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

Control of a spin qubit in a lateral GaAs quantum dot based on symmetry of gating potential

Pavle Stipsić Faculty of Physics, University of Belgrade, Studentski trg 12, 11001 Belgrade, Serbia    Marko Milivojević NanoLab, QTP Center, Faculty of Physics, University of Belgrade, Studentski trg 12, 11001 Belgrade, Serbia Department of Theoretical Physics and Astrophysics, Faculty of Science, P. J. Šafárik University, Park Angelinum 9, 040 01 Košice, Slovak Republic
Abstract

We study the influence of quantum dot symmetry on the Rabi frequency and phonon-induced spin relaxation rate in a single-electron GaAs spin qubit. We find that anisotropic dependence on the magnetic field direction is independent of the choice of the gating potential. Also, we discover that relative orientation of the quantum dot, with respect to the crystallographic frame, is relevant in systems with 𝐂1v{\bf C}_{1{\rm v}}, 𝐂2v{\bf C}_{2{\rm v}}, or 𝐂n{\bf C}_{n} (n4rn\neq 4r) symmetry. To demonstrate the important impact of the gating potential shape on the spin qubit lifetime, we compare the effects of an infinite-wall equilateral triangle, square, and rectangular confinement with the known results for the harmonic potential. In the studied cases, enhanced spin qubit lifetime is revealed, reaching almost six orders of magnitude increase for the equilateral triangle gating.

PACS numbers: 81.07.Ta, 71.70.Ej, 72.10.Di, 76.30.−v, 31.15.Hz

preprint: AIP/123-QED

I Introduction

Every quantum two-level system can act as the quantum bit, a basic unit of quantum information processing NC10 ; BdV . Among different solid-state implementations of the qubit system LdV98 ; OTW+99 ; NPT99 ; YTB+12 , single-electron spin in a semiconductor quantum dot (QD) can be used to achieve the task. In order to manipulate spins of charge carriers embedded inside a semiconductor material electrically, through electric dipole spin resonance (EDSR) R60 , the presence of spin-orbit interaction (SOI) is obligatory.

Besides its positive effect in EDSR based schemes GBL06 ; BL07 ; NKN+07 ; R08 ; BSO+11 ; KSW+14 ; TYO+18 ; KLS19 ; SKT+19 , SOI enables the electron-phonon coupling mediated transitions between the qubit states KN00 ; PF05 ; S17 ; LLH+18 , affecting the spin qubit lifetime. To suppress the coupling to phonons, different approaches like the optimization of the QD design CBG+07 ; MSL16 or control of the system size CM07 were suggested. The observed anisotropy of the spin relaxation rate on the in-plane magnetic field orientation FAT05 offered another playground for fine-tuning of the spin qubit’s desired properties. In circular QDs, this is the only degree of freedom accessible in the optimization of the spin qubit, while for the elliptical confining potential SKS+14 ; MSL16 ; OS07 ; AMR+08 orientation of the QD potential with respect to the crystallographic frame can be used as the tuning parameter.

Evidently, different symmetry of the gating potential PRC15 is the main reason for the observed behavior. But to what extent can the potential symmetry alter the basic properties of the electrically controlled spin qubit? To address this question, we have performed a general analysis valid for the lateral GaAs QD system with 𝐂nv{\bf C}_{n{\rm v}} or 𝐂n{\bf C}_{n} symmetry of the gating potential. Besides the expected anisotropy on the magnetic field orientation, we were able to find potential symmetries for which the QD orientation with respect to the crystallographic frame can act as another control parameter of the spin qubit characteristics. With our theory, we offer a simple and efficient way to determine the impact of the gating potential on the Rabi frequency and spin relaxation rate. This is shown in the example of anisotropic and isotropic harmonic potential, as well as for the infinite-wall equilateral triangle, square, and rectangular potential.

This paper is organized as follows. In Section II we define a single-electron GaAs spin qubit model. In Section III we define the dipole moment of the electrically controlled spin qubit that describes both the Rabi frequency and SOI-induced spin relaxation rate mediated by acoustic phonons. In Section IV we present the main results of the paper: analytical expressions for the dipole moment in the case of the gating potential with 𝐂nv{\bf C}_{n{\rm v}} or 𝐂n{\bf C}_{n} symmetry. In Section V, to illustrate the impact of the gating potential on the spin qubit lifetime we use the obtained expressions to compare the influence of the harmonic confinement with an infinite-wall equilateral triangle, square, and rectangular potential. In Section VI we give our conclusions.

II Dynamics of the lateral QD

We start with the Hamiltonian describing the lateral dynamics of a single-electron in the GaAs material

H=H0+Hz+Hso=px2+py22m+V(x,y)+Hz+Hso,H=H_{0}+H_{\rm z}+H_{\rm so}=\frac{p_{x}^{2}+p_{y}^{2}}{2m^{*}}+V(x,y)+H_{\rm z}+H_{\rm so}, (1)

where pxp_{x} and pyp_{y} are the momentum operators, mm^{*} is the effective mass (m=0.067mem^{*}=0.067m_{e} for GaAs, mem_{e} is the electron mass), while V(x,y)V(x,y) is the gating potential used to localize the electron in a QD. In the lateral system, symmetries that can be present are the nn-fold rotational symmetry and the vertical mirror plane symmetry σv\sigma_{\rm v}. For simplicity, we assume that σv\sigma_{\rm v} coincides with the yzyz plane of the QD coordinate frame (see FIG. 1). Thus, we assume a general form of the orbital Hamiltonian H0H_{0} that has a 𝐂nv{\bf C}_{n{\rm v}} or 𝐂n{\bf C}_{n} (n=n=\infty also) symmetry. Due to the symmetry, eigenenergies and eigenvectors of H0H_{0} can be classified according to the irreducible representations (IRs) of a given point-group symmetry.

Besides H0H_{0}, in Eq. (1) the Zeeman term HzH_{\rm z} appears, describing the coupling of spin and magnetic field:

Hz=gμB𝑩𝒔,H_{\rm z}=g\mu_{B}{\bm{B}}\cdot{\bm{s}}, (2)

where gg is the effective Landé factor (g0.44g\approx-0.44 for GaAs), μB\mu_{B} is the Bohr magneton, 𝒔=1/2𝝈{\bm{s}}=1/2{\bm{\sigma}} is the electron’s spin, and 𝑩=B𝒏{\bm{B}}=B{\bm{n}} is the in-plane magnetic field forming an angle α\alpha with the crystallographic [100] axis. In Eq. (1) we have neglected the orbital effects of the in-plane magnetic field. This is a reasonable assumption for the magnetic field strength weaker than a few T{\rm T} RSF11 . In the case of the magnetic field applied in the zz direction, orbital effects would be much more pronounced RSF11 .

Eigenstates of H0+HzH_{0}+H_{\rm z} can be written in a direct product form |Ψi±=|Ψi|±|{\Psi_{i}\pm}\rangle=|{\Psi_{i}}\rangle\otimes{|{\pm}\rangle}, where |Ψi|{\Psi_{i}}\rangle corresponds to the eigenvectors of the Hamiltonian H0H_{0} with an energy ϵi\epsilon_{i}, while |±|{\pm}\rangle represents eigenvectors of HzH_{\rm z} with spin projection parallel or antiparallel to the magnetic field direction and an eigenenergy ±gμBB/2\pm g\mu_{B}B/2, respectively. The effect of HzH_{\rm z} on the eigenspectra of H0H_{0} can be seen as the splitting of H0H_{0} eigenenergies into two branches with an energy difference |g|μBB|g|\mu_{B}B. In this work, we assume that |g|μBB|g|\mu_{B}B is much weaker than the energy difference between the ground and the first excited state of the orbital Hamiltonian H0H_{0}.

Refer to caption
Figure 1: A schematic view of the GaAs lateral QD. The yy axis of the QD reference frame coincides with the vertical mirror plane symmetry σv\sigma_{\rm v}. We define the angle between the chosen xx axis and the crystallographic [100][100] axis as φ\varphi. The magnetic field is aligned along the 𝐧{\bf n} direction, forming an angle α\alpha with the [100][100] direction.

Besides H0H_{0} (HzH_{\rm z}) that acts trivially in the spin (orbital) space, the SOI Hamiltonian does not commute with H0+HzH_{0}+H_{\rm z}. It consists of two terms, Dresselhaus DR and Rashba RA : the Dresselhaus term exists due to the bulk inversion asymmetry of the structure, while the Rashba term is present when an electric field perpendicular to the growth direction is applied. The form of spin-orbit coupling is dependent on the structure’s symmetry. For GaAs, having the zincblende structure, the SOI Hamiltonian is equal to

Hso=2αd(pycsypxcsx)+2αr(pxcsypycsx),\displaystyle H_{\rm so}=2\alpha_{\rm d}(p_{y}^{\rm c}s_{y}-p_{x}^{\rm c}s_{x})+2\alpha_{\rm r}(p_{x}^{\rm c}s_{y}-p_{y}^{\rm c}s_{x}),\;\; (3)

where αr\alpha_{\rm r} and αd\alpha_{\rm d} are Rashba and Dresselhaus coupling constants, while pxcp_{x}^{\rm c} and pycp_{y}^{\rm c} are momentum operators in the [100] and [010] crystallographic directions, respectively. The electron spin is locked to the crystal momentum, since the potential trap confines electron of the crystal. Thus, an electron in a QD inherits the features of the crystal for which the crystal momentum is only appropriately defined. However, we have the choice to define the xx axis of our coordinate frame independently on the crystallographic [100] direction. Assuming that the angle between them is φ\varphi, pxcp_{x}^{\rm c} and pycp_{y}^{\rm c} should be written in terms of momentum operators in the chosen frame: pxc=pxcosφpysinφp_{x}^{\rm c}=p_{x}\cos{\varphi}-p_{y}\sin{\varphi}, pyc=pxsinφ+pycosφp_{y}^{\rm c}=p_{x}\sin{\varphi}+p_{y}\cos{\varphi}.

The spin-orbit Hamiltonian can be written in a different form using the Rashba lr=2/2mαrl_{\rm r}=\hbar^{2}/2m\alpha_{r} and Dresselhaus ld=2/2mαdl_{\rm d}=\hbar^{2}/2m\alpha_{d} precession lengths,

Hso=(pycsypxcsxmld+pxcsypycsxmlr).\displaystyle H_{\rm so}=\hbar\Big{(}\frac{p_{y}^{\rm c}s_{y}-p_{x}^{\rm c}s_{x}}{m^{*}l_{\rm d}}+\frac{p_{x}^{\rm c}s_{y}-p_{y}^{\rm c}s_{x}}{m^{*}l_{\rm r}}\Big{)}. (4)

To compare the ratio of the spin-orbit precession length and the orbital confinement length ll, we redefine lrl_{\rm r} and ldl_{\rm d} in terms of the overall spin-orbit length lsol_{\rm so} and the spin-orbit angle ν\nu:

ld1=lso1sinν,lr1=lso1cosν.l_{\rm d}^{-1}=l_{\rm so}^{-1}\sin{\nu},\;l_{\rm r}^{-1}=l_{\rm so}^{-1}\cos{\nu}. (5)

Since we assume no doping of the GaAs material SL05 , lsol_{\rm so} can be considered constant. Moreover, the relation lsoll_{\rm so}\gg l RSF11 ; RPS+14 is satisfied in GaAs QDs, meaning that SOI can be treated as a perturbation.

Without SOI, qubit states can be defined as |Ψ0±=|Ψ0|±|{\Psi_{0}\pm}\rangle=|{\Psi_{0}}\rangle\otimes|{\pm}\rangle, where |Ψ0|{\Psi_{0}}\rangle corresponds to the ground state of the spin-independent Hamiltonian H0H_{0}. Because SOI can be treated on the level of a perturbation, we calculate first-order corrections of the qubit states due to spin-orbit coupling. Since it is known that the standard perturbation technique badly incorporates the spin-orbit-induced corrections CWL04 ; BSF10 , we follow the procedure explained in Ref. MSL16 : the Hamiltonian HH is transformed using the unitary operator U=exp(i𝐧so𝒔)U={\rm exp}({\rm i}{\bf n}_{\rm so}\cdot{\bm{s}}), defined with the help of the position-dependent spin-orbit vector 𝐧so=lso1(r1sinν+r2cosν,r1cosνr2sinν,0){\bf n}_{\rm so}=l_{\rm so}^{-1}(r_{1}\sin{\nu}+r_{2}\cos{\nu},-r_{1}\cos{\nu}-r_{2}\sin{\nu},0):

UHU=H0+Hz+Hsoeff.UHU^{{\dagger}}=H_{0}+H_{\rm z}+H_{\rm so}^{\rm eff}. (6)

The unitary operator UU does not change the orbital and Zeeman Hamiltonian. On the other hand, the SOI Hamiltonian HsoH_{\rm so} is transformed into

Hsoeff=gμB(𝐧so×𝐁)𝒔24mlso2(1+2lzszcos2ν),H_{\rm so}^{\rm eff}=g\mu_{B}({\bf n}_{\rm so}\times{\bf B})\cdot{\bm{s}}-\frac{\hbar^{2}}{4m^{*}l_{\rm so}^{2}}\Big{(}1+2l_{z}s_{z}\cos{2\nu}\Big{)}, (7)

where lz=i(r1r2r2r1)l_{z}=-{\rm i}(r_{1}\partial_{r_{2}}-r_{2}\partial_{r_{1}}) is the orbital angular momentum. Using HsoeffH_{\rm so}^{\rm eff}, the first-order correction of the qubit states can be written as

δ|Ψ0σ=Ui0,σ′′Ψiσ′′|Hsoeff|Ψ0σϵ0ϵi+σσ′′2gμBB|Ψiσ′′,\delta|{\Psi_{0}\sigma^{\prime}}\rangle=U\sum_{i\neq 0,\sigma^{\prime\prime}}\frac{\langle{\Psi_{i}\sigma^{\prime\prime}}|H_{\rm so}^{\rm eff}|{\Psi_{0}\sigma^{\prime}}\rangle}{\epsilon_{0}-\epsilon_{i}+\frac{\sigma^{\prime}-\sigma^{\prime\prime}}{2}g\mu_{B}B}|{\Psi_{i}\sigma^{\prime\prime}}\rangle, (8)

where the sum over i0i\neq 0 corresponds to all orbital eigenvectors |Ψi|{\Psi_{i}}\rangle different from the ground state |Ψ0|{\Psi_{0}}\rangle, while σ′′=±\sigma^{\prime\prime}=\pm.

The lateral QD model is valid if the electron dynamics in the zz direction is suppressed; i.e., an electron is always in the ground state. Thus, we assume that confinement length in the zz direction is much stronger than in the xyxy plane. The Hamiltonian describing the quantum confinement in the zz direction is equal to H(z)=pz2/2m+V(z)H(z)=p_{z}^{2}/2m^{*}+V(z), where V(z)=eE0zV(z)=eE_{0}z for z0z\geq 0 and V(z)=V(z)=\infty for z<0z<0. To this Hamiltonian corresponds the following ground state (for z>0z>0SS03

Ψ0(z)=1.4261χAi(χz2.3381),\Psi_{0}(z)=1.4261\sqrt{\chi}{\rm Ai}(\chi z-2.3381), (9)

where Ai{\rm Ai} is the Airy function, while χ=(2meE0/2)1/3\chi=(2m^{*}eE_{0}/\hbar^{2})^{1/3} is the inverse of the characteristic length z0=1.5587/χz_{0}=1.5587/\chi in the zz direction.

In order to simplify the notation, in the rest of the paper we assume that ||{\uparrow}\rangle and ||{\downarrow}\rangle represent SOI corrected qubit states in the xyxy-plane, while |Ψ=|Ψ0(z)|{\Psi_{\uparrow}}\rangle=|{\uparrow}\rangle\Psi_{0}(z) and |Ψ=|Ψ0(z)|{\Psi_{\downarrow}}\rangle=|{\downarrow}\rangle\Psi_{0}(z) correspond to wavefunctions of the qubit states in three dimensions.

III Rabi frequency and phonon induced spin relaxation rate

Electrical control of the spin qubit is possible by applying the in-plane oscillating electric field 𝐄cosωt{\bf E}\cos{\omega t}, resulting in the Rabi Hamiltonian HR=e𝐄𝐫cos(ωt)H_{\rm R}=e{\bf E}\cdot{\bf r}\cos(\omega t). The Rabi frequency, measuring the speed of the single-qubit rotations, is equal to Ω=e/|𝐄|𝐫||\Omega=e/\hbar|{\bf E}\cdot\langle{\uparrow}|{\bf r}|{\downarrow}\rangle|, where

𝐝=|𝐫|{\bf d}_{\uparrow\downarrow}=\langle{\uparrow}|{\bf r}|{\downarrow}\rangle (10)

is the dipole moment (in ee units), present due to the SOI induced spin mixing mechanism. Misalignment of the applied field direction and the dipole moment leads to a trivial suppression of the Rabi frequency. Since it is beneficial to increase the Rabi frequency as much as possible, the electric field should be applied in the direction of the dipole moment. Thus, for fixed |𝐄||{\bf E}|, the maximal value max(Ω)=Ω\max(\Omega)=\Omega_{\uparrow\downarrow} of the Rabi frequency

Ω=e|𝐄||𝐝|\Omega_{\uparrow\downarrow}=\frac{e}{\hbar}|{\bf E}||{\bf d}_{\uparrow\downarrow}| (11)

is completely dependent on the strength of the dipole moment.

Since spin-phonon interaction in semiconductor QDs is irrelevant KN00 , unlike donor-bound electrons in direct band-gap semiconductors LKD+16 , only electron-phonon-induced transition between the qubit states should be considered in the study of spin relaxation. Electron-phonon coupling is triggered by the SOI-induced admixture mechanism, being highly dependent on the symmetry of the gating potential LKD+16 . We determine the rate of spin relaxation at T=0T=0 from the Fermi golden rule,

Γ=2πν𝐪|Mν(𝐪)|2|Ψ|ei𝐪𝐫c|Ψ|2δ(ϵων𝐪),\Gamma_{\uparrow\downarrow}=\frac{2\pi}{\hbar}\sum_{\nu{\bf q}}|M_{\nu}({\bf q})|^{2}|\langle{\Psi_{\uparrow}}|{\rm e}^{{\rm i}{\bf q}\cdot{\bf r}_{\rm c}}|{\Psi_{\downarrow}}\rangle|^{2}\delta(\epsilon_{\uparrow\downarrow}-\hbar\omega_{\nu{\bf q}}), (12)

assuming the dominant contribution of acoustic phonons, having an energy ων𝐪\hbar\omega_{\nu{\bf q}}, equal to the level separation between the qubit states, ϵ=|g|μBB\epsilon_{\uparrow\downarrow}=|g|\mu_{B}B. For magnetic field strengths up to a few T, relevant for this work, the linear dependence of phonon frequencies on the crystal wave vector length can be used, ων𝐪=cν|𝐪|\omega_{\nu{\bf q}}=c_{\nu}|{\bf q}|, giving us |𝐪|=|g|μBB/cν|{\bf q}|=|g|\mu_{B}B/\hbar c_{\nu} coorframe .

The geometric factor |Mν(𝐪)|2|M_{\nu}({\bf q})|^{2} is dependent on the phonon mode, longitudinal (LA) or transverse (TA). The longitudinal geometric factor CBG+06

|MLA(𝐪)|2=D22ρcLAV|𝐪|+32π2(eh14)2ϵ2ρcLAV(3qxqyqz)2|𝐪|7|M_{\rm LA}({\bf q})|^{2}=\frac{\hbar D^{2}}{2\rho c_{{\rm LA}}V}|{\bf q}|+\frac{32\pi^{2}\hbar(eh_{14})^{2}}{\epsilon^{2}\rho c_{{\rm LA}}V}\frac{(3q_{x}q_{y}q_{z})^{2}}{|{\bf q}|^{7}} (13)

depends on both DD and h14h_{14}, representing the deformation and piezoelectric constant, respectively. On the other hand, the transverse geometric factor CBG+06

|MTA(𝐪)|2\displaystyle|M_{\rm TA}({\bf q})|^{2} =\displaystyle= 232π2(eh14)2ϵ2ρcTAV\displaystyle 2\frac{32\pi^{2}\hbar(eh_{14})^{2}}{\epsilon^{2}\rho c_{{\rm TA}}V}
×|qx2qy2+qx2qz2+qy2qz2|𝐪|5(3qxqyqz)2|𝐪|7|\displaystyle\times\left|\frac{q_{x}^{2}q_{y}^{2}+q_{x}^{2}q_{z}^{2}+q_{y}^{2}q_{z}^{2}}{|{\bf q}|^{5}}-\frac{(3q_{x}q_{y}q_{z})^{2}}{|{\bf q}|^{7}}\right|

is dependent on the piezoelectric constant solely. Other parameters for the GaAs material are CWL04 ; MSL16 cLA=5290m/sc_{\rm LA}=5290{\rm m}/{\rm s}, cTA=2480m/sc_{\rm TA}=2480{\rm m}/{\rm s}, ρ=5300kg/m3\rho=5300{\rm kg}/{\rm m^{3}}, D=7eVD=7{\rm eV}, eh14=1.4×109eV/meh_{14}=1.4\times 10^{9}{\rm eV}/{\rm m}, and ϵ=12.9\epsilon=12.9.

Finally, in Eq. (12) both the lateral and the zz-direction confinement enter the relaxation rate through the scattering matrix element |Ψ|ei𝐪𝐫c|Ψ|2|\langle{\Psi_{\uparrow}}|{\rm e}^{{\rm i}{\bf q}\cdot{\bf r}_{\rm c}}|{\Psi_{\downarrow}}\rangle|^{2}. We employ the dipole approximation ei𝐪𝐫c1+i𝐪𝐫c{\rm e}^{{\rm i}{\bf q}\cdot{\bf r}_{c}}\approx 1+{\rm i}{\bf q}\cdot{\bf r}_{c}, justified for magnetic field strengths below a few T.

To summarize, the phonon-induced relaxation rate can be divided into three separate channels: the deformation phonons Γdef\Gamma_{\uparrow\downarrow}^{\rm def}, the longitudinal piezoelectric phonons Γpiez,LA\Gamma_{\uparrow\downarrow}^{\rm piez,LA}, and the transverse piezoelectric phonons Γpiez,TA\Gamma_{\uparrow\downarrow}^{\rm piez,TA}. In GaAs QDs, Γpiez,TA\Gamma_{\uparrow\downarrow}^{\rm piez,TA} is the dominant relaxation channel, being two orders of magnitude stronger than Γpiez,LA+Γdef\Gamma_{\uparrow\downarrow}^{\rm piez,LA}+\Gamma_{\uparrow\downarrow}^{\rm def} in the dipole approximation regime. Thus, we can identify the total relaxation rate with Γpiez,TA\Gamma_{\uparrow\downarrow}^{\rm piez,TA} CSZ+18 :

Γ\displaystyle\Gamma_{\uparrow\downarrow} =\displaystyle= 256π(eh14)2(|g|μBB)3105cTA5ρ4ϵ2(1+733KTA2z02)|𝐝|2,\displaystyle\frac{256\pi(eh_{14})^{2}(|g|\mu_{B}B)^{3}}{105c_{\rm TA}^{5}\rho\hbar^{4}\epsilon^{2}}(1+\frac{7}{33}K_{\rm TA}^{2}z_{0}^{2})|{\bf d}_{\uparrow\downarrow}|^{2},\;\;\;\;\;\; (15)

where KTA=|g|μBB/cTAK_{\rm TA}=|g|\mu_{B}B/\hbar c_{\rm TA}. We assume a typical confinement length l=10l=10nm RSF11 ; RPS+14 of the GaAs QD in an experimental setup and magnetic field up to a few T (see Section II). Since confinement in the zz direction is much stronger than in the xyxy plane, z0lz_{0}\ll l, we conclude that 7KTA2z02/337K_{\rm TA}^{2}z_{0}^{2}/33 is much weaker than 1. In other words, the influence of the confinement in the zz direction can be neglected.

Note that Γ\Gamma_{\uparrow\downarrow} is squarely dependent on the absolute value of the dipole moment, meaning that the knowledge of the dipole moment is sufficient to fully explain the behavior of both the Rabi frequency and the spin relaxation rate.

Refer to caption
Figure 2: A schematic view of the first-order perturbation correction of the qubit states |A0±|{A_{0}\pm}\rangle in the case of 𝐂nv(𝐂v){\bf C}_{n{\rm v}}({\bf C}_{\infty{\rm v}}) (left) and 𝐂2v{\bf C}_{2{\rm v}} (right) symmetry. In the first case, states that correct the qubit states have twofold orbital degeneracy and transform according to the IR E1E_{1}. These states are split by the Zeeman energy |g|μBB|g|\mu_{B}B. The transition between the SOI uncorrected qubit states and the |Ea,bi|{E_{a,b}^{i}}\rangle states is enabled by the xx and yy terms from HsoeffH_{\rm so}^{\rm eff}. In the second case, orbital states involved in the qubit states correction transform according to IRs A1A_{1} and B1B_{1}; the transition is triggered by the terms yy and xx from HsoeffH_{\rm so}^{\rm eff}, respectively.

IV Analytical expression for the dipole moment

Based on the previous conclusion, we come to the main objective: to derive symmetry allowed expression for the dipole moment. The results can be divided into three cases, according to the system’s group symmetry: (1) 𝐂nv{\bf C}_{n{\rm v}} (n3n\geq 3) and 𝐂v{\bf C}_{\infty{\rm v}}, (2) 𝐂2v{\bf C}_{2{\rm v}} and 𝐂1v{\bf C}_{1{\rm v}}, and (3) 𝐂n{\bf C}_{n} and 𝐂{\bf C}_{\infty}.

IV.1 Dipole moment for systems with 𝐂nv{\bf C}_{n{\rm v}} (n3n\geq 3) or 𝐂v{\bf C}_{\infty{\rm v}} symmetry

To find the SOI-induced perturbative correction of the qubit states, we first rewrite the unitarily transformed SOI Hamiltonian in the coordinate frame of the potential,

Hsoeff\displaystyle H_{\rm so}^{\rm eff} =\displaystyle= gμBBsz(x(sin(ν+φ)sinα+cos(νφ)cosα)\displaystyle g\mu_{B}Bs_{z}\Big{(}x\big{(}\sin{(\nu+\varphi)}\sin{\alpha}+\cos{(\nu-\varphi)}\cos{\alpha}\big{)} (16)
+y(cos(ν+φ)sinα+sin(νφ)cosα)),\displaystyle+y\big{(}\cos{(\nu+\varphi)}\sin{\alpha}+\sin{(\nu-\varphi)}\cos{\alpha}\big{)}\Big{)},

and neglect the second term in Eq. (7), assuming magnetic field strengths >μ>\muT needed to appropriately define the qubit states. For simplicity, we define two factors,

vx\displaystyle v_{x} =\displaystyle= sin(ν+φ)sinα+cos(νφ)cosα,\displaystyle\sin(\nu+\varphi)\sin{\alpha}+\cos(\nu-\varphi)\cos{\alpha}, (17)
vy\displaystyle v_{y} =\displaystyle= cos(ν+φ)sinα+sin(νφ)cosα,\displaystyle\cos{(\nu+\varphi)}\sin{\alpha}+\sin{(\nu-\varphi)}\cos{\alpha}, (18)

with whose help HsoeffH_{\rm so}^{\rm eff} can be written in a more compact form.

The Hamiltonian HsoeffH_{\rm so}^{\rm eff} is in the orbital space dependent on the coordinates xx and yy that transform according to the IR E1E_{1}. Their symmetry behavior restricts the states that can appear in the perturbative correction of the qubit states. It is simple to check that only states transforming according to the IR E1E_{1} are allowed. This is illustrated in the left-hand panel of FIG. 2.

We label the ground state of the orbital Hamiltonian as |A0|{A_{0}}\rangle, since the ground state in quantum mechanical systems is of the maximal possible symmetry LCC08 and it should transform according to the A0A_{0} IR, representing the objects invariant under all group symmetry operations (see Table 1). We write two complex conjugate basis vectors of the two-dimensional IR E1E_{1} as |Eai|{E_{a}^{i}}\rangle and |Ebi|{E_{b}^{i}}\rangle, where ii labels the energy level. Also, we define the energy difference between the excited level and the ground state as ϵi=ϵexiϵgr\epsilon^{i}=\epsilon_{ex}^{i}-\epsilon_{gr}.

Table 1: For 𝐂nv{\bf C}_{n{\rm v}} and 𝐂v{\bf C}_{\infty{\rm v}} symmetry groups, tables of matrices of the corresponding IRs are given JB67 , tabulated on the generators CnC_{n}(RβR_{\beta}) and σv\sigma_{\rm v}, where CnC_{n}(RβR_{\beta}) represents a rotation for the angle 2π/n2\pi/n (β\beta) around the zz axis. In the 𝐂nv{\bf C}_{n{\rm v}} case, two-dimensional IRs exist if n3n\geq 3. In both cases, two-dimensional IRs are written in a complex conjugate basis.
𝐂nv{\bf C}_{n{\rm v}} IR mm Cn{C}_{n} σv\sigma_{\rm v}
A0/B0A_{0}/B_{0} 0 11 ±1\pm 1
EmE_{m} (0,n2)(0,\frac{n}{2}) (ei2πnm00ei2πnm)\left(\begin{array}[]{cc}{\rm e}^{{\rm i}\frac{2\pi}{n}m}&0\\ 0&{\rm e}^{-{\rm i}\frac{2\pi}{n}m}\\ \end{array}\right) (0110)\left(\begin{array}[]{cc}0&1\\ 1&0\\ \end{array}\right)
An2/Bn2A_{\frac{n}{2}}/B_{\frac{n}{2}} n2\frac{n}{2} -1 ±1\pm 1
𝑪v{\bm{C}}_{\infty{\rm v}} IR mm RβR_{\beta} σv\sigma_{\rm v}
A0/B0A_{0}/B_{0} 0 11 ±1\pm 1
EmE_{m} 1,2,1,2,... (eiβm00eiβm)\left(\begin{array}[]{cc}{\rm e}^{{\rm i}\beta m}&0\\ 0&{\rm e}^{-{\rm i}\beta m}\\ \end{array}\right) (0110)\left(\begin{array}[]{cc}0&1\\ 1&0\\ \end{array}\right)

Due to the negative gg factor, the lowest qubit state |A0+=|A0|+|{A_{0}+}\rangle=|{A_{0}}\rangle\otimes{|{+}\rangle} is parallel to the magnetic field direction, while |A0=|A0||{A_{0}-}\rangle=|{A_{0}}\rangle\otimes{|{-}\rangle} is the qubit state with spin projection antiparallel to the magnetic field direction. The first-order perturbative correction to the qubit states is written as |δA0±|{\delta A_{0}\pm}\rangle. Thus, we can write the SOI corrected qubit states as |=|A0±+|δA0±|{\uparrow\downarrow}\rangle=|{A_{0}\pm}\rangle+|{\delta A_{0}\pm}\rangle, where the normalization factor is omitted as the correction is small. Correspondingly, the dipole moment is equal to

𝐝=j=x,y|𝐫𝒆j|𝒆j\displaystyle{\bf d}=\sum_{j=x,y}\langle{\uparrow}|{\bf r}\cdot{\bm{e}}_{j}|{\downarrow}\rangle{\bm{e}}_{j} =\displaystyle= j=x,y(A0+|𝐫𝒆j|δA0𝒆j\displaystyle\sum_{j=x,y}\Big{(}\langle{A_{0}+}|{\bf r}\cdot{\bm{e}}_{j}|{\delta A_{0}-}\rangle{\bm{e}}_{j} (19)
+δA0+|𝐫𝒆j|A0𝒆j).\displaystyle+\langle{\delta A_{0}+}|{\bf r}\cdot{\bm{e}}_{j}|{A_{0}-}\rangle{\bm{e}}_{j}\Big{)}.

Since lsoll_{\rm so}\gg l, we approximate the unitary operator UU with I2I_{2}, where I2I_{2} is the identity 2×22\times 2 matrix. After noticing that ±|sz|=1/2\langle{\pm}|s_{z}|{\mp}\rangle=-1/2, ±|sz|±=0\langle{\pm}|s_{z}|{\pm}\rangle=0, we find the SOI-induced corrections of the qubit states

|δA0±\displaystyle|{\delta A_{0}\pm}\rangle =\displaystyle= |g|μBB2lsoi(Eai|xvx+yvy|A0ϵi±|g|μBB|Eai\displaystyle\frac{|g|\mu_{B}B}{2l_{\rm so}}\sum_{i}\Big{(}\frac{\langle{E_{a}^{i}}|xv_{x}+yv_{y}|{A_{0}}\rangle}{\epsilon^{i}\pm|g|\mu_{B}B}|{E_{a}^{i}\mp}\rangle (20)
+Ebi|xvx+yvy|A0ϵi±|g|μBB|Ebi).\displaystyle+\frac{\langle{E_{b}^{i}}|xv_{x}+yv_{y}|{A_{0}}\rangle}{\epsilon^{i}\pm|g|\mu_{B}B}|{E_{b}^{i}\mp}\rangle\Big{)}.

Additionally, transition dipole matrix elements are labeled as

Xi=Eai|x|A0,Yi=Eai|y|A0.X^{i}=\langle{E_{a}^{i}}|x|{A_{0}}\rangle,\;Y^{i}=\langle{E_{a}^{i}}|y|{A_{0}}\rangle. (21)

Since the Zeeman splitting is much smaller than the orbital excitation energies, |g|μBBϵi|g|\mu_{B}B\ll\epsilon^{i}, the approximation ϵi±|g|μBBϵi\epsilon^{i}\pm|g|\mu_{B}B\approx\epsilon^{i} can be made. Thus, Eq. (20) is transformed into

|δA0±=|g|μBB2lsoi\displaystyle|{\delta A_{0}\pm}\rangle=\frac{|g|\mu_{B}B}{2l_{\rm so}}\sum_{i} (Xivx+Yivyϵi|Eai\displaystyle\Big{(}\frac{X^{i}v_{x}+Y^{i}v_{y}}{\epsilon^{i}}|{E_{a}^{i}\mp}\rangle
+\displaystyle+ (Xi)vx+(Yi)vyϵi|Ebi),\displaystyle\frac{(X^{i})^{*}v_{x}+(Y^{i})^{*}v_{y}}{\epsilon^{i}}|{E_{b}^{i}\mp}\rangle\Big{)}, (22)

where (Xi)(X^{i})^{*} and (Yi)(Y^{i})^{*} are the complex conjugates of XiX^{i} and YiY^{i}, respectively. Components of the dipole moment can now be written in a more compact form

dx\displaystyle d_{x} =\displaystyle= 2|g|μBBlsoi|Xi|2vx+Re(Xi(Yi))vyϵi,\displaystyle\frac{2|g|\mu_{B}B}{l_{\rm so}}\sum_{i}\frac{|X^{i}|^{2}v_{x}+{\rm Re}(X^{i}(Y^{i})^{*})v_{y}}{\epsilon^{i}},
dy\displaystyle d_{y} =\displaystyle= 2|g|μBBlsoi|Yi|2vy+Re(Xi(Yi))vxϵi,\displaystyle\frac{2|g|\mu_{B}B}{l_{\rm so}}\sum_{i}\frac{|Y^{i}|^{2}v_{y}+{\rm Re}(X^{i}(Y^{i})^{*})v_{x}}{\epsilon^{i}}, (23)

where Re(Xi(Yi)){\rm Re}(X^{i}(Y^{i})^{*}) stands for the real part of Xi(Yi)X^{i}(Y^{i})^{*}. Potential dependent parameters that enter Eq. (23) are the transition dipole matrix elements and the excitation energies. Besides them, dipole moment components are dependent on the spin-orbit angle ν\nu, magnetic field angle α\alpha, and the angle φ\varphi between the [100] crystallographic direction and the xx axis.

A further simplification of Eq. (23) stems from the existence of the vertical mirror symmetry σv\sigma_{\rm v}, requiring that Re(Xi(Yi)){\rm Re}(X^{i}(Y^{i})^{*}) must be zero. This can be proven in a few simple steps. First, we deduce from the matrix of an IR E1E_{1}, representing the vertical mirror plane, that σv\sigma_{\rm v} transforms one IR vector into the other, E1(σv)|Ea,bi=|Eb,aiE_{1}(\sigma_{\rm v})|{E_{a,b}^{i}}\rangle=|{E_{b,a}^{i}}\rangle. Furthermore, yy remains unchanged, while xx acquires a minus sign, leading to the following behavior of the transition matrix elements XiX^{i} and YiY^{i} under vertical mirror plane symmetry:

Xiσv(Xi),Yiσv(Yi).\displaystyle X^{i}\xrightarrow{\sigma_{\rm v}}-(X^{i})^{*},Y^{i}\xrightarrow{\sigma_{\rm v}}(Y^{i})^{*}. (24)

From the previous relations, we conclude that the term Re(Xi(Yi)){\rm Re}(X^{i}(Y^{i})^{*}) transforms into Re(Xi(Yi))-{\rm Re}(X^{i}(Y^{i})^{*}), meaning that this object does not obey the symmetry of a system and must vanish.

Additionally, rotational symmetry of a system imposes that matrix elements |Xi|2|X^{i}|^{2} and |Yi|2|Y^{i}|^{2} are equal. This can be concluded from the action of the rotation CnC_{n} for an angle βn=2π/n\beta_{n}=2\pi/n around the zz axis, being the element of the group symmetry. An element CnC_{n} leaves the vector |A0|{A_{0}}\rangle unchanged and adds a phase exp(iβn){\rm exp}({{\rm i}\beta_{n}}) to the vector |Eai|{E_{a}^{i}}\rangle. Also, it transforms xx and yy to xcosβn+ysinβnx\cos{\beta_{n}}+y\sin{\beta_{n}} and xsinβn+ycosβn-x\sin{\beta_{n}}+y\cos{\beta_{n}}. Thus, XiX^{i} and YiY^{i} are transformed into exp(iβn)(Xicosβn+Yisinβn){\rm exp}(-{\rm i}\beta_{n})(X^{i}\cos{\beta_{n}}+Y^{i}\sin{\beta_{n}}) and exp(iβn)(Xisinβn+Yicosβn){\rm exp}(-{\rm i}\beta_{n})(-X^{i}\sin{\beta_{n}}+Y^{i}\cos{\beta_{n}}), respectively. Correspondingly,

|Xi|2Cn|Xi|2cos2βn+|Yi|2sin2βn,\displaystyle|X^{i}|^{2}\xrightarrow{C_{n}}|X^{i}|^{2}\cos^{2}{\beta_{n}}+|Y^{i}|^{2}\sin^{2}{\beta_{n}},
|Yi|2Cn|Xi|2sin2βn+|Yi|2cos2βn,\displaystyle|Y^{i}|^{2}\xrightarrow{C_{n}}|X^{i}|^{2}\sin^{2}{\beta_{n}}+|Y^{i}|^{2}\cos^{2}{\beta_{n}}, (25)

where we have neglected the Re(Xi(Yi)){\rm Re}(X^{i}(Y^{i})^{*}) term, which was previously proven to equal to zero. Since |Xi|2|X^{i}|^{2} and |Yi|2|Y^{i}|^{2} must remain unchanged under the group symmetry operations, we conclude that the relation |Xi|2=|Yi|2|X^{i}|^{2}=|Y^{i}|^{2} must hold. Thus, we have obtained a general relation for the dipole moment in the case of the potential symmetry 𝐂nv{\bf C}_{n{\rm v}} (n3n\geq 3):

𝐝𝐂nv=2|g|μBBlso(i|Xi|2ϵi)(vx𝐞x+vy𝐞y).{\bf d}_{\uparrow\downarrow}^{{\bf C}_{n{\rm v}}}=\frac{2|g|\mu_{B}B}{l_{\rm so}}\Big{(}\sum_{i}\frac{|X^{i}|^{2}}{\epsilon^{i}}\Big{)}(v_{x}{\bf e}_{x}+v_{y}{\bf e}_{y}). (26)

In these situations, the absolute value of the dipole moment |𝐝𝐂nv|2(1+sin2αsin2ν)|{\bf d}_{\uparrow\downarrow}^{{\bf C}_{n{\rm v}}}|^{2}\sim(1+\sin{2\alpha}\sin{2\nu}) is independent of the orientation of the potential with respect to the crystallographic frame.

Analogous analysis can be conducted in the 𝐂v{\bf C}_{\infty{\rm v}} case. Since the matrix form of the IRs A0A_{0} and E1E_{1} (see Table 1) for this symmetry group is the same as for 𝐂nv{\bf C}_{n{\rm v}}, the procedure is exactly the same if the change βnβ\beta_{n}\rightarrow\beta in the previous discussion is made.

As an example, we implement the derived formula (26) in the case of the isotropic two-dimensional harmonic confinement Viho(x,y)=1/2mω2(x2+y2)V^{\rm iho}(x,y)=1/2m^{*}\omega^{2}(x^{2}+y^{2}) with 𝐂v{\bf C}_{\infty{\rm v}} symmetry, assuming only one excited level in the perturbative correction of the qubit states. With the help of the states ψ0\psi_{0} and ψ1\psi_{1}, corresponding to the ground and the first excited states of the one-dimensional harmonic oscillator, we can define the ground state |A0|{A_{0}}\rangle and two complex conjugate eigenstates |Ea|{E_{a}}\rangle and |Eb|{E_{b}}\rangle of the degenerate level: |A0=ψ0(x)ψ0(y)|{A_{0}}\rangle=\psi_{0}(x)\psi_{0}(y), |Ea=(ψ0(x)ψ1(y)+iψ1(x)ψ0(y))/2|{E_{a}}\rangle=(\psi_{0}(x)\psi_{1}(y)+{\rm i}\psi_{1}(x)\psi_{0}(y))/\sqrt{2}, and |Eb=(ψ0(x)ψ1(y)iψ1(x)ψ0(y))/2|{E_{b}}\rangle=(\psi_{0}(x)\psi_{1}(y)-{\rm i}\psi_{1}(x)\psi_{0}(y))/\sqrt{2}. In this case, the squared norm of the transition matrix element is equal to |X|2=/4mω|X|^{2}=\hbar/4m^{*}\omega. Using the energy difference of the ground and the first excited energy level ϵ=ω\epsilon=\hbar\omega and the confinement length l=/mωl=\sqrt{\hbar/m^{*}\omega}, an expression for the dipole moment is obtained MSL16 :

𝐝iho=|g|μBBml42lso2(vx𝐞x+vy𝐞y).{\bf d}^{\rm iho}_{\uparrow\downarrow}=\frac{|g|\mu_{B}Bm^{*}l^{4}}{2l_{\rm so}\hbar^{2}}\Big{(}v_{x}{\bf e}_{x}+v_{y}{\bf e}_{y}\Big{)}. (27)

IV.2 Dipole moment for systems with 𝐂2v{\bf C}_{2{\rm v}} or 𝐂1v{\bf C}_{1{\rm v}} symmetry

As the next step, we discuss potentials with 𝐂2v{\bf C}_{2{\rm v}} symmetry. In this case, coordinates xx and yy transform according to the IRs B1B_{1} and A1A_{1}, respectively. Their symmetry behavior imposes the following: x(y)x(y) couples the ground state |A0|{A_{0}}\rangle with states transforming according to the IR B1(A1)B_{1}(A_{1}) (see the right-hand panel of FIG. 2). Thus, the SOI-induced corrections of the qubit states are

|δA0±\displaystyle|{\delta A_{0}\pm}\rangle =\displaystyle= |g|μBB2lsoi(B1i|xvx|A0ϵB1i|B1i\displaystyle\frac{|g|\mu_{B}B}{2l_{\rm so}}\sum_{i}\Big{(}\frac{\langle{B_{1}^{i}}|xv_{x}|{A_{0}}\rangle}{\epsilon_{B_{1}}^{i}}|{B_{1}^{i}\mp}\rangle (28)
+A1i|yvy|A0ϵA1i|A1i),\displaystyle+\frac{\langle{A_{1}^{i}}|yv_{y}|{A_{0}}\rangle}{\epsilon_{A_{1}}^{i}}|{A_{1}^{i}\mp}\rangle\Big{)},

where ϵB1i(ϵA1i)\epsilon_{B_{1}}^{i}(\epsilon_{A_{1}}^{i}) is the energy difference between the energy level transforming according to the IR B1(A1)B_{1}(A_{1}) and the ground-state energy. We define the transition matrix elements as

Xi=B1i|x|A0,Yi=A1i|y|A0,X^{i}=\langle{B_{1}^{i}}|x|{A_{0}}\rangle,\;Y^{i}=\langle{A_{1}^{i}}|y|{A_{0}}\rangle, (29)

and obtain the formula for the dipole moment,

𝐝𝐂2v=|g|μBBlsoi(|Xi|2ϵB1ivx𝐞x+|Yi|2ϵA1ivy𝐞y).{\bf d}^{{\bf C}_{2{\rm v}}}_{\uparrow\downarrow}=\frac{|g|\mu_{B}B}{l_{\rm so}}\sum_{i}\Big{(}\frac{|X^{i}|^{2}}{\epsilon_{B_{1}}^{i}}v_{x}{\bf e}_{x}+\frac{|Y^{i}|^{2}}{\epsilon_{A_{1}}^{i}}v_{y}{\bf e}_{y}\Big{)}. (30)

In this case, anisotropy of the dipole moment appears since it is not forbidden that i|Xi|2/ϵB1i\sum_{i}|X^{i}|^{2}/\epsilon_{B_{1}}^{i} differs from i|Yi|2/ϵA1i\sum_{i}|Y^{i}|^{2}/\epsilon_{A_{1}}^{i}.

The anisotropy of the dipole moment can be illuminated using the example of the anisotropic two-dimensional harmonic potential Vaho(x,y)=1/2m(ωx2x2+ωy2y2)V^{\rm aho}(x,y)=1/2m^{*}(\omega_{x}^{2}x^{2}+\omega_{y}^{2}y^{2}), with different confinement lengths lx=/mωxl_{x}=\sqrt{\hbar/m^{*}\omega_{x}} and ly=/mωyl_{y}=\sqrt{\hbar/m^{*}\omega_{y}} along the xx and yy directions. We set l=lxl=l_{x} and ly=kll_{y}=kl, where k<1k<1 is the measure of anisotropy. We assume two excited orbital states in the perturbative correction: one of type A1A_{1} and one of type B1B_{1}. In this case we define the ground state |A0=ψ0(x)ψ0(y)|{A_{0}}\rangle=\psi_{0}(x)\psi_{0}(y) and two excited orbital states |A1=ψ0(x)ψ1(y)|{A_{1}}\rangle=\psi_{0}(x)\psi_{1}(y) and |B1=ψ1(x)ψ0(y)|{B_{1}}\rangle=\psi_{1}(x)\psi_{0}(y), where ψ0/1(x/y)\psi_{0/1}(x/y) represents the ground or first excited state (subscript 0 or 1, respectively) of the one-dimensional harmonic oscillator problem in the xx or yy direction. The obtained result

𝐝aho=|g|μBBml42lso2(vx𝐞x+k4vy𝐞y){\bf d}^{\rm aho}_{\uparrow\downarrow}=\frac{|g|\mu_{B}Bm^{*}l^{4}}{2l_{\rm so}\hbar^{2}}\Big{(}v_{x}{\bf e}_{x}+k^{4}v_{y}{\bf e}_{y}\Big{)} (31)

is again consistent with Ref. MSL16 .

In the case of the 𝐂1v{\bf C}_{1{\rm v}} symmetry, using a similar analysis as in the previous case, we obtain the expression for the dipole moment

𝐝𝐂1v=|g|μBBlsoi(|Xi|2ϵB0ivx𝐞x+|Yi|2ϵA0ivy𝐞y),{\bf d}^{{\bf C}_{1{\rm v}}}_{\uparrow\downarrow}=\frac{|g|\mu_{B}B}{l_{\rm so}}\sum_{i}\Big{(}\frac{|{X^{i}}|^{2}}{\epsilon_{B_{0}}^{i}}v_{x}{\bf e}_{x}+\frac{|Y^{i}|^{2}}{\epsilon_{A_{0}}^{i}}v_{y}{\bf e}_{y}\Big{)}, (32)

where Xi=B0i|x|A0X^{i}=\langle{B_{0}^{i}}|x|{A_{0}}\rangle, Yi=A0i|y|A0Y^{i}=\langle{A_{0}^{i}}|y|{A_{0}}\rangle, and ϵA0/B0i\epsilon_{A_{0}/B_{0}}^{i} is the energy difference between the nondegenerate energy level transforming according to the IR A0/B0A_{0}/B_{0} and the ground-state energy.

IV.3 Dipole moment for systems with 𝐂n{\bf C}_{n} or 𝐂{\bf C}_{\infty} symmetry

In the case of 𝐂n{\bf C}_{n} symmetry, all IRs AmA_{m} (m(n/2,n/2]m\in(-n/2,n/2]) are one dimensional and represent an element of symmetry CnsC_{n}^{s} (s=0,1,,n1)(s=0,1,...,n-1) as ei2πms/n{\rm e}^{{\rm i}2\pi ms/n}. Besides the geometric symmetry, the time-reversal symmetry Θ\Theta should be included also TIMrev . Time-reversal Θ\Theta changes the sign of the quantum number mm labeling the IR vector |Am|{A_{m}}\rangle, since it acts as a complex conjugation in the orbital space:

Θ|Am=|Am.\Theta|{A_{m}}\rangle=|{A_{-m}}\rangle. (33)

The eigenproblem of the Hamiltonian H|Am=ϵm|AmH|{A_{m}}\rangle=\epsilon_{m}|{A_{m}}\rangle, when combined with the commutation relation [Θ,H0]=0[\Theta,H_{0}]=0, gives us

H|Am=ϵm|Am,H|{A_{-m}}\rangle=\epsilon_{m}|{A_{-m}}\rangle, (34)

stating that, for n3n\geq 3, vectors |Am|{A_{m}}\rangle and |Am|{A_{-m}}\rangle are eigenstates of the degenerate level ϵm\epsilon_{m}. To this degenerate level corresponds the reducible representation AmAmA_{m}\oplus A_{-m} (except for m=n/2m=n/2). The representation AmAmA_{m}\oplus A_{-m} is equivalent to the IR EmE_{m} of the 𝐂nv{\bf C}_{n{\rm v}} group (see Table 1) if the generator σv\sigma_{\rm v} is neglected. In other words, Eq. (23) for the dipole moment is valid also in this case, since it is obtained without assuming the presence of vertical mirror symmetry. In this case vectors |Eai|{E_{a}^{i}}\rangle and |Ebi|{E_{b}^{i}}\rangle coincide with |A1i|{A_{1}^{i}}\rangle and |A1i|{A_{-1}^{i}}\rangle, respectively.

A further simplification of Eq. (23) appears for systems whose symmetry element is π/2\pi/2 rotation. This happens if the relation n=4rn=4r (rr\in\mathds{N}) is satisfied. Since Re(Xi(Yi))=0{\rm Re}(X^{i}(Y^{i})^{*})=0 and |Xi|2=|Yi|2|X^{i}|^{2}=|Y^{i}|^{2} in this case, Eq. (26) is relevant. Using the same reasoning it can be concluded that Eq. (26) is valid in the 𝐂{\bf C}_{\infty} case also.

Finally, the dipole moment components for the 𝐂2{\bf C}_{2} symmetry are equal to

(𝐝𝐂2)x\displaystyle({\bf d}^{{\bf C}_{2}}_{\uparrow\downarrow})_{x} =\displaystyle= |g|μBBlsoi|Xi|2vx+Re(Xi(Yi))vyϵA1i,\displaystyle\frac{|g|\mu_{B}B}{l_{\rm so}}\sum_{i}\frac{|{X^{i}}|^{2}v_{x}+{\rm Re}(X^{i}(Y^{i})^{*})v_{y}}{\epsilon_{A_{1}}^{i}},
(𝐝𝐂2)y\displaystyle({\bf d}^{{\bf C}_{2}}_{\uparrow\downarrow})_{y} =\displaystyle= |g|μBBlsoi|Yi|2vy+Re(Xi(Yi))vxϵA1i,\displaystyle\frac{|g|\mu_{B}B}{l_{\rm so}}\sum_{i}\frac{|{Y^{i}}|^{2}v_{y}+{\rm Re}(X^{i}(Y^{i})^{*})v_{x}}{\epsilon_{A_{1}}^{i}}, (35)

where Xi=A1i|x|A0X^{i}=\langle{A_{1}^{i}}|x|{A_{0}}\rangle, Yi=A1i|y|A0Y^{i}=\langle{A_{1}^{i}}|y|{A_{0}}\rangle, and ϵA1i\epsilon_{A_{1}}^{i} is the energy difference between the level transforming according to the IR A1A_{1} and the ground-state energy.

To conclude, anisotropy of the potential orientation with respect to the crystallographic frame is present in systems without the π/2\pi/2 group element (n4rn\neq 4r, rr\in\mathds{N}); isotropic behavior is present if a rotation for π/2\pi/2 is the group element, i.e., if n=4rn=4r (r)r\in\mathds{N}) or n=n=\infty.

V Applications: infinite-wall equilateral triangle, square, and rectangular potential

The results presented in the previous Section fully explain the dependence of the Rabi frequency and spin relaxation rate on the spin-orbit angle, magnetic field direction, and the relative orientation of the gating potential with respect to the crystallographic frame.

However, symmetry arguments alone cannot provide us with a qualitative estimation of the spin relaxation rate, corresponding to the phonon-allowed spin qubit lifetime. Since Γ\Gamma_{\uparrow\downarrow} is known for the harmonic gating MSL16 , we wish to compare the phonon-induced spin relaxation rate of other confinement potentials with the known values. To this end, we analyze the spin qubit confined inside the infinite-wall equilateral triangle, square, and rectangular gating potential (see FIG. 3):

Vtqd\displaystyle V^{\rm tqd} =\displaystyle= {0,x[y3a3,ay33],y[a36,a33],otherwise,\displaystyle\left\{\begin{array}[]{rl}0,&x\in[\frac{y\sqrt{3}-a}{3},\frac{a-y\sqrt{3}}{3}],y\in[\frac{-a\sqrt{3}}{6},\frac{a\sqrt{3}}{3}]\\ \infty,&{\rm otherwise},\\ \end{array}\right. (38)
Vrqd\displaystyle V^{\rm rqd} =\displaystyle= {0,x[a2,a2],y[b2,b2],otherwise.\displaystyle\left\{\begin{array}[]{rl}0,&x\in[-\frac{a}{2},\frac{a}{2}],y\in[-\frac{b}{2},\frac{b}{2}]\\ \infty,&{\rm otherwise}.\\ \end{array}\right. (41)

In the first case, Eq. (38), the potential has 𝐂3v{\bf C}_{3{\rm v}} symmetry and the corresponding eigenvectors of the spin-independent Hamiltonian H0H_{0} transform according to the one-dimensional IRs A0A_{0} and B0B_{0} and two-dimensional E1E_{1} IR of the 𝐂3v{\bf C}_{3{\rm v}} group. The set of eigenenergies ϵp,qtqd\epsilon_{p,q}^{\rm tqd} and eigenvectors ψp,qA0\psi^{A_{0}}_{p,q}, ψp,qB0\psi^{B_{0}}_{p,q}, and ψp,qE1±\psi^{E_{1}\pm}_{p,q} LB85 are dependent on two parameters pp and qq that have different sets of allowed values for each IR. Their concrete form is given in Appendix A.

In the second case, Eq. (41), the symmetry of the potential is dependent on the ratio k=b/a(0,1]k=b/a\in(0,1]: if k=1k=1, the symmetry of the problem is 𝐂4v{\bf C}_{4{\rm v}}; otherwise, 𝐂2v{\bf C}_{2{\rm v}} is the symmetry of the spin-independent Hamiltonian H0H_{0}. In both situations, eigenenergies and eigenvalues can be found by using the separation of variables. The set of eigenenergies ϵp,qrqd\epsilon_{p,q}^{\rm rqd} and eigenvectors ψp,qrqd\psi_{p,q}^{\rm rqd} in this case is

ϵp,qrqd\displaystyle\epsilon_{p,q}^{\rm rqd} =\displaystyle= 2π22ma2(p2+q2k2),\displaystyle\frac{\hbar^{2}\pi^{2}}{2m^{*}a^{2}}(p^{2}+\frac{q^{2}}{k^{2}}), (42)
ψp,qrqd\displaystyle\psi_{p,q}^{\rm rqd} =\displaystyle= 2aksin[pπa(x+a2)]sin[qπak(y+ka2)],\displaystyle\frac{2}{a\sqrt{k}}\sin{\big{[}\frac{p\pi}{a}(x+\frac{a}{2})\big{]}}\sin{\big{[}\frac{q\pi}{ak}(y+\frac{ka}{2})\big{]}},\;\;\; (43)

defined using the two independent parameters p1p\geq 1 and q1q\geq 1 that take integer values. However, these solutions do not have any definite symmetry CCM+05 . Therefore, they need to be symmetrized to apply the general results from Section IV. Symmetry-adapted eigenfunctions can be found in Appendix B.

Refer to caption
Figure 3: Infinite-wall equilateral triangle (left) and rectangular (right) gating potential. In both cases, potential is zero inside the area of the polygon; otherwise it is \infty.

After calculating the transition dipole matrix element and the excitation energies for two excited states in the perturbative correction check , we obtain the desired results

𝐝tqd\displaystyle{\bf d}_{\uparrow\downarrow}^{\rm tqd} =\displaystyle= 324226352π8|g|μBBma4lso2(vx𝒆x+vy𝒆y),\displaystyle\frac{3^{24}}{2^{26}35^{2}\pi^{8}}\frac{|g|\mu_{B}Bm^{*}a^{4}}{l_{\rm so}\hbar^{2}}(v_{x}{\bm{e}}_{x}+v_{y}{\bm{e}}_{y}), (44)
𝐝rqd\displaystyle{\bf d}_{\uparrow\downarrow}^{\rm rqd} =\displaystyle= 2935π6|g|μBBma4lso2(vx𝒆x+k4vy𝒆y),\displaystyle\frac{2^{9}}{3^{5}\pi^{6}}\frac{|g|\mu_{B}Bm^{*}a^{4}}{l_{\rm so}\hbar^{2}}(v_{x}{\bm{e}}_{x}+k^{4}v_{y}{\bm{e}}_{y}), (45)

where the first result corresponds to the infinite-wall equilateral triangle potential, while the second one is valid for both the infinite-wall square, k=1k=1, and rectangular, k1k\neq 1, potentials. Dipole moment constants 324/226352π83.6×1043^{24}/2^{26}35^{2}\pi^{8}\approx 3.6\times 10^{-4} and 29/35π62.2×1032^{9}/3^{5}\pi^{6}\approx 2.2\times 10^{-3} from Eqs. (44) and (45) suggest a much weaker dipole moment when compared to the harmonic gating of the same confinement length [see Eqs. (27) and (31)].

Using the relation Γ|𝐝|2\Gamma_{\uparrow\downarrow}\approx|{\bf d}_{\uparrow\downarrow}|^{2}, we conclude that square and rectangular confined QDs have a relaxation rate that is four orders of magnitude weaker than the harmonic potential; in the equilateral triangle case, a decrease of almost six orders of magnitude is observed. Thus, our result indicates a significant influence of the gating potential on the spin qubit lifetime and a beneficial role of the equilateral triangle confinement.

VI Conclusions

We have investigated the influence of the gating potential symmetry on the Rabi frequency and phonon-induced spin relaxation rate in a single-electron GaAs quantum dot. Our results suggest that, independently of the symmetry of the gating potential, both the Rabi frequency and spin relaxation rate are dependent on the orientation of the magnetic field and the spin-orbit angle. Additionally, in systems with 𝐂1v{\bf C}_{1{\rm v}}, 𝐂2v{\bf C}_{2{\rm v}}, and 𝐂n{\bf C}_{n} (n4rn\neq 4r) symmetry, orientation of the quantum dot potential with respect to the crystallographic reference frame is another degree of freedom that can be used to tune the desired properties of the system. The validity of the approach is confirmed on the known results for the isotropic and anisotropic harmonic potential. Additionally, we have compared the spin qubit lifetime in the case of an infinite-wall rectangular, square and equilateral triangle gating with the harmonic confinement. Our results indicate the enhanced lifetime of the spin qubit, reaching an almost six-orders-of-magnitude increase in the case of the equilateral triangle gating. In the end, we emphasize that in the regime of strong electric field, nonlinear effects KGS12 ; RGP15 ; VP18 cannot be fully explained by the symmetry of the gating potential, thus placing the conclusions of our work in the weak driving regime solely.

Acknowledgements.
We thank Nenad Vukmirović for fruitful discussion. This research is funded by the Serbian Ministry of Science (Project ON171035) and the National Scholarship Programme of the Slovak Republic (ID 28226).

Appendix A Particle in the infinite-wall equilateral triangle potential: eigenenergies and eigenvectors

Here we summarize the results from Ref. LB85 regarding the Schrödinger equation solution of the particle in the infinite-wall equilateral triangle potential, having 𝐂3v{\bf C}_{3{\rm v}} symmetry. Due to the symmetry, eigenvectors transform according to the one-dimensional IRs A0A_{0} and B0B_{0} and the two-dimensional IR E1E_{1}. The concrete forms of eigenenergies and eigenstates,

ϵp,qtqd=82π23ma2(p2+pq+q2),\epsilon_{p,q}^{\rm tqd}=\frac{8\hbar^{2}\pi^{2}}{3m^{*}a^{2}}(p^{2}+pq+q^{2}), (46)
ψp,qA0(x,y)\displaystyle\psi^{A_{0}}_{p,q}(x,y) =\displaystyle= cos[2πqax]sin[2π(2p+q)a3y]\displaystyle\cos{\big{[}\frac{2\pi q}{a}x\big{]}}\sin{\big{[}\frac{2\pi(2p+q)}{a\sqrt{3}}y\big{]}}
cos[2πpax]sin[2π(p+2q)a3y]\displaystyle-\cos{\big{[}\frac{2\pi p}{a}x\big{]}}\sin{\big{[}\frac{2\pi(p+2q)}{a\sqrt{3}}y\big{]}}
cos[2π(p+q)ax]sin[2π(pq)a3y],\displaystyle-\cos{\big{[}\frac{2\pi(p+q)}{a}x\big{]}}\sin{\big{[}\frac{2\pi(p-q)}{a\sqrt{3}}y\big{]}},
q\displaystyle q =\displaystyle= 0,1,2,,p=q+1,q+2,\displaystyle 0,1,2,...,\;\;\;p=q+1,q+2,... (47)
ψp,qB0(x,y)\displaystyle\psi^{B_{0}}_{p,q}(x,y) =\displaystyle= sin[2πqax]sin[2π(2p+q)a3y]\displaystyle\sin{\big{[}\frac{2\pi q}{a}x\big{]}}\sin{\big{[}\frac{2\pi(2p+q)}{a\sqrt{3}}y\big{]}}
sin[2πpax]sin[2π(p+2q)a3y]\displaystyle-\sin{\big{[}\frac{2\pi p}{a}x\big{]}}\sin{\big{[}\frac{2\pi(p+2q)}{a\sqrt{3}}y\big{]}}
+sin[2π(p+q)ax]sin[2π(pq)a3y],\displaystyle+\sin{\big{[}\frac{2\pi(p+q)}{a}x\big{]}}\sin{\big{[}\frac{2\pi(p-q)}{a\sqrt{3}}y\big{]}},
q\displaystyle q =\displaystyle= 1,2,3,,p=q+1,q+2,\displaystyle 1,2,3,...,\;\;\;p=q+1,q+2,... (48)
ψp,qE1±(x,y)\displaystyle\psi^{E_{1}\pm}_{p,q}(x,y) =\displaystyle= ψp,qB0(x,y)±iψp,qA0(x,y),\displaystyle\psi^{B_{0}}_{p,q}(x,y)\pm{\rm i}\psi^{A_{0}}_{p,q}(x,y),
q\displaystyle q =\displaystyle= 13,23,43,53,,p=q+1,q+2,\displaystyle\frac{1}{3},\frac{2}{3},\frac{4}{3},\frac{5}{3},...,\;\;\;p=q+1,q+2,...\;\; (49)

are dependent on two parameters pp and qq that have different allowed values for each IR.

Refer to caption
Figure 4: Infinite-wall equilateral triangle potential with the point group symmetry 𝐂3v{\bf C}_{3{\rm v}}. Inside the equilateral triangle potential is 0, otherwise it is \infty.

Note that the coordinate frame used to derive the previous equations (see FIG. 4) differs from the frame used in our work (see the left-hand panel of FIG. 3). To adapt the eigenfunction from Eqs. (47)-(49) to our case, a suitable change of coordinates xx+a/2x\rightarrow x+a/2 and yy+a3/6y\rightarrow y+a\sqrt{3}/6 should be made.

xxxxxxxxx

Appendix B Particle in the infinite-wall square and rectangular potential: eigenvectors

The infinite-wall square potential has 𝐂4v{\bf C}_{4{\rm v}} symmetry with the corresponding IRs A0/B0A_{0}/B_{0}, A2/B2A_{2}/B_{2}, and E1E_{1}. Eigenvectors that transform according to the given IRs and the set of allowed quantum numbers are

ψp,qA0(x,y)\displaystyle\psi^{A_{0}}_{p,q}(x,y) =\displaystyle= cos[pπax]cos[qπay]+cos[qπax]cos[pπay]\displaystyle\cos{\big{[}\frac{p\pi}{a}x\big{]}}\cos{\big{[}\frac{q\pi}{a}y\big{]}}+\cos{\big{[}\frac{q\pi}{a}x\big{]}}\cos{\big{[}\frac{p\pi}{a}y\big{]}}
q\displaystyle q =\displaystyle= 1,3,5,,p=q,q+2,q+4,\displaystyle 1,3,5,...,\;\;\;p=q,q+2,q+4,... (50)
ψp,qB0(x,y)\displaystyle\psi^{B_{0}}_{p,q}(x,y) =\displaystyle= sin[pπax]sin[qπay]sin[qπax]sin[pπay]\displaystyle\sin{\big{[}\frac{p\pi}{a}x\big{]}}\sin{\big{[}\frac{q\pi}{a}y\big{]}}-\sin{\big{[}\frac{q\pi}{a}x\big{]}}\sin{\big{[}\frac{p\pi}{a}y\big{]}}
q\displaystyle q =\displaystyle= 2,4,6,,p=q+2,q+4,\displaystyle 2,4,6,...,\;\;\;p=q+2,q+4,... (51)
ψp,qA2(x,y)\displaystyle\psi^{A_{2}}_{p,q}(x,y) =\displaystyle= cos[pπax]cos[qπay]cos[qπax]cos[pπay]\displaystyle\cos{\big{[}\frac{p\pi}{a}x\big{]}}\cos{\big{[}\frac{q\pi}{a}y\big{]}}-\cos{\big{[}\frac{q\pi}{a}x\big{]}}\cos{\big{[}\frac{p\pi}{a}y\big{]}}
q\displaystyle q =\displaystyle= 1,3,5,,p=q+2,q+4,\displaystyle 1,3,5,...,\;\;\;p=q+2,q+4,... (52)
ψp,qB2(x,y)\displaystyle\psi^{B_{2}}_{p,q}(x,y) =\displaystyle= sin[pπax]sin[qπay]+sin[qπax]sin[pπay]\displaystyle\sin{\big{[}\frac{p\pi}{a}x\big{]}}\sin{\big{[}\frac{q\pi}{a}y\big{]}}+\sin{\big{[}\frac{q\pi}{a}x\big{]}}\sin{\big{[}\frac{p\pi}{a}y\big{]}}
q\displaystyle q =\displaystyle= 2,4,6,,p=q,q+2,q+4,\displaystyle 2,4,6,...,\;\;\;p=q,q+2,q+4,... (53)

xxxxxxxxxxxxxxxxxxxxxxxxxx

ψp,qE1±(x,y)\displaystyle\psi^{E_{1}\pm}_{p,q}(x,y) =\displaystyle= cos[pπax]sin[qπay]±isin[qπax]cos[pπay]\displaystyle\cos{\big{[}\frac{p\pi}{a}x\big{]}}\sin{\big{[}\frac{q\pi}{a}y\big{]}}\pm{\rm i}\sin{\big{[}\frac{q\pi}{a}x\big{]}}\cos{\big{[}\frac{p\pi}{a}y\big{]}}
p\displaystyle p =\displaystyle= 1,3,5,,q=p+1,p+3,\displaystyle 1,3,5,...,\;\;\;q=p+1,p+3,... (54)

In the case of the infinite-wall rectangular potential 𝐂2v{\bf C}_{2{\rm v}} symmetry is relevant. Eigenfunctions transforming according to the IRs A0/B0A_{0}/B_{0} and A1/B1A_{1}/B_{1} and the corresponding set of quantum numbers are

ψp,qA0(x,y)\displaystyle\psi^{A_{0}}_{p,q}(x,y) =\displaystyle= cos[pπax]cos[qπkay]\displaystyle\cos{\big{[}\frac{p\pi}{a}x\big{]}}\cos{\big{[}\frac{q\pi}{ka}y\big{]}}
q\displaystyle q =\displaystyle= 1,3,5,,p=q,q+2,q+4,\displaystyle 1,3,5,...,\;\;\;p=q,q+2,q+4,... (55)
ψp,qB0(x,y)\displaystyle\psi^{B_{0}}_{p,q}(x,y) =\displaystyle= sin[pπax]sin[qπkay]\displaystyle\sin{\big{[}\frac{p\pi}{a}x\big{]}}\sin{\big{[}\frac{q\pi}{ka}y\big{]}}
q\displaystyle q =\displaystyle= 2,4,6,,p=q,q+2,q+4,\displaystyle 2,4,6,...,\;\;\;p=q,q+2,q+4,... (56)
ψp,qA1(x,y)\displaystyle\psi^{A_{1}}_{p,q}(x,y) =\displaystyle= cos[pπax]sin[qπkay]\displaystyle\cos{\big{[}\frac{p\pi}{a}x\big{]}}\sin{\big{[}\frac{q\pi}{ka}y\big{]}}
p\displaystyle p =\displaystyle= 1,3,5,,q=p+1,p+3,p+5,\displaystyle 1,3,5,...,\;\;\;q=p+1,p+3,p+5,...\;\;\; (57)
ψp,qB1(x,y)\displaystyle\psi^{B_{1}}_{p,q}(x,y) =\displaystyle= sin[pπax]cos[qπkay]\displaystyle\sin{\big{[}\frac{p\pi}{a}x\big{]}}\cos{\big{[}\frac{q\pi}{ka}y\big{]}}
q\displaystyle q =\displaystyle= 1,3,5,,p=q+1,q+3,q+5,\displaystyle 1,3,5,...,\;\;\;p=q+1,q+3,q+5,...\;\;\; (58)

In both cases, eigenenergies are given in Eq. (42) (k=1k=1 in the 𝐂4v{\bf C}_{4{\rm v}} case and k1k\neq 1 for the 𝐂2v{\bf C}_{2{\rm v}} symmetry).

References

  • (1) M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information, Cambridge University Press, (2010).
  • (2) C. H. Bennett and D. P. DiVincenzo, Nature 404, 247-255 (2000).
  • (3) D. Loss and D. P. DiVincenzo, Phys. Rev. A 57, 120 (1998).
  • (4) T. P. Orlando, J. E. Mooij, L. Tian, C. H. van der Wal, L. S. Levitov, S. Lloyd, and J. J. Mazo, Phys. Rev. B 60, 15398 (1999).
  • (5) Y. Nakamura, Yu. A. Pashkin, and J. S. Tsai, Nature 398, 786-788 (1999).
  • (6) M. Yamamoto, S. Takada, C. Bäuerle, K. Watanabe, A. D. Wieck, and S. Tarucha, Nat. Nanotechnol. 7, 247-251 (2012).
  • (7) E. I. Rashba, Sov. Phys. Solid State 2, 1109 (1960).
  • (8) V. N. Golovach, M. Borhani, and D. Loss, Phys. Rev. B 74, 165319 (2006).
  • (9) D. V. Bulaev and D. Loss, Phys. Rev. Lett. 98, 097202 (2007).
  • (10) K. C. Nowack, F. H. L. Koppens, Yu. V. Nazarov, and L. M. K. Vandersypen, Science 318, 1430-1433 (2007).
  • (11) E. I. Rashba, Phys. Rev. B 78, 195302 (2008).
  • (12) R. Brunner, Y.-S. Shin, T. Obata, M. Pioro-Ladrière, T. Kubo, K. Yoshida, T. Taniyama, Y. Tokura, and S. Tarucha, Phys. Rev. Lett. 107, 146801 (2011).
  • (13) E. Kawakami, P. Scarlino, D. R. Ward, F. R. Braakman, D. E. Savage, M. G. Lagally, Mark Friesen, S. N. Coppersmith, M. A. Eriksson, and L. M. K. Vandersypen, Nat. Nanotechnol. 9, 666-670 (2014).
  • (14) K. Takeda, J. Yoneda, T. Otsuka, T. Nakajima, M. R. Delbecq, G. Allison, Y. Hoshi, N. Usami, K. M. Itoh, S. Oda, T. Kodera, and S. Tarucha, npj Quantum Inf. 4, 54 (2018).
  • (15) D. V. Khomitsky, E. A. Lavrukhina, and E. Ya. Sherman, Phys. Rev. B 99, 014308 (2019).
  • (16) S. Studenikin, M. Korkusinski, M. Takahashi, J. Ducatel, A. Padawer-Blatt, A. Bogan, D. Guy Austing, L. Gaudreau, P. Zawadzki, A. Sachrajda, Y. Hirayama, L. Tracy, J. Reno, and T. Hargett, Commun. Phys. 2, 159 (2019).
  • (17) A. V. Khaetskii and Y. V. Nazarov, Phys. Rev. B 61, 12639 (2000); A. V. Khaetskii and Y. V. Nazarov, Phys. Rev. B 64, 125316 (2001).
  • (18) P. Stano and J. Fabian, Phys. Rev. B 72, 155410 (2005).
  • (19) V. N. Stavrou, J. Phys.: Condens. Matter 29, 485301 (2017); V. N. Stavrou, J. Phys.: Condens. Matter 30, 455301 (2018).
  • (20) Z.-H. Liu, R. Li, X. Hu and J. Q. You, Sci. Rep. 8, 2302 (2018).
  • (21) J. I. Climente, A. Bertoni, G. Goldoni, M. Rontani, and E. Molinari, Phys. Rev. B 75, 081303(R) (2007).
  • (22) O. Malkoc, P. Stano, and D. Loss, Phys. Rev. B 93, 235413 (2016).
  • (23) D. Chaney and P. A. Maksym, Phys. Rev. B 75, 035323 (2007).
  • (24) V. I. Fal’ko, B. L. Altshuler, and O. Tsyplyatyev, Phys. Rev. Lett. 95, 076603 (2005).
  • (25) P. Scarlino, E. Kawakami, P. Stano, M. Shafiei, C. Reichl, W. Wegscheider, and L. M. K. Vandersypen, Phys. Rev. Lett. 113, 256802 (2014).
  • (26) O. Olendski and T. V. Shahbazyan, Phys. Rev. B 75, 041306(R) (2007).
  • (27) S. Amasha, K. MacLean, I. P. Radu, D. M. Zumbühl, M. A. Kastner, M. P. Hanson, and A. C. Gossard, Phys. Rev. Lett. 100, 046803 (2008).
  • (28) J. Planelles, F. Rajadell, and J. I. Climente, Phys. Rev. B 92, 041302(R) (2015).
  • (29) M. Raith, P. Stano, and J. Fabian, Phys. Rev. B 83, 195318 (2011).
  • (30) G. Dresselhaus, Phys. Rev. 100, 580 (1955).
  • (31) E. I. Rashba, Fiz. Tv. Tela (Leningrad) 2, 1224 (1960); Sov. Phys. Solid State 2, 1109 (1960).
  • (32) E. Ya. Sherman and D. J. Lockwood, Phys. Rev. B 72, 125340 (2005).
  • (33) M. Raith, T. Pangerl, P. Stano, and J. Fabian, Phys. Status Solidi b 251, 1924 (2014).
  • (34) J. L. Cheng, M. W. Wu, and C. Lü, Phys. Rev. B 69, 115318 (2004).
  • (35) F. Baruffa, P. Stano, and J. Fabian, Phys. Rev. Lett. 104, 126401 (2010).
  • (36) R. de Sousa and S. Das Sarma, Phys. Rev. B 68, 155330 (2003).
  • (37) X. Linpeng, T. Karin, M. V. Durnev, R. Barbour, M. M. Glazov, E. Ya. Sherman, S. P. Watkins, S. Seto, and K.-M. C. Fu, Phys. Rev. B 94, 125401 (2016).
  • (38) Note that both the wave vector 𝐪{\bf q} and the electron coordinate 𝐫c{\bf r}_{c} are written in the crystallographic reference frame.
  • (39) J. I. Climente, A. Bertoni, G. Goldoni, and E. Molinari, Phys. Rev. B 74, 035313 (2006).
  • (40) L. C. Camenzind, L. Yu, P. Stano, J. D. Zimmerman, A. C. Gossard, D. Loss, and D. M. Zumbühl, Nat. Commun. 9, 3454 (2018).
  • (41) E. Lijnen, L. F. Chibotaru, and A. Ceulemans, Phys. Rev. E 77, 016702 (2008).
  • (42) L. Jansen and M. Boon, Theory of Finite Groups: Applications in Physics, North Holland, Amsterdam, (1967); M. Damnjanović, O simetriji u kvantnoj nerelativističkoj fizici, Fizički fakultet, Beograd (2000). http://www.ff.ac.rs/Katedre/QMF/SiteQMF/pdf/sknf2e.pdf
  • (43) In the case of 𝐂nv/𝐂v{\bf C}_{n{\rm v}}/{\bf C}_{\infty{\rm v}} groups, time-reversal Θ\Theta is the symmetry of the system. However, it can be checked that Θ\Theta can be safely neglected. Time-reversal in the case of vectors from one-dimensional IRs has a trivial action. In the case of the two-dimensional IRs EmE_{m} time-reversal transforms one vector of the IR into the other; in other words, it has the same behavior as the vertical mirror plane symmetry σv\sigma_{\rm v} and can be ignored.
  • (44) W.-K. Li and S. M. Blinder, J. Math. Phys. 26, 2784 (1985).
  • (45) L. F. Chibotaru, A. Ceulemans, M. Morelle, G. Teniers, C. Carballeira, and V. V. Moshchalkov, J. Math. Phys. 46, 095108 (2005).
  • (46) We have explicitly checked that other states appearing in the perturbative expansion can be safely ignored.
  • (47) D. V. Khomitsky, L. V. Gulyaev, and E. Ya. Sherman, Phys. Rev. B 85, 125312 (2012).
  • (48) J. Romhányi, G. Burkard, and A. Pályi, Phys. Rev. B 92, 054422 (2015).
  • (49) M. T. Veszeli and A. Pályi, Phys. Rev. B 97, 235433 (2018).