Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Available online at www.sciencedirect.

com

ScienceDirect
Materials Today: Proceedings 2 (2015) 4247 – 4255

nanoFIS 2014 - Functional Integrated nanoSystems

Synthesis of graphene-layer nanosheet coatings by PECVD


Reinhard Kaindla,*, Georg Jakopica, Roland Reselb, Jeanine Pichlerc, Alexander Fiana,
Evelin Fisslthalerd, Werner Groggere, Bernhard C. Bayerf,g, Roland Fischerh, Wolfgang
Waldhausera
a
JOANNEUM RESEARCH, MATERIALS – Institute for Surface Technologies and Photonics, Leobner Straße 94, A-8712 Niklasdorf, Austria
b
Graz University of Technology, Institute of Solid State Physics, Petersgasse 16/II, A-8010, Austria
c
Graz University of Technology, Institute for Chemistry and Technology of Materials, Stremayrgasse 9, A-8010 Graz, Austria
d
Graz Centre for Electron Microscopy, Steyrergasse 17/III, A-8010 Graz, Austria
e
Graz University of Technology, Institute for Electron Microscopy and Nanoanalysis, Steyrergasse 17/III, A-8010 Graz, Austria
f
Faculty of Physics, University of Vienna, Boltzmanngasse 5, A-1090 Vienna, Austria
g
Department of Engineering, University of Cambridge, Cambridge CB3 0FA, UK
h
Graz University of Technology, Institute of Anorganic Chemistry, Stremayrgasse 9/IV, A-8010 Graz, Austria

Abstract

Plasma enhanced chemical vapor deposition (PECVD) is a potentially industrially scalable route for graphene synthesis. Thin
carbon coatings were deposited by PECVD from C2H2 and post-deposition irradiated by Ar+. Irradiation with 400 eV accelerated
Ar+, deposition temperatures of 400 °C and post deposition annealing up to 800 °C result in crystal growth and graphitization, i.e.
increased fraction of C-C sp2 bonding. The same effect is achieved by deposition on copper and nickel foils, and interestingly on
NaCl. For irradiated samples a two-layer graphene-layer nanosheet morphology is indicated from our characterization, with a
variation of film properties across the film thickness.
Copyright © 2014 Elsevier Ltd. All rights reserved.
© 2015 Elsevier
Selection Ltd. All rights
and peer-review reserved.
under responsibility of Conference Committee of nanoFIS 2014 - Functional Integrated nanoSystems.
Selection and peer-review under responsibility of Conference Committee of nanoFIS 2014 - Functional Integrated nanoSystems.
Keywords: graphene; graphite; physical vapor deposition; plasma enhanced chemical vapor deposition

* Corresponding author. Tel.: +43-316-876-2303; fax: +43-316-876-9-2303.


E-mail address: reinhard.kaindl@joanneum.at

2214-7853 © 2015 Elsevier Ltd. All rights reserved.


Selection and peer-review under responsibility of Conference Committee of nanoFIS 2014 - Functional Integrated nanoSystems.
doi:10.1016/j.matpr.2015.09.010
4248 Reinhard Kaindl et al. / Materials Today: Proceedings 2 (2015) 4247 – 4255

1. Introduction

Two-dimensional (2D) crystalline materials such as graphene have excellent physical and chemical properties, e.g.
low absorbance of visible light, high mechanical flexibility, high thermal and electrical conductivity and a high
Young’s modulus [1-4]. In order to bring graphene from the laboratory to the market-place, large-scale production
methods are however necessary. Using catalytic chemical vapor deposition (CVD) up to 30-inch graphene films have
been manufactured, employing a roll-to-roll transfer process from the metallic catalyst substrates, and thus enabling
touch-screen device manufacture [5]. Alternatively however, plasma-enhanced chemical vapor deposition (PECVD)
and physcial vapor deposition (PVD) such as magnetron sputtering or pulsed laser deposition (PLD) are potentially
attractive for large scale graphene synthesis. PECVD and PVD could enable graphene deposition directly on a larger
variety of substrates than CVD, since no catalytically active substrate is required in PECVD and PVD. Also the
temperature range for successful film deposition is commonly lower in PECVD and PVD compared to classic thermal
CVD.
For instance, single and few-layer graphene was synthesized by PECVD on nickel and copper substrates and at
relatively low temperatures down to 475 °C [6-10]. Few reports exist about graphene synthesis by magnetron
sputtering: Single and double layer films were deposited directly on MgO(111) at growth temperatures around 727 °C
[11,12]. Metal-incorporated, graphitic microporous carbon, containing highly ordered graphene layers, has been
obtained from a thin amorphous carbon film by a metal nanoparticle sputtering process at ambient substrate
temperature [13]. Nano-graphene structures were deposited by near-infrared PLD ablation of graphite on Si [14]. Few-
layer graphene was fabricated on metal substrates with an energetic carbon source supplied by PLD [15]. Ni-catalysed
solid state re-crystallisation of PVD carbon/Nickel bilayer stacks led upon vacuum annealing to controlled high quality
graphene growth [7,16,17]. Thus industrially compatible PECVD and PVD techniques could have the potential to
grow graphene with properties similar to current state-of-the-art catalytic CVD, but to date a detailed understanding
of the relevant PECVD/PVD growth parameters remains elusive.
In this study nanometer-thin carbon coatings deposited by PECVD methods were investigated. It is shown that, by
carefully choosing deposition parameters, substrate materials and post-deposition treatments by Ar+-ion irradiation
coatings containing nanosheets of graphene- and graphite layers can be synthesized. In accordance with the
recommendations for two-dimensional carbon materials nomenclature [18] the term graphene-layer nanosheet
coatings is introduced instead of “graphene-like” and “graphite-like”, which was used in previous studies [19-21] as
an analogy to diamond-like carbon coatings (DLC). Graphene-layer nanosheet coatings contain single-atom sheets of
hexagonally arranged, sp2 bonded carbon atoms in a partly 3D ordered graphitic and turbostratic faulted structure. The
coatings were deposited on nickel, copper, Si(100), Si/SiO2, MgO(100) and NaCl substrates and analyzed by
spectroscopic ellipsometry, Raman spectroscopy, X-Ray reflectivity, four-point conductivity measurements, X-ray
photoelectron spectroscopy and high-resolution transmission electron microscopy.

2. Experimental

2.1. Coating deposition

The coatings were deposited on Si(100) and Si/SiO2 wafer (100 nm thermal oxide), 0.1 mm thick nickel and copper
foil (Alfa Aesar Puratronic, 99.999%), MgO(100), 1 mm thick polished NaCl windows (Edmund Optics), and on
standard glass sample holder. All substrates were additionally cleaned with acetone and isopropanol before deposition.
Most of the substrates were plasma treated by 60 seconds Ar+ irradiation in the energy range 800 eV, using a linear
anode layer source, prior to PECVD. Deposition of the PECVD coatings was done from C2H2 and C2H2-Ar mixtures
at an acceleration voltage of 3 kV, gas flow 20 sccm C2H2 and 15 sccm C2H2 and 5 sccm Ar, respectively. Base and
deposition pressure in the chamber were in the range 4•10 -3 and 3•10-2 Pa. Substrate temperatures ranged from room
temperature to 400 °C. After deposition, selected coatings were irradiated by Ar+ at variable energy from ~400 up to
800 eV. Selected coatings were also annealed in Ar atmosphere at atmospheric pressure and up to 800 °C. Thickness
of the deposited coatings varied from below 10 up to 100 nm.
Reinhard Kaindl et al. / Materials Today: Proceedings 2 (2015) 4247 – 4255 4249

The PECVD grown films were further compared to benchmark Cu-catalyzed CVD graphene, which has been
transferred to Si/SiO2 using standard polymer scaffold/catalyst wet etch transfers (without any additional cleaning
applied) [22].

2.2. Raman spectroscopy

Confocal Raman spectra were obtained with a HORIBA JOBIN YVON LabRam-HR 800 Raman micro-
spectrometer. The sample was excited by the 532 nm emission line of a 30 mW Nd-YAG-laser under an OLYMPUS
100u objective (N.A. = 0.9). The size and power of the laser spot on the surface were approximately 1 μm and 0.5
mW, respectively. The scattered light was dispersed by a grating with 1800 lines/mm and collected by a 1024 u 256
open electrode CCD detector. The spectral resolution, determined by measuring the Rayleigh line, was about 1.4 cm-1.
Third order polynomial and convoluted Gauss-Lorentz functions were applied for background correction, peak fitting
and determination of full width at half maximum (FWHM) and intensity ratio of the D- and G-peak [I(D)/I(G)]. The
wavenumber accuracy of about 0.5 cm-1 was achieved by adjusting the zero-order position of the grating and regularly
checked by a Neon spectral calibration lamp.

2.3. Ellipsometry

The ellipsometric data have been derived using a J.A.Woollam Co. Inc. VASE ellipsometer with rotating analyser,
Si/GaAs tandem diode detector, equipped with phase compensator (auto retarder) to minimize psi/delta errors by input
polarization optimization. Radiation source is a Xe high pressure short arc lamp with subsequent Czerny-Turner
scanning monochromator. The beam diameter was approx. 3 mm; the angles of incidence used for data acquisition
were 65°, 70° and 75°, spectral range from 300 nm to 1300 nm. The optical constants have been calculated by fitting
the experimental data with electrodynamic model based thin film calculations using a generalized multi oscillator
model including short- and long wavelength poles and offset to build the dielectric function.

2.4. X-Ray Photoelectron Spectroscopy (XPS)

The bonding states of the thin carbon layers were identified by XPS, using a Multiprobe Surface Analysis System
from Omicron Nanotechnology. Al-KD1-radiation (1486.7 eV) from a DAR400 x-ray source was monochromatized
with an XM500 quartz crystal monochromator (energy width 0.15 eV); the photoelectron energies were measured in
an EA125 hemispherical electron analyzer with an overall spectrometer- resolution of 0.5 eV. Ar+ sputter cleaning or
heating prior to the measurement was avoided in order to identify the original surface chemistry. The fraction of each
chemical compound in the film was determined by deconvolution of the detailed element peaks with a
Gaussian/Lorentzian function.

2.5. X-Ray Reflectivity (XRR)

All the measurements were conducted with a PANalytical Empyrean diffractometer using a sealed copper tube (λKα
= 0.154 nm). Additionally to the X-ray tube a monochromatizing multilayer mirror is installed at the primary side.
The incident slit was chosen to a divergence of (1/32)° throughout all measurements and a fixed beam-mask of 10
mm.
The diffracted beam optics of the device consisted of an anti-scatter slit of 100 μm in front of a Soller slit with a
divergence of 0.02 rad and a PIXcel solid state detector. The evaluation and fitting of the experimental data from the
detector unit was executed with the X’PERT Reflectivity software.

2.6. Electrical characterisation

Sheet resistance was determined by I/V-characteristics using a 4-point-prober from Jandel Engineering Ltd. with
currents forced in the range of several nA.
4250 Reinhard Kaindl et al. / Materials Today: Proceedings 2 (2015) 4247 – 4255

2.7. Transmission Electron Microscopy (TEM)

After PECVD, the NaCl substrate was dissolved in distilled water in order to float off the PECVD carbon film
without further impairment, and to pick it up with a copper TEM grid. For high resolution transmission electron
microscopy (HRTEM) investigations a Tecnai TF20 equipped with a field emission gun and featuring an acceleration
voltage of 200 kV was used (FEI Company). This microscope is equipped with a High Resolution Gatan Imaging
Filter including UltraScan CCD camera (2048 pixels x 2048 pixels; Gatan Inc.) that was used to capture the
micrographs. The experiments were carried out under such illumination conditions that no changes in the structure of
the irradiated areas during the image acquisition were detected. Finally, the resulting images were evaluated and
analyzed with Digital Micrograph (Gatan Inc.).

3. Results and discussion

Fig. 1. Refraction index of a 52 nm thick as deposited (as dep.) PECVD carbon coating on Si(100) wafer (solid line), a 10 nm thick coating irradiated
by 400 eV Ar+ for 0.5 minutes (stippled-dotted), a 103 nm thick coating irradiated for 1 minute (dotted) and a 92 nm thick coating irradiated for 40
minutes (fine stippled). All coatings were deposited at an acceleration voltage of 3 kV. Coating thickness was also determined by ellipsometry after
Ar+ etching. For comparison, reconstructed nx (diamonds) and nz (squares) of graphite [23], modelled n of transferred graphene flakes on Si/SiO2
(circles) [24], and modelled n calculated from measured ellipsometric data of transferred CVD graphene on SiO2/Si wafer is displayed (coarse
stippled line).

In Figure 1 the refraction index n of four PECVD carbon coatings on Si/SiO2 wafers is displayed. The coatings
were deposited at room temperature at 3 kV accelerating voltage and irradiated after deposition by Ar + at 400 eV.
From previous experiments by the author’s group it is known that 3 kV acceleration voltages result in amorphous
carbon coatings with a more sp2-rich, graphitic binding structure [25,26]. It was expected that Ar + radiation would
preferably etch the at room temperature thermodynamically more unstable sp3 coordinated clusters and chains in the
coating while stabilizing the graphitic domains. The wavelength dependent refraction index n of such treated coatings
seem to support this assumption. In contrast to the as deposited, all irradiated coating show a local maximum of n in
the UV-range followed by more or less constant or decreasing values in the visible to near infrared (NIR) range,
independent of thickness. Neither reconstructed nx within nor nz perpendicular to the basal plane of graphite show the
local maximum in the UV range. The modelled n of graphene flakes and from a measured reference CVD transferred
onto Si/SiO2 in contrast trend toward n of the irradiated carbon coatings, although the absolute values are different
(Fig. 1). We note that the employed models are highly sensitivity to the effective optical thickness of thin layers,
Reinhard Kaindl et al. / Materials Today: Proceedings 2 (2015) 4247 – 4255 4251

which can deviate from “geometrical” thickness. It was found that the optical thickness of graphene changed 30%
from flake to flake [24]. For the reference CVD graphene curve in Fig. 1 a optical thickness of graphene of 0.63 nm
was used, which was determined from XRR measurements, whereas ref. [24] calculated with a fixed thickness of
0.335 nm. For the transferred CVD graphene also possible effects from polymer residues from transfer could further
influence the extracted absolute n values [27,28].
To sum up, n of as deposited PECVD carbon coatings on Si(100) wafer and nx and nz of graphite changes rather
continous with the wavelength in the range 300 – 1000 nm whereas Ar+ irradiated coatings, transferred CVD graphene
and graphene flakes on Si/SiO2 wafer show a local maximum in the UV and rather constant n in the visible and near
infrared range.

Fig. 2. Full width at half maximum (FWHM) of the G-peak (diamonds) and intensity ratio of the D- and the G-peak I(D)/I(G) (squares) plotted
against post deposition Ar+ ion irradiation energy for PECVD coatings deposited at room temperature on Si/SiO2 wafers.

Fig. 3. FWHM(G) (diamonds) and I(D)/I(G) (squares) for PECVD coatings deposited in the same run at 400 °C on five different substrates.

The effects of Ar+ irradiation to PECVD coatings can be demonstrated by Raman spectroscopic data (Figure 2).
Full width at half maximum (FWHM) of the G-peak around 1580 cm-1 and intensity ratios of the D-peak around 1360
cm-1 divided by the G-peak [I(D)/I(G)] are plotted against post deposition Ar+ irradiation energy. Room temperature
deposited PECVD coatings irradiated at lower Ar+ energy around 400 eV yielded lower FWHM(G) and higher
I(D)/IG) than those irradiated at 800 eV. From amorphous carbon to nanocrystalline graphite I(D)/I(G) first increases
4252 Reinhard Kaindl et al. / Materials Today: Proceedings 2 (2015) 4247 – 4255

(from zero) but then from nanocrystalline graphite to high quality graphene I(D)/I(G) again decreases and vanishes to
zero [29]. These two regimes with a turning point were also observed in single-layer graphene samples exposed to
pulses of oxygen plasma [30]. In the regime from amorphous carbon to nanocrystalline graphite FWHM(G) is
normally inverse correlated with the in-plane correlation length or in other words size of ordered crystal domains and
occurrence of a D-mode implies existence of sp2 bonded carbon in six-membered rings [25,26,31,32]. PECVD
amorphous hydrogenated carbon (a-C:H) films for example showed decreasing FWHM(G) and increasing I(D)/I(G),
related with an increasing size of graphitic clusters and an decreasing population of sp2 sites arranged in chains [33].
Therefore Fig. 2 indicates that for our PECVD films larger crystallites and a higher fraction of sp2 bonded carbon are
obtained at the lower post-deposition irradiation energies of 400 eV.

Fig. 4. X-ray photoelectron spectra of C 1s of PECVD carbon coatings on copper (blue) and Si/SiO2 wafer (red).

Coatings were also deposited at elevated temperatures up to 400 °C and a part of the coatings were annealed after
deposition in Ar atmosphere up to 800 °C. This resulted in decreasing FWHM(G) down to 100 cm-1 and increasing
I(D)/I(G) to >1, suggesting temperature induced crystal growth and graphitization in the regime from amorphous
carbon to nanocrystalline graphite. Reducing the partial pressure of C2H2 by 25 % and adding Ar instead caused an
average reduction of the coatings FWHM(G) from 148 down to 118 cm-1 and a slight decrease in I(D)/I(G) from 1.28
to 1.09. This indicates a comparably weak influence of Ar gas species at low partial pressures to the structure of
coatings deposited from C2H2 precursor gas.
The influence of substrate material was evaluated by depositing carbon coatings on glass, Si/SiO2 wafer, copper
and nickel foils and NaCl at 400 °C in C2H2/Ar atmosphere, followed by Ar+ irradiation with energy of 400 eV.
FWHM(G) is highest for the glass (136 cm-1) and interestingly lowest for the NaCl (105 cm-1) substrate; the highest
I(D)/I(G) of 1.37 was determined for copper substrate, the lowest (1.13) for Si/SiO2 wafer (Figure 3). Combined, this
suggests larger crystallite size for NaCl, copper and nickel as well as higher fraction of sp2 bonding for carbon coatings
deposited on these materials.
The C 1s XPS spectra of two PECVD coatings which were deposited on copper and Si/SiO2 wafer at 400 °C in
C2H2 atmosphere and irradiated by Ar+ with energy of 800 eV are displayed in Figure 4. The thickness of the carbon
coatings, estimated by the intensity ratios of C 1s compared to Cu 2p3 and Si 2p peaks, is around 5 – 6 nm for the
copper and 3 nanometer for the Si/SiO2 substrate. For the coating on copper three peaks at 284.5, 285.4 and 288.2 eV
were deconvoluted, with the majority intensity ~284.5 eV. Typical binding energies for sp2 and sp3 carbon in graphite
and graphene are ~284 and 284.8 eV [34], respectively, while in graphene oxide amounts of sp3, C-O and C=O
Reinhard Kaindl et al. / Materials Today: Proceedings 2 (2015) 4247 – 4255 4253

bondings at 288 and 289 eV can be observed. It is therefore indicated that the surface of the coating on copper is
relatively enriched in sp3 carbon and partially oxidized.
Four peaks at 283, 284.1, 285.8 and 287.7 were deconvoluted for the coating on Si/SiO2. The lowest energy peak
at 283 eV can be assigned to C-Si and is probably attributed to the pre-deposition Ar+ plasma treatment. The intense
peak around 284 eV suggests dominating sp2 carbon. Again oxidation is indicated by the peak at 288 eV and another
one ~286 eV, a C-O-C component typically observed in samples that have been exposed to air.
Raman spectra of these very thin coatings could not be obtained because of low signal-to-noise ratios. However, a
higher C-C sp2 content was estimated for thicker coatings on copper in contrast to Si/SiO2 based on Raman (see above).
This apparent contradiction could be explained by the much higher sampling volume of Raman spectroscopy around
>5 μm3 [35] whereas XPS detects only electrons that escape from the top 0 to 10 nm of the material being analyzed
[36]. This suggests a different C-C sp2/sp3 ratio in the bulk and surface of our Ar +-irradiated PECVD carbon coatings.
On Cu a sp3-enriched layer is indicated on top of a sp2-rich “bulk”. In contrast, for Si/SiO2, a C-C sp2 enriched surface
layer is indicated. It has been shown previously that such sp2-rich layers commonly exist on the surface of amorphous
hydrogenated (a-C:H) and tetrahedral (ta-C) carbon films [37,38]. Consistently, the surface resistivity of the coatings
deposited on Si/SiO2 and NaCl in this study (Table 1) is in the range of a-C:H carbon films (~7.5 kΩ/sq [38]).

Fig. 5. (a) High-resolution transmission electron microscope image. (b) The distance of the parallel planes of carbon atoms within the area

marked by a red rectangle in (a) is 0.35 nm, which corresponds to the typical graphite lattice plane distance in crystallographic c-direction.

In order to gain some information about thickness and density of the carbon layer deposited on Si/SiO2 wafer XRR-
scans were performed. The thickness of the room temperature as-deposited coating on Si/SiO2 wafer of about 51 nm
and density of 1.8 g/cm3 corresponds to thickness measurements from stylus profilometry and its density is in the
range expected for a-C:H PECVD coatings [39]. A coating deposited at 400 °C and irradiated by Ar+ ions accelerated
to 400 eV could not be sufficiently fitted with a single carbon layer model. Better agreement between measured XRR-
profile and the fit was achieved with a 5 nm carbon top layer 1 and a 93 nm thick carbon layer 2 with very low mass
densities. This could be explained by void formation, oriented nano-sized graphene particles and/or multi-layer
4254 Reinhard Kaindl et al. / Materials Today: Proceedings 2 (2015) 4247 – 4255

graphene growth in a near turbostratic mode on the surface, which are induced by elevated deposition temperatures
and post-deposition Ar ion treatments. X-ray diffraction and XRR studies of carbon films deposited at temperatures
up to 800 °C by pulsed laser ablation [14] describe a textured, open film structure with aggregation of aromatic planes
and voids and density dropdown. Near turbostratic growth and low layer density was found on the surface of multi-
layer graphene/4H-SiC systems [40] However, a partial coverage of the surface by water deposits or other
contaminants from sample handling and storage cannot be totally excluded.
A strong indication for the co-existence of oriented graphene nanosheets and turbostratic areas in PECVD coatings
deposited at elevated temperatures and irradiated by Ar+ is the HRTEM image shown in Figure 5. The coating with
an average thickness of 70 nm was deposited on a NaCl crystal, which was dissolved prior to the TEM investigations.
In the HRTEM image small areas with parallel planes of carbon atoms that are aligned vertically with respect to the
substrate can be observed. The measured distance between the lattice planes is about 0.35 nm, which is in accordance
with the typical lattice plane distance of graphite (0.34 nm) [41]; the size of the ordered domains is in the range of
several nanometers (inset shows measurement of lattice plane distance at the area indicated by the red rectangle in
Fig. 5). These carbon nanosheets are surrounded by rather unstructured, probably turbostratic areas with a maximum
diameter of about 10 nm.

Table 1. Sheet resistance of PECVD carbon coatings on SiO2/Si, glass, and NaCl, and reference values for graphene and
amorphous carbon. FLG – few-layer graphene; MLG – monolayer graphene
Coating/substrate Deposition (gas Sheet resistance [Ω/sq]
composition)
PECVD/glass PECVD (C2H2-Ar) 6•104 – 1•107
PECVD/NaCl PECVD (C2H2-Ar) 8•104 – 2•105
PECVD/SiO2 PECVD (C2H2-Ar) 1•104 – 2•105
FLG [6] PECVD (C2H2-Ar) 4•103 – 9•103
MLG [42] CVD 468 – 770
a-C:H [43,44] CVD, PVD 3.7•104 – 1•106

4. Conclusions

In summary, we investigated thin carbon coatings deposited by PECVD from C 2H2 and irradiated by Ar+ after
deposition. We find that irradiation with Ar+ accelerated to 400 eV, elevated deposition temperatures of 400 °C and
post deposition annealing in Ar atmosphere up to 800 °C result in crystal growth and graphitization, i.e. increased
fraction of C-C sp2 bonding. The same effect is achieved by deposition on copper and nickel foils, and interestingly
NaCl. For irradiated samples a layered nanosheet structure with graphene-layers in the bulk and sp2 or sp3 enriched
top layers depending upon the substrate material is indicated, as well as a variation of film properties across the
thickness.

References

[1] H. Gwon, H.-S. Kim, K.U. Lee, D.-H. Seo, Y.C. Park, Y.-S. Lee, B.T. Ahn, K. Kang, Energy & Env. Sci. 4 (2011) 1277
[2] S. Pang, Y. Hernandez, X. Feng, K. Mullen, Adv. Mater. 23 (2011) 2779-95
[3] S. Stankovich, D.A. Dikin, G.H. Dommett, K.M. Kohlhaas, E.J. Zimney, E.A. Stach, R.D. Piner, S.T. Nguyen, R.S. Ruoff, Nature 442 (2006)
282-6
[4] A.H. Castro Neto, F. Guinea, N.M.R. Peres, K.S. Novoselov, A.K. Geim, Rev. Mod. Phys. 81 (2009) 109-162
[5] S. Bae, H. Kim, Y. Lee, X. Xu, J.-S. Park, Y. Zheng, J. Balakrishnan, T. Lei, H. Ri Kim, Y.I. Song, Y.-J. Kim, K.S. Kim, B. Ozyilmaz, J.-H.
Ahn, B.H. Hong, S. Iijima, Nat Nano 5 (2010) 574-578
[6] K.-J. Peng, C.-L. Wu, Y.-H. Lin, Y.-J. Liu, D.-P. Tsai, Y.-H. Pai, G.-R. Lin, J. Mater. Chem. C 1 (2013) 3862-3870
[7] L. Baraton, Z. He, C.S. Lee, J.L. Maurice, C.S. Cojocaru, A.F. Gourgues-Lorenzon, Y.H. Lee, D. Pribat, Nanotechnology 22 (2011) 085601
[8] J.L. Qi, W.T. Zheng, X.H. Zheng, X. Wang, H.W. Tian, Appl. Surf. Sci. 257 (2011) 6531-6534
[9] S.-H. Chan, S.-H. Chen, W.-T. Lin, M.-C. Li, Y.-C. Lin, C.-C. Kuo, Nanoscale Res. Lett. 8 (2013) 285
[10] B.-J. Lee, T.-W. Lee, S. Park, H.-Y. Yu, J.-O. Lee, S.-H. Lim, G.-H. Jeong, Mater. Lett. 65 (2011) 1127-1130
[11] J.A. Kelber, M. Zhou, S. Gaddam, F.L. Pasquale, L.M. Kong, P.A. Dowben, ECS Transactions 45 (2012) 49-61
Reinhard Kaindl et al. / Materials Today: Proceedings 2 (2015) 4247 – 4255 4255

[12] J.A. Kelber, S. Gaddam, C. Vamala, S. Eswaran, P.A. Dowben, Direct graphene growth on MgO(111) by physical vapor deposition:
interfacial chemistry and band gap formation, 1 ed.; Drouhin, H.-J. M., et al., (eds.) SPIE, San Diego, California, USA, 2011, Vol. 8100, pp
81000Y-13
[13] A.N. Banerjee, B.-K. Min, S.W. Joo, Appl. Surf. Sci. 268 (2013) 588-600
[14] E. Cappelli, S. Orlando, M. Servidori, C. Scilletta, Appl. Surf. Sci. 254 (2007) 1273-1278
[15] A.T.T. Koh, Y.M. Foong, D.H.C. Chua, Diamond Relat. Mater. 25 (2012) 98-102
[16] R.S. Weatherup, C. Baehtz, B. Dlubak, B.C. Bayer, P.R. Kidambi, R. Blume, R. Schloegl, S. Hofmann, Nano Lett. 13 (2013) 4624-4631
[17] G. Pan, M. Heath, D. Horsell, M.L. Wears, arXiv:1209.0489 [cond-mat.mtrl-sci] (2013)
[18] A. Bianco, H.-M. Cheng, T. Enoki, Y. Gogotsi, R.H. Hurt, N. Koratkar, T. Kyotani, M. Monthioux, C.R. Park, J.M.D. Tascon, J. Zhang,
Carbon 65 (2013) 1-6
[19] R. Kaindl, R. Resel, J. Pichler, R. Fischer, G. Jakopic, W. Waldhauser, Graphene- and graphite-like coatings by PVD and PECVD,
Presented at CBC 2014 – 3rd Austrian Symposium on Carbon Based Coatings, Seggauberg, Leibnitz, (2014)
[20] R. Kaindl, J. Pichler, R. Fischer, G. Jakopic, W. Waldhauser, Mitt. Oest. Mineral. Ges. 159 (2013) 70
[21] R. Kaindl, J. Pichler, R. Fischer, G. Jakopic, W. Waldhauser, Deposition of ultra-thin graphene-like coatings by PECVD methods Presented
at Graphene 2013, Bilbao, Spain, (2013)
[22] P.R. Kidambi, B.C. Bayer, R. Blume, Z.J. Wang, C. Baehtz, R.S. Weatherup, M.G. Willinger, R. Schloegl, S. Hofmann, Nano Lett. 13
(2013) 4769-4778
[23] V.G. Kravets, A.N. Grigorenko, R.R. Nair, P. Blake, S. Anissimova, K.S. Novoselov, A.K. Geim, Phys. Rev. B: Condens. Matter 81 (2010)
155413
[24] J.W. Weber, V.E. Calado, M.C.M. van de Sanden, Appl. Phys. Lett. 97 (2010) 091904
[25] M. Kahn, M. Cekada, R. Berghauser, W. Waldhauser, C. Bauer, C. Mitterer, E. Brandstätter, Diamond Relat. Mater. 17 (2008) 1647-1651
[26] M. Kahn, M. Čekada, T. Schöberl, R. Berghauser, C. Mitterer, C. Bauer, W. Waldhauser, E. Brandstätter, Thin Solid Films 517 (2009)
6502-6507
[27] M. Kratzer, B.C. Bayer, P.R. Kidambi, A. Matković, R. Gajić, A. Cabrero-Vilatela, R.S. Weatherup, A.W. Hofmann, C. Teichert, Appl.
Phys. Lett. 106 (2015) 103101
[28] A. Matković, U. Ralević, M. Chhikara, M.M. Jakovljević, D. Jovanović, G. Bratina, R. Gajić, J. Appl. Phys. 114 (2013) 093505
[29] A.C. Ferrari, Solid State Commun. 143 (2007) 47-57
[30] I. Childres, L.A. Jauregui, J. Tian, Y.P. Chen, New J. Phys. 13 (2011) 025008
[31] J. Robertson, Mater. Sci. Eng., R 37 (2002) 129-281
[32] A.C. Ferrari, and J. Robertson, Phys. Rev. B: Condens. Matter 64 (2001) 075414-1-075414-13
[33] M. Kahn, S. Paskvale, M. Čekada, T. Schöberl, W. Waldhauser, C. Mitterer, P. Pelicon, E. Brandstätter, Diamond Relat. Mater. 19 (2010)
1245-1248
[34] J.C. Lascovich, and S. Scaglione, Appl. Surf. Sci. 78 (1994) 17-23
[35] L. Markwort, B. Kip, E. DaSilva, B. Rousell, Appl. Spectrosc. 49 (1995) 1411-1430
[36] S. Hüfner, Photoelectron Spectroscopy: Principles and Applications, Springer Science & Business Media, 2003
[37] C.A. Davis, K.M. Knowles, G.A.J. Amaratunga, Surf. Coat. Technol. 76–77, Part 1 (1995) 316-321
[38] S.S. Tinchev, arXiv:1311.0605 [cond-mat.mtrl-sci] (2013)
[39] A.C. Ferrari, A. Libassi, B.K. Tanner, V. Stolojan, J. Yuan, L.M. Brown, S.E. Rodil, B. Kleinsorge, J. Robertson, Phys. Rev. B: Condens.
Matter 62 (2000) 11089-11103
[40] J. Hass, R. Feng, J.E. Millan-Otoya, X. Li, M. Sprinkle, P.N. First, C. Berger, W.A.d. Heer, E.H. Conrad, Phys. Rev. B: Condens. Matter 75
(2007) 214109
[41] R.M. Hazen, R.T. Downs, A.P. Jones, L. Kah, Rev. Mineral. Geochem. 75 (2013) 7-46
[42] H. Ning, Gratom-M1-Cu Certificate of Quality, Bluestone Global Tech, Wappingers Falls, NY 12590, 2013, p 2
[43] S. Miyagawa, S. Nakao, J. Choi, M. Ikeyama, Y. Miyagawa, Nucl. Instrum. Methods Phys. Res., Sect. B 242 (2006) 346-348
[44] R.U.R. Sagar, X. Zhang, C. Xiong, Y. Yu, Carbon 76 (2014) 64-70

You might also like