Next Article in Journal
Green Catalysis in Nanomaterials—Photocatalysis and Electrocatalysis
Previous Article in Journal
Photoelectrochemical-Type Photodetectors Based on Ball Milling InSe for Underwater Optoelectronic Devices
Previous Article in Special Issue
LAB-to-FAB Transition of 2D FETs: Available Strategies and Future Trends
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Tungsten Diselenide Nanoparticles Produced via Femtosecond Ablation for SERS and Theranostics Applications

by
Andrei Ushkov
1,
Dmitriy Dyubo
1,
Nadezhda Belozerova
1,2,
Ivan Kazantsev
3,
Dmitry Yakubovsky
1,
Alexander Syuy
1,3,
Gleb V. Tikhonowski
3,4,
Daniil Tselikov
1,4,
Ilya Martynov
1,
Georgy Ermolaev
3,
Dmitriy Grudinin
3,
Alexander Melentev
1,
Anton A. Popov
4,
Alexander Chernov
1,
Alexey D. Bolshakov
1,5,6,7,
Andrey A. Vyshnevyy
1,3,
Aleksey Arsenin
1,3,7,
Andrei V. Kabashin
8,
Gleb I. Tselikov
3 and
Valentyn Volkov
3,*
1
Moscow Center for Advanced Studies, Kulakova Str. 20, Moscow 123592, Russia
2
Frank Laboratory of Neutron Physics, Joint Institute for Nuclear Research, Dubna 141980, Russia
3
Emerging Technologies Research Center, XPANCEO, Internet City, Emmay Tower, Dubai, United Arab Emirates
4
MEPhI, Institute of Engineering Physics for Biomedicine (PhysBio), Moscow 115409, Russia
5
Faculty of Physics, St. Petersburg State University, Universitetskaya Emb. 7–9, St. Petersburg 199034, Russia
6
Center for Nanotechnologies, Alferov University, Khlopina 8/3, St. Petersburg 194021, Russia
7
Laboratory of Advanced Functional Materials, Yerevan State University, Yerevan 0025, Armenia
8
National Center for Scientific Research, LP3, Aix-Marseille University, CNRS, 13288 Marseille, France
*
Author to whom correspondence should be addressed.
Nanomaterials 2025, 15(1), 4; https://doi.org/10.3390/nano15010004
Submission received: 19 November 2024 / Revised: 16 December 2024 / Accepted: 19 December 2024 / Published: 24 December 2024

Abstract

:
Due to their high refractive index, record optical anisotropy and a set of excitonic transitions in visible range at a room temperature, transition metal dichalcogenides have gained much attention. Here, we adapted a femtosecond laser ablation for the synthesis of WSe2 nanoparticles (NPs) with diameters from 5 to 150 nm, which conserve the crystalline structure of the original bulk crystal. This method was chosen due to its inherently substrate-additive-free nature and a high output level. The obtained nanoparticles absorb light stronger than the bulk crystal thanks to the local field enhancement, and they have a much higher photothermal conversion than conventional Si nanospheres. The highly mobile colloidal state of produced NPs makes them flexible for further application-dependent manipulations, which we demonstrated by creating substrates for SERS sensors.

1. Introduction

Tungsten diselenide (WSe2) belongs to a family of van der Waals materials that have attracted exceptional attention since the isolation of graphene in 2004 [1] due to their extraordinary electrical, optical and magnetic properties. Being a transition metal dichalcogenide (TMDC), WSe2 has a layered crystalline structure, where neighboring hexagonal, closely packed monolayers of tungsten and selenium are stacked by weak van der Waals (vdW) forces (see Figure 1a). The unique material structure of TMDCs leads to their outstanding mechanical properties (exfoliation down to atomically thin flakes [2,3]), electronic (electronic structure engineering [4]) and optical (high refractive index and giant anisotropy [5,6]) characteristics, which drive extensive TMDC research for medical [7,8,9], electronic [10,11], optical [12,13,14,15] and space [16] scientific communities.
Among a set of common transition metals of TMDCs, the tungsten atom is the heaviest, which sufficiently impacts the WSe2 properties. For example, the WSe2 monolayer valence band splitting is three times larger than that for the MoS2 monolayer, which makes it promising for valleytronic and spintronic applications [17,18]. Unlike most TMDC materials [19,20,21,22], which exhibit n-type conductivity without doping, WSe2 is hole-conducting and is among the most frequently used p-type 2D materials [23,24]. In contrast to WSe2’s single- and few-layered state, its bulk form is much less elaborated in terms of applications, despite the abundance of shapes already reported: nanosheets assembled into nanoflowers [25], nanotubes [26], vertically aligned nanosheets [27] and quantum dots [28]. Notably, bulk WSe2, however, has a set of advantageous characteristics for optics, inherited from the TMDC material family: a high refractive index n > 4 in the visible and infrared spectral ranges, a prominent (20–30%) optical anisotropy and a set of excitonic transitions in the visible range at room temperature. Importantly, WSe2 has a resonant excitonic absorption peak at 750 nm, which is extremely interesting for photoheating and theranostic applications in the NIR-I transparency window of biological tissues. The key issue of exploiting the high refractive index of WSe2 in photonic nanostructures is the efficiency and output quality of fabrication methods.
Femtosecond laser material processing is a high-throughput manufacturing tool employed in science and industry and offers the advantage of a flexible, precise treatment of a wide range of materials [29,30,31].
For van der Waals materials, naturally occurring in a layered plane geometry, their shape transformation into spherical nanoparticles is a challenging task [32]. It was already shown that the femtosecond pulsed laser ablation in liquid (PLAL) approach allows the synthesis of spherical vdW NPs, which inherit the crystalline and optical properties of the bulk target, demonstrate an onion-like or polycrystalline structure and have a broad size distribution from 5 to 250 nm [33,34,35]. For the first time, it was demonstrated for MoS2 and WS2 materials [33,34], then for MoSe2 as well [35]. Interestingly, in the case of MoS2 NPs, the laser fragmentation can be employed to further tune their atomic composition and yield oxidized MoO3−x NPs [34]. In the case of the MoSe2 target, the femtosecond ablation revealed the 2H/1T phase transition in NPs [35].
In our work, we study the possibilities and features of the synthesis of spherical WSe2 nanoparticles by the femtosecond laser ablation method. Previously, we have successfully used PLAL for preparing NPs of Si [36], TiN [37], MoS2 [33,34] and other materials, including core-satellite nanocomposites [38,39,40]. Here, we demonstrate that PLAL yields crystalline WSe2 NPs with a nearly spherical shape and diameters ranging from ∼10 nm to ∼150 nm. Importantly, prepared NPs conserve the high refractive index and excitonic response of an initial bulk WSe2 target. The latter is illustrated via the photoheating experiments with a tunable Ti–sapphire laser in a wavelength range from 700 nm to 870 nm, which passes through the WSe2 A-exciton position (∼750 nm). We believe that the proposed high-yield and substrate-free technique will unlock new optical and theranostic applications of WSe2 NPs.

