Home Physical Sciences Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2
Article Open Access

Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2

  • Mikhail K. Tatmyshevskiy , Georgy A. Ermolaev ORCID logo EMAIL logo , Dmitriy V. Grudinin , Aleksandr S. Slavich , Nikolay V. Pak , Marwa A. El-Sayed , Alexander Melentev , Elena Zhukova , Roman I. Romanov , Dmitry I. Yakubovsky , Andrey A. Vyshnevyy , Sergey M. Novikov , Aleksey V. Arsenin and Valentyn S. Volkov
Published/Copyright: December 8, 2025

Abstract

van der Waals transition metal dichalcogenides, distinguished by a high refractive index and giant optical anisotropy, are promising materials for integrated photonic devices. However, their superior optical properties are nowadays limited to exfoliated samples with only a micrometer scale, whereas industrial integration requires at least cm-scale dimensions. Here, we resolve this problem for MoTe2 by demonstrating that chemical vapor deposition synthesis can provide an identical optical response to the benchmark exfoliated samples in a broad spectral range (250–5,000 nm). It allows us to show high-performance waveguiding properties of MoTe2 with a subwavelength footprint of ∼λ/8 for telecommunication wavelengths. Therefore, our findings reveal MoTe2 as an ideal platform for the next-generation nanophotonics.

1 Introduction

The goal of miniaturizing integrated photonics to the level of electronics has been pursued for decades [1], [2], [3]. With the potential to overcome the diffraction limit, the most recent attempts used plasmonic nanostructures that support surface plasmon polaritons [4], [5], [6]. However, this approach faces a significant setback due to high ohmic losses, which limit the applications of plasmonic components and prevent the creation of chip-scale plasmonic integrated circuits [5], [7]. This highlights the urgent need for alternative lossless materials with exceptionally high refractive index n to overcome the limitations of the plasmonic approach [8], [9], [10].

This challenge establishes a new branch of all-dielectric and integrated nanophotonics focusing on high-n materials [11], [12], [13], [14]. Further studies [15], [16], [17], [18], [19], [20], [21], [22], [23] reveal that van der Waals (vdW) materials, in particular, transition metal dichalcogenides (TMDCs) [24], [25], are ideal platforms for all-dielectric nanophotonics thanks to their record values of refractive index n ∼ 4 and pronounced excitons. However, these records are traditionally demonstrated for exfoliated samples, which are unsuitable for mass production because of their limited area with dimensions of only ∼10–100 µm [8], [26], [27]. On the other hand, alternative methods of large-scale vdW synthesis, such as molecular beam epitaxy (MBE) and chemical vapor deposition (CVD), have another problem: the refractive index of synthesized films is significantly lower than that of their exfoliated counterparts [28], [29], [30], or the refractive index does not change, but extinction appears [31], [32]. One might expect that in the transparency region, the refractive index is less sensitive to sample-specific variations. In reality, however, CVD-grown samples demonstrate a much lower refractive index compared to the etalon refractive index of exfoliated samples, as seen in Figure 1a–g. This difference originates from the fact that defects diminish materials’ resonances, such as excitons, oscillator strength, which significantly influences the refractive index due to Kramers–Kronig relations. At the same time, exfoliated vdW materials demonstrate the record-high refractive index in the transparency region, overcoming even the classical high-n materials, including Si, GaP, and TiO2 (Figures 1g and h). In Figure 1h, we plot this trade-off between transparency and refractive index for a range of materials. It is important to note that Figure 1h is designed as a materials-selection chart for nanophotonic applications. Therefore, the x-axis, labeled “Optical bandgap, E g opt ,” represents a practical “transparency edge” rather than the fundamental quasiparticle bandgap. We define this edge as the photon energy at which the material’s extinction coefficient, k, exceeds a threshold of 0.001, corresponding to a material absorption loss of approximately 350 dB/cm at telecommunication wavelengths and also the limit of detection for traditional spectroscopic techniques like spectroscopic ellipsometry. This metric reflects the practical limit of useful transparency for compact on-chip devices. The y-axis (“Refractive index, n g opt ”) shows the refractive index of each material measured at its respective transparency edge. This approach explains, for example, why silicon with fundamental bandgap of ∼1.1 eV is plotted at ∼1.33 eV [33], [34], the energy at which its weak indirect absorption surpasses our defined threshold of k = 0.001. In contrast, for van der Waals materials like hBN with strong excitonic features, the optical response is dominated by these resonances, which define the practical absorption limit at energies below the quasiparticle bandgap (∼5.9 eV) [35]. Meanwhile, the value of ∼4.9 eV in Figure 1h reflects the practical transparency edge of hBN [10], [36]. Still, from Figure 1h, we can clearly see that van der Waals materials show a higher refractive index in a given transparency region than classical high-refractive index materials. As a result, vdW materials offer a better mode size diffraction limit λ/2n, where λ is the wavelength and n is the refractive index, owing to the superior trade-off between optical bandgap and refractive index (Figure 1g). Therefore, finding vdW materials for which CVD or MBE films exhibit similar optical responses to exfoliated samples is in high demand.

Figure 1: 
Refractive index of vdW materials. (a) Schematics of CVD and mechanical exfoliation fabrication methods of vdW materials. Comparison of the CVD and exfoliated refractive index for (b) MoS2, (c) hBN, (d) PdSe2, (e) SnS2, and (f) SnSe2. Optical constants are adopted from several reports [8], [10], [37], [38]. (g) Comparison of the refractive index of vdW materials with other high refractive index materials. Optical constants are adopted from the RefractiveIndex.Info database [39]. (h) Comparison of the maximum refractive index of vdW materials with other established highly refractive materials. (i) Comparison of the mode size, determined by diffraction limit λ/2n, for vdW materials and other established highly refractive materials. The dashed lines in panels (h) and (i) are the linear fitting of the data to demonstrate the linear correlations.
Figure 1:

Refractive index of vdW materials. (a) Schematics of CVD and mechanical exfoliation fabrication methods of vdW materials. Comparison of the CVD and exfoliated refractive index for (b) MoS2, (c) hBN, (d) PdSe2, (e) SnS2, and (f) SnSe2. Optical constants are adopted from several reports [8], [10], [37], [38]. (g) Comparison of the refractive index of vdW materials with other high refractive index materials. Optical constants are adopted from the RefractiveIndex.Info database [39]. (h) Comparison of the maximum refractive index of vdW materials with other established highly refractive materials. (i) Comparison of the mode size, determined by diffraction limit λ/2n, for vdW materials and other established highly refractive materials. The dashed lines in panels (h) and (i) are the linear fitting of the data to demonstrate the linear correlations.

