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

Valley-Filling Instability and Critical Magnetic Field for Interaction-Enhanced Zeeman Response in Doped WSe2 Monolayers

Fengyuan Xuan Centre for Advanced 2D Materials, National University of Singapore, 6 Science Drive 2, Singapore 117546    Su Ying Quek [email protected] Department of Physics, National University of Singapore, 2 Science Drive 3, Singapore 117551 Centre for Advanced 2D Materials, National University of Singapore, 6 Science Drive 2, Singapore 117546 NUS Graduate School Integrative Sciences and Engineering Programme, National University of Singapore Department of Materials Science and Engineering, National University of Singapore
Abstract

Carrier-doped transition metal dichalcogenide (TMD) monolayers are of great interest in valleytronics due to the large Zeeman response (g-factors) in these spin-valley-locked materials, arising from many-body interactions. We develop an ab initio approach based on many-body perturbation theory to compute the interaction-enhanced g-factors in carrier-doped materials. We show that the g-factors of doped WSe2 monolayers are enhanced by screened exchange interactions resulting from magnetic-field-induced changes in band occupancies. Our interaction-enhanced g-factors gg^{*} agree well with experiment. Unlike traditional valleytronic materials such as silicon, the enhancement in g-factor vanishes beyond a critical magnetic field BcB_{c} achievable in standard laboratories. We identify ranges of gg^{*} for which this change in g-factor at BcB_{c} leads to a valley-filling instability and Landau level alignment, which is important for the study of quantum phase transitions in doped TMDs. We further demonstrate how to tune the g-factors and optimize the valley-polarization for the valley Hall effect.

Valleytronics, the control and manipulation of the valley degree of freedom (valley pseudospin), is being actively considered as the next new paradigm for information processing. The field of valleytronics dates back to investigations on traditional semiconductors such as siliconOhkawa ; Sham , but the ability to exploit valley polarizations in these materials has been limited NatRevMat . A major impetus for the renaissance of valleytronics is the recent discovery that H-phase transition metal dichalcogenide (TMD) semiconductor monolayers (MLs) are excellent candidates for valleytronics applications NatRevMat ; RMP2018 . The spin-valley locking effect in these MLs Xiao2012 leads to long lifetimes for spin- and valley-polarization, while individual valleys can be probed and controlled using circularly-polarized light, paving the way to use the valley pseudospin for information processing. The valley Zeeman response in TMD MLs Xuan2020 ; Rohlfing2020 ; PRB2020 ; PRB2020 ; NC2020 ; Marie2020 ; 2018Ensslin ; LL2019 ; Mak ; 2020PRL ; MakPRL ; jump ; NanoLett ; NatMater ; LL2017 ; 2018PRBR is also significantly larger than in traditional semiconductors Roth ; Yafet ; Janak ; Fang ; Lak ; ExpPRB ; Shayegan .

When an external magnetic field is applied normal to the TMD ML, the energies of the valleys shift in equal magnitude and opposite directions. This Zeeman effect is quantified by the orbital and spin magnetic moments, which contribute to the Landé g-factors. In TMD MLs, the intrinsic Landé g-factors are about 66 times larger Xuan2020 ; Rohlfing2020 ; PRB2020 ; Marie2020 than that in silicon, where only the spin magnetic moment dominates Roth ; Yafet . The larger g-factors in TMD MLs allow for greater control in tuning the energetics of the valley pseudospins, and results in a larger valley-polarized current, which is important for observations of the valley Hall effect NatRevMat . Besides the Zeeman effect, an external magnetic field also results in a quantization of states to form Landau levels (LLs).

Much of the current research on valleytronics seeks to understand how to manipulate the valley pseudospins in TMDs NatRevMat ; RMP2018 . It has been found that carrier doping can dramatically enhance the g-factors in TMD monolayers 2020PRL ; MakPRL ; jump ; NanoLett ; NatMater ; LL2017 ; 2018PRBR , opening up the possibility to tune the valley pseudospin by gating in a magnetic field. This enhancement in g-factors has been attributed to many-body interactions 2020PRL ; MakPRL ; jump ; NanoLett ; NatMater ; LL2017 ; 2018PRBR ; Janak ; ExpPRB , but a quantitative understanding is lacking. To interpret the g-factor enhancement in doped TMDs, experimentalists have typically relied on existing theoretical literature dating back to the 1960s-1970sJanak ; Ando . However, there are two shortcomings of these theoretical approaches. Firstly, they are not ab initio methods and cannot provide quantitative predictions. Secondly, these studies all focused on silicon or III-V semiconductors, which are very different in nature from the TMD monolayers. It is important to question if the experimental observations on g-factors in doped TMDs serve only to validate in a new material what was already known for silicon, or if it is possible to observe novel phenomena not known before in conventional valleytronics materials.

In this work, we develop an ab initio approach based on many body perturbation theory to compute the interaction-enhanced Landé g-factors in carrier-doped systems. We predict that the larger intrinsic g-factors in TMD MLs enable the observation of a critical magnetic field BcB_{c} above which the interaction-induced enhancement in the g-factors vanishes in doped TMDs. We identify ranges of the enhanced g-factor genhg^{*}_{\text{enh}} for which the discontinuous change in g-factor at BcB_{c} results in a LL alignment and valley-filling instability for BBcB\gtrsim B_{c}. Such a phenomenon has not been observed or predicted for silicon and other conventional valleytronic materials. Our computed interaction-enhanced g-factors for hole-doped ML WSe2 agree well with experiment 2020PRL ; jump ; NatMater and can be tuned by dielectric screening. The predicted valley-filling instability for BBcB\gtrsim B_{c} provides theoretical insights into recent experimental observations jump of a pronounced Landau level-filling instability at a critical magnetic field which closely matches our predicted values. The associated alignment of LLs is of interestjump to investigate quantum phase transitions in these doped TMDs Braz ; Donk ; Miserev ; Roch ; Nature1999 ; Science2000 . The recent observation of fractional quantum Hall states associated with non-abelian anyons in ML WSe2 FQHWSe2 highlights the potential of creating pseudo-spinors from aligned LLs for topological quantum computing applications QCRMP .

RESULTS

Interaction-enhanced g-factors

For a many-electron system described by a static mean-field Hamiltonian H|n𝐤=En𝐤|n𝐤H\ket{n\mathbf{k}}=E_{n\mathbf{k}}\ket{n\mathbf{k}}, it has been shown that an out-of-plane magnetic field BB results in the following expression for the LLs at the KK valley Xuan2020 :

ϵN,K=En𝐊+(N+12)2mμBBgn𝐊IμBB,\displaystyle\epsilon_{N,K}=E_{n\mathbf{K}}+(N+\frac{1}{2})\frac{2}{m^{*}}\mu_{B}B-g^{I}_{n\mathbf{K}}\mu_{B}B, (1)

where nn is the corresponding band index, En𝐊E_{n\mathbf{K}} is the mean-field single-particle energy at KK, mm^{*} is the valley effective mass, μB\mu_{B} is the Bohr magneton and N=0,1,2,N=0,1,2,... is the LL index. The total intrinsic single band g-factor consists of the orbital and spin contribution, gn𝐊I=gn𝐊orbgssz,n𝐊g^{I}_{n\mathbf{K}}=g^{\text{orb}}_{n\mathbf{K}}-g_{s}s_{z,n\mathbf{K}}, where gsg_{s} is taken to be 2.02.0 and sz,n𝐊s_{z,n\mathbf{K}} is the spin quantum number. The orbital g-factor is defined using the orbital magnetic moment gn𝐊orbμB=mn𝐊zg^{\text{orb}}_{n\mathbf{K}}\mu_{B}=m^{z}_{n\mathbf{K}} Xuan2020 (see Methods).

However, the above static mean-field description does not account for the energy-dependent electron self-energy Σ(E)\Sigma(E) that is necessary for a many-body description of the quasiparticle (QP) energies. Σ(E)\Sigma(E) represents the change in the energy of the bare particle due to the interaction of the particle with itself via the interacting many-body system. The change in self-energy as the QP energy shifts with BB leads to an effective renormalized g-factor, gn𝐊g^{*}_{n\mathbf{K}}, defined as (En𝐤QP,1En𝐤QP)=gn𝐊μBB(E_{n\mathbf{k}}^{\text{QP},1}-E_{n\mathbf{k}}^{\text{QP}})=g^{*}_{n\mathbf{K}}\mu_{B}B, where

En𝐤QP=En𝐤+Σ(En𝐤QP)En𝐤QP,1=En𝐤+Σ(En𝐤QP,1)+gn𝐊IμBB,\displaystyle\begin{split}E_{n\mathbf{k}}^{\text{QP}}=E_{n\mathbf{k}}&+\Sigma(E_{n\mathbf{k}}^{\text{QP}})\\ E_{n\mathbf{k}}^{\text{QP},1}=E_{n\mathbf{k}}&+\Sigma(E_{n\mathbf{k}}^{\text{QP},1})+g^{I}_{n\mathbf{K}}\mu_{B}B,\end{split} (2)

Thus we have,

gn𝐊Ign𝐊=1dΣ(E)dE.\displaystyle\frac{g^{I}_{n\mathbf{K}}}{g^{*}_{n\mathbf{K}}}=1-\frac{d\Sigma(E)}{dE}. (3)

Such a renormalization effect is missing in previous first principles calculations of g-factors in TMDs Xuan2020 ; Rohlfing2020 ; PRB2020 .

Table 1: GWGW single band and exciton g-factors of undoped ML WSe2 at the KK valley. X0X0 and D0D0 refer to the lowest energy spin-allowed and spin-forbidden optical transitions Xuan2020 ; Rohlfing2020 ; PRB2020 . For the single-band g-factors, cc and vv refer to the frontier conduction and valence bands, respectively, while \uparrow and \downarrow refer to spin up and spin down bands at KK.
Renormalized (gg^{*}) Intrinsic (gIg^{I})
gcg_{c\uparrow} -4.36 -5.50
gcg_{c\downarrow} -2.05 -2.59
gvg_{v\uparrow} -6.63 -8.31
gvg_{v\downarrow} -4.48 -5.62
gX0g_{X0} -4.54
gD0g_{D0} -9.16

The electron self-energy in this work is computed within the GWGW approximation Hedin (see Methods), which uses the first order term in the perturbative expansion of Σ\Sigma in terms of the screened Couloumb interaction WW. Table 1 shows the renormalized and intrinsic g-factors computed for undoped monolayer WSe2. We see that the magnitudes of the renormalized g-factors are reduced by 20\sim 20% compared to the intrinsic GWGW g-factors, because Σ(E)E\frac{\partial\Sigma(E)}{\partial E} is in general negative Louie . The exciton g-factors deduced using the renormalized g-factors are in good agreement with experiment 2d2015 ; NC2019 ; NatPhys2015 ; 2d2019 ; Liu2019 .

Refer to caption
Fig. 1: Valence band g-factor in hole-doped ML WSe2. a Schematic figure of the energy dispersion of hole-doped WSe2 ML in the presence of an out-of-plane magnetic field, in the mixed polarized regime where the g-factor is enhanced. b Valence band g-factor |g||g^{*}| in hole-doped ML WSe2 as a function of hole density. Red squares: Calculated results; Dark blue and light blue circles: Experimental data from Ref. jump, , NatMater, , respectively; Purple and green dotted lines: Experimental data from Ref. Marie2020, , 2020PRL, , respectively (hole densities are given in a range only).

In contrast to the undoped system, doped systems have partially occupied bands. Thus, if the band occupancies are also changing in response to the magnetic field, there is an additional term in dΣ(E)dE\frac{d\Sigma(E)}{dE}:

dΣ(E)dE=Σ(E)E+Σ(E)ffE\displaystyle\frac{d\Sigma(E)}{dE}=\frac{\partial\Sigma(E)}{\partial E}+\frac{\partial\Sigma(E)}{\partial f}\frac{\partial f}{\partial E} (4)

where ff is the Fermi-Dirac distribution function. The second term in Eq (4) can be simplified to give (see Methods):

Σ(E)ffE|m|2πW¯nkF(E=EF),\displaystyle\begin{split}\frac{\partial\Sigma(E)}{\partial f}\frac{\partial f}{\partial E}\approx\frac{\lvert m^{*}\rvert}{2\pi}\bar{W}_{nk_{F}}(E=E_{F}),\end{split} (5)

where nn is the band index of the frontier doped band, and EFE_{F} and kFk_{F} are respectively the Fermi energy and Fermi wave vector. This term comes from the screened exchange contribution to the self-energy. W¯nkF\bar{W}_{nk_{F}} (defined in Methods) is an effective quasi-2D screened Coulomb potential which can be evaluated completely from first principles. Since W¯nkF\bar{W}_{nk_{F}} is positive, the second term of Eq. (4) leads to an enhancement effect for the g-factor. For an ideal 2D fermion gas, this second term reduces to the term dΣ(E)/dEd\Sigma(E)/dE derived in Ref. Janak, (see also Ref. Quinn, ) where the first term in Eq. (4) is ignored. We note that our numerical results for W¯vkF\bar{W}_{vk_{F}} as a function of hole density differ from those computed for an ideal 2D fermion gasstern ; Janak , although the corresponding numerical results for the bare Coulomb potential match well with the ideal 2D case (see Supplementary Figure 1). This observation implies that an ab initio non-local description of the dielectric function of the quasi-2D system is important for a quantitative prediction of the renormalized g-factors.

The g-factors in this work are all computed for the valence band at KK, and henceforth, the subscript v𝐊v\mathbf{K} is omitted. The magnitudes of our computed valence band g-factors |g||g^{*}| are plotted as red squares in Fig. 1b for ML WSe2 with different hole densities. Our predicted g-factors agree well with those deduced from multiple experiments jump ; 2020PRL ; Marie2020 ; NatMater on hole-doped ML WSe2. The renormalized g-factor for the undoped system is labeled g0g^{*}_{0}. Due to interaction-induced enhancement, the g-factor increases significantly once hole carriers are introduced. This enhancement reduces as the hole density is increased as expected from the density dependence of the many-body Coulomb interactions (see Supplementary Figure 1).

Critical magnetic field

Refer to caption
Fig. 2: Magnetic field dependence of g-factor. a Schematic figure illustrating the Zeeman effect as BB is increased with fixed hole density. The pink and blue shaded regions illustrate the magnitude of hole density due to the constant density of states in the quadratic band in two dimensions. As BB increases for BBcB\leq B_{c}, the absolute Fermi level position is unchanged, but the band occupancies change. For B>BcB>B_{c}, the Fermi level shifts with the band at KK and the band occupancies remain unchanged. b-c Plots of b |g||g^{*}| and c valley Zeeman split EZE_{Z} (labeled in a) as a function of magnetic field BB for monolayer WSe2 with hole density 5.2×1012cm25.2\times 10^{12}\text{cm}^{-2}.

A subtle but important point is that Eq. (5) applies only when the band occupancies are changing with BB, and the Fermi level EFE_{F} is fixed. In electrostatic gating experiments, the carrier concentration is fixed rather than the absolute Fermi level. However, the Zeeman shifts in KK and KK^{\prime} are equal in magnitude but opposite in sign (Fig. 2a, B<BcB<B_{c}), and the density of states for the quadratic bands in 2D is independent of energy. So for BB small enough that both valleys have carriers (mixed polarized regime; Fig. 1a), EFE_{F} is fixed while the band occupancies change and both terms in Eq. (4) will apply, leading to the interaction-enhanced g-factors, which we label as genhg^{*}_{\text{enh}}. However, above a critical magnetic field Bc|EF|/(|genh|μB)B_{c}\sim|E_{F}|/(|g^{*}_{\text{enh}}|\mu_{B}), only one valley has carriers (Fig. 2a, B>BcB>B_{c}) (see Supplementary Note for a more precise expression for BcB_{c}). As BB increases beyond BcB_{c}, a constant hole density is maintained when EFE_{F} shifts with the bands without changing the band occupancies. Thus, for B>BcB>B_{c}, only the first term in Eq. (4) applies, similar to the undoped case, leading to an abrupt drop in gg^{*} at B=BcB=B_{c} (Fig. 2b), with a corresponding piecewise-linear Zeeman split EZE_{Z} (Fig. 2c). For a hole density of 5.2×1012cm25.2\times 10^{12}\text{cm}^{-2}, Bc36B_{c}\approx 36 T, and we have g11.0g^{*}\approx-11.0 for B<BcB<B_{c} and g6.6g^{*}\approx-6.6 for B>BcB>B_{c}.

This abrupt drop in gg^{*} at a critical magnetic field has never been reported or predicted before in traditional valleytronic materials such as silicon. Indeed, BcB_{c} is inversely related to |genh||g^{*}_{\text{enh}}|, and it is the large intrinsic g-factors and hence large renormalized g-factors for TMDs that allow for BcB_{c} to be small enough to be reached in standard laboratories. For the same hole density of 5.2×1012cm25.2\times 10^{12}\text{cm}^{-2}, we predict BcB_{c} in silicon to be 400\sim 400 T. The larger intrinsic g-factors for TMD MLs arise from the large orbital g-factors, which consist of a valley term, an orbital term, and a cross term that involves coupling between the phase-winding of the Bloch states and the parent atomic orbitals. Xuan2020

Refer to caption
Fig. 3: Critical magnetic field and Zeeman split. (a) Critical magnetic field as a function of hole density. Blue circles: Experimentally derived from the onset of the fully-polarized regime (Ref. jump, ). Red squares: Predicted in this work. In the yellow-shaded area, the renormalized valence band g-factor is the value of gg^{*} in the undoped TMD ML, g0g^{*}_{0}. In the cyan-shaded area, the valence band g-factor is enhanced by interactions. (b) 2D plot for EZE_{Z} as a function of hole density and external magnetic field.

Since BcB_{c} is the value of BB characterizing the onset of the fully polarized regime, BcB_{c} can be deduced using optical measurements of the exciton and polaron energies for KK and KK^{\prime}jump . In Fig. 3a, we plot these values of BcB_{c} (blue circles) and compare them with our predicted values (red squares). The predicted dependence of BcB_{c} on the hole density agrees well with experiment.

How can one maximize the concentration of valley-(and spin-)polarized carriers in the TMD ML? As the hole concentration increases, BcB_{c} increases, giving a larger range of BB for which gg^{*} is enhanced by interactions (Fig. 3a). However, the magnitude of genhg^{*}_{\text{enh}} decreases as the hole concentration increases (Fig. 1). These competing effects imply that for any given BB field B0B_{0}, there is an optimal hole concentration ρ0\rho_{0} which maximizes the Zeeman split EZE_{Z} (Fig. 3b). This optimal hole concentration ρ0\rho_{0} corresponds to the hole concentration for which Bc=B0B_{c}=B_{0} (Fig. 3), and yields a maximum concentration of valley-polarized carriers. These predictions are useful for realizations of the valley Hall effect and other applications where a high concentration of valley-polarized carriers is desired.

LL Alignment and Valley Filling Instability

Refer to caption
Fig. 4: LL fan diagram and valley-filling instability. a LL fan diagram for ML WSe2 valence band with a hole density of 5.2×1012cm25.2\times 10^{12}\text{cm}^{-2}. The LLs are labeled by the LL index NN and valley KK or KK^{\prime}. The slopes of the plots decrease in magnitude when BB increases beyond BcB_{c}, leading to a crossing between 0K0K^{\prime} and 5K5K (purple circle). b Schematic figure for LLs at B1B_{1} and B2B_{2} as marked in (d). The blue and pink shading represent the hole populations in KK and KK^{\prime}, respectively. c Hole occupancies of LLs for a constant hole density of 5.2×1012cm25.2\times 10^{12}\text{cm}^{-2}. The blue/pink shading refer to hole occupations in the nearest LLs, as demarcated by the black/red dashed lines. Each LL is fully occupied before the next LL lower in energy, resulting in the zigzag-shaped fine-structure at EEFE\sim E_{F}. d Zoom-in figure of (c) showing the valley-filling instability as indicated by the isolated pink triangle. The hole population in 5K5K is transferred to 0K0K^{\prime} for B>BXB>B_{X}, until BB is large enough (yellow circle) that LLs 0K0K to 4K4K carry all the holes. e Range of interaction-enhanced |genh||g^{*}_{\text{enh}}| for which a valley-filling instability exists, with corresponding ΔB\Delta B.

The abrupt change in gg^{*} at B=BcB=B_{c} also has other interesting implications. As BB increases beyond BcB_{c}, the decrease in |g||g^{*}| results in a decrease in magnitude of the slopes of the LL fan diagrams (Fig. 4a), leading to a crossing between the energies of 0K0K^{\prime} and NKNK (purple circle, N=5N=5 in Fig. 4a). If NKNK has carriers, this LL alignment results in a valley-filling instability, where the hole population is transferred back and forth between the two valleys for small changes in BB.

LLs are filled with holes from higher energy to lower energy. In Fig. 4c, the blue and pink shading indicate the filling of the LLs with holes. If NKNK is fully occupied, the plot is shaded from ϵNK+ωc2\epsilon_{NK}+\frac{\hbar\omega_{c}}{2} down to ϵNKωc2\epsilon_{NK}-\frac{\hbar\omega_{c}}{2}. At B=BcB=B_{c}, an integer number of LLs (55 in Fig. 4c) are completely filled with holes while 0K0K^{\prime} is completely empty. As BB decreases slightly below BcB_{c}, the LL degeneracy decreases, and the LL with the next lower energy (0K0K^{\prime} in Fig. 4c) is required to contain the holes, leading to a symmetric zigzag fine structure about EFE_{F}. As BB increases slightly above BcB_{c}, the LL degeneracy increases. As long as ϵ5K>ϵ0K\epsilon_{5K}>\epsilon_{0K^{\prime}}, only the KK valley is filled with holes (B=B1B=B_{1} in Fig. 4b), and the blue shaded area represents the constant hole density. But when BB is slightly larger than BXB_{X} (Fig. 4d) where the energies of 0K0K^{\prime} and 5K5K cross, holes will start to fill the KK^{\prime} valley again (B=B2B=B_{2} in Fig. 4b), until the BB field is large enough that the LLs 0K0K to 4K4K can contain all the holes (yellow circle, Fig. 4d) and the system becomes fully polarized again. This represents a valley-filling instability, where KK^{\prime} is depleted of holes from B=BcB=B_{c} to B=BXB=B_{X}, and filled again up to BX+ΔBB_{X}+\Delta B. In practice, when holes begin to fill 0K0K^{\prime}, the mixed polarized regime is reached and gg^{*} becomes enhanced, leading to a change in the slope of the fan diagram that is expected to result in a LL alignment not just for B=BXB=B_{X} but also for BB up to BX+ΔBB_{X}+\Delta B.

Our predictions provide important theoretical insights into a recent experiment on doped ML WSe2jump , where optical absorption plots showed a pronounced signature of the peak positions changing from one inter-LL transition to another over a small range of BB close to the onset of the fully-polarized regime in the experiment. This is consistent with the highest occupied LL in the KK^{\prime} valley being emptied and partially filled with holes at BBcB\sim B_{c} in our predictions. The authors of Ref. jump, attributed this observation to the oscillatory g-factors predicted for traditional semiconductors such as silicon Ando . However, in this theory, the changes in the g-factors are directly related to the position of the Fermi level relative to the LLs, and the g-factors therefore have an “oscillatory” Ando dependence on BB rather than a pronounced change at one particular value of BB as seen in the experiment. Furthermore, such a pronounced instability was not observed in experiments Fang ; Lak ; ExpPRB ; Shayegan on the g-factors in doped silicon and other traditional valleytronic materials for which these oscillatory g-factors were predicted. Thus, this pronounced instability observed in doped ML WSe2jump is in fact a manifestation of the valley-filling instabilities that are predicted here to emerge specifically for doped TMDs. Our conclusion is further supported by the fact that the measured values of BcB_{c} and BXB_{X} are respectively 3232 and 3838 T for |genh|11|g^{*}_{\text{enh}}|\sim 11jump , close to our predicted values of 3131 and 3737 T for the same |genh||g^{*}_{\text{enh}}| (see also Supplementary Table 1).

The alignment of LLs is of interest to investigate quantum phase transitions in these doped TMDs jump ; Braz ; Donk ; Miserev ; Roch ; Nature1999 ; Science2000 . Given that the LLs are expected to align for BB between BXB_{X} and BX+ΔBB_{X}+\Delta B, it is interesting to predict how large ΔB\Delta B can be and how sensitive ΔB\Delta B is to fluctuations in genhg^{*}_{\text{enh}}. Not all values of genhg^{*}_{\text{enh}} will result in the instability (see Supplementary Figure 2 and Supplementary Note). In particular, if the energies of 0K0K^{\prime} and NKNK cross at BXB_{X}, the valley-filling instability only occurs if NKNK is occupied with holes. Fig. 4e plots the ranges of genhg^{*}_{\text{enh}} for which a valley-filling instability will occur, as well as the corresponding ΔB\Delta B values. We see that an optimal range of |genh||g^{*}_{\text{enh}}| for LL alignment is 10.4<|genh|<11.210.4<|g^{*}_{\text{enh}}|<11.2. Here, ΔB\Delta B is quite large (8\sim 8T for |genh|10.4|g^{*}_{\text{enh}}|\sim 10.4) and is also fairly robust to changes in genhg^{*}_{\text{enh}}. The corresponding values of BcB_{c} and BXB_{X} fall within 3030 to 4040 T (Supplementary Table 1), well within the reach of experiments.

The alignment of LLs in different valleys can in principle be achieved for B<BcB<B_{c} if genhg^{*}_{\text{enh}} can be tuned such that 0K0K^{\prime} matches exactly with NKNK for some NN. However, once genhg^{*}_{\text{enh}} deviates slightly from this value, due to fluctuations in the hole density or dielectric environment (see Fig. 5), the LLs are no longer aligned. Our predictions above enable the alignment of LLs while allowing for some fluctuations in genhg^{*}_{\text{enh}}.

Refer to caption
Fig. 5: Tunability of |genh||g^{*}_{\text{enh}}| using dielectric screening. Ab initio valence band g-factor |genh||g^{*}_{\text{enh}}| as a function of background dielectric constant ϵmed\epsilon^{\text{med}} for a hole density of 5.18×1012cm25.18\times 10^{12}\text{cm}^{-2}.

Tunability of interaction-enhanced g-factor

We further note that in addition to electrostatic gating which changes the carrier concentration and thus genhg^{*}_{\text{enh}} (Fig. 1), genhg^{*}_{\text{enh}} can also be tuned by dielectric screening (Fig. 5). The tunability of genhg^{*}_{\text{enh}} with the background dielectric constant can be understood from the fact that genhg^{*}_{\text{enh}} is related to the effective quasi-2D screened Coulomb potential at the Fermi surface (Eq. 5). This tunability of genhg^{*}_{\text{enh}} provides a handle to control the valley-polarized current, BcB_{c} and ΔB\Delta B.

DISCUSSION

In summary, our ab initio calculations show that many-body interactions in doped TMD MLs enhance the g-factors compared to the undoped MLs, up to a critical magnetic field BcB_{c} above which the g-factors revert to those in the undoped systems. Such a phenomenon has not been predicted or observed in silicon and other traditional valleytronic materials, because the corresponding BcB_{c} would be much larger due to the smaller g-factors in these materials.

The enhancement in g-factors arises from the effect of a magnetic-field-induced change in occupancies on the screened exchange interactions. This effect is only present in the mixed-polarized regime (B<BcB<B_{c}). As the carrier concentration increases, gg^{*} decreases and BcB_{c} increases, so that for any value of the magnetic field B0B_{0}, the valley-polarization is maximized when the carrier concentration is such that Bc=B0B_{c}=B_{0}. This prediction has implications for maximizing the valley- and spin-polarized current for the valley Hall effect.

The computed interaction-enhanced g-factors agree well with experiment for different doping concentrations. We further identify the values of genhg^{*}_{\text{enh}} and corresponding ranges of BB that lead to a valley-filling instability and expected LL alignment, which are of interest for the investigation of quantum phase transitions in doped TMDs Braz ; Donk ; Miserev ; Roch ; Nature1999 ; Science2000 . Recent observations of fractional quantum Hall states associated with non-abelian anyons in ML WSe2 FQHWSe2 suggest that the creation of pseudo-spinors from a linear combination of valley-aligned LLs can be useful for topological quantum computing applications QCRMP .

METHODS

Calculation of intrinsic g-factor

The orbital component of the intrinsic g-factor gn𝐊orbg^{\text{orb}}_{n\mathbf{K}} is defined as gn𝐊orbμB=mn𝐊zg^{\text{orb}}_{n\mathbf{K}}\mu_{B}=m^{z}_{n\mathbf{K}} Xuan2020 :

𝐦n𝐊=ie2𝐤un𝐤|×[H𝐤En𝐤]|𝐤un𝐤|𝐤=𝐊.\displaystyle\mathbf{m}_{n\mathbf{K}}=-\frac{ie}{2\hbar}\expectationvalue{\times[H_{\mathbf{k}}-E_{n\mathbf{k}}]}{\partial_{\mathbf{k}}u_{n\mathbf{k}}}|_{\mathbf{k}=\mathbf{K}}. (6)

We use the PBE exchange-correlation functional PBE1996 for the DFT mean-field calculations QE and the details follow those in Ref. Xuan2020, . For GWGW calculations of the intrinsic g-factors, we use a non-uniform sampling method sub of the Brillouin Zone starting with a 12×1212\times 12 k-grid as implemented in the BerkeleyGW code BGW . The energy-dependence of the dielectric function is treated within the generalized plasmon pole (GPP) model Louie . An energy cutoff of 3535 Ry with 40004000 empty bands is used for the reciprocal space expansion of the dielectric matrix. The intrinsic single-band g-factor reduces by only 0.20.2 when an energy cutoff of 22Ry with 200200 empty bands is used.

Calculation of renormalized g-factor

The renormalized g-factor gg^{*} is computed from the intrinsic g-factor gIg^{I} and dΣ(E)dE\frac{d\Sigma(E)}{dE} using Eq. 3. As the dependence of En𝐤QPE_{n\mathbf{k}}^{\text{QP}} on BB is no longer linear, we generalize Eq. (2) to the case where En𝐤QPE_{n\mathbf{k}}^{\text{QP}} refers to the quasiparticle energies in the presence of a BB-field, and the applied BB represents a small increment in BB.

We approximate gIg^{I} using the value in the undoped system. The intrinsic single band g-factors from DFT calculations do not change when ML WSe2 is doped with holes.

The first term of Eq. 4 can be obtained directly from the BerkeleyGW outputBGW .

Σ=iGW\Sigma=iGW can be partitioned Louie into the dynamical non-local screened-exchange (SEX) and Coulomb-hole (COH) interaction terms Σ=ΣSEX+ΣCOH\Sigma=\Sigma^{\text{SEX}}+\Sigma^{\text{COH}}. Only the screened exchange term depends on the occupancies ff and contributes to the second term of Eq. 4. The screened exchange energy ΣSEX\Sigma^{\text{SEX}} in our ab initio plane-wave calculation can be written as (see Supplemental Note):

Σn𝐊SEX(E)=m1N𝐪Ω𝐪𝐆𝐆fm𝐊𝐪n𝐊|ei(𝐪+𝐆)𝐫|m𝐊𝐪m𝐊𝐪|ei(𝐪+𝐆)𝐫|n𝐊×ϵ𝐆𝐆1(𝐪,EEm𝐊𝐪)v𝐪+𝐆=m1(2π)2BZd2qW¯m𝐪fm𝐊𝐪,\displaystyle\begin{split}\Sigma_{n\mathbf{K}}^{\text{SEX}}(E)&=-\sum_{m}\frac{1}{N_{\mathbf{q}}\Omega}\sum_{\mathbf{qGG^{\prime}}}f_{m\mathbf{K}-\mathbf{q}}\bra{n\mathbf{K}}e^{i(\mathbf{q}+\mathbf{G})\cdot\mathbf{r}}\ket{m\mathbf{K}-\mathbf{q}}\bra{m\mathbf{K}-\mathbf{q}}e^{-i(\mathbf{q}+\mathbf{G}^{\prime})\cdot\mathbf{r}}\ket{n\mathbf{K}}\\ &\times\epsilon^{-1}_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{q},E-E_{m\mathbf{K}-\mathbf{q}})v_{\mathbf{q}+\mathbf{G}}\\ &=-\sum_{m}\frac{1}{(2\pi)^{2}}\int_{BZ}d^{2}q\bar{W}_{m\mathbf{q}}f_{m\mathbf{K}-\mathbf{q}},\end{split} (7)

where we define the quasi-2D screened Coulomb potential:

W¯m𝐪(E)=1L𝐆𝐆n𝐊|ei(𝐪+𝐆)𝐫|m𝐊𝐪m𝐊𝐪|ei(𝐪+𝐆)𝐫|n𝐊×ϵ𝐆𝐆1(𝐪,EEm𝐊𝐪)v𝐪+𝐆.\displaystyle\begin{split}\bar{W}_{m\mathbf{q}}(E)&=\frac{1}{L}\sum_{\mathbf{GG^{\prime}}}\bra{n\mathbf{K}}e^{i(\mathbf{q}+\mathbf{G})\cdot\mathbf{r}}\ket{m\mathbf{K}-\mathbf{q}}\bra{m\mathbf{K}-\mathbf{q}}e^{-i(\mathbf{q}+\mathbf{G}^{\prime})\cdot\mathbf{r}}\ket{n\mathbf{K}}\\ &\times\epsilon^{-1}_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{q},E-E_{m\mathbf{K}-\mathbf{q}})v_{\mathbf{q}+\mathbf{G}}.\end{split} (8)

Here, Ω\Omega is the cell volume, N𝐪N_{\mathbf{q}} is the number of q-points and v𝐪v_{\mathbf{q}} is the Coulomb potential with the slab Coulomb truncation scheme applied trunc . W¯m𝐪\bar{W}_{m\mathbf{q}} is an effective quasi-2D screened Coulomb potential defined in valley KK and LL is the height of the supercell for ML WSe2. The second term in Eq. (4) is then given by (see Supplemental Note):

Σ(E)ffE=m1(2π)2BZd2qW¯m𝐪fm𝐊𝐪E|m|2πW¯nkF(E=EF).\displaystyle\begin{split}\frac{\partial\Sigma(E)}{\partial f}\frac{\partial f}{\partial E}&=-\sum_{m}\frac{1}{(2\pi)^{2}}\int_{BZ}d^{2}q\bar{W}_{m\mathbf{q}}\frac{\partial f_{m\mathbf{K}-\mathbf{q}}}{\partial E}\approx\frac{|m^{*}|}{2\pi}\bar{W}_{nk_{F}}(E=E_{F}).\end{split} (9)

We compute gv𝐊g^{*}_{v\mathbf{K}} by evaluating W¯vkF\bar{W}_{vk_{F}} ab initio using the random phase approximation for the dielectric matrix. Due to the partial occupancies, we calculate the dielectric matrix using a dense reciprocal space sampling of 120×120120\times 120, a 22Ry GG-vector cut off and 2929 bands. gg^{*} is unchanged when we use instead 44Ry and 299299 bands, and reduces by 3%\sim 3\% when a 240×240240\times 240 k-mesh is used. Care is taken to include the effect of spin-orbit splitting at the valleys. For the effective mass, we use our DFT value of m=0.48mem^{*}=-0.48m_{e}, which agrees well with the experimentally deduced value for hole-doped WSe2 LL2016 . If electronic screening is ignored, the effective quasi-2D bare Coulomb potential V¯mq\bar{V}_{mq} is defined by:

V¯mq(E)=1L𝐆𝐆n𝐤|ei(𝐪+𝐆)𝐫|m𝐤𝐪m𝐤𝐪|ei(𝐪+𝐆)𝐫|n𝐤δ𝐆𝐆v𝐪+𝐆.\displaystyle\bar{V}_{mq}(E)=\frac{1}{L}\sum_{\mathbf{GG^{\prime}}}\bra{n\mathbf{k}}e^{i(\mathbf{q}+\mathbf{G})\cdot\mathbf{r}}\ket{m\mathbf{k}-\mathbf{q}}\bra{m\mathbf{k}-\mathbf{q}}e^{-i(\mathbf{q}+\mathbf{G}^{\prime})\cdot\mathbf{r}}\ket{n\mathbf{k}}\delta_{\mathbf{G}\mathbf{G}^{\prime}}v_{\mathbf{q}+\mathbf{G}}. (10)

Our first principles results for V¯vkF\bar{V}_{vk_{F}} (Supplementary Figure 1) agrees with the analytically-derived 2D Coulomb potential. At low doping densities, V¯vkF\bar{V}_{vk_{F}} is very large, which would change the sign of gg^{*} compared to the intrinsic g-factor, indicating that screening is important for a meaningful description of gg^{*}.

Background dielectric constant

A uniform background dielectric constant (ϵmed\epsilon^{\text{med}}) can be simply added to the dielectric function of the system to obtain the total dielectric function: ϵ(𝐫,𝐫,ω)=ϵWSe2(𝐫,𝐫,ω)+ϵmed1\epsilon(\mathbf{r},\mathbf{r}^{\prime},\omega)=\epsilon^{\text{WSe}_{2}}(\mathbf{r},\mathbf{r}^{\prime},\omega)+\epsilon^{\text{med}}-1. In our first principles calculation, the dielectric function is expanded in a plane wave basis: ϵ(𝐫,𝐫,ω)=𝐪𝐆𝐆ei(𝐪+𝐆)𝐫ϵ𝐆𝐆(𝐪,ω)ei(𝐪+𝐆)𝐫\epsilon(\mathbf{r},\mathbf{r}^{\prime},\omega)=\sum_{\mathbf{qGG^{\prime}}}e^{i(\mathbf{q}+\mathbf{G})\cdot\mathbf{r}}\epsilon_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{q},\omega)e^{-i(\mathbf{q}+\mathbf{G}^{\prime})\cdot\mathbf{r}^{\prime}}. Thus we approximate the effect of screening by a dielectric medium by modifying the static dielectric matrix as follows:

ϵ𝐆𝐆(𝐪,0)=ϵ𝐆𝐆WSe2+(ϵmed1)δ𝐆𝐆.\displaystyle\epsilon_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{q},0)=\epsilon^{\text{WSe}_{2}}_{\mathbf{G}\mathbf{G}^{\prime}}+(\epsilon^{\text{med}}-1)\delta_{\mathbf{G}\mathbf{G}^{\prime}}. (11)

ACKNOWLEDGEMENTS

This work is supported by the NUS Provost’s Office, the Ministry of Education (MOE 2017-T2-2-139) and the National Research Foundation (NRF), Singapore, under the NRF medium-sized centre programme. Calculations were performed on the computational cluster in the Centre for Advanced 2D Materials and the National Supercomputing Centre, Singapore.

References

  • (1) Ohkawa, F. J. & Uemura, Y. Theory of valley splitting in an NN‐channel (100) inversion layer of Si I. Formulation by extended zone effective mass theory. J. Phys. Soc. Jpn 43, 907 (1977).
  • (2) Sham, L. J., Allen Jr, S. J., Kamgar, A. & Tsui, D. C. Valley–valley splitting in inversion layers on a high-index surface of silicon. Phys. Rev. Lett. 40, 472 (1978).
  • (3) Schaibley, J. R. et al. Valleytronics in 2D materials. Nat. Rev. Mater. 1, 16055 (2016).
  • (4) Wang, G. et al. Colloquium: Excitons in atomically thin transition metal dichalcogenides. Rev. Mod. Phys. 90, 021001 (2018).
  • (5) Xiao, D., Liu, G.-B., Feng, W., Xu, X. & Yao, W. Coupled spin and valley physics in monolayers of MoS2 and other group-VI dichalcogenides. Phys. Rev. lett. 108, 196802 (2012).
  • (6) Xuan, F. & Quek, S. Y. Valley Zeeman effect and Landau levels in two-dimensional transition metal dichalcogenides. Phys. Rev. Res. 2, 033256 (2020).
  • (7) Deilmann, T., Krüger, P. & Rohlfing, M. Ab Initio studies of exciton g factors: monolayer transition metal dichalcogenides in magnetic fields. Phys. Rev. Lett. 124, 226402 (2020).
  • (8) Woźniak, T., Faria Junior, P. E., Seifert, G., Chaves, A. & Kunstmann, J. Exciton gg factors of van der Waals heterostructures from first-principles calculations. Phys. Rev. B 101, 235408 (2020).
  • (9) Förste, J. et al. Exciton gg-factors in monolayer and bilayer WSe2 from experiment and theory. Nat. Commun. 11, 4539 (2020).
  • (10) Robert, C. et al. Measurement of conduction and valence bands g-factors in a transition metal dichalcogenide monolayer. Phys. Rev. Lett. 126, 067403 (2021).
  • (11) Pisoni, R. et al. Interactions and magnetotransport through spin-valley coupled Landau levels in monolayer MoS2. Phys. Rev. Lett. 121, 247701 (2018).
  • (12) Smoleński, T. et al. Interaction-induced Shubnikov–de Haas oscillations in optical conductivity of monolayer MoSe2. Phys. Rev. Lett. 123, 097403 (2019).
  • (13) Wang, Z., Shan, J. & Mak, K. F. Valley- and spin-polarized Landau levels in monolayer WSe2. Nat. Nanotechnol. 12, 144 (2017).
  • (14) Liu, E. et al. Landau-quantized excitonic absorption and luminescence in a monolayer valley semiconductor. Phys. Rev. Lett. 124, 097401 (2020).
  • (15) Wang, Z., Mak, K. F. & Shan, J. Strongly interaction-enhanced valley magnetic response in monolayer WSe2. Phys. Rev. Lett. 120, 066402 (2018).
  • (16) Li, J. et al. Spontaneous valley polarization of interacting carriers in a monolayer semiconductor. Phys. Rev. Lett. 125, 147602 (2020).
  • (17) Lin, J. et al. Determining interaction enhanced valley susceptibility in spin-valley-locked MoS2. Nano Lett. 19, 1736 (2019).
  • (18) Gustafsson, M. V. et al. Ambipolar Landau levels and strong band-selective carrier interactions in monolayer WSe2. Nat. Mater. 17, 411 (2018).
  • (19) Movva, H. C. P. et al. Density-dependent quantum Hall states and Zeeman splitting in monolayer and bilayer WSe2. Phys. Rev. Lett. 118, 247701 (2017).
  • (20) Larentis, S. et al. Large effective mass and interaction-enhanced Zeeman splitting of KK-valley electrons in MoSe2. Phys. Rev. B 97, 201407(R) (2018).
  • (21) Roth, L. M. gg factor and donor spin-lattice relaxation for electrons in germanium and silicon. Phys. Rev. 118, 1534 (1960).
  • (22) Yafet, Y. gg factors and spin-lattice relaxation of conduction electrons. Solid State Phys. 14, 1-98 (1963).
  • (23) Janak, J. F. gg factor of the two-dimensional interacting electron gas. Phys. Rev. 178, 1416 (1969).
  • (24) Fang, F. F. & Stiles, P. J. Effects of a tilted magnetic field on a two-dimensional electron gas. Phys. Rev. 174, 823 (1968).
  • (25) Lakhani, A. A. & Stiles, P. J. Experimental study of oscillatory values of gg^{*} of a two-dimensional electron gas. Phys. Rev. Lett. 31, 25 (1973).
  • (26) Nicholas, R. J., Haug, R. J., Klitzing, K. v. & Weimann, G. Exchange enhancement of the spin splitting in a GaAs-GaxAl1-xAs heterojunction. Phys. Rev. B 37, 1294 (1988).
  • (27) Papadakis, S. J., De Poortere, E. P. & Shayegan, M. Anomalous spin splitting of two-dimensional electrons in an AlAs quantum well. Phys. Rev. B 59, R12743 (1999).
  • (28) Ando, T. & Uemura, Y. Theory of oscillatory g factor in an MOS inversion layer under strong magnetic fields. J. Phys. Soc. Japan 37, 1044 (1974).
  • (29) Braz, J.E.H., Amorim, B. & Castro, E. V. Valley-polarized magnetic state in hole-doped monolayers of transition-metal dichalcogenides. Phys. Rev. B 98, 161406 (2018).
  • (30) van der Donk, M. & Peeters, F. M. Rich many-body phase diagram of electrons and holes in doped monolayer transition metal dichalcogenides. Phys. Rev. B 98, 115432 (2018).
  • (31) Miserev, D., Klinovaja, J., & Loss, D. Exchange intervalley scattering and magnetic phase diagram of transition metal dichalcogenide monolayers. Phys. Rev. B 100, 014428 (2019).
  • (32) Roch, J. G. et al. First-order magnetic phase transition of mobile electrons in monolayer MoS2. Phys. Rev. Lett. 124, 187602 (2020).
  • (33) Piazza, V. et al. First-order phase transitions in a quantum Hall ferromagnet. Nature 402, 638 (1999).
  • (34) De Poortere, E. P., Tutuc, E., Papadakis, S. J. & Shayegan, M. Resistance spikes at transitions between quantum Hall ferromagnets. Science 290, 1546 (2000).
  • (35) Shi, Q. et al. Odd- and even-denominator fractional quantum Hall states in monolayer WSe2. Nat. Nanotechnol. 15, 569 (2020).
  • (36) Nayak, C., Simon, S. H., Stern, A., Freedman, M. & Das Sarma, S. Non-Abelian anyons and topological quantum computation. Rev. Mod. Phys. 80, 1083 (2008).
  • (37) Hedin, L. New method for calculating the one-particle green’s function with application to the electron-gas problem. Phys. Rev. 139, A796 (1965).
  • (38) Hybertsen, M. S. & Louie, S. G. Electron correlation in semiconductors and insulators: band gaps and quasiparticle energies. Phys. Rev. B 34, 5390 (1986).
  • (39) Wang, G. et al. Magneto-optics in transition metal diselenide monolayers. 2D Mater. 2, 034002 (2015).
  • (40) Li, Z. et al. Emerging photoluminescence from the dark-exciton phonon replica in monolayer WSe2. Nat. Commun. 10, 2469 (2019).
  • (41) Srivastava, A. et al. Valley Zeeman effect in elementary optical excitations of monolayer WSe2. Nat. Phys. 11, 141 (2015).
  • (42) Koperski, M. et al. Orbital, spin and valley contributions to Zeeman splitting of excitonic resonances in MoSe2, WSe2 and WS2 Monolayers. 2D Mater. 6, 015001 (2019).
  • (43) Liu, E. et al. Gate tunable dark trions in monolayer WSe2. Phys. Rev. Lett. 123, 027401 (2019).
  • (44) Ting, C. S., Lee, T. K. & Quinn, J. J. Effective mass and gg factor of interacting electrons in the surface inversion Layer of silicon. Phys. Rev. Lett. 34, 870 (1974).
  • (45) Stern, F. Polarizability of a two-dimensional electron gas. Phys. Rev. Lett. 18, 546 (1967).
  • (46) Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 77, 3865 (1996).
  • (47) Giannozzi, P. et al. QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials. J. Phys.: Condens. Matter 21, 395502 (2009).
  • (48) da Jornada, F. H., Qiu, D. Y. & Louie, S. G. Nonuniform sampling schemes of the Brillouin zone for many-electron perturbation-theory calculations in reduced dimensionality. Phys. Rev. B 95, 035109 (2017).
  • (49) Deslippe, J. et al. BerkeleyGW: a massively parallel computer package for the calculation of the quasiparticle and optical properties of materials and nanostructures. Comput. Phys. Commun. 183, 1269 (2012).
  • (50) Ismail-Beigi, S. Truncation of periodic image interactions for confined systems. Phys. Rev. B 73, 233103 (2006).
  • (51) Fallahazad, B. et al. Shubnikov–de Haas oscillations of high-mobility holes in monolayer and bilayer WSe2: Landau level degeneracy, effective mass, and negative compressibility. Phys. Rev. Lett. 116, 086601 (2016).

Supplementary Information for: Valley-Filling Instability and Critical Magnetic Field for Interaction-Enhanced Zeeman Response in Doped WSe2 Monolayers

Refer to caption
Fig. 6: Quasi-2D screened Coulomb potential W¯vkF\bar{W}_{vk_{F}} and bare Coulomb potential V¯vkF\bar{V}_{vk_{F}} as a function of hole density. a Red squares: first principles results for W¯vkF\bar{W}_{vk_{F}} in atomic units computed using the valence band wavefunctions at KK. Dashed line: screened Coulomb potential for an ideal 2D fermion gas stern using a dielectric constant ϵ\epsilon of 15.615.6 medeps for WSe2. b Black squares: first principles results for V¯vkF\bar{V}_{vk_{F}} in atomic units computed using the valence band wavefunctions at KK. Blue curve: analytical 2D bare Coulomb potential.
Refer to caption
Fig. 7: LL fan diagram for ML WSe2 valence band with a hole density of 5.2×1012cm25.2\times 10^{12}\text{cm}^{-2}. Here, genh=8.0g^{*}_{\text{enh}}=-8.0 and no valley-filling instability occurs.
Table 2: Values of BcB_{c}, BXB_{X} and ΔB\Delta B (T) for given |genh||g^{*}_{\text{enh}}|.
|genh||g^{*}_{\text{enh}}| BcB_{c} BXB_{X} BX+ΔBB_{X}+\Delta B ΔB\Delta B
[10.4,11.2][10.4,11.2] [38.5,31.2][38.5,31.2] [38.5,37.4][38.5,37.4] [46.2,37.4][46.2,37.4] [7.7,0][7.7,0]
[12.5,13.7][12.5,13.7] [16.7,10.0][16.7,10.0] [16.7,11.7][16.7,11.7] [19.5,11.7][19.5,11.7] [2.8,0][2.8,0]

I Supplementary Note

I.1 Derivation of the expression for Σ(E)ffE\frac{\partial\Sigma(E)}{\partial f}\frac{\partial f}{\partial E}

The screened exchange term in our ab initio GWGW self-energy can be written as:

Σn𝐊SEX(E)=m1N𝐪Ω𝐪𝐆𝐆fm𝐊𝐪n𝐊|ei(𝐪+𝐆)𝐫|m𝐊𝐪m𝐊𝐪|ei(𝐪+𝐆)𝐫|n𝐊×ϵ𝐆𝐆1(𝐪,EEm𝐊𝐪)v𝐪+𝐆=m1L1N𝐪S𝐪𝐆𝐆fm𝐊𝐪n𝐊|ei(𝐪+𝐆)𝐫|m𝐊𝐪m𝐊𝐪|ei(𝐪+𝐆)𝐫|n𝐊×ϵ𝐆𝐆1(𝐪,EEm𝐊𝐪)v𝐪+𝐆=m1(2π)2d2q1L𝐆𝐆fm𝐊𝐪n𝐊|ei(𝐪+𝐆)𝐫|m𝐊𝐪m𝐊𝐪|ei(𝐪+𝐆)𝐫|n𝐊×ϵ𝐆𝐆1(𝐪,EEm𝐊𝐪)v𝐪+𝐆=m1(2π)2BZd2qW¯m𝐪fm𝐊𝐪,\displaystyle\begin{split}\Sigma^{\text{SEX}}_{n\mathbf{K}}(E)&=-\sum_{m}\frac{1}{N_{\mathbf{q}}\Omega}\sum_{\mathbf{qGG^{\prime}}}f_{m\mathbf{K}-\mathbf{q}}\bra{n\mathbf{K}}e^{i(\mathbf{q}+\mathbf{G})\cdot\mathbf{r}}\ket{m\mathbf{K}-\mathbf{q}}\bra{m\mathbf{K}-\mathbf{q}}e^{-i(\mathbf{q}+\mathbf{G^{\prime}})\cdot\mathbf{r^{\prime}}}\ket{n\mathbf{K}}\\ &\times\epsilon^{-1}_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{q},E-E_{m\mathbf{K}-\mathbf{q}})v_{\mathbf{q}+\mathbf{G}}\\ &=-\sum_{m}\frac{1}{L}\frac{1}{N_{\mathbf{q}}S}\sum_{\mathbf{qGG^{\prime}}}f_{m\mathbf{K}-\mathbf{q}}\bra{n\mathbf{K}}e^{i(\mathbf{q}+\mathbf{G})\cdot\mathbf{r}}\ket{m\mathbf{K}-\mathbf{q}}\bra{m\mathbf{K}-\mathbf{q}}e^{-i(\mathbf{q}+\mathbf{G^{\prime}})\cdot\mathbf{r^{\prime}}}\ket{n\mathbf{K}}\\ &\times\epsilon^{-1}_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{q},E-E_{m\mathbf{K}-\mathbf{q}})v_{\mathbf{q}+\mathbf{G}}\\ &=-\sum_{m}\frac{1}{(2\pi)^{2}}\int d^{2}q\frac{1}{L}\sum_{\mathbf{GG^{\prime}}}f_{m\mathbf{K}-\mathbf{q}}\bra{n\mathbf{K}}e^{i(\mathbf{q}+\mathbf{G})\cdot\mathbf{r}}\ket{m\mathbf{K}-\mathbf{q}}\bra{m\mathbf{K}-\mathbf{q}}e^{-i(\mathbf{q}+\mathbf{G^{\prime}})\cdot\mathbf{r^{\prime}}}\ket{n\mathbf{K}}\\ &\times\epsilon^{-1}_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{q},E-E_{m\mathbf{K}-\mathbf{q}})v_{\mathbf{q}+\mathbf{G}}\\ &=-\sum_{m}\frac{1}{(2\pi)^{2}}\int_{BZ}d^{2}q\bar{W}_{m\mathbf{q}}f_{m\mathbf{K}-\mathbf{q}},\end{split} (12)

where LL is the cell height, SS is the cell area, Ω\Omega is the cell volume, N𝐪N_{\mathbf{q}} is the number of q points.

In the above expression, the inverse dielectric function is written in a reciprocal space basis

ϵ1(𝐫,𝐫,E)=1Ω𝐪𝐆𝐆ei(𝐪+𝐆)𝐫ϵ𝐆𝐆1(𝐪,EEm𝐊𝐪)ei(𝐪+𝐆)𝐫,\displaystyle\epsilon^{-1}(\mathbf{r},\mathbf{r}^{\prime},E)=\frac{1}{\Omega}\sum_{\mathbf{qGG^{\prime}}}e^{i(\mathbf{q}+\mathbf{G})\cdot\mathbf{r}}\epsilon^{-1}_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{q},E-E_{m\mathbf{K}-\mathbf{q}})e^{-i(\mathbf{q}+\mathbf{G^{\prime}})\cdot\mathbf{r^{\prime}}}, (13)

and W¯m𝐪(E)\bar{W}_{m\mathbf{q}}(E) is defined as

W¯m𝐪(E)=1L𝐆𝐆n𝐊|ei(𝐪+𝐆)𝐫|m𝐊𝐪m𝐊𝐪|ei(𝐪+𝐆)𝐫|n𝐊×ϵ𝐆𝐆1(𝐪,EEm𝐊𝐪)v𝐪+𝐆.\displaystyle\begin{split}\bar{W}_{m\mathbf{q}}(E)&=\frac{1}{L}\sum_{\mathbf{GG^{\prime}}}\bra{n\mathbf{K}}e^{i(\mathbf{q}+\mathbf{G})\cdot\mathbf{r}}\ket{m\mathbf{K}-\mathbf{q}}\bra{m\mathbf{K}-\mathbf{q}}e^{-i(\mathbf{q}+\mathbf{G^{\prime}})\cdot\mathbf{r^{\prime}}}\ket{n\mathbf{K}}\\ &\times\epsilon^{-1}_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{q},E-E_{m\mathbf{K}-\mathbf{q}})v_{\mathbf{q}+\mathbf{G}}.\end{split} (14)

Σ(E)ffE\frac{\partial\Sigma(E)}{\partial f}\frac{\partial f}{\partial E} can be simplified as:

Σ(E)ffE=m1(2π)2d2qW¯m𝐪fm𝐊𝐪E=m1(2π)2d2qW¯m𝐪δ(Em𝐊𝐪EF)=1(2π)2d2qW¯n𝐪δ(En𝐪EF)=1(2π)2q𝑑q𝑑ϕqW¯n𝐪δ(E𝐪EF)=1(2π)202πq𝑑q𝑑ϕqW¯n𝐪|dE𝐪/dq||q=kF|m|2πW¯nkF(E=EF),\displaystyle\begin{split}\frac{\partial\Sigma(E)}{\partial f}\frac{\partial f}{\partial E}&=-\sum_{m}\frac{1}{(2\pi)^{2}}\int d^{2}q\bar{W}_{m\mathbf{q}}\frac{\partial f_{m\mathbf{K}-\mathbf{q}}}{\partial E}\\ &=\sum_{m}\frac{1}{(2\pi)^{2}}\int d^{2}q\bar{W}_{m\mathbf{q}}\delta(E_{m\mathbf{K}-\mathbf{q}}-E_{F})\\ &=\frac{1}{(2\pi)^{2}}\int d^{2}q\bar{W}_{n\mathbf{q}}\delta(E_{n\mathbf{q}}-E_{F})\\ &=\frac{1}{(2\pi)^{2}}\int qdqd\phi_{q}\bar{W}_{n\mathbf{q}}\delta(E_{\mathbf{q}}-E_{F})\\ &=\frac{1}{(2\pi)^{2}}\int_{0}^{2\pi}qdqd\phi_{q}\frac{\bar{W}_{n\mathbf{q}}}{|dE_{\mathbf{q}}/dq|}|_{q=k_{F}}\\ &\approx\frac{|m^{*}|}{2\pi}\bar{W}_{nk_{F}}(E=E_{F}),\end{split} (15)

where the delta function picks up the integrand at the Fermi surface and nn is the band index of the frontier doped band. We set E=EFE=E_{F}, which is equivalent to taking the approximation that ϵ𝐆𝐆1(𝐤𝐅,EKEF)ϵ𝐆𝐆1(𝐤𝐅,0)\epsilon^{-1}_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{k_{F}},E_{K}-E_{F})\approx\epsilon^{-1}_{\mathbf{G}\mathbf{G}^{\prime}}(\mathbf{k_{F}},0), where EKE_{K} is the energy at the band extremum in KK. This approximation is valid because EKEFE_{K}-E_{F} is on the order of meVs, and the dielectric function is fairly constant in this energy range. LiYang

I.2 Critical magnetic field BcB_{c} and condition for valley-filling instability

The condition for the valley-filling instability is given by ΔB>0\Delta B>0. In the following, we derive expressions for BcB_{c}, BXB_{X}, and ΔB\Delta B. Let NKNK be the KK-valley LL so that the energy of 0K0K^{\prime} lies in between those of NKNK and (N+1)K(N+1)K when B<BcB<B_{c}. This is equivalent to the condition that

Nm<|genh|<N+1m\displaystyle\frac{N}{m^{*}}<|g^{*}_{\text{enh}}|<\frac{N+1}{m^{*}} (16)

BcB_{c} is defined as the minimum value of BB at which 0K0K^{\prime} is just completely empty with holes. This implies that NKNK must also be completely occupied with holes at B=BcB=B_{c}. All the hole population originally in 0K0K^{\prime} for B<BcB<B_{c} has been transferred to NKNK at B=BcB=B_{c}. Thus, at B=BcB=B_{c}, we have:

ϵ0K+ωc2EF=EF(ϵNKωc2)\displaystyle\epsilon_{0K^{\prime}}+\frac{\hbar\omega_{c}}{2}-E_{F}=E_{F}-(\epsilon_{NK}-\frac{\hbar\omega_{c}}{2}) (17)

So,

ϵNK+ϵ0K2=EF\displaystyle\frac{\epsilon_{NK}+\epsilon_{0K^{\prime}}}{2}=E_{F} (18)

where EFE_{F} is the position of the Fermi level in the absence of the magnetic field. This condition leads to the expression:

Bc=|m||EF|(N+1)μB.\displaystyle B_{c}=\frac{|m^{*}||E_{F}|}{(N+1)\mu_{B}}. (19)

This expression for BcB_{c} gives results very similar to Bc|EF|/(|genh|μB)B_{c}\sim|E_{F}|/(|g^{*}_{\text{enh}}|\mu_{B}) given in the main text.

BXB_{X} is obtained by finding the value of BB at which ϵNK=ϵ0K\epsilon_{NK}=\epsilon_{0K^{\prime}}, for B>BcB>B_{c}:

BX=(|genh||g0|)BcN/|m||g0|,\displaystyle B_{X}=\frac{(|g^{*}_{\text{enh}}|-|g^{*}_{0}|)B_{c}}{N/|m^{*}|-|g^{*}_{0}|}, (20)

where g0g^{*}_{0} is the renormalized g-factor for B>BcB>B_{c}, i.e. the renormalized g-factor of the undoped system, and genhg^{*}_{\text{enh}} refers to the interaction-enhanced g-factor. BX+ΔBB_{X}+\Delta B is the value of BB at which the KK-valley LLs from index 0 to (N1)(N-1) can just contain all the holes, and can be obtained as the value of BB for the crossing point indicated by the yellow circle in Figure 4d of the main text. Thus, we have:

BX+ΔB=(1+1N)Bc.\displaystyle B_{X}+\Delta B=(1+\frac{1}{N})B_{c}. (21)

This gives:

ΔB=[N/|m||genh|N/|m||g0|+1N]Bc.\displaystyle\Delta B=[\frac{N/|m^{*}|-|g^{*}_{\text{enh}}|}{N/|m^{*}|-|g^{*}_{0}|}+\frac{1}{N}]B_{c}. (22)

The instability condition for genhg^{*}_{\text{enh}} can be obtained from ΔB>0\Delta B>0. Together with Eq. 16, we have

Nm<|genh|<N+1m|g0|N.\displaystyle\frac{N}{m^{*}}<|g^{*}_{\text{enh}}|<\frac{N+1}{m^{*}}-\frac{|g^{*}_{0}|}{N}. (23)

This expression also implies a condition on NN given by

Nm<N+1m|g0|N.\displaystyle\frac{N}{m^{*}}<\frac{N+1}{m^{*}}-\frac{|g^{*}_{0}|}{N}. (24)

Using our prediction of |g0|=6.6|g^{*}_{0}|=6.6 and m=0.48m^{*}=-0.48, the minimum NN is 44, giving the minimum |genh||g^{*}_{\text{enh}}| for a valley-filling instability to be 8.338.33.

References

  • (1) Stern, F. Polarizability of a two-dimensional electron gas. Phys. Rev. Lett. 18, 546 (1967).
  • (2) Laturia, A., Van de Put, M. L. & Vandenberghe, W. G. Dielectric properties of hexagonal boron nitride and transition metal dichalcogenides: from monolayer to bulk. NPJ 2D Mater. Appl. 2, 6 (2018).
  • (3) Liang, Y. & Yang, L. Carrier plasmon induced nonlinear band gap renormalization in two-dimensional semiconductors. Phys. Rev. Lett. 114, 063001 (2015).