Valley-Filling Instability and Critical Magnetic Field for Interaction-Enhanced Zeeman Response in Doped WSe2 Monolayers
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 agree well with experiment. Unlike traditional valleytronic materials such as silicon, the enhancement in g-factor vanishes beyond a critical magnetic field achievable in standard laboratories. We identify ranges of for which this change in g-factor at 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 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 above which the interaction-induced enhancement in the g-factors vanishes in doped TMDs. We identify ranges of the enhanced g-factor for which the discontinuous change in g-factor at results in a LL alignment and valley-filling instability for . 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 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 , it has been shown that an out-of-plane magnetic field results in the following expression for the LLs at the valley Xuan2020 :
(1) |
where is the corresponding band index, is the mean-field single-particle energy at , is the valley effective mass, is the Bohr magneton and is the LL index. The total intrinsic single band g-factor consists of the orbital and spin contribution, , where is taken to be and is the spin quantum number. The orbital g-factor is defined using the orbital magnetic moment Xuan2020 (see Methods).
However, the above static mean-field description does not account for the energy-dependent electron self-energy that is necessary for a many-body description of the quasiparticle (QP) energies. 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 leads to an effective renormalized g-factor, , defined as , where
(2) |
Thus we have,
(3) |
Such a renormalization effect is missing in previous first principles calculations of g-factors in TMDs Xuan2020 ; Rohlfing2020 ; PRB2020 .
Renormalized () | Intrinsic () | |
-4.36 | -5.50 | |
-2.05 | -2.59 | |
-6.63 | -8.31 | |
-4.48 | -5.62 | |
-4.54 | ||
-9.16 |
The electron self-energy in this work is computed within the approximation Hedin (see Methods), which uses the first order term in the perturbative expansion of in terms of the screened Couloumb interaction . 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 % compared to the intrinsic g-factors, because 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 .

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 :
(4) |
where is the Fermi-Dirac distribution function. The second term in Eq (4) can be simplified to give (see Methods):
(5) |
where is the band index of the frontier doped band, and and are respectively the Fermi energy and Fermi wave vector. This term comes from the screened exchange contribution to the self-energy. (defined in Methods) is an effective quasi-2D screened Coulomb potential which can be evaluated completely from first principles. Since 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 derived in Ref. Janak, (see also Ref. Quinn, ) where the first term in Eq. (4) is ignored. We note that our numerical results for 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 , and henceforth, the subscript is omitted. The magnitudes of our computed valence band g-factors 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 . 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

A subtle but important point is that Eq. (5) applies only when the band occupancies are changing with , and the Fermi level is fixed. In electrostatic gating experiments, the carrier concentration is fixed rather than the absolute Fermi level. However, the Zeeman shifts in and are equal in magnitude but opposite in sign (Fig. 2a, ), and the density of states for the quadratic bands in 2D is independent of energy. So for small enough that both valleys have carriers (mixed polarized regime; Fig. 1a), 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 . However, above a critical magnetic field , only one valley has carriers (Fig. 2a, ) (see Supplementary Note for a more precise expression for ). As increases beyond , a constant hole density is maintained when shifts with the bands without changing the band occupancies. Thus, for , only the first term in Eq. (4) applies, similar to the undoped case, leading to an abrupt drop in at (Fig. 2b), with a corresponding piecewise-linear Zeeman split (Fig. 2c). For a hole density of , T, and we have for and for .
This abrupt drop in at a critical magnetic field has never been reported or predicted before in traditional valleytronic materials such as silicon. Indeed, is inversely related to , and it is the large intrinsic g-factors and hence large renormalized g-factors for TMDs that allow for to be small enough to be reached in standard laboratories. For the same hole density of , we predict in silicon to be 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

Since is the value of characterizing the onset of the fully polarized regime, can be deduced using optical measurements of the exciton and polaron energies for and jump . In Fig. 3a, we plot these values of (blue circles) and compare them with our predicted values (red squares). The predicted dependence of 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, increases, giving a larger range of for which is enhanced by interactions (Fig. 3a). However, the magnitude of decreases as the hole concentration increases (Fig. 1). These competing effects imply that for any given field , there is an optimal hole concentration which maximizes the Zeeman split (Fig. 3b). This optimal hole concentration corresponds to the hole concentration for which (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

The abrupt change in at also has other interesting implications. As increases beyond , the decrease in results in a decrease in magnitude of the slopes of the LL fan diagrams (Fig. 4a), leading to a crossing between the energies of and (purple circle, in Fig. 4a). If 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 .
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 is fully occupied, the plot is shaded from down to . At , an integer number of LLs ( in Fig. 4c) are completely filled with holes while is completely empty. As decreases slightly below , the LL degeneracy decreases, and the LL with the next lower energy ( in Fig. 4c) is required to contain the holes, leading to a symmetric zigzag fine structure about . As increases slightly above , the LL degeneracy increases. As long as , only the valley is filled with holes ( in Fig. 4b), and the blue shaded area represents the constant hole density. But when is slightly larger than (Fig. 4d) where the energies of and cross, holes will start to fill the valley again ( in Fig. 4b), until the field is large enough that the LLs to can contain all the holes (yellow circle, Fig. 4d) and the system becomes fully polarized again. This represents a valley-filling instability, where is depleted of holes from to , and filled again up to . In practice, when holes begin to fill , the mixed polarized regime is reached and 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 but also for up to .
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 close to the onset of the fully-polarized regime in the experiment. This is consistent with the highest occupied LL in the valley being emptied and partially filled with holes at 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 rather than a pronounced change at one particular value of 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 and are respectively and T for jump , close to our predicted values of and T for the same (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 between and , it is interesting to predict how large can be and how sensitive is to fluctuations in . Not all values of will result in the instability (see Supplementary Figure 2 and Supplementary Note). In particular, if the energies of and cross at , the valley-filling instability only occurs if is occupied with holes. Fig. 4e plots the ranges of for which a valley-filling instability will occur, as well as the corresponding values. We see that an optimal range of for LL alignment is . Here, is quite large (T for ) and is also fairly robust to changes in . The corresponding values of and fall within to T (Supplementary Table 1), well within the reach of experiments.
The alignment of LLs in different valleys can in principle be achieved for if can be tuned such that matches exactly with for some . However, once 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 .

Tunability of interaction-enhanced g-factor
We further note that in addition to electrostatic gating which changes the carrier concentration and thus (Fig. 1), can also be tuned by dielectric screening (Fig. 5). The tunability of with the background dielectric constant can be understood from the fact that is related to the effective quasi-2D screened Coulomb potential at the Fermi surface (Eq. 5). This tunability of provides a handle to control the valley-polarized current, and .
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 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 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 (). As the carrier concentration increases, decreases and increases, so that for any value of the magnetic field , the valley-polarization is maximized when the carrier concentration is such that . 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 and corresponding ranges of 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 is defined as Xuan2020 :
(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 calculations of the intrinsic g-factors, we use a non-uniform sampling method sub of the Brillouin Zone starting with a 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 Ry with empty bands is used for the reciprocal space expansion of the dielectric matrix. The intrinsic single-band g-factor reduces by only when an energy cutoff of Ry with empty bands is used.
Calculation of renormalized g-factor
The renormalized g-factor is computed from the intrinsic g-factor and using Eq. 3. As the dependence of on is no longer linear, we generalize Eq. (2) to the case where refers to the quasiparticle energies in the presence of a -field, and the applied represents a small increment in .
We approximate 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.
can be partitioned Louie into the dynamical non-local screened-exchange (SEX) and Coulomb-hole (COH) interaction terms . Only the screened exchange term depends on the occupancies and contributes to the second term of Eq. 4. The screened exchange energy in our ab initio plane-wave calculation can be written as (see Supplemental Note):
(7) |
where we define the quasi-2D screened Coulomb potential:
(8) |
Here, is the cell volume, is the number of q-points and is the Coulomb potential with the slab Coulomb truncation scheme applied trunc . is an effective quasi-2D screened Coulomb potential defined in valley and is the height of the supercell for ML WSe2. The second term in Eq. (4) is then given by (see Supplemental Note):
(9) |
We compute by evaluating 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 , a Ry -vector cut off and bands. is unchanged when we use instead Ry and bands, and reduces by when a 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 , 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 is defined by:
(10) |
Our first principles results for (Supplementary Figure 1) agrees with the analytically-derived 2D Coulomb potential. At low doping densities, is very large, which would change the sign of compared to the intrinsic g-factor, indicating that screening is important for a meaningful description of .
Background dielectric constant
A uniform background dielectric constant () can be simply added to the dielectric function of the system to obtain the total dielectric function: . In our first principles calculation, the dielectric function is expanded in a plane wave basis: . Thus we approximate the effect of screening by a dielectric medium by modifying the static dielectric matrix as follows:
(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 ‐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 factors of van der Waals heterostructures from first-principles calculations. Phys. Rev. B 101, 235408 (2020).
- (9) Förste, J. et al. Exciton -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 -valley electrons in MoSe2. Phys. Rev. B 97, 201407(R) (2018).
- (21) Roth, L. M. factor and donor spin-lattice relaxation for electrons in germanium and silicon. Phys. Rev. 118, 1534 (1960).
- (22) Yafet, Y. factors and spin-lattice relaxation of conduction electrons. Solid State Phys. 14, 1-98 (1963).
- (23) Janak, J. F. 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 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 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


I Supplementary Note
I.1 Derivation of the expression for
The screened exchange term in our ab initio self-energy can be written as:
(12) |
where is the cell height, is the cell area, is the cell volume, is the number of q points.
In the above expression, the inverse dielectric function is written in a reciprocal space basis
(13) |
and is defined as
(14) |
can be simplified as:
(15) |
where the delta function picks up the integrand at the Fermi surface and is the band index of the frontier doped band. We set , which is equivalent to taking the approximation that , where is the energy at the band extremum in . This approximation is valid because is on the order of meVs, and the dielectric function is fairly constant in this energy range. LiYang
I.2 Critical magnetic field and condition for valley-filling instability
The condition for the valley-filling instability is given by . In the following, we derive expressions for , , and . Let be the -valley LL so that the energy of lies in between those of and when . This is equivalent to the condition that
(16) |
is defined as the minimum value of at which is just completely empty with holes. This implies that must also be completely occupied with holes at . All the hole population originally in for has been transferred to at . Thus, at , we have:
(17) |
So,
(18) |
where is the position of the Fermi level in the absence of the magnetic field. This condition leads to the expression:
(19) |
This expression for gives results very similar to given in the main text.
is obtained by finding the value of at which , for :
(20) |
where is the renormalized g-factor for , i.e. the renormalized g-factor of the undoped system, and refers to the interaction-enhanced g-factor. is the value of at which the -valley LLs from index to can just contain all the holes, and can be obtained as the value of for the crossing point indicated by the yellow circle in Figure 4d of the main text. Thus, we have:
(21) |
This gives:
(22) |
The instability condition for can be obtained from . Together with Eq. 16, we have
(23) |
This expression also implies a condition on given by
(24) |
Using our prediction of and , the minimum is , giving the minimum for a valley-filling instability to be .
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).