In this work, we demonstrate that MoTe2 exhibits a unique combination of qualities, including a slight difference (of less than 1 % in the transparency range) between CVD MoTe2 film and exfoliated MoTe2 optical properties alongside the highest refractive index among all known TMDCs (MoS2, WS2, MoSe2, and WSe2) and transparency in the near-infrared region, as measured using broadband spectroscopic ellipsometry. Furthermore, we demonstrate that MoTe2 exhibits the best light-guiding properties with the footprint of ∼λ/8 using near-field optical microscopy. All these properties make MoTe2 a technological material for postsilicon nanophotonics.

2 Results

Figure 2a shows the crystal structure of a commercially synthesized CVD-grown (see Methods in Supplementary Information) 2H–MoTe2. Optical and scanning electron microscopy images in Figures 2b and c confirm the uniformity of our CVD MoTe2 film. The purpose of these images is to provide a visual inspection of the MoTe2 film at both microscale and macroscale. In addition, we performed atomic-force microscopy measurements to determine the film thickness (Figure 2d). Figure 2d shows that the thickness of our CVD-grown MoTe2 is 3.5 nm. To confirm the film’s crystal phase and chemical composition, we applied Raman spectroscopy and X-ray photoelectron spectroscopy (XPS). The Raman spectrum (Figure 2e) shows that our film has a prominent peak at about 234 cm−1, corresponding to E 2 g 1 in-plane vibrational modes, and two less intensive peaks corresponding to A 1g (170 cm−1) and B 2 g 1 (290 cm−1) out-of-plane vibrational modes, respectively. These peak positions agree with previously reported Raman spectra of MoTe2 [40]. Figures 2f and g shows the XPS spectra of Mo3d and Te3d core levels, respectively. The dominant doublets with 3d5/2 peaks positions at 228.4 eV and 573.2 eV are associated with the MoTe2 compound [41]. In addition, XPS spectra show the presence of MoOx and TeOx signals (the doublets with higher energies represent Mo and Te oxidized state [42], [43]), indicating surface oxidation (Figures 2f and g). Surface oxidation is a common phenomenon for many TMDCs, including MoTe2, when exposed to ambient conditions [44]. Hence, the Mo and Te oxidized states do not show the quality of the film. In contrast, the calculated Mo/Te ratio confirms the stoichiometry of our MoTe2 sample of a 33.5:66.5 ratio, which is very close to the expected 1:2 (see Supplementary Information).

Figure 2: 
Structural and morphological characterization of MoTe2 film. (a) Crystal lattice structure of 2H–MoTe2, (b) optical image, (c) SEM image, (d) AFM image and height profile, (e) Raman spectrum, (f and g) XPS spectra of CVD MoTe2 film.
Figure 2:

Structural and morphological characterization of MoTe2 film. (a) Crystal lattice structure of 2H–MoTe2, (b) optical image, (c) SEM image, (d) AFM image and height profile, (e) Raman spectrum, (f and g) XPS spectra of CVD MoTe2 film.

To determine the optical constants of our MoTe2 CVD film, we measure spectral ellipsometry and reflectance over a broad spectral range (λ = 250–1,700 nm for ellipsometry and λ = 450–5,000 nm for reflectance). The measured spectra for ellipsometry parameters Ψ and Δ are presented in Figures 3a and b, while the reflectance spectrum is shown in the inset of Figure 3d. In order to ensure consistency in the overlap range 450–1,700 nm of ellipsometry and reflectance, we use both spectra (from ellipsometry and reflectance) at the same time for fitting the optical properties of MoTe2. The model for processing ellipsometry data consists of four layers: the bottom 0.5 mm-thick layer of silicon, a layer of 286 nm SiO2, a t MoTe2 nm-thick MoTe2 layer, and a t oxide nm-thick oxide layer, with the total thickness t MoTe2 + t oxide = 3.5 nm as determined from the AFM profile (Figure 2d). The best-fit procedure yielded a thickness of t MoTe2 = 3.0 nm for the MoTe2 layer and t oxide = 0.5 nm for the oxide. The minor influence of this sub-nanometer oxide on the extracted optical constants, confirmed by a comparative analysis in Supplementary Information, is a direct consequence of the overwhelming optical signature of the high-index MoTe2. In the optical thin regime (tλ), the contribution of a layer to the total reflection is approximately proportional to the product of its thickness, t, and its optical contrast [45], [46], [47], related to n 2 n env 2 , where n env = n air 2 + n sub 2 / 2 1.3 (n air = 1 and n sub = n SiO2 ≈ 1.45) is the refractive index of the effective surrounding medium [48]. Given the large disparity in both thickness (t MoTe2 = 3.0 nm and t oxide = 0.5 nm) and, more critically, refractive index (n MoTe2 ≈ 4.5 and n oxide ≈ 2.0), the contribution of the MoTe2 layer dominates that of the oxide by a factor of t MoTe 2 n MoTe 2 2 n env 2 / t oxide n oxide 2 n env 2 50 . Consequently, the fitting procedure is overwhelmingly sensitive to the optical properties of MoTe2, with the thin oxide representing only a minor perturbation to the overall optical response. For modeling the optical constants of MoTe2, we used the Tauc–Lorentz oscillators model, which has proven to describe the optical response of vdW materials extremely well [49]. The resulting in-plane refractive index n and extinction coefficient k of our CVD film are shown in Figures 3c and d. From Figures 3c and d, we notice that the optical constants of CVD MoTe2 almost coincide with the optical constants of exfoliated MoTe2 [50], especially in the telecommunication range (1,260–1,675 nm). Interestingly, neglecting an oxide layer in the optical model also gives similar optical properties for MoTe2 (see Supplementary Information). Additionally, we note that the refractive index of MoTe2 for these wavelengths is the largest among classical TMDCs (MoS2, WS2, MoSe2, and WSe2), as illustrated on the example of λ = 1,550 nm in the inset of Figure 3c.

