Tailoring Photoluminescence From MoS2 Monolayers by Mie-Resonant Metasurfaces

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Article

Cite This: ACS Photonics 2019, 6, 1002−1009 pubs.acs.org/journal/apchd5

Tailoring Photoluminescence from MoS2 Monolayers by Mie-


Resonant Metasurfaces
Tobias Bucher,*,†,‡ Aleksandr Vaskin,†,‡ Rajeshkumar Mupparapu,†,‡ Franz J. F. Löchner,†,‡
Antony George,‡,§ Katie E. Chong,∥ Stefan Fasold,†,‡ Christof Neumann,§ Duk-Yong Choi,⊥
Falk Eilenberger,†,‡,#,○ Frank Setzpfandt,†,‡ Yuri S. Kivshar,∥ Thomas Pertsch,†,‡,#,○
Andrey Turchanin,§,‡,□ and Isabelle Staude†,‡

Institute of Applied Physics, Friedrich Schiller University Jena, 07745 Jena, Germany

Abbe Center of Photonics, Friedrich Schiller University Jena, 07745 Jena, Germany
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

§
Institute of Physical Chemistry, Friedrich Schiller University Jena, 07743 Jena, Germany

Nonlinear Physics Centre, Research School of Physics and Engineering, The Australian National University, Canberra, ACT 2601,
Downloaded via PRINCETON UNIV on September 18, 2023 at 19:17:29 (UTC).

Australia

Laser Physics Centre, Research School of Physics and Engineering, The Australian National University, Canberra, ACT 2601,
Australia
#
Fraunhofer-Institute for Applied Optics and Precision Engineering IOF, 07745 Jena, Germany

Max Planck School of Photonics

Jena Center for Soft Matter (JCSM), 07743 Jena, Germany
*
S Supporting Information

ABSTRACT: We experimentally investigate coupling of the photo-


luminescence (PL) from monolayers of MoS2 to Mie-resonant
metasurfaces consisting of silicon nanocylinders. By a systematic variation
of the nanocylinder diameter, we sweep the metasurface resonances over
the excitonic emission band of monolayer MoS2. We observe strong
enhancement, as well as spectral and directional reshaping of the
emission. By a comprehensive optical characterization, we unveil the
different physical factors, including electronic, photonic, and mechanical
influences, responsible for the observed PL changes. Importantly, we
show that by geometrical tuning of the nanocylinder resonances, the
emission can be tailored from occurring under very large angles to being directed out of the substrate plane. Our results
highlight the need and potential of controlling not only the photonic, but also electronic and mechanical environmental factors
for tailoring PL from TMD monolayers by integrating them in nanophotonic architectures.
KEYWORDS: light-emitting metasurfaces, Mie resonances, dielectric nanoantennas, transition metal dichalcogenides, 2D materials,
excitonic emission

V an der Waals stacked transition metal dichalcogenide


(TMD) crystals attracted massive research attention over
the past years due to their unique physical properties when
photonic applications by integrating them with resonant
photonic nanostructures.
The interaction of light with localized emitters such as
thinned down to few- or monolayer sheets.1−4 In the few-layer fluorescent molecules, nitrogen vacancy centers in diamond,
regime, their mechanical, electronic, and optical properties not and semiconductor quantum dots placed in proximity to
only remarkably deviate from the properties of respective bulk photonic and plasmonic nanostructures has been widely
materials, but also strongly depend on the number of layers investigated.15−19 Similar studies have been extended to
forming the crystal stack. Importantly, it was shown TMD monolayers with plasmonic nanostructures for enhanc-
theoretically and experimentally that some TMDs (MX2, ing their interaction with light.20−25 Recently, plasmonic
where M = Mo, W and X = S, Se) transition from indirect to nanostructures attracted further interest owing to their
direct band gap semiconductors when approaching the capabilities of routing bichiral emission of TMD monolayers,
monolayer phase,5,6 show strong excitonic effects7−9 as well targeting valleytronics applications.26−28 However, plasmonic
as large spin−orbit splitting,10 and exhibit circular dichroism at
the K/K′ valleys.11−14 These unique properties make TMDs Received: December 23, 2018
versatile emitters, which can be harnessed for light emitting Published: March 5, 2019

© 2019 American Chemical Society 1002 DOI: 10.1021/acsphotonics.8b01771


ACS Photonics 2019, 6, 1002−1009
ACS Photonics


Article

nanostructures may also induce strong local strain29,30 and can SAMPLE FABRICATION
mediate significant injection of hot carriers into TMD We fabricated hybrid structures of 1L-MoS2 supported by
monolayers, thus, quenching the emission31,32 and altering silicon nanocylinder metasurfaces using a poly(methyl
the crystalline phase.33 Therefore, alternative material plat- methacrylate) (PMMA) assisted wet-transfer process,41,42 as
forms for nanophotonic applications based on TMD depicted in Figure 2a. First, 1L-MoS2 flakes were synthesized
monolayers are highly desirable.
Resonant dielectric metasurfaces34,35 are promising candi-
dates in this respect. These metasurfaces consist of two-
dimensional arrays of dielectric nanoresonators with a size,
geometry, and spacing such that multipolar Mie-type
resonances are combined with little absorption in the visible
regime. Single dielectric nanoresonators provide only moderate
emission enhancement of emitters placed in their proximity as
compared to their plasmonic counterparts. However, their
dense arrangement in a resonant dielectric metasurface can
result in a strong collective response, which can be tailored to
simultaneously enhance the light−matter interaction and
efficiently manipulate emission directivity.36,37
Previous work has been done on integrating TMD
monolayers with resonant dielectric nanostructures. In
particular, Cihan et al.38 highlight the role of multipolar
electric and magnetic resonances of single silicon nanowires to
control the MoS2 emission directionality and spectrum,
whereas Chen et al.39 show enhanced directional emission
from WSe2 monolayers integrated with a multiresonant silicon
waveguide-grating structure. Furthermore, Zhang et al.40 have
considered Fano resonances in dielectric nanohole arrays to
simultaneously enhance the absorption and emission of MoS2
monolayers while achieving unidirectional emission.
In this work we explore the potential of resonant all-
dielectric metasurfaces to manipulate the light interaction with
molybdenum disulfide monolayers (1L-MoS2) by alteration of
their photonic and electric environment as well as their
surrounding topography. We realize a hybrid system
constituted of 1L-MoS2 placed on top of silicon nanocylinder
metasurfaces by means of a wet transfer process. Figure 1

Figure 2. (a) Scheme of the PMMA-assisted wet-transfer process. (b)


True-color optical microscope image of 1L-MoS2 flakes as CVD-
grown on the sacrificial thermal-oxide capped silicon wafer. (c) Top-
view scanning electron micrograph of a silicon metasurface on glass
used as target substrate in the fishing step. (d) True-color optical
microscope image of a silicon metasurface after fishing and removal of
the PMMA transfer-layer. (e) Cross-sectional scanning electron
micrograph of the sample shown in (d) and for a region covered by
the monolayer molybdenum disulfide flake.

by chemical vapor deposition (CVD) from solid precursors


using silicon wafers capped with a 300 nm thermal-oxide layer
as substrates43 (see Methods for details on the growth
Figure 1. Sketch of a 1L-MoS2 flake (the crystal structure is shown in process). Our growth process results in densely spaced 1L-
the inset) placed on top of a silicon nanocylinder metasurface on a
glass substrate.
MoS2 flakes with edge lengths typically reaching up to 60 μm
and is widely scalable for large area coverage. Figure 2b shows
an optical microscope image of as-grown 1L-MoS2 flakes that
displays a conceptual sketch of the resulting structures. are clearly visible as triangles (single crystal) or clusters. The
Photoluminescence (PL) spectroscopy, microscopy, and formation of clustered flakes can occur due to merging of
back-focal-plane imaging were performed to investigate the adjacent 1L-MoS2 flakes by either forming grain boundaries
interaction of the resonant dielectric nanostructures with the (polycrystalline 1L-MoS2 flake) or an overlapping region
1L-MoS2. Verification of the monolayer phase as well as further (bilayer MoS2). The latter case was excluded in our
characterization was done by means of Raman spectroscopy experiments by observing the visual change in contrast in
and second-harmonic measurements. The multidimensional optical microscope images for bilayer MoS2 as compared to
analysis allows us to isolate different environmental factors that 1L-MoS2. Note that the color impression of the 1L-MoS2
may influence the emission characteristics of the 1L-MoS2. flakes in Figure 2b originates from thin film interferences in the
1003 DOI: 10.1021/acsphotonics.8b01771
ACS Photonics 2019, 6, 1002−1009
ACS Photonics Article

Figure 3. (a) Spatial PL mappings of 1L-MoS2 on silicon nanocylinder metasurfaces shown for different samples of varying diameter D. The
metasurface areas are indicated by white dashed lines. (b) PL spectra of the same 1L-MoS2 flakes, as shown in (a), measured on top of the silicon
nanocylinder metasurface (red curves) and on bare substrate (blue curves), with each measurement position being indicated by a blue circle in (a),
respectively. The values for the bare substrate are multiplied by 3 for better visibility. For each metasurface, the relative emission enhancement
PLmetasurf/PLsubstr (green curves) as well as the transmittance spectrum (dotted black curves) are also plotted. The black dashed line denotes the
spectral position of the resonance with significant quadrupole contribution of the nanocylinders. Note that the enhancement curves were smoothed
by a three-pixel binning of the wavelength resolution of the spectrometer.

thermal-oxide capping of the growth substrate, where the blue sequently, the transfer layer can be fished from the surface by
color is determined by the thickness of the layer. Next, silicon catching it with the target substrate, that is, the fabricated
nanocylinder metasurfaces were fabricated on standard micro- metasurface sample. In the final step, the PMMA layer was
scope coverslips (borosilicate glass) using an electron-beam removed in a critical point dryer using acetone as solvent and
lithography based process (see Methods for details on the liquid carbon dioxide as supercritical fluid.42,46 In Figure 2d, a
fabrication process). Our samples consist of square arrays of top-view optical microscope image of a fabricated silicon
silicon nanocylinders with constant height H = 188 nm, lattice metasurface (dark square) after the successful transfer of 1L-
constant Λ = 560 nm, and varying diameter D = 200−350 nm. MoS2 flakes (cyan triangles) is shown. Note that the structural
Note that in contrast to previous works employing nano- integrity of the 1L-MoS2 flakes is largely preserved during the
cylinder arrays,44 the nanocylinders are designed such that they transfer. Further, we analyzed the placement of the 1L-MoS2
support optical resonances throughout the visible and near- flakes on the silicon metasurface by taking an SEM of a cross-
infrared range including the emission band of 1L-MoS2 with a section through the hybrid structure, milled using a focused
center wavelength of 660 nm. In order to assess the multipolar ion beam (see Figure 2e). To improve the quality of the cross-
composition of the considered metasurface resonances, we sectional cut, we deposited platinum on top and below the
performed numerical simulations of the metasurface trans- transferred 1L-MoS2 flake by means of electron-beam-induced
mittance, showing very good agreement with experimentally deposition (visible as grainy regions on top and in between the
measured spectra (see Supporting Information). Then, a silicon nanocylinders). The 1L-MoS2 flake is lying flat on top
multipole decomposition of the local currents inside the of the metasurface providing physical contact of the 1L-MoS2
nanoresonators was performed,45 revealing that the resonances flakes with the nanocylinders, while it is freely suspended in the
overlapping with the emission band of 1L-MoS 2 are region between adjacent nanocylinders. The suspended layer
characterized by electric dipolar as well as significant electric appears thicker in the cross section image than expected from a
and magnetic quadrupolar contributions (see Figure S4b). 1L-MoS2 flake, which is likely due to organic residue and Si/
Therefore, in the following we refer to them as resonances with SiO2 debris accumulated during the focused ion beam milling
significant quadrupole contribution. A scanning electron process. The monolayer phase of the transferred MoS2 flakes
micrograph (SEM) of a typical fabricated sample is shown in was verified by Raman spectroscopy and second-harmonic
Figure 2c. generation microscopy (see Supporting Information for
Finally, for transferring the 1L-MoS2 from the growth details). All measurements in this study were performed
under ambient conditions.


substrate to the target metasurfaces, a silicon wafer with as-
grown 1L-MoS2 flakes was spin-coated with PMMA and
immersed into a weak potassium hydroxide (KOH) solution to PHOTOLUMINESCENCE ENHANCEMENT
separate the silicon oxide layer and the PMMA. Eventually, the To investigate the effect of the metasurfaces on the emission
wafer sinks down leaving the PMMA layer (plus superficially properties of 1L-MoS2 first, we compare the A-excitonic
embedded 1L-MoS2 flakes) floating on the surface. Sub- emission of single flakes of 1L-MoS2 situated on top of a
1004 DOI: 10.1021/acsphotonics.8b01771
ACS Photonics 2019, 6, 1002−1009
ACS Photonics Article

Figure 4. (a) Relative spectral weights wX/T of exciton (red curves) and trion (blue curves) species in 1L-MoS2 situated on silicon nanocylinder
metasurfaces (solid lines) and glass substrate (dotted lines). (b) Enhancement factors of excitonic (red curve) and trionic (blue curve)
contributions. The lines were added as guides to the eye.

metasurface with that of flakes placed on the bare substrate. To performed additional measurements and data analysis in order
exclude possible variations of the emission properties between to check for their individual contributions to the observed
different flakes, we consider only flakes which are partly changes in PL. First, we quantified the reshaping by employing
situated on top of the metasurface and partly on the substrate. a multi-Lorentzian fitting scheme to the measured PL spectra
PL microscopy and spectroscopy measurements were from 1L-MoS2 supported by bare substrate, as well as by
performed using a commercially available confocal laser- silicon nanocylinder metasurfaces of varying diameters. For all
scanning microscope setup (PicoQuant, MicroTime200). For samples, two dominant contributions can be identified, and
excitation, a 532 nm pulsed laser with a repetition rate of 80 resulting fit parameters of a double-Lorentzian model show
MHz and pulse lengths of 100 ps was focused on the sample peak positions at 655 nm ± 2 nm and 670 nm ± 3 nm and full-
using a 40x/0.65 NA objective resulting in an estimated spot widths at half-maximum (fwhm) of 17 nm ± 2 nm and 38 nm
diameter of 2r = 2λ/(NA·π) ≈ 0.52 μm. The average laser ± 7 nm, respectively. The well reproduced peak positions and
power in front of the objective was ≈5 μW. The same objective widths throughout all samples suggest a global reason common
was used to collect the PL in reflection configuration. In the to all samples leading to the observed reshaping. In agreement
detection path the reflected laser light was blocked using a 550 with previously reported values,5 we attribute the peak at the
nm long-pass filter. An additional 685 nm ± 35 nm bandpass shorter wavelength to neutral A-excitonic (X) emission of 1L-
filter was used for acquiring spatial PL mappings. Measured PL MoS2. Due to the red-shift of the center wavelength and the
mappings of 1L-MoS2 flakes transferred onto four different broadening of the peak width, we attribute the second peak to
silicon nanocylinder metasurfaces with varying diameter (240, negatively charged A-trionic (T) emission induced by
260, 275, and 300 nm, as measured from SEMs) are shown in unintentional n-doping mediated by the substrate material.47,48
Figure 3a. Each mapping shows a fairly uniformly enhanced PL Even though strain modulations can lead to similar changes
signal in the area of the silicon nanocylinder metasurfaces (red-shift and broadening) in the emission spectrum of 1L-
(inside white dashed lines). Typical enhancement factors range TMDs,49 we exclude macroscopic strain as dominant cause for
from 5 to 8. the occurrence of the second peak by Raman spectroscopy (see
Further, we measured PL spectra of the same four 1L-MoS2 Supporting Information). We quantified the intensity con-
flakes. Each measurement position is indicated by a blue circle tributions of the exciton species 0X and of the trion species 0T
in the respective PL mapping in Figure 3a and was chosen to by integrating the respective Lorentzian contributions within
exclude grain boundaries or damaged regions of the 1L-MoS2 their fwhm’s. Their relative spectral weight in the overall
flakes. The resulting spectra are presented in Figure 3b. When observed emission is defined as wX/T = 0X/T/(0X + 0T). It
situated on a bare substrate (blue curves), we observe emission was measured for 1L-MoS2 on bare substrates as well as on top
centered at around 668 nm ± 5 nm. In contrast to 1L-MoS2 on of different silicon nanocylinder metasurfaces and is shown in
bare substrate, 1L-MoS2 placed on top of the silicon Figure 4a. For 1L-MoS2 on bare substrates, we observe a
nanocylinder metasurfaces (red curves) shows not only dominant trion species, whereas for 1L-MoS2 supported by the
enhanced PL but also a spectral reshaping of the emission metasurfaces both exciton and trion species are of the same
band. We observe a spectral broadening and a shift of the order of brightness. Across all samples, the spectral weight of
emission band maxima to shorter wavelengths around 655 nm the exciton species gets increased up to wX(metasurface) =
± 1 nm. The shift becomes eminent as a pronounced peak in 0.65 and a relative enhancement of the excitonic species up to
the relative PL enhancement spectra (green curve), defined as wX(metasurface)/wX(substrate) = 14 is observed, which can be
the ratio PLmetasurf/PLsubstr. Additionally, the flat but non- explained by the fact that the metasurfaces provide a regular
vanishing tails of the relative enhancement spectra are related pattern of pillars supporting the 1L-MoS2 wherever they are
to a uniformly enhanced PL in the whole range of the emission positioned directly above individual nanocylinders but
band. Note that the enhancement curves were smoothed using suspending the 1L-MoS2 in between nanocylinders of adjacent
a three-pixel binning of the wavelength resolution of the unit cells. Hence, the freestanding 1L-MoS2 is not efficiently n-
spectrometer for better visibility in the plot. doped and exhibits dominant neutral excitonic emission. From
Next, we aim to identify the physical origin of the purely geometric considerations, an even higher relative
enhancement and spectral reshaping by considering photonic enhancement of the excitonic species would be expected,
influences from the resonant metasurfaces, strain modulations pointing toward residual doping of the freestanding 1L-MoS2
from the nanostructured topography, and electronic influences due to sample impurities. We further calculated the absolute
by unintentional doping from the substrate materials. We enhancement factors ηX/T = 0 metasurf
X/T /0 substr
X/T of the exciton

1005 DOI: 10.1021/acsphotonics.8b01771


ACS Photonics 2019, 6, 1002−1009
ACS Photonics Article

Figure 5. Measured back-focal plane images of the photoluminescence emitted by 1L-MoS2 flakes on top of silicon nanocylinder metasurfaces with
diameters D = 240, 260, 275, and 300 nm collected with a 660 nm ± 5 nm bandpass filter.

(red curve) and trion (blue curve) populations mediated by diameter and supported resonances is apparent. We attribute
the silicon nanocylinder metasurfaces, as shown in Figure 4b. A this to a combination of two factors. First, the radiative rate
drastic enhancement of the absolute exciton emission up to a enhancement is expected to be rather low for the considered
factor of 170 is observed, strongly exceeding the aforemen- sample geometry.52 Second, slight differences in the 1L-MoS2
tioned relative enhancement up to a factor of 14 due to the sample quality might be introduced during the wet transfer
repopulation of exciton and trion species. However, as shown process, making a comparison of the absolute enhancement
by Shin et al.,50 the surface roughness of amorphous substrates values across many flakes and samples difficult. Note that slight
can strongly influence the radiation efficiency of 1L-MoS2. By changes in the weak photoluminescence signal emitted from
random microscopic strain modulation on the amorphous 1L-MoS2 on bare substrate can lead to stronger variations in
surface, the band structure of the supported 1L-MoS2 can the calculated enhancement factors as the emission from 1L-
locally be modified resulting in a transition from a direct to an MoS2 on bare substrate is used as reference.
indirect band gap material. Accordingly, freestanding 1L-MoS2
shows an up to 2 orders of magnitude higher radiation
efficiency, leading to the strong enhancement observed in our
■ DIRECTIONAL SHAPING OF EMISSION
While the resonant-photonic influence of the metasurface on
experiments. For the trionic emission we observe an absolute the emission spectra is weak, its influence on the 1L-MoS2
enhancement of up to 11 despite the relative depopulation of emission pattern remains to be investigated. We analyzed the
the trion contribution in 1L-MoS2 situated on top of the directionality of emission by back-focal plane (BFP) imaging of
metasurfaces. Following the explanation above, this might the PL from the hybrid structures consisting of 1L-MoS2 on
partly suggest a better surface quality of the nanostructured silicon nanocylinder metasurfaces. BFP images were recorded
silicon compared to the amorphous glass substrate which, as a using a home-built setup with a 532 nm continuous wave
result of the etching process during the fabrication of the excitation laser and a 100×/0.73NA focusing objective. The
silicon nanostructures, is expected to have a high surface average laser power at the sample position was 0.8 mW. The
roughness. Further, we also need to consider the photonic same objective was used for collecting the signal in reflection.
influence of the substrate itself which, due to its higher The signal was then coupled into a dedicated lens system for
refractive index, provides a higher density of photonic states imaging the back-focal plane of the objective onto a charge-
compared to air. Consequently, the emission efficiency can be coupled device (CCD) camera. The collected light was filtered
increased but, more importantly, the emitted radiation is using a 660 nm ± 5 nm bandpass filter. The experimental
mainly directed into the substrate halfspace.51 This effect is results for the same four metasurfaces as presented above are
reduced for 1L-MoS2 situated some distance away when being shown in Figure 5. For all measurements we observe a 4-fold
supported by the metasurfaces and, hence, leading to an symmetric pattern in the measured BFP images of 1L-MoS2 on
increased emission into the air halfspace, that is, the collection top of silicon nanocylinder metasurfaces reflecting the
side in our experiments. We also checked for a possible symmetry of the underlying nanostructures. However, strong
photonic influence on the observed change of the PL spectra of changes occur for the different nanocylinder diameters with
1L-MoS2 mediated by the resonant silicon metasurfaces. respect to the actual angular intensity distributions. To further
However, from Figure 4b, no pronounced systematic depend- understand the measurement results, we analyzed the BFP
ency of the observed enhancement factors on the nanocylinder pattern regarding the influence of the resonators (nanocylinder
1006 DOI: 10.1021/acsphotonics.8b01771
ACS Photonics 2019, 6, 1002−1009
ACS Photonics Article

diameter) and the lattice (square array arrangement). Starting containing the MoO3 and the substrates was heated to the
with the influence of the lattice, we analyzed the presence of growth temperature of 750 °C at a rate of 40 °C min−1 and
propagating lattice modes for different emission directions. For held at that temperature for 15 min. The heating of the first
a fixed wavelength (λ = 660 nm), as defined by the bandpass zone containing the sulfur was controlled such that it reaches
filter used for the BFP measurements, we calculated the 200 °C at the same time when the second zone reaches 750
number of propagating lattice modes for different emission °C. After the growth, the two zones were allowed to cool down
directions (θ,φ) and respective points in the back-focal plane slowly by turning off the furnace until the second zone reached
(kx,ky) = k0(sin(θ)cos(φ),sin(θ)sin(φ)). Regions of the back- a temperature of 350 °C, followed by a rapid cool down to
focal plane with a constant number of propagating lattice room temperature by opening the tube furnace.
modes accurately describe the underlying pattern in the Fabrication of Silicon Metasurfaces. For fabrication of
measured BFP images (see Supporting Information for the silicon metasurfaces, we first deposited thin films of
details). However, this simple analytical consideration cannot hydrogenated amorphous silicon (a-Si:H) with a thickness of
make any predictions for the amount of light which is emitted 188 nm on standard microscope coverslips (borosilicate glass)
into the directions (θ,φ) due to the presence of the resonant using plasma-enhanced chemical vapor deposition at a
scatterers. To this end, the actual farfield pattern from each temperature of 250 °C. Afterward, the substrates were spin-
unit cell needs to be considered, which will be dominated by coated with the negative-tone electron-beam resist maN-2403.
the scattering properties of the nanocylinders. As seen in the The nanocylinders were then defined by electron-beam
measurements, for increasing nanocylinder diameter D, the PL lithography in combination with inductively coupled plasma
gets redirected more strongly toward the collection objective, etching of the silicon thin film using the developed electron-
until the main lobe of the emission is directed normal to the beam resist as an etch mask. As etch gases, we used sulfur
metasurface by sweeping the resonance with significant hexafluoride (SF6, 1.8 sccm) and fluoroform (CHF3, 50 sccm).
quadrupole contribution of the nanocylinders into spectral Etching was performed at 20 °C with 10 mTorr at 500 W
overlap with the 1L-MoS2 emission band maximum. Hence, induction power and 15 W bias power. Finally, residual resist
the radiation pattern directed into the air half space can be and organic solvent residue left on the sample were removed
tailored from emission under very large angles (off-resonant using an oxygen plasma.
case) to emission directed preferentially out of the substrate
plane (resonant case). ■ ASSOCIATED CONTENT

■ *
S Supporting Information
CONCLUSION The Supporting Information is available free of charge on the
We have demonstrated experimentally the integration of CVD- ACS Publications website at DOI: 10.1021/acsphoto-
grown 1L-MoS2 with silicon nanocylinder metasurfaces and nics.8b01771.
have studied the influences of the electronic and photonic Raman (Figure S1) and second-harmonic (Figure S2)
environment on the spectral and directional properties of the analysis of transferred 1L-MoS2 flakes. Discussion of the
emitted photoluminescence. The observed spectral reshaping lattice influence on measured back-focal plane images
and strong enhancement of the measured photoluminescence (Figure S3). Numerically calculated metasurface trans-
spectra have been attributed to the charge transfer mediated mittance and corresponding multipole decomposition
repopulation of excitons and trions as well as the freestanding (Figure S4) (PDF)


nature of 1L-MoS2 on silicon nanocylinder metasurfaces. Most
importantly, by tuning a metasurface resonance with significant AUTHOR INFORMATION
quadrupole contribution to spectrally overlap with the Corresponding Author
emission band maximum of 1L-MoS2, a preferential emission *E-mail: tobias.bucher@uni-jena.de.
out of the metasurface plane could be achieved. By this, we
have demonstrated an efficient platform for accessing the ORCID
strong emission from freestanding TMD monolayers being Tobias Bucher: 0000-0003-1335-0432
directed out of the substrate plane allowing for light collection Aleksandr Vaskin: 0000-0002-3014-1002
with simple low-NA optics, a crucial step for integrating TMD Franz J. F. Löchner: 0000-0002-8003-7916
monolayers in large-scale photonic applications. Stefan Fasold: 0000-0001-5330-2465


Yuri S. Kivshar: 0000-0002-3410-812X
METHODS Isabelle Staude: 0000-0001-8021-572X
CVD Growth of Monolayer MoS2. 1L-MoS2 flakes were Notes
The authors declare no competing financial interest.


grown by chemical vapor deposition on thermally oxidized
silicon substrates (Sil’tronix, oxide thickness 300 nm, rough-
ness < 0.2 nm RMS). After cleaning the growth substrates by ACKNOWLEDGMENTS
ultrasonication in acetone for 5 min followed by washing in Financial support by the Thuringian State Government within
isopropanol, they were placed face down in a two-zone tube its ProExcellence initiative (ACP2020) and by the German
furnace with a tube diameter of 55 mm. Two alumina boats, Research Foundation (STA 1426/2-1) is gratefully acknowl-
one containing 200 mg of sulfur powder (99.98%, Sigma- edged. This research was partly funded by the German
Aldrich), the other containing ∼1 μg MoO3 powder (99.97%, Ministry of Education and Research (BMBF) under the project
Sigma-Aldrich), were placed in the first and second zone of the identifier 13N14147; responsibility for the content of this work
tube furnace, respectively, such that the substrates were facing resides with the authors. Furthermore, the authors acknowl-
the MoO3 precursors. Under atmospheric pressure and a edge their participation in the Erasmus Mundus NANOPHI
constant argon flow of 50 cm3 min−1, the second zone project, Contract Number 2013 5659/002-001. Y.S.K.
1007 DOI: 10.1021/acsphotonics.8b01771
ACS Photonics 2019, 6, 1002−1009
ACS Photonics Article

acknowledges support from the Humboldt Foundation. F.J.F.L. (17) Beams, R.; Smith, D.; Johnson, T. W.; Oh, S.-H.; Novotny, L.;
has been funded by the German Research Foundation (DFG) Vamivakas, A. N. Nanoscale Fluorescence Lifetime Imaging of an
through the International Research Training Group (IRTG) Optical Antenna with a Single Diamond NV Center. Nano Lett. 2013,
2101. This research is supported by an Australian Government 13, 3807−3811.
(18) Bauch, M.; Toma, K.; Toma, M.; Zhang, Q.; Dostalek, J.
Research Training Program (RTP) Scholarship. This work was
Plasmon-Enhanced Fluorescence Biosensors: a Review. Plasmonics
performed in part at the ACT node of the Australian National 2014, 9, 781−799.
Fabrication Facility, a company established under the National (19) Krasnok, A.; Caldarola, M.; Bonod, N.; Alú, A. Spectroscopy
Collaborative Research Infrastructure Strategy to provide and Biosensing with Optically Resonant Dielectric Nanostructures.
nano- and microfabrication facilities for Australian researchers. Adv. Opt. Mater. 2018, 6, 1701094.
This project has also received funding from the joint European (20) Najmaei, S.; Mlayah, A.; Arbouet, A.; Girard, C.; Léotin, J.;
Unions Horizon 2020 and DFG research and innovation Lou, J. Plasmonic Pumping of Excitonic Photoluminescence in
programme FLAG-ERA under a Grant TU149/9-1. We thank Hybrid MoS2-Au Nanostructures. ACS Nano 2014, 8, 12682−12689.
Stephanie Höppener and Ulrich S. Schubert for enabling our (21) Goodfellow, K. M.; Beams, R.; Chakraborty, C.; Novotny, L.;
Raman spectroscopy and microscopy studies at the JCSM. Vamivakas, A. N. Integrated nanophotonics based on nanowire


plasmons and atomically thin material. Optica 2014, 1, 149−152.
(22) Butun, S.; Tongay, S.; Aydin, K. Enhanced Light Emission from
REFERENCES Large-Area Monolayer MoS2 Using Plasmonic Nanodisc Arrays. Nano
(1) Zeng, H.; Cui, X. An optical spectroscopic study on two- Lett. 2015, 15, 2700−2704.
dimensional group-VI transition metal dichalcogenides. Chem. Soc. (23) Kern, J.; Trügler, A.; Niehues, I.; Ewering, J.; Schmidt, R.;
Rev. 2015, 44, 2629−2642. Schneider, R.; Najmaei, S.; George, A.; Zhang, J.; Lou, J.; Hohenester,
(2) Mak, K. F.; Shan, J. Photonics and optoelectronics of 2D U.; Michaelis de Vasconcellos, S.; Bratschitsch, R. Nanoantenna-
semiconductor transition metal dichalcogenides. Nat. Photonics 2016, Enhanced LightMatter Interaction in Atomically Thin WS2. ACS
10, 216. Photonics 2015, 2, 1260−1265.
(3) Schaibley, J. R.; Yu, H.; Clark, G.; Rivera, P.; Ross, J. S.; Seyler, (24) Lee, B.; Park, J.; Han, G. H.; Ee, H.-S.; Naylor, C. H.; Liu, W.;
K. L.; Yao, W.; Xu, X. Valleytronics in 2D materials. Nat. Rev. Mater. Johnson, A. C.; Agarwal, R. Fano Resonance and Spectrally Modified
2016, 1, 16055. Photoluminescence Enhancement in Monolayer MoS2 Integrated
(4) Manzeli, S.; Ovchinnikov, D.; Pasquier, D.; Yazyev, O. V.; Kis, A. with Plasmonic Nanoantenna Array. Nano Lett. 2015, 15, 3646−3653.
2D transition metal dichalcogenides. Nat. Rev. Mater. 2017, 2, 17033. (25) Chen, H.; Yang, J.; Rusak, E.; Straubel, J.; Guo, R.; Myint, Y.
(5) Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F. Atomically W.; Pei, J.; Decker, M.; Staude, I.; Rockstuhl, C.; Lu, Y.; Kivshar, Y. S.;
Thin MoS2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010, Neshev, D. Manipulation of photoluminescence of two-dimensional
105, 136805. MoSe2 by gold nanoantennas. Sci. Rep. 2016, 6, 22296.
(6) Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.-Y.; (26) Chen, H.; Liu, M.; Xu, L.; Neshev, D. N. Valley-selective
Galli, G.; Wang, F. Emerging Photoluminescence in Monolayer MoS2. directional emission from a transition-metal dichalcogenide mono-
Nano Lett. 2010, 10, 1271−1275. layer mediated by a plasmonic nanoantenna. Beilstein J. Nanotechnol.
(7) Mak, K. F.; He, K.; Lee, C.; Lee, G. H.; Hone, J.; Heinz, T. F.; 2018, 9, 780−788.
Shan, J. Tightly bound trions in monolayer MoS2. Nat. Mater. 2013, (27) Gong, S.-H.; Alpeggiani, F.; Sciacca, B.; Garnett, E. C.; Kuipers,
12, 207. L. Nanoscale chiral valley-photon interface through optical spin-orbit
(8) Shi, H.; Yan, R.; Bertolazzi, S.; Brivio, J.; Gao, B.; Kis, A.; Jena, coupling. Science 2018, 359, 443−447.
D.; Xing, H. G.; Huang, L. Exciton Dynamics in Suspended (28) Sun, L.; Wang, C.-Y.; Krasnok, A.; Choi, J.; Shi, J.; Gomez-Diaz,
Monolayer and Few-Layer MoS2 2D Crystals. ACS Nano 2013, 7, J. S.; Zepeda, A.; Gwo, S.; Shih, C.-K.; Alù, A.; Li, X. Separation of
1072−1080. valley excitons in a MoS2 monolayer using a subwavelength
(9) Ugeda, M. M.; Bradley, A. J.; Shi, S.-F.; da Jornada, F. H.; Zhang, asymmetric groove array. Nat. Photonics 2019, 13, 180−184.
Y.; Qiu, D. Y.; Ruan, W.; Mo, S.-K.; Hussain, Z.; Shen, Z.-X.; Wang, (29) Sun, Y.; Liu, K.; Hong, X.; Chen, M.; Kim, J.; Shi, S.; Wu, J.;
F.; Louie, S. G.; Crommie, M. F. Giant bandgap renormalization and Zettl, A.; Wang, F. Probing Local Strain at MX2-Metal Boundaries
excitonic effects in a monolayer transition metal dichalcogenide with Surface Plasmon-Enhanced Raman Scattering. Nano Lett. 2014,
semiconductor. Nat. Mater. 2014, 13, 1091. 14, 5329−5334.
(10) Molina-Sánchez, A.; Sangalli, D.; Hummer, K.; Marini, A.; (30) Hwang, D. Y.; Suh, D. H. Evolution of a high local strain in
Wirtz, L. Effect of spin-orbit interaction on the optical spectra of rolling up MoS2 sheets decorated with Ag and Au nanoparticles for
single-layer, double-layer, and bulk MoS2. Phys. Rev. B: Condens. surface-enhanced Raman scattering. Nanotechnology 2017, 28,
Matter Mater. Phys. 2013, 88, 045412. 025603.
(11) Cao, T.; Wang, G.; Han, W.; Ye, H.; Zhu, C.; Shi, J.; Niu, Q.; (31) Bhanu, U.; Islam, M. R.; Tetard, L.; Khondaker, S. I.
Tan, P.; Wang, E.; Liu, B.; Feng, J. Valley-selective circular dichroism Photoluminescence quenching in gold - MoS2 hybrid nanoflakes.
of monolayer molybdenum disulphide. Nat. Commun. 2012, 3, 887. Sci. Rep. 2015, 4, 5575.
(12) Zeng, H.; Dai, J.; Yao, W.; Xiao, D.; Cui, X. Valley polarization (32) Li, J.; Ji, Q.; Chu, S.; Zhang, Y.; Li, Y.; Gong, Q.; Liu, K.; Shi, K.
in MoS2 monolayers by optical pumping. Nat. Nanotechnol. 2012, 7, Tuning the photo-response in monolayer MoS2 by plasmonic nano-
490. antenna. Sci. Rep. 2016, 6, 23626.
(13) Mai, C.; Barrette, A.; Yu, Y.; Semenov, Y. G.; Kim, K. W.; Cao, (33) Kang, Y.; Najmaei, S.; Liu, Z.; Bao, Y.; Wang, Y.; Zhu, X.;
L.; Gundogdu, K. Many-Body Effects in Valleytronics: Direct Halas, N. J.; Nordlander, P.; Ajayan, P. M.; Lou, J.; Fang, Z.
Measurement of Valley Lifetimes in Single-Layer MoS2. Nano Lett. Plasmonic Hot Electron Induced Structural Phase Transition in a
2014, 14, 202−206. MoS2 Monolayer. Adv. Mater. 2014, 26, 6467−6471.
(14) Xiao, J.; Ye, Z.; Wang, Y.; Zhu, H.; Wang, Y.; Zhang, X. (34) Decker, M.; Staude, I. Resonant dielectric nanostructures: a
Nonlinear optical selection rule based on valley-exciton locking in low-loss platform for functional nanophotonics. J. Opt. 2016, 18,
monolayer WS2. Light: Sci. Appl. 2015, 4, No. e366. 103001.
(15) Bharadwaj, P.; Deutsch, B.; Novotny, L. Optical Antennas. Adv. (35) Kruk, S.; Kivshar, Y. Functional Meta-Optics and Nano-
Opt. Photonics 2009, 1, 438−483. photonics Governed by Mie Resonances. ACS Photonics 2017, 4,
(16) Barth, M.; Schietinger, S.; Schrüder, T.; Aichele, T.; Benson, O. 2638−2649.
Controlled coupling of NV defect centers to plasmonic and photonic (36) Vaskin, A.; Bohn, J.; Chong, K. E.; Bucher, T.; Zilk, M.; Choi,
nanostructures. J. Lumin. 2010, 130, 1628−1634. D.-Y.; Neshev, D. N.; Kivshar, Y. S.; Pertsch, T.; Staude, I. Directional

1008 DOI: 10.1021/acsphotonics.8b01771


ACS Photonics 2019, 6, 1002−1009
ACS Photonics Article

and Spectral Shaping of Light Emission with Mie-Resonant Silicon


Nanoantenna Arrays. ACS Photonics 2018, 5, 1359−1364.
(37) Liu, S.; et al. Light-Emitting Metasurfaces: Simultaneous
Control of Spontaneous Emission and Far-Field Radiation. Nano Lett.
2018, 18, 6906−6914.
(38) Cihan, A. F.; Curto, A. G.; Raza, S.; Kik, P. G.; Brongersma, M.
L. Silicon Mie resonators for highly directional light emission from
monolayer MoS2. Nat. Photonics 2018, 12, 284−290.
(39) Chen, H.; Nanz, S.; Abass, A.; Yan, J.; Gao, T.; Choi, D.-Y.;
Kivshar, Y. S.; Rockstuhl, C.; Neshev, D. N. Enhanced Directional
Emission from Monolayer WSe2 Integrated onto a Multiresonant
Silicon-Based Photonic Structure. ACS Photonics 2017, 4, 3031−3038.
(40) Zhang, X.; Choi, S.; Wang, D.; Naylor, C. H.; Johnson, A. T.
C.; Cubukcu, E. Unidirectional Doubly Enhanced MoS2 Emission via
Photonic Fano Resonances. Nano Lett. 2017, 17, 6715−6720.
(41) Turchanin, A.; Beyer, A.; Nottbohm, C. T.; Zhang, X.; Stosch,
R.; Sologubenko, A.; Mayer, J.; Hinze, P.; Weimann, T.; Gölzhäuser,
A. One Nanometer Thin Carbon Nanosheets with Tunable
Conductivity and Stiffness. Adv. Mater. 2009, 21, 1233−1237.
(42) Winter, A.; George, A.; Neumann, C.; Tang, Z.; Mohn, M. J.;
Biskupek, J.; Masurkar, N.; Reddy, A. L. M.; Weimann, T.; Hübner,
U.; Kaiser, U.; Turchanin, A. Lateral heterostructures of two-
dimensional materials by electron-beam induced stitching. Carbon
2018, 128, 106−116.
(43) George, A.; Neumann, C.; Kaiser, D.; Mupparapu, R.; Lehnert,
T.; Hübner, U.; Tang, Z.; Winter, A.; Kaiser, U.; Staude, I.;
Turchanin, A. Controlled growth of transition metal dichalcogenide
monolayers using Knudsen-type effusion cells for the precursors. J.
Phys. Mater. 2019, 2, 016001.
(44) Chaste, J.; Missaoui, A.; Huang, S.; Henck, H.; Ben Aziza, Z.;
Ferlazzo, L.; Naylor, C.; Balan, A.; Johnson, A. T. C.; Braive, R.;
Ouerghi, A. Intrinsic Properties of Suspended MoS2 on SiO2/Si Pillar
Arrays for Nanomechanics and Optics. ACS Nano 2018, 12, 3235−
3242.
(45) Grahn, P.; Shevchenko, A.; Kaivola, M. Electromagnetic
multipole theory for optical nanomaterials. New J. Phys. 2012, 14,
093033.
(46) Angelova, P.; et al. A Universal Scheme to Convert Aromatic
Molecular Monolayers into Functional Carbon Nanomembranes.
ACS Nano 2013, 7, 6489−6497.
(47) Ross, J. S.; Wu, S.; Yu, H.; Ghimire, N. J.; Jones, A. M.;
Aivazian, G.; Yan, J.; Mandrus, D. G.; Xiao, D.; Yao, W.; Xu, X.
Electrical control of neutral and charged excitons in a monolayer
semiconductor. Nat. Commun. 2013, 4, 1474.
(48) Scheuschner, N.; Ochedowski, O.; Kaulitz, A.-M.; Gillen, R.;
Schleberger, M.; Maultzsch, J. Photoluminescence of freestanding
single- and few-layer MoS2. Phys. Rev. B: Condens. Matter Mater. Phys.
2014, 89, 125406.
(49) Schmidt, R.; Niehues, I.; Schneider, R.; Drüppel, M.; Deilmann,
T.; Rohlfing, M.; de Vasconcellos, S. M.; Castellanos-Gomez, A.;
Bratschitsch, R. Reversible uniaxial strain tuning in atomically thin
WSe2. 2D Mater. 2016, 3, 021011.
(50) Shin, B. G.; Han, G. H.; Yun, S. J.; Oh, H. M.; Bae, J. J.; Song,
Y. J.; Park, C.-Y.; Lee, Y. H. Indirect Bandgap Puddles in Monolayer
MoS2 by Substrate-Induced Local Strain. Adv. Mater. 2016, 28, 9378−
9384.
(51) Novotny, L.; Hecht, B. Principles of Nano-Optics; Cambridge
University Press, 2012.
(52) Bouchet, D.; Mivelle, M.; Proust, J.; Gallas, B.; Ozerov, I.;
Garcia-Parajo, M. F.; Gulinatti, A.; Rech, I.; De Wilde, Y.; Bonod, N.;
Krachmalnicoff, V.; Bidault, S. Enhancement and Inhibition of
Spontaneous Photon Emission by Resonant Silicon Nanoantennas.
Phys. Rev. Appl. 2016, 6, 064016.

1009 DOI: 10.1021/acsphotonics.8b01771


ACS Photonics 2019, 6, 1002−1009

You might also like