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

Light-hole spin confined in germanium

Patrick Del Vecchio Department of Engineering Physics, École Polytechnique de Montréal, Montréal, C.P. 6079, Succ. Centre-Ville, Montréal, Québec, Canada H3C 3A7    Oussama Moutanabbir [email protected] Department of Engineering Physics, École Polytechnique de Montréal, Montréal, C.P. 6079, Succ. Centre-Ville, Montréal, Québec, Canada H3C 3A7
Abstract

The selective confinement of light holes (LHs) in a tensile-strained germanium (Ge) quantum well is studied by mapping the electronic structure of Ge1-xSnx/Ge/Ge1-xSnx heterostructures as a function of Sn content, residual strain, and Ge well thickness. It is shown that above 12at.%12\,\text{at.}\% Sn and below 0.4%0.4\% residual compressive strain in the barriers, the tensile strain in Ge becomes sufficiently large to yield a valence band edge with LH-like character, thus forming a quasi two-dimensional LH gas in Ge. The LH ground state has a larger in-plane effective mass than that of heavy holes (HHs) in Si1-yGey/Ge/Si1-yGey quantum wells. Moreover, LHs in optimal Ge1-xSnx/Ge/Ge1-xSnx heterostructures are found to exhibit a strong gg-tensor anisotropy, with the in-plane component one order of magnitude larger than that of HHs in typical planar systems. Two of three structure-inversion-asymmetry Rashba parameters, both of which are critical in electric-dipole-spin-resonance experiments, are effectively 10 times the size of the cubic Rashba parameter in HH quantum wells. In the regime of LH selective confinement, every layer of the heterostructure is of direct bandgap, which can be relevant for efficient optical photon-spin qubit interfaces. This work discusses the broad landscape of the characteristics of LH spin confined in Ge to guide the design and implementation of LH spin-based devices.

I Introduction

Hole spins in group-IV planar gated quantum dots are promising candidates for robust and scalable qubits [1, 2, 3, 4, 5, 6, 7]. Developing these qubits has been so far exclusively based on heavy-hole (HH) states, as the materials currently used are restricted to compressively strained germanium (Ge) heterostructures. Interestingly, the advent of the germanium/germanium-tin (Ge/Ge1-xSnx) material system provides an additional degree of freedom to tailor the valence band character in tensile-strained Ge, thus paving the way to implement silicon-compatible platforms for light-hole (LH) spin qubits [8, 9]. These LH qubits share many of the advantages benefiting the HH ones but also bring about other attractive characteristics pertaining to LHs and Ge/Ge1-xSnx heterostructures. These include a strong Rashba-type spin-orbit interaction (SOI) [9] and an efficient coupling with proximity-induced superconductivity [10] in addition to the bandgap directness [11] relevant to coupling with optical photons. These characteristics can expand the functionalities of hole spin qubits by facilitating the implementation of hybrid superconductor-semiconductor devices and photon-spin interfaces.

Although Ge1-xSnx semiconductors have been the subject of extensive studies in recent years, research in this area has mainly been focused on integrated photonics and optoelectronics leveraging Ge1-xSnx strain- and composition-dependent bandgap [11] leaving their spin properties practically unexplored [12, 13, 14, 9]. As a matter of fact, studies on hole spin in Ge/Ge1-xSnx are still conspicuously missing in the literature. Herein, this work addresses the dynamics of LH spin confined in Ge1-xSnx/Ge/Ge1-xSnx heterostructures and elucidates the key parameters affecting its behavior as a function of strain, well thickness, Sn content, and magnetic field orientation. First, the article describes and discusses the electronic structure of tensile-strained Ge on Ge1-xSnx. Second, the parameters defining the band alignment of the Ge1-xSnx/Ge/Ge1-xSnx quantum well are introduced and the criteria for LH confinement in Ge are established. The third section outlines the Hamiltonian of the in-plane motion of LHs for out-of-plane and in-plane magnetic fields yielding LH parameters such as the effective mass, the gg-tensor, and the Rashba-SOI parameters. LH-HH mixing within the LH ground state is investigated in the fourth section. It is important to note that the focus here is on LH properties in the planar system without the effects of electrostatic in-plane confinement introduced in quantum dot systems. [9]

II LH quantum well in Ge1-xSnx/Ge/Ge1-xSnx

II.1 Ge1-xSnx/Ge/Ge1-xSnx heterostructure

Before delving into the details of the electronic structure of Ge/Ge1-xSnx, it is instructive to examine the strain-related behavior of bulk Ge. Figure 1(a) outlines the effect of tensile strain on the band structure of bulk Ge calculated by 8-band (full lines) and 4-band (dashed lines) kpk\cdot p theory. Here, the calculations assume a bi-isotropic biaxial strain in the (001)(001)-plane without any shear deformations, which is expected for an ideal [001][001]-oriented epitaxial growth [15, 8]. Under these conditions, the fourfold degeneracy of the VB edge at the Γ\Gamma point is lifted yielding two spin-degenerate LH and HH bands. In the case of tensile strain, the VB edge is LH-like, whereas in the broadly studied compressively strained Ge, the VB edge is of the HH type. As discussed in the following, there is a threshold of tensile strain beyond which it becomes possible to control and selectively manipulate spin 1/2 LHs instead of spin 3/2 HHs.

The proposed heterostructure consists of a Ge1-xSnx/Ge/Ge1-xSnx quantum well grown on silicon [8], as illustrated in Fig. 1(b). Thick Ge1-xSnx buffer layers with an increasing Sn content from the Ge virtual substrate (VS) up to the bottom Ge1-xSnx barrier prevents the propagation of defects and dislocations near the Ge quantum well [8]. The lattice mismatch between Ge1-xSnx and Ge (with Sn contents above 10at.%{\sim}10\,\text{at.}\,\%) is leveraged to achieve high tensile strain in the coherently grown Ge quantum well. A top Ge1-xSnx barrier is then grown on the tensile-strained Ge layer. As shown in the following, careful engineering of the lattice strain and Sn content leads to a direct bandgap Ge LH quantum well.

Refer to caption
Figure 1: (a) Bulk dispersion near the Γ\Gamma point (𝐤2.5nm1\|\mathbf{k}\|\leq 2.5\,\text{nm}^{-1} in both directions) of relaxed Ge (left) and tensile-strained Ge (right) computed by the 8-band (full lines) and the 4-band (dashed lines) kpk\cdot p frameworks. The nonzero components of the strain tensor are εxx=εyy=2%\varepsilon_{xx}=\varepsilon_{yy}=2\% and εzz=2(c12/c11)εxx1.3%\varepsilon_{zz}=-2(c_{12}/c_{11})\varepsilon_{xx}\approx-1.3\%. The conduction band was obtained from an effective mass approximation in the 4-band computations. (b) Schematic illustration of the proposed Ge1-xSnx/Ge/Ge1-xSnx heterostructure. (c) Band alignment profile of a Ge1-xSnx/Ge/Ge1-xSnx quantum well for selected barrier parameters. In this case, ΔE1,2,3\Delta E_{1,2,3} and LBO are positive. εxx\varepsilon_{xx} in each layer is written in parentheses.

II.2 Band alignment and LH confinement

The energy band alignment of the Ge1-xSnx/Ge/Ge1-xSnx heterostructure is computed within the assumption that the in-plane lattice constant aa_{\parallel} is the same in each layer, which is consistent with a pseudomorphic epitaxial growth. The in-plane components εxx\varepsilon_{xx} and εyy\varepsilon_{yy} of the strain tensor in a material ii are given by εxx=εyy=a/a0i1\varepsilon_{xx}=\varepsilon_{yy}=a_{\parallel}/a_{0}^{i}-1, where a0ia_{0}^{i} is the lattice constant of the fully relaxed material ii. The out-of-plane strain component εzz\varepsilon_{zz} then follows immediately from the relation εzz=2(c12/c11)εxx\varepsilon_{zz}=-2(c_{12}/c_{11})\varepsilon_{xx}, where c11c_{11} and c12c_{12} are the material elastic constants (see Table 1) and εkl=0\varepsilon_{kl}=0 if klk\neq l [15]. Three parameters determine completely the band alignment of the Ge1-xSnx/Ge/Ge1-xSnx heterostructure: the Sn content xx, the in-plane strain εBR\varepsilon_{\text{BR}} in the barriers, and the in-plane strain in the Ge well εW\varepsilon_{\text{W}}. Since the latter can be counted for from the condition that aa_{\parallel} is constant along the growth direction zz, only the barrier composition and strain (xx, εBR\varepsilon_{\text{BR}}) are required to evaluate the band alignment.

A typical band alignment is displayed in Fig. 1(c) for x=0.13x=0.13 and εBR=0.125%\varepsilon_{\text{BR}}=-0.125\%. In this instance, HHs and electrons are pushed away from Ge and form a continuum of states in the Ge0.87Sn0.13 barriers. Meanwhile, LHs are selectively confined in Ge, thereby forming a LH quantum well. Here, the combination of large strain in Ge (εW=1.95%\varepsilon_{\text{W}}=1.95\%) with small εBR\varepsilon_{\text{BR}} pushes the LH ground state (LH1) above the HH continuum, leading to the possibility, at very low hole density, to populate only LH1 and thus to create a pure quasi two-dimensional LH gas in Ge. Such system is achievable only in a specific region of the parameter space (xx, εBR\varepsilon_{\text{BR}}), depending on four energy offsets [see Fig. 1(c)]:

ΔE1\displaystyle\Delta E_{1} =max(LH)max(HH),\displaystyle=\max(\text{LH})-\max(\text{HH}), (1)
ΔE2\displaystyle\Delta E_{2} =ELH1max(HH),\displaystyle=E_{\text{LH}1}-\max(\text{HH}), (2)
ΔE3\displaystyle\Delta E_{3} =ELH1min(LH),\displaystyle=E_{\text{LH}1}-\min(\text{LH}), (3)
LBO=max(LH)min(LH)=ΔE1ΔE2+ΔE3.\displaystyle\begin{split}\text{LBO}&=\max(\text{LH})-\min(\text{LH})\\ &=\Delta E_{1}-\Delta E_{2}+\Delta E_{3}.\end{split} (4)

Here, max(LH)\max(\text{LH}) and min(LH)\min(\text{LH}) are the energies at the bottom and top of the LH quantum well, respectively, and max(HH)\max(\text{HH}) is the energy at the edge of the HH continuum. The zero energy point is placed on the ground LH subband (i.e., ELH1=0E_{\text{LH1}}=0). Band offsets ΔE1\Delta E_{1} and LBO (LH band offset) do not depend on the well thickness ww.

Figure 2(a) presents a two-dimensional map of band offsets LBO and ΔE1\Delta E_{1} with xx and εBR\varepsilon_{\text{BR}} as independent parameters. The corresponding strain in the Ge well εW\varepsilon_{\text{W}} is also shown (black dotted lines) only for the tensile strain regime (εW>0\varepsilon_{\text{W}}>0). The Ge indirect-to-direct transition occurs at εW=1.68%\varepsilon_{\text{W}}=1.68\% (solid black line) according to the parameterization described in Appendix A. Similarly, the Ge1-xSnx barriers exhibit bandgap directness above the dashed-dot blue line. Constant LBO are indicated by the solid red curves, where LBO=0\text{LBO}=0 corresponds to a completely flat LH profile along the growth direction. Finally, dashed red curves indicate constant ΔE1\Delta E_{1}, where ΔE1=0\Delta E_{1}=0 corresponds to the LH band edge in Ge sitting at the same energy as the HH band edge in the barriers. As discussed in the following, a large and positive ΔE1\Delta E_{1} allows LH1 to emerge from the continuum for sufficiently thick quantum wells.

Refer to caption
Figure 2: (a) Two-dimensional contour map of band offsets ΔE1\Delta E_{1} and LBO, strain in Ge, and bandgap directness as a function of xx and εBR\varepsilon_{\text{BR}} at T=300KT=300\,\text{K}. Contour lines for LBO and ΔE1\Delta E_{1} are given at 0, 5050, and 100meV100\,\text{meV}. The star indicates the system from Fig. 1(c). (b) Minimal well thickness w0w_{0} (solid lines) required for a LH-like valence band edge as a function of εBR\varepsilon_{\text{BR}} at fixed Sn content in the barriers. Dashed lines are the Ge critical growth thickness hch_{c} estimated from the People and Bean relation. Solid circles indicate where w0=hcw_{0}=h_{c} for a given xx. For x=0.10x=0.10, hc46nmh_{c}\approx 46\,\text{nm} at εBR=0\varepsilon_{\text{BR}}=0.

LHs are confined in Ge if and only if LBO>0\text{LBO}>0. However, depending on ΔE1\Delta E_{1} and the well thickness ww, these confined LHs could be situated within the HH continuum (negative ΔE2\Delta E_{2}). In the region ΔE10\Delta E_{1}\leq 0, LHs can never emerge from the continuum since the bottom of the LH well is below the HH band edge. However, if ΔE1>0\Delta E_{1}>0, there is a minimal QW thickness w0w_{0} for which LH1 emerges from the continuum. This lower bound depends on both Sn content and lattice strain (i.e., xx and εBR\varepsilon_{\text{BR}}) in the barriers and is plotted in Fig. 2(b) (solid lines). At w=w0w=w_{0}, the energy required for LH1 to escape the well is exactly LBOΔE1\text{LBO}-\Delta E_{1} (i.e., the strain-induced HH-LH splitting in the barriers). w0w_{0}\to\infty on the ΔE1=0\Delta E_{1}=0 curve, whereas w0=0w_{0}=0 if the barriers are fully relaxed. In addition to the lower bound w0<ww_{0}<w, the maximum strain energy that the Ge QW can accommodate also introduces an upper bound on ww. This upper bound is given by the Ge critical growth thickness hch_{c}, beyond which misfit dislocations start to appear at the interfaces and tensile strain in Ge is strongly suppressed. The dashed lines in Fig. 2(b) show an estimation of hch_{c} based on the People and Bean formula [16, 17]. Here, this model is applied for the critical thickness of a Ge layer with equilibrium lattice constant a0Gea0(0)a_{0}^{\text{Ge}}\equiv a_{0}(0) pseudomorphically grown on Ge1-xSnx with lattice constant a0(x)(1+εBR)a_{0}(x)\left(1+\varepsilon_{\text{BR}}\right).

III LH spin properties

III.1 Effective masses and spin parameters

In this section, the LH subband effective mass, the out-of-plane and in-plane gg factors, and the Rashba parameters are computed as a function of Sn content in the Ge1-xSnx barriers (xx) and well thickness ww. These parameters give important information on how LHs move in the plane and how they respond to magnetic fields. Despite being intrinsic to a given subband (here, the focus is given to the lowest subband LH1), there is generally a significant influence from neighboring levels through inter-subband couplings. Moreover, the wavefunction spread across interfaces can also influence subband parameters. In heterostructures such as Ge/Si1-yGey quantum wells, where both types of holes are confined into the same layer and the band offsets are large, intersubband coupling is significant to an energy scale comprising only a few tens of subbands due to quantization effects. For instance, it is often a reasonable approximation to include the coupling to only 1 or 2 LH subbands for a HH ground state [18], or around 50 LH subbands when band offsets are taken into account [19]. In Ge/Ge1-xSnx quantum wells, intersubband couplings must include 102{\sim}10^{2} subbands due to the neighboring continuum. Moreover, small LBOs leading to a sizable spread of LH1 into the barriers require an accurate description of the subband envelopes. To address these effects, a numerical approach is employed instead of a variational method for the envelope problem [9]. Moreover, due to the nearby continuum, the coupling from the 200200 closest subbands to LH1 are taken into account. Our implementation of kpk\cdot p theory and how strain is incorporated into the model is described in Appendix B.

From the point of view of 8-band kpk\cdot p theory, a subband such as LH1 always consists of the superposition of spin 1/21/2 CB electron, LH, and split-off (SO) hole envelopes. At 𝐤=0\mathbf{k}_{\parallel}=0, this can be written as

|η,σ=|12,σ2c|c+|32,σ2|+σ|12,σ2|s,\Ket{\eta,\sigma}=\Ket{\frac{1}{2},\frac{\sigma}{2}}_{c}\Ket{c}+\Ket{\frac{3}{2},\frac{\sigma}{2}}\Ket{\ell}+\sigma\Ket{\frac{1}{2},\frac{\sigma}{2}}\Ket{s}, (5)

where σ=±1\sigma=\pm 1 is a pseudo-spin quantum number and z|c,,s=ψc,,s(z)\Braket{z}{c,\ell,s}=\psi_{c,\ell,s}(z) are the CB, LH, and SO envelope functions, respectively. To avoid any confusion, “LH” subbands (e.g., LH1) are designated as η\eta subbands to distinguish them from their LH envelope component. The kets |32,σ2|j,m\Ket{\frac{3}{2},\frac{\sigma}{2}}\equiv\ket{j,m} and so on are the bulk Bloch states. The additional contribution from HHs away from 𝐤=0\mathbf{k}_{\parallel}=0 is investigated in the next section. An η\eta subband is normalized according to

1=τ={c,,s}τ|τ,1=\sum_{\tau=\{c,\ell,s\}}{\Braket{\tau}{\tau}}, (6)

with |>c|c\Braket{\ell}{\ell}>\Braket{c}{c} and |>s|s\Braket{\ell}{\ell}>\Braket{s}{s} for a level such as LH1. For instance, the SO contribution s|s\Braket{s}{s} in LH1 is typically smaller than 10%10\% for the range of barrier and well parameters considered here, but it plays an important role in the effective mass calculations, as discussed in the following. The CB contribution c|c\Braket{c}{c} in LH1 is around 5%5\%, where the envelopes |c\Ket{c} are antisymmetric with respect to the center of the well and have their maximal amplitude near the interfaces.

For an out-of-plane magnetic field 𝐁=B𝐞z\mathbf{B}=B\mathbf{e}_{z}, the in-plane motion of |η,σ\Ket{\eta,\sigma} is described by an effective two-dimensional Hamiltonian (see Appendix C)

Heff=α0γK2+α0λ2g2σz+iβ1(Kσ+h.c.)iβ2(K+3σ+h.c.)+iβ3(KK+Kσ+h.c.),\begin{split}H_{\text{eff}}^{\perp}&=\alpha_{0}\gamma K_{\parallel}^{2}+\frac{\alpha_{0}}{\lambda^{2}}\frac{g_{\perp}}{2}\sigma_{z}\\ &+i\beta_{1}\left(K_{-}\sigma_{+}-\text{h.c.}\right)-i\beta_{2}\left(K_{+}^{3}\sigma_{+}-\text{h.c.}\right)\\ &+i\beta_{3}\left(K_{-}K_{+}K_{-}\sigma_{+}-\text{h.c.}\right),\end{split} (7)

where α0=2/(2m0)\alpha_{0}=\hbar^{2}/(2m_{0}) with m0m_{0} the free electron mass, 𝐊=𝐤+e𝐀/\mathbf{K}=\mathbf{k}+e\mathbf{A}/\hbar is the mechanical wavevector, 𝐀=B/2(y𝐞x+x𝐞y)\mathbf{A}=B/2(-y\mathbf{e}_{x}+x\mathbf{e}_{y}) is the vector potential such that 𝐁=×𝐀\mathbf{B}=\nabla\times\mathbf{A}, and 𝐤i\mathbf{k}\to-i\nabla is the canonical wavevector. Notably, K2={K,K+}/2=Kx2+Ky2K_{\parallel}^{2}=\{K_{-},K_{+}\}/2=K_{x}^{2}+K_{y}^{2}, 1/λ2=[K,K+]/2=eB/1/\lambda^{2}=[K_{-},K_{+}]/2=eB/\hbar, and K±=Kx±iKyK_{\pm}=K_{x}\pm iK_{y}. The effective parameters in (7) are the following: γ=m0/m\gamma=m_{0}/m^{*} is the in-plane inverse effective mass, gg_{\perp} is the out-of-plane gg factor, and β1,2,3\beta_{1,2,3} represent the three types of Rashba splittings. The first is linear in KK, whereas the last two are cubic in KK. Rashba parameters arise from space inversion asymmetry. When a small DC electric field 𝐄=Ez𝐞z\mathbf{E}=E_{z}\mathbf{e}_{z} is applied to an otherwise symmetric well, all β\beta parameters behave linearly with EzE_{z} and involve only odd powers of the field:

βi=αiEz+O(Ez3),\beta_{i}=\alpha_{i}E_{z}+O\left(E_{z}^{3}\right), (8)

where i=1,2,3i=1,2,3. For an in-plane magnetic field 𝐁=B(𝐞xcosϕ+𝐞ysinϕ)\mathbf{B}=B(\mathbf{e}_{x}\cos\phi+\mathbf{e}_{y}\sin\phi) with vector potential 𝐀=B(𝐞xsinϕ𝐞ycosϕ)z\mathbf{A}=B(\mathbf{e}_{x}\sin\phi-\mathbf{e}_{y}\cos\phi)z, the in-plane motion is instead described by

Heff=α0γk2+α0λ2g2(eiϕσ++h.c.)+iβ1(kσ+h.c.)iβ2(k+3σ+h.c.)+iβ3(kk+kσ+h.c.).\begin{split}H_{\text{eff}}^{\parallel}&=\alpha_{0}\gamma k_{\parallel}^{2}+\frac{\alpha_{0}}{\lambda^{2}}\frac{g_{\parallel}}{2}\left(e^{-i\phi}\sigma_{+}+\text{h.c.}\right)\\ &+i\beta_{1}\left(k_{-}\sigma_{+}-\text{h.c.}\right)-i\beta_{2}\left(k_{+}^{3}\sigma_{+}-\text{h.c.}\right)\\ &+i\beta_{3}\left(k_{-}k_{+}k_{-}\sigma_{+}-\text{h.c.}\right).\end{split} (9)

The in-plane gg factor is given by

g2={c|zgkz|c2+|u+|}2{13α0c|zP|++|s},\begin{split}\frac{g_{\parallel}}{2}&=\Im\left\{\Braket{c}{z{\color[rgb]{0,0,0}g}k_{z}}{c}-2\Braket{+}{{\color[rgb]{0,0,0}u_{+}^{\prime}}}{-}\right\}\\ &-\sqrt{2}\Re\left\{\frac{1}{\sqrt{3}\alpha_{0}}\Braket{c}{zP}{+}+\Braket{-}{s}\right\},\end{split} (10)

where

|±=|±2±1/2|s,\displaystyle\ket{\pm}=\ket{\ell}\pm 2^{\pm 1/2}\ket{s}, (11)
u±={zγ3,kz}±[zκ,kz],\displaystyle{\color[rgb]{0,0,0}u_{\pm}^{\prime}}=\{z{\color[rgb]{0,0,0}\gamma_{3}},k_{z}\}{\color[rgb]{0,0,0}\pm}[z{\color[rgb]{0,0,0}\kappa},k_{z}], (12)

with γ3\gamma_{3} a Luttinger parameter, κ\kappa the bulk hole gg-factor, and PP the so-called Kane momentum matrix element (see Appendix A). In the context of heterostructures, material parameters such as γ3\gamma_{3}, κ\kappa, or PP are operators that act on envelope functions. Importantly, they do not commute with kzk_{z} and are diagonal in position basis. For example, γ3|z=γ3(z)|z\gamma_{3}\ket{z}=\gamma_{3}(z)\ket{z}, where the function γ3(z)\gamma_{3}(z) gives the value of γ3\gamma_{3} at coordinate zz. In (10), the z=0z=0 coordinate is chosen such that

z=τ={c,,s}τ|z|τ=0.\braket{z}=\sum_{\tau=\{c,\ell,s\}}{\Braket{\tau}{z}{\tau}}=0. (13)
Refer to caption
Figure 3: LH1 subband parameters as a function of the well thickness and xx for fully relaxed Ge1-xSnx barriers (εBR=0\varepsilon_{\text{BR}}=0). (a) Inverse effective mass γ\gamma. (b) Out-of-plane gg factor. (c) In-plane gg factor. (d), (e), and (f) are α1\alpha_{1}, α2\alpha_{2}, and α3\alpha_{3} Rashba parameters, respectively. The calculations were carried out for a well thickness in the 551515 nm range. The data displayed here are for thicknesses separated by a 2 nm step.

This ensures, when setting kx=ky=0k_{x}=k_{y}=0, that gg_{\parallel} is gauge independent and corresponds to that at the subband edge. This is because the quantum numbers k±k_{\pm} are generally gauge dependent, and if z=0\braket{z}=0, taking the expectation value on both sides of K±=k±ize±iϕ/λ2K_{\pm}=k_{\pm}\mp ize^{\pm i\phi}/\lambda^{2} gives K±=k±\Braket{K_{\pm}}=k_{\pm}, and thus associates to k±k_{\pm} the gauge independent quantity K±\Braket{K_{\pm}}. Equation (10) reduces to the well-known |g|=4κ|g_{\parallel}|=4{\color[rgb]{0,0,0}\kappa} in the special case of 4-band Luttinger Hamiltonian with LBO\text{LBO}\to\infty. The in-plane effective mass γ\gamma, the gg-factor components, and the three Rashba parameters αi\alpha_{i} are plotted as a function of xx and ww in Fig. 3 panels (a), (b)–(c), and (d)–(f) respectively. A negative gg means that the spin-down level (σ=1\sigma=-1) is closer to the bandgap than the spin-up level.

III.2 LH-HH mixing

In the vicinity of 𝐤=0\mathbf{k}_{\parallel}=0, η\eta subbands acquire a small HH component in addition to the three terms in (5), resulting in a η\eta-HH mixed state |ψ,𝐤\ket{\psi,\mathbf{k}_{\parallel}}. For LH1, this can be written as (up to a normalization constant and to first order in kk):

|ψ,𝐤=|η,σiσα0kσ|32,3σ2lTlx|hlELH1ElH+,\ket{\psi,\mathbf{k}_{\parallel}}=\ket{\eta,\sigma}-i\sigma\alpha_{0}k_{-\sigma}\Ket{\frac{3}{2},\frac{3\sigma}{2}}\sum_{l}{\frac{{\color[rgb]{0,0,0}T_{l}^{\text{x}}}\ket{h_{l}}}{E_{\text{LH1}}-E_{l}^{\text{H}}}}+\dots, (14)

where |32,3σ2\Ket{\frac{3}{2},\frac{3\sigma}{2}} is the HH bulk Bloch state, |hl\ket{h_{l}} is the ll-th HH envelope with energy ElHE_{l}^{\text{H}} at 𝐤=0\mathbf{k}_{\parallel}=0, and the coefficients TlxT_{l}^{\text{x}} are

Tlx=hl|(P2α0|c3iu+|),\displaystyle{\color[rgb]{0,0,0}T_{l}^{\text{x}}}=\bra{h_{l}}\left(\frac{P}{\sqrt{2}\alpha_{0}}\ket{c}-\sqrt{3}i{\color[rgb]{0,0,0}u}_{+}\ket{-}\right), (15)
u±={γ3,kz}±[κ,kz].\displaystyle{\color[rgb]{0,0,0}u}_{\pm}=\{{\color[rgb]{0,0,0}\gamma_{3}},k_{z}\}\pm[{\color[rgb]{0,0,0}\kappa},k_{z}]. (16)

To first order in kk, mixing is stronger between η\eta and HH subbands with opposite parity (from kzk_{z} terms in TlxT_{l}^{\text{x}}) and with the same spin component sign (i.e., |η,σ\ket{\eta,\sigma} couples with |32,3σ2\Ket{\frac{3}{2},\frac{3\sigma}{2}}). The HH contribution ρ\rho in the mixed subband |ψ,𝐤\Ket{\psi,\mathbf{k}_{\parallel}} is given by the sum of the absolute square of each coefficient in front of HH terms. By symmetry, only even powers of kk must appear in ρ\rho:

ρ=ak2+O(k4),\rho=ak_{\parallel}^{2}+O(k^{4}), (17)

where aa can be found from (14):

a=α02l|Tlx|2(ELH1ElH)2.a=\alpha_{0}^{2}\sum_{l}{\frac{|{\color[rgb]{0,0,0}T_{l}^{\text{x}}}|^{2}}{\left(E_{\text{LH1}}-E_{l}^{\text{H}}\right)^{2}}}. (18)

The expression ρak2\rho\approx ak_{\parallel}^{2} is valid for small 𝐤\mathbf{k}_{\parallel} such that ρ1\rho\ll 1. In general, ρ\rho lies in the interval [0,1][0,1] with ρ=0\rho=0 (ρ=1\rho=1) corresponding to a pure η\eta (HH) subband.

ρ\rho as a function of kxk_{x} is displayed in Fig. 4(a) for w=6nmw=6\,\text{nm} and w=10nmw=10\,\text{nm} at x=0.13x=0.13. When kxk_{x} is small, the parabolic term in (17) fits well the numerically computed ρ\rho. Mixing decreases with increasing energy splitting between LH1 and the HH continuum, as indicated by smaller ρ\rho at the larger well thickness w=10nmw=10\,\text{nm}. This remains true for different Sn compositions, as illustrated in Fig. 4(b) where aa is plotted as a function of ww and xx for εBR=0\varepsilon_{\text{BR}}=0.

Refer to caption
Figure 4: (a) HH contribution ρ\rho as a function of the wavevector kxk_{x} for a quantum well thickness w=6nmw=6\,\text{nm} and w=10nmw=10\,\text{nm}. Dashed lines correspond to the quadratic approximation ρak2\rho\approx ak_{\parallel}^{2}. (b) Coefficient aa as a function of ww and xx. The data displayed here are for thicknesses separated by a 2 nm step.

IV Discussion

The preceding results demonstrate that the lattice mismatch between Ge1-xSnx alloys and Ge provides an additional degree of freedom to engineer the tensile strain required to confine LHs in Ge. According to Fig. 2(a), the region of interest, as defined by the parameters (x,εBR)(x,\varepsilon_{\text{BR}}), lies in the range where xx is above 0.120.12 and the compressive strain in the barriers εBR\varepsilon_{\text{BR}} is below 0.4%-0.4\% (i.e., |εBR|0.4%|\varepsilon_{\text{BR}}|\lesssim 0.4\%). In this range, all band offsets ΔEi\Delta E_{i} are positive and the Ge layer is of direct bandgap. Ge1-xSnx layers at Sn content in the proposed range have already been demonstrated experimentally [11]. However, the addition of a highly tensile strained Ge layer on top of strain-relaxed Ge1-xSnx is still under development. For instance, the authors in Ref. [8] reported a 1.65%1.65\% tensile-strained Ge quantum well on partially relaxed Ge0.854Sn0.146 barriers with a residual strain to εBR0.54%\varepsilon_{\text{BR}}\approx-0.54\%. This system would be located near the crossing between the 0meV0\,\text{meV} ΔE1\Delta E_{1} line and the 50meV50\,\text{meV} LBO line in Fig. 2(a), very close to the optimal region of interest mentioned earlier. Strain relaxation in the barriers is necessary to enhance confinement in Ge, while relaxing the criterion of minimal well thickness w0w_{0} required for a LH-like valence band edge (ΔE2>0\Delta E_{2}>0). The ideal amount of strain relaxation for a given barrier Sn content can be estimated from Fig. 2(b). For example, a barrier with x=0.14x=0.14 does not allow a LH-like valence band edge for |εBR|>0.542%|\varepsilon_{\text{BR}}|>0.542\% compressive strain. The additional requirement w0<hcw_{0}<h_{c}, where hch_{c} is the critical thickness of Ge, further reduces the range of |εBR||\varepsilon_{\text{BR}}| to around <0.5%<0.5\% compressive strain. Reducing the amount of Sn in the barriers relaxes the limit imposed by the critical thickness hch_{c}, but at the cost of a smaller LBO. In contrast, increasing xx to 0.180.18 for instance slightly increases the range for |εBR||\varepsilon_{\text{BR}}| to around <0.525%<0.525\%, and increases significantly the LBO and ΔE1\Delta E_{1} [Fig. 2(a)] but at the expense of a smaller hch_{c} and a narrower window w0<w<hcw_{0}<w<h_{c}, but still in the range typically achievable in epitaxial growth experiments.

The in-plane effective mass γ\gamma [Fig. 3(a)] shows a strong dependence on both xx and ww, with small γ\gamma expected from the general rule that LHs are heavier in the plane than HHs. There is also an interesting feature where the dispersion changes from a hole-like (γ<0\gamma<0) to an electron-like (γ>0\gamma>0) curvature at 𝐤=0\mathbf{k}_{\parallel}=0. In the hole-like regime, the valence band edge is formed by a single valley located at 𝐤=0\mathbf{k}_{\parallel}=0. In contrast, in the electron-like regime, the valence band edge consists of four valleys, each located a distance k0k_{0}^{*} from 𝐤=0\mathbf{k}_{\parallel}=0 along the four equivalent 110\left<110\right> crystallographic directions in the QW plane. For instance, for a 5nm5\,\text{nm} well at x=0.13x=0.13 and εBR=0\varepsilon_{\text{BR}}=0, the valley minima are located at k00.0813nm1k_{0}^{*}\approx 0.0813\,\text{nm}^{-1} away from 𝐤=0\mathbf{k}_{\parallel}=0. For larger wavevectors, the dispersion goes away from the bandgap as required, owing to k4k^{4}-terms or higher that are not taken in account by the effective Hamiltonians (7) and (9). According to Fig. 3(a), electron-like subbands occur for small ww and large xx. This effect takes place for two reasons. The first is when the LH subband anticrosses a neighboring HH subband such that the curvature is inverted at 𝐤=0\mathbf{k}_{\parallel}=0. This is typical in systems where the ground state is HH-like and the first LH subband is allowed to mix strongly with the first excited HH subband [20], or when the LH subband is close to a HH continuum (e.g., when ww is small). When a LH is far from neighboring HH levels (e.g., when tensile strain is large), mixing decreases, as illustrated in Fig. 4(b), and becomes too weak to invert the curvature. The second reason for a curvature change is related to the anticrossing between the LH and the SO bands in the bulk dispersion of Ge for kz>0k_{z}>0 [Fig. 1(a)]. This anticrossing results in a curvature sign change of the LH band at some point kzk_{z}^{*} such that the bulk energy dispersion E(kx,ky;kz)E(k_{x},k_{y};k_{z}) has a hole-like (electron-like) curvature when kzk_{z} is fixed to a value smaller (larger) than kzk_{z}^{*} and kx,yk_{x,y} are close to zero. For a quantum well along the zz direction, the reciprocal-space envelope functions ψ~(kz)\tilde{\psi}(k_{z}) become wider for thinner wells and thus get a larger contribution from the electron-like regions in kk-space, resulting in a dispersion with inverted curvature at 𝐤=0\mathbf{k}_{\parallel}=0.

A major difference with η\eta subbands is the large gg_{\parallel} compared to HH systems [21, 22, 23, 6, 7]. A comparison between gg_{\parallel} and gg_{\perp} reveals an anisotropy for well thicknesses away from 10nm{\sim}10\,\text{nm}. Both components have a stronger dependence on ww than on xx but have opposite behavior with ww due to how they couple with neighboring subbands. For the out-of-plane component, g2κg_{\perp}\sim 2{\color[rgb]{0,0,0}\kappa} for large ww because the coupling with neighboring levels becomes weaker as LH1 gets further from the continuum. In contrast, the in-plane gg-factor does not depend on couplings with neighboring HH levels [c.f. (10)] and is instead more influenced by the spatial distribution of the envelopes across the layers. Thus, for large (small) ww, g4κg_{\parallel}\sim 4{\color[rgb]{0,0,0}\kappa} with κ\kappa being that of Ge (Ge1-xSnx).

Another peculiar feature associated with η\eta subbands is the absence of direct connection between γ\gamma and gg_{\perp} in contrast with HHs (i.e., see Eq. (5) in Ref. [24]). Perturbation theory gives the following for γ\gamma and gg_{\perp}:

γ\displaystyle\gamma =Γη+C+D,\displaystyle={\color[rgb]{0,0,0}\Gamma^{\eta}}+{\color[rgb]{0,0,0}C}+{\color[rgb]{0,0,0}D}, (19a)
g2\displaystyle\frac{g_{\perp}}{2} =Gη2C+D,\displaystyle=\frac{{\color[rgb]{0,0,0}G^{\eta}}}{2}-{\color[rgb]{0,0,0}C}+{\color[rgb]{0,0,0}D}, (19b)

where Γη\Gamma^{\eta}, GηG^{\eta} are described in Appendix B, whereas CC and DD are described in Appendix C. For HHs, a similar expansion would give

γH\displaystyle\gamma^{\text{H}} =ΓH+C,\displaystyle={\color[rgb]{0,0,0}\Gamma^{\text{H}}}+{\color[rgb]{0,0,0}C^{\prime}}, (20a)
gH2\displaystyle\frac{g_{\perp}^{\text{H}}}{2} =GH2+C.\displaystyle=\frac{{\color[rgb]{0,0,0}G^{\text{H}}}}{2}+{\color[rgb]{0,0,0}C^{\prime}}. (20b)

In the latter case, one can combine the equations for γH\gamma^{\text{H}} and gHg_{\perp}^{\text{H}} to eliminate the CC^{\prime} term, resulting in an expression involving only the mass and the gg-factor [24, 25]:

gH2=GH2ΓH+γH.\frac{g_{\perp}^{\text{H}}}{2}=\frac{{\color[rgb]{0,0,0}G^{\text{H}}}}{2}-{\color[rgb]{0,0,0}\Gamma^{\text{H}}}+\gamma^{\text{H}}. (21)

However, the result is different for η\eta subbands due to the additional DD term in (19). The latter is also related to the nonzero β1\beta_{1} coefficient of η\eta subbands [9].

Rashba parameters follow the general behavior αi0\alpha_{i}\to 0 as ww increases. This is caused by a reduced sensitivity of the wavefunction to electric fields when the level does not spread as extensively into the barriers. Although the QW is characterized by relatively small LBOs (100meV\lesssim 100\,\text{meV}) and small out-of-plane LH effective masses, the device operation can comfortably sustain realistic DC electric fields along the growth direction regardless of the well thickness and LBO without inducing any envelope leak into the barriers. In devices where space inversion symmetry needs to be broken, such as in electric dipole spin resonance (EDSR) experiments, this should not be an issue as the relevant Rashba parameter α3\alpha_{3} for EDSR, (which is proportional to γ2+γ3{\color[rgb]{0,0,0}\gamma_{2}}+{\color[rgb]{0,0,0}\gamma_{3}}) is one order of magnitude larger than the α2\alpha_{2} Rashba parameter (proportional to γ2γ3{\color[rgb]{0,0,0}\gamma_{2}}-{\color[rgb]{0,0,0}\gamma_{3}}), thus requiring smaller out-of-plane fields [9].

V Conclusions

This work demonstrates how Ge1-xSnx/Ge/Ge1-xSnx heterostructures can be tailored to achieve a selective confinement of LHs in Ge while pushing HHs in to the Ge1-xSnx barriers. For a sufficiently large Sn content (x>0.12x>0.12), small residual compressive strain in the barriers (|εBR|<0.4%|\varepsilon_{\text{BR}}|<0.4\%) and a well thickness w>w0w>w_{0}, the LH ground state emerges from within the HH continuum, thus yielding a pure LH-like valence band edge (ΔE2>0\Delta E_{2}>0). This regime also corresponds to a direct bandgap in both the well and its barriers, owing to the large tensile strain in the Ge well and the high Sn content in the barriers. Satisfying the condition of ΔE2>0\Delta E_{2}>0 imposes a threshold for residual strain in the barriers beyond which a LH-like VB edge becomes virtually impossible.

The in-plane effective mass, the out-of-plane and in-plane gg-factor, and the Rashba parameters α1,2,3\alpha_{1,2,3} were computed by explicitly taking in account the spread of the LH envelopes into the barriers and the coupling with the neighboring HH continuum. Small inverse effective masses γ\gamma are obtained. A peculiar sign change in γ\gamma appearing for small well thicknesses (w7nmw\lesssim 7\,\text{nm}) is observed and attributed to the proximity of the LH to the HH continuum (larger LH-HH mixing) and the contribution of the SO band in the LH spinor. An increasingly strong anisotropy in the gg-factor components is also observed for well thicknesses away from 10nm{\sim}10\,\text{nm}. Most notably, the in-plane component of the gg-tensor is significantly larger than what is expected in HH systems. A nonzero linear Rashba parameter α1\alpha_{1} was obtained, as anticipated for LH systems, with an α3\alpha_{3} coefficient one order of magnitude larger than α2\alpha_{2}.

Acknowledgments. O.M. acknowledges support from NSERC Canada (Discovery, SPG, and CRD Grants), Canada Research Chairs, Canada Foundation for Innovation, Mitacs, PRIMA Québec, Defence Canada (Innovation for Defence Excellence and Security, IDEaS), the European Union’s Horizon Europe research and innovation program under Grant Agreement No 101070700 (MIRAQLS), the US Army Research Office Grant No. W911NF-22-1-0277, and the Air Force Office of Scientific and Research Grant No. FA9550-23-1-0763.

Appendix A Parameterization of Ge1-xSnx

The material parameters for the Ge1-xSnx alloy were calculated in the full composition range by interpolating the parameters from pure Ge and pure Sn:

A(x)=(1x)AGe+xASnx(1x)b.A(x)=(1-x)A^{\text{Ge}}+xA^{\text{Sn}}-x(1-x)b. (22)
Table 1: Input parameters with bowings for the 8-band kpk\cdot p model.
Parameter Germanium Tin Bowing
Lattice constant
a0a_{0} (Å, 300K300\,\text{K}) 5.6579565.657956a 6.4894176.489417b  0.083-0.083c*
Energy gaps
EgΓ0E_{g\Gamma}^{0} (eV) 0.89810.8981b 0.413-0.413b 2.462.46d*
EgL0E_{g\text{L}}^{0} (eV) 0.7400.740p 0.1000.100c 1.231.23k*
αΓ\alpha_{\Gamma} (104eV/K10^{-4}\,\text{eV}/\,\text{K}) 6.8426.842e 7.94-7.94d
αL\alpha_{\text{L}} (104eV/K10^{-4}\,\text{eV}/\,\text{K}) 4.5614.561e
βΓ\beta_{\Gamma} (K) 398398e 1111d
βL\beta_{\text{L}} (K) 210210e
Δ\Delta (eV) 0.2900.290b 0.7700.770f 0.100-0.100c
Ev,avgE_{v,\text{avg}} (eV) 0 0.690.69g
Elastic constants
c11c_{11} (GPa) 124124b 69.069.0b
c12c_{12} (GPa) 41.341.3b 29.329.3b
c44c_{44} (GPa) 68.368.3b 36.236.2b
Deformation potentials{\dagger}
acΓa_{c\Gamma} (eV) 8.24-8.24h 6.00-6.00l
acLa_{c\text{L}} (eV) 1.54-1.54h 2.14-2.14m
ava_{v} (eV) 1.241.24h 1.581.58m
bb (eV) 2.86-2.86i 2.7-2.7n
Effective mass and spin parameters
mcΓLm_{c\Gamma}^{\text{{\color[rgb]{0,0,0}L}}} (m0m_{0}) 0.03860.0386 0.057-0.057
γ1L\gamma_{1}^{\text{{\color[rgb]{0,0,0}L}}} 13.3813.38j
γ2L\gamma_{2}^{\text{{\color[rgb]{0,0,0}L}}} 4.244.24j See Eq. (25)
γ3L\gamma_{3}^{\text{{\color[rgb]{0,0,0}L}}} 5.695.69j
κL\kappa^{\text{{\color[rgb]{0,0,0}L}}} 3.413.41f 11.84-11.84f
gLg^{\text{{\color[rgb]{0,0,0}L}}} 2.77-2.77 86.686.6
  • {\dagger}

    The convention a=acava=a_{c}-a_{v} is used.

  • *

    See Eq. (23)

  • a

    Reference [26]

  • b

    Reference [27]

  • c

    Reference [28]

  • d

    Reference [29]

  • e

    Reference [30]

  • f

    Reference [31]

  • g

    Reference [32]

  • h

    Reference [15]

  • i

    Reference [33]

  • j

    Reference [34]

  • k

    Reference [35]

  • l

    Reference [36]

  • m

    Reference [37]

  • n

    Reference [38]

  • p

    Reference [39]

Here, 0x10\leq x\leq 1 is the alloy fraction and bb is a bowing constant for the parameter AA if necessary. For temperature-dependent quantities such as the lattice constant or the bandgaps, we apply (22) on the temperature-dependent parameters of the alloy’s constituents. For instance, the bandgap of Ge1-xSnx at the Γ\Gamma point, EgΓ(x,T)E_{g\Gamma}(x,T), is given by

EgΓ(x,T)=(1x)EgΓGe(T)+xEgΓSn(T)x(1x)bΓ,E_{g\Gamma}(x,T)=(1-x)E_{g\Gamma}^{\text{Ge}}(T)+xE_{g\Gamma}^{\text{Sn}}(T)-x(1-x)b_{\Gamma}, (23)

where

EgΓGe(T)=EgΓ0,GeαΓGeT2βΓGe+TE_{g\Gamma}^{\text{Ge}}(T)=E_{g\Gamma}^{0,\text{Ge}}-\frac{\alpha_{\Gamma}^{\text{Ge}}T^{2}}{\beta_{\Gamma}^{\text{Ge}}+T} (24)

and similar equations for Sn and the L valley. The average VB energy Ev,avgE_{v,\text{avg}} is given by Ev,avg=EvΔ/3E_{v,\text{avg}}=E_{v}-\Delta/3, where EvE_{v} is the VB edge energy and Δ\Delta is the spin-orbit splitting [15].

The electron effective mass mcΓLm_{c\Gamma}^{\text{{\color[rgb]{0,0,0}L}}} and the electron gLg^{\text{{\color[rgb]{0,0,0}L}}} factor were calculated following the approach in Ref. [31] to make them consistent with the bandgaps and spin-orbit couplings listed in Table 1. The Luttinger parameters γ1,2,3L\gamma_{1,2,3}^{\text{{\color[rgb]{0,0,0}L}}} were interpolated between pure Ge and Ge0.80Sn0.20 using the data from Ref. [40], giving

γiL=γiL,Ge(1x0.2)+γiL,GeSn(x0.2)bix0.2(1x0.2),\gamma_{i}^{\text{{\color[rgb]{0,0,0}L}}}=\gamma_{i}^{\text{{\color[rgb]{0,0,0}L},Ge}}\left(1-\frac{x}{0.2}\right)+\gamma_{i}^{\text{{\color[rgb]{0,0,0}L},GeSn}}\left(\frac{x}{0.2}\right)-b_{i}\frac{x}{0.2}\left(1-\frac{x}{0.2}\right), (25)

with γiL,Ge\gamma_{i}^{\text{{\color[rgb]{0,0,0}L},Ge}} listed in Table 1, γ1L,GeSn=29.2108\gamma_{1}^{\text{{\color[rgb]{0,0,0}L},GeSn}}=29.2108, γ2L,GeSn=12.2413\gamma_{2}^{\text{{\color[rgb]{0,0,0}L},GeSn}}=12.2413, γ3L,GeSn=13.7387\gamma_{3}^{\text{{\color[rgb]{0,0,0}L},GeSn}}=13.7387, and b1=20.3391b_{1}=20.3391, b2=9.6609b_{2}=9.6609, b3=9.8187b_{3}=9.8187.

The Kane momentum matrix element PP, which couples the two conduction bands to the six valence bands, is known to sometimes cause spurious solutions, often appearing as levels within the bandgap or with violently oscillating envelopes [41, 42, 43]. The approach employed here to eliminate spurious solutions is described in Ref. [41] and consists of rescaling PP:

P2=322mcΓL(2EgΓ+1EgΓ+Δ)1.P^{2}=\frac{3\hbar^{2}}{2m_{c\Gamma}^{\text{{\color[rgb]{0,0,0}L}}}}\left(\frac{2}{E_{g\Gamma}}+\frac{1}{E_{g\Gamma}+\Delta}\right)^{-1}. (26)

The remote band contributions in the Luttinger parameters are then re-ajusted [34] according to Pα0EpP\equiv\sqrt{\alpha_{0}E_{p}} given in (26):

γ1=γ1LEp3Eg\displaystyle{\color[rgb]{0,0,0}\gamma_{1}}=\gamma_{1}^{\text{{\color[rgb]{0,0,0}L}}}-\frac{E_{p}}{3E_{g}} (27)
γ2,3=γ2,3LEp6Eg\displaystyle{\color[rgb]{0,0,0}\gamma_{2,3}}=\gamma_{2,3}^{\text{{\color[rgb]{0,0,0}L}}}-\frac{E_{p}}{6E_{g}} (28)
κ=κLEp6Eg\displaystyle{\color[rgb]{0,0,0}\kappa}=\kappa^{\text{{\color[rgb]{0,0,0}L}}}-\frac{E_{p}}{6E_{g}} (29)
g=gL+2Ep3EgΔEg+Δ\displaystyle{\color[rgb]{0,0,0}g}=g^{\text{{\color[rgb]{0,0,0}L}}}+\frac{2E_{p}}{3E_{g}}\frac{\Delta}{E_{g}+\Delta} (30)
m0mcΓ=m0mcΓL2Ep3Eg3Eg/2+ΔEg+Δ=0.\displaystyle\frac{m_{0}}{{\color[rgb]{0,0,0}m_{c\Gamma}}}=\frac{m_{0}}{m_{c\Gamma}^{\text{{\color[rgb]{0,0,0}L}}}}-\frac{2E_{p}}{3E_{g}}\frac{3E_{g}/2+\Delta}{E_{g}+\Delta}=0. (31)

Appendix B kpk\cdot p framework

Our implementation of kpk\cdot p theory is based on the model presented in Ref. [44], which is an extension of standard kpk\cdot p frameworks [34, 45] for heterostructures with finite energy band offsets at the interfaces (i.e., when a proper ordering between material parameter operators and kzk_{z} is critical).

The 8-band kpk\cdot p Hamiltonian can be written as a sum of different contributions:

H=Hk+HSO+Hε+V.H=H_{k}+H_{\text{SO}}+H_{\varepsilon}+V. (32)

The first term, HkH_{k}, depends on the mechanical wavevector K=k+eA/\textbf{K}=\textbf{k}+e\textbf{A}/\hbar. It automatically includes the Zeeman Hamiltonian through the relation K×K=eB/(i)\textbf{K}\times\textbf{K}=e\textbf{B}/(i\hbar). In the Cartesian basis

X={|S+,|S,|X+,|Y+,|Z+,|X,|Y,|Z},\begin{split}\mathcal{B}_{X}=&\left\{\ket{S+},\ket{S-},\right.\\ &\left.\ket{X+},\ket{Y+},\ket{Z+},\ket{X-},\ket{Y-},\ket{Z-}\right\},\end{split} (33)

HkH_{k} is given by

Hk=[Hcck12×2Hcvk12×2Hcvk12×2Hvvk]+[HB00HB13×3],\begin{split}H_{k}&=\begin{bmatrix}H_{cc}^{k}&1_{2\times 2}\otimes H_{cv}^{k}\\ 1_{2\times 2}\otimes H_{cv}^{k\dagger}&1_{2\times 2}\otimes H_{vv}^{k}\end{bmatrix}\\ &+\begin{bmatrix}H_{B}&0\\ 0&H_{B}\otimes 1_{3\times 3}\end{bmatrix},\end{split} (34)

with

Hcck=Eg+αKαAKα+iα02αβγϵαβγKα(gg0)Kβσγ,\displaystyle\begin{split}H_{cc}^{k}&=E_{g}+\sum_{\alpha}K_{\alpha}AK_{\alpha}\\ &+\frac{i\alpha_{0}}{2}\sum_{\alpha\beta\gamma}\epsilon_{\alpha\beta\gamma}K_{\alpha}(g-g_{0})K_{\beta}\sigma_{\gamma},\end{split} (35)
Hcvk=[iPKxiPKyiPKz],\displaystyle H_{cv}^{k}=\begin{bmatrix}iPK_{x}&iPK_{y}&iPK_{z}\end{bmatrix}, (36)
(Hvvk)i,j={αKαMKα+Ki(LM)Kii=jKiN+Kj+KjNKiij,\displaystyle\left(H_{vv}^{k}\right)_{i,j}=\begin{cases}\sum_{\alpha}K_{\alpha}MK_{\alpha}+K_{i}(L-M)K_{i}&i=j\\ K_{i}N_{+}K_{j}+K_{j}N_{-}K_{i}&i\neq j\end{cases}, (37)
HB=iα0g02αβγϵαβγKαKβσγ,\displaystyle H_{B}=\frac{i\alpha_{0}g_{0}}{2}\sum_{\alpha\beta\gamma}\epsilon_{\alpha\beta\gamma}K_{\alpha}K_{\beta}\sigma_{\gamma}, (38)

where ϵαβγ\epsilon_{\alpha\beta\gamma} is the Levi-Civita tensor, A=α0m0/mcΓ=0A=\alpha_{0}m_{0}/m_{c\Gamma}=0 [see (31)] and g02g_{0}\approx 2 is the free electron gg-factor. The term HBH_{B} arises from the interaction of the free electron spin with the magnetic field [44].

The second term in (32), HSOH_{\text{SO}}, includes the effects of the SO band and is proportional to the spin-orbit gap Δ\Delta:

HSO=Δ3[1000000001000000000i000100i0000i000001i0000010i00000ii00001i0000].H_{\text{SO}}=\frac{\Delta}{3}\begin{bmatrix}1&0&0&0&0&0&0&0\\ 0&1&0&0&0&0&0&0\\ 0&0&0&-i&0&0&0&1\\ 0&0&i&0&0&0&0&-i\\ 0&0&0&0&0&-1&i&0\\ 0&0&0&0&-1&0&i&0\\ 0&0&0&0&-i&-i&0&0\\ 0&0&1&i&0&0&0&0\end{bmatrix}. (39)

Strain is incorporated by means of the Bir-Pikus formalism [46], resulting in the term

Hε=[Hccε0012×2Hvvε],H_{\varepsilon}=\begin{bmatrix}H_{cc}^{\varepsilon}&0\\ 0&1_{2\times 2}\otimes H_{vv}^{\varepsilon}\end{bmatrix}, (40)

with

Hccε=12×2acTr{ε},\displaystyle H_{cc}^{\varepsilon}=1_{2\times 2}a_{c}\text{Tr}\{\varepsilon\}, (41)
(Hvvε)i,j={mTr{ε}+(lm)εiii=jnεijij.\displaystyle\left(H_{vv}^{\varepsilon}\right)_{i,j}=\begin{cases}m\text{Tr}\{\varepsilon\}+(l-m)\varepsilon_{ii}&i=j\\ n\varepsilon_{ij}&i\neq j\end{cases}. (42)

Finally, the potential energy V=Ev,avg+eEzzV=E_{v,\text{avg}}+eE_{z}z. The parameters LL, MM, N±N_{\pm}, ll, mm, and nn are related to the usual Luttinger parameters and deformation potentials by

[LMN++α0Nα0]=α0[1400120000330033][γ1γ2γ3κ],\displaystyle\begin{bmatrix}L\\ M\\ N_{+}+\alpha_{0}\\ N_{-}-\alpha_{0}\end{bmatrix}=-\alpha_{0}\begin{bmatrix}1&4&0&0\\ 1&-2&0&0\\ 0&0&3&3\\ 0&0&3&-3\end{bmatrix}\begin{bmatrix}\gamma_{1}\\ \gamma_{2}\\ \gamma_{3}\\ \kappa\end{bmatrix}, (43)
[lmn]=[120110003][avbd].\displaystyle\begin{bmatrix}l\\ m\\ n\end{bmatrix}=\begin{bmatrix}1&2&0\\ 1&-1&0\\ 0&0&\sqrt{3}\end{bmatrix}\begin{bmatrix}a_{v}\\ b\\ d\end{bmatrix}. (44)

A change of basis from X\mathcal{B}_{X} to the so-called angular momentum basis J\mathcal{B}_{J} brings the Hamiltonian HH in a 2×22\times 2 block diagonal matrix (each block being 4×44\times 4) when evaluated with B=0B=0 and kx=ky=0k_{x}=k_{y}=0:

H0\displaystyle H_{0} UXJH(kx=ky=0,B=0)UXJ\displaystyle\equiv U_{X\leftarrow J}^{\dagger}H(k_{x}=k_{y}=0,B=0)U_{X\leftarrow J}
=[H+00H],\displaystyle=\begin{bmatrix}H_{+}&0\\ 0&H_{-}\end{bmatrix},

where

J={|12,12c,|32,12,|12,12,|32,32,|12,12c,|32,12,|12,12,|32,32},\displaystyle\begin{split}\mathcal{B}_{J}=&\left\{\Ket{\frac{1}{2},\frac{1}{2}}_{c},\Ket{\frac{3}{2},\frac{1}{2}},\Ket{\frac{1}{2},\frac{1}{2}},\Ket{\frac{3}{2},\frac{3}{2}},\right.\\ &\left.\Ket{\frac{1}{2},-\frac{1}{2}}_{c},\Ket{\frac{3}{2},-\frac{1}{2}},\Ket{\frac{1}{2},-\frac{1}{2}},\Ket{\frac{3}{2},-\frac{3}{2}}\right\},\end{split} (45)
UXJ=[1000000000001000000s20s6s30000is20is6is300s23s3000000s6s30000s20is6is30000is200000s23s30],\displaystyle U_{X\leftarrow J}=\begin{bmatrix}1&0&0&0&0&0&0&0\\ 0&0&0&0&1&0&0&0\\ 0&0&0&-s_{2}&0&s_{6}&-s_{3}&0\\ 0&0&0&-is_{2}&0&-is_{6}&is_{3}&0\\ 0&s_{23}&-s_{3}&0&0&0&0&0\\ 0&-s_{6}&-s_{3}&0&0&0&0&s_{2}\\ 0&-is_{6}&-is_{3}&0&0&0&0&-is_{2}\\ 0&0&0&0&0&s_{23}&s_{3}&0\end{bmatrix}, (46)
Hσ=V+Δ3+[EgΓ000000Δ00]+[kzAkzi2/3PkziσPkz/30α0kzγ+kz8σα0kzγ2kz0α0kzγ1kz0α0kzγkz]+Tr{ε}[ac000av00av0av]+bδε[000012σ0001],\displaystyle\begin{split}H_{\sigma}&=V+\frac{\Delta}{3}+\begin{bmatrix}E_{g\Gamma}&0&0&0\\ &0&0&0\\ \dagger&&-\Delta&0\\ &&&0\end{bmatrix}\\ &+\begin{bmatrix}k_{z}Ak_{z}&i\sqrt{2/3}Pk_{z}&-i\sigma Pk_{z}/\sqrt{3}&0\\ &-\alpha_{0}k_{z}\gamma_{+}k_{z}&\sqrt{8}\sigma\alpha_{0}k_{z}\gamma_{2}k_{z}&0\\ \dagger&&-\alpha_{0}k_{z}\gamma_{1}k_{z}&0\\ &&&-\alpha_{0}k_{z}\gamma_{-}k_{z}\end{bmatrix}\\ &+\text{Tr}\{\varepsilon\}\begin{bmatrix}a_{c}&0&0&0\\ &a_{v}&0&0\\ \dagger&&a_{v}&0\\ &&&a_{v}\end{bmatrix}+b\delta\varepsilon\begin{bmatrix}0&0&0&0\\ &-1&\sqrt{2}\sigma&0\\ \dagger&&0&0\\ &&&1\end{bmatrix},\end{split} (47)

with s2=1/2s_{2}=1/\sqrt{2}, s3=1/3s_{3}=1/\sqrt{3}, s6=1/6s_{6}=1/\sqrt{6}, s23=2/3s_{23}=\sqrt{2/3}, σ=±1\sigma=\pm 1 the pseudo-spin index, γ±=γ1±2γ2\gamma_{\pm}=\gamma_{1}\pm 2\gamma_{2}, and δε=εxxεzz\delta\varepsilon=\varepsilon_{xx}-\varepsilon_{zz} (all shear components in the strain tensor vanish). Since H+H_{+} and HH_{-} differ only by a minus sign in the LH-SO and the CB-SO coupling elements, both share the same energy spectrum (Kramers’ degeneracy) and the eigenstates of HH_{-} are the time-reversed eigenstates of H+H_{+}. Additionally, HσH_{\sigma} is itself block diagonal: one 3×33\times 3 block representing a CB-LH-SO superposition (or a η\eta level) and one 1×11\times 1 block representing a pure HH level (or H for short). Eigenstates of HσH_{\sigma} are thus either of type η\eta or H (with pseudo-spin σ\sigma):

|η,σ=|12,σ2c|c+|32,σ2|+σ|12,σ2|s,\displaystyle\Ket{\eta,\sigma}=\Ket{\frac{1}{2},\frac{\sigma}{2}}_{c}\Ket{c}+\Ket{\frac{3}{2},\frac{\sigma}{2}}\Ket{\ell}+\sigma\Ket{\frac{1}{2},\frac{\sigma}{2}}\Ket{s}, (48a)
|H,σ=|32,3σ2|h.\displaystyle\Ket{\text{H},\sigma}=\Ket{\frac{3}{2},\frac{3\sigma}{2}}\Ket{h}. (48b)

The energies and eigenstates of H+H_{+} are computed for each set of quantum well parameters (xx, εBR\varepsilon_{\text{BR}} and ww) with the substitution kzizk_{z}\to-i\partial_{z}, without any assumptions on the shape of the envelopes. We choose a grid spacing of 0.01nm0.01\,\text{nm} for the finite differences and keep the 200200 subbands that are the closest to LH1 (N=NηN=N_{\eta} + NH=200N_{\text{H}}=200). We found that N=200N=200 is large enough for the effective parameters in (7) and (9) to converge.

We diagonalize HH away from kx=ky=0k_{x}=k_{y}=0 and B=0B=0 by first projecting HH onto the orthonormal basis (48). This brings HH to a 4×44\times 4 block-matrix form (with total dimension 2N×2N2N\times 2N), where each block consists of all the subbands of one aforementioned type (H or η\eta, spin up/down) and couplings thereof. Taking the basis ordering 0={|H+,|η+,|η,|H}\mathcal{B}_{0}=\{\ket{\text{H}+},\ket{\eta+},\ket{\eta-},\ket{\text{H}-}\} (and bold characters to emphasize that we are in basis 0\mathcal{B}_{0}), the Hamiltonian HH when B is perpendicular to the plane is given by

H=𝐄0+α0[𝐌γK2+12λ2𝐌g+(i𝐌1K+𝐌2K2+h.c.)],\begin{split}\textbf{H}=\mathbf{E}_{0}&+\alpha_{0}\left[\mathbf{M}_{\gamma}K_{\parallel}^{2}+\frac{1}{2\lambda^{2}}\mathbf{M}_{g}\right.\\ &+\left.\left(i\mathbf{M}_{1}K_{-}+\mathbf{M}_{2}K_{-}^{2}+\text{h.c.}\right)\right],\end{split} (49)

where E0=diag{EH,Eη,Eη,EH}\textbf{E}_{0}=\text{diag}\{\textbf{E}^{\text{H}},\textbf{E}^{\eta},\textbf{E}^{\eta},\textbf{E}^{\text{H}}\} are the energies associated to H0H_{0} (EH,η\textbf{E}^{\text{H},\eta} are also diagonal) and

𝐌γ\displaystyle\mathbf{M}_{\gamma} =diag{𝚪H,𝚪η,𝚪η,𝚪H},\displaystyle=\text{diag}\{\boldsymbol{\Gamma}^{\text{H}},\boldsymbol{\Gamma}^{\eta},\boldsymbol{\Gamma}^{\eta},\boldsymbol{\Gamma}^{\text{H}}\}, (50a)
𝐌g\displaystyle\mathbf{M}_{g} =diag{𝐆H,𝐆η,𝐆η,𝐆H},\displaystyle=\text{diag}\{\mathbf{G}^{\text{H}},\mathbf{G}^{\eta},-\mathbf{G}^{\eta},-\mathbf{G}^{\text{H}}\}, (50b)
𝐌1\displaystyle\mathbf{M}_{1} =[0𝐓x0000𝐓η0000𝐓x0000],\displaystyle=\begin{bmatrix}0&\mathbf{T}^{\text{x}}&0&0\\ 0&0&\mathbf{T}^{\eta}&0\\ 0&0&0&\mathbf{T}^{\text{x}\dagger}\\ 0&0&0&0\end{bmatrix}, (50c)
𝐌2\displaystyle\mathbf{M}_{2} =[00𝛍0000𝛍𝛅0000𝛅00],\displaystyle=\begin{bmatrix}0&0&\boldsymbol{\upmu}&0\\ 0&0&0&\boldsymbol{\upmu}^{\dagger}\\ \boldsymbol{\updelta}^{\dagger}&0&0&0\\ 0&\boldsymbol{\updelta}&0&0\end{bmatrix}, (50d)

with (assuming A=0A=0 and g0=2g_{0}=2):

Γl,lH=hl|γ1+γ2|hl,\displaystyle\Gamma_{l,l^{\prime}}^{\text{H}}=-\bra{h_{l}}\gamma_{1}+\gamma_{2}\ket{h_{l^{\prime}}}, (51a)
Gl,lH=hl|6κ|hl,\displaystyle G_{l,l^{\prime}}^{\text{H}}=-\bra{h_{l}}6\kappa\ket{h_{l^{\prime}}}, (51b)
Γj,jη=13+j|γ1+γ2|+j23j|γ12γ2|j,\displaystyle\Gamma_{j,j^{\prime}}^{\eta}=-\frac{1}{3}\bra{+_{j}}\gamma_{1}+\gamma_{2}\ket{+_{j^{\prime}}}-\frac{2}{3}\bra{-_{j}}\gamma_{1}-2\gamma_{2}\ket{-_{j^{\prime}}}, (51c)
Gj,jη=cj|g|cj2+j|κ|+j43(+j|+jj|j),\displaystyle\begin{split}G_{j,j^{\prime}}^{\eta}=&\bra{c_{j}}g\ket{c_{j^{\prime}}}-2\bra{+_{j}}\kappa\ket{+_{j^{\prime}}}\\ &-\frac{4}{3}\left(\braket{+_{j}}{+_{j^{\prime}}}-\braket{-_{j}}{-_{j^{\prime}}}\right),\end{split} (51d)
Tj,jη=16α0(cj|P|+j++j|P|cj)+icj|[g/2,kz]|cji(+j|u+|jj|u|+j),\displaystyle\begin{split}T_{j,j^{\prime}}^{\eta}=&\frac{1}{\sqrt{6}\alpha_{0}}\left(\bra{c_{j}}P\ket{+_{j^{\prime}}}+\bra{+_{j}}P\ket{c_{j^{\prime}}}\right)\\ &+i\bra{c_{j}}[g/2,k_{z}]\ket{c_{j^{\prime}}}\\ &-i\left(\bra{+_{j}}u_{+}\ket{-_{j^{\prime}}}-\bra{-_{j}}u_{-}\ket{+_{j^{\prime}}}\right),\end{split} (51e)
Tl,jx=hl|(P2α0|cj3iu+|j),\displaystyle T_{l,j}^{\text{x}}=\bra{h_{l}}\left(\frac{P}{\sqrt{2}\alpha_{0}}\ket{c_{j}}-\sqrt{3}iu_{+}\ket{-_{j}}\right), (51f)
μl,j=32hl|γ2+γ3|+j,\displaystyle\mu_{l,j}=\frac{\sqrt{3}}{2}\bra{h_{l}}\gamma_{2}+\gamma_{3}\ket{+_{j}}, (51g)
δl,j=32hl|γ2γ3|+j.\displaystyle\delta_{l,j}=\frac{\sqrt{3}}{2}\bra{h_{l}}\gamma_{2}-\gamma_{3}\ket{+_{j}}. (51h)

Here, Tl,jxT_{l,j}^{\text{x}} really is the same as in (15) but with the explicit dependence on the η\eta subband index jj. When B is in-plane, the Hamiltonian HH is given by

𝐇=𝐄0+α0{𝐌γk2+1λ4𝐌γ′′+[i(𝐌1+2λ2𝐌2eiϕ1λ2𝐌γeiϕ)k+𝐌2k21λ2𝐌1eiϕe2iϕλ4𝐌2′′+h.c.]},\begin{split}\mathbf{H}=\mathbf{E}_{0}&+\alpha_{0}\left\{\mathbf{M}_{\gamma}k_{\|}^{2}+\frac{1}{\lambda^{4}}\mathbf{M}_{\gamma}^{\prime\prime}\right.\\ &+\left.\left[i\left(\mathbf{M}_{1}+\frac{2}{\lambda^{2}}\mathbf{M}_{2}^{\prime}e^{-i\phi}-\frac{1}{\lambda^{2}}\mathbf{M}_{\gamma}^{\prime}e^{i\phi}\right)k_{-}\right.\right.\\ &+\left.\left.\mathbf{M}_{2}k_{-}^{2}-\frac{1}{\lambda^{2}}\mathbf{M}_{1}^{\prime}e^{-i\phi}-\frac{e^{-2i\phi}}{\lambda^{4}}\mathbf{M}_{2}^{\prime\prime}+\text{h.c.}\right]\right\},\end{split} (52)

where the Mi\textbf{M}_{i} with primes are defined similarly to those without primes [c.f. (50)] but with

Γl,lpH=hl|zp(γ1+γ2)|hl,\displaystyle\Gamma_{l,l^{\prime}}^{p\mathrm{H}}=-\bra{h_{l}}z^{p}(\gamma_{1}+\gamma_{2})\ket{h_{l^{\prime}}}, (53a)
Γj,jpη=13+j|zp(γ1+γ2)|+j23j|zp(γ12γ2)|j,\displaystyle\begin{split}\Gamma_{j,j^{\prime}}^{p\eta}=&-\frac{1}{3}\bra{+_{j}}z^{p}(\gamma_{1}+\gamma_{2})\ket{+_{j^{\prime}}}\\ &-\frac{2}{3}\bra{-_{j}}z^{p}(\gamma_{1}-2\gamma_{2})\ket{-_{j^{\prime}}},\end{split} (53b)
Tj,jη=16α0(cj|zP|+j++j|zP|cj)+icj|[zg/2,kz]|cji(+j|u+|jj|u|+j)+12[sj|j+j|sj],\displaystyle\begin{split}T_{j,j^{\prime}}^{\prime\eta}=&\frac{1}{\sqrt{6}\alpha_{0}}\left(\bra{c_{j}}zP\ket{+_{j^{\prime}}}+\bra{+_{j}}zP\ket{c_{j^{\prime}}}\right)\\ &+i\bra{c_{j}}\left[zg/2,k_{z}\right]\ket{c_{j^{\prime}}}\\ &-i\left(\bra{+_{j}}u_{+}^{\prime}\ket{-_{j^{\prime}}}-\bra{-_{j}}u_{-}^{\prime}\ket{+_{j^{\prime}}}\right)\\ &+\frac{1}{\sqrt{2}}\left[\braket{s_{j}}{-_{j^{\prime}}}+\braket{-_{j}}{s_{j^{\prime}}}\right],\end{split} (53c)
Tl,jx=hl|[zP2α0|cj3(iu+|j+12|sj)],\displaystyle T_{l,j}^{\prime\text{x}}=\bra{h_{l}}\left[\frac{zP}{\sqrt{2}\alpha_{0}}\ket{c_{j}}-\sqrt{3}\left(iu_{+}^{\prime}\ket{-_{j}}+\frac{1}{\sqrt{2}}\ket{s_{j}}\right)\right], (53d)
μl,jp=32hl|zp(γ2+γ3)|+j,\displaystyle\mu_{l,j}^{p}=\frac{\sqrt{3}}{2}\bra{h_{l}}z^{p}(\gamma_{2}+\gamma_{3})\ket{+_{j}}, (53e)
δl,jp=32hl|zp(γ2γ3)|+j,\displaystyle\delta_{l,j}^{p}=\frac{\sqrt{3}}{2}\bra{h_{l}}z^{p}(\gamma_{2}-\gamma_{3})\ket{+_{j}}, (53f)

where p=1p=1 corresponds to one prime and p=2p=2 corresponds to two primes. We point out that in basis 0\mathcal{B}_{0}, the strain components and the SO energy Δ\Delta do not appear explicitly in H, since they are already taken in account by the energies EH,η\textbf{E}^{\text{H},\eta} and the envelope functions associated with H0H_{0}. This is because HH does not contain any terms such as εiikj\varepsilon_{ii}k_{j} or Δkj\Delta\cdot k_{j}.

Appendix C Perturbative expansion of γ\gamma and gg_{\perp}

A perturbative expansion of the kpk\cdot p Hamiltonian for small 𝐊{\color[rgb]{0,0,0}\mathbf{K}}_{\parallel} yields explicit formulas for the effective parameters appearing in (7) and (9). This is obtained by means of a Schrieffer-Wolff transformation (SWT) [34]. To this end, the basis 0\mathcal{B}_{0} is convenient since H is exactly diagonal when K\textbf{K}_{\parallel} and BB are zero (see Appendix B for the notation). The zero-th order terms in the SWT are directly the energies E0\textbf{E}_{0}. The first-order terms are given by the diagonal elements in the nonzero blocks of the matrices Mi\textbf{M}_{i}. For instance, the first-order contribution to the effective mass of η\eta subbands are the diagonal entries of 𝚪η\boldsymbol{\Gamma}^{\eta}. Similarly, the Rashba parameter β1=α0Tj,jη\beta_{1}=\alpha_{0}T_{j,j}^{\eta}. The in-plane gg factor stems from the M1\textbf{M}_{1}^{\prime} term in (52), which has a nonvanishing block in the (η+,η)(\eta+,\eta-) subspace, thus yielding g=2Tj,jηg_{\parallel}=-2T_{j,j}^{\prime\eta} [c.f. (10)]. Besides β1\beta_{1} and gg_{\parallel}, the effective parameters in (7) and (9) require a second- or third-order SWT to be described exactly. In particular, γ\gamma and gg_{\perp} are exactly described by second-order perturbation only. For the jj-th η\eta subband, the terms CC and DD in (19) correspond to the second-order corrections, and are given by

C=α0l|Tl,jx|2EjηElH,\displaystyle{\color[rgb]{0,0,0}C}=\alpha_{0}\sum_{l}{\frac{\left|{\color[rgb]{0,0,0}T_{l,j}^{\text{x}}}\right|^{2}}{E_{j}^{\eta}-E_{l}^{\text{H}}}}, (54)
D=α0jj|Tj,jη|2EjηEjη.\displaystyle{\color[rgb]{0,0,0}D}=\alpha_{0}\sum_{j^{\prime}\neq j}{\frac{|{\color[rgb]{0,0,0}T_{j,j^{\prime}}^{\eta}|^{2}}}{E_{j}^{\eta}-E_{j^{\prime}}^{\eta}}}. (55)

For the ll-th HH subband, the CC^{\prime} term in (20) is

C=α0j|Tl,jx|2ElHEjη.{\color[rgb]{0,0,0}C^{\prime}}=\alpha_{0}\sum_{j}{\frac{\left|{\color[rgb]{0,0,0}T_{l,j}^{\text{x}}}\right|^{2}}{E_{l}^{\text{H}}-E_{j}^{\eta}}}. (56)

References