2. Materials and Methods

2.1. Sample Preparation

Femtosecond pulsed laser ablation in liquids (PLAL) was used to synthesize WSe2 nanoparticles (NPs). A high-purity, synthetically grown bulk WSe2 crystal (2D Semiconductors Inc., Scottsdale, AZ, USA) with dimensions of 1 × 4 × 4 mm served as the target material. The crystal was placed inside a 10 mL glass vessel filled with 2 mL of DI water. The height of the water column above the crystal surface was 10 mm and was kept constant in all experiments. A femtosecond Yb:KGW laser system (TETA-10, Avesta, Moscow, Russia) operating at a wavelength of 1033 nm was employed for the ablation process. The laser delivered pulses with a duration of 270 fs, a pulse energy of 100 μ J , and a repetition rate of 1 kHz. The laser beam was focused by a lens with a working distance of 100 mm. The lens-to-crystal distance was carefully adjusted to between 95–105 mm to maximize the efficiency of nanoparticle synthesis. Beam scanning in the XY plane across the target surface was achieved using two mutually perpendicular linear translation motorized stages, which perform a cyclic sweep of a square pattern at a speed of 5 mm/s. A standard ablation duration was set at 10 min. After ablation, the prepared colloidal solutions of WSe2 NPs were collected and analyzed without further purification or chemical treatment, except differential centrifugation.

2.2. Sample Characterization

The crystalline structure of single NPs was characterized by the high-resolution TEM system (JEM 2010; JEOL, Tokyo, Japan) at 200 kV. Drops of 2 µL of sample colloids were drop-casted on carbon-coated TEM copper grids and dried at ambient conditions.
A scanning SEM system (MAIA 3; Tescan, Brno, Czech Republic) with an integrated EDX detector (X-act; Oxford Instruments, Abingdon, UK) was used to examine the shape of large NPs at an inclined view and EDX characterization of the bulk crystal ablation target. The samples were prepared by dropping 2 µL of colloids on a clean Si substrate and drying at ambient conditions.
For temperature-dependent Raman characterization, a 20 µL drop of a sample colloid was drop-casted on a cover glass substrate instead of a metallic substrate to avoid the heat transfer to the substrate and dried at ambient conditions for 1 h.
For SERS measurements, 20 µL drops of sample colloids were drop-casted on aluminum surface and spin-coated to obtain thin films. As-prepared SERS substrates were covered with 2 µL of dye of various concentrations 10 4 10 8 M and dried under ambient conditions for 1 h. Raman spectroscopy was performed using Horiba LabRAM HR Evolution setup with excitation wavelengths of 532 nm and 633 nm, a diffraction grating of 600 lines/mm and a microscope objective of 100×, NA 0.9. A good reproducibility of spectra was observed during the measurements.

3. Results and Discussion

3.1. Laser Ablation Synthesis of WSe2 NPs

Bulk tungsten diselenide has a layered structure typical for TMDCs materials. Inside every layer, tungsten atoms have a trigonal prismatic coordination with regard to selenium; therefore, the hexagonal lattice visible from the top is composed of selenium, tungsten and, again, selenium sandwich planes (see Figure 1a). The distance between neighboring sandwich layers is ≈ 3.14   Å [41].
The PLAL setup is schematically shown in Figure 1d and in Supplementary Note S1. The bulk crystal is placed at the bottom of the glass vessel, filled with DI water. The free surface of the liquid is 1 cm above the top surface of the crystal. Femtosecond laser pulses (1030 nm, 280 fs, 100 μ J ) are focused on the bulk crystal surface via the lens (focal distance 100 mm) and generate cavity bubbles with ionized plasma. The plasma cooling and cavity bubbles deflation leads to the plasma condensation into crystalline NPs of different size. The prepared colloid demonstrates a good stability (during at least one month after the synthesis) due to the significant charge of WSe2 NPs with an electrostatic potential reaching 30 40 mV (see Supplementary Note S2).
Bulk WSe2 crystal for PLAL was purchased from 2D Semiconductors Inc. The energy-dispersive X-ray spectroscopy (EDX) of the crystal (Figure 1b) confirms the stoichiometry quality, W (32.52 at.%), Se (67.48 at.%), whereas the selected-area electron diffraction (SAED) of microscopic WSe2 flake demonstrates a typical hexagonal pattern of monocrystalline TMDCs (Figure 1c).

3.2. Morphology of WSe2 Nanoparticles

The structural characteristics of WSe2 nanoparticles synthesized by pulsed laser diffraction (PLAL) in liquids are presented in Figure 2. Transmission electron microscopy (TEM) image (Figure 2a) shows the distribution of the nanoparticles. Most of them have a near-spherical morphology. To further confirm that, we performed a tilted scanning electron microscopy (SEM) (Figure 2b) at an inclination angle of 45 degrees and found that NPs retain a round shape in the image. The EDX spectrum (Figure 2c) confirms that NPs conserve the atomic composition of the original crystal: W (12.95 at.%) and Se (29.88 at.%). Some additional peaks correspond to the TEM grid copper Cu (57.17 at.%). The SAED pattern (Figure 2d) reveals distinct diffraction rings, which correspond to the interplanar spacings of crystalline WSe2. The distances corresponding to the planes (0 0 4), (1 0 0), (0 0 6), (1 0 4), (1 0 6) and (1 1 0) are in a good agreement with theoretically calculated values (see Supplementary Note S3). High-resolution TEM (HRTEM) of a single WSe2 NP (Figure 2e) further confirms its polycrystalline structure; there are visible zones with different orientations of crystallites. The differential centrifugation of the synthesized colloid was performed in order to separate various average NPs sizes (see the next Section). The fact that the Raman spectra of the as-prepared colloids (Figure 2f) demonstrate a characteristic WSe2 peak at 250 cm−1 and, generally, repeat spectral features of the bulk crystal indicates the quality of WSe2 NPs in a broad range of sizes.

