Open Access Article
Tsung-Hsien Leea,
Sheng-Lung Choub,
Tzu-Ping Huangb,
Chak-Ming Liub,
Chih-Hao Chinb,
Meng-Yeh Linb,
Hui-Fen Chen*c and
Yu-Jong Wu
*ab
aDepartment of Applied Chemistry and Institute for Molecular Science, National Yang Ming Chiao Tung University, Hsinchu 30010, Taiwan
bNational Synchrotron Radiation Research Center, 101 Hsin-Ann Road, Hsinchu Science Park, Hsinchu 30076, Taiwan. E-mail: yjwu@nsrrc.org.tw
cDepartment of Medicinal and Applied Chemistry, Kaohsiung Medical University, 100, Shih-Chuan 1st Road, Kaohsiung 80708, Taiwan. E-mail: hfchen@kmu.edu.tw
First published on 6th October 2025
We report a comprehensive photoluminescence (PL) and photoluminescence excitation (PLE) study of monolayer MoS2 and its van der Waals heterostructures with hBN and graphene under deep-ultraviolet (DUV) excitation. Using synchrotron-based VUV/UV spectroscopy, we reveal that while pristine MoS2 exhibits only A-exciton emission at ∼660 nm under visible excitation, broadband near-infrared emission (750–900 nm) emerges at cryogenic temperatures under DUV excitation in MoS2/hBN and MoS2/graphene heterostructures. This emission indicates a nonlocal excitation–emission mechanism facilitated by interfacial energy transfer from the UV-absorbing layers. In MoS2/hBN, a broad UV band near 350 nm also appears under 200 nm excitation and is attributed to impurity-related defect luminescence in hBN. The interfacial processes are governed by temperature-sensitive radiative channels involving defect-bound states or localized excitons in MoS2. Our results highlight the crucial role of interlayer coupling and spectral sensitization in enabling new radiative pathways in 2D heterostructures, offering novel strategies for tailoring light emission in layered optoelectronic systems.
To address these limitations, various strategies have been developed to enhance and modulate the PL response of MoS2, including plasmonic coupling, defect passivation, dielectric environment tuning, and heterostructure engineering. Among these, constructing vertical van der Waals heterostructures with other 2D materials has proven particularly effective in modifying excitonic behavior, charge dynamics, and interfacial energy transfer.19–29 For instance, graphene, with its high carrier mobility and gapless electronic structure, can efficiently quench MoS2 emission via interlayer charge transfer.30–35 In contrast, hexagonal boron nitride (hBN), a wide-bandgap (∼6 eV), atomically flat insulator, offers a clean dielectric interface that preserves or enhances intrinsic PL of MoS2 by suppressing nonradiative pathways.36–40
Beyond the visible regime, deep-ultraviolet (DUV) excitation offers access to higher-energy transitions in MoS2 and its heterostructures, potentially enabling new radiative channels not observable under conventional visible excitation. However, the PL behavior of MoS2-based heterostructures under vacuum ultraviolet (VUV) excitation remains largely unexplored, particularly with respect to interfacial energy transfer and defect-assisted emission at low temperatures. In this study, we investigate the photophysical response of monolayer MoS2 and its heterostructures with hBN and graphene under synchrotron-based VUV/UV excitation. Using temperature-dependent PL and photoluminescence excitation (PLE) spectroscopy, we reveal the emergence of broadband near-infrared emission in MoS2-based heterostructures upon deep-UV excitation. Our results show that this emission arises from an interlayer excitation–emission decoupling process, facilitated by energy transfer from hBN or graphene into MoS2. These findings offer new insights into interfacial energy transfer mechanisms and highlight a novel route for controlling the optical response of MoS2 through far-UV sensitization and heterostructure engineering.
Photoabsorption and photoluminescence measurements were carried out at beamline BL03 of the Taiwan Light Source (TLS-BL03).41–45 Synchrotron radiation was dispersed by a monochromator to provide photon energies ranging from 4 to 40 eV. The incident photon flux was monitored using a gold mesh (∼90% transmission) and recorded with a Keithley 6512 electrometer. A LiF or quartz window was employed as a cutoff filter to eliminate unwanted high-order high-energy photons from the beam.
