into a collective squeezed spin state. Such measurements cay of each spatial mode. The importance of thermal motion are done primarily using optical Faraday rotation in paraffin- grows as the degree of squeezing increases, as the latter re- coated cells [14–16, 68] and recently also in the presence of lies on squeezing in higher-spatial modes. To see this, we buffer gas [32]. The duration of the probe pulse and the spa- plot in Fig. 3d the decay of squeezing in the buffer gas cell tial profile of the probe beam determine the spatial profile of with a wide probe beam and with the initial state extremely 2 the squeezed spin state and hence its lifetime. squeezed hxˆ (r, t = 0)i ≪ 1/4. The squeezing rapidly de- We shall employ the same two cells and geometry from the cays, as a power law, until only the lowest-order mode remains previous section [see Fig. 2a]. Given a probe pulse duration squeezed. This indicates the practical difficulty in achieving 2 much shorter than w0/D and assuming the measurement sen- and maintaining a high degree of squeezing. An interesting sitivity surpasses the standard quantum limit, a squeezed state behavior is apparent for the case of a large beam in a coated 2 is formed, with initial spin variance hxˆG(0)i ≤ 1/4, where cell [Fig. 3c, w0 = 8 mm]. Here, the significant overlap with xˆG(t) is the measured spin operator [defined in Appendix D the uniform produces a certain degree of squeezing that is es- as a weighted integral over the local operator xˆ(r, t)]. The pecially long lived. state is remeasured (validated) after some dephasing time t These results demonstrate the significance of accounting [see Fig. 3a]. In this type of experiments, narrow beams are for many diffusion modes when considering fragile nonclas- often preferable, as the local intensity affects the measurement sical states or high-fidelity operations. For example, the sensitivity, and since narrow beams simplify the use of optical presented calculations for the 25-dB squeezing require 1000 cavities. Considering the different diffusion modes un(r) with modes to converge. their decay rates Γn, we use Eq. (10) to calculate the evolution in the dark of the spin variance C. Coupling of alkali-metal spins to noble-gas spins 2 X (G) 2 −Γnt 2 2 hxˆG(t)i = ( |In | e ) (hxˆG(0)i−1/4)+1/4. (12) n Lastly, we consider collisional spin exchange between two Figures 3b and 3c present the calculated evolution. As ex- atomic species [2, 7, 33, 35, 36, 39, 69]. When the two species pected, a narrow probe beam squeezes the atoms with which experience different wall couplings, their spin dynamics is de- it overlaps, which are spanned in a superposition of diffu- termined by different diffusion-mode bases. Therefore mutual sion modes [the first low-order modes in the buffer gas cell spin exchange, which is due to a local coupling (atom-atom are visualized in Fig. 1d], leading to multimode temporal dy- collisions), depends on the mode overlap between these bases. namics. The measured squeezing decreases due to atoms dif- Here we consider the coupling of alkali-metal spins to fusing out of the beam, as manifested by the exponential de- noble-gas spins, such as -3, for potential applications 6

(a) (d) w /R>>1 with a collective rate J via spin-exchange collisions [33]. Un- 0 like the noble-gas spins, the alkali-metal spins are strongly af- 20 thermal motion fected by the cell walls, and consequently their low-order dif- in the dark 10 fusion modes ua (r) are different. This mode mismatch, be- t m tween ua (r) and ub (r), leads to fractional couplings c J, squeezing varification 0 m n mn R 3 a∗ b squeezing [dB] -6 -4 -2 0 10 10 10 10 where cmn = V d r um (r)un(r) are the overlap coeffi- t/T (b) (c) w cients. In particular, |cm0|J are the couplings to the uniform w =1mm w =1mm 0 0 (long lived) mode of the noble-gas spins. Usually, no anti- 6 w =2mm 6 w =2mm 0 0 relaxation coating is used in these experiments, thus |cm0|<1. w =4mm w =4mm 4 0 4 0 w =8mm w =8mm Here we demonstrate a calculation for a spherical cell of 0 0 radius R, for which the radial mode bases ua (r) and ub (r) 2 2 dominated by m n wall coupling Γ Γ squeezing [dB] squeezing [dB] and associated decay rates am and bn are presented in 0 0 Appendix C, alongside the first c values for an uncoated -2 -1 0 1 2 3 -4 -3 -2 -1 0 m0 10 10 10 10 10 10 10 10 10 10 10 cell (Table II). The calculation includes the first m, n ≤ t t [ms] [ms] 70 modes [71]. As the initial state, we consider the dou- bly excited (Fock) state of the alkali-metal spins |ψ i = Figure 3. Lifetime of spin squeezing. (a) Experimental sequence 0 √1 (P α a† )2|0i |0i = |2i |0i , where |0i |0i is the comprising a short measurement (squeezing) pulse, followed by de- 2 m m m a b a b a b phasing in the dark due to thermal motion for