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

Spin current generation due to differential rotation

Takumi Funato1,2, Shunichiro Kinoshita3,4, Norihiro Tanahashi4, Shin Nakamura4, and Mamoru Matsuo2,5,6,7 1Center for Spintronics Research Network, Keio University, Yokohama 223-8522, Japan 2Kavli Institute for Theoretical Sciences, University of Chinese Academy of Sciences, Beijing, 100190, China. 3Department of Physics, College of Humanities and Sciences, Nihon University, Tokyo 156-8550, Japan 4Department of Physics, Chuo University, Tokyo 112-8551, Japan 5CAS Center for Excellence in Topological Quantum Computation, University of Chinese Academy of Sciences, Beijing 100190, China 6RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan 7Advanced Science Research Center, Japan Atomic Energy Agency, Tokai, 319-1195, Japan
Abstract

We study nonequilibrium spin dynamics in differentially rotating systems, deriving an effective Hamiltonian for conduction electrons in the comoving frame. In contrast to conventional spin current generation mechanisms that require vorticity, our theory describes spins and spin currents arising from differentially rotating systems regardless of vorticity. We demonstrate the generation of spin currents in differentially rotating systems, such as liquid metals with Taylor-Couette flow. Our alternative mechanism will be important in the development of nanomechanical spin devices.

pacs:
Valid PACS appear here

Introduction.— Generating and controlling spin currents is a key challenge in spintronics [1]. Recent advancements of nanofabrication have enabled us to utilize mechanical motions in materials for spin transport [2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18]. In particular, the gyromagnetic effect [19, 20, 21], the conversion of mechanical angular momentum to electron spin, is increasingly significant in this context. Since Barnett, Einstein, and de Haas discovered [19, 20, 21], the gyromagnetic effect has been observed across a vast range of rotational speeds from a few Hz to 102210^{22} Hz in both magnetic and non-magnetic materials [22, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34]. Moreover, this effect defies the constraints of spin-orbit interaction strength to generate spin current [35, 36, 37, 38, 39, 40]. A prime example of the spin-orbit free mechanism relying on the gyromagnetic effect is the generation of spin currents in Cu thin films [36], conventionally considered unsuitable due to weak spin-orbit interactions, using non-uniform rotations in surface acoustic waves. This revelation opens new avenues in material selection for micro/nanomechanical spin devices.

Traditionally, spin density in steady rigid body rotation is generated through the spin-rotation coupling Hsr=𝒔𝛀H_{\rm sr}=-\bm{s}\cdot\bm{\Omega}, where 𝒔\bm{s} is spin and 𝛀\bm{\Omega} is angular velocity [41], known as the Barnett effect [19]. If this interaction can be localized, spin density gradient and spin current via diffusion may result. Previously, such ‘localization’ was embodied by the spin-vorticity coupling Hsv=𝒔𝝎H_{\rm sv}=-\bm{s}\cdot\bm{\omega}, in which the rigid angular velocity 𝛀\bm{\Omega} is promoted to the vorticity 𝝎=(1/2)×𝒗\bm{\omega}=(1/2)\bm{\nabla}\times\bm{v} of the velocity field 𝒗\bm{v} of the lattice. Here, a coupling with differential rotation 𝛀(𝒓)=𝒓×𝒗/r2\bm{\Omega}(\bm{r})=\bm{r}\times\bm{v}/r^{2} provides another way to localize the interaction, though it has been overlooked in conventional theories. This new coupling enables spin current generation in systems with non-uniform rotational motion but without vorticity, and it may expand our understanding of spin transport driven by non-uniform rotation.

In this study, we investigate the non-equilibrium spin dynamics in differentially rotating systems within a microscopic theory. By mapping into a comoving frame, we construct an effective Hamiltonian for conduction electrons in these systems, demonstrating the emergence of effective gauge fields. Furthermore, we derive microscopic expressions for the spin density and spin current of conduction electrons driven by these emergent gauge fields. By applying this to a liquid metal and a non-magnetic metallic cantilever as examples of differentially rotating systems, we estimate the concrete amount of the spin current. In particular, we show that even in cases such as Taylor-Couette flow where the vorticity-gradient is zero, spin currents can be generated due to the differential rotation. Consequently, we uncover mechanisms of angular momentum transfer that have not been captured by traditional frameworks, specifically those involving temporally and spatially modulated lattices transferring momentum to conduction electron spins. Our results will contribute to the rapidly expanding field of non-equilibrium spin physics in nanomechanical systems.

Emergent Gauge Fields in Comoving Frame.—We consider the free electron system subject to momentum scattering and spin-orbit scattering due to the impurities. In the inertial laboratory frame the Hamiltonian is given by

H^\displaystyle\hat{H}^{\prime} =d3xψ^(𝒙){22m2ϵF+Vimp(𝒙,t)\displaystyle=\int d^{3}x\hat{\psi}^{\dagger}(\bm{x})\biggl{\{}-\frac{\hbar^{2}}{2m}\bm{\nabla}^{2}-\epsilon_{F}+V^{\prime}_{\text{imp}}(\bm{x},t)
+λso𝝈[Vimp(𝒙,t)×(i)]}ψ^(𝒙),\displaystyle\quad+\lambda_{\text{so}}\bm{\sigma}\cdot[\bm{\nabla}V^{\prime}_{\text{imp}}(\bm{x},t)\times(-i\hbar\bm{\nabla})]\biggr{\}}\hat{\psi}(\bm{x}), (1)

where ψ^(𝒙)\hat{\psi}(\bm{x}) is the electron field operator, ϵF\epsilon_{F} is the Fermi energy, 𝝈=(σx,σy,σz)\bm{\sigma}=(\sigma^{x},\sigma^{y},\sigma^{z}) are the Pauli matrices, and λso\lambda_{\text{so}} is the strength of the spin-orbit interaction. The third term represents the impurity scattering and the fourth term represents the spin-orbit scattering. Here, Vimp(𝒙,t)=ju(𝒙𝒓j(t))V^{\prime}_{\text{imp}}(\bm{x},t)=\sum_{j}u(\bm{x}-\bm{r}^{\prime}_{j}(t)) is the total impurity potential, where u(𝒙𝒓j(t))u(\bm{x}-\bm{r}^{\prime}_{j}(t)) is a single impurity potential due to the jj-th impurity located at the position 𝒓j(t)\bm{r}^{\prime}_{j}(t). It is worth noting that the electrons are subject to the moving impurities because we suppose the total system is differentially rotating. To characterize the differential rotation of the system, we introduce a rotation angle Φ(𝒙,t)\Phi(\bm{x},t) around the zz-axis, which is chosen as a rotation axis. When we take a cylindrical coordinate system, the coordinate transformation from the laboratory frame 𝒓=(r,φ,z)\bm{r}^{\prime}=(r^{\prime},\varphi^{\prime},z^{\prime}) to the rotating frame 𝒓=(r,φ,z)\bm{r}=(r,\varphi,z) can be written as r=rr=r^{\prime}, z=zz=z^{\prime}, and φ=φΦ(𝒓,t)\varphi=\varphi^{\prime}-\Phi(\bm{r}^{\prime},t). Note that Φ\Phi is independent of φ\varphi, i.e., φΦ=0\partial_{\varphi}\Phi=0 because of axisymmetry. Supposing Φ(𝒙,t)=0\Phi(\bm{x},t)=0 at an initial time t=0t=0, the position of the jj-th impurity at tt is given by 𝒓j(t)=z[Φ(𝒓j,t)]𝒓j\bm{r}^{\prime}_{j}(t)=\mathcal{R}_{z}[\Phi(\bm{r}_{j},t)]\bm{r}_{j}, where z\mathcal{R}_{z} denotes rotation around the zz-axis and 𝒓j\bm{r}_{j} is the position at the initial time.

Now, we define a generator of the differential rotation with angle Φ(𝒙,t)\Phi(\bm{x},t) as

Q^Φ(t)=d3xΦ(𝒙,t)ψ^(𝒙)Jzψ^(𝒙),\hat{Q}_{\Phi}(t)=\int d^{3}x\Phi(\bm{x},t)\hat{\psi}^{\dagger}(\bm{x})J^{z}\hat{\psi}(\bm{x}), (2)

where JzJ^{z} is the total angular momentum operator acting on coordinates and spin space as Jz=iφ+σz/2J^{z}=-i\hbar\partial_{\varphi}+\hbar\sigma^{z}/2. Note that JzJ^{z} and Φ(𝒙,t)\Phi(\bm{x},t) are commutative. For an arbitrary state vector in the laboratory frame |Ψ(t)|\Psi^{\prime}(t)\rangle, the state vector in the rotating frame is given by

|Ψ(t)=exp[iQ^Φ(t)]|Ψ(t).|\Psi(t)\rangle=\exp\left[\frac{i}{\hbar}\hat{Q}_{\Phi}(t)\right]|\Psi^{\prime}(t)\rangle. (3)

