DOI:
10.1039/D1MH01186C
(Communication)
Mater. Horiz., 2022,
9, 1089-1098
Bright excitonic multiplexing mediated by dark exciton transition in two-dimensional TMDCs at room temperature†
Received
27th July 2021
, Accepted 19th January 2022
First published on 20th January 2022
Abstract
2D-semiconductors with strong light–matter interaction are attractive materials for integrated and tunable optical devices. Here, we demonstrate room-temperature wavelength multiplexing of the two-primary bright excitonic channels (Ab-, Bb-) in monolayer transition metal dichalcogenides (TMDs) arising from a dark exciton mediated transition. We present how tuning dark excitons via an out-of-plane electric field cedes the system equilibrium from one excitonic channel to the other, encoding the field polarization into wavelength information. In addition, we demonstrate how such exciton multiplexing is dictated by thermal-scattering by performing temperature dependent photoluminescence measurements. Finally, we demonstrate experimentally and theoretically how excitonic mixing can explain preferable decay through dark states in MoX2 in comparison with WX2 monolayers. Such field polarization-based manipulation of excitonic transitions can pave the way for novel photonic device architectures.
New concepts
In this work, we demonstrate strong wavelength multiplexing of the two-primary bright excitonic channels in MoX2 based monolayers, exploiting the dark excitation of the Ad-transition. This is realized by control over the out-of-plane component of the excitation electric field, encoding the polarization information into the two bright spectral channels at room temperature. Particularly, substantial out-of-plane excitation is used to generate high energy Ad-excitons, which leads to dominant Bb-emission. In contrast, purely in-plane dipole excitation results in a predominant Ab-emission. We further establish the pivotal role of thermal energy and CB ordering in supporting or suppressing the multiplexing for MoX2 and WX2, respectively. Finally, state of the art ab initio GW-BSE calculations were carried out demonstrating a rich excitonic exchange interaction that strongly support the presented PL multiplexing phenomena, temperature dependent PL spectra and the observed deviation between Mo- and W-based TMDCs. Our new findings shine light on the fundamental optical properties of monolayer TMDs, while paving a novel route for TMDs based photonic devices, e.g. spectral bits for optical computing, multifunctional light sources and more.
|
Introduction
Excitons are fundamental optically excited, Coulomb-bound electron–hole pairs in semiconductors. Ever-present, excitons play a major role in the optical properties of reduced dimensionality semiconductors such as quantum wells (1D and 2D) and monolayer transition-metal dichalcogenides (TMDs), due to their reduced dielectric screening and significant spatial confinement.1–4 In particular, in comparison with other 2D semiconductors, monolayer TMDs feature robust bound excitons with exceptionally large binding energy,2–4 making monolayer TMDs an attractive platform for adequate light–matter interaction investigations and nanophotonic applications at room temperature.5–8
Monolayer TMDs, typically presented as MX2 (where M = Mo, W; X = S, Se), have non-centrosymmetric hexagonal crystal structure, resulting in a direct bandgap at the ±K valley.9,10 In addition, strong spin–orbit coupling (SOC) removes the valence band (VB) and conduction band (CB) spin degeneracy.11 For the former, the spin splitting is on the order of hundreds of meV, spectrally separating two distinct excitonic channels generated from the higher and lower VB, termed the A- and B-excitons, respectively.2,12 Compared with its VB counterpart, the CB splitting is considerably smaller- on the scale of a few meV.10,13 While small, this splitting governs the optical selection rules, separating the excitonic transition into distinct bright and dark channels with orthogonal transition dipole orientations.14,15 Thus, the associated excitonic transitions can be classified as either optically bright Ab-, Bb- or optically dark Ad-, Bd- excitons (Fig. 1a). Note that the brightness of these states is also determined by symmetry and momentum selection rules as extensively discussed before for this class of materials.16–18
 |
| Fig. 1 Photoluminescence selection of Ab- and Bb-exciton in monolayer MoS2 induced by in- and out-of-plane excitation. (a) – Electronic band structure of monolayer MoS2 at the K valley. Bands color code represents the spin degree of freedom. The colored (dashed) arrows represent bright (dark) exciton transitions, respectively. (b) Schematic illustration of the Ab- and Bb-exciton wavelength multiplexing in monolayer MoS2 on Si/SiO2 substrate. The method is based on tuning the out-of-plane component of the exciting electric field to select exciton channel. (c) In-plane excitation (marked by dashed line Ab) followed by emission from the Ab-excitonic channel marked by the red arrow. On the left, an illustration of an in-plane dipole transition (Ab-) with its radiation profile excited by in plane polarized field. (d) Photoluminescence spectra of MoS2 excited by mostly in-plane polarized light showing a predominant emission of the Ab-excitonic channel marked by λ1. (e) Out-of-plane excitation (marked by the dashed line Ad) followed by Bb-excitonic channel marked by the orange arrow. On the left an illustration of an out-of-plane dipole (Ad-) transition with its radiation profile excited by an out-of-plane polarized field. (f) Photoluminescence spectra of MoS2 excited through dark exciton results in the predominant emission of the Bb excitonic channel marked as λ2. | |
Bright excitons correspond to optically-allowed electron–hole transitions with the same spin orientation. These transitions strongly couple to light via in-plane (x–y direction) dipoles,19 making the Ab-, Bb-channels the dominant features in conventional photoluminescence (PL) spectrum. Interestingly, the main PL emission channel in TMDs arises from the Ab-exciton,2,20 where to date, limited increase of the Bb-exciton emission signal was realized by several factors such as pump intensity,21 substrate,22 sample quality,23 and bright-to-bright exciton mixing24 (see ESI† note 1). Nonetheless, an all-optical way enabling strong modulation between the two bright excitonic channels can pave the way for exciting tunable multi-wavelength devices.
The two low-lying and spin-split dark excitons, Ad- and Bd-, are commonly considered optically forbidden due to the implied violation of the spin selection rule. However, while the Bd-, is noncontroversial strictly dipole forbidden,14,25 recent group theory analysis suggests that the restriction upon Ad- can be dismantled by considering out-of-plane dipole transitions.14,18,25–27 This dark channel was experimentally demonstrated in various ways, such as surface plasmons polaritons coupling28,29 and applied in-plane magnetic fields mixing.30–32 Yet, due to the limited strength of its out-of-plane (z-direction) dipoles, it does not yield significant PL at room temperature in comparison with its bright counterparts. Nevertheless, the existence of these dark excitonic states has an acute implication on the material optical signature, arising from the different ordering between the Ad- and Ab- excitons.33–36 For example, in different TMDs with chemical compound such as Mo- (W-), the PL is quenched (enhanced) at elevated temperature due to the thermal-population of the dark (bright) excitonic state.36,37 Although the delicate thermal stabilization and the Ab-, Bb-exchange interaction has been studied in different TMDs, the possible coupling between the Ad-channel to the Bb-excitonic channel was not demonstrated.
In this work, we report a novel approach to manipulate the bright excitonic channels in MoX2-based monolayers, exploiting the excitation of the Ad-state at room temperature. Through control over the out-of-plane component of the excitation electric field, one can select whether to uplift the excitonic population of either the Ab- or the Bb-states, ultimately, encoding the polarization information into the two bright spectral channels (Fig. 1b). Particularly, substantial out-of-plane excitation is used to generate Ad-excitons (Fig. 1e), followed by a surprisingly strong Bb-emission (Fig. 1f) mediated by an ascribed excitonic exchange interaction. In contrast, purely in-plane dipole excitation populates mostly the Ab-excitonic state (Fig. 1c), resulting in its predominant emission (Fig. 1d). By combining the robust optical excitation, thermal role and inherent excitonic exchange interactions, we accomplish polarization information multiplexing into the spectral Ab- and Bb-excitonic channels. Finally, we deduce by ab initio calculations and experimental observation exciton mixing of high-energy Ad- and low-lying Bb-states, explaining the higher probability for their coupling in monolayer MoSe2 compared to WSe2, which enables (inhibits) the multiplexing operation in Mo- (W-) based monolayer TMDs.
Results and discussion
Emergence of Bb-exciton mediated by dark excitonic transition
The dark excitonic transition depicts the selective excitation of Ad-excitons through the application of out-of-plane polarization. In monolayer MoX2, the formation of this dark exciton mediates the recombination of the Bb-channel, which we theoretically rationalize by an efficient intravalley exciton exchange interactions, based on ab initio calculations of the electron–hole transitions, within the GW and Bethe Salpeter equation (GW-BSE) approach and as recently shown for bright A–B exciton coupling.24 Realization of substantial electric field polarization in the z-direction (propagation axis) can be achieved by employing a high numerical aperture (NA) in a high-confocal microscope. Here, the electrical field spatial distribution of the in- and out-of-plane polarizations varies significantly along the propagation axis (Fig. 3a and b), promoting a focal plane based selective excitation of in- and out-of plane dipoles, generating bright and dark excitons, respectively.38,39 To experimentally demonstrate the principle of the mechanism, a CVD grown MoS2 monolayer was transferred onto a 300 nm SiO2/Si substrate. Continuously excited by an off-resonance 532 nm linearly polarized laser, PL measurements were carried using a high NA objective of 0.75 at different focal planes in a confocal microscope. As a result, a significant wavelength multiplexing of the measured PL spectra (i.e. anti-correlated change in the magnitude of the two predominant peaks at around 670 and 620 nm) involving the Ab- and Bb-channels, was achieved by the continues change in the focal plane (Fig. 2a). In comparison, for a low NA objective of 0.25, where the electric field polarization in the z-direction is considerably lower, the transition between the two excitonic states was not observed (Fig. 2b). This implies that the multiplexing occurs due to the alteration of selective excitation of the in- and out-of plane dipoles.
 |
| Fig. 2 In-plane and out-of-plane PL dependent excitation. (a) Measured PL spectra of MoS2 for different focal planes, taken with high NA (0.75) objective showing a clear transition between Ab- and Bb-emission as function of the focal plane position. (b) Measured PL spectra of MoS2 for different focal planes, taken with a low NA (0.25) objective showing no excitonic transition. (c) Schematic illustration of the experimental setup, the objective is moved to different focal planes (dashed red line), where the z-axis origin is determined to be at the focal plane overlapping with the MoS2 monolayer. | |
To demonstrate the connection between the dark exciton transition and the Bb-emission, we calculated the spatial distribution of the electric field by numerically solving the Debye–Wolf integral for a linearly x-polarized Gaussian beam.40 The electric field amplitude components of |Ex| and |Ez| for NA of 0.75 are depicted in Fig. 3a and b respectively, showing a ratio of 3
:
1 for the max amplitude of |Ex|:|Ez| at the focal point. For comparison, the same simulation was performed for NA of 0.25 (Fig. 3c and d), exhibiting a significantly higher ratio of 10
:
1 between the electric fields of |Ex| and |Ez|. The relation between the integrated z-field intensity Iz of an area of 2λ around the optical axis and the Bb-exciton emission (integrated fitted Gaussian Bb-exciton PL peak, normalized to the integrated maximum of the set) at different focal planes are presented in Fig. 3e. A clear correlation between the out-of-plane electric field polarization and the Bb-exciton emission is demonstrated, where the maximum signal of the Bb-exciton is achieved at the focal plane.
 |
| Fig. 3 PL dependence on In-plane and out-of-plane excitation. (a and b) Calculated electric field in the xz-plane of a propagating Gaussian beam for a high NA of 0.75 objective. The max amplitude ratio received for |Ex|:|Ez| is approximately 3 : 1 indicating that a significant contribution to out-of-plane dipole excitation is feasible. The white lines indicate the spatial amplitude distribution at different focal planes. (c and d) Similar calculated electric fields for a low NA of 0.25 objective. Here, we find a considerably lower ratio between the |Ex|:|Ez| of around 10 : 1. (e) Correlation between the integrated z-field intensity Iz for an area of 2λ around the optical axis and the Bb-excitonic contribution (integrated Gaussian signal) from the general spectrum. Both Iz and the Bb-contribution were normalized to the set value at z = 0. We note that the Bb-contribution set was slightly shifted by 0.17 (μm) to have better overlapping with the field intensity. | |
Considering that the dark transition prompts the population of the Ad-excitonic state, the relation to the observed B-emission feature can be explained by several excitonic interaction routs, which occur on a faster timescale than the Ad-lifetime,30 such as intravalley excitonic exchange,24 intervalley momentum-forbidden excitonic interaction33 and the formation of charged excitons.41–44 The latter, however, can be excluded due to the lack of alteration in the B-signal under gate bias (see ESI† note 2); while an interplay between the neutral and charged excitons can be observed for the A-exciton regime. Additionally, the population of the Ab- seems to increase by the application of positive voltage with no apparent change of the B-emission, indicating a less effective bright-to-bright intravalley excitonic exchange.20,24 While intervalley interaction has been shown to be prudent in the optical behavior of monolayer TMDs, it is mostly emphasized in the context of the Bb-bleaching.34,45 Thus, we connect the increase of the B-emission feature to the Bb-exciton via an unexplored dark mediated excitonic transition, as will be shown below.
As aforementioned, out-of-plane excitation promotes the excitonic population of the higher energetic dark state in monolayer MoS2, which is followed by the strong Bb-excitonic emission. Nonetheless, such dark states can decay non-radiatively through intravalley scattering, leading to the bleaching of the Bb-excitonic state and increase in Ab-emission.24,34,46,47 Since the intravalley scattering time τbd occurs on a faster time scale than the radiative recombination τbd ≪ τAd, τAb, τBb,48,49 thermal equilibrium has a major role in determining the population distribution within the excitonic levels. We therefore expect that at low (high) temperatures, a strongly populated lower (higher) excitonic level, will result in strong Ab- (Bb-) exciton emission. Evidently, strong Ab-exciton emission was recently observed in monolayer MoSe2 at low temperatures, albeit excitation was induced by out-of-plane polarization.25
Thermal dependent photoluminescence
To study the thermal dependency of the above excitonic multiplexing, we conducted PL measurements at temperatures range of 125–298 K using a THMS600 cooling system (LinKam Ltd.) and an objective with NA of 0.45 to account for the increased working distance. Fig. 4a and b present the PL measurements at different focal planes for the two temperatures extrema i.e. at 298 K and at 125 K, showing a substantial reduction in the Bb-excitonic emission channel at low temperature. We first consider a simplified excitonic picture, which does not include exchange interactions between the A and B transitions, where both the Ad- and Bb-excitonic energy levels lie above the Ab- and Bd-levels, respectively; thus, thermal scattering may impose higher populations of the Bb- and Ad-states, as schematically demonstrated in Fig. 4c (see ESI† note 3). However, such simplified explanation is not enough to understand the origin of the modulation without an exchange-driven interaction, as we further explore below, since thermal energy alone cannot surpass the 150 meV energetic barrier between the excitonic species (Fig. 4c). With that in mind, the ratio between the Bb- and Ab-emission can be naively described by an Arrhenius relation for the intravalley scattering between the A- and B-states. Here, under an assumption that the Ad- emission is negligible with respect to the intravalley scattering, the excitonic population ratio can be represented as B/A ∝ exp(−Δ/kBT). Where Δ is the thermal activation energy, kB is the Boltzmann constant and A and B correspond to the population of the Ab- and Bb-excitonic population, respectively. This relation corresponds to the linear behavior between ln(Bb/Ab) (integrated intensities of the Bb- and Ab-emission) taken at z = 0 and 1/kBT (Fig. 4d). Indeed, the extracted activation energy of ∼15 meV is not sufficient to explain direct transitions between the A and B exciton onset, and demands a more complicated dynamical effects, as we elaborate below.
 |
| Fig. 4 Temperature dependent PL modulation. Measured normalized PL colormap of monolayer MoS2 taken with a NA of 0.45 for different focal planes at 298 K (a) and 125 K (b). The spectra were normalized for the maximum value of the measurement set at a given temperature. The thermal energy kBT is marked on the bottom right in each spectrum. The Bb-excitonic emission is observed at high temperatures, whereas at low temperatures the exciton emission is quenched. (c) Schematic of intravalley non-mixed excitonic states in MoS2. Dashed lines represent dark transitions, band color represent correlating CB band transition (upper CB-orange; lower CB- green). Here the Kex depicts the excitonic mixing of the dark (Ad)-to-bright (Bb-) exchange. (d and e) Relative PL integrated intensity between the Ab- and Bb-excitonic emission for MoS2 and MoSe2, respectively. The intensities are taken in a natural logarithm as a function of (kBT)−1 marked by the black squares. The results were fitted with the thermal scattering model presented in the text marked by the red line with a slope of Δ ≅ 15 and Δ ≅ 20 (me V) for MoS2 and MoSe2, respectively. | |
To examine the generality of this phenomenon, similar temperature dependent PL measurements were conducted for monolayer MoSe2 placed on a 300 nm SiO2/Si substrate (see ESI† note 3). Similar multiplexing of the Bb- and Ab-channels was observed, while MoSe2 presented larger Bb/Ab excitonic emission ratio that can be attributed to its rapid valley depolarization, stronger Bb-momentum dipole transition50 and comparable stronger mixing rates.24 Additionally, the temperature dependent PL spectra of MoSe2 are in correspondence with the Arrhenius relation of Δ ≅ 20 me V (Fig. 4e). These results demonstrate the importance of thermal energy for realizing strong Bb-excitonic emission mediated by the dark exciton transition. The discrepancy between the observed thermal barrier of ∼10–20 meV and the significantly larger energy difference between the Ad and the Bb energies (>100 meV) suggest that the excited Ad excitons involve high energetic level states, i.e. X2, corresponding to the broadband excitation in our setup (2.33 eV). Such high energetic excitation is sufficient to directly populate Ad excitons with high enough energy to mediate the strong Bb emission, as will be shown below. We also note that the similar observed response for different MoX2 monolayers suggests a common mechanism that originates from their profound exciton mixing, as we show in the next sections. It is also important to note that similar trend in modulation of Bb/Ab exciton emission ratio is expected due to chromatic aberration, which similarly separates frequencies into different focal plains.51 Importantly, the chromatic aberration in the spectral range of the Ab- and Bb-was found to be minor in our setup (see ESI† note 5).
We further compare between PL measurements of MoSe2 (Fig. 5a) and WSe2 (Fig. 5b), both placed on a 300 nm SiO2/Si substrate and measured with a NA of 0.75. Interestingly, a significantly weaker multiplexing of the Bb/Ab ratio is observed for WSe2 in comparison with MoSe2. These findings point to exciton coupling mechanisms that depend in both the energy and the coupling probabilities between the participating excitons, as we discuss below.
 |
| Fig. 5 Excitonic mixing in TMDCs and PL excitonic modulation. (a and b) Room temperature PL spectra taken with a NA of 0.75 at different focal planes for MoSe2 and WSe2, respectively. The observed excitonic modulation is profound in MoSe2, while very small in WSe2. (c–f) Computed GW-BSE absorption spectra and the respective contributing exciton transitions for MoS2, MoSe2, WS2, and WSe2 monolayers, respectively. Different colors represent the fraction of the normalized contributions from Ab-, Ad-, Bb-, Bd-transitions as defined in the text. Comparing between the examined systems, MoS2 and MoSe2 (c and d) have significantly larger mixing between bright and dark transitions than WS2 and WSe2 (e and f), where hardly any mixing occurs. In addition, while in the Mo-based monolayers the Bb-state appears close in energy and slightly below the second A excitation, in the W-based systems they appear around higher A excitations, reducing the probability to be occupied by optically excited A states. | |
Dark-to-bright exciton mixing in MoSe2 and WSe2
To understand the mechanism dominating the observed dark-to-bright exciton transitions, we analyze plausible exciton interaction mechanisms. First, we consider single-exciton state coupling due to intravalley exchange interactions, as was recently demonstrated in MoS2.24 We preformed many-body perturbation theory calculations within the GW-BSE approximation, while explicitly including spinor wavefunctions of the electron and hole states. This method allows a close inspection of the various electron–hole transitions contributing to the excitonic states in the energy range of the explored A and B peaks. In particular, we study the exciton mixing level of the Ab, Ad, Bb, and Bd states, for MoS2, MoSe2, WS2, and WSe2 (Fig. 5c–f). For MoS2 and MoSe2 we find that A exciton states strongly mix dark and bright electron–hole transitions, and slightly mix with B states, in good correspondence with previous observations.24 Importantly, higher-energy mixed (dark–bright) A exciton states are also found slightly above the spin-split bright B peak in both systems (full data of state mixing is given in the ESI† note 6). Our results point to possible decay channels which can eventually explain the modulated Bb emission. In this scenario, in both MoS2 and MoSe2, excited-state absorption can populate mixed Ad- and Ab-exciton states, which in turn couple to mixed Bd- and Bb-states. Such coupling is allowed due to intervalley exchange mixing, as shown by Guo et al.24 and in our present calculations. In stark contrast, the excitonic picture in WSe2 and WS2 reveals a very different mixing nature: the computed low-energy A excitations are well separated and do not mix dark and bright nature. In addition, due to the large SOC energy split, the Bb exciton energetically coincide with much higher excited A states, lowering the decay probability to the low-energy Bb-exciton.
Considering the relatively low effective thermal barrier that is extracted from the temperature dependent PL measurements, it is plausible that excited-state A excitons thermally couple to low-lying B excitons via absorption or emission of phonons with energies at the scale of ∼10–60 meV before they optically decay. Such relaxation mechanism can be associated with previously-observed phonon-assisted dark exciton decay processes34,52–54 with exciton coupling allowed due to exchange mixing. A second scenario, although less plausible, involves an upconversion mechanism from the lowest, X1 exciton state, partially dominated by Ad transitions, to the higher X2 exciton state, dominated by Bb transitions. Such exciton–exciton coupling must overcome energy differences at the order of 150 meV in Mo-based monolayers, and 300 meV in the W-based ones. This channel may be associated to a recently-suggested upconversion process between bright A and B states in WSe2, observed in PLE measurements, and related with non-linear processes such as phonon-assisted Auger recombination.46 To explore this further, PL resonant excitation measurements were carried out with monolayer MoS2 (ESI† Fig. S8). Here, under excitation of 633 nm at different focal planes the B-excitonic signal was not observed, making the upconversion process a less plausible exchange route compared to the coherent exchange interaction in the 2s region. This indicates the pivotal role of the exchange interaction and specifically, the part of the dark-to-bright exchange channels, arising from out-of-plane excited Ad-related states for the amplification of the Bb-signal. We thus attribute these bright-to-dark exchange channels as the predominant mechanism for the multiplexing action in MoX2 TMDs monolayers. The computed exciton mixing is much weaker for the case of WS2/WSe2, offering a possible explanation for our experimental findings of reduced Bb/Ab transition probabilities (Fig. 5b). Finally, there are still several key questions that one has to address in order to get a quantitative understating regarding the quantum efficiency of such dark to bright excitonic transition. In particular, the selective high-energy excitonic absorption (dark and bright) as a function of light polarization (longitudinal vs. transverse) and their correlation with the resulting PL emission.
Conclusions
In summary, we demonstrate a new approach for bright excitons multiplexing of the Ab- and Bb-excitonic channels in monolayer TMDs at room temperature. As a proof of concept, dark excitonic mediated transitions were realized by a confocal microscope with high NA. Thus, selective excitation of high energy Ad-excitons were observed at different focal planes, mediating the emission of the Bb-exciton channel via dark-to-bright exchange interactions. Furthermore, we establish the important role of excitonic thermal scattering and excitonic ordering. Finally, GW-BSE calculations were employed to unveil profound exciton state mixing and coupling in MoX2 compared with WX2, which ultimately enables the multiplexing operation. These findings shed light on the fundamental optical properties of monolayer TMDs, while paving a novel route for TMDs based photonic devices, e.g. spectral bits for optical computing, multifunctional and tunable light sources and more.
Materials and methods
Sample preparation
MoS2 and MoSe2 monolayers were grown by a space confined CVD approach, as previously reported.55,56 In a typical growth process, MoO3 (99.5%, Sigma Aldrich) and Sulfur (or Selenium) powders (99.95%, Sigma Aldrich) were used as the metal and the chalcogen precursors, respectively. A ceramic boat containing (∼3.5 mg) of MoO3 powder was placed at the center of a 1-inch diameter CVD furnace. A piece of 300 nm thermally grown SiO2 coated Si substrate was mounted on top of the same boat with its polished surface facing upwards. Few small pieces of mica sheets were cut and placed above the target substrate. The Sulfur (or Se, ∼350 mg) boat was placed ∼22 cm upstream from the MoO3 source-growth substrate. The quartz tube was initially purged with ∼ 250 sccm of highly pure Ar (5N) carrier gas for 10 minutes. Thereafter, the furnace temperature was ramped to 750 °C at 15 °C per minute with 30 sccm of Ar flow and the calchogen was separately heated to 150 °C or 240 °C, for S and Se, respectively. The growth time was kept for 5–10 minutes at 750 °C. MoS2 and MoSe2 samples were then wet-transfer to the SiO2/Si substrate: The TMDs monolayers where spin coated with 0.5 wt% polystyrene (PS) for 60 seconds at 4000 rpm, followed with two baking steps at 90 °C for 35 minutes and 120 °C for 15 minutes. The samples were then placed in DI for 5 minutes until the PS-TMDs film was malleable enough to peel from the substrate. The film was taken out with the Si/SiO2 substrate and baked with the same previous conditions. Finally, toluene was used to dissolve the PS film. The highly crystalline grown CVD WSe2 sample was commercially acquired (2D semiconductors) and was mechanically transferred to a SiO2/Si substrate.
Optical measurements.
PL measurements were conducted using a WITec alpha 300R confocal Raman microscopy system; with excitation power intensity of 0.5 mW from a single fiber mode 532 nm linear polarized laser. The PL was acquired using a 600 g mm−1 grating. Measurements at different focal planes were conducted by moving the objective position in z-axis. The objectives used in the experiment include: 10x/0.25, 50x/0.75 by Zeiss and 50x/0.45 by Olympus. Thermal measurements were conducted using a LinKam THMS600 temperature-controlled chamber. For convenience, the experimental zero position of the focal plane is set according to the strongest received Bb-exciton signal (Fig. 2c).
Theoretical calculations
We obtain the starting point electron and hole wavefunctions from density functional theory (DFT)57 calculations within the PBE approximation,58 using the Quantum Espresso package.59 We use a 2D periodic unit cell with lattice vectors of 3.29 Å for MoSe2 and WSe2 and 3.16 Å for MoS2 and WS2, with a 15 Å distance between repeating unit cells in the out-of-plane direction. We apply a plane-wave basis and norm-conserving pseudopotentials. A 30 × 30 × 1 k-point grid was used to calculate the self-consistent charge density with a 70 Ry wave function cutoff. Spin polarization was included using fully-relativistic corrections. We compute the quasiparticle band structure using the GW method within the BerkeleyGW software58 using the Hybertsen–Louie generalized plasmon-pole (HL-GPP) model.60,61 We used a screening energy cutoff of 25 Ry for the reciprocal lattice components of the dielectric matrix and included 4000 spinor bands in the summation. We employed the nonuniform neck subsampling (NNS) scheme62 to sample the Brillouin zone. In this scheme, we use a 6 × 6 × 1 uniform q-grid and include an additional set of 10 q-points in the Voronoi cell around q = 0. A truncated Coulomb interaction was used to prevent spurious interactions between periodic images of the 2D sheet.63 We solve the Bethe Salpeter equation (BSE) for the spinor electron and hole wavefunctions within the Tamm–Dancoff approximation (TDA), as implemented in the BerkeleyGW code.60,64 The BSE matrix elements were calculated on a uniform 24 × 24 × 1 k-grid and then interpolated to a 48 × 48 × 1 fine k-grid, with 6 empty and 6 occupied bands included in the transition matrix.
Data availability
The data that support the findings of this study are available from the corresponding author upon reasonable request.
Author contributions
S. K. B. C. and E.K. conceived the experimental concept. S.K. and S.M. performed the experimental work. B. C. and S. S. performed optical calculations. P. M. and A. P. performed CVD growth of 2D samples. S. R. A. and T. A. performed the ab initio calculations. V. K., A. I., S. R. A., E. H. and E. K. supervised the work. All authors participated in the data analysis and in the writing of the manuscript.
Conflicts of interest
The authors declare no competing interests.
Acknowledgements
E. K. gratefully acknowledge the Israel Science Foundation (ISF), grant 1567/18, for financial assistance and the RBNI for the nanofabrication facilities. E. K. thanks the Taub fellowship for leadership in science and technology, supported by the Taub Foundation and the Alon fellowship. E. H. gratefully acknowledge the Israel Science Foundation (ISF), the US Air Force Office of Scientific Research (FA9550-18-1-0208) through their program on Photonic Metamaterials and the Israel Ministry of Science, Technology and Space. S. R. A. acknowledges support from the Israel Science Foundation (ISF) Grant No. 1208/19 and an Alon Fellowship. P. K. M. and A. I. acknowledge the generous support from the Israel Science Foundation, projects # 2549/17 and 2171/17. This research used computational resources of the National Energy Research Scientific Computing Center (NERSC). This research used computational resources of the National Energy Research Scientific Computing Center (NERSC) and of the ChemFarm Computational Cluster at the Weizmann Institute.
References
-
G. D. Scholes and G. Rumbles, Materials For Sustainable Energy: A Collection of Peer-Reviewed Research and Review Articles from Nature Publishing Group, 2011, 12–25 Search PubMed
.
- K. F. Mak, C. Lee, J. Hone, J. Shan and T. F. Heinz, Phys. Rev. Lett., 2010, 105, 136805 CrossRef PubMed
.
- K. He, N. Kumar, L. Zhao, Z. Wang, K. F. Mak, H. Zhao and J. Shan, Phys. Rev. Lett., 2014, 113, 026803 CrossRef CAS PubMed
.
- A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C.-Y. Chim, G. Galli and F. Wang, Nano Lett., 2010, 10, 1271–1275 CrossRef CAS PubMed
.
- K. Rong, B. Wang, A. Reuven, E. Maguid, B. Cohn, V. Kleiner, S. Katznelson, E. Koren and E. Hasman, Nat. Nanotechnol., 2020, 15, 927–933 CrossRef CAS PubMed
.
- M. Koperski, M. R. Molas, A. Arora, K. Nogajewski, A. O. Slobodeniuk, C. Faugeras and M. Potemski, Nanophotonics, 2017, 6, 1289–1308 CAS
.
- F. Xia, H. Wang, D. Xiao, M. Dubey and A. Ramasubramaniam, Nat. Photonics, 2014, 8, 899–907 CrossRef CAS
.
- D. Xiao, G.-B. Liu, W. Feng, X. Xu and W. Yao, Phys. Rev. Lett., 2012, 108, 196802 CrossRef PubMed
.
- W. Zhao, Z. Ghorannevis, L. Chu, M. Toh, C. Kloc, P.-H. Tan and G. Eda, ACS Nano, 2013, 7, 791–797 CrossRef CAS PubMed
.
- A. Kormányos, G. Burkard, M. Gmitra, J. Fabian, V. Zólyomi, N. D. Drummond and V. Fal’ko, 2D Mater., 2015, 2, 022001 CrossRef
.
- X. Xu, W. Yao, D. Xiao and T. F. Heinz, Nat. Phys., 2014, 10, 343–350 Search PubMed
.
- K. Kośmider, J. W. González and J. Fernández-Rossier, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 88, 245436 CrossRef
.
- G.-B. Liu, W.-Y. Shan, Y. Yao, W. Yao and D. Xiao, Phys. Rev. B: Condens. Matter Mater. Phys., 2013, 88, 085433 CrossRef
.
- J. Echeverry, B. Urbaszek, T. Amand, X. Marie and I. Gerber, Phys. Rev. B, 2016, 93, 121107 CrossRef
.
- A. Slobodeniuk and D. Basko, 2D Mater., 2016, 3, 035009 CrossRef
.
- T. Mueller and E. Malic, npj 2D Mater. Appl., 2018, 2, 29 CrossRef
.
- J. Xiao, M. Zhao, Y. Wang and X. Zhang, Nanophotonics, 2017, 6, 1309–1328 CAS
.
- Z. Ye, T. Cao, K. O’brien, H. Zhu, X. Yin, Y. Wang, S. G. Louie and X. Zhang, Nature, 2014, 513, 214–218 CrossRef CAS PubMed
.
- J. A. Schuller, S. Karaveli, T. Schiros, K. He, S. Yang, I. Kymissis, J. Shan and R. Zia, Nat. Nanotechnol., 2013, 8, 271 CrossRef CAS PubMed
.
- Y.-C. Wu, S. Samudrala, A. McClung, T. Taniguchi, K. Watanabe, A. Arbabi and J. Yan, ACS Nano, 2020, 14, 10503–10509 CrossRef CAS PubMed
.
- S. Dubey, S. Lisi, G. Nayak, F. Herziger, V.-D. Nguyen, T. Le Quang, V. Cherkez, C. Gonzalez, Y. J. Dappe and K. Watanabe, ACS Nano, 2017, 11, 11206–11216 CrossRef CAS PubMed
.
- M. Buscema, G. A. Steele, H. S. van der Zant and A. Castellanos-Gomez, Nano Res., 2014, 7, 561–571 CrossRef
.
- K. M. McCreary, A. T. Hanbicki, S. V. Sivaram and B. T. Jonker, APL Mater., 2018, 6, 111106 CrossRef
.
- L. Guo, M. Wu, T. Cao, D. M. Monahan, Y.-H. Lee, S. G. Louie and G. R. Fleming, Nat. Phys., 2019, 15, 228–232 Search PubMed
.
- G. Wang, C. Robert, M. Glazov, F. Cadiz, E. Courtade, T. Amand, D. Lagarde, T. Taniguchi, K. Watanabe and B. Urbaszek, Phys. Rev. Lett., 2017, 119, 047401 CrossRef CAS PubMed
.
- M. R. Molas, A. O. Slobodeniuk, T. Kazimierczuk, K. Nogajewski, M. Bartos, P. Kapuıfmmode, K. Oreszczuk, K. Watanabe, T. Taniguchi, C. Faugeras, P. Kossacki, D. M. Basko and M. Potemski, Phys. Rev. Lett., 2019, 123, 096803 CrossRef CAS PubMed
.
- L. M. Schneider, S. S. Esdaille, D. A. Rhodes, K. Barmak, J. C. Hone and A. Rahimi-Iman, Sci. Rep., 2020, 10, 1–6 CrossRef PubMed
.
- K.-D. Park, T. Jiang, G. Clark, X. Xu and M. B. Raschke, Nat. Nanotechnol., 2018, 13, 59–64 CrossRef CAS PubMed
.
- Y. Zhou, G. Scuri, D. S. Wild, A. A. High, A. Dibos, L. A. Jauregui, C. Shu, K. De Greve, K. Pistunova and A. Y. Joe, Nat. Nanotechnol., 2017, 12, 856 CrossRef CAS PubMed
.
- X.-X. Zhang, T. Cao, Z. Lu, Y.-C. Lin, F. Zhang, Y. Wang, Z. Li, J. C. Hone, J. A. Robinson and D. Smirnov, Nat. Nanotechnol., 2017, 12, 883–888 CrossRef CAS PubMed
.
- C. Robert, B. Han, P. Kapuscinski, A. Delhomme, C. Faugeras, T. Amand, M. Molas, M. Bartos, K. Watanabe and T. Taniguchi, Nat. Commun., 2020, 11, 1–8 Search PubMed
.
- Z. Lu, D. Rhodes, Z. Li, D. Van Tuan, Y. Jiang, J. Ludwig, Z. Jiang, Z. Lian, S.-F. Shi and J. Hone, 2D Mater., 2019, 7, 015017 CrossRef
.
- E. Malic, M. Selig, M. Feierabend, S. Brem, D. Christiansen, F. Wendler, A. Knorr and G. Berghäuser, Phys. Rev. Mater., 2018, 2, 014002 CrossRef CAS
.
- M. Selig, F. Katsch, R. Schmidt, S. M. de Vasconcellos, R. Bratschitsch, E. Malic and A. Knorr, Phys. Rev. Res., 2019, 1, 022007 CrossRef CAS
.
- M. R. Molas, C. Faugeras, A. Slobodeniuk, K. Nogajewski, M. Bartos, D. Basko and M. Potemski, 2D Mater., 2017, 4, 021003 CrossRef
.
- X.-X. Zhang, Y. You, S. Y. F. Zhao and T. F. Heinz, Phys. Rev. Lett., 2015, 115, 257403 CrossRef PubMed
.
- G.-H. Peng, P.-Y. Lo, W.-H. Li, Y.-C. Huang, Y.-H. Chen, C.-H. Lee, C.-K. Yang and S.-J. Cheng, Nano Lett., 2019, 19, 2299–2312 CrossRef CAS PubMed
.
- M. Brotons-Gisbert, R. Proux, R. Picard, D. Andres-Penares, A. Branny, A. Molina-Sánchez, J. F. Sánchez-Royo and B. D. Gerardot, Nat. Commun., 2019, 10, 1–10 CrossRef CAS PubMed
.
- R. Scott, J. Heckmann, A. V. Prudnikau, A. Antanovich, A. Mikhailov, N. Owschimikow, M. Artemyev, J. I. Climente, U. Woggon and N. B. Grosse, Nat. Nanotechnol., 2017, 12, 1155–1160 CrossRef CAS PubMed
.
- E. Wolf, Proc. R. Soc. London, Ser. A, 1959, 253, 349–357 Search PubMed
.
- K. F. Mak, K. He, C. Lee, G. H. Lee, J. Hone, T. F. Heinz and J. Shan, Nat. Mater., 2013, 12, 207–211 CrossRef CAS PubMed
.
- J. Förste, N. V. Tepliakov, S. Y. Kruchinin, J. Lindlau, V. Funk, M. Förg, K. Watanabe, T. Taniguchi, A. S. Baimuratov and A. Högele, Nat. Commun., 2020, 11, 1–8 Search PubMed
.
- K. Shinokita, X. Wang, Y. Miyauchi, K. Watanabe, T. Taniguchi, S. Konabe and K. Matsuda, Phys. Rev. B, 2019, 99, 245307 CrossRef CAS
.
- Brill, ACS Appl. Mater. Interfaces, 2021, 13(27), 32590–32597 CrossRef CAS PubMed
.
- G. Berghäuser, I. Bernal-Villamil, R. Schmidt, R. Schneider, I. Niehues, P. Erhart, S. M. de Vasconcellos, R. Bratschitsch, A. Knorr and E. Malic, Nat. Commun., 2018, 9, 1–8 CrossRef PubMed
.
- M. Manca, M. Glazov, C. Robert, F. Cadiz, T. Taniguchi, K. Watanabe, E. Courtade, T. Amand, P. Renucci and X. Marie, Nat. Commun., 2017, 8, 1–7 CrossRef PubMed
.
- Z. Wang, A. Molina-Sanchez, P. Altmann, D. Sangalli, D. De Fazio, G. Soavi, U. Sassi, F. Bottegoni, F. Ciccacci and M. Finazzi, Nano Lett., 2018, 18, 6882–6891 CrossRef CAS PubMed
.
- K. F. Mak, K. He, J. Shan and T. F. Heinz, Nat. Nanotechnol., 2012, 7, 494–498 CrossRef CAS PubMed
.
- L. Yang, N. A. Sinitsyn, W. Chen, J. Yuan, J. Zhang, J. Lou and S. A. Crooker, Nat. Phys., 2015, 11, 830–834 Search PubMed
.
- M. Yang, C. Robert, Z. Lu, D. Van Tuan, D. Smirnov, X. Marie and H. Dery, Phys. Rev. B, 2020, 101, 115307 CrossRef CAS
.
- P. Lasch, A. Hermelink and D. Naumann, Analyst, 2009, 134, 1162–1170 RSC
.
- M. Selig, G. Berghäuser, A. Raja, P. Nagler, C. Schüller, T. F. Heinz, T. Korn, A. Chernikov, E. Malic and A. Knorr, Nat. Commun., 2016, 7, 1–6 Search PubMed
.
- D. Christiansen, M. Selig, G. Berghäuser, R. Schmidt, I. Niehues, R. Schneider, A. Arora, S. M. de Vasconcellos, R. Bratschitsch and E. Malic, Phys. Rev. Lett., 2017, 119, 187402 CrossRef PubMed
.
- Z. Li, T. Wang, C. Jin, Z. Lu, Z. Lian, Y. Meng, M. Blei, M. Gao, T. Taniguchi and K. Watanabe, ACS Nano, 2019, 13, 14107–14113 CrossRef CAS PubMed
.
- P. Mohapatra, S. Deb, B. Singh, P. Vasa and S. Dhar, Appl. Phys. Lett., 2016, 108, 042101 CrossRef
.
- P. K. Mohapatra, K. Ranganathan and A. Ismach, Adv. Mater. Interfaces, 2020, 7, 2001549 CrossRef CAS
.
- W. Kohn and L. J. Sham, Phys. Rev., 1965, 140, A1133 CrossRef
.
- J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed
.
- P. Giannozzi, O. Andreussi, T. Brumme, O. Bunau, M. B. Nardelli, M. Calandra, R. Car, C. Cavazzoni, D. Ceresoli and M. Cococcioni, J. Phys.: Condens. Matter, 2017, 29, 465901 CrossRef CAS PubMed
.
- J. Deslippe, G. Samsonidze, D. A. Strubbe, M. Jain, M. L. Cohen and S. G. Louie, Comput. Phys. Commun., 2012, 183, 1269–1289 CrossRef CAS
.
- M. S. Hybertsen and S. G. Louie, Phys. Rev. Lett., 1985, 55, 1418 CrossRef CAS PubMed
.
- H. Felipe, D. Y. Qiu and S. G. Louie, Phys. Rev. B, 2017, 95, 035109 CrossRef
.
- S. Ismail-Beigi, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 233103 CrossRef
.
- M. Rohlfing and S. G. Louie, Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 62, 4927 CrossRef CAS
.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d1mh01186c |
‡ These authors contributed equally. |
|
This journal is © The Royal Society of Chemistry 2022 |
Click here to see how this site uses Cookies. View our privacy policy here.