3.3. Optical Response of WSe2 Nanoparticles

The PLAL approach inherently produces a polydisperse colloid. Remarkably, the colloid remained stable for over a month with negligible aggregation/precipitation, due to a significant negative zeta potential 35 mV (see Supplementary Note S2). Given the high refractive index of WSe2 in the visible spectrum, pronounced changes in optical properties may be achieved due to variations in nanoparticle size. To obtain solutions with different average diameters of NPs, the initial colloid underwent differential centrifugation. Particle size distributions obtained from the manual analysis of TEM images and dynamic light scattering (DLS) measurements confirmed log-normal size distributions, with the average particle size shifting to smaller values with increasing revolutions per minute (rpm). Taking into account that the size distribution of initial colloid is log-normal as well, it leads to different mass concentrations of samples and, consequently, to different color saturations (transparencies) of the solutions at different centrifugation speeds (Figure 3e). The most transparent solutions were obtained for lowest and highest centrifugation speeds (see Figure 3e). Moreover, the difference in solutions’ hues indicates the difference in their extinction properties (see Figure 3b). Firstly, solutions with smaller NPs possess a more rapid extinction fall at a shorter wavelength ( 1 / λ 4 ), typical for the Rayleigh scattering regime. Secondly, the excitonic peak, located at 750 nm for colloids with NPs of smaller average size, is red-shifted for larger NPs (see the inset in Figure 3b). The presence of the excitonic peak, which originates from the bulk crystal (see Figure 3d), indicates NPs’ crystallinity. The polycrystalline nature of NPs is taken into account in extinction calculations by using a homogenized dielectric function, ε a v = 2 ε o / 3 + ε e / 3 , where ε o and ε e are in-plane and out-of-plane bulk dielectric permittivities, correspondingly. The peak shifting is caused by the fact that for WSe2 NPs large enough (i.e., 130 nm in diameter), the total extinction peak consists of two components (see Figure 3c): a material excitonic feature (visible as a peak in electrical total extinction contribution) and a magnetic dipole resonance (visible in magnetic dipole channel). For WSe2 NPs large enough, the magnetic dipole resonance, as well as other magnetic and electric multipoles, shifts towards longer wavelengths with increasing the particle size, which is an important feature of the Mie scattering regime. Thus, the synthesized particles might have applications not only as Rayleigh scatterers, but as Mie-resonators as well after a more methodical separation of large WSe2 NPs.

3.4. Photothermal Applications of WSe2 Nanoparticles

WSe2, as an efficient absorber in visible and near-IR bands, has gained much attention as a perspective thin film solar cell absorber [42,43,44]. In this regard, we examine the photothermal properties of WSe2 NPs in visible range at a wavelength of 532 nm, which is readily accessible in our Raman setup and is close to the peak wavelength of solar irradiance at Earth’s surface [45].
Figure 4a demonstrates a temperature-dependent shift in range from 25 ℃ up to 300 ℃ of Raman E 2 g 1 peak of WSe2 NPs, drop-casted on a cover glass substrate instead of a metallic substrate to avoid the heat transfer to the substrate. The sample temperature was monitored via the LINKAM temperature control stage; Raman excitation was set 20 μ W at 532 nm. We assume a linear dependence of the Raman peak position ω ( T ) on temperature T due to the relatively small temperature range examined in the study:
ω T = ω 0 + χ 1 T ,
where ω 0 is E 2 g 1 peak position at 25 ℃ and χ 1 0.013 is the first-order temperature coefficient, obtained from the linear fitting of experimental data from Figure 4a. The obtained χ 1 is within a 10% deviation tolerance with values reported in recent works on multilayered WSe2 flakes [46,47] (see Supplementary Note S4). Then, we use Equation (1) to retrieve the local sample temperature from temperature-dependent Raman spectra. The excitation laser 532 nm with different powers was focused into a diffraction-limited spot via 100× microscope objective, thus revealing the laser-induced heating efficiency (see Figure 4b). Interestingly, at a quite low irradiation power of 0.7 mW, WSe2 NPs are heated by almost 350 ℃ and, therefore, have more than four times higher photothermal efficiency in comparison with WSe2 bulk crystal, which originates from an enhanced nanoresonators absorption.
Photothermal antibacterial and cancer therapy nowadays attracts growing attention as a minimally invasive nanomedicine approach. Due to the bacteria’s inability to adapt and elaborate anti-thermal mechanisms, efficient nanosensitizers are highly required for successive method implementation. Thermal agents should possess a sensitivity to the NIR-I optical window, as well as have a symmetrical (spherical) shape and small (less than 100 nm) size to facilitate their uptake by biological cells. We examine the phototermal response of synthesized crystalline WSe2 NPs in the NIR-I optical window (700–980 nm). A collimated beam of 830 nm laser source passes through a cuvette with 1 mL of heated colloid. The solution temperature as a function of time was recorded by a thermal imaging camera. Consequent heating–cooling cycles of the same WSe2 colloid confirmed a good reproducibility of a photothermal conversion coefficient [48] η 0.44 (see Figure 4c). The photothermal response of WSe2 NPs was compared with Si NPs with the same extinction of water solutions at 830 nm (see Figure 4d). Due to the presence of excitonic transitions in a visible range, which are red-shifted as compared with those in Si, WSe2 NPs possess a higher absorption component in the total extinction cross-section in near-IR (see Supplementary Note S5), which results in a 4-fold enhancement of the photothermal efficiency (see Figure 4e).
Photoheating experiments analogous to those presented in Figure 4e were performed with a tunable titanium-sapphire laser source for wavelengths in the range from 700 nm to 870 nm, which includes the WSe2 A-exciton (see Figure 4f.) The experimentally observed photothermal conversion efficiency η is 0.48 for the shortest utilized wavelength of 700 nm, reaches the maximum value of 0.55 close to the excitonic resonance at 750 nm and drops down at longer wavelengths. Colloidal extinction was measured using transmission through a cuvette with a deionized water as a baseline; the colloidal absorption curve was obtained by multiplying the values of extinction by η (see Supplementary Note S6). This approach allows retrieving colloidal optical absorption properties from photoheating experiments.