The Schrödinger equation in the laboratory frame, it|Ψ(t)=H^|Ψ(t)i\hbar\partial_{t}|\Psi^{\prime}(t)\rangle=\hat{H}^{\prime}|\Psi^{\prime}(t)\rangle, yields

it|Ψ(t)\displaystyle i\hbar\frac{\partial}{\partial t}|\Psi(t)\rangle =(eiQ^Φ/H^eiQ^Φ/Q^tΦ)|Ψ(t)\displaystyle=(e^{i\hat{Q}_{\Phi}/\hbar}\hat{H}^{\prime}e^{-i\hat{Q}_{\Phi}/\hbar}-\hat{Q}_{\partial_{t}\Phi})|\Psi(t)\rangle (4)
=H^T|Ψ(t),\displaystyle=\hat{H}_{T}|\Psi(t)\rangle,

where Q^tΦ=d3xtΦ(𝒙,t)ψ^(𝒙)Jzψ^(𝒙)\hat{Q}_{\partial_{t}\Phi}=\int d^{3}x\partial_{t}\Phi(\bm{x},t)\hat{\psi}^{\dagger}(\bm{x})J^{z}\hat{\psi}(\bm{x}) and Q^Φ\hat{Q}_{\Phi} commute because of φΦ=0\partial_{\varphi}\Phi=0. The Hamiltonian HTH_{T} which governs dynamics in the rotating frame. The density operator in the rotating frame, ρ^(t)\hat{\rho}(t), is given by ρ^(t)=eiQ^Φ/ρ^(t)eiQ^Φ/\hat{\rho}(t)=e^{i\hat{Q}_{\Phi}/\hbar}\hat{\rho}^{\prime}(t)e^{-i\hat{Q}_{\Phi}/\hbar}, where ρ^(t)\hat{\rho}^{\prime}(t) is the density operator in the laboratory frame. The time evolution of ρ^(t)\hat{\rho}(t) is determined by itρ^(t)=[H^T,ρ^(t)]i\hbar\partial_{t}\hat{\rho}(t)=[\hat{H}_{T},\hat{\rho}(t)]. Assuming that the single impurity potential u(𝒙)u(\bm{x}) is isotropic and its typical range aa, such that u(𝒙)0u(\bm{x})\simeq 0 for |𝒙|a|\bm{x}|\gg a, is much smaller than a typical scale of the gradient of the differential rotation, i.e., a|Φ|1a|{\bm{\nabla}}\Phi|\ll 1, the Hamiltonian in the rotating frame can be rewritten as

H^T=d3xψ^(𝒙){12m(i𝑨sJz)2As,0JzϵF\displaystyle\hat{H}_{T}=\int d^{3}x\hat{\psi}^{\dagger}(\bm{x})\biggl{\{}\frac{1}{2m}(-i\hbar{\bm{\nabla}}-{\bm{A}}_{s}J^{z})^{2}-A_{s,0}J^{z}-\epsilon_{F}
+Vimp(𝒙)+λso𝝈[Vimp(𝒙)×(i𝑨sJz)]}ψ^(𝒙),\displaystyle+V_{\text{imp}}(\bm{x})+\lambda_{\text{so}}\bm{\sigma}\cdot[{\bm{\nabla}}V_{\text{imp}}(\bm{x})\times(-i\hbar{\bm{\nabla}}-{\bm{A}}_{s}J^{z})]\biggr{\}}\hat{\psi}(\bm{x}), (5)

where the time and spatial derivatives of the rotation angle are denoted by

As,μ(𝒙,t)=(tΦ(𝒙,t),Φ(𝒙,t))(μ=0,x,y,z).\displaystyle A_{s,\mu}(\bm{x},t)=\Bigl{(}\partial_{t}\Phi(\bm{x},t),{\bm{\nabla}}\Phi(\bm{x},t)\Bigr{)}\quad(\mu=0,x,y,z). (6)

We call As,μ(𝒙,t)A_{s,\mu}(\bm{x},t) “emergent gauge field” in this paper. In the rotating frame, the effects of the differential rotation are represented by the emergent gauge fields, whereas the impurity potential given by Vimp(𝒙)=ju(𝒙𝒓j)V_{\text{imp}}(\bm{x})=\sum_{j}u(\bm{x}-\bm{r}_{j}) does not depend on time under the assumption a|Φ|1a|{\bm{\nabla}}\Phi|\ll 1.

Setup.—We present the Fourier representation of the total Hamiltonian in the rotating frame to facilitate calculations: H^T=H^0+H^imp+H^so+H^(t)\hat{H}_{T}=\hat{H}_{0}+\hat{H}_{\text{imp}}+\hat{H}_{\text{so}}+\hat{H}^{\prime}(t), where H^(t)\hat{H}^{\prime}(t) is the contribution of the emergent gauge field, and we treat it as a perturbation. The first term H^0=𝒌ϵ𝒌ψ^𝒌ψ^𝒌\hat{H}_{0}=\sum_{\bm{k}}\epsilon_{\bm{k}}\hat{\psi}_{\bm{k}}^{\dagger}\hat{\psi}_{\bm{k}} represents the kinetic term, where ϵ𝒌=2k2/2mϵF\epsilon_{\bm{k}}=\hbar^{2}k^{2}/2m-\epsilon_{F} is the kinetic energy, and ψ^𝒌\hat{\psi}_{\bm{k}} is the Fourier component of the electron annihilation operator. The second and third terms describe the momentum scattering and the spin-orbit scattering due to the impurities, respectively. These are expressed as H^imp=𝒌𝒌V𝒌𝒌ψ^𝒌ψ^𝒌\hat{H}_{\text{imp}}=\sum_{\bm{k}\bm{k}^{\prime}}V_{\bm{k}-\bm{k}^{\prime}}\hat{\psi}_{\bm{k}}^{\dagger}\hat{\psi}_{\bm{k}^{\prime}} and H^so=iλso𝒌𝒌V𝒌𝒌(𝒌×𝒌)ψ^𝒌𝝈ψ^𝒌\hat{H}_{\text{so}}=i\hbar\lambda_{\text{so}}\sum_{\bm{k}\bm{k}^{\prime}}V_{\bm{k}-\bm{k}^{\prime}}(\bm{k}\times\bm{k}^{\prime})\cdot\hat{\psi}_{\bm{k}}^{\dagger}\bm{\sigma}\hat{\psi}_{\bm{k}^{\prime}}, where V𝒌V_{\bm{k}} denotes the Fourier component of the impurity potential Vimp(𝒙)V_{\text{imp}}(\bm{x}). We assume a short-range impurity potential, i.e., u(𝒙𝒓j)=uiδ(𝒙𝒓j)u(\bm{x}-\bm{r}_{j})=u_{\text{i}}\delta(\bm{x}-\bm{r}_{j}), where uiu_{\text{i}} is the strength of the impurity potential defined by ui=d3xu(𝒙)u_{\text{i}}=\int d^{3}xu(\bm{x}) in general. While, the perturbed part, denoted by H^(t)=H^s+(Lz)\hat{H}^{\prime}(t)=\hat{H}_{s}+\order{L^{z}} with Lz=iφL^{z}=-i\hbar\partial_{\varphi} being the orbital angular momentum, represents the effect of the emergent gauge fields. The Hamiltonian H^s\hat{H}_{s} incorporates the electron spin, given by

H^s\displaystyle\hat{H}_{s} =22m𝒌𝒌𝒒ψ^𝒌+(𝒌σzδ𝒌𝒌12𝑨s,𝒌𝒌)ψ^𝒌𝑨s(𝒒)\displaystyle=-\frac{\hbar^{2}}{2m}\sum_{\bm{k}\bm{k}^{\prime}\bm{q}}\hat{\psi}_{\bm{k}_{+}}^{\dagger}\left(\bm{k}\sigma^{z}\delta_{\bm{k}\bm{k}^{\prime}}-\frac{1}{2}\bm{A}_{s,\bm{k}-\bm{k}^{\prime}}\right)\hat{\psi}_{\bm{k}^{\prime}_{-}}\cdot\bm{A}_{s}(\bm{q})
2s^(𝒒)As,0(𝒒),\displaystyle\quad-\frac{\hbar}{2}\hat{s}(\bm{q})A_{s,0}(\bm{q}), (7)

where 𝑨s,𝒒\bm{A}_{s,\bm{q}} is the Fourier component of the emergent gauge fields, and 𝒌±=𝒌±𝒒/2\bm{k}_{\pm}=\bm{k}\pm\bm{q}/2 are defined.

To define the spin-current operator, we consider the temporal modulation of the zz-polarized spin density, ts^(𝒒)=i𝒒𝒋^s(𝒒)+𝒯^𝒒\partial_{t}\hat{s}(\bm{q})=-i\bm{q}\cdot\hat{\bm{j}}_{s}(\bm{q})+\hat{\mathcal{T}}_{\bm{q}}, where s^(𝒒)=𝒌ψ^𝒌σzψ^𝒌+\hat{s}(\bm{q})=\sum_{\bm{k}}\hat{\psi}^{\dagger}_{\bm{k}_{-}}\sigma^{z}\hat{\psi}_{\bm{k}_{+}} is the spin-density operator, and 𝒯^𝒒\hat{\mathcal{T}}_{\bm{q}} describes the spin torque due to the spin-orbit interaction of the impurities. The spin-current density operator polarized in the zz-direction is defined by 𝒋^s(𝒒)=𝒌𝒌ψ^𝒌𝒋s,𝒌𝒌ψ^𝒌+\hat{\bm{j}}_{s}(\bm{q})=\sum_{\bm{k}\bm{k}^{\prime}}\hat{\psi}^{\dagger}_{\bm{k}_{-}^{\prime}}\bm{j}_{s,\bm{k}^{\prime}\bm{k}}\hat{\psi}_{\bm{k}_{+}}, where the matrix elements 𝒋s,𝒌𝒌\bm{j}_{s,\bm{k}^{\prime}\bm{k}} is given by

𝒋s,𝒌𝒌=δ𝒌𝒌𝒗𝒌σz+λsoV𝒌𝒌[𝒆z×(𝒌𝒌)]𝑨s,𝒌𝒌2m,\displaystyle\bm{j}_{s,\bm{k}^{\prime}\bm{k}}=\delta_{\bm{k}^{\prime}\bm{k}}\bm{v}_{\bm{k}}\sigma^{z}+\lambda_{\text{so}}V_{\bm{k}^{\prime}-\bm{k}}[\bm{e}_{z}\times(\bm{k}^{\prime}-\bm{k})]-\frac{\hbar\bm{A}_{s,\bm{k}^{\prime}-\bm{k}}}{2m}, (8)

where 𝒗𝒌=𝒌/m\bm{v}_{\bm{k}}=\hbar\bm{k}/m is the velocity and 𝒆z\bm{e}_{z} is the unit vector in zz-direction.

