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

Many-body theory of phonon-induced spin relaxation and decoherence

Jinsoo Park Department of Applied Physics and Materials Science, California Institute of Technology, Pasadena, CA 91125, USA.    Yao Luo Department of Applied Physics and Materials Science, California Institute of Technology, Pasadena, CA 91125, USA.    Jin-Jian Zhou School of Physics, Beijing Institute of Technology, Beijing 100081, China.    Marco Bernardi [email protected] Department of Applied Physics and Materials Science, California Institute of Technology, Pasadena, CA 91125, USA.
Abstract

First-principles calculations enable accurate predictions of electronic interactions and dynamics. However, computing the electron spin dynamics remains challenging. The spin-orbit interaction causes various dynamical phenomena that couple with phonons, such as spin precession and spin-flip ee-ph scattering, which are difficult to describe with current first-principles calculations. In this work, we show a rigorous framework to study phonon-induced spin relaxation and decoherence, by computing the spin-spin correlation function and its vertex corrections due to ee-ph interactions. We apply this approach to a model system and develop corresponding first-principles calculations of spin relaxation in GaAs. Our vertex-correction formalism is shown to capture the Elliott-Yafet, Dyakonov-Perel, and strong-precession mechanisms - three independent spin decoherence regimes with distinct physical origins thereby unifying their theoretical treatment and calculation. Our method is general and enables quantitative studies of spin relaxation, decoherence, and transport in a wide range of materials and devices.

I Introduction

Linear response theory provides a microscopic understanding of the response of a system to external perturbations and computes the associated correlation functions  [1, 2, 3, 4, 5, 6]. First-principles calculations of electronic interactions [7, 8, 9, 10, 11, 12, 13, 14] complement this formalism, enabling precise predictions of materials properties and transport coefficients without resorting to empirical models or fitting parameters. In this context, electron-phonon (ee-ph) interactions are particularly important as they govern a wide range of phenomena such as charge transport [15], superconductivity [16], spin transport [17, 18, 19] and spin decoherence [20, 21, 22].
The Boltzmann transport equation (BTE) is widely used to study the response to an external electric field [23, 24]. The field drives the electronic populations fn𝒌f_{n\bm{k}}, for states with band nn and crystal momentum 𝒌\bm{k}, away from the equilibrium Fermi-Dirac distribution fn𝒌0f^{0}_{n\bm{k}}, while the ee-ph interactions dissipate electron energy and act to restore equilibrium, resulting in a steady-state current e(fn𝒌fn𝒌0)vn𝒌e(f_{n\bm{k}}-f^{0}_{n\bm{k}})v_{n\bm{k}}, where ee is the electron charge and vn𝒌v_{n\bm{k}} is the band velocity [23]. In the many-body formalism, the BTE at low electric field is formally equivalent to the ladder vertex-correction to the dc conductivity [4]. In that framework, one determines the current-current correlation function, with vertex corrections from the ee-ph interactions obtained by summing over ladder diagrams, and computes the conductivity from the dissipative part of the susceptibility. A key factor making this approach equivalent to the BTE is that the electron velocity is band-diagonal in the Bloch basis, m𝒌|v^|n𝒌=δnm𝒌En𝒌/\bra{m\bm{k}}\hat{v}\ket{n\bm{k}}=\delta_{nm}\partial_{\bm{k}}E_{n\bm{k}}/\hbar [4].
However, studying the response to an external field of an arbitrary operator that couples with phonons is more difficult. The matrix representation of an operator A^\hat{A} is in general nondiagonal in the Bloch basis, A^nm𝒌=m𝒌|A^|n𝒌\hat{A}_{nm\bm{k}}\!=\!\bra{m\bm{k}}\hat{A}\ket{n\bm{k}}, and can mix states in different bands. The BTE cannot be applied in this case because due to its population-based formalism it neglects such off-diagonal (inter-band) components. A framework treating the response of non-diagonal operators coupled with ee-ph interactions is still missing.
An important example is spin relaxation and decoherence, where spin-orbit coupling (SOC) makes the spin operators non-diagonal in the band index, and phonons can change the electron spin through ee-ph interactions [19]. Theories of spin decoherence focus on two distinct models - the Elliott-Yafet (EY) mechanism [25, 26], where ee-ph collisions rotate the spin direction, and the Dyakonov-Perel (DP) mechanism [27], where spin precession in the SOC field induces a motional narrowing of the spin. The dominant mechanism depends on the system - typically, EY dominates in centrosymmetric and DP in non-centrosymmetric materials. Spin relaxation exhibits opposite trends in these two mechanisms, with spin relaxation times proportional to the ee-ph relaxation times in EY, and inversely proportional in DP. We have recently shown that EY spin relaxation can be computed from first-principles in the spin relaxation time approximation (sRTA) [28]-the spin counterpart of the transport RTA for charge transport [1, 4] - but spin precession and the DP mechanism are neglected in the sRTA.
Here we show a many-body approach to compute the susceptibility for an arbitrary non-diagonal operator coupled to ee-ph interactions. Our diagrammatic approach, based on the Kubo formula with vertex corrections to the susceptibility in an external injection field, calculates an effective phonon-dressed operator and its renormalized dynamics. We derive a Bethe-Salpeter equation (BSE) for the vertex corrections, and specializing to the spin operator, we use the vertex corrections to compute spin relaxation and precession. We show that the vertex corrections can capture spin decoherence due to both the EY and DP mechanisms and can also model the strong-precession regime, a third mechanism distinct from EY and DP. We find these three mechanisms in the exact solution of a two-level system, and also identify them in a real material, GaAs, using first-principles calculations. Combined with first-principles ee-ph calculations, our method is poised to advance microscopic understanding of phonon-induced spin decoherence [29], with applications ranging from solid-state qubits to quantum materials with spin Hall effect, valley-dependent spin physics, and Rashba effect.
The paper is organized as follows: In Sec. II, we derive the BSE for the phonon-dressed vertex, discuss its physical interpretation, and calculate the susceptibility in response to an injection field. In Sec. III-IV, we apply this formalism to study spin dynamics in a model two-level system and in a real material, GaAs, discussing spin relaxation due to the EY, DP, and strong-precession mechanisms.

II Theory

We derive a self-consistent BSE for the vertex correction to the susceptibility due to ee-ph interactions, focusing on a general vector observable 𝐀^\mathbf{\hat{A}}. We then present a physical interpretation of the vertex corrections and the renormalized dynamics of the operator. We employ atomic units and set =1\hbar=1.

II.1 Interacting Green’s function

We consider an unperturbed Hamiltonian H0H_{0} diagonal in the Bloch basis, n𝒌|H0|n𝒌=εn𝒌δnn\bra{n^{\prime}\bm{k}}H_{0}\ket{n\bm{k}}=\varepsilon_{n\bm{k}}\delta_{nn^{\prime}}. The interacting imaginary-time Green’s function 𝒢(iωa)\mathcal{G}(i\omega_{a}) is written using the Dyson equation as [1]

𝒢(iωa)1=𝒢(0)(iωa)1Σ(iωa),\mathcal{G}(i\omega_{a})^{-1}=\mathcal{G}^{(0)}(i\omega_{a})^{-1}-\Sigma(i\omega_{a}), (1)

where ωa\omega_{a} are fermionic Matsubara frequencies, 𝒢(0)(iωa)\mathcal{G}^{(0)}(i\omega_{a}) is the non-interacting Green’s function, and Σ(iωa)\Sigma(i\omega_{a}) is the lowest order (Fan-Migdal) ee-ph self-energy [1, 15, 30], whose band- and 𝒌\bm{k}-dependent expression is

Σnn𝒌(iωa)=1βNqVucmm𝒒ν,iqc[gnmν(𝒌,𝒒)]gnmν(𝒌,𝒒)×𝒟ν𝒒(iqc)𝒢mm𝒌+𝒒(iωa+iqc).\begin{split}\Sigma_{nn^{\prime}\bm{k}}&(i\omega_{a})=-\frac{1}{\beta N_{q}V_{\text{uc}}}\sum_{mm^{\prime}\bm{q}\nu,iq_{c}}\left[g_{n^{\prime}m^{\prime}\nu}(\bm{k},\bm{q})\right]^{*}g_{nm\nu}(\bm{k},\bm{q})\\ &\times\mathcal{D}_{\nu\bm{q}}(iq_{c})\mathcal{G}_{mm^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+iq_{c}).\end{split} (2)

Here, β=1/kBT\beta=1/{k_{B}T} at temperature TT, NqN_{q} is the number of 𝒒\bm{q}-points in the summation, VucV_{\text{uc}} is the unit cell volume, qcq_{c} is the bosonic Matsubara frequency of the phonon, and 𝒟ν𝒒(iqc)=2ων𝒒/((iqc)2ων𝒒2)\mathcal{D}_{\nu\bm{q}}(iq_{c})=2\omega_{\nu\bm{q}}/((iq_{c})^{2}-\omega_{\nu\bm{q}}^{2}) is the non-interacting phonon Green’s function for a phonon with mode index ν\nu, wave-vector 𝒒\bm{q}, and energy ων𝒒\omega_{\nu\bm{q}}. The ee-ph matrix elements gnmν(𝒌,𝒒)g_{nm\nu}(\bm{k},\bm{q}) quantify the probability amplitude for an electron in a Bloch state |ψn𝒌\ket{\psi_{n\bm{k}}}, with band index nn and crystal momentum 𝒌\bm{k}, to scatter into a final state |ψm𝒌+𝒒\ket{\psi_{m\bm{k}+\bm{q}}} by emitting or absorbing a phonon [15, 31],

gnmν(𝒌,𝒒)=ψm𝒌+𝒒|ν𝒒V^|ψn𝒌,g_{nm\nu}(\bm{k},\bm{q})\!=\!\bra{\psi_{m\bm{k}+\bm{q}}}\partial_{\nu\bm{q}}\hat{V}\!\ket{\psi_{n\bm{k}}}\!, (3)

where ν𝒒V^\partial_{\nu\bm{q}}\hat{V} is the perturbation to the potential acting on an electron due to a given phonon mode (ν,𝒒)(\nu,\bm{q}).

Refer to caption
Figure 1: (a) Bare bubble diagram without the vertex correction. (b) Bubble diagram including the vertex correction. (c) Bethe-Salpeter equation for the vertex corrections Λ\Lambda from electron-phonon interactions within the ladder approximation. The wavy line is the phonon propagator and the red dots are the ee-ph matrix elements gnmν(𝒌,𝒒)g_{nm\nu}(\bm{k},\bm{q}).

II.2 Kubo formula and correlation function

We consider a complex vector operator 𝐀^\mathbf{\hat{A}}, with matrix elements in the direction α\alpha written as Anm𝒌α=m𝒌|A^α|n𝒌A^{\alpha}_{nm\bm{k}}=\bra{m\bm{k}}\hat{A}^{\alpha}\ket{n\bm{k}}. We derive the 𝐀^𝐀^\mathbf{\hat{A}}-\mathbf{\hat{A}} correlation function with a procedure analogous to the derivation of the dc conductivity in the ladder approximation [4]. Here, the operator 𝐀^\mathbf{\hat{A}} is in general non-diagonal in the band index, leading to matrix elements Anm𝒌αA^{\alpha}_{nm\bm{k}}, so the derivation for the diagonal case given in Ref. [4] needs to be extended to non-diagonal operators and vertex corrections.
We first derive the correlation function in imaginary time and frequency, and then extend it to real frequencies via analytic continuation. The retarded correlation function for the operator 𝐀^\mathbf{\hat{A}} can be obtained from the Kubo formula [1]

χαβ(𝒑,iνb)=0β𝑑τeiνbτTτA^α(𝒑,τ)A^β(𝒑,0),\chi_{\alpha\beta}(\bm{p},i\nu_{b})=\int_{0}^{\beta}d\tau e^{i\nu_{b}\tau}\left<T_{\tau}\hat{A}^{\alpha}(\bm{p},\tau)\hat{A}^{\beta}(-\bm{p},0)\right>, (4)

where 𝒑\bm{p} is a wave-vector, νb\nu_{b} is a bosonic Matsubara frequency, τ\tau is imaginary time ranging from 0 to β=1/kBT\beta=1/{k_{B}T} at temperature TT, and TτT_{\tau} is the imaginary time-ordering operator. Here we focus on the 𝒑0\bm{p}\rightarrow 0 limit, so we drop 𝒑\bm{p} from the equations. This correlation function can be expressed as a sum of bubble diagrams PP as [1]

χαβ(iνb)=1βiωaP(iωa,iωa+iνb).\chi_{\alpha\beta}(i\nu_{b})=\frac{1}{\beta}\sum_{i\omega_{a}}P(i\omega_{a},i\omega_{a}+i\nu_{b}). (5)

Let us consider the bare bubble diagram that includes the electron self-energy only in the electron propagator 𝒢\mathcal{G}, as shown in Fig. 1(a):

χαβ(iνb)=1βVuciωaTr[𝒢(iωa)A^α𝒢(iωa+iνb)A^β],\chi_{\alpha\beta}(i\nu_{b})=\frac{1}{\beta V_{\text{uc}}}\!\sum_{i\omega_{a}}\Tr[\mathcal{G}(i\omega_{a})\hat{A}^{\alpha}\mathcal{G}(i\omega_{a}+i\nu_{b})\hat{A}^{\beta}\big{]}, (6)

where the trace is evaluated over the band and momentum indices. In this expression, the operator 𝐀^\mathbf{\hat{A}} can be regarded as the bare vertex of the correlation function. For the velocity operator, Eq. (6) leads to the well-known Drude conductivity [4, 1].
In this work, the corrections to the vertex originate from the ee-ph interactions, which couple electronic states with different band and crystal momenta. Figure 1(b) shows the correlation function including the vertex correction Λ\Lambda,

χαβ(iνb)=1βVuciωaTr[𝒢(iωa)A^α𝒢(iωa+iνb)×A^βΛβ(iωa,iωa+iνb)],\begin{split}\chi_{\alpha\beta}(i\nu_{b})=\frac{1}{\beta V_{\text{uc}}}\!\sum_{i\omega_{a}}\Tr&\big{[}\mathcal{G}(i\omega_{a})\hat{A}^{\alpha}\mathcal{G}(i\omega_{a}+i\nu_{b})\\ &\times\hat{A}^{\beta}\Lambda^{\beta}(i\omega_{a},i\omega_{a}+i\nu_{b})\big{]},\end{split} (7)

where A^βΛβ(iωa,iωa+iνb)\hat{A}^{\beta}\Lambda^{\beta}(i\omega_{a},i\omega_{a}+i\nu_{b}) is the phonon-dressed vertex for the operator A^\hat{A} in the Cartesian direction β\beta. Note that the vertex correction Λβ(iωa,iωa+iνb){\Lambda}^{\beta}(i\omega_{a},i\omega_{a}+i\nu_{b}) is a complex-valued vector that contains information about the operator dynamics renormalized by the ee-ph interactions.

