Next Article in Journal
Deposition Mechanism and Characterization of Plasma-Enhanced Atomic Layer-Deposited SnOx Films at Different Substrate Temperatures
Previous Article in Journal
Reduced Graphene Oxide-Extracellular Matrix Scaffolds as a Multifunctional and Highly Biocompatible Nanocomposite for Wound Healing: Insights into Characterization and Electroconductive Potential
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Deposition of ZnO Nanowires on CVD-Graphene/Copper Substrates

1
Group of Nanomaterials and Renewable Energies, Laboratory of Solid State Physics, Faculty of Sciences Dhar El Mahraz, Sidi Mohammed Ben Abdellah University, P.O. Box 1796, Atlas Fez 30000, Morocco
2
National Institute of Materials Physics, Atomistilor 405A, 077125 Magurele, Romania
*
Author to whom correspondence should be addressed.
Nanomaterials 2022, 12(16), 2858; https://doi.org/10.3390/nano12162858
Submission received: 28 July 2022 / Revised: 11 August 2022 / Accepted: 16 August 2022 / Published: 19 August 2022

Abstract

:
ZnO nanostructures were electrochemically synthesized on Cu and on chemical vapor deposited (CVD)-graphene/Cu electrodes. The deposition was performed at different electrode potentials ranging from −0.8 to −1.2 V, employing a zinc nitrate bath, and using voltametric and chronoamperometric techniques. The effects of the electrode nature and of the working electrode potential on the structural, morphological, and optical properties of the ZnO structures were investigated. It was found that all the samples crystallize in hexagonal wurtzite structure with a preferential orientation along the c-axis. Scanning electron microscopy (SEM) images confirm that the presence of a graphene covered electrode led to the formation of ZnO nanowires with a smaller diameter compared with the deposition directly on copper surface. The photoluminescence (PL) measurements revealed that the ZnO nanowires grown on graphene/Cu exhibit stronger emission compared to the nanowires grown on Cu. The obtained results add another possibility of tailoring the properties of such nanostructured films according to the specific functionality required.

1. Introduction

Quasi-one-dimensional ZnO nanostructures, such as nanowires, nanorods, and nanotubes are seen to be interesting subjects of research since these can be used as building blocks for a wide range of applications [1,2,3,4]. ZnO is a semiconducting material (usually n type) with a set of properties which include high transparency, wide direct bandgap (3.37 eV), large exciton binding energy of 60 meV, and an electron mobility which can reach 155 cm2 V−1s−1 [5,6,7]. The material is rather abundant, non-toxic, and radiation resistant, these qualities making it an interesting material for numerous electronic and optoelectronic applications such as sensors [8], transistors [1,9], supercapacitors [10,11], photovoltaic devices [12], and photocatalytic structures [13]. A large number of reports dealing with the synthesis of ZnO nanostructures is available, simple, low-cost preparation methods such as hydrothermal method [14], spin-coating [15], electrochemical deposition [2,16], and chemical deposition [17,18] being preferred. It was found that for this particular oxide the synthesis parameters are directly influencing the micromorphology, ZnO being a polymorphic material. Platelets, rods, tubes, wires, prisms, and tubes are obtained.
Low dimensional carbon structures have opened new opportunities for applications due to their specific properties. In particular, graphene, a monolayer of carbon atoms with a 2D hexagonal crystal lattice has a wide range of excellent properties, such as high electron mobility [19], optical transparency [20], and increased mechanical and chemical stability [21] and is therefore suitable for electronic or optoelectronic applications [22,23,24]. Different techniques have been employed for graphene fabrication including processes such as chemical vapor deposition (CVD) [25], annealing SiC substrates [26], liquid-phase exfoliation (LPE), and mechanical exfoliation [27,28], each process producing different amounts of material with different quality parameters. Among these methods, the growth by CVD employs metallic substrates as catalysts and a hydrocarbon gas as carbon source (CH4). The process is based on carbon surface adsorption on the metal substrate with low carbon solubility (Cu). CVD growth of graphene results in single or few layers of material deposition, usually into a 2-dimensional polycrystalline film. The graphene CVD process on metal substrates has shown good results in producing single-layer graphene on a large surface with high quality in terms of structure and purity [29,30].
In recent years, ZnO nanostructures/CVD-graphene hybrids and nanocomposites have been synthesized to improve the performance of photocatalyst [31] and optoelectronic applications such as supercapacitor, sensor, light-emitting diodes, and photodetectors [11,32,33,34,35,36]. In particular, Brent et al. [37] used a hydrothermal floating technique to deposit ZnO nanorods on CVD-graphene to investigate the effects of the microstructure and interface on the performance of ZnO nanorods/CVD-graphene UV photodetectors. Ju et al. [38] constructed membrane-based flexible photodetectors for imaging applications, based on a structure of Ag nanowires/ZnO nanowires/graphene in which ZnO nanowires were grown by CVD, mixed with Ag nanowires, and used as a film. Ge et al. [39] reported the fabrication of ZnO nanorods on CVD-graphene surface by hydrothermal method for simultaneous electrochemical sensing. Hak et al. [40] demonstrated the synthesis of vertically aligned ZnO NWs on CVD-graphene/Cu by using self-catalyzed vapor-phase transport technique and developed graphene/ZnO NR/graphene heterojunction devices. Usually, the fabrication of this type of composites takes place in two steps: a first step for the growth of graphene on metal substrate and a second step for the deposition of ZnO nanostructures on graphene surface. Various techniques were employed for depositing ZnO on graphene such as hydrothermal technique and self-catalyzed vapor-phase transport. There are few published papers on the growth of ZnO nanowires on graphene-covered substrates by employing electrochemical deposition [41,42,43]. Electrodeposition is an interesting technique for preparing thin semiconductor films since it is not based on expensive equipment and provides good control over parameters such as thickness, morphology, crystallite size, and aspect ratio of the deposit. By tailoring the process parameters, such as bath composition, temperature, deposition time, potential, or current, deposits can be obtained with a wide range of morphologic or structural characteristics [2,44,45]. In this way, nanostructures which include nanowires, nanorods, prisms or platelets can be fabricated in a controlled manner. Using CVD-graphene on Cu substrates can improve the control on the morphology, density, and orientation of the ZnO nanostructures in an electrodeposition process.
In this work, an approach in fabricating tailored ZnO nanowires on Cu and CVD-graphene/Cu films working electrodes is described. A CVD process was employed for fabricating graphene on copper substrates. Different electrode potentials were used during the electrochemical deposition process. A wide range of techniques, including scanning electron microscopy (SEM), X-ray diffraction (XRD), Raman spectroscopy, as well as reflectance (R%) and photoluminescence (PL) spectroscopy, were used to study the effect of the electrode potential and graphene surface on the structural, morphological, and luminescence properties of ZnO/Cu and ZnO/G/Cu films.

2. Materials and Methods

2.1. Synthesis of Graphene, ZnO/Cu, and ZnO/G/Cu Films

Graphene (G) was prepared on polycrystalline copper (Cu) substrates (99.99+% purity from Goodfellow Cambridge Limited, Huntingdon, England) of 25 µm thickness, by employing the chemical vapor deposition method (CVD). An As-One Rapid Thermal Processor system produced by Annealsys (Montpellier, France) was employed. The Cu substrates (0.5 cm × 2 cm) were cleaned by ultrasonication in water with detergent, deionized water, acetone, and isopropanol respectively. A typical procedure was employed, starting by placing the Cu substrates in the deposition chamber and annealing it at 900 °C in a H2 (10 sccm) atmosphere for 15 min at a pressure of 10 mbar. Further, the graphene was grown on the metal surface using a gas reaction mixture of H2:CH4 with a flow ratio of 10:25 being injected in as carbon (C) source. This process was performed for 15 min at a temperature of 900 °C and a pressure of 10 mbar and was followed by one of rapid cooling in Ar atmosphere. The samples were then removed from the deposition chamber and used as working electrodes as prepared.
ZnO electrodeposition was performed using an aqueous solution with a composition of Zn(NO3)2 · 6H2O and 100 mM KNO3. The bath temperature was kept constant using a thermostat and a double wall electrochemical cell during the process. For ZnO nanostructures electrodeposition, a three-electrodes system was used, consisting of a Pt mesh woven as a counter electrode (CE), a potassium chloride saturated Red-Rod electrode as reference (RE), and Cu or G/Cu (0.5 cm × 2 cm) as working electrodes (WE). The electrochemical deposition processes were performed using a Voltalab 10 PGZ 100 potentiostat/galvanostat controlled by computer with Volta Master 4 Electrochemical Software from Radiometer Analytical SAS (Lyon, France). ZnO nanowires growth on Cu and CVD-G/Cu was carried out at different potential values versus the RE at a temperature of 90 °C, these conditions being found to be optimal in order to achieve the highest density of vertically aligned wires. The electrochemical polarization measurements were performed in order to understand the electrodeposition process of ZnO on Cu and G/Cu. The ZnO NWs/Cu and ZnO NWs/G/Cu samples were obtained using potentiostatic deposition, with applied potentials: −0.8 V, −0.9 V, −1 V, −1.1 V, −1.2 V vs. RE without stirring. After the electrodeposition, the as-deposited films were rinsed with distilled water and dried. The preparation procedure is shown in Figure 1.

2.2. Characterization

The structural characterization of the films was performed by X-ray diffraction measurements performed with a Bruker D8 Advance X-ray diffractometer (Bruker AXS GmbH, Karlsruhe, Germany) using CuKα radiation (λ = 1.54178 Å) in the 20–70° range. The surface morphology of ZnO nanostructures/Cu and ZnO/Graphene/Cu was analyzed using a Gemini 500 field emission scanning electron microscope from Zeiss (Oberkochen, Germany). The Raman spectra of the ZnO/Cu and ZnO/G/Cu films were recorded at room temperature using a LabRAM HR Evolution Raman spectrometer (Horiba Jobin-Yvon: Edison, NJ, USA), with a He–Ne laser functioning at 633 nm being focused by an Olympus 100× objective on the surface of the films. The reflectance spectra were recorded using a PerkinElmer Lambda 45 UV–Vis spectrophotometer(Waltham, Massachusetts, USA) and the photoluminescence measurements were performed using an Edinburgh FL 920 PL spectrometer (Edinburgh Instruments, Livingston, UK), with 325 nm UV excitation at room temperature.

3. Results and Discussion

3.1. Electrochemical Study

The electrochemical polarization curves recorded for ZnO deposition on Cu and G/Cu presented in Figure 2a show both similarities and differences between the two types of working electrodes. It is known that the electrochemical deposition of ZnO is a two-step process with a first step being the nitrate to nitrite reduction resulting in local OH ions production, the second step being the precipitation of Zn(OH)2 which further decomposes into ZnO [46,47]. The electrochemical current measurements are therefore not a direct measurement of the deposition process but an indirect one through the rate of nitrate reduction. The polarization curves estimate the evolution of the reduction potential on Cu and graphene/Cu working electrodes, respectively. Figure 2a illustrates the differences between the Cu and graphene/Cu working electrodes in ZnO nanowires electrodeposition. Two observations can be made, the deposition on copper electrode generates a higher current and a wide peak at about −550 mV. This may be related to a pronounced catalytic effect of the copper substrates on the reduction reaction taking place. Furthermore, the conclusion is supported by the fact that the main deposition range starts at a more anodic potential for copper working electrodes, the difference being of about 100 mV.
For the electrodeposition of ZnO nanowires on Cu and graphene/Cu using chronoamperometry, several different potentials from −0.8 V to −1.2 V were selected using a deposition bath containing 5 mM of Zn(NO3)2·6H2O and 100 mM KNO3. Chronoamperometric curves presented in Figure 2b,c for the two working electrodes and for the full range of deposition potentials show a behavior of a nucleation–growth mechanism, a tendency which is more pronounced for the graphene-covered electrode. As was observed already from the electrochemical polarization curves, the current density is higher for the case of simple copper electrodes—i.e., a higher nitrate reduction rate at similar electrode potential. A current decrease can be observed after an initial sharp peak in the first seconds due to the formation of the double layer and due to the nucleation process [48]. For the graphene-covered electrode the process is slower, a slight decrease in current being observed throughout the experiment.
The near constant current region corresponds to a stable deposition regime, the deposition rate depending on bath composition and on the electrode potential. When increasing the applied potential, an increase in current density is observed.

3.2. Structural Properties

The X-ray diffraction patterns of ZnO/Cu and ZnO/G/Cu films deposited at applied potentials of −0.8 V, −0.9 V, −1 V, and −1.1 V are shown in Figure 3a,b. The characteristic peaks at 31.89°, 34.56°, 36.4°, 47.57°, 56.6°, and 62.9°, correspond to the (100), (002), (101), (102), (110), and (103), crystal planes of hexagonal wurtzite structure of ZnO (JCPDS NO.36-1451). These planes were observed in all samples with different relative intensities.
In all cases, two peaks at 43° and 51° were observed, which correspond to the (111) and (200) planes of the Cu substrate [29]. Furthermore, the intensities of (002) peaks for all samples were the strongest (Figure 4a), different from the powder standard, which indicates that the ZnO nanowires have a preferred orientation along the c-axis direction, usual for this particular morphology and for these experimental parameters. However, it is easy to notice that the ratio between the (002) peak and the other peaks is higher for the graphene-covered electrodes. For the samples obtained at −0.9 V deposition potential on graphene covered electrode, other peaks are almost invisible showing an almost perfect orientation of the nanowires. Corroborating this observation to the electrochemical data, one can conclude that on graphene, the nucleation process is slower but also the nuclei are preferentially oriented along the (002) direction.
Based on the full-width-at-half-maximum (FWHM) data for the (101), (002), and (100) diffraction peaks (Table 1), the average crystallite size of the ZnO nanowires was calculated using the Debye–Scherrer formula, as given in Equation (1) [49]:
D   = k   λ β cos θ    
where, k = 0.89 is Scherer’s factor, λ is the X-ray wavelength (λ = 1.54178 Å, CuKα radiation), θ is the Bragg angle, and β the line broadening at half height of the maximum intensity (FWHM), measured in radians. The mean crystallite size of the ZnO nanowires corresponding to the FWHM of the most intense peak at 34.56° showed that ZnO/G/Cu films (66.19 nm, 58.35 nm, 48.57 nm, 49.76 nm) was smaller than synthesized on Cu (68.9 nm, 65.65 nm, 56.14 nm, 57.13 nm) (Figure 4b). Additionally, the FWHM value of the highest intense peak (002) of ZnO grown on G/Cu was a little smaller than ZnO/Cu, which suggests that the layer of graphene impacted the growth of ZnO nanowires most probably by decreasing the diameter. It is also important to point out that, as the deposition potential is more electronegative, the crystallite size decreases as a consequence of the induced change in deposition mechanisms from a reaction rate limitation to a diffusion rate limitation.
The preferential orientation was estimated from the XRD measurements by calculating the texture coefficient for a (hkl) plan (TC (hkl)) by using the Equation (2) [50]:
TC ( h k l ) = I ( h k l ) / I 0 ( h k l ) 1 N 1 N I ( h k l ) / I 0 ( h k l )
I(hkl) is the measured intensity, I0(hkl) is the standard intensity from the (JCPDS NO.36-1451) and n is the number of reflections; three diffraction (N = 3) planes being chosen (100), (002), (101). Figure 5a,b shows the texture coefficient (TC) of (100), (002), and (101) planes for ZnO/Cu and ZnO/G/Cu at a different applied potential. The high TC of (002) values in all samples confirms that the ZnO nanowires exhibit a preferential orientation growth along the c-axis. Moreover, the (002) texture is higher in the graphene covered electrode samples than in the samples grown on simple copper electrodes.

3.3. Morphology Analysis

Figure 6 and Figure 7 show SEM images of ZnO nanowires grown by electrochemical deposition on Cu and CVD–graphene/Cu electrodes at different electrode potential values. According to SEM images of ZnO nanowires/Cu (Figure 6a–e), it can be observed that ZnO nanowires are dense and homogeneous, completely coating the Cu electrode. In both cases, one can notice a well-crystalized material. In addition, the SEM images of ZnO/Cu at −0.8 V and −0.9 V (Figure 6a,b) show the presence of nanowires with hexagonal and planar face at their top. The analysis of SEM images at −1 V, −1.1 V, and −1.2 V reveals the presence of nanowires with a pyramid face at their top. This is the result of a change induced by the difference between a reaction rate-limited process at less electronegative potential and a diffusion rate-limited process at more electronegative potential.
In the case of CVD-graphene/Cu electrodes, (Figure 7a) the ZnO nanowires obtained follow the same tendencies in terms of diameter/crystallite size. It is interesting to note that when the deposition takes place in the range of potentials of −0.8 V to −1.1 V (Figure 7a–d), a high density of wires can be observed, this decreasing in the case of an electrode potential of −1.2 V (Figure 7e). This behavior is again consistent with a nucleation growth model and a diffusion rate limited process. The cross-section SEM images (e.g., Figure 7f) confirm the formation of ZnO nanowires/graphene/Cu film at −0.9 V. In this particular case, the length of the ZnO nanowires is 1.5–2 μm.
The average diameter distribution was determined by measuring ZnO nanowires from SEM images using Image J software (Figure 8). This figure shows nanowire diameter with a size distribution centered between 20 and 300 nm. Moreover, the average size in diameter of ZnO nanowires grown on G/Cu is lower compared to those of ZnO grown on Cu substrates (Figure 8k). Furthermore, as the potential becomes more electronegative the average diameter becomes smaller in both cases, copper and graphene covered electrodes. In addition, according to the obtained XRD results, the nanowires grown on the graphene surface are more vertically aligned than ZnO/Cu (see Figure 5). As a result, the graphene surface plays an important role in the growth, orientation, density, and diameters of electrodeposited ZnO nanostructures.

3.4. Raman Study

The uniformity and structural quality of the electrodeposited nanostructures were also analyzed from the Raman spectroscopy measurements at room temperature. Figure 9 shows the Raman spectra of the ZnO/G/Cu films deposited at different potentials. For all Raman spectra of ZnO/G/Cu, three peaks are observed at 436 cm−1, 1308 cm−1, and 1600 cm−1. The first peak at 436 cm−1 ( E 2 H i g h mode) is associated to oxygen vibration of ZnO [50]. The second peak at 1308 cm−1 corresponds to the D band of graphene attributed to layer disorder due to the level of defects. The third peak at 1600 cm−1 corresponds to the graphene G band originating from in-plane stretching vibration of sp2 carbon and is a double degenerate phonon mode (E2g symmetry) [51,52]. The spectra of ZnO/G/Cu deposited at −1.1 V, includes a new peak at 2656 cm−1 assigned to 2D band of graphene. According to the literature, the 2D band is related to the number of graphene sheets being visible only in this case due to the lowest nanowire density.
Comparing the Raman spectra between the samples grown on the two types of electrodes (Figure 10a,b), the growth of ZnO on graphene shifts the position of the peak E 2 H i g h to higher frequencies (from 435 to 438 cm−1) and modifies its intensity.
Furthermore, the full width at half maximum (FWHM) of the peak at 438 cm−1 ( E 2 H i g h mode) decreases in the ZnO/Cu when compared to ZnO/G/Cu samples most probably due to the covalent interaction of the ZnO with the graphene layers (Figure 10a,b). According to data analysis from literature, Biroju et al. [53] noted that lower E 2 h i g h in ZnO wires, as is the case for the ZnO deposited on graphene covered electrodes, indicate a better crystalline structure.
The crystallite size of graphene used for ZnO electrodeposition can be obtained from Raman spectroscopy measurements. Thus, the ID/IG ratio obtained from the Raman spectra on ZnO/G/Cu samples was used to calculate the average crystallite size (La) of graphene layer (Table 2), using the Equation (3) [54]:
L a ( nm ) = ( 2.4 × 10 10 ) λ 4 ( I D I G ) 1 ,
where λ is the laser excitation (633 nm).

3.5. Optical Properties

The UV–Vis total reflectance spectra of ZnO/Cu and ZnO/G/Cu are shown in Figure 11, a blue shift being observed in the band for the graphene-covered electrodes.
The inset of Figure 11 shows the estimation of the band gap of samples by calculating the slopes of reflectance (R) spectra based on Kubelka–Munk function F(R), using Equation (4):
F ( R ) = ( 1 R ) 2 2 R
The band gap energy of ZnO/Cu was found to be 3.15 eV, which is slightly less than the band gap of ZnO/G/Cu (3.21 eV) (inset Figure 11). This can be induced by the interaction between ZnO nanowires and graphene layer and to the concentration of point defects in the material. Nanakkalt et al. [55] reported that such a blue shift in the absorption band and in the energy band gap can be associated to the differences in concentration of point defects induced by the presence of graphene. In any case, these two films have a significantly lower band gap than previously reported for ZnO (3.37 eV) [7].

3.6. Photoluminescence

In Figure 12, the photoluminescence (PL) spectra of ZnO NWs/Cu and ZnO/G/Cu obtained using a 325 nm excitation are presented. The typical ZnO PL spectra exhibit two distinct types of bands [44,56]. The near band to band emission (NBE) is located around ~376–378 nm (UV emission) and is attributed to free exciton recombination while the other bands are related to various point defects. In the present case, the broad visible emission band in the range of 580–602 nm can be assigned to oxygen-related defects. The PL spectra plotted show that in all the cases, the wires grown on graphene present a more intense excitonic peak when compared to ZnO nanostructures grown directly on copper at the same electrode potential. The observation is related to a lower concentration of defects and can be correlated to the results obtained from the diffraction measurements. Defect concentration is also related to the deposition potential, the luminescence of the exciton peak being more intense for the samples grown at less electronegative potential. The results are consistent with the fact that for a lower growth rate, there is a lower concentration of defects in the material. For the case of the samples grown at the most electronegative potential, the excitonic peak almost disappears, being dwarfed by the defect-related bands. As expected, the only case where the excitonic peak is more intense than the defect-related band is at the lowest deposition rate, i.e., at the least electronegative potential (−0.8 V). On the other hand, the PL intensity ratio of the UV emission to the visible emission (Iuv/Ivis) (Tabel 3), meaning the ratio between the intensity of the exciton luminescence and the defect related luminescence, was found for all samples below one except for the structures grown on graphene at −800 mV. This particular case is related to a high-quality material with strong near band-to-band emission and lower concentration of defects. As a comparison, the film of ZnO/G/Cu grown at applied potential of −0.9 V shows the lowest Iuv/Ivis value compared to the other films, which suggests that there may be more defects induced during the growth such as oxygen vacancies (VO) [57]. In terms of potential applications, Ahn et al. [58] noted that the sensitivity of a gas sensor increases linearly with the concentration of defects in the analyzed samples.
In Figure 13a,b, the Gaussian fitting for the PL spectra of ZnO NWs grown on Cu and ZnO NWs grown on graphene/Cu are presented. Black lines correspond to the experimental data and color lines represent the fitted data. The broad visible emission bands were deconvoluted into three PL peaks. The green emission peak located around 531–535 nm is attributed to the ionized oxygen interstitial and/or complex of VOZni. The green-yellow emission peak around 588–601 nm corresponding to the defects induced by oxygen vacancies (VO) situated in the ZnO lattice is the luminescence produced by the recombination of photogenerated holes with electrons of the VO [59]. The yellow-orange emission peak at 635–639 nm is usually attributed to the presence of O interstitials [60,61]. Due to the relative intensities of the three peaks, the center of the visible PL bands emission shows a blue shift for ZnO/G/Cu with respect to the ZnO/Cu, a possible cause for this behavior being the interaction between ZnO nanowires and the graphene.
The emissions spectra of all samples were also characterized using the CIE 1931 chromaticity diagram (Figure 14). The color coordinates of ZnO/Cu were estimated and summarized in Table 3, corresponding to yellow-orange emission. The obtained color coordinates for ZnO/G/Cu display a shift from yellowish-orange to greenish-yellow compared with ZnO/Cu.

4. Conclusions

Electrochemical deposition was employed to deposit ZnO nanostructures on copper and graphene/copper working electrodes. The graphene-covered electrodes were obtained by chemical vapor deposition.
Electrochemical deposition of ZnO is a two-step process, i.e., reduction of nitrate ions to nitrite and precipitation of ZnO. The electrochemical measurements show a higher current density in the case of the copper electrodes while the deposition on graphene takes place at more electronegative potentials. The process is one of nucleation/growth, with a higher nucleation rate in the case of copper substrates. The morphological, structural, and optical measurements revealed the fact that the higher structural quality/lower defect concentration can be observed for the samples deposited on graphene, with a higher texture coefficient and intense luminescence given by the exciton recombination.
The sample of ZnO/G/Cu prepared at −0.8 V with the ZnO nanowires diameter of 50–100 nm shows a stronger UV emission at 378 nm as compared to that of ZnO/Cu obtained at the same deposition potential. ZnO/graphene/Cu nanohybrid heterostructures are interesting for a wide range of optoelectronic applications and various sensors.
The use of G/Cu as a substrate is a straightforward preparation procedure, removing the need for a graphene transfer procedure.

Author Contributions

Conceptualization, I.B., E.M. and I.E.; methodology, I.B., E.M. and I.E.; validation, E.M., A.J. and I.E.; formal analysis, I.B., E.M. and I.E.; investigation, I.B., E.M., M.E. and I.E.; resources, E.M. and I.E.; data curation, I.B. and E.M.; writing—original draft preparation, I.B. and E.M.; writing—review and editing, E.M. and I.E.; visualization, E.M. and I.E.; supervision, E.M. and I.E.; project administration, E.M.; funding acquisition, E.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Romanian Ministry of Foreign Affairs and Agence Universitaire de la Francophonie for the Eugen Ionescu research and mobility grants and Romanian Ministry of Research, Innovation, and Digitization, in the frame of Core Program PN19-03 (contract No. 21N/08.02.2019) and PN-III-P4-ID-PCE-2020-2291 (Contract no. PCE 45/02/2021). Fee for open access publication was supported from the project 35PFE/2021 funded by the Romanian Ministry of Research, Innovation and Digitization.

Acknowledgments

I.B. acknowledges the Romanian Ministry of Foreign Affairs and the Agence Universitaire de la Francophonie for the Eugen Ionescu research and mobility grant at the National Institute of Materials Physics. E.M. acknowledges the Romanian Ministry of Research, Innovation and Digitization in the frame of Core Program PN19-03 (contract No. 21N/08.02.2019), PN-III-P4-ID-PCE-2020-2291 (Contract no. PCE 45/02/2021) and project 35PFE/2021 for the financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Florica, C.; Costas, A.; Kuncser, A.; Preda, N.; Enculescu, I. High Performance FETs Based on ZnO Nanowires Synthesized by Low Cost Methods. Nanotechnology 2016, 27, 475303. [Google Scholar] [CrossRef] [PubMed]
  2. Matei, E.; Enculescu, M.; Preda, N.; Enculescu, I. ZnO Morphological, Structural and Optical Properties Control by Electrodeposition Potential Sweep Rate. Mater. Chem. Phys. 2012, 134, 988–993. [Google Scholar] [CrossRef]
  3. Badre, C.; Pauporté, T. Nanostructured ZnO-Based Surface with Reversible Electrochemically Adjustable Wettability. Adv. Mater. 2009, 21, 697–701. [Google Scholar] [CrossRef]
  4. Chang, S.P.; Chang, S.J.; Lu, C.Y.; Li, M.J.; Hsu, C.L.; Chiou, Y.Z.; Hsueh, T.J.; Chen, I.C. A ZnO Nanowire-Based Humidity Sensor. Superlattices Microstruct. 2010, 47, 772–778. [Google Scholar] [CrossRef]
  5. Galdámez-Martinez, A.; Santana, G.; Güell, F.; Martínez-Alanis, P.R.; Dutt, A. Photoluminescence of ZnO Nanowires: A Review. Nanomaterials 2020, 10, 857. [Google Scholar] [CrossRef] [PubMed]
  6. Djurišić, A.B.; Leung, Y.H. Optical Properties of ZnO Nanostructures. Small 2006, 2, 944–961. [Google Scholar] [CrossRef]
  7. Janotti, A.; Van de Walle, C.G. Fundamentals of zinc oxide as a semiconductor. Rep. Prog. Phys. 2009, 72, 126501. [Google Scholar] [CrossRef]
  8. Shetti, N.P.; Bukkitgar, S.D.; Reddy, K.R.; Reddy, C.V.; Aminabhavi, T.M. ZnO-based nanostructured electrodes for electrochemical sensors and biosensors in biomedical applications. Biosens. Bioelectron. 2019, 141, 111417. [Google Scholar] [CrossRef]
  9. Liu, X.; Lin, P.; Yan, X.; Kang, Z.; Zhao, Y.; Lei, Y.; Li, C.; Du, H.; Zhang, Y. Enzyme-Coated Single ZnO Nanowire FET Biosensor for Detection of Uric Acid. Sens. Actuators B Chem. 2013, 176, 22–27. [Google Scholar] [CrossRef]
  10. Ahmed, F.; Almutairi, G.; AlOtaibi, B.; Kumar, S.; Arshi, N.; Hussain, S.G.; Umar, A.; Ahmad, N.; Aljaafari, A. Binder-Free Electrode Based on ZnO Nanorods Directly Grown on Aluminum Substrate for High Performance Supercapacitors. Nanomaterials 2020, 10, 1979. [Google Scholar] [CrossRef]
  11. Rafique, A.; Bianco, S.; Fontana, M.; Pirri, C.F.; Lamberti, A. Flexible Wire-Based Electrodes Exploiting Carbon/ZnO Nanocomposite for Wearable Supercapacitors. Ionics 2017, 23, 1839–1847. [Google Scholar] [CrossRef]
  12. Li, D.; Zhao, L.; Wu, R.; Ronning, C.; Lu, J.G. Temperature-Dependent Photoconductance of Heavily Doped ZnO Nanowires. Nano Res. 2011, 4, 1110–1116. [Google Scholar] [CrossRef]
  13. Smazna, D.; Shree, S.; Polonskyi, O.; Lamaka, S.; Baum, M.; Zheludkevich, M.; Faupel, F.; Adelung, R.; Mishra, Y.K. Mutual Interplay of ZnO Micro-and Nanowires and Methylene Blue during Cyclic Photocatalysis Process. J. Environ. Chem. Eng. 2019, 7, 103016. [Google Scholar] [CrossRef]
  14. Opoku, C.; Dahiya, A.S.; Poulin-Vittrant, G.; Camara, N.; Alquier, D. Source-Gating Effect in Hydrothermally Grown ZnO Nanowire Transistors. Phys. Status Solidi (A) 2016, 213, 2438–2445. [Google Scholar] [CrossRef]
  15. Poongodi, G.; Anandan, P.; Kumar, R.M.; Jayavel, R. Studies on Visible Light Photocatalytic and Antibacterial Activities of Nanostructured Cobalt Doped ZnO Thin Films Prepared by Sol–Gel Spin Coating Method. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 148, 237–243. [Google Scholar] [CrossRef] [PubMed]
  16. Florica, C.; Matei, E.; Costas, A.; Molares, M.E.T.; Enculescu, I. Field Effect Transistor with Electrodeposited ZnO Nanowire Channel. Electrochim. Acta 2014, 137, 290–297. [Google Scholar] [CrossRef]
  17. Florica, C.; Preda, N.; Enculescu, M.; Zgura, I.; Socol, M.; Enculescu, I. Superhydrophobic ZnO Networks with High Water Adhesion. Nanoscale Res. Lett. 2014, 9, 385. [Google Scholar] [CrossRef]
  18. Preda, N.; Enculescu, M.; Zgura, I.; Socol, M.; Matei, E.; Vasilache, V.; Enculescu, I. Superhydrophobic Properties of Cotton Fabrics Functionalized with ZnO by Electroless Deposition. Mater. Chem. Phys. 2013, 138, 253–261. [Google Scholar] [CrossRef]
  19. Torrisi, F.; Hasan, T.; Wu, W.; Sun, Z.; Lombardo, A.; Kulmala, T.S.; Hsieh, G.-W.; Jung, S.; Bonaccorso, F.; Paul, P.J. Inkjet-Printed Graphene Electronics. ACS Nano 2012, 6, 2992–3006. [Google Scholar] [CrossRef]
  20. Hwang, T.; Kwon, H.-Y.; Oh, J.-S.; Hong, J.-P.; Hong, S.-C.; Lee, Y.; Ryeol Choi, H.; Jin Kim, K.; Hossain Bhuiya, M.; Nam, J.-D. Transparent Actuator Made with Few Layer Graphene Electrode and Dielectric Elastomer, for Variable Focus Lens. Appl. Phys. Lett. 2013, 103, 023106. [Google Scholar] [CrossRef]
  21. Panahi-Sarmad, M.; Razzaghi-Kashani, M. Actuation Behavior of PDMS Dielectric Elastomer Composites Containing Optimized Graphene Oxide. Smart Mater. Struct. 2018, 27, 085021. [Google Scholar] [CrossRef]
  22. Beams, R.; Bharadwaj, P.; Novotny, L. Electroluminescence from Graphene Excited by Electron Tunneling. Nanotechnology 2014, 25, 055206. [Google Scholar] [CrossRef] [PubMed]
  23. Shao, Y.; Wang, J.; Wu, H.; Liu, J.; Aksay, I.A.; Lin, Y. Graphene Based Electrochemical Sensors and Biosensors: A Review. Electroanalysis 2010, 22, 1027–1036. [Google Scholar] [CrossRef]
  24. Park, H.; Chang, S.; Jean, J.; Cheng, J.J.; Araujo, P.T.; Wang, M.; Bawendi, M.G.; Dresselhaus, M.S.; Bulović, V.; Kong, J. Graphene Cathode-Based ZnO Nanowire Hybrid Solar Cells. Nano Lett. 2013, 13, 233–239. [Google Scholar] [CrossRef] [PubMed]
  25. Zhang, Y.I.; Zhang, L.; Zhou, C. Review of Chemical Vapor Deposition of Graphene and Related Applications. Acc. Chem. Res. 2013, 46, 2329–2339. [Google Scholar] [CrossRef]
  26. Emtsev, K.V.; Bostwick, A.; Horn, K.; Jobst, J.; Kellogg, G.L.; Ley, L.; McChesney, J.L.; Ohta, T.; Reshanov, S.A.; Röhrl, J. Towards Wafer-Size Graphene Layers by Atmospheric Pressure Graphitization of Silicon Carbide. Nat. Mater. 2009, 8, 203–207. [Google Scholar] [CrossRef]
  27. Paton, K.R.; Varrla, E.; Backes, C.; Smith, R.J.; Khan, U.; O’Neill, A.; Boland, C.; Lotya, M.; Istrate, O.M.; King, P. Scalable Production of Large Quantities of Defect-Free Few-Layer Graphene by Shear Exfoliation in Liquids. Nat. Mater. 2014, 13, 624–630. [Google Scholar] [CrossRef]
  28. Hernandez, Y.; Nicolosi, V.; Lotya, M.; Blighe, F.M.; Sun, Z.; De, S.; McGovern, I.T.; Holland, B.; Byrne, M.; Gun’Ko, Y.K. High-Yield Production of Graphene by Liquid-Phase Exfoliation of Graphite. Nat. Nanotechnol. 2008, 3, 563–568. [Google Scholar] [CrossRef]
  29. Reina, A.; Jia, X.; Ho, J.; Nezich, D.; Son, H.; Bulovic, V.; Dresselhaus, M.S.; Kong, J. Large Area, Few-Layer Graphene Films on Arbitrary Substrates by Chemical Vapor Deposition. Nano Lett. 2009, 9, 30–35. [Google Scholar] [CrossRef]
  30. Suk, J.W.; Kitt, A.; Magnuson, C.W.; Hao, Y.; Ahmed, S.; An, J.; Swan, A.K.; Goldberg, B.B.; Ruoff, R.S. Transfer of CVD-Grown Monolayer Graphene onto Arbitrary Substrates. ACS Nano 2011, 5, 6916–6924. [Google Scholar] [CrossRef]
  31. Yu, M.; Ma, Y.; Liu, J.; Li, X.; Li, S.; Liu, S. Sub-Coherent Growth of ZnO Nanorod Arrays on Three-Dimensional Graphene Framework as One-Bulk High-Performance Photocatalyst. Appl. Surf. Sci. 2016, 390, 266–272. [Google Scholar] [CrossRef]
  32. Dong, X.; Cao, Y.; Wang, J.; Chan-Park, M.B.; Wang, L.; Huang, W.; Chen, P. Hybrid structure of zinc oxide nanorods and three dimensional graphene foam for supercapacitor and electrochemical sensor applications. RSC Adv. 2012, 2, 4364–4369. [Google Scholar] [CrossRef]
  33. Amiri, M.H.; Namdar, N.; Mashayekhi, A.; Ghasemi, F.; Sanaee, Z.; Mohajerzadeh, S. Flexible Micro Supercapacitors Based on Laser-Scribed Graphene/ZnO Nanocomposite. J. Nanoparticle Res. 2016, 18, 237. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Li, X.; Wang, L.; Yi, X.; Wu, D.; Zhu, H.; Wang, G. Enhanced light emission of GaN-based diodes with a NiOx/graphene hybrid electrode. Nanoscale 2012, 4, 5852–5855. [Google Scholar] [CrossRef]
  35. Zhang, Y.; Li, X.; Wang, L.; Li, X.; Yi, X.; Zhang, N.; Li, J.; Zhu, H.; Wang, G. Annealed InGaN green light-emitting diodes with graphene transparent conductive electrodes. J. Appl. Phys. 2012, 111, 114501. [Google Scholar] [CrossRef]
  36. Wu, G.; Tang, L.; Deng, G.; Liu, L.; Hao, Q.; Yuan, S.; Wang, J.; Wei, H.; Zhao, Y.; Yue, B.; et al. Transparent dual-band ultraviolet photodetector based on graphene/p-GaN/AlGaN heterojunction. Opt. Express 2022, 30, 21349–21361. [Google Scholar] [CrossRef]
  37. Brent, C.; Qingfeng, L.; Jianwei, L.; Maogang, G.; Dan, E.; Matthew, C.; Alex, S.; Judy, W. Facile zinc oxide nanowire growth on graphene via a hydrothermal floating method: Towards Debye length radius nanowires for ultraviolet photodetection. J. Mater. Chem. C 2017, 5, 10087–10093. [Google Scholar]
  38. Ju, D.; Liu, X.; Zheng feng, Z.; Wang, S.; Liu, S.; Gu, Y.; Chang, J.; Qingquan, L.; Zou, Y. Solution processed membrane-based wearable ZnO/graphene Schottky UV photodetectors with imaging application. Nanotechnology 2019, 30, 375701. [Google Scholar] [CrossRef]
  39. Ge, C.; Hongji, L.; Mingji, L.; Cuiping, L.; Xiaoguo, W.; Baohe, Y. Synthesis of a ZnO nanorod/CVD graphene composite for simultaneous sensing of dihydroxybenzene isomers. Carbon 2015, 95, 1–9. [Google Scholar] [CrossRef]
  40. Cho, H.D.; Kim, D.Y.; Lee, J.-K. ZnO Nanorod/Graphene Hybrid-Structures Formed on Cu Sheet by Self-Catalyzed Vapor-Phase Transport Synthesis. Nanomaterials 2021, 11, 450. [Google Scholar] [CrossRef]
  41. Cembrero, J.; Pruna, A.; Pullini, D.; Busquets-Mataix, D. Effect of Combined Chemical and Electrochemical Reduction of Graphene Oxide on Morphology and Structure of Electrodeposited ZnO. Ceram. Int. 2014, 40, 10351–10357. [Google Scholar] [CrossRef]
  42. George, L.; Thomas, S.; Nalini, S.; Jayaraj, M.K. Electrodeposition of ZnO Nanostructures on Graphene for Optoelectronic Applications. In Proceedings of the AIP Conference Proceedings. AIP Conf. Proc. 2019, 2162, 020014. [Google Scholar]
  43. Yin, Z.; Wu, S.; Zhou, X.; Huang, X.; Zhang, Q.; Boey, F.; Zhang, H. Electrochemical Deposition of ZnO Nanorods on Transparent Reduced Graphene Oxide Electrodes for Hybrid Solar Cells. Small 2010, 6, 307–312. [Google Scholar] [CrossRef]
  44. Matei, E.; Enculescu, I. Electrodeposited ZnO Films with High UV Emission Properties. Mater. Res. Bull. 2011, 46, 2147–2154. [Google Scholar] [CrossRef]
  45. Thomas, M.A.; Sun, W.W.; Cui, J.B. Mechanism of Ag Doping in ZnO Nanowires by Electrodeposition: Experimental and Theoretical Insights. J. Phys. Chem. C 2012, 116, 6383–6391. [Google Scholar] [CrossRef]
  46. Izaki, M.; Omi, T. Transparent Zinc Oxide Films Prepared by Electrochemical Reaction. Appl. Phys. Lett. 1996, 68, 2439–2440. [Google Scholar] [CrossRef]
  47. Peulon, S.; Lincot, D. Cathodic Electrodeposition from Aqueous Solution of Dense or Open-Structured Zinc Oxide Films. Adv. Mater. 1996, 8, 166–170. [Google Scholar] [CrossRef]
  48. Matei, E.; Costas, A.; Florica, C.; Enculescu, M.; Pintilie, I.; Pintilie, L.; Enculescu, I. Electrical Properties of Templateless Electrodeposited ZnO Nanowires. Mater. Sci. Semicond. Process. 2016, 42, 364–372. [Google Scholar] [CrossRef]
  49. Kumar, V.; Som, S.; Yousif, A.; Singh, N.; Ntwaeaborwa, O.M.; Kapoor, A.; Swart, H.C. Effect of annealing on the structural, morphological and photoluminescence properties of ZnO thin films prepared by spin coating. J. Colloid Interface Sci. 2014, 428, 8–15. [Google Scholar] [CrossRef]
  50. Xu, C.X.; Sun, X.W.; Zhang, X.H.; Ke, L.; Chua, S.J. Photoluminescent Properties of Copper-Doped Zinc Oxide Nanowires. Nanotechnology 2004, 15, 856. [Google Scholar] [CrossRef]
  51. Park, S.H.; Bak, S.M.; Kim, K.H.; Jegal, J.P.; Lee, S.I.; Lee, J.; Kim, K.B. Solid-State Microwave Irradiation Synthesis of High-Quality Graphene Nanosheets under Hydrogen Containing Atmosphere. J. Mater. Chem. 2011, 21, 680–686. [Google Scholar] [CrossRef]
  52. Efthimiopoulos, I.; Mayanna, S.; Stavrou, E.; Torode, A.; Wang, Y. Extracting the Anharmonic Properties of the G-Band in Graphene Nanoplatelets. J. Phys. Chem. C 2020, 124, 4835–4842. [Google Scholar] [CrossRef]
  53. Biroju, R.K.; Tilak, N.; Rajender, G.; Dhara, S.; Giri, P.K. Catalyst Free Growth of ZnO Nanowires on Graphene and Graphene Oxide and Its Enhanced Photoluminescence and Photoresponse. Nanotechnology 2015, 26, 145601. [Google Scholar] [CrossRef] [PubMed]
  54. Cancado, L.G.; Takai, K.; Enoki, T.; Endo, M.; Kim, Y.A.; Mizusaki, H.; Jorio, A.; Coelho, L.N.; Magalhães-Paniago, R.; Pimenta, M.A. General equation for the determination of the crystallite size La of nanographite by Raman spectroscopy. Appl. Phys. Lett. 2006, 88, 163106. [Google Scholar] [CrossRef]
  55. Nanakkal, A.R.; Alexander, L.K. Photocatalytic Activity of Graphene/ZnO Nanocomposite Fabricated by Two-step Electrochemical Route. J. Chem. Sci. 2017, 129, 95–102. [Google Scholar] [CrossRef]
  56. Hsu, J.W.P.; Tallant, D.R.; Simpson, R.L.; Missert, N.A.; Copeland, R.G. Luminescent Properties of Solution-Grown ZnO Nanorods. Appl. Phys. Lett. 2006, 88, 252103. [Google Scholar] [CrossRef]
  57. Mahmood, K.; Park, S.B.; Sung, H.J. Enhanced photoluminescence, Raman spectra and field-emission behavior of indium-doped ZnO nanostructures. J. Mater. Chem. C 2013, 1, 3138–3149. [Google Scholar] [CrossRef]
  58. Ahn, M.W.; Park, K.S.; Heo, J.H.; Park, J.G.; Kim, D.W.; Choi, K.J.; Lee, J.H.; Hong, S.H. Gas sensing properties of defect-controlled ZnO-nanowire gas sensor. Appl. Phys. Lett. 2008, 93, 263103. [Google Scholar] [CrossRef]
  59. Boukhoubza, I.; Khenfouch, M.; Achehboune, M.; Mothudi, B.M.; Zorkani, I.; Jorio, A. Graphene Oxide/ZnO Nanorods/Graphene Oxide Sandwich Structure: The Origins and Mechanisms of Photoluminescence. J. Alloys Compd. 2019, 797, 1320–1326. [Google Scholar] [CrossRef]
  60. Zhao, Q.X.; Klason, P.; Willander, M.; Zhong, H.M.; Lu, W.; Yang, J.H. Deep-Level Emissions Influenced by O and Zn Implantations in ZnO. Appl. Phys. Lett. 2005, 87, 211912. [Google Scholar] [CrossRef]
  61. Djurišić, A.B.; Leung, Y.H.; Tam, K.H.; Ding, L.; Ge, W.K.; Chen, H.Y.; Gwo, S. Green, Yellow, and Orange Defect Emission from ZnO Nanostructures: Influence of Excitation Wavelength. Appl. Phys. Lett. 2006, 88, 103107. [Google Scholar] [CrossRef]
Figure 1. Synthesis process of CVD-graphene, ZnO/Cu, and ZnO/graphene/Cu at different applied potentials.
Figure 1. Synthesis process of CVD-graphene, ZnO/Cu, and ZnO/graphene/Cu at different applied potentials.
Nanomaterials 12 02858 g001
Figure 2. (a) Electrochemical polarization curve for ZnO electrodeposition on Cu and G/Cu. Current versus time curves at different deposition potentials for (b) ZnO/Cu and (c) ZnO/G/Cu.
Figure 2. (a) Electrochemical polarization curve for ZnO electrodeposition on Cu and G/Cu. Current versus time curves at different deposition potentials for (b) ZnO/Cu and (c) ZnO/G/Cu.
Nanomaterials 12 02858 g002
Figure 3. X-ray diffraction patterns of (a) ZnO/Cu and (b) ZnO/G/Cu electrodeposited at different applied potentials.
Figure 3. X-ray diffraction patterns of (a) ZnO/Cu and (b) ZnO/G/Cu electrodeposited at different applied potentials.
Nanomaterials 12 02858 g003
Figure 4. (a) The intensity of the (101), (002), and (100) peaks and (b) the average crystallite size of ZnO/Cu and ZnO/G/Cu prepared at different applied potentials.
Figure 4. (a) The intensity of the (101), (002), and (100) peaks and (b) the average crystallite size of ZnO/Cu and ZnO/G/Cu prepared at different applied potentials.
Nanomaterials 12 02858 g004
Figure 5. Texture coefficient (TC) of (a) ZnO/Cu and (b) ZnO/G/Cu at different applied potentials.
Figure 5. Texture coefficient (TC) of (a) ZnO/Cu and (b) ZnO/G/Cu at different applied potentials.
Nanomaterials 12 02858 g005
Figure 6. SEM images of ZnO nanowires grown on Cu substrate at different potentials: (a) −0.8 V, (b) −0.9V, (c) −1V, (d) −1.1V, (e) −1.2 V vs. RE. The insets show the SEM images at high magnification.
Figure 6. SEM images of ZnO nanowires grown on Cu substrate at different potentials: (a) −0.8 V, (b) −0.9V, (c) −1V, (d) −1.1V, (e) −1.2 V vs. RE. The insets show the SEM images at high magnification.
Nanomaterials 12 02858 g006
Figure 7. SEM images of ZnO nanowires grown on graphene/Cu at different potentials (a) −0.8 V, (b) −0.9 V, (c) −1 V, (d) −1.1 V, (e) −1.2 V vs. RE. (f) Cross-section of ZnO/G/Cu obtained at −0.9 V vs. RE. The insets show the SEM images at higher magnification.
Figure 7. SEM images of ZnO nanowires grown on graphene/Cu at different potentials (a) −0.8 V, (b) −0.9 V, (c) −1 V, (d) −1.1 V, (e) −1.2 V vs. RE. (f) Cross-section of ZnO/G/Cu obtained at −0.9 V vs. RE. The insets show the SEM images at higher magnification.
Nanomaterials 12 02858 g007
Figure 8. Diameter size distribution of ZnO nanowires deposited on Cu (ae) and G/Cu (fj) at different potentials, (k) average diameter size of ZnO nanowires as function of deposition potential for both types of electrodes.
Figure 8. Diameter size distribution of ZnO nanowires deposited on Cu (ae) and G/Cu (fj) at different potentials, (k) average diameter size of ZnO nanowires as function of deposition potential for both types of electrodes.
Nanomaterials 12 02858 g008
Figure 9. Raman spectra and images of the ZnO/G/Cu films deposited at −0.8 V, −0.9 V, −1 V, and −1.1 V vs. RE.
Figure 9. Raman spectra and images of the ZnO/G/Cu films deposited at −0.8 V, −0.9 V, −1 V, and −1.1 V vs. RE.
Nanomaterials 12 02858 g009
Figure 10. Raman spectra showing enlarged area of E 2 H i g h (438 cm−1) mode positions and FWHM of (a) ZnO/Cu and (b) ZnO/G/Cu films deposited at different potentials.
Figure 10. Raman spectra showing enlarged area of E 2 H i g h (438 cm−1) mode positions and FWHM of (a) ZnO/Cu and (b) ZnO/G/Cu films deposited at different potentials.
Nanomaterials 12 02858 g010
Figure 11. Modified Kubelka–Munk function for direct band gap energy (F(R) × E)2 versus photon energy (Inset) and UV–Vis total reflectance spectra of ZnO/Cu and ZnO/G/Cu.
Figure 11. Modified Kubelka–Munk function for direct band gap energy (F(R) × E)2 versus photon energy (Inset) and UV–Vis total reflectance spectra of ZnO/Cu and ZnO/G/Cu.
Nanomaterials 12 02858 g011
Figure 12. (af) PL spectra of ZnO/Cu and ZnO/G/Cu prepared at different applied potentials.
Figure 12. (af) PL spectra of ZnO/Cu and ZnO/G/Cu prepared at different applied potentials.
Nanomaterials 12 02858 g012
Figure 13. Gaussian fitting of the PL spectra of (a) ZnO/Cu, (b) ZnO/G/Cu films obtained using different applied potentials (black and color lines represent the experimental and the fitted data, respectively).
Figure 13. Gaussian fitting of the PL spectra of (a) ZnO/Cu, (b) ZnO/G/Cu films obtained using different applied potentials (black and color lines represent the experimental and the fitted data, respectively).
Nanomaterials 12 02858 g013
Figure 14. The CIE chromaticity diagram for (a) ZnO/Cu, (b) ZnO/G/Cu films.
Figure 14. The CIE chromaticity diagram for (a) ZnO/Cu, (b) ZnO/G/Cu films.
Nanomaterials 12 02858 g014
Table 1. Values of 2θ, FHWM, crystallite size, and average crystallite size D of ZnO/Cu and ZnO/G/Cu at different applied potentials.
Table 1. Values of 2θ, FHWM, crystallite size, and average crystallite size D of ZnO/Cu and ZnO/G/Cu at different applied potentials.
ZnO/CuZnO/G/Cu
Potential (V)(hkl) Peak
(Degree)
FHWM (Rad)Crystallite Size nmAverage Crystallite Size D (nm)
(Degree)
FHWM (Rad)Crystallite Size nmAverage Crystallite Size D (nm)
−0.8 (100)31.990.15655.2068.9031.980.14463.8067.03
(002)34.60.10086.60 34.570.09885.20
(101)36.450.15064.90 36.480.18552.10
−0.9(100)31.990.14662.3065.6531.980.23240.3058.38
(002)34.590.11284.10 34.570.09189.90
(101)36.500.16350.50 36.470.22044.70
−1 (100)31.990.19747.5056.1431.990.26436.6048.57
(002)34.590.12963.70 34.580.12267.40
(101)36.500.17357.00 36.480.24041.70
−1.1(100)31.990.18053.1057.1331.990.18430.1049.83
(002)34.600.13061.20 34.580.12080.20
(101)36.500.17357.00 36.480.23139.10
Table 2. Intensity ratio ID/IG and crystallite size (La) of G for samples of ZnO/G/Cu.
Table 2. Intensity ratio ID/IG and crystallite size (La) of G for samples of ZnO/G/Cu.
Potential (V)ID/IGLa (nm)
−0.8 1.3827.92
−0.91.3029.64
−11.2929.87
−1.11.2030.03
Table 3. Chromaticity coordinates and the PL intensities ratio of the UV emission to the visible emission for ZnO/Cu and ZnO/G/Cu films.
Table 3. Chromaticity coordinates and the PL intensities ratio of the UV emission to the visible emission for ZnO/Cu and ZnO/G/Cu films.
CIEIuv/Ivis
SamplesPotential (V)xy
ZnO/Cu−0.80.450.470.43
−0.90.470.480.66
−10.480.460.54
−1.10.460.460.81
−1.20.450.46-
ZnO/G/Cu−0.80.470.442.09
−0.90.470.480.40
−10.460.460.64
−1.10.370.460.87
−10.440.47-
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Boukhoubza, I.; Matei, E.; Jorio, A.; Enculescu, M.; Enculescu, I. Electrochemical Deposition of ZnO Nanowires on CVD-Graphene/Copper Substrates. Nanomaterials 2022, 12, 2858. https://doi.org/10.3390/nano12162858

AMA Style

Boukhoubza I, Matei E, Jorio A, Enculescu M, Enculescu I. Electrochemical Deposition of ZnO Nanowires on CVD-Graphene/Copper Substrates. Nanomaterials. 2022; 12(16):2858. https://doi.org/10.3390/nano12162858

Chicago/Turabian Style

Boukhoubza, Issam, Elena Matei, Anouar Jorio, Monica Enculescu, and Ionut Enculescu. 2022. "Electrochemical Deposition of ZnO Nanowires on CVD-Graphene/Copper Substrates" Nanomaterials 12, no. 16: 2858. https://doi.org/10.3390/nano12162858

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

Article Metrics

Back to TopTop