Calculation of Spin Current.—We now compute the spin current induced by the emergent gauge fields. The statistical average of the spin density and spin current is given by

j^μ(𝒒,ω)=dϵ2πi𝒌𝒌Tr[jsμ,𝒌𝒌G𝒌+,𝒌<(ϵ+,ϵ)],\displaystyle\langle\hat{j}_{\mu}(\bm{q},\omega)\rangle=\int^{\infty}_{-\infty}\frac{d\epsilon}{2\pi i}\sum_{\bm{k}\bm{k}^{\prime}}\Tr\left[j_{s\mu,\bm{k}^{\prime}\bm{k}}G^{<}_{\bm{k}_{+},\bm{k}^{\prime}_{-}}(\epsilon_{+},\epsilon_{-})\right], (9)

where ϵ±=ϵ±ω/2\epsilon_{\pm}=\epsilon\pm\omega/2, js0,𝒌𝒌=σzδ𝒌𝒌j_{s0,\bm{k}^{\prime}\bm{k}}=\sigma^{z}\delta_{\bm{k}^{\prime}\bm{k}}, and the trace is taken for the spin space. Here, the four-vector j^μ=(s^,𝒋^s)\hat{j}_{\mu}=(\hat{s},\hat{\bm{j}}_{s}) represents the spin density and spin current operators. The function G𝒌+,𝒌<(ϵ+,ϵ)G^{<}_{\bm{k}_{+},\bm{k}^{\prime}_{-}}(\epsilon_{+},\epsilon_{-}) is the lesser component of the nonequilibrium path-ordered Green function, defined by G𝒌,𝒌(t,t)=iTKψ^𝒌+(t)ψ^𝒌(t)G_{\bm{k},\bm{k}^{\prime}}(t,t^{\prime})=-i\langle T_{K}\hat{\psi}_{\bm{k}_{+}}(t)\hat{\psi}^{\dagger}_{\bm{k}^{\prime}_{-}}(t^{\prime})\rangle, where TKT_{K} is a path-ordering operator, ψ^(t)=U^(t)ψ^(t)U^(t)\hat{\psi}(t)=\hat{U}^{\dagger}(t)\hat{\psi}(t)\hat{U}(t) is the Heisenberg representation with U^(t)=Texp[(i/)tH^T(τ)𝑑τ]\hat{U}(t)=T\text{exp}[-(i/\hbar)\int^{t}_{-\infty}\hat{H}_{T}(\tau)d\tau] and TT being time-ordering operator, and =tr(ρ^)\langle\cdots\rangle=\text{tr}(\hat{\rho}\cdots) represents the expectation value with the density operator ρ^\hat{\rho}.

Assuming that the characteristic energy scales of the momentum scattering and the spin-orbit scattering due to the impurities are much smaller than the Fermi energy, i.e., niuiϵFn_{\text{i}}u_{\text{i}}\ll\epsilon_{F} and 2λso2kF41\hbar^{2}\lambda_{\text{so}}^{2}k_{F}^{4}\ll 1, we treat them in the Born approximation. With the uniformly random distribution of impurities, we perform the average of their positions to obtain the retarded/advanced Green function: g𝒌r/a(ϵ)=1/(ϵϵ𝒌±iγ)g^{r/a}_{\bm{k}}(\epsilon)=1/(\epsilon-\epsilon_{\bm{k}}\pm i\hbar\gamma), where γ=πniui2ν0(1+22λso2kF4/3)\hbar\gamma=\pi n_{\text{i}}u_{\text{i}}^{2}\nu_{0}(1+2\hbar^{2}\lambda_{\text{so}}^{2}k_{F}^{4}/3) is the damping constant calculated with the density of state per spin at Fermi level ν0=mkF/2π22\nu_{0}=mk_{F}/2\pi^{2}\hbar^{2}. We assume that γϵF\hbar\gamma\ll\epsilon_{F}. This condition is well-satisfied when uiν01u_{\text{i}}\nu_{0}\lesssim 1.

The spin-current density in linear response to the emergent gauge fields is expressed as j^μ(𝒒,ω)=Kμν(ω)As,ν(𝒒,ω)\langle\hat{j}_{\mu}(\bm{q},\omega)\rangle=K_{\mu\nu}(\omega)A_{s,\nu}(\bm{q},\omega), where Kμν(ω)K_{\mu\nu}(\omega) is the response function. It is presumed that the time and spatial variation of the differential rotation are much slower than the electron mean-free path l=vFτl=v_{F}\tau and momentum relaxation time τ=1/2γ\tau=1/2\gamma, respectively, i.e., l|Φ|1l|\bm{\nabla}\Phi|\ll 1 and τ|tΦ|1\tau|\partial_{t}\Phi|\ll 1, where vF=kF/mv_{F}=\hbar k_{F}/m is the Fermi velocity and kF=2mϵF/2k_{F}=\sqrt{2m\epsilon_{F}/\hbar^{2}} is the Fermi wavenumber. In terms of Fourier space, conditions lq1lq\ll 1 and τω1\tau\omega\ll 1 hold. By including the ladder vertex corrections due to the impurities, and using the relations vFq/2γϵF\hbar v_{F}q/2\ll\hbar\gamma\ll\epsilon_{F} and ω/2γϵF\hbar\omega/2\ll\hbar\gamma\ll\epsilon_{F}, the response function is calculated as

Kμν(ω)=δμ0δν0σc2e2D\displaystyle K_{\mu\nu}(\omega)=\delta_{\mu 0}\delta_{\nu 0}\frac{\hbar\sigma_{c}}{2e^{2}D}
+iω4π𝒌v𝒌,μTr[σzg+rσz(v𝒌,ν+Λνs)ga],\displaystyle\qquad+i\omega\frac{\hbar}{4\pi}\sum_{\bm{k}}v_{\bm{k},\mu}\Tr[\sigma^{z}g^{r}_{+}\sigma^{z}(v_{\bm{k},\nu}+\Lambda_{\nu}^{s})g^{a}_{-}], (10)

where σc=nee2τ/m\sigma_{c}=n_{e}e^{2}\tau/m is the Drude conductivity with ne=4ϵFν0/3n_{e}=4\epsilon_{F}\nu_{0}/3 being the number density of the electrons and e(>0)e(>0) being the elementary charge, and D=vF2τ/3D=v_{F}^{2}\tau/3 is the diffusion constant. We set v𝒌,0=1v_{\bm{k},0}=1 and v𝒌,i=ki/mv_{\bm{k},i}=\hbar k_{i}/m. Here, Λνs\Lambda^{s}_{\nu} describes the three-point vertex corrections, and g±r/a=g𝒌±r/a(±ω/2)g^{r/a}_{\pm}=g^{r/a}_{\bm{k}_{\pm}}(\pm\omega/2) are specified. The first term of the response function represents the spin susceptibility for the rigid rotation [42], known as the Barnett effect.

Performing straightforward calculation, we derive the rotation-induced spin density and spin current:

s^(𝒒,ω)\displaystyle\langle\hat{s}(\bm{q},\omega)\rangle =iωσc2e2Dτs1Dq2iω+τs1Φ(𝒒,ω),\displaystyle=-i\omega\frac{\hbar\sigma_{c}}{2e^{2}D}\frac{\tau_{s}^{-1}}{Dq^{2}-i\omega+\tau_{s}^{-1}}\Phi(\bm{q},\omega), (11)
𝒋^s(𝒒,ω)\displaystyle\langle\hat{\bm{j}}_{s}(\bm{q},\omega)\rangle =iωσc2e2τs1Dq2iω+τs1i𝒒Φ(𝒒,ω),\displaystyle=i\omega\frac{\hbar\sigma_{c}}{2e^{2}}\frac{\tau_{s}^{-1}}{Dq^{2}-i\omega+\tau_{s}^{-1}}i{\bm{q}}\Phi(\bm{q},\omega), (12)

where τs=9τ/82λso2kF4\tau_{s}=9\tau/8\hbar^{2}\lambda_{\text{so}}^{2}k_{F}^{4} is the spin-relaxation time. Combining these results in the real space, we obtain Fick’s law, 𝒋s=Ds{\bm{j}}_{s}=-D\bm{\nabla}s. This implies that our spin current is a diffusive flow produced by the gradient of the spin density, in which the impurity scattering governs the diffusion.

Now, we focus on long-term dynamics such that time scales are longer than the period of the rotation, ωΩ\omega\lesssim\Omega. If the spin relaxation is much faster than typical scales of the angular velocity and the spatial variation of the differential rotation, i.e., Dq2,ωΩτs1Dq^{2},\omega\lesssim\Omega\ll\tau_{s}^{-1}, which is well-satisfied in metals, the rotation-induced spin current reduces to the following form in the real space:

𝒋s(𝒙,t)=σc2e2tΦ(𝒙,t).\displaystyle\bm{j}_{s}(\bm{x},t)=-\frac{\hbar\sigma_{c}}{2e^{2}}\bm{\nabla}\partial_{t}\Phi(\bm{x},t). (13)

In addition, the rotation-induced spin density reduces to

s(𝒙,t)=σc2e2DtΦ(𝒙,t),s(\bm{x},t)=\frac{\hbar\sigma_{c}}{2e^{2}D}\partial_{t}\Phi(\bm{x},t), (14)

which is the Barnett effect generalized to differential rotations. The susceptibility given by σc/2e2D\hbar\sigma_{c}/2e^{2}D is identical to that of the Barnett effect for rigid rotations. These results suggest that the spin density and current are polarized along the rotation axis and the spin current is driven in the direction of the spatial gradient of the angular velocity. By contrast, if the spin relaxation is so slow that τs1ωΩ\tau_{s}^{-1}\ll\omega\lesssim\Omega, the spin density (11) as well as the spin current (12) vanish, which implies that the spin relaxation is necessary to generate the spin current and spin density. Despite this fact, the magnitude of the spin current (13) is independent of the spin relaxation time.