Photoabsorption spectra were obtained by scanning photon energies over the 350–105 nm range with a 1 nm bandwidth. The VUV/UV light was transmitted perpendicularly through the thin film sample and subsequently detected via luminescence conversion using a sodium salicylate-coated window. The converted luminescence signal was collected by a photon-counting photomultiplier tube (PMT). The beam spot size was approximately 2 × 2 mm2, allowing for reference signal (I0) acquisition by shifting the beam to an uncoated region of the same optical window. Photoluminescence (PL) measurements were performed using the synchrotron radiation from TLS-BL03 as the excitation source. Emission signals were dispersed by a monochromator (Jobin-Yvon iHR320, 1 nm resolution) and detected by a photon-counting PMT (Hamamatsu R943-02). Photoluminescence excitation (PLE) spectra were recorded by monitoring the emission intensity at a fixed wavelength while scanning the excitation photon energy. Both PL and PLE spectra were corrected for the spectral response of the detection system and the excitation source. The temperature-dependent PL measurements were carried out using a closed-cycle helium cryostat system (ARS, DE-202) equipped with optical access and a temperature controller (Lakeshore 331), allowing precise regulation from 10 K to 300 K. The excitation power density during VUV-PL measurements was estimated to be ∼1 × 1012 photons per s per cm2 based on beamline calibration, and photobleaching behavior under prolonged exposure is attributed to UV-induced formation of sulfur vacancies in MoS2.
Additional PL measurements under visible-light excitation were performed using a commercial spectrometer (FS5, Edinburgh Instruments). Raman spectra were collected using a home-built confocal micro-Raman spectroscopy system equipped with a 532 nm DPSS laser (50 mW), a monochromator (iHR320, HORIBA) with a 1200 lines per mm diffraction grating, and a thermoelectrically cooled CCD detector (Newton DU970P-BVF, ANDOR). The system provided a spectral resolution of 1.5 cm−1 around 550 nm. The atomic force microscope (AFM) images were acquired with an AFM (NTEGRA II, NT-MDT) in contact mode under ambient condition. A silicon cantilever (NSG30, YFQ Technology Co., Ltd) with an Au-coated reflective side and a nominal spring constant of 22 N m−1 was used. The line scan rate was 0.4 Hz. Images were acquired over a 20 × 20 μm2 area. The spatial resolution along the X and Y-axes was 4.6 nm and 0.67 nm along the Z-axis.
In the MoS2/hBN heterostructure, both the E12g and A1g modes exhibited slight shifts to 385.9 and 405.9 cm−1, respectively, resulting in a frequency separation of 20.0 cm−1. This value closely matches that of monolayer MoS2, suggesting that the MoS2 film maintains its monolayer nature when supported on hBN. The relatively small lattice mismatch and excellent lattice compatibility between MoS2 and hBN, both possessing hexagonal crystal symmetry, effectively minimize interfacial strain during the transfer process.38,40 Additionally, Raman signals from the hBN layer were not observed, likely due to the intrinsically weak Raman activity of single-layer hBN.44 However, complementary UV absorption measurements confirmed the presence of hBN in the heterostructure, as discussed in Section 3.2. In contrast, the MoS2/graphene heterostructure exhibited additional Raman features characteristic of graphene, including the G band (∼1589 cm−1) and the 2D band (∼2700 cm−1). The observation of the weak D band (∼1350 cm−1) indicates that the graphene layer exhibits less defects. The E12g and A1g modes of MoS2 in this heterostructure were observed at 384.9 and 405.8 cm−1, respectively, resulting in a frequency separation of 20.9 cm−1. This separation is consistent with monolayer MoS2, suggesting that both the MoS2 and graphene layers in the heterostructure are single layers.