II.3 Bethe-Salpeter equation for the phonon-dressed vertex

The leading correction to the vertex is obtained by summing over ladder diagrams, which can be viewed as an abstract form of charge conservation in the presence of ee-ph scattering [1, 4]. The vertex correction Λnn𝒌α\Lambda^{\alpha}_{nn^{\prime}\bm{k}} satisfies the self-consistent BSE, shown diagrammatically in Fig. 1(c) and written as

Ann𝒌αΛnn𝒌α(iωa,iωa+iνb)=Ann𝒌α1βNqVucmmll𝒒ν,iqc[gnmν(𝒌,𝒒)]gnmν(𝒌,𝒒)𝒟ν𝒒(iqc)×𝒢ml𝒌+𝒒(iωa+iqc)𝒢lm𝒌+𝒒(iωa+iνb+iqc)×All𝒌αΛll𝒌+𝒒α(iωa+iqc,iωa+iνb+iqc).\begin{split}A^{\alpha}_{nn^{\prime}\bm{k}}&\Lambda^{\alpha}_{nn^{\prime}\bm{k}}(i\omega_{a},i\omega_{a}+i\nu_{b})=A^{\alpha}_{nn^{\prime}\bm{k}}\\ -&\frac{1}{\beta N_{q}V_{\text{uc}}}\sum_{mm^{\prime}ll^{\prime}\bm{q}\nu,iq_{c}}\left[g_{n^{\prime}m^{\prime}\nu}(\bm{k},\bm{q})\right]^{*}g_{nm\nu}(\bm{k},\bm{q})\mathcal{D}_{\nu\bm{q}}(iq_{c})\\ &\times\mathcal{G}_{ml\bm{k}+\bm{q}}(i\omega_{a}+iq_{c})\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+i\nu_{b}+iq_{c})\\ &\times A^{\alpha}_{ll^{\prime}\bm{k}}\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+iq_{c},i\omega_{a}+i\nu_{b}+iq_{c}).\end{split} (8)

The kernel of this BSE [32] is the ee-ph interaction [gnmν(𝒌,𝒒)]gnmν(𝒌,𝒒)𝒟ν𝒒(iqc)\left[g_{n^{\prime}m^{\prime}\nu}(\bm{k},\bm{q})\right]^{*}g_{nm\nu}(\bm{k},\bm{q})\mathcal{D}_{\nu\bm{q}}(iq_{c}).
Following Mahan [1] and Ref. [4], we first sum over the bosonic Matsubara frequency iqciq_{c} in Eq. (8). This summation, defined as S(iωa,iωa+iνb)S(i\omega_{a},i\omega_{a}+i\nu_{b}), reads:

S(iωa,iωa+iνb)=llSll(iωa,iωa+iνb)=1βlliqc𝒟ν𝒒(iqc)Λll𝒌+𝒒α(iωa+iqc,iωa+iνb+iqc)×𝒢ml𝒌+𝒒(iωa+iqc)𝒢lm𝒌+𝒒(iωa+iνb+iqc).\begin{split}S&(i\omega_{a},i\omega_{a}+i\nu_{b})=\sum_{ll^{\prime}}S_{ll^{\prime}}(i\omega_{a},i\omega_{a}+i\nu_{b})\\ =&\frac{1}{\beta}\sum_{ll^{\prime}iq_{c}}\,\mathcal{D}_{\nu\bm{q}}(iq_{c})\,\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+iq_{c},i\omega_{a}+i\nu_{b}+iq_{c})\\ &\times\mathcal{G}_{ml\bm{k}+\bm{q}}(i\omega_{a}+iq_{c})\,\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+i\nu_{b}+iq_{c}).\end{split} (9)

As usual, the summation is done by constructing a contour integral along a circle at infinity:

dz2πinB(z)𝒟ν𝒒(z)Λll𝒌+𝒒α(iωa+z,iωa+iνb+z)×𝒢ml𝒌+𝒒(iωa+z)𝒢lm𝒌+𝒒(iωa+iνb+z),\begin{split}\oint\frac{dz}{2\pi i}&n_{\rm B}(z)\mathcal{D}_{\nu\bm{q}}(z)\,\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+z,i\omega_{a}+i\nu_{b}+z)\\ \times\,&\mathcal{G}_{ml\bm{k}+\bm{q}}(i\omega_{a}+z)\,\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+i\nu_{b}+z),\end{split} (10)

where nBn_{\rm B} are Bose-Einstein occupations. The integrand has poles at z=iqcz=iq_{c}, z=±ων𝒒z=\pm\omega_{\nu\bm{q}}, and branch cuts along z=iωaz=-i\omega_{a} and z=iωaiνbz=-i\omega_{a}-i\nu_{b} [1, 4]. Employing Cauchy’s residue theorem, we obtain

Sll(iωa,iωa+iνb)=Nν𝒒Λll𝒌+𝒒α(iωa+ων𝒒,iωa+iνb+ων𝒒)×𝒢ml𝒌+𝒒(iωa+ων𝒒)𝒢lm𝒌+𝒒(iωa+iνb+ων𝒒)[Nν𝒒+1]Λll𝒌+𝒒α(iωaων𝒒,iωa+iνbων𝒒)×𝒢ml𝒌+𝒒(iωaων𝒒)𝒢lm𝒌+𝒒(iωa+iνbων𝒒)dε2πif(ε)2ων𝒒(εiωa)2ων𝒒2𝒢lm𝒌+𝒒(ε+iνb)×[Λll𝒌+𝒒α(ε+iη,ε+iνb)𝒢ml𝒌+𝒒(ε+iη)Λll𝒌+𝒒α(εiη,ε+iνb)𝒢ml𝒌+𝒒(εiη)]dε2πif(ε)2ων𝒒(εiωaiνb)2ων𝒒2𝒢ml𝒌+𝒒(εiνb)×[Λll𝒌+𝒒α(εiνb,ε+iη)𝒢lm𝒌+𝒒(ε+iη)Λll𝒌+𝒒α(εiνb,εiη)𝒢lm𝒌+𝒒(εiη)],\begin{split}S_{ll^{\prime}}(i&\omega_{a},i\omega_{a}+i\nu_{b})\\ =-&N_{\nu\bm{q}}\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+\omega_{\nu\bm{q}},i\omega_{a}+i\nu_{b}+\omega_{\nu\bm{q}})\\ &\times\mathcal{G}_{ml\bm{k}+\bm{q}}(i\omega_{a}+\omega_{\nu\bm{q}})\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+i\nu_{b}+\omega_{\nu\bm{q}})\\ -&[N_{\nu\bm{q}}+1]\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(i\omega_{a}-\omega_{\nu\bm{q}},i\omega_{a}+i\nu_{b}-\omega_{\nu\bm{q}})\\ &\times\mathcal{G}_{ml\bm{k}+\bm{q}}(i\omega_{a}-\omega_{\nu\bm{q}})\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+i\nu_{b}-\omega_{\nu\bm{q}})\\ -&\int\frac{d\varepsilon^{\prime}}{2\pi i}f(\varepsilon^{\prime})\frac{2\omega_{\nu\bm{q}}}{(\varepsilon^{\prime}-i\omega_{a})^{2}-\omega_{\nu\bm{q}}^{2}}\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}(\varepsilon^{\prime}+i\nu_{b})\\ &\times[\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(\varepsilon^{\prime}+i\eta,\varepsilon^{\prime}+i\nu_{b})\mathcal{G}_{ml\bm{k}+\bm{q}}(\varepsilon^{\prime}+i\eta)\\ &-\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(\varepsilon^{\prime}-i\eta,\varepsilon^{\prime}+i\nu_{b})\mathcal{G}_{ml\bm{k}+\bm{q}}(\varepsilon^{\prime}-i\eta)]\\ -&\int\frac{d\varepsilon^{\prime}}{2\pi i}f(\varepsilon^{\prime})\frac{2\omega_{\nu\bm{q}}}{(\varepsilon^{\prime}-i\omega_{a}-i\nu_{b})^{2}-\omega_{\nu\bm{q}}^{2}}\mathcal{G}_{ml\bm{k}+\bm{q}}(\varepsilon^{\prime}-i\nu_{b})\\ &\times[\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(\varepsilon^{\prime}-i\nu_{b},\varepsilon^{\prime}+i\eta)\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}(\varepsilon^{\prime}+i\eta)\\ &-\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(\varepsilon^{\prime}-i\nu_{b},\varepsilon^{\prime}-i\eta)\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}(\varepsilon^{\prime}-i\eta)]\,,\end{split} (11)

where Nν𝒒=nB(ων𝒒)N_{\nu\bm{q}}\!=\!n_{\rm B}(\omega_{\nu\bm{q}}) are temperature dependent phonon occupations, f(ε)f(\varepsilon) is the Fermi-Dirac distribution function, and η\eta is a positive infinitesimal.
The leading contribution to Sll(iωa,iωa+iνb)S_{ll^{\prime}}(i\omega_{a},i\omega_{a}+i\nu_{b}) comes from the combination of retarded and advanced Green’s functions, GRG^{R} and GAG^{A}, while terms of O([GR]2,[GA]2)O([G^{R}]^{2},[G^{A}]^{2}) can be neglected at low electron density [1, 4]. Therefore, after the analytic continuations iωaεiηi\omega_{a}\rightarrow\varepsilon-i\eta and iωa+iνbε+ν+iηi\omega_{a}+i\nu_{b}\rightarrow\varepsilon+\nu+i\eta, and using the identity 1x+iη=P1xiπδ(x)\frac{1}{x+i\eta}=P\frac{1}{x}-i\pi\delta(x), we obtain Sll(εiη,ε+iη)S_{ll^{\prime}}(\varepsilon-i\eta,\varepsilon+i\eta) in limit of ν0\nu\rightarrow 0,

Sll(εiη,ε+iη)=[Nν𝒒+f(ε+ων𝒒)]Λαll𝒌+𝒒(ε+ων𝒒)×Gml𝒌+𝒒R(ε+ων𝒒)Glm𝒌+𝒒A(ε+ων𝒒)[Nν𝒒+1f(εων𝒒)]Λαll𝒌+𝒒(εων𝒒)×Gml𝒌+𝒒R(εων𝒒)Glm𝒌+𝒒A(εων𝒒),\begin{split}S_{ll^{\prime}}(\varepsilon-&i\eta,\varepsilon+i\eta)\\ =-[&N_{\nu\bm{q}}+f(\varepsilon+\omega_{\nu\bm{q}})]\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(\varepsilon+\omega_{\nu\bm{q}})\\ &\times G_{ml\bm{k}+\bm{q}}^{R}(\varepsilon+\omega_{\nu\bm{q}})G_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}^{A}(\varepsilon+\omega_{\nu\bm{q}})\\ -[&N_{\nu\bm{q}}+1-f(\varepsilon-\omega_{\nu\bm{q}})]\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(\varepsilon-\omega_{\nu\bm{q}})\\ &\times G_{ml\bm{k}+\bm{q}}^{R}(\varepsilon-\omega_{\nu\bm{q}})G_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}^{A}(\varepsilon-\omega_{\nu\bm{q}}),\end{split} (12)

where the index A (R) stands for advanced (retarded) function, and Λα(ε)Λα(εiη,ε+iη)\Lambda^{\alpha}(\varepsilon)\equiv\Lambda^{\alpha}(\varepsilon-i\eta,\varepsilon+i\eta).
Using this result, we write the self-consistent BSE for the phonon-dressed vertex 𝐀^Λ\mathbf{\hat{A}}\Lambda at energy ε\varepsilon as:

Ann𝒌αΛnn𝒌α(ε)=Ann𝒌α+1NqVucmmll𝒒ν[gnmν(𝒌,𝒒)]gnmν(𝒌,𝒒)All𝒌+𝒒α×[(Nν𝒒+f(ε+ων𝒒))Λll𝒌+𝒒α(ε+ων𝒒)×Gml𝒌+𝒒R(ε+ων𝒒)Glm𝒌+𝒒A(ε+ων𝒒)+(Nν𝒒+1f(εων𝒒))Λll𝒌+𝒒α(εων𝒒)×Gml𝒌+𝒒R(εων𝒒)Glm𝒌+𝒒A(εων𝒒)].\begin{split}&A^{\alpha}_{nn^{\prime}\bm{k}}\Lambda^{\alpha}_{nn^{\prime}\bm{k}}(\varepsilon)=A^{\alpha}_{nn^{\prime}\bm{k}}\\ &+\frac{1}{N_{q}V_{\text{uc}}}\sum_{mm^{\prime}ll^{\prime}\bm{q}\nu}\left[g_{n^{\prime}m^{\prime}\nu}(\bm{k},\bm{q})\right]^{*}g_{nm\nu}(\bm{k},\bm{q})A^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}\\ &\times\bigg{[}(N_{\nu\bm{q}}+f(\varepsilon+\omega_{\nu\bm{q}}))\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(\varepsilon+\omega_{\nu\bm{q}})\\ &\times G_{ml\bm{k}+\bm{q}}^{R}(\varepsilon+\omega_{\nu\bm{q}})G_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}^{A}(\varepsilon+\omega_{\nu\bm{q}})\\ &+(N_{\nu\bm{q}}+1-f(\varepsilon-\omega_{\nu\bm{q}}))\Lambda^{\alpha}_{ll^{\prime}\bm{k}+\bm{q}}(\varepsilon-\omega_{\nu\bm{q}})\\ &\times G_{ml\bm{k}+\bm{q}}^{R}(\varepsilon-\omega_{\nu\bm{q}})G_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}^{A}(\varepsilon-\omega_{\nu\bm{q}})\bigg{]}.\end{split} (13)

By solving Eq. (13), we obtain the phonon-dressed vertex Ann𝒌αΛnn𝒌α(ε)A^{\alpha}_{nn^{\prime}\bm{k}}\Lambda^{\alpha}_{nn^{\prime}\bm{k}}(\varepsilon) and its dependence on band, crystal momentum and energy.
In the weak scattering regime, where the electron spectral function has a well-defined quasiparticle peak [5] and the off-diagonal self-energy can be neglected [33, 34], the Green’s function becomes band-diagonal and the self-energies can be evaluated on-shell. Then the product of the retarded and advanced Green’s functions, GRGAG^{R}G^{A}, can be approximated as [35]

Gm𝒌+𝒒R(ε)Gm𝒌+𝒒A(ε)=Gm𝒌+𝒒A(ε)Gm𝒌+𝒒R(ε)Gm𝒌+𝒒R(ε)1Gm𝒌+𝒒A(ε)1πδ(εεm𝒌+𝒒)+πδ(εεm𝒌+𝒒)iP1εεm𝒌+𝒒+iP1εεm𝒌+𝒒i(Σm𝒌+𝒒RΣm𝒌+𝒒A)+i(εm𝒌+𝒒εm𝒌+𝒒),\begin{split}&G^{R}_{m\bm{k}+\bm{q}}(\varepsilon)G^{A}_{m^{\prime}\bm{k}+\bm{q}}(\varepsilon)\\ &=\frac{G_{m^{\prime}\bm{k}+\bm{q}}^{A}(\varepsilon)-G_{m\bm{k}+\bm{q}}^{R}(\varepsilon)}{G_{m\bm{k}+\bm{q}}^{R}(\varepsilon)^{-1}-G_{m^{\prime}\bm{k}+\bm{q}}^{A}(\varepsilon)^{-1}}\\ &\approx\frac{\pi\delta(\varepsilon\!-\!\varepsilon_{m^{\prime}\bm{k}+\bm{q}})\!+\!\pi\delta(\varepsilon\!-\!\varepsilon_{m\bm{k}+\bm{q}})\!-\!iP\frac{1}{\varepsilon-\varepsilon_{m^{\prime}\bm{k}+\bm{q}}}\!+\!iP\frac{1}{\varepsilon-\varepsilon_{m\bm{k}+\bm{q}}}}{i(\Sigma^{R}_{m\bm{k}+\bm{q}}-\Sigma^{A}_{m^{\prime}\bm{k}+\bm{q}})+i(\varepsilon_{m\bm{k}+\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}})},\end{split} (14)

a function that is strongly peaked at electron energies ε=εm𝒌+𝒒\varepsilon\!=\!\varepsilon_{m\bm{k}+\bm{q}} and ε=εm𝒌+𝒒\varepsilon\!=\!\varepsilon_{m^{\prime}\bm{k}+\bm{q}}. Therefore, we can further simplify the full-frequency BSE in Eq. (13) to a double-pole ansatz, which evaluates the vertex corrections only at these two energies:

Ann𝒌αΛnn𝒌α(ε)=Ann𝒌α+2πNqVucmm𝒒ν[gnmν(𝒌,𝒒)]gnmν(𝒌,𝒒)×12[{(Nν𝒒+fm𝒌+𝒒)(δ(ε+ων𝒒εm𝒌+𝒒)iπP1ε+ων𝒒εm𝒌+𝒒)+(Nν𝒒+1fm𝒌+𝒒)(δ(εων𝒒εm𝒌+𝒒)iπP1εων𝒒εm𝒌+𝒒)}×Amm𝒌+𝒒αΛmm𝒌+𝒒α(εm𝒌+𝒒)i(Σm𝒌+𝒒RΣm𝒌+𝒒A)+i(εm𝒌+𝒒εm𝒌+𝒒)+{(Nν𝒒+fm𝒌+𝒒)(δ(ε+ων𝒒εm𝒌+𝒒)+iπP1ε+ων𝒒εm𝒌+𝒒)+(Nν𝒒+1fm𝒌+𝒒)(δ(εων𝒒εm𝒌+𝒒)+iπP1εων𝒒εm𝒌+𝒒)}×Amm𝒌+𝒒αΛmm𝒌+𝒒α(εm𝒌+𝒒)i(Σm𝒌+𝒒RΣm𝒌+𝒒A)+i(εm𝒌+𝒒εm𝒌+𝒒)],\begin{split}&A^{\alpha}_{nn^{\prime}\bm{k}}\Lambda^{\alpha}_{nn^{\prime}\bm{k}}(\varepsilon)=A^{\alpha}_{nn^{\prime}\bm{k}}+\frac{2\pi}{N_{q}V_{\text{uc}}}\sum_{mm^{\prime}\bm{q}\nu}\left[g_{n^{\prime}m^{\prime}\nu}(\bm{k},\bm{q})\right]^{*}g_{nm\nu}(\bm{k},\bm{q})\\ &\times\frac{1}{2}\bigg{[}\{(N_{\nu\bm{q}}+f_{m\bm{k}+\bm{q}})(\delta(\varepsilon+\omega_{\nu\bm{q}}-\varepsilon_{m\bm{k}+\bm{q}})-\frac{i}{\pi}P\frac{1}{\varepsilon+\omega_{\nu\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}}})\\ &+(N_{\nu\bm{q}}+1-f_{m\bm{k}+\bm{q}})(\delta(\varepsilon-\omega_{\nu\bm{q}}-\varepsilon_{m\bm{k}+\bm{q}})-\frac{i}{\pi}P\frac{1}{\varepsilon-\omega_{\nu\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}}})\}\times\frac{A^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}\Lambda^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}(\varepsilon_{m\bm{k}+\bm{q}})}{i(\Sigma^{R}_{m\bm{k}+\bm{q}}-\Sigma^{A}_{m^{\prime}\bm{k}+\bm{q}})+i(\varepsilon_{m\bm{k}+\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}})}\\ &+\{(N_{\nu\bm{q}}+f_{m^{\prime}\bm{k}+\bm{q}})(\delta(\varepsilon+\omega_{\nu\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}})+\frac{i}{\pi}P\frac{1}{\varepsilon+\omega_{\nu\bm{q}}-\varepsilon_{m\bm{k}+\bm{q}}})\\ &+(N_{\nu\bm{q}}+1-f_{m^{\prime}\bm{k}+\bm{q}})(\delta(\varepsilon-\omega_{\nu\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}})+\frac{i}{\pi}P\frac{1}{\varepsilon-\omega_{\nu\bm{q}}-\varepsilon_{m\bm{k}+\bm{q}}})\}\times\frac{A^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}\Lambda^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}(\varepsilon_{m^{\prime}\bm{k}+\bm{q}})}{i(\Sigma^{R}_{m\bm{k}+\bm{q}}-\Sigma^{A}_{m^{\prime}\bm{k}+\bm{q}})+i(\varepsilon_{m\bm{k}+\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}})}\bigg{]},\end{split} (15)

where ε\varepsilon equals εn𝒌\varepsilon_{n\bm{k}} or εn𝒌\varepsilon_{n^{\prime}\bm{k}}, and fm𝒌+𝒒f(εm𝒌+𝒒)f_{m\bm{k}+\bm{q}}\equiv f(\varepsilon_{m\bm{k}+\bm{q}}).
We have tested the consistency of this theory by deriving a Ward identity [36, 4, 1] relating the self-energy and vertex corrections (see Appendix A). This result guarantees that ee-ph diagrams are taken into account consistently in the self-energy and in our BSE.

II.4 The dressed vertex and its interpretation

Table 1: Summary of the formalism for charge transport and spin decoherence.
Charge transport (Ref. [4]) Spin decoherence
Operator vn𝒌v_{n\bm{k}} (diagonal) snm𝒌s_{nm\bm{k}} (non-diagonal)
External field \mathcal{F} Vector potential (𝐀\mathbf{A}) Magnetic field (𝐁\mathbf{B})
Injection field ˙\dot{\mathcal{F}} 𝐄(ν)=iν𝐀(ν)\mathbf{E}(\nu)=-i\nu\mathbf{A}(\nu) 𝐁˙(ν)=iν𝐁(ν)\dot{\mathbf{B}}(\nu)=-i\nu\mathbf{B}(\nu)
Vertex correction Λ\Lambda Λn𝒌α(εn𝒌)\Lambda_{n\bm{k}}^{\alpha}(\varepsilon_{n\bm{k}}) Λnn𝒌α(εn𝒌),Λnn𝒌α(εn𝒌)\Lambda_{nn^{\prime}\bm{k}}^{\alpha}(\varepsilon_{n\bm{k}}),\Lambda_{nn^{\prime}\bm{k}}^{\alpha}(\varepsilon_{n^{\prime}\bm{k}})
Renormalized dynamics τ\tau, ω\omega τn𝒌(tr)α=τn𝒌e-phΛn𝒌α(ϵn𝒌)\tau_{n\bm{k}}^{(\text{tr})\alpha}=\tau^{\text{e-ph}}_{n\bm{k}}\Lambda_{n\bm{k}}^{\alpha}(\epsilon_{n\bm{k}}) 11τnn𝒌α(εn𝒌)+iωnn𝒌α(εn𝒌)=Λnn𝒌α(εn𝒌)i(Σn𝒌RΣn𝒌A)+i(εn𝒌εn𝒌)\frac{1}{\frac{1}{\tau_{nn^{\prime}\bm{k}}^{\alpha}(\varepsilon_{n\bm{k}})}+i{\omega}_{nn^{\prime}\bm{k}}^{\alpha}(\varepsilon_{n\bm{k}})}=\frac{\Lambda^{\alpha}_{nn^{\prime}\bm{k}}(\varepsilon_{n\bm{k}})}{i(\Sigma^{R}_{n\bm{k}}-\Sigma^{A}_{n^{\prime}\bm{k}})+i(\varepsilon_{n\bm{k}}-\varepsilon_{n^{\prime}\bm{k}})}

We focus on the dressed operator divided by the band energy difference, a key term in Eq. (15):

Amm𝒌+𝒒αΛmm𝒌+𝒒α(εm𝒌+𝒒)i(Σm𝒌+𝒒RΣm𝒌+𝒒A)+i(εm𝒌+𝒒εm𝒌+𝒒).\frac{A^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}\Lambda^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}(\varepsilon_{m^{\prime}\bm{k}+\bm{q}})}{i(\Sigma^{R}_{m\bm{k}+\bm{q}}\!-\!\Sigma^{A}_{m^{\prime}\bm{k}+\bm{q}})\!+\!i(\varepsilon_{m\bm{k}+\bm{q}}\!-\!\varepsilon_{m^{\prime}\bm{k}+\bm{q}})}. (16)

This ratio describes the renormalized dynamics associated with the operator 𝐀^\mathbf{\hat{A}} in the presence of ee-ph interactions. This dynamics is obtained by dividing Eq. (16) by the bare operator expectation value Amm𝒌+𝒒αA^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}, obtaining