The absence of τs\tau_{s} from the long-term dynamics of the spin density and the spin current is explained as follows. In the response function (10), the first term that originates from the spin-rotation coupling As,0JzA_{s,0}J^{z} in (5) is principal, while the other terms including the spin-orbit coupling are suppressed by τsω1\tau_{s}\omega\ll 1. This means that the spin density is determined only by the susceptibility of the Barnett effect and the angular velocity. The gradient of this spin density produces the spin current due to the diffusion caused by the impurity scattering, as shown. Thus, the spin density and current are independent of τs\tau_{s}.

The spin-orbit interaction contributes only to the transient process that is necessary to drive the system to the final steady state, but it does not contribute to long-term dynamics. Indeed, for ωΩ\omega\gtrsim\Omega, (11) and (12) provide the following spin transport equation:

st+𝒋s=sτs+σc2e2DτstΦ,\frac{\partial s}{\partial t}+{\bm{\nabla}}\cdot{\bm{j}}_{s}=-\frac{s}{\tau_{s}}+\frac{\hbar\sigma_{c}}{2e^{2}D\tau_{s}}\partial_{t}\Phi, (15)

which describes the transient process with a time-scale ωτs1\omega\simeq\tau_{s}^{-1}. We expect to obtain similar diffusive spin currents as long as there are not only the spin-orbit interactions as presented here but also other interactions that can produce transient processes satisfying Ωτs1τ1\Omega\ll\tau_{s}^{-1}\ll\tau^{-1}.

Taylor-Couette Flow.—As an explicit example, let us consider a two-dimensional steady flow with concentric circular streamlines. In this case, the flow velocity is parallel to the φ\varphi-direction, 𝒗=(0,vφ,0)\bm{v}=(0,v_{\varphi},0), satisfying the following Navier-Stokes equation: r2vφ+(rvφ)/rvφ/r2=0\partial_{r}^{2}v_{\varphi}+(\partial_{r}v_{\varphi})/r-v_{\varphi}/r^{2}=0. The general solution is vφ=c1/r+c2rv_{\varphi}=c_{1}/r+c_{2}r with integration constants c1c_{1} and c2c_{2} determined by boundary conditions. The first term represents irrotational flow, while the second term represents rigid-rotation flow. We consider the two infinitely long coaxial cylinders of radii r1r_{1} and r2r_{2} (r2>r1r_{2}>r_{1}), and the inner and outer cylinders are rotating at constant angular velocities Ω1\Omega_{1} and Ω2\Omega_{2}, respectively. Under these boundary conditions, vφ(r1)=r1Ω1v_{\varphi}(r_{1})=r_{1}\Omega_{1} and vφ(r2)=r2Ω2v_{\varphi}(r_{2})=r_{2}\Omega_{2}, the constants are obtained as c1=(Ω1Ω2)r12r22/(r22r12)c_{1}=(\Omega_{1}-\Omega_{2})r_{1}^{2}r_{2}^{2}/(r_{2}^{2}-r_{1}^{2}) and c2=(Ω2r22Ω1r12)/(r22r12)c_{2}=(\Omega_{2}r_{2}^{2}-\Omega_{1}r_{1}^{2})/(r_{2}^{2}-r_{1}^{2}). This concentric steady flow, known as the Taylor-Couette flow [43], induces the steady differential rotation with angular velocity Ω(r)=c1/r2+c2\Omega(r)=c_{1}/r^{2}+c_{2}, leading to the generation of spin current (see Fig. 1(a)):

𝒋s(r)=𝒆rσce2r12r22r22r12Ω1Ω2r3,\displaystyle\bm{j}_{s}(r)=\bm{e}_{r}\frac{\hbar\sigma_{c}}{e^{2}}\frac{r_{1}^{2}r_{2}^{2}}{r_{2}^{2}-r_{1}^{2}}\frac{\Omega_{1}-\Omega_{2}}{r^{3}}, (16)

where 𝒆r\bm{e}_{r} being the unit vector in the rr-direction. Notably, since the vorticity in this system is constant, ×𝒗=2c2𝒆z\bm{\nabla}\times\bm{v}=2c_{2}\bm{e}_{z}, the conventional spin currents owing to the spin-vorticity coupling, which require the vorticity gradient [44, 35] or time-dependent vorticity [45], do not appear. On the other hand, our theory predicts the generation of spin current even in vorticity-free cases c2=0c_{2}=0.

To estimate the magnitude of the spin current, we assume that the radii of the two cylinders are much larger than the gap between them d=r2r1d=r_{2}-r_{1}, i.e., r1,r2dr_{1},r_{2}\gg d, and only the outer cylinder is rotating, Ω1=0\Omega_{1}=0 and Ω20\Omega_{2}\neq 0, for simplicity. Under this assumption, the spin current is approximated as js=σcΩ2/2e2dj_{s}=\hbar\sigma_{c}\Omega_{2}/2e^{2}d. We consider (Ga,In)Sn as the fluid with the electric conductivity σc=3.26×106(Ωm)1\sigma_{c}=3.26\times 10^{6}(\Omega\cdot\text{m})^{-1} [38]. Set d1μd\sim 1\mum and Ω2102\Omega_{2}\sim 10^{2}kHz, the magnitude of the spin current in charge current units is estimated as ejs1.07×102Am-2ej_{s}\sim 1.07\times 10^{2}\text{A$\cdot$m${}^{-2}$}.

Torsional Oscillation of Cantilever.—As another example, we focus on the torsional oscillation of a cantilever, wherein one of the ends is securely fixed while external forces are exerted on the opposite end. These forces induce only a twisting motion in the cantilever, not bending or other deformations. In this case, the angular velocity of the system varies along the rotation axis rather than the radial direction. The distortion angle φ(z,t)\varphi(z,t) of the cantilever dictates the subsequent equation of motion: Cz2φ=ρmIt2φC\partial_{z}^{2}\varphi=\rho_{m}I\partial_{t}^{2}\varphi, where CC is the torsional rigidity, ρm\rho_{m} is the mass density, and II is the moment of inertia of the cross-section about its center of mass. By solving the equation of motion under the boundary conditions φ(0,t)=0\varphi(0,t)=0 and zφ(l,t)=0\partial_{z}\varphi(l,t)=0 and considering the initial conditions φ(l,0)=φ0\varphi(l,0)=\varphi_{0} and tφ(z,0)=0\partial_{t}\varphi(z,0)=0, we derive the solution as φn(z,t)=φ0sinknzcosωnt\varphi_{n}(z,t)=\varphi_{0}\sin k_{n}z\cos\omega_{n}t, where kn=(2n1)π/2lk_{n}=(2n-1)\pi/2l and ωn=vkn\omega_{n}=vk_{n} with the integer n1n\geq 1 and the velocity v=C/ρmIv=\sqrt{C/\rho_{m}I}. The spin current, driven by the nn-th torsional oscillation of cantilever, flows along the zz-direction as given by (see Fig. 1(b))

𝒋s,n(z,t)=𝒆zσcφ0v2e2kn2cosknzsinωnt.\displaystyle\bm{j}_{s,n}(z,t)=\bm{e}_{z}\frac{\hbar\sigma_{c}\varphi_{0}v}{2e^{2}}k_{n}^{2}\cos k_{n}z\sin\omega_{n}t. (17)

The mechanism under investigation in this study represents a universal phenomenon, irrespective of material choice, and fundamentally distinct from the previous theory [46] that focus solely on magnetic materials.

Finally, we estimate the magnitude of the spin current driven by the torsional oscillation. For a plate-shaped cantilever with width aa, thickness bb and length ll, the quantities CC and II are calculated as Cμab3/3C\simeq\mu ab^{3}/3 and Ia3b/12I\simeq a^{3}b/12 (aba\gg b) with Lamé constant μ\mu. The magnitude of the total spin current in charge current units is denoted by Jn=eabjs,n(0,0)J_{n}=eabj_{s,n}(0,0), while that attributed to the first torsional oscillation mode is given by J1=(π2σcφ0/4e)(b/l)2μ/ρmJ_{1}=(\pi^{2}\hbar\sigma_{c}\varphi_{0}/4e)(b/l)^{2}\sqrt{\mu/\rho_{m}}. We consider that the cantilever is composed of copper with weak spin-orbit interaction. By using the charge conductivity σc=6.45×107Ω1m1\sigma_{c}=6.45\times 10^{7}\,\Omega^{-1}\text{m}^{-1}, the Lamé constant μ=48.3GHz\mu=48.3\,\text{GHz} and the mass density ρm=8.96×103kg/m3\rho_{m}=8.96\times 10^{3}\,\text{kg}/\text{m}^{3}, the total spin current is estimated as J10.15μJ_{1}\sim 0.15\,\muA for φ00.01\varphi_{0}\sim 0.01 and b/l=1/4b/l=1/4.

Refer to caption
Figure 1: Schematic illustration showing the generation of spin current due to (a) the Taylor-Couette flow in a liquid metal and (b) torsional motion of a cantilever.

Conclusion and discussion.— We studied non-equilibrium spin dynamics in differentially rotating systems within a microscopic theory. We obtained a Hamiltonian with emergent gauge fields by using a mapping to the comoving frame of the differential rotation. We estimated the spin currents generated by differential rotation in a liquid metal and a non-magnetic metallic nanomechanical system for experimental reference. Our mechanism produces spin currents even in vorticity-free systems. Hence, this mechanism of spin current generation is novel and distinct from the known mechanisms based on the spin-vorticity coupling.

