Open Access
5 November 2015 Femtosecond spectroscopy on MoS2 flakes from liquid exfoliation: surfactant independent exciton dynamics
Daniele Vella, Victor Vega-Mayoral, Christoph Gadermaier, Natasa Vujicic, Tetiana Borzda, Peter Topolovsek, Matej Prijatelj, Iacopo Tempra, Eva Arianna Aurelia Pogna, Giulio Cerullo
Author Affiliations +
Abstract
Ionic surfactants, which are widely used to stabilize nanomaterials in dispersions, can drastically alter the nanomaterial’s photophysical properties. Here, we use femtosecond optical spectroscopy to study the dynamics of excitons and charges in few-layer flakes of the two-dimensional semiconductor MoS2. We compare samples obtained via exfoliation in water with different amounts of adsorbed sodium cholate, obtained by repeated washing of the dried flakes. We find that the femtosecond dynamics is remarkably stable against the surfactant adsorption, with a slight increase of the initial exciton quenching occurring during the first few picoseconds as the only appreciable effect.

1.

Introduction

In the last five years, most of the know-how acquired by the carbon community and culminating in the discovery and exploitation of graphene1 has been transferred to isolate and engineer transition metal dichalcogenide (TMD) layered materials. Semiconducting TMDs, such as MoS2, WS2, MoSe2, and WSe2, can also be exfoliated into two-dimensional (2-D) layers of single unit cell thickness, suitable for logic devices and optoelectronics applications.24 A peculiarity of the semiconducting TMDs is the crossover from indirect to direct gap semiconductor when going from bulk or few-layer material to monolayers.5 This is a result of the gradual change of the band structure as a function of the number of layers due to quantum confinement.6

Liquid exfoliation has been proven to produce stable dispersions of flakes of 2-D TMDs via a procedure that involves ultrasonication and centrifugation. However, processing nanomaterials in liquid dispersion can crucially alter their optical response and their electronic properties.7 For example, during continuous wave excitation of single wall carbon nanotubes (SWNTs) covered with sodium cholate, chemisorption of oxygen leads to hole doping and photoluminescence quenching. In low pH solution, protonation of the oxygen absorbed on an SWNT with a sodium dodecyl sulfate shell bleaches the optical absorption and consequently quenches the luminescence by Auger recombination.810 Moreover, it localizes electrons on the SWNT and increases the number of holes in the valence band (VB).11 The surfactant surrounding SWNT can induce photoluminescence blinking via trapping photoexcited charges followed by a slow recombination. This process also changes the emission spectrum via Stark effect.12,13

Femtosecond pump-probe spectroscopy represents a versatile technique to study the electron dynamics in semiconducting nanomaterials. It has elucidated, for instance, the dynamics of charge carriers, excitons, and phonons14,15 or the Stark effect induced by photogenerated and trapped charge carriers.16 Here, we use femtosecond pump-probe spectroscopy to study few-layer MoS2 from liquid exfoliation in water and sodium cholate17 with different amounts of adsorbed surfactant controlled by a multistep washing protocol. We exploit the sensitivity of the pump-probe signal to the change of the electronic excited state population18 in order to investigate the effect of the surfactant on the carrier and exciton dynamics of MoS2. For different amounts of surfactant, we obtain the same shape and relaxation behavior of the pump-probe spectra. This work demonstrates the optical stability of MoS2 against surfactant adsorption, which sets it apart from other semiconducting nanomaterials.

2.

Experimental Details

2.1.

Sample Preparation

An initial solution of 4.5 g sodium cholate and 15 g of MoS2 in powder form (Sigma Aldrich) dissolved in 1 l of deionized (DI) water was ultrasonicated for 100 h, using an Elmasonic P ultrasonic bath at 37 kHz and maximum power. This solution, which contains the flakes of MoS2, was used as a starting point for producing all the samples.4

In order to obtain films of few-layer MoS2 flakes with high surfactant content (SC), 10 ml of ethanol was added to 10 ml of the main solution, which was then spray-deposited on a quartz substrate (kept at 150°C). In parallel, different washing steps have been done to produce samples with medium and low SC.

MoS2 with medium SC was obtained by centrifugation of a part of the initial solution in a Cole Parmer Power Spin BX centrifuge for 60 min at 11,000 rpm, followed by the addition of clean DI water and ultrasonication in a bath. This procedure was repeated three times to remove a part of the surfactant from the solution and from the MoS2 surface. At a later stage, ethanol was added as described above, and the mixture was sprayed on quartz. Figure 1(a) shows a transmission electron microscope image of a typical few-layer MoS2 flake obtained by this process. To further reduce the concentration of sodium cholate on the surface of MoS2, we introduced an additional step in the procedure for the medium SC samples: after each centrifugation, the solution was washed with DI water and HCl (0.1 ml of 1 M HCl per 50 ml of MoS2 dispersion). HCl reacts with sodium cholate and forms sodium chloride and cholic acid, which is soluble in ethanol. This step is expected to remove most of the sodium cholate from the flake surface. Such obtained low SC dispersion is again deposited on a quartz substrate.

Fig. 1

(a) Transmission electron microscope image of few layers of MoS2 from liquid exfoliation in deionized water and sodium cholate. (b) Absorption spectra of few-layer MoS2 ensembles from liquid exfoliation with different surfactant content (SC) and a reference sample exfoliated via a surfactant-free process normalized to the B-exciton resonance. Inset shows actual optical density. Bright-field mages from an optical microscope of MoS2 deposited on a quartz substrate: (c) low SC, (d) medium SC, (e) high SC, and (f) reference. Scale bar: 1 mm.

JNP_10_1_012508_f001.png

As a reference sample, we used MoS2 obtained in crystalline form (Wolfram Camp Mining Ltd.), expanded in hydrazine19 and exfoliated, without surfactant, in the organic solvent N-methyl-2-pyrrolidone by ultrasonication. The films were deposited on quartz substrates by a Langmuir-Blodgett process.20

2.2.

Pump-Probe Spectroscopy

The laser source was a commercial regeneratively amplified mode-locked Ti:sapphire laser (Coherent Libra), yielding pulses at 1 kHz repetition rate, with 800 nm center wavelength and 100 fs duration. The pump pulses were generated by frequency doubling the fundamental in a thin β barium borate crystal. The pump energy was 140 nJ energy per pulse, with 150μm spot size. Broadband white-light continuum pulses generated in a sapphire plate, covering the 420 to 750 nm wavelength range, were used as probe pulses. The pump and probe beams were polarized parallel to each other. After passing through the sample, the probe beam was focused onto the entrance slit of a spectrometer equipped with a 1024-pixel back-thinned CCD sensor S7030-1006 and electronics specially designed for fast read-out times and low noise.21 By recording the transmitted probe spectrum with (Ton) and without (Toff) the pump as a function of wavelength λ and pump-probe delay τ, we obtain the differential transmission (ΔT/T) spectrum as

Eq. (1)

ΔT/T(λ,τ)=[Ton(λ,τ)Toff(λ,τ)]/Toff(λ,τ).

The delay between pump and probe pulse was varied by moving a translation stage, which changes the length of the optical path of one beam with respect to the other.

2.3.

Absorption Spectroscopy

The optical absorption spectra were recorded in transmission using a V-570 Jasco spectrometer with 0.5 nm spectral resolution, equipped with a single monochromator covering the spectral range from 190 to 2500 nm (6.52 to 0.49 eV).

2.4.

Raman Spectroscopy

Raman spectroscopy of MoS2 was performed with an NT-MDT NTEGRA SPECTRA confocal Raman microscope in backscattering geometry with spectral resolution of 0.7cm1. We used the excitation line at 488 nm and a 20× objective (N.A. 0.4) to focus onto a spot size of 20μm. The Raman signals were detected with a CCD array at 80°C. We used a laser power of 5 mW.

3.

Results and Discussion

The normalized absorption spectra in Fig. 1(b) show the characteristic exciton resonances A to D at 1.8, 2.0, 2.8, and 3.0 eV. The resonances A and B are ascribed to excitons associated with the direct gap transition at the K point.22,23 The splitting originates primarily from the splitting of the VB due to spin-orbit coupling. The C exciton is ascribed either to a transition from a band below the VB toward the conduction band (CB) or to a transition from the VB into a nested region24 of the CB. The D exciton is commonly ascribed to impurities.22

Recently, it has been shown that for surfactant dispersions of MoS2, the characteristics of the absorption spectrum provide a metric for the average flake size and thickness.25 Since it has been formally proven only for aqueous dispersion, we do not extract any quantitative characterization from our absorption spectra; however, the general trends should be valid also for dried films. Specifically, the ratio between the B exciton peak and the absorption at 3.6 eV is a measure of the typical flake size, and the position of the A exciton peak is a measure of the typical flake thickness (number of layers). The samples differ very little in both respects, the most significant difference being that the reference sample, probably due to the different exfoliation protocol, seems to consist of somewhat smaller flakes. Repeated washing, on the other hand, does not appreciably affect the flake size and thickness distribution.

To evaluate the homogeneity of the films on the relevant scale of 150 microns (the spot size used for the pump-probe experiment), we show optical images of the deposited MoS2 [Figs. 1(c)1(f)]. The darker hue of green seen for the medium SC film, matched by a higher optical density and stronger Raman and pump-probe signals (see below), is due to a higher film thickness. For the used high concentration, our simple spraying technique gives little control over the spraying speed and hence the film thickness. A significant inhomogeneity is seen only for the high SC sample. The high surfactant concentration may favor lateral aggregation of flakes or have an influence on the drying dynamics of the film. However, the apparent islands and voids are small enough to mostly average out over the spot size of the pump-probe experiment.

To find out whether the inhomogeneity in the film thickness is also related to an inhomogeneity in the flake characteristics, we measured Raman spectra (depicted in Fig. 2) on each sample on four randomly selected spots with a distance of several millimeters and a spot size of 20 microns, much smaller than the characteristic length scale of the inhomogeneity. The distance between the peaks E2g1 (in-plane mode) and A1g (out-of-plane mode) is a robust measure of the flake thickness.26 We obtain Δω=24.5cm1, corresponding to an average flake thickness of four layers for the high SC sample, and Δω=25cm1, corresponding to five or more layers for all the others.

Fig. 2

Raman spectra for excitation at 488 nm in four different points of the sample: (a) low SC, (b) medium SC, (c) high SC, and (d) reference.

JNP_10_1_012508_f002.png

Given the similar electronic structure of the flakes within the same film, we can use the Raman intensity as an indication of the amount of material present.27 The thicker sample (medium SC) shows a similar Raman intensity at all spots, while for the others, the intensity varies strongly. This confirms the observation in the optical micrographs that on a length scale of 20 microns, the thicker sample is quite homogeneous, while the others are not. However, the flake thickness and hence the electronic structure is more homogeneous, and over the spot size of the pump-probe experiment (150 microns), the inhomogeneity averages out.

The ΔT/T spectra of the reference sample are shown in Fig. 3(a) for different pump-probe delays. As previously observed on MoS2, there are six salient features in the visible spectral range:28 three photobleaching peaks of the exciton absorption resonances A0, B0, and C0, and three photoinduced absorption peaks A, B, and C ascribed to photogenerated charges, one for each of the excitons. The exciton bleaching features show a fast decay during the first few picoseconds, followed by a slower decay on the few-100 ps scale. The photoinduced absorption peaks grow during the first few picoseconds and then decay with a characteristic time similar to the slow exciton decay. Charge photogeneration occurs via a combination of a direct excitation mechanism (i.e., within our instrumental resolution of 100fs), and via dissociation of hot excitons, on a time scale of a few hundred femtoseconds to few picoseconds.28 This process can be seen in Fig. 3(a) as a fast decrease of the bleaching features A0, B0, and C0 and a corresponding growth of the charge features A, B, and C during the first few picoseconds. Together with the spectral change of shape—relative increase of the absorption and decrease of the bleaching features—the peaks also undergo a blueshift with increasing delay. In a more comprehensive study with different excitation fluences, it was found that the more the excited states the more the spectra are shifted to the red.28 This has been explained as Stark effect due to photoexcited charges,16,29 inter-excitonic interaction,30 or band gap renormalization.31 Additionally, at short delays, the A and C photoinduced absorption peaks appear artificially redshifted due to the strong overlap with the A0 and C0 bleaching peaks, respectively. The fast decay of the bleaching peaks diminishes this effect with increasing pump-probe delay, resulting in an additional apparent temporal blueshift of these two photoinduced absorption features. After 10ps, there is no significant shift of the spectra anymore and all features decay with a time constant of several hundred picoseconds, as has already been observed14 on individual flakes of MoS2. Comparing the reference sample with those of different SC [Figs. 3(b) and 3(c)], one finds a trend toward slightly faster decay for higher SC. Additionally, for higher SC, the C exciton bleaching is weaker.

Fig. 3

(a) ΔT/T spectra of the reference sample at selected pump-probe delays. Comparison of ΔT/T spectra at (b) 500 fs and (c) 900 ps pump-probe delays for different samples.

JNP_10_1_012508_f003.png

The trend toward faster initial exciton decay for higher SC observed in the ΔT/T spectra is confirmed by the dynamics at representative probe energies, close to the respective spectral peaks depicted in Figs. 4(a)4(f). The timescale of the delayed charge formation [Figs. 4(a), 4(c), and 4(e)] corresponds to that of the initial exciton decay [Figs. 4(b), 4(d), and 4(f)], confirming the origin of these charges from the dissociation of excitons. Both the slow (few hundred picoseconds) decay component as well as the direct and delayed charge photogeneration yields do not vary appreciably between the different samples.

Fig. 4

ΔT/T dynamics for different samples at selected probe energies: (a) 1.74, (b) 1.84, (c) 1.92, (d) 2.02, (e) 2.34, and (f) 2.69 eV.

JNP_10_1_012508_f004.png

In conclusion, we have shown that the exciton and charge dynamics in few-layer MoS2 flakes is robust against adsorption of the anionic surfactant sodium cholate. This is in stark contrast to other nanomaterials such as carbon nanotubes32 or C60 fullerenes,33 where anionic surfactants generally lead to strong exciton quenching. This robustness highlights the viability of surfactant-assisted liquid processing (exfoliation, deposition, printing, etc) of MoS2 and other semiconducting TMDs.

Acknowledgments

D.V., V.V.-M., T.B., N.V., I.T., and E.A.A.P. performed the experiments under the guidance of G.C. P.T. and M.P. prepared the samples. D.V. and C.G. analyzed the data and wrote the paper with critical input from all authors. We appreciate the stimulating discussions with J. Strle. The research leading to these results has received funding from LASERLAB-EUROPE (grant agreement no. 284464, EC’s Seventh Framework Programme), the Marie-Curie ITN “MoWSeS,” the ERC Advanced Grant “Trajectory,” the Graphene Flagship (contract no. CNECT-ICT-604391), and the Center of Excellence in Nanoscience and Nanotechnology, Slovenia. The research was carried out in the context of the COST “Nanospectroscopy.”

References

1. 

K. S. Novoselov et al., “Two-dimensional atomic crystals,” Proc. Nat. Acad. Sci., 102 10451 –10453 (2005). http://dx.doi.org/10.1073/pnas.0502848102 JPCBFK 1520-6106 Google Scholar

2. 

Q. H. Wang et al., “Electronics and optoelectronics of two-dimensional transition metal dichalcogenides,” Nat. Nanotech., 7 699 –712 (2012). http://dx.doi.org/10.1038/nnano.2012.193 NNAABX 1748-3387 Google Scholar

3. 

B. Radisavljevic et al., “Single-layer MoS2 transistors,” Nat. Nanotech., 6 147 –150 (2011). http://dx.doi.org/10.1038/nnano.2010.279 NNAABX 1748-3387 Google Scholar

4. 

J. N. Coleman et al., “Two-dimensional nanosheets produced by liquid exfoliation of layered materials,” Science, 331 568 –571 (2011). http://dx.doi.org/10.1126/science.1194975 SCIEAS 0036-8075 Google Scholar

5. 

K. F. Mak et al., “Atomically thin MoS2: a new direct-gap semiconductor,” Phys. Rev. Lett., 105 136805 (2010). http://dx.doi.org/10.1103/PhysRevLett.105.136805 PRLTAO 0031-9007 Google Scholar

6. 

A. Splendiani et al., “Emerging photoluminescence in monolayer MoS2,” Nano Lett., 10 1271 –1275 (2010). http://dx.doi.org/10.1021/nl903868w NALEFD 1530-6984 Google Scholar

7. 

E. A. Obraztsova et al., “Effect of environment on ultrafast photoexcitation kinetics in single-wall carbon nanotubes,” Phys. Status Solidi B, 247 2831 –2834 (2010). http://dx.doi.org/10.1002/pssb.201000238 Google Scholar

8. 

F. Wang et al., “Observation of rapid Auger recombination in optically excited semiconducting carbon nanotubes,” Phys. Rev. B, 70 241403 (2004). http://dx.doi.org/10.1103/PhysRevB.70.241403 Google Scholar

9. 

A. L. Efros and M. Rosen, “Random telegraph signal in the photoluminescence intensity of a single quantum dot,” Phys. Rev. Lett., 78 1110 –1113 (1997). http://dx.doi.org/10.1103/PhysRevLett.78.1110 PRLTAO 0031-9007 Google Scholar

10. 

P. A. Frantsuzov and R. A. Marcus, “Explanation of quantum dot blinking without the long-lived trap hypothesis,” Phys. Rev. B, 72 155321 (2005). http://dx.doi.org/10.1103/PhysRevB.72.155321 Google Scholar

11. 

M. S. Strano et al., “Reversible, band-gap-selective protonation of single-walled carbon nanotubes in solution,” J. Phys. Chem. B, 107 6979 –6985 (2003). http://dx.doi.org/10.1021/jp027664a Google Scholar

12. 

O. Kiowski et al., “Single-walled carbon nanotubes show stable emission and simple photoluminescence spectra with weak excitation sidebands at cryogenic temperatures,” Phys. Rev. B, 76 075422 (2007). http://dx.doi.org/10.1103/PhysRevB.76.075422 Google Scholar

13. 

K. Matsuda et al., “Photoluminescence intermittency in an individual single-walled carbon nanotube at room temperature,” Appl. Phys. Lett., 86 123116 (2005). http://dx.doi.org/10.1063/1.1894609 APPLAB 0003-6951 Google Scholar

14. 

H. Shi et al., “Exciton dynamics in suspended monolayer and few-layer MoS2 crystals,” ACS Nano, 7 1072 –1080 (2013). http://dx.doi.org/10.1021/nn303973r ANCAC3 1936-0851 Google Scholar

15. 

T. Koyama et al., “Transient absorption kinetics associated with higher exciton states in semiconducting single-walled carbon nanotubes: relaxation of excitons and phonons,” J. Phys. Chem. C, 117 20289 –20299 (2013). http://dx.doi.org/10.1021/jp406650t JPCCCK 1932-7447 Google Scholar

16. 

G. Soavi et al., “Ultrafast charge photogeneration in semiconducting carbon nanotubes,” J. Phys. Chem. C, 117 10849 –10855 (2013). http://dx.doi.org/10.1021/jp404009z JPCCCK 1932-7447 Google Scholar

17. 

R. J. Smith et al., “Large-scale exfoliation of inorganic layered compounds in aqueous surfactant solutions,” Adv. Mater., 23 3944 –3948 (2011). http://dx.doi.org/10.1002/adma.201102584 ADVMEW 0935-9648 Google Scholar

18. 

J. Cabanillas-Gonzalez, G. Grancini and G. Lanzani, “Pump-probe spectroscopy in organic semiconductors: monitoring fundamental processes of relevance in optoelectronics,” Adv. Mater., 23 5468 –5485 (2011). http://dx.doi.org/10.1002/adma.201102015 ADVMEW 0935-9648 Google Scholar

19. 

J. Zheng et al., “High yield exfoliation of two-dimensional chalcogenides using sodium naphthalenide,” Nat. Commun., 5 2995 (2014). http://dx.doi.org/10.1038/ncomms3995 NCAOBW 2041-1723 Google Scholar

20. 

W. M. R. Divigalpitiya, R. F. Frindt and S. R. Morrison, “Thin oriented films of molybdenum disulfide,” Thin Solid Films, 186 177 –192 (1990). http://dx.doi.org/10.1016/0040-6090(90)90511-B THSFAP 0040-6090 Google Scholar

21. 

D. Polli, L. Lüer and G. Cerullo, “High-time-resolution pump-probe system with broadband detection for the study of time-domain vibrational dynamics,” Rev. Sci. Instrum., 78 103108 (2007). http://dx.doi.org/10.1063/1.2800778 RSINAK 0034-6748 Google Scholar

22. 

R. F. Frindt and A. D. Yoffe, “Physical properties of layer structures: optical properties and photoconductivity of thin crystals of molybdenum disulphide,” Proc. R. Soc. A, 273 69 –83 (1963). http://dx.doi.org/10.1098/rspa.1963.0075 Google Scholar

23. 

D. Y. Qiu, F. H. da Jornada and S. G. Louie, “Optical spectrum of MoS2: many-body effects and diversity of exciton states,” Phys. Rev. Lett., 111 216805 (2013). http://dx.doi.org/10.1103/PhysRevLett.111.216805 PRLTAO 0031-9007 Google Scholar

24. 

D. Kozawa et al., “Photocarrier relaxation pathway in two-dimensional semiconducting transition metal dichalcogenides,” Nat. Commun., 5 4543 (2014). http://dx.doi.org/10.1038/ncomms5543 NCAOBW 2041-1723 Google Scholar

25. 

C. Backes et al., “Edge and confinement effects allow in situ measurement of size and thickness of liquid-exfoliated nanosheets,” Nat. Commun., 5 4576 (2014). http://dx.doi.org/10.1038/ncomms5576 NCAOBW 2041-1723 Google Scholar

26. 

H. Li et al., “From bulk to monolayer MoS2: evolution of Raman scattering,” Adv. Funct. Mater., 22 1385 –1390 (2012). http://dx.doi.org/10.1002/adfm.201102111 AFMDC6 1616-301X Google Scholar

27. 

C. Lee et al., “Anomalous lattice vibrations of single- and few-layer MoS2,” ACS Nano, 4 2695 –2700 (2010). http://dx.doi.org/10.1021/nn1003937 Google Scholar

28. 

T. Borzda et al., “Charge photogeneration in few-layer MoS2,” Adv. Funct. Mater., 25 3351 –3358 (2015). http://dx.doi.org/10.1002/adfm.201500709 AFMDC6 1616-301X Google Scholar

29. 

C. Gadermaier et al., “Long lived charged states in single-walled carbon nanotubes,” Nano Lett., 6 301 –305 (2006). http://dx.doi.org/10.1021/nl051970t NALEFD 1530-6984 Google Scholar

30. 

S. Sim et al., “Exciton dynamics in atomically thin MoS2: interexcitonic interaction and broadening kinetics,” Phys. Rev. B, 88 075434 (2013). http://dx.doi.org/10.1103/PhysRevB.88.075434 Google Scholar

31. 

G. Tränkle et al., “General relation between band gap renormalization and carrier density in two-dimensional electron-hole plasmas,” Phys. Rev. B, 36 6712 –6714 (1987). http://dx.doi.org/10.1103/PhysRevB.36.6712 Google Scholar

32. 

C. Georgi et al., “Photoinduced luminescence blinking and bleaching in individual single-walled carbon nanotubes,” Chem. Phys. Chem., 9 1460 –1464 (2008). http://dx.doi.org/10.1002/cphc.200800179 Google Scholar

33. 

A. F. Clements et al., “Photophysical properties of C60 colloids suspended in water with Triton X-100 surfactant: excited-state properties with femtosecond resolution,” J. Phys. Chem. A, 113 6437 –6445 (2009). http://dx.doi.org/10.1021/jp8102518 JPCAFH 1089-5639 Google Scholar

Biography

Daniele Vella received his MSc degree in biomedical engineering from the University of Pisa in 2013. In October 2013, he joined the Department of Complex Matter at the Jozef Stefan Institute (JSI), Slovenia, as a PhD student. His research interests include optical and electronic properties of nanomaterials with specific attention to femtosecond spectroscopy on thin films and devices based on two-dimensional semiconducting transition metal dichalcogenides.

Christoph Gadermaier has studied applied physics and, after two years as visiting scientist at the Politecnico di Milano, Italy, obtained his PhD in 2002 from Graz University of Technology, Austria, on ultrafast photoexcitation dynamics in conjugated polymers. After two years as a postdoc at the Politecnico di Milano and two years at JSI, Slovenia, he became research associate at JSI and associate professor at the JSI International Post-Graduate School.

Biographies for the other authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Daniele Vella, Victor Vega-Mayoral, Christoph Gadermaier, Natasa Vujicic, Tetiana Borzda, Peter Topolovsek, Matej Prijatelj, Iacopo Tempra, Eva Arianna Aurelia Pogna, and Giulio Cerullo "Femtosecond spectroscopy on MoS2 flakes from liquid exfoliation: surfactant independent exciton dynamics," Journal of Nanophotonics 10(1), 012508 (5 November 2015). https://doi.org/10.1117/1.JNP.10.012508
Published: 5 November 2015
Lens.org Logo
CITATIONS
Cited by 6 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Molybdenum

Excitons

Picosecond phenomena

Femtosecond phenomena

Liquids

Spectroscopy

Absorption

Back to Top