Next Article in Journal
Advances in Sustainable Development: Technological and Economic Overview
Next Article in Special Issue
Quantitative Evaluation of the Effects of Heat Island on Building Energy Simulation: A Case Study in Wuhan, China
Previous Article in Journal
Mapping Methodology for Self and Mutual Inductance Profiles in IPT Systems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Comparison between Solution-Based Synthesis Methods of ZrO2 Nanomaterials for Energy Storage Applications

1
CENIMAT|i3N, Department of Materials Science, School of Science and Technology, NOVA University Lisbon and CEMOP/UNINOVA, 2829-516 Caparica, Portugal
2
Physics Department & I3N, Aveiro University, Campus Universitário de Santiago, 3810-193 Aveiro, Portugal
*
Authors to whom correspondence should be addressed.
Energies 2022, 15(17), 6452; https://doi.org/10.3390/en15176452
Submission received: 26 July 2022 / Revised: 27 August 2022 / Accepted: 31 August 2022 / Published: 3 September 2022
(This article belongs to the Special Issue Solutions towards Zero Carbon Buildings)

Abstract

:
The present study is focused on the synthesis of zirconium dioxide (ZrO2) nanomaterials using the hydrothermal method assisted by microwave irradiation and solution combustion synthesis. Both synthesis techniques resulted in ZrO2 powders with a mixture of tetragonal and monoclinic phases. For microwave synthesis, a further calcination treatment at 800 °C for 15 min was carried out to produce nanopowders with a dominant monoclinic ZrO2 phase, as attested by X-ray diffraction (XRD) and Raman spectroscopy. The thermal behavior of the ZrO2 nanopowder was investigated by in situ XRD measurements. From the scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images, the presence of near spherical nanoparticles was clear, and TEM confirmed the ZrO2 phases that comprised the calcinated nanopowders, which include a residual tetragonal phase. The optical properties of these ZrO2 nanopowders were assessed through photoluminescence (PL) and PL excitation (PLE) at room temperature (RT), revealing the presence of a broad emission band peaked in the visible spectral region, which suffers a redshift in its peak position, as well as intensity enhancement, after the calcination treatment. The powder resultant from the solution combustion synthesis was composed of plate-like structures with a micrometer size; however, ZrO2 nanoparticles with different shapes were also observed. Thin films were also produced by solution combustion synthesis and deposited on silicon substrates to produce energy storage devices, i.e., ZrO2 capacitors. The capacitors that were prepared from a 0.2 M zirconium nitrate-based precursor solution in 2-methoxyethanol and annealed at 350 °C exhibited an average dielectric constant (κ) of 11 ± 0.5 and low leakage current density of 3.9 ± 1.1 × 10−7 A/cm2 at 1 MV/cm. This study demonstrates the simple and cost-effective aspects of both synthesis routes to produce ZrO2 nanomaterials that can be applied to energy storage devices, such as capacitors.

