Coatings: Zno Nanostructured Thin Films Via Supersonic Plasma Jet Deposition
Coatings: Zno Nanostructured Thin Films Via Supersonic Plasma Jet Deposition
Article
ZnO Nanostructured Thin Films via Supersonic
Plasma Jet Deposition
Chiara Carra 1 , Elisa Dell’Orto 1 , Vittorio Morandi 2 and Claudia Riccardi 1, *
1 Dipartimento di Fisica “Giuseppe Occhialini”, Università degli Studi di Milano-Bicocca, Piazza della Scienza 3,
I-20126 Milan, Italy; [email protected] (C.C.); [email protected] (E.D.)
2 Institute for Microelectronics and Microsystems, Italian National Research Council, Bologna Unit,
Via Gobetti 101, I-40129 Bologna, Italy; [email protected]
* Correspondence: [email protected]; Tel.: +39-0264482314
Received: 15 July 2020; Accepted: 8 August 2020; Published: 13 August 2020
Abstract: Zinc Oxide nanostructured thin films were grown by a novel plasma assisted vapour
deposition method, which aims to combine the versatility of deposition processes that are mediated
by plasma with the capability to control particles diffusion and nucleation. For this purpose, the
proposed approach spatially separates into two different vacuum chambers the creation of zinc oxide
from a metalorganic precursor from the actual film growth, thanks to the extraction of a supersonic
jet of plasma seeded by the precursor fragments. The characterization of the reactor in different
plasma conditions has been carried out by means of optical emission spectroscopy (OES). ZnO
films with different degrees of purity, thickness uniformity, as well as different morphologies can
be obtained varying the deposition parameters. The samples profiles have been collected in order
to evaluate deposition rates and films uniformity. The as-prepared as well as annealed thin films
were characterized by attenuated total reflection Fourier transform infrared spectroscopy (ATR-FTIR)
to evaluate their chemical composition and purity. According to Raman analyses, the annealed
samples are high-purity wurtzite-type crystalline zinc oxide films. Atomic force microscopy (AFM)
and scanning electron microscopy (SEM) confirm a surface morphology characterized by columnar
structures.
Keywords: nanostructured thin films; zinc acetylacetonate; ZnO; plasma deposition and supersonic
plasma jet
1. Introduction
Nanotechnology proves to be extremely attractive for successful applications in various fields,
such as electronics, photovoltaics, environmental protection, and chemical industry. In particular,
thin-films science and technology play a crucial role in the development of these modern and advanced
technologies [1,2]. Depending on the nanostructures features, thin films show very different electrical
and optical properties from their bulk crystalline counterparts. The downscaling process in thin film
fabrication has two major effects; firstly, the exposed surface appears to be much greater than the
thickness of the structures, thus allowing for a better efficiency in surface processes and chemical
reactivity; secondly, the dimensional quantum confinement alters the band structures of the material
leading to new and exotic optical and electrical characteristics [3]. To maximize the advantages of
this technology, it is possible to study and develop techniques to deposit nanostructured thin films
of desired characteristics by acting directly on the growth processes controlling the clusterization.
The zinc oxide, object of this work, is a semiconductor having a direct band gap of 3.10–3.37 eV
at room temperature [4]. High-quality ZnO bulk crystals have a large exciton binding energy
of ∼60 meV and a effective total radiative and nonradiative lifetime of excitons of the order of
1 ns [5]. As a result, ZnO posses an intense near-band-edge excitonic emission at room or higher
temperatures and it holds great interest as active material. Additionally, ZnO has a high-energy
radiation stability and amenability to wet chemical etching, making it a suitable candidate for
space application and fabrication of microelectronics devices, respectively. Recently, zinc oxide
nanostructures with controlled dimension, size, porosity, crystal facets, and mesoscale architectures
have raised great interest in various applications in sensoring material, photocatalysts, solar cells,
and light-emitting applications [6–9]. Among other growth methods, such as sol-gel processes,
molecular-beam epitaxy, and RF magnetron sputtering, the plasma-assisted chemical vapor deposition
is particularly interesting because of its high growth rate and large-scale applications. The high vacuum
conditions guarantee a lower surface damage and the low temperature better adhesion and smaller
grain size. Moreover, the Plasma Assisted Supersonic Jet Deposition has proven to be successful for
the growth of high-quality ZnO films with interesting nanostructured morphology.
near the Mach disk of the shock wave. Films characteristics, such as thickness, uniformity, purity, and
morphology, can be tuned by changing plasma and jet parameters.
The main vacuum vessel, a stainless-steel cylinder (410 mm height, 160 mm inner radius), has been
converted into two chambers, fabricating a cylindrical plasma chamber inside (95 mm height, 62.5 mm
inner radius) connected to the upper lid of the vessel. The experimental set-up is shown in Figure 1.
The chambers are connected through a circular converging nozzle, having a diameter of 6.9 mm,
and it accelerates the gas to a sonic speed. The substrates are fixed on an aluminum movable holder
placed at the desired distance from the nozzle, along the axis of the jet. A turbomolecular pump and
a rotary pump provide the high vacuum conditions in the chambers (the limit pressure inside the
deposition chamber is near 10−5 Pa). The gate valve of the main pumping group can be partially closed,
reducing the effective pumping speed, and allowing to set a desired pressure inside the deposition
chamber (Pd ) with negligible variations to the plasma chamber pressure (Pp ). The pressure ratio
between the two chambers is defined as R = Pp /Pd and it can be varied from 40 to 1 to model the
supersonic jet geometry and energy. During the measurements, pressure in the chambers is monitored
using two capacitance pressure gauges and a full range gauge, with an adequate pressure reading
precision in the pressure ranges considered (1–10 Pa in the plasma chamber and 0.01–10 Pa in the
deposition chamber). The metalorganic precursor is injected in the plasma chamber upon reaching
a sufficiently high level power of the plasma discharge. The solid precursor, held in an aluminum
tank outside the main vacuum vessel, is heated through a copper heating wire wrapped around
the tank, powered by a variable autotransformer. The precursor temperature can be monitored by
means of a thermocouple and varied to obtain different flows. A copper pipe, whose opening is
controlled through a Swagelok on–off valve, allows the sublimed precursor inlet. The dissociation of
the precursor occurs in a plasma of argon and oxygen: the first provides a stable plasma discharge
and the second favors the precursor oxidation. Gases are injected directly in the plasma chamber
through two different micrometric valves, in order to independently regulate their flows. The flow
regimes in the plasma chamber at pressure of 4–8 Pa can be considered continuous, since the Knudsen
number is k n < 0.01. Diffusion acts efficiently and the different species are well-mixed in the plasma.
While inside the zone of silence in the deposition chamber, k n is calculated from the estimated gas
mean free path, the collisional frequency of the supersonic jet, and the length of the Mach disk at the
chosen distance from the orifice. The pressure range for the calculation was observed experimentally
and it is between 0.13–0.25 Pa. The estimated Knudsen number is 0.04 < k n < 0.08, therefore the flow
regime in the supersonic jet region is transitional, guaranteeing an uniform jet over the substrates.
The inductively coupled source is a two- and three-quarter loop planar antenna, made of copper wire
and cooled by deionized water [12]. The antenna is placed inside the lower lid of the plasma chamber,
shielded by a Teflon scaffold and covered by an alumina disk to reduce sputtering. Feeding the antenna
through a 13.56 MHz radiofrequency power generator (PFG 1600 RF, Huttinger Elektronic, Ditzingen,
Germany) at 450 W, an uniform and stable plasma discharge generates inside the plasma chamber.
The supply RF signal is connected to the external edge of the coil with an L-type matching box along
the transmission line and the internal end is connected to the ground. Inductive sources generate quite
large plasma densities and relatively small electron temperatures at moderate pressures, in the range
of 1–100 Pa. This favors dissociation of external, weakly bound chemical groups in organic precursors,
partially preserving the backbone structures [13].
Coatings 2020, 10, 788 4 of 16
Figure 1. Schematic representation of the Plasma Assisted Supersonic Jet Deposition (PA-SJD) setup.
The plasma chamber and the deposition chamber are connected by the converging nozzle.
2.2. Diagnostics
Thermogravimetric and differential scanning calorimetry analyses of precursor were conducted
through a TGA/DSC1 STARe system (Mettler Toledo, Columbus, OH, USA), provided by a 10−6 g
resolution balance. The analyses were conducted at the same time by warming the precursor aluminum
tank with heating rate of 5 ◦ C/min at atmospheric pressure. The experiment was conducted in a
nitrogen atmosphere in order to prevent unwanted oxidation reactions. The discharge emission spectra
were recorded by means of an optical spectrometer (S2000, Ocean Insight, Largo, FL, USA), equipped
with a 10 µm slit and a 2048 pixels CCD, having a spectral band extending from 180 nm to 850 nm
and a resolution of 0.3 nm. Emission spectra were captured directly imaging the plasma region
through a quartz view-port on the top flange with an IR enhanced optical fiber (FC-IR050-2, Avantes,
Apeldoorn, The Netherlands). In order to obtain the real relative intensities of emission lines at different
wavelengths, the system was calibrated with both a deuterium and halogen lamp (Avalight-DHc,
Avantes, Apeldoorn, The Netherlands) to correct for the device sensitivity. The exposure time was
optimized to match the CCD sensitivity, avoiding overcounts. Each spectra was averaged over multiple
acquisitions for noise reduction and a dark spectrum with comparable statistics was subtracted.
For what concerns thin films, the thickness analyzes were carried out through a Dektak 8 Stylus
Profilometer (Veeco, Plainview, NY, USA), equipped with a diamond stylus and capable to provide
a vertical resolution of 1 Å. The applied stylus tracking force was 15 mg. The profilometer could
measure small vertical features ranging in height from 10 nm to 1 mm. Surface chemical composition of
samples was investigated by means of a Fourier Transform Infrared (FTIR) spectrometer (Nicolet iS10,
Thermo Fisher Scientific, Waltham, MA, USA), equipped with a ATR sampling accessory (Smart iTR,
Thermo Fisher Scientific, Waltham, MA, USA). For each spectrum, 64 coadditions scans with a spectral
resolution of 2 cm−1 were recorded. Confocal micro-Raman scattering were performed using a
LabRAM (Jobin Yvon, HORIBA, Kyoto, Kyoto, Japan) spectrometer and a He-Ne laser at 632.8 nm
as excitation source, focusing the beam on a circular spot of 1 µm in diameter through the optics
of a microscope (Olympus, Shinjuku City, Tokyo, Japan) and an objective with magnification 100×
and Numeric Aperture 0.90. The scattered light was detected by means of a Peltier- cooled silicon
charge-coupled device cooled at 200 K, after a notch filter (Kaiser Optical Systems, Ann Arbor, MI, USA)
with 15 nm of bandwidth and a monochromator with a spectral resolution of 1 cm−1 . Atomic force
microscopy (AFM) measurements were carried out using a Solver P47-PRO (NT-MDT, Moscow, Russia),
in semi-contact (tapping) mode on dry samples using HA-NC (High Accuracy Non Contact) silicon
Coatings 2020, 10, 788 5 of 16
tips (NT-MDT, Moscow, Russia) of typical spring constant 3.5 N/m. The Scanning Electron Microscope
(SEM) was a ZEISS 1530 (Jena, Germany) instrument equipped with a Schottky emitter and operated
at a primary beam energy of 10 keV.
Table 1. Summary of the bond energies between carbon, oxygen, hydrogen, and zinc in zinc
acetylacetonate hydrate [15].
Bond Energies
Bond C-C C-O C-H Zn-O O-H C=C C=O
eV 3.576 3.7 4.26 2.861 4.768 6.239 8.271
Oxygen radicals are produced in the plasma because of the dissociation of oxygen molecules:
O2 + e− −→ 2O + e−
The final component of the neutral chemistry that needs to be considered is the oxidation of the
atomic zinc released by the decomposition of zinc acetylacetonate. However, thermal decomposition
of Zn(acac)2 may be safely ignored, as no deposition is observed if plasma is not ignited [16]. This only
leaves Zn as a potential deposition precursor. Thus, the deposition process can be simply described as
a two-step process consisting of atomic zinc adsorption followed by oxidation:
Although other species could participate in the oxidation step, atomic oxygen is greatly abundant
under plasma operation and has the highest reactivity. To evaluate the temperature at which precursor
had to be heated to guarantee a sufficient flux in the plasma chamber, it was necessary to preliminary
measure the vapor pressure curves of the zinc acetylacetonate, as there were no previous reported
measures in literature. The system to measure the precursor vapor pressure curves is a reservoir
which contains the substance, maintained under vacuum by a rotary pump and provided with two
inlets, one for the thermocouple and another for the pressure gauge. The tank was heated, in order
Coatings 2020, 10, 788 6 of 16
to reproduce a thermostat, by means of a heat resistant wire wrapped around it, carrying a current
generated by a variable autotransformer. The temperature was evaluated through the thermocouple,
while the pressure in the container was measured by a capacitive pressure sensor. The system was
connected to the pumping units through a valve. The sample was degassed before proceeding
with measures by heating, cooling and, consequently, pumping of the residual gases. It should be
emphasized that the measurement of vapor pressure curves appears very difficult at less than 1 kPa,
as the data are often influenced by important systematic errors due to the sensitivity of the vacuum
gauges and the degassing of the tank walls. Additionally, literature lacks in vapor pressures data
and, where present, they can show differences of a few tens of percent. The measurements were
conducted by heating the sample and by measuring the corresponding pressure inside the reservoir.
Data exhibited a certain repeatability during the different sessions. Vapor pressure curves obtained
are fitted with the Clausius–Clapeyron equation, in order to estimate the sublimation enthalpy of the
compound. Clausius–Clapeyron equation can be rewritten arrested at the first order in a way more
useful for the purposes of this work:
b
P = cost + e a+ T (1)
where P is the pressure of the gas, T is its temperature, a is a parameter depending on vapor nature,
and b is defined as −E0 /R , being E0 the sublimation enthalpy and R the gas constant [17]. In Figure 2
is shown the fit of a typical vapor pressure curve of Zn(acac)2 hydrate, obtained heating the tank at
2 ◦ C/min. While Figure 3 shows the curve after the dehydration of the compound, carried out by
pumping the gaseous content of the tank (presumably only due to water vapor) after heating at 60 ◦ C
in vacuum conditions. Therefore curve is attributable to the anhydrous form of zinc acetylacetonate,
and it is realized by increasing the temperature with a heating rate of 2 ◦ C/min. From parameter b of
Equation (1) is therefore calculated the molar latent heat of sublimation [18] for zinc acetylacetonate
hydrate and anhydrous forms: respectively, 111 kJ/mol and 23.5 kJ/mol.
Figure 4. Zinc acetylacetonate TGA in nitrogen atmosphere at atmospheric pressure. Heating rate:
5 ◦ C/min.
The hydration water (approximately 8% of the total mass) is lost between 53 ◦ C and 100 ◦ C.
A further decomposition of the compound lasts up to 214 ◦ C, due to parallel reactions including
phase transition and the loss via thermal decomposition of acetylacetonate ligands [19]. The DSC
analysis was performed to measure the heat flow difference coming from the sample as the temperature
increases, making possible the identification of the temperatures at which exothermic or endothermic
processes occur for zinc acetylacetonate. Figure 5 shows the heat flow absorbed by the sample while
increasing temperature.
Figure 5. Zinc acetylacetonate DSC in nitrogen atmosphere at atmospheric pressure. Heating rate:
5 ◦ C/min.
Three main endothermic peaks are visible having different characteristics, indicative of their
originating process. In general, broad and asymmetrical peaks, such as those at 95 ◦ C and 170 ◦ C,
are associated with relatively slow processes, such as vapor production. Regarding zinc acetylacetonate
hydrate, the peak at 95 ◦ C may, therefore, be indicative of hydration water loss. The symmetrical
Coatings 2020, 10, 788 8 of 16
narrow peak at 131 ◦ C is associated with the melting of the anhydrous residue, confirmed by studies
conducted on anhydrous zinc acetylacetonate [20]; at 170 ◦ C is located the peak originated by its
vaporization. In literature, reheat experiments have been conducted in order to validate peaks nature.
It has been shown that the peak at lower temperature was absent during the reheating process, meaning
that the peak was due to an irreversible process and could be attributed to water loss. The peak at
131 ◦ C was also observed by reheating the sample, confirming its reversibility, thus being compatible
with melting. The fusion enthalpy per mole of anhydrous zinc acetylacetonate is evaluated by the
area under the 131 ◦ C peak, giving a value of 16.5 kJ/mol, in good agreement with the literature
(17 kJ/mol). By adding the substance vaporization enthalpy value calculated from the peak area
at 170 ◦ C (9.6 kJ/mol), the sublimation enthalpy of one mole of anhydrous zinc acetylacetonate is
26.1 kJ, in great agreement with the result from vapor pressure curve of anhydrous zinc acetylacetonate
(23.5 kJ/mol). The molar sublimation enthalpy of anhydrous zinc acetylacetonate is added to the
contribution of hydration water loss characteristic peak, centered at 95 ◦ C, in order to obtain the molar
enthalpy of the hydrate compound. The calculated value is 114 kJ per mole of zinc acetylacetonate
hydrate, in great agreement with the value measured from vapor pressure curve of 111 kJ/mol. The
previous results allowed to tune the precursor flux from 0.3 to 1 g/h.
0 5 10 15 20 25
3 3
Power Density (10 W/m )
Figure 6. E–H mode transition at 8 Pa, with plasma composition of 40% Ar and 60% O2 .
Figure 7. Carbon monoxide, hydrogen and oxygen peaks relative intensity versus precursor
temperature.
πθ
ρ(r, x ) = ρ(0, x ) cos2 θ cos2 ( ) (2)
2φ
with ρ particle density, r distance from the point where the measure is taken to the point of incidence of
the flow axis, x distance between the nozzle and the substrate, φ constant equal to 1.5 for the considered
gas mixture, and θ angle given by:
r
θ = arctan (3)
x − x0
with x0 distance between the nozzle and the point from which the flux lines would originate if it was a
point source (Figure 8). The estimated values of the collected data led to calculation of the maximum
thickness of the deposit.
Coatings 2020, 10, 788 10 of 16
In Figure 9, examples of typical deposit profiles with respect to growth rates are displayed.
Deposition rates at 9 mm from the nozzle, with a pressure in the plasma chamber of 8 Pa and a 40%
Ar and 60% O2 plasma and a precursor deposition temperature of 100 ◦ C are of about 30 nm/min.
This suggests that the growth rate can be tuned varying the precursor temperature, in order to obtain
the more desired thickness. Only a slightly major precursor flux is obtained by halving the total
pressure in the plasma chamber, presumably with consequences on film morphology. The most
thick deposit is 475 nm high with a deposition rate of 34 nm/min. Defining the aspect ratio as the
proportionality between film thickness and its width at half maximum, it can be appreciated that
the aspect ratio increases while increasing the deposition rate. Thus, by augmenting the quantity of
clusters deposited, the film profile becomes sharper.
500
19.4 nm/min
25.0 nm/min
400
30.2 nm/min
Thickness (nm)
300
200
100
-8 -6 -4 -2 0 2 4 6 8
Figure 9. Data and fit of deposits distribution centered at the jet axis as a function of the growth rate.
an efficient precursor fragmentation, indeed the higher Ar pressure is the more -CHx groups are
retained in films [27]. Moreover, at low oxygen concentration, CO absorption bands are not present.
Therefore augmenting the quantity of oxygen in the plasma, a better purity is achieved, and it is
almost total when oxygen constitutes the 75% of the plasma. Finally, it could be observed a strong
broad band absorption in the spectrum of the most pure sample. This is in good agreement with a
hypothesis of polycrystalline films, in which the light absorption is dominated by electron-scattering
mechanisms that are linked to grain boundaries in the materials and directly proportional to a power
of the incident light wavelength.
Figure 10. ATR-FTIR analysis of samples deposited (in red) with different plasma compositions at 4 Pa
total treatment pressure with respect to the precursor spectrum (in green).
modes are neither Raman nor IR active. As derived from the Raman tensors, all of the Raman active
modes can be observed in backscattering geometry, except for the E1 (LO) mode. Table 2 shows the
frequency and the symmetry of the fundamental optical modes in wurtzite ZnO [28,29]. Figure 11
shows the Raman spectrum of an annealed thin film in a Ar 1.6 Pa + O2 2.4 Pa mixture. The measured
absorption is in excellent agreement with peaks enlisted in Table 2. Therefore, the nature of the
deposited films can be reasonably attributed to zinc oxide in its wurtzite structure.
Table 2. Frequency and simmetry of the fundamental optical modes in wurtzite ZnO.
Figure 11. Raman spectrum of the thickest film deposited, to overwhelm the silicon substrate noise.
AFM and SEM analyses are performed to investigate the film surface morphology. Figure 12 shows
the results of an AFM analysis performed on a thin film deposited at Pp = 8 Pa. The morphological
structure of the film is displayed in a planar and 3D view, grains are visible on sample surface.
The surface roughness is calculated as the root mean square of the height distribution [30] and reported
in Table 3. The maximum height is 22 nm over the surface.
Figure 12. Example of Atomic Force Microscopy (AFM) results of a deposited film.
Coatings 2020, 10, 788 13 of 16
Figure 13 shows SEM images performed on annealed samples. The samples are deposited in two
different total treatment pressure in the plasma chamber (the first Pp = 8 Pa, the second Pp = 4 Pa)
keeping fixed the Ar/O2 ratio in the mixture. In both cases, films show a uniform distribution of
columnar structures with impervious grain boundaries. The obliqueness of the structures is due to
shadowing effect related to the geometry of the supersonic jet. Columnar structures are well defined,
since their first deposition steps and increase their dimensions as the film grows. Structures are larger
for sample deposited with a lower pressure in the plasma chamber. This is in agreement with the major
precursor flux at lower pressures behaviour, as previously discussed. For films deposited at 8 Pa, the
columns show a diameter that ranges from about 15 nm to 25 nm while growing in height. For samples
at lower pressure, the diameters range from about 18 nm to 40 nm. The polycrystalline columns and
the porosity of the films could be of great interest in application such as sensor systems, for the large
surface area to volume ratio, and solar cells, since a favourite current flow direction enhances power
conversion efficiency [31].
Table 3. Surface roughness (RMS) and grains size calculated from the AFM analysis.
(a) Thin film deposited from (b) Thin film deposited from
a 3.2 Pa Ar + 4.8 Pa O2 a 1.6 Pa Ar + 2.4 Pa O2
starting mixture. starting mixture.
Figure 13. SEM micrographs of the columnar structures in two samples different in the starting plasma
pressure of 8 Pa in panel (a) and 4 Pa in panel (b).
it was possible to relate the heating temperature to the pressure of the precursor, thus permitting the
precursor injection with a controllable flux.
OES revealed the transition between the capacitively coupled mode (E mode) and the inductively
coupled mode (H mode), the latter suitable for depositions, being characterized by a major quantity of
electrons and reactive species. OES results indicate that the optimal total pressure in the first chamber
is in the 4–8 Pa range and the pressure in the second chamber between 0.1–0.4 Pa, in order to obtain
a supersonic expansion length large enough to allow depositions at distances of 3–25 mm from the
nozzle, along the jet axis. Because of the nature of the expansion, a decrease in the deposition rate
was expected by moving away from the nozzle. Therefore, the distance of the substrate from the
nozzle was set at 9 mm. The total pressure and gas composition variation influences film morphology
and chemical structure. The films characteristics were evaluated by means of contact profilometry,
ATR-FTIR spectroscopy, Raman spectroscopy, AFM and SEM. The thickness of films were less than
500 nm and the maximun growth rate was 34 nm/min. It was demonstrated that an increase in
aspect ratio is observed increasing the deposition rate. ATR-FTIR was employed in order to obtain
information about the film purity. Many CHx and COx impurities were observed in several samples,
a complete combustion occurs only at high oxygen quantities. ATR-FTIR spectra after the thermal
treatment showed the excellent removal of any impurity. The Raman spectroscopy confirmed that the
zinc oxide is coordinated in a wurtzite structure, the most common and stable form for applications.
The wurtzite structure is non-centrosymmetric, so crystals have piezoelectricity properties. This is
a basic prerequisite for applications, such as Surface Acoustic Wave (SAW) devices and actuators.
AFM measurements revealed the presence of grains with radii of the order of tens nanometers on the
film surface, having roughness of the order of nanometers. Thin films morphology is organized in
loose packed columns, as assessed by the SEM micrographs, with largest structures obtained at lower
pressures in the plasma chamber. Because of these regular and columnar structures, deposited films
are promising for a future study on their applicability as sensors, both gas sensors and biosensors,
photocatalysts for contaminants degradation, purification and solar cells. Further investigations will
be directed to the conductivity, gas sensitivity and selectivity, piezoresponse, and transparency of the
nanostructured ZnO thin films.
Author Contributions: Conceptualization, C.C. and C.R.; methodology, E.D.; validation, C.C., C.R. and E.D.;
formal analysis, C.C., C.R. and V.M.; investigation, C.C., C.R. and V.M.; resources, C.R.; data curation, E.D.;
writing—original draft preparation, C.C.; writing—review and editing, C.C. and C.R.; visualization, E.D.;
supervision, C.C. and C.R.; project administration, C.R.; funding acquisition, C.R. All authors have read and
agreed to the published version of the manuscript.
Funding: This research was partially funded by CARIPLO Fundation (“Produzione di film sottili e nanostrutture
gerarchiche di ossidi, semiconduttori e metalli mediante un plasma supersonico reattivo” Project, 2010–2013,
No. 2010.0623).
Acknowledgments: We gratefully acknowledge the technical support of Alessandro Mietner in the device
development and experiment execution.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Zanini, S.; Grimoldi, E.; Riccardi, C. Development of controlled releasing surfaces by plasma deposited
multilayers. Mater. Chem. Phys. 2013, 138, 850–855.
2. Zanini, S.; Polissi, A.; Maccagni, E.A.; Dell’Orto, E.C.; Liberatore, C.; Riccardi, C. Development of antibacterial
quaternary ammonium silane coatings on polyurethane catheters. J. Colloid Interface Sci. 2015, 451, 78–84.
3. Guozhong, C. Nanostructures and Nanomaterials: Synthesis, Properties and Applications; World Scientific:
Singapore, 2004.
4. Klingshirn, C.F.; Waag, A.; Hoffmann, A.; Geurts, J. Zinc Oxide: From Fundamental Properties towards Novel
Applications; Springer Science & Business Media: Berlin, Germany, 2010; Volume 120.
5. Özgür, Ü.; Alivov, Y.I.; Liu, C.; Teke, A.; Reshchikov, M.A.; Doğan, S.; Avrutin, V.; Cho, S.J.; Morkoç, H. A
comprehensive review of ZnO materials and devices. J. Appl. Phys. 2005, 98, 041301, doi:10.1063/1.1992666.
Coatings 2020, 10, 788 15 of 16
6. Lu, H.; Wang, S.; Zhao, L.; Li, J.; Dong, B.; Xu, Z. Hierarchical ZnO microarchitectures assembled by
ultrathin nanosheets: Hydrothermal synthesis and enhanced photocatalytic activity. J. Mater. Chem. 2011,
21, 4228–4234.
7. Zhang, H.; Sun, J.; Dagle, V.L.; Halevi, B.; Datye, A.K.; Wang, Y. Influence of ZnO facets on Pd/ZnO catalysts
for methanol steam reforming. ACS Catal. 2014, 4, 2379–2386.
8. Gao, G.; Yu, L.; Vinu, A.; Shapter, J.G.; Batmunkh, M.; Shearer, C.J.; Yin, T.; Huang, P.; Cui, D. Synthesis of
ultra-long hierarchical ZnO whiskers in a hydrothermal system for dye-sensitised solar cells. RSC Adv. 2016,
6, 109406–109413.
9. Ramadan, R.; Torres-Costa, V.; Martín-Palma, R.J. Fabrication of Zinc Oxide and Nanostructured Porous
Silicon Composite Micropatterns on Silicon. Coatings 2020, 10, 529.
10. Dell’Orto, E.; Caldirola, S.; Sassella, A.; Morandi, V.; Riccardi, C. Growth and properties of nanostructured
titanium dioxide deposited by supersonic plasma jet deposition. Appl. Surf. Sci. 2017, 425, 407–415.
11. Caldirola, S.; Roman, H.; Riccardi, C. Ion dynamics in a supersonic jet: Experiments and simulations.
Phys. Rev. E 2016, 93, 033202.
12. Biganzoli, I.; Fumagalli, F.; Di Fonzo, F.; Barni, R.; Riccardi, C. A supersonic plasma jet source for controlled
and efficient thin film deposition. J. Mod. Phys. 2012, 3, 1626.
13. Friedrich, J. The Plasma Chemistry of Polymer Surfaces: Advanced Techniques for Surface Design; John Wiley &
Sons: Hoboken, NJ, USA, 2012.
14. Baxter, J.B.; Aydil, E.S. Metallorganic chemical vapor deposition of ZnO nanowires from zinc acetylacetonate
and oxygen. J. Electrochem. Soc. 2008, 156, H52.
15. Vovna, V.; Korochentsev, V.; Dotsenko, A. Electronic structures and photoelectron spectra of zinc (II)
bis-β-diketonates. Russ. J. Coord. Chem. 2012, 38, 36–43.
16. Wolden, C.A. The role of oxygen dissociation in plasma enhanced chemical vapor deposition of zinc oxide
from oxygen and diethyl zinc. Plasma Chem. Plasma Process. 2005, 25, 169–192.
17. Macknick, A.; Prausnitz, J. Vapor pressures of heavy liquid hydrocarbons by a group-contribution method.
Ind. Eng. Chem. Fundam. 1979, 18, 348–351.
18. Lucchesi, C.A.; Lewis, W. Latent heat of sublimation of terephthalic acid from differential thermal analysis
data. J. Chem. Eng. Data 1968, 13, 389–391.
19. Musić, S.; Šarić, A.; Popović, S. Formation of nanosize ZnO particles by thermal decomposition of zinc
acetylacetonate monohydrate. Ceram. Int. 2010, 36, 1117–1123.
20. Coates, P.D. A Study of the Preparation and Pyrolysis of β-diketonate and Carboxylate Precursors to
Semiconducting Zinc Oxide Films. Ph.D. Thesis, Durham University, Durham, UK, 1994.
21. Barni, R.; Zanini, S.; Riccardi, C. Diagnostics of reactive RF plasmas. Vacuum 2007, 82, 217–219.
22. Zanini, S.; Riccardi, C.; Orlandi, M.; Grimoldi, E. Characterisation of SiOx Cy Hz thin films deposited by
low-temperature PECVD. Vacuum 2007, 82, 290–293.
23. Ashkenas, H.; Sherman, F.S. Structure and Utilization of Supersonic Free Jets in Low Density Wind Tunnels;
NTRS—NASA Technical Reports Server: Washington, DC, USA, 1965.
24. Zanini, S.; Massini, P.; Mietta, M.; Grimoldi, E.; Riccardi, C. Plasma treatments of PET meshes for fuel–water
separation applications. J. Colloid Interface Sci. 2008, 322, 566–571.
25. Hlaing Oo, W.; McCluskey, M.; Lalonde, A.; Norton, M. Infrared spectroscopy of ZnO nanoparticles
containing CO2 impurities. Appl. Phys. Lett. 2005, 86, 073111.
26. Pholnak, C.; Sirisathitkul, C.; Suwanboon, S.; Harding, D.J. Effects of precursor concentration and reaction
time on sonochemically synthesized ZnO nanoparticles. Mater. Res. 2014, 17, 405–411.
27. Zanini, S.; Ziano, R.; Riccardi, C. Stable Poly (Acrylic Acid) films from acrylic acid/argon plasmas: Influence
of the mixture composition and the reactor geometry on the thin films chemical structures. Plasma Chem.
Plasma Process. 2009, 29, 535.
28. Damen, T.C.; Porto, S.; Tell, B. Raman effect in zinc oxide. Phys. Rev. 1966, 142, 570.
29. Cuscó, R.; Alarcón-Lladó, E.; Ibanez, J.; Artús, L.; Jiménez, J.; Wang, B.; Callahan, M.J. Temperature dependence
of Raman scattering in ZnO. Phys. Rev. B 2007, 75, 165202.
Coatings 2020, 10, 788 16 of 16
30. Grimoldi, E.; Zanini, S.; Siliprandi, R.; Riccardi, C. AFM and contact angle investigation of growth and
structure of pp-HMDSO thin films. Eur. Phys. J. D 2009, 54, 165–172.
31. Trifiletti, V.; Ruffo, R.; Turrini, C.; Tassetti, D.; Brescia, R.; Di Fonzo, F.; Riccardi, C.; Abbotto, A. Dye-sensitized
solar cells containing plasma jet deposited hierarchically nanostructured TiO2 thin photoanodes. J. Mater.
Chem. A 2013, 1, 11665–11673.
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).