3.5. WSe2 NPs for SERS

Figure 5a,b presents Raman spectra of the rhodamine 6G (R6G) and crystal violet (CV), respectively, in the concentration range of 10−4–10−8 M, which were enhanced by means of WSe2 substrates.
The Raman spectra of R6G molecules acquired from the WSe2 substrate exhibit strong lines at 614, 776, 1130, 1185, 1315, 1364, 1510, 1574 and 1649 cm−1, respectively, with excitation at 532 nm. All peaks obtained are in good agreement with earlier Raman R6G reports [49,50].
A pronounced line on 614 cm−1 corresponds to the C–C–C ring in-plane bending vibration. The C–H out-of-plane bending vibration for R6G is observed at 776 cm−1, while the C–H in-plane bending vibration is observed at 1130 cm−1. The C–O–C stretching frequency appears at 1187 cm−1. Peaks at 1315, 1364, 1510, 1574 and 1649 cm−1 correspond to the aromatic C–C stretch of the R6G molecule [51].
The SERS spectrum of CV contains peaks at 1619 and 1588 cm−1, which correspond to the C–C stretching vibration of the phenyl ring. The peak at 1372 cm−1 is the C–C center stretching vibration, and those at 1175 and 804 cm−1 are C–H bending vibrations. The radical-ring skeletal vibration and C–N bending vibration occur at 915 and 423 cm−1, respectively [52,53].
As expected, the SERS intensity was gradually weakened with decreasing R6G and CV concentration. When the concentration goes down to 1 × 10 7 M, the R6G and CV peaks can still be detected in range of 400–1700 cm−1 with feeble intensity (see Figure 5), demonstrating that the detection limit of the WSe2 substrate has been reached. Current works [54,55] approve the high efficiency of SERS sensors based on WSe2 structures, with LOD values in the range of 10 5 10 8 M.
We perform the analogous SERS experiments on substrates with crystalline MoS2 NPs, synthesized by the same PLAL approach. Interestingly, nanoparticles from both TMDCs demonstrate a pronounced SERS effect (see Figure 5). The LOD values for MoS2 are also moderate ( 10 7 M), which might be caused by the SERS chemical mechanism (CM) [56]. In Figure 5d, MoS2 peaks (marked by asterisks) at low dye concentrations of 10 6 10 8 M were observed in the range of 370–470 cm−1 of CV spectra. The MoS2 modes were identified as E 2 g 1 at 385 cm−1, A1g at 409 cm−1 and 2LA(M) at 460 cm−1. The appearance of these spectral lines is caused by a CV signal weakness at low dye concentrations.
The obtained structures have scope for improvement. For example, considering particles with a larger specific surface area and more intense absorption at excitation wavelengths may be attractive for SERS. Moreover, the Raman signal can be enhanced [54,57,58,59] by means of creating hybrid structures (TMDC–TMDC) or core–shell/ core–satellite structures (TMDCs-Au NPs) with a lower detection threshold.
In the broader context, PLAL-synthesized colloids from different TMDCs allow to find appropriate optical properties (i.e., refractive index, excitonic absorption), which, considering a high stability of NP solutions [33], may be desirable in applications.

4. Conclusions

In conclusion, we adapted the femtosecond laser ablation method for the production of WSe2 NPs from bulk crystal. Importantly, the synthesized NPs retain the crystalline (polycrystalline) structure of the target and inherit the high refractive index and excitonic absorption. The high photothermal response is four times stronger than that of conventional Si scatterers, which is attractive for medical purposes like cancer therapy [60], whereas the high refractive index and variety of accessible sizes unlock WSe2 NPs for Mie-tronics [61]. SERS sensing with WSe2 nanoparticles substrate was also demonstrated. Generally, highly symmetrical, nano-sized, crystalline and highly refractive NPs are promising for various applications in postsilicon nanotechnologies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15010004/s1, Figure S1: The scheme of the experimental femtosecond ablation setup. Mirror M1 is fixed, whereas mirrors M2 and M3 have kinematic mounts for optical path alignment; Figure S2: Measured zeta potential of WSe2 colloidal NPs. (a) Zeta potential distribution in every colloid; (b) Variations of the most probable zeta potential with centrifuge rotational speed; Table S1: Iinterplanar spacings dexp, determined from SAED pattern of WSe2 NPs, and calculated dhkl distances between consecutive crystallographic planes with Miller indices hkl; Figure S3: Comparison of experimental first-order temperature coefficients χ 1 from the literature and in the present work. Dashed lines denote linear fitting of experimental data (color circles); Figure S4: Comparison of theoretical extinction and absorption curves of WSe2 and Si water colloids, calculated via Mie theory. The vertical dashed line corresponds to the 830 nm laser diode line used for photoheating experiments; Figure S5: (a) Definitions of “heating” and “cooling” curves and their fitting using Eqs. S2,S3; (b) The same thermogram as in (a), re-calculated in the adiabatic representation via Eq. S4; See Refs. [46,47,48].

Author Contributions

Conceptualization, A.D.B., A.A. and V.V.; methodology, G.V.T., A.V.K. and A.U.; validation, A.A., A.A.P. and A.C.; formal analysis, I.K., G.E. and A.A.V.; investigation, D.D., I.M., D.Y., A.S., G.I.T., D.T., N.B. and A.U.; resources, A.C., D.G. and A.M.; data curation, I.K., A.U. and N.B.; writing—original draft preparation, A.U., D.D. and I.K.; writing—review and editing, A.D.B. and A.A.V.; visualization, A.U.; supervision, A.D.B., A.A. and V.V.; project administration, A.U.; funding acquisition, A.A., A.D.B. and V.V.; All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Higher Education of the Russian Federation, agreement 075-15-2024-560.

Data Availability Statement

The original contributions presented in the study are included in the article and Supplementary Materials; further inquiries can be directed to the corresponding author.

Acknowledgments

The authors acknowledge Valentyn Solovey for the graphical illustration of the PLAL setup.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
TMDCTransition metal dichalcogenide
NPNanoparticle
PLALPulsed laser ablation in liquid
SEMScanning electron microscope
TEMTransmission electron microscope
SAEDSelected area electron diffraction
EDXEnergy-dispersive X-ray spectroscopy
SERSSurface-enhanced Raman scattering
R6GRhodamine 6G
CVCrystal violet