In addition to the characteristic first-order E12g and A1g modes of MoS2 near 380 and 405 cm−1, prominent Raman features are observed in the 500–800 cm−1 region for the MoS2/hBN and MoS2/graphene heterostructures. These additional peaks correspond to second-order and combination modes such as 2LA(M), E12g + LA(M), and A1g + LA(M), which are notably enhanced under 532 nm excitation. Although monolayer MoS2 typically exhibits weak second-order Raman bands, their relative intensity becomes significantly stronger in the heterostructures. This enhancement may arise from resonance Raman scattering, where the 532 nm excitation lies near the A-exciton absorption of MoS2 (∼1.88 eV), thereby increasing the scattering cross-section for multi-phonon processes. Moreover, the presence of atomically thin hBN or graphene beneath the MoS2 layer can modify the local dielectric environment and introduce interfacial strain or phonon coupling, which may further promote phonon-assisted scattering pathways. The observation of these pronounced high-order modes suggests strong exciton–phonon interactions and underscores the influence of interfacial effects in modifying the vibrational response of 2D heterostructures. On the other hand, the reduced frequency separation (Δ) = 20.9 cm−1 in MoS2/graphene, compared to monolayer MoS2, arises from the interplay of charge transfer and strain effects. The interfacial electron transfer that selectively softens A1g, and modest tensile strain that shifts E12g more strongly than A1g. In MoS2/hBN, where charge transfer is strongly suppressed, (Δ) ≈ 20.0 cm−1 instead reflects near-intrinsic monolayer behavior. These results align with previous reports on doping- and strain-induced Raman shifts in MoS2.32,48,49
AFM measurements were performed on the transferred MoS2 thin films deposited on MgF2, as well as on MoS2/hBN and MoS2/graphene heterostructures, as shown in Fig. 2. The topography scans (20 μm × 20 μm) reveal continuous coverage across the scanned areas, with only minor corrugations and transfer-related wrinkles that are commonly observed in mechanically laminated 2D films. The root-mean-square (RMS) surface roughness values are 1.41 nm for MoS2/MgF2, 1.40 nm for MoS2/hBN/MgF2, and 1.58 nm for MoS2/graphene/MgF2, respectively. These values confirm that the films are smooth and laterally uniform, with no evidence of cracks or large voids across the probed regions. Given that the optical excitation beam diameter (∼2 mm) is much larger than the AFM scan window, the absence of large-area defects ensures that the probed regions in our spectroscopic measurements are representative of well-formed van der Waals heterointerfaces. Because the commercial films were supplied on PMMA-coated carrier substrates, it was not possible to perform Raman characterization of the as-grown MoS2 prior to transfer, nor to capture a clear step-height profile for thickness verification in our AFM scans. Nevertheless, the observed morphology is fully consistent with monolayer MoS2 coverage and supports the conclusion that the anomalous Raman mode separation observed on MgF2 arises from substrate-induced strain and interfacial coupling, rather than from multilayer contamination.
![]() | ||
| Fig. 2 AFM topography (20 μm × 20 μm) of transferred MoS2 thin films on (A) MgF2, (B) hBN/MgF2, and (C) graphene/MgF2 substrates. | ||
Overall, the Raman and AFM analysis confirm the successful fabrication of high-quality MoS2/hBN and MoS2/graphene heterostructures while preserving the crystallinity of the MoS2 films. The observed variations in peak positions and frequency separations reflect the subtle influence of different substrates and heterointerfaces on the vibrational and electronic properties of the MoS2 layers.
![]() | ||
| Fig. 3 Absorption spectra of MoS2, MoS2/graphene, and MoS2/hBN thin films on MgF2 substrates at 295 K. | ||
In the deep-UV region (125–300 nm), the absorption increases steadily without distinct peaks,52,53 consistent with transitions to higher-lying conduction band states.50 These states lie well above the quasiparticle gap and involve more delocalized interband transitions, possibly extending to the continuum. The QSGW calculations show that these higher-energy states are separated from the K-point conduction band minimum by several electron volts and thus are only accessible with deep-UV excitation sources.50 Importantly, the presence of substantial absorption at wavelengths below 300 nm (∼4.1 eV and above) provides an energetic window for exciting hot carriers or higher-lying excitonic states, which are often not probed in conventional visible-light photoluminescence studies. These VUV transitions set the stage for the observation of broadband emission under deep-UV excitation, as will be discussed later.
Photoluminescence measurements under different excitation conditions revealed a strong dependence on both temperature and photon energy, as shown in Fig. 4. Under room-temperature excitation with visible light (e.g., 430 nm), a sharp emission band centered around 670 nm was detected, corresponding to the radiative recombination of the A-exciton. The PLE spectrum monitored at 670 nm shows a broad response from 400 to 500 nm, consistent with A- and B-exciton resonances and the C-exciton absorption shoulder.11,12,14,15 No additional emission was observed when excitation wavelengths below 300 nm were used at room temperature, indicating that high-energy excitation alone is insufficient to produce radiative recombination under thermalized conditions. However, under cryogenic conditions (10 K), excitation with UV light at 260 nm and 300 nm induced a distinct, broad emission band arising near 690 nm, extending toward 900 nm. This broadband luminescence is fundamentally different from the A-exciton emission observed under visible excitation. It cannot be reproduced by any excitation wavelength in the visible or UV range at room temperature. To further elucidate the nature of the broadband emission observed under VUV excitation, temperature-dependent PL measurements were conducted from 10 K to 300 K under 300 nm excitation, as shown in Fig. 5. At 10 K, the emission was most intense, but as the temperature increased, the PL intensity rapidly decreased and was nearly quenched above 250 K. This thermal quenching behavior suggests that the broadband luminescence originates from defect-bound states or localized excitons that become thermally delocalized at elevated temperatures. The activation of nonradiative recombination pathways and thermal escape of carriers from shallow defect traps at higher temperatures effectively suppress the radiative recombination channels responsible for this emission.54
![]() | ||
| Fig. 5 PL spectra of MoS2 thin films on MgF2 substrates at various temperatures under 300 nm excitation. | ||
The emergence of broadband emission under UV excitation and low temperature is attributed to defect-assisted recombination and high-energy carrier relaxation processes. Excitation at energies above ∼4.1 eV may promote electrons to higher conduction bands, followed by phonon scattering or trapping at sub-gap defect states, which enables radiative recombination pathways not otherwise accessible. At low temperatures, nonradiative decay channels are suppressed, and thermal detrapping is reduced, allowing these defect-related luminescence channels to dominate. The absence of this emission at room temperature is likely due to thermal escape of carriers from shallow traps and activation of nonradiative recombination channels, as commonly reported in temperature-dependent studies of MoS2.54–56 The thermal quenching of broadband emission is attributed to thermally activated delocalization of carriers from shallow defect states, enhanced phonon-assisted nonradiative decay, and increased carrier–phonon scattering at elevated temperatures. These processes reduce the population of radiatively active localized states and suppress defect-assisted emission, consistent with previous studies on low-temperature PL behavior in MoS2. In contrast, pure MgF2 substrates under identical excitation conditions do not exhibit such broadband emission, confirming that the observed luminescence originates from the MoS2 layer itself rather than the substrate or its intrinsic defects. The broadband emission is attributed to radiative recombination via defect-bound states or localized excitons stabilized at cryogenic temperatures. While the spectral shape and thermal behavior strongly suggest defect-assisted processes, further investigation (e.g., time-resolved PL or low-temperature near-field imaging) is needed to definitively distinguish between static defect trapping and confinement-induced localization.
Additionally, we observed that prolonged exposure of the MoS2 thin films to ultraviolet excitation leads to a gradual decrease in PL intensity over time. The observed photobleaching behavior under continuous UV excitation may be associated with the photoinduced desorption of sulfur atoms from the MoS2 lattice. The photon energies used in this study, particularly below 300 nm, are sufficient to break Mo–S bonds, potentially leading to the formation of sulfur vacancies. These defect sites can initially enhance defect-related emission,24,29 but may subsequently serve as nonradiative recombination centers as their density increases, thereby reducing the overall PL intensity.
In the deep-UV region (∼200 nm and below), the MoS2/hBN and MoS2/graphene samples show significant increases in absorbance compared to pristine MoS2. This enhancement is not solely attributed to MoS2 but rather originates from the intrinsic optical transitions of the constituent hBN and graphene layers. In our previous work, monolayer hBN exhibits a strong absorption band at ∼6.1 eV (∼203 nm),44 corresponding to the π → π* transition near the K point of the Brillouin zone, which is associated with direct excitons in monolayer hBN. Similarly, the enhanced deep-UV absorbance in the MoS2/graphene heterostructure can be partly attributed to the π → π* transitions of graphene, which exhibit broad absorption features extending into the deep-UV regime due to its semi-metallic band structure with a vanishing density of states at the Dirac point.42,43 It is important to note that the observed increase in absorbance below 220 nm may also be partially influenced by increased light scattering due to surface roughness or non-uniform stacking in the heterostructures. This possibility arises from the imperfect planar interfaces formed during the layer-transfer process.
