Open Access Article
Marleen Hußmann
a,
Mira Kreßler
b,
Patryk Kusch
*b and
Siegfried Eigler
*a
aInstitute of Chemistry and Biochemistry, Freie Universität Berlin, 14195 Berlin, Germany. E-mail: siegfried.eigler@fu-berlin.de
bInstitute of Physics, Freie Universität Berlin, 14195 Berlin, Germany. E-mail: patryk.kusch@fu-berlin.de
First published on 4th November 2025
In the design of nanoscale materials, hybrid van der Waals heterostructures that integrate the excitonic landscape of atomically thin transition metal dichalcogenide (TMDC) semiconductors with molecular electric dipoles offer enhanced control over light–matter interactions and charge carrier dynamics. Even minor deviations in homogeneity can profoundly affect their optoelectronic properties and, consequently, device performance, necessitating stringent quality control capable of probing structural and compositional divergences down to the nanoscale. However, the reliable characterization of such complex, multilayered systems, remains challenging due to the interplay of chemical, structural, and optical inhomogeneities across different length scales. In this study, we examine a trilayer heterostructure consisting of chemical vapor deposition (CVD) graphene (G), a self-assembled layer of Rhodamine 6G (R6G), and a transferred monolayer MoS2 (G/R6G/MoS2), incorporating regions of a tri- and multilayer MoS2 as well. Comprehensive structural and optical characterization was performed to identify possible inhomogeneities, employing photoluminescence (PL) spectroscopy, Raman spectroscopy, Kelvin probe force microscopy (KPFM), and scattering-type scanning near-field optical microscopy (s-SNOM). Analytical methods indicate that the TMDC layer has almost uniform molecular coverage and preserved crystallinity. Importantly, near-field optical imaging demonstrates the propagation of exciton-polaritons in MoS2, with a clear redshift of the polariton wavelength upon R6G integration, signifying substantial modulation of the local dielectric environment and excitonic response. These findings underscore the tunability of hybrid 2D molecular–inorganic interfaces and their promise for advanced applications in nanophotonic devices, excitonic circuitry, and quantum optoelectronics.
In addition, molecules can form monolayer structures as well, once they self-align (self-assembling molecules)14 or face an external force.15 By combining these molecular monolayers and TMDCs new possibilities for controlling and modifying physical and chemical phenomena at the atomic scale becomes possible by tailoring interfacial electronic states, engineering new quantum effects, and enhancing catalytic activity.16–19 Herein, molecular layers can introduce dipole fields that shift the electronic band alignment of TMDCs, influence exciton recombination dynamics, and modulate spin–orbit interactions.20 Such engineered heterostructures hold promise for applications for fundamental research and technological applications in areas such as energy harvesting, optoelectronics, quantum computing and neuromorphic devices.21 Moreover, incorporating molecular films into heterostructures like G/MoS2 enhance even further their complexity, chemical and electronic tunability, enabling precise control over electronic, optical, and interfacial properties (i.e. charge transfer,22 dipole interactions,23 and surface reactivity24).25
However, the fabrication of these hybrid systems remains a significant challenge, requiring high control over molecular deposition, interfacial cleanliness, and stability under ambient conditions. Achieving uniform and reproducible molecular coverage on TMDC surfaces while preserving the intrinsic properties of both components is challenging and crucial for realizing their full potential. Indeed, molecular layers can undergo structural rearrangements26 or electronic modifications upon interaction with the TMDC substrate,27 further complicating the design and fabrication of these systems. As challenging as the fabrication itself, is their characterization and hence understanding of such mutual influence between the molecular monolayer and the TMDC for optimizing their functional properties. The interaction at the interface can significantly alter charge transfer dynamics, exciton lifetimes and the local electronic structure, raising the question about advanced characterization techniques for their detailed investigation.28
In our approach, we address these challenges by introducing rhodamine 6G (R6G), a well-characterized laser dye, as an additional molecular layer between graphene (G) and MoS2 that form a trilayered system (G/R6G/MoS2, Fig. 1A). The sample was produced by dip-coating CVD graphene monolayer in a solution of R6G with subsequent mechanical transfer of MoS2 on top (Fig. 1B). We verify the formation of a homogenous R6G film and demonstrate the successful transfer of the TMDC layer using complementary micro- and nanoscale optical and spectroscopic experiments. Photoluminescence (PL) and Raman spectroscopy (Raman) provide the standard microscopic information. Kelvin probe force microscopy (KPFM) supported by scanning near-field optical microscopy (SNOM) amplifies the nanoscopic field, providing much more detailed and comprehensive insights into the prepared heterostructure. To emphasize the capabilities of such structures, we present an image showing propagating polaritons within the heterostructure in real space. Compared to polaritons in bare MoS2 a change of the polariton wavelength is revealed, underlining the impact of the present molecular film.