Figure 3: 
Ellipsometry measurements of MoTe2. (a,b) Experimental (solid lines) and calculated (dashed line) spectra of ellipsometric parameters Ψ and Δ, measured from MoTe2 thin film, (c) in-plane refractive index, and (d) in-plane extinction coefficients of bulk [50] and CVD (this work) MoTe2. Tabulated optical constants of MoTe2 are collected in Supplementary Information.
Figure 3:

Ellipsometry measurements of MoTe2. (a,b) Experimental (solid lines) and calculated (dashed line) spectra of ellipsometric parameters Ψ and Δ, measured from MoTe2 thin film, (c) in-plane refractive index, and (d) in-plane extinction coefficients of bulk [50] and CVD (this work) MoTe2. Tabulated optical constants of MoTe2 are collected in Supplementary Information.

Due to the atomic thickness of the film (3.0 nm) and the standard reflection-based ellipsometry configuration used, our measurements do not have sufficient sensitivity to reliably extract the out-of-plane optical constants n c [51]. However, we can infer the expected anisotropy by referencing the recent comprehensive study on exfoliated MoTe2 [50]. Given that our CVD film shows an in-plane refractive index nearly identical to that of exfoliated samples, it is reasonable to expect that it possesses a similarly large optical anisotropy. Moreover, such optical anisotropy can be advantageous for certain nanophotonics applications, for example, optical anisotropy can expand the parameter space for achieving single-mode operation and can be used as a design parameter to optimize mode confinement and reduce crosstalk [9].

Certainly, this finding of equal refractive index of CVD-grown and exfoliated MoTe2 does not necessarily imply that all CVD-synthesized MoTe2 will exhibit identical optical constants to exfoliated material. The underlying physical mechanism enabling the “high-quality” of our specific CVD MoTe2 sample, in terms of matching optical constants to exfoliated samples, can be attributed to the thermodynamics of MoTe2 between 2H and 1T’ phases [40]. Unlike other TMDCs, MoTe2 has a relatively small energy difference (ΔE ≈ 45 meV) between its 2H (semiconducting) and 1T’ (metallic) phases [52]. Consequently, the 2H phase appears in CVD only within the very narrow synthesis parameters since even a small concentration of defects leads to the phase transition from 2H to 1T’. Therefore, the thermodynamics of MoTe2 ensures the high quality of 2H–MoTe2 because otherwise CVD results in 1T’-MoTe2, which is in stark contrast with other TMDCs like MoS2, where a much larger energy difference (ΔE ≈ 650 meV) [52] means that even a large concentration of defects cannot induce this phase transition. As a result, one can synthesize 2H–MoS2 under a wide range of conditions, leading to films with varying defect densities: a “bad” synthesis still yields a 2H film, albeit a low-quality one [28]. We believe it is one of the primary reasons why 2H–MoTe2 received considerably less attention compared to other TMDCs despite having the superior optical properties. However, the complexity of the 2H–MoTe2 synthesis results in low defect density films, once achieved. Since the 2H–MoTe2 and 1T’-MoTe2 phases are nearly degenerate in energy, this delicate balance is easily tipped by external factors, including strain [52], doping [53], and, most importantly, defects [40]. Therefore, the system effectively “refuses” to form a highly defective 2H–MoTe2 film [40]. Hence, the challenge of synthesizing 2H–MoTe2 is actually a great advantage for its practical applications, and this self-limiting growth kinetics of MoTe2 are the primary reason for reproducing the refractive index in both CVD-grown and exfoliated MoTe2.

From the optical constants in Figure 3, we can evaluate the defect concentration from optical absorption. Assuming a simplistic model of light absorbance in the sub-bandgap region being proportional to the density of defects [54], we can estimate, to the first order approximation, light absorbance α as α = σN defects, where α = 4πk/λ (k is the extinction coefficient), σ is the absorption cross section for a single defect site, and N defects is the density of defect sites. The typical value of σ ranges from 10−17 to 10−15 cm2 [55]. Let us take the smallest value of σ = 10−17 cm2 to obtain the upper limit of N defects. For calculations, we take λ = 1,550 nm, which gives N defects = 4πk/(λσ) ≈ 1.3·1020 atoms/cm3. To calculate the percentage of defects, we need the total density of atoms in MoTe2. 2H–MoTe2 has a hexagonal crystal structure with two formula units (MoTe2) per unit cell. The dimensions of the unit cell are a = 0.352 nm and c = 1.391 nm [56]. Using these values, the total density of atoms in MoTe2 is given by N total = (atoms in unit cell)/V unit cell = 6/(( 3 /2)a 2 c) ≈ 4·1022 atoms/cm3. Therefore, the upper boundary of defects percentage in our sample is given by N defects/N total ≈ 0.3 %, which is an order of magnitude less than 3 % required for the phase transition from 2H to 1T’. This value is also in agreement with XPS data that, indicating the stoichiometry of our MoTe2 is 33.5:66.5, as (33.5–0.1 %):(66.5 + 0.2 %) ≈ 0.502, which is within 0.4 % accuracy of the ideal stoichiometry ratio of 0.5. These defects’ optical losses have a negligible impact on waveguide mode propagation direction, scattering losses, and confinement in planar waveguides since they are determined mostly by the refractive index.

To demonstrate the waveguiding properties of MoTe2, we study planar MoTe2 waveguides using scattering near-field optical microscopy (s-SNOM). The s-SNOM measurements were performed on a thicker (≈235 nm) MoTe2 flake prepared by mechanical exfoliation. The purpose of this experiment was to demonstrate the excellent waveguiding properties stemming from MoTe2 intrinsic high refractive index. For the measurements, we used a standard telecommunication wavelength range of 1,500–1,600 nm to excite the waveguiding mode inside the MoTe2 planar waveguide by focusing the light on the tip of an s-SNOM. Excited mode travels inside the waveguide and scatters on the edge. That scattering field interferes with the background signal, thus giving us the oscillating pattern (Figure 4a). To retrieve guiding mode properties, we studied their fast Fourier transform (Figure 4b). It requires taking into account the frequency shift [57]:

n eff = n obs cos α sin β ,

where n eff is the effective mode index, n obs is the observed effective mode index, α is the angle between the illumination wavevector k and its projection on the sample surface k , and β is the angle between k and the edge of the planar waveguide, where scattering occurs. Given the extracted refractive index components n ab and frequency shift, we calculate the energy (E = hc/λ)-momentum (q = 1/λ) dispersion relation of the waveguide mode using the transfer matrix method [58] (see Supplementary Information). Plotting experimental effective indices on this map (Figure 4c) shows excellent agreement between experiment and theory.