To explore the influence of supporting 2D materials on the optical response of MoS2, we examined the PL and PLE spectra of MoS2/hBN and MoS2/graphene heterostructures at 300 and 10 K. As comparison of PL spectra upon excitation with 430 nm in Fig. 4, pristine monolayer MoS2 exhibits the highest PL intensity centered around 660 nm. When interfaced with graphene, the PL intensity is markedly quenched, suggesting efficient nonradiative energy or charge transfer from MoS2 to graphene due to high carrier mobility and zero bandgap of graphene.31,32,35 In contrast, the MoS2/hBN heterostructures maintains a PL intensity comparable to or slightly higher than pristine MoS2, indicating minimal interfacial quenching. Upon deep-UV excitation, as shown in Fig. 6, both heterostructures exhibit broadband near-infrared PL centered above 750 nm when excited at short wavelengths, in contrast to pristine MoS2, which shows no detectable emission under the same excitation conditions. This broadband luminescence is most prominent in the MoS2/hBN sample when excited at 200 nm, and also appears in MoS2/graphene under 260 nm excitation. Notably, no emission is observed from the underlying MgF2 substrate under any conditions. Fig. 5 depicts the corresponding PLE spectra, monitored at 750–785 nm emission wavelengths, further confirming that the excitation profiles of the heterostructures differ markedly from that of MoS2 alone. In the MoS2/hBN sample, a sharp PLE peak near 200 nm is observed, consistent with the known deep-UV absorption band of monolayer hBN.44 Similarly, the MoS2/graphene sample shows enhanced excitation response around 260 nm, aligning with the strong absorption of graphene in the far-UV regime.43 These results clearly indicate that the deep-UV excitation leading to broadband emission in MoS2 is enabled only when a second absorbing 2D layer (hBN or graphene) is present. Under the same excitation, pristine MoS2 does not exhibit this emission, implying that the heterostructure architecture plays a critical role in facilitating this process.
In addition to low-temperature results, PL spectra were also acquired at room temperature to evaluate the thermal stability of the broadband emission in MoS2-based heterostructures. As shown in Fig. 7, the broad emission band centered around 350 nm observed upon 200 nm excitation is attributed to defect-related luminescence from the hBN layer. This emission likely arises from impurity-induced states, such as those associated with oxygen or carbon atoms incorporated during the CVD growth process.57 These defects form donor–acceptor pairs that facilitate radiative recombination at energies significantly lower than the intrinsic excitonic emission of hBN. The absence of this band under longer-wavelength excitation further supports its origin from deep-level states that require high-energy photons for activation. The emission intensity gradually decreases with longer excitation wavelengths and becomes negligible beyond 300 nm. This indicates that excitation near the π–π* transition of hBN (∼6.2 eV) remains an effective pathway to generate broadband luminescence from the MoS2 layer even at ambient conditions, albeit at a reduced efficiency compared to 10 K.
![]() | ||
| Fig. 7 PL spectra of MoS2/hBN thin films on MgF2 substrates at 300 K under excitation with various wavelengths. | ||
Similarly, in the MoS2/graphene sample (Fig. 8), detectable broadband emission is observed at room temperature under deep-UV excitation. Excitation at 220–130 nm yields a relatively flat emission profile, with intensity centered in the 400–800 nm range. The signal strength diminishes with increasing excitation wavelength and becomes negligible at 300 nm. This suggests that high-energy excitation of graphene's π-electron system can still trigger emissive states in the MoS2 layer, although with significant thermal quenching compared to cryogenic measurements. Although MoS2-based heterostructures exhibit broadband emission features under deep-UV excitation, the PLQY measured at 430 nm excitation shows no significant enhancement compared to pristine MoS2. This is likely due to the absorption being dominated by the top MoS2 layer, limiting interlayer excitation dynamics, and the instrumental cutoff of the commercial spectrometer beyond 850 nm, which restricts detection of the far-IR emission components.
![]() | ||
| Fig. 8 PL spectra of MoS2/graphene thin films on MgF2 substrates at 300 K under excitation with various wavelengths. | ||
Additionally, from the Raman spectra (Fig. 1), we evaluated the wavenumber positions of the E12g and A1g modes across the sample and constructed a correlative plot to extract strain and doping values. Using the linear decoupling method described in previous studies,58–60 we find that the MoS2 in the heterostructures exhibits a tensile strain of approximately 0.20–0.25% and an n-type doping level of (0.9–1.1) × 1012 cm−2. In contrast, we also used the shift of the A exciton peak (Fig. 4) to estimate the biaxial strain. However, strain estimates derived from A exciton PL peak shifts (∼0.05–0.07%) are lower than those obtained from Raman analysis (∼0.20–0.25%), likely due to differences in how these two techniques respond to local doping, substrate dielectric screening, and exciton binding energy. Such variation is consistent with previous reports where PL and Raman diagnostics yield complementary but non-identical strain values in monolayer TMDCs.
Overall, while both MoS2/hBN and MoS2/graphene heterostructures retain their ability to produce broadband PL under vacuum ultraviolet excitation at room temperature, the emission intensity is notably reduced due to increased nonradiative losses. These results reinforce the role of high-energy absorbing layers (hBN or graphene) in sensitizing the MoS2 emission response beyond its intrinsic excitation range, and highlight the temperature dependence of this nonlocal excitation–emission process.
In the MoS2/hBN heterostructure, the broadband emission is efficiently triggered by ∼200 nm excitation, resonant with the π–π* transitions of hBN centered at ∼6.1 eV. Similarly, in the MoS2/graphene system, the emission is most pronounced under ∼260 nm excitation, aligning with the broad π-electron absorption continuum of graphene. These observations imply an excitation–emission decoupling mechanism, where the absorption of high-energy photons occurs in one material (donor), and the emission is generated in another (acceptor).
Two primary mechanisms may account for this interfacial coupling: nonradiative energy transfer and carrier tunneling.61–65 Nonradiative Förster resonance energy transfer (FRET) can occur between the UV-absorbing layer (hBN or graphene) and MoS2 if the donor emission overlaps with the acceptor absorption and the separation is within the Förster radius (∼1–10 nm). Alternatively, photoexcited hot carriers or excitons in hBN or graphene may tunnel across the interface into localized or defect-related states in MoS2, followed by radiative recombination. At cryogenic temperatures, these defect-bound excitons become stabilized due to suppressed thermal delocalization, resulting in strong broadband luminescence. This emission rapidly quenches at higher temperatures (above ∼250 K), consistent with thermally activated nonradiative decay and carrier escape from shallow trap states. The higher emission intensity in MoS2/hBN compared to MoS2/graphene is likely due to the atomically smooth, insulating nature of hBN, which minimizes interfacial electronic dissipation and promotes longer exciton lifetimes.
The contrasting effects of hBN and graphene on the photoluminescence behavior of MoS2 originate from their fundamentally different electronic structures. hBN is a wide-bandgap (∼6.1 eV) insulator with no available electronic states near the Fermi level or the MoS2 conduction/valence bands. As such, it suppresses interlayer charge transfer, preserving the radiative recombination in MoS2. This insulating character allows energy absorbed in hBN (e.g., through π–π transitions) to be nonradiatively transferred into MoS2, populating localized or defect-bound states that emit in the near-infrared under cryogenic conditions.
In contrast, graphene exhibiting the gapless electronic structure and high density of available states around the Fermi level enable ultrafast interlayer charge transfer. Under visible excitation, this leads to quenching of the MoS2 exciton emission. However, under deep-UV excitation, graphene absorbs high-energy photons and can generate hot carriers, some of which may tunnel across the interface into MoS2. These carriers may become trapped at defect sites, giving rise to broadband luminescence. Compared to hBN, the PL intensity in MoS2/graphene is weaker due to competing nonradiative dissipation in the conductive graphene layer.
The findings demonstrate that heterostructure assembly can be used to activate otherwise inaccessible optical channels in MoS2. By tailoring the spectral absorption properties of adjacent 2D layers and optimizing the interfacial contact, excitation–emission decoupling can be employed to extend the functional spectral range of TMDCs. This principle offers a new paradigm for engineering light–matter interactions in low-dimensional systems, enabling the design of devices where absorption and emission are spatially and energetically separated.
Future work could focus on time-resolved spectroscopy to further unravel the ultrafast dynamics of the energy transfer processes, as well as theoretical modeling of interlayer exciton formation in MoS2-based heterostructures. Moreover, the role of interfacial strain, dielectric screening, and stacking order in modulating the radiative behavior warrants further exploration. Ultimately, these results open new avenues for designing hybrid 2D systems with tunable optical functionalities driven by van der Waals interface physics.
| This journal is © The Royal Society of Chemistry 2025 |