The schematical fabrication process is illustrated in Fig. 1B. R6G was selected for its advantageous properties, including its absorption and emission range, high quantum yield, remarkable stability, commercial availability and cost efficiency.32,33 The maximum of the absorption of R6G32 at 530 nm overlaps perfectly with the standard green laser (532 nm) leading to strong photoluminescence32 (PL) from 530–630 nm (2.3–2.0 eV), which does not overlap with the PL of monolayer MoS2
34 from 600–750 nm (2.07–1.65 eV). At the same time the R6G absorption is energetically higher enabling a channel for electron transfer from R6G to MoS2. Thus, R6G serves as an efficient charge mediator in the heterostructure due to its ability to facilitate rapid energy and charge transfer at the interface with both graphene and MoS2.12,22,35 Its partial two-dimensional nature ensures seamless integration with MoS2, forming clean van der Waals interfaces essential for efficient optoelectronic coupling.35 Additionally, the strong π–π interactions between graphene and R6G enhance dye adsorption, promoting the formation of homogenous molecular layers and effective fluorescence quenching.36
CVD-grown graphene was chosen over mechanically exfoliated graphene due to its scalable production of high-quality, defect-free films, confirmed by the absence of the D-peak in Raman spectra (Fig. S1).37 Growth was optimized using a gas mixture of Ar, H2, and CH4, followed by CO2 etching and a dry transfer method using a polydimethylsiloxane (PDMS) stamp to minimize residues and improve cleanliness.38,39 Key parameters such as the cooling rate while growing or applied pressure during transfer significantly influenced crack formation and surface integrity. The graphene was transferred onto Si/SiO2 substrates using an x,y,z-micromanipulator, followed by annealing under vacuum at 130 °C to remove water and PDMS residues. R6G was adsorbed from a 1 µmol L−1 aqueous solution for 30 s and rinsed. For the top layer, mechanically exfoliated MoS2 flakes were used to ensure high purity and to enable precise positioning on selected graphene regions. We note that it is challenging to alignment layers and reach transfer consistency due to PDMS elasticity and pressure sensitivity, for details see SI. The final heterostructure is presented in Fig. 1, visualized by optical microscopy, and atomic force microscopy (AFM). Due to the fabrication process, graphene decorated with R6G covers the entire AFM-imaged area. The thickness of the R6G layer is around 4 nm, estimated from AFM topography cross section at the marked area (2) in Fig. 1C and E, which is at the crossover of a hole, where graphene together with R6G is missing, and R6G covered graphene subtracting a typical graphene height of 1 nm.40 The transferred MoS2 exhibits regions of varying thickness—monolayer, trilayer, and multilayer—which were identified and confirmed through Raman spectroscopy and AFM height analysis. Given the inherent complexity and multi-component nature of the G/R6G/MoS2 heterostructure, a comprehensive understanding of its optical and electronic interactions is crucial. Each layer in the heterostructure can be individually analyzed through its distinct spectroscopic signatures. This leads to intriguing energetic and optical characteristics that arise from the combined yet distinct properties of each individual component within the heterostructure. Spatially resolved characterization techniques such as Raman spectroscopy and photoluminescence (PL) mapping are employed to elucidate the intricate local variations within the heterostructure. This lays the groundwork for further exploration of its functional properties.
In the G/R6G/MoS2 (G/R6G/MoS2) heterostructure, three main areas are differentiated for further interpretation, which differ in the thickness of MoS2, G/R6G/monolayer MoS2 (G/R6G/1L-MoS2), G/R6G/trilayer MoS2 (G/R6G/3L-MoS2) and G/R6G/bulk MoS2 (G/R6G/bulk-MoS2). Areas of those heterostructures are directly visible in the optical microscope images and confirmed by AFM height measurements (Fig. 1D). For comparison the surrounding structure of simple G/R6G is analyzed, too. Representative PL spectra of each heterostructure (zero to bulk layer MoS2, Fig. 2B) convey their PL evolution (2.33 eV to 1.7 eV) within the sample emitting in the visible range. By additional pattern control of shape and design of the sample, the distinct PL signals of each layer concerning their intensity as well as their shift may serve as unique optical output signal by using only one wavelength (532 nm) switching from a green (R6G) to red (1L-MoS2) or completely suppressed emission (bulk MoS2).
Here, G/R6G (green) is dominated by the broad PL of R6G at 2.26 eV maximum. The heterostructure G/R6G/MoS2 exhibits a strong PL signal arising from the MoS2. The introduction of the intermediate R6G layer mitigates this quenching through energy transfer, partial electronic decoupling between graphene and MoS2, and furthermore induced n-type doping to R6G. This results in a 6-fold increased MoS2 PL response within the heterostructure compared to values of G/MoS2,41 and 3-fold increased PL compared to bare MoS2. This behavior indicates that quenching is effectively suppressed, effectively decoupling MoS2 electronically from graphene (Fig. 2A). Additionally, a blueshift is observed, resulting from the n-type doping to R6G, by reducing the number of bound electrons in trions. The extent of these effects depends on the molecular alignment, thickness, and spatial position of R6G, as well as its interaction with the MoS2. The heterostructure, further, shows increasing PL influence of MoS2 with reduced layer thickness and concurrently a strong decrease of PL of R6G. The layer-dependent population shift from the trion at 1.85 eV (bulk MoS2) to exciton at 1.89 eV (monolayer MoS2), as marked in Fig. 2B, suggests a most efficient electron transfer to R6G in the heterostructure with monolayer MoS2.
Raman spectroscopy allows the characterization of the individual components of the heterostructure, as well as identifying inhomogeneities, strain and doping by recording the G (∼1580 cm−1) and 2D (∼2700 cm−1) peaks of the CVD-grown graphene and Raman modes of MoS2 are the E2g (∼385 cm−1) and A1g (∼405 cm−1) modes for multilayer and A′1, E′ for monolayers.42–44 Those Raman modes are termed as E2g and A1g in this manuscript. Exemplary MoS2 spectra of the different areas of the heterostructure are shown in Fig. 2C. Spectra of the bare components are in the SI (Fig. S1). The Raman signals of R6G are located between 600 and 1650 cm−1.45 Those Raman signals of R6G are overshadowed in the G/R6G as well as the heterostructure by its strong PL, which appears around 2.25 eV when using a 532 nm excitation source. This suggests that no significant electron transfer (i.e., PL quenching) occurs to the CVD graphene, likely due to the applied amount and/or random orientation of R6G molecules.46,47 Also, the defect related Raman mode of graphene (D peak at ∼1350 cm−1) does not arise in the entire heterostructure, which confirms the nonexistence of covalent modification (Fig. S1) or formation of other defects. The Raman modes of MoS2 give information about the number of layers by determining the difference between the A1g and E2g Raman peak positions of MoS2 (Fig. 2C and 3C). In that way, it is confirmed that the upper region (optical image, Fig. 1A and Raman map, Fig. 3C) corresponds to 1L-MoS2 (ΔA1g/E2g = 19 cm−1, Fig. 2C, red), the middle region to 3L-MoS2 (ΔA1g/E2g = 23.5 cm−1, Fig. 2C, rosé), and the lower region to bulk MoS2 (ΔA1g/E2g = 25 cm−1, Fig. 2C, beige).31 Especially in MoS2, both Raman signals E2g and A1g are additionally sensitive to strain and doping, respectively, showing peak broadening, intensity changes or shifts.48,49 As described by Cao et al. for G/MoS2, graphene is p-doped, while for the G/R6G/1L-MoS2 heterostructure the shift to higher frequencies compared to single 1L-MoS2 (Fig. 2C) of about 0.9 cm−1 confirms the above mentioned n-doping of MoS2 to R6G layer as indicated also by the blue-shifted PL of the heterostructure (Fig. 2a).41 Li et al. analyzed the undoped layer dependent Raman evolution of MoS2 with equally broadened distance of A1g and E2g in the trilayer compared to the monolayer.42 However, in G/R6G/3L-MoS2 the E2g is massively shifted of around 3 cm−1 compared to a pure trilayer MoS2 while A1g almost remains, suggesting an increased stacking-induced intralayer changes, which may origin in the neighbored R6G structures.43
![]() | ||
| Fig. 3 Merged 2D Raman maps (each 10 µm × 10 µm) of the integral of (A) MoS2 peak around 385 cm−1, (B) MoS2 peak around 405 cm−1and (C) the peak position difference of A1g − E2g. | ||
From PL and Raman spectra recorded on different areas of the heterostructure, we observe the presence of interesting features, like unexpected intensity modulation and blue shift of MoS2, indicating n-type doping induced by the presence of R6G and confirming the electron transfer. However, questions remain regarding the quality and homogeneity of the molecular R6G layer over the total structure, and how it influences the optical and electronic properties of the heterostructure. Spatially resolved micro-Raman and micro-PL spectroscopy provide powerful tools to investigate these aspects with insights into local strain, doping levels, layer uniformity, and interfacial interactions, enabling a detailed understanding of structure–property relationships within the complete heterostructure.
To obtain an initial estimate of molecular coverage, we measure the PL intensity, which is sensitive to both the thickness and orientation of the molecular layer and provides insight into the influence of R6G on its optical properties.
51 and R6G
33 (∼2.255 eV, Fig. 4C) offer the opportunity to observe spectral changes, which commonly correlate with structural modifications. The specific areas of the PL peaks, not intensities, are plotted as z-coordinate of the mapping, where bright colors correspond to a larger area of Lorentzian fits and thus to stronger signals (Fig. 4).
The distribution of R6G (Fig. 4C) across the monolayer MoS2 highlights the formation of a continuous molecular film approximately 4 nm thickness, with some variations in emission intensity across the sample. These local variations, which correlate with features in the PL maps of A exciton and A− trion (Fig. 4B and A) and the AFM image (Fig. 1C). Minor inhomogeneities likely originate from the formation of small R6G agglomerates, or a localized missing molecular coverage of graphene. Despite these small-scale differences, the overall uniformity of the molecular film underscores the robustness of our fabrication approach. Only a few isolated regions show deviations in PL intensity, as marked in Fig. 4. In the inner part of the marked area (1), structural inhomogeneities are responsible for the diminishing A exciton intensity (darker lines within), which are conversely stronger signals for the A− trion, since merged AFM images (Fig. 1C) show an almost smooth area except for long crossing wrinkles, which are known to redshift PL due to strain by decreasing energy of the direct energy band in monolayer MoS2.52–54 The increase of the R6G signal in the outer part of (1) may be reduced to a loose connection of the MoS2 to the heterostructure in the fabrication process during its transfer. In the marked area (2), previously discussed prominent R6G agglomeration up to 30 nm tall pockets (see Raman maps (Fig. 3), AFM image (Fig. 1C) and microscope picture (Fig. 1A), locally amplify its PL spectra, while MoS2 signals stay unchanged. Remarkably, the A− trion is strongly increased in the monolayer in the marked area (3). By examining the microscope picture, it can be attributed to the missing graphene underneath not quenching the n-doping effect of the bottom molecules to the upper MoS2.55 However, a more detailed interpretation of smaller features like the hole in the graphene lattice, as seen in the optical image (Fig. 1A) and AFM (Fig. 1C), is not interpretable in the macroscopic PL measurements.
Although the discussed inhomogeneities are confined to small areas and can be further minimized by refining the fabrication process, they offer unique opportunities to probe and understand the interfacial optical and electronic properties of the heterostructure in more detail by high-resolution nanoimaging techniques for a deeper exploration of these localized features.
In our measurements, KPFM reveals a remarkably constant CPD across each distinct MoS2 region—monolayer, trilayer, and bulk—confirming the electronic homogeneity of the heterostructure at the microscale (Fig. 5). This uniformity strongly suggests a continuous and overall, even distribution of the R6G molecules beneath the MoS2 as significant variations in R6G thickness or density would lead to observable CPD fluctuations. Further, the results confirm the hypothesis of a loose connection in the marked area (1) of Fig. 4C to be the origin of the PL increase. Only isolated structural features, which effectively alter the height, such as wrinkles in the monolayer MoS2 (marked area (1) in Fig. 4C) or agglomerates of R6G (marked area (2) in Fig. 4C) that extend into the trilayer region as seen in AFM image (Fig. 1D), introduce localized CPD increase. Specifically, wrinkles reaching heights of 25–30 nm are associated with CPD shifts of approximately 200 mV, while R6G agglomerates in the range of 10–15 nm in height result in CPD increases of 50–150 mV. The most pronounced CPD change, approximately 400 mV, is observed in areas where the graphene layer is absent and R6G is trapped in a pocket beneath MoS2. In such configurations, the R6G is no longer electronically coupled to graphene, and the resulting local environment favors more efficient electron removal.
![]() | ||
| Fig. 5 Merged KPFM image with magnification on missing graphene (lower layer) within heterostructure G/R6G/MoS2. | ||
Interestingly, the CPD values in regions with bilayer graphene (Fig. 1A, indicated as 2L) are nearly identical to those with monolayer graphene, suggesting that the electronic coupling between R6G and graphene does not significantly change with increasing graphene thickness. This observation supports the hypothesis of n-type doped graphene by R6G, as previously reported by Yu et al., who correlated the work function of mono- and bilayer graphene with field-effect transistor behavior.58 Together with PL and Raman measurements, the CPD measurements suggest an electron transfer from MoS2 to graphene via R6G. Furthermore, regions consisting of trilayer MoS2 atop the G/R6G stack exhibit CPD values indistinguishable from those measured on G/R6G regions without MoS2, suggesting a reduced influence of the R6G layer on the MoS2 surface potential as its thickness increases. This behavior implies a thickness-dependent screening effect, where the doping and charge transfer effects of R6G become less pronounced with increasing MoS2 layer number.
Altogether, these KPFM results highlight the excellent electronic uniformity of the heterostructure and demonstrate the sensitivity of the technique to subtle structural features. The consistent CPD across each MoS2 thickness region underscores the stable integration of the R6G layer, while localized CPD variations provide valuable insight into nanoscale deviations such as molecular agglomerates or structural irregularities. These findings illustrate the capability of KPFM to resolve and distinguish even minor heterogeneities within an otherwise highly homogeneous and well-structured system.
To support our findings, we also imaged the heterostructure using scattering-type scanning near-field optical microscopy (s-SNOM) of the O2 amplitude in Fig. 6. O3 and O4, the third and fourth function show same but noisier behavior due to lower modulations (see Fig. S2). For experimental details, see the SI. In the visible spectral range, the s-SNOM signal is primarily governed by the local tip–sample interaction and is approximately proportional to both the tip–sample distance and the complex dielectric function of the sample. Consequently, the resulting near-field images provide qualitative nanoscale maps of local dielectric variations.59
![]() | ||
| Fig. 6 Merged s-SNOM images. The inset highlight the polaritons in G/R6G/bulk-MoS2 by adapting the intensity scale. Layer transitions are marked by black lines. | ||
When external factors such as doping, strain, or defects are introduced into the heterostructure, the local dielectric function is altered, often resulting in a spectral shift.59,60 This shift manifests as contrast changes in the near-field amplitude images. Fig. 6 shows the amplitude image of the heterostructure. As expected, the near-field signal amplitude systematically varies when moving across regions of different thicknesses—specifically from monolayer, to trilayer, and to bulk regions of the 2D material. Within each individual region, the signal remains relatively uniform, indicating a high degree of homogeneity in the deposited R6G film.
Interestingly, in areas where the CPD signal varies (inset of Fig. 5), we also observe corresponding changes in the near-field amplitude (Fig. 6). This correlation suggests that doping not only modifies the local electronic properties (as detected by KPFM) but also induces changes in the optical response through alterations in the dielectric function. In regions devoid of MoS2, the dielectric environment is different, leading to observable changes in near-field amplitude. This complementary behavior between near-field optical imaging and KPFM highlights the dual sensitivity of these techniques to local doping levels—affecting both electronic and optical properties—at nanometer resolution.
Of particular interest is the observation of propagating exciton–polaritons within the bulk region of the MoS2 heterostructure. As shown in previous studies, the propagation characteristics of phonon polaritons can be strongly influenced by adjacent thin molecular films, where nanoscale molecular layers were demonstrated to alter the polariton dispersion via coupling to molecular vibrational modes—highlighting the general sensitivity of polaritonic systems to their dielectric environment.61 Similarly, the propagation of exciton polaritons, including their wavelength and dispersion relation, has been reported to depend sensitively on the surrounding dielectric environment in prior work.62 We observe propagating polaritons with a wavelength of 345 nm in the bulk region, whereas for bare bulk MoS2, their wavelength is 357 nm, as reported in ref. 61 and confirmed by our measurements, corresponding to a reduction of about 4%. As result, the wavevector (k = 2π/λ) increase from 1.76 × 10−2 cm−1 (bare bulk) to 1.82 × 10−2 cm−1 (heterostructure). Achieving such a shift requires the effective dielectric constant of the environment to increase from 2.1 (SiO2) to 2.9 (calculated using the dispersion relation for TM modes from ref. 61 for MoS2), indicating that the molecular overlayer modifies the dielectric surroundings and thereby alters the polariton dispersion. Because polaritons are highly sensitive to changes in the dielectric landscape, even an ultrathin molecular film of only 4 nm thickness can significantly affect their propagation. To support this finding, we modeled the dispersion of the fundamental TM0 mode in MoS2 using an anisotropic slab-waveguide equation that includes an effective dielectric boundary to account for a thin spacer layer between the flake and the SiO2 substrate (see Fig. S3). For a fixed 4 nm spacer, increasing its permittivity shifts the TM0 dispersion to higher in-plane wavevectors, indicating stronger optical confinement in the MoS2 layer. To achieve the observed decrease in the wavelength of the TM mode, and thus an increase of the wavevector, εsp needs to be 2.15. Moreover, according to the topography images (Fig. 1), the position of Raman modes (Fig. 3) and PL signals (Fig. 4), height variations, strain, or doping play no major role, that further supports the results of the modeling.
This highlights the potential of molecular layers to actively tailor, modulate, and guide polaritonic modes in two-dimensional materials, opening new opportunities for nanoscale photonic device engineering.
Our findings provide direct experimental evidence that incorporated molecular layers can serve as an effective tool to modulate the local dielectric environment by the nature of the molecular layer, excitonic lifetimes, and polaritonic dispersion in TMDC-based systems. The observed shift in polariton wavelength by in the hybrid heterostructure, compared to bare MoS2, underscores the strong interfacial coupling and the potential to engineer light–matter interactions at the molecular scale. This work highlights the promising avenue of incorporating molecular dipole fields into 2D material stacks to create tunable, reconfigurable platforms for quantum optics, neuromorphic computing, and low-dimensional optoelectronics.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5nr02928g.
| This journal is © The Royal Society of Chemistry 2025 |