Λmm𝒌+𝒒α(εm𝒌+𝒒)i(Σm𝒌+𝒒RΣm𝒌+𝒒A)+i(εm𝒌+𝒒εm𝒌+𝒒).\frac{\Lambda^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}(\varepsilon_{m^{\prime}\bm{k}+\bm{q}})}{i(\Sigma^{R}_{m\bm{k}+\bm{q}}\!-\!\Sigma^{A}_{m^{\prime}\bm{k}+\bm{q}})\!+\!i(\varepsilon_{m\bm{k}+\bm{q}}\!-\!\varepsilon_{m^{\prime}\bm{k}+\bm{q}})}. (17)

The physical meaning of this ratio can be understood by analyzing the simple case of the velocity operator. As the velocity operator is band-diagonal and satisfies vmm𝒌+𝒒α=vm𝒌+𝒒αδmmv^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}=v^{\alpha}_{m\bm{k}+\bm{q}}\delta_{mm^{\prime}}, the band energy difference in the denominator vanishes, so the denominator is purely real because Σm𝒌+𝒒A=(Σm𝒌+𝒒R)\Sigma^{A}_{m^{\prime}\bm{k}+\bm{q}}=(\Sigma^{R}_{m^{\prime}\bm{k}+\bm{q}})^{*}. Thus Eq. (16) for the velocity operator becomes

vm𝒌+𝒒αΛmm𝒌+𝒒α(εm𝒌+𝒒)i(Σm𝒌+𝒒RΣm𝒌+𝒒A)=vm𝒌+𝒒ατm𝒌+𝒒e-phΛmm𝒌+𝒒α(εm𝒌+𝒒),\frac{v^{\alpha}_{m\bm{k}+\bm{q}}\Lambda^{\alpha}_{mm\bm{k}+\bm{q}}(\varepsilon_{m\bm{k}+\bm{q}})}{i(\Sigma^{R}_{m\bm{k}+\bm{q}}-\Sigma^{A}_{m\bm{k}+\bm{q}})}=v^{\alpha}_{m\bm{k}+\bm{q}}\tau^{\text{e-ph}}_{m\bm{k}+\bm{q}}\Lambda^{\alpha}_{mm\bm{k}+\bm{q}}(\varepsilon_{m\bm{k}+\bm{q}}), (18)

where we used τm𝒌+𝒒e-ph=1/|2ImΣm𝒌+𝒒|\tau^{\text{e-ph}}_{m\bm{k}+\bm{q}}=1/|2\imaginary\Sigma_{m\bm{k}+\bm{q}}| for the ee-ph collision time. This equation gives the renormalized ee-ph mean free path, and dividing by the bare velocity we obtain the renormalized relaxation time, also known as the transport relaxation time [4],

τm𝒌+𝒒α(tr)τm𝒌+𝒒e-phΛmm𝒌+𝒒α(εm𝒌+𝒒)=Λmm𝒌+𝒒α(εm𝒌+𝒒)i(Σm𝒌+𝒒RΣm𝒌+𝒒A).\tau_{m\bm{k}+\bm{q}}^{\alpha(\text{tr})}\equiv\tau^{\text{e-ph}}_{m\bm{k}+\bm{q}}\,\Lambda^{\alpha}_{mm\bm{k}+\bm{q}}(\varepsilon_{m\bm{k}+\bm{q}})=\frac{\Lambda^{\alpha}_{mm\bm{k}+\bm{q}}(\varepsilon_{m\bm{k}+\bm{q}})}{{i(\Sigma^{R}_{m\bm{k}+\bm{q}}-\Sigma^{A}_{m\bm{k}+\bm{q}}})}. (19)

For a non-diagonal operator, both the vertex correction and the operator expectation value are complex, so the ratio in Eq. (17) cannot be represented by a single real quantity with units of time as in Eq. (19). To extend the vertex correction to non-diagonal operators, we generalize this formalism by defining the renormalized microscopic relaxation times τmm𝒌+𝒒α(ε)\tau_{mm^{\prime}\bm{k}+\bm{q}}^{\alpha}(\varepsilon) and introducing the precession frequencies ωmm𝒌+𝒒α(ε)\omega_{mm^{\prime}\bm{k}+\bm{q}}^{\alpha}(\varepsilon):

11τmm𝒌+𝒒α(ε)+iωmm𝒌+𝒒α(ε)Λmm𝒌+𝒒α(ε)i(Σm𝒌+𝒒RΣm𝒌+𝒒A)+i(εm𝒌+𝒒εm𝒌+𝒒),\begin{split}&\frac{1}{\frac{1}{\tau_{mm^{\prime}\bm{k}+\bm{q}}^{\alpha}(\varepsilon)}+i{\omega}_{mm^{\prime}\bm{k}+\bm{q}}^{\alpha}(\varepsilon)}\\ &\equiv\frac{\Lambda^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}(\varepsilon)}{i(\Sigma^{R}_{m\bm{k}+\bm{q}}\!-\!\Sigma^{A}_{m^{\prime}\bm{k}+\bm{q}})\!+\!i(\varepsilon_{m\bm{k}+\bm{q}}\!-\!\varepsilon_{m^{\prime}\bm{k}+\bm{q}})},\end{split} (20)

where ε\varepsilon equals εm𝒌+𝒒\varepsilon_{m\bm{k}+\bm{q}} or εm𝒌+𝒒\varepsilon_{m^{\prime}\bm{k}+\bm{q}}. This way, without the vertex correction, the renormalized relaxation time reduces to the (non-diagonal) ee-ph collision time, τmm𝒌+𝒒e-ph=1/|ImΣm𝒌+𝒒+ImΣm𝒌+𝒒|\tau^{\text{e-ph}}_{mm^{\prime}\bm{k}+\bm{q}}=1/|\imaginary\Sigma_{m\bm{k}+\bm{q}}+\imaginary\Sigma_{m^{\prime}\bm{k}+\bm{q}}|, and the renormalized precession frequency reduces to the bare operator rotation frequency, ωB=(εm𝒌+𝒒+ReΣm𝒌+𝒒)(εm𝒌+𝒒+ReΣm𝒌+𝒒)\omega_{\textrm{B}}=(\varepsilon_{m\bm{k}+\bm{q}}+\real\Sigma_{m\bm{k}+\bm{q}})-(\varepsilon_{m^{\prime}\bm{k}+\bm{q}}+\real\Sigma_{m^{\prime}\bm{k}+\bm{q}}), with Amm𝒌+𝒒α(t)eiωBtA^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}(t)\propto e^{i\,{\omega}_{\textrm{B}}t}.

II.5 Vertex correction to the susceptibility

We derive the vertex-corrected susceptibility in response to an external field for the generic observable 𝐀^\mathbf{\hat{A}}. Suppose that the complex operator A^α\hat{A}^{\alpha} couples to a vector field α\mathcal{F}^{\alpha}, with perturbation Hamiltonian H=A^ααH^{\prime}=-\hat{A}^{\alpha}\mathcal{F}^{\alpha}. The susceptibility is defined as the response function in

A^α(ν)=χαβ(ν)β(ν),\langle\hat{A}^{\alpha}(\nu)\rangle=\chi_{\alpha\beta}(\nu)\mathcal{F}^{\beta}(\nu), (21)

where \mathcal{F} is the external field along the direction β\beta, and A^α(ν)\langle\hat{A}^{\alpha}(\nu)\rangle is the response of the system along α\alpha at frequency ν\nu due to the applied field.
To study relaxation and dissipation, we rewrite the response of the system as

A^α(ν)=σαβ(ν)˙β(ν),\langle\hat{A}^{\alpha}(\nu)\rangle=\sigma_{\alpha\beta}(\nu)\dot{\mathcal{F}}^{\beta}(\nu), (22)

thus expressing it in terms of the susceptibility σαβ\sigma_{\alpha\beta} to the “injection field” at frequency ν\nu, and ˙β(ν)=iνβ(ν)\dot{\mathcal{F}}^{\beta}(\nu)=-i\nu\mathcal{F}^{\beta}(\nu). The injection field produces a nonequilibrium electron distribution with an injection rate equal to the inverse relaxation time of 𝐀^\mathbf{\hat{A}} [37]. From Eqs. (21)-(22), we obtain

σαβ(ν)=χαβ(ν)iν.\sigma_{\alpha\beta}(\nu)=\frac{\chi_{\alpha\beta}(\nu)}{-i\nu}. (23)

When \mathcal{F} is the vector potential 𝐀\mathbf{A}, the injection field becomes the electric field 𝐄(ν)=iν𝐀(ν)\mathbf{E}(\nu)=-i\nu\mathbf{A}(\nu), the observable of interest is the current operator Anm𝒌α=eδnmvn𝒌αA^{\alpha}_{nm\bm{k}}=e\delta_{nm}v^{\alpha}_{n\bm{k}}, and σαβ(ν)\sigma_{\alpha\beta}(\nu) is the frequency-dependent conductivity tensor. When \mathcal{F} is the magnetic field 𝐁\mathbf{B}, the injection field is its time derivative, 𝐁˙(ν)=iν𝐁(ν)\dot{\mathbf{B}}(\nu)=-i\nu\mathbf{B}(\nu), and the observable is the electron magnetic moment Anm𝒌α=gμBsnm𝒌αA^{\alpha}_{nm\bm{k}}=g\mu_{B}s^{\alpha}_{nm\bm{k}}, which is proportional to the spin matrix snm𝒌αs^{\alpha}_{nm\bm{k}} [37, 38]. These results are summarized in Table 1.
We write the correlation function with vertex correction [see Eq. (7)] as a contour integral along a circle at infinity [1, 4],

χαβ(iνb)=1Vucdz2πif(z)Tr[𝒢(z)A^α𝒢(z+iνb)A^βΛβ(z,z+iνb)],\begin{split}\chi_{\alpha\beta}(i\nu_{b})=&-\frac{1}{V_{\text{uc}}}\oint\frac{dz}{2\pi i}f(z)\Tr[\\ \hfil&\mathcal{G}(z)\hat{A}^{\alpha}\mathcal{G}(z+i\nu_{b})\hat{A}^{\beta}\Lambda^{\beta}(z,z+i\nu_{b})\big{]},\end{split} (24)

which has branch cuts along z=iνbz=-i\nu_{b} and z=0z=0, and poles at z=iωaz=i\omega_{a}, and thus

χαβ(iνb)=1Vucdε2πif(ε)Tr[𝒢(ε+iη)A^α𝒢(ε+iνb)A^βΛβ(ε+iη,ε+iνb)+𝒢(εiη)A^α𝒢(ε+iνb)A^βΛβ(εiη,ε+iνb)𝒢(εiνb)A^α𝒢(ε+iη)A^βΛβ(εiνb,ε+iη)+𝒢(εiνb)A^α𝒢(εiη)A^βΛβ(εiνb,εiη)].\begin{split}&\chi_{\alpha\beta}(i\nu_{b})=\frac{1}{V_{\text{uc}}}\int\frac{d\varepsilon}{2\pi i}f(\varepsilon)\Tr[\\ \hfil&-\mathcal{G}(\varepsilon+i\eta)\hat{A}^{\alpha}\mathcal{G}(\varepsilon+i\nu_{b})\hat{A}^{\beta}\Lambda^{\beta}(\varepsilon+i\eta,\varepsilon+i\nu_{b})\\ &+\mathcal{G}(\varepsilon-i\eta)\hat{A}^{\alpha}\mathcal{G}(\varepsilon+i\nu_{b})\hat{A}^{\beta}\Lambda^{\beta}(\varepsilon-i\eta,\varepsilon+i\nu_{b})\\ &-\mathcal{G}(\varepsilon-i\nu_{b})\hat{A}^{\alpha}\mathcal{G}(\varepsilon+i\eta)\hat{A}^{\beta}\Lambda^{\beta}(\varepsilon-i\nu_{b},\varepsilon+i\eta)\\ &+\mathcal{G}(\varepsilon-i\nu_{b})\hat{A}^{\alpha}\mathcal{G}(\varepsilon-i\eta)\hat{A}^{\beta}\Lambda^{\beta}(\varepsilon-i\nu_{b},\varepsilon-i\eta)\bigg{]}.\end{split} (25)

After the analytic continuation iνbν+iηi\nu_{b}\rightarrow\nu+i\eta, we obtain the retarded correlation function to leading order by neglecting the terms GRGRG^{R}G^{R} and GAGAG^{A}G^{A} [1, 4]:

χαβ(ν)=1Vucdε2πi(f(ε)f(ε+ν))Tr[GR(ε)A^αGA(ε+ν)A^βΛβ(iωaiη,iωa+iνb+iη)]1NkVucnm𝒌dε2πi(f(ε)f(ε+ν))Anm𝒌αAmn𝒌β×Λmn𝒌β(εiη,ε+ν+iη)πδ(εεn𝒌)+πδ(ε+νεm𝒌)i(Σm𝒌RΣn𝒌A)+i(εm𝒌+νεn𝒌),\begin{split}&\chi_{\alpha\beta}(\nu)\\ &=\frac{1}{V_{\text{uc}}}\int\frac{d\varepsilon}{2\pi i}(f(\varepsilon)-f(\varepsilon+\nu))\Tr[\\ \hfil&~{}~{}~{}~{}G^{R}(\varepsilon)\hat{A}^{\alpha}G^{A}(\varepsilon+\nu)\hat{A}^{\beta}\Lambda^{\beta}(i\omega_{a}-i\eta,i\omega_{a}+i\nu_{b}+i\eta)\big{]}\\ &\approx\frac{1}{N_{k}V_{\text{uc}}}\sum_{nm\bm{k}}\int\frac{d\varepsilon}{2\pi i}(f(\varepsilon)-f(\varepsilon+\nu))A^{\alpha}_{nm\bm{k}}A^{\beta}_{mn\bm{k}}\\ &\times\!\Lambda^{\beta}_{mn\bm{k}}(\varepsilon\!-\!i\eta,\varepsilon\!+\!\nu\!+\!i\eta)\frac{\pi\delta(\varepsilon\!-\!\varepsilon_{n\bm{k}})\!+\!\pi\delta(\varepsilon\!+\!\nu\!-\!\varepsilon_{m\bm{k}})}{i(\Sigma^{R}_{m\bm{k}}\!-\!\Sigma^{A}_{n\bm{k}})\!+\!i(\varepsilon_{m\bm{k}}\!+\!\nu\!-\!\varepsilon_{n\bm{k}})},\end{split} (26)

where NkN_{k} is the number of 𝒌\bm{k}-points, and we used Eq. (14) in the last equality. This equation characterizes the frequency-dependent response of the system [39].
We focus on the dc limit ν0\nu\rightarrow 0, where the driving field is static. The susceptibility with respect to the injection field becomes

limν0σαβ(ν)=limν01νImχαβ(ν)=1NkVucRenm𝒌Anm𝒌αAmn𝒌β×12[(dfn𝒌dε)Λmn𝒌β(εn𝒌)+(dfm𝒌dε)Λmn𝒌β(εm𝒌)]i(Σm𝒌RΣn𝒌A)+i(εm𝒌εn𝒌).\begin{split}&\lim_{\nu\rightarrow 0}\sigma_{\alpha\beta}(\nu)=-\lim_{\nu\rightarrow 0}\frac{1}{\nu}\imaginary\chi_{\alpha\beta}(\nu)\\ &=\frac{1}{N_{k}V_{\text{uc}}}\real\sum_{nm\bm{k}}A^{\alpha}_{nm\bm{k}}A^{\beta}_{mn\bm{k}}\\ &~{}~{}~{}~{}~{}\times\frac{\frac{1}{2}[(-\frac{df_{n\bm{k}}}{d\varepsilon})\Lambda^{\beta}_{mn\bm{k}}(\varepsilon_{n\bm{k}})+(-\frac{df_{m\bm{k}}}{d\varepsilon})\Lambda^{\beta}_{mn\bm{k}}(\varepsilon_{m\bm{k}})]}{{i(\Sigma^{R}_{m\bm{k}}-\Sigma^{A}_{n\bm{k}})+i(\varepsilon_{m\bm{k}}-\varepsilon_{n\bm{k}})}}.\end{split} (27)

This static susceptibility has both band-diagonal (n=mn=m) and off-diagonal (nmn\neq m) contributions. The band-diagonal contribution

σαβ(d)(0)=1NkVucn𝒌Ann𝒌αAnn𝒌βτn𝒌e-phΛnn𝒌β(εn𝒌)(dfn𝒌dε)\sigma_{\alpha\beta}^{\rm(d)}(0)=\frac{1}{N_{k}V_{\text{uc}}}\sum_{n\bm{k}}A^{\alpha}_{nn\bm{k}}A^{\beta}_{nn\bm{k}}\tau^{\text{e-ph}}_{n\bm{k}}\Lambda^{\beta}_{nn\bm{k}}(\varepsilon_{n\bm{k}})(-\frac{df_{n\bm{k}}}{d\varepsilon}) (28)

is the only contribution to the static susceptibility for a band-diagonal operator. For example, for the velocity operator, Eq. (28) becomes the well-known electrical conductivity tensor within the BTE, which has been studied extensively using both empirical and first-principles calculations [23, 24, 40, 41, 31] (note that solving exactly the BTE is equivalent to computing the velocity vertex correction [4]). For non-diagonal operators, our formalism introduces an off-diagonal contribution in Eq. (27):

σαβ(nd)(0)=1N𝒌VucRenm𝒌Anm𝒌αAmn𝒌β×12[(dfn𝒌dε)Λmn𝒌β(εn𝒌)+(dfm𝒌dε)Λmn𝒌β(εm𝒌)]i(Σm𝒌RΣn𝒌A)+i(εm𝒌εn𝒌).\begin{split}&\sigma_{\alpha\beta}^{\rm(nd)}(0)=\frac{1}{N_{\bm{k}}V_{\text{uc}}}\real\sum_{n\neq m\bm{k}}A^{\alpha}_{nm\bm{k}}A^{\beta}_{mn\bm{k}}\\ &~{}~{}~{}~{}~{}\times\frac{\frac{1}{2}[(-\frac{df_{n\bm{k}}}{d\varepsilon})\Lambda^{\beta}_{mn\bm{k}}(\varepsilon_{n\bm{k}})+(-\frac{df_{m\bm{k}}}{d\varepsilon})\Lambda^{\beta}_{mn\bm{k}}(\varepsilon_{m\bm{k}})]}{{i(\Sigma^{R}_{m\bm{k}}-\Sigma^{A}_{n\bm{k}})+i(\varepsilon_{m\bm{k}}-\varepsilon_{n\bm{k}})}}.\end{split} (29)

For the velocity operator, this term enables studies of charge transport in the presence of inter-band coherence [42]; for the spin operator, this contribution is essential to describe how ee-ph interactions modify spin precession.

II.6 Renormalized relaxation time

We derive an expression for the renormalized macroscopic relaxation time of an operator 𝐀^\mathbf{\hat{A}} due to ee-ph interactions. Using Eq. (28), the average relaxation time for the band-diagonal components Ann𝒌αA^{\alpha}_{nn\bm{k}} is

ταβ=n𝒌Ann𝒌αAnn𝒌βτn𝒌e-phΛnn𝒌β(εn𝒌)(dfn𝒌dε)n𝒌Ann𝒌αAnn𝒌β(dfn𝒌dε).\begin{split}\tau_{\alpha\beta}=\frac{\sum_{n\bm{k}}A^{\alpha}_{nn\bm{k}}A^{\beta}_{nn\bm{k}}\tau^{\text{e-ph}}_{n\bm{k}}\Lambda^{\beta}_{nn\bm{k}}(\varepsilon_{n\bm{k}})(-\frac{df_{n\bm{k}}}{d\varepsilon})}{\sum_{n\bm{k}}A^{\alpha}_{nn\bm{k}}A^{\beta}_{nn\bm{k}}\big{(}-\frac{df_{n\bm{k}}}{d\varepsilon}\big{)}}.\end{split} (30)

For the velocity operator, this equation gives the well-known Drude dc electrical conductivity, while for the spin operator one obtains the phonon-dressed macroscopic spin relaxation time, as discussed below. These results generalize the linear response treatment for band-diagonal operators presented in Ref. [4] and extend it to non-diagonal operators.

III Spin relaxation and decoherence

We now specialize to the non-diagonal spin operator, and apply our formalism to study phonon-induced spin relaxation and decoherence. The BSE for the phonon-dressed spin vertex - called hereafter spin-phonon BSE - is a key result obtained from Eq. (13) by replacing Ann𝒌αA^{\alpha}_{nn^{\prime}\bm{k}} with the spin operator snn𝒌αs^{\alpha}_{nn^{\prime}\bm{k}}. In matrix form and using a compact notation, the spin-phonon BSE can be written in a way that clearly matches the diagram in Fig. 1(c):

𝒔𝚲𝒌(ε)=𝒔𝒌+1NqVucν𝒒±gν𝒌𝒒[GA𝒔𝚲GR]𝒌+𝒒,ε±ων𝒒gν𝒌𝒒F±(T),\begin{split}\bm{s}\bm{\Lambda}_{\bm{k}}(\varepsilon)\!=\!\bm{s}_{\bm{k}}+\frac{1}{N_{q}V_{\text{uc}}}\!\sum_{\nu\bm{q}\pm}&\textbf{g}_{\nu\bm{k}\bm{q}}^{\dagger}\!\left[{G}^{A}\bm{s}\bm{\Lambda}{G}^{R}\right]_{\!\!\begin{smallmatrix}\bm{k}+\bm{q},~{}~{}\\ \varepsilon\pm\omega_{\nu\bm{q}}\end{smallmatrix}}\!\!\!\textbf{g}_{\nu\bm{k}\bm{q}}\,F_{\pm}(T),\end{split} (31)

where 𝒔𝚲𝒌(ε)=𝒔nn𝒌𝚲nn𝒌(ε)\bm{s}\bm{\Lambda}_{\bm{k}}(\varepsilon)=\bm{s}_{nn^{\prime}\bm{k}}\bm{\Lambda}_{nn^{\prime}\bm{k}}(\varepsilon) is the phonon-dressed spin vertex, F±(T)=Nν𝒒+12±[f(ε±ων𝒒)12]F_{\pm}(T)=N_{\nu\bm{q}}+\frac{1}{2}\pm[f(\varepsilon\pm\omega_{\nu\bm{q}})-\frac{1}{2}] is a thermal occupation factor at temperature TT, and [𝐠ν𝒌𝒒]nm=gnmν(𝒌,𝒒)\left[\mathbf{g}_{\nu\bm{k}\bm{q}}\right]_{nm}=g_{nm\nu}(\bm{k},\bm{q}) are ee-ph matrix elements [31].
In the weak scattering regime, this spin-phonon BSE can be rewritten using the double-pole ansatz discussed above (where ε\varepsilon equals εn𝒌\varepsilon_{n\bm{k}} or εn𝒌\varepsilon_{n^{\prime}\bm{k}}):

snn𝒌αΛnn𝒌α(ε)=snn𝒌α+2πNqVucmm𝒒ν[gnmν(𝒌,𝒒)]gnmν(𝒌,𝒒)×12[{(Nν𝒒+fm𝒌+𝒒)(δ(ε+ων𝒒εm𝒌+𝒒)iπP1ε+ων𝒒εm𝒌+𝒒)+(Nν𝒒+1fm𝒌+𝒒)(δ(εων𝒒εm𝒌+𝒒)iπP1εων𝒒εm𝒌+𝒒)}×smm𝒌+𝒒αΛmm𝒌+𝒒α(εm𝒌+𝒒)i(Σm𝒌+𝒒RΣm𝒌+𝒒A)+i(εm𝒌+𝒒εm𝒌+𝒒)+{(Nν𝒒+fm𝒌+𝒒)(δ(ε+ων𝒒εm𝒌+𝒒)+iπP1ε+ων𝒒εm𝒌+𝒒)+(Nν𝒒+1fm𝒌+𝒒)(δ(εων𝒒εm𝒌+𝒒)+iπP1εων𝒒εm𝒌+𝒒)}×smm𝒌+𝒒αΛmm𝒌+𝒒α(εm𝒌+𝒒)i(Σm𝒌+𝒒RΣm𝒌+𝒒A)+i(εm𝒌+𝒒εm𝒌+𝒒)].\begin{split}&s^{\alpha}_{nn^{\prime}\bm{k}}\Lambda^{\alpha}_{nn^{\prime}\bm{k}}(\varepsilon)=s^{\alpha}_{nn^{\prime}\bm{k}}+\frac{2\pi}{N_{q}V_{\text{uc}}}\sum_{mm^{\prime}\bm{q}\nu}\left[g_{n^{\prime}m^{\prime}\nu}(\bm{k},\bm{q})\right]^{*}g_{nm\nu}(\bm{k},\bm{q})\\ &\times\frac{1}{2}\bigg{[}\{(N_{\nu\bm{q}}+f_{m\bm{k}+\bm{q}})(\delta(\varepsilon+\omega_{\nu\bm{q}}-\varepsilon_{m\bm{k}+\bm{q}})-\frac{i}{\pi}P\frac{1}{\varepsilon+\omega_{\nu\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}}})\\ &+(N_{\nu\bm{q}}+1-f_{m\bm{k}+\bm{q}})(\delta(\varepsilon-\omega_{\nu\bm{q}}-\varepsilon_{m\bm{k}+\bm{q}})-\frac{i}{\pi}P\frac{1}{\varepsilon-\omega_{\nu\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}}})\}\times\frac{s^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}\Lambda^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}(\varepsilon_{m\bm{k}+\bm{q}})}{i(\Sigma^{R}_{m\bm{k}+\bm{q}}-\Sigma^{A}_{m^{\prime}\bm{k}+\bm{q}})+i(\varepsilon_{m\bm{k}+\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}})}\\ &+\{(N_{\nu\bm{q}}+f_{m^{\prime}\bm{k}+\bm{q}})(\delta(\varepsilon+\omega_{\nu\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}})+\frac{i}{\pi}P\frac{1}{\varepsilon+\omega_{\nu\bm{q}}-\varepsilon_{m\bm{k}+\bm{q}}})\\ &+(N_{\nu\bm{q}}+1-f_{m^{\prime}\bm{k}+\bm{q}})(\delta(\varepsilon-\omega_{\nu\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}})+\frac{i}{\pi}P\frac{1}{\varepsilon-\omega_{\nu\bm{q}}-\varepsilon_{m\bm{k}+\bm{q}}})\}\times\frac{s^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}\Lambda^{\alpha}_{mm^{\prime}\bm{k}+\bm{q}}(\varepsilon_{m^{\prime}\bm{k}+\bm{q}})}{i(\Sigma^{R}_{m\bm{k}+\bm{q}}-\Sigma^{A}_{m^{\prime}\bm{k}+\bm{q}})+i(\varepsilon_{m\bm{k}+\bm{q}}-\varepsilon_{m^{\prime}\bm{k}+\bm{q}})}\bigg{]}.\end{split} (32)

