Giant enhancement of exciton diffusion near an electronic Mott insulator
Abstract
Bose-Fermi mixtures naturally appear in various physical systems. In semiconductor heterostructures, such mixtures can be realized, with bosons as excitons and fermions as dopant charges. However, the complexity of these hybrid systems challenges the comprehension of the mechanisms that determine physical properties such as mobility. In this study, we investigate interlayer exciton diffusion in an H-stacked WSe2/WS2 heterobilayer. Our measurements are performed in the dilute exciton density limit at low temperatures to examine how the presence of charges affects exciton mobility. Remarkably, for charge doping near the Mott insulator phase, we observe a giant enhancement of exciton diffusion of three orders of magnitude compared to charge neutrality. We attribute this observation to mobile valence holes, which experience a suppressed moiré potential due to the electronic charge order in the conduction band, and recombine with any conduction electron in a non-monogamous manner. This new mechanism emerges for sufficiently large fillings in the vicinity of correlated generalized Wigner crystal and Mott insulating states. Our results demonstrate the potential to characterize correlated electron states through exciton diffusion and provide insights into the rich interplay of bosons and fermions in semiconductor heterostructures.
Layered transition metal dichalcogenides (TMDs) have become an interesting platform to study collective emerging electronic phenomena, including Mott insulators (?, ?, ?), generalized Wigner crystals (?, ?), density waves (?), fractional Chern insulators (?, ?), quasi-exciton condensation (?), superconductivity (?, ?), and kinetic ferromagnetism (?). One remarkable direction is to use optical excitons to probe various electronic orders (?). Since excitons can be generally considered as bosons, hybrid electron-exciton systems provide a natural platform for investigating Bose-Fermi mixtures. Bose-Fermi mixtures are ubiquitous in many-body physics, from solutions of fermionic 3He in bosonic superfluid 4He (?), and quark-meson models in QCD physics (?, ?), to ultracold atoms (?, ?, ?, ?).
Recent experimental implementations of Bose-Fermi mixtures in TMDs (?, ?, ?, ?, ?, ?) have motivated the search for exotic physics such as tunable particle interactions (?, ?), and topological superconductivity (?, ?). A key challenge is to understand how excitons behave in correlated electronic environments. Measuring the transport of excitons immersed in an electron gas has the potential to address this challenge. Although this technique has been proven effective in exploring the rich physics of TMD systems (?, ?, ?, ?, ?, ?, ?, ?, ?), the high exciton occupation considered in these demonstrations prevents probing fermionic correlations. Here, we explore the limit in which a very dilute concentration of bosons (excitons) diffuses in a 2D fermionic (electronic) gas to investigate how the rich phases of the fermionic many-body system affect the dynamics of bosons.
Our system consists of a WSe2/WS2 moiré heterostructure where we employ space- and time-resolved techniques to study diffusing interlayer excitons (IXs) immersed in a 2D fermionic electron gas. Our measurements reveal that the exciton diffusion coefficient is highly sensitive to changes in the electronic filling of the system. We analyze exciton dynamics within the various exotic electronic states that are realized by our system, uncovering a rich landscape dominated by polarons, generalized Wigner crystals, and a Mott-insulating state. Depending on the electronic state, we observe dramatic variations in the mobility of the diffusing species, with changes up to three orders of magnitude. Our results challenge the common assumption that electrons and holes forming IXs always move together in a monogamous manner, particularly near the Mott-insulating state. With the help of an effective model Hamiltonian, an interplay of two different channels for the diffusion of excitonic species is demonstrated. These findings present a novel optical approach for exploring complex quantum states in condensed matter systems.
1 Physical system

The experimental setup and moiré system are depicted in Fig. 1a. The system consists of an H-stacked WS2/WSe2 bilayer heterostructure hosting a triangular moiré lattice (see Supplementary Material). Optically exciting this system creates IX that form between spatially separated electrons and holes residing in different moiré registries (?). The nature and dynamics of IX is determined by the state of the electron gas (Fig. 1a). The spatially resolved photoluminescence (PL) emission, which is broader than the diffraction-limited optical excitation, encodes the information of IX diffusion (Fig. 1b-c). The excitation pump is resonant with the WSe2 intralayer exciton in all our measurements (upper panel of Fig. 1d). Upon ultra-fast electron transfer (in the order of femtoseconds (?)), the IX form, diffuse, and optically recombine (lower panel of Fig. 1d). As mentioned earlier, in this work we focus on the ultra-low exciton density regime where exciton-exciton interactions are negligible. Hence, the system can be treated as individual bosons moving within a gas of fermions.
To identify this regime, we look at the collected PL spectrum of IX as a function of the electron filling () for different pump intensities. Details on the calibration of are provided in Supplementary Note 1. Figure 1e-f displays the normalized PL spectra for two pump intensities (80 nW/m2 and 5 nW/m2). For both exciton densities, the charge neutral (CN) region is dominated by a low-energy IX (X1) which transitions into a high-energy exciton (X2) at . The transition is benchmarked by a meV gap in the PL emission, and it has been recently demonstrated to arise from the strong on-site exciton-electron repulsions owing to the formation of an electronic Mott-insulating state at (?, ?, ?). Although the system’s response in the Mott insulating state is similar for both pump intensities, there is a clear difference at low electron densities: in the ultra-low exciton density case (panel f) an additional gap of meV is observed near . This redshift corresponds to the transition from X1 to attractive polaron (AP), as corroborated by the diffusion measurements discussed later. The AP formation is facilitated in such an H-stacked system where the hole can bind with multiple nearest neighboring electrons (?); more details regarding the AP formation can be found in Supplementary Note 2. Understanding the precise nature of the AP in the presence of a lattice is an open problem and will be an interesting future research direction both experimentally and theoretically. Here, we identify the AP with the dressed charge complex comprised of X1 and dopant charges. It is important to note that AP formation occurs in both cases when the system is doped, however, a complete transition into AP can only be observed when the ratio of X1 to electron density is small; in other words, when each exciton can be dressed by surrounding electrons. This observation is validated by even lower pump intensity ( 5 nW/) measurements, as shown in Supplementary Note 3. All the presented data has been taken in this low-pump intensity regime unless stated otherwise.

In this low-pump intensity regime, we identify four distinct spectral features as a function of electron doping as shown in Fig. 1f. The extracted intensity and energy values are shown in Fig. 2a-b, and they can be distinguished as follows:
-
1.
Charge-neutral region: dominated by X1 emission,
-
2.
Moderate-doping region (): largely dominated by AP, except at certain fractional fillings (e.g. 1/3, 2/3, and 6/7) that host correlated states.
-
3.
Mott-insulating region (): dominated by X1. The strong electron localization inhibits the formation of APs in this region.
-
4.
High-doping region (): dominated by X2 due to double occupancy of moiré sites (doublon-hole pair).
Having identified the four different doping regimes, we now measure the diffusion dynamics of the quasiparticles of each regime and their dependence on the density of the 2D electron gas.
2 Exciton diffusion mechanisms
We extract the diffusion length from the spatially-resolved PL emission by employing a Gaussian fitting routine and analyze it as a function of . Considering that the measured PL profile corresponds to a convolution between quasiparticle diffusion and laser intensity distribution, the diffusion length is given by , where and correspond to the half-widths of the PL and laser profiles, respectively. More details about the fitting procedure are given in SM. The obtained values for the diffusion length are displayed as blue markers in Fig. 2c. In the CN region, the diffusion length is less than 200 nm; over an order of magnitude lower than previously reported measurements in structures without a moiré lattice (?, ?). Such a suppression in diffusion length indicates that the IX kinetic energy is quenched by the presence of a strong moiré potential in our system. Further suppression of is observed upon doping the system, i.e. in the region dominated by AP. Remarkably, a significant enhancement of is noticed as the system is further doped to , followed by a significant drop in the regime dominated by X2. For further clarity, spectrally resolved diffusion is shown in Supplementary Note 4.
A complete analysis of the IX mobility requires the measurement of its lifetime (), as the modulations in diffusion can originate either from the tunneling rate variation or the quasiparticle’s lifetime. We measure this quantity by using a femtosecond pulsed laser with a repetition rate of kHz while the signal is collected in a superconducting single photon detector (see Methods). The observed lifetime as a function of is shown with green markers in Fig. 2c. We notice a trend of lifetime reduction with doping, and modulations at some fractional fillings. In agreement with previous reports (?), decreases by over an order of magnitude, from s in the CN region to ns in the Mott region. The reduction in and enhancement of in the Mott region suggests the presence of highly mobile particles with low effective mass.
Assuming that the dynamics is diffusive, we can use the diffusion coefficient () to study different regimes. We observe a remarkable three-orders of magnitude enhancement of the mobility at the Mott-insulating region, see Fig. 2d. This intriguing behavior can be counter-intuitive at first, as high mobility contrasts with the insulating nature of the electron gas. To comprehend this anomalous behavior, we analyze how the changes in lead to the renormalization of the diffusing particle’s effective mass (). This effect originates from the Coulomb potential () exerted by the electron gas on the diffusing particles. The addition of and the moiré potential, together with changes in the mass of the diffusing particles determine the large variations in . An effective model to describe this behavior (?) and the extracted values of for each doping region are presented in Supplementary Note 5.

The four physical situations depicted in Fig. 2e describe how the electrons affect . With the mentioned model, we estimate the energy dispersion of the diffusing particles for each case (Fig. 2f). For the CN region, as , the electron and hole experience a similar potential landscape in their respective layers, favoring their motion together as a bound particle. As expected for moiré-trapped excitons, we observe a diffusion coefficient several orders of magnitude lower than the values reported for IX in the absence of moiré potential (?). In the AP region, where remains comparatively negligible, we detect an increase in the effective mass . The formation of heavier quasiparticles resulting from the dopant electrons dressing the injected optical excitons leads to the observed reduction in mobility. The decreasing trend in reverses at . This phenomenon is discussed in more detail in the next section.
The dramatic three-orders-of-magnitude increase in mobility near the Mott region can be attributed to the stronger effective potential in the electron layer and the suppressed potential in the hole layer, as depicted in Fig. 2e. This significant variation in effective potential results from the crystallization of electrons, which generates with the same period and phase as the electron moiré lattice but out of phase with the hole moiré lattice as holes occupy different atomic registry. Additionally, due to the absence of any vacancies in the electron layer, the hole in the other layer is free to hop into any moiré site. Hence, in the Mott region, the hole is not bound to a specific electron and can recombine with any electron in the lattice in a non-monogamous way. This non-monogamous hopping, combined with the reduced effective potential depth, leads to a decrease in , accounting for the orders-of-magnitude enhancement in . However, beyond , the excited electrons become mobile again, allowing the electron and hole to move together, in contrast to the previous non-monogamous hole dynamics. This is pictorially shown in the last panel of Fig. 2e, where an additional electron on top of the electron lattice and a hole in the other layer form a diffusing exciton. Interestingly, in this X2 region, both the hole and the electron encounter shallower potentials due to the out-of-phase in each layer. This shallower potential results in orders-of-magnitude increased mobility of X2 compared to X1. However, the monogamous motion manifests as a reduction in mobility compared to the holes in the Mott region. This discussion elucidates the observed strong variation in across the four different fermionic regions.
3 Diffusion at generalized Wigner crystal states
The remarkable sensitivity of the mobility to the fermionic state of the system is further manifested at certain fractional fillings that host generalized Wigner crystals due to the long-range repulsive interactions. Before focusing on these specific filling factors, we reiterate that the general trend of is reversed at . This observation can be understood by analyzing the occupation of the nearest neighbor (NN) sites of the injected IX below and above . Figure 3 illustrates how, in the two regimes, the NN filling determines the interplay between the two diffusion mechanisms. At , where all the NN sites are vacant, only monogamous motion can take place (Fig. 3a), while for , non-monogamous motion exclusively determines the exciton mobility due to the complete occupation of the moiré sites (Fig. 3b). For any other filling, the system experiences an interplay between the two diffusion mechanisms.
To elucidate, we make a direct comparison of the available diffusion channels at and . The monogamous motion of the exciton, allowed for , is blocked by the filled NN sites in the electron layer at (Fig. 3c and e). Hence, one would expect lower mobility at compared to . However, the experimental results (Fig. 2d) show a local minimum but an overall increase in at compared to . To understand this, we note that while the occupied NN sites hinder the monogamous motion of electron and hole, they simultaneously create non-monogamous channels that facilitate the diffusion of holes, as shown by the arrows in the bottom layer of Fig. 3e. Therefore, the increasing trend in is a consequence of the dominant non-monogamous hole diffusion. This effect gets stronger at as the occupation of both NN and next-nearest neighbor (NNN) sites (Fig. 3f) increases the number of diffusion channels in the hole layer. It should be noted that the local minima in at and result from the inhibition of the monogamous motion of IXs because the electrons’ wavefunctions are more localized.
The interplay of the two discussed diffusion mechanisms determines the non-trivial dependence of in the explored range of . In this picture, the dominant mechanism changes at . The monogamous motion of quasiparticles (X1 or AP) determines the mobility in the region 0 1/2, while it is the non-monogamous hole diffusion for 1/2 1. (Fig. 3d) acts as the turning point, exhibiting the lowest mobility as both processes are effectively suppressed.

4 Pump intensity phase map and temperature dependence
Our physical picture of the diffusion mechanisms is limited to a regime of low exciton density at cryogenic temperatures. Additional physical processes affecting the dynamics are expected to emerge for higher densities and temperatures. To gain insights into these more complex regimes, we map the dependence of the diffusion length as a function of and pump intensity. As illustrated in Figure 4a, the enhancement of the mobility near the Mott region is evident for pump intensities less than 100 nW/, indicated by the white dashed line. Beyond this pump intensity, the sharp peak associated with the Mott insulating region becomes less distinctive and shifts to lower , indicating a regime where additional processes such as double exciton occupancies and exciton-exciton scattering become relevant. The monotonic increase in with pump intensity at charge neutrality (CN) suggests that these scattering processes are enhanced with the exciton density. For pump intensities greater than 1 Wm2, the system enters a non-linear repulsive regime where exciton-exciton dipole repulsion dominates the diffusion dynamics. Figure 4a highlights these different regimes.
The reduction of mobility due to the formation of AP is more sensitive to exciton density. Hence, the features in this region start to fade away at around 50 nW/m2. In this case, the exciton to polaron transition is blurred out due to the larger population of excitons compared to the dopant electrons in the lattice.
Another crucial parameter for is the temperature (). Figure 4b shows that the diffusion peak at the Mott insulating state vanishes for K. Although the Mott gap persists over the full measured temperature range (see Supplementary Note 6), no significant dependence of with is observed beyond 200 K. This is attributed to increased phonon-assisted scattering mechanisms that dominate the diffusion dynamics at higher temperatures (?). Additionally, the signatures of AP in the voltage regime 1 V 2 V are suppressed above 60 K, as the AP cannot form at elevated temperatures due to its low binding energy.
5 Outlook
In conclusion, we unveil the two mechanisms affecting the mobility of a dilute exciton population in an electron-doped moiré system. The demonstrated breakdown of conventional monogamous exciton diffusion and clear signatures of electronic correlations evidence the potential of this technique to probe complex states of matter.
In perspective, a broad range of physical effects can be explored by measuring the mobility of embedded particles in van der Waals structures. For example, the nature of exotic phases of matter, including kinetic magnetism (?, ?), exciton condensation (?), and anomalous quantum Hall regimes (?, ?), could be studied with this technique. In particular, incorporating polarization resolution and fields into this technique offers the possibility to study fractional Chern insulators (?), Mott-moiré excitons (?), spin polarons (?), and tunable electron-exciton interactions (?, ?). Moreover, the microseconds-long exciton lifetimes allow one to conceive time-resolved diffusion experiments, capable of extracting valuable insights into the temporal dynamics of fermionic correlated states.
References and Notes
- 1. Y. Tang, et al., Simulation of Hubbard model physics in WSe2/WS2 moiré superlattices. Nature 579 (7799), 353–358 (2020).
- 2. Y. Shimazaki, et al., Strongly correlated electrons and hybrid excitons in a moiré heterostructure. Nature 580 (7804), 472–477 (2020).
- 3. Y. Shimazaki, et al., Optical signatures of periodic charge distribution in a Mott-like correlated insulator state. Physical Review X 11 (2), 021027 (2021).
- 4. Y. Xu, et al., Correlated insulating states at fractional fillings of moiré superlattices. Nature 587 (7833), 214–218 (2020).
- 5. E. C. Regan, et al., Mott and generalized Wigner crystal states in WSe2/WS2 moiré superlattices. Nature 579 (7799), 359–363 (2020).
- 6. Y. Zeng, et al., Exciton density waves in Coulomb-coupled dual moiré lattices. Nat. Mater. 22 (2), 175–179 (2023).
- 7. H. Park, et al., Observation of fractionally quantized anomalous Hall effect. Nature 622 (7981), 74–79 (2023).
- 8. Y. Zeng, et al., Thermodynamic evidence of fractional Chern insulator in moiré MoTe2. Nature 622 (7981), 69–73 (2023).
- 9. L. Ma, et al., Strongly correlated excitonic insulator in atomic double layers. Nature 598 (7882), 585–589 (2021).
- 10. Y. Xia, et al., Unconventional superconductivity in twisted bilayer WSe2 (2024).
- 11. Y. Guo, et al., Superconductivity in twisted bilayer WSe2 (2024).
- 12. L. Ciorciaro, et al., Kinetic magnetism in triangular moiré materials. Nature 623 (7987), 509–513 (2023).
- 13. L. Du, et al., Moiré photonics and optoelectronics. Science 379 (6639), eadg0014 (2023).
- 14. C. Ebner, D. O. Edwards, The low temperature thermodynamic properties of superfluid solutions of 3He in 4He. Phys. Rep. 2 (2), 77–154 (1971).
- 15. B.-J. Schaefer, J. Wambach, The phase diagram of the quark–meson model. Nucl. Phys. A 757 (3-4), 479–492 (2005).
- 16. P. Achenbach, et al., The present and future of QCD. Nuclear Physics A 1047, 122874 (2024).
- 17. K. Günter, T. Stöferle, H. Moritz, M. Köhl, T. Esslinger, Bose-Fermi mixtures in a three-dimensional optical lattice. Physical Review Letters 96 (18), 180402 (2006).
- 18. F. Schreck, et al., Sympathetic cooling of bosonic and fermionic lithium gases towards quantum degeneracy. Physical Review A 64 (1), 011402 (2001).
- 19. M.-G. Hu, et al., Bose polarons in the strongly interacting regime. Physical review letters 117 (5), 055301 (2016).
- 20. M. Delehaye, I. F. Barbut, S. Laurent, A mixture of bose and fermi superfluids. Science 345 (6200), 1035–1038 (2014).
- 21. Z. Lian, et al., Valley-polarized excitonic Mott insulator in WS2/WSe2 moiré superlattice. Nat. Phys. 20 (1), 34–39 (2023).
- 22. B. Gao, et al., Excitonic Mott insulator in a Bose-Fermi-Hubbard system of moiré WS2/WSe2 heterobilayer. Nat. Commun. 15 (1), 2305 (2024).
- 23. R. Xiong, et al., Correlated insulator of excitons in WSe2/WS2 moiré superlattices. Science 380 (6647), 860–864 (2023).
- 24. H. Park, et al., Dipole ladders with large Hubbard interaction in a moiré exciton lattice. Nat. Phys. 19 (9), 1286–1292 (2023).
- 25. L. Ma, et al., Strongly correlated excitonic insulator in atomic double layers. Nature 598 (7882), 585–589 (2021).
- 26. Z. Zhang, et al., Correlated interlayer exciton insulator in heterostructures of monolayer WSe2 and moiré WS2/WSe2. Nat. Phys. 18 (10), 1214–1220 (2022).
- 27. C. Kuhlenkamp, M. Knap, M. Wagner, R. Schmidt, A. m. c. Imamoğlu, Tunable Feshbach Resonances and Their Spectral Signatures in Bilayer Semiconductors. Phys. Rev. Lett. 129, 037401 (2022), doi:10.1103/PhysRevLett.129.037401.
- 28. I. Schwartz, et al., Electrically tunable Feshbach resonances in twisted bilayer semicondsowinuctors. Science 374 (6565), 336–340 (2021).
- 29. C. Zerba, C. Kuhlenkamp, A. Imamoğlu, M. Knap, Realizing topological superconductivity in tunable Bose-Fermi mixtures with transition metal dichalcogenide heterostructures. Physical Review Letters 133 (5), 056902 (2024).
- 30. E. Y. Andrei, et al., The marvels of moiré materials. Nat. Rev. Mater. 6 (3), 201–206 (2021).
- 31. E. Malic, R. Perea-Causin, R. Rosati, D. Erkensten, S. Brem, Exciton transport in atomically thin semiconductors. nature communications 14 (1), 3430 (2023).
- 32. J. Choi, et al., Moiré potential impedes interlayer exciton diffusion in van der Waals heterostructures. Sci. Adv. 6 (39), eaba8866 (2020).
- 33. J. Wang, et al., Diffusivity Reveals Three Distinct Phases of Interlayer Excitons in Heterobilayers. Phys. Rev. Lett. 126, 106804 (2021), doi:10.1103/PhysRevLett.126.106804.
- 34. C. Jin, et al., Imaging of pure spin-valley diffusion current in WS2-WSe2 heterostructures. Science 360 (6391), 893–896 (2018).
- 35. Z. Sun, et al., Excitonic transport driven by repulsive dipolar interaction in a van der Waals heterostructure. Nat. Photonics 16 (1), 79–85 (2022).
- 36. L. Yuan, et al., Twist-angle-dependent interlayer exciton diffusion in WS2-WSe2 heterobilayers. Nat. Mater. 19 (6), 617–623 (2020).
- 37. E. Wietek, et al., Nonlinear and Negative Effective Diffusivity of Interlayer Excitons in Moiré-Free Heterobilayers. Phys. Rev. Lett. 132, 016202 (2024), doi:10.1103/PhysRevLett.132.016202.
- 38. F. Tagarelli, et al., Electrical control of hybrid exciton transport in a van der Waals heterostructure. Nat. Photonics 17 (7), 615–621 (2023).
- 39. A. Rossi, et al., Anomalous interlayer exciton diffusion in WS2/WSe2 moiré heterostructure. ACS nano (2024).
- 40. X. Wang, et al., Intercell moiré exciton complexes in electron lattices. Nat. Mater. 22 (5), 599–604 (2023).
- 41. X. Hong, et al., Ultrafast charge transfer in atomically thin MoS2/WS2 heterostructures. Nat. Nanotechnol. 9 (9), 682–686 (2014).
- 42. Z. Li, et al., Interlayer Exciton Transport in MoSe2/WSe2 Heterostructures. ACS Nano 15 (1), 1539–1547 (2021), doi:10.1021/acsnano.0c08981.
- 43. J. Choi, et al., Moiré potential impedes interlayer exciton diffusion in van der Waals heterostructures. Sci. Adv. 6 (39), eaba8866 (2020).
- 44. C. Lagoin, F. Dubin, Key role of the moiré potential for the quasicondensation of interlayer excitons in van der Waals heterostructures. Physical Review B 103 (4), L041406 (2021).
- 45. A. Chernikov, M. M. Glazov, Chapter Three - Exciton diffusion in 2D van der Waals semiconductors, in 2D Excitonic Materials and Devices, P. B. Deotare, Z. Mi, Eds. (Elsevier), vol. 112 of Semiconductors and Semimetals, pp. 69–110 (2023), doi:https://doi.org/10.1016/bs.semsem.2023.09.001.
- 46. H. Yang, Y.-H. Zhang, Exciton- and light-induced ferromagnetism from doping a moiré Mott insulator. Phys. Rev. B. 110 (4) (2024).
- 47. W. Yao, Q. Niu, Berry phase effect on the exciton transport and on the exciton Bose-Einstein condensate. Phys. Rev. Lett. 101 (10), 106401 (2008).
- 48. T.-S. Huang, Y.-Z. Chou, C. L. Baldwin, F. Wu, M. Hafezi, Mott-moiré excitons. Phys. Rev. B 107, 195151 (2023), doi:10.1103/PhysRevB.107.195151.
- 49. Z. Tao, et al., Observation of spin polarons in a frustrated moiré Hubbard system. Nature Physics pp. 1–5 (2024).
- 50. L. Yuan, et al., Twist-angle-dependent interlayer exciton diffusion in WS2–WSe2 heterobilayers. Nature materials 19 (6), 617–623 (2020).
- 51. B. Amin, T. P. Kaloni, U. Schwingenschlögl, Strain engineering of WS 2, WSe 2, and WTe 2. Rsc Advances 4 (65), 34561–34565 (2014).
Acknowledgement
The authors acknowledge fruitful discussions with Ming Xie, Ajit Srivastava, and Angel Rubio. YZ acknowledges support from the National Science Foundation under Award No. DMR-2145712. M.K. acknowledges support from the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy–EXC–2111–390814868 and from the European Research Council (ERC) under the European Unions Horizon 2020 research and innovation programme (Grant Agreement No. 851161).
Methods
Device fabrication
The hBN encapsulated WSe2/WS2 heterostructures were fabricated using a dry-transfer method reported in the literature (?). All flakes were exfoliated from bulk crystals onto Si/SiO2 (285 nm) and identified by their optical contrast. The top/bottom gates and TMD contact are made of few-layer graphene. The flakes were picked sequentially with a polymer stamp and released onto a Si/SiO2 (90 nm) substrate. Later, electrodes consisting of 10 nm of chromium and 70 nm of gold were patterned on the substrate. They were fabricated using standard electron-beam lithography techniques and thermal evaporation. The sample was annealed at 300∘ C for 2 hr.
Optical Measurements
All the measurements are performed in a dilution refrigerator at 3.5K unless stated otherwise. The sample is excited using a Ti:Sapphire laser tuned at 733 nm, resonant to the WSe2 intralayer exciton, and focused to diffraction limit with an 80 microscope objective. Spatially resolved images are collected in a CCD camera (Princeton Instruments Blaze HRX) coupled to a spectrograph (Princeton Instruments SP2750). The total magnification of the optical setup is . For lifetime measurements, we excite the sample with a 100 fs pulsed Ti:Sapphire laser. We use a pulse picker to achieve a repetition rate low enough to detect the optical decay of the long-lived IX ( kHz). In this case, we use a superconducting nanowire single-photon detector and an event timer module to obtain the time-correlated PL signal.
Competing interests
The authors declare no competing interests.
Data availability
All of the data that support the findings of this study are reported in the main text and Supplementary Material. Source data are available from the corresponding authors on reasonable request.
Supplementary materials
Supplementary Text
Supplementary Materials for
Giant enhancement of exciton diffusion near an electronic Mott insulator
Pranshoo Upadhyay1,2,∗,
Daniel G. Suárez-Forero1,∗,†,
Tsung-Sheng Huang1,∗, Mahmoud Jalali Mehrabad1,
Beini Gao1,
Supratik Sarkar1,2,
Deric Session1, Kenji Watanabe3,
Takashi Taniguchi3,
You Zhou4,5, Michael Knap6,7,
Mohammad Hafezi1,‡
1Joint Quantum Institute, University of Maryland, College Park, MD 20742, USA.
2Department of Electrical and Computer Engineering, University of Maryland,
College Park, MD 20742, USA.
3Research Center for Materials Nanoarchitectonics, National Institute for Materials Science,
1-1 Namiki, Tsukuba 305-0044, Japan.
4Department of Materials Science and Engineering, University of Maryland,
College Park, MD 20742, USA.
5Maryland Quantum Materials Center, College Park, Maryland 20742, USA.
6Technical University of Munich, TUM School of Natural Sciences,
Physics Department, 85748 Garching, Germany.
7Munich Center for Quantum Science and Technology (MCQST),
Schellingstr. 4, 80799 München, Germany.
Corresponding authors. Email: † [email protected], ‡ [email protected]
∗These authors contributed equally to this work.
This PDF file includes:
Supplementary Note 1. Device structure and optical characterization
Supplementary Note 2. Formation of AP and its dependence on pump intensity
Supplementary Note 3. Ultra-low power excitation regime
Supplementary Note 4. Spectrally resolved diffusion
Supplementary Note 5. Doping-dependent renormalization of moiré potential
Supplementary Note 6. Temperature dependence of Mott gap
Supplementary Note 1. Device structure and optical characterization
We fabricate a transition metal dichalcogenide (TMD) hetero-bilayer sample for diffusion measurements. WS2 and WSe2 are the two monolayer TMDs used for this purpose. After aligning their edges, they are stacked to achieve 0o twist angle, i.e., 2H stacking order. The sample is encapsulated with hBN ( 35 nm) and gated on both sides for independent control of the out-of-plane electric field and doping. Figure S1a shows an optical micrograph of the sample. We characterize the device using reflection and photoluminescence (PL) measurements. Figures S1b and c present the reflection contrast spectrum of WSe2 intralayer exciton and PL map of the interlayer exciton (IX) measured on the bilayer region, respectively.
Both measurements show distinctive features at the charge-neutral (CN) region ( = 0), the Mott-insulating region ( = 1), and the band-insulating region ( = 2). Identifying these regions is useful to calibrate the gate voltage () to . However, for better accuracy, we track the peak intensity variations of the IX species with doping as shown in Fig. 2a (Main text). Here, is benchmarked by the emergence of the peak associated with the attractive polaron (AP). The presence of that peak coincides with a decrease in the X1 intensity. is identified by the complete extinction of the X1 intensity. It is important to mention that a proper calibration of can only be performed at ultra-low pump intensities (a few nW/ or lower), because it relies on the features of the polaron gap, only present at these low exciton densities.

Supplementary Note 2. Formation of AP and its dependence on pump intensity
Due to the H-stacking order in this system, electrons and holes localize in different moiré registries as discussed in detail in recent literature (?). This lateral separation of electrons and holes is the key to the formation of interlayer AP in the system. In this configuration, a hole in the WSe2 layer can bind with three sites in the electron layer with equal probability. This allows the hole in the bottom layer (WSe2) to pair with a cloud of electrons in the top layer (WS2). Due to the complexity involved in understanding the precise nature of APs, specifically in the moiré system, we consider a trion-like charge complex, without affecting the conclusion of this work, to explain our observations. In Fig. S2a, we illustrate this exemplary case of AP where an embedded X1 binds with one doped electron. It is important to note that the additional electrons in AP, i.e. other than the optically excited one, should be from the opposite valley to avoid short-range repulsive interactions (Fig. S2b). This is not expected for R-stacked heterobilayers, because in that case, electrons and holes are confined in the same registries, leading to weak repulsive interaction between X1 and doped electrons (?).
It should be further noted that AP formation can be distinctively observed in PL only for ultra-low excitation power. When the density of excitons is lower than the density of doped electrons in the system, we can have each exciton associated with at least one electron, forming AP. However, here, the density of doped electrons must be small as at higher densities close to 1 the repulsive interactions become strong and inhibit the formation of AP. We illustrate the dependence of AP on exciton and electron density in Fig. S3. For the case of 1 nW/, we have a clear transition from X1 to AP upon doping. This gap is observed to reduce and eventually vanish with excitation pump intensity. Additionally, AP transitions to X1 again at high electron doping close to 1.


Supplementary Note 3. Ultra-low power excitation regime
Based on the above discussion regarding AP formation, we determine the dilute exciton density regime. A complete polaron gap ( 16 meV) is used as a benchmark for achieving the dilute regime. In the main text, we have shown a complete polaron gap for 5 nW/. In Fig. S4, we demonstrate that any pump intensity below 5 nW/ will have a comparable spectrum in the entire range of doping. This confirms our regime of operation.

Supplementary Note 4. Spectrally resolved diffusion
As discussed in the main manuscript, the surge in diffusion length () near the Mott-insulating region is a novel observation that warrants a detailed analysis. In Fig. S5a, we show spectrally resolved diffusion collected by using gratings and separating the spectrum into two regions: one below 860 nm (X2) and one above (X1 or AP). We apply the Gaussian fitting routine and extract the diffusion length. As shown by the red (X1 or AP) and purple (X2) markers, diffusion near the Mott region is solely from X1. Although both X1 and X2 have similar intensity in that region (Fig. S5b), X2 demonstrates negligible diffusion. However, for , the diffusion is dominated by X2 and results in a peak near = 2.

Supplementary Note 5. Doping-dependent renormalization of moiré potential
To model the variations in diffusivity with doping, we solve a simplified 1D lattice Hamiltonian: Here, is the mass of the exciton, trion, or hole, depending on the regime of consideration, in the absence of the moiré potential. is the total potential resulting from adding the moiré potential () and the Coulomb interactions () which accounts for the modification of the superlattice due the charge order, and is the moiré lattice spacing. This Hamiltonian, whose eigenstates in real space are the Mathieu functions, has been successfully used to model interlayer excitons in TMD heterobilayers (?). Although this model does not capture the geometric details of the moiré lattice, it provides the effective mass () of diffusing particles. , determined by the values of and in each regime, is a critical parameter due to its inverse proportionality to the mobility. Extracting the variation in allows us to understand the rich behavior shown in Fig. 2d of the main text.
To evaluate the renormalization of the effective moiré potential () we consider certain parameters: 100 meV in each layer (?), the mass of electron () and hole () in WS2 and WSe2 is considered to be 0.5 and 0.4 , respectively (?). Here, is the free electron mass. The list of , and values for the diffusing particle in the corresponding fermionic region are listed in Table S1. For the charge-neutral (CN) region, for X1 is obtained by adding of both layers. It can be noted that in the 1D picture, we do not account for the rotational motion of electrons and hence, direct addition is considered to be a valid approximation. For the AP region, can be assumed to remain the same as in the CN region. For our calculations, we consider the simplest scenario where an excited X1 binds to one doped electron (trion-like). Developing a full theoretical model for the AP is an open problem in the presence of a lattice and would be an interesting future direction. Here, we assume a trion picture where we can approximate its mass as 2 + . This increase in leads to a much stronger localization effect in the presence of the moiré lattice. This effect is captured by the increase in indicating slower diffusion of AP than X1.
In the Mott region, electrons in WS2 crystallize due to strong electron-electron repulsion. However, holes in WSe2 experience a weak effective potential because of the addition of an out-of-phase periodic potential ( meV) exerted by the doped electron lattice (WS2) on the holes. The obtained is out-of-phase compared to of the hole layer because the electrons and holes reside in different moiré registries, which are laterally displaced. This leads to a strong suppression in ( 12 meV), allowing for non-monogamous hole diffusion.
For X2 diffusion, for the electron and hole is calculated separately and then added. The potential for the hole layer remains roughly the same as in the Mott region, as the hole still experiences the strong due to the electron lattice. However, for the electron layer, there are strong repulsive on-site interactions between the doped electron and the excited electron. This effectively increases the energy of the excited electron, which then sees a shallower potential. We approximate as - , where represents the on-site electron-electron repulsions. For H-stacking, this can be approximated as the exciton-electron repulsion, which is observed from PL to be around 46 meV. Therefore, of 54 meV for the electron layer and 12 meV for the hole layer give an overall of 66 meV experienced by X2. This leads to an increase in the effective mass () compared to the Mott region, but it remains orders of magnitude smaller than in the CN region.
Fermionic Region (Diffusing particle) | () | (meV) | () |
---|---|---|---|
CN Region (X1) | 0.9 | 200 | 182 |
Low doping (AP) | 1.4 | 200 | 2800 |
Mott (hole) | 0.4 | 12 | 0.4 |
High Doping (X2) | 0.9 | 66 | 4.6 |
Supplementary Note 6. Temperature dependence of Mott gap
We perform temperature-dependent PL measurements to verify the presence of the Mott gap in the entire range of temperatures used for measurements. Figure S6 demonstrates the Mott gap even at 260 K, although the gap reduction is observed with temperature.