Figure 4: 
Near-field study of MoTe2 planar waveguides. (a) Near-field images, amplitude Amp(E), and phase Arg(E) for 1,500 and 1,600 nm. Other wavelengths are presented in Supplementary Information. (b) Fast Fourier Transform (FFT) of the complex near-field signal from panel (a). Green arrows mark the peak associated with the waveguide modes. (c) Transfer matrix calculations for planar MoTe2 waveguide. The experimental (q = 1/λ, E = hc/λ) data points (blue and green circles) show good agreement with the calculated dispersion. (d) Comparison of waveguide widths derived from multiple materials, each optimized to achieve the same effective mode index in the fundamental mode. The waveguides were fully encapsulated in SiO2 and had square cross sections. Their widths were optimized so that the fundamental mode would have the same effective refractive index. The inset shows a schematic image of the encapsulated waveguides. (e) Relation of crosstalk length and distance between cores of the waveguides for high-refractive index materials at 1,550 nm wavelength. The width is optimized for each point.
Figure 4:

Near-field study of MoTe2 planar waveguides. (a) Near-field images, amplitude Amp(E), and phase Arg(E) for 1,500 and 1,600 nm. Other wavelengths are presented in Supplementary Information. (b) Fast Fourier Transform (FFT) of the complex near-field signal from panel (a). Green arrows mark the peak associated with the waveguide modes. (c) Transfer matrix calculations for planar MoTe2 waveguide. The experimental (q = 1/λ, E = hc/λ) data points (blue and green circles) show good agreement with the calculated dispersion. (d) Comparison of waveguide widths derived from multiple materials, each optimized to achieve the same effective mode index in the fundamental mode. The waveguides were fully encapsulated in SiO2 and had square cross sections. Their widths were optimized so that the fundamental mode would have the same effective refractive index. The inset shows a schematic image of the encapsulated waveguides. (e) Relation of crosstalk length and distance between cores of the waveguides for high-refractive index materials at 1,550 nm wavelength. The width is optimized for each point.

To assess MoTe2 as a high refractive index photonic material, we compare MoTe2 waveguiding properties with other high refractive vdW materials, including MoS2, WS2, MoSe2, and WSe2. We consider waveguides operating at standard telecommunication wavelength λ = 1,550 nm. To enable a fair comparison, the waveguide widths were individually optimized so that the fundamental mode in each case exhibited the same effective refractive index. Therefore, although the physical dimensions vary, the mode confinement condition was held constant across all materials. For material analysis, we found the widths of square-shaped waveguides encapsulated in SiO2 with a given exponential decay χ = 0.89 μm−1 of the evanescent tail of the mode field outside the waveguide core, defined as:

χ = 2 π λ n eff 2 n Si O 2 2 .

Figure 4d confirms that MoTe2 waveguide core is the smallest compared to Si and other TMDCs. The results of waveguide crosstalk studies ( L ct = λ / 2 n o n e , with n o and n e being effective indices of odd and even supermodes of a pair of parallel waveguides) in Figure 4e show that MoTe2 demonstrates the highest integration density.

3 Conclusions

In summary, we have reported the broadband (250–5,000 nm) optical constants of CVD MoTe2 and shown that its refractive index is in good agreement with the refractive index of exfoliated MoTe2, opening an avenue for industrial-scale use of MoTe2 in photonics. Moreover, MoTe2 transparency and a high refractive index of ∼4.5 in the near-infrared region result in high-performance waveguides with a footprint of ∼λ/8. Combined with the fact that CVD MoTe2 is like exfoliated MoTe2 in terms of optical properties, this material is a promising platform for nanophotonics.

Looking forward, the availability of scalable, high-quality MoTe2 opens numerous avenues for advanced nanophotonics devices [26], such as δ-waveguides [59] (see Supplementary Information) and integrated nanophotonics [9]. Furthermore, high-index MoTe2 nanostructures are ideal candidates for realizing high-Q Mie-resonant metasurfaces [16] for applications in sensing [14], flat optics [60], and enhanced light–matter interactions [11]. Finally, as MoTe2 is known to host defect-based quantum emitters in the telecommunication spectral range [61], our work could establish a viable path toward integrating single-photon sources into high-index photonic circuits for scalable photonic quantum information technologies.

Supplementary information

Supplementary Information contains sections Materials and Methods, Additional Figures, and tabulated optical constants.


Corresponding author: Georgy A. Ermolaev, Emerging Technologies Research Center, XPANCEO, Internet City, Emmay Tower, Dubai, United Arab Emirates, E-mail: 

Mikhail K. Tatmyshevskiy, Georgy A. Ermolaev, and Dmitriy V. Grudinin contributed equally to this work.


Funding source: RSF

Award Identifier / Grant number: 25-29-00437

Acknowledgments

MKT and NVP gratefully acknowledge the financial support from the RSF (grant No. 25-29-00437).

  1. Research funding: This research was supported by the RSF (grant No. 25-29-00437).

  2. Author contributions: MKT, GAE, and DVG contributed equally to this work. GAE, AAV, SMN, AVA, and VSV suggested and directed the project. MKT, GAE, DVG, ASS, NVP, MAES, AM, EZ, RIR, and DIY performed the measurements and analyzed the data. DVG and AAV provided theoretical support. MKT, GAE, and DVG wrote the original manuscript. All authors have accepted responsibility for the entire content of this manuscript and consented to its submission to the journal, reviewed all the results, and approved the final version of the manuscript.

  3. Conflict of interest: Authors state no conflicts of interest.

  4. Data availability: The datasets generated during and/or analyzed during the current study are available from the corresponding author upon reasonable request.

References

[1] S. I. Bozhevolnyi, V. S. Volkov, E. Devaux, J.-Y. Laluet, and T. W. Ebbesen, “Channel plasmon subwavelength waveguide components including interferometers and ring resonators,” Nature, vol. 440, no. 7083, pp. 508–511, 2006, https://doi.org/10.1038/nature04594.Search in Google Scholar PubMed

[2] M. Smit, J. van der Tol, and M. Hill, “Moore’s law in photonics,” Laser Photon. Rev., vol. 6, no. 1, pp. 1–13, 2012, https://doi.org/10.1002/lpor.201100001.Search in Google Scholar

[3] C. Zhang et al.., “Integrated photonics beyond communications,” Appl. Phys. Lett., vol. 123, no. 23, p. 230501, 2023, https://doi.org/10.1063/5.0184677.Search in Google Scholar

[4] D. Y. Fedyanin, D. I. Yakubovsky, R. V. Kirtaev, and V. S. Volkov, “Ultralow-loss CMOS copper plasmonic waveguides,” Nano Lett., vol. 16, no. 1, pp. 362–366, 2016, https://doi.org/10.1021/acs.nanolett.5b03942.Search in Google Scholar PubMed

[5] D. I. Yakubovsky et al.., “Optical nanoimaging of surface plasmon polaritons supported by ultrathin metal films,” Nano Lett., vol. 23, no. 20, pp. 9461–9467, 2023, https://doi.org/10.1021/acs.nanolett.3c02947.Search in Google Scholar PubMed

[6] D. C. Look and K. D. Leedy, “ZnO plasmonics for telecommunications,” Appl. Phys. Lett., vol. 102, no. 18, p. 182107, 2013, https://doi.org/10.1063/1.4804984.Search in Google Scholar

[7] G. V. Naik, V. M. Shalaev, and A. Boltasseva, “Alternative plasmonic materials: Beyond gold and silver,” Adv. Mater., vol. 25, no. 24, pp. 3264–3294, 2013, https://doi.org/10.1002/adma.201205076.Search in Google Scholar PubMed

[8] G. A. Ermolaev et al.., “Giant optical anisotropy in transition metal dichalcogenides for next-generation photonics,” Nat. Commun., vol. 12, no. 1, p. 854, 2021, https://doi.org/10.1038/s41467-021-21139-x.Search in Google Scholar PubMed PubMed Central

[9] A. A. Vyshnevyy et al.., “Van der Waals Materials for Overcoming Fundamental Limitations in Photonic Integrated Circuitry,” Nano Lett., vol. 23, no. 17, pp. 8057–8064, 2023, https://doi.org/10.1021/acs.nanolett.3c02051.Search in Google Scholar PubMed

[10] D. V. Grudinin et al.., “Hexagonal boron nitride nanophotonics: A record-breaking material for the ultraviolet and visible spectral ranges,” Mater. Horizons, vol. 10, no. 7, pp. 2427–2435, 2023, https://doi.org/10.1039/d3mh00215b.Search in Google Scholar PubMed

[11] A. I. Kuznetsov, A. E. Miroshnichenko, M. L. Brongersma, Y. S. Kivshar, and B. Luk’yanchuk, “Optically resonant dielectric nanostructures,” Science, vol. 354, no. 6314, p. aag2472, 2016, https://doi.org/10.1126/science.aag2472.Search in Google Scholar PubMed

[12] J. B. Khurgin, “Expanding the photonic palette: Exploring high index materials,” ACS Photonics, vol. 9, no. 3, pp. 743–751, 2022, https://doi.org/10.1021/acsphotonics.1c01834.Search in Google Scholar

[13] G. Ermolaev et al.., “Giant and tunable excitonic optical anisotropy in single-crystal halide perovskites,” Nano Lett., vol. 23, no. 7, pp. 2570–2577, 2023, https://doi.org/10.1021/acs.nanolett.2c04792.Search in Google Scholar PubMed

[14] V. Maslova, G. Ermolaev, E. S. Andrianov, A. V. Arsenin, V. S. Volkov, and D. G. Baranov, “The influence of shot noise on the performance of phase singularity-based refractometric sensors,” Nanophotonics, vol. 14, no. 14, pp. 2463–2472, 2025, https://doi.org/10.1515/nanoph-2025-0101.Search in Google Scholar PubMed PubMed Central

[15] G. Zograf, A. Y. Polyakov, M. Bancerek, T. J. Antosiewicz, B. Küçüköz, and T. O. Shegai, “Combining ultrahigh index with exceptional nonlinearity in resonant transition metal dichalcogenide nanodisks,” Nat. Photonics, vol. 18, no. 7, pp. 751–757, 2024, https://doi.org/10.1038/s41566-024-01444-9.Search in Google Scholar

[16] R. Verre, D. G. Baranov, B. Munkhbat, J. Cuadra, M. Käll, and T. Shegai, “Transition metal dichalcogenide nanodisks as high-index dielectric Mie nanoresonators,” Nat. Nanotechnol., vol. 14, no. 7, pp. 679–684, 2019, https://doi.org/10.1038/s41565-019-0442-x.Search in Google Scholar PubMed

[17] G. A. Ermolaev et al.., “Broadband optical properties of monolayer and bulk MoS2,” npj 2D Mater. Appl., vol. 4, no. 21, pp. 1–6, 2020, https://doi.org/10.1038/s41699-020-0155-x.Search in Google Scholar

[18] G. A. Ermolaev et al.., “Wandering principal optical axes in van der Waals triclinic materials,” Nat. Commun., vol. 15, no. 1, p. 1552, 2024, https://doi.org/10.1038/s41467-024-45266-3.Search in Google Scholar PubMed PubMed Central

[19] V. Maslova, P. Lebedev, and D. G. Baranov, “Topological phase singularities in light reflection from non‐hermitian uniaxial media,” Adv. Opt. Mater., vol. 12, no. 6, p. 2303263, 2024, https://doi.org/10.1002/adom.202303263.Search in Google Scholar

[20] V. A. Maslova and N. S. Voronova, “Spatially-indirect and hybrid exciton–exciton interaction in MoS2 homobilayers,” 2D Mater., vol. 11, no. 2, p. 025006, 2024, https://doi.org/10.1088/2053-1583/ad1a6c.Search in Google Scholar

[21] H. Ling, R. Li, and A. R. Davoyan, “All van der Waals Integrated Nanophotonics with Bulk Transition Metal Dichalcogenides,” ACS Photonics, vol. 8, no. 3, pp. 721–730, 2021, https://doi.org/10.1021/acsphotonics.0c01964.Search in Google Scholar

[22] Y. Feng et al.., “Visible to mid-infrared giant in-plane optical anisotropy in ternary van der Waals crystals,” Nat. Commun., vol. 14, no. 1, p. 6739, 2023, https://doi.org/10.1038/s41467-023-42567-x.Search in Google Scholar PubMed PubMed Central

[23] G. Ma et al.., “Excitons enabled topological phase singularity in a single atomic layer,” ACS Nano, vol. 17, no. 18, pp. 17751–17760, 2023, https://doi.org/10.1021/acsnano.3c02478.Search in Google Scholar PubMed

[24] M. Nørgaard, T. Yezekyan, S. Rolfs, C. Frydendahl, N. A. Mortensen, and V. A. Zenin, “Near-field refractometry of van der Waals crystals,” Nanophotonics, vol. 14, no. 14, pp. 2473–2483, 2025, https://doi.org/10.1515/nanoph-2025-0117.Search in Google Scholar PubMed PubMed Central

[25] P. G. Zotev et al.., “Nanophotonics with multilayer van der Waals materials,” Nat. Photonics, vol. 19, no. 8, pp. 788–802, 2025, https://doi.org/10.1038/s41566-025-01717-x.Search in Google Scholar

[26] P. G. Zotev et al.., “Van der Waals Materials for Applications in Nanophotonics,” Laser Photon. Rev., vol. 17, no. 8, p. 2200957, 2023, https://doi.org/10.1002/lpor.202200957.Search in Google Scholar

[27] H. Ling, J. B. Khurgin, and A. R. Davoyan, “Atomic-Void van der Waals Channel Waveguides,” Nano Lett., vol. 22, no. 15, pp. 6254–6261, 2022, https://doi.org/10.1021/acs.nanolett.2c01819.Search in Google Scholar PubMed

[28] G. A. Ermolaev et al.., “Optical constants and structural properties of epitaxial MoS2 monolayers,” Nanomaterials, vol. 11, no. 6, p. 1411, 2021, https://doi.org/10.3390/nano11061411.Search in Google Scholar PubMed PubMed Central

[29] K. M. Islam, R. Synowicki, T. Ismael, I. Oguntoye, N. Grinalds, and M. D. Escarra, “In‐plane and Out‐of‐Plane optical properties of monolayer, few‐layer, and thin‐film MoS2 from 190 to 1700 nm and their application in photonic device design,” Adv. Photonics Res., vol. 2, no. 5, p. 2000180, 2021, https://doi.org/10.1002/adpr.202000180.Search in Google Scholar

[30] G. A. Ermolaev et al.., “Broadband optical properties of atomically thin PtS2 and PtSe2,” Nanomaterials, vol. 11, no. 12, p. 3269, 2021, https://doi.org/10.3390/nano11123269.Search in Google Scholar PubMed PubMed Central

[31] G. A. Ermolaev et al.., “Unveiling the broadband optical properties of Bi2Te3: Ultrahigh refractive index and promising applications,” Appl. Phys. Lett., vol. 125, no. 24, p. 241101, 2024, https://doi.org/10.1063/5.0219511.Search in Google Scholar

[32] G. A. Ermolaev et al.., “Broadband optical properties of Bi2Se3,” Nanomaterials, vol. 13, no. 9, p. 1460, 2023, https://doi.org/10.3390/nano13091460.Search in Google Scholar PubMed PubMed Central

[33] C. M. Herzinger, B. Johs, W. A. McGahan, J. A. Woollam, and W. Paulson, “Ellipsometric determination of optical constants for silicon and thermally grown silicon dioxide via a multi-sample, multi-wavelength, multi-angle investigation,” J. Appl. Phys., vol. 83, no. 6, pp. 3323–3336, 1998, https://doi.org/10.1063/1.367101.Search in Google Scholar

[34] C. Schinke et al.., “Uncertainty analysis for the coefficient of band-to-band absorption of crystalline silicon,” AIP Adv., vol. 5, no. 6, p. 067168, 2015, https://doi.org/10.1063/1.4923379.Search in Google Scholar

[35] G. Cassabois, P. Valvin, and B. Gil, “Hexagonal boron nitride is an indirect bandgap semiconductor,” Nat. Photonics, vol. 10, no. 4, pp. 262–266, 2016, https://doi.org/10.1038/nphoton.2015.277.Search in Google Scholar

[36] B. Choi et al.., “Natural hyperbolicity of hexagonal boron nitride in the deep ultraviolet,” arXiv:2507.13271, 2025.Search in Google Scholar

[37] A. Slavich et al.., “Multifunctional van der Waals PdSe2 for light detection, guiding and modulation,” Nat. Commun., vol. 16, no. 1, p. 9201, 2025, https://doi.org/10.1038/s41467-025-64247-8.Search in Google Scholar PubMed PubMed Central

[38] G. A. Ermolaev et al.., “Broadband optical constants and nonlinear properties of SnS2 and SnSe2,” Nanomaterials, vol. 12, no. 1, p. 141, 2021, https://doi.org/10.3390/nano12010141.Search in Google Scholar PubMed PubMed Central

[39] M. N. Polyanskiy, “Refractiveindex.info database of optical constants,” Sci. Data, vol. 11, no. 1, p. 94, 2024, https://doi.org/10.1038/s41597-023-02898-2.Search in Google Scholar PubMed PubMed Central

[40] J. P. Fraser et al.., “Selective phase growth and precise-layer control in MoTe2,” Commun. Mater., vol. 1, no. 1, p. 48, 2020, https://doi.org/10.1038/s43246-020-00048-4.Search in Google Scholar

[41] Q. Wang et al.., “Precise layer control of MoTe2 by ozone treatment,” Nanomaterials, vol. 9, no. 5, p. 756, 2019, https://doi.org/10.3390/nano9050756.Search in Google Scholar PubMed PubMed Central

[42] T. Thomas et al.., “Room temperature ammonia sensing of α-MoO3 nanorods grown on glass substrates,” Thin Solid Films, vol. 722, p. 138575, 2021, https://doi.org/10.1016/j.tsf.2021.138575.Search in Google Scholar

[43] F. Xie, W. C. H. Choy, C. Wang, X. Li, S. Zhang, and J. Hou, “Low‐temperature solution‐processed hydrogen molybdenum and vanadium bronzes for an efficient hole-transport layer in organic electronics,” Adv. Mater., vol. 25, no. 14, pp. 2051–2055, 2013, https://doi.org/10.1002/adma.201204425.Search in Google Scholar PubMed

[44] B. J. Kim et al.., “Phase controlled metalorganic chemical vapor deposition growth of wafer-scale molybdenum ditelluride,” ACS Nanosci. Au, vol. 5, no. 1, pp. 1–8, 2025, https://doi.org/10.1021/acsnanoscienceau.4c00050.Search in Google Scholar PubMed PubMed Central

[45] S. Yoo and Q.-H. Park, “Spectroscopic ellipsometry for low-dimensional materials and heterostructures,” Nanophotonics, vol. 11, no. 12, pp. 2811–2825, 2022, https://doi.org/10.1515/nanoph-2022-0039.Search in Google Scholar PubMed PubMed Central

[46] G.-H. Jung, S. Yoo, and Q.-H. Park, “Measuring the optical permittivity of two-dimensional materials without a priori knowledge of electronic transitions,” Nanophotonics, vol. 8, no. 2, pp. 263–270, 2018, https://doi.org/10.1515/nanoph-2018-0120.Search in Google Scholar

[47] K. R. Lee, J. Youn, and S. Yoo, “Deterministic reflection contrast ellipsometry for thick multilayer two-dimensional heterostructures,” Nanophotonics, vol. 13, no. 8, pp. 1417–1424, 2024, https://doi.org/10.1515/nanoph-2023-0753.Search in Google Scholar PubMed PubMed Central

[48] H. T. Nguyen-Truong, “Exciton binding energy and screening length in two-dimensional semiconductors,” Phys. Rev. B, vol. 105, no. 20, p. L201407, 2022, https://doi.org/10.1103/physrevb.105.l201407.Search in Google Scholar

[49] G. A. Ermolaev, D. I. Yakubovsky, Y. V. Stebunov, A. V. Arsenin, and V. S. Volkov, “Spectral ellipsometry of monolayer transition metal dichalcogenides: Analysis of excitonic peaks in dispersion,” J. Vac. Sci. Technol. B, vol. 38, no. 1, p. 014002, 2020, https://doi.org/10.1116/1.5122683.Search in Google Scholar

[50] L. A. Bereznikova et al.., “Artificial intelligence guided search for van der Waals materials with high optical anisotropy,” Mater. Horizons, vol. 12, no. 6, pp. 1953–1961, 2025, https://doi.org/10.1039/d4mh01332h.Search in Google Scholar PubMed

[51] Z. Xu et al.., “Optical detection of the susceptibility tensor in two-dimensional crystals,” Commun. Phys., vol. 4, no. 1, p. 215, 2021, https://doi.org/10.1038/s42005-021-00711-3.Search in Google Scholar

[52] K.-A. N. Duerloo, Y. Li, and E. J. Reed, “Structural phase transitions in two-dimensional Mo- and W-dichalcogenide monolayers,” Nat. Commun., vol. 5, no. 1, p. 4214, 2014, https://doi.org/10.1038/ncomms5214.Search in Google Scholar PubMed

[53] Y. Wang et al.., “Structural phase transition in monolayer MoTe2 driven by electrostatic doping,” Nature, vol. 550, no. 7677, pp. 487–491, 2017, https://doi.org/10.1038/nature24043.Search in Google Scholar PubMed

[54] C. Roxlo et al.., “Catalytic defects at molybdenum disulfide “edge” planes,” Solid State Ionics, vol. 22, no. 1, pp. 97–104, 1986, https://doi.org/10.1016/0167-2738(86)90063-9.Search in Google Scholar

[55] P. Tonndorf et al.., “Single-photon emission from localized excitons in an atomically thin semiconductor,” Optica, vol. 2, no. 4, p. 347, 2015, https://doi.org/10.1364/optica.2.000347.Search in Google Scholar

[56] E. Uesugi, X. Miao, H. Ota, H. Goto, and Y. Kubozono, “Transistor properties of exfoliated single crystals of 2H−Mo(S1−xTex)2 (0<x<1),” Phys. Rev. B, vol. 95, no. 24, p. 245310, 2017.10.1103/PhysRevB.95.245310Search in Google Scholar

[57] F. Hu et al.., “Imaging exciton-polariton transport in MoSe2 waveguides,” Nat. Photonics, vol. 11, no. 6, pp. 356–360, 2017, https://doi.org/10.1038/nphoton.2017.65.Search in Google Scholar

[58] N. C. Passler and A. Paarmann, “Generalized 4 × 4 matrix formalism for light propagation in anisotropic stratified media: Study of surface phonon polaritons in polar dielectric heterostructures,” J. Opt. Soc. Am. B, vol. 34, no. 10, p. 2128, 2017, https://doi.org/10.1364/josab.34.002128.Search in Google Scholar

[59] M. Lee et al.., “Wafer-scale δ waveguides for integrated two-dimensional photonics,” Science, vol. 381, no. 6658, pp. 648–653, 2023, https://doi.org/10.1126/science.adi2322.Search in Google Scholar PubMed

[60] Z. Li et al.., “Meta-optics achieves RGB-achromatic focusing for virtual reality,” Sci. Adv., vol. 7, no. 8, p. eabe4458, 2021, https://doi.org/10.1126/sciadv.abe4458.Search in Google Scholar PubMed PubMed Central

[61] H. Zhao, M. T. Pettes, Y. Zheng, and H. Htoon, “Site-controlled telecom-wavelength single-photon emitters in atomically-thin MoTe2,” Nat. Commun., vol. 12, no. 1, p. 6753, 2021, https://doi.org/10.1038/s41467-021-27033-w.Search in Google Scholar PubMed PubMed Central


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/nanoph-2025-0468).


Received: 2025-09-14
Accepted: 2025-11-24
Published Online: 2025-12-08

© 2025 the author(s), published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Frontmatter
  2. Reviews
  3. Light-driven micro/nanobots
  4. Tunable BIC metamaterials with Dirac semimetals
  5. Large-scale silicon photonics switches for AI/ML interconnections based on a 300-mm CMOS pilot line
  6. Perspective
  7. Density-functional tight binding meets Maxwell: unraveling the mysteries of (strong) light–matter coupling efficiently
  8. Letters
  9. Broadband on-chip spectral sensing via directly integrated narrowband plasmonic filters for computational multispectral imaging
  10. Sub-100 nm manipulation of blue light over a large field of view using Si nanolens array
  11. Tunable bound states in the continuum through hybridization of 1D and 2D metasurfaces
  12. Integrated array of coupled exciton–polariton condensates
  13. Disentangling the absorption lineshape of methylene blue for nanocavity strong coupling
  14. Research Articles
  15. Demonstration of multiple-wavelength-band photonic integrated circuits using a silicon and silicon nitride 2.5D integration method
  16. Inverse-designed gyrotropic scatterers for non-reciprocal analog computing
  17. Highly sensitive broadband photodetector based on PtSe2 photothermal effect and fiber harmonic Vernier effect
  18. Online training and pruning of multi-wavelength photonic neural networks
  19. Robust transport of high-speed data in a topological valley Hall insulator
  20. Engineering super- and sub-radiant hybrid plasmons in a tunable graphene frame-heptamer metasurface
  21. Near-unity fueling light into a single plasmonic nanocavity
  22. Polarization-dependent gain characterization in x-cut LNOI erbium-doped waveguide amplifiers
  23. Intramodal stimulated Brillouin scattering in suspended AlN waveguides
  24. Single-shot Stokes polarimetry of plasmon-coupled single-molecule fluorescence
  25. Metastructure-enabled scalable multiple mode-order converters: conceptual design and demonstration in direct-access add/drop multiplexing systems
  26. High-sensitivity U-shaped biosensor for rabbit IgG detection based on PDA/AuNPs/PDA sandwich structure
  27. Deep-learning-based polarization-dependent switching metasurface in dual-band for optical communication
  28. A nonlocal metasurface for optical edge detection in the far-field
  29. Coexistence of weak and strong coupling in a photonic molecule through dissipative coupling to a quantum dot
  30. Mitigate the variation of energy band gap with electric field induced by quantum confinement Stark effect via a gradient quantum system for frequency-stable laser diodes
  31. Orthogonal canalized polaritons via coupling graphene plasmon and phonon polaritons of hBN metasurface
  32. Dual-polarization electromagnetic window simultaneously with extreme in-band angle-stability and out-of-band RCS reduction empowered by flip-coding metasurface
  33. Record-level, exceptionally broadband borophene-based absorber with near-perfect absorption: design and comparison with a graphene-based counterpart
  34. Generalized non-Hermitian Hamiltonian for guided resonances in photonic crystal slabs
  35. A 10× continuously zoomable metalens system with super-wide field of view and near-diffraction–limited resolution
  36. Continuously tunable broadband adiabatic coupler for programmable photonic processors
  37. Diffraction order-engineered polarization-dependent silicon nano-antennas metagrating for compact subtissue Mueller microscopy
  38. Lithography-free subwavelength metacoatings for high thermal radiation background camouflage empowered by deep neural network
  39. Multicolor nanoring arrays with uniform and decoupled scattering for augmented reality displays
  40. Permittivity-asymmetric qBIC metasurfaces for refractive index sensing
  41. Theory of dynamical superradiance in organic materials
  42. Second-harmonic generation in NbOI2-integrated silicon nitride microdisk resonators
  43. A comprehensive study of plasmonic mode hybridization in gold nanoparticle-over-mirror (NPoM) arrays
  44. Foundry-enabled wafer-scale characterization and modeling of silicon photonic DWDM links
  45. Rough Fabry–Perot cavity: a vastly multi-scale numerical problem
  46. Classification of quantum-spin-hall topological phase in 2D photonic continuous media using electromagnetic parameters
  47. Light-guided spectral sculpting in chiral azobenzene-doped cholesteric liquid crystals for reconfigurable narrowband unpolarized light sources
  48. Modelling Purcell enhancement of metasurfaces supporting quasi-bound states in the continuum
  49. Ultranarrow polaritonic cavities formed by one-dimensional junctions of two-dimensional in-plane heterostructures
  50. Bridging the scalability gap in van der Waals light guiding with high refractive index MoTe2
  51. Ultrafast optical modulation of vibrational strong coupling in ReCl(CO)3(2,2-bipyridine)
  52. Chirality-driven all-optical image differentiation
  53. Wafer-scale CMOS foundry silicon-on-insulator devices for integrated temporal pulse compression
  54. Monolithic temperature-insensitive high-Q Ta2O5 microdisk resonator
  55. Nanogap-enhanced terahertz suppression of superconductivity
  56. Large-gap cascaded Moiré metasurfaces enabling switchable bright-field and phase-contrast imaging compatible with coherent and incoherent light
  57. Synergistic enhancement of magneto-optical response in cobalt-based metasurfaces via plasmonic, lattice, and cavity modes
  58. Scalable unitary computing using time-parallelized photonic lattices
  59. Diffusion model-based inverse design of photonic crystals for customized refraction
  60. Wafer-scale integration of photonic integrated circuits and atomic vapor cells
  61. Optical see-through augmented reality via inverse-designed waveguide couplers
  62. One-dimensional dielectric grating structure for plasmonic coupling and routing
  63. MCP-enabled LLM for meta-optics inverse design: leveraging differentiable solver without LLM expertise
  64. Broadband variable beamsplitter made of a subwavelength-thick metamaterial
  65. Scaling-dependent tunability of spin-driven photocurrents in magnetic metamaterials
  66. AI-based analysis algorithm incorporating nanoscale structural variations and measurement-angle misalignment in spectroscopic ellipsometry
Downloaded on 22.1.2026 from https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2025-0468/html
Scroll to top button