This BSE for the phonon-dressed spin vertex, used in this work to study spin dynamics, should not be confused with the widely used BSE for excitons and optical spectra [2], which is entirely unrelated.
The vertex corrections Λnn𝒌α\Lambda^{\alpha}_{nn^{\prime}\bm{k}} obtained by solving the BSE govern spin dynamics as they renormalize spin relaxation and precession [29]. The macroscopic spin relaxation times are obtained using the thermal average in Eq. (30),

ταβ(s)=n𝒌snn𝒌αsnn𝒌βτn𝒌e-phΛnn𝒌β(εn𝒌)(dfn𝒌dε)n𝒌snn𝒌αsnn𝒌β(dfn𝒌dε).\tau^{(s)}_{\alpha\beta}=\frac{\sum_{n\bm{k}}s^{\alpha}_{nn\bm{k}}s^{\beta}_{nn\bm{k}}\tau^{\text{e-ph}}_{n\bm{k}}\Lambda^{\beta}_{nn\bm{k}}(\varepsilon_{n\bm{k}})(-\frac{df_{n\bm{k}}}{d\varepsilon})}{\sum_{n\bm{k}}s^{\alpha}_{nn\bm{k}}s^{\beta}_{nn\bm{k}}\,\big{(}\!-\frac{df_{n\bm{k}}}{d\varepsilon}\big{)}}. (33)

For α=β\alpha=\beta along the external magnetic field, Eq. (33) gives the longitudinal spin relaxation time, usually called T1T_{1}, along the direction α\alpha, while for a perpendicular magnetic field one obtains the transverse spin relaxation time, T2T_{2} [43]. The renormalized microscopic spin relaxation times (τnn𝒌α\tau_{nn^{\prime}\bm{k}}^{\alpha}) and spin precession rates (ωnn𝒌α\omega_{nn^{\prime}\bm{k}}^{\alpha}), which are matrices in Bloch basis, are computed from the vertex corrections Λnn𝒌α\Lambda^{\alpha}_{nn^{\prime}\bm{k}} using

11τnn𝒌α(ε)+iωnn𝒌α(ε)Λnn𝒌α(ε)i(Σn𝒌RΣn𝒌A)+i(εn𝒌εn𝒌).\frac{1}{\frac{1}{\tau_{nn^{\prime}\bm{k}}^{\alpha}(\varepsilon)}+i{\omega}_{nn^{\prime}\bm{k}}^{\alpha}(\varepsilon)}\equiv\frac{\Lambda^{\alpha}_{nn^{\prime}\bm{k}}(\varepsilon)}{i(\Sigma^{R}_{n\bm{k}}-\Sigma^{A}_{n^{\prime}\bm{k}})+i(\varepsilon_{n\bm{k}}-\varepsilon_{n^{\prime}\bm{k}})}. (34)

The diagonal components with n=nn\!=\!n^{\prime} give the renormalized microscopic spin relaxation times, τnn𝒌β=τn𝒌e-phΛnn𝒌β(εn𝒌)\tau^{\beta}_{nn\bm{k}}=\tau^{\text{e-ph}}_{n\bm{k}}\,\Lambda^{\beta}_{nn\bm{k}}(\varepsilon_{n\bm{k}}), entering Eq. (33).

IV Results

We apply our formalism to study spin relaxation and decoherence. We first present analytic results for a two-level model system, and then focus on first-principles calculations on a real material, GaAs. Application to a wider range of materials is presented in our companion paper [29].

IV.1 Two-level system with optical phonon scattering

We study spin dynamics in a two-level system to understand different phonon-induced spin relaxation mechanisms. In our model, the electron spins undergo phonon-induced spin-flip transitions together with spin precession in the SOC field modified by the ee-ph interactions. We solve the spin-phonon BSE for this system and derive analytic expressions for the vertex corrections and spin relaxation times. Our analysis sheds light on phonon-dressed operators and their renormalized dynamics, providing a starting point to understand phonon-induced spin relaxation in real materials with complex band structures, phonon dispersions, and ee-ph interactions.
Consider a periodic two-level system where each level is spin degenerate (see Fig. 2). The Hilbert space consists of four Bloch states, which are eigenstates of the Hamiltonian:

H|n=εn|nH\!\ket{n}=\varepsilon_{n}\!\ket{n} (35)

where |n\ket{n} is the nn-th energy eigenstate. The two lowest-energy states |1\ket{1} and |2\ket{2} are degenerate (ε1=ε2\varepsilon_{1}=\varepsilon_{2}) and differ only in their spin part. The other two states, |3\ket{3} and |4\ket{4}, are higher in energy by ωO\omega_{O} and are perturbed by an internal magnetic field along x^\hat{x} due to SOC. This field causes a small Zeeman splitting, ΔωO\Delta\ll\omega_{O}, such that ε3,4=ε1+ωO±Δ2\varepsilon_{3,4}=\varepsilon_{1}+\omega_{O}\pm\frac{\Delta}{2}.
We separate the space-dependent part |ψn\ket{\psi_{n}} and the spin-dependent part |χn\ket{\chi_{n}} of the two eigenstates as |n=|ψn|χn\ket{n}=\ket{\psi_{n}}\otimes\ket{\chi_{n}}. The two lowest states have an identical space-dependent part |ψ\ket{\psi} and are spin polarized along z^\hat{z}:

|1=|ψ(10),|2=|ψ(01),\ket{1}=\ket{\psi}\otimes\left(\begin{matrix}1\\ 0\end{matrix}\right),~{}\ket{2}=\ket{\psi}\otimes\left(\begin{matrix}0\\ 1\end{matrix}\right), (36)

with the following spin matrix elements along zz:

1|s^z|1=2|s^z|2=12,1|s^z|2=0.\begin{split}&\bra{1}\hat{s}_{z}\ket{1}=-\bra{2}\hat{s}_{z}\ket{2}=\frac{1}{2},\\ &\bra{1}\hat{s}_{z}\ket{2}=0.\end{split} (37)

Above, s^=σ^/2\hat{s}=\hat{\sigma}/2 is the spin operator and σ^\hat{\sigma} are Pauli matrices. The two upper bands have an identical space-dependent part |ϕ\ket{\phi} and are spin polarized along x^\hat{x}:

|3=|ϕ12(11),|4=|ϕ12(11),\ket{3}=\ket{\phi}\otimes\frac{1}{\sqrt{2}}\left(\begin{matrix}1\\ 1\end{matrix}\right),~{}\ket{4}=\ket{\phi}\otimes\frac{1}{\sqrt{2}}\left(\begin{matrix}1\\ -1\end{matrix}\right), (38)

with spin matrix elements

3|s^z|3=4|s^z|4=0,3|s^z|4=12.\begin{split}&\bra{3}\hat{s}_{z}\ket{3}=\bra{4}\hat{s}_{z}\ket{4}=0,\\ &\bra{3}\hat{s}_{z}\ket{4}=\frac{1}{2}.\end{split} (39)

The space-dependent part of the two lower states is orthogonal to that of the upper states, and the spin matrix elements between the two sets of states are zero.
In our model, an electron can scatter between the lower and upper levels by emitting or absorbing an optical phonon. These transitions are associated with ee-ph matrix elements gnm=m|ΔV^|ng_{nm}=\bra{m}\Delta\hat{V}\ket{n}, where ΔV^\Delta\hat{V} is the perturbation potential due to the optical phonon. We assume that this perturbation potential has the form

ΔV^=ΔV^(𝒓)(abba),\Delta\hat{V}=\Delta\hat{V}(\bm{r})\otimes\left(\begin{matrix}a&b\\ b&a\end{matrix}\right), (40)

where V^(𝒓)\hat{V}(\bm{r}) is the space-dependent part and (abba)\left(\begin{smallmatrix}a&b\\ b&a\end{smallmatrix}\right) the spin-dependent part of the perturbation, with aa and bb real numbers. Due to the presence of SOC, the spin-dependent part (abba)\left(\begin{smallmatrix}a&b\\ b&a\end{smallmatrix}\right) is different from the identity matrix. We consider a small spin-mixing bb, where a2+b2=1a^{2}+b^{2}=1, so that each phonon collision has a small probability b21b^{2}\ll 1 to flip the zz-component of the spin  [25]. The ee-ph matrix elements become gnm=g0χm|(abba)|χng_{nm}=g_{0}\bra{\chi_{m}}\left(\begin{smallmatrix}a&b\\ b&a\end{smallmatrix}\right)\ket{\chi_{n}}, where g0=ϕ|ΔV^(𝒓)|ψg_{0}=\bra{\phi}\Delta\hat{V}(\bm{r})\ket{\psi}. Thus the spin-dependent ee-ph matrix elements are

g13=g23=12g0(a+b),g14=g24=12g0(ab).\begin{split}g_{13}=g_{23}=\frac{1}{\sqrt{2}}g_{0}(a+b),\\ g_{14}=-g_{24}=\frac{1}{\sqrt{2}}g_{0}(a-b).\end{split} (41)
Refer to caption
Figure 2: Schematic of the energy levels, scattering rates, and spin orientations of the four Bloch states in our model.

IV.2 Spin-phonon BSE and spin relaxation times

We construct the spin-phonon BSE for our two-level system. There are four non-zero spin matrix elements in our model: s11zs^{z}_{11}, s22zs^{z}_{22}, s34zs^{z}_{34}, and s43zs^{z}_{43}, where snmz=m|s^z|ns^{z}_{nm}=\bra{m}\hat{s}_{z}\ket{n}. Using Eq. (32), the diagonal matrix elements have only one vertex correction, while the off-diagonal matrix elements have two energy-dependent vertex corrections. Assuming a small Zeeman splitting Δ\Delta, we can neglect the energy dependence of the off-diagonal vertex corrections, and define Λnmz=Λnmz(εn)\Lambda^{z}_{nm}=\Lambda^{z}_{nm}(\varepsilon_{n}). We thus have four unknown vertex corrections: Λ11z\Lambda^{z}_{11}, Λ22z\Lambda^{z}_{22}, Λ34z\Lambda^{z}_{34}, and Λ43z\Lambda^{z}_{43}. Taking the ee-ph scattering rates of the two lower and upper states to be Γ1\Gamma_{1} and Γ3\Gamma_{3}, respectively, the spin-phonon BSE becomes

Λ11z=1+(12b2)(Γ1Λ34ziΔ+Γ3+Γ1Λ43ziΔ+Γ3)Λ22z=1(12b2)(Γ1Λ34ziΔ+Γ3+Γ1Λ43ziΔ+Γ3)Λ34z=1+(12b2)(Γ3Λ11zΓ1+Γ3Λ22zΓ1)Λ43z=1+(12b2)(Γ3Λ11zΓ1+Γ3Λ22zΓ1).\begin{split}&\Lambda^{z}_{11}=1+\left(\frac{1}{2}-b^{2}\right)\left(\frac{\Gamma_{1}\Lambda^{z}_{34}}{-i\Delta+\Gamma_{3}}+\frac{\Gamma_{1}\Lambda^{z}_{43}}{i\Delta+\Gamma_{3}}\right)\\ -&\Lambda^{z}_{22}=-1-\left(\frac{1}{2}-b^{2}\right)\left(\frac{\Gamma_{1}\Lambda^{z}_{34}}{-i\Delta+\Gamma_{3}}+\frac{\Gamma_{1}\Lambda^{z}_{43}}{i\Delta+\Gamma_{3}}\right)\\ &\Lambda^{z}_{34}=1+\left(\frac{1}{2}-b^{2}\right)\left(\frac{\Gamma_{3}\Lambda^{z}_{11}}{\Gamma_{1}}+\frac{\Gamma_{3}\Lambda^{z}_{22}}{\Gamma_{1}}\right)\\ &\Lambda^{z}_{43}=1+\left(\frac{1}{2}-b^{2}\right)\left(\frac{\Gamma_{3}\Lambda^{z}_{11}}{\Gamma_{1}}+\frac{\Gamma_{3}\Lambda^{z}_{22}}{\Gamma_{1}}\right).\end{split} (42)

Note that this equation treats the phonon-induced spin flips and the spin precessional dynamics self-consistently.
The first two lines in Eq. (42) give Λ11z=Λ22z\Lambda^{z}_{11}=\Lambda^{z}_{22}, while from the third and fourth lines we obtain Λ34z=Λ43z\Lambda^{z}_{34}=\Lambda^{z}_{43}. Therefore, the solution of the spin-phonon BSE is

Λ11z=Λ22z=Γ32+Δ2+Γ3Γ1(12b2)Γ324b2(1b2)+Δ2Λ34z=Λ43z=(Γ32+Δ2)(Γ1+Γ3(12b2))(Γ324b2(1b2)+Δ2)Γ1.\begin{split}&\Lambda^{z}_{11}=\Lambda^{z}_{22}=\frac{\Gamma_{3}^{2}+\Delta^{2}+\Gamma_{3}\Gamma_{1}(1-2b^{2})}{\Gamma_{3}^{2}4b^{2}(1-b^{2})+\Delta^{2}}\\ &\Lambda^{z}_{34}=\Lambda^{z}_{43}=\frac{(\Gamma_{3}^{2}+\Delta^{2})(\Gamma_{1}+\Gamma_{3}(1-2b^{2}))}{(\Gamma_{3}^{2}4b^{2}(1-b^{2})+\Delta^{2})\Gamma_{1}}.\end{split} (43)

The state-dependent spin relaxation times τ11z\tau^{z}_{11}, τ22z\tau^{z}_{22}, τ34z\tau^{z}_{34}, and τ43z\tau^{z}_{43} are obtained using Eq. (20):

τ11z=τ22z=Γ32+Δ2+Γ3Γ1(12b2)(Γ324b2(1b2)+Δ2)Γ1τ34z=τ43z=(Γ32+Δ2)(Γ1+Γ3(12b2))(Γ324b2(1b2)+Δ2)Γ1Γ3.\begin{split}&\tau^{z}_{11}=\tau^{z}_{22}=\frac{\Gamma_{3}^{2}+\Delta^{2}+\Gamma_{3}\Gamma_{1}(1-2b^{2})}{(\Gamma_{3}^{2}4b^{2}(1-b^{2})+\Delta^{2})\Gamma_{1}}\\ &\tau^{z}_{34}=\tau^{z}_{43}=\frac{(\Gamma_{3}^{2}+\Delta^{2})(\Gamma_{1}+\Gamma_{3}(1-2b^{2}))}{(\Gamma_{3}^{2}4b^{2}(1-b^{2})+\Delta^{2})\Gamma_{1}\Gamma_{3}}.\end{split} (44)

Figure 3(a) shows the vertex correction for the lowest state, Λ11z\Lambda_{11}^{z} in Eq. (43), and Fig. 3(b) shows the spin relaxation time τ11z{\tau_{11}^{z}} from Eq. (44), both plotted as a function of the ee-ph collision time τ1=1/Γ1\tau_{1}=1/\Gamma_{1}. These quantities show three distinct regimes with a qualitatively different dependence on the ee-ph collision time, which correspond to the EY, DP and strong-precession regimes. Our formalism encompasses these three regimes [29] because it can capture both spin-flip scattering and spin precession. The physics of these regimes is discussed below.

Refer to caption
Figure 3: (a) The vertex correction Λ11z\Lambda^{z}_{11} as a function of the ee-ph collision time τ1\tau_{1} for the model system in Fig. 2. (b) Spin relaxation time as a function of the ee-ph collision time τ1\tau_{1} for our model system. We show the full solution of the spin-phonon BSE in Eqs. (43)-(44)) (black curve) and approximate results for the Elliott-Yafet (blue, Eqs. (45)-(46)), Dyakonov-Perel (red, Eqs. (49)-(50)), and strong-precession regimes (green, Eqs. (51)-(52)). These results are obtained by setting τ3=τ1\tau_{3}=\tau_{1} and b=0.02b=0.02.

IV.3 Elliott-Yafet regime

We first focus on the EY regime, where spin relaxation occurs primarily through spin-flip transitions. In this regime, the spin-flip probability is small since b1b\ll 1, and the spin precession rate is much smaller than the scattering rate, Δ2Γ3b\Delta\ll 2\Gamma_{3}b. Using these conditions in Eq. (43), we obtain the following vertex corrections in the EY regime:

Λ11z=Λ22z=τ1+τ34b21τ1Λ34z=Λ43z=τ1+τ34b21τ3,\begin{split}\Lambda^{z}_{11}=\Lambda^{z}_{22}=\frac{\tau_{1}+\tau_{3}}{4b^{2}}\frac{1}{\tau_{1}}\\ \Lambda^{z}_{34}=\Lambda^{z}_{43}=\frac{\tau_{1}+\tau_{3}}{4b^{2}}\frac{1}{\tau_{3}},\end{split} (45)

where τ1=1/Γ1\tau_{1}=1/\Gamma_{1} and τ3=1/Γ3\tau_{3}=1/\Gamma_{3} are state-dependent ee-ph collision times. These vertex corrections determine the spin relaxation times via Eq. (34):

τ11z=τ22z=τ34z=τ43z=τ1+τ34b2.\tau^{z}_{11}=\tau^{z}_{22}=\tau^{z}_{34}=\tau^{z}_{43}=\frac{\tau_{1}+\tau_{3}}{4b^{2}}. (46)

These results, shown in Fig. 3, approximate well the solution of the spin-phonon BSE in the EY regime.
In the conventional theory of EY spin relaxation, the spin relaxation times are proportional to 1/b21/b^{2} and to the ee-ph collision times (here, τ1\tau_{1} and τ3\tau_{3}[28, 44, 45]. Our results in Eq. (46) are consistent with that trend, although the spin relaxation times are proportional to the average of the ee-ph collision times of the two levels, (τ1+τ3)/2(\tau_{1}+\tau_{3})/2, and not to their individual values. This difference is a result of considering both forward- and back-scattering processes between the two electronic levels in the spin-phonon BSE [4], different from the simpler sRTA [28, 44, 45] which neglects electron back-scattering.

IV.4 Dyakonov-Perel regime

Next, we discuss the regime where spin precession is important and governs spin relaxation. Here, the spin-flip probability is still small (b1b\ll 1) but the internal magnetic field due to SOC is significant, such that Δ2Γ3b\Delta\gg 2\Gamma_{3}b. Inserting these conditions in Eqs. (43)-(44), the vertex corrections become

Λ11z=Λ22z=Γ32+Γ3Γ1+Δ2Δ2,Λ34z=Λ43z=(Γ32+Δ2)(Γ1+Γ3)Δ2Γ1,\begin{split}&\Lambda^{z}_{11}=\Lambda^{z}_{22}=\frac{\Gamma_{3}^{2}+\Gamma_{3}\Gamma_{1}+\Delta^{2}}{\Delta^{2}},\\ &\Lambda^{z}_{34}=\Lambda^{z}_{43}=\frac{(\Gamma_{3}^{2}+\Delta^{2})(\Gamma_{1}+\Gamma_{3})}{\Delta^{2}\Gamma_{1}},\end{split} (47)

and for the spin relaxation times we obtain

τ11z=τ22z=Γ32+Γ3Γ1+Δ2Δ2Γ1,τ34z=τ43z=(Γ32+Δ2)(Γ1+Γ3)Δ2Γ1Γ3.\begin{split}&\tau^{z}_{11}=\tau^{z}_{22}=\frac{\Gamma_{3}^{2}+\Gamma_{3}\Gamma_{1}+\Delta^{2}}{\Delta^{2}\Gamma_{1}},\\ &\tau^{z}_{34}=\tau^{z}_{43}=\frac{(\Gamma_{3}^{2}+\Delta^{2})(\Gamma_{1}+\Gamma_{3})}{\Delta^{2}\Gamma_{1}\Gamma_{3}}.\end{split} (48)

These results describe spin relaxation governed by spin precession and renormalized by the ee-ph interactions.
In the DP regime, the ee-ph scattering rates are much greater than the bare spin precession rate Δ\Delta, and thus Γ1,3Δ\Gamma_{1,3}\gg\Delta. The vertex corrections in the DP regime become

Λ11z=Λ22z=Γ32+Γ3Γ1Δ2Γ1Γ1Λ34z=Λ43z=Γ32+Γ3Γ1Δ2Γ1Γ3,\begin{split}&\Lambda^{z}_{11}=\Lambda^{z}_{22}=\frac{\Gamma_{3}^{2}+\Gamma_{3}\Gamma_{1}}{\Delta^{2}\Gamma_{1}}\Gamma_{1}\\ &\Lambda^{z}_{34}=\Lambda^{z}_{43}=\frac{\Gamma_{3}^{2}+\Gamma_{3}\Gamma_{1}}{\Delta^{2}\Gamma_{1}}\Gamma_{3},\end{split} (49)

while for the spin relaxation times we obtain

τ11z=τ22z=τ34z=τ43z=Γ32+Γ3Γ1Δ2Γ1.\tau^{z}_{11}=\tau^{z}_{22}=\tau^{z}_{34}=\tau^{z}_{43}=\frac{\Gamma_{3}^{2}+\Gamma_{3}\Gamma_{1}}{\Delta^{2}\Gamma_{1}}. (50)

These results, shown in Fig. 3, are an excellent approximation to the BSE solution in the DP regime.
A hallmark of DP relaxation is the inverse proportionality between the spin relaxation and ee-ph collision times [19], a trend captured by our treatment of the DP regime [see Eq. (50)]. Our formalism shows in Eq. (49) that this trend originates from the inverse-square scaling of the vertex corrections with the ee-ph collision times, ΛΓ1,321/τ1,32\Lambda\sim\Gamma_{1,3}^{2}\sim 1/\tau_{1,3}^{2}. Note that in the EY regime, rescaling the ee-ph collision times by a constant factor has no effect on the vertex corrections, which depend only on the ratio τ1/τ3\tau_{1}/\tau_{3}, and thus the EY spin relaxation times are proportional to the ee-ph collision times. The situation is different in the DP regime, where rescaling the ee-ph collision times changes the vertex corrections. These scaling trends can be employed in real materials to identify the dominant microscopic mechanisms for spin relaxation and decoherence [29].

IV.5 Strong-precession regime

A third, distinct regime is realized when the Zeeman splitting is much greater than the ee-ph scattering rates, ΔΓ1,3\Delta\gg\Gamma_{1,3}, such that the spins precess much faster than the rate of ee-ph collisions. In this strong-precession regime [43, 46, 47], the vertex corrections become

Λ11z=Λ22z=1,Λ34z=Λ43z=1+Γ3Γ1\begin{split}&\Lambda^{z}_{11}=\Lambda^{z}_{22}=1,\\ &\Lambda^{z}_{34}=\Lambda^{z}_{43}=1+\frac{\Gamma_{3}}{\Gamma_{1}}\end{split} (51)

and can be approximated as Λ1\Lambda\approx 1. Therefore, the vertex corrections are less important than in the EY or DP regimes. It follows that in the strong-precession regime the spin relaxation times are proportional to the ee-ph collision times:

τ11z=τ22z=1Γ1,τ34z=τ43z=1Γ1+1Γ3.\begin{split}&\tau^{z}_{11}=\tau^{z}_{22}=\frac{1}{\Gamma_{1}},\\ &\tau^{z}_{34}=\tau^{z}_{43}=\frac{1}{\Gamma_{1}}+\frac{1}{\Gamma_{3}}.\end{split} (52)

These results, shown in Fig. 3, approximate well the BSE solution in the strong-precession regime. Their physical interpretation is interesting: in the strong-precession regime, the spins precess for many full cycles between phonon collisions, randomizing the spin direction. Consequently, the spin relaxation times become equal to the ee-ph collision times, and thus the vertex corrections Λ1\Lambda\approx 1 can be neglected.

IV.6 Uncovering the three regimes in GaAs

To identify the three spin relaxation mechanisms in a real material, we study spin relaxation in GaAs using our first-principles implementation of the spin-phonon BSE [29]. We focus on spin relaxation for conduction band electrons in GaAs at room temperature (300 K), and investigate how the spin relaxation times depend on the ee-ph collision times.
We compute the ground state and band structures of GaAs with density functional theory (DFT), using a plane-wave basis in the Quantum ESPRESSO code [48] and employing the HSE06 hybrid functional [49] to obtain an accurate band gap and electronic structure. We use fully relativistic norm-conserving pseudopotentials, generated in the local-density approximation (LDA) with Pseudo Dojo [50], together with a kinetic energy cutoff of 72 Ry and a relaxed lattice constant of 5.60 Å. The phonon energies and perturbation potentials are computed using density functional perturbation theory (DFPT) [7]. Using our perturbo code [31], we compute the ee-ph matrix elements on coarse Brillouin zone grids with 8×8×88\times 8\times 8 𝒌\bm{k} and 𝒒\bm{q} points, following which we generate spinor Wannier functions with the Wannier90 code [51] and use them in perturbo [31], with a method we developed in Ref. [28], to jointly interpolate the ee-ph matrix elements and spin matrices. The long-range quadrupole ee-ph interactions [10, 52, 11, 53] are included to fully account for ee-ph coupling with long-wavelength phonons. We interpolate the ee-ph matrix elements to fine Brillouin zone grids with up to 200×200×200200\times 200\times 200 𝒌\bm{k} and 𝒒\bm{q} points, and a 10 meV Gaussian broadening for the energy-conserving delta functions in the ee-ph scattering rates [41]. The spin relaxation times in Eq. (30) are computed using the tetrahedron integration method [54], assuming a non-degenerate electron concentration of 1016cm310^{16}~{}\text{cm}^{-3}. The spin-phonon BSE in Eq. (32) is solved with an augmented iterative approach described in Ref. [29].

Refer to caption
Figure 4: First-principles calculations of (a) vertex corrections and (b) spin relaxation times for electron spins in GaAs, computed at room temperature and plotted as a function of the average ee-ph collision time. The arrows indicate the values for the real material, GaAs (black dot), while the other values are obtained by rescaling the ee-ph interaction strength, as explained in the text. The vertical dotted lines are placed at the inflection points of the spin relaxation times, and separate the EY, DP, and strong-precession regimes.

Using our first-principles spin-phonon BSE, we compute the spin relaxation time for electron spins in GaAs, and obtain a value 51 ps at 300 K in excellent agreement with the experimental value of 42 ps [55]. We also obtain an average ee-ph collision time of 410 fs, computing using

τe-ph=n𝒌τnn𝒌e-ph(εn𝒌)(dfn𝒌dε)n𝒌(dfn𝒌dε)\langle\tau^{\text{e-ph}}\rangle=\frac{\sum_{n\bm{k}}\tau^{\text{e-ph}}_{nn\bm{k}}(\varepsilon_{n\bm{k}})(-\frac{df_{n\bm{k}}}{d\varepsilon})}{\sum_{n\bm{k}}\,\big{(}\!-\frac{df_{n\bm{k}}}{d\varepsilon}\big{)}} (53)

and an average vertex correction Λz171\langle\Lambda^{z}\rangle\!\approx\!171, computed from

Λz=n𝒌|snn𝒌z|2Λnn𝒌z(εn𝒌)(dfn𝒌dε)n𝒌|snn𝒌z|2(dfn𝒌dε).\langle\Lambda^{z}\rangle=\frac{\sum_{n\bm{k}}\absolutevalue{s^{z}_{nn\bm{k}}}^{2}\Lambda^{z}_{nn\bm{k}}(\varepsilon_{n\bm{k}})(-\frac{df_{n\bm{k}}}{d\varepsilon})}{\sum_{n\bm{k}}\absolutevalue{s^{z}_{nn\bm{k}}}^{2}\,\big{(}\!-\frac{df_{n\bm{k}}}{d\varepsilon}\big{)}}. (54)

These “real-material” values for GaAs are shown with a dot in Fig. 4(a) and (b), where we also show spin relaxation times and vertex corrections obtained by artificially varying the average ee-ph collision time (by rescaling the ee-ph matrix elements). The resulting trends show that the spin relaxation times are inversely proportional to the ee-ph collision time, placing GaAs in the DP spin-relaxation regime.
Upon decreasing the average ee-ph collision time τe-ph\langle\tau^{\text{e-ph}}\rangle in Fig. 4(a), the system evolves from the DP to the EY regime, and the vertex correction first increases and then saturates to a maximal value of \sim1900 in the short ee-ph collision time limit. As a consequence, the spin relaxation time in Fig. 4(b) peaks at τe-ph0.09\langle\tau^{\text{e-ph}}\rangle\!\approx\!0.09 ps and then decreases linearly at shorter ee-ph collision times. This crossover from the DP to the EY mechanism, obtained here by artificially tuning the ee-ph collision time in GaAs, is consistent with the results from our model system.
Conversely, when the ee-ph collision time is increased, the system transitions from the DP to the strong-precession regime: Fig. 4(a) shows that the vertex correction first decreases as τe-ph2\langle\tau^{\text{e-ph}}\rangle^{-2} and then plateaus to a minimal value close to unity for long ee-ph collision times. Accordingly, the spin relaxation time in Fig. 4(b) reaches a minimum for τe-ph3\langle\tau^{\text{e-ph}}\rangle\approx 3 ps and then increases linearly at longer ee-ph collision times.
In summary, real materials exhibit the same spin relaxation mechanisms and vertex correction trends as our two-level system. These regimes for phonon-induced spin dynamics are summarized in Table 2. The presence of three distinct spin-relaxation regimes is general, and we expect it to be valid beyond the case of a simple semiconductor (GaAs) studied here.

Table 2: Summary of the Elliott-Yafet, Dyakonov-Perel, and strong-precession regimes.
EY DP Strong-precession
Spin-flip Spin precession Spin precession
Λ\Lambda 1/b2\sim{1}/{b^{2}} 1/(τe-ph)2\sim{1}/{(\tau^{\text{e-ph}})^{2}} 1\sim 1
τs\tau^{s} τe-ph/b2\sim{\tau^{\text{e-ph}}}/{b^{2}} 1/τe-ph\sim{1}/{\tau^{\text{e-ph}}} τe-ph\sim\tau^{\text{e-ph}}

V Discussion

Our formalism can capture all three of the EY, DP and strong-precession regimes in a unified framework. The reason can be inferred from the diagrammatic representation of the spin-phonon BSE in Fig. 1(c). In the diagrams, the EY relaxation is due to ee-ph scattering in the presence of spin mixing, which originates from the ee-ph interactions [gnmν(𝒌,𝒒)]\left[g_{n^{\prime}m^{\prime}\nu}(\bm{k},\bm{q})\right]^{*} and gnmν(𝒌,𝒒)g_{nm\nu}(\bm{k},\bm{q}) and the wiggly line in the kernel of the BSE. The DP and strong-precession mechanisms are included by virtue of the electron propagators with two different band indices, 𝒢ml𝒌+𝒒𝒢lm𝒌+𝒒\mathcal{G}_{ml\bm{k}+\bm{q}}\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}, placed between the wiggly line and the spin vertex. These propagators take into account spin precession (due to the SOC field) between ee-ph collisions. This elegant formalism captures a wide range of spin physics in a single diagram.
Although our discussion has focused on T1T_{1} spin relaxation times, the spin decoherence times T2T_{2} at finite magnetic fields can also be computed, as we plan to show in future work. Finally, the approach presented in this work for the phonon-dressed vertex is general and goes beyond spin relaxation. It can be employed to study the dynamics of any observable that couples with phonons, for which we also expect to find the three phonon-induced relaxation regimes discussed above for spin dynamics.

VI Conclusion

We have formulated a theory for the vertex corrections from ee-ph interactions to the susceptibility of a non-diagonal operator. The key result is a self-consistent BSE to calculate the phonon-dressed vertex, which encodes the dynamics of the operator coupling with phonons. When applied to spin, this approach enables quantitative calculations of spin relaxation and decoherence [29]. We have shown that our spin-phonon BSE captures both spin-flip transitions and spin precession, unifying the treatment of three spin decoherence mechanisms (EY, DP, and strong-precession) conventionally treated with separate heuristic models. By leveraging efficient workflows for first-principles ee-ph calculations [31], our method enables quantitative studies of spin relaxation and decoherence in a wide range of bulk and two-dimensional materials, as we show in the companion paper [29]. These advances open new avenues for understanding spin relaxation and decoherence in spintronics, magnetism, multiferroics, quantum materials and quantum technologies.

Acknowledgements.
This work was supported by the National Science Foundation under Grants No. DMR-1750613 and QII-TAQS 1936350, which provided for method development, and Grant No. OAC-2209262, which provided for code development. J.P. acknowledges support by the Korea Foundation for Advanced Studies.

Appendix A Ward identity

We derive a Ward identity for our BSE in Eq. (13). The Ward identity relates the vertex corrections with the electron self-energy [4, 1, 36], and guarantees that diagrams are taken into account consistently in the self-energy and in the BSE for the vertex [4, 1, 36].
For a system with Hamiltonian HH, the operator A^\hat{A} is related to the the negative derivative of the Hamiltonian with respect to the external field \mathcal{F}, A^=H\hat{A}=-\gradient_{\mathcal{F}}H. We compute the change of the Fan-Migdal self-energy in Eq. (2) with respect to \mathcal{F}, by taking the derivative

Σnn𝒌(iωa)=1βNqVucmmll𝒒ν,iqc[gnmν(𝒌,𝒒)]gnmν(𝒌,𝒒)×𝒟ν𝒒(iqc)(𝒢mm𝒌(iωa+iqc)).\begin{split}&{\gradient_{\mathcal{F}}\Sigma_{nn^{\prime}\bm{k}}(i\omega_{a})}\\ &~{}~{}=-\frac{1}{\beta N_{q}V_{\text{uc}}}\sum_{mm^{\prime}ll^{\prime}\bm{q}\nu,iq_{c}}\left[g_{n^{\prime}m^{\prime}\nu}(\bm{k},\bm{q})\right]^{*}g_{nm\nu}(\bm{k},\bm{q})\\ &~{}~{}~{}\times\mathcal{D}_{\nu\bm{q}}(iq_{c})\left(\gradient_{\mathcal{F}}\mathcal{G}_{mm^{\prime}\bm{k}}(i\omega_{a}+iq_{c})\right).\end{split} (55)

Employing the matrix identity

𝒢mm𝒌=ll𝒢ml𝒌[(𝒢1)]ll𝒌𝒢lm𝒌=ll𝒢ml𝒌(Hll𝒌Σll𝒌)𝒢lm𝒌,\begin{split}{\gradient_{\mathcal{F}}\mathcal{G}_{mm^{\prime}\bm{k}}}&=-\sum_{ll^{\prime}}\mathcal{G}_{ml\bm{k}}\left[{\gradient_{\mathcal{F}}(\mathcal{G}^{-1}})\right]_{ll^{\prime}\bm{k}}\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}}\\ &=-\sum_{ll^{\prime}}\mathcal{G}_{ml\bm{k}}\left(-\gradient_{\mathcal{F}}H_{ll^{\prime}\bm{k}}-\gradient_{\mathcal{F}}\Sigma_{ll^{\prime}\bm{k}}\right)\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}},\end{split} (56)

we obtain a self-consistent equation for the self-energy derivatives:

Σnn𝒌(iωa)=1βNqVucmmll𝒒ν,iqc[gnmν(𝒌,𝒒)]gnmν(𝒌,𝒒)×𝒟ν𝒒(iqc)𝒢ml𝒌+𝒒(iωa+iqc)𝒢lm𝒌+𝒒(iωa+iqc)×(Hll𝒌+Σll𝒌(iωa+iqc)).\begin{split}&{\gradient_{\mathcal{F}}\Sigma_{nn^{\prime}\bm{k}}(i\omega_{a})}\\ &~{}~{}=-\frac{1}{\beta N_{q}V_{\text{uc}}}\sum_{mm^{\prime}ll^{\prime}\bm{q}\nu,iq_{c}}\left[g_{n^{\prime}m^{\prime}\nu}(\bm{k},\bm{q})\right]^{*}g_{nm\nu}(\bm{k},\bm{q})\\ &~{}~{}~{}\times\mathcal{D}_{\nu\bm{q}}(iq_{c})\mathcal{G}_{ml\bm{k}+\bm{q}}(i\omega_{a}+iq_{c})\mathcal{G}_{l^{\prime}m^{\prime}\bm{k}+\bm{q}}(i\omega_{a}+iq_{c})\\ &~{}~{}~{}\times\left(\gradient_{\mathcal{F}}H_{ll^{\prime}\bm{k}}+\gradient_{\mathcal{F}}\Sigma_{ll^{\prime}\bm{k}}(i\omega_{a}+iq_{c})\right).\end{split} (57)

Comparing Eq. (57) and Eq. (8) in the iνb0i\nu_{b}\to 0 limit, we discover the Ward identity expressed in terms of the Hamiltonian, vertex corrections, and self-energy:

(H)𝚲(iωa,iωa)=H+Σ(iωa),(\gradient_{\mathcal{F}}H)\bm{\Lambda}(i\omega_{a},i\omega_{a})=\gradient_{\mathcal{F}}H+{\gradient_{\mathcal{F}}\Sigma(i\omega_{a})}, (58)

where (H)𝚲(iωa,iωa)=(Hnn𝒌α)Λnn𝒌α(iωa,iωa)(\gradient_{\mathcal{F}}H)\bm{\Lambda}(i\omega_{a},i\omega_{a})=(\frac{\partial H_{nn^{\prime}\bm{k}}}{{\mathcal{\partial F^{\alpha}}}}){\Lambda}^{\alpha}_{nn^{\prime}\bm{k}}(i\omega_{a},i\omega_{a}). Equation (58) can be equivalently expressed in terms of the operator matrix elements Ann𝒌αA^{\alpha}_{nn^{\prime}\bm{k}} ,

Ann𝒌αΛnn𝒌α(iωa,iωa)=Ann𝒌αΣnn𝒌(iωa)α.A^{\alpha}_{nn^{\prime}\bm{k}}\Lambda^{\alpha}_{nn^{\prime}\bm{k}}(i\omega_{a},i\omega_{a})=A^{\alpha}_{nn^{\prime}\bm{k}}-\frac{\partial\Sigma_{nn^{\prime}\bm{k}}(i\omega_{a})}{\partial\mathcal{F}^{\alpha}}. (59)

Our expression for the Ward identity in Eqs. (58)-(59) is consistent with the Ward identity for the velocity operator derived in Ref. [4]:

vnn𝒌αΛnn𝒌α(iωa,iωa)=vnn𝒌α+Σnn𝒌(iωa)kα,v^{\alpha}_{nn\bm{k}}\Lambda^{\alpha}_{nn\bm{k}}(i\omega_{a},i\omega_{a})=v^{\alpha}_{nn\bm{k}}+\frac{\partial\Sigma_{nn\bm{k}}(i\omega_{a})}{\partial k^{\alpha}}, (60)

as the velocity operator is defined as v^=H\hat{v}=\gradient_{\mathcal{F}}H with =𝒌\mathcal{F}=\bm{k}. This result further validates our BSE for the phonon-dressed vertex.

References