Paper The following article is Open access

Quantum Brownian motion for magnets

, and

Published 21 March 2022 © 2022 The Author(s). Published by IOP Publishing Ltd on behalf of the Institute of Physics and Deutsche Physikalische Gesellschaft
, , Citation J Anders et al 2022 New J. Phys. 24 033020 DOI 10.1088/1367-2630/ac4ef2

Download Article PDF
DownloadArticle ePub

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

1367-2630/24/3/033020

Abstract

Spin precession in magnetic materials is commonly modelled with the classical phenomenological Landau–Lifshitz–Gilbert (LLG) equation. Based on a quantized three-dimensional spin + environment Hamiltonian, we here derive a spin operator equation of motion that describes precession and includes a general form of damping that consistently accounts for memory, coloured noise and quantum statistics. The LLG equation is recovered as its classical, Ohmic approximation. We further introduce resonant Lorentzian system–reservoir couplings that allow a systematic comparison of dynamics between Ohmic and non-Ohmic regimes. Finally, we simulate the full non-Markovian dynamics of a spin in the semi-classical limit. At low temperatures, our numerical results demonstrate a characteristic reduction and flattening of the steady state spin alignment with an external field, caused by the quantum statistics of the environment. The results provide a powerful framework to explore general three-dimensional dissipation in quantum thermodynamics.

Export citation and abstract BibTeX RIS

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

The continued miniaturisation of critical components in consumer electronics and neighbouring technologies will require a deeper understanding of thermal noise and general thermodynamic principles beyond the classical macroscopic world. Quantum thermodynamics [13] has emerged as a field addressing the conceptual challenges related to the exchange of energy and information at the nanoscale. Recent advances include studies of heat transport in quantum systems [411], the characterisation of memory effects in their dynamics [1219], and clarification of the impact of quantum coherence and correlation on thermodynamic processes [2024]. The establishment of a generalised thermodynamic framework, valid for nanoscale systems that strongly couple to environmental modes, is well under way [2533], and for magnetic molecules an environment-induced renormalisation of the anisotropy has been predicted [34]. Two open quantum systems models have served as the workhorse for many of these conceptual studies; the Caldeira–Leggett model for quantum Brownian motion [8, 35, 36] and the spin-boson model of a spin (or many spins) coupled to a one-dimensional harmonic bath [5, 6, 33, 3739]. These describe a very wide range of physical situations extending to studies of quantum effects in bio-chemical reactions [40], where they are used to model exciton–phonon interactions [41].

Until now few nanoscale technologies have required the use of advanced open quantum systems techniques. But advances in engineering magnetic materials for magnetic hard drives at unprecedented length and time-scales [42] are likely to require a more detailed picture of spin dynamics including memory and quantum signatures. Here we introduce a three-dimensional open quantum system model to characterise the quantum Brownian motion of spins in magnetic materials.

Magnetic behaviour has been studied extensively based on the classical phenomenological Landau–Lifshitz–Gilbert (LLG) equation [4347]

Equation (1)

which is routinely solved with micromagnetic and atomistic simulations. Here M is the magnetic moment, γ is the gyromagnetic ratio and B eff is an effective magnetic field which includes the external field [107], exchange and anisotropy effects, as well as stochastic magnetic noise $\boldsymbol{b}\propto \sqrt{T}$ stemming from an environment at temperature T that was added by Brown [108] in 1963 [48]. The final term on the right of (1) is the so-called 'Gilbert damping' term and the positive constant ηG is the damping parameter [109], which is often rewritten as ηG = η/| M ||γ| with a unit-free η.

Gilbert damping is not derived from microscopic principles, but chosen as the simplest term that could serve to align the magnetic moment with the applied field [43]. As we will see, it contains no memory which is increasingly seen as a limitation [49, 50]. Advances in engineering magnetic materials at the nanoscale and manipulating them on ultrafast timescales also indicate that a theory beyond the classical LLG equation is required [47]. Early attempts have pursued a path integral derivation of a quantum spin dynamics equation [51], as well as other conceptually related classical and quantum derivations [52, 53]. These derivations were not directly applied to the calculation of magnetization dynamics or steady states, nor have they been connected to recent generalizations of Gilbert damping that include inertial terms [49, 50, 5459] or provided an assessment of quantum effects.

Here we go further and develop a comprehensive and quantum-thermodynamically consistent theory suitable to describe the quantum dynamics of spins in magnetic materials including non-Markovian damping, coloured noise and quantum zero-point fluctuations. Unlike the conceptually pioneering Caldeira–Leggett model that has few experimental realisations, the developed three-dimensional quantum spin model is directly applicable to atomistic spin dynamics simulations [47, 60], ultrafast magnetism experiments [61], and systems exhibiting anisotropic damping [62]. The theory of magnetic materials developed here is analogous to macroscopic Quantum ElectroDynamics (QED), which successfully describes quantum electromagnetism in dielectric materials [6365]. Instead of trying to give a fully microscopic description that accounts for every interaction in the material, macroscopic QED characterizes electromagnetic materials in terms of frequency dependent susceptibilities, and can predict both the Lamb shift and the Casimir effect [64, 66]. This success motivates the development of our effective medium theory for magnetic materials, which we hope can help match experimental trends that go beyond the LLG paradigm [67], without necessarily requiring more detailed microscopic models. Previous extensions of the LLG equation have considered similar Hamiltonians to ours [68], but invoked the Born–Markov approximation which removes potentially important temporal correlations. Furthermore, the theory developed here provides the framework for examining non-classical noise power spectra, such as those used recently in semiclassical simulations of magnetic materials [60], that have improved the match between simulations and experimental data.

This paper 1 is organised as follows: in section 1 the general quantum spin dynamics equation for spin operator precession in three dimensions is derived. For the simplest, Ohmic, coupling this equation is found to reduce to the memory-free LLG equation. In section 2 we introduce Lorentzian couplings as a systematic method for exploring non-Markovian dynamical regimes in general open quantum systems, including spins. Finally, in section 3 we detail a numerical method to simulate non-Markovian dynamics, and present results for a single classical spin that illustrate the differences between spin dynamics and steady states arising with non-trivial memory, coloured noise, and quantum bath statistics in comparison to those obtained with the memory-free LLG equation. Conclusions and open questions are discussed in section 4.

1. Quantum spin dynamics equation

1.1. System + environment Hamiltonian

We begin by introducing the quantized Hamiltonian describing the different contributions to the total energy of the system, consisting of spins as well as environmental degrees of freedom (e.g. electrons and phonons), given by

Equation (2)

where ${\hat{H}}_{\mathrm{S}}$ is the bare spin Hamiltonian operator which captures the spin energy in external fields and interactions between spins, ${\hat{H}}_{\mathrm{R}}$ is the environmental or reservoir Hamiltonian, and ${\hat{V}}_{\text{int}}$ is the interaction between the spins and the reservoir.

We choose ${\hat{H}}_{\mathrm{S}}$ as the sum of the interaction with a homogeneous external field B ext [110] and the exchange interaction between three-dimensional spin vector operators ${\hat{\boldsymbol{S}}}^{(n)}\!=({\hat{S}}_{1}^{(n)},{\hat{S}}_{2}^{(n)},{\hat{S}}_{3}^{(n)})$ at sites n of a lattice [111],

Equation (3)

Here ${\mathcal{J}}^{(nm)}$ is the exchange tensor for spin pairs (n, m) [112], which can include the Dzyaloshinskii–Moriya interaction [47]. It is straightforward to include additional energetic terms in the bare spin Hamiltonian, such as energies associated with magnetic anisotropy. Instead of the magnetic moment M used in equation (1), we will here work with the spin angular momentum S proportional to M , M = γ S , where γ is the gyromagnetic ratio. In the following we will assume the gyromagnetic ratio |γ| = |ge μB/ |= 1.76 × 1011 s−1 T−1 for an electron, with γ taking the negative sign.

The reservoir Hamiltonian is commonly modelled as a set of harmonic oscillators [35, 63], and we here follow the continuous reservoir approach by Huttner and Barnett [63], taking the reservoir Hamiltonian as

Equation (4)

It describes a continuous frequency reservoir at each lattice site n, where ${\hat{\mathbf{\Pi }}}_{\omega }^{(n)}\!$ and ${\hat{\boldsymbol{X}}}_{\omega }^{(n)}\!$ are (three-dimensional) momentum and position operators of the reservoir oscillator with frequency ω. The position operators ${\hat{\boldsymbol{X}}}_{\omega }^{(n)}\!$ physically represent variations in the environment to which the spin at site n responds, see illustration figure 1, as for example, in magnon–phonon mediated loss [69]. Unlike most system + environment Hamiltonians which assume one-dimensional coupling, we here take the spin–reservoir interaction to be of the three-dimensional form

Equation (5)

This coupling allows angular momentum transfer as well as energy transfer between the spins and the environment. Here ${\mathcal{C}}_{\omega }^{(n)}$ is a three-dimensional coupling tensor and a function of frequency ω. At each ω, the coupling tensor determines the strength of the coupling of each spin to its reservoir oscillators at frequency ω, thus acting as a frequency filter. As we shall see, the choice of the coupling ${\mathcal{C}}_{\omega }^{(n)}$ will determine the damping of the spin dynamics as well as the stochastic noise experienced by the spins.

Figure 1.

Figure 1. Illustration of Hamiltonian model. In addition to precessing in an external field B ext, and coupling to its spin neighbours m with strength ${\mathcal{J}}^{(nm)}$, each spin ${\hat{\boldsymbol{S}}}^{(n)}\!$ couples to a continuum of environmental modes (phonons, electrons) at frequency ω with a coupling function ${\mathcal{C}}_{\omega }^{(n)}$. All environmental modes are assumed to be thermal at the same temperature T.

Standard image High-resolution image

For readers concerned about time-reversal symmetry of ${\hat{V}}_{\text{int}}$ in (5), we note that ${\hat{\boldsymbol{X}}}_{\omega }^{(n)}\!$ should be interpreted as an effective magnetic field seen by the spins due to their interaction with the environment, which has the same time symmetry as ${\hat{\boldsymbol{S}}}^{(n)}\!$ [113].

1.2. Equations of motion

Having set up the full Hamiltonian (2) of the spins and environment degrees of freedom allows the study of the spins' reduced state dynamics ${\hat{\rho }}_{\mathrm{S}}(t)={\mathrm{t}\mathrm{r}}_{\mathrm{R}}\left[{\hat{\rho }}_{\mathrm{S}\mathrm{R}}(t)\right]$ of the total state

Equation (6)

where the reservoir state ${\hat{\rho }}_{\mathrm{R}}={\text{e}}^{-\beta {\hat{H}}_{\mathrm{R}}}/\mathrm{t}\mathrm{r}[{\text{e}}^{-\beta {\hat{H}}_{\mathrm{R}}}]$ is thermal at some inverse temperature β = 1/kB T. In what follows it will be more convenient to work instead in the Heisenberg picture where the state is stationary, ${\hat{\rho }}_{\mathrm{S}\mathrm{R}}(0)$, while the time dependence of an operator $\hat{O}(t)$ is governed by the commutator $\mathrm{d}\hat{O}(t)/\mathrm{d}t=(\mathrm{i}/\hslash )[\hat{H},\hat{O}(t)]$. Expectation values at time t can then be obtained as

Equation (7)

Using the standard commutation relations for the spin operators (and orbital angular momentum operators in general), $[{\hat{S}}_{j}^{(n)},{\hat{S}}_{k}^{(m)}]=\mathrm{i}\hslash {\delta }_{mn}{\sum }_{l}{{\epsilon}}_{jkl}{\hat{S}}_{l}^{(n)}$ and for the position/momentum operators, $[{{\hat{X}}_{\omega ,j}}^{(n)},{{\hat{{\Pi}}}_{{\omega }^{\prime },k}}^{(m)}]=\mathrm{i}\hslash {\delta }_{nm}{\delta }_{jk}\delta (\omega -{\omega }^{\prime })$, we obtain the following equations of motion for the spin operators ${\hat{\boldsymbol{S}}}^{(n)}\!(t)$

Equation (8)

where ${\hspace{6pt}\bar{\hspace{-5pt}\mathcal{J}}}^{(nm)}=(1/2)[{\mathcal{J}}^{(nm)}+{({\mathcal{J}}^{(mn)})}^{\text{T}}]$ is the symmetrized exchange tensor and ${\hat{\boldsymbol{B}}}_{\text{env}}^{(n)}={\int }_{0}^{\infty }\!\mathrm{d}\omega \,{\mathcal{C}}_{\omega }^{(n)}{\hat{\boldsymbol{X}}}_{\omega }^{(n)}\!$ is a magnetic field operator formed from the reservoir oscillators at site n.

In turn, the equations of motion for these operators are

Equation (9)

i.e. the reservoir oscillators are driven by the motion of the spins, with the (transposed) coupling tensors ${\mathcal{C}}_{\omega }^{(n)\,\text{T}}$ governing the degree of driving for each of the continuum of oscillators. We assume retarded boundary conditions so that the reservoir responds only to the past behaviour of the spins. The retarded Green function for $({\partial }_{t}^{2}+{\omega }^{2}){G}_{\omega }=\delta (t-{t}^{\prime })$ is Gω (tt') = Θ(tt')sin(ω(tt'))/ω and can be used to solve equation (9) exactly, giving

Equation (10)

Here, ${\hat{\boldsymbol{a}}}_{\omega }^{(n)}\!$ and ${\hat{\boldsymbol{a}}}_{\omega }^{(n)\,{\dagger}}\!$ are (vectors of) bosonic ladder operators with their components obeying $[{\hat{a}}_{\omega ,j}^{(n)},{\hat{a}}_{{\omega }^{\prime },k}^{(m)\,{\dagger}}]={\delta }_{nm}\,{\delta }_{jk}\,\delta (\omega -{\omega }^{\prime })$. Classically these correspond to the two integration constants for the differential equation (9) which set the initial amplitude and velocity of the oscillator. The evolution begins at time t0, at which the reservoir operators are assumed to be independent of the spin operators and the initial constants are applied.

Substituting the reservoir solutions (10) into the equations of motion for the spins (8), we obtain the first result: the Heisenberg–Langevin equation that governs three-dimensional quantum spin dynamics under the influence of memory and coloured quantum noise is

Equation (11)

where h.c. indicates the Hermitian conjugate of the preceding expression. The term ${\hat{\boldsymbol{b}}}^{(n)}\!(t)$ is a Hermitian magnetic noise operator for site n,

Equation (12)

which plays the role of the stochastic noise first described by Brown [48]. Here it arises from the spin's interaction with its reservoir. As we will see below, the bath noise can be coloured and contain quantum zero-point fluctuations. In addition to the coloured noise ${\hat{\boldsymbol{b}}}^{(n)}\!$, a kernel tensor ${\mathcal{K}}^{(n)}\!(t-{t}^{\prime })$ appears in equation (11), which captures the damping of the spins. It arises from the coupling tensor ${\mathcal{C}}_{\omega }$ and is given by

Equation (13)

where τ = tt'. Here Θ is the Heaviside function which makes the spin's dynamics at time t, see (11), a function of the spin's state at previous times t0t' < t [114].

The three-dimensional spin dynamics equation (11) describes the evolution of spin operators with a general tensorial damping kernel and explicitly includes memory of the past dynamics (non-Markovianity). This contrasts with previous derivations of quantum spin dynamics in the form of a master equation [68] which can be solved numerically. But to obtain the master equation a range of simplifying assumptions were made, including weak spin-environment coupling, as well as the Markov and secular approximations. These approximations are quite strong and may not always be justified for a given spin system. We further remark that a different method of including (classical) coloured noise is based on the Miyazaki–Seki approach [70]. Similar to (8), this model assumes that the equation of motion of the spins is coupled to a stochastic equation of motion for the lattice enabling transfer of energy and angular momentum between the lattice and the spin systems. Different coupling potentials, such as harmonic and Morse potentials, have been considered and the spin–lattice coupling impact on the magnetisation has been characterised [71].

The above spin + environment Hamiltonian and its resulting equations of motion share many similarities with the well-known Caldeira–Leggett model for harmonic quantum Brownian motion [26, 35] and the spin-boson model [33, 37, 39]. A key difference is the three-dimensional nature of the reservoir interaction in (5) which leads to the cross product in (11). The spin-boson model is recovered as a special case for spin 1/2 operators and rank-1 coupling tensors, see appendix A1.

1.3. Fluctuation–dissipation relation and ${\hat{\boldsymbol{S}}}^{2}$

The presence of the kernel in (11) gives rise to spin damping that can significantly differ from Gilbert damping, as explored further in section 2. Previous generalisations to Gilbert damping have considered the spins' interactions with the lattice and electron motion [54, 7173], and included inertial terms [49] and further memory terms within a damping kernel [5558]. First measurements have recently confirmed the presence of inertial corrections [50, 59]. But these studies have not considered how the noise may have to change from the standard white magnetic noise that is included in the effective magnetic field in the LLG equation (1).

A separate strand of reasoning has argued that to extend the LLG equation to the quantum setting requires that the stochastic noise included in the effective field B eff should have a quantum distribution rather than the classical Boltzmann one [60, 7477]. These studies partially reproduce experimentally measured magnetization over temperature curves. But the noise considered vanishes at low temperatures, whereas the full Bose–Einstein distribution famously has non-zero quantum fluctuations even at T = 0 K. Other pioneering investigations have discussed the effect of coloured noise and non-Markovian effects on the ultrafast demagnetization rate [70], while not requiring the fluctuation–dissipation theorem (FDT) to hold and not including quantum effects.

The quantum spin dynamics equation (11) brings these pieces, the damping kernel, quantum statistics and coloured noise, together in a quantum thermodynamically consistent framework. To see this we now prove that the reservoir's two signatures—the stochastic field ${\hat{\boldsymbol{b}}}^{(n)}\!$ and memory kernel ${\mathcal{K}}^{(n)}\!$—always fulfil the (quantum) FDT, for any choice of coupling tensor ${\mathcal{C}}_{\omega }^{(n)}$.

The reservoir power spectrum $\tilde{\mathcal{P}}(\omega )$ is defined as the (quantum symmetrised [115]) expectation value of the autocorrelation function of the magnetic noise ${\hat{\boldsymbol{b}}}^{(n)}\!(t)$ in the thermal reservoir state ${\hat{\rho }}_{\mathrm{R}}={\text{e}}^{-\beta {\hat{H}}_{\mathrm{R}}}/\mathrm{t}\mathrm{r}[{\text{e}}^{-\beta {\hat{H}}_{\mathrm{R}}}]$, and then taking the Fourier transform from the time to the frequency domain [116], i.e.

Equation (14)

where {..} is the anti-commutator. Inserting (12), one finds that the only non-trivial contributions to this expression come from the bosonic ladder operator expectation values ${\langle \left\{{\hat{a}}_{\omega ,l}^{(n)},{\hat{a}}_{{\omega }^{\prime },{l}^{\prime }}^{(m)\,{\dagger}}\right\}\rangle }_{\beta }={\delta }_{nm}{\delta }_{l{l}^{\prime }}\delta (\omega -{\omega }^{\prime })\mathrm{coth}(\beta \hslash \omega /2)$. Meanwhile by equation (13) the square of the coupling tensor, ${\mathcal{C}}_{\omega }^{(n)}{\mathcal{C}}_{\omega }^{(n)\,\text{T}}$, is related to the imaginary part of the Fourier transform of the damping kernel ${\mathcal{K}}^{(n)}\!(\tau )$, i.e.

Equation (15)

Hence one obtains for all choices of the interaction tensor ${\mathcal{C}}_{\omega }^{(n)}$ in equation (5) the quantum FDT

Equation (16a)

Equation (16b)

where the last line is the well-known classical high-temperature (or low frequency) approximation, and the first line is the low-temperature limit where the Bose–Einstein distribution of the reservoir oscillator modes becomes important [117].

Finally, the spin dynamics equation (11) leaves the square of the spin operators a constant of motion, see appendix A2,

where ${\hat{\boldsymbol{B}}}^{(n)}={\boldsymbol{B}}_{\text{ext}}+{\hat{\boldsymbol{B}}}_{\text{env}}^{(n)}+\frac{1}{\gamma }{\sum }_{m\ne n}{\hspace{6pt}\bar{\hspace{-5pt}\mathcal{J}}}^{(nm)}{\hat{\boldsymbol{S}}}^{(m)}\!$ is an effective magnetic field at site n see equation (8), which commutes with all components of ${\hat{\boldsymbol{S}}}^{(n)}\!$. This confirms that the spin length $\vert \hat{\boldsymbol{S}}\vert $ is a constant in time, e.g. for a spin-1/2 $\hat{\boldsymbol{S}}=\frac{\hslash }{2}\hat{\boldsymbol{\sigma }}$ with ${\hat{\sigma }}_{j}$ the Pauli matrices, one has a constant ${\hat{\boldsymbol{S}}}^{2}=\frac{{\hslash }^{2}}{4}{\hat{\boldsymbol{\sigma }}}^{2}=\frac{3{\hslash }^{2}}{4}\,{1}_{2}$. This is consistent with micromagnetics, where the LLG equation—modified to include the effects of exchange, crystalline anisotropy, magnetostatics, and damping—conserves the total angular momentum, S 2 at every point [45].

2. Coupling functions

The general quantum spin dynamics equation (11) is specified completely by the coupling tensor ${\mathcal{C}}_{\omega }$ in (5) that sets the interaction between spins and reservoirs. Here we will show that the standard LLG equation (1) arises as a special case of (11), for a particular choice of coupling ${\mathcal{C}}_{\omega }$. We will further introduce a class of Lorentzian coupling functions which allow a systematic exploration of spin dynamics behaviours beyond the LLG equation.

For simplicity, from here on we will consider a single spin (hence dropping all lattice site superscripts) and use isotropic coupling tensors ${\mathcal{C}}_{\omega }={C}_{\omega }\,{1}_{3}$ with Cω a scalar function, and similarly for the corresponding kernels $\mathcal{K}(\tau )=K(\tau )\,{1}_{3}$ and power spectra $\tilde{\mathcal{P}}(\omega )=\tilde{P}(\omega ){1}_{3}$. These choices are appropriate for 3D materials in which all spins couple to the environment in the same manner in all spatial directions. For other materials, such as 2D layers, non-isotropic coupling tensors can be considered.

2.1. Ohmic coupling and recovery of LLG equation

In the open quantum systems literature, coupling functions that are linear in frequency are referred to as 'Ohmic', while those proportional to higher and lower powers of ω are called super- and sub-Ohmic, respectively. For the magnetic system considered here, Ohmic coupling means

Equation (17)

where ηG is a positive constant with units of kg A−2 m−2 s−1, as in equation (1), and epsilon+ is an infinitesimal positive constant which we take to zero at the end of the calculation. The corresponding kernel (13) is close to instantaneous,

Equation (18)

from which we can see that a positive value of epsilon+ ensures the response kernel KOhm is causal. For Ohmic coupling (17), the damping term in (11) reduces to

Equation (19)

i.e. Ohmic coupling results in a damping term proportional to a first order time derivative of the system variable [118]. In the magnetic case this is $\hat{\boldsymbol{S}}$, while for quantum Brownian motion it would be the oscillator position [35]. Inserting (19) into the general quantum spin dynamics equation (11), we see that it recovers (the quantum version of) the LLG equation (1) with ηG being identified as the damping parameter. This becomes the well-known classical equation in the limit of large spins. Consequently we will also refer to the Ohmic coupling function (17) as 'LLG coupling'.

Given a coupling function it is straightforward to find the corresponding power spectrum (16a) governing the time correlations of the quantum noise. For ${C}_{\omega }^{\mathsf{\text{Ohm}}}$ the quantum power spectrum is

Equation (20)

where we have taken the limit epsilon+ → 0. The Ohmic quantum power spectrum (20) is proportional to the damping parameter ηG, and tends to a linear (diverging) function of frequency in the low temperature limit ℏωkB T, see figures 2(g) and (h). However, power spectra diverging with frequency are unphysical. Coupling any system to environmental modes must go to zero at large enough frequencies, such as the interaction of spins with lattice phonons or the effect of conduction electrons scattering off the spins. A common way of integrating this physical information into Ohmic coupling is to introduce a cut-off, i.e. an upper frequency to which the coupling grows linearly, and after which it is set to 0 or decays to 0 in either algebraic or exponential form [38]. A different approach was taken in [60, 74] for the modelling of spin dynamics, where a semi-quantum Ohmic power spectrum similar to (20) was chosen, but with the zero-point noise subtracted ensuring that at T = 0 K it vanishes for all frequencies.

Figure 2.

Figure 2. Comparison of coupling functions, memory kernels and power spectra: top panels show (a) coupling function cω , (c) time dependent damping kernel k(tt'), and magnetic noise power spectrum $\tilde{p}(\omega )$ at (e) T = 200 K, and (g) T = 1 K, for parameter Set 1 (33a) (solid blue). These are for Lorentzian coupling, see Section 2.2, and given in unit-free form, see Section 2.4. The yellow shading indicates the range of typical spin dynamics frequencies ω ∈ [0, 2.5ωL]. Bottom panels show the same quantities for Lorentzian coupling with parameter Set 2 (33b) (solid red). In (a) and (b) the Lorentzian coupling functions are compared to the LLG (Ohmic) approximation (magenta dash–dotted, equation (17)). In (e)–(h) the Lorentzian power spectrum is compared to the quantum LLG approximation (cyan squares, equation (20)), and its high temperature white noise limit (magenta dash–dotted, equation (21)).

Standard image High-resolution image

In the high temperature limit kB Tℏω, equation (20) reduces to the classical power spectrum

Equation (21)

which is frequency independent (white) noise. Thus Ohmic coupling plus the high temperature approximation to the power spectrum recovers the LLG equation (1) commonly used to simulate magnetic materials over a wide range of temperatures.

As a first step to unravel how the dynamics predicted by the quantum spin equation (11) can deviate from that predicted by the LLG equation (1), one can expand the spin vector (operator) at time t' in equation (11) around the end point of integration to arbitrary order,

Equation (22)

The 0th order coefficient, κ0, is dropped here as it multiplies $\hat{\boldsymbol{S}}(t)\times \hat{\boldsymbol{S}}(t)$ which is immediately 0 for a classical spin vector S (t). For quantum spin operators, this cross-product is an anti-Hermitian operator, ${(\hat{\boldsymbol{S}}\times \hat{\boldsymbol{S}})}^{{\dagger}}=-\hat{\boldsymbol{S}}\times \hat{\boldsymbol{S}}$, which then cancels in the Hermitian equation of motion (11). The higher expansion coefficients κm>0 are proportional to the mth one-sided moment of the memory kernel

Equation (23)

where we have assumed that the dynamics has been running for some time longer than the kernel decay time, for which one can replace the initial time t0 by −.

For the Ohmic kernel only the first moment is non-zero and corresponds to the (negative) damping parameter,

Equation (24)

while ${\kappa }_{m > 1}^{\mathsf{\text{Ohm}}}=0$. This complete lack of higher moments, which would maintain a certain degree of memory in the dynamics, shows that Ohmic coupling dynamics can only be an approximation to any real dynamics. For example within magnetism, the memory-free form of damping (19) is known as Gilbert damping [43], and is almost universally used to describe magnetization dynamics through the LLG equation (1). However, 'inertial' corrections to such dynamics have been proposed [49] and their presence was recently confirmed experimentally [50].

2.2. Lorentzian coupling

Here we provide a tool to systematically study dynamics beyond the Ohmic case, allowing one to include memory and coloured noise effects in a manner consistent with the quantum fluctuation dissipation theorem (16a). We consider the class of Lorentzian coupling functions (similar to [78])

Equation (25)

where A is a coupling amplitude, with the following properties: (i) for small ω, ${C}_{\omega }^{\mathsf{\text{Lor}}}$ grows linearly with ω and can be approximated by an Ohmic coupling function, (ii) at large ω, ${C}_{\omega }^{\mathsf{\text{Lor}}}$ smoothly decays to zero, and (iii) at some intermediate frequency ω0, ${C}_{\omega }^{\mathsf{\text{Lor}}}$ has a resonant peak with some width Γ. This peak characterises the confined range and relative strength of system–environment interaction with two parameters. Alternative 'peaks' such as Gaussians or top hat functions could be considered, but here we chose the Lorentzian shape due to the fact that many expressions can be solved analytically and, as we will demonstrate in section 3, Lorentzian couplings allow us to efficiently simulate non-Markovian dynamics.

We call the above functions 'Lorentzian coupling' since the corresponding damping kernel in the frequency domain is the widely studied Lorentzian response

Equation (26)

where the imaginary part is obtained from (15) and the real part is determined using the usual Kramers–Kronig relations. In the time-domain the Lorentzian memory kernel is

Equation (27)

where ${\omega }_{1}=\sqrt{{\omega }_{0}^{2}-\frac{{{\Gamma}}^{2}}{4}}$ and Γ/2 can now be interpreted as the kernel decay rate. For the coupling function (25), the collective response of the environment is thus equivalent to a single harmonic oscillator of resonant frequency ω0 and damping rate Γ/2 [79]. From the quantum FDT (16a) it follows that the corresponding power spectrum is

Equation (28)

which takes its largest values at frequencies close to ω0 and tends to zero as ω−3 at large ω. This power spectrum differs from classical Ohmic noise (21) in two important respects. Firstly, the quantum mechanical treatment means that the low temperature noise is not proportional to temperature, and does not vanish at zero temperature. Secondly, even in the high temperature limit, the noise spectrum is frequency dependent (coloured), unlike the white noise of equation (21). The presented theory thus captures both of these aspects in a consistent quantum thermodynamic framework.

Unlike Ohmic coupling (17) which depends on a single parameter ηG, the Lorentzian coupling function (25), and hence its kernel and spectrum, depends on three parameters. These allow a systematic study of different regimes of the environment response. Specifically, the memory time of the environment can be continuously varied by changing ω0 and Γ, which can lead to very different magnetic behaviour. Beyond the spin dynamics explored here, the proposed Lorentzian coupling may also be a useful tool for the characterisation of quantum Brownian motion of a variety of systems, including oscillators and free particles.

2.3. Two coupling regimes

To better understand the relation between the Ohmic and Lorentzian coupling functions, one may consider their kernel moments κm in expansion (22). In contrast to the Ohmic case, for the Lorentzian kernel (27) all κm are non-zero and given by

Equation (29)

The first relevant two moments are

Equation (30)

and when comparing to the Ohmic case, one finds that the first Lorentzian moment can be identified with minus the damping parameter ηG, i.e. ${\kappa }_{1}^{\mathsf{\text{Lor}}}=-\frac{A{\Gamma}}{{\omega }_{0}^{4}}=-{\eta }_{\mathrm{G}}$. For a material with a given damping parameter ηG this fixes one of the Lorentzian parameters, i.e.

Equation (31)

which now only depends on the two parameters ω0 and Γ. For a specific material these may be approximately determined through information contained in the density of states of the environment to which the spins couple [80].

Inserting expansion (22) with Lorentzian moments (29) in the quantum spin dynamics equation (11) one can distinguish two different dynamical situations.

Ohmic regime. When the resonant frequency ω0 and the damping rate Γ of the reservoir coupling is much larger than the spin operators' typical frequency of motion, ω0ω, each successive term in expansion (22) is smaller by an extra factor of (ω/ω0). In the limit of infinite ω0 but finite damping parameter ηG, the Lorentzian damping term in (11) thus tends to the Ohmic one (19), $-{\gamma }^{2}{\eta }_{\mathrm{G}}\hat{\boldsymbol{S}}(t)\times {\partial }_{t}\hat{\boldsymbol{S}}(t)$.

Non-Ohmic regime. When the environmental frequency ω0 is comparable to typical spin motion frequencies, ω0ω the Ohmic approximation to the Lorentzian kernel begins to fail. The first deviation in (11) is a new term proportional to ${\kappa }_{2}^{\mathsf{\text{Lor}}}$. I.e. in addition to the Gilbert damping term, one adds the term ${\gamma }^{2}{\kappa }_{2}^{\mathsf{\text{Lor}}}\hat{\boldsymbol{S}}(t)\times {\partial }_{t}^{2}\hat{\boldsymbol{S}}(t)$ containing a second time derivative of the spin operator $\hat{\boldsymbol{S}}$. By analogy with the classical equation of motion for a massive body, this is known as an 'inertial' modification to the spin dynamics [49]. As the ratio κ2/κ1 has the dimensions of time, one may introduce an 'inertial timescale' τin [49], which for the Lorentzian is

Equation (32)

A large inertial time τin indicates the presence of non-Markovian dynamics, i.e. dynamics that has a certain degree of memory. For a high quality factor resonance ω0 ≫ Γ the inertial time is (half) the kernel's decay time, τd = 2/Γ. For increasing resonance width Γ, the inertial time τin decreases and memory effects become less important. In magnetic systems this timescale determines the time over which nutation oscillations are observed in the precession of the spin. Such inertial corrections to standard magnetism have very recently been observed for the first time in ultrafast experiments on thin films [50]. Curiously, the inertial timescale, see equation (32), becomes negative when Γ > ω0, a fact that may be the subject of future investigation.

Similarly to the kernel expansion (22), one can also expand the Lorentzian power spectrum in frequency, see appendix A3. Generally, only the odd moments κ2m+1 appear in the power spectrum. This implies that if a kernel only has non-trivial first (κ1) and second (κ2) moments, while higher moments vanish (κm>2 = 0) then the power spectrum will still be given by the (quantum) Ohmic one (20), with ηG = −κ1. When third or higher moments are non-zero, then the power spectrum of the noise will deviate from the Ohmic case at all temperatures.

To summarise, here we have demonstrated that Lorentzian coupling functions, kernels and power spectra provide a systematic framework to explore system dynamics that arises from inertial terms and other memory effects, while recovering the standard Ohmic limit whenever the Lorentzian resonance frequency ω0 is much larger than the typical system frequencies.

2.4. Unit-free variables and Lorentzian parameters

In section 3 we perform semi-classical simulations of the dynamics of equation (11), for the Lorentzian kernel of section 2.2. For this purpose we re-write expressions in the operator equation (11) in terms of a unit-free set of quantities. The time coordinate t is replaced with the unit-free coordinate ωL t, where ωL = |γ|| B ext| is the Larmor frequency. In addition the spin operator $\hat{\boldsymbol{S}}$ with largest eigenvalue S0 is re-written in terms of a unit-free operator $\hat{\boldsymbol{s}}$ with largest eigenvalue 1, $\hat{\boldsymbol{S}}=\mathrm{sign}(\gamma ){S}_{0}\hat{\boldsymbol{s}}$. The sign of the gyromagnetic ratio is included in the definition of $\hat{\boldsymbol{s}}$ so that $\hat{\boldsymbol{s}}$ aligns with the magnetic field, whatever the sign of γ.

From equation (11) we can see that the damping kernel has dimensions of magnetic field squared divided by angular momentum, which leads us to identify a unit-free damping kernel k(tt') through $K(t-{t}^{\prime })=\vert {\boldsymbol{B}}_{\text{ext}}{\vert }^{2}{S}_{0}^{-1}k(t-{t}^{\prime })$. Similarly, the unit-free coupling function cω is defined through ${C}_{\omega }=\vert {\boldsymbol{B}}_{\text{ext}}\vert {S}_{0}^{-1/2}{c}_{\omega }$ and the unit-free spectral functions $\tilde{p}$ through $\tilde{P}=\hslash \vert {\boldsymbol{B}}_{\text{ext}}{\vert }^{2}{S}_{0}^{-1}{\omega }_{\mathrm{L}}^{-1}\tilde{p}$.

Looking at the Lorentzian kernel K(tt') in (26), the pulling of dimensions can be achieved by setting the kernel amplitude to $A=\vert {\boldsymbol{B}}_{\text{ext}}{\vert }^{2}{S}_{0}^{-1}\alpha $ where α now is a frequency. For the simulations we choose the Lorentzian parameters ω0, Γ and α to be independent of spin length S0. Through (5) this implies an interaction energy scaling of ${\hat{V}}_{\text{int}}\propto {S}_{0}\sqrt{A}\propto \sqrt{{S}_{0}}$ which sets the scaling of the interaction versus self-energy to ${\hat{V}}_{\text{int}}/{\hat{H}}_{\mathrm{S}}\propto 1/\sqrt{{S}_{0}}$. Apart from it being implied by dimensional analysis, such scaling is heuristically plausible in many physical contexts. E.g. it is similar to the increasing ratio of the surface (where reservoir interaction occurs) to volume (self-energy) for decreasing system sizes. For microscopic systems for which ${\hat{V}}_{\text{int}}$ is no longer small in comparison to ${\hat{H}}_{\mathrm{S}}$ a thermodynamic treatment beyond the weak coupling limit [26, 27, 81], a limit tacitly assumed in standard thermodynamics, may be required.

Similarly for Ohmic coupling leading to the LLG equation (1), the above scaling choice amounts to choosing ηGA ∝ 1/S0 implying that $\eta ={\eta }_{\mathrm{G}}{\gamma }^{2}{S}_{0}=\alpha {\Gamma}{\omega }_{\mathrm{L}}^{2}/{\omega }_{0}^{4}$ is assumed to be independent of S0. In physical situations where this assumption is not justified, one may instead choose η and α to depend on the spin length S0.

For a spin in an external field B ext the typical frequency of the dynamics is set by the Larmor frequency, ωL. In the simulations discussed in section 3 we will use the following two sets of Lorentzian parameters, all expressed in terms of ωL,

Equation (33a)

Equation (33b)

In the second rows we have also listed the equivalent unit-free Gilbert damping η, the inertial timescale τin, and the memory kernel decay time τd for the Lorentzian kernel. Figures 25, show plots obtained with Lorentzian coupling functions (25) with parameter Sets 1 and 2, which are shown in blue and red, respectively. The Ohmic LLG approximation is shown in magenta when the classical reservoir (21) is considered, and in cyan when the quantum reservoir (20) is considered. The external field is set to B ext = 10 T e z with e z the unit-vector in z-direction throughout.

Figure 3.

Figure 3. Sample of stochastic short-time spin dynamics for different couplings and noises. Stochastic short-time dynamics of sz (blue in top row & red in bottom row), sx (green) and | s | (black dashed) according to equation (11) with Lorentzian coupling function ${C}_{\omega }^{\mathsf{\text{Lor}}}$, see (25), for a classical spin initially in state s = (−1, 0, 0). All traces in the four panels are generated from the same sample of Gaussian noise, enabling a direct comparison. Shown are the dynamics for Set 1 (top row) and Set 2 (bottom row), and two spin + temperature pairs: S0 = 1/2 and T = 1 K (left column), and S0 = 200/2 and T = 200 K (right column). Also shown in all four panels are the sz -dynamics according to the LLG equation (1) with damping parameter η ≈ 0.02 for two types of noise: the classical flat white noise power spectrum equation (21) (magenta) and the quantum noise power spectrum equation (20) (cyan). All cyan and magenta plots are off-set by +0.05 and +0.1, respectively, to avoid overlapping. The external magnetic field is set to B ext = (0, 0, 10 T) setting the timescale to ${\omega }_{\mathrm{L}}^{-1}\approx 0.57\times 1{0}^{-12}\enspace \text{s}$, and the simulation time interval is $\mathrm{d}t=0.15{\omega }_{\mathrm{L}}^{-1}$.

Standard image High-resolution image
Figure 4.

Figure 4. Ensemble averaged spin relaxation dynamics. Averaged dynamics ⟨sz ⟩, averaged over 500 stochastic traces up to time ${t}_{\mathrm{max}}=2\pi \times 48{\omega }_{\mathrm{L}}^{-1}$. Shown are the dynamics according to equation (11) with Lorentzian coupling functions (25) for parameter Set 1 (blue) and Set 2 (red). Also shown is the ensemble averaged dynamics according to the LLG equation (1) with damping parameter η ≈ 0.02 for two types of noise: classical white noise (magenta) and full quantum noise (cyan). The two panels show two spin + temperature pairs: S0 = 1/2 and T = 1 K (left), and S0 = 200/2 and T = 200 K (right). Note that blue, cyan and magenta curves lie on top of each other in (b). As discussed in section 2.4, the plots assume that both A and ηG scale as $\propto {S}_{0}^{-1}$ with spin size S0, making α and η the same for the two spin sizes. The external magnetic field is set to B ext = (0, 0, 10 T) and the simulation time interval is $\mathrm{d}t=0.15{\omega }_{\mathrm{L}}^{-1}$.

Standard image High-resolution image
Figure 5.

Figure 5. Steady state ⟨sz ⟩ of a spin interacting with a (quantum) reservoir at temperature T. For a spin in an external magnetic field B ext = (0, 0, 10 T) and interacting with a thermal reservoir, the time averaged value of ⟨sz ⟩ is obtained by integrating equation (11) for Lorentzian parameter Set 1 (blue crosses) and Set 2 (red dots), as well as by integrating the LLG equation (1) with damping parameter η = 0.02 for two types of noise: classical white noise (magenta dash–dotted, Eq. (21)) and full quantum noise (cyan dashed, Eq. (20). The spin lengths are S0 = 1/2 (panels (a) and (b)), and S0 = 200/2 (panel (c)). For the small spin in panel (a), the three curves with quantum noise (cyan, blue, red) start at ⟨sz ⟩-values in the range 0.2–0.4 at T = 0 K. For both S0 values, the steady state for the LLG equation with classical noise (magenta) coincide well with ${\langle {s}_{z}\rangle }_{\mathsf{\text{stat phys}}}$ (black), derived from classical statistical mechanics for a thermal distribution, see appendix A5. Note that all curves lie on top of each other in panel (c). Panel (b) shows the same plot as panel (a)—but with the y-axis 'rescaled' as m(T) = ⟨sz (T)⟩/⟨sz (0)⟩ so that all plots start at 1 at T = 0 K. While the magenta curve remains the same as in (a), the rescaled blue, red and cyan curves now show a flattened decay behaviour that somewhat resembles the corrected magnetization curves for real materials analysed in [67]. Error bars for the four simulations (blue, red, cyan, magenta) are indicated at (a) T = 1 K and (c) T = 200 K. The simulation time interval is $\mathrm{d}t=0.15{\omega }_{\mathrm{L}}^{-1}$.

Standard image High-resolution image

Parameter Set 1 has been chosen to have a resonant frequency ω0 much larger than the characteristic spin precession frequency ωL (Ohmic regime). Consequently we can truncate the series given in equation (22) to leading order and recover the Ohmic form (19) typically considered in magnetism theory. The validity of this approximation is demonstrated in the top row of figure 2. Figure 2(a) shows that the Lorentzian coupling function ${c}_{\omega }^{\mathsf{\text{Lor}}}$ is well approximated by LLG (Ohmic) coupling for the relevant frequency range, while figure 2(c) shows that the kernel is approximately instantaneous on the timescale ${\omega }_{\mathrm{L}}^{-1}$, in line with Ohmic damping for which ${\kappa }_{m > 1}^{\mathsf{\text{Ohm}}}=0$. Figure 2(e) shows that at high temperature (T = 200 K) and for relevant frequencies ωωL, the power spectrum ${\tilde{p}}^{\mathsf{\text{Lor}}}$ is well approximated by the quantum Ohmic (LLG) power spectrum (20), and its classical limit (21). Figure 2(g) shows that at lower temperatures (T = 1 K) quantum noise becomes important. Here the Ohmic approximation (20) remains valid, while its classical limit (21) is invalid.

Parameter Set 2 is chosen such that it has a resonant frequency ω0 comparable to the precession frequency ωL (non-Ohmic regime). Here it is inaccurate to truncate the series (22) and the damping will be fundamentally non-Ohmic. To directly compare with Ohmic dynamics generated by Lorentzian coupling with Set 1, both parameter sets have been chosen to correspond to the same unit-free Gilbert damping parameter, η = 0.02. The failure of the Ohmic approximation is demonstrated in the bottom row of figure 2. Figure 2(b) shows that the linear approximation to the coupling function fails in the relevant frequency range. For this set of parameters the damping kernel now exhibits significant memory and figure 2(d) shows that the response persists over a timescale of several ${\omega }_{\mathrm{L}}^{-1}$. This memory kernel implies, through the FDT (16a), a coloured quantum noise power spectrum ${\tilde{p}}_{\mathsf{\text{qu}}}^{\mathsf{\text{Lor}}}$. As shown in figure 2(h), this coloured Lorentzian ${\tilde{p}}_{\mathsf{\text{qu}}}^{\mathsf{\text{Lor}}}$ (red) differs from the LLG counterpart, ${\tilde{p}}_{\mathsf{\text{qu}}}^{\mathsf{\text{Ohm}}}$ (cyan). Furthermore, figure 2(f) shows that in the high temperatures ${\tilde{p}}_{\mathsf{\text{qu}}}^{\mathsf{\text{Lor}}}$ (red) also differs very significantly from the LLG power spectrum ${\tilde{p}}_{\mathsf{\text{cl}}}^{\mathsf{\text{Ohm}}}$ (magenta). The presence of both memory and coloured quantum noise for the Lorentzian with parameter Set 2 are both signatures of a thermostat that substantially deviates from the classical Ohmic assumptions and, as we will see in the next section, leads to markedly different short time dynamics and steady state of sz .

3. Semi-classical spin dynamics simulations

The general spin dynamics equation (11) is an operator equation for quantum spins in a lattice, each interacting with neighbouring spins and with a bosonic reservoir. Solving the quantum dynamics using, for example, Lorentzian coupling, kernel and spectrum, is rather difficult without approximations, even numerically, and such exploration is left for future work.

To make progress here, we will numerically solve the full non-Markovian dynamics for a semi-classical version of equation (11), while including coloured quantum noise and memory effects arising from the coupling to the environment. In this approximation, the quantum spin operator $\hat{\boldsymbol{S}}$ is replaced by a classical spin vector S , and the quantum stochastic noise field vector $\hat{\boldsymbol{b}}$ is replaced by a stochastic classical field vector b with statistics that obey the quantum FDT (16a). This semi-classical approach is currently used in the theory of molecular and ionic dynamics [82, 83], and was perhaps first applied by Koch et al [84] to include the effects of quantum fluctuations in Josephson junctions. It has been justified through an expansion of a forward–backward path integral [85, 86] (note the remark of Caldeira and Leggett on p 589 of [35]), and is valid when the potential energy can be expanded in the path integral to first order in the deviations from the average path. The validity of applying this approach to the decay of metastable states was investigated in detail in [87].

Here we simulate a single spin allowing us to illustrate the impact of memory effects and the reservoir's quantum statistics on the spin dynamics and steady state. The simulation details presented below can readily be extended to multiple interacting spins and could be integrated in sophisticated atomistic spin dynamics simulations such as those used in [47, 60].

3.1. How to simulate coloured noise and memory kernel

Here we detail how to efficiently simulate non-Markovian dynamics that arises as a result of Lorentzian coupling (25) for the example of spins vectors. Numerical implementation of (11) requires both—the integration of the kernel with the spin state of previous time steps and the inclusion of coloured noise as follows.

Dropping the spin index and for simplicity assuming any isotropic kernel $\mathcal{K}(\tau )={1}_{3}\,K(\tau )$, the three vector components bj (t) for j = 1, 2, 3 of the magnetic noise (12) are implemented as [88]

Equation (34)

where ξj (t') is standard white Gaussian noise for the jth component, which is delta correlated ⟨ξj (t) ξk (t')⟩ = δjk δ(tt'). The 'coloured noise' comes from choosing F(tt') as the Fourier transform of the square root of the power spectrum associated with the kernel through (16a), i.e.

Equation (35)

which can be implemented using a fast Fourier transform. To simulate the effect of a Lorentzian damping kernel (26) we numerically integrate [89] the following set of first order coupled differential equations for the spin vector S and two dummy vectors V and W :

Equation (36)

The integrated values of the dummy vectors and the spin are separated by the time step dt. Solving these equations is equivalent to solving the integro-differential equation (11) for a Lorentzian kernel, see appendix A4, but is numerically more straightforward to implement.

3.2. Single trajectories for different couplings and noises

We wish to illustrate on a single trajectory level, the differences between the dynamics predicted by (11) with either an approximately Ohmic (Set 1) or non-Ohmic (Set 2) Lorentzian coupling function, as well as the dynamics predicted by the standard LLG equation. At first, because the dynamics is intrinsically stochastic, trajectories will naturally differ and cannot readily be compared. However, looking at the noise generation in equations (34) and (35), one can see that the same white noise ξj (t) for j = 1, 2, 3 may be used as a seed to create comparable 'stochastic' noise for different power spectra $\tilde{P}(\omega )$.

Figure 3 shows the stochastic short time dynamics of a single classical spin for two pairs of spin length and temperature, S0 = 1/2 at T = 1 K (left panel) for a single electron, and S0 = 200/2 at T = 200 K (right panel) for a mesoscopic cluster of spins with a combined larger effective spin.

The dynamics is obtained according to equation (11) for Lorentzian coupling (25) with S0-scaling $A=\vert {\boldsymbol{B}}_{\text{ext}}{\vert }^{2}{S}_{0}^{-1}\alpha $, for parameter sets Set 1 (top panel, blue) and Set 2 (bottom panel, red), and with the quantum coloured noise given by (28). For comparison we also show the short time dynamics according to the LLG equation (1) with S0-scaling ${\eta }_{\mathrm{G}}=A{\Gamma}/{\omega }_{0}^{4}=\vert {\boldsymbol{B}}_{\text{ext}}{\vert }^{2}{S}_{0}^{-1}{\omega }_{\mathrm{L}}^{-2}\eta $ with the Gilbert damping parameter η = 0.02 common to both Lorentzian parameter sets, see (33a). That implies that the top and bottom LLG plots are identical. For the standard LLG equation two types of noise are considered—high-temperature classical noise (magenta) see equation (21), and quantum noise (cyan) see equation (20). Since the same white noise time series is used as a seed for producing the stochastic noise for all traces, we can compare them directly. We will here focus on sz , the component of s = sign(γ) S /S0 aligned with the external field B ext.

Three features stand out in figure 3: (i) as expected from section 2, the dynamics generated with equation (11) with Lorentzian Set 1 (top, blue) closely matches the dynamics obtained with the LLG equation (1) with quantum noise (cyan) for both spin–temperature pairs, (ii) the quantum statistics of the reservoir (cyan) at low temperatures (left) introduces differences to the LLG dynamics compared to the LLG dynamics obtained with classical noise (magenta), and (iii) memory effects that are present for Lorentzian Set 2 (bottom, red) result in significantly different dynamics from that arising with the memory-free Lorentzian Set 1 (top, blue).

We remark that due to the spin/temperature ratio being the same for the two spin–temperature pairs, the LLG equation with classical noise (magenta) integrates to exactly the same dynamics in left and right panel, see appendix A7. This scaling relation ceases to be true for the LLG equation with quantum noise (cyan). Another difference to note is that in figures 3(a)–(d) the dynamics for Lorentzian parameter Set 1 (blue) varies more rapidly in time than for Lorentzian parameter Set 2 (red). This is due to the high frequency content of Set 1's power spectrum, see figures 2(e) and (g).

Finally, the spin component sx (green) and the spin-vector length | s | (black) are shown for the Lorentzian coupling with Set 1 and Set 2 in the top and bottom panels of figure 3, respectively. The plots of | s | show that the numerical integration of equation (11) indeed leads to a constant spin-vector length | s | = 1, i.e. no renormalisation is required.

3.3. Ensemble-averaged ⟨sz ⟩ trajectories

Figure 4 shows the ensemble averaged ⟨sz ⟩ over time, averaged over 500 stochastic trajectories. We now highlight two important features in figure 4. Firstly, at low temperatures (left) the quantum statistics of the reservoir (blue, red, cyan) results in a much depleted value of ⟨sz ⟩, roughly at around 0.28, in comparison to that obtained with the LLG equation with classical noise (magenta), ca 0.85. This indicates that for this particular choice of spin length and temperature the quantum character of the reservoir strongly affects the value of ⟨sz ⟩, as further discussed below, and the classical high-temperature limit taken in (16b) would not be appropriate. For the high temperature T = 200 K + larger spin pair S0 = 200/2 (right), the difference between classical and quantum statistics of the reservoir can be neglected and ⟨sz ⟩ settles at 0.85 independent of whether the spin dynamics integration was done for equation (11) with either Set 1 (blue) or Set 2 (red), or for equation (1) with either classical (magenta) or quantum noise (cyan).

Secondly, for the large spin–temperature pair (right), there is clear evidence of a much quicker relaxation to steady state (by a factor of a third) for Lorentzian Set 2 (red) compared to the other plots (blue, cyan, magenta). This is a non-Markovian effect that arises because the memory kernel for Set 2 has an appreciable memory over time, see figure 2(d), while the other memory kernels are (close to) instantaneous. This quicker equilibration occurs because the non-Markovian kernel leads to a smoother dynamics which in turn is more quickly sampled by the dynamical system.

3.4. Steady state ⟨sz ⟩ as a function of temperature

Figure 5 shows the average steady state spin value ⟨sz ⟩ as a function of temperature T, found by time-averaging a single trajectory sz over late times, from 0.75 tmax to ${t}_{\mathrm{max}}=2\pi \times 7200{\omega }_{\mathrm{L}}^{-1}$. There are two key observations to make in figure 5. Firstly, for both the small spin (left) and the large spin (right) the steady state ⟨sz ⟩ obtained with the LLG equation and classical noise (21) matches the standard statistical physics prediction ${\langle {s}_{z}\rangle }_{\mathsf{\text{stat phys}}}=\mathrm{coth}\left(\frac{{S}_{0}{\omega }_{\mathrm{L}}}{{k}_{\mathrm{B}}T}\right)-\frac{{k}_{\mathrm{B}}T}{{S}_{0}{\omega }_{\mathrm{L}}}$, see appendix A5.

Secondly, for simulations that include the full quantum noise (cyan, blue, red) in the dynamics of the small spin at low temperatures (left), we observe reduced ⟨sz ⟩ values in the range 0.2–0.4 at T = 0 K, i.e. well below the classical value of 1. This arises because the power spectrum ${\tilde{P}}_{\mathsf{\text{qu}}}^{\mathsf{\text{Lor}}}$, given through the quantum FDT (16a), includes quantum fluctuations which remain even for T → 0 K. The steady state curves with quantum noise also show a characteristic 'flattening' compared to the steep decay of the steady state with temperature for classical noise (magenta). Qualitatively speaking, this quantum zero point noise, when compared to classical noise, is as if thermal noise is 'on' even at 0 K. I.e. taking the Larmor frequency as the relevant frequency, and setting ${\tilde{P}}_{\mathsf{\text{qu}}}^{\mathsf{\text{Ohm}}}(0\enspace \text{K})={\tilde{P}}_{\mathsf{\text{cl}}}^{\mathsf{\text{Ohm}}}({T}_{\mathsf{\text{cl}}})$ for the Ohmic coupling, for example, defines a classical temperature of ${T}_{\text{cl}}=\frac{\hslash {\omega }_{\mathrm{L}}}{2{k}_{\mathrm{B}}}$ = 6.7 K 'equivalent' to the quantum zero-temperature case. For S0 = 1/2 the classical statistical physics steady state value at Tcl is ${\langle {s}_{z}\rangle }_{\mathsf{\text{stat phys}}}\approx 0.3$. This indeed is of comparable size to the ⟨sz ⟩ values obtained with quantum noise at T = 0 K. The corresponding steady state value for the large spin (S0 = 200/2) is ${\langle {s}_{z}\rangle }_{\mathsf{\text{stat phys}}}\approx 0.995\approx 1$, see figure 5(c). Generally, for the classical temperature ${T}_{\text{cl}}$ one obtains ${\langle {s}_{z}\rangle }_{\mathsf{\text{stat phys}}}=\mathrm{coth}\left(\frac{2{S}_{0}}{\hslash }\right)-\frac{\hslash }{2{S}_{0}}$, which only depends on the spin length S0 while being independent of the field strength | B ext|. With increasing S0 this function rises very sharply from $\approx 0.3$ to 1. For example for spin length S0 = 5/2, the quantum zero temperature value is ${\langle {s}_{z}\rangle }_{\mathsf{\text{stat phys}}}\approx 0.8$ and its decay with increasing temperate is shown in figure 6(b) in appendix A6.

The middle panel, figure 5(b), gives an alternative illustration of the steady state value for the small spin as a function of environment temperature T. It shows the same plot as panel (a), but with the y-axis rescaled as m(T) = ⟨sz (T)⟩/⟨sz (0)⟩. All plots now start at 1 at T = 0 K, independent of whether the dynamics was integrated with quantum or classical noise. Interestingly, the overarching behaviour of the resulting curves bears some resemblance with heuristically rescaled magnetization curves that match experimental data [67, 90]. Running high-end atomistic simulations of equation (11), instead of (1), for multiple interacting classical spins would answer if the quantum power spectrum's impact on their low temperature magnetisation behaviour as well as their Curie temperature is the reason for the apparent rescaling.

We note that larger numerical uncertainties arise for the quantum noise because an additional scale is present in comparison to classical noise, see appendix A7. Error bars obtained from an ensemble of simulations are indicated at one temperature value in (a) and (c) for all four ⟨sz ⟩ curves in figure 5.

4. Conclusions and open questions

We have derived a general quantum spin dynamics equation, (11), capable of describing three-dimensional precession and damping. The terms arising from the reservoir interaction are treated in a quantum thermodynamically consistent manner, by tracing the origin of both the memory kernel, $\mathcal{K}(\tau )$, and the stochastic noise, $\hat{\boldsymbol{b}}(t)$, to a single coupling function, Cω . Secondly, Lorentzian coupling functions were proposed and shown to provide a systematic means to investigate different dynamical regimes—from Ohmic to non-Ohmic dynamics which is subject to memory and coloured noise. We showed that only in the Ohmic regime, the standard LLG equation with Gilbert damping, widely used in magnetism, is recovered. Finally, we provided details of how to include Lorentzian memory and coloured noise in numerical simulations of open system dynamics. For the example of a single spin vector, we illustrated that a non-Ohmic Lorentzian kernel leads to a faster equilibration time of ⟨sz ⟩ in comparison to the Ohmic (LLG) regime. We also discussed the steady state differences that arise when the full quantum thermostat with quantum zero-point noise is employed, in contrast to classical white noise.

The above three ingredients provide a complete framework for the simulation of damped three-dimensional precession including memory and coloured noise. It can readily be adapted in atomistic spin dynamics simulations [47, 60] that solve the dynamics of millions of interacting spins.

The theory presented here will be a useful tool for investigating non-Markovian behaviour, opening up a number of avenues for future research at the intersection of quantum thermodynamics, magnetic materials and beyond. For example, it is an open question to clarify the connection between the three-dimensional precession described by the spin equation (11) and rotational Brownian motion. The orientation of a non-symmetric rotating body behaves analogously to the three-dimensional spin vector, and the motion and viscosity of a gas surrounding a rotating body simultaneously act on its motion while obeying the FDT as discussed in recent work [91, 92].

Within magnetism, for particular materials of interest, detailed models of the coupling functions Cω can be developed that are based on an understanding of the interactions between spins, phonons, and electrons in the material [80]. Coupling to optical modes may further be included to describe, for example, whispering gallery photon–magnon coupling which leads to an effective Gilbert damping term that can take either sign [93]. A direct experimental characterisation of a material's damping kernel $\mathcal{K}$ that determines memory and noise in (11) may be attempted, for example with high field experiments such as those recently reported in [50]. Of particular interest are dynamical features beyond the inertial kernel approximation, which will also modify the noise spectrum at larger temperatures.

While we here discussed scalar couplings to the environment in depth, equation (11) does hold for any real 3 × 3 matrix ${\mathcal{C}}_{\omega }$ describing the spin–environment interaction in three dimensions. Anisotropic coupling tensors can be implemented, suitable for describing magnetization dynamics within thin films [62], where one direction is coupled differently to environmental modes than the other two. One simplification of our three-dimensional model is to choose a coupling tensor such that spins interact with only one-dimensional environmental modes. This reduces the theory to the spin-boson model, see appendix A1, whose quantum thermodynamic properties have been discussed very extensively, recently for example in [33].

Microscopic heat transport in spin systems can also be analysed by allowing non-equilibrium situations where individual reservoir modes at frequencies ω and for spins n are thermal—but at different temperatures. This will result in spin dynamics that shuffles energy from one reservoir mode to another, and could result in two- and more-temperature models. For example, the possibility of different phonon modes, each with their own temperature, to couple with different strengths to electrons has recently been analysed in [94] for a magnetic system excited by an ultra-short laser pulse. Furthermore, spin-torque effects arising, e.g. from spin-polarised leads can be included, as demonstrated in [95]. Furthermore, in deriving the FDT we have assumed bosonic environmental modes but it would be insightful to identify changes to the properties of equation (11) that arise when the spins couple directly to electrons, or fermionic modes in general [14, 15, 96].

Beyond the quantum character of the reservoir, it will be important to numerically solve the full quantum dynamics according to equation (11), including spin operators interacting with neighbouring spin operators. Advanced quantum numerical methods such as hierarchical equations of motion [97], and the recently proposed time-evolving matrix product operators method [98] will be required to efficiently describe the time evolution of even just a single quantum spin coupled to a non-Markovian environment. For multiple interacting spins at low temperatures one can expect entanglement between the spins being present during the short-time dynamics, and even in steady state [99101]. Unfortunately, evaluating such properties will very quickly become a numerically hard problem, requiring advanced numerical techniques such as density-matrix renormalisation group [102] to find realistic approximate solutions. Vice versa, solving (11) within the classical spin vector approximation while including a full quantum power spectrum for the environmental modes, may prove insightful and numerically tractable in the context of finding suitable models for noise in quantum computing hardware, such as superconducting qubits that are held in the mK range [103]. The results may also inform implementations of Young's double slit experiment with a levitated single magnetic domain nanoparticle using the Einstein–de Haas effect [104, 105].

Acknowledgments

We thank Karen Livesey, Richard Evans, Marco Berritta, Stefano Scali, Federico Cerisola, Luis Correa, James Cresser, Somayyeh Nemati, Claudia Clarke, Ian Ford and Rob Hicken for inspiring discussions, and Carsten Henkel and Richard Evans for comments on a draft of this manuscript. SARH thanks Paul Kinsler for pointing out the stupidity of numerically solving an integro-differential equation when an ODE will do. SARH also acknowledges funding from the Royal Society and TATA (RPG-2016-186). CRJS and JA acknowledge support and funding from the EPSRC Centre for Doctoral Training in Electromagnetic Metamaterials EP/L015331/1. JA acknowledges funding from EPSRC (EP/R045577/1) and the Royal Society.

Data availability statement

The data shown in figures 26 are available upon reasonable request to JA, janet@qipc.org.

: Appendix

A1. Recovery of the spin-boson model

The spin-boson model is recovered as a special case of the three-dimensional Hamiltonian (2), and hence its dynamics is also given by equation (11). To see this one may drop the site index n, and choose the external field as

Equation (37)

for some angle θ. Taking spin 1/2 operators $\hat{\boldsymbol{S}}=(\hslash /2)\boldsymbol{\sigma }$ and the coupling tensor as ${({\mathcal{C}}_{\omega })}_{jk}={C}_{\omega }\,{\delta }_{jk}{\delta }_{j1}$ with a scalar coupling function Cω , one recovers from (2) the one-dimensional spin-boson Hamiltonian

Equation (38)

where ${\hat{H}}_{\mathrm{R}}^{2\mathrm{D}}$ is a decoupled two-dimensional reservoir that can be dropped from the dynamics.

A2.  ${\hat{\boldsymbol{S}}}^{2}$ is a constant of motion of equation (11)

To evaluate the derivative of ${({\hat{\boldsymbol{S}}}^{(n)}\!(t))}^{2}$ we abbreviate the effective magnetic field operator at site n in equation (8), as

Equation (39)

Using the definition of ${\hat{\boldsymbol{B}}}_{\text{env}}^{(n)}$ given below equation (8), it is clear that ${\hat{\boldsymbol{B}}}^{(n)}$ is a Hermitian operator that commutes with the spin operators ${\hat{\boldsymbol{S}}}^{(n)}\!$. Dropping the site index and time argument to avoid index overload, we find from (11)

Equation (40)

where the final equality follows because of the symmetry of the operator ${\hat{S}}_{j}{\hat{S}}_{k}+{\hat{S}}_{k}{\hat{S}}_{j}$ under exchange of the indices jk.

A3. Lorentzian power spectrum expansion

Similar to the damping kernel term expansion (22), in moments (29) and time-derivatives, the Lorentzian power spectrum (28) can be expanded in powers of frequency ω, as

Equation (41)

where we have kept the quantum coth unexpanded. The ${\kappa }_{m}^{\mathsf{\text{Lor}}}$ are the same coefficients as those given in (29). For small frequencies ω the first term in the series (41) dominates and the power spectrum takes the (quantum) Ohmic form

Equation (42)

where comparison with (20) again shows that $-{\kappa }_{1}^{\text{Lor}}$ is the effective Gilbert damping constant.

Beyond the Ohmic regime, one can see in (41) that only the odd moments ${\kappa }_{2m+1}^{\mathsf{\text{Lor}}}$ contribute. Therefore the inertial term ${\kappa }_{2}^{\mathsf{\text{Lor}}}$, which is the first deviation of the damping kernel from Ohmic behaviour, does not change the quantum fluctuations in (16a). Only when the third order time derivative of the spin operator contributes significantly to equation (11), will memory effects begin to colour the spectrum away from the (quantum) Ohmic form (20).

A4. Set of equations for kernel simulation

Here we show that the simulation of the kernel in equation (11) can be achieved by numerically integrating a set of first order coupled differential equations. We assume a single spin and rewrite equation (11) as

Equation (43)

where we have defined $\boldsymbol{V}(t)=\gamma {\int }_{{t}_{0}}^{\infty }\!\mathrm{d}{t}^{\prime }\,K(t-{t}^{\prime })\,\boldsymbol{S}({t}^{\prime })$. Furthermore we define$\boldsymbol{W}(t)=\frac{\mathrm{d}\boldsymbol{V}(t)}{\mathrm{d}t}$, and note that we may choose the initial values as V(t0)=0 and W (t0 )=0 for the Lorentzian kernel (27). Expressing K(tt') through its Fourier transform $\tilde{K}(\omega )$, choosing a Lorentzian kernel (26) and considering the expression $\boldsymbol{Z}(t){:=}\frac{\mathrm{d}\boldsymbol{W}(t)}{\mathrm{d}t}+{\Gamma}\boldsymbol{W}(t)+{\omega }_{0}^{2}\boldsymbol{V}(t)$, we obtain

Equation (44)

Rearranging gives

Equation (45)

as stated in the main text.

A5. Statistical physics prediction for ⟨sz ⟩ as a function of temperature

For a classical spin S of length S0 = n ℏ/2 in an external field B ext = B e z the thermal average ⟨sz ⟩ is determined by the Boltzmann distribution for the Hamiltonian H = −γ S B ext at inverse temperature β = 1/kB T,

Equation (46)

with ${Z}_{a}={\int }_{-{S}_{0}}^{+{S}_{0}}\!\mathrm{d}{S}_{z}\,{\mathrm{e}}^{a{S}_{z}}$ where a = βγB. This gives

Equation (47)

and hence

Equation (48)

Equation (49)

where ωL = |γB| and ${s}_{z}=\mathrm{sign}(\gamma )\,\frac{{S}_{z}}{{S}_{0}}$. In the magnetism literature, sometimes a reduced temperature experienced by a spin with n ≠ 1 is defined as Tred = T/n, i.e. the temperature is effectively reduced in comparison to the temperature experienced by a spin with n = 1.

A6. Steady state ⟨sz ⟩ plot for spin S0 = 5/2

As discussed in the main text, the impact of the quantum zero-point noise on the steady state ⟨sz ⟩ value at T = 0 K is very highly spin length dependent. For some materials a fundamental spin value of S0 = 1/2 will not be appropriate. For example iron (III) has five electrons in the outer d shell, and then from Hund's rules the spin is maximized to S = 5/2, and the orbital angular momentum is zero, L = 0. Therefore J = S, Landé g-factor equals 2, and the gyromagnetic ratio remains the electron gyromagnetic ratio. Figure 6(b) shows the steady state ⟨sz ⟩ plot as a function of temperature for spin S0 = 5/2, next to those for spin S0 = 1/2 (a) and S0 = 200/2 (c). The ⟨sz ⟩ value is below 1, at $\approx 0.8$, but the reduction is far less severe than for the spin-1/2.

Figure 6.

Figure 6. Same notation and Lorentzian parameters as in figure 5 but here also showing spin length S0 = 5/2 in panel (b). Panels (a) and (c) show ⟨sz ⟩ for S0 = 1/2 and S0 = 200/2, respectively, as in main text. Each curve is obtained by, for each temperature T, time-averaging over the late times of a single stochastic trajectory, from 0.75tmax to ${t}_{\mathrm{max}}=2\pi \times 8000{\omega }_{\mathrm{L}}^{-1}$.

Standard image High-resolution image

A7. Scales for classical and quantum thermostats

Here we establish the set of scales determining the dynamics described by equations (1) and (11) with either classical or quantum power spectra.

For a single spin, i.e. ignoring exchange terms etc, one may rescale the LLG equation (1) using M = γ S = |γ|S0 s with spin length | S | = S0. One obtains

Equation (50)

where the spin length S0 and ηG appear together, setting the first scale. Furthermore the effective field including classical stochastic noise with power spectrum (21), is given through (34) and (35) by

Equation (51)

Here we have introduced an S0 so that ηG appears together with it, and we find the second scale to be given by T/S0. The third scale is clearly set by the strength of the external field, B ext. We note that if one chooses the same S0 ηG value for different spin lengths, i.e. assumes ηG scales as 1/S0, as we have done in the main text, then only two scales are left, B ext and T/S0.

However, for the quantum Ohmic power spectrum the components of the stochastic noise can be written as

Equation (52)

Clearly, in the quantum case the temperature T now appears separately from spin length S0, thus introducing an additional scale in comparison to the classical case. Moreover the fact that the frequency integration for the stochastic field does not simplify as in (51) means that relaxation to the steady state at low temperatures (where the coth x cannot be approximated as 1/x) will be much more noisy than in the high temperature case. Thus in our simulations, this additional scale leads to larger uncertainties in the steady state results, as seen in figure 5(a).

Finally, we note that for the integration of the quantum Ohmic power spectrum in (52) we have introduced a frequency cut-off ωc by hand, which is necessary at low temperatures to avoid the integral diverging. At low temperatures, this cut-off will set an additional, somewhat artificial, scale of the problem. Importantly, such cut-off is not required for the Lorentzian coupling since the power spectrum (28) decays at high frequencies, even at low T.

Footnotes

  • An early version of this paper was posted under a different name, "Versatile three-dimensional quantum spin dynamics equation with guaranteed fluctuation-dissipation link", arxiv:2009.00600v1 (2020).

Please wait… references are loading.