Introduction

Periodically driven systems enable the realization of various nonequilibrium quantum states and their control. Periodically driven systems are realized by a time-periodic field, and their properties in a nonequilibrium steady state can be described by the Floquet theory1,2,3,4,5, in which the effective Hamiltonian is independent of time. In fact, various theoretical predictions, such as the light-induced anomalous-Hall effect (AHE)6,7,8 and the Floquet time crystal9,10,11, are confirmed by experiments. Then, since the effective Hamiltonian of the Floquet theory depends on parameters of the driving field, its properties can be controlled by tuning the driving field. This is called Floquet engineering3,4,5. For example, it is possible to change the magnitude, sign, and bond anisotropy of exchange interactions of Mott insulators12,13,14,15,16. The Floquet engineering has been studied in many fields of physics, including condensed-matter, cold-atom, and optical physics.

Although there are many studies of the AHE of periodically driven systems, the Floquet engineering of the spin-Hall effect (SHE) is still lacking. The SHE is the key phenomenon in spintronics17,18,19,20,21. In the SHE, an electron spin current, a flow of the spin angular momentum, is generated by an electric field perpendicular to it22,23,24. This is the spin version of the AHE, in which an electron charge current is generated25,26. A significant difference between the AHE and SHE is about time-reversal symmetry (TRS): TRS is broken in the AHE, whereas it holds in the SHE. Since TRS can be broken by circularly polarized light27, there should be some essential differences between those of a periodically driven system. It is highly desirable to investigate the intrinsic SHE of a periodically driven multiorbital metal because the intrinsic SHE, the SHE intrinsic to the electronic structure, can be engineered by the driving field and several multiorbital metals, such as Pt, have the huge SHE28,29.

Here we show that in a multiorbital metal driven by circularly polarized light, the charge current generated by the AHE can be reversed by changing the helicity of light, whereas the spin current generated by the SHE remains unchanged. This is demonstrated by constructing a theory of pump-probe measurements of the AHE and SHE of a periodically driven t2g-orbital metal coupled to a heat bath and evaluating their conductivities numerially. This significant difference between the AHE and SHE results from the difference in TRS and thus should hold in many periodically driven systems. We also show that spin–orbit coupling (SOC) is vital for the SHE of the periodically driven multiorbital metal, whereas it is unnecessary for the AHE. This property is distinct from that of non-driven metals.

Results and discussion

Periodically driven t 2g-orbital metal

We consider a t2g-orbital metal coupled to a heat bath in the presence of a field A(t) (Fig. 1a):

$$H(t)={H}_{{{{{{{{\rm{s}}}}}}}}}(t)+{H}_{{{{{{{{\rm{sb}}}}}}}}}+{H}_{{{{{{{{\rm{b}}}}}}}}}.$$
(1)

(Note that in the t2g-orbital metal, such as Sr2RuO4, electrons occupy the t2g orbitals, i.e., the dyz, dzx, and dxy orbitals). First, Hs(t) is the system Hamiltonian, the Hamiltonian of t2g-orbital electrons with A(t),

$${H}_{{{{{{{{\rm{s}}}}}}}}}(t)=\mathop{\sum}\limits_{{{{{{{{\bf{k}}}}}}}}}\mathop{\sum}\limits_{a,b={d}_{yz},{d}_{zx},{d}_{xy}}\mathop{\sum}\limits_{\sigma ,{\sigma }^{{\prime} }=\uparrow ,\downarrow }{\bar{\epsilon }}_{ab}^{\sigma {\sigma }^{{\prime} }}({{{{{{{\bf{k}}}}}}}},t){c}_{{{{{{{{\bf{k}}}}}}}}a\sigma }^{{{{\dagger}}} }{c}_{{{{{{{{\bf{k}}}}}}}}b{\sigma }^{{\prime} }}.$$
(2)

Here \({c}_{{{{{{{{\bf{k}}}}}}}}a\sigma }^{{{{\dagger}}} }\) and ckaσ are the creation and annihilation operators, respectively, of an electron for orbital a with momentum k and spin σ, and

$${\bar{\epsilon }}_{ab}^{\sigma {\sigma }^{{\prime} }}({{{{{{{\bf{k}}}}}}}},t)=[{\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}},t)-\mu {\delta }_{a,b}]{\delta }_{\sigma ,{\sigma }^{{\prime} }}+{\xi }_{ab}^{\sigma {\sigma }^{{\prime} }},$$
(3)

where ϵab(k, t), μ, and \({\xi }_{ab}^{\sigma {\sigma }^{{\prime} }}\) are the kinetic energy with the Peierls phase factors due to A(t), the chemical potential, and SOC, respectively (see Methods). Throughout this paper, we use the unit  = 1, kB = 1, and alc = 1, where alc is the lattice constant. In addition to Hs(t), we have considered Hb and Hsb, the Hamiltonian of a Büttiker-type heat bath30,31,32,33 at temperature Tb and the system-bath coupling Hamiltonian (see Methods). This is because a nonequilibrium steady state can be realized due to the damping coming from the second-order perturbation of Hsb32,34.

Fig. 1: Set-up of the anomalous-Hall or spin-Hall effect and electronic properties of our model.
figure 1

a Set-up of the anomalous-Hall or spin-Hall effect for our model of Sr2RuO4 driven by circularly polarized light in the presence of the coupling to a heat bath. In Sr2RuO4, Ru ions form the square lattice; at each ion, four electrons occupy the Ru t2g orbitals (i.e., the dyz, dzx, and dxy orbitals). In the pump-probe measurements of the anomalous-Hall and spin-Hall effects, the probe field induces the charge and spin currents, respectively, perpendicular to it, and the pump field, a field of left- or right-circularly polarized light, periodically drives Sr2RuO4. The nonequilibrium steady state is realized because of the coupling to the heat bath. b The finite hopping processes of electrons in t2g orbitals on the square lattice. The dyz, dzx, and dxy represent these orbitals. t1, t2, and t3 are the nearest-neighbor hopping integrals, and t4 and t5 are the next nearest-neighbor ones. c The Fermi surface obtained for the non-driven case of our model in the quarter of the Brillouin zone. The other parts are reproducible by using the rotational symmetry.