References

  1. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.E.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed]
  2. Hossen, M.F.; Shendokar, S.; Aravamudhan, S. Defects and Defect Engineering of Two-Dimensional Transition Metal Dichalcogenide (2D TMDC) Materials. Nanomaterials 2024, 14, 410. [Google Scholar] [CrossRef] [PubMed]
  3. Falin, A.; Holwill, M.; Lv, H.; Gan, W.; Cheng, J.; Zhang, R.; Qian, D.; Barnett, M.R.; Santos, E.J.; Novoselov, K.S.; et al. Mechanical properties of atomically thin tungsten dichalcogenides: WS2, WSe2, and WTe2. ACS Nano 2021, 15, 2600–2610. [Google Scholar] [CrossRef] [PubMed]
  4. Zhao, Y.H.; Yang, F.; Wang, J.; Guo, H.; Ji, W. Continuously tunable electronic structure of transition metal dichalcogenides superlattices. Sci. Rep. 2015, 5, 8356. [Google Scholar] [CrossRef]
  5. Ermolaev, G.A.; Stebunov, Y.V.; Vyshnevyy, A.A.; Tatarkin, D.E.; Yakubovsky, D.I.; Novikov, S.M.; Baranov, D.G.; Shegai, T.; Nikitin, A.Y.; Arsenin, A.V.; et al. Broadband optical properties of monolayer and bulk MoS2. NPJ 2D Mater. Appl. 2020, 4, 21. [Google Scholar] [CrossRef]
  6. Ermolaev, G.; Grudinin, D.; Stebunov, Y.; Voronin, K.V.; Kravets, V.; Duan, J.; Mazitov, A.; Tselikov, G.; Bylinkin, A.; Yakubovsky, D.; et al. Giant optical anisotropy in transition metal dichalcogenides for next-generation photonics. Nat. Commun. 2021, 12, 854. [Google Scholar] [CrossRef]
  7. Lin, T.; Zhong, L.; Guo, L.; Fu, F.; Chen, G. Seeing diabetes: Visual detection of glucose based on the intrinsic peroxidase-like activity of MoS2 nanosheets. Nanoscale 2014, 6, 11856–11862. [Google Scholar] [CrossRef]
  8. Fusco, L.; Gazzi, A.; Peng, G.; Shin, Y.; Vranic, S.; Bedognetti, D.; Vitale, F.; Yilmazer, A.; Feng, X.; Fadeel, B.; et al. Graphene and other 2D materials: A multidisciplinary analysis to uncover the hidden potential as cancer theranostics. Theranostics 2020, 10, 5435. [Google Scholar] [CrossRef]
  9. Anju, S.; Mohanan, P. Biomedical applications of transition metal dichalcogenides (TMDCs). Synth. Met. 2021, 271, 116610. [Google Scholar] [CrossRef]
  10. Jiang, D.; Liu, Z.; Xiao, Z.; Qian, Z.; Sun, Y.; Zeng, Z.; Wang, R. Flexible electronics based on 2D transition metal dichalcogenides. J. Mater. Chem. A 2022, 10, 89–121. [Google Scholar] [CrossRef]
  11. Kang, T.; Tang, T.W.; Pan, B.; Liu, H.; Zhang, K.; Luo, Z. Strategies for controlled growth of transition metal dichalcogenides by chemical vapor deposition for integrated electronics. ACS Mater. Au 2022, 2, 665–685. [Google Scholar] [CrossRef] [PubMed]
  12. Verre, R.; Baranov, D.G.; Munkhbat, B.; Cuadra, J.; Käll, M.; Shegai, T. Transition metal dichalcogenide nanodisks as high-index dielectric Mie nanoresonators. Nat. Nanotechnol. 2019, 14, 679–683. [Google Scholar] [CrossRef] [PubMed]
  13. Munkhbat, B.; Küçüköz, B.; Baranov, D.G.; Antosiewicz, T.J.; Shegai, T.O. Nanostructured transition metal dichalcogenide multilayers for advanced nanophotonics. Laser Photonics Rev. 2023, 17, 2200057. [Google Scholar] [CrossRef]
  14. Ermolaev, G.; Grudinin, D.; Voronin, K.; Vyshnevyy, A.; Arsenin, A.; Volkov, V. Van der Waals materials for subdiffractional light guidance. Photonics 2022, 9, 744. [Google Scholar] [CrossRef]
  15. Stuhrenberg, M.; Munkhbat, B.; Baranov, D.G.; Cuadra, J.; Yankovich, A.B.; Antosiewicz, T.J.; Olsson, E.; Shegai, T. Strong light–matter coupling between plasmons in individual gold bi-pyramids and excitons in mono-and multilayer WSe2. Nano Lett. 2018, 18, 5938–5945. [Google Scholar] [CrossRef]
  16. Roy, S.; Bermel, P. Tungsten-disulfide-based ultrathin solar cells for space applications. IEEE J. Photovoltaics 2022, 12, 1184–1191. [Google Scholar] [CrossRef]
  17. Xiao, D.; Liu, G.B.; Feng, W.; Xu, X.; Yao, W. Coupled spin and valley physics in monolayers of MoS 2 and other group-VI dichalcogenides. Phys. Rev. Lett. 2012, 108, 196802. [Google Scholar] [CrossRef]
  18. Yuan, H.; Bahramy, M.S.; Morimoto, K.; Wu, S.; Nomura, K.; Yang, B.J.; Shimotani, H.; Suzuki, R.; Toh, M.; Kloc, C.; et al. Zeeman-type spin splitting controlled by an electric field. Nat. Phys. 2013, 9, 563–569. [Google Scholar] [CrossRef]
  19. Quereda, J.; Ghiasi, T.S.; You, J.S.; van den Brink, J.; van Wees, B.J.; van der Wal, C.H. Symmetry regimes for circular photocurrents in monolayer MoSe2. Nat. Commun. 2018, 9, 3346. [Google Scholar] [CrossRef]
  20. Lundt, N.; Cherotchenko, E.; Iff, O.; Fan, X.; Shen, Y.; Bigenwald, P.; Kavokin, A.; Höfling, S.; Schneider, C. The interplay between excitons and trions in a monolayer of MoSe2. Appl. Phys. Lett. 2018, 112, 031107. [Google Scholar] [CrossRef]
  21. Liu, Y.; Hu, X.; Wang, T.; Liu, D. Reduced binding energy and layer-dependent exciton dynamics in monolayer and multilayer WS2. ACS Nano 2019, 13, 14416–14425. [Google Scholar] [CrossRef] [PubMed]
  22. Blackburn, J.L.; Zhang, H.; Myers, A.R.; Dunklin, J.R.; Coffey, D.C.; Hirsch, R.N.; Vigil-Fowler, D.; Yun, S.J.; Cho, B.W.; Lee, Y.H.; et al. Measuring photoexcited free charge carriers in mono-to few-layer transition-metal dichalcogenides with steady-state microwave conductivity. J. Phys. Chem. Lett. 2019, 11, 99–107. [Google Scholar] [CrossRef] [PubMed]
  23. Zhao, P.; Kiriya, D.; Azcatl, A.; Zhang, C.; Tosun, M.; Liu, Y.S.; Hettick, M.; Kang, J.S.; McDonnell, S.; Kc, S.; et al. Air stable p-doping of WSe2 by covalent functionalization. ACS Nano 2014, 8, 10808–10814. [Google Scholar] [CrossRef] [PubMed]
  24. Cheng, Q.; Pang, J.; Sun, D.; Wang, J.; Zhang, S.; Liu, F.; Chen, Y.; Yang, R.; Liang, N.; Lu, X.; et al. WSe2 2D p-type semiconductor-based electronic devices for information technology: Design, preparation, and applications. InfoMat 2020, 2, 656–697. [Google Scholar] [CrossRef]
  25. Sokolikova, M.S.; Sherrell, P.C.; Palczynski, P.; Bemmer, V.L.; Mattevi, C. Direct solution-phase synthesis of 1T’WSe2 nanosheets. Nat. Commun. 2019, 10, 712. [Google Scholar] [CrossRef]
  26. Xu, K.; Wang, F.; Wang, Z.; Zhan, X.; Wang, Q.; Cheng, Z.; Safdar, M.; He, J. Component-controllable WS2(1-x)Se2x nanotubes for efficient hydrogen evolution reaction. ACS Nano 2014, 8, 8468–8476. [Google Scholar] [CrossRef]
  27. Sierra-Castillo, A.; Haye, E.; Acosta, S.; Bittencourt, C.; Colomer, J.F. Synthesis and characterization of highly crystalline vertically aligned WSe2 nanosheets. Appl. Sci. 2020, 10, 874. [Google Scholar] [CrossRef]
  28. Azam, A.; Yang, J.; Li, W.; Huang, J.K.; Li, S. Tungsten diselenides (WSe2) quantum dots: Fundamental, properties, synthesis and applications. Prog. Mater. Sci. 2023, 132, 101042. [Google Scholar] [CrossRef]
  29. Ding, Y.; Liu, L.; Wang, C.; Li, C.; Lin, N.; Niu, S.; Han, Z.; Duan, J. Bioinspired near-full transmittance MgF2 window for infrared detection in extremely complex environments. ACS Appl. Mater. Interfaces 2023, 15, 30985–30997. [Google Scholar] [CrossRef]
  30. Li, Z.; Lin, J.; Jia, X.; Li, X.; Li, K.; Wang, C.; Sun, K.; Ma, Z. High efficiency femtosecond laser ablation of alumina ceramics under the filament induced plasma shock wave. Ceram. Int. 2024, 50, 47472–47484. [Google Scholar] [CrossRef]
  31. Li, Z.; Lin, J.; Wang, C.; Li, K.; Jia, X.; Wang, C.; Duan, J. Damage performance of alumina ceramic by femtosecond laser induced air filamentation. Opt. Laser Technol. 2025, 181, 111781. [Google Scholar] [CrossRef]
  32. Ye, F.; Musselman, K.P. Synthesis of low dimensional nanomaterials by pulsed laser ablation in liquid. APL Mater. 2024, 12, 050602. [Google Scholar] [CrossRef]
  33. Tselikov, G.I.; Ermolaev, G.A.; Popov, A.A.; Tikhonowski, G.V.; Panova, D.A.; Taradin, A.S.; Vyshnevyy, A.A.; Syuy, A.V.; Klimentov, S.M.; Novikov, S.M.; et al. Transition metal dichalcogenide nanospheres for high-refractive-index nanophotonics and biomedical theranostics. Proc. Natl. Acad. Sci. USA 2022, 119, e2208830119. [Google Scholar] [CrossRef]
  34. Chernikov, A.S.; Tselikov, G.I.; Gubin, M.Y.; Shesterikov, A.V.; Khorkov, K.S.; Syuy, A.V.; Ermolaev, G.A.; Kazantsev, I.S.; Romanov, R.I.; Markeev, A.M.; et al. Tunable optical properties of transition metal dichalcogenide nanoparticles synthesized by femtosecond laser ablation and fragmentation. J. Mater. Chem. C 2023, 11, 3493–3503. [Google Scholar] [CrossRef]
  35. Ye, F.; Ayub, A.; Karimi, R.; Wettig, S.; Sanderson, J.; Musselman, K.P. Defect-Rich MoSe2 2H/1T Hybrid Nanoparticles Prepared from Femtosecond Laser Ablation in Liquid and Their Enhanced Photothermal Conversion Efficiencies. Adv. Mater. 2023, 35, 2301129. [Google Scholar] [CrossRef]
  36. Al-Kattan, A.; Nirwan, V.P.; Popov, A.; Ryabchikov, Y.V.; Tselikov, G.; Sentis, M.; Fahmi, A.; Kabashin, A.V. Recent advances in laser-ablative synthesis of bare Au and Si nanoparticles and assessment of their prospects for tissue engineering applications. Int. J. Mol. Sci. 2018, 19, 1563. [Google Scholar] [CrossRef]
  37. Popov, A.A.; Tselikov, G.; Dumas, N.; Berard, C.; Metwally, K.; Jones, N.; Al-Kattan, A.; Larrat, B.; Braguer, D.; Mensah, S.; et al. Laser-synthesized TiN nanoparticles as promising plasmonic alternative for biomedical applications. Sci. Rep. 2019, 9, 1194. [Google Scholar] [CrossRef]
  38. Popov, A.A.; Swiatkowska-Warkocka, Z.; Marszalek, M.; Tselikov, G.; Zelepukin, I.V.; Al-Kattan, A.; Deyev, S.M.; Klimentov, S.M.; Itina, T.E.; Kabashin, A.V. Laser-ablative synthesis of ultrapure magneto-plasmonic core-satellite nanocomposites for biomedical applications. Nanomaterials 2022, 12, 649. [Google Scholar] [CrossRef]
  39. Al-Kattan, A.; Tselikov, G.; Metwally, K.; Popov, A.A.; Mensah, S.; Kabashin, A.V. Laser ablation-assisted synthesis of plasmonic Si@ Au core-satellite nanocomposites for biomedical applications. Nanomaterials 2021, 11, 592. [Google Scholar] [CrossRef]
  40. Zavidovskiy, I.A.; Martynov, I.V.; Tselikov, D.I.; Syuy, A.V.; Popov, A.A.; Novikov, S.M.; Kabashin, A.V.; Arsenin, A.V.; Tselikov, G.I.; Volkov, V.S.; et al. Leveraging Femtosecond Laser Ablation for Tunable Near-Infrared Optical Properties in MoS2-Gold Nanocomposites. Nanomaterials 2024, 14, 1961. [Google Scholar] [CrossRef]
  41. Voß, D.; Krüger, P.; Mazur, A.; Pollmann, J. Atomic and electronic structure of WSe2 from ab initio theory: Bulk crystal and thin film systems. Phys. Rev. B 1999, 60, 14311. [Google Scholar] [CrossRef]
  42. Neilson, K.M.; Hamtaei, S.; Nassiri Nazif, K.; Carr, J.M.; Rahimisheikh, S.; Nitta, F.U.; Brammertz, G.; Blackburn, J.L.; Hadermann, J.; Saraswat, K.C.; et al. Toward Mass Production of Transition Metal Dichalcogenide Solar Cells: Scalable Growth of Photovoltaic-Grade Multilayer WSe2 by Tungsten Selenization. ACS Nano 2024, 18, 24819–24828. [Google Scholar] [CrossRef] [PubMed]
  43. Nassiri Nazif, K.; Nitta, F.U.; Daus, A.; Saraswat, K.C.; Pop, E. Efficiency limit of transition metal dichalcogenide solar cells. Commun. Phys. 2023, 6, 367. [Google Scholar] [CrossRef]
  44. Rahman, M.A. Performance analysis of WSe2-based bifacial solar cells with different electron transport and hole transport materials by SCAPS-1D. Heliyon 2022, 8, e09800. [Google Scholar] [CrossRef]
  45. Wang, L.; Yu, J. Chapter 1—Principles of photocatalysis. In S-scheme Heterojunction Photocatalysts; Yu, J., Zhang, L., Wang, L., Zhu, B., Eds.; Interface Science and Technology; Elsevier: Amsterdam, The Netherlands, 2023; Volume 35, pp. 1–52. [Google Scholar] [CrossRef]
  46. Easy, E.; Gao, Y.; Wang, Y.; Yan, D.; Goushehgir, S.M.; Yang, E.H.; Xu, B.; Zhang, X. Experimental and computational investigation of layer-dependent thermal conductivities and interfacial thermal conductance of one-to three-layer WSe2. ACS Appl. Mater. Interfaces 2021, 13, 13063–13071. [Google Scholar] [CrossRef]
  47. Li, Z.; Wang, Y.; Jiang, J.; Liang, Y.; Zhong, B.; Zhang, H.; Yu, K.; Kan, G.; Zou, M. Temperature-dependent Raman spectroscopy studies of 1–5-layer WSe2. Nano Res. 2020, 13, 591–595. [Google Scholar] [CrossRef]
  48. Abu Serea, E.S.; Orue, I.; Garcia, J.A.; Lanceros-Mendez, S.; Reguera, J. Enhancement and tunability of plasmonic-magnetic hyperthermia through shape and size control of Au: Fe3O4 janus nanoparticles. ACS Appl. Nano Mater. 2023, 6, 18466–18479. [Google Scholar] [CrossRef]
  49. He, X.N.; Gao, Y.; Mahjouri-Samani, M.; Black, P.; Allen, J.; Mitchell, M.; Xiong, W.; Zhou, Y.; Jiang, L.; Lu, Y. Surface-enhanced Raman spectroscopy using gold-coated horizontally aligned carbon nanotubes. Nanotechnology 2012, 23, 205702. [Google Scholar] [CrossRef]
  50. Sil, S.; Kuhar, N.; Acharya, S.; Umapathy, S. Is chemically synthesized graphene ‘really’a unique substrate for SERS and fluorescence quenching? Sci. Rep. 2013, 3, 3336. [Google Scholar] [CrossRef]
  51. Watanabe, H.; Hayazawa, N.; Inouye, Y.; Kawata, S. DFT vibrational calculations of rhodamine 6G adsorbed on silver: Analysis of tip-enhanced Raman spectroscopy. J. Phys. Chem. B 2005, 109, 5012–5020. [Google Scholar] [CrossRef]
  52. Meng, W.; Hu, F.; Jiang, X.; Lu, L. Preparation of silver colloids with improved uniformity and stable surface-enhanced Raman scattering. Nanoscale Res. Lett. 2015, 10, 34. [Google Scholar] [CrossRef] [PubMed]
  53. Lai, K.; Zhang, Y.; Du, R.; Zhai, F.; Rasco, B.A.; Huang, Y. Determination of chloramphenicol and crystal violet with surface enhanced Raman spectroscopy. Sens. Instrum. Food Qual. Saf. 2011, 5, 19–24. [Google Scholar] [CrossRef]
  54. Jebakumari, K.E.; Murugasenapathi, N.; Palanisamy, T. Engineered two-dimensional nanostructures as SERS substrates for biomolecule sensing: A review. Biosensors 2023, 13, 102. [Google Scholar] [CrossRef] [PubMed]
  55. Lv, Q.; Tan, J.; Wang, Z.; Yu, L.; Liu, B.; Lin, J.; Li, J.; Huang, Z.H.; Kang, F.; Lv, R. Femtomolar-level molecular sensing of monolayer tungsten diselenide induced by heteroatom doping with long-term stability. Adv. Funct. Mater. 2022, 32, 2200273. [Google Scholar] [CrossRef]
  56. Le Ru, E.; Etchegoin, P. Principles of Surface-Enhanced Raman Spectroscopy: And Related Plasmonic Effects; Elsevier: Amsterdam, The Netherlands, 2008. [Google Scholar]
  57. Majumdar, D.; Jana, S.; Ray, S.K. Gold nanoparticles decorated 2D-WSe2 as a SERS substrate. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2022, 278, 121349. [Google Scholar] [CrossRef]
  58. Chen, M.; Liu, D.; Du, X.; Lo, K.H.; Wang, S.; Zhou, B.; Pan, H. 2D materials: Excellent substrates for surface-enhanced Raman scattering (SERS) in chemical sensing and biosensing. TrAC Trends Anal. Chem. 2020, 130, 115983. [Google Scholar] [CrossRef]
  59. Tang, X.; Hao, Q.; Hou, X.; Lan, L.; Li, M.; Yao, L.; Zhao, X.; Ni, Z.; Fan, X.; Qiu, T. Exploring and Engineering 2D Transition Metal Dichalcogenides toward Ultimate SERS Performance. Adv. Mater. 2024, 36, 2312348. [Google Scholar] [CrossRef]
  60. Zograf, G.P.; Petrov, M.I.; Makarov, S.V.; Kivshar, Y.S. All-dielectric thermonanophotonics. Adv. Opt. Photonics 2021, 13, 643–702. [Google Scholar] [CrossRef]
  61. Kivshar, Y. The rise of Mie-tronics. Nano Lett. 2022, 22, 3513–3515. [Google Scholar] [CrossRef]
Figure 1. (a) Crystal structure of bulk WSe2; (b) EDX characterization of bulk WSe2 crystal; (c) SAED characterization of microscopic WSe2 flake. Yellow arrows with indices denote reciprocal lattice vectors. (d) Schematic view of PLAL.
Figure 1. (a) Crystal structure of bulk WSe2; (b) EDX characterization of bulk WSe2 crystal; (c) SAED characterization of microscopic WSe2 flake. Yellow arrows with indices denote reciprocal lattice vectors. (d) Schematic view of PLAL.
Nanomaterials 15 00004 g001
Figure 2. (a) Typical TEM image of the synthesized WSe2 NPs; (b) SEM photographs at inclined view revealing the spherical shape of NPs; (c) EDX spectrum from the synthesized NPs showing the elemental composition of WSe2 NPs. Copper signal is from the TEM grid. (d) SAED on synthesized WSe2 NPs with the most visible dhkl lines; (e) TEM image of a single nanoparticle showing its polycrystalline structure; (f) Raman spectra of bulk WSe2 crystal and synthesized NPs, separated by centrifugation at different rotation speeds. Excitation wavelength was 532 nm.
Figure 2. (a) Typical TEM image of the synthesized WSe2 NPs; (b) SEM photographs at inclined view revealing the spherical shape of NPs; (c) EDX spectrum from the synthesized NPs showing the elemental composition of WSe2 NPs. Copper signal is from the TEM grid. (d) SAED on synthesized WSe2 NPs with the most visible dhkl lines; (e) TEM image of a single nanoparticle showing its polycrystalline structure; (f) Raman spectra of bulk WSe2 crystal and synthesized NPs, separated by centrifugation at different rotation speeds. Excitation wavelength was 532 nm.
Nanomaterials 15 00004 g002
Figure 3. Differential centrifugation of WSe2 NPs. (a) Size distributions and average sizes of nanoparticles, obtained at different rotational speeds and measured by counting on TEM image (violet curve) and by dynamic light scattering spectroscopy (blue curve). (b) Measured extinction spectra for WSe2 colloids with various NP average diameter D . (c) Calculated extinction spectra (total and contributions from electric and magnetic dipole channels) for a spherical WSe2 NP with a homogenized isotropic dielectric function, obtained from bulk WSe2 crystal data in (d) as ε a v = 2 ε o / 3 + ε e / 3 . (d) Optical constants of bulk WSe2; (e) Image of bottles with centrifugated WSe2 colloidal solutions in DI water.
Figure 3. Differential centrifugation of WSe2 NPs. (a) Size distributions and average sizes of nanoparticles, obtained at different rotational speeds and measured by counting on TEM image (violet curve) and by dynamic light scattering spectroscopy (blue curve). (b) Measured extinction spectra for WSe2 colloids with various NP average diameter D . (c) Calculated extinction spectra (total and contributions from electric and magnetic dipole channels) for a spherical WSe2 NP with a homogenized isotropic dielectric function, obtained from bulk WSe2 crystal data in (d) as ε a v = 2 ε o / 3 + ε e / 3 . (d) Optical constants of bulk WSe2; (e) Image of bottles with centrifugated WSe2 colloidal solutions in DI water.
Nanomaterials 15 00004 g003
Figure 4. Photoheating response of WSe2 NPs. (a) Temperature-dependent Raman spectra at 532 nm excitation of WSe2 NPs. (b) Laser-induced heating (532 nm irradiation) and E 2 g 1 peak position shift of WSe2 NPs and bulk crystal. (c) Dynamics of laser-induced heating of colloids by laser diode 830 nm, 1 W. (d) Measured extinction spectra of WSe2 and Si water colloids, normalized at photoheating wavelength 830 nm. (e) Time-resolved photoheating of WSe2 and Si colloids irradiated by the 830 nm laser; both heating (laser is on) and cooling (laser is off) steps of the experiment are shown. (f) Optical extinction and absorption curves of WSe2 colloid from (d), prepared for the photoheating experiment with a tunable laser source. Absorption curve is calculated using photothermal conversion coefficients, indicated with black arrows and obtained experimentally from the photoheating experiments at different laser wavelengths (see the main text).
Figure 4. Photoheating response of WSe2 NPs. (a) Temperature-dependent Raman spectra at 532 nm excitation of WSe2 NPs. (b) Laser-induced heating (532 nm irradiation) and E 2 g 1 peak position shift of WSe2 NPs and bulk crystal. (c) Dynamics of laser-induced heating of colloids by laser diode 830 nm, 1 W. (d) Measured extinction spectra of WSe2 and Si water colloids, normalized at photoheating wavelength 830 nm. (e) Time-resolved photoheating of WSe2 and Si colloids irradiated by the 830 nm laser; both heating (laser is on) and cooling (laser is off) steps of the experiment are shown. (f) Optical extinction and absorption curves of WSe2 colloid from (d), prepared for the photoheating experiment with a tunable laser source. Absorption curve is calculated using photothermal conversion coefficients, indicated with black arrows and obtained experimentally from the photoheating experiments at different laser wavelengths (see the main text).
Nanomaterials 15 00004 g004
Figure 5. SERS spectra of (a) rhodamine 6G (R6G) and (b) crystal violet (CV) in the concentration range 10 4 10 8 M adsorbed on WSe2 NPs substrate; (c) R6G and (d) CV in the concentration range 10 4 10 8 M adsorbed on MoS2 NPs substrate. Peaks marked by asterisks (*) correspond to MoS2.
Figure 5. SERS spectra of (a) rhodamine 6G (R6G) and (b) crystal violet (CV) in the concentration range 10 4 10 8 M adsorbed on WSe2 NPs substrate; (c) R6G and (d) CV in the concentration range 10 4 10 8 M adsorbed on MoS2 NPs substrate. Peaks marked by asterisks (*) correspond to MoS2.
Nanomaterials 15 00004 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ushkov, A.; Dyubo, D.; Belozerova, N.; Kazantsev, I.; Yakubovsky, D.; Syuy, A.; Tikhonowski, G.V.; Tselikov, D.; Martynov, I.; Ermolaev, G.; et al. Tungsten Diselenide Nanoparticles Produced via Femtosecond Ablation for SERS and Theranostics Applications. Nanomaterials 2025, 15, 4. https://doi.org/10.3390/nano15010004

AMA Style

Ushkov A, Dyubo D, Belozerova N, Kazantsev I, Yakubovsky D, Syuy A, Tikhonowski GV, Tselikov D, Martynov I, Ermolaev G, et al. Tungsten Diselenide Nanoparticles Produced via Femtosecond Ablation for SERS and Theranostics Applications. Nanomaterials. 2025; 15(1):4. https://doi.org/10.3390/nano15010004

Chicago/Turabian Style

Ushkov, Andrei, Dmitriy Dyubo, Nadezhda Belozerova, Ivan Kazantsev, Dmitry Yakubovsky, Alexander Syuy, Gleb V. Tikhonowski, Daniil Tselikov, Ilya Martynov, Georgy Ermolaev, and et al. 2025. "Tungsten Diselenide Nanoparticles Produced via Femtosecond Ablation for SERS and Theranostics Applications" Nanomaterials 15, no. 1: 4. https://doi.org/10.3390/nano15010004

APA Style

Ushkov, A., Dyubo, D., Belozerova, N., Kazantsev, I., Yakubovsky, D., Syuy, A., Tikhonowski, G. V., Tselikov, D., Martynov, I., Ermolaev, G., Grudinin, D., Melentev, A., Popov, A. A., Chernov, A., Bolshakov, A. D., Vyshnevyy, A. A., Arsenin, A., Kabashin, A. V., Tselikov, G. I., & Volkov, V. (2025). Tungsten Diselenide Nanoparticles Produced via Femtosecond Ablation for SERS and Theranostics Applications. Nanomaterials, 15(1), 4. https://doi.org/10.3390/nano15010004

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop