This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

The semiclassical theory for spin dynamics in a disordered system

Tsung-Wei Chen [email protected] Department of Physics, National Sun Yat-sen University, Kaohsiung 80424, Taiwan    Hsiu-Chuan Hsu [email protected] Graduate Institute of Applied Physics, National Chengchi University, Taipei 11605, Taiwan Department of Computer Science, National Chengchi University, Taipei 11605, Taiwan
Abstract

We investigate the Drude model of spin dynamics in two-dimensional spin-orbit coupled systems. In the absence of an applied electric field, the spin aligns with the k-dependent effective magnetic field. The influence of disorder (the momentum relaxation time τ\tau) on the system is considered. In the presence of an electric field, the change in momentum causes a change in the effective magnetic field. The in-plane spin can precess around the successive change in the orientation of the effective magnetic field. We find that up to the linear order of the electric field, the spin orientation undergoes Larmor-like and non-Larmor like precession. Furthermore, we find that the non-Larmor motion over a very short time (tτt\ll\tau) exactly equals the result obtained from the Kubo formula. This means that the Kubo formula only captures the system’s response over a very short evolution. The spin-Hall conductivity for Larmor and non-Larmor precession in the Rashba system is analytically calculated. We find that the intrinsic spin–Hall current is not a universal constant and correctly drops to zero when the Rashba spin-orbit coupling drops to zero. We also calculate the time-averaged Larmor and non-Larmor spin-Hall conductivities (SHCs) for the k-cubic Rashba system and compare them to experimental values. The time-averaged Larmor SHC vanishes and the non-Larmor SHC is given by 2.1(q/8π)2.1(q/8\pi) which is very close to the experimental value 2.2(q/8π)2.2(q/8\pi) by Wunderlich et al. [Phys. Rev. Lett. 94, 047204 (2005)].

pacs:
71.70.Ej, 72.10.-d, 72.25.Dc, 73.43.Cd

I Introduction

When applying an external electric field in a spin-orbit coupled system, lateral spin motion is generated and spin accumulates at the edges of the sample, which is known as the spin-Hall effect Hirsch1999 . The spin-Hall effect has been extensively studied theoretically and through experiments over the past decades Sinova2015 . The mechanism of the intrinsic spin–Hall effect stems from the Berry curvature Mura2003 ; Chang2010 or the adiabatic evolution of spin around the effective magnetic field Sinova2004 . A challenging issue is how to define the spin current because each component of spin is not a conserved quantity. In Ref. Shi2006 , a generic definition of spin current (SZXN spin current) is proposed, and the spin current is composed of the conventional definition of spin current and the spin-torque dipole current. The conventional and spin-torque dipole currents were found to be related to the Berry phase in k-linear systems Shen2004 ; Chen2014 , while no such relationship exists for spin-orbit coupled systems with higher order momentum. Finally, the spin-Hall conductivity obtained from the SZXN spin current in response to the electric field is a universal constant in the k-linear system, regardless of the strength of the spin-orbit coupling, such as in pure Rashba  Rashba1984 ; Sinova2004 and Rashba–Dresselhaus systems  Dress1955 ; Chen2006 . This makes it difficult to compare theoretical results with experimental observations.

In semiclassical theory, the lateral motion of spin would be attributed to the force exerted on the spin. From this point of view, the intrinsic force due to the k-linear spin-orbit coupling was found to be related to the conventional definition of spin current Shen2005 . In the presence of disorder, the spin force in the Rashba–Dresselhaus system vanishes by accounting for the vertex correction Chen2009 . However, in the Drude model regime, an effective Lorentz force can be generated by the crystal field Eugene2007 , and it was shown to be related to the spin-Hall effect. Importantly, the spin-Hall conductivity generated by the effective Lorentz force is proportional to the charge conductance, and the resulting spin current fits experimental values observed for metals Eugene2007 , which sheds light on the evaluation of the observable spin motion in the Drude model regime. It also worths noting that the semiclassical analysis of spin dynamics in the intrinsic Luttinger system was investigated by using adiabatic approximation Jiang2005 .

In this study, we attempt to develop spin dynamics for two-dimensional spin-orbit coupled systems in the Drude model regime. The spin dynamics is treated as semiclassical motion governed by the Heisenberg equation of motion. The spin precession is ascribed to the rotation around the successive change in the effective magnetic field. To preserve the magnitude of the spin, the spin precession up to the first order of the electric field is comprised of Larmor and non-Larmor precessions. The former is a periodic motion and the latter is an exponential decay function. The spin-Hall conductivity (SHC) was also calculated by using the two components.

We find that in the regime t/τ1t/\tau\ll 1, the non-Larmor SHC is exactly equal to the result obtained from the Kubo formula. The formalism of the Kubo formula does not consider the Larmor precession. At finite τ\tau, we find that both the Larmor SHC and non-Larmor SHC vanish when the elapsed time is much larger than the relaxation time τ\tau. In the strong disorder limit (finite τ\tau and vanishing Rashba coupling), both the Larmor and non-Larmor SHCs vanish independently over a finite time. In the intrinsic limit (infinite τ\tau compared to the Larmor frequency), the Larmor SHC exactly cancels the non-Larmor SHC in the Kubo formula regime (t/τ1t/\tau\ll 1), which solves the problem of determining the universal constant of the SHC in previous studies. We also calculated the time-averaged spin-Hall conductivity for the k-cubic Rashba system Wund2005 . The result is close to the experimental value. We note that the time-averaged quantum Hall conductivity is also investigated in Liu2019 .

The remainder of this paper is organized as follows. In Sec. II, the influence of the momentum relaxation time on the spin-orbit coupled system is investigated. The spin motion up to the first order of the electric field is composed of Larmor and non-Larmor precessions. The boundary condition is also determined by using preservation of the spin magnitude. The analytical solutions for the Larmor and non-Larmor precessions are presented in Sec. III. The relationship between the non-Larmor precession and the Kubo formula is discussed. In Sec. IV, the spin-Hall conductivities for both Larmor and non-Larmor precessions are calculated. In the intrinsic Rashba system, the resulting spin-Hall conductivity drops to zero when the Rashba coupling drops to zero. We also calculate the Larmor and non-Larmor spin-Hall conductivities for the k-cubic Rashba system. The resulting spin-Hall conductivity is shown to be close to the experimental value. Finally, the conclusion is presented in Sec. V.

II 2D Hamiltonian and Drude Model

The two-dimensional spin-orbit coupled Hamiltonian under consideration is given by

H0=2k22m+2𝝈𝛀0,H_{0}=\frac{\hbar^{2}k^{2}}{2m}+\frac{\hbar}{2}\bm{\sigma}\cdot\bm{\Omega}_{0}, (1)

where 𝝈=(σx,σy,0)\bm{\sigma}=(\sigma_{x},\sigma_{y},0) with 𝝈\bm{\sigma} been Pauli matrices and 𝛀0=(2dx/,2dy/,0)\bm{\Omega}_{0}=(2d_{x}/\hbar,2d_{y}/\hbar,0) is the Larmor frequency. The terms dxd_{x} and dyd_{y} depend on the spin-orbit interaction and two-dimensional momentum 𝐤=(kx,ky)\mathbf{k}=(k_{x},k_{y}). There is no term with dzd_{z}, which implies that the spin is forced to lie along the plane. For example, in the Rashba–Dresselhaus system, we have dx=αky+βkxd_{x}=\alpha k_{y}+\beta k_{x} and dy=αkxβkyd_{y}=-\alpha k_{x}-\beta k_{y}. For the k-cubic Rashba-Dresselhaus system Bulaev2005 , the Hamiltonian is given by H0=ϵk+iα(σ+k3σk+3)β(σ+kk+k+σk+kk+)H_{0}=\epsilon_{k}+i\alpha(\sigma_{+}k_{-}^{3}-\sigma_{-}k_{+}^{3})-\beta(\sigma_{+}k_{-}k_{+}k_{-}+\sigma_{-}k_{+}k_{-}k_{+}). In this case, σ±=(σx±iσy)/2\sigma_{\pm}=(\sigma_{x}\pm i\sigma_{y})/2 and k±=kx±ikyk_{\pm}=k_{x}\pm ik_{y}. The corresponding dxd_{x} and dyd_{y} are defined as dx=[αsin(3ϕ)+βcos(ϕ)]k3d_{x}=[\alpha\sin(3\phi)+\beta\cos(\phi)]k^{3}, and dy=[αcos(3ϕ)+βsin(ϕ)]k3d_{y}=[-\alpha\cos(3\phi)+\beta\sin(\phi)]k^{3}, where ϕ=tan1(ky/kx)\phi=\tan^{-1}(k_{y}/k_{x}). The eigenenergies and eigenvectors are obtained by solving H0|n𝐤=En𝐤|n𝐤H_{0}|n\mathbf{k}\rangle=E_{n\mathbf{k}}|n\mathbf{k}\rangle, which is given by

Enk=ϵknd,E_{nk}=\epsilon_{k}-nd, (2)

where d=dx2+dy2d=\sqrt{d_{x}^{2}+d_{y}^{2}} and n=±n=\pm is the energy band index. In the two-dimensional spin-orbit coupled systems, it is easy to show that the spin z component in the unperturbed Hamiltonian Eq. (1) is zero, i.e., n𝐤|σz|n𝐤\langle{n\mathbf{k}}|\sigma_{z}|{n\mathbf{k}}\rangle=0.

We consider a simple model in which the momentum and the Fermi-Dirac distribution eventually approach a finite displacement through the collision between electrons and impurities. In this paper, an electric charge is denoted as q-q and q>0q>0. The equation of motion of momentum in the presence of a constant electric field is given by

d𝐤tdt=q𝐄+(𝐤𝐤t)τ,\hbar\frac{d\mathbf{k}_{t}}{dt}=-q\mathbf{E}+\frac{\hbar(\mathbf{k}-\mathbf{k}_{t})}{\tau}, (3)

where 𝐤\mathbf{k} is the electron’s momentum before the electric field is switched on, and τ\tau is the momentum relaxation time. Eq. (3) can be exactly solved, and the result is given by Jone1985

𝐤t=𝐤q𝐄t,\mathbf{k}_{t}=\mathbf{k}-\frac{q\mathbf{E}}{\hbar}t^{\ast}, (4)

where

t=τ(1et/τ).t^{\ast}=\tau\left(1-e^{-t/\tau}\right). (5)

Note that t(0)=0t^{\ast}(0)=0 and t()=τt^{\ast}(\infty)=\tau. When τ\tau is very large (the clean limit), we have et/τ=1t/τ+t2/2τ2+e^{-t/\tau}=1-t/\tau+t^{2}/2\tau^{2}+\cdots and t(t/τ1)tt^{\ast}(t/\tau\ll 1)\rightarrow t. This implies that 𝐤t=𝐤q𝐄t/\mathbf{k}_{t}=\mathbf{k}-q\mathbf{E}t/\hbar if tτt\ll\tau, which is the same result as when impurities are not present. We refer to the limit t/τ1t/\tau\ll 1 as the clean limit. In this limit t/τ1t/\tau\ll 1, equilibrium cannot be achieved.

As time tt is much larger than τ\tau, the momentum has a finite displacement of τq𝐄/-\tau q\mathbf{E}/\hbar. The Fermi-Dirac distribution fn𝐤f_{n\mathbf{k}} in 𝐤\mathbf{k}-space is also displaced by the electric field such that (tτt\gg\tau)

fn𝐤t=fn𝐤τq𝐄/=fn𝐤τq𝐄fn𝐤𝐤+.\begin{split}f_{n\mathbf{k}_{t}}&=f_{n\mathbf{k}-\tau q\mathbf{E}/\hbar}\\ &=f_{n\mathbf{k}}-\tau\frac{q\mathbf{E}}{\hbar}\cdot\frac{\partial f_{n\mathbf{k}}}{\partial\mathbf{k}}+\cdots.\end{split} (6)

The first order correction to the distribution function is proportional to the relaxation time which gives the Drude’s result of charge conductivity Jone1985 . Furthermore, we note that the quantum kinetic equations including spin-orbit coupling has been investigated in Ref. Bryksin2006 , in which the first order correction of distribution function is related to the spin density matrix. However, in this paper, the first order correction of distribution function is irrelevant in the linear response regime. This is because the spin precession around the effective magnetic field is also caused by the applied electric field and moreover the unperturbed spin z component is zero [see Eqs. (37) and (38)].

The Heisenberg equation of motion of spin 𝐬0n𝐤|eiH0t/𝝈eiH0t/|n𝐤\mathbf{s}^{0}\equiv\langle{n\mathbf{k}}|e^{iH_{0}t/\hbar}\bm{\sigma}e^{-iH_{0}t/\hbar}|{n\mathbf{k}}\rangle is given by

t𝐬0=𝛀0×𝐬0,\frac{\partial}{\partial t}\mathbf{s}^{0}=\bm{\Omega}_{0}\times\mathbf{s}^{0}, (7)

where 𝛀0\bm{\Omega}_{0} depends on momentum 𝐤\mathbf{k}. In the presence of an applied electric field, the change in momentum would lead to the change in the effective magnetic field and the spin orientation varies due to the precession. Semiclassically, we assume that the effect of disorder and the applied electric field are included in the effective magnetic field and the spin dynamics is still governed by the Heisenberg equation of motion. Thus, the resulting Hamiltonian can be written as

H(t)=2𝐤t22m+2𝝈𝛀(t),H(t)=\frac{\hbar^{2}\mathbf{k}_{t}^{2}}{2m}+\frac{\hbar}{2}\bm{\sigma}\cdot\bm{\Omega}(t), (8)

where 𝐤t\mathbf{k}_{t} is given by Eq. (4), 𝛀(t)=2𝐝(t)/\bm{\Omega}(t)=2\mathbf{d}(t)/\hbar and 𝐝(t)=(dx(𝐤t),dy(𝐤t),0)\mathbf{d}(t)=(d_{x}(\mathbf{k}_{t}),d_{y}(\mathbf{k}_{t}),0). We require that the Hamiltonian Eq. (8) satisfies the Schrodinger equation

it|ψ(t)=H(t)|ψ(t).i\hbar\frac{\partial}{\partial t}|\psi(t)\rangle=H(t)|\psi(t)\rangle. (9)

The time-dependent spin 𝐬(t)\mathbf{s}(t) is defined as 𝐬(t)ψ(t)|𝝈|ψ(t)\mathbf{s}(t)\equiv\langle\psi(t)|\bm{\sigma}|\psi(t)\rangle which is the averaged spin in the presence of disorder and an electric field. Apply the first derivative of time to 𝐬(t)\mathbf{s}(t), we have

t𝐬(t)=tψ(t)|𝝈|ψ(t)+ψ(t)|𝝈|tψ(t)=ψ(t)|1i[𝝈,H(t)]|ψ(t),\begin{split}\frac{\partial}{\partial t}\mathbf{s}(t)&=\langle\frac{\partial}{\partial t}\psi(t)|\bm{\sigma}|\psi(t)\rangle+\langle\psi(t)|\bm{\sigma}|\frac{\partial}{\partial t}\psi(t)\rangle\\ &=\langle\psi(t)|\frac{1}{i\hbar}[\bm{\sigma},H(t)]|\psi(t)\rangle,\end{split} (10)

where Eq. (8) was used. By using the commutator for Pauli matrices [σi,σj]=2iϵijkσk[\sigma_{i},\sigma_{j}]=2i\epsilon_{ijk}\sigma_{k}, we obtain the Heisenberg equation of motion for the averaged spin,

t𝐬(t)=𝛀(t)×𝐬(t),\frac{\partial}{\partial t}\mathbf{s}(t)=\bm{\Omega}(t)\times\mathbf{s}(t), (11)

where [𝝈,𝐤t2]=0[\bm{\sigma},\mathbf{k}_{t}^{2}]=0 was used. We emphasize that the equation of motion for momentum is derived by using semiclassical Drude model (Eq.(4)), not the expectation value of 𝐤\mathbf{k} with respect to the state |ψ(t)|\psi(t)\rangle, i.e., 𝐤tψ(t)|𝐤|ψ(t)\mathbf{k}_{t}\neq\langle\psi(t)|\mathbf{k}|\psi(t)\rangle.

We now return to the momentum dependence of the effective magnetic field. The influence of the disorder in the presence of an electric field will lead to a change in the effective magnetic field, that is, di(t)=di(𝐤t)d_{i}(t)=d_{i}(\mathbf{k}_{t}). Therefore, up to the linear order of the electric field,

di(𝐤t)=di(kaqEat)=diqEatdika+o(Ea2).\begin{split}d_{i}(\mathbf{k}_{t})&=d_{i}(k_{a}-\frac{qE_{a}}{\hbar}t^{\ast})\\ &=d_{i}-\frac{qE_{a}}{\hbar}t^{\ast}\frac{\partial d_{i}}{\partial k_{a}}+o(E_{a}^{2}).\end{split} (12)

Similar to the Fermi-Dirac distribution in the presence of disorder, when tt\rightarrow\infty, the 𝐝\mathbf{d} vector changes its direction from did_{i} to another fixed direction di(τeEa/)(di/ka)d_{i}-(\tau eE_{a}/\hbar)(\partial d_{i}/\partial k_{a}). We define 𝛀(t)=2𝐝(t)/=𝛀0+𝛀(t)\bm{\Omega}(t)=2\mathbf{d}(t)/\hbar=\bm{\Omega}_{0}+\bm{\Omega}^{\prime}(t), where 𝛀0=(2dx/,2dy/,0)\bm{\Omega}_{0}=(2d_{x}/\hbar,2d_{y}/\hbar,0) and

𝛀(t)=qEat𝛀0ka,\bm{\Omega}^{\prime}(t)=-\frac{qE_{a}}{\hbar}t^{\ast}\frac{\partial\bm{\Omega}_{0}}{\partial k_{a}}, (13)

which also lies on the 2D plane. The change in the effective magnetic field will cause the spin to precess around the new direction of the effective magnetic field. The spin’s z component will be non-zero at this time. The trajectory of 𝐝(t)\mathbf{d}(t) in dd-space is a straight line up to the first order of the electric field. The resulting angular velocity of 𝐝\mathbf{d} measured from the origin must vary with time. Therefore, we cannot use a rotating frame with fixed spin because the resulting Hamiltonian does not commute at different times. Instead, we use the perturbative method defined in Eq. (11). The unitarity is broken and the magnitude of the spin is preserved only up to the first order of an electric field. Namely, we have

𝐬(t)=𝐬0(t)+𝐬(t)+o(Ea2),\mathbf{s}(t)=\mathbf{s}^{0}(t)+\mathbf{s}^{\prime}(t)+o(E_{a}^{2}), (14)

where 𝐬0\mathbf{s}^{0} and 𝐬\mathbf{s}^{\prime} correspond to the solutions without an electric field and with an electric field, respectively. Inserting this result into Eq. (11),

t𝐬=𝛀0×𝐬+𝛀×𝐬0,\frac{\partial}{\partial t}\mathbf{s}^{\prime}=\bm{\Omega}_{0}\times\mathbf{s}^{\prime}+\bm{\Omega}^{\prime}\times\mathbf{s}^{0}, (15)

where the following unperturbed Eq. (7) was used. The solution of Eq. (7) (using |𝐬0|=1|\mathbf{s}^{0}|=1) is given by

sx0=cosΘcosθ0sinΘsinθ0sin(Ω0t),sy0=sinΘcosθ0+cosΘsinθ0sin(Ω0t),sz0=sinθ0cos(Ω0t),\begin{split}&s^{0}_{x}=\cos\Theta\cos\theta_{0}-\sin\Theta\sin\theta_{0}\sin(\Omega_{0}t),\\ &s^{0}_{y}=\sin\Theta\cos\theta_{0}+\cos\Theta\sin\theta_{0}\sin(\Omega_{0}t),\\ &s^{0}_{z}=-\sin\theta_{0}\cos(\Omega_{0}t),\end{split} (16)

where the angle θ0\theta_{0} is the angle between 𝛀0\bm{\Omega}_{0} and 𝐬0\mathbf{s}^{0}, and Θ\Theta is defined as tanΘ=dy/dx\tan\Theta=d_{y}/d_{x}. Eq. (7) implies that 𝐬0\mathbf{s}^{0} behaves as the spin in the intrinsic system H0H_{0}, where the Larmor frequency is 𝛀0\bm{\Omega}_{0}. The spin should be aligned with 𝛀0\bm{\Omega}_{0} without applying an electric field. In this sense, we can assume that θ0=0\theta_{0}=0, and we have

sx0=cosΘ=dxd,sy0=sinΘ=dyd,sz0=0.\begin{split}&s^{0}_{x}=\cos\Theta=\frac{d_{x}}{d},\\ &s^{0}_{y}=\sin\Theta=\frac{d_{y}}{d},\\ &s^{0}_{z}=0.\end{split} (17)

To obtain results for different energy bands, we can simply use the replacement 𝐬0(n)𝐬0\mathbf{s}_{0}\rightarrow(-n)\mathbf{s}_{0} in Eq. (17), where n=±n=\pm is the band index. Substituting Eq. (17) with the appropriate band index into Eq. (15), we have (λaqEa\lambda_{a}\equiv qE_{a})

tsz=λatnΩ0Γaz+syΩ0xsxΩ0y,\frac{\partial}{\partial t}s_{z}^{\prime}=-\frac{\lambda_{a}t^{\ast}}{\hbar}n\Omega_{0}\Gamma^{z}_{a}+s_{y}^{\prime}\Omega_{0x}-s_{x}^{\prime}\Omega_{0y}, (18)

for the spin-z component, and

tsx=Ω0ysz,tsy=Ω0xsz,\begin{split}\frac{\partial}{\partial t}s_{x}^{\prime}&=\Omega_{0y}s_{z}^{\prime},\\ \frac{\partial}{\partial t}s_{y}^{\prime}&=-\Omega_{0x}s_{z}^{\prime},\end{split} (19)

for the spin x and y components, where Γaz\Gamma^{z}_{a} is defined as

Γaz=1d2(dxdykadydxka).\Gamma_{a}^{z}=\frac{1}{d^{2}}\left(d_{x}\frac{\partial d_{y}}{\partial k_{a}}-d_{y}\frac{\partial d_{x}}{\partial k_{a}}\right). (20)

We observe that the term 𝛀(t)×𝐬0\bm{\Omega}^{\prime}(t)\times\mathbf{s}^{0} in Eq. (15) behaves as a torque such that 𝐬\mathbf{s}^{\prime} not only has a Larmor rotation (periodic rotation) but also a non-Larmor rotation (non-periodic motion). Therefore, similar to the derivations in a previous study, the spin will be composed of non-Larmor (N) and Larmor (L) precession terms.

𝐬=𝚺L+𝚺N\mathbf{s}^{\prime}=\bm{\Sigma}^{L}+\bm{\Sigma}^{N} (21)

The term 𝚺L\bm{\Sigma}^{L} is called the Larmor component and is governed by

t𝚺L=𝛀0×𝚺L.\frac{\partial}{\partial t}\bm{\Sigma}^{L}=\bm{\Omega}_{0}\times\bm{\Sigma}^{L}. (22)

Eq. (22) is similar to Eq. (7); however, 𝚺N\bm{\Sigma}^{N} contains the response of the electric field, and we expect that the direction of 𝚺L\bm{\Sigma}^{L} is not parallel to the static Larmor frequency 𝛀0\bm{\Omega}_{0}. On the other hand, the term 𝚺N\bm{\Sigma}^{N} is called the non-Larmor component, which satisfies

t𝚺N=𝛀0×𝚺N+𝛀×𝐬0.\frac{\partial}{\partial t}\bm{\Sigma}^{N}=\bm{\Omega}_{0}\times\bm{\Sigma}^{N}+\bm{\Omega}^{\prime}\times\mathbf{s}^{0}. (23)

It must be emphasized that Eqs. (22) and (23) hold only if the magnitude of spin 𝐬(t)\mathbf{s}(t) is valid up to the second order of electric field. The magnitude of spin is given by |𝐬|2=|𝐬0+𝐬+o(Ea2)|2=|𝐬0|2+2𝐬0𝐬+o(Ea2)|\mathbf{s}|^{2}=|\mathbf{s}^{0}+\mathbf{s}^{\prime}+o(E_{a}^{2})|^{2}=|\mathbf{s}^{0}|^{2}+2\mathbf{s}^{0}\cdot\mathbf{s}^{\prime}+o(E_{a}^{2}). To preserve the magnitude of spin up to the second order of EaE_{a}, we require that

𝐬0𝐬=0\mathbf{s}^{0}\cdot\mathbf{s}^{\prime}=0 (24)

at any time tt. Furthermore because the time dependence of the Larmor component is different from that of non-Larmor components, Eq. (24) implies that

𝐬0𝚺L=0,𝐬0𝚺N=0.\begin{split}&\mathbf{s}^{0}\cdot\bm{\Sigma}^{L}=0,\\ &\mathbf{s}^{0}\cdot\bm{\Sigma}^{N}=0.\end{split} (25)

On the other hand, at t=0t=0, the precession of spin begins from the in-plane and ends as an out-of-plane orientation. In this sense, the z component sz(0)s_{z}^{\prime}(0) must be zero, although its rate of change is non-zero. This can be described by

sz(0)=0,(szt)t=0=Ω0(cosΘsy(0)sinΘsx(0))0,\begin{split}&s^{\prime}_{z}(0)=0,\\ &\left(\frac{\partial s_{z}^{\prime}}{\partial t}\right)_{t=0}=\Omega_{0}\left(\cos\Theta s_{y}^{\prime}(0)-\sin\Theta s_{x}^{\prime}(0)\right)\neq 0,\end{split} (26)

where Eq. (15) at t=0t=0 was used. We note that a similar derivation was proposed in Ref. Paul2018 , in which the effective magnetic field is replaced by a pseudo-magnetic field derived from the spin force. In the following section, we solve 𝐬\mathbf{s}^{\prime} by acquiring the boundary conditions for Eqs. (25) and (26).

III Larmor and Non-Larmor precessions

The solutions of Eq. (22) would be the same as Eq. (16). That is, we have

𝚺xL=ΣL[cosΘcosθsinΘsinθsin(Ω0t)],𝚺yL=ΣL[sinΘcosθ+cosΘsinθsin(Ω0t)],𝚺zL=ΣL[sinθcos(Ω0t)].\begin{split}&\bm{\Sigma}^{L}_{x}=\Sigma^{L}\left[\cos\Theta\cos\theta-\sin\Theta\sin\theta\sin(\Omega_{0}t)\right],\\ &\bm{\Sigma}^{L}_{y}=\Sigma^{L}\left[\sin\Theta\cos\theta+\cos\Theta\sin\theta\sin(\Omega_{0}t)\right],\\ &\bm{\Sigma}^{L}_{z}=\Sigma^{L}\left[-\sin\theta\cos(\Omega_{0}t)\right].\end{split} (27)

However, the difference is that 𝚺L\bm{\Sigma}^{L} is in linear order with the applied electric field in the disordered system. The angle θ\theta between 𝚺L\bm{\Sigma}^{L} and 𝛀0\bm{\Omega}_{0} depends on the electric field and relaxation time τ\tau. First, by using the requirement in Eq. (25), we have ΣxLcosΘ+ΣyLsinΘ=0\Sigma^{L}_{x}\cos\Theta+\Sigma^{L}_{y}\sin\Theta=0, which implies that ΣLcosθ=0\Sigma^{L}\cos\theta=0. Because ΣL0\Sigma^{L}\neq 0, we obtain θ=π/2\theta=\pi/2, which means that 𝚺L\bm{\Sigma}^{L} is always perpendicular to 𝛀0\bm{\Omega}_{0}. Eq. (27) then becomes

𝚺xL=ΣLsinΘsin(Ω0t),𝚺yL=ΣLcosΘsin(Ω0t),𝚺zL=ΣLcos(Ω0t).\begin{split}&\bm{\Sigma}^{L}_{x}=-\Sigma^{L}\sin\Theta\sin(\Omega_{0}t),\\ &\bm{\Sigma}^{L}_{y}=\Sigma^{L}\cos\Theta\sin(\Omega_{0}t),\\ &\bm{\Sigma}^{L}_{z}=-\Sigma^{L}\cos(\Omega_{0}t).\end{split} (28)

Applying the time derivative to the z component of Eq. (23), and noting that dt/dt=et/τdt^{\ast}/dt=e^{-t/\tau}, we obtain

2t2ΣzN+Ω02ΣzN+nλaΩ0Γazet/τ=0,\frac{\partial^{2}}{\partial t^{2}}\Sigma^{N}_{z}+\Omega_{0}^{2}\Sigma^{N}_{z}+\frac{n\lambda_{a}}{\hbar}\Omega_{0}\Gamma^{z}_{a}e^{-t/\tau}=0, (29)

where Ω0=2d/\Omega_{0}=2d/\hbar was used. The solution of Eq. (29) is given by

ΣzN=Anet/τ,\Sigma^{N}_{z}=A_{n}e^{-t/\tau}, (30)

where AnA_{n} is the dimensionless quantity

An=nλaΩ0Γaz11τ2+Ω02.A_{n}=-\frac{n\lambda_{a}}{\hbar}\Omega_{0}\Gamma^{z}_{a}\frac{1}{\frac{1}{\tau^{2}}+\Omega_{0}^{2}}. (31)

Eq. (30) plays an important role in the spin-Hall effect. We will return to this point in the next section. We now use Eq. (26), which is ΣzN(0)+ΣzL(0)=0\Sigma^{N}_{z}(0)+\Sigma^{L}_{z}(0)=0. Thus, AnΣL=0A_{n}-\Sigma^{L}=0, and the resulting z component of 𝚺L\bm{\Sigma}^{L} is given by

ΣzL=Ancos(Ω0t)\Sigma^{L}_{z}=-A_{n}\cos(\Omega_{0}t) (32)

and the x and y components of 𝚺L\bm{\Sigma}^{L} are given by

ΣxL=AnsinΘsin(Ω0t),ΣyL=AncosΘsin(Ω0t).\begin{split}\Sigma^{L}_{x}=&-A_{n}\sin\Theta\sin(\Omega_{0}t),\\ \Sigma^{L}_{y}=&A_{n}\cos\Theta\sin(\Omega_{0}t).\end{split} (33)

Substituting Eqs. (30) and (31) into Eq. (23) for the x and y components, we obtain

ΣxN=AnΩ0τ(et/τ1τ2+Ω02Ω02)sinΘ,ΣyN=AnΩ0τ(et/τ1τ2+Ω02Ω02)cosΘ,\begin{split}\Sigma^{N}_{x}&=-A_{n}\Omega_{0}\tau\left(e^{-t/\tau}-\frac{\frac{1}{\tau^{2}}+\Omega_{0}^{2}}{\Omega_{0}^{2}}\right)\sin\Theta,\\ \Sigma^{N}_{y}&=A_{n}\Omega_{0}\tau\left(e^{-t/\tau}-\frac{\frac{1}{\tau^{2}}+\Omega_{0}^{2}}{\Omega_{0}^{2}}\right)\cos\Theta,\end{split} (34)

Arbitrary time-independent constants CxC_{x} and CyC_{y} can be added to ΣxN\Sigma^{N}_{x} and ΣyN\Sigma^{N}_{y}, respectively. The two constants must obey the condition dxCydyCx=0d_{x}C_{y}-d_{y}C_{x}=0. Furthermore, the preserved magnitude of spin |𝐬(t)|=1+o(Ea2)|\mathbf{s}(t)|=1+o(E_{a}^{2}) (see Eq. (25)) implies that dxCx+dyCy=0d_{x}C_{x}+d_{y}C_{y}=0. The only solution is Cx=Cy=0C_{x}=C_{y}=0 if d0d\neq 0. On the other hand, it can be shown that Eqs. (30), (32), (33), and (34) satisfy the second boundary condition of Eq. (26).

When tt\rightarrow\infty, the momentum has a constant shift and the effective magnetic field does not change its direction over time. The system achieves an equilibrium state. We find that the in-plane components ΣxN\Sigma^{N}_{x} and ΣyN\Sigma^{N}_{y} do not change with time, and the non-Larmor spin z component ΣzN\Sigma^{N}_{z} vanishes. Only the 𝚺L\bm{\Sigma}^{L} Larmor components survive in the system. In the next section, we calculate the Larmor SHC and non-Larmor SHC in the Rashba system. We close this section by discussing the relationship between our results and the Kubo formula in the intrinsic system.

If the energy gap is much larger than the broadening of the energy band due to the disorder, the system is said to be within the intrinsic limit, that is, 2d/τ2d\gg\hbar/\tau. This implies that the intrinsic limit is given by Ω0τ1\Omega_{0}\tau\gg 1 when 𝐤0\mathbf{k}\neq 0. When τ\tau is finite, the intrinsic limit implies that the system is in the strong spin-orbit coupling regime and cannot be zero unless τ\tau\rightarrow\infty. Therefore, the intrinsic limit leads to the result that the system displays a physical evolution only for a very short time compared to a finite τ\tau. In the following, we will demonstrate that at the intrinsic limit (and clean limit), the non-Larmor spin z component ΣzN\Sigma^{N}_{z} is exactly equal to the result obtained from the Kubo formula.

If we take the intrinsic limit Ω0τ1\Omega_{0}\tau\gg 1, then ΣiN\Sigma^{N}_{i} is proportional to (et/τ1)Ω0τ(e^{-t/\tau}-1)\Omega_{0}\tau. Then, taking the clean limit t/τ1t/\tau\ll 1, we have (et/τ1)Ω0τΩ0t(e^{-t/\tau}-1)\Omega_{0}\tau\approx-\Omega_{0}t. Therefore, considering the limit Ω0τ1\Omega_{0}\tau\gg 1 and then t/τ1t/\tau\ll 1, we have

ΣxN+AnΩ0tsinΘ,ΣyNAnΩ0tcosΘ.AnnλaΩ0Γaz.\begin{split}\Sigma^{N}_{x}&\approx+A_{n}\Omega_{0}t\sin\Theta,\\ \Sigma^{N}_{y}&\approx-A_{n}\Omega_{0}t\cos\Theta.\\ A_{n}&\approx-\frac{n\lambda_{a}}{\hbar\Omega_{0}}\Gamma_{a}^{z}.\end{split} (35)

Note that in the limit Ω0τ1\Omega_{0}\tau\gg 1 with t/τ1t/\tau\ll 1, the non-Larmor spin x and y components exhibit no physical growth in time. Furthermore, Eq. (35) can be exactly derived from the Heisenberg equation of motion Chen2019 . If we substitute Eq. (35) back into Eq. (23), we obtain

ΣzNt0,ΣzN=An=nλaΩ0Γaz.\begin{split}&\frac{\partial\Sigma^{N}_{z}}{\partial t}\approx 0,\\ &\Sigma^{N}_{z}=A_{n}=-\frac{n\lambda_{a}}{\hbar\Omega_{0}}\Gamma^{z}_{a}.\end{split} (36)

Interestingly, because of the linear time dependence of the spin x and y components, the response of the non-Larmor spin z component is a constant only for a very short time. It has been shown that Eq. (36) is exactly the same as the result obtained from the Kubo formula in the intrinsic case Chen2014 ; Chen2019 . The spin-Hall current from the Kubo formula using the result in Eq. (36) is a universal constant in the Rashba system. The universal constant also leads to the problem that the spin x and y components have no physical growth with time, as has been demonstrated in Refs  Chen2006 ; Chen2019 ; Dim2005 ; Chalaev2005 . Nevertheless, as discussed in the above results, the regime of the Kubo formula in the intrinsic spin-orbit coupled system is valid only for a very short time limit. In particular, the short time limit corresponds to the time when the spin adiabatically aligns its orientation with the effective magnetic field. Consequently, the change in the magnitude of the effective magnetic field (and thus the spin in that direction) yields a non-zero spin z component in order to preserve the spin magnitude Sinova2004 .

IV time-dependent spin-Hall effect

The spin-Hall current deduced from the spin-dynamics would be semiclassically obtained by using the conventional definition of the spin current, which is the simple multiplication of spin (J/2)sz(J\hbar/2)s_{z} and the lateral velocity kx/m\hbar k_{x}/m in which the electric field is applied in y-direction, where J=1J=1 for the Rashba–Dresselhaus system and J=3J=3 for the k-cubic Rashba system. The time-dependent spin-Hall current is given by

𝒥xz=n𝐤fn𝐤tJ2szkxm=n𝐤(fn𝐤tqEyfn𝐤ky)J2(sz0+sz)kxm.\begin{split}\mathcal{J}^{z}_{x}&=\sum_{n\mathbf{k}}f_{n\mathbf{k}_{t}}\frac{J\hbar}{2}s_{z}\frac{\hbar k_{x}}{m}\\ &=\sum_{n\mathbf{k}}(f_{n\mathbf{k}}-t^{\ast}\frac{qE_{y}}{\hbar}\cdot\frac{\partial f_{n\mathbf{k}}}{\partial k_{y}})\frac{J\hbar}{2}(s^{0}_{z}+s_{z}^{\prime})\frac{\hbar k_{x}}{m}.\\ \end{split} (37)

We note that sz0=0s^{0}_{z}=0 (see Eq. (17)) in the present systems under consideration. Up to the first order of the applied electric field, the time-dependent spin-Hall current Eq. (37) becomes

𝒥xz=n𝐤fn𝐤J2szkxm=σxyzEy,\mathcal{J}^{z}_{x}=\sum_{n\mathbf{k}}f_{n\mathbf{k}}\frac{J\hbar}{2}s^{\prime}_{z}\frac{\hbar k_{x}}{m}=\sigma^{z}_{xy}E_{y}, (38)

The linear response of the spin z component szs^{\prime}_{z} is given by sz=ΣzL(t)+ΣzN(t)s_{z}^{\prime}=\Sigma^{L}_{z}(t)+\Sigma^{N}_{z}(t) [see Eqs. (30) and (32)]. The spin-Hall conductivity σxyz\sigma^{z}_{xy} can now be written as the Larmor component σxyL\sigma^{L}_{xy} and non-Larmor component σxyN\sigma^{N}_{xy}, that is,

σxyz=σxyL+σxyN.\sigma^{z}_{xy}=\sigma^{L}_{xy}+\sigma^{N}_{xy}. (39)
Refer to caption
Figure 1: (color online) Numerical results of Larmor (σxyL\sigma^{L}_{xy}) and non-Larmor (σxyN\sigma^{N}_{xy}) spin-Hall conductivities in the presence of disorder with finite τ\tau. When the Rashba coupling α\alpha approaches zero (corresponding to a strong disorder), both SHCs drop to zero independently.
Refer to caption
Figure 2: (color online) Numerical values of σxyL\sigma^{L}_{xy} and σxyN\sigma^{N}_{xy} for the Rashba system with finite τ\tau. As time increases beyond τ\tau, σxyN\sigma^{N}_{xy} vanishes, while σxyL\sigma^{L}_{xy} survives but decays over time.
Refer to caption
Figure 3: (color online)Numerical values of the Larmor spin-Hall conductivity (σxyL\sigma^{L}_{xy}) for the Rashba system in the intrinsic limit. At finite time tt, when the Rashba coupling drops to zero, σxyL\sigma^{L}_{xy} approaches the universal constant but is opposite to σxyN\sigma^{N}_{xy} in sign.

IV.1 k-linear Rashba system

Considering the k-linear Rashba system, it can be shown that for the Larmor SHC,

σxyL=αq4πmkFkF+𝑑kk2cos(2αkt/)1τ2+(2αk)2.\sigma^{L}_{xy}=\frac{\alpha q}{4\pi m}\int_{k^{-}_{F}}^{k^{+}_{F}}dk\frac{k^{2}\cos(2\alpha kt/\hbar)}{\frac{1}{\tau^{2}}+(\frac{2\alpha k}{\hbar})^{2}}. (40)

For the non-Larmor component,

σxyN=αq4πmet/τkFkF+𝑑kk21τ2+(2αk)2,\sigma^{N}_{xy}=-\frac{\alpha q}{4\pi m}e^{-t/\tau}\int_{k^{-}_{F}}^{k^{+}_{F}}dk\frac{k^{2}}{\frac{1}{\tau^{2}}+(\frac{2\alpha k}{\hbar})^{2}}, (41)

where the Fermi momenta at the two bands kF±k_{F}^{\pm} are given by

kF±k0=±1+1+μR,k0mα2,\frac{k_{F}^{\pm}}{k_{0}}=\pm 1+\sqrt{1+\frac{\mu}{\mathcal{E}_{R}}},~{}k_{0}\equiv\frac{m\alpha}{\hbar^{2}}, (42)

where μ\mu is the Fermi energy. We define the Rashba energy R\mathcal{E}_{R} as R=2k02/2m=mα2/22\mathcal{E}_{R}=\hbar^{2}k_{0}^{2}/2m=m\alpha^{2}/2\hbar^{2}. Equation (41) is then exactly the same as the spin-Hall conductivity obtained from the Kubo formula in the presence of disorder Loss2004 , except for the time dependence et/τe^{-t/\tau}. We find that in the presence of a strong disorder (finite τ\tau and vanishing Rashba coupling), the Larmor and non-Larmor SHCs vanish independently, as shown in Fig. 1. Furthermore, when τ\tau is finite, the time evolution of the Larmor and non-Larmor SHCs is shown in Fig. 2. We find that the non-Larmor SHC approaches zero faster than the Larmor SHC. The Larmor SHC survives but decays over time.

Consider the intrinsic limit τ\tau\rightarrow\infty in the k-linear Rashba system, where the time tt can be considered finite in this limit. It can be shown that

σxyL=q8π4αk0t[sin(2αkF+t)sin(2αkFt)],σxyN=q8π,\begin{split}&\sigma^{L}_{xy}=\frac{q}{8\pi}\frac{\hbar}{4\alpha k_{0}t}\left[\sin\left(\frac{2\alpha k_{F}^{+}}{\hbar}t\right)-\sin\left(\frac{2\alpha k_{F}^{-}}{\hbar}t\right)\right],\\ &\sigma^{N}_{xy}=-\frac{q}{8\pi},\end{split} (43)

where et/τ1e^{-t/\tau}\approx 1 for σxyN\sigma^{N}_{xy}. The universal constant in σxyN\sigma^{N}_{xy} is exactly the same as the result obtained from the Kubo formula. When the two bands are occupied, kF+k_{F}^{+} is not equal to kFk_{F}^{-}, and the oscillating term in σyxL\sigma^{L}_{yx} is always finite. As time progresses, σxyL0\sigma^{L}_{xy}\rightarrow 0, which is the same as the result with finite τ\tau. When the Rashba coupling is very small (for finite tt),

sin(2αkF+t)sin(2αkFt)=2αt(kF+kF)+o(α6t3)=4αk0t+o(α6t3),\begin{split}&\sin\left(\frac{2\alpha k_{F}^{+}t}{\hbar}\right)-\sin\left(\frac{2\alpha k_{F}^{-}t}{\hbar}\right)\\ &=\frac{2\alpha t}{\hbar}\left(k_{F}^{+}-k_{F}^{-}\right)+o(\alpha^{6}t^{3})\\ &=\frac{4\alpha k_{0}t}{\hbar}+o(\alpha^{6}t^{3}),\end{split} (44)

where Eq. (42) was used. We also note that kF±k_{F}^{\pm} is the order of α\alpha. Substituting Eq. (44) into Eq. (43),

σxyz=σxyL+σxyN=q8πq8π=0.\sigma^{z}_{xy}=\sigma^{L}_{xy}+\sigma^{N}_{xy}=\frac{q}{8\pi}-\frac{q}{8\pi}=0. (45)

Therefore, we find that when the Larmor motion is considered, we can solve the problem caused by the Kubo formula. The numerical values for the Larmor SHC in Eq. (43) are shown in Fig. 3. We find that at finite tt, vanishingly small Rashba coupling indeed exhibits a region where σxyLq/8π\sigma^{L}_{xy}\rightarrow q/8\pi.

In short, the spin dynamics preserve the spin magnitude up to the first order of the electric field. In this perturbation method, spin motion is composed of Larmor and non-Larmor precessions. The Larmor SHC does not have the same time-dependent function as the non-Larmor SHC, and thus, in general, they cannot cancel each other. Furthermore, in the intrinsic system, the spin-Hall conductivity from the Kubo formula is equal to the non-Larmor SHC over a short time as compared to τ\tau. In the Rashba system, the spin-Hall conductivity is generally not a universal constant, and we have demonstrated that when Rashba coupling vanishes, the spin-Hall conductivity also vanishes.

Refer to caption
Figure 4: (color online) Numerical values of the time-averaged spin-Hall conductivities vs. the parameter \ell for the k-cubic Rashba system. The experimental value 2.2(e/8π)2.2(e/8\pi) (green dotted line) corresponds to the value 6\ell\approx 6 (the vertical dotted line). The Larmor/non-Larmor SHC is shown by blue dashed / dash-dotted line. The total SHC is shown by red solid line.

IV.2 k-cubic Rashba system

We close this section by calculating the spin-Hall conductivity in the k-cubic Rashba system, in which the experimental value is 2.2(q/8π)2.2(q/8\pi) in the clean limit Wund2005 . In the intrinsic two-dimensional k-cubic Rashba system, the SHC is shown to be 9(q/8π)9(q/8\pi) when the length scale 2/2mα\hbar^{2}/2m\alpha is much larger than 4πnh\sqrt{4\pi n_{h}}  Loss2005 , where nhn_{h} is the 2D hole density. The reduction of the SHC in the intrinsic k-cubic Rashba system has been investigated by many authors. In Ref. Bern2005 , the SHC is ascribed to the finite thickness of the quantum well, and the conventional definition of the spin current yields a value of 1.9(q/8π)1.9(q/8\pi). By further taking into account the spin-torque dipole current in the quantum well with a finite thickness, the authors in Ref. Zhang2008 demonstrated that the SHC resulting from the SZXN spin current yields a value of 1.2(q/8π)1.2(q/8\pi), which is also very close to the experimental value.

However, for the above two cases, the structure inversion asymmetric (SIA) Rashba term is neglected, although experiments were performed under strong SIA Rashba coupling. As indicated in Ref. Zhang2008 , the inclusion of the SIA Rashba term would lead to a value of 10(q/8π)-10(q/8\pi), which is higher than the experimental value, and the sign of SHC changes. We also note that the spin-Hall effect would be due to the edge spin accumulation Nomura2005 and the magnetization at the edge of the sample would be in non-equilibrium Bleibaum2006 . To compare the experimental values, we used the time-averaged formula for the time-dependent parts tauavg ,

f(t)¯=Ω02π02π/Ω0f(t)𝑑t.\overline{f(t)}=\frac{\Omega_{0}}{2\pi\ell}\int_{0}^{2\pi\ell/\Omega_{0}}f(t)dt. (46)

That is, the system’s response should have vanishingly small Larmor SHC in average. Equation (46) implies that the time-averaged quantity is a function of kk. This is because each kk point has its own Larmor frequency. Nonetheless, since the spins do not interact with each other, there is no physical constraint demanding that the parameter should depend on k. Each k point should be averaged by the same period. The quantity \ell is the parameter that can be tuned to fit the experimental value, which is the effective number of rotations from the initial spin state back to the initial spin state. ecause the wave function of spin should rotate 4π4\pi in order to return to its original wave function Sak2017 , and we expect =2\ell=2 for k-linear spin-orbit coupled system. For k-cubic Rashba system, the rotational symmetry in momentum is three-fold in the wave function, and we expect =2×3=6\ell=2\times 3=6 (see Appendix A).

By using Eq. (46), the Larmor and non-Larmor SHCs can be written as (the detail is shown in Appendix B)

σxyL¯=9q8π2msin(2π)2πkFkF+𝑑k2αk4(τ)2+(2αk3)2,σxyN¯=9q8π2m12πkFkF+𝑑k2αk4(τ)2+(2αk3)22αk3(/τ)(1e2π(/τ)/2αk3),\begin{split}&\overline{\sigma^{L}_{xy}}=-\frac{9q}{8\pi}\frac{\hbar^{2}}{m}\frac{\sin(2\pi\ell)}{2\pi\ell}\int_{k_{F}^{-}}^{k_{F}^{+}}dk\frac{2\alpha k^{4}}{\left(\frac{\hbar}{\tau}\right)^{2}+\left(2\alpha k^{3}\right)^{2}},\\ &\overline{\sigma^{N}_{xy}}=\frac{9q}{8\pi}\frac{\hbar^{2}}{m}\frac{1}{2\pi\ell}\int_{k_{F}^{-}}^{k_{F}^{+}}dk\frac{2\alpha k^{4}}{\left(\frac{\hbar}{\tau}\right)^{2}+\left(2\alpha k^{3}\right)^{2}}\frac{2\alpha k^{3}}{(\hbar/\tau)}\left(1-e^{-2\pi\ell(\hbar/\tau)/2\alpha k^{3}}\right),\end{split} (47)

The experimental value of the spin splitting is approximately 102eV10^{-2}eV at the Fermi momentum: kF0.35nm1k_{F}\approx 0.35nm^{-1} Wund2005 . The estimated Rashba coupling is approximately 0.12eVnm30.12eVnm^{-3}. The quasi-particle lifetime broadening is /τ1.2×103eV\hbar/\tau\approx 1.2\times 10^{-3}eV. The experimental value of the hole mass is m=0.27m0m=0.27m_{0}. On the other hand, according to the observation of SHC in the clean two-dimensional hole gas Wund2005 , the hole concentration (nhn_{h}) is nh=2×1012cm2n_{h}=2\times 10^{12}cm^{-2}. The Fermi wave vectors can be extracted according to the equations Loss2005

kF±\displaystyle k_{F}^{\pm} =\displaystyle= 1222mα[11(2mα2)24πnh]\displaystyle\mp\frac{1}{2}\frac{\hbar^{2}}{2m\alpha}\left[1-\sqrt{1-\left(\frac{2m\alpha}{\hbar^{2}}\right)^{2}4\pi n_{h}}\right]
+\displaystyle+ 12(22mα)2[11(2mα)24πnh]+3πnh.\displaystyle\sqrt{\frac{-1}{2}\left(\frac{\hbar^{2}}{2m\alpha}\right)^{2}\left[1-\sqrt{1-\left(\frac{2m\alpha}{\hbar}\right)^{2}4\pi n_{h}}\right]+3\pi n_{h}}.

The numerical values of the Fermi wave vectors are kF+=0.41nm1k_{F}^{+}=0.41nm^{-1} and kF=0.29nm1k_{F}^{-}=0.29nm^{-1}. The numerical results of Eq. (47) are showed in Fig. 4.We find that the total SHC oscillates and follows non-Lamor SHC which is attributed to the adiabatic evolution of ΣzN\Sigma^{N}_{z}. The experimental value 2.2(q/8π)2.2(q/8\pi) corresponds to the parameter value 6\ell\approx 6 as expected. The contribution from Larmor SHC is vanishingly small. By using =6\ell=6, we have σxyN¯=2.1(q/8π)\overline{\sigma^{N}_{xy}}=2.1(q/8\pi).

V Conclusion

We solved the spin dynamics of two-dimensional spin-orbit coupled systems in the Drude model regime, in which the momentum relaxation time is taken into account. By considering the change in the effective magnetic field induced by the applied electric field, the spin will precess around the successive change of the effective magnetic field. The spin dynamics were investigated by using the perturbation method. The spin response to the linear order of the electric field is found to be composed of Larmor and non-Larmor precessions. The Larmor motion is the precession occurring around the unperturbed effective magnetic field, and the non-Larmor motion of spin is due to the extra torque induced by the electric field. The time-dependent spin-Hall conductivity (SHC) of the Rashba system was then calculated. In the presence of disorder, when time grows larger than the relaxation time, we found that both the Larmor and non-Larmor SHCs drop to zero. However, the non-Larmor SHC decays faster than the Larmor SHC. This occurs because the presence of the disorder forces the system to achieve equilibrium, and the effective magnetic field eventually does not change over time. On the other hand, we found that the non-Larmor motion in the short time limit is exactly equal to the result obtained from the Kubo formula in the intrinsic case. In the intrinsic Rashba system, we found that the spin-Hall conductivity is not a universal constant. Furthermore, the spin-Hall conductivity vanishes when the Rashba coupling vanishes. We also calculated the Larmor and non-Larmor SHCs for the k-cubic Rahsba system. To meaningfully compare our results with the experimental value, we used the time-averaged spin-Hall conductivity. By comparing the calculated results to the experimental results, we found that the 6×2π6\times 2\pi rotation for the spin wave function and the preservation of the magnitude of spin lead to the experimental result.

Acknowledgements.
T.-W.Chen would like to thank D.-W. Chiou for valuable discussions. This work was supported by the Ministry of Science and Technology of Taiwan under Grants No. 108-2112-M-110-009 and 109-2112-M-110-006, and 108-2112-M-004-002-MY2.

Appendix A effective number of rotations

The parameter \ell is the effective number of rotations of the spin to get back to the same initial state. For the unperturbed k-cubic Rashba model,

H0=2k22m+iα(k3σ+k+3σ)=(ϵkiαk3e3iϕiαk3e3iϕϵk)\begin{split}H_{0}&=\frac{\hbar^{2}k^{2}}{2m}+i\alpha(k_{-}^{3}\sigma_{+}-k_{+}^{3}\sigma_{-})\\ &=\left(\begin{array}[]{cc}\epsilon_{k}&i\alpha k^{3}e^{-3i\phi}\\ -i\alpha k^{3}e^{3i\phi}&\epsilon_{k}\end{array}\right)\end{split} (49)

where ϵk=2k2/2m\epsilon_{k}=\hbar^{2}k^{2}/2m, σ±=(σx±iσy)/2\sigma_{\pm}=(\sigma_{x}\pm i\sigma_{y})/2, k±=kx±ikyk_{\pm}=k_{x}\pm ik_{y} and ϕ=tan1(ky/kx)\phi=\tan^{-1}(k_{y}/k_{x}). The unperturbed wave function |ψ0(t)|\psi_{0}(t)\rangle is assumed to evolve according to the Schrodinger equation

it|ψ0(t)=H0|ψ0(t)i\hbar\frac{\partial}{\partial t}|\psi_{0}(t)\rangle=H_{0}|\psi_{0}(t)\rangle (50)

The eigenstates of Eq. (49) obeying H0|±=E±|±H_{0}|\pm\rangle=E_{\pm}|\pm\rangle are given by

|±𝐤=12(±e3iϕi),\displaystyle|{\pm\mathbf{k}}\rangle=\frac{1}{\sqrt{2}}\begin{pmatrix}&\pm e^{3i\phi}&\\ &i&\end{pmatrix}, (51)

and the corresponding eigenenergies are E±𝐤=ϵkαk3E_{\pm\mathbf{k}}=\epsilon_{k}\mp\alpha k^{3}, where ϵk=2k2/2m\epsilon_{k}=\hbar^{2}k^{2}/2m. For the initial state |ψi|\psi_{i}\rangle with a wave vector kk, after time tt, the time evolution of the state is given by

|ψ0(t)=eiH0t/|ψi=eiϵkt/(e+iΩ0t/2|+𝐤+𝐤|ψi+eiΩ0t/2|𝐤𝐤|ψi),\begin{split}&|\psi_{0}(t)\rangle\\ &=e^{-iH_{0}t/\hbar}|\psi_{i}\rangle\\ &=e^{-i\epsilon_{k}t/\hbar}\left(e^{+i\Omega_{0}t/2}|{+\mathbf{k}}\rangle\langle{+\mathbf{k}}|\psi_{i}\rangle+e^{-i\Omega_{0}t/2}|{-\mathbf{k}}\rangle\langle{-\mathbf{k}}|\psi_{i}\rangle\right),\end{split} (52)

where Ω0=2αk3/\Omega_{0}=2\alpha k^{3}/\hbar is the Larmor frequency. For a latter time t+2π/Ω0t+2\pi\ell/\Omega_{0}, the state becomes

|ψ0(t+2πΩ0)=eiϵk(t+2π/Ω0)/×(eiΩ0t/2|+𝐤()+𝐤|ψi+eiΩ0t/2|𝐤()𝐤|ψi),\begin{split}&|\psi_{0}(t+\frac{2\pi\ell}{\Omega_{0}})\rangle=e^{-i\epsilon_{k}(t+2\pi\ell/\Omega_{0})/\hbar}\\ &\times\left(e^{i\Omega_{0}t/2}|{+\mathbf{k}(\ell)}\rangle\langle{+\mathbf{k}}|\psi_{i}\rangle+e^{-i\Omega_{0}t/2}|{-\mathbf{k}(\ell)}\rangle\langle{-\mathbf{k}}|\psi_{i}\rangle\right),\end{split} (53)

where

|±𝐤()=12(±e3i(ϕ±π/3)ie±iπ).\displaystyle|{\pm\mathbf{k}(\ell)}\rangle=\frac{1}{\sqrt{2}}\begin{pmatrix}&\pm e^{3i(\phi\pm\pi\ell/3)}&\\ &ie^{\pm i\pi\ell}&\end{pmatrix}. (54)

For the spinor |ψ0(t+2πΩ0)|\psi_{0}(t+\frac{2\pi\ell}{\Omega_{0}})\rangle to get back to the same state as |ψ0(t)|\psi_{0}(t)\rangle with the dynamical phase exp(iϵk(t+2π/Ω0)t/)\exp(i\epsilon_{k}(t+2\pi\ell/\Omega_{0})t/\hbar), it is required that \ell to be an even integer and a multiple of 33. Thus, \ell is given by the least common multiple of 22 and 33, which is 66. The fitting parameter \ell reflects the three-fold symmetry of the Hamiltonian and the eigenstates.

Appendix B SHC for k-cubic Rashba system

The unperturbed Hamiltonian for the k-cubic Rashba system is given by Eq. (49) which can be written as

H0=2k22m+σxdx+σydy.H_{0}=\frac{\hbar^{2}k^{2}}{2m}+\sigma_{x}d_{x}+\sigma_{y}d_{y}. (55)

The angle ϕ\phi is defined as tanϕ=ky/kx\tan\phi=k_{y}/k_{x} and dxd_{x} and dyd_{y} can be written as

dx=αk3sin(3ϕ)=α(3kx2kyky3),dy=αk3cos(3ϕ)=α(3ky2kxkx3).\begin{split}&d_{x}=\alpha k^{3}\sin(3\phi)=\alpha(3k_{x}^{2}k_{y}-k_{y}^{3}),\\ &d_{y}=-\alpha k^{3}\cos(3\phi)=\alpha(3k_{y}^{2}k_{x}-k_{x}^{3}).\end{split} (56)

Substituting Eq. (56) into Eq. (20), and noting that d=αk3d=\alpha k^{3}, then

Γyz=1d2(dxdykydydxky)=3cosϕk.\Gamma^{z}_{y}=\frac{1}{d^{2}}\left(d_{x}\frac{\partial d_{y}}{\partial k_{y}}-d_{y}\frac{\partial d_{x}}{\partial k_{y}}\right)=\frac{3\cos\phi}{k}. (57)

The non-Larmor SHC is given by

σxyN=n𝐤fn𝐤32(+nqΩ01(τ)2+Ω02Γyz)ϵkkxet/τ.\sigma^{N}_{xy}=\sum_{n\mathbf{k}}f_{n\mathbf{k}}\frac{3\hbar}{2}\left(+\frac{nq\Omega_{0}}{\hbar}\frac{1}{\left(\frac{\hbar}{\tau}\right)^{2}+\Omega_{0}^{2}}\Gamma_{y}^{z}\right)\frac{\partial\epsilon_{k}}{\hbar\partial k_{x}}e^{-t/\tau}. (58)

Substituting Eq. (57) into Eq. (58) and noting that Ω0=2αk3/\Omega_{0}=2\alpha k^{3}/\hbar, we obtain

σxyN=9q8π2mkFkF+𝑑k2αk4(τ)2+(2αk3)2et/τ.\sigma^{N}_{xy}=\frac{9q}{8\pi}\frac{\hbar^{2}}{m}\int_{k_{F}^{-}}^{k_{F}^{+}}dk\frac{2\alpha k^{4}}{\left(\frac{\hbar}{\tau}\right)^{2}+\left(2\alpha k^{3}\right)^{2}}e^{-t/\tau}. (59)

For the Larmor SHC, et/τe^{-t/\tau} is replaced by cos(Ω0t)\cos(\Omega_{0}t) and an overall negative sign is added, that is,

σxyL=9q8π2mkFkF+𝑑k2αk4(τ)2+(2αk3)2cos(2αk3t).\sigma^{L}_{xy}=-\frac{9q}{8\pi}\frac{\hbar^{2}}{m}\int_{k_{F}^{-}}^{k_{F}^{+}}dk\frac{2\alpha k^{4}}{\left(\frac{\hbar}{\tau}\right)^{2}+\left(2\alpha k^{3}\right)^{2}}\cos\left(\frac{2\alpha k^{3}}{\hbar}t\right). (60)

References

  • (1) M. I. D’yakonov, and V. I. Perel, Phys. Lett. A 35, 459 (1971); J. E. Hirsch, Phys. Rev. Lett. 83, 1834 (1999).
  • (2) J. Sinova, S. O. Valenzuela, J. Wunderlich, C. H. Back, T. Jungwirth, Rev. Mod .Phys. 87, 1213 (2015).
  • (3) S. Murakami, N. Nagaosa, and S. C. Zhang, Science 301, 1348 (2003); Phys. Rev. B 69, 235206 (2004).
  • (4) D. Xiao, M.-C. Chang, Q. Niu, Rev. Mod. Phys. 82, 1959 (2010).
  • (5) J. Sinova, D. Culcer, Q. Niu, N. A. Sinitsyn, T. Jungwirth, and A. H. MacDonald, Phys. Rev. Lett. 92, 126603 (2004).
  • (6) J. Shi, P. Zhang, D. Xiao, and Q. Niu, Phys. Rev. Lett. 96, 076604 (2006).
  • (7) S.-Q. Shen, Phys. Rev. B 70, 081311(R) (2004).
  • (8) T.-W. Chen, J.-H. Li, and C.-D. Hu, Phys. Rev. B 90, 195202 (2014).
  • (9) Y. A. Bychkov and E. I. Rashba, J. Phys. C 17, 6039 (1984).
  • (10) G. Dresselhaus, Phys. Rev. 100, 580 (1955).
  • (11) T.-W. Chen, C.-M. Huang, G. Y. Guo, Phys. Rev. B 73, 235309 (2006); T.-W. Chen, G. Y. Guo, Phys. Rev. B 79, 125301 (2009).
  • (12) S.-Q. Shen, Phys. Rev. Lett. 95, 187203 (2005); W. Zawadzki, Phys. Rev. Lett. 99, 179701 (2007); K.Y. Bliokh, Europhys. Lett. 72, 7(2005).
  • (13) T.-W. Chen, H.-C. Hsu, G. Y. Guo, Phys. Rev. B 80, 165302 (2009).
  • (14) E. M. Chudnovsky, Phys. Rev. Lett. 99, 206601 (2007); V. Ya. Kravchenko, Phys. Rev. Lett. 100, 199703 (2008); E. M. Chudnovsky, Phys. Rev. Lett. 100, 199704 (2008).
  • (15) Z. F. Jiang, R. D. Li, S.-C. Zhang and W. M. Liu, Phys. Rev. B 72, 045201, (2005).
  • (16) J. Wunderlich, B. Kaestner, J. Sinova, and T. Jungwirth, Phys. Rev. Lett. 94, 047204 (2005).
  • (17) D. Liu, S. Chen, X. Meng and N. Ma and X. Zhao, Physica B: Condens. Matt. 561, 155 (2019).
  • (18) D. V. Bulaev and D. Loss, Phys. Rev. Lett. 95, 076805 (2005).
  • (19) W. Jones and Norman H. March, Theoretical Solid State Physics, Vol. 2: Non-Equilibrium and Disorder (Dover, 1985).
  • (20) V. V. Bryksin and P. Kleinert, Phys. Rev. B 73, 165313 (2006). P. Kleinert and V. V. Bryksin, J. Phys.: Condens. Matter 18, 7497 (2006).
  • (21) B. Paul, T. K. Ghosh, Phys. Lett. A 379, 728 (2015).
  • (22) T.-W. Chen, Z.-Y. Huang, D.-W. Chiou, Phys. Lett. A 383, 2504 (2019).
  • (23) O. V. Dimitrova, Phys. Rev. B 71, 245327 (2005).
  • (24) O. Chalaev and D. Loss, Phys. Rev. B 71, 245318 (2005).
  • (25) J. Schliemann and D. Loss, Phys. Rev. B 69, 165315 (2004).
  • (26) J. Schliemann and D. Loss, Phys. Rev. B 71, 085308 (2005).
  • (27) B. A. Bernevig and S.-C. Zhang, Phys. Rev. Lett. 95, 016801 (2005).
  • (28) P. Zhang, Z. Wang, J. Shi, D. Xiao, and Q. Niu, Phys. Rev. B 77, 075304 (2008).
  • (29) K. Nomura, J. Wunderlich, J. Sinova, B. Kaestner, A. H. MacDonald, and T. Jungwirth, Phys. Rev. B 72, 245330 (2005).
  • (30) O. Bleibaum and S. Wachsmuth, Phys. Rev. B 74, 195330 (2006).
  • (31) In this study we focus on the bulk SHC and the boundary condition of a finite sample is not taken into account. On the contrary, the observed spin-Hall current is due to the spin current near edges of sample. Furthermore, various k points have different Larmor frequencies, which should be considered in the time-averaged SHC. Therefore, the time average accounting for the finite size effect, i.e., (1/T)0Tf(t)𝑑t(1/T)\int_{0}^{T}f(t)dt is not used in this paper.
  • (32) J. J. Sakurai, and Jim Napolitano, Modern Quantum Mechanics (Cambridge University Press, 2017).