The parameters of Hs(t) are chosen to reproduce the electronic structure of Sr2RuO435. The hopping integrals on the square lattice are parametrized by t1, t2, t3, t4, and t5 (Fig. 1b)36, and μ is determined from the condition ne = 4, where ne is the electron number per site; the value of μ is fixed at that determined in the non-driven case. We set (t1, t2, t3, t4, t5) = (0.675, 0.09, 0.45, 0.18, 0.03) (eV)36 and ξ = 0.17 eV37, where ξ is the coupling constant of SOC, in order that the Fermi surface (Fig. 1c) is consistent with that observed experimentally38.

Theory of pump-probe measurements of the SHE and AHE

The SHE and AHE of a periodically driven system are detectable by pump-probe measurements. In the pump-probe measurements39, a system is periodically driven by the pump field Apump(t), and its properties are analyzed by the probe field Aprob(t). Thus, we set A(t) = Apump(t) + Aprob(t) and treat the effects of Apump(t) in the Floquet theory and those of Aprob(t) in the linear-response theory33,40; in our analyses, Apump(t) is chosen to be

$${{{{{{{{\bf{A}}}}}}}}}_{{{{{{{{\rm{pump}}}}}}}}}(t) \, { = } \, ^{t}({A}_{0}\cos {{\Omega }}t\,{A}_{0}\sin ({{\Omega }}t+\delta )),$$
(4)

where Ω = 2π/T and T is the period of Apump(t). The anomalous-Hall and spin-Hall conductivities \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}(t,{t}^{{\prime} })\) and \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}(t,{t}^{{\prime} })\) are defined as

$${\sigma }_{yx}^{{{{{{{{\rm{Q}}}}}}}}}(t,{t}^{{\prime} })=\frac{1}{i\omega }\frac{\delta \langle {j}_{{{{{{{{\rm{Q}}}}}}}}}^{y}(t)\rangle }{\delta {A}_{{{{{{{{\rm{prob}}}}}}}}}^{x}({t}^{{\prime} })},$$
(5)

where \(\langle {j}_{{{{{{{{\rm{C}}}}}}}}}^{y}(t)\rangle\) and \(\langle {j}_{{{{{{{{\rm{S}}}}}}}}}^{y}(t)\rangle\) are the expectation values of the charge and spin current density operators, respectively. In our AHE or SHE, we have considered the charge or spin current, respectively, generated along the y-axis with the probe field applied along the x-axis (Fig. 1a). (Note that our SHE is different from the SHE of light, in which the helicity-dependent transverse shift of light at an interface is induced41,42,43). Then, the charge and spin current operators \({J}_{{{{{{{{\rm{C}}}}}}}}}^{y}(t)=N{j}_{{{{{{{{\rm{C}}}}}}}}}^{y}(t)\) and \({J}_{{{{{{{{\rm{S}}}}}}}}}^{y}(t)=N{j}_{{{{{{{{\rm{S}}}}}}}}}^{y}(t)\), where N is the number of sites, are determined from the continuity equations (see Methods)28,29,44,45:

$${J}_{{{{{{{{\rm{Q}}}}}}}}}^{y}(t)=\mathop{\sum}\limits_{{{{{{{{\bf{k}}}}}}}}}\mathop{\sum}\limits_{a,b}\mathop{\sum}\limits_{\sigma }{v}_{ab\sigma }^{({{{{{{{\rm{Q}}}}}}}})y}({{{{{{{\bf{k}}}}}}}},t){c}_{{{{{{{{\bf{k}}}}}}}}a\sigma }^{{{{\dagger}}} }(t){c}_{{{{{{{{\bf{k}}}}}}}}b\sigma }(t),$$
(6)

where \({v}_{ab\sigma }^{({{{{{{{\rm{C}}}}}}}})y}({{{{{{{\bf{k}}}}}}}},t)=(-e)\frac{\partial {\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}},t)}{\partial {k}_{y}}\), \({v}_{ab\sigma }^{({{{{{{{\rm{S}}}}}}}})y}({{{{{{{\bf{k}}}}}}}},t)=\frac{1}{2}{{{{{{{\rm{sgn}}}}}}}}(\sigma )\frac{\partial {\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}},t)}{\partial {k}_{y}}\), and \({{{{{{{\rm{sgn}}}}}}}}(\sigma )=1\) or − 1 for σ =  or , respectively. By combining Eq. (6) with Eq. (5) and using a method of Green’s functions34,44,46,47, we can express \({\sigma }_{yx}^{{{{{{{{\rm{Q}}}}}}}}}(t,{t}^{{\prime} })\) in terms of electron Green’s functions (see Methods).

To analyze the SHE and AHE in the nonequilibrium steady state, we consider the time-averaged dc anomalous-Hall and spin-Hall conductivities \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\),

$${\sigma }_{yx}^{{{{{{{{\rm{Q}}}}}}}}}=\mathop{\lim }\limits_{\omega \to 0}{{{{{{{\rm{Re}}}}}}}}\int\nolimits_{0}^{T}\frac{d{t}_{{{{{{{{\rm{av}}}}}}}}}}{T}\int\nolimits_{-\infty }^{\infty }d{t}_{{{{{{{{\rm{rel}}}}}}}}}{e}^{i\omega {t}_{{{{{{{{\rm{rel}}}}}}}}}}{\sigma }_{yx}^{{{{{{{{\rm{Q}}}}}}}}}(t,{t}^{{\prime} }),$$
(7)

where \({t}_{{{{{{{{\rm{rel}}}}}}}}}=t-{t}^{{\prime} }\) and \({t}_{{{{{{{{\rm{av}}}}}}}}}=(t+{t}^{{\prime} })/2\)33. Since we can calculate Eq. (7) in a way similar to that for charge transport of single-orbital systems32,33,40, we present the final result here (for the derivation, see Supplementary Note 1):

$${\sigma }_{yx}^{{{{{{{{\rm{Q}}}}}}}}}= \frac{1}{N}\mathop{\sum}\limits_{{{{{{{{\bf{k}}}}}}}}}\mathop{\sum}\limits_{a,b,c,d}\mathop{\sum}\limits_{\sigma ,{\sigma }^{{\prime} }}\int\nolimits_{-{{\Omega }}/2}^{{{\Omega }}/2}\frac{d{\omega }^{{\prime} }}{2\pi }\mathop{\sum }\limits_{m,l,n,q=-\infty }^{\infty }{[{v}_{ab\sigma }^{({{{{{{{\rm{Q}}}}}}}})y}({{{{{{{\bf{k}}}}}}}})]}_{ml} \\ \times {[{v}_{cd{\sigma }^{{\prime} }}^{({{{{{{{\rm{C}}}}}}}})x}({{{{{{{\bf{k}}}}}}}})]}_{nq}\left\{\frac{\partial {[{G}_{b\sigma c{\sigma }^{{\prime} }}^{{{{{{{{\rm{R}}}}}}}}}({{{{{{{\bf{k}}}}}}}},{\omega }^{{\prime} })]}_{ln}}{\partial {\omega }^{{\prime} }}{[{G}_{d{\sigma }^{{\prime} }a\sigma }^{ < }({{{{{{{\bf{k}}}}}}}},{\omega }^{{\prime} })]}_{qm}\right. \\ \left.-{[{G}_{b\sigma c{\sigma }^{{\prime} }}^{ < }({{{{{{{\bf{k}}}}}}}},{\omega }^{{\prime} })]}_{ln}\frac{\partial {[{G}_{d{\sigma }^{{\prime} }a\sigma }^{{{{{{{{\rm{A}}}}}}}}}({{{{{{{\bf{k}}}}}}}},{\omega }^{{\prime} })]}_{qm}}{\partial {\omega }^{{\prime} }}\right\},$$
(8)

where \({[{v}_{ab\sigma }^{({{{{{{{\rm{Q}}}}}}}})\nu }({{{{{{{\bf{k}}}}}}}})]}_{mn}\) (Q = C or S, ν = y or x) and \({[{G}_{a\sigma b{\sigma }^{{\prime} }}^{r}({{{{{{{\bf{k}}}}}}}},{\omega }^{{\prime} })]}_{mn}\) (r = R, A, or < ) are given by

$${[{v}_{ab\sigma }^{({{{{{{{\rm{Q}}}}}}}})\nu }({{{{{{{\bf{k}}}}}}}})]}_{mn}=\int\nolimits_{0}^{T}\frac{dt}{T}{e}^{i(m-n){{\Omega }}t}{v}_{ab\sigma }^{({{{{{{{\rm{Q}}}}}}}})\nu }({{{{{{{\bf{k}}}}}}}},t),$$
(9)
$${\left[{G}_{a\sigma b{\sigma }^{{\prime} }}^{r}({{{{{{{\bf{k}}}}}}}},{\omega }^{{\prime} })\right]}_{mn} = \int\nolimits_{-\infty }^{\infty }d{t}_{{{{{{{{\rm{rel}}}}}}}}}{e}^{i({\omega }^{{\prime} }+\frac{m+n}{2}{{\Omega }}){t}_{{{{{{{{\rm{rel}}}}}}}}}}\int\nolimits_{0}^{T}\frac{d{t}_{{{{{{{{\rm{av}}}}}}}}}}{T}\\ \times {e}^{i(m-n){{\Omega }}{t}_{{{{{{{{\rm{av}}}}}}}}}}{G}_{a\sigma b{\sigma }^{{\prime} }}^{r}({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} }),$$
(10)

respectively; the three Green’s functions are determined from the Dyson equation with the damping Γ due to the system-bath coupling (see Methods). (For the energy dispersion of our model, see Supplementary Note 2.)

Helicity-independent \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and helicity-dependent \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\)

We evaluate \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) numerically. (For the details of the numerical calculations, see Methods.) We set Γ = 0.03 eV and Tb = 0.05 eV; Γ is chosen to be smaller than Tb because the system is supposed to be well described by the Fermi liquid. To study how \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) are affected by the helicity of light, we consider the Apump(t)’s for δ = 0 and π [Eq. (4)], ALCP(t) and ARCP(t), which correspond to the cases of the left- and right-circularly polarized light, respectively. We show how \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) depend on a dimensionless quantity u = eA0 = eE0/Ω. Note that the u dependence at fixed Ω corresponds to the dependence on E0, the amplitude of the electric field.

\({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) have the different helicity dependences. Figure 2a shows the dependence of \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) on u = eA0 for Apump(t) = ALCP(t) or ARCP(t) at Ω = 6 eV. The \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) for Apump(t) = ALCP(t) is the same as that for Apump(t) = ARCP(t). This property holds even at Ω = 4 and 2 eV (Fig. 2b, c). Note that Ω = 6, 4, and 2 eV correspond to Ω > W, Ω ≈ W, and Ω < W, respectively, where W( ≈ 4eV) is the bandwidth in the non-driven case. Meanwhile, \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\)’s for Apump(t) = ALCP(t) and ARCP(t) are opposite in sign and the same in magnitude at Ω = 6, 4, and 2 eV (Fig. 2d–f). Although such helicity-dependent \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) was experimentally shown in graphene8, its origin may be unexplored. Note that the difference between the u dependences of \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) can be qualitatively understood by considering the dominant terms of the Bessel functions due to the Peierls phase factors (see Supplementary Note 3 and Supplementary Fig. 1).

Fig. 2: Helicity dependences of the spin-Hall and anomalous-Hall conductivities.
figure 2

a, b, c The dependences of the spin-Hall conductivity \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) on the dimensionless quantity u = eA0 in the case of left- or right-circularly polarized light (LCP or RCP) at Ω = 6, 4, and 2 eV, where Ω is the frequency of light. The red and blue curves correspond to those in the case of left- or right-circularly polarized light, respectively. d, e, f The dependences of the anomalous-Hall conductivity \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) on u = eA0 in the case of left- or right-circularly polarized light at Ω = 6, 4, and 2 eV. The same notations as those in a, b, c are used.

This difference between \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) comes from the difference in TRS. Under the time-reversal operation Trev, time t, momentum k, and spin σ are changed as follows: (t, k, σ) → (− t, − k, − σ), where − σ =  or for σ =  or , respectively. The spin current and charge current are expressed as \({{{{{{{{\bf{J}}}}}}}}}_{{{{{{{{\rm{S}}}}}}}}}=\frac{1}{2}({{{{{{{{\bf{J}}}}}}}}}_{\uparrow }-{{{{{{{{\bf{J}}}}}}}}}_{\downarrow })\) and JC = (− e)(J + J), where J and J are the contributions from the spin-up and spin-down electrons, respectively. Thus, (JS, JC) → (JS, − JC) is obtained as a result of Trev because (J, J) → (− J, − J) is satisfied under Trev (Fig. 3a, b). (This is the reason why TRS is broken in the AHE and not broken in the SHE.) Meanwhile, the right- and left-circularly polarized light fields are connected by Trev because ARCP(− t) = ALCP(t). Namely, replacing ALCP(t) by ARCP(t) corresponds to applying Trev. Thus, the helicity-independent \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and the helicity-dependent \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) result from JS → JS and JC → − JC, respectively, under Trev.

Fig. 3: Time-reversal symmetry of the charge current and spin current.
figure 3

a, b The charge currents and the spin currents before and after the time-reversal operation Trev. The charge current JC and the spin current JS are JC = (− e)(J + J) and JS = (1/2)(J − J), where J and J are the spin-up and spin-down electron currents, respectively. As a result of Trev, J and J become − J and − J, respectively. Thus, JC changes its sign (a), whereas JS remains the same (b). Namely, JC breaks time-reversal symmetry, but JS does not.

The same helicity dependences hold in many periodically driven multiorbital metals. The spin current and charge current are of the same form for some transition metals (e.g., Pt and Au)28 and transition-metal oxides. Then, the similar SHE and AHE can be realized using circularly polarized light. Thus, the above arguments are applicable to many transition-metal oxides and transition metals driven by circularly polarized light.

SOC-dependent \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and SOC-independent \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\)

There is another difference between \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\). Figure 4a compares the u dependence of \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) with SOC to that without SOC. In the absence of SOC, \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}=0\). This is because there is no spin-dependent term in the Hamiltonian except for SOC. The spin-dependent term, such as SOC, is needed to obtain the finite difference between the spin-up and spin-down electron currents. Meanwhile, the u dependence of \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) with SOC is almost the same as that without SOC (Fig. 4b). This is because a spin-independent electron current can be generated by using the kinetic energy terms with the Peierls phase factors6,33 and a multiorbital mechanism48 using SOC does not contribute to \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) in the presence of spin degeneracy, which is not lifted by the Peierls phase factors. Note that in periodically driven systems, \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) can be finite even without orbital degrees of freedom6,33 because the Peierls phase factors can lead to the terms odd with respect to momentum in the energy dispersion (see Supplementary Note 2).

Fig. 4: Spin–orbit coupling dependences of the spin-Hall and anomalous-Hall conductivities.
figure 4

a, b The dependences of the spin-Hall and anomalous-Hall conductivities \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) on the dimensionless quantity u = eA0 in the case of left-circularly polarized light at Ω = 6 eV with and without spin–orbit coupling. Here Ω is the frequency of light. The red and yellow curves correspond to those with and without spin–orbit coupling, respectively.

These results suggest that in periodically driven multiorbital metals, SOC is vital for the SHE, whereas it is unnecessary for the AHE. This suggestion may be valid as long as the effects of the driving field can be treated as the Peierls phase factors and there is no magnetic order. In addition, this is distinct from the property of non-driven multiorbital metals where SOC is vital for the SHE and AHE24,25,26,28,29,48. In contrast, the multiorbital nature is required for the SHE of periodically driven systems, whereas it is unnecessary for the AHE.

Implications and experimental realization

We discuss some implications of our results. First, the difference between the helicity dependences of \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) can be used to distinguish the spin current and charge current without ambiguity. Since that difference results from the symmetry of Trev, the same helicity dependences should hold in many periodically driven systems. In addition, the similar arguments enable us to distinguish two currents, one of which breaks TRS (and the other does not), in not only Hall effects, but also other transport phenomena. Thus, our results have revealed the core physics discipline about the relations between TRS and transport properties of periodically driven systems. Then, our theory can be extended to the SHE and AHE of other multiorbital metals and other transport phenomena. For example, a combination of it and first-principles calculations enables us to systematically search the SHE and AHE of periodically driven multiorbital metals. Thus, our results provide the first step towards the Floquet engineering of spintronics phenomena, including the SHE, of periodically driven multiorbital metals.

Finally, we comment on experimental realization. In our theory, interaction effects and heating effects are neglected. For Sr2RuO4, electron-electron interactions cause the orbital-dependent damping and mass enhancement35,49. Since these effects are quantitative29, the interaction effects may not change our results at least qualitatively. The differences in the helicity dependence and the SOC dependence will hold because those interaction effects do not break TRS. In general, the periodic driving makes the system to heat up50. However, for the periodically driven open system, such as our system, a nonequilibrium steady state can be realized due to Γ32,33,51 at times larger than τ(= /2Γ) ≈ 11fs = O(10fs). In fact, the AHE predicted theoretically in a periodically driven open system6 is experimentally realized7,8. For Sr2RuO4, in which alc ≈ 0.39 nm35, u(= ealcA0) = 0.3 at Ω = 2, 4, or 6 eV corresponds to E0 = A0/Ω ≈ 15, 31, or 46 MVcm−1, respectively. Since the pump field of the order of 10 MVcm−1 is experimentally accessible52, we conclude that the predicted properties of \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\) and \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) could be observed in the pump-probe measurements of the SHE and AHE in periodically driven Sr2RuO4.

Methods

Tight-binding Hamiltonian with SOC

We have chosen the following tight-binding Hamiltonian for t2g-orbital electrons as Hs(t):

$${H}_{{{{{{{{\rm{s}}}}}}}}}(t)= \mathop{\sum}\limits_{i,j}\mathop{\sum}\limits_{a,b={d}_{yz},{d}_{zx},{d}_{xy}}\mathop{\sum}\limits_{\sigma =\uparrow ,\downarrow }[{t}_{ij}^{ab}(t)-\mu {\delta }_{i,j}{\delta }_{a,b}]{c}_{ia\sigma }^{{{{\dagger}}} }{c}_{jb\sigma }\\ +\mathop{\sum}\limits_{i}\mathop{\sum}\limits_{a,b={d}_{yz},{d}_{zx},{d}_{xy}}\mathop{\sum}\limits_{\sigma ,{\sigma }^{{\prime} }=\uparrow ,\downarrow }{\xi }_{ab}^{\sigma {\sigma }^{{\prime} }}{c}_{ia\sigma }^{{{{\dagger}}} }{c}_{ib{\sigma }^{{\prime} }},$$
(11)

where \({t}_{ij}^{ab}(t)\)’s are the hopping integrals with the Peierls phase factors due to A(t), \({t}_{ij}^{ab}(t)={t}_{ij}^{ab}{e}^{-ie({{{{{{{{\bf{R}}}}}}}}}_{i}-{{{{{{{{\bf{R}}}}}}}}}_{j})\cdot {{{{{{{\bf{A}}}}}}}}(t)}\), and \({\xi }_{ab}^{\sigma {\sigma }^{{\prime} }}\) is the coupling constant of the SOC for t2g-orbital electrons. The finite elements of \({\xi }_{ab}^{\sigma {\sigma }^{{\prime} }}={({\xi }_{ba}^{{\sigma }^{{\prime} }\sigma })}^{* }\) are given by \({\xi }_{{d}_{yz}{d}_{zx}}^{\uparrow \uparrow }={\xi }_{{d}_{zx}{d}_{xy}}^{\uparrow \downarrow }=i\xi /2\), \({\xi }_{{d}_{yz}{d}_{xy}}^{\uparrow \downarrow }=-\xi /2\), \({\xi }_{{d}_{xy}{d}_{yz}}^{\uparrow \downarrow }=\xi /2\), and \({\xi }_{{d}_{xy}{d}_{zx}}^{\uparrow \downarrow }={\xi }_{{d}_{yz}{d}_{zx}}^{\downarrow \downarrow }=-i\xi /2\). By using the Fourier coefficients of the operators, we can write Eq. (11) as Eq. (2) with Eq. (3), in which ϵab(k, t) is given by \({\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}},t)={\sum }_{j}{t}_{ij}^{ab}(t){e}^{-i{{{{{{{\bf{k}}}}}}}}\cdot ({{{{{{{{\bf{R}}}}}}}}}_{i}-{{{{{{{{\bf{R}}}}}}}}}_{j})}\).

Büttiker-type heat bath

Hsb and Hb in Eq. (1) are given by

$${H}_{{{{{{{{\rm{sb}}}}}}}}}=\mathop{\sum}\limits_{i}\mathop{\sum}\limits_{p}\mathop{\sum}\limits_{a={d}_{yz},{d}_{zx},{d}_{xy}}\mathop{\sum}\limits_{\sigma =\uparrow ,\downarrow }{V}_{pa\sigma }({c}_{ia\sigma }^{{{{\dagger}}} }{b}_{ip}+{b}_{ip}^{{{{\dagger}}} }{c}_{ia\sigma }),$$
(12)
$${H}_{{{{{{{{\rm{b}}}}}}}}}=\mathop{\sum}\limits_{i}\mathop{\sum}\limits_{p}({\epsilon }_{p}-{\mu }_{{{{{{{{\rm{b}}}}}}}}}){b}_{ip}^{{{{\dagger}}} }{b}_{ip},$$
(13)

where bip and \({b}_{ip}^{{{{\dagger}}} }\) are the annihilation and creation operators, respectively, of a bath’s fermion at site i for mode p, Vpaσ is the coupling constant, and ϵp and μb are the energy and chemical potential of a bath’s fermion. Note that μb is chosen in order that there is no current between the system and bath. The heat bath is supposed to be in equilibrium at temperature Tb. The main effect of the heat bath is the damping appearing in electron Green’s functions32,33.

Charge current and spin current operators

We derive the charge current and spin current operators using the continuity equations. Theories using these operators derived in that way succeed in describing the SHE observed in non-driven multiorbital metals28,53.

First, we derive the charge current operator JC(t). JC(t) is supposed to satisfy the continuity equation44,

$$\frac{d{\rho }_{j}(t)}{dt}+\nabla \cdot {{{{{{{{\bf{j}}}}}}}}}_{j}^{({{{{{{{\rm{C}}}}}}}})}(t)=0,$$
(14)

where \({\rho }_{j}(t)=(-e){\sum }_{a}{\sum }_{\sigma }{c}_{ja\sigma }^{{{{\dagger}}} }(t){c}_{ja\sigma }(t)\) and \({\sum }_{j}{{{{{{{{\bf{j}}}}}}}}}_{j}^{({{{{{{{\rm{C}}}}}}}})}(t)={{{{{{{{\bf{J}}}}}}}}}_{{{{{{{{\rm{C}}}}}}}}}(t)\). Using Eq. (14), we have

$$\mathop{\sum}\limits_{j}{{{{{{{{\bf{R}}}}}}}}}_{j}\frac{d{\rho }_{j}(t)}{dt}=-\mathop{\sum}\limits_{j}{{{{{{{{\bf{R}}}}}}}}}_{j}\nabla \cdot {{{{{{{{\bf{j}}}}}}}}}_{j}^{({{{{{{{\rm{C}}}}}}}})}(t)={{{{{{{{\bf{J}}}}}}}}}_{{{{{{{{\rm{C}}}}}}}}}(t),$$
(15)

where we have omitted the surface contributions. By combining it with the Heisenberg equation, we can write Eq. (15) as

$${{{{{{{{\bf{J}}}}}}}}}_{{{{{{{{\rm{C}}}}}}}}}(t)=i[{H}_{{{{{{{{\rm{s}}}}}}}}}(t),\mathop{\sum}\limits_{j}{{{{{{{{\bf{R}}}}}}}}}_{j}{\rho }_{j}(t)].$$
(16)

(Note that there is no contribution from Hsb because the bath’s chemical potential is chosen in order that there is no current between the system and bath). After some calculations, we obtain

$${{{{{{{{\bf{J}}}}}}}}}_{{{{{{{{\rm{C}}}}}}}}}(t) =i\mathop{\sum}\limits_{i,j}\mathop{\sum}\limits_{a,b}\mathop{\sum}\limits_{\sigma }(-e){t}_{ij}^{ab}(t)({{{{{{{{\bf{R}}}}}}}}}_{j}-{{{{{{{{\bf{R}}}}}}}}}_{i}){c}_{ia\sigma }^{{{{\dagger}}} }(t){c}_{jb\sigma }(t)\\ =-e\mathop{\sum}\limits_{{{{{{{{\bf{k}}}}}}}}}\mathop{\sum}\limits_{a,b}\mathop{\sum}\limits_{\sigma }\frac{\partial {\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}},t)}{\partial {{{{{{{\bf{k}}}}}}}}}{c}_{{{{{{{{\bf{k}}}}}}}}a\sigma }^{{{{\dagger}}} }(t){c}_{{{{{{{{\bf{k}}}}}}}}b\sigma }(t).$$
(17)

Similarly, we derive the spin current operator JS(t). We suppose that JS(t) satisfies

$$\frac{d{S}_{j}^{z}(t)}{dt}+\nabla \cdot {{{{{{{{\bf{j}}}}}}}}}_{j}^{({{{{{{{\rm{S}}}}}}}})}(t)=0,$$
(18)

where \({S}_{j}^{z}(t)={\sum }_{a}{\sum }_{\sigma }\frac{1}{2}{{{{{{{\rm{sgn}}}}}}}}(\sigma ){c}_{ja\sigma }^{{{{\dagger}}} }(t){c}_{ja\sigma }(t)\) and \({\sum }_{j}{{{{{{{{\bf{j}}}}}}}}}_{j}^{({{{{{{{\rm{S}}}}}}}})}(t)={{{{{{{{\bf{J}}}}}}}}}_{{{{{{{{\rm{S}}}}}}}}}(t)\). In a way similar to the derivation of JC(t), JS(t) is given by

$${{{{{{{{\bf{J}}}}}}}}}_{{{{{{{{\rm{S}}}}}}}}}(t) =i[{H}_{{{{{{{{\rm{s}}}}}}}}}(t),\mathop{\sum}\limits_{j}{{{{{{{{\bf{R}}}}}}}}}_{j}{S}_{j}^{z}(t)]\\ =\frac{1}{2}\mathop{\sum}\limits_{{{{{{{{\bf{k}}}}}}}}}\mathop{\sum}\limits_{a,b}\mathop{\sum}\limits_{\sigma }{{{{{{{\rm{sgn}}}}}}}}(\sigma )\frac{\partial {\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}},t)}{\partial {{{{{{{\bf{k}}}}}}}}}{c}_{{{{{{{{\bf{k}}}}}}}}a\sigma }^{{{{\dagger}}} }(t){c}_{{{{{{{{\bf{k}}}}}}}}b\sigma }(t).$$
(19)

Anomalous-Hall and spin-Hall conductivities as functions of time

We express \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}(t,{t}^{{\prime} })\) and \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}(t,{t}^{{\prime} })\) in terms of the electron Green’s functions. Using Eq. (6), we have

$$\langle {j}_{{{{{{{{\rm{C}}}}}}}}}^{y}(t)\rangle =\frac{-i}{N}\mathop{\sum}\limits_{{{{{{{{\bf{k}}}}}}}}}\mathop{\sum}\limits_{a,b}\mathop{\sum}\limits_{\sigma }{v}_{ab\sigma }^{({{{{{{{\rm{C}}}}}}}})y}({{{{{{{\bf{k}}}}}}}},t){G}_{b\sigma a\sigma }^{ < }({{{{{{{\bf{k}}}}}}}};t,t),$$
(20)
$$\langle {j}_{{{{{{{{\rm{S}}}}}}}}}^{y}(t)\rangle =\frac{-i}{N}\mathop{\sum}\limits_{{{{{{{{\bf{k}}}}}}}}}\mathop{\sum}\limits_{a,b}\mathop{\sum}\limits_{\sigma }{v}_{ab\sigma }^{({{{{{{{\rm{S}}}}}}}})y}({{{{{{{\bf{k}}}}}}}},t){G}_{b\sigma a\sigma }^{ < }({{{{{{{\bf{k}}}}}}}};t,t),$$
(21)

where \({G}_{b{\sigma }^{{\prime} }a\sigma }^{ < }({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} })\) is the lesser Green’s function34,44,46,47,

$${G}_{b{\sigma }^{{\prime} }a\sigma }^{ < }({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} })=i\langle {c}_{{{{{{{{\bf{k}}}}}}}}a\sigma }^{{{{\dagger}}} }({t}^{{\prime} }){c}_{{{{{{{{\bf{k}}}}}}}}b{\sigma }^{{\prime} }}(t)\rangle .$$
(22)

By substituting Eqs. (20) and (21) into Eq. (5), we can express \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}(t,{t}^{{\prime} })\) and \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}(t,{t}^{{\prime} })\) as follows:

$${\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}(t,{t}^{{\prime} })={\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}(1)}(t,{t}^{{\prime} })+{\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}(2)}(t,{t}^{{\prime} }),$$
(23)
$${\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}(t,{t}^{{\prime} })={\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}(1)}(t,{t}^{{\prime} })+{\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}(2)}(t,{t}^{{\prime} }),$$
(24)

where

$${\sigma }_{yx}^{{{{{{{{\rm{Q}}}}}}}}(1)}(t,{t}^{{\prime} })=\frac{-1}{\omega N}\mathop{\sum}\limits_{{{{{{{{\bf{k}}}}}}}}}\mathop{\sum}\limits_{a,b}\mathop{\sum}\limits_{\sigma }\frac{\delta {v}_{ab\sigma }^{({{{{{{{\rm{Q}}}}}}}})y}({{{{{{{\bf{k}}}}}}}},t)}{\delta {A}_{{{{{{{{\rm{prob}}}}}}}}}^{x}({t}^{{\prime} })}{G}_{b\sigma a\sigma }^{ < }({{{{{{{\bf{k}}}}}}}};t,t),$$
(25)
$${\sigma }_{yx}^{{{{{{{{\rm{Q}}}}}}}}(2)}(t,{t}^{{\prime} })=\frac{-1}{\omega N}\mathop{\sum}\limits_{{{{{{{{\bf{k}}}}}}}}}\mathop{\sum}\limits_{a,b}\mathop{\sum}\limits_{\sigma }{v}_{ab\sigma }^{({{{{{{{\rm{Q}}}}}}}})y}({{{{{{{\bf{k}}}}}}}},t)\frac{\delta {G}_{b\sigma a\sigma }^{ < }({{{{{{{\bf{k}}}}}}}};t,t)}{\delta {A}_{{{{{{{{\rm{prob}}}}}}}}}^{x}({t}^{{\prime} })}.$$
(26)

Then, using the Dyson equation of Green’s functions and the Langreth rule33,47, we obtain

$$ \frac{\delta {G}_{b\sigma a\sigma }^{ < }({{{{{{{\bf{k}}}}}}}};t,t)}{\delta {A}_{{{{{{{{\rm{prob}}}}}}}}}^{x}({t}^{{\prime} })}\\ = -\mathop{\sum}\limits_{c,d}\mathop{\sum}\limits_{{\sigma }^{{\prime} }}{v}_{cd{\sigma }^{{\prime} }}^{({{{{{{{\rm{C}}}}}}}})x}({{{{{{{\bf{k}}}}}}}},{t}^{{\prime} })\left[{G}_{b\sigma c{\sigma }^{{\prime} }}^{{{{{{{{\rm{R}}}}}}}}}({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} }){G}_{d{\sigma }^{{\prime} }a\sigma }^{ < }({{{{{{{\bf{k}}}}}}}};{t}^{{\prime} },t)\right.\\ \quad \left.+{G}_{b\sigma c{\sigma }^{{\prime} }}^{ < }({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} }){G}_{d{\sigma }^{{\prime} }a\sigma }^{{{{{{{{\rm{A}}}}}}}}}({{{{{{{\bf{k}}}}}}}};{t}^{{\prime} },t)\right],$$
(27)

where \({G}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{R}}}}}}}}}({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} })\) and \({G}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{A}}}}}}}}}({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} })\) are the retarded and advanced Green’s functions34,44,46,47, respectively,

$${G}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{R}}}}}}}}}({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} })=-i\theta (t-{t}^{{\prime} })\langle \{{c}_{{{{{{{{\bf{k}}}}}}}}a\sigma }(t),{c}_{{{{{{{{\bf{k}}}}}}}}b{\sigma }^{{\prime} }}^{{{{\dagger}}} }({t}^{{\prime} })\}\rangle ,$$
(28)
$${G}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{A}}}}}}}}}({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} })=i\theta ({t}^{{\prime} }-t)\langle \{{c}_{{{{{{{{\bf{k}}}}}}}}a\sigma }(t),{c}_{{{{{{{{\bf{k}}}}}}}}b{\sigma }^{{\prime} }}^{{{{\dagger}}} }({t}^{{\prime} })\}\rangle .$$
(29)

Combining Eq. (27) with Eq. (26), we have

$${\sigma }_{yx}^{{{{{{{{\rm{Q}}}}}}}}(2)}(t,{t}^{{\prime} })= \frac{1}{\omega N}\mathop{\sum}\limits_{{{{{{{{\bf{k}}}}}}}}}\mathop{\sum}\limits_{a,b,c,d}\mathop{\sum}\limits_{\sigma ,{\sigma }^{{\prime} }}{v}_{ab\sigma }^{({{{{{{{\rm{Q}}}}}}}})y}({{{{{{{\bf{k}}}}}}}},t){v}_{cd{\sigma }^{{\prime} }}^{({{{{{{{\rm{C}}}}}}}})x}({{{{{{{\bf{k}}}}}}}},{t}^{{\prime} })\\ \times \left[{G}_{b\sigma c{\sigma }^{{\prime} }}^{{{{{{{{\rm{R}}}}}}}}}({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} }){G}_{d{\sigma }^{{\prime} }a\sigma }^{ < }({{{{{{{\bf{k}}}}}}}};{t}^{{\prime} },t)\right.\\ \left.+ \, {G}_{b\sigma c{\sigma }^{{\prime} }}^{ < }({{{{{{{\bf{k}}}}}}}};t,{t}^{{\prime} }){G}_{d{\sigma }^{{\prime} }a\sigma }^{{{{{{{{\rm{A}}}}}}}}}({{{{{{{\bf{k}}}}}}}};{t}^{{\prime} },t)\right].$$
(30)

Dyson equation of Green’s functions

The Green’s functions of our periodically driven system are determined from the Dyson equation in a matrix form:

$$G={G}_{0}+{G}_{0}{{\Sigma }}G,$$
(31)

where G, G0, and Σ are the matrices of the Green’s functions with Hsb, those without Hsb, and the self-energies due to the second-order perturbation of Hsb, respectively,

$$\begin{array}{r}G=\left(\begin{array}{ll}{G}^{{{{{{{{\rm{R}}}}}}}}}&{G}^{{{{{{{{\rm{K}}}}}}}}}\\ 0&{G}^{{{{{{{{\rm{A}}}}}}}}}\end{array}\right),{G}_{0}=\left(\begin{array}{ll}{G}_{0}^{{{{{{{{\rm{R}}}}}}}}}&{G}_{0}^{{{{{{{{\rm{K}}}}}}}}}\\ 0&{G}_{0}^{{{{{{{{\rm{A}}}}}}}}}\end{array}\right),{{\Sigma }}=\left(\begin{array}{ll}{{{\Sigma }}}^{{{{{{{{\rm{R}}}}}}}}}&{{{\Sigma }}}^{{{{{{{{\rm{K}}}}}}}}}\\ 0&{{{\Sigma }}}^{{{{{{{{\rm{A}}}}}}}}}\end{array}\right).\end{array}$$
(32)

The superscripts R, A, and K denote the retarded, advanced, and Keldysh components, respectively. For example, the matrix GR as a function of k and ω is given by \({G}^{{{{{{{{\rm{R}}}}}}}}}=({[{G}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{R}}}}}}}}}({{{{{{{\bf{k}}}}}}}},\omega )]}_{mn})\) for a, b = dyz, dzx, dxy, \(\sigma ,{\sigma }^{{\prime} }=\uparrow ,\downarrow\), and m, n = −,   , 0,   , . The retarded, advanced, and Keldysh components are related to the lesser one through the identity, such as

$${G}^{ < }=\frac{1}{2}({G}^{{{{{{{{\rm{K}}}}}}}}}-{G}^{{{{{{{{\rm{R}}}}}}}}}+{G}^{{{{{{{{\rm{A}}}}}}}}}).$$
(33)

By treating the effects of Hsb in the second-order perturbation theory, we can express the retarded, advanced, and Keldysh self-energies as follows:

$${[{{{\Sigma }}}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{R}}}}}}}}}({{{{{{{\bf{k}}}}}}}},\omega )]}_{mn}=-i{\delta }_{m,n}{\delta }_{a,b}{\delta }_{\sigma ,{\sigma }^{{\prime} }}{{\Gamma }},$$
(34)
$${[{{{\Sigma }}}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{A}}}}}}}}}({{{{{{{\bf{k}}}}}}}},\omega )]}_{mn}=+i{\delta }_{m,n}{\delta }_{a,b}{\delta }_{\sigma ,{\sigma }^{{\prime} }}{{\Gamma }},$$
(35)
$${[{{{\Sigma }}}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{K}}}}}}}}}({{{{{{{\bf{k}}}}}}}},\omega )]}_{mn}=-2i{\delta }_{m,n}{\delta }_{a,b}{\delta }_{\sigma ,{\sigma }^{{\prime} }}{{\Gamma }}\tanh \frac{\omega +m{{\Omega }}}{2T_{{{\rm{b}}}}}.$$
(36)

In deriving them, we have omitted the real parts and replaced \(\pi {\sum }_{p}{V}_{pa\sigma }{V}_{pb{\sigma }^{{\prime} }}\delta (\omega +m{{\Omega }}-{\epsilon }_{p}+{\mu }_{{{{{{{{\rm{b}}}}}}}}})\) by \({{\Gamma }}{\delta }_{a,b}{\delta }_{\sigma ,{\sigma }^{{\prime} }}\) for simplicity. Such simplification may be sufficient because the main effect of Hsb is the relaxation towards the nonequilibrium steady state due to the damping32,33. Then, using the matrix relation G−1G = 1 and Eq. (32), we have

$${({G}^{{{{{{{{\rm{R}}}}}}}}})}^{-1}={({G}^{-1})}^{{{{{{{{\rm{R}}}}}}}}},$$
(37)
$${({G}^{{{{{{{{\rm{A}}}}}}}}})}^{-1}={({G}^{-1})}^{{{{{{{{\rm{A}}}}}}}}},$$
(38)
$${G}^{{{{{{{{\rm{K}}}}}}}}}=-{G}^{{{{{{{{\rm{R}}}}}}}}}{({G}^{-1})}^{{{{{{{{\rm{K}}}}}}}}}{G}^{{{{{{{{\rm{A}}}}}}}}},$$
(39)

where

$${G}^{-1}=\left(\begin{array}{ll}{({G}^{-1})}^{{{{{{{{\rm{R}}}}}}}}}&{({G}^{-1})}^{{{{{{{{\rm{K}}}}}}}}}\\ 0&{({G}^{-1})}^{{{{{{{{\rm{A}}}}}}}}}\end{array}\right).$$
(40)

Therefore, the retarded and advanced Green’s functions with Hsb are obtained by calculating the inverse matrices of \({({G}^{-1})}^{{{{{{{{\rm{R}}}}}}}}}\) and \({({G}^{-1})}^{{{{{{{{\rm{A}}}}}}}}}\), respectively,

$${[{({G}^{-1})}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{R}}}}}}}}}({{{{{{{\bf{k}}}}}}}},\omega )]}_{mn} = \, (\omega +\mu +m{{\Omega }}+i{{\Gamma }}){\delta }_{m,n}{\delta }_{a,b}{\delta }_{\sigma ,{\sigma }^{{\prime} }}\\ -{\xi }_{ab}^{\sigma {\sigma }^{{\prime} }}{\delta }_{m,n}-{[{\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}})]}_{mn}{\delta }_{\sigma ,{\sigma }^{{\prime} }},$$
(41)
$${[{({G}^{-1})}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{A}}}}}}}}}({{{{{{{\bf{k}}}}}}}},\omega )]}_{mn} = \, (\omega +\mu +m{{\Omega }}-i{{\Gamma }}){\delta }_{m,n}{\delta }_{a,b}{\delta }_{\sigma ,{\sigma }^{{\prime} }}\\ -{\xi }_{ab}^{\sigma {\sigma }^{{\prime} }}{\delta }_{m,n}-{[{\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}})]}_{mn}{\delta }_{\sigma ,{\sigma }^{{\prime} }},$$
(42)

where

$${[{\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}})]}_{mn}=\int\nolimits_{0}^{T}\frac{dt}{T}{e}^{i(m-n){{\Omega }}t}{\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}},t).$$
(43)

The expressions of \({[{\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}})]}_{mn}\) for our model are provided in Supplementary Note 2; as shown there, \({[{\epsilon }_{ab}({{{{{{{\bf{k}}}}}}}})]}_{mn}\) includes the Bessel functions of the first kind as a function of u = eA0. After obtaining these Green’s functions, we can obtain the Keldysh Green’s function with Hsb by combining Eq. (39) with the following equation:

$${[{({G}^{-1})}_{a\sigma b{\sigma }^{{\prime} }}^{{{{{{{{\rm{K}}}}}}}}}({{{{{{{\bf{k}}}}}}}},\omega )]}_{mn}=2i{{\Gamma }}{\delta }_{m,n}{\delta }_{a,b}{\delta }_{\sigma ,{\sigma }^{{\prime} }}{{\Gamma }}\tanh \frac{\omega +m{{\Omega }}}{2T_{{{\rm{b}}}}}.$$
(44)

We finally obtain the lesser Green’s function with Hsb using the three Green’s functions obtained and Eq. (33).

Numerical calculations

We numerically calculated Eq. (8) for Q = C or S, \({\sigma }_{yx}^{{{{{{{{\rm{C}}}}}}}}}\) or \({\sigma }_{yx}^{{{{{{{{\rm{S}}}}}}}}}\), in the following way. The momentum summation was calculated by dividing the Brillouin zone into a Nx × Ny mesh and setting Nx = Ny = 100. The frequency integral was done by using \(\int\nolimits_{-{{\Omega }}/2}^{{{\Omega }}/2}d{\omega }^{{\prime} }F({\omega }^{{\prime} })\approx \mathop{\sum }\nolimits_{s = 0}^{W-1}{{\Delta }}{\omega }^{{\prime} }F({\omega }_{s}^{{\prime} })\), where \({\omega }_{s}^{{\prime} }=-{{\Omega }}/2+s{{\Delta }}{\omega }^{{\prime} }\) and \({\omega }_{W}^{{\prime} }={{\Omega }}/2\), and setting \({{\Delta }}{\omega }^{{\prime} }=0.005\) eV. The frequency derivatives of the Green’s functions was approximated by using \(\frac{\partial F({\omega }^{{\prime} })}{\partial {\omega }^{{\prime} }}\approx \frac{F({\omega }^{{\prime} }+{{\Delta }}{\omega }^{{\prime} })-F({\omega }^{{\prime} }-{{\Delta }}{\omega }^{{\prime} })}{2{{\Delta }}{\omega }^{{\prime} }}\). The summations over the Floquet indices, \(\mathop{\sum }\nolimits_{m,l,n,q = -\infty }^{\infty }\), was replaced by \(\mathop{\sum }\nolimits_{m,l,n,q = -{n}_{\max }}^{{n}_{\max }}\), and \({n}_{\max }\) was fixed at \({n}_{\max }=2\) for Ω = 6 and 4 eV or \({n}_{\max }=3\) for Ω = 2 eV.