Distinctions between our mechanism and those proposed in other literature [45, 44, 35] can be summarized as follows. In terms of spin transport equations, the source terms that violate the conservation of the spin density differ in the mechanisms. The spin current is proportional to the spatial gradient of these source terms. For our case, the spin transport equation is given in (15) and the source term is proportional to Ω\Omega, the angular velocity of the orbital rotation around a fixed axis. On the other hand, the source term given in [45] and those in [44, 35] are proportional to tω~\partial_{t}\tilde{\omega} and ω~\tilde{\omega}, respectively, where ω~\tilde{\omega} is the vorticity. These distinctions can be detected experimentally. In an irrotational flow, the vorticity is zero, preventing the spin current generation due to the spin-vorticity coupling. Creating an irrotational flow akin to a differentially rotating system allows us to detect spin currents specific to our mechanism. One may achieve such flow in the Taylor-Couette flow by taking appropriate boundary conditions that realize c2=0c_{2}=0 with c10c_{1}\neq 0. Furthermore, we can distinguish the mechanisms even in the case of c20c_{2}\neq 0. Since the vorticity in this system is uniform and time-independent, neither the source term in [45] nor that in [44, 35] can contribute to the spin current. As a result, the Taylor-Couette flow can generate the spin current in our mechanism but not in the other ones. Exploring these experiments would be highly valuable.

We have comments on possible pictures that come from (13). If we interpret As,0=tΦA_{s,0}=\partial_{t}\Phi as a “chemical potential” for the spin density, (13) suggests that the diffusive spin current is produced as a result of the position-dependent “chemical potential” for the spin density. Also, rewriting (13) as 𝒋s(𝒙,t)/ne=[(/2)tΦ]τ/m\bm{j}_{s}(\bm{x},t)/n_{e}=[-(\hbar/2)\bm{\nabla}\partial_{t}\Phi]\tau/m, we may interpret that a “thermodynamic force” (/2)tΦ-(\hbar/2)\bm{\nabla}\partial_{t}\Phi is acting on the electrons depending on their spin, since τ/m\tau/m is the mobility of the electron and 𝒋s(𝒙,t)/ne\bm{j}_{s}(\bm{x},t)/n_{e} can be understood as the mean velocity of the electrons. These interpretations should be further refined by future studies.

Acknowledgements.—We would like to thank D. Oue and Y. Nozaki for the valuable and informative discussion. This work was partially supported by JST CREST Grant No. JPMJCR19J4, Japan. We acknowledge JSPS KAKENHI for Grants (Nos. JP21H01800, JP21H04565, JP23H01839, JP21H05186, JP19K03659, JP19H05821, JP18K03623 and JP21H05189). The work was supported in part by the Chuo University Personal Research Grant. The authors thank RIKEN iTHEMS NEW working group for providing the genesis of this collaboration.

References

  • Maekawa et al. [2017] S. Maekawa, S. O. Valenzuela, E. Saitoh, and T. Kimura, Spin current, Vol. 22 (Oxford University Press, 2017).
  • Uchida et al. [2011] K.-i. Uchida, T. An, Y. Kajiwara, M. Toda, and E. Saitoh, Surface-acoustic-wave-driven spin pumping in Y3Fe5O12/Pt hybrid structure, Appl. Phys. Lett. 99, 212501 (2011).
  • Weiler et al. [2012] M. Weiler, H. Huebl, F. S. Goerg, F. D. Czeschka, R. Gross, and S. T. B. Goennenwein, Spin Pumping with Coherent Elastic Waves, Phys. Rev. Lett. 108, 176601 (2012).
  • Deymier et al. [2014] P. A. Deymier, J. O. Vasseur, K. Runge, A. Manchon, and O. Bou-Matar, Phonon-magnon resonant processes with relevance to acoustic spin pumping, Phys. Rev. B 90, 224421 (2014).
  • Puebla et al. [2020] J. Puebla, M. Xu, B. Rana, K. Yamamoto, S. Maekawa, and Y. Otani, Acoustic ferromagnetic resonance and spin pumping induced by surface acoustic waves, J. Phys. D: Appl. Phys. 53, 264002 (2020).
  • Al Misba et al. [2020] W. Al Misba, M. M. Rajib, D. Bhattacharya, and J. Atulasimha, Acoustic-Wave-Induced Ferromagnetic-Resonance-Assisted Spin-Torque Switching of Perpendicular Magnetic Tunnel Junctions with Anisotropy Variation, Phys. Rev. Applied 14, 014088 (2020).
  • Camara et al. [2019] I. Camara, J.-Y. Duquesne, A. Lemaître, C. Gourdon, and L. Thevenard, Field-Free Magnetization Switching by an Acoustic Wave, Phys. Rev. Applied 11, 014045 (2019).
  • Tateno et al. [2020] S. Tateno, G. Okano, M. Matsuo, and Y. Nozaki, Phys. Rev. B 102, 104406 (2020).
  • Kawada et al. [2021] T. Kawada, M. Kawaguchi, T. Funato, H. Kohno, and M. Hayashi, Acoustic spin Hall effect in strong spin-orbit metals, Sci. Adv. 7, eabd9697 (2021).
  • Sasaki et al. [2017] R. Sasaki, Y. Nii, Y. Iguchi, and Y. Onose, Nonreciprocal propagation of surface acoustic wave in Ni / LiNbO 3, Phys. Rev. B 95, 020407 (2017).
  • Tateno and Nozaki [2020] S. Tateno and Y. Nozaki, Highly nonreciprocal spin waves excited by magnetoelastic coupling in a Ni/Si bilayer, Phys. Rev. Applied 13, 034074 (2020).
  • Xu et al. [2020] M. Xu, K. Yamamoto, J. Puebla, K. Baumgaertl, B. Rana, K. Miura, H. Takahashi, D. Grundler, S. Maekawa, and Y. Otani, Nonreciprocal surface acoustic wave propagation via magneto-rotation coupling, Science Advances 6, eabb1724 (2020).
  • Küß et al. [2021] M. Küß, M. Heigl, L. Flacke, A. Hörner, M. Weiler, A. Wixforth, and M. Albrecht, Nonreciprocal magnetoacoustic waves in dipolar-coupled ferromagnetic bilayers, Phys. Rev. Appl. 15, 034060 (2021).
  • Matsumoto et al. [2022] H. Matsumoto, T. Kawada, M. Ishibashi, M. Kawaguchi, and M. Hayashi, Large surface acoustic wave nonreciprocity in synthetic antiferromagnets, Appl. Phys. Express 15, 063003 (2022).
  • Lyons et al. [2023] T. P. Lyons, J. Puebla, K. Yamamoto, R. S. Deacon, Y. Hwang, K. Ishibashi, S. Maekawa, and Y. Otani, Acoustically Driven Magnon-Phonon Coupling in a Layered Antiferromagnet, Phys. Rev. Lett. 131, 196701 (2023).
  • Chowdhury et al. [2017] P. Chowdhury, A. Jander, and P. Dhagat, Nondegenerate Parametric Pumping of Spin Waves by Acoustic Waves, IEEE Magnetics Letters 8, 1 (2017).
  • Alekseev et al. [2020] S. G. Alekseev, S. E. Dizhur, N. I. Polzikova, V. A. Luzanov, A. O. Raevskiy, A. P. Orlov, V. A. Kotov, and S. A. Nikitov, Magnons parametric pumping in bulk acoustic waves resonator, Applied Physics Letters 117, 072408 (2020).
  • Liao et al. [2023] L. Liao, J. Puebla, K. Yamamoto, J. Kim, S. Maekawa, Y. Hwang, Y. Ba, and Y. Otani, Valley-Selective Phonon-Magnon Scattering in Magnetoelastic Superlattices, Phys. Rev. Lett. 131, 176701 (2023).
  • Barnett [1915] S. J. Barnett, Phys. Rev. 6, 239 (1915).
  • A. Einstein and de Haas [1915] A. Einstein and W. J. de Haas, Proc. KNAW 18, 696 (1915).
  • Scott [1962] G. Scott, Review of gyromagnetic ratio experiments, Reviews of Modern Physics 34, 102 (1962).
  • [22] T. M. Wallis, J. Moreland, and P. Kabos, Einstein–de Haas effect in a NiFe film deposited on a microcantilever, Applied Physics Letters 89, 122502.
  • [23] G. Zolfagharkhani, A. Gaidarzhy, P. Degiovanni, S. Kettemann, P. Fulde, and P. Mohanty, Nanomechanical detection of itinerant electron spin flip, Nature Nanotechnology 3, 720.
  • [24] K. Harii, Y.-J. Seo, Y. Tsutsumi, H. Chudo, K. Oyanagi, M. Matsuo, Y. Shiomi, T. Ono, S. Maekawa, and E. Saitoh, Spin Seebeck mechanical force, Nature Communications 10, 2616.
  • [25] K. Mori, M. G. Dunsmore, J. E. Losby, D. M. Jenson, M. Belov, and M. R. Freeman, Einstein–de Haas effect at radio frequencies in and near magnetic equilibrium, Physical Review B 102, 054415.
  • Chudo et al. [2014] H. Chudo, M. Ono, K. Harii, M. Matsuo, J. Ieda, R. Haruki, S. Okayasu, S. Maekawa, H. Yasuoka, and E. Saitoh, Appl. Phys. Express 7, 063004 (2014).
  • Chudo et al. [2015] H. Chudo, K. Harii, M. Matsuo, J. Ieda, M. Ono, S. Maekawa, and E. Saitoh, J. Phys. Soc. Jpn. 84, 043601 (2015).
  • [28] M. Arabgol and T. Sleator, Observation of the Nuclear Barnett Effect, Physical Review Letters 122, 177202.
  • Chudo et al. [2021] H. Chudo, M. Matsuo, S. Maekawa, and E. Saitoh, Phys. Rev. B 103, 174308 (2021).
  • Wood et al. [2017] A. Wood, E. Lilette, Y. Fein, V. Perunicic, L. Hollenberg, R. Scholten, and A. Martin, Magnetic pseudo-fields in a rotating electron–nuclear spin system, Nature Physics 13, 1070 (2017).
  • Hirohata et al. [2018] A. Hirohata, Y. Baba, B. A. Murphy, B. Ng, Y. Yao, K. Nagao, and J.-y. Kim, Magneto-optical detection of spin accumulation under the influence of mechanical rotation, Scientific reports 8, 1974 (2018).
  • [32] C. Dornes, Y. Acremann, M. Savoini, M. Kubli, M. J. Neugebauer, E. Abreu, L. Huber, G. Lantz, C. A. F. Vaz, H. Lemke, E. M. Bothschafter, M. Porer, V. Esposito, L. Rettig, M. Buzzi, A. Alberca, Y. W. Windsor, P. Beaud, U. Staub, D. Zhu, S. Song, J. M. Glownia, and S. L. Johnson, The ultrafast Einstein–de Haas effect, Nature 565, 209.
  • [33] M. Ganzhorn, S. Klyatskaya, M. Ruben, and W. Wernsdorfer, Quantum Einstein-de Haas effect, Nature Communications 7, 11443.
  • Adamczyk et al. [2017] L. Adamczyk, J. Adkins, G. Agakishiev, M. Aggarwal, Z. Ahammed, N. Ajitanand, I. Alekseev, D. Anderson, R. Aoyama, A. Aparin, et al., Global λ\lambda hyperon polarization in nuclear collisions, Nature 548 (2017).
  • Takahashi et al. [2016] R. Takahashi, M. Matsuo, M. Ono, K. Harii, H. Chudo, S. Okayasu, J. Ieda, S. Takahashi, S. Maekawa, and E. Saitoh, Nature Phys 12, 52 (2016).
  • Kobayashi et al. [2017] D. Kobayashi, T. Yoshikawa, M. Matsuo, R. Iguchi, S. Maekawa, E. Saitoh, and Y. Nozaki, Phys. Rev. Lett. 119, 077202 (2017).
  • Takahashi et al. [2020] R. Takahashi, H. Chudo, M. Matsuo, K. Harii, Y. Ohnuma, S. Maekawa, and E. Saitoh, Nat Commun 11, 3009 (2020).
  • Tabaei Kazerooni et al. [2020] H. Tabaei Kazerooni, A. Thieme, J. Schumacher, and C. Cierpka, Phys. Rev. Applied 14, 014002 (2020).
  • Tabaei Kazerooni et al. [2021] H. Tabaei Kazerooni, G. Zinchenko, J. Schumacher, and C. Cierpka, Phys. Rev. Fluids 6, 043703 (2021).
  • Tateno et al. [2021] S. Tateno, Y. Kurimune, M. Matsuo, K. Yamanoi, and Y. Nozaki, Phys. Rev. B 104, L020404 (2021).
  • Hehl and Ni [1990] F. W. Hehl and W.-T. Ni, Inertial effects of a dirac particle, Physical Review D 42, 2045 (1990).
  • Funato and Kohno [2018] T. Funato and H. Kohno, J. Phys. Soc. Jpn. 87, 073706 (2018).
  • Taylor [1923] G. I. Taylor, Philos. Trans. R. Soc. London, Ser. A 223, 289 (1923).
  • Matsuo et al. [2017] M. Matsuo, Y. Ohnuma, and S. Maekawa, Phys. Rev. B 96, 020401 (2017).
  • Matsuo et al. [2013] M. Matsuo, J. Ieda, K. Harii, E. Saitoh, and S. Maekawa, Phys. Rev. B 87, 180402 (2013).
  • Fujimoto and Matsuo [2020] J. Fujimoto and M. Matsuo, Magnon current generation by dynamical distortion, Phys. Rev. B 102, 020406 (2020).

I Transformation to differentially rotating frame

By using the total angular momentum operator acting on coordinates and spin space,

Jzi(𝒙×)zI+12σz,,J^{z}\equiv-i\hbar(\bm{x}\times\bm{\nabla})_{z}\otimes I+1\otimes\frac{\hbar}{2}\sigma^{z},, (18)

we define the following operator:

Q^Φ(t)=d3xΦ(𝒙,t)ψ^(𝒙)Jzψ^(𝒙),\hat{Q}_{\Phi}(t)=\int d^{3}x\Phi(\bm{x},t)\hat{\psi}^{\dagger}(\bm{x})J^{z}\hat{\psi}(\bm{x}), (19)

which is a generator of the differential rotation with angle Φ(𝒙,t)\Phi(\bm{x},t) around the zz-axis. Note that we assume the rotation angle is axisymmetric, i.e., Φ/φ=0\partial\Phi/\partial\varphi=0. The canonical commutation relation {ψ^(𝒙),ψ^(𝒚)}=δ(𝒙𝒚)\{\hat{\psi}(\bm{x}),\hat{\psi}^{\dagger}(\bm{y})\}=\delta(\bm{x}-\bm{y}) yields

[Q^Φ(t),ψ^(𝒙)]=Φ(𝒙,t)Jzψ^(𝒙),[Q^Φ(t),ψ^(𝒙)]=Φ(𝒙,t)(Jzψ^(𝒙)).[\hat{Q}_{\Phi}(t),\hat{\psi}(\bm{x})]=-\Phi(\bm{x},t)J^{z}\hat{\psi}(\bm{x}),\quad[\hat{Q}_{\Phi}(t),\hat{\psi}^{\dagger}(\bm{x})]=\Phi(\bm{x},t)(J^{z}\hat{\psi}(\bm{x}))^{\dagger}. (20)

We have

exp[iQ^Φ(t)]ψ^(𝒙)exp[iQ^Φ(t)]\displaystyle\exp\left[\frac{i}{\hbar}\hat{Q}_{\Phi}(t)\right]\hat{\psi}(\bm{x})\exp\left[-\frac{i}{\hbar}\hat{Q}_{\Phi}(t)\right] =eiΦ(𝒙,t)Jz/ψ^(𝒙)\displaystyle=e^{-i\Phi(\bm{x},t)J^{z}/\hbar}\hat{\psi}(\bm{x}) (21)
=eiΦ(𝒙,t)σz/2ψ^(𝒙)(𝒙eΦ(𝒙,t)(𝒙×)z𝒙).\displaystyle=e^{-i\Phi(\bm{x},t)\sigma^{z}/2}\hat{\psi}(\bm{x}^{\prime})\qquad(\bm{x}^{\prime}\equiv e^{-\Phi(\bm{x},t)(\bm{x}\times\bm{\nabla})_{z}}\bm{x}).

Note that it can be explicitly written in matrix form as

eΦ(𝒙,t)(𝒙×)z𝒙=z1[Φ(𝒙,t)]𝒙,z(φ)(cosφsinφ0sinφcosφ0001).e^{-\Phi(\bm{x},t)(\bm{x}\times\bm{\nabla})_{z}}\bm{x}=\mathcal{R}^{-1}_{z}[\Phi(\bm{x},t)]\bm{x},\quad\mathcal{R}_{z}(\varphi)\equiv\matrixquantity(\cos\varphi&-\sin\varphi&0\cr\sin\varphi&\cos\varphi&0\cr 0&0&1). (22)

The free part of the Hamiltonian (namely, the kinetic term) transforms as

eiQ^Φ/H^0eiQ^Φ/\displaystyle e^{i\hat{Q}_{\Phi}/\hbar}\hat{H}_{0}e^{-i\hat{Q}_{\Phi}/\hbar} =H^0+i01𝑑λeiλQ^Φ/[Q^Φ,H^0]eiλQ^Φ/\displaystyle=\hat{H}_{0}+\frac{i}{\hbar}\int^{1}_{0}d\lambda e^{i\lambda\hat{Q}_{\Phi}/\hbar}[\hat{Q}_{\Phi},\hat{H}_{0}]e^{-i\lambda\hat{Q}_{\Phi}/\hbar} (23)
=H^0+i[Q^Φ,H^0]122[Q^Φ,[Q^Φ,H^0]]i3!3[Q^Φ,[Q^Φ,[Q^Φ,H^0]]]+\displaystyle=\hat{H}_{0}+\frac{i}{\hbar}[\hat{Q}_{\Phi},\hat{H}_{0}]-\frac{1}{2\hbar^{2}}[\hat{Q}_{\Phi},[\hat{Q}_{\Phi},\hat{H}_{0}]]-\frac{i}{3!\hbar^{3}}[\hat{Q}_{\Phi},[\hat{Q}_{\Phi},[\hat{Q}_{\Phi},\hat{H}_{0}]]]+\cdots
=12md3𝒙[iψ^(𝒙)Φ(𝒙,t)Jzψ^(𝒙)][iψ^(𝒙)Φ(𝒙,t)Jzψ^(𝒙)],\displaystyle=\frac{1}{2m}\int d^{3}\bm{x}[-i\hbar\bm{\nabla}\hat{\psi}(\bm{x})-\bm{\nabla}\Phi(\bm{x},t)J_{z}\hat{\psi}(\bm{x})]^{\dagger}\cdot[-i\hbar\bm{\nabla}\hat{\psi}(\bm{x})-\bm{\nabla}\Phi(\bm{x},t)J_{z}\hat{\psi}(\bm{x})],

where we have used the following equations:

[Q^Φ,H^0]=i22md3𝒙Φ(𝒙,t)[(iψ^(𝒙))Jzψ^(𝒙)+(Jzψ^(𝒙))(iψ^(𝒙))],[\hat{Q}_{\Phi},\hat{H}_{0}]=\frac{i\hbar^{2}}{2m}\int d^{3}\bm{x}\bm{\nabla}\Phi(\bm{x},t)\cdot\left[(-i\bm{\nabla}\hat{\psi}(\bm{x}))^{\dagger}J_{z}\hat{\psi}(\bm{x})+(J_{z}\hat{\psi}(\bm{x}))^{\dagger}(-i\bm{\nabla}\hat{\psi}(\bm{x}))\right], (24)
[Q^Φ,[Q^Φ,H^0]]=2md3𝒙|Φ(𝒙,t)|2(Jzψ^(𝒙))Jzψ^(𝒙),[\hat{Q}_{\Phi},[\hat{Q}_{\Phi},\hat{H}_{0}]]=-\frac{\hbar^{2}}{m}\int d^{3}\bm{x}|\bm{\nabla}\Phi(\bm{x},t)|^{2}(J_{z}\hat{\psi}(\bm{x}))^{\dagger}J_{z}\hat{\psi}(\bm{x}), (25)
[Q^Φ,[Q^Φ,[Q^Φ,H^0]]]=0.[\hat{Q}_{\Phi},[\hat{Q}_{\Phi},[\hat{Q}_{\Phi},\hat{H}_{0}]]]=0. (26)

The impurity potential part is

eiQ^Φ/[d3xψ^(𝒙)Vimp(𝒙,t)ψ^(𝒙)]eiQ^Φ/\displaystyle e^{i\hat{Q}_{\Phi}/\hbar}\left[\int d^{3}x\hat{\psi}^{\dagger}(\bm{x})V^{\prime}_{\text{imp}}(\bm{x},t)\hat{\psi}(\bm{x})\right]e^{-i\hat{Q}_{\Phi}/\hbar} =d3xψ^(𝒙)Vimp(𝒙,t)ψ^(𝒙)\displaystyle=\int d^{3}x\hat{\psi}^{\dagger}(\bm{x}^{\prime})V^{\prime}_{\text{imp}}(\bm{x},t)\hat{\psi}(\bm{x}^{\prime}) (27)
=jd3xψ^(𝒙)u(𝒙𝒓j(t))ψ^(𝒙)(𝒓j(t)=z[Φ(𝒓j,t)]𝒓j)\displaystyle=\sum_{j}\int d^{3}x\hat{\psi}^{\dagger}(\bm{x}^{\prime})u(\bm{x}-\bm{r}^{\prime}_{j}(t))\hat{\psi}(\bm{x}^{\prime})\quad(\bm{r}^{\prime}_{j}(t)=\mathcal{R}_{z}[\Phi(\bm{r}_{j},t)]\bm{r}_{j})
=jd3xψ^(𝒙)u(z[Φ(𝒙,t)]𝒙z[Φ(𝒓j,t)]𝒓j)ψ^(𝒙)\displaystyle=\sum_{j}\int d^{3}x\hat{\psi}^{\dagger}(\bm{x})u(\mathcal{R}_{z}[\Phi(\bm{x},t)]\bm{x}-\mathcal{R}_{z}[\Phi(\bm{r}_{j},t)]\bm{r}_{j})\hat{\psi}(\bm{x})
jd3xψ^(𝒙)u(𝒙𝒓j)ψ^(𝒙)=d3xψ^(𝒙)Vimp(𝒙)ψ^(𝒙),\displaystyle\simeq\sum_{j}\int d^{3}x\hat{\psi}^{\dagger}(\bm{x})u(\bm{x}-\bm{r}_{j})\hat{\psi}(\bm{x})=\int d^{3}x\hat{\psi}^{\dagger}(\bm{x})V_{\text{imp}}(\bm{x})\hat{\psi}(\bm{x}),

where we have changed a variable of integration as 𝒙z[Φ(𝒙,t)]𝒙\bm{x}\to\mathcal{R}_{z}[\Phi(\bm{x},t)]\bm{x} in the third line. Because we assume that a single impurity potential u(𝒙)u(\bm{x}) is isotropic and has compact support in |𝒙|<a|\bm{x}|<a, we have

u(z[Φ(𝒙,t)]𝒙z[Φ(𝒓j,t)]𝒓j)=u(𝒙z[Φ(𝒓j,t)Φ(𝒙,t)]𝒓j)u(𝒙𝒓j)u(\mathcal{R}_{z}[\Phi(\bm{x},t)]\bm{x}-\mathcal{R}_{z}[\Phi(\bm{r}_{j},t)]\bm{r}_{j})=u(\bm{x}-\mathcal{R}_{z}[\Phi(\bm{r}_{j},t)-\Phi(\bm{x},t)]\bm{r}_{j})\simeq u(\bm{x}-\bm{r}_{j}) (28)

for |𝒙𝒓j|<a1/|Φ||\bm{x}-\bm{r}_{j}|<a\ll 1/|\bm{\nabla}\Phi|. In a similar manner, the spin-orbit interaction part is

eiQ^Φ/[λsod3xψ^(𝒙)𝝈[Vimp(𝒙,t)×(i)]ψ^(𝒙)]eiQ^Φ/\displaystyle e^{i\hat{Q}_{\Phi}/\hbar}\left[\lambda_{\text{so}}\int d^{3}x\hat{\psi}^{\dagger}(\bm{x})\bm{\sigma}\cdot[\bm{\nabla}V^{\prime}_{\text{imp}}(\bm{x},t)\times(-i\hbar\bm{\nabla})]\hat{\psi}(\bm{x})\right]e^{-i\hat{Q}_{\Phi}/\hbar} (29)
=λsod3xψ^(𝒙)𝝈[Vimp(𝒙)×(iΦ(𝒙,t)Jz)]ψ^(𝒙).\displaystyle=\lambda_{\text{so}}\int d^{3}x\hat{\psi}^{\dagger}(\bm{x})\bm{\sigma}\cdot[\bm{\nabla}V_{\text{imp}}(\bm{x})\times(-i\hbar\bm{\nabla}-\bm{\nabla}\Phi(\bm{x},t)J^{z})]\hat{\psi}(\bm{x}).

II Definition of spin current operator

In this section, we define the spin-current density operator through the continuum equation with respect to the spin density. The zz-polarized spin density operator is defined by

s^(𝒙)ψ^(𝒙)σzψ^(𝒙).\displaystyle\hat{s}(\bm{x})\equiv\hat{\psi}^{\dagger}(\bm{x})\sigma^{z}\hat{\psi}(\bm{x}). (30)

The time derivative of the zz-polarized spin density operator in the Heisenberg representation is given by

ts^(𝒙,t)=𝒋^s(𝒙,t)+𝒯^(𝒙,t).\displaystyle\partialderivative{t}\hat{s}(\bm{x},t)=-\bm{\nabla}\cdot\hat{\bm{j}}_{s}(\bm{x},t)+\hat{\mathcal{T}}(\bm{x},t). (31)

The first term is the divergence of the spin-current density:

𝒋^s(𝒙)=ψ^(𝒙)[σz2im+λso𝒆z×V2m𝑨s]ψ^(𝒙).\displaystyle\hat{\bm{j}}_{s}(\bm{x})=\hat{\psi}^{\dagger}(\bm{x})\quantity[\frac{\hbar\sigma^{z}}{2im}\overleftrightarrow{\bm{\nabla}}+\lambda_{\text{so}}\bm{e}_{z}\times\bm{\nabla}V-\frac{\hbar}{2m}\bm{A}_{s}]\hat{\psi}(\bm{x}). (32)

The second term represents the spin torque due to the impurity spin-orbit interaction:

𝒯^(𝒙)=iλsoψ^(𝒙)[σz𝒆z×(V×i)]ψ^(𝒙).\displaystyle\hat{\mathcal{T}}(\bm{x})=-\frac{i}{\hbar}\lambda_{\text{so}}\hat{\psi}^{\dagger}(\bm{x})\quantity[\sigma^{z}\bm{e}_{z}\times\quantity(\bm{\nabla}V\times\frac{\hbar}{i}\bm{\nabla})]\hat{\psi}(\bm{x}). (33)

III calculation of spin density and spin current density

Refer to caption
Figure 2: The diagrams representing the (a) response functions and the (b) three-point vertices. The black circles represent the vertices and the shaded region represents the ladder vertex corrections. The dotted lines represent the impurity potential and the crosses represent the impurities.

In this section, we demonstrate the calculation of the spin density and spin-current density driven by the differential rotation in linear response to the emergent gauge fields. The response functions of spin density and spin-current density corresponding to the diagrams shown in Fig. 2(a) are expressed as

Kμν(ω)=δμ0δν0σc2e2D+iω4π𝒌v𝒌,μtr[σzg+rσz(v𝒌,ν+Λνs)ga],\displaystyle K_{\mu\nu}(\omega)=\delta_{\mu 0}\delta_{\nu 0}\frac{\hbar\sigma_{c}}{2e^{2}D}+i\omega\frac{\hbar}{4\pi}\sum_{\bm{k}}v_{\bm{k},\mu}\text{tr}[\sigma^{z}g^{r}_{+}\sigma^{z}(v_{\bm{k},\nu}+\Lambda_{\nu}^{s})g^{a}_{-}], (34)

where the three-point vertices Λνs\Lambda^{s}_{\nu} are shown in Fig. 2(b). Up to the second order in the wavenumber 𝒒\bm{q} and frequency ω\omega, the response functions are calculated as

K00(ω)\displaystyle K_{00}(\omega) =σc2e2D+iω2π[1+Λ0s]I0=σc2e2DDq2+τs1Dq2iω+τs1,\displaystyle=\frac{\hbar\sigma_{c}}{2e^{2}D}+i\omega\frac{\hbar}{2\pi}\Bigl{[}1+\Lambda_{0}^{s}\Bigr{]}I_{0}=\frac{\hbar\sigma_{c}}{2e^{2}D}\frac{Dq^{2}+\tau_{s}^{-1}}{Dq^{2}-i\omega+\tau_{s}^{-1}}, (35)
K0j(ω)\displaystyle K_{0j}(\omega) =iω2π[1+Λ0s]Ij=iωσc2e2iqjDq2iω+τs1,\displaystyle=i\omega\frac{\hbar}{2\pi}\Bigl{[}1+\Lambda_{0}^{s}\Bigr{]}I_{j}=-i\omega\frac{\hbar\sigma_{c}}{2e^{2}}\frac{iq_{j}}{Dq^{2}-i\omega+\tau_{s}^{-1}}, (36)
Ki0(ω)\displaystyle K_{i0}(\omega) =iω2πIi[1+Λ0s]=iωσc2e2iqiDq2iω+τs1,\displaystyle=i\omega\frac{\hbar}{2\pi}I_{i}\Bigl{[}1+\Lambda_{0}^{s}\Bigr{]}=-i\omega\frac{\hbar\sigma_{c}}{2e^{2}}\frac{iq_{i}}{Dq^{2}-i\omega+\tau_{s}^{-1}}, (37)
Kij(ω)\displaystyle K_{ij}(\omega) =iω2π[Iij+IiΛjs]=iωσc2e2(δijDqiqjDq2iω+τs1),\displaystyle=i\omega\frac{\hbar}{2\pi}\Bigl{[}I_{ij}+I_{i}\Lambda_{j}^{s}\Bigr{]}=i\omega\frac{\hbar\sigma_{c}}{2e^{2}}\left(\delta_{ij}-\frac{Dq_{i}q_{j}}{Dq^{2}-i\omega+\tau_{s}^{-1}}\right), (38)

where the Latin indices represent the spacial directions, i.e., i=x,y,zi=x,y,z. Here, the integrations IμνI_{\mu\nu} are defined by

Iμν=𝒌v𝒌,μv𝒌,νg+rga,\displaystyle I_{\mu\nu}=\sum_{\bm{k}}v_{\bm{k},\mu}v_{\bm{k},\nu}g^{r}_{+}g^{a}_{-}, (39)

and given by

I0=I00=πν0γ[1τ(Dq2iω)],\displaystyle I_{0}=I_{00}=\frac{\pi\nu_{0}}{\hbar\gamma}[1-\tau(Dq^{2}-i\omega)], (40)
Ii=Ii0=I0i=Diqiπν0γ,\displaystyle I_{i}=I_{i0}=I_{0i}=-Diq_{i}\frac{\pi\nu_{0}}{\hbar\gamma}, (41)
Iij=vF23πν0γ[δij+δijτ(iω35Dq2)65τDqiqj].\displaystyle I_{ij}=\frac{v_{F}^{2}}{3}\frac{\pi\nu_{0}}{\hbar\gamma}\left[\delta_{ij}+\delta_{ij}\tau\left(i\omega-\frac{3}{5}Dq^{2}\right)-\frac{6}{5}\tau Dq_{i}q_{j}\right]. (42)

Note that the response functions satisfy the following identities:

iωK00+iqjK0j=iωσc2e2Dτs1Dq2iω+τs1,iωKi0+iqjKij=iωσc2e2τs1Dq2iω+τs1iqi.-i\omega K_{00}+iq_{j}K_{0j}=-i\omega\frac{\hbar\sigma_{c}}{2e^{2}D}\frac{\tau_{s}^{-1}}{Dq^{2}-i\omega+\tau_{s}^{-1}},\quad-i\omega K_{i0}+iq_{j}K_{ij}=i\omega\frac{\hbar\sigma_{c}}{2e^{2}}\frac{\tau_{s}^{-1}}{Dq^{2}-i\omega+\tau_{s}^{-1}}iq_{i}. (43)

The spin density is given by

s^(𝒒,ω)\displaystyle\langle\hat{s}(\bm{q},\omega)\rangle =K00As,0+K0jAs,j\displaystyle=K_{00}A_{s,0}+K_{0j}A_{s,j}
=σc2e2Dτs1Dq2iω+τs1As,0+σc2e2iqjDq2iω+τs1(iωAs,jiqjAs,0)\displaystyle=\frac{\hbar\sigma_{c}}{2e^{2}D}\frac{\tau_{s}^{-1}}{Dq^{2}-i\omega+\tau_{s}^{-1}}A_{s,0}+\frac{\hbar\sigma_{c}}{2e^{2}}\frac{iq_{j}}{Dq^{2}-i\omega+\tau_{s}^{-1}}(-i\omega A_{s,j}-iq_{j}A_{s,0})
=σc2e2Dτs1Dq2iω+τs1(iωΦ).\displaystyle=\frac{\hbar\sigma_{c}}{2e^{2}D}\frac{\tau_{s}^{-1}}{Dq^{2}-i\omega+\tau_{s}^{-1}}(-i\omega\Phi). (44)

The spin-current density is given by

j^s,i(𝒒,ω)\displaystyle\langle\hat{j}_{s,i}(\bm{q},\omega)\rangle =Ki0As,0+KijAs,j\displaystyle=K_{i0}A_{s,0}+K_{ij}A_{s,j}
=iωσc2e2τs1Dq2iω+τs1As,i\displaystyle=i\omega\frac{\hbar\sigma_{c}}{2e^{2}}\frac{\tau_{s}^{-1}}{Dq^{2}-i\omega+\tau_{s}^{-1}}A_{s,i}
+iωσc2e21Dq2iω+τs1(iωAs,iiqiAs,0)+iωσc2e2iDqjDq2iω+τs1(iqiAs,jiqjAs,i)\displaystyle\quad+i\omega\frac{\hbar\sigma_{c}}{2e^{2}}\frac{1}{Dq^{2}-i\omega+\tau_{s}^{-1}}(-i\omega A_{s,i}-iq_{i}A_{s,0})+i\omega\frac{\hbar\sigma_{c}}{2e^{2}}\frac{iDq_{j}}{Dq^{2}-i\omega+\tau_{s}^{-1}}(iq_{i}A_{s,j}-iq_{j}A_{s,i})
=iωσc2e2τs1Dq2iω+τs1iqiΦ.\displaystyle=i\omega\frac{\hbar\sigma_{c}}{2e^{2}}\frac{\tau_{s}^{-1}}{Dq^{2}-i\omega+\tau_{s}^{-1}}iq_{i}\Phi. (45)

We note that (As,0,As,i)=(iωΦ,iqiΦ)(A_{s,0},A_{s,i})=(-i\omega\Phi,iq_{i}\Phi).

III.1 Ladder vertex corrections

In this section, we calculate ladder vertex corrections due to the impurity scattering and spin-orbit scattering. First, we define the elementary vertex fabf_{ab} shown in Fig. 3(a) corresponding to the coupling to the single impurity:

fab=δab+iλso(𝒌×𝒌)𝝈ab,\displaystyle f_{ab}=\delta_{ab}+i\hbar\lambda_{\text{so}}(\bm{k}\times\bm{k}^{\prime})\cdot\bm{\sigma}_{ab}, (46)

where the Latin indices a,ba,b describe the spin space. The proper four-point vertex Γ0\Gamma^{0} shown in the first term on the right-hand side of Fig. 3(b) is calculated as

Γab,cd0=niui2fadfcbFS=πν0(γ0δadδcb+13γso𝝈ad𝝈cb),\displaystyle\Gamma^{0}_{ab,cd}=n_{i}u_{i}^{2}\langle f_{ad}f_{cb}\rangle_{\text{FS}}=\frac{\hbar}{\pi\nu_{0}}\left(\gamma_{0}\delta_{ad}\delta_{cb}+\frac{1}{3}\gamma_{\text{so}}\bm{\sigma}_{ad}\cdot\bm{\sigma}_{cb}\right), (47)

where FS\langle\cdots\rangle_{\text{FS}} means averaging over at the Fermi surface, γ0=πniui2ν0\hbar\gamma_{0}=\pi n_{i}u_{i}^{2}\nu_{0} is the damping due to the impurity scattering, and γso=2λso2kF4γ0\gamma_{\text{so}}=\hbar^{2}\lambda_{\text{so}}^{2}k_{F}^{4}\gamma_{0} is the damping due to the spin-orbit scattering. The four-point vertex shown in Fig. 3(b) is determined by the following Dyson equation:

Γab,cd(𝒒)\displaystyle\Gamma_{ab,cd}(\bm{q}) =Γab,cd0+Γab,ef0I0(𝒒)Γfe,cd(𝒒)\displaystyle=\Gamma^{0}_{ab,cd}+\Gamma^{0}_{ab,ef}I_{0}(\bm{q})\Gamma_{fe,cd}(\bm{q})
=Γc(𝒒)δabδcd+Γs(𝒒)𝝈ab𝝈cd,\displaystyle=\Gamma_{c}(\bm{q})\delta_{ab}\delta_{cd}+\Gamma_{s}(\bm{q})\bm{\sigma}_{ab}\cdot\bm{\sigma}_{cd}, (48)

where a,,fa,\ldots,f are the spin indices, and

Γc(𝒒)\displaystyle\Gamma_{c}(\bm{q}) =4πν0τ21Dq2iω,\displaystyle=\frac{\hbar}{4\pi\nu_{0}\tau^{2}}\frac{1}{Dq^{2}-i\omega}, (49)
Γs(𝒒)\displaystyle\Gamma_{s}(\bm{q}) =4πν0τ21ττsDq2iω+τs1.\displaystyle=\frac{\hbar}{4\pi\nu_{0}\tau^{2}}\frac{1-\frac{\tau}{\tau_{s}}}{Dq^{2}-i\omega+\tau_{s}^{-1}}. (50)

Therefore, the three-point vertices Λνs\Lambda^{s}_{\nu} are calculated by

σabαΛνs(𝒒)\displaystyle\sigma^{\alpha}_{ab}\Lambda^{s}_{\nu}(\bm{q}) =σdcαΓab,cd(𝒒)Iν,\displaystyle=\sigma^{\alpha}_{dc}\Gamma_{ab,cd}(\bm{q})I_{\nu}, (51)

and given by

Λ0s(𝒒)\displaystyle\Lambda^{s}_{0}(\bm{q}) =1τ(Dq2iω+τs1)1,\displaystyle=\dfrac{1}{\tau(Dq^{2}-i\omega+\tau^{-1}_{s})}-1, (52)
Λjs(𝒒)\displaystyle\Lambda^{s}_{j}(\bm{q}) =Diqjτ(Dq2iω+τs1).\displaystyle=-\dfrac{Diq_{j}}{\tau(Dq^{2}-i\omega+\tau^{-1}_{s})}. (53)
Refer to caption
Figure 3: (a) The elementary vertex due to the single impurity. (b) The four-point vertex in the ladder approximation.