duration t, and a veri- vacuum state with all spins pointing downwards. We take fication pulse. The same probe beam is used for both pulses. In the the initial excitation to be spatially uniform, for which the P b calculations, the Gaussian distribution is initially squeezed by the coefficients αm = cm0 satisfy αmum(r) = u (r) = 2 m 0 measurement to spin variance of hxˆG(0)i = 0.05 (7-dB squeezing). (4πR3/3)−1/2. We calculate the transfer of this excitation We take the same cell geometry as in Fig. 2 (R = 1 cm, L = 3 cm). via spin exchange to the uniform mode ˆb of the noble-gas (b) Degree of spin squeezing vs time in the buffer gas cell, calculated 0 −1/2 ˆ† 2 using Eq. (12). Spin squeezing exhibits multiexponential decay as- spins, i.e., to the state |0ia|2ib = 2 (b0) |0ia|0ib. sociated with multiple diffusion modes. Larger probe beams lead to Figure 4 displays the exchange fidelity F = 2 longer squeezing lifetimes, as the beam overlaps better with lower- max |hψ(t)| |0ia|2ib| as a function of both spin-exchange order modes. (c) Spin squeezing in the coated cell. In a coated cell, rate J and quality of coating N. As N increases, the initial the decay rate of the uniform diffusion mode, dominated by wall cou- uniform excitation matches better the lower-order modes of pling, is substantially lower than that of higher-order modes. There- the alkali-metal spins, which couple better to the uniform fore, ensuring a significant overlap of the probe beam with the uni- modes of the noble-gas spins. Indeed we find that the form mode is even more important in coated cells for maximizing the squeezing lifetime. The dotted line in (b) and (c) is a single exponen- exchange fidelity grows with increasing J and N. tial decay hxˆ2(t)i = hxˆ2(0)ie−2Γwt + (1 − e−2Γwt)/4, shown for 2 2 reference with Γw = π D/w0 and w0 = 4 mm; note the difference in time-scales between (b) and (c). (d) An extremely squeezed state V. DISCUSSION relies more on higher-order spatial modes and thus loses its squeez- 2 2 ing degree rapidly (time is normalized by Tw = R /π D). The cal- We have presented a fully quantum model, based on a culation is initialized with a uniform distribution of squeezing and Bloch-Heisenberg-Langevin formalism, for the effects of dif- includes the first 1000 radial modes, required for convergence. fusion on the collective spin states in a thermal gas. The model is valid when the atomic mean free path is much shorter than the apparatus typical dimension. This is often the case in quantum optics [34]. The nuclear spins of noble gases are for warm alkali-metal-vapor systems, even when a buffer gas well protected by the enclosing complete electronic shells and is not deliberately introduced, as the out-gassing of a spin- thus sustain many collisions with other atoms and with the preserving wall coating can lead to mean free paths on the cell walls. Their lifetime typically reaches minutes and hours order of millimeters [66, 67]. [5, 6, 70]. In an alkali-metal–noble-gas mixture, the We have mostly focused on highly polarized spin ensem- acts as a buffer both for itself and for the alkali-metal atoms, bles, typically used to study nonclassical phenomena that em- so that both species diffuse, and their collective spin states can ploy the transverse component of the spin. It is important to be described by our Bloch-Heisenberg-Langevin model. note that Eqs. (3) and (7) hold generally and can be applied As the noble-gas spins do not relax by wall collisions, their to unpolarized systems as well. For example, the presented b lowest-order diffusion mode u0 (r) is that associated with the analysis of spin noise spectra holds for unpolarized vapor [ac- characteristic (extremely) long life time. Higher-order modes counting for suitable spin statistics in Eq. (11) using Eq. (D2)] b 2 un(r) decay due to diffusion with typical rates Γwalln = and is thus applicable to nonclassical experiments done in 2 2 2 n π D/R , where R is the length scale of the system. For that regime [32]. Our model can also describe other space- −1 −1 typical systems, Γwall is of the order of (1 ms) −(1 sec) . dependent phenomena, such as the dynamics in the presence Consequently, to enjoy the long lifetimes of noble-gas spins, of nonuniform driving fields [29]. one should employ solely the uniform mode. The presented model agrees with existing mean-field de- The alkali-metal spins couple locally to the noble-gas spins scriptions of diffusion of atomic spins. It further agrees with 7

gases and particularly to warm alkali-metal vapors. One such 1 example is a recent demonstration of transfer of quantum cor- 1 relations by the diffusion of alkali-metal atoms between dif- 0.8 ferent spatial regions [29]. Other examples involve a single 0.3 active region, e.g., when spin squeezing is performed using a 0.6 small probe beam over a long probing time, with the goal of 0.1 coupling efficiently to the uniform diffusion mode in a coated 0.4 cell [17, 21]. The resulting spatio-temporal dynamics can be 0.03 described using our model in order to assess the obtainable de- 0.2 gree of squeezing. In particular, our model predicts that high 0.01 buffer gas pressure would improve the lifetime of squeezed states when small probe beams are employed (e.g., when us- 0 0 ing optical cavities or when high probe intensities are needed), 100 101 102 103 thus encouraging the realization of such experiments. We thank Eugene Polzik for fruitful discussions and in- Figure 4. Excitation exchange between polarized alkali-metal and sights. We acknowledge financial support by a European Re- noble-gas spins. Shown is the exchange fidelity of the doubly ex- search Council starting investigator grant (Q-PHOTONICS cited (Fock) states |2ia|0ib and |0ia|2ib. We assume a spherical cell Grant No. 678674), the Israel Science Foundation, the Pazy containing potassium and helium-3. The quality N of the wall coat- Foundation, the Minerva Foundation with funding from the ing for the alkali metal is varied between no coating (N ≤ 1) and Federal German Ministry for Education and Research, and the perfect coating (N → ∞). The noble-gas spins do not couple to the Laboratory in Memory of Leon and Blacky Broder. cell walls. The exchange fidelity approaches 1 when J  Γwall, Γa, as then the spin exchange is efficient for many diffusion modes; here Γwall is the contribution of wall collisions to the relaxation rate of the alkali-metal spin (i.e., the typical diffusion rate to the walls), and Γa is the contribution of atomic collisions. The calculations are per- formed for a cell radius R = 5 mm and with 1 atm of helium-3. The Appendix A: Diffusion-induced noise 2 diffusion constants are Da = 0.35 cm /s for the potassium (mean 2 2 free path λa = 50 nm) so that Γwall = π Da/R = 1/(70 ms), 2 In the main text, we formulate the dynamics of a collective and Db = 0.7 cm /s for the helium (mean free path λb = 20 nm). −1 spin operator as driven from local density fluctuations. For The additional homogeneous decay of the alkali metal is Γa ≈ 6 s [7]. The wall coating plays a significant role, since for Nλa/R > 1 deriving Eq. (3), we use the Lagrangian version of Eq. (2), (i.e., N > 105) the diffusion modes of the potassium and helium where the noise is defined for each particle individually spins match. 2 (a) √ ∂na/∂t = D∇ na + ∇[η (t) na]. (A1) models employing the dissipation-fluctuation theorem to de- rive the spin noise spectrum from the decay associated with Here η(a)(t) is a white Gaussian process with vanishing mean 0 diffusion. Importantly, it extends all these models by describ- (a) (a) (a ) 0 hη ic = 0 and with correlations hηi (t)ηj (t )ic = ing quantum correlations and explicitly deriving the quantum 0 2Dδijδaa0 δ(t − t ). Substituting these into Eq. (1) provides noise of the Brownian motion. The suggested model assumes the definition for the quantum noise components as λ  R,L (Fickian diffusion) and thus does not hold for the special case of small, low-pressure, coated cells where the atomic motion is predominantly ballistic (λ & R,L) [21, 63]. Nevertheless, it may still provide a qualitative description Na ˆ X (a) (a) of the effect of wall collisions on the uniform spin distribu- fµ(r, t) = ˆsµ (t)∇[η na(r, t)]. (A2) tion across both diffusion regimes [63]. Our model lays the a=1 groundwork for treatments of such systems by considering non-Markovian motional dynamics. Following the lines of Ref. [47], we consider an alternative, Our results highlight the multimode nature of the dynam- equivalent definition ics. As exemplified for the applications considered in Sec. IV, one often needs to account for multiple diffusion modes, with √ fˆ (r, t) = ∇[ˆs (r, t)η(r, t)/ n], (A3) the high-order modes introducing additional quantum noise or µ µ reducing fidelities. As a rule of thumb, if ε is the allowed in- fidelity or excess quantum noise, then one should include the as also provided in the main text. According to both defini- first ∼ ε−1 modes in the calculations. tions, fˆ is a stochastic Gaussian process (linear operations on Since thermal motion is inherent to gas-phase systems, our a Gaussian process accumulate to a Gaussian process) with model could be beneficial to many studies of nonclassical spin a vanishing mean. Consequently, the equivalence of the two 8 definitions is a result of the equality of the noise correlations coupling is stochastic and Markovian, thus resulting in an ex- ponential decay of the scattered spin, and that the noise due to ˆ ˆ0 X X (a) (a) hfµfν ic =h ∇i( ˆsµ naηi )× diffusion in the bulk vanishes within a thin boundary layer at i a the wall. This leads to the scattering described by Eq. (6). The X 0 X (a0) (a0) accompanying noise processes for atoms a and a0 satisfy the ∇j( ˆsν na0 ηj )ic j a0 relations 0 X (a) (a) 0 0 0 =2D(∇ · ∇ )( ˆsµ ˆsν na)δ(r − r )δ(t − t ) µ ν 0 −1/N −1/N (a) $δ(t − t ) [w ˆ (t), wˆ 0 (t )] = i e (1−e )ˆs δ 0 , a a a µνξ ξ aa v¯ P (a) (a0) (B1) 0 aa0 ˆsµ naˆsν na0 0 0 1/N −1 =2D(∇ · ∇ ) P δ(r − r )δ(t − t ) where µ, ν = x, y, z. Here $ = (e − 1) λ/3 is the 0 na0 √a effective correlation distance of the wall-scattering noise, de- 0 0 0 0 0 =2D(∇ · ∇ )(ˆsµˆsν / nn )δ(r − r )δ(t − t ) fined such that the commutation relations of the spin oper- X √ X √ ators are conserved for all diffusion modes, i.e., for the en- =h ∇ (ˆs η / n) × ∇0 (ˆs0 η0 / n0)i , i µ i j ν j c tire cell (bulk and boundary). It changes monotonically from i j $ = e−1/N λ/3 for spin-destructing walls (N  1) to (A4) $ = Nλ/3 for spin-preserving walls (N  1). The continuous operator wˆ (r, t) used in the main text to where we used the identity na(r, t)na0 (r, t) = describe the noise due to interactions with the cell walls is δaa0 na(r, t)na0 (r, t). Here and henceforth, we use tags P to abbreviate the coordinates (r0, t0) for a field, i.e., defined as wˆ = a wˆ ana. It is the analog of wˆ a(t), like ˆs is F 0 = F (r0, t0) and F = F (r, t). to ˆsa. It vanishes for positions r the distance of which from The quantum noise, the commutation relations of which are the boundary is larger than $, and its commutation relations shown in Eq. (4), conserves the spin commutation relations are 0 0 [ˆsµ(r, t),ˆsν (r , t)] = iξµνˆsξδ(r − r ). This can be seen from

µ 0ν −1/N −1/N h[w ˆ , wˆ ]ic =iµνξe (1 − e )$/v¯ × 0 0 0 0 h[ˆsµ(r, t + dt),ˆsν (r , t + dt)] − [ˆsµ(r, t),ˆsν (r , t)]ic (A5) ˆsξ(r, t)δ(r − r )δ(t − t ). (B2) 0 = iξµν δ(r − r )hˆsξ(r, t + dt) − ˆsξ(r, t)ic The last expression is defined only for coordinates r and r0 and then on the cell boundary and vanishes elsewhere. As an example, for a rectangular cell with a wall at x = L/2, we shall define 0 2 0 iξµν D(∇ + ∇ ) [ˆsξδ(r − r )]dt (A6) coordinates on the boundary r⊥ = yyˆ + zzˆ and substitute 2 0 δ(r − r0) = 1 δ(y − y0)δ(z − z0) x = x0 = L/2 = iξµν D(∇ ˆsξ)δ(r − r )dt, $ at . For a spherical cell with a wall at |r| = R, we use δ(r − r0) = 0 where the last equality stems from (∇ + ∇0)δ(r − r0) = 0. 1 δ(Ω−Ω ) $ R2 , where Ω is the angular position of coordinate r. In Sec. III, we focus on highly polarized ensembles, where Using wˆ (r, t), the scattering matrix for the spin-density op- the dynamics is described by the bosonic annihilation oper- erator becomes Sˆs = e−1/N · ˆs + wˆ . We write Eq. (7) for the ator aˆ, under the Holstein-Primakoff approximation. Under spin density operator using the noise field wˆ . In addition, wˆ is these conditions, the thermal noise operating on the bosonic defined only on the boundary, such that (nˆ · ∇)wˆ | ∝ ˆ √ boundary excitations becomes f = ∇(ˆaη/ n). In addition, the same δ0(0) and therefore vanishes. † † 0 0 conditions ensure that aˆ aˆ = 0, aˆ(r, t)ˆa (r , t) = δ(r − r ), Finally, under the Holstein-Primakoff approximation, we † 0 0 0 and aˆ(r, t)ˆa (r , t)δ(r − r ) = nδ(r − r ), thus providing use Eq. (B2) to find the noise operating on aˆ due to wall scat- √ √ tering. The operator wˆ =w ˆ /p2|s | becomes a vacuum ˆˆ0† 0 † 0 0 0 − z hff ic = h∇aˆη/ n ∇ (ˆa ) η / n ic † 0 0 noise, satisfying hwˆ wˆ ic = 0, and for r and r on the cell = −2D∇2δ(r − r0)δ(t − t0), (A7) boundary,

ˆ† ˆ0 0† −1/N −1/N 0 0 and hf f ic = 0. Therefore, the noise becomes a vacuum hwˆwˆ ic = 2e (1 − e )$/vδ¯ (r − r )δ(t − t ). (B3) noise and conserves the commutation relations of the bosonic operators. We denote the correlations of the diffusion noise in Considering a general spin distribution, the noise due to the the bulk as C(r, r0) = −2D∇2δ(r − r0). walls exists only in a volume of order $S, where S is the cell surface area, while the noise due to diffusion in the bulk exists in the entire volume V . The ratio of the two scales as 0† ˆˆ0† Appendix B: Model for wall coupling hwˆwˆ )ic/hff )ic ∝ $S/V ∝ λ/R, where R is the typical dimension of the cell. Consequently, in our considered diffu- We adopt a simplified model for describing the scattering sive regime λ  R, the diffusion noise dominates over that of of atoms off the cell walls. The model assumes that the wall the wall scattering for nonuniform spin distributions. 9

N /L Appendix C: Solving the diffusion-relaxation 10 Bloch-Heisenberg-Langevin equations cot(kL/2) 1 8 k (N) n The diffusion-relaxation equation in the Bloch-Heisenberg- 0.3 6 Langevin formalism, in the limit of a highly polarized spin 0.1 gas, is presented in Sec. III. Here we first solve Eqs. (8) and 4 (9) for a simplified 1D case by following the method described 0.03 in the main text. We provide explicit expressions for the 2 0.01 mode-specific noise sources due to motion in the bulk and at 0 0 the boundary. Finally, we provide tabulated solutions for the 0 1 2 3 4 5 three-dimensional cases of rectangular, cylindrical, and spher- kL/2 ical cells. Consider a 1D cell with a single spatial coordinate −L/2 ≤ Figure 5. Graphical solutions for the Robin boundary condition. Here we solve the 1D equations for a system of length L = 1 cm x ≤ L/2. The functions uk(x) that solve the Helmholtz equation ∂2u /∂x2 + k2u = 0 are the relaxation-diffusion and mean free path of λ = 0.5 µm (characteristic of 100 Torr of k k buffer gas) and for different values of N. modes, where the decay rates Γ introduced in the main text 2 + + + are Γ = Dk . These solutions are uk = Ak cos(k x) u− = A− sin(k−x) and k k , composing symmetric and anti- equations in the absence of wall-induced fluctuations. There- symmetric modes. The annihilation operator decomposes into 2 P ± ± fore pˆ(x, t) is defined such that ∇ pˆ(x, t) = 0. a superposition aˆ(x, t) = k,± aˆk (t)uk (x). To further sim- plify the example, we take only symmetric spin distribution and symmetric noise into consideration, i.e., we keep only the + In our 1D symmetric case, pˆ(x, t) =p ˆ(t) is uniform. modes uk and omit the “+” superscript. Note that a physi- cal noise is random and generally has no defined symmetry, Writing the full boundary equation for aˆ provides pˆ(t) = −1/N but it can be decomposed into components with well-defined wˆ(t)/(1 − e ). We decompose pˆ(t) into the modes to ob- R L/2 symmetry. tain pˆn(t) = −L/2 pˆ(t)un(x)dx = 2An sin(knL/2)ˆp(t)/kn. The bulk diffusion equation becomes ∂aˆk/∂t = i[H, aˆk] − Substituting this in Eq. (8) provides the equation for the homo- L/2 ˆ Dk2aˆ + R fuˆ dx. We break the boundary equation into geneous mode operators hn. k −L/2 k ˆ a homogeneous part, where the noise is omitted, and an inho- In the case of a magnetic Zeeman Hamiltonian H = iω0Sz, mogeneous part, which includes the noise. The former can be we find decomposed into the different modes and is simplified to the ˆ ˆ ˆ ˆ 1+e−1/N ∂hn/∂t = −iω0hn − Γnhn + fn − iω0pˆn − ∂pˆn/∂t, (C2) algebraic equation cot(kL/2) = 2 1−e−1/N λk [72]. For general values of N, this is a Robin boundary condi- the solutions of which are tion, which can be solved numerically or graphically as pre- ˆ −(iω0+Γn)tˆ sented in Fig. 5. The discrete solutions kn define a complete hn = e hn (0) + t and orthonormal set of discrete modes u = A cos(k x), R −(iω0+Γn)(t−τ) ˆ ∂ n n n 0 e (fn(τ) − (iω0 + ∂τ )ˆpn(τ))dτ. spanning all symmetric spin distributions in the 1D cell, and (C3) L/2 ˆ R ∗ Substituting into aˆn(t) =p ˆn(t) + hn(t) and differentiating −L/2 umundx = δnm. These provide the discrete decay t rates Γn. with respect to provides the evolution of the annihilation op- For example, in the Dirichlet case of destructive walls erators of the spin modes (Nλ/L  1), k = (2n + 1)π/L. The annihilation operators n ∂aˆ /∂t = −(iω + Γ )ˆa + fˆ + fˆw, (C4) R L/2 ∗ n 0 n n n n of the various modes are aˆn(t) = −L/2 un(x)ˆa(x, t)dx, and ˆ R L/2 ∗ ˆ where the noise operators are fn(t) = −L/2 un(x)f(x, t)dx and wˆ(t) = [w ˆ(L/2, t) +w ˆ(−L/2, t)]/2. Z L/2 2A Γ sin(k L/2) ˆ fˆw = Γ u∗ (x)ˆp(x, t)dx = n n n wˆ The treatment of fn as a bulk source term operating on in- n n n −1/N −L/2 (1 − e )kn dependent modes is a common technique [73]. It differs, how- (C5) ever, from the treatment of the noise at the boundaries. We is the quantum noise due to wall collisions. Finally, we can deal with this term by defining auxiliary fields combine the two noise terms and obtain the total, mode- specific, noise operator X ˆ aˆ(x, t) =p ˆ(x, t) + hn(t)un(x), (C1) t n Z ± 2 ˆ −[iω0+D(kn ) ](t−τ) ˆ ˆw Wn = e [fn(τ) + fn (τ)]dτ, (C6) as we desire to use pˆ(x, t) to imbue the wall noise as a source 0 ˆ acting on the modes aˆn, while hn solves the homogeneous appearing in Eq. (10). 10

cell shape rectangular cylindrical spherical symmetric: (+) symmetry angular: n spherical: `, p anti-symmetric: (−) 0 ≤ r ≤ R 0 ≤ ρ ≤ R coordinate range −L/2 ≤ x ≤ L/2 0 ≤ θ ≤ π 0 ≤ ϕ ≤ 2π 0 ≤ ϕ ≤ 2π + 2 1+e−1/N + cot(k L/2) = λk −1/N −1/N n 3 1−e−1/N n Jn(kνnR) 2 1+e j`(kn`R) 2 1+e boundary equation −1/N − J0 (k R) = 3 −1/N λkνn − j0 (k R) = 3 −1/N λkn` − tan(k−L/2) = 2 1+e λk− n νn 1−e ` n` 1−e n 3 1−e−1/N n + + + un (x) = An cos(kn x) inϕ un(r) − − − uνn(ρ, ϕ) = AνnJn(kνnρ)e un`p(r, θ, ϕ) = An`pj`(kn`r)Y`p(θ, ϕ) un (x) = An sin(kn x)

Table I. Solutions of the diffusion-relaxation modes for rectangular, cylindrical, and spherical cells. Jn(x) is the nth Bessel function of the first 2 kind, j`(x) is the `th spherical Bessel function of the first kind, and Y`p(θ, ϕ) are the spherical harmonics. The decay rates satisfy Γn = Dkn.

cmn n = 0 n = 1 n = 2 n = 3 n = 4 illustrated in Fig. 2a, we consider a cylindrical cell with ra- m = 0 0.780 0.609 -0.126 0.058 -0.033 dius R and length L, with the cylinder axis along xˆ. The spins m = 1 -0.390 0.652 0.622 -0.158 0.079 are polarized along zˆ, parallel to an external applied magnetic m = 2 0.260 -0.274 0.647 0.627 -0.173 field B = Bzˆ. We use ρ and ϕ as the cylindrical coordinates, m = 3 -0.195 0.182 -0.256 0.644 0.629 m = 4 0.156 -0.139 0.1680 -0.246 0.643 and x as the axial coordinate. A linearly polarized probe beam travels along xˆ with a R 3 ∗ 2 2 Table II. Overlap coefficients, cmn = V d rAm(r)Bn(r), of the Gaussian intensity profile IG(r) = I0 exp(−2ρ /w0), where first five spherically symmetric modes, i.e., ` = p = 0. We take w0 is the beam waist radius. We assume a negligible beam Am(r) to be the diffusion modes of a spherical cell with radius R = divergence within the cell and require the normalization 1 and destructive walls, and Bn(r) to be the modes in the same cell R 2 3 −2 2 −4R2/w2 I (r)d r = 1, so that (I ) = πLw (1 − e 0 )/4. but with spin-conserving walls. V G 0 0 The probe frequency is detuned from the atomic transition, such that the probe is not depleted and does not induce addi- Under the influence of the noise sources Wˆ n and the dissi- tional spin decay. pation Γn, the spin operators of the diffusion modes obey the The linear polarization of the probe rotates due to the Fara- fluctuation-dissipation theorem, and their commutation rela- day effect, with the rotation angle proportional to the spin ˆ ˆw ˆ† ˆw† tions are conserved, resulting from h(fn0 +fn0 )(fn+fn )ic = projection along the beam propagation direction. Therefore, 0 ˆ† ˆw† ˆ ˆw 2Γnδn0nδ(t−t ) and h(fn0 +fn0 )(fn +fn )ic = 0. Note that measurement of the rotation angle provides a measurement of the conservation of local commutation relations is already pre- ˆsx weighted by its overlap with the beam profile. Precisely, ˆ R 3 sented in Appendices A and B (where f applies for the bulk the operator xˆG(t) = V d rIG(r)ˆx(r, t), where xˆ(r, t) = and wˆ for the boundary) without the mode decomposition. [ˆa(r, t) +a ˆ†(r, t)]/2, is measured in this scheme [17]. Notably, however, it also holds for the nonlocal (diffusion) We identify the atomic diffusion modes in the cylindrical modes. cell as un(r). Note that in Table I the modes require sev- For completeness, we provide in Table I the diffusion- eral labels, which we replace here with a single label n for relaxation modes for rectangular, cylindrical, and spherical brevity. We decompose the spin operator and the probe in- cells. Various applications, such as those involving collisional P tensity profile using the modes xˆ(r, t) = n xˆn(t)un(r) and (local) coupling between two spin ensembles, also require the P (G) † R 3 ∗ IG(r) = n In un(r), where xˆn(t) = (ˆan +a ˆn)/2 and overlap coefficients cmn = d rAm(r)Bn(r) between dif- V I(G) = R d3rI (r)u∗ (r). Using these, we express the mea- fusion modes Am(r) and Bn(r). These are presented in Table n V G n P (G) II for spherically-symmetric modes, where Am(r) are modes sured spin operator as xˆG(t) = n In xˆn(t). for highly destructive walls (N . 1), and Bn(r) are for inert We calculate the spin noise spectrum from its formal defi- walls (N  L/λ). These conditions are typical for a mixture nition of alkali-metal vapor and noble gas, as discussed in section IV. Z T Z T 1 0 2πif(τ−τ 0) 0 Sxx(f) = lim xˆG(τ)ˆxG(τ )e dτdτ T →∞ T 0 0 (D1) Appendix D: Faraday rotation measurement setup utilizing the temporal evolution of the modes as given by ˆ † ˆ Eq. (10), and including the noise properties Wn0 (t)Wn(t) = ˆ ˆ † −2Γnt In Sec. IV, we consider two experimental setups where the 0 and Wn0 (t)Wn(t) = (1 − e )δn0n derived from transverse component of a polarized spin ensemble is mea- Appendix C. The spin noise spectral density appearing in sured by means of the Faraday rotation. This scheme is com- Eq. (11) holds for both polarized and unpolarized ensembles, mon in alkali-metal spin measurements [14, 32, 74, 75]. As with 11

[26] O. Firstenberg, M. Shuker, A. Ben-Kish, D. R. Fredkin, ( N. Davidson, and A. Ron, Phys. Rev. A 76, 013818 (2007). 1 polarized P˜ = (D2) [27] O. Firstenberg, P. London, D. Yankelev, R. Pugatch, M. Shuker, 2(S + 1)/3 unpolarized and N. Davidson, Phys. Rev. Lett. 105, 183602 (2010), arXiv:1008.2654 [quant-ph]. where S is the single-particle spin magnitude. [28] Y. Xiao, I. Novikova, D. F. Phillips, and R. L. Walsworth, Phys. For the considered geometry, the standard quantum limit is Rev. Lett. 96, 043601 (2006). [1−exp(−2R2/w2)]2 ˆ2 nVbeam 0 [29] J. Sun, X. Zhang, W. Qu, E. E. Mikhailov, I. Novikova, H. Shen, hSxi ≥ Nbeam/4 = 2 2 , where Nbeam = 4 1−exp(−4R /w0 ) R 3 2 R 2 3 and Y. Xiao, Phys. Rev. Lett. 123, 203604 (2019). n[ V IG(r)d r] / V IG(r)d r is the number of atoms in the [30] H. Bao, J. Duan, S. Jin, X. Lu, P. Li, W. Qu, M. Wang, 2 beam, and Vbeam = πLw0 is the beam volume [76]. I. Novikova, E. E. Mikhailov, K.-F. Zhao, et al., Nature 581, 159 (2020). [31] H. Bao, M. Wang, P. Li, W. Qu, E. Mikhailov, I. Novikova, H. Shen, and Y. Xiao, in 2016 Proceedings of Progress in Elec- tromagnetic Research Symposium (PIERS) (IEEE, 2016) pp. ∗ [email protected] 3974–3974. [1] W. Happer, Rev. Mod. Phys. 44, 169 (1972). [32] J. Kong, R. Jiménez-Martínez, C. Troullinou, V. G. Lu- [2] W. Happer and A. C. Tam, Phys. Rev. A 16, 1877 (1977). civero, G. Toth, and M. W. Mitchell, Nat. Commun. 11, [3] O. Katz and O. Firstenberg, Nat. Commun. 9, 2074 (2018). 10.1038/s41467-020-15899-1 (2020). [4] M. V. Balabas, T. Karaulanov, M. P. Ledbetter, and D. Budker, [33] O. Katz, R. Shaham, and O. Firstenberg, arXiv preprint , Phys. Rev. Lett. 105, 070801 (2010). arXiv:1905.12532v2 (2019). [5] T. G. Walker and W. Happer, Rev. Mod. Phys. 69, 629 (1997). [34] O. Katz, R. Shaham, E. S. Polzik, and O. Firstenberg, Phys. [6] T. R. Gentile, P. J. Nacher, B. Saam, and T. G. Walker, Rev. Rev. Lett. 124, 043602 (2020). Mod. Phys. 89, 045004 (2017). [35] A. T. Dellis, M. Loulakis, and I. K. Kominis, Phys. Rev. A 90, [7] W. Happer, Y.-Y. Jau, and T. Walker, Optically Pumped Atoms 032705 (2014). (WILEY-VCH, 2010). [36] K. Mouloudakis, M. Loulakis, and I. K. Kominis, Phys. Rev. [8] J. M. Brown, S. J. Smullin, T. W. Kornack, and M. V. Romalis, Research 1, 033017 (2019). Phys. Rev. Lett. 105, 151604 (2010), 1006.5425. [37] K. Mouloudakis and I. K. Kominis, arXiv preprints , [9] D. Sheng, S. Li, N. Dural, and M. V. Romalis, Phys. Rev. Lett. arXiv:2004.11790v2 (2020). 110, 160802 (2013). [38] G. Vasilakis, V. Shah, and M. V. Romalis, Phys. Rev. Lett. 106, [10] D. Budker and M. Romalis, Nat. Phys. 3, 227 (2007). 143601 (2011). [11] D. Budker and D. F. J. Kimball, Optical Magnetometry (Cam- [39] D. Roy, L. Yang, S. A. Crooker, and N. A. Sinitsyn, Sci. Rep. bridge University Press, Cambridge, 2013). 5, 9573 (2015). [12] S. A. Crooker, D. G. Rickel, A. V. Balatsky, and D. L. Smith, [40] S. Seltzer and M. V. Romalis, J. Appl. Phys. 106, 114905 Nature (London) 431, 49 (2004). (2009). [13] I. M. Bloch, Y. Hochberg, E. Kuflik, and T. Volansky, J. High [41] V. G. Lucivero, N. D. McDonough, N. Dural, and M. V. Roma- Energy Phys. 2020 (1), 167. lis, Phys. Rev. A 96, 062702 (2017). [14] B. Julsgaard, A. Kozhekin, and E. S. Polzik, Nature (London) [42] N. A. Sinitsyn and Y. V. Pershin, Reports on Progress in Physics 413, 400 (2001). 79, 106501 (2016). [15] J. F. Sherson, H. Krauter, R. K. Olsson, B. Julsgaard, K. Ham- [43] G. E. Katsoprinakis, A. T. Dellis, and I. K. Kominis, Phys. Rev. merer, I. Cirac, and E. S. Polzik, Nature (London) 443, 557 A 75, 042502 (2007). (2006). [44] P. Glasenapp, N. A. Sinitsyn, L. Yang, D. G. Rickel, D. Roy, [16] K. Jensen, W. Wasilewski, H. Krauter, T. Fernholz, B. M. A. Greilich, M. Bayer, and S. A. Crooker, Phys. Rev. Lett. 113, Nielsen, M. Owari, M. B. Plenio, A. Serafini, M. M. Wolf, and 156601 (2014). E. S. Polzik, Nat. Phys. 7, 13 (2011). [45] V. G. Lucivero, R. Jiménez-Martínez, J. Kong, and M. W. [17] K. Hammerer, A. S. Sørensen, and E. S. Polzik, Rev. Mod. Mitchell, Phys. Rev. A 93, 053802 (2016). Phys. 82, 1041 (2010). [46] V. G. Lucivero, A. Dimic, J. Kong, R. Jiménez-Martínez, and [18] M. D. Eisaman, A. André, F. Massou, M. Fleischhauer, A. S. M. W. Mitchell, Phys. Rev. A 95, 041803 (2017). Zibrov, and M. D. Lukin, Nature (London) 438, 837 (2005). [47] D. S. Dean, Journal of Physics A: Mathematical and General [19] T. Peyronel, O. Firstenberg, Q.-Y. Liang, S. Hofferberth, A. V. 29, L613 (1996). Gorshkov, T. Pohl, M. D. Lukin, and V. Vuletic,´ Nature (Lon- [48] C. H. Volk, J. G. Mark, and B. C. Grover, Phys. Rev. A 20, 2381 don) 488, 57 (2012). (1979). [20] A. V. Gorshkov, J. Otterbach, M. Fleischhauer, T. Pohl, and [49] T. M. Kwon, J. G. Mark, and C. H. Volk, Phys. Rev. A 24, 1894 M. D. Lukin, Phys. Rev. Lett. 107, 133602 (2011). (1981). [21] J. Borregaard, M. Zugenmaier, J. M. Petersen, H. Shen, G. Vasi- [50] P. A. Heimann, I. A. Greenwood, and J. H. Simpson, Phys. Rev. lakis, K. Jensen, E. S. Polzik, and A. S. Sørensen, Nature Com- A 23, 1209 (1981). munications 7, 11356 (2016), arXiv:1501.03916 [quant-ph]. [51] Z. Wu, W. Happer, and J. M. Daniels, Phys. Rev. Lett. 59, 1480 [22] A. Kastler, J. Opt. Soc. Am. 47, 460 (1957). (1987). [23] F. Masnou-Seeuws and M.-A. Bouchiat, J. Phys. 28, 406 [52] E. B. Alexandrov, M. V. Balabas, D. Budker, D. English, D. F. (1967). Kimball, C. H. Li, and V. V. Yashchuk, Phys. Rev. A 66, 042903 [24] Z. Wu, S. Schaefer, G. D. Cates, and W. Happer, Phys. Rev. A (2002). 37, 1161 (1988). [53] M. T. Graf, D. F. Kimball, S. M. Rochester, K. Kerner, [25] S. Li, P. Vachaspati, D. Sheng, N. Dural, and M. V. Romalis, C. Wong, D. Budker, E. B. Alexandrov, M. V. Balabas, and Phys. Rev. A 84, 061403 (2011). V. V. Yashchuk, Phys. Rev. A 72, 023401 (2005). 12

[54] B. Driehuys, G. D. Cates, W. Happer, H. Mabuchi, B. Saam, (2016). M. S. Albert, and A. Wishnia, Phys. Lett. A 184, 88 (1993). [67] A. Hatakeyama, T. Kuroda, N. Sekiguchi, and K. Ishikawa, [55] B. Driehuys, G. D. Cates, and W. Happer, Phys. Rev. Lett. 74, Appl. Phys. B 125, 133 (2019). 4943 (1995). [68] W. Wasilewski, K. Jensen, H. Krauter, J. J. Renema, M. V. Bal- [56] S. R. Breeze, S. Lang, I. Moudrakovski, C. I. Ratcliffe, J. A. abas, and E. S. Polzik, Phys. Rev. Lett. 104, 133601 (2010). Ripmeester, B. Simard, and G. Santyr, J. Appl. Phys. 86, 4040 [69] O. Katz, O. Peleg, and O. Firstenberg, Phys. Rev. Lett. 115, (1999). 113003 (2015). [57] M. F. Hsu, G. D. Cates, I. Kominis, I. A. Aksay, and [70] C. Gemmel, W. Heil, S. Karpuk, K. Lenz, C. Ludwig, D. M. Dabbs, Applied Physics Letters 77, 2069 (2000), Y. Sobolev, K. Tullney, M. Burghoff, W. Kilian, S. Knappe- https://doi.org/10.1063/1.1312855. Grüneberg, W. Müller, A. Schnabel, F. Seifert, L. Trahms, and [58] For long-lived solutions of the diffusion equation, the flux to- S. Baeßler, Eur. Phys. J. D 57, 303 (2010). wards the wall nˆ · ∇ˆs is of the same order as ˆs/R. [71] Excess decay and noise due to the modes m, n ≥ 70 are intro- [59] This limit is obtained only when λ  R, which is also neces- duced along the lines of Eq. (S4) in Ref. [33]. sary for the validity of the diffusion equation. [72] Here in one dimension, we use the relation D = λv¯ between [60] M. N. Ozisik, Boundary value problems of heat conduction the diffusion coefficient and the kinetic parameters. In three di- (Courier Corporation, 2002). mensions, this becomes D = λv/¯ 3, such that cot (kL/2) = −1/N [61] T. Holstein and H. Primakoff, Phys. Rev. 58, 1098 (1940). 2 1+e λk. $ has similar dimensionality dependence. 3 1−e−1/N [62] C. Kittel, Quantum Theory of Solids (John Wiley & Sons, New [73] D. A. Steck, Quantum and Atom Optics (Department of York, 1987). Physics, University of Oregon, Eugene, Oregon, 2007). [63] Y. Tang, Y. Wen, L. Cai, and K. Zhao, Phys. Rev. A 101, 013821 [74] V. B. Braginsky and F. Y. Khalili, Rev. Mod. Phys. 68, 1 (1996). (2020). [75] S. Appelt, A. B. Baranga, C. J. Erickson, M. V. Romalis, A. R. [64] M. C. D. Tayler, T. F. Sjolander, A. e. Pines, and D. Budker, J. Young, and W. Happer, Phys. Rev. A 58, 1412 (1998). Magn. Reson. 270, 35 (2016). [76] V. Shah, G. Vasilakis, and M. V. Romalis, Phys. Rev. Lett. 104, [65] R. Pugatch, O. Firstenberg, M. Shuker, and N. Davidson, Phys. 013601 (2010). Rev. Lett. 102, 150602 (2009). [66] N. Sekiguchi and A. Hatakeyama, Appl. Phys. B 122, 81