Quantum thermodynamic derivation of the energy resolution limit in magnetometry

Iannis K. Kominis Department of Physics, University of Crete, Heraklion 70013, Greece
Abstract

It was recently demonstrated that a multitude of realizations of several magnetic sensing technologies satisfy the energy resolution limit, which connects a quantity composed by the variance of the magnetic field estimate, the sensor volume and the measurement time, and having units of action, with Planck-constant-over-2-pi\hbarroman_ℏ. A first-principles derivation of this limit is still elusive. We here present such a derivation based on quantum thermodynamic arguments. We show that the energy resolution limit is a result of quantum thermodynamic work necessarily associated with quantum measurement and Landauer erasure, the work being exchanged with the magnetic field. We apply these considerations to atomic magnetometers, diamond magnetometers, and SQUIDs, spanning an energy resolution limit from 100superscript100Planck-constant-over-2-pi10^{0}\hbar10 start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT roman_ℏ to 107superscript107Planck-constant-over-2-pi10^{7}\hbar10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_ℏ. This connection between quantum thermodynamics and magnetometry can help advance quantum sensing technologies towards even more sensitive devices.

Magnetic fields convey useful information in diverse physical settings, therefore it is no surprise that quantum sensing [1] of magnetic fields is one of the pillars of the second quantum revolution [2]. Whether classical or quantum, magnetometers are characterized by several figures of merit like bandwidth [3, 4, 5, 6], dynamic range [7, 8, 9, 10], sensor size and scalability [11, 12, 13, 14], accuracy [15], or the ability to operate in harsh environments [16, 17].

Magnetic sensitivity stands out as a prominent sensor characteristic, since the resolution of any measurement is limited by the intrinsic noise of the sensor. Advances in magnetic sensitivity brought about by superconducting sensors [18, 19] and optical pumping magnetometers [20, 21] spurred numerous applications, like sensing magnetic fields produced by the human brain [22, 23, 24, 25, 26] or heart [27, 28], materials characterization [29, 30], even table-top probes of new physics [31, 32, 33, 34, 35]. Therefore, understanding the fundamental limitations to magnetic sensitivity is crucial for pushing magnetic sensors towards optimal performance, and thus unraveling new applications.

By analyzing a large body of published work [36], it was demonstrated that tens of different realizations of several magnetic sensing technologies appear to have a unifying property, namely they all seem to satisfy the so-called energy resolution limit [37]. This limit states that (δB)2Vτ/2μ0greater-than-or-approximately-equalssuperscript𝛿𝐵2𝑉𝜏2subscript𝜇0Planck-constant-over-2-pi(\delta B)^{2}V\tau/2\mu_{0}\gtrapprox\hbar( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V italic_τ / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⪆ roman_ℏ, where (δB)2superscript𝛿𝐵2(\delta B)^{2}( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the variance of the magnetic field estimate, V𝑉Vitalic_V the sensor volume, τ𝜏\tauitalic_τ the measurement time, and μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT the magnetic permeability of vacuum. The left-hand side of this bound has units of action. The fact that the right-hand side roughly equals Planck-constant-over-2-pi\hbarroman_ℏ is aesthetically pleasing when discussing a fundamental limit. Surprisingly, as the authors in [36] pointed out, a first-principles derivation of the energy resolution limit (ERL) is still elusive.

The ERL contains the expression (δB)2V/2μ0superscript𝛿𝐵2𝑉2subscript𝜇0(\delta B)^{2}V/2\mu_{0}( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, reminiscent of the magnetic energy within the sensor volume V𝑉Vitalic_V. However, (δB)2/2μ0superscript𝛿𝐵22subscript𝜇0(\delta B)^{2}/2\mu_{0}( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is not the actual magnetic energy density, since it contains (δB)2superscript𝛿𝐵2(\delta B)^{2}( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT instead of B2superscript𝐵2B^{2}italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Various attempts [36] to derive the ERL based on quantum speed limits or energy-time uncertainty relations do not work, since instead of (δB)2superscript𝛿𝐵2(\delta B)^{2}( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, they involve the expression δ(B2)=2BδB𝛿superscript𝐵22𝐵𝛿𝐵\delta(B^{2})=2B\delta Bitalic_δ ( italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) = 2 italic_B italic_δ italic_B, further leading to the counterintuitive result that δB𝛿𝐵\delta Bitalic_δ italic_B is suppressed when increasing B𝐵Bitalic_B.

We will here derive the ERL based on quantum thermodynamic arguments. In particular, we will connect the ERL to the quantum work performed during quantum measurement and/or Landauer erasure of information. To this end, we treat the magnetic field as an integral part of the quantum thermodynamic environment of the sensor. When the quantum thermodynamic work accompanying the process of measurement is exchanged with the magnetic field energy, it leads to magnetic field fluctuations.

The field of quantum thermodynamics [38, 39, 40, 41, 42, 43, 44] has unified quantum information and quantum measurements with thermodynamic processes. The understanding of the physical nature of information [45, 46], and the energy cost of information erasure [47] greatly inspired the development of quantum information science. Few works, however, have so far considered the connection of quantum thermodynamics with quantum metrology [48, 49, 50, 51]. The current work falls in this direction.

The idea that the magnetic or electric field itself can act as a source/sink of quantum thermodynamic work is not new [48, 52]. We here push this idea further, towards understanding the ERL in magnetometry in a general way independent of the specific technology realization. The crux of the matter is the following. Let uB=B2/2μ0subscript𝑢𝐵superscript𝐵22subscript𝜇0u_{B}=B^{2}/2\mu_{0}italic_u start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT = italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT be the magnetic energy density. If the field fluctuates by δB𝛿𝐵\delta Bitalic_δ italic_B, where δB=0expectation𝛿𝐵0\braket{\delta B}=0⟨ start_ARG italic_δ italic_B end_ARG ⟩ = 0, then the field energy within the volume V𝑉Vitalic_V of the sensor will change by VuB+δBuB=(δB)2V/2μ0𝑉expectationsubscript𝑢𝐵𝛿𝐵subscript𝑢𝐵superscript𝛿𝐵2𝑉2subscript𝜇0V\braket{u_{B+\delta B}-u_{B}}=(\delta B)^{2}V/2\mu_{0}italic_V ⟨ start_ARG italic_u start_POSTSUBSCRIPT italic_B + italic_δ italic_B end_POSTSUBSCRIPT - italic_u start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG ⟩ = ( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. If the cause of this fluctuation is the exchange of work W𝑊Witalic_W between sensor and field, it will be W=VuB+δBuB𝑊𝑉expectationsubscript𝑢𝐵𝛿𝐵subscript𝑢𝐵W=V\braket{u_{B+\delta B}-u_{B}}italic_W = italic_V ⟨ start_ARG italic_u start_POSTSUBSCRIPT italic_B + italic_δ italic_B end_POSTSUBSCRIPT - italic_u start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT end_ARG ⟩, and the corresponding field fluctuation will be δB=2μ0W/V𝛿𝐵2subscript𝜇0𝑊𝑉\delta B=\sqrt{2\mu_{0}W/V}italic_δ italic_B = square-root start_ARG 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_W / italic_V end_ARG. Finally, using the relation W=(δB)2V/2μ0𝑊superscript𝛿𝐵2𝑉2subscript𝜇0W=(\delta B)^{2}V/2\mu_{0}italic_W = ( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and quantum speed limits to connect the exchanged work W𝑊Witalic_W with the time during which the exchange takes place, we will arrive at the ERL.

To proceed formally and find W𝑊Witalic_W we use the approach of [53], which establishes the minimum energy cost of measurement and Landauer erasure. The authors consider a system 𝒮𝒮{\cal S}caligraphic_S, a meter {\cal M}caligraphic_M, and a thermal bath {\cal B}caligraphic_B. A measurement is performed on 𝒮𝒮{\cal S}caligraphic_S, meaning that 𝒮𝒮{\cal S}caligraphic_S and {\cal M}caligraphic_M become entangled. Then follows a projective measurement on {\cal M}caligraphic_M. Finally, to make the process cyclic, the information in 𝒮+𝒮{\cal S+M}caligraphic_S + caligraphic_M is erased. The authors show that the combined work cost (done by {\cal M}caligraphic_M on {\cal B}caligraphic_B) of measurement and erasure is bounded below:

Wmeas+WeraskBT,subscript𝑊meassubscript𝑊erassubscript𝑘𝐵𝑇W_{\rm meas}+W_{\rm eras}\geq k_{B}T{\cal I},italic_W start_POSTSUBSCRIPT roman_meas end_POSTSUBSCRIPT + italic_W start_POSTSUBSCRIPT roman_eras end_POSTSUBSCRIPT ≥ italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T caligraphic_I , (1)

where T𝑇Titalic_T is the bath temperature, with which the meter is in thermal contact, and {\cal I}caligraphic_I the mutual information between 𝒮𝒮{\cal S}caligraphic_S and {\cal M}caligraphic_M. This information satisfies 0H0𝐻0\leq{\cal I}\leq H0 ≤ caligraphic_I ≤ italic_H, where H𝐻Hitalic_H is the Shannon entropy of the possible measurement results. If this work is exchanged between sensor and magnetic field, it will be (δB)2V/2μ0kBTsuperscript𝛿𝐵2𝑉2subscript𝜇0subscript𝑘𝐵𝑇(\delta B)^{2}V/2\mu_{0}\geq k_{B}T{\cal I}( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≥ italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T caligraphic_I. If the exchange takes place during time τ𝜏\tauitalic_τ, we can use the Margolus-Levitin quantum speed limit [54, 55], kBTτπ2subscript𝑘𝐵𝑇𝜏𝜋2Planck-constant-over-2-pik_{B}T{\cal I}\tau\geq{\pi\over 2}\hbaritalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T caligraphic_I italic_τ ≥ divide start_ARG italic_π end_ARG start_ARG 2 end_ARG roman_ℏ, and finally we arrive at the ERL:

(δB)2Vτ2μ0π2superscript𝛿𝐵2𝑉𝜏2subscript𝜇0𝜋2Planck-constant-over-2-pi{{(\delta B)^{2}V\tau}\over{2\mu_{0}}}\geq{\pi\over 2}\hbardivide start_ARG ( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V italic_τ end_ARG start_ARG 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ≥ divide start_ARG italic_π end_ARG start_ARG 2 end_ARG roman_ℏ (2)

To apply the speed limit for the exchanged work the magnetic field is treated not as an external parameter, but as a physical degree of freedom. This can be done by e.g. quantizing the field (like in [36]), and considering transitions moving energy kBTsubscript𝑘𝐵𝑇k_{B}T{\cal I}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T caligraphic_I from states where this energy is localized in the sensor degrees of freedom, to orthogonal states where it is localized in the field.

We will now specify the above for the case of optical pumping magnetometers, so the system 𝒮𝒮{\cal S}caligraphic_S doing the sensing is an atomic vapor of spin-1/2 atoms (hyperfine structure is not relevant in this discussion). In the standard magnetometry framework the atomic spins are first spin-polarized by an optical pumping pulse, then precess due to the magnetic field, and finally are probed e.g. by a light beam. The magnetic field B𝐵Bitalic_B points along the z𝑧zitalic_z-axis, and the eigenstates of σzsubscript𝜎𝑧\sigma_{z}italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT are the computational basis states |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ and |1ket1\ket{1}| start_ARG 1 end_ARG ⟩. The atoms are spin-polarized along the x𝑥xitalic_x-axis, their initial state being |ψ0=(|0+|1)/2ketsubscript𝜓0ket0ket12\ket{\psi_{0}}=(\ket{0}+\ket{1})/\sqrt{2}| start_ARG italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ = ( | start_ARG 0 end_ARG ⟩ + | start_ARG 1 end_ARG ⟩ ) / square-root start_ARG 2 end_ARG. The magnetic field evolves |ψ0ketsubscript𝜓0\ket{\psi_{0}}| start_ARG italic_ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ into |ψχ=(|0+eiχ|1)/2ketsubscript𝜓𝜒ket0superscript𝑒𝑖𝜒ket12\ket{\psi_{\chi}}=(\ket{0}+e^{i\chi}\ket{1})/\sqrt{2}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT end_ARG ⟩ = ( | start_ARG 0 end_ARG ⟩ + italic_e start_POSTSUPERSCRIPT italic_i italic_χ end_POSTSUPERSCRIPT | start_ARG 1 end_ARG ⟩ ) / square-root start_ARG 2 end_ARG, where χBproportional-to𝜒𝐵\chi\propto Bitalic_χ ∝ italic_B. A measurement in the eigenbasis of σxsubscript𝜎𝑥\sigma_{x}italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT conveys information about B𝐵Bitalic_B.

However, in our analysis, as measurement in the quantum thermodynamic context of [53] we consider the dephasing produced by atomic collisions. In particular, in the SERF regime of interest here, spin destruction (SD) collisions [56] are the fundamental mechanism for spin decoherence. In a binary SD collision one atom is the system 𝒮𝒮{\cal S}caligraphic_S and another atom is the meter {\cal M}caligraphic_M. During the collision, the two atomic spins become entangled, and further SD collisions act as a projective measurement on the meter atom, along the same lines described in [57] for spin-exchange collisions. Thus, SD collisions can be seen as performing an unobserved measurement of |ψχketsubscript𝜓𝜒\ket{\psi_{\chi}}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT end_ARG ⟩ in the computational basis, pushing |ψχketsubscript𝜓𝜒\ket{\psi_{\chi}}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT end_ARG ⟩ towards ρ=12|00|+12|11|𝜌12ket0bra012ket1bra1\rho={1\over 2}\ket{0}\bra{0}+{1\over 2}\ket{1}\bra{1}italic_ρ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG | start_ARG 0 end_ARG ⟩ ⟨ start_ARG 0 end_ARG | + divide start_ARG 1 end_ARG start_ARG 2 end_ARG | start_ARG 1 end_ARG ⟩ ⟨ start_ARG 1 end_ARG |. The measurement time τ𝜏\tauitalic_τ is the duration of this process, i.e. the SD spin-relaxation time.

Information erasure (e.g. by an optical pumping pulse) renders the whole process cyclic. Nevertheless, information erasure does not incur an energy cost, since the optical pumping photons are scattered into the light field having practically zero temperature [58]. Hence in our case, the thermodynamic work entering (1) is solely due to the aforementioned measurement.

The Shannon entropy of the decohered state ρ𝜌\rhoitalic_ρ is H=ln2𝐻2H=\ln 2italic_H = roman_ln 2. We still need to determine the mutual information {\cal I}caligraphic_I. When a system atom collides with a meter atom, their spins become entangled. The entanglement is not maximal, because the atom’s spin phase change in SD collisions is small [59], unlike the case of spin-exchange collisions [60]. Hence it might appear that Hmuch-less-than𝐻{\cal I}\ll Hcaligraphic_I ≪ italic_H. However, after many such binary collisions, the system state is fully decohered. We consider all those collisions as one process mapping |ψχketsubscript𝜓𝜒\ket{\psi_{\chi}}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT end_ARG ⟩ into ρ𝜌\rhoitalic_ρ along the measurement time τ𝜏\tauitalic_τ, during which the total mutual information =H𝐻{\cal I}=Hcaligraphic_I = italic_H. It is interesting to note that this information of 1 bit pertains to the whole vapor, as individual atoms cannot be addressed. Experiments with hot vapors measure the ensemble spin, and (apart from relaxation effects attributed to binary collisions) the measurements reflect single-atom information, the signals being amplified by N𝑁Nitalic_N, the number of atoms.

To apply the bound kBTτπ2subscript𝑘𝐵𝑇𝜏𝜋2Planck-constant-over-2-pik_{B}T{\cal I}\tau\geq{\pi\over 2}\hbaritalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T caligraphic_I italic_τ ≥ divide start_ARG italic_π end_ARG start_ARG 2 end_ARG roman_ℏ, we crucially note that the temperature T𝑇Titalic_T should be the spin temperature Tssubscript𝑇𝑠T_{s}italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. While the translational degrees of freedom are indeed governed by the thermodynamic temperature, which is the room temperature or higher, the spin temperature can be widely different. It is known that spin-exchange collisions lead to a spin-temperature distribution [60, 61]. Moreover, at low spin polarization pertinent to understanding sensor noise, spin-destruction collisions obey the same spin dynamics as spin-exchange collisions, and thus also lead to a spin-temperature distribution [60, 61]. Without optical pumping, the spin-temperature at equilibrium is infinite (diagonal density matrix). However, it is known from spin-noise spectroscopy [62, 63, 64, 65, 66] that there are spontaneous spin-polarization fluctuations of order 1/N1𝑁1/\sqrt{N}1 / square-root start_ARG italic_N end_ARG around this equilibrium, hence Tssubscript𝑇𝑠T_{s}italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is finite.

To find Tssubscript𝑇𝑠T_{s}italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT when the magnetic field is B𝐵Bitalic_B we relate μB/2kBTs𝜇𝐵2subscript𝑘𝐵subscript𝑇𝑠\mu B/2k_{B}T_{s}italic_μ italic_B / 2 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, where μ𝜇\muitalic_μ the atom’s magnetic moment, with the spin-noise polarization of the vapor along the z𝑧zitalic_z axis, which is on the order of 1/N1𝑁1/\sqrt{N}1 / square-root start_ARG italic_N end_ARG. Consider the spin-temperature density matrix ρ=eH/kBTs/Tr{eH/kBTs}𝜌superscript𝑒𝐻subscript𝑘𝐵subscript𝑇𝑠Trsuperscript𝑒𝐻subscript𝑘𝐵subscript𝑇𝑠\rho=e^{-H/k_{B}T_{s}}/{\rm Tr}\{e^{-H/k_{B}T_{s}}\}italic_ρ = italic_e start_POSTSUPERSCRIPT - italic_H / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT / roman_Tr { italic_e start_POSTSUPERSCRIPT - italic_H / italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT end_POSTSUPERSCRIPT }, where the magnetic Hamiltonian is H=μB2σz𝐻𝜇𝐵2subscript𝜎𝑧H=-{{\mu B}\over 2}\sigma_{z}italic_H = - divide start_ARG italic_μ italic_B end_ARG start_ARG 2 end_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. Setting 2kBTs=μNB2subscript𝑘𝐵subscript𝑇𝑠𝜇𝑁𝐵2k_{B}T_{s}=\mu\sqrt{N}B2 italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_μ square-root start_ARG italic_N end_ARG italic_B, and expanding ρ12(1+σz/N)𝜌121subscript𝜎𝑧𝑁\rho\approx{1\over 2}(1+\sigma_{z}/\sqrt{N})italic_ρ ≈ divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 + italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT / square-root start_ARG italic_N end_ARG ), it follows that indeed σz=Tr{ρσz}1/Nexpectationsubscript𝜎𝑧Tr𝜌subscript𝜎𝑧1𝑁\braket{\sigma_{z}}={\rm Tr}\{\rho\sigma_{z}\}\approx 1/\sqrt{N}⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG ⟩ = roman_Tr { italic_ρ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT } ≈ 1 / square-root start_ARG italic_N end_ARG.

This point is further substantiated: an atom in the state |ψχketsubscript𝜓𝜒\ket{\psi_{\chi}}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT end_ARG ⟩ has zero energy, given that Hσzproportional-to𝐻subscript𝜎𝑧H\propto\sigma_{z}italic_H ∝ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT. When the atom’s state is finally projected to either |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ or |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ by SD collisions, the field provides the atom with energy ±μB/2plus-or-minus𝜇𝐵2\pm\mu B/2± italic_μ italic_B / 2. Projection to either |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ or |1ket1\ket{1}| start_ARG 1 end_ARG ⟩ has probability 1/2, thus the total energy exchanged with the field is on average zero, apart from an uncertainty μBN/2𝜇𝐵𝑁2\mu B\sqrt{N}/2italic_μ italic_B square-root start_ARG italic_N end_ARG / 2 (binomial distribution with equal probabilities). This is the same quantity derived previously as kBTssubscript𝑘𝐵subscript𝑇𝑠k_{B}T_{s}italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT.

Finally, this result can be generalized by considering the time-dependent relaxation dynamics of the vapor, starting from the fully spin-polarized state |ψχ=(|0+eiχ|1)/2ketsubscript𝜓𝜒ket0superscript𝑒𝑖𝜒ket12\ket{\psi_{\chi}}=(\ket{0}+e^{i\chi}\ket{1})/\sqrt{2}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT end_ARG ⟩ = ( | start_ARG 0 end_ARG ⟩ + italic_e start_POSTSUPERSCRIPT italic_i italic_χ end_POSTSUPERSCRIPT | start_ARG 1 end_ARG ⟩ ) / square-root start_ARG 2 end_ARG. For this state, the spin polarization along 𝐳^^𝐳\mathbf{\hat{z}}over^ start_ARG bold_z end_ARG, given by σzexpectationsubscript𝜎𝑧\braket{\sigma_{z}}⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG ⟩, is exactly zero. During the relaxation transient, we can write dσzt=(1|σ+t|)dξ𝑑subscriptexpectationsubscript𝜎𝑧𝑡1subscriptexpectationsubscript𝜎𝑡𝑑𝜉d\braket{\sigma_{z}}_{t}=(1-|\braket{\sigma_{+}}_{t}|)d\xiitalic_d ⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = ( 1 - | ⟨ start_ARG italic_σ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT end_ARG ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT | ) italic_d italic_ξ, where dξ𝑑𝜉d\xiitalic_d italic_ξ is a Gaussian noise process of variance 1/Nτ1𝑁𝜏1/N\tau1 / italic_N italic_τ, and |σ+|=et/τsubscript𝜎superscript𝑒𝑡𝜏|\sigma_{+}|=e^{-t/\tau}| italic_σ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT | = italic_e start_POSTSUPERSCRIPT - italic_t / italic_τ end_POSTSUPERSCRIPT is the transverse spin decaying with time constant τ𝜏\tauitalic_τ. That is, the spin polarization along 𝐳^^𝐳\mathbf{\hat{z}}over^ start_ARG bold_z end_ARG is described by a nonlinear noise term, reflecting its dependence on the spin state along the transient, during which atoms are gradually projected from the sate |ψχketsubscript𝜓𝜒\ket{\psi_{\chi}}| start_ARG italic_ψ start_POSTSUBSCRIPT italic_χ end_POSTSUBSCRIPT end_ARG ⟩ to |0ket0\ket{0}| start_ARG 0 end_ARG ⟩ or |1ket1\ket{1}| start_ARG 1 end_ARG ⟩. After |σ+|subscript𝜎|\sigma_{+}|| italic_σ start_POSTSUBSCRIPT + end_POSTSUBSCRIPT | has decayed to zero, SD collisions still drive spin fluctuations along 𝐳^^𝐳\mathbf{\hat{z}}over^ start_ARG bold_z end_ARG, thus dσztτ=dξ𝑑subscriptexpectationsubscript𝜎𝑧much-greater-than𝑡𝜏𝑑𝜉d\braket{\sigma_{z}}_{t\gg\tau}=d\xiitalic_d ⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG ⟩ start_POSTSUBSCRIPT italic_t ≫ italic_τ end_POSTSUBSCRIPT = italic_d italic_ξ. The average variance of σztsubscriptexpectationsubscript𝜎𝑧𝑡\braket{\sigma_{z}}_{t}⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG ⟩ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT during the relaxation time τ𝜏\tauitalic_τ is given by 1Nτ0τ(1et/τ)2𝑑t1𝑁𝜏superscriptsubscript0𝜏superscript1superscript𝑒𝑡𝜏2differential-d𝑡{1\over{N\tau}}\int_{0}^{\tau}(1-e^{-t/\tau})^{2}dtdivide start_ARG 1 end_ARG start_ARG italic_N italic_τ end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_τ end_POSTSUPERSCRIPT ( 1 - italic_e start_POSTSUPERSCRIPT - italic_t / italic_τ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_d italic_t, which translates to an average uncertainty δσz0.5/N𝛿expectationsubscript𝜎𝑧0.5𝑁\delta\braket{\sigma_{z}}\approx 0.5/\sqrt{N}italic_δ ⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG ⟩ ≈ 0.5 / square-root start_ARG italic_N end_ARG. In other words, it is the late time noise that dominates the result, which apart from the factor of 0.5, was found in the main text by considering the equilibrium spin-temperature state instead of the relaxation transient. It is noted that fluctuations in σzexpectationsubscript𝜎𝑧\braket{\sigma_{z}}⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG ⟩ due to external probing of the spins are in addition to those considered here, which are due to the “internal” measurements due to SD collisions. An optimized probing, however, would be designed so as to not significantly decrease the relaxation time τ𝜏\tauitalic_τ, so our result is approximately valid even in this more general case. Overall,

kBTs=μNBsubscript𝑘𝐵subscript𝑇𝑠𝜇𝑁𝐵k_{B}T_{s}=\mu\sqrt{N}Bitalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_μ square-root start_ARG italic_N end_ARG italic_B (3)

To proceed, we set BδB𝐵𝛿𝐵B\approx\delta Bitalic_B ≈ italic_δ italic_B, i.e. work at small magnetic fields close to the sought after intrinsic sensor noise. Parenthetically, it would be interesting to investigate the ERL at higher magnetic fields. There we anticipate rich phenomenology, since higher magnetic fields produced by degrees of freedom external to the sensor (e.g. a current-carrying coil) might open new channels for the exchange of thermodynamic work between the sensor and its environment. Thus the relation (δB)2V/2μ0kBTsuperscript𝛿𝐵2𝑉2subscript𝜇0subscript𝑘𝐵𝑇(\delta B)^{2}V/2\mu_{0}\geq k_{B}T{\cal I}( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≥ italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T caligraphic_I, which formally leads to δBBproportional-to𝛿𝐵𝐵\delta B\propto\sqrt{B}italic_δ italic_B ∝ square-root start_ARG italic_B end_ARG in this discussion of atomic sensors, cannot be simply extrapolated to arbitrarily high fields.

Using the bound kBTsτπ/2subscript𝑘𝐵subscript𝑇𝑠𝜏𝜋Planck-constant-over-2-pi2k_{B}T_{s}{\cal I}\tau\geq\pi\hbar/2italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT caligraphic_I italic_τ ≥ italic_π roman_ℏ / 2, we find δB(π/2ln(2))(/μNτ)𝛿𝐵𝜋22Planck-constant-over-2-pi𝜇𝑁𝜏\delta B\geq(\pi/2\ln(2))(\hbar/\mu\sqrt{N}\tau)italic_δ italic_B ≥ ( italic_π / 2 roman_ln ( 2 ) ) ( roman_ℏ / italic_μ square-root start_ARG italic_N end_ARG italic_τ ). The relaxation time due to SD collisions is given by 1/τ=nσsdv¯1𝜏𝑛subscript𝜎sd¯𝑣1/\tau=n\sigma_{\rm sd}\overline{v}1 / italic_τ = italic_n italic_σ start_POSTSUBSCRIPT roman_sd end_POSTSUBSCRIPT over¯ start_ARG italic_v end_ARG, where is n=N/V𝑛𝑁𝑉n=N/Vitalic_n = italic_N / italic_V the atom number density, σsdsubscript𝜎sd\sigma_{\rm sd}italic_σ start_POSTSUBSCRIPT roman_sd end_POSTSUBSCRIPT the SD cross section, and v¯¯𝑣\overline{v}over¯ start_ARG italic_v end_ARG the relative velocity of the colliding partners. Hence

δBπ2ln2σsdv¯NμV𝛿𝐵𝜋22Planck-constant-over-2-pisubscript𝜎sd¯𝑣𝑁𝜇𝑉\delta B\geq{\pi\over{2\ln 2}}{{\hbar\sigma_{\rm sd}\overline{v}\sqrt{N}}\over% {\mu V}}italic_δ italic_B ≥ divide start_ARG italic_π end_ARG start_ARG 2 roman_ln 2 end_ARG divide start_ARG roman_ℏ italic_σ start_POSTSUBSCRIPT roman_sd end_POSTSUBSCRIPT over¯ start_ARG italic_v end_ARG square-root start_ARG italic_N end_ARG end_ARG start_ARG italic_μ italic_V end_ARG (4)

Using (4) we can finally obtain the ERL, which reads

(δB)2Vτ2μ0(π28(ln2)2σsdv¯μ0μ2)superscript𝛿𝐵2𝑉𝜏2subscript𝜇0Planck-constant-over-2-pisuperscript𝜋28superscript22Planck-constant-over-2-pisubscript𝜎sd¯𝑣subscript𝜇0superscript𝜇2{{(\delta B)^{2}V\tau}\over{2\mu_{0}}}\geq\hbar\Big{(}{{\pi^{2}}\over{8(\ln 2)% ^{2}}}{{\hbar\sigma_{\rm sd}\overline{v}}\over{\mu_{0}\mu^{2}}}\Big{)}divide start_ARG ( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V italic_τ end_ARG start_ARG 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ≥ roman_ℏ ( divide start_ARG italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 8 ( roman_ln 2 ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG roman_ℏ italic_σ start_POSTSUBSCRIPT roman_sd end_POSTSUBSCRIPT over¯ start_ARG italic_v end_ARG end_ARG start_ARG italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) (5)

The atom’s magnetic moment is μ=μB/q𝜇subscript𝜇𝐵𝑞\mu=\mu_{B}/qitalic_μ = italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT / italic_q, where μBsubscript𝜇𝐵\mu_{B}italic_μ start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT is the Bohr magneton and q=[(S(S+1)+I(I+1)]/S(S+1)q=[(S(S+1)+I(I+1)]/S(S+1)italic_q = [ ( italic_S ( italic_S + 1 ) + italic_I ( italic_I + 1 ) ] / italic_S ( italic_S + 1 ) is the nuclear slowing down factor with S=1/2𝑆12S=1/2italic_S = 1 / 2 and I𝐼Iitalic_I the nuclear spin [56]. Using the SD cross sections [56] and the respective relative velocities v¯¯𝑣\overline{v}over¯ start_ARG italic_v end_ARG, a number density n=1014cm3𝑛superscript1014superscriptcm3n=10^{14}~{}{\rm cm^{-3}}italic_n = 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and volume V=10cm3𝑉10superscriptcm3V=10~{}{\rm cm^{3}}italic_V = 10 roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, we list in Table I the values of δB𝛿𝐵\delta Bitalic_δ italic_B of (4) and the respective ERL of (5) for K41superscriptK41{}^{41}{\rm K}start_FLOATSUPERSCRIPT 41 end_FLOATSUPERSCRIPT roman_K, Rb87superscriptRb87{}^{87}{\rm Rb}start_FLOATSUPERSCRIPT 87 end_FLOATSUPERSCRIPT roman_Rb and Cs133superscriptCs133{}^{133}{\rm Cs}start_FLOATSUPERSCRIPT 133 end_FLOATSUPERSCRIPT roman_Cs.

Table 1: Atomic Magnetometer ERL. The numbers reflect the magnetic field noise and corresponding ERL for an atomic vapor of atom number density 1014cm3superscript1014superscriptcm310^{14}~{}{\rm cm^{-3}}10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT and volume 10cm310superscriptcm310~{}{\rm cm^{3}}10 roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT.
Atom K41superscriptK41{}^{41}{\rm K}start_FLOATSUPERSCRIPT 41 end_FLOATSUPERSCRIPT roman_K Rb87superscriptRb87{}^{87}{\rm Rb}start_FLOATSUPERSCRIPT 87 end_FLOATSUPERSCRIPT roman_Rb Cs133superscriptCs133{}^{133}{\rm Cs}start_FLOATSUPERSCRIPT 133 end_FLOATSUPERSCRIPT roman_Cs
δB𝛿𝐵\delta Bitalic_δ italic_B (1017Tsuperscript1017T10^{-17}~{}{\rm T}10 start_POSTSUPERSCRIPT - 17 end_POSTSUPERSCRIPT roman_T) 2 10 1000
Predicted ERL (Planck-constant-over-2-pi\hbarroman_ℏ) 4 25 6054

For a comparison, the authors in [67] used a cesium vapor of volume V=1cm3𝑉1superscriptcm3V=1~{}{\rm cm^{3}}italic_V = 1 roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT and number density 2×1013cm32superscript1013superscriptcm32\times 10^{13}~{}{\rm cm^{-3}}2 × 10 start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT, which by (4) lead to δB1014T𝛿𝐵superscript1014T\delta B\approx 10^{-14}~{}{\rm T}italic_δ italic_B ≈ 10 start_POSTSUPERSCRIPT - 14 end_POSTSUPERSCRIPT roman_T. Since this noise is distributed within the bandwidth 1/τ1𝜏1/\tau1 / italic_τ, the corresponding spectral density is 10pG/Hz10pGHz10~{}{\rm pG/\sqrt{Hz}}10 roman_pG / square-root start_ARG roman_Hz end_ARG, whereas the authors measured a noise level of 400pG/Hz400pGHz400~{}{\rm pG/\sqrt{Hz}}400 roman_pG / square-root start_ARG roman_Hz end_ARG and, under the premise of a theoretical optimization, project noise 2pG/Hz2pGHz2~{}{\rm pG/\sqrt{Hz}}2 roman_pG / square-root start_ARG roman_Hz end_ARG.

We will now elaborate further on the bound (4), rewriting it as

δB23κμ0μNV,𝛿𝐵23𝜅subscript𝜇0𝜇𝑁𝑉\delta B\geq{2\over 3}{\kappa\mu_{0}\mu\sqrt{N}\over V},italic_δ italic_B ≥ divide start_ARG 2 end_ARG start_ARG 3 end_ARG divide start_ARG italic_κ italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_μ square-root start_ARG italic_N end_ARG end_ARG start_ARG italic_V end_ARG , (6)

in order to exhibit the magnetic field produced within a spherical magnetized volume of a randomly spin-polarized vapor. Only now, this dipolar field is seen to be amplified by the factor κ=(3π/4ln2)σsdv¯/μ0μ2𝜅3𝜋42Planck-constant-over-2-pisubscript𝜎sd¯𝑣subscript𝜇0superscript𝜇2\kappa=({{3\pi}/{4\ln 2}}){{\hbar\sigma_{\rm sd}\overline{v}}/{\mu_{0}\mu^{2}}}italic_κ = ( 3 italic_π / 4 roman_ln 2 ) roman_ℏ italic_σ start_POSTSUBSCRIPT roman_sd end_POSTSUBSCRIPT over¯ start_ARG italic_v end_ARG / italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. On the one hand we can argue, like in [36], that this amplification is forced by the uncertainty relation. Given that the uncertainty ΔσxΔsubscript𝜎𝑥\Delta\sigma_{x}roman_Δ italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT is negligible for a vapor optically pumped along the x𝑥xitalic_x-axis, the uncertainty ΔσyΔsubscript𝜎𝑦\Delta\sigma_{y}roman_Δ italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT will be determined by the field δB𝛿𝐵\delta Bitalic_δ italic_B driving spin precession from the x𝑥xitalic_x axis to the y𝑦yitalic_y-axis, i.e. Δσy=δϕσxΔsubscript𝜎𝑦𝛿italic-ϕexpectationsubscript𝜎𝑥\Delta\sigma_{y}=\delta\phi\braket{\sigma_{x}}roman_Δ italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT = italic_δ italic_ϕ ⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_ARG ⟩ where δϕ=μδBτ/𝛿italic-ϕ𝜇𝛿𝐵𝜏Planck-constant-over-2-pi\delta\phi=\mu\delta B\tau/\hbaritalic_δ italic_ϕ = italic_μ italic_δ italic_B italic_τ / roman_ℏ. Since ΔσyΔsubscript𝜎𝑦\Delta\sigma_{y}roman_Δ italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT and ΔσzΔsubscript𝜎𝑧\Delta\sigma_{z}roman_Δ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT have to satisfy the uncertainty relation ΔσyΔσz|σx|Δsubscript𝜎𝑦Δsubscript𝜎𝑧expectationsubscript𝜎𝑥\Delta\sigma_{y}\Delta\sigma_{z}\geq|\braket{\sigma_{x}}|roman_Δ italic_σ start_POSTSUBSCRIPT italic_y end_POSTSUBSCRIPT roman_Δ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT ≥ | ⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_x end_POSTSUBSCRIPT end_ARG ⟩ |, it follows by setting Δσz=1/NΔsubscript𝜎𝑧1𝑁\Delta\sigma_{z}=1/\sqrt{N}roman_Δ italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT = 1 / square-root start_ARG italic_N end_ARG that δB𝛿𝐵\delta Bitalic_δ italic_B is given by (6).

There is yet another interpretation for the factor κ𝜅\kappaitalic_κ, interesting in its own right. We can consider the vacuum permeability μ0subscript𝜇0\mu_{0}italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT replaced by κμ0𝜅subscript𝜇0\kappa\mu_{0}italic_κ italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, as if the spins have a tendency to align, and κ𝜅\kappaitalic_κ represents the relative permeability constant. But indeed, binary collisions do correlate atoms, as was shown in [57] for spin-exchange collisions. It was shown that the negativity of the two-atom spin state scales as sinϕse2superscriptsubscriptitalic-ϕse2\sin\phi_{\rm se}^{2}roman_sin italic_ϕ start_POSTSUBSCRIPT roman_se end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, where ϕse1subscriptitalic-ϕse1\phi_{\rm se}\approx 1italic_ϕ start_POSTSUBSCRIPT roman_se end_POSTSUBSCRIPT ≈ 1 is the spin-exchange phase change. Spin-destruction collisions incur a spin phase change ϕ1much-less-thanitalic-ϕ1\phi\ll 1italic_ϕ ≪ 1. Since at low spin polarization pertinent to the study of sensitivity limits (see also discussion in End Matter) the dynamics of spin-exchange and spin-destruction are the same [61], we can write that the entanglement between two atoms after an SD collision will scale like sinϕ2ϕ2superscriptitalic-ϕ2superscriptitalic-ϕ2\sin\phi^{2}\approx\phi^{2}roman_sin italic_ϕ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≈ italic_ϕ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, since now ϕ1much-less-thanitalic-ϕ1\phi\ll 1italic_ϕ ≪ 1. Hence any two colliding partners will share a small correlation. The upside is that this correlation can be shared by many more atoms. Imagine a ”bath” atom experiencing consecutive collisions with many ”system” atoms. These will all be slightly correlated, and will reside in a ”correlation volume” Vcsubscript𝑉𝑐V_{c}italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Consider two spins inside this volume, which are about to collide. Their interaction energy is ϵ=κμ0μ2/V~italic-ϵ𝜅subscript𝜇0superscript𝜇2~𝑉\epsilon=\kappa\mu_{0}\mu^{2}/\tilde{V}italic_ϵ = italic_κ italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / over~ start_ARG italic_V end_ARG, where V~~𝑉\tilde{V}over~ start_ARG italic_V end_ARG is the volume defined by the two spins. This interaction reorients the spins in a time scale /ϵPlanck-constant-over-2-piitalic-ϵ\hbar/\epsilonroman_ℏ / italic_ϵ. For the aforementioned correlation to be maintained, this time should be equal or larger than V~/σsdv¯~𝑉subscript𝜎sd¯𝑣\tilde{V}/\sigma_{\rm sd}\overline{v}over~ start_ARG italic_V end_ARG / italic_σ start_POSTSUBSCRIPT roman_sd end_POSTSUBSCRIPT over¯ start_ARG italic_v end_ARG, which is the time required for their collision to correlate them. From this requirement we obtain again κ=σsdv¯/μ0μ2𝜅Planck-constant-over-2-pisubscript𝜎sd¯𝑣subscript𝜇0superscript𝜇2\kappa=\hbar\sigma_{\rm sd}\overline{v}/\mu_{0}\mu^{2}italic_κ = roman_ℏ italic_σ start_POSTSUBSCRIPT roman_sd end_POSTSUBSCRIPT over¯ start_ARG italic_v end_ARG / italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. We can also estimate the correlation volume Vcsubscript𝑉𝑐V_{c}italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, which will contain Ncsubscript𝑁𝑐N_{c}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT atoms. We can write κ=Ncϕ𝜅subscript𝑁𝑐italic-ϕ\kappa=\sqrt{N_{c}}\phiitalic_κ = square-root start_ARG italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG italic_ϕ, since the spin variance scales as ϕ2superscriptitalic-ϕ2\phi^{2}italic_ϕ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and thus the uncertainty as ϕitalic-ϕ\phiitalic_ϕ, while for ϕ1italic-ϕ1\phi\approx 1italic_ϕ ≈ 1 we would get an enhancement κ=Nc𝜅subscript𝑁𝑐\kappa=\sqrt{N_{c}}italic_κ = square-root start_ARG italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG, retrieving the case of fully polarized spins. The phase ϕitalic-ϕ\phiitalic_ϕ is related to the SD relaxation time and the collision rate 1/T1𝑇1/T1 / italic_T, i.e. it is 1/τ=ϕ2/T1𝜏superscriptitalic-ϕ2𝑇1/\tau=\phi^{2}/T1 / italic_τ = italic_ϕ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_T as discussed in [60]. The collision rate is 1/T=v¯/n1/31𝑇¯𝑣superscript𝑛131/T=\overline{v}/n^{-1/3}1 / italic_T = over¯ start_ARG italic_v end_ARG / italic_n start_POSTSUPERSCRIPT - 1 / 3 end_POSTSUPERSCRIPT. Putting everything together we find Nc=(v¯/μ0μ2)2σsdn2/3subscript𝑁𝑐superscriptPlanck-constant-over-2-pi¯𝑣subscript𝜇0superscript𝜇22subscript𝜎sdsuperscript𝑛23N_{c}=(\hbar\overline{v}/\mu_{0}\mu^{2})^{2}\sigma_{\rm sd}n^{-2/3}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = ( roman_ℏ over¯ start_ARG italic_v end_ARG / italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_μ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT roman_sd end_POSTSUBSCRIPT italic_n start_POSTSUPERSCRIPT - 2 / 3 end_POSTSUPERSCRIPT. For a K41superscriptK41{}^{41}{\rm K}start_FLOATSUPERSCRIPT 41 end_FLOATSUPERSCRIPT roman_K number density of n=1014cm3𝑛superscript1014superscriptcm3n=10^{14}~{}{\rm cm}^{-3}italic_n = 10 start_POSTSUPERSCRIPT 14 end_POSTSUPERSCRIPT roman_cm start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT it is Nc109subscript𝑁𝑐superscript109N_{c}\approx 10^{9}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT, and Vc=Nc/n0.01mm3subscript𝑉𝑐subscript𝑁𝑐𝑛0.01superscriptmm3V_{c}=N_{c}/n\approx 0.01~{}{\rm mm}^{3}italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_n ≈ 0.01 roman_mm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT. For Cs133superscriptCs133{}^{133}{\rm Cs}start_FLOATSUPERSCRIPT 133 end_FLOATSUPERSCRIPT roman_Cs these numbers would be Nc1013subscript𝑁𝑐superscript1013N_{c}\approx 10^{13}italic_N start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 10 start_POSTSUPERSCRIPT 13 end_POSTSUPERSCRIPT, and Vc0.1cm3subscript𝑉𝑐0.1superscriptcm3V_{c}\approx 0.1~{}{\rm cm}^{3}italic_V start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≈ 0.1 roman_cm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT.

We finally note that the correlation described by κ𝜅\kappaitalic_κ is not inconsistent with a spin-temperature distribution. We can rescale Tssubscript𝑇𝑠T_{s}italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT as given by Eq. (3) by κ𝜅\kappaitalic_κ, i.e. TsTs/κsubscript𝑇𝑠subscript𝑇𝑠𝜅T_{s}\rightarrow T_{s}/\kappaitalic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT → italic_T start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT / italic_κ, and immediately arrive at expression (6) by considering the magnetic field produced by the noise polarization of this vapor, which will now read σzκ/Nexpectationsubscript𝜎𝑧𝜅𝑁\braket{\sigma_{z}}\approx\kappa/\sqrt{N}⟨ start_ARG italic_σ start_POSTSUBSCRIPT italic_z end_POSTSUBSCRIPT end_ARG ⟩ ≈ italic_κ / square-root start_ARG italic_N end_ARG. Stated differently, this correlation embodied in κ𝜅\kappaitalic_κ reflects a slightly cooler spin temperature, or equivalently, a higher noise polarization.

The physical picture that emerges is that the atomic vapor will behave like a “squashy” spin medium exhibiting spin fluctuations, split in a number of “domains” not unlike ferromagnetic systems. Such domains exhibit intra-domain correlations, but not inter-domain correlation. Their contributions to the noise field add in quadrature, hence the total noise field is still given by (6). The vapor as a whole continuously exchanges energy with its self-field, setting an unavoidable noise level for measuring externally applied magnetic fields. Towards an experimental verification of the noise δB𝛿𝐵\delta Bitalic_δ italic_B, we propose the use of a miniaturized cesium cell of volume e.g. 1mm31superscriptmm31~{}{\rm mm^{3}}1 roman_mm start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT, in order to boost the previous estimates of δB𝛿𝐵\delta Bitalic_δ italic_B by a factor of 104superscript10410^{4}10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT. To avoid wall relaxation the cell should have buffer gas in order to avoid wall relaxation [68, 69], or anti-relaxation coating [70]. Also, the cell should be surrounded by a superconducting flux transformer in order to alleviate Johnson noise that would dominate the signal induced by the changing flux produced by δB𝛿𝐵\delta Bitalic_δ italic_B in an ordinary coil.

The possibility that spin uncertainty in an atomic vapor creates a fluctuating field consistent with the ERL was discussed in [36] and references therein, but largely dismissed, mainly on grounds of angular momentum and energy conservation. The physical picture painted here differs in several subtle ways from [36]: (i) We claim that (a) when starting from a spin-polarized state, which during the time τ𝜏\tauitalic_τ relaxes to the equilibrium state exhibiting spin-noise, energy will be exchanged between atomic spins and magnetic field, and (b) this process will continue even if the atoms are left alone in their equilibrium state. While the authors in [36] refer to quantum-measurement induced spin uncertainty, we refer to spontaneous spin noise that exists in such vapors whether there is or isn’t an external measurement limited by quantum uncertainty. (ii) There is no issue with angular momentum conservation, since in SD collisions angular momentum is taken up by translational angular momentum of the colliding atoms. Similarly, there is no issue with energy conservation, since spin dynamics are driven by collisions, the translational energy of which is supplied by an external energy source. A tiny fraction of this energy is transformed into magnetic interactions through SD and other spin-dependent collisions. (iii) The authors in [36] postulated such magnetic field fluctuations emanating from spin uncertainty, and showed they are consistent with the ERL. We started from the opposite direction, postulating the thermodynamic work exchanged between sensor and field, and arriving at these spin fluctuations and their self-interaction, connecting the thermodynamic work to the specific measurement process going on during SD spin relaxation. (iv) The authors in [36] considered two different noise sources, one stemming from the uncertainty relation discussed above, and another from the field produced by randomly polarized spins. Then they elaborated on the difficulties of adding those two terms in quadrature. We unify these into a single noise term interpreted with the novel type of correlations manifested as an enhanced permeability of the vapor. (v) The quantum speed limits did not work in the considerations of [36], because they concerned the total energy B2V/2μ0superscript𝐵2𝑉2subscript𝜇0B^{2}V/2\mu_{0}italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Here they concern the energy exchanged between atoms and field, which is (δB)2V/2μ0superscript𝛿𝐵2𝑉2subscript𝜇0(\delta B)^{2}V/2\mu_{0}( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT.

A pending question is whether the ERL is a hard fundamental limit. In fact, there are measurements claiming an ERL below Planck-constant-over-2-pi\hbarroman_ℏ [71, 72, 73]. We note that the magnetic field fluctuations in (4) were derived assuming a spin-temperature equilibrium state of the atomic vapor. Such thermal equilibrium state might not be the case, thus a basic assumption of the derivation of the quantum thermodynamic work in [53] would break down in the first place. For example, for a spin-squeezed equilibrium state, the spin fluctuations would be suppressed by some squeezing factor ξ1𝜉1\xi\leq 1italic_ξ ≤ 1 [74]. Correspondingly, the ERL in (5) would be violated by the factor ξ2superscript𝜉2\xi^{2}italic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. In this respect, the claims in [71, 72, 73] seem plausible. Given that the Heisenberg limit allows ξ2=1/Nsuperscript𝜉21𝑁\xi^{2}=1/Nitalic_ξ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 / italic_N, where N𝑁Nitalic_N is the atom number, one could imagine an ultra-low ERL of /NPlanck-constant-over-2-pi𝑁\hbar/Nroman_ℏ / italic_N. This conclusion tacitly assumes that the relaxation time τ𝜏\tauitalic_τ is not ξ𝜉\xiitalic_ξ-dependent. However, the physics of spin relaxation in correlated vapors is far from being understood [75, 76, 77]. Thus, notwithstanding the claims [71, 72, 73], we cannot make strong statements about the general validity of the ERL.

We will, however, provide two more demonstrations of the quantum thermodynamic ERL, for the SQUID and the diamond sensor [78, 79]. Going back to the relation (δB)2V/2μ0=kBTsuperscript𝛿𝐵2𝑉2subscript𝜇0subscript𝑘𝐵𝑇(\delta B)^{2}V/2\mu_{0}=k_{B}T{\cal I}( italic_δ italic_B ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_V / 2 italic_μ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T caligraphic_I, these sensors should satisfy kBTτπ2subscript𝑘𝐵𝑇𝜏𝜋2Planck-constant-over-2-pik_{B}T{\cal I}\tau\geq{\pi\over 2}\hbaritalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T caligraphic_I italic_τ ≥ divide start_ARG italic_π end_ARG start_ARG 2 end_ARG roman_ℏ. Regarding diamond sensors, the electron spin relaxes by the end of the measurement, hence as in the case of atomic vapors, the extracted information is =ln22{\cal I}=\ln 2caligraphic_I = roman_ln 2. We use the numbers of [80], where a room-temperature diamond sensor has spin relaxation time τ1μs𝜏1𝜇s\tau\approx 1~{}{\rm\mu s}italic_τ ≈ 1 italic_μ roman_s. We find an optimal ERL of kBTτ3×107subscript𝑘𝐵𝑇𝜏3superscript107Planck-constant-over-2-pik_{B}T{\cal I}\tau\approx 3\times 10^{7}\hbaritalic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T caligraphic_I italic_τ ≈ 3 × 10 start_POSTSUPERSCRIPT 7 end_POSTSUPERSCRIPT roman_ℏ. The magnetic sensitivity reported in [80] is δB=300pT/Hz𝛿𝐵300pTHz\delta B=300~{}{\rm pT}/\sqrt{\rm Hz}italic_δ italic_B = 300 roman_pT / square-root start_ARG roman_Hz end_ARG, and the measurement volume estimated as V=600μm×4650μm2𝑉600𝜇m4650𝜇superscriptm2V=600~{}{\rm\mu m\times 4650~{}\mu m^{2}}italic_V = 600 italic_μ roman_m × 4650 italic_μ roman_m start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Thus, the measured ERL reads 109superscript109Planck-constant-over-2-pi10^{9}\hbar10 start_POSTSUPERSCRIPT 9 end_POSTSUPERSCRIPT roman_ℏ, so a 30-fold improvement is possible. The authors in [80] state δB𝛿𝐵\delta Bitalic_δ italic_B is a factor of 6 away from the photon shot-noise limit. If this saturates the ERL, the sensing volume would need to be revised by 20%. In any case, such considerations illuminate the constructive role the quantum-thermodynamic ERL can play in understanding and optimizing magnetic sensors.

For SQUIDs the temperature T𝑇Titalic_T is on the order of 1K1superscriptK1~{}^{\circ}~{}{\rm K}1 start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPT roman_K. What is the information {\cal I}caligraphic_I in this case? The SQUID could be seen as distinguishing between a flux nΦ0𝑛subscriptΦ0n\Phi_{0}italic_n roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and a flux (n+1)Φ0𝑛1subscriptΦ0(n+1)\Phi_{0}( italic_n + 1 ) roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, where Φ0subscriptΦ0\Phi_{0}roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the flux quantum.

Table 2: SQUID ERL. Comparison between measurements and quantum thermodynamic prediction
Publication [81] [82] [83] [84] [85]
p𝑝pitalic_p (106superscript10610^{-6}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT) 0.045 5.5 0.084 0.6 1.3
T(KT({\rm{}^{\circ}K}italic_T ( start_FLOATSUPERSCRIPT ∘ end_FLOATSUPERSCRIPT roman_K) 4.2 1.5 0.29 4.2 4.2
τ𝜏\tauitalic_τ (105ssuperscript105s10^{-5}~{}{\rm s}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT roman_s) 0.5 0.013 5.0 0.5 0.5
Predicted ERL (Planck-constant-over-2-pi\hbarroman_ℏ) 2.1 1.6 2.6 24 50
Measured ERL (Planck-constant-over-2-pi\hbarroman_ℏ) 6.3 1.6 1.7 35 121

These two altrenatives would indeed correspond to 1 bit of information. However, in the flux-locked-loop operating mode [18, 19], the actual flux noise (the excursions of the flux around the operating point (n+1/2)Φ0𝑛12subscriptΦ0(n+1/2)\Phi_{0}( italic_n + 1 / 2 ) roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) is on the order of pΦ0𝑝subscriptΦ0p\Phi_{0}italic_p roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, hence the gained information is plnp𝑝𝑝{\cal I}\approx-p\ln pcaligraphic_I ≈ - italic_p roman_ln italic_p, where p1much-less-than𝑝1p\ll 1italic_p ≪ 1. We thus expect (plnp)kBTτπ2𝑝𝑝subscript𝑘𝐵𝑇𝜏𝜋2Planck-constant-over-2-pi(-p\ln p)k_{B}T\tau\geq{\pi\over 2}\hbar( - italic_p roman_ln italic_p ) italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T italic_τ ≥ divide start_ARG italic_π end_ARG start_ARG 2 end_ARG roman_ℏ. To demonstrate this, we use the results of five SQUID papers [81, 82, 83, 84, 85] reporting (i) the flux noise as a fraction of Φ0subscriptΦ0\Phi_{0}roman_Φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT per Hz1/2superscriptHz12{\rm Hz}^{1/2}roman_Hz start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT, from which we extract the parameter p𝑝pitalic_p given a 1 Hz resolution bandwidth, and (ii) the measurement bandwidth, from which we extract τ𝜏\tauitalic_τ. The authors in [81, 82, 83, 84, 85] explicitly report the measured ERL, with which we compare the predicted ERL, i.e. the expression (plnp)kBTτ𝑝𝑝subscript𝑘𝐵𝑇𝜏(-p\ln p)k_{B}T\tau( - italic_p roman_ln italic_p ) italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T italic_τ. It is seen at Table II that the measured ERL is higher and of the same order compared to the quantum thermodynamic ERL (except the case of [83], where they coincide within the experimental errors of [83]). Thus regarding SQUIDs, the quantum thermodynamic ERL appears to work for a range from 11001Planck-constant-over-2-pi100Planck-constant-over-2-pi1\hbar-100\hbar1 roman_ℏ - 100 roman_ℏ.

Concluding, we have presented a first-principles derivation for the energy resolution limit in magnetic sensing, specified for atomic spin magnetometers, diamond sensors and SQUIDs. We demonstrate the applicability of the ERL for starkly different physical parameter values relevant to three starkly different sensor technologies, over a range of seven orders of magnitude in Planck-constant-over-2-pi\hbarroman_ℏ. The understanding of the physical principles behind the ERL could have multifaceted consequences for the effort to advance quantum sensing technologies.

References

  • [1] C. L. Degen, F. Reinhard, and P. Cappellaro, Quantum sensing, Rev. Mod. Phys. 89, 035002 (2017).
  • [2] I. H. Deutsch, Harnessing the power of the second quantum revolution, PRX Quantum 1, 020101 (2020).
  • [3] G.E. Katsoprinakis, D. Petrosyan, and I. K. Kominis, High frequency atomic magnetometer by use of electromagnetically induced transparency, Phys. Rev. Lett. 97, 230801 (2006).
  • [4] R. Jiménez-Martínez, W. C. Griffith, S. Knappe, J. Kitching, and M. Prouty, High-bandwidth optical magnetometer, J. Opt. Soc. Am. B 29, 3398 (2012).
  • [5] R. Li, F. N. Baynes, A. N. Luiten, and C. Perrella, Continuous high-sensitivity and high-bandwidth atomic magnetometer, Phys. Rev. Applied 14, 064067 (2020).
  • [6] S. Li, J. Liu, M.Jin, K. T. Akiti, P. Dai, Z. Xu, and T. E.-T. Nwodom, A kilohertz bandwidth and sensitive scalar atomic magnetometer using an optical multipass cell, Measurement 190, 110704 (2022).
  • [7] G. Waldherr, J. Beck, P. Neumann, R. S. Said, M. Nitsche, M. L. Markham, D. J. Twitchen, J. Twamley, F. Jelezko, and J. Wrachtrup, High-dynamic-range magnetometry with a single nuclear spin in diamond, Nature Nanotech. 7, 105 (2012).
  • [8] N. M. Nusran, M. U. Momeen, and M. V. G. Dutt, High-dynamic-range magnetometry with a single electronic spin in diamond, Nature Nanotech. 7, 109 (2012).
  • [9] H. Clevenson, L. M. Pham; C. Teale; K. Johnson, D. Englund, and D. Braje, Robust high-dynamic-range vector magnetometry with nitrogen-vacancy centers in diamond, Appl. Phys. Lett. 112, 252406 (2018)
  • [10] K. K. G. Kurian, S. S. Sahoo, P. K. Madhu, and G. Rajalakshmi, Single-beam room-temperature atomic magnetometer with large bandwidth and dynamic range, Phys. Rev. Applied 19, 054040 (2023).
  • [11] P. D. Schwindt, S. Knappe, V. Shah, L. Hollberg, J. Kitching, L.-A. Liew, and J. Moreland, Chip-scale atomic magnetometer, Appl. Phys. Lett. 85, 6409 (2004).
  • [12] V. Shah, S. Knappe, P. D. D. Schwindt, and J. Kitching, Subpicotesla atomic magnetometry with a microfabricated vapour cell, Nat. Photonics 1, 649 (2007).
  • [13] Y. Sebbag, E. Talker, A. Naiman, Y. Barash, and U. Levy Demonstration of an integrated nanophotonic chip-scale alkali vapor magnetometer using inverse design. Light Sci. Appl. 10, 54 (2021).
  • [14] F. Gotardo, B. J. Carey, H. Greenall, G. I. Harris, E. Romero, D. Bulla, E. M. Bridge, J. S. Bennett, S. Foster, and W. P. Bowen, Waveguide-integrated chip-scale optomechanical magnetometer, Opt. Express 31, 37663 (2023).
  • [15] Z. D. Grujić, P. A. Koss, G. Bison, and A. Weis, A sensitive and accurate atomic magnetometer based on free spin precession, Europ. Phys. J. D 69, 1 (2015).
  • [16] K.-M. C. Fu, G. Z. Iwata, A. Wickenbrock, and D. Budker, Sensitive magnetometry in challenging environments, AVS Quantum Sci. 2, 044702 (2020).
  • [17] X. Bai, K. Wen, D. Peng, S. Liu, and L. Luo, Atomic magnetometers and their application in industry, Front. Phys. 11, 1212368. (2023).
  • [18] J. Clarke, SQUID fundamentals, in W. H. Dordrecht, SQUlD Sensors: Fundamentals, Fabrication and Applications (Kluwer Academic Press, 1996)
  • [19] R. L. Fagaly, SQUID instruments and applications, Rec. Sci. Instr. 77, 101101 (2006).
  • [20] I. K. Kominis, T. W. Kornack, J. C. Allred, and M. V. Romalis, A subfemtotesla multichannel atomic magnetometer, Nature 422, 596 (2003).
  • [21] D. Budker and M. V. Romalis, Optical magnetometry, Nature Phys. 3, 227 (2007).
  • [22] M. Hämäläinen, R. Hari, R. J. Ilmoniemi, J. Knuutila, and O. V. Lounasmaa, Magnetoencephalography - theory, instrumentation, and applications to noninvasive studies of the working human brain, Rev. Mod. Phys. 65, 413 (1993).
  • [23] H. Xia, A. B.-A. Baranga, D. Hoffman, and M. V. Romalis, Magnetoencephalography with an atomic magnetometer, Appl. Phys. Lett. 89, 211104 (2006).
  • [24] T. H. Sander, J. Preusser, R. Mhaskar, J. Kitching, L. Trahms, and S. Knappe, Magnetoencephalography with a chip-scale atomic magnetometer, Biomed. Opt. Express 3, 981 (2012).
  • [25] J. Sheng, S. Wan, Y. Sun, R. Dou, Y. Guo, K. Wei, K. He; J. Qin, and J.-H. Gao, Magnetoencephalography with a Cs-based high-sensitivity compact atomic magnetometer, Rev. Sci. Instrum. 88, 094304 (2017).
  • [26] R. Zhang, W. Xiao, Y. Ding, Y. Feng, X. Peng, L. Shen, C. Sun, T. Wu, Y. Wu, Y. Yang, Z. Zheng, X. Zhang, J. Chen, and H. Guo, Recording brain activities in unshielded Earth’s field with optically pumped atomic magnetometers, Sci. Adv. 6, eaba8792 (2020).
  • [27] G. Bison, R. Wynands, and A. Weis, A laser-pumped magnetometer for the mapping of human cardiomagnetic fields, Appl. Phys. B 76, 325 (2003).
  • [28] J. Kim, I. Savukov, and S. Newman, Magnetocardiography with a 16-channelf iber-coupled single-cell Rb optically pumped magnetometer, Appl.Phys. Lett. 114, 143702 (2019).
  • [29] M. V. Romalis and H. B. Dang, Atomic magnetometers for materials characterization, Mater. Today 14, 258 (2011).
  • [30] F. Fabre, A. Finco, A. Purbawati, A. Hadj-Azzem, N. Rougemaille, J. Coraux, I. Philip, and V. Jacques, Characterization of room-temperature in-plane magnetization in thin flakes of CrTe2subscriptCrTe2{\rm CrTe_{2}}roman_CrTe start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT with a single-spin magnetometer, Phys. Rev. Materials 5, 034008 (2021).
  • [31] G. Vasilakis, J. M. Brown, T. W. Kornack, and M. V. Romalis, Limits on new long range nuclear spin-dependent forces set with a K3Hesuperscript3KHe{\rm K-^{3}He}roman_K - start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_He comagnetometer, Phys. Rev. Lett. 103, 261801 (2009).
  • [32] M. Smiciklas, J. M. Brown, L. W. Cheuk, S. J. Smullin, and M. V. Romalis, New test of local Lorentz invariance using a Ne21RbKsuperscriptNe21RbK{\rm{}^{21}Ne-Rb-K}start_FLOATSUPERSCRIPT 21 end_FLOATSUPERSCRIPT roman_Ne - roman_Rb - roman_K comagnetometer, Phys. Rev. Lett. 107, 171604 (2011).
  • [33] S. Pustelny, D. F. Jackson Kimball, C. Pankow, M. P. Ledbetter, P. Wlodarczyk, P. Wcislo, M. Pospelov, J. R. Smith, J. Read, W. Gawlik, and D. Budker, The global network of optical magnetometers for exotic physics (GNOME): A novel scheme to search for physics beyond the Standard Model, Ann. Phys. (Berlin) 525, 659 (2013).
  • [34] A. O. Sushkov, Quantum science and the search for axion dark matter, PRX Quantum 4, 020101 (2023).
  • [35] D. F. J. Kimball, D. Budker, T. E. Chupp, A. A. Geraci, S. Kolkowitz, J. T. Singh, and A. O. Sushkov, Probing fundamental physics with spin-based quantum sensors, Phys. Rev. A 108, 010101 (2023).
  • [36] M. W. Mitchell and S. P. Alvarez, Colloquium: Quantum limits to the energy resolution of magnetic field sensors, Rev. Mod. Phys. 92, 021001 (2020).
  • [37] D. Robbes, Highly sensitive magnetometers - a review, Sens. Actuators A 129, 86 (2006).
  • [38] J. Gemmer, M. Michel and G. Mahler, Quantum Thermodynamics: Emergence of Thermodynamic Behavior Within Composite Quantum Systems, Lect. Notes Phys. 784 (Springer, Berlin Heidelberg 2009).
  • [39] R. Kosloff, Quantum thermodynamics: A dynamical viewpoint, Entropy 15, 2100 (2013).
  • [40] J. M. R. Parrondo, J. M. Horowitz, and T. Sagawa., Thermodynamics of information, Nat. Phys. 11, 131 (2015).
  • [41] S. Vinjanampathya, J. Anders, Quantum thermodynamics, Contemp. Phys. 57, 545 (2016).
  • [42] J. Goold, M. Huber, A. Riera, L. del Rio and P. Skrzypczyk, The role of quantum information in thermodynamics - a topical review, J. Phys. A: Math. Theor. 49 143001 (2016).
  • [43] S. Deffner and S. Campbell, Quantum Thermodynamics (Morgan & Claypool Publishers, 2019)
  • [44] A. Tuncer and O. E. Mustecaplioglu, Turk. J. Phys. 44, 404 (2020).
  • [45] R. Landauer, Irreversibility and heat generation in the computing process, IBM J. Res. Dev. 5, 183 (1961).
  • [46] C. H. Bennett, The thermodynamics of computation - a review, Int. J. Theor. Phys. 21, 905 (1982).
  • [47] M. B. Plenio and V. Vitelli, The physics of forgetting: Landauer’s erasure principle and information theory, Contemp. Phys. 42, 25 (2001).
  • [48] P. Kammerlander and J. Anders, Coherence and measurement in quantum thermodynamics, Sci. Rep. 6, 22174 (2016).
  • [49] P. Lipka-Bartosik and R. Demkowicz-Dobrzański, Thermodynamic work cost of quantum estimation protocols, J. Phys. A: Math. Theor. 51, 474001 (2018).
  • [50] S. Bhattacharjee, U. Bhattacharya, W. Niedenzu, V. Mukherjee and A. Dutta, Quantum magnetometry using two-stroke thermal machines, New J. Phys. 22 013024 (2020).
  • [51] Y. Chu and J. Cai, Thermodynamic Principle for Quantum Metrology, Phys. Rev. Lett. 128, 200501 (2022).
  • [52] J. V. Koski, V. F. Maisi, T. Sagawa, and J. P. Pekola, Experimental observation of the role of mutual information in the nonequilibrium dynamics of a Maxwell demon, Phys. Rev. Lett. 113, 030601 (2014).
  • [53] T. Sagawa and M. Ueda, Minimal energy cost for thermodynamic information processing: measurement and information erasure, Phys. Rev. Lett. 102, 250602 (2009).
  • [54] N. Margolus and L. B. Levitin, The maximum speed of dynamical evolution, Physica D 120, 188 (1998).
  • [55] S, Deffner and S. Campbell, Quantum speed limits: from Heisenberg’s uncertainty principle to optimal quantum control, J. Phys. A: Math. Theor. 50 453001 (2017).
  • [56] J. C. Allred, R. N. Lyman, T. W. Kornack, and M. V. Romalis, High-sensitivity atomic magnetometer unaffected by spin-exchange relaxation, Phys. Rev. Lett. 89, 130801 (2002).
  • [57] K. Mouloudakis and I.K. Kominis, Spin-exchange collisions in hot vapors creating and sustaining bipartite entanglement, Phys. Rev. A Lett. 103, L010401 (2021).
  • [58] S. M. Rochester, K. Szymański, M. Raizen, S. Pustelny, M. Auzinsh, and D. Budker, Efficient polarization of high-angular-momentum systems, Phys. Rev. A 94, 043416 (2016).
  • [59] S. Kadlecek, T. Walker, D. K. Walter, C. Erickson, and W. Happer, Spin-axis relaxation in spin-exchange collisions of alkali-metal atoms, Phys. Rev. A 63, 052717 (2001).
  • [60] W. Happer and W. A. van Wijngaarden, An optical pumping primer, Hyperfine Int. 38, 435 (1987).
  • [61] S. Appelt, A. B.-A. Baranga, C. J. Erickson, M. V. Romalis, A. R. Young, and W. Happer, Theory of spin-exchange optical pumping of 3-He and 129-Xe, Phys. Rev. A 58, 1412 (1998).
  • [62] S. A. Crooker, D. G. Rickel, A. V. Balatsky, and D. L. Smith, Spectroscopy of spontaneous spin noise as a probe of spin dynamics and magnetic resonance, Nature 43, 49 (2004).
  • [63] G. E. Katsoprinakis, A. T. Dellis, and I. K. Kominis, Measurement of transverse spin-relaxation rates in a rubidium vapor by use of spin-noise spectroscopy, Phys. Rev. A 75, 042502 (2007).
  • [64] E. B. Aleksandrov and V. S. Zapasskii, Spin noise spectroscopy, J. Phys.: Conf. Series 324, 012002 (2011).
  • [65] N. A. Sinitsyn and Y. V. Pershin, The theory of spin noise spectroscopy: a review, Rep. Prog. Phys. 79, 106501 (2016).
  • [66] K. Mouloudakis, F. Vouzinas, A. Margaritakis, A. Koutsimpela, G. Mouloudakis, V. Koutrouli, M. Skotiniotis, G. P. Tsironis, M. Loulakis, M. W. Mitchell, G. Vasilakis, and I. K. Kominis, Interspecies spin-noise correlations in hot atomic vapors, Phys. Rev. A 108, 052822 (2023).
  • [67] M. P. Ledbetter, I. M. Savukov, V. M. Acosta, D. Budker, and M. V. Romalis, Spin-exchange-relaxation-free magnetometry with Cs vapor, Phys. Rev. A 77, 033408 (2008).
  • [68] L.-A. Liew, S. Knappe, J. Moreland, H. Robinson, L. Hollberg, and J. Kitching, Microfabricated alkali atom vapor cells, Appl. Phys. Lett. 84, 2694 (2004).
  • [69] M. Hasegawa, R. K. Chutani, C. Gorecki, R. Boudot, P. Dziuban, V. Giordano, S. Clatot, and L. Mauri, Microfabrication of cesium vapor cells with buffer gas for MEMS atomic clocks, Sens. Actuators A 167, 594 (2011).
  • [70] R. Straessle, M. Pellaton, C. Affolderbach, Y. Pétremand, D. Briand, G. Mileti, and N. F. de Rooij, Microfabricated alkali vapor cell with anti-relaxation wall coating, Appl. Phys. Lett. 105, 043502 (2014).
  • [71] A. Vinante, C. Timberlake, D. Budker, D. F. Jackson Kimball, A. O. Sushkov, and H. Ulbricht, Surpassing the energy resolution limit with ferromagnetic torque sensors, Phys. Rev. Lett. 127, 070801 (2021).
  • [72] S. P. Alvarez , P. Gomez, S. Coop, R. Zamora-Zamora, C. Mazzinghi, and M. W. Mitchell, Single-domain Bose condensate magnetometer achieves energy resolution per bandwidth below Planck-constant-over-2-pi\hbarroman_ℏ, Proc. Nat. Acad. USA 119, e2115339119 (2022).
  • [73] F. Ahrens, W. Ji, D. Budker, C. Timberlake, H. Ulbricht, and A. Vinante, Levitated ferromagnetic magnetometer with energy resolution well below Planck-constant-over-2-pi\hbarroman_ℏ, arXiv:2401:03774.
  • [74] G. Tóth C. Knapp, O. Gühne, and H. J. Briegel, Optimal spin squeezing inequalities detect bound entanglement in spin models, Phys. Rev. Lett. 99, 250405 (2007).
  • [75] M. Auzinsh, D. Budker, D. F. Kimball, S. M. Rochester, J. E. Stalnaker, A. O. Sushkov, and V. V. Yashchuk, Can a quantum nondemolition measurement improve the sensitivity of an atomic magnetometer?, Phys. Rev. Lett. 93, 173002 (2004).
  • [76] I. K. Kominis, Sub-shot-noise magnetometry with a correlated spin-relaxation dominated alkali-metal vapor, Phys. Rev. Lett. 100, 073002 (2008).
  • [77] J. Kong, R. Jiménez-Martínez, C. Troullinou, V. G. Lucivero, Géza Tóth, and M. W. Mitchell, Measurement-induced, spatially-extended entanglement in a hot, strongly-interacting atomic system, Nat. Commun. 11, 2415 (2020).
  • [78] G. Balasubramanian, I. Y. Chan, R. Kolesov, M. Al-Hmoud, J. Tisler, C. Shin, C. Kim, A. Wojcik, P. R. Hemmer, A. Krueger, T. Hanke, A. Leitenstorfer, R. Bratschitsch, F. Jelezko, and J. Wrachtrup, Nanoscale imaging magnetometry with diamond spins under ambient conditions, Nature 455, 648 (2008).
  • [79] H. Zheng, Z. Sun, G. Chatzidrosos, C. Zhang, K. Nakamura, H. Sumiya, T. Ohshima, J. Isoya, J. Wrachtrup, A. Wickenbrock, and D. Budker, Microwave-free vector magnetometry with nitrogen-vacancy centers along a single axis in diamond, Phys. Rev. Appl. 13, 044023 (2020).
  • [80] R. L. Patel, L. Q. Zhou, A. C. Frangeskou,G. A. Stimpson, B. G. Breeze, A. Nikitin, M. W. Dale, E. C. Nichols,W. Thornley,B. L. Green,M. E. Newton, A. M. Edmonds, M. L. Markham, D. J. Twitchen, and G. W. Morley, Subnanotesla magnetometry with a fiber-coupled diamond sensor, Phys. Rev. Appl. 14, 044058 (2020).
  • [81] M. Schmelz, V. Zakosarenko, T. Schönau, S. Anders, S. Linzen,R. Stolz, and H.-G. Meyer, Nearly quantum limited nanoSQUIDs based on cross-type Nb/AlOxsubscriptOx{\rm O_{x}}roman_O start_POSTSUBSCRIPT roman_x end_POSTSUBSCRIPT/Nb junctions, Supercond. Sci. Technol. 30, 014001 (2017).
  • [82] R. T. Wakai and D. J. Van Harlingen, Signal and white noise properties of edge junction dc SQUIDs, Appl. Phys. Lett. 52, 1182 (1988).
  • [83] D. D. Awschalom, J. R. Rozen, M. B. Ketchen, W. J. Gallagher, A. W. Kleinsasser, R. L. Sandstrom, and B. Bumble, Low-noise modular microsusceptometer using nearly quantum limited dc SQUIDs, Appl. Phys. Lett. 53, 2108 (1988).
  • [84] M. Schmelz, V. Zakosarenko, A. Chwala, T. Schönau, R. Stolz, S. Anders, S. Linzen, and H.-G. Meyer, Thin-film-based ultralow noise SQUID magnetometer, IEEE Transact. Appl. Supercond. 26, 1600804 (2016).
  • [85] M. Schmelz, R. Stolz, V. Zakosarenko, T. Schönau, S. Anders, L. Fritzsch, M. Mück, and H.-G. Meyer, Field-stable SQUID magnetometer with sub-fT Hz1/2superscriptHz12{\rm Hz}^{-1/2}roman_Hz start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT resolution based on sub-micrometer cross-type Josephson tunnel junctions, Supercond. Sci. Technol. 24 065009 (2011).