1. Introduction

Research in nanotechnology has grown over in recent years since it opens up possibilities for the design and fabrication of novel materials, as well as flexible and smart devices at the nanometer scale [1,2]. In this regard, nanostructures based on metal oxides have been extensively studied due to their unique physical and chemical properties and great potential in a myriad of applications, including electronics, energy conversion and storage, biomedicine, catalysis and sensing [3,4,5,6,7,8].
ZrO2 is a metal oxide that exhibits three crystalline phases under atmospheric pressure, depending on the synthesis route: monoclinic, which is thermodynamically the most stable form [9], tetragonal and cubic [10,11,12]. The cubic form of ZrO2 is stable at high temperatures (usually above 2370 °C), while the tetragonal is stable within the range of 1170–2370 °C, and the monoclinic phase stabilizes below 1170 °C [4,9].
ZrO2 exhibits excellent properties, such as good optical transparency in the visible and near-infrared spectral range [13,14], high thermal and chemical stabilities, mechanical strength and fracture toughness, low absorption of light, high index of refraction [15], high corrosion resistance [16], high ionic stability at high temperature, it is biocompatible [17] and presents relatively low leakage current [18] and high breakdown field [19]. Despite this, ZrO2 presents a wide band gap value (theoretically estimated as ~5.42 eV for monoclinic, ~6.40 eV for tetragonal and ~5.55 eV for cubic phases [14,20,21]). The introduction of defects in its structure, for instance, through doping with cations or the reduction of particle size below a critical value, is reported to allow the stabilization of high-temperature phases (i.e., cubic and/or tetragonal phases) at RT [22,23].
Applications for this material range from ceramic industry, gas sensing, solar fuel cells and biomedical/catalytic fields to its incorporation in large-scale integrated circuits, owing to a high dielectric constant and as a gate dielectric in metal-oxide-semiconductor devices (MOS) [22,24,25,26]. In this regard, ZrO2 has emerged as a potential alternative to replace low-κ SiO2 in electronic devices [27]. It is reported that the dielectric constant of ZrO2 films ranges from κ = 20 in the monoclinic phase to κ = 37–47 for the cubic and tetragonal phases, whereas in the case of amorphous ZrO2, the dielectric constant values are between 14 and 25 [14].
Recently ZrO2-based materials have been explored in the field of energy. Optimization of the light absorption towards sunlight in zirconia has shown enormous potential for its use in solar receivers since this refractory material possesses enhanced mechanical/thermal properties and oxidation resistance, being a competitive material to substitute the typically used silicon carbide (SiC) [28,29]. Other applications include refractory/metal production [30]. In addition, it has demonstrated high energy storage performance, for instance, by employing ferroelectric ZrO2 thin films as energy storage capacitors [31] or antiferroelectric ZrO2, a material for dynamic random access memories (DRAM) [32]. Moreover, ZrO2 capacitors display ultra-high power density and ultra-fast charge/discharge rates, yet their energy storage densities are relatively lower than the traditional electrical energy-storage devices, such as supercapacitors and batteries [33]. Nevertheless, these devices can play a key role in energy saving and thus reducing the carbon dioxide (CO2) footprint. When associated with renewable energy sources, capacitors can have a real impact on climate change, influencing almost all sectors of human life, including the chemical/petrochemical, automotive, textiles, wood/paper and construction sectors, whereas, for the latter, with the advent of the zero carbon buildings, it can reduce the global greenhouse gas emissions emitting buildings. The future of humanity relies on addressing the availability of using carbon-neutral energy sources and finding solutions to lower the CO2 footprint, industrial residues and human contaminants with direct impact in public health and on the environment.
Most of the applications for ZrO2 depend on its dimensions. It is well known that when the size of metal oxide materials is reduced to the nanoscale level, different properties are found from their bulk counterparts [34]. Novel characteristics owed to quantum size effects may occur when the cluster size is smaller than the Bohr exciton radius. Moreover, the properties of the surface start to play a relevant role that may overrule the bulk ones. Hence, the synthesis method is of paramount importance to control not only the size of ZrO2 nanomaterials but also the crystalline phase, morphology and lattice defects that highly influence the physico-chemical properties of the produced nanomaterials [35].
ZrO2 nanomaterials have already been obtained with different morphologies, such as thin films [14], nanoparticles [36], nanobelts [37], nanowires [38] and nanotubes [39], mainly through chemical approaches [40]. The most common methods for the synthesis of ZrO2 nanostructures include co-precipitation [41], sol-gel [42] and solution combustion [35], sonochemical-assisted methods [43,44], chemical vapor deposition [45], emulsion processing routes [46,47], as well as hydrothermal [10,48] and microwave syntheses [49,50]. Lately, a continuous increase in research work has been observed related to fast continuous methods for controlling the synthesis of ZrO2 nanomaterials [51,52,53,54,55]. Continuous hydrothermal flow synthesis (CHFS) processes are preferred due to the possibility of tailoring and obtaining unique nanoparticle properties that cannot be achieved using more conventional batch hydrothermal or other synthesis methods. These processes have shown potential in manufacturing reproducible and homogeneous nanomaterials, with high throughput, in addition to being energy efficient and eco-friendly, since for this last case, water is typically used rather than toxic organic solvents. Annealing treatments can also be avoided by employing these processes with superheated water conditions (typically in the range of 200−400 °C) [55,56,57]. Even though these systems offer several benefits, their high cost is a downside [57].
Several studies have reported the production of ZrO2 nanomaterials using sol-gel, particularly solution combustion synthesis, which is included in the sol-gel method [58,59,60]. Solution combustion synthesis is an attractive technique for the preparation of ZrO2 nanopowders and thin films, owing to its simplicity, energy and time savings, cost-effectiveness, versatility, higher purity compared with conventional sol-gel methods, low synthesis temperatures, and compatibility with flexible substrates and large scale production [61,62,63].
The synthesis of ZrO2 nanomaterials using the hydrothermal method assisted by microwave irradiation has also been growing exponentially over recent years [49,64,65,66], mainly due to its several advantages, such as volumetric heating (the entire volume of solution is evenly heated, instead of relying on heat diffusion processes across the reaction vessels), and the reaction times can be shorter as it is possible to synthesize nanostructures in just a few minutes. Furthermore, it also allows accurate control of particle morphology and size by adjusting the microwave parameters [50,67]. Upon optimization of the synthesis parameters, such as temperature, pH, time and zirconium oxide precursors (e.g., zirconyl chloride, zirconyl hydroxide, zirconyl nitrate hydrate and zirconium alkoxides), different ZrO2 phases can be obtained [68].
Solution-based ZrO2 nanomaterials are a highly appealing alternative to physical methods due to their process simplicity, high throughput, reduced equipment cost, since no vacuum-based systems are required [69,70], and the possibility to fabricate optoelectronic devices, even at low temperatures and by using green solvents [70]. As indicated, a broad range of solution-based synthesis methods have been developed to prepare ZrO2 nanomaterials; however, solution combustion synthesis and hydrothermal synthesis/microwave irradiation are typically preferred, and for that reason, these techniques are presented in this work. The same zirconium precursor (zirconium (IV) oxynitrate hydrate) was used in both synthesis techniques.
Several studies have already reported the dielectric properties of the ZrOx films produced by different solution-based processes. For instance, Seon et al. fabricated ZrO2 films via a non-hydrolytic sol-gel route at low temperatures, using 2-methoxyethanol (2-ME) as a solvent, with further annealing at 300 °C. The solution precursors chosen were zirconium chloride and zirconium isopropoxide, which were prepared in equimolar amounts, and acted, respectively, as a metal halide and a metal alkoxide. A breakdown voltage greater than 4 MV cm−1, a high dielectric constant (near 10) and a low leakage current density of 5 × 10−8 A cm−2 at a field of 1 MV cm−1 were obtained [71]. Another study by Gong et al. showed that by employing a low-temperature annealing treatment at 160 °C, high-quality amorphous ZrO2 dielectric films could be produced via a low-cost solution process. In this study, a solution was prepared using zirconium (IV) acetylacetonate as zirconium precursor and N, N-dimethylformamide as a solvent. Hydrolysis and condensation reactions occurred during stirring of the solution for 32 h at 90 °C. The films exhibited a leakage current of 3.6 × 10−5 A cm−2 at −3 V, a capacitance of ~117.1 nF cm−2 and a dielectric constant of 7.8, both at 1 KHz [72]. Wang et al. also fabricated high-κ ZrO2-dielectric films using a lightware (LW) irradiation-induced chloride-based low-temperature solution route. Results demonstrated a great capacitance of 270 nF cm−2, a high dielectric constant of 14.1 (at 100 Hz) and a low leakage current of 7.6 × 10−8 A cm−2, under 2 MV/cm. The superior performance was attributed to the effective formation of the metal-oxygen (M-O) framework that enabled the elimination of oxygen defects [73]. Another study by Jung et al. demonstrated the fabrication of ultrathin ZrOx films by deep ultraviolet irradiation, which revealed a leakage current density as low as 10−11 A cm−2 at 1 MV/cm, a capacitance of 260 nF cm−2 (at 1 MHz), high dielectric constant (22) and good breakdown voltage (around 6 MV/cm) [74]. In addition, Luo et al. prepared high-quality ZrO2 films using an oxygen-doped precursor solution (ODS). The ODS-ZrO2 films showed a low leakage current density of 10−7 A cm−2 (at 2 MV/cm), high breakdown electric field (7.0 MV/cm) and dielectric constant of 19.5 [75]. However, the challenge still relies on the fabrication of metal-oxide films with a high-quality surface (a smooth surface with a dense network) at a low temperature and through a simple approach to guarantee a low leakage current density and high breakdown field [75].
In this study, ZrO2 nanomaterials were produced in the form of powders or thin films through solution-based processes, i.e., a hydrothermal method assisted by microwave irradiation and solution combustion synthesis. The microwave synthesized powder was further calcinated at 800 °C for 15 min under atmospheric conditions. The ZrO2 nanomaterials were characterized by XRD, Raman spectroscopy, SEM coupled with energy dispersive X-ray spectroscopy (EDS) and focused ion beam (FIB) and TEM. The thermal behavior of the nanopowder produced under microwave irradiation was investigated through in situ XRD, and these powders had their optical properties assessed through PL and PLE at RT. The ZrOx thin films produced by the solution combustion synthesis were further tested as capacitors.

2. Materials and Methods

2.1. Hydrothermal Synthesis of ZrO2 Nanoparticles Assisted by Microwave Irradiation

The ZrO2 nanoparticle synthesis route was adapted from ref. [4]. In a typical synthesis, 50 mL of an aqueous (aq.) solution of 0.2 M of zirconium (IV) oxynitrate hydrate (Sigma-Aldrich, 99.9%, CAS: 14985-18-3, ZrO(NO3)2·xH2O) is mixed with 50 mL of an aq. solution of 0.4 M of sodium hydroxide (Labchem, CAS: 1310-73-2, NaOH). The reagents were used without any further purification. The 100 mL solution was left to stir for 30 min. The molar ratio of zirconium precursor and sodium hydroxide was kept at 1:2. Microwave synthesis was then carried out with a CEM microwave digestion system, Matthews, NC, USA (MARS one), and the applied microwave parameters were 1000 W, 230 ± 10 °C and 25 min. Afterwards, the previous solution was equally distributed into Teflon vessels of 75 mL (each vessel containing 20 mL of solution). Subsequently, the centrifugation of the resultant nanopowder was performed for 3 min at 4750 rpm and washed three times alternately with deionized water and isopropyl alcohol (IPA). Finally, the nanopowder was dried in a desiccator at 60 °C for 5 h. The yield was around 0.77 g of nanopowder/batch.
After microwave synthesis, the dried ZrO2 nanopowder was further calcinated in an alumina ceramic crucible at 800 °C for 15 min using a Nabertherm furnace under atmospheric conditions. The calcination treatment aimed to guarantee the formation of the mostly thermodynamically stable ZrO2 phase, i.e., the monoclinic phase.

2.2. Solution Combustion Synthesis

A solution with a concentration of 0.2 M of zirconium (IV) oxynitrate hydrate (Sigma-Aldrich, St. Louis, MO, USA, 99.9%, CAS: 14985-18-3) was prepared in 2-methoxyethanol (2-ME, ACROS Organics, 99%, C₃H₈O₂), and it was left to stir at room temperature for 2 h. Urea (Sigma-Aldrich, 98%, CO(NH2)2) was used as fuel for the combustion reaction and continued stirring for a minimum of 1 h. The molar proportion between urea and the zirconium oxide precursor was 5:3 to ensure the redox stoichiometry of the reaction. The powder was produced considering 10 mL of the combustion solution with the same concentration and transferred to an alumina ceramic crucible for further annealing in an air furnace at 350 °C for 1 h.
As for the thin films, and prior to the ZrOx deposition, a 2.5 × 2.5 cm p-type single crystal 100-oriented silicon substrate (resistivity ≅ 1–2 Ω∙cm) was cleaned with acetone for 10 min in an ultrasonic cleaning bath at 60 °C. This cleaning process was repeated with IPA, followed by deionized water and then the substrates were dried using nitrogen flow. The adhesion properties of the thin films to the substrate were enhanced by exposing the substrate to UV-ozone treatment (PSD-UV Novascan system-Ames, IA, USA) for 15 min at a 5 cm distance from the UV lamp. Prior to solution casting, the precursor solution was filtered through a 0.22 µm polytetrafluoroethylene (PTFE) syringe filter and one layer was deposited by spin coating on the silicon (Si) substrate at 2000 rpm for 35 s. To evaporate the solvent and improve the densification, the thin films were later annealed at 350 °C on a hot plate for 30 min.
The metal-insulator-semiconductor (MIS) capacitors were fabricated considering the structure Al/ZrOx/p-Si. The schematic of the MIS devices’ fabrication is depicted in Figure 1. To optimize the ohmic contact, the capacitors were completed by thermal evaporation of an 80 nm aluminum bottom electrode on the Si substrate and 80 nm Al top electrodes by applying a circular shadow mask, which produced capacitors with an area of 0.2 mm2.

2.3. Characterization Techniques

XRD experiments were carried out using a PANalytical’s X’Pert PRO MPD diffractometer (Almelo, The Netherlands) equipped with an X’Celerator 1D detector and using CuKα radiation (λ = 1.540598 Å). XRD data were recorded from 20° to 80° 2θ range with a step of 0.05°. The produced ZrO2 nanopowders were measured in the Bragg–Brentano configuration. The crystalline phases in the produced nanopowders were identified by comparison with the International Centre for Diffraction Data (ICDD). The simulated monoclinic phase corresponds to ICDD file No. 00-037-1484, the simulated tetragonal phase to ICDD file No. 00-050-1089 and the simulated cubic phase to ICDD file No. 00-049-1642. In situ XRD experiments were performed in an MRI chamber from Bruker (Massachusetts, United States of America) at atmospheric conditions using a Pt–Rh foil as the heating element. The in situ diffractograms were recorded in the same 2θ range (20° to 80°) at temperatures of 30, 100, 200, 300, 400, 500, 600, 700 and 800 °C. The ZrO2 nanopowder was kept at each temperature for 15 min, during which 15 consecutive scans were performed between 27° and 33° with a scanning step of 0.017° to investigate the evolution of the monoclinic and tetragonal phases. The temperature was increased at a rate of 50 °C/min.
The crystallinity of the produced ZrOx thin films on silicon substrates annealed at 350 °C was also investigated using the same XRD equipment in grazing incidence mode (GIXRD, X’Pert PRO PANalytical) with a step of 0.2°/min and an angle of incidence of the X-ray beam fixed at 0.5° in the range of 10–60° (2θ).
SEM images were acquired using a Hitachi Regulus 8220 Scanning Electron Microscope (Mito, Japan) equipped with energy dispersive X-ray spectroscopy (EDS) equipment. For the FIB experiments, a Carl Zeiss AURIGA CrossBeam (FIB-SEM) workstation (Carl Zeiss MicrosPANI/TiO2 copy GmbH, Oberkochen, Germany) was used. The inner structure of the ZrO2 capacitor was observed by FIB, where Ga+ ions were accelerated to 30 kV at 50 pA, and the etching depth was maintained at around 500 nm. TEM observations were performed with a Hitachi HF5000 field-emission transmission electron microscope (Mito, Japan) operated at 200 kV. A drop of the sonicated dispersions was deposited onto 200-mesh lacey-carbon copper grids and allowed to dry before observation. The average particle size and standard deviation were calculated from the dimensions of 50 nanoparticles with ImageJ software based on TEM images.
Raman spectroscopy measurements were conducted with an inVia Qontor confocal Raman microscope from Renishaw (Kingswood, UK). A 50 mW green diode operated at 532 nm was used as the excitation source, with an exposure time of 10 s and settings of 3 and 7 accumulations for ZrO2 nanopowders before and after annealing treatment, respectively. The laser beam was focused with a long working distance (8.2 mm) 50 × Olympus objective and a 100 × Olympus objective with 0.35 mm working distance, respectively, for the samples before and after annealing treatment. The Raman spectra were recorded in the range of 110–800 cm−1 (as an extended scan). Several scans on different points of the ZrO2 nanopowder’s surface were recorded, and the present results are based on their average. Possible fluctuations of the Raman equipment were avoided with a previous calibration with a silicon wafer (521 cm−1 peak). All Raman measurements were performed at ambient conditions.
RT PL and PLE measurements were carried out in the microwave synthesized ZrO2 nanopowders, before and after calcination, using a Fluorolog-3 Horiba Scientific set-up with a double additive grating Gemini 180 monochromator (1200 gr/mm and 2 × 180 mm) in the excitation and a triple grating iHR550 spectrometer in the emission (1200 gr/mm and 550 mm). A 450 W Xe lamp was used as the excitation source, and different excitation wavelengths were explored. The PLE data was obtained by monitoring the maximum of the PL emission.
The electrical and dielectric properties such as capacitance-voltage (CV), capacitance-frequency (Cf) and current-voltage (IV) measurements of the MIS capacitors ZrOx-based were investigated by using a semiconductor parameter analyzer (Keysight B1500A) with a probe station (Cascade EPS150 Triax).

3. Results and Discussion

3.1. Structural and Optical Characterization of the ZrO2 Nanopowders Produced under Microwave Irradiation

3.1.1. Raman Spectroscopy Measurements

ZrO2 nanopowder was synthesized through a hydrothermal method assisted by microwave irradiation. Raman spectroscopy measurements were performed to investigate the crystalline phase of the synthesized ZrO2 nanoparticles [76]. A further calcination process was also carried out at 800 °C for 15 min. Figure 2a,b show the Raman spectra in the 110–800 cm−1 range for the synthesized ZrO2 nanopowders before and after the calcination treatment, respectively. Regarding the as-synthesized ZrO2 powder, as seen in Figure 2a, broader Raman bands were observed when a comparison was made with the calcinated material (Figure 2b) [77]. Nevertheless, some of the peak positions of the vibrational modes occurred at the same frequencies, meaning that contributions of the same ZrO2 polymorph were present in both materials. Tetragonal zirconia (D4h point group) was expected to exhibit six Raman active vibrational modes at the Γ point of the first Brillouin zone with A1g + 2B1g + 3Eg symmetries [78]. The frequencies of these modes were reported as 146 cm−1 (Eg), 270 cm−1 (A1g), 318 cm−1 (B1g), 458 cm−1 (Eg), 602 cm−1 (B1g) and 648 cm−1 (Eg) [79]. For the case of the monoclinic ZrO2 (C2h point group), group theory analysis predicted 18 Raman active modes, 9Ag + 9Bg [80], with frequencies reported at 179 cm−1 (Ag), 190 cm−1 (Ag), 224 cm−1 (Bg), 305 cm−1 (Ag), 334 cm−1 (Bg), 348 cm−1 (Ag), 381 cm−1 (Bg), 476 cm−1 (Ag), 505 cm−1 (Bg), 536 cm−1 (Bg), 556 cm−1 (Ag), 616 cm−1 (Bg), 637 cm−1 (Ag) and a non-indexed peak at 757 cm−1 [77,81]. By comparing the spectra of the two materials in Figure 2, we were able to identify that the vibrational modes matched those reported for both crystalline phases [77,81]. In particular, the bandwidth of the Raman lines in the as-synthesized material was likely to be due to the overlap of the vibrational signatures from the tetragonal and monoclinic ZrO2, as clearly identified by the presence of the 146 cm−1 and 265 cm−1 resonances and the bandshape asymmetry observed at 458 cm−1 and 648 cm−1, which were assigned to the tetragonal phase. In contrast with Figure 2a, no bands related to the tetragonal ZrO2 are visible in Figure 2b; however, its presence cannot be excluded. In addition, another band at 501 cm−1 (Bg) is visible, assigned to monoclinic ZrO2 [77,81]. Cubic zirconia exhibits a characteristic broad band mode centered at 625 cm−1 due to disordered oxygen sub-lattice. This vibrational mode is not visible in Figure 2a,b. Moreover, the absence of the cubic phase was already expected since it is stable only at annealing temperatures near to ZrO2 melting point or at RT by doping processes [82]. No other bands were detected [83].

3.1.2. X-ray Diffraction

XRD measurements were also performed to confirm the crystalline phase/phases obtained from Raman spectroscopy measurements. As seen in Figure 3a, the as-synthesized ZrO2 nanopowder presents a mixture of ZrO2 tetragonal and monoclinic phases, considering the ICDD card numbers 00-037-1484 and 00-050-1089 for the ZrO2 monoclinic and tetragonal phases, respectively.
Calcination treatment after microwave synthesis is an important step in transforming an amorphous to crystalline phase or a single/mixture of crystalline phases into a specific crystalline phase, inducing optical and structural’ modifications to the material [13,84]. In this regard, XRD diffractograms of the ZrO2 nanopowder were recorded at different temperatures to infer the thermal behavior of the ZrO2 nanomaterials and thus analyze the phase transition from tetragonal to monoclinic, Figure 3a. As can be observed, the intensity of the diffraction maximum at 30.176°, characteristic of the tetragonal phase, decreases with increasing temperature. At 800 °C, the peaks detected in the diffractogram of the ZrO2 nanopowder (Figure 3b) were fully assigned to the monoclinic ZrO2 phase (ICDD 00-037-1484). However, when performing XRD studies from 27° and 33° (Figure 3c) with a fixed temperature of 800 °C but different annealing times (each minute was recorded up to 15 min), it is evident that the conversion to monoclinic is accomplished at 800 °C with a minimal presence of the tetragonal phase. No other crystalline phases or impurities were detected with either of these two techniques.
The phase transformation of ZrO2 tetragonal to monoclinic at 800 °C has been previously reported in the literature. Horti et al. [13] prepared ZrO2 nanoparticles by chemical co-precipitation method and further annealed at different temperatures for 2 h, whose phase transition of tetragonal to monoclinic was seen to occur at the calcination temperature of 800 °C, with an evident mixture of both phases. In another study [85], ZrO2 nanoparticles were produced by conventional and ultrasound-assisted precipitation in an alkaline medium followed by calcination at different temperatures ranging from 400 °C to 900 °C for 3 h. At 800 °C, a mixture of three ZrO2 phases was reported, i.e., tetragonal-monoclinic-cubic phases. The present study has the advantage of having monoclinic ZrO2 as the dominant phase after a fast calcination treatment of 15 min instead of hours.

3.1.3. Electron Microscopy

Figure 4 shows the SEM images of the ZrO2 nanoparticles synthesized under microwave irradiation with a further calcination treatment at 800 °C for 15 min. It can be seen that the particles display a nearly spherical shape (Figure 4a), and from the EDS measurements, a homogeneous distribution of Zr (Figure 4b) and O (Figure 4c) is evident. No impurities were detected by EDS.
The atomic percentages (at.%) of Zr and O were also estimated from EDS spectra for the ZrO2 nanopowders before and after the calcination treatment (Table 1). The atomic percentage values were close to those expected for pure zirconium oxide (the Zr/O ratio was 1:2 for the material before calcination and 1:2.7 after calcination).
TEM measurements were also carried out for the ZrO2 nanoparticles produced under microwave irradiation before and after calcination (Figure 5). The TEM images of the ZrO2 nanoparticles before calcination demonstrate the presence of heterogeneous particles with an average particle size of 6.7 ± 1.9 nm. Abnormal large particles were also observed (Figure 5a,b). In terms of particle shape, microwave synthesis resulted in different nanostructures, including nanospheres, nanosquares and irregular-shaped nanoparticles. The heterogeneity detected is consistent with the XRD and Raman spectroscopy results, where it was shown that after microwave synthesis, there was a mixture of ZrO2 phases, with the presence of both tