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

Valley transport driven by dynamic lattice distortion

Yuya Ominato Kavli Institute for Theoretical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China.    Daigo Oue Kavli Institute for Theoretical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China. The Blackett Laboratory, Department of Physics, Imperial College London, Prince Consort Road, Kensington, London SW7 2AZ, United Kingdom    Mamoru Matsuo Kavli Institute for Theoretical Sciences, University of Chinese Academy of Sciences, Beijing 100190, China. CAS Center for Excellence in Topological Quantum Computation, University of Chinese Academy of Sciences, Beijing 100190, China RIKEN Center for Emergent Matter Science (CEMS), Wako, Saitama 351-0198, Japan Advanced Science Research Center, Japan Atomic Energy Agency, Tokai 319-1195, Japan
Abstract

Angular momentum conversion between mechanical rotation and the valley degree of freedom in 2D Dirac materials is investigated theoretically. Coupling between the valley and vorticity of dynamic lattice distortions is derived by applying the kpk\cdot p method to 2D Dirac materials with an inertial effect. Lattice strain effects are also incorporated. Valley transfer and valley-dependent carrier localization are predicted using the dynamic lattice distortions. The transport properties are found to be controllable, allowing the system to be insulating and to generate pulsed charge current. Our formalism offers a route toward mechanical manipulation of valley dynamics in 2D Dirac materials.

I Introduction

Valleytronics is an emerging field that uses the valley degrees of freedom, i.e., the local extrema of the electronic band structure, with the aim of developing innovative approaches to information processing, optoelectronic devices, and quantum computation Schaibley et al. (2016); Krasnok et al. (2018); Vitale et al. (2018). In 2D Dirac materials such as gapped graphene Semenoff (1984); Xiao et al. (2007) and monolayer transition-metal dichalcogenides (TMDCs) Xiao et al. (2012), the electrons carry valley-dependent orbital angular momentum that originates from the Berry curvature caused by inversion symmetry breaking Xiao et al. (2007, 2010). As a result, the valleys in these systems can be identified based on the intrinsic orbital angular momentum of the electrons, as illustrated in Fig. 1 (a). This means that the valleys can be controlled using an external field coupled to the orbital angular momentum. In fact, it has been demonstrated experimentally that circularly polarized light beams Cao et al. (2012); Mak et al. (2012); Zeng et al. (2012); Sallen et al. (2012); Guddala et al. (2021) and magnetic fields Li et al. (2014); Aivazian et al. (2015); MacNeill et al. (2015); Srivastava et al. (2015) can be used to manipulate the valley degree of freedom. Since these demonstrations, more versatile valley control methods have been desired. One candidate method taken from familiar phenomena is the gyroscopic effect, by which the angular momentum is coupled with mechanical rotation.

Refer to caption
Figure 1: (a) Intrinsic valley-dependent orbital angular momentum for two inequivalent valleys. (b) 2D Dirac materials on a substrate in the presence of Love-type SAWs. The sum of the pseudo-valley Zeeman coupling (pVZC), and the valley-vorticity coupling (VVC) is shown for three cases. (i) The carriers are trapped at the bottom of UU and the SAW conveys the valley degree of freedom. (ii) The carriers are localized when the pVZC exceeds the Fermi energy. Both the K+K_{+} and KK_{-} valleys are localized at the same location. (iii) The carriers are localized when the VVC exceeds the Fermi energy. Here, the K+K_{+} and KK_{-} valley carriers are localized at different locations. (c) The dynamic lattice distortion leads to two effective magnetic fields: the strain-induced pseudo-magnetic field 𝑩s\bm{B}^{s} and the vorticity field 𝝎\bm{\omega}.

This gyroscopic coupling technique has been observed and used in a variety of systems. For example, gyroscopes are used to measure the gravitomagnetic field precisely in a curved space near the Earth Everitt et al. (2011). The quantum version of the gyroscopic effect is called the gyromagnetic effect, where both the orbital and spin angular momenta are coupled to the mechanical rotation. The gyromagnetic effect was discovered during an attempt to explore the origins of magnetism Einstein and de Haas (1915); Barnett (1915, 1935) and led to the discovery of electron spin angular momentum before quantum mechanics was established.

While the original targets in these gyromagnetic experiments were limited to ferromagnetic materials, gyromagnetic coupling itself is a universal phenomenon that occurs even in nonmagnetic materials. Indeed, gyromagnetic coupling has been observed in various branches of physics, including spintronics Chudo et al. (2014); Imai et al. (2018, 2019); Takahashi et al. (2016, 2020); Kazerooni et al. (2020, 2021); Kobayashi et al. (2017); Harii et al. (2019); Okano et al. (2019); Kurimune et al. (2020a, b); Mori et al. (2020); Tateno et al. (2020, 2021), ultrafast demagnetization processes Dornes et al. (2019), and quark-gluon many-body systems Adamczyk et al. (2017). These studies have shown that the coupling between angular momentum and rotation emerges universally in a variety of systems and continues to have a tremendous impact on a wide range of topics in fundamental physics.

In this work, we propose an alternative valley manipulation technique that uses gyroscopic coupling between the valley and the local rotational motion (or vorticity) excited by dynamic distortion in 2D Dirac materials. Figure 1 (b) shows schematic images of systems in which 2D Dirac materials are placed on a substrate and dynamic lattice distortion is then applied. We consider a Love wave, which is a type of horizontally polarized surface acoustic wave (SAW). This SAW leads to two effective magnetic fields: the strain-induced pseudo-magnetic field BzsB_{z}^{s} and the vorticity field ωz\omega_{z} as shown in Fig. 1 (c).

Both of these effective fields couple to the valley orbital angular momentum, thus allowing the valley to be manipulated using the SAW. We also consider electron doping of these 2D materials. In the presence of a standing SAW, valley transfer and dynamic valley polarization can be realized, depending on the Fermi energy εF\varepsilon_{F} and the strengths of BzsB^{s}_{z} and ωz\omega_{z}.

We also discuss the charge transport when an external DC electric field is applied to 2D Dirac materials in the presence of the SAW. We find that the conductivity is suppressed by the SAW and that a pulsed charge current is generated; this current can be used as a broadband microwave source. Our findings presented here pave the way towards valley device applications for 2D Dirac materials using the SAW devices Delsing et al. (2019).

The paper is organized as follows. The emergent gauge fields induced by the dynamic lattice distortion are introduced in Sec. II. The mechanism and procedure to incorporate the emergent gauge fields are explained. Finally, two kinds of the Zeeman like coupling are derived. The valley manipulation using the SAW are proposed in Sec. III. Furthermore, pulsed current generation and microwave radiation based on the proposed valley manipulation mechanism are discussed as application examples in Sec. IV and V, respectively. A conclusion is given in Sec. VI.

II Gauge fields induced by the dynamic lattice distortion

We consider 2D Dirac materials in which the dynamic lattice distortion is induced by the SAW in the substrate. The dynamic lattice distortion leads to two kinds of emergent gauge fields, the velocity fields due to the inertial effect Matsuo et al. (2017) and the strain-induced gauge field Suzuura and Ando (2002); Manes (2007); Vozmediano et al. (2010). They originate from the different mechanism. In the following subsections, we describe how each gauge field arises and how it is incorporated into the low-energy effective Hamiltonian of 2D Dirac materials.

II.1 Velocity field due to the inertial effect

In this subsection, we focus on the inertial effect. The period of the SAW is about a nanosecond and the typical lifetime of electrons would be about a picosecond, so that an adiabatic approximation is valid, where electrons adiabatically follow the dynamic lattice distortion. As a result, the effect of the dynamic lattice distortion is incorporated as the inertial effect Matsuo et al. (2017). Starting from the generally covariant Dirac Lagrangian, which governs the dynamics of the spin-1/2 particle in curved spacetime, the inertial effect is incorporated into the non-relativistic Hamiltonian as a U(1)U(1) potential

H=(𝒑+e𝑨v)22m+V(𝒓),\displaystyle H=\frac{(\bm{p}+e\bm{A}^{v})^{2}}{2m}+V(\bm{r}), (1)

where V(𝒓)V(\bm{r}) is a periodic potential of a crystal and 𝑨v\bm{A}^{v} is the emergent gauge field

𝑨v=me(u˙x,u˙y,u˙z),\displaystyle\bm{A}^{v}=-\frac{m}{e}(\dot{u}_{x},\dot{u}_{y},\dot{u}_{z}), (2)

with the velocity field of the lattice 𝒖˙\dot{\bm{u}}. The detailed derivation is explained in the Appendix. Therefore, the inertial effect is incorporated into the low-energy effective model as conventional electromagnetic field.

Based on the discussion in the above paragraph, the low-energy effective Hamiltonian for 2D Dirac materials, e.g., gapped graphene Semenoff (1984); Xiao et al. (2007) and TMDCs Xiao et al. (2012), is given by

H2D=v(ξσxπx+σyπy)+Δσz,\displaystyle H_{2D}=v(\xi\sigma_{x}\pi_{x}+\sigma_{y}\pi_{y})+\Delta\sigma_{z}, (3)

where vv is the velocity, 𝝈\bm{\sigma} are Pauli matrices that describe the pseudo spin, ξ=±1\xi=\pm 1 specifies the states at the K+K_{+} and KK_{-} valleys, 𝝅\bm{\pi} is the kinetic momentum, and Δ\Delta is an asymmetric potential breaking the inversion symmetry. The effect of the dynamic lattice distortion can be incorporated as emergent gauge fields. We then substitute 𝝅=𝒑+e𝑨v\bm{\pi}=\bm{p}+e\bm{A}^{v} into the above.

II.2 Strain-induced gauge field

In addition to the velocity field, the dynamic lattice distortion also leads to the strain-induced gauge field given by Suzuura and Ando (2002); Manes (2007); Vozmediano et al. (2010)

ξ𝑨s=ξE0ev(uxxuyy,2uxy),\displaystyle\xi\bm{A}^{s}=\xi\frac{E_{0}}{ev}(u_{xx}-u_{yy},-2u_{xy}), (4)

with the strain tensors uij=(jui+iuj)/2u_{ij}=(\partial_{j}u_{i}+\partial_{i}u_{j})/2 and the material parameter E0E_{0}. There are two ways to derive the strain-induced gauge field, starting from the tight-binding model and argument based on the symmetry. In addition to the velocity field 𝑨v\bm{A}^{v}, the strain-induced gauge field ξ𝑨s\xi\bm{A}^{s} has to be incorporated into the kinetic momentum 𝝅\bm{\pi} as 𝝅=𝒑+eξ𝑨s+e𝑨v\bm{\pi}=\bm{p}+e\xi\bm{A}^{s}+e\bm{A}^{v}. Note here that both the velocity field and the strain-induced gauge field are induced by the dynamic lattice distortion, but it is not always the case that both fields will be generated. For example, when a rigid body rotation is considered, only the velocity field due to the inertial effect arises.

II.3 Hamiltonian near the conduction band bottom

Although the electronic states in 2D Dirac materials is obtained by solving Eq. (3), it is convenient to reduce the two-band model to a one-band model to examine the physics near the conduction band bottom. Using the Schrieffer-Wolff transformation Schrieffer and Wolff (1966), the two-band model given in Eq. (3) can be reduced to a one-band model. The Schrieffer-Wolff transformation is a widely used technique that perturbatively incorporates the effect of off-diagonal matrix elements and systematically reduces the size of the matrix to be solved. Consequently, the effective Hamiltonian near the conduction band bottom is given by

H2Deff=𝝅22m+lγBzs2ξlωz2,\displaystyle H^{\mathrm{eff}}_{2D}=\frac{\bm{\pi}^{2}}{2m^{\ast}}+l\frac{\gamma B^{s}_{z}}{2}-\xi l\frac{\omega_{z}}{2}, (5)

where m=Δ/v2m^{\ast}=\Delta/v^{2} is the effective mass, ξl=ξm/m\xi l=\xi\hbar m/m^{\ast} is the intrinsic valley-dependent orbital angular momentum Xiao et al. (2007, 2010); Koshino and Ando (2010); Koshino (2011), γ=e/m\gamma=e/m is the gyromagnetic ratio, Bzs=xAysyAxsB^{s}_{z}=\partial_{x}A^{s}_{y}-\partial_{y}A^{s}_{x} is the strain-induced pseudo-magnetic field, and ωz=xu˙yyu˙x\omega_{z}=\partial_{x}\dot{u}_{y}-\partial_{y}\dot{u}_{x} is the vorticity field. The first term is the conventional kinetic energy term with the effective mass mm^{\ast}. The second term describes the coupling between the valley magnetic moment and the pseudo-magnetic field, which we call the pseudo valley Zeeman coupling (pVZC). The second term is independent of the valley index because the strain field preserves the time-reversal symmetry. The third term describes the coupling between the valley orbital angular momentum and the vorticity field, which we call the valley-vorticity coupling (VVC). The VVC is similar to the valley Zeeman coupling, which describes the coupling between the valley magnetic moment and a magnetic field Li et al. (2014); Aivazian et al. (2015); MacNeill et al. (2015); Srivastava et al. (2015). The derivation of the VVC is one of the main results reported in this paper. Note here that the acoustoelectric effect in the TMDC was discussed previously in a system similar to our setup Kalameitsev et al. (2019); Sonowal et al. (2020); however, the coupling between the dynamic lattice distortion and the valley orbital angular momentum is discussed for the first time in this work.

III Valley manipulation using SAW

In our setup, the Love wave is excited in the substrate by an external force, which leads to the dynamic lattice distortion in 2D Dirac materials. The detailed properties of the Love wave are explained in the Appendix. The lattice displacement vector in 2D Dirac materials is given by

𝒖=(0,u0sin(kx)sin(ωt),0),\displaystyle\bm{u}=(0,u_{0}\sin(kx)\sin(\omega t),0), (6)

which means that the vorticity field and the pseudo-magnetic field are given by

𝝎=(0,0,ω0cos(kx)cos(ωt))\displaystyle\bm{\omega}=(0,0,\omega_{0}\cos(kx)\cos(\omega t)) (7)

with ω0=u0ω2/cs\omega_{0}=u_{0}\omega^{2}/c_{s} and

𝑩s=(0,0,B0ssin(kx)sin(ωt))\displaystyle\bm{B}^{s}=(0,0,B_{0}^{s}\sin(kx)\sin(\omega t)) (8)

with B0s=ω0E0/evcsB_{0}^{s}=\omega_{0}E_{0}/evc_{s}, respectively. As shown in the Appendix, the dispersion relation is approximately linear, so that we use the dispersion relation ω=csk\omega=c_{s}k, where csc_{s} is the SAW velocity. The sum of the pVZC and the VVC can then be treated as a periodic potential

U=lω02[Rsin(kx)sin(ωt)ξcos(kx)cos(ωt)],\displaystyle U=l\frac{\omega_{0}}{2}\left[R\sin(kx)\sin(\omega t)-\xi\cos(kx)\cos(\omega t)\right], (9)

where the dimensionless material parameter RR is introduced

R=γE0evcs.\displaystyle R=\frac{\gamma E_{0}}{evc_{s}}. (10)

Different valley manipulations are possible, depending on the value of RR. Note that the dynamic lattice distortion induced here is not directly related to the phonon modes in the 2D Dirac materials. We expect that the effect due to the pVZC and the VVC can be detected even in the presence of the thermally activated phonon modes. Indeed, phenomena related to our proposal, such as strain induced Landau level formation in graphene Nigge et al. (2019) and spin manipulation using the SAW device Weiler et al. (2011); Kobayashi et al. (2017); Kurimune et al. (2020a), have been observed at room temperature.

We consider three cases showing characteristic carrier distribution in the electron-doped system.

  • (i) When the Fermi energy is sufficiently smaller than both the pVZC and the VVC (i.e., when εF|lγB0s|,|lω0|\varepsilon_{F}\ll|l\gamma B_{0}^{s}|,|l\omega_{0}|), the carriers are trapped at the bottom of potential and are transferred in the opposite direction for each valley, as shown in Fig. 1 (b)(i). This is a promising candidate mechanism to convey the valley information. This case can be realized using any value of RR.

  • (ii) When the Fermi energy is sufficiently larger than the VVC but smaller than the pVZC (i.e., when |lω0|εF|lγB0s||l\omega_{0}|\ll\varepsilon_{F}\ll|l\gamma B_{0}^{s}|), the carriers are then localized at the time ωt=(N+1/2)π,(N=0,±1,)\omega t=(N+1/2)\pi,\ (N=0,\pm 1,\cdots), as shown in Fig 1 (b)(ii). In this case, the spatial distribution of the carriers is almost independent of the valley. This case can be realized when R1R\gg 1.

  • (iii) When the Fermi energy is sufficiently larger than the pVZC but smaller than the VVC (i.e., when |lγB0s|εF|lω0||l\gamma B_{0}^{s}|\ll\varepsilon_{F}\ll|l\omega_{0}|), the carriers are then localized at the time ωt=Nπ,(N=0,±1,)\omega t=N\pi,\ (N=0,\pm 1,\cdots), as shown in Fig. 1 (b)(iii). In contrast to the second case, each valley’s carrier is localized in a different location separated by π/q\pi/q. This case can be realized when R1R\ll 1.

The valley transfer shown in Fig. 1 (b)(i) and the carrier localizations shown in Fig. 1 (b)(ii) and (iii) are detectable experimentally using time-resolved optical diffraction measurements. Because the carrier density pattern is periodic, diffraction patterns will appear under coherent illumination. The periods shown in Fig. 1 (b)(ii) and (iii) are different and thus the corresponding diffraction patterns are also different. Furthermore, the valley dependence of the carriers is also detectable using Kerr rotation microscopy, because the valleys carry an orbital magnetic moment Lee et al. (2016). The experimental signatures are listed in Table. 1.

Table 1: Experimental signatures to detect the valley manipulations discussed in this work.
Case (i) (ii) (iii)
Optical diffraction \checkmark \checkmark \checkmark
Kerr rotation \checkmark \checkmark
Pulsed current \checkmark \checkmark

Next, we estimate the strengths of the vorticity field and the pseudo-magnetic field. Using the parameters of graphene from the previous study Suzuura and Ando (2002), the relationship for the strengths of the vorticity field and the pseudo-magnetic field is given as ω0/γ103×B0s\omega_{0}/\gamma\sim 10^{-3}\times B^{s}_{0}, and when the parameters of MoS2\mathrm{MoS}_{2} from the preious study Rostami et al. (2015) are used, the relationship is given as ω0/γB0s\omega_{0}/\gamma\sim B_{0}^{s}. The experimentally feasible parameters for the SAW are u0100pmu_{0}\approx 100\ \mathrm{pm}, ct103m/sc_{t}\approx 10^{3}\ \mathrm{m/s}, and ω30GHz\omega\approx 30\ \mathrm{GHz}, meaning the strength of the vorticity field can be estimated to be ω0/γ0.01T\omega_{0}/\gamma\sim 0.01\ \mathrm{T}.

IV Manipulation of transport properties

In addition to the spatial distribution of the carriers, the carrier transport properties are also modulated by the SAW. Here, we discuss their longitudinal transport properties along the xx direction in the three cases described above based on the semiclassical Boltzmann transport theory. In case (i), the carriers are trapped at the bottom of the potential UU at any time, which means that the system becomes insulating. In cases (ii) and (iii), conversely, the system undergoes repeated carrier localization and delocalization when the amplitude of UU exceeds and falls below εF\varepsilon_{F}, respectively, because of the time dependence of UU. To investigate the longitudinal transport properties in the latter cases, we use semiclassical Boltzmann transport theory. In the following discussion, we focus on the second case and omit the VVC for simplicity, which corresponds to the assumption that R1R\gg 1. This simplification does not affect the main results given below because the effect of the VVC is negligible due to the condition |lω0|εF|l\omega_{0}|\ll\varepsilon_{F}. As shown in Eq. (9), although there is a phase difference between the pVZC and the VVC, their qualitative effects on the longitudinal transport properties are almost the same. Therefore, the results are also applicable to the third case.

The semiclassical Boltzmann kinetic equation is given by

fξt+𝒓˙fξ𝒓+𝒑˙fξ𝒑=(fξt)coll,\displaystyle\frac{\partial f_{\xi}}{\partial t}+\dot{\bm{r}}\cdot\frac{\partial f_{\xi}}{\partial\bm{r}}+\dot{\bm{p}}\cdot\frac{\partial f_{\xi}}{\partial\bm{p}}=\left(\frac{\partial f_{\xi}}{\partial t}\right)_{\mathrm{coll}}, (11)

where fξf_{\xi} is the distribution function for the valleys K+K_{+} and KK_{-}, and the equations of motion are given by

𝒓˙\displaystyle\dot{\bm{r}} =εξ,𝒑𝒑𝒑˙×𝛀ξ,\displaystyle=\frac{\partial\varepsilon_{\xi,\bm{p}}}{\partial\bm{p}}-\dot{\bm{p}}\times\bm{\Omega}_{\xi}, (12)
𝒑˙\displaystyle\dot{\bm{p}} =e𝑬+m𝒓˙×𝝎e𝒓˙×ξ𝑩s,\displaystyle=-e\bm{E}+m\dot{\bm{r}}\times\bm{\omega}-e\dot{\bm{r}}\times\xi\bm{B}^{s}, (13)

where εξ,𝒑\varepsilon_{\xi,\bm{p}} is the energy band in the presence of the pVZC, 𝛀ξ=𝒑×iuξ,𝒑|𝒑|uξ,𝒑\bm{\Omega}_{\xi}=\hbar\nabla_{\bm{p}}\times i\langle u_{\xi,\bm{p}}|\nabla_{\bm{p}}|u_{\xi,\bm{p}}\rangle is the Berry curvature for each valley, and |uξ,𝒑|u_{\xi,\bm{p}}\rangle is the periodic part of the Bloch wave function. The second term in Eq. (12) is the anomalous velocity Xiao et al. (2010), the second term in Eq. (13) is the Coriolis force that originates from the vorticity field of the lattice Matsuo et al. (2017), and the third term in Eq. (13) is the valley-dependent Lorentz force that originates from the pseudo-magnetic field.

Next, we consider the charge transport along the xx-direction in the presence of the external electric field 𝑬=(Ex,0,0)\bm{E}=(E_{x},0,0). We use the following relaxation time approximation:

(fξt)coll=fξfξ(0)τ,\displaystyle\left(\frac{\partial f_{\xi}}{\partial t}\right)_{\mathrm{coll}}=-\frac{f_{\xi}-f^{(0)}_{\xi}}{\tau}, (14)

where τ\tau is the valley independent relaxation time and fξ(0)=1/(1+exp[(εξ,𝒑εF)/kBT])f^{(0)}_{\xi}=1/(1+\exp[(\varepsilon_{\xi,\bm{p}}-\varepsilon_{F})/k_{B}T]) is the Fermi distribution function. The above expression of the relaxation time approximation is based on the three assumptions. First, the valley quantum number is treated as a conserved quantity. This assumption is reasonable if the sample is sufficiently clean that the effect of the atomic scale scatterers is negligible and there is no intervalley scattering process. In addition, the SAW rarely leads to intervalley scattering because the wavelength of the SAW is much larger than the lattice constant. Next, the relaxation time approximation is valid even in the presence of the time-dependent potential because the SAW period is much larger than the lifetime of the electron. In our setup, the SAW period is about a nanosecond, and the lifetime of the electron would be about a picosecond. Finally, the relaxation time is independent of the valley quantum number. This assumption is reasonable if the scatterers have no characteristic feature that breaks the valley degeneracy.

Assuming a weak vorticity field and a weak pseudo-magnetic field (i.e., ω0m|Ωξ,z|1\omega_{0}m|\Omega_{\xi,z}|\ll 1, ω0τ1\omega_{0}\tau\ll 1, and ωcτ1\omega_{c}\tau\ll 1 with ωc=eB0s/m\omega_{c}=eB^{s}_{0}/m^{\ast}), the anomalous velocity, the Coriolis force, and the Lorentz force then only give corrections for the longitudinal conductivity and have almost no effect on the qualitative behavior of the longitudinal conductivity. Based on this assumption, we can omit these three terms. Note here that there is no Hall voltage in the current setup because the net vorticity field is zero. The SAW dynamics are much slower than the dynamics of the electrons, which means that the adiabatic approximation is valid and it can be assumed that the steady state is achieved at each moment. Finally, the charge current can be calculated by deriving the steady state of the distribution function at each moment and using snapshot energy bands. We define δfξ\delta f_{\xi} as fξ=fξ(0)+δfξf_{\xi}=f^{(0)}_{\xi}+\delta f_{\xi}, and δfξ\delta f_{\xi} is given by

δfξ=τeExvx(fξεξ,𝒑),\displaystyle\delta f_{\xi}=\tau eE_{x}v_{x}\left(\frac{\partial f_{\xi}}{\partial\varepsilon_{\xi,\bm{p}}}\right), (15)

where vx=εξ,𝒑/pxv_{x}=\partial\varepsilon_{\xi,\bm{p}}/\partial p_{x}. Using δfξ\delta f_{\xi}, the charge current at T=0T=0 can be given by

jx=ξ=±1e2τExd𝒑(2π)2vx2δ(εFεξ,𝒑).\displaystyle j_{x}=\sum_{\xi=\pm 1}e^{2}\tau E_{x}\int\frac{d\bm{p}}{(2\pi\hbar)^{2}}v_{x}^{2}\delta(\varepsilon_{F}-\varepsilon_{\xi,\bm{p}}). (16)
Refer to caption
Figure 2: (a)-(c) Snapshot energy bands at several values of ωt\omega t. (d) Charge current as a function of both εF\varepsilon_{F} and ωt\omega t. (e) Charge current as a function of ωt\omega t at several values of εF\varepsilon_{F}. We set |lγB0s|/2=1.5ε0|l\gamma B^{s}_{0}|/2=1.5\varepsilon_{0}, where ε0=2q2/2m\varepsilon_{0}=\hbar^{2}q^{2}/2m^{\ast}. The unit of jxj_{x} is given by j0=(n0e2τ/m)Exj_{0}=(n_{0}e^{2}\tau/m^{\ast})E_{x} with carrier density n0=ε0m/2π2n_{0}=\varepsilon_{0}m^{\ast}/2\pi\hbar^{2}.

Figure 2 (a)-(c) show snapshot energy bands as a function of pxp_{x} with a fixed py=0p_{y}=0 at several values of ωt\omega t. The pVZC is treated as a spatially periodic potential in each case, which means that the energy gap opens and the group velocity decreases as the amplitude BzsB_{z}^{s} increases. Therefore, the charge current is suppressed by the SAW. Figure 2 (d) shows the charge current as a function of ωt\omega t and the Fermi energy εF\varepsilon_{F}. At a fixed εF\varepsilon_{F}, the charge current changes with the period π\pi and reaches a maximum value at ωt=Nπ,(N=0,±1,)\omega t=N\pi,\ (N=0,\pm 1,\cdots). The charge current is almost zero around ωt=(N+1/2)π,(N=0,±1,)\omega t=(N+1/2)\pi,\ (N=0,\pm 1,\cdots) when the Fermi energy is sufficiently small in comparison to the pVZC. Figure 2 (e) shows the charge current as a function of ωt\omega t at several fixed εF\varepsilon_{F}. One can see that the sharp pulsed current is generated when εF|lγB0s|\varepsilon_{F}\ll|l\gamma B_{0}^{s}|. At a finite temperature, the pulsed current is subject to thermal broadening. The sharp pulsed current can still be generated in a condition that the thermal broadening energy is much smaller than the pVZC (i.e., kBT|lγB0s|k_{B}T\ll|l\gamma B_{0}^{s}|). Using the parameters for graphene with energy gap Δ=0.1eV\Delta=0.1\ \mathrm{eV} Zhou et al. (2007, 2008), the pVZC is estimated as |lγB0s/2| 30meV|l\gamma B_{0}^{s}/2|\sim\ 30\mathrm{meV}, which is the same order of the thermal broadening energy at room temperature. Therefore, the condition to generate the sharp pulsed current is expected to be experimentally feasible. Within the quadratic approximation, the Fermi energy εF\varepsilon_{F} is proportional to the carrier density nn (i.e., εF=2π2n/m\varepsilon_{F}=2\pi\hbar^{2}n/m^{\ast}), so that one can replace the Fermi energy in Fig. 2 with the carrier density by multiplying the factor m/2π2m^{\ast}/2\pi\hbar^{2}.

To summarize the discussion of the transport properties above, when the Fermi energy is sufficiently smaller than both the pVZC and the VVC, the system becomes insulating, but when the Fermi energy is larger than either the pVZC or the VVC, the conductivity is then suppressed and a pulsed current can be generated, as listed in Table. 1.

V Broadband microwave sources

The charge current obtained above can be divided into an AC component and a DC component using the form jx(t)=jxac(t)+jxdcj_{x}(t)=j^{\mathrm{ac}}_{x}(t)+j^{\mathrm{dc}}_{x}, with jxdc=jx(π/2ω)j^{\mathrm{dc}}_{x}=j_{x}(\pi/2\omega). The DC component jxdcj^{\mathrm{dc}}_{x} corresponds to the minimum value of the charge current and becomes finite when εF/ε00.54\varepsilon_{F}/\varepsilon_{0}\gtrsim 0.54. The oscillating current jxac(t)j^{\mathrm{ac}}_{x}(t) can be used for broadband and tunable microwave sources, which is useful for device applications Hu et al. (2021). A schematic image of the microwave radiation in our setup is shown in Fig. 3(a).

Figure 3 (b) shows the Fermi energy dependence of the full width at half maximum (FWHM) value of the pulse. When the Fermi energy increases, the FWHM also increases, and the range of Fourier components decreases. There is a kink structure of the FWHM at εF/ε00.54\varepsilon_{F}/\varepsilon_{0}\approx 0.54 because jxdcj^{\mathrm{dc}}_{x} becomes finite and increases above this point.

The averaged radiation intensity of the microwave in the upper half space II is given by

I=Z02ϵ+ϵ++ϵ0π/ωd(ωt)π[jxac(t)]2,\displaystyle I=\frac{Z_{0}}{2}\frac{\sqrt{\epsilon_{+}}}{\sqrt{\epsilon_{+}}+\sqrt{\epsilon_{-}}}\int^{\pi/\omega}_{0}\frac{d(\omega t)}{\pi}\left[j^{\mathrm{ac}}_{x}(t)\right]^{2}, (17)

where Z0377ΩZ_{0}\approx 377\ \Omega is the characteristic impedance of free space, and ϵ+\epsilon_{+} and ϵ\epsilon_{-} are permitivity of free space and substrate, respectively. Figure 3 (c) shows the averaged intensity as a function of the Fermi energy. The averaged radiation intensity becomes a local maximum at εF/ε00.54\varepsilon_{F}/\varepsilon_{0}\approx 0.54 above which jxdcj^{\mathrm{dc}}_{x} becomes finite. Setting the charge current as j01A/mj_{0}\sim 1\ \mathrm{A}/\mathrm{m} and the permitivity as ϵ±1\epsilon_{\pm}\sim 1, the averaged radiation intensity is estimated as I0.1mW/cm2I\sim 0.1\ \mathrm{mW}/\mathrm{cm}^{2} in the energy range considered here.

Refer to caption
Figure 3: (a) Schematic image of the microwave radiation produced by jxac(t)j_{x}^{\mathrm{ac}}(t). (b) Pulse FWHM as a function of the Fermi energy. The inset shows the Fourier components of the pulse at εF/ε0=0.35\varepsilon_{F}/\varepsilon_{0}=0.35. (c) Averaged radiation intensity as a function of the Fermi energy. The unit is set as I0=Z02ϵ+ϵ++ϵj02I_{0}=\frac{Z_{0}}{2}\frac{\sqrt{\epsilon_{+}}}{\sqrt{\epsilon_{+}}+\sqrt{\epsilon_{-}}}j_{0}^{2}.

VI Conclusion

In this work, we have presented an investigation of the low-energy effective Hamiltonian of 2D Dirac materials in the presence of the dynamic lattice distortions. We have derived the valley-vorticity coupling (VVC), which is regarded as an inertial effect. In addition to the VVC, we also incorporated the strain-induced pseudo-magnetic field into the analysis. We have proposed mechanisms for both valley transfer and valley-dependent carrier localization using an SAW.

We have also investigated the electronic transport properties in the presence of the SAW. The conductivity along the direction of propagation of the SAW is suppressed by the dynamic lattice distortions. Periodic modulation of the conductivity results in a pulsed charge current when an electrostatic field is applied along the SAW. This pulsed charge current act as a broadband microwave source and can be used for optoelectronic device applications. These results pave the way toward the realization of next-generation valleytronics device applications using the SAW devices.

Acknowledgements.— The authors are grateful to Y. Nozaki and A. Yamakage providing valuable comments. D.O. is funded by the President’s PhD Scholarships at Imperial College London. This work was supported by the Priority Program of Chinese Academy of Sciences under Grant No. XDB28000000, and by JSPS KAKENHI for Grants (Nos. JP20H01863 and JP21H04565) from MEXT, Japan.

Appendix

VI.1 Inertial effect due to dynamic lattice distorsions

Let us consider the inertial effects due to dynamic lattice distortions such as elastic motion in the presence of SAWs. Previous studies have shown that the strain field caused by these lattice distortions leads to an strain-induced gauge field Suzuura and Ando (2002); Manes (2007); Vozmediano et al. (2010). In the following subsequent discussion, we omit the the strain-induced gauge field and focus on the inertial effect.

The generally covariant Dirac Lagrangian, which governs the dynamics of the spin-1/2 particle in curved spacetime, is Brill and Wheeler (1957); Hehl et al. (1976)

=Ψ¯[iγa^ea^(x)μ(pμ𝒜μ)mc]Ψ,\displaystyle\mathcal{L}=\bar{\Psi}\left[i\gamma^{\hat{a}}e_{\hat{a}}{}^{\mu}(x)\left(p_{\mu}-\mathcal{A}_{\mu}\right)-mc\right]\Psi, (A.1)

where mm and cc represent the mass of an electron and the speed of light, respectively. The spin connection 𝒜μ\mathcal{A}_{\mu} is given by 𝒜μ=iωa^b^μ[γa^,γb^]/8\mathcal{A}_{\mu}=i\hbar\omega_{\hat{a}\hat{b}\mu}[\gamma^{\hat{a}},\gamma^{\hat{b}}]/8, where \hbar represents the Planck’s constant, ωa^b^\omega_{\hat{a}\hat{b}} is related to the tetrad ea^e_{\hat{a}} as dea^=ωa^b^eb^de_{\hat{a}}=\omega_{\hat{a}}{}^{\hat{b}}\wedge e_{\hat{b}}, and γa^\gamma^{\hat{a}} is the gamma matrix. Note that μ\mu in Eq. (A.1) represents the vector indices in curved spacetime and the hatted indices a^,b^\hat{a},\hat{b} represent the local Lorentz indices.

We can describe the relationship between a local rest frame on the lattice and an inertial frame in terms of the lattice displacement vector 𝒖\bm{u} and the lattice velocity field 𝒖˙(x)(|𝒖˙|/c1)\dot{\bm{u}}(x)\,\,(|\dot{\bm{u}}|/c\ll 1), d𝒓=d𝒓+𝒖˙(x)dtd\bm{r}^{\prime}=d\bm{r}+\dot{\bm{u}}(x)dt. We can explicitly write the tetrad e0^=01,e0^=iu˙i/c,ej^=00,ej^=iδji.e_{\hat{0}}{}^{0}=1,e_{\hat{0}}{}^{i}=-\dot{u}^{i}/c,e_{\hat{j}}{}^{0}=0,e_{\hat{j}}{}^{i}=\delta^{i}_{j}. As a result, the generally covariant Dirac Lagrangian (A.1) leads to the Dirac Hamiltonian in the local rest frame Matsuo et al. (2017)

HD\displaystyle H_{\rm D} =\displaystyle= βmc2+(c𝜶𝒖˙(x))𝒑𝚺𝝎(x)2,\displaystyle\beta mc^{2}+(c{\bm{\alpha}}-\dot{\bm{u}}(x))\cdot\bm{p}-{\bm{\Sigma}}\cdot\frac{{\bm{\omega}}(x)}{2}, (A.2)

where β=γ0\beta=\gamma^{0} and αi=γ0γi\alpha_{i}=\gamma^{0}\gamma^{i} are the Dirac matrices, and Σa=2ϵabc[γb,γc]\Sigma_{a}=\frac{\hbar}{2}\epsilon_{abc}[\gamma^{b},\gamma^{c}] is the spin operator, and 𝝎(x)=×𝒖˙(x){\bm{\omega}}(x)=\nabla\times\dot{\bm{u}}(x) is the vorticity field of the lattice distortion. Here, two inertial effects on the spinor field due to the lattice velocity field 𝒖˙(x)\dot{\bm{u}}(x) are included in the Hamiltonian. First, the velocity operator of the Dirac spinor in the inertial frame c𝜶c{\bm{\alpha}} is replaced with c𝜶𝒖˙(x)c{\bm{\alpha}}-\dot{\bm{u}}(x). Second, we include the spin-vorticity coupling 𝚺𝝎(x)/2-{\bm{\Sigma}}\cdot{\bm{\omega}}(x)/2.

The lowest order of the Foldy-Wouthuysen-Tani transformation Foldy and Wouthuysen (1950); Tani (1951) for Eq. (A.2) leads to the Pauli-Schrödinger equation for an electron in the local rest frame on the lattice

iψt=Hψ,H=(𝒑+e𝑨v(x))22m𝒔𝝎(x)2,\displaystyle i\hbar\frac{\partial\psi}{\partial t}=H^{\prime}\psi,\,\,H^{\prime}=\frac{(\bm{p}+e\bm{A}^{v}(x))^{2}}{2m}-\bm{s}\cdot\frac{{\bm{\omega}}(x)}{2}, (A.3)

where 𝒔\bm{s} is the spin angular momentum of the electron, which obeys the commutation relation [si,sj]=iϵijksk[s_{i},s_{j}]=i\hbar\epsilon_{ijk}s_{k}. This result shows that the inertial effects can be introduced into the conventional Hamiltonian in an inertial frame via the following two emergent gauge fields: the U(1) potential

𝑨v(x)=me𝒖˙(x)=me(u˙x(x),u˙y(x),u˙z(x)),\displaystyle\bm{A}^{v}(x)=-\frac{m}{e}\dot{\bm{u}}(x)=-\frac{m}{e}\left(\dot{u}_{x}(x),\dot{u}_{y}(x),\dot{u}_{z}(x)\right), (A.4)

and the spin-vorticity coupling, or the SU(2) scalar potential, 𝒔𝝎(x)/2-\bm{s}\cdot{\bm{\omega}}(x)/2. Consequently, the inertial effect of the U(1) potential 𝑨v\bm{A}^{v} is incorporated into the low energy effective Hamiltonian by the Peierls substitution 𝒑𝒑+e𝑨v\bm{p}\to\bm{p}+e\bm{A}^{v} as the conventional electromagnetic field. In the main text, we omit the spin-vorticity coupling, based on the assumption that its effect is small when compared with the orbital effect that originates from the Peierls substitution. This assumption is valid for the 2D Dirac materials because the intrinsic orbital angular momentum is larger than the spin angular momentum. Adding the periodic potential of the lattice, which is omitted in the above discussion, gives the non-relativistic Hamiltonian Eq. (1) used in the main text.

VI.2 Surface acoustic wave: Love wave

In this section, we explain the Love wave, a horizontally polarized (transverse) surface acoustic wave. The equation of motion for an isotropic elastic body is given by Love (1911); Officer (1974); Landau and Lifshitz (1986)

2𝒖t2=ct2Δ𝒖+(c2ct2)(𝒖),\displaystyle\frac{\partial^{2}\bm{u}}{\partial t^{2}}=c_{t}^{2}\Delta\bm{u}+(c_{\ell}^{2}-c_{t}^{2})\nabla(\nabla\cdot\bm{u}), (B.1)

where 𝒖\bm{u} is the displacement vector, and ctc_{t} and cc_{\ell} are the velocities of the transverse and longitudinal waves, respectively. In the following, we focus on transverse waves oscillate in the yy-direction (i.e., 𝒖=(0,uy,0)\bm{u}=(0,u_{y},0)). From the transversality condition (i.e., 𝒖=0\nabla\cdot\bm{u}=0), the second term on the right-hand side vanishes.

We consider a slab medium M1M_{1} with a thickness of HH on a semi-infinite medium M2M_{2}, as shown in Fig. 4. We assume solutions of Eq. (B.1) of the form

ua(x,z,t)=ha(z)ei(kxωt),\displaystyle u_{a}(x,z,t)=h_{a}(z)e^{i(kx-\omega t)}, (B.2)

where ua(x,z,t)(a=1,2)u_{a}(x,z,t)~{}(a=1,2) is the yy-component of the displacement vector in medium MaM_{a}. Substituting Eq. (B.2) into Eq. (B.1), we obtain the differential equation for ha(z)h_{a}(z)

(k2ω2ca2)ha(z)=2ha(z)z2,\displaystyle\left(k^{2}-\frac{\omega^{2}}{c_{a}^{2}}\right)h_{a}(z)=\frac{\partial^{2}h_{a}(z)}{\partial z^{2}}, (B.3)

where cac_{a} is the velocity of transverse waves in each medium. Since we consider surface waves, we assume a solution which vanishes at negative infinity (i.e., limzh2(z)=0\lim_{z\to-\infty}h_{2}(z)=0). Therefore, the solutions ha(z)h_{a}(z) are written as

h1(z)=Aeiqz+Beiqz,\displaystyle h_{1}(z)=Ae^{iqz}+Be^{-iqz}, (B.4)
h2(z)=Ceκz,\displaystyle h_{2}(z)=Ce^{\kappa z}, (B.5)

where qq and κ\kappa are positive numbers given by

q=ω2c12k2,\displaystyle q=\sqrt{\frac{\omega^{2}}{c_{1}^{2}}-k^{2}}, (B.6)
κ=k2ω2c22.\displaystyle\kappa=\sqrt{k^{2}-\frac{\omega^{2}}{c_{2}^{2}}}. (B.7)

The velocities c1c_{1} and c2c_{2} have to satisfy c1<c2c_{1}<c_{2} since both qq and κ\kappa are positive numbers.

The values of qq and κ\kappa are determined by finding solutions under appropriate boundary conditions. Here, we apply the free surface boundary condition at z=0z=0, where the stress on the surface vanishes, and the continuities of displacement and stress at the boundary between M1M_{1} and M2M_{2}:

h1(z)z|z=0=0,\displaystyle\frac{\partial h_{1}(z)}{\partial z}\Big{|}_{z=0}=0, (B.8)
h1(H)=h2(H),\displaystyle h_{1}(-H)=h_{2}(-H), (B.9)
ρ1c12h1(z)z|z=H=ρ2c22h2(z)z|z=H,\displaystyle\rho_{1}c_{1}^{2}\frac{\partial h_{1}(z)}{\partial z}\Big{|}_{z=-H}=\rho_{2}c_{2}^{2}\frac{\partial h_{2}(z)}{\partial z}\Big{|}_{z=-H}, (B.10)

where ρa\rho_{a} is the density of MaM_{a}. Substituting Eq. (B.4) and (B.5) into the above equations, we obtain an equation system for the coefficients A,B,A,B, and CC,

AB=0,\displaystyle A-B=0, (B.11)
AeiqH+BeiqH=CeκH,\displaystyle Ae^{-iqH}+Be^{iqH}=Ce^{-\kappa H}, (B.12)
iρ1c12q(AeiqHBeiqH)=ρ2c22κCeκH.\displaystyle i\rho_{1}c_{1}^{2}q\left(Ae^{-iqH}-Be^{iqH}\right)=\rho_{2}c_{2}^{2}\kappa Ce^{-\kappa H}. (B.13)

To obtain nontrivial solutions, we should have

κq=ρ1c12ρ2c22tan(qH).\displaystyle\frac{\kappa}{q}=\frac{\rho_{1}c_{1}^{2}}{\rho_{2}c_{2}^{2}}\tan\left(qH\right). (B.14)

Substituting Eq. (B.6) and (B.7) into Eq. (B.14) to find ω\omega and kk that satisfy the boundary conditions, we obtain the dispersion relation. Figure 5 (a) and (b) show the dispersion relation and the amplitude of the Love waves, respectively. The phase velocity of the Love waves csc_{s} is also obtained from the dispersion relation. Using the obtained displacement uau_{a} and phase velocity csc_{s}, the intensity of the Love waves IsI_{s} is given by

Is=H0𝑑z12csρ1|u˙1|2+H𝑑z12csρ2|u˙2|2.\displaystyle I_{s}=\int^{0}_{-H}dz\frac{1}{2}c_{s}\rho_{1}|\dot{u}_{1}|^{2}+\int^{-H}_{-\infty}dz\frac{1}{2}c_{s}\rho_{2}|\dot{u}_{2}|^{2}. (B.15)

The 2D Dirac material follows the excited Love wave on the substrate, which dynamically distorts the lattice of the 2D material. By superimposing of the traveling-wave-type solutions (B.2), we can get the dynamic lattice distortion of the standing-wave type, which is employed in the main text.

Refer to caption
Figure 4: The slab medium M1M_{1} of thickness HH is stacked on the semi-infinite medium M2M_{2}.
Refer to caption
Figure 5: (a) The red, green, and blue curves represent the dispersion relation of the Love waves. The dotted lines are ω=c1k\omega=c_{1}k and ω=c2k\omega=c_{2}k. We set the parameters as indicated in the figure legends. (b) The amplitude of the Love wave ha(z)h_{a}(z) at k=30μm1k=30\mu\mathrm{m}^{-1}, where a=1a=1 for 1<z<0-1<z<0 and a=2a=2 for z<1z<-1. The red, green, and blue curves correspond to the red, green, and blue branches in the left panel, respectively.

VI.3 Microwave radiation

In our setup, we have an oscillating electric current in the xx-direction; hence, the magnetic (electric) field emitted by the current should oscillate in the yy- (xx-) direction. Integrating Maxwell’s equations gives the electric field continuity and the magnetic field discontinuity at boundaries. In the frequency domain, we can write

limδh0[Ex(z,ω)]z=δhz=+δh=0,\displaystyle\lim_{\delta h\rightarrow 0}\big{[}E_{x}(z,\omega)\big{]}_{z=-\delta h}^{z=+\delta h}=0, (C.1)
limδh0[Hy(z,ω)]z=δhz=+δh=Jx(ω),\displaystyle\lim_{\delta h\rightarrow 0}\big{[}H_{y}(z,\omega)\big{]}_{z=-\delta h}^{z=+\delta h}=-J_{x}(\omega), (C.2)

where the electric current on the 2D material JJ is given in [A/m]\mathrm{[A/m]}. We have the current but no incoming electromagnetic waves. Only outgoing waves are present, e.g.,

Ex(z,ω)\displaystyle E_{x}(z,\omega) ={Ex+(ω)e+i(ω/v+)zz>0,Ex(ω)ei(ω/v)zz<0,\displaystyle=\begin{cases}E_{x}^{+}(\omega)e^{+i(\omega/v_{+})z}&z>0,\\ E_{x}^{-}(\omega)e^{-i(\omega/v_{-})z}&z<0,\end{cases} (C.3)

where v±=c/ϵ±v_{\pm}=c/\sqrt{\epsilon_{\pm}} is the speed of light in each medium with each permittivity ϵ±\epsilon_{\pm}. The magnetic field is associated with the electric field via the characteristic impedance Kong (1986),

Ex±(ω)=±Z0ϵ±Hy±(ω),\displaystyle E_{x}^{\pm}(\omega)=\pm\frac{Z_{0}}{\sqrt{\epsilon_{\pm}}}H_{y}^{\pm}(\omega), (C.4)

where Z0μ0/ϵ0377ΩZ_{0}\equiv\sqrt{\mu_{0}/\epsilon_{0}}\approx 377\ \Omega is the characteristic impedance of free space. We rearrange Eqs. (C.1, C.2) in a matrix form,

(11ϵ+ϵ)(Ex+(ω)Ex(ω))=(0Z0Jx(ω)).\displaystyle\begin{pmatrix}1&-1\\ \sqrt{\epsilon_{+}}&\sqrt{\epsilon_{-}}\end{pmatrix}\begin{pmatrix}E_{x}^{+}(\omega)\\ E_{x}^{-}(\omega)\end{pmatrix}=\begin{pmatrix}0\\ Z_{0}J_{x}(\omega)\end{pmatrix}. (C.5)

Inverting, we can find

Ex±(ω)\displaystyle E_{x}^{\pm}(\omega) =Z0Jx(ω)ϵ++ϵ,Hy±(ω)=±ϵ±Jx(ω)ϵ++ϵ,\displaystyle=\frac{Z_{0}J_{x}(\omega)}{\sqrt{\epsilon_{+}}+\sqrt{\epsilon_{-}}},\quad H_{y}^{\pm}(\omega)=\frac{\pm\sqrt{\epsilon_{\pm}}J_{x}(\omega)}{\sqrt{\epsilon_{+}}+\sqrt{\epsilon_{-}}}, (C.6)

The radiation intensities in the upper and lower media are calculated from the Poynting vector as following:

dw±dω\displaystyle\frac{\mathrm{d}w_{\pm}}{\mathrm{d}\omega} =±12Re[{Ex±(ω)}Hy±(ω)],\displaystyle=\frac{\pm 1}{2}\operatorname{Re}\left[\{E_{x}^{\pm}(\omega)\}^{*}H_{y}^{\pm}(\omega)\right], (C.7)
=12ϵ±ϵ++ϵZ0|Jx(ω)|2.\displaystyle=\frac{1}{2}\frac{\sqrt{\epsilon_{\pm}}}{\sqrt{\epsilon_{+}}+\sqrt{\epsilon_{-}}}Z_{0}|J_{x}(\omega)|^{2}. (C.8)

This is the radiation intensity between an frequency interval [ω,ω+dω][\omega,\omega+\mathrm{d}\omega]. By integrating this quantity, we can obtain averaged radiation intensity. Remind that the integration over the frequency is equivalent to taking the average in the time domain (Eq. (17) in the main text).

References

  • Schaibley et al. (2016) J. R. Schaibley, H. Yu, G. Clark, P. Rivera, J. S. Ross, K. L. Seyler, W. Yao,  and X. Xu, Nature Reviews Materials 1, 1 (2016).
  • Krasnok et al. (2018) A. Krasnok, S. Lepeshov,  and A. Alú, Optics express 26, 15972 (2018).
  • Vitale et al. (2018) S. A. Vitale, D. Nezich, J. O. Varghese, P. Kim, N. Gedik, P. Jarillo-Herrero, D. Xiao,  and M. Rothschild, Small 14, 1801483 (2018).
  • Semenoff (1984) G. W. Semenoff, Physical Review Letters 53, 2449 (1984).
  • Xiao et al. (2007) D. Xiao, W. Yao,  and Q. Niu, Physical review letters 99, 236809 (2007).
  • Xiao et al. (2012) D. Xiao, G.-B. Liu, W. Feng, X. Xu,  and W. Yao, Physical review letters 108, 196802 (2012).
  • Xiao et al. (2010) D. Xiao, M.-C. Chang,  and Q. Niu, Reviews of modern physics 82, 1959 (2010).
  • Cao et al. (2012) T. Cao, G. Wang, W. Han, H. Ye, C. Zhu, J. Shi, Q. Niu, P. Tan, E. Wang, B. Liu, et al., Nature communications 3, 1 (2012).
  • Mak et al. (2012) K. F. Mak, K. He, J. Shan,  and T. F. Heinz, Nature nanotechnology 7, 494 (2012).
  • Zeng et al. (2012) H. Zeng, J. Dai, W. Yao, D. Xiao,  and X. Cui, Nature nanotechnology 7, 490 (2012).
  • Sallen et al. (2012) G. Sallen, L. Bouet, X. Marie, G. Wang, C. Zhu, W. Han, Y. Lu, P. Tan, T. Amand, B. Liu, et al., Physical Review B 86, 081301 (2012).
  • Guddala et al. (2021) S. Guddala, Y. Kawaguchi, F. Komissarenko, S. Kiriushechkina, A. Vakulenko, K. Chen, A. Alù, V. M. Menon,  and A. B. Khanikaev, Nature Communications 12, 1 (2021).
  • Li et al. (2014) Y. Li, J. Ludwig, T. Low, A. Chernikov, X. Cui, G. Arefe, Y. D. Kim, A. M. Van Der Zande, A. Rigosi, H. M. Hill, et al., Physical review letters 113, 266804 (2014).
  • Aivazian et al. (2015) G. Aivazian, Z. Gong, A. M. Jones, R.-L. Chu, J. Yan, D. G. Mandrus, C. Zhang, D. Cobden, W. Yao,  and X. Xu, Nature Physics 11, 148 (2015).
  • MacNeill et al. (2015) D. MacNeill, C. Heikes, K. F. Mak, Z. Anderson, A. Kormányos, V. Zólyomi, J. Park,  and D. C. Ralph, Physical review letters 114, 037401 (2015).
  • Srivastava et al. (2015) A. Srivastava, M. Sidler, A. V. Allain, D. S. Lembke, A. Kis,  and A. Imamoğlu, Nature Physics 11, 141 (2015).
  • Everitt et al. (2011) C. F. Everitt, D. DeBra, B. Parkinson, J. Turneaure, J. Conklin, M. Heifetz, G. Keiser, A. Silbergleit, T. Holmes, J. Kolodziejczak, et al., Physical Review Letters 106, 221101 (2011).
  • Einstein and de Haas (1915) A. Einstein and W. de Haas, KNAW Proc. 18 I, 696 (1915).
  • Barnett (1915) S. J. Barnett, Physical Review 6, 239 (1915).
  • Barnett (1935) S. J. Barnett, Reviews of Modern Physics 7, 129 (1935).
  • Chudo et al. (2014) H. Chudo, M. Ono, K. Harii, M. Matsuo, J. Ieda, R. Haruki, S. Okayasu, S. Maekawa, H. Yasuoka,  and E. Saitoh, Applied Physics Express 7, 063004 (2014).
  • Imai et al. (2018) M. Imai, Y. Ogata, H. Chudo, M. Ono, K. Harii, M. Matsuo, Y. Ohnuma, S. Maekawa,  and E. Saitoh, Applied Physics Letters 113, 052402 (2018).
  • Imai et al. (2019) M. Imai, H. Chudo, M. Ono, K. Harii, M. Matsuo, Y. Ohnuma, S. Maekawa,  and E. Saitoh, Applied Physics Letters 114, 162402 (2019).
  • Takahashi et al. (2016) R. Takahashi, M. Matsuo, M. Ono, K. Harii, H. Chudo, S. Okayasu, J. Ieda, S. Takahashi, S. Maekawa,  and E. Saitoh, Nature Physics 12, 52 (2016).
  • Takahashi et al. (2020) R. Takahashi, H. Chudo, M. Matsuo, K. Harii, Y. Ohnuma, S. Maekawa,  and E. Saitoh, Nature Communications 11, 3009 (2020).
  • Kazerooni et al. (2020) H. T. Kazerooni, A. Thieme, J. Schumacher,  and C. Cierpka, Physical Review Applied 14, 014002 (2020).
  • Kazerooni et al. (2021) H. T. Kazerooni, G. Zinchenko, J. Schumacher,  and C. Cierpka, Physical Review Fluids 6, 043703 (2021).
  • Kobayashi et al. (2017) D. Kobayashi, T. Yoshikawa, M. Matsuo, R. Iguchi, S. Maekawa, E. Saitoh,  and Y. Nozaki, Physical review letters 119, 077202 (2017).
  • Harii et al. (2019) K. Harii, Y.-J. Seo, Y. Tsutsumi, H. Chudo, K. Oyanagi, M. Matsuo, Y. Shiomi, T. Ono, S. Maekawa,  and E. Saitoh, Nat. Commun. 10, 1 (2019).
  • Okano et al. (2019) G. Okano, M. Matsuo, Y. Ohnuma, S. Maekawa,  and Y. Nozaki, Physical review letters 122, 217701 (2019).
  • Kurimune et al. (2020a) Y. Kurimune, M. Matsuo,  and Y. Nozaki, Physical review letters 124, 217205 (2020a).
  • Kurimune et al. (2020b) Y. Kurimune, M. Matsuo, S. Maekawa,  and Y. Nozaki, Physical Review B 102, 174413 (2020b).
  • Mori et al. (2020) K. Mori, M. Dunsmore, J. Losby, D. Jenson, M. Belov,  and M. Freeman, Phys. Rev. B 102, 054415 (2020).
  • Tateno et al. (2020) S. Tateno, G. Okano, M. Matsuo,  and Y. Nozaki, Physical Review B 102, 104406 (2020).
  • Tateno et al. (2021) S. Tateno, Y. Kurimune, M. Matsuo, K. Yamanoi,  and Y. Nozaki, Physical Review B 104, L020404 (2021).
  • Dornes et al. (2019) C. Dornes, Y. Acremann, M. Savoini, M. Kubli, M. J. Neugebauer, E. Abreu, L. Huber, G. Lantz, C. A. Vaz, H. Lemke, et al., Nature 565, 209 (2019).
  • Adamczyk et al. (2017) L. Adamczyk, J. Adkins, G. Agakishiev, M. Aggarwal, Z. Ahammed, N. Ajitanand, I. Alekseev, D. Anderson, R. Aoyama, A. Aparin, et al., Nature (London) 548 (2017).
  • Delsing et al. (2019) P. Delsing, A. N. Cleland, M. J. A. Schuetz, J. Knörzer, G. Giedke, J. Ignacio Cirac, K. Srinivasan, M. Wu, K. C. Balram, C. Bäuerle, T. Meunier, C. J. B. Ford, P. V. Santos, E. Cerda-Méndez, H. Wang, H. J. Krenner, E. D. S. Nysten, M. Weiß, G. R. Nash, L. Thevenard, C. Gourdon, P. Rovillain, M. Marangolo, J.-Y. Duquesne, G. Fischerauer, W. Ruile, A. Reiner, B. Paschke, D. Denysenko, D. Volkmer, A. Wixforth, H. Bruus, M. Wiklund, J. Reboud, J. M. Cooper, Y. Fu, M. S. Brugger, F. Rehfeldt,  and C. Westerhausen, J. Phys. D Appl. Phys. 52, 353001 (2019).
  • Matsuo et al. (2017) M. Matsuo, Y. Ohnuma,  and S. Maekawa, Physical Review B 96, 020401 (2017).
  • Suzuura and Ando (2002) H. Suzuura and T. Ando, Physical review B 65, 235412 (2002).
  • Manes (2007) J. L. Manes, Physical Review B 76, 045430 (2007).
  • Vozmediano et al. (2010) M. A. Vozmediano, M. Katsnelson,  and F. Guinea, Physics Reports 496, 109 (2010).
  • Schrieffer and Wolff (1966) J. R. Schrieffer and P. A. Wolff, Physical Review 149, 491 (1966).
  • Koshino and Ando (2010) M. Koshino and T. Ando, Physical Review B 81, 195431 (2010).
  • Koshino (2011) M. Koshino, Physical Review B 84, 125427 (2011).
  • Kalameitsev et al. (2019) A. Kalameitsev, V. Kovalev,  and I. Savenko, Physical review letters 122, 256801 (2019).
  • Sonowal et al. (2020) K. Sonowal, A. V. Kalameitsev, V. M. Kovalev,  and I. G. Savenko, Phys. Rev. B 102, 235405 (2020).
  • Nigge et al. (2019) P. Nigge, A. Qu, É. Lantagne-Hurtubise, E. Marsell, S. Link, G. Tom, M. Zonno, M. Michiardi, M. Schneider, S. Zhdanovich, et al., Science advances 5, eaaw5593 (2019).
  • Weiler et al. (2011) M. Weiler, L. Dreher, C. Heeg, H. Huebl, R. Gross, M. S. Brandt,  and S. T. Gönnenwein, Physical review letters 106, 117601 (2011).
  • Lee et al. (2016) J. Lee, K. F. Mak,  and J. Shan, Nature nanotechnology 11, 421 (2016).
  • Rostami et al. (2015) H. Rostami, R. Roldán, E. Cappelluti, R. Asgari,  and F. Guinea, Physical Review B 92, 195402 (2015).
  • Zhou et al. (2007) S. Y. Zhou, G.-H. Gweon, A. Fedorov, d. First, PN, W. De Heer, D.-H. Lee, F. Guinea, A. Castro Neto,  and A. Lanzara, Nature materials 6, 770 (2007).
  • Zhou et al. (2008) S. Zhou, D. Siegel, A. Fedorov, F. E. Gabaly, A. Schmid, A. Neto, D.-H. Lee,  and A. Lanzara, Nature Materials 7, 259 (2008).
  • Hu et al. (2021) H. Hu, X. Lin,  and Y. Luo, Prog. Electromagn. Res. 171, 75 (2021).
  • Brill and Wheeler (1957) D. R. Brill and J. A. Wheeler, Reviews of Modern Physics 29, 465 (1957).
  • Hehl et al. (1976) F. W. Hehl, P. Von der Heyde, G. D. Kerlick,  and J. M. Nester, Reviews of Modern Physics 48, 393 (1976).
  • Foldy and Wouthuysen (1950) L. L. Foldy and S. A. Wouthuysen, Physical Review 78, 29 (1950).
  • Tani (1951) S. Tani, Progress of Theoretical Physics 6, 267 (1951).
  • Love (1911) A. E. H. Love, Some problems of geodynamics (Cambridge University Press, 1911).
  • Officer (1974) C. B. Officer, Introduction to Theoretical Geophysics (Springer, 1974).
  • Landau and Lifshitz (1986) L. D. Landau and E. M. Lifshitz, Theory of Elasticity: Volume 7 (Theoretical Physics) (Butterworth-Heinemann, 1986).
  • Kong (1986) J. A. Kong, Electromagnetic wave theory (John Wiley & Sons, Inc., NewYork, 1986).