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

thanks: [email protected]

Excitons in Atomically Thin TMD in Electric and Magnetic Fields

Jack N. Engdahl School of Physics, University of New South Wales, Sydney 2052, Australia    Harley D. Scammell School of Mathematical and Physical Sciences, University of Technology Sydney, Ultimo, NSW 2007, Australia    Dmitry K. Efimkin School of Physics and Astronomy, Monash University, Victoria 3800, Australia    Oleg P. Sushkov School of Physics, University of New South Wales, Sydney 2052, Australia
Abstract

The magnetic field dependence of photoabsorption provides direct insights into the band structure of semiconductors. It is perhaps surprising that there is a large discrepancy between electron, hole, and reduced mass reported in the recent literature. Motivated by this puzzle we reconsider excitonic magneto-absorption and find that the commonly employed perturbative approach, namely for computing the diamagnetic shift, is inadequate to account for the parameter ranges considered in existing data. In particular, we develop the theory for strong magnetic field and, upon analysis of the data, arrive at the set of exciton parameters different to what has been estimated perturbatively in the literature. Only s-wave excitons are visible in photoluminescence as the spectral weight of p-wave states is too small, this limits the amount of information that can be extracted about the underlying band structure. To overcome this, we propose to study p-wave states by mixing them with s-wave states by external in-plane electric field and show that a moderate DC electric field would provide sufficient mixing to brighten p-wave states. We calculate energies of the p-wave states including the effects of valley-orbital splitting and the orbital Zeeman shift, and show that this provides direct information on the electron-hole mass asymmetry.

I Introduction

Monolayer (1L) Transition Metal Dichalcogenides (TMDs) are a class of two dimensional (2D) semiconducting materials that feature a direct band gap at the K points of the Brillouin zone, quadratic band structure and large band gaps Wang et al. (2018); Scharf et al. (2019). The strong Coulomb interaction in 1L TMDs results in huge exciton binding energies which causes excitons to dominate the absorption spectrum even at room temperature Wang et al. (2018); Scharf et al. (2019); Chen et al. (2019); Meckbach et al. (2018); Qiu et al. (2016, 2019). 1L TMD semiconductors are 2D systems with hexagonal Brillouin zones, however they lack inversion symmetry. Both the conduction and valence bands are spin split by the spin-orbit interaction and the spin and valley degrees of freedom are coupled Kormányos et al. (2015); Stier et al. (2016); Rostami et al. (2013); Wang et al. (2018); Scharf et al. (2019). The unique properties of 1L TMDs render them an ideal platform for the study of exciton physics  Chen et al. (2019); Wu et al. (2015); Liu et al. (2021); Goldstein et al. (2020); Liu et al. (2020); Bange et al. (2023); Zhu et al. (2023) and dynamical screening effects, particularly in regards to quasiparticle band gap renormalization Gao et al. (2016); Liang and Yang (2015); Qiu et al. (2019, 2016); Zibouche et al. (2021); Liu et al. (2019). Further, 1L TMDs are ideal for the development of optoelectronic devices Wang et al. (2012); Mueller and Malic (2018); Taffelli et al. (2021), including photosensor devices  Yin et al. (2012); Velusamy et al. (2015); Chang et al. (2014); Zhang et al. (2013); Lopez-Sanchez et al. (2013); Gonzalez Marin et al. (2019) and the fabrication of vertical and lateral heterostructures with advanced optical performance, with such devices providing an excellent environment for light-matter coupling in the form of exciton-polaritons.  Dufferwiel et al. (2018); Zhao et al. (2023). As such, accurate determination of the effective Hamiltonian parameters of 1L TMDs would be beneficial for modelling these technologies, and several others, including single photon emitters Sortino et al. (2021); Schuler et al. (2020); Srivastava et al. (2015); Tonndorf et al. (2015).

Exciton physics in 2D TMD systems is highly sensitive to reduced (band) mass, characteristic screening length and the external dielectric constant and hence the appropriate analysis of exciton resonances can be used to determine these quantities. Further to this, we will show how the electron and hole band mass asymmetry can also be extracted from excitonic spectroscopy, providing a near complete picture of the underlying band structure (effective Hamiltonian).

In the present work we consider excitons in magnetic and electric fields. While our conclusions are generic, to be specific we concentrate on WSe2. There are two major messages of our work. (i) At realistic out-of plane magnetic fields a simple perturbation theory approach in magnetic field is not sufficient and one needs an exact solution to analyse existing data. Developing the solution and performing analysis of the data we determine parameters of the system which differ significantly from those obtained with simple perturbation theory. (ii) While s-wave TMD excitons have been observed in a number of photoexcitation experiments, p-wave excitons are invisible because of the extremely small spectral weight. Application of a moderate in-plane electric field makes p-wave exciton states visible in photoexcitation. We predict properties of the p-wave states: energies, photoexcitation probabilities, valley-orbital level splitting and orbital Zeeman effect due to the electron-hole mass-asymmetry. Experimental studies of p-wave states are absolutely feasible and they will allow the study of these effects and to establish rather precisely the mass-asymmetry.

(i) Measurements of diamagnetic energy shifts of exciton s-states provides a route to determine the effective parameters of 1L TMDs. The perturbation theory formula for the diamagnetic shift reads Landau and Lifshitz (2007),

δE=e28μr2B2,\delta E=\frac{e^{2}}{8\mu}\langle r^{2}\rangle B^{2}, (1)

with e=|e|e=|e| the elementary charge, μ\mu the reduced mass, r2\langle r^{2}\rangle the square of the radius of the exciton in a given quantum state, and BB the out-of-plane magnetic field strength. This formula is used to fit experimental data on TMD excitons Walck and Reinecke (1998); Chen et al. (2019); Stier et al. (2016). We will show that this formula, obtained from a perturbation theory in BB, is insufficient at strong BB-fields. And instead, we will derive the appropriate strong field behaviour of the diamagnetic shift. This was previously shown for a 2D Coulomb problem in magnetic field in the strong field and weak field limits MacDonald and Ritchie (1986); Laird et al. (2022) but in this work we consider a Keldysh potential and present a solution for arbitrary magnetic field, similar to Ref. Kezerashvili and Spiridonova (2021) however we extend the method to account for arbitrary orbital angular momentum and particle-hole mass asymmetry which is essential for real systems. Using our non-perturbative approach, we can provide accurate fitting to available experimental data, and in doing so, establish the key parameters for the low-energy TMD Hamiltonian.

(ii) Physics of p-wave exciton states is more rich than that of s-wave states. It includes the valley-orbital effect (the splitting of l=±1l=\pm 1 states without magnetic field) and the orbital Zeeman effect (the splitting of l=±1l=\pm 1 states in magnetic field). These phenomena are intimately related to the valley anomalous Hall effect in electron transport. Unfortunately the p-wave states are invisible in photoexcitation; these are “dark” states. We propose to make the states visible by applying an in-plane electric field. The electric field should be sufficuently strong to brighten the p-wave states relative to the s-wave states, but still remain weak enough such that bound states are not destroyed. We find the appropriate electric field strength is E13E\sim 1-3V/μ\mum. We calculate values of the aforementioned effects in combined electric and magnetic fields.

The rest of the paper is organised as follows: Section II discusses the diamagnetic shift and existing photoluminescence experimental data. Section III walks through the derivation of the effective Hamiltonian for exciton in 1L TMD. In Section IV we present our results for s-wave excitons. We compare our theory to experimental results and extract reduced mass, dielectric constant and characteristic screening length. In Section V we consider excitons in combined in-plane electric and out-of-plane magnetic fields and show how such measurements can allow the study of anomalous effects in 1L TMDs. We summarise our conclusions in Section VI.

II Background: The Diamagnetic Shift

To stress the shortcomings of Eq.(1), we present Fig. 1, taken from experimental paper Ref. Chen et al. (2019), which plots the excitonic energies against magnetic field in 1L WSe2 encapsulated in hBN. The data shows a combination of the diamagnetic shift, which is quadratic in B, and a linear in B valley Zeeman effect that we discuss in Appendix A. The valley Zeeman effect is not exciton state specific, it is common for all exciton states. To exclude the valley Zeeman effect the data should be symmetrised in B. It is clear by inspection of Fig. 1 that the 3s and 4s diamagnetic shifts are not quadratic at high BB and hence cannot be described by Eq.(1); the perturbation theory is insufficient. A strong field approach is necessary.

There is an interesting theoretical aspect of the problem that we would like to stress. In the weak B-field limit the expression for the diamagnetic energy shift, Eq.(1), is independent of the dimensionality of the exciton. However, in strong BB-field the exciton dynamics strongly depends on dimensionality. In the three-dimensional (3D) case the exciton wave function is like a needle aligned with the magnetic field, and the Larmor circle oscillates along the needle resulting in a series of 1D Coulomb levels built on each Landau level  Landau and Lifshitz (2007). In the case of a 2D exciton the zz-confinement does not allow this oscillation. The exciton remains confined to the xx-yy plane at any field MacDonald and Ritchie (1986). Hence, diamagnetic shifts in strong B are very different in 3D and 2D cases.

Refer to caption

Figure 1: Exciton energy of s-wave excitons in 1L WSe2 encapsulated in hBN in a magnetic field as measured using photoluminescence, taken from Ref. Chen et al. (2019). The colour plot shows the intensity of photoluminescence spectra for 1s, 2s, 3s and 4s excitons. A particular valley is selected using the circularly polarised light. There is a linear valley Zeeman effect that causes an offset between B<0B<0 and B>0B>0 results. The superimposed green lines are plots of the photoluminescence spectra at B=0B=0 T and B=31B=-31 T.

III Hamiltonian for Exciton in Magnetic Field

In the presence of a magnetic field, the formation of an exciton can be described by the following Hamiltonian

H^=(𝐩^e+ec𝐀e)22me+(𝐩^hec𝐀h)22mh+V(𝐫e𝐫h)\begin{split}\hat{H}=\frac{\left(\hat{\mathbf{p}}_{\mathrm{e}}+\frac{e}{c}{\mathbf{A}}_{\mathrm{e}}\right)^{2}}{2m_{\mathrm{e}}}+\frac{\left(\hat{\mathbf{p}}_{\mathrm{h}}-\frac{e}{c}{\mathbf{A}}_{\mathrm{h}}\right)^{2}}{2m_{\mathrm{h}}}+V(\mathbf{r}_{e}-\mathbf{r}_{h})\end{split} (2)

Here 𝐩^e(h)\hat{\mathbf{p}}_{\mathrm{e}(\mathrm{h})} is the momentum operator for the electron (hole), and me(h)m_{\mathrm{e}(\mathrm{h})} is its effective mass. In the cylindrical gauge, the magnetic field B can be described by the vector potential 𝐀e(h)=[𝐁×𝐫e(h)]/2\mathbf{A}_{\mathrm{e}(\mathrm{h})}=[\mathbf{B}\times\mathbf{r}_{\mathrm{e}(\mathrm{h})}]/2. If TMD monolayer is deposited at the insulating substrate, the Coulomb attraction is accurately described by the Rytova-Keldysh potential given by

V(𝐫)=πe22ϵr0[H0(rr0)Y0(rr0)].V(\mathbf{r})=-\frac{\pi e^{2}}{2\epsilon r_{0}}\left[H_{0}\left(\frac{r}{r_{0}}\right)-Y_{0}\left(\frac{r}{r_{0}}\right)\right]. (3)

Here H0H_{0} and Y0Y_{0} are the Struve and Bessel functions of the second kind respectively and r0r_{0} is the characteristic screening length. The length r0=2πα/ϵr_{0}=2\pi\alpha/\epsilon is determined by the polarizability of the TMD monolayer α\alpha and effective dielectric constant ϵ\epsilon of the surrounding media.

It is instructive to introduce the center of mass 𝐑=(me𝐫e+mh𝐫h)/(me+mh)\mathbf{R}=(m_{\mathrm{e}}\mathbf{r}_{\mathrm{e}}+m_{\mathrm{h}}\mathbf{r}_{\mathrm{h}})/(m_{\mathrm{e}}+m_{\mathrm{h}}) and relative 𝐫=𝐫e𝐫h\mathbf{r}=\mathbf{r}_{\mathrm{e}}-\mathbf{r}_{\mathrm{h}} coordinates. The corresponding momenta are 𝐏=i𝐑{\mathbf{P}}=-i\hbar\frac{\partial}{\partial{\mathbf{R}}} and 𝐩=i𝐫{\mathbf{p}}=-i\hbar\frac{\partial}{\partial{\mathbf{r}}}. For the remainder of this work we shall use units =c=1\hbar=c=1. The Lamb transformation Lamb (1952) H^UH^U{\hat{H}}\to U^{{\dagger}}{\hat{H}}U governed by the unitary operator U=eie2𝐁[𝐑×𝐫]U=e^{i\frac{e}{2}\mathbf{B}\cdot[\mathbf{R}\times\mathbf{r}]} further simplifies Hamiltonian Eq.(2) as

H^\displaystyle\hat{H} =\displaystyle= 𝐩22μ+μωc2𝐫28+V(𝐫)+νωc2[𝐫×𝐩]z\displaystyle\frac{\mathbf{p}^{2}}{2\mu}+\frac{\mu\omega_{\mathrm{c}}^{2}\mathbf{r}^{2}}{8}+V(\mathbf{r})+\frac{\nu\omega_{\mathrm{c}}}{2}[\mathbf{r}\times\mathbf{p}]_{z} (4)
+\displaystyle+ eBM[𝐫×𝐏]z+𝐏22M\displaystyle\frac{eB}{M}[\mathbf{r}\times\mathbf{P}]_{z}+\frac{\mathbf{P}^{2}}{2M}

with

M=me+mh,μ=mememe+mh,ν=memhme+mh.M=m_{e}+m_{h},\quad\mu=\frac{m_{\mathrm{e}}m_{\mathrm{e}}}{m_{\mathrm{e}}+m_{\mathrm{h}}},\quad\nu=\frac{m_{\mathrm{e}}-m_{\mathrm{h}}}{m_{\mathrm{e}}+m_{\mathrm{h}}}. (5)

We have also introduced the cyclotron frequency ωc=eB/μ\omega_{c}=eB/\mu. We restrict ourselves only to the optically active excitons, i.e. excitons with zero center of mass momentum 𝐏=0\mathbf{P}=0. Besides, the fourth term is determined by the orbital momentum for the relative motion L^z=[𝐫×𝐩]z\hat{L}_{z}=[\mathbf{r}\times\mathbf{p}]_{z}, which is a good quantum number L^l\hat{L}\to l. It describes the orbital Zeeman shift of excitonic levels and is nonzero only for non-s-wave states.

The excitonic states are shaped by the interplay between the attractive Rytova-Keldysh potential and the harmonic confinement due to the magnetic filed. They smoothly evolve from the 2D hydrogenic-like series to the equidistant set of Landau levels. It is convenient to calculate the magnetic field dependence of excitonic states using the momentum space representation, i.e., ψ(𝐫)ψ𝐩\psi(\mathbf{r})\to\psi_{\bf p}. The eigenvalue problem with Hamiltonian (4) transforms to the integro-differential equation given by

[𝐩22μμωc28𝐩2+lνωc2]ψ𝐩+𝐩V𝐩𝐩ψ𝐩=ϵψ𝐩,\left[\frac{\mathbf{p}^{2}}{2\mu}-\frac{\mu\omega^{2}_{c}}{8}\nabla_{\mathbf{p}}^{2}+\frac{l\nu\omega_{\mathrm{c}}}{2}\right]\psi_{\mathbf{p}}+\sum_{\mathbf{p}^{\prime}}V_{\mathbf{p}-\mathbf{p}^{\prime}}{\psi_{\mathbf{p}^{\prime}}}=\epsilon\psi_{\mathbf{p}}, (6)

where V𝐩=2πe2/εp(1+r0p)V_{\bf p}=-2\pi e^{2}/\varepsilon p(1+r_{0}p) is the Fourier transform of the Rytova-Keldysh potential. It can be further simplified if we split the orbital part of the wave function ψ𝐩=ψpleilθp\psi_{\bf p}=\psi_{p}^{l}e^{il\theta_{p}} and rescale its radial part as ψpl=χpl/p\psi_{p}^{l}=\chi_{p}^{l}/\sqrt{p}, allowing the radial Hamiltonian to be written in a form that is strictly Hermitian. The resulting eigenvalue problem is given by

[p22μ+μωc28(d2dp2+l21/4p2)+lνωc2]χpl\displaystyle\left[\frac{p^{2}}{2\mu}+\frac{\mu\omega_{\mathrm{c}}^{2}}{8}\left(-\frac{d^{2}}{dp^{2}}+\frac{l^{2}-1/4}{p^{2}}\right)+\frac{l\nu\omega_{\mathrm{c}}}{2}\right]\chi_{p}^{l}
+0ppdp2πVl(p,p)χpl=ϵχpl.\displaystyle+\int_{0}^{\infty}\frac{\sqrt{p\;p^{\prime}}dp^{\prime}}{2\pi}\;V_{l}(p,p^{\prime})\chi_{p^{\prime}}^{l}=\epsilon\chi_{p}^{l}\ . (7)

Here Vl(p,p)=V𝐩𝐩eilφφV_{l}(p,p^{\prime})=\langle V_{\mathbf{p}-\mathbf{p}^{\prime}}e^{il\varphi}\rangle_{\varphi} is the multipole moment of interactions and involves averaging over the relative polar angle φ\varphi between momenta 𝐩\mathbf{p} and 𝐩\mathbf{p}^{\prime}. We solve Eq.(III) numerically using the linear algebra LAPACK package. Further details of these calculations are presented in Appendix B.

IV s-wave Excitons

Refer to caption
Refer to caption
Figure 2: (a) The magnetic field dependence of s-wave excitonic states in WSe2 monolayer encapsulated in hBN. Our theoretical curves (solid lines) are an excellent match to the experimental data from Ref. Chen et al. (2019) (asterisks). Dashed lines show the theory with the set of parameters obtained by fitting Chen et al. (2019) the ground excitonic state with the perturbative expression, Eq.(1). (b) The first derivative of excitonic energies EnE_{n} (solid lines) and their asymptotic values En=ωc(nr+1/2)E^{\infty}_{n}=\omega_{\mathrm{c}}(n_{r}+1/2) (dashed lines) with respect to the magnetic field. The 1s state remains in the parabolic regime while the Rydberg states indicate transition to the quantized harmonic regime even at experimentally realistic BB field.

We wish to compare the measured magnetic field dependence of energies EXE_{\mathrm{X}} for the s-wave excitonic resonances Chen et al. (2019) with the computed excitonic energies ϵ\epsilon. They are connected as EX=Δ+ϵE_{\mathrm{X}}=\Delta+\epsilon where Δ1.9\Delta\approx 1.9 eV is the band gap in the TMD monolayer. Hence we have 3 fitting parameters: the reduced electron-hole mass μ\mu, the dielectric constant ε\varepsilon, and the screening length r0r_{0}. As presented in panel a of Fig. 2, our theoretical curves (solid lines) provide an excellent fit of the experimental data (asterisks) for all resolved excitonic states (1s - 4s)111We reiterate that the experimental points in Fig. 2 correspond to that of Fig. 1, yet averaged over +B+B and B-B. This averaging removes the linear valley Zeeman effect. The fitting parameters are

μ=0.175m0,ε=3.9,r0=4.5nm/ε.\displaystyle\mu=0.175m_{0}\ ,\ \ \ \varepsilon=3.9\ ,\ \ \ r_{0}=4.5nm/\varepsilon\ . (8)

We make special mention of our fitted value ε=3.9\varepsilon=3.9, which is within the expected range for hBN Dean et al. (2010). With these parameters, the ground state binding energy at zero magnetic field is found to |ϵ1s|=173|\epsilon_{1s}|=173meV.

The original experimental work Chen et al. (2019) has used the simple parabolic approximation (1) for the diamagnetic shift. This has led to a different set of parameters: μ=0.22\mu=0.22 m0, ε=4.5\varepsilon=4.5, and r0=4.5r_{0}=4.5 nm/ε\varepsilon. The theoretical curves (dashed lines in Fig. 2a) calculated with this set of parameters fit the data well only for the ground excitonic state, but the discrepancy for the excited excitonic states is evident. Our conclusion aligns well with a recent unpublished work 222David de la Fuente Pico, Jesper Levinson, Meera M. Parish and Francesca Maria Marchetti, private communication..

Refer to caption
Refer to caption
Figure 3: (a) Root mean square radius of 1s, 2s, 3s, and 4s excitons as a function of magnetic field. The dashed red line is the magnetic length lB=e/Bl_{B}=\sqrt{e/B}. (b) Electric dipole radial matrix element rnp,nsr_{np,ns} for n=2n=2, n=3n=3 and n=4n=4 as a function of magnetic field.

The behaviour of the eigenenergies of the Rydberg states strays from the parabolic diamagnetic approximation and it can be deduced from Eq.(III) that in the large BB limit, BB\to\infty, the eigenergies approach those of 2D Landau levels. In this case the eigenenergy is quantized in radial quantum number nrn_{r} and angular momentum ll as ϵnr,l=ωc(2nr+|l|+νl+1)/2\epsilon_{n_{r},l}=\omega_{c}\left(2n_{r}+|l|+\nu l+1\right)/2 and is therefore linear in magnetic field BB. The radial quantum number nrn_{r} is related to the principle quantum number nn as nr=n|l|1n_{r}=n-|l|-1. In panel b of Fig. 2 we present the calculated first derivative of the excitonic energy calculated from Eq.(III) with respect to magnetic field. Solid lines in Fig. 2b correspond to solid lines in Fig. 2a. Dashed lines in Fig. 2b correspond to asymptotic values of the derivatives (BB\to\infty) determined by Landau levels in 2D. While in the experimental region 0<B<300<B<30T the derivative of the 1s state is practically linear in BB in accordance with perturbation theory, for higher states deviations from linearity are evident. Thus it is clear that the Rydberg states transition from the diamagnetic regime towards the Landau level regime at experimentally realistic BB, rendering the quadratic diamagnetic approximation Eq.(1) invalid for the Rydberg excitons.

The region of validity of the diamagnetic approximation is also illustrated by considering the size of the exciton, or the root mean square (rms) radius rrmsr_{rms}. The size of the exciton may be calculated directly from the wavefunction.

|r2|=ψ|r2|ψ=0((ψplp)2p+(ψpl)2l2p)dp2π\langle|r^{2}|\rangle=\langle\psi|r^{2}|\psi\rangle=\int_{0}^{\infty}\left(\left(\frac{\partial\psi_{p}^{l}}{\partial p}\right)^{2}p+(\psi_{p}^{l})^{2}\frac{l^{2}}{p}\right)\frac{dp}{2\pi} (9)

The exciton size rrms=ψ|r2|ψr_{rms}=\sqrt{\langle\psi|r^{2}|\psi\rangle} is plotted against magnetic field in panel a of Fig. 3. The 1s exciton radius is 2\sim 2 nm and remains smaller than the magnetic length lB=e/Bl_{B}=\sqrt{e/B} over the entire range of magnetic field, however this is not true for the Rydberg excitons. Thus, to accurately model the Rydberg states over this range of BB the exact Hamiltonian (III) is required.

V Detection of p-wave excitons via mixing due to in-plane electric field

Refer to caption
Refer to caption
Figure 4: Relative excitation probabilities of 2s2s^{\prime}, 2p+2p_{+}^{\prime}, and 2p2p_{-}^{\prime} exciton states in different valleys of WSe2 versus external in-plane electric field E. Linear polarization of the exciting laser is assumed. Panel a corresponds to zero out-of-plane magnetic field, B=0B=0. Panel b corresponds to B=30B=30T. Solid lines correspond to τ=+1\tau=+1 valley and dashed lines correspond to τ=1\tau=-1 valley. For B=0 the p±p_{\pm}^{\prime} curves in different valleys are identical, but what is the p+p_{+}^{\prime} curve for one valley is the pp_{-}^{\prime} curve for the other valley. For B0B\neq 0 all the curves are different.

The physics of p-wave exciton states is richer than that of s-wave states. It includes the valley-orbital effect (the splitting of l=±1l=\pm 1 states without magnetic field) and the orbital Zeeman effect (the splitting of l=±1l=\pm 1 states in magnetic field). Unfortunately the spectral weight of direct photoexcitation of a p-wave exciton calculated in Appendix C is too small for direct detection, 103\sim 10^{-3} in relative units compared with the s-wave state. However, one can mix s and p states by applying in-plane electric field and hence excite the p state due to admixture of the s state.

Refer to caption
Refer to caption
Figure 5: Eigenenergies of 2s2s^{\prime}, 2p+2p_{+}^{\prime}, and 2p2p_{-}^{\prime} exciton states in different valleys of WSe2 versus external in-plane electric field E. Panel a corresponds to zero out-of-plane magnetic field, B=0B=0. Panel b corresponds to B=30B=30T. Solid lines correspond to τ=+1\tau=+1 valley and dashed lines correspond to τ=1\tau=-1 valley.

There is a recent experiment where the p-wave states have been observed using this method Zhu et al. (2023). However, in experiment Zhu et al. (2023) the electric field was relatively strong and an interpretation in terms of simple electron-hole bound states (excitons) is questionable. The major part of Ref. Zhu et al. (2023) data corresponds to events above the exciton ionisation limit where the observed resonances correspond to doorway scattering states 333O. P. Sushkov, J. N. Engdahl and D. K. Efimkin, to be published. Contrary to this in the present work we consider a relatively weak electric field when the description in terms of bound excition states is correct.

Energies of np-states with or without magnetic field can be calculated using Eq.(III). The np level is always slightly lower than the ns-level. However, Eq.(III) is missing an important effect: the valley-orbital level splitting between the states p+p_{+} and pp_{-} that correspond to angular momentum l=+1l=+1 and l=1l=-1, where this splitting exists even at B=0. Using the language of Dirac equation one can say that Eq.(III) corresponds to the “non-relativistic approximation” to the Dirac equation that at B=0B=0 gives degenerate p±p_{\pm} states, and the p+p_{+} - pp_{-} splitting is the 1st “relativistic” correction to this equation, The splitting is discussed in Appendix D, it is small, from only a few to several meV, and if it is positive in one valley, Δϵτ=ϵϵ+>0\Delta\epsilon_{\tau}=\epsilon_{-}-\epsilon_{+}>0, it is negative in the other valley.

Having in mind that energies of nsns, np+np_{+} and npnp_{-} states with the same n are close and assuming that the electric field E=ExE=E_{x} is not too strong we can restrict our analysis to a 3-level approximation with the effective Hamiltonian

H^=[ϵs,0vs,pvs,p+vp,sϵp,0+Δϵτ/20vp+,s0ϵp+,0Δϵτ/2]\displaystyle\hat{H}=\begin{bmatrix}\epsilon_{s,0}&v_{s,p_{-}}&v_{s,p_{+}}\\ v_{p_{-},s}&\epsilon_{p_{-},0}+\Delta\epsilon_{\tau}/2&0\\ v_{p_{+},s}&0&\epsilon_{p_{+},0}-\Delta\epsilon_{\tau}/2\end{bmatrix} (10)

Here ϵa,0\epsilon_{a,0} is the eigenenergy of the state with angular momentum lal_{a} calculated with Eq.(III). Δϵτ\Delta\epsilon_{\tau} is the valley-orbital splitting, and va,b=eExxa,bv_{a,b}=eE_{x}x_{a,b}, where

xa,b=a|x|b=ψaeilaφpx(ψbeilbφ)pdp2π=rps/2.x_{a,b}=\langle a|x|b\rangle=\int\psi_{a}^{*}e^{-il_{a}\varphi}\frac{\partial}{\partial p_{x}}\left(\psi_{b}e^{il_{b}\varphi}\right)\frac{pdp}{2\pi}=r_{ps}/2\ . (11)

In panel b of Fig. 3 we present plots of the electric dipole radial matrix element rnp,nsr_{np,ns} for n=2,3,4n=2,3,4 as a function of magnetic field.

Before diagonalization of (10) the energies ϵs,0\epsilon_{s,0}, ϵp+,0\epsilon_{p_{+},0} and ϵp,0\epsilon_{p_{-},0} must be calculated from Eq.(III). From the fit of the s-wave states we know all the parameters of this equation except for ν\nu, the particle-hole asymmetry term defined in (5). Here we take ν=0.1\nu=-0.1, which is obtained from the effective electron and hole masses determined from DFT calculations in Ref. Kormányos et al. (2015). However, armed with our formalism, we suggest that analysis of future experiments can establish ν\nu, providing vital information on the underlying band structure. The valley-orbital splitting Δϵτ\Delta\epsilon_{\tau} is discussed and calculated in Appendix D. In principle it depends on magnetic field, but for n=2 this dependence is negligible over the range of magnetic field that we consider. The splitting depends on the velocity in the effective Dirac equation, see Appendix D. We approximate the velocity from the band gap and the effective mass as v2Δ/4μv^{2}\approx\Delta/4\mu, which results in a valley-orbital splitting Δϵτ±3.8\Delta\epsilon_{\tau}\approx\pm 3.8meV. Finally we reiterate the point that we already have made in Section II: the linear in B valley Zeeman effect that is not exciton specific and can be determined from the 1s state data is subtracted from all our results. After this preparatory procedure, diagonalization of Eq.(10) is straightforward. We use nomenclature ss^{\prime}, p+p_{+}^{\prime} and pp_{-}^{\prime} to denote the states that originate from ss, p+p_{+} and pp_{-} at E=B=0E=B=0. The photoexcitation probability is proportional to the weight of the bare s-wave state in the corresponding wave function. In Fig.4 we plot the probabilities versus electric field for n=2 states. The probabilities are normalized such that P2s=1P_{2s}=1 at E=0E=0. As previosuly mentioned, the approach used in the present work fails at strong electric fields. This is due to the effective potential of the Keldysh potential combined with the strong electric field failing to permit exciton bound states. A separate analysis of strong electric fields 444O. P. Sushkov, J. N. Engdahl and D. K. Efimkin, to be published shows that for n=2 states in WSe2 the approach employed in the present work is valid up to E=34E=3-4V/μ\mum, above this field bound n=2 exciton states do not exist. This is where we terminate our plots in Fig.4. Here the probability of the pp^{\prime}-state excitation can be up to 20% of that of the s-wave. The relative probabilities in Fig.4 assume linear polarization of the excitation laser that does not separately select single valleys. Panel a in Fig.4 displayes probabilities at B=0. Plots for different valleys are identical, but what is the p+p_{+}^{\prime} curve for one valley is the pp_{-}^{\prime} curve for the other valley. Panel b in Fig.4 displayes probabilities at B=30T.

In Fig.5 we plot energies of 2s2s^{\prime}, 2p+2p_{+}^{\prime}, and 2p2p_{-}^{\prime} states versus external in-plane electric field E for two values of magnetic field, B=0B=0 and B=30B=30T. These plots depend on the valley-orbital energy splitting between the p±p_{\pm} levels, Δϵτ\Delta\epsilon_{\tau}, and also on the mass asymmetry parameter ν\nu-term in Eq.(4). In Fig.5 we assume values that follow from DFT calculationsKormányos et al. (2015) and indirect analysis of ARPES dataNguyen et al. (2019). Measurements of these energy levels would shed light on true values of these parameters.

VI Conclusion

We have shown that the perturbation theory approximation for diamagnetic shift of exciton energy levels is not sufficient to describe existing data. We develop the theory for strong magnetic field, reanalyse the data for monolayer WSe2, and arrive to the set of exciton parameters presented in Eq.(8). This set is different from what was known in literature, although the reduced mass is similar to that obtained from DFT calculations.

Only s-wave excitons are visible in photoluminescence, the spectral weight of p-wave states is too small at just 103\sim 10^{-3} relative to the s-wave. We propose to study p-wave states by mixing them with s-wave states by application of external in-plane electric field and show that a moderate electric field E3V/μmE\approx 3V/\mu m results in the intensity about 20% relative to s-wave. We calculate energy levels in combined electric and magnetic fields and demonstrate that our proposal opens a path to experimentally study the valley-orbital level splitting and the electron-hole mass asymmetry.

Acknowledgements

We acknowledge discussions with Zeb Krix, Francesca Marchetti and Jesper Levinson. This work was supported by the Australian Research Council Centre of Excellence in Future Low- Energy Electronics Technologies (CE170100039).

Appendix A Reduction of the Dirac like Hamiltonian to the low energy Hamiltonian, valley Zeeman g-factor.

Start from the 2D Dirac Hamiltonian, Eq.(12), for a particle in 1L TMD Rostami et al. (2013), truncated to order quadratic in momentum, 𝐪=ir{\bf q}=-i\partial_{r}. Parameters α\alpha and β\beta define particle-hole mass asymmetry, γ1\gamma\approx 1 determines spin splitting of the conduction band (γ=1\gamma=1 gives only splitting of the valence band) and λ\lambda determines the magnitude of this spin splitting. Δ0\Delta_{0} is the band gap in the absence of spin-orbit coupling and vv is Fermi-Dirac velocity. Note that in this notation the valley and spin indices are τ=±1\tau=\pm 1 and s=±1s=\pm 1.

H^=Δ02σz+λτsγσz2+v𝐪στ+|𝐪|24m0(α+βσz)\hat{H}=\frac{\Delta_{0}}{2}\sigma_{z}+\lambda\tau s\frac{\gamma-\sigma_{z}}{2}+v\mathbf{q}\cdot\mathbf{\sigma_{\tau}}+\frac{|\mathbf{q}|^{2}}{4m_{0}}(\alpha+\beta\sigma_{z}) (12)

We can write this as a 2x2 eigensystem, where q±=τqx±iqyq_{\pm}=\tau q_{x}\pm iq_{y}, στ=(τσx,σy)\sigma_{\tau}=(\tau\sigma_{x},\sigma_{y}), λ1=(λτs/2)(γ1)\lambda_{1}=(\lambda\tau s/2)(\gamma-1) and λ2=(λτs/2)(γ+1)\lambda_{2}=(\lambda\tau s/2)(\gamma+1).

H^=[Δ02+λ1+q24m0(α+β)vqvq+Δ02+λ2+q24m0(αβ)]\hat{H}=\\ \begin{bmatrix}\frac{\Delta_{0}}{2}+\lambda_{1}+\frac{q^{2}}{4m_{0}}(\alpha+\beta)&vq_{-}\\ vq_{+}&\frac{-\Delta_{0}}{2}+\lambda_{2}+\frac{q^{2}}{4m_{0}}(\alpha-\beta)\end{bmatrix} (13)
H^[ψ1ψ2]=ϵ[ψ1ψ2]\displaystyle\hat{H}\begin{bmatrix}\psi_{1}\\ \psi_{2}\end{bmatrix}=\epsilon\begin{bmatrix}\psi_{1}\\ \psi_{2}\end{bmatrix} (14)

Excitons that we consider have binding energy much smaller than the band gap. For Dirac equation this corresponds to non-relativistic limit. Hence we follow the standard procedure of derivation of Pauli equation from Dirac equation Beresteckij et al. (2008). We explicitly write out the upper and lower component equations, substituting ψ2\psi_{2} and ψ1\psi_{1} into the equations respectively and then taking the limit ϵ±Δ\epsilon\approx\pm\Delta. For the positive energy energy solution ψ2ψ1\psi_{2}\ll\psi_{1} and for the negative energy energy solution ψ1ψ2\psi_{1}\ll\psi_{2}. Hence we arrive to the following energies.

ϵ1=Δ02+λτs12(γ1)+q124m0(α+β)+v2q12Δ0λτs12(γ+1)ϵ2=Δ02+λτs22(γ+1)+q224m0(αβ)v2q22Δ0+λτs22(γ1)\epsilon_{1}=\frac{\Delta_{0}}{2}+\frac{\lambda\tau s_{1}}{2}(\gamma-1)+\frac{q_{1}^{2}}{4m_{0}}(\alpha+\beta)+\frac{v^{2}q_{1}^{2}}{\Delta_{0}-\frac{\lambda\tau s_{1}}{2}(\gamma+1)}\\ \epsilon_{2}=\frac{-\Delta_{0}}{2}+\frac{\lambda\tau s_{2}}{2}(\gamma+1)+\frac{q_{2}^{2}}{4m_{0}}(\alpha-\beta)-\frac{v^{2}q_{2}^{2}}{\Delta_{0}+\frac{\lambda\tau s_{2}}{2}(\gamma-1)} (15)

Now, to go to the hole description instead of the electron description, we perform charge conjugation on the negative energy solution, where 𝐪𝟐𝐪𝟐\mathbf{q_{2}}\to\mathbf{-q_{2}}, s2s2s_{2}\to-s_{2} and ϵ2ϵ2,+\epsilon_{2}\to-\epsilon_{2,+}.

ϵ2\displaystyle\epsilon_{2} =\displaystyle= Δ02+λτs22(γ+1)q224m0(αβ)\displaystyle\frac{\Delta_{0}}{2}+\frac{\lambda\tau s_{2}}{2}(\gamma+1)-\frac{q_{2}^{2}}{4m_{0}}(\alpha-\beta) (16)
+\displaystyle+ v2q22Δ0λτs22(γ1)\displaystyle\frac{v^{2}q_{2}^{2}}{\Delta_{0}-\frac{\lambda\tau s_{2}}{2}(\gamma-1)}

In presence of a magnetic field we first perform gauge replacement 𝐪𝐪+e𝐀\mathbf{q}\to\mathbf{q}+e\mathbf{A} in the Hamiltonian (12) and then repeat the above procedure. Hence, we arrive at the following.

ϵ1=Δ02+λτs12(γ1)+|𝐪𝟏+e𝐀𝟏|24m0(α+β)+v2((𝐪𝟏+e𝐀𝟏)2+τeB)Δ0λτs12(γ+1)ϵ2=Δ02+λτs22(γ+1)|𝐪𝟐e𝐀𝟐|24m0(αβ)+v2((𝐪𝟐e𝐀𝟐)2τeB)Δ0λτs22(γ1)\epsilon_{1}=\frac{\Delta_{0}}{2}+\frac{\lambda\tau s_{1}}{2}(\gamma-1)+\frac{|\mathbf{q_{1}}+e\mathbf{A_{1}}|^{2}}{4m_{0}}(\alpha+\beta)\\ +\frac{v^{2}((\mathbf{q_{1}}+e\mathbf{A_{1}})^{2}+\tau eB)}{\Delta_{0}-\frac{\lambda\tau s_{1}}{2}(\gamma+1)}\\ \epsilon_{2}=\frac{\Delta_{0}}{2}+\frac{\lambda\tau s_{2}}{2}(\gamma+1)-\frac{|\mathbf{q_{2}}-e\mathbf{A_{2}}|^{2}}{4m_{0}}(\alpha-\beta)\\ +\frac{v^{2}((\mathbf{q_{2}}-e\mathbf{A_{2}})^{2}-\tau eB)}{\Delta_{0}-\frac{\lambda\tau s_{2}}{2}(\gamma-1)} (17)

These equation represent effective single particle Hamiltonians for electron and hole.

Finally, let us consider an excitation of electron from the valence band to the conduction band via E1 optical transition. The excitation creates an electron and a hole. The spin of the electron is se=s1=ss_{e}=s_{1}=s and the spin of the hole is sh=s1=ss_{h}=-s_{1}=-s. The Hamiltonian of the system is the combined Hamiltonian of the electron and hole. The linear Zeeman term may be denoted as H^Z\hat{H}_{Z}.

H^Z\displaystyle\hat{H}_{Z} =\displaystyle= v2τeB(1Δ0+λτs2(γ1)1Δ0λτs2(γ+1))\displaystyle-v^{2}\tau eB\left(\frac{1}{\Delta_{0}+\frac{\lambda\tau s}{2}(\gamma-1)}-\frac{1}{\Delta_{0}-\frac{\lambda\tau s}{2}(\gamma+1)}\right) (18)
\displaystyle\approx v2τeBλΔ0(Δ0λ)\displaystyle-\frac{v^{2}\tau eB\lambda}{\Delta_{0}(\Delta_{0}-\lambda)}

Here in the second line we take the small conduction band spin-orbit-splitting limit γ1\gamma\approx 1. The v2v^{2}-terms in Eqs.(17) contribute to the effective masses of electron and hole. It is easy to check that the following relation is valid.

v2Δ0(Δ0λ)=12m0(2Δ0λ)(m0μ+β).\frac{v^{2}}{\Delta_{0}(\Delta_{0}-\lambda)}=\frac{1}{2m_{0}(2\Delta_{0}-\lambda)}\left(\frac{m_{0}}{\mu}+\beta\right). (19)

Note also that in the language of Eq. (13) Δ0\Delta_{0} refers to the average band gap for the two spin split band pairs, meaning for A-excitons the band gap is Δ=Δ0λ\Delta=\Delta_{0}-\lambda. Thus we may rewrite the valley Zeeman term as

H^ZτμBBλ2Δ+λ(m0μ+β)=gτμBB{\hat{H}}_{Z}\simeq-\tau\frac{\mu_{B}B\lambda}{2\Delta+\lambda}\left(\frac{m_{0}}{\mu}+\beta\right)=-g\tau\mu_{B}B (20)

The value of the spin-orbit constant according to DFT calculationsKormányos et al. (2015) and experimental measurements Nguyen et al. (2019) is about λ0.24\lambda\approx 0.24eV. The value of the reduced mass μ\mu is known, Eq.(8). The value of the Fermi-Dirac velocity extracted from ARPES data for a similar hBN/WSe2/hBN device Nguyen et al. (2019) is v0.4v\approx 0.4eVnm. Comparing this with Eq.(19) one finds that the β\beta-term in Eqs.(19),(20) is small compared to m0/μm_{0}/\mu. Hence, using Eq.(20) we find the theoretical prediction for the valley Zeeman g-factor, g0.34g\approx 0.34. On the other hand experimental data from Ref. Chen et al. (2019) presented in our Fig.1 gives g=2.1g=2.1, a dramatic disagreement between the theory and the experiment. Such a disagreement for the valley-Zeeman g-factors in TMD materials is known in literature Woźniak et al. (2020); Rybkovskiy et al. (2017). There are also claims in literature Woźniak et al. (2020) that account of multiple bands (up to 200 bands) can bring theory to agreement with experiment. The dramatic disagreemt is a very interesting problem, but it is beyond the scope of the current work since it is irrelevant to questions considered here.

Appendix B Numerical Methods

We solve Eq. III through linear algebra methods. The angular integration in the potential in Eq.  III is performed with 200200 points, so Δθ=2π/200\Delta\theta=2\pi/200. The radial momentum is discretized with 500 grid points, the grid step is Δp=0.01/aB\Delta p=0.01/a_{B}, where aB1.2a_{B}\approx 1.2 nm is effective Bohr radius. Hence the Hamiltonian in Eq.  III is a 500×500500\times 500 matrix. There is a small pitfall with this method, Eq.(III) is singular for s-wave states at p0p\to 0. If Eq.(III) for l=0l=0 is written symbolically as

H^=𝐩22μ+e2B2r28μc2+V(𝐫)\hat{H}=\frac{\mathbf{p}^{2}}{2\mu}+\frac{e^{2}B^{2}r^{2}}{8\mu c^{2}}+V(\mathbf{r}) (21)

is is clear that the singularity is due to r2r^{2} divergence of (21) at large r. Hence we have to be careful with discretization at small p. We resolve this by adding a regularization factor to the matrix elements at the smallest momentum. A simple test that the discretization is correct is that eigenenergies and eigenfunctions of Eq.(III) at V=0V=0 coincide with that of Eq.(21) obtained by the conventional analytic method.

Appendix C s-wave and p-wave relative photoexcitatin spectral weights

Firstly, consider the Hamiltonian for an insulator without electromagnetic field. In the Hamiltonian (12) we can disregard the second and the fourth terms, but we must add the trigonal warping  Rostami et al. (2013) tt-term that is responsible for excitation of a p-wave state. Hence we arrive at

H^=\displaystyle\hat{H}= Δ2σz+v𝐪στ+t𝐪στσx𝐪στ\displaystyle\frac{\Delta}{2}\sigma_{z}+v\mathbf{q}\cdot\mathbf{\sigma_{\tau}}+t\mathbf{q}\cdot\sigma_{\tau}^{*}\sigma_{x}\mathbf{q}\cdot\sigma_{\tau}^{*} (22)

In matrix form it is

H^=[Δ2vq+tq+2vq++tq2Δ2]\displaystyle\hat{H}=\begin{bmatrix}\frac{\Delta}{2}&vq_{-}+tq_{+}^{2}\\ vq_{+}+tq_{-}^{2}&\frac{-\Delta}{2}\end{bmatrix} (23)

It is convenient to work in the gauge where the scalar potential is zero. Hence the vecor potential potential and the electric field of the photon are related as 𝐀=𝐄0/ωeiωt{\bf A}={\bf E}_{0}/\omega e^{-i\omega t} To be specific we consider light linearly polarized along x.

A+=A=τAx=τA=E0ω.E=E0eiωtA_{+}=A_{-}=\tau A_{x}=\tau A=\frac{E_{0}}{\omega}\;.\;E=E_{0}e^{-i\omega t} (24)

Interaction with photon arises from the standard gauge replacement 𝐪𝐪+e𝐀{\bf q}\to{\bf q}+e{\bf A} in the Hamiltonian (23). Here 𝐀{\bf A} is vector potential of the photon. This give the following interaction Hamiltonian linear in 𝐀{\bf A}.

H^int\displaystyle\hat{H}_{int} =\displaystyle= [0evA+2etq+A+evA++2etqA0]\displaystyle\begin{bmatrix}0&evA_{-}+2etq_{+}A_{+}\\ evA_{+}+2etq_{-}A_{-}&0\end{bmatrix} (25)
\displaystyle\to eτA[0v+2tq+v+2tq0]\displaystyle e\tau A\begin{bmatrix}0&v+2tq_{+}\\ v+2tq_{-}&0\end{bmatrix}

The matrix element, M, for the electron exitation from the valence (lower) band to the conduction (upper) band is

MeτA=[10]H^inteτA[01]=v+2tq+\displaystyle\frac{M}{e\tau A}=\begin{bmatrix}1\\ 0\end{bmatrix}^{\dagger}\frac{\hat{H}_{int}}{e\tau A}\begin{bmatrix}0\\ 1\end{bmatrix}=v+2tq_{+} (26)

This is written in terms of plane waves. We need to combine the plane wave decomposition to the exciton wave function ψq\psi_{q}

MMψqd2q(2π)2M\to\int M\psi_{q}\frac{d^{2}q}{(2\pi)^{2}} (27)

At this point, recall q+=τqx+iqyq_{+}=\tau q_{x}+iq_{y} and ψq(q,ϕ)=ψqleilϕ\psi_{q}(q,\phi)=\psi_{q}^{l}e^{il\phi}. Below the superscript (+)(+) and ()(-) for p-wave refers to l=+1l=+1 and l=1l=-1 respectively.

MseτA=vψq,sqdq2π\frac{M_{s}}{e\tau A}=\int v\psi_{q,s}q\frac{dq}{2\pi} (28)
Mp(+)eτA=tψq,pq2(τ1)dq2π\frac{M_{p}^{(+)}}{e\tau A}=\int t\psi_{q,p}q^{2}(\tau-1)\frac{dq}{2\pi} (29)
Mp()eτA=tψq,pq2(τ+1)dq2π\frac{M_{p}^{(-)}}{e\tau A}=\int t\psi_{q,p}q^{2}(\tau+1)\frac{dq}{2\pi} (30)

We see that with linearly polarized light the chirality of the excited p-wave state is valley dependent. The Fermi-Dirac velocity is approximately vΔ/4μ0.45v\approx\sqrt{\Delta/4\mu}\approx 0.45 eVnm, see discussion in Appendix A. Hence, taking t=0.93×102t=-0.93\times 10^{-2} eVnm2 from DFT calculations from Ref. Kormányos et al. (2015), we calculate the magnitude of the matrix elements. Here we take n=2n=2 such that we consider 22s and 22p.

|M2seτA|6.2×102\displaystyle\left|\frac{M_{2s}}{-e\tau A}\right|\propto 6.2\times 10^{-2}
|M2peτA|3.4×103\displaystyle\left|\frac{M_{2p}}{-e\tau A}\right|\propto 3.4\times 10^{-3} (31)

The spectral weight is proportional to |M2||M^{2}|. Hence the ratio of 2s and 2p spectral weights is

w02pw02s=|M2p|2|M2s|23×103\frac{w_{0\to 2p}}{w_{0\to 2s}}=\frac{|M_{2p}|^{2}}{|M_{2s}|^{2}}\approx 3\times 10^{-3} (32)

Signatures of p-wave excitons are thus faint compared with s-wave excitons in such systems. We find this value is approximately independent of magnetic field strength for the fields considered in this work.

Appendix D Valley-orbital splitting

To calculate the valley-orbital we use the same method that is used for calculation of spin-orbit interaction in positronium Beresteckij et al. (2008). So, first we consider scattering of electron from hole. The scattering amplitude is given by the following diagram.

{feynman}\vertexpp\vertexpp^{\prime}\vertexpp\vertexpp^{\prime}\vertexelectronelectron\vertexholehole\vertex\vertex\vertex\vertex\vertex\vertex\diagramqq
Figure 6: Electron-hole scattering amplitude

The amplitude is

M=UqΨp|ΨpelectronΨp|Ψphole.M=-U_{q}\langle\Psi^{*}_{p^{\prime}}|\Psi_{p}\rangle_{electron}\langle\Psi^{*}_{p}|\Psi_{p^{\prime}}\rangle_{hole}. (33)

Here Uq=2πe2εq(1+r0q)U_{q}=\frac{2\pi e^{2}}{\varepsilon q(1+r_{0}q)} is the Fourier transform of repulsive Keldysh potential and Ψp\Psi_{p} is the two component Dirac-like eigenfunction of electron/hole discussed in Appendix A.

electron:\displaystyle electron:\ Ψp[1vp+Δ]\displaystyle\Psi_{p}\approx\begin{bmatrix}1\\ \frac{vp_{+}}{\Delta}\end{bmatrix}
hole:\displaystyle hole:\ \ \ \ \ \ \ Ψp=[vpΔ1]\displaystyle\Psi_{p}=\begin{bmatrix}-\frac{vp_{-}}{\Delta}\\ 1\end{bmatrix} (34)

The common sign (-) in Eq.(33) indicates that the interaction is attractive. Direct evaluation of (33) gives

M\displaystyle M =Uq(1+2v2Δ2pp++)\displaystyle=-U_{q}(1+2\frac{v^{2}}{\Delta^{2}}p_{-}^{\prime}p_{+}+...)
=Uq(1+2v2Δ2(p2+𝐪𝐩iτ[𝐩×𝐪]z))\displaystyle=-U_{q}(1+2\frac{v^{2}}{\Delta^{2}}(p^{2}+\mathbf{q}\cdot\mathbf{p}-i\tau[\mathbf{p}\times\mathbf{q}]_{z})...) (35)

Here 𝐪=𝐩𝐩{\bf q}={\bf p}^{\prime}-{\bf p}, we keep only the pp+iτ[𝐩×𝐪]zp_{-}^{\prime}p_{+}\to-i\tau[\mathbf{p}\times\mathbf{q}]_{z} term relevant for the valley-orbit interaction. It is easy to check that this term in (D) is the matrix element 𝐩|Hvo|𝐩\langle{\bf p}^{\prime}|H_{vo}|{\bf p}\rangle of the effective valley-orbit Hamiltonian

Hvo=2τv2Δ2Ur×𝐩,\displaystyle H_{vo}=-2\tau\frac{v^{2}}{\Delta^{2}}\nabla U_{r}\times\mathbf{p}, (36)

where UrU_{r} is repulsive Keldysh potential in coordinate representation. Noting that Ur×𝐩=1/rrUrl\nabla U_{r}\times\mathbf{p}=1/r\partial_{r}U_{r}l, the valley-orbital splitting between the p+p_{+} and pp_{-} states is

Δϵ=4τv2Δ20dUrdrψr2𝑑r,\Delta\epsilon=-4\tau\frac{v^{2}}{\Delta^{2}}\int_{0}^{\infty}\frac{dU_{r}}{dr}\psi_{r}^{2}dr, (37)

where the p-wave radial function in the coordinate space is normalised as

0ψr2r𝑑r=1\int_{0}^{\infty}\psi_{r}^{2}rdr=1 (38)

The wave function comes from the Fourier transform of ψp\psi_{p} from Eq. (6), ψ𝐩=ψpeilθp.\psi_{\bf p}=\psi_{p}e^{il\theta_{p}}.

ψr=0ψpJ1(pr)pdp2π.\psi_{r}=\int_{0}^{\infty}\psi_{p}J_{1}(pr)\frac{pdp}{\sqrt{2\pi}}. (39)

Substituting in the Keldysh interaction, v2=Δ/4μv^{2}=\Delta/4\mu and Δ=1.9\Delta=1.9 eV we find splitting |Δϵ|=3.8|\Delta\epsilon|=3.8 meV at B=0B=0 T. Note that the sign of this splitting is dependent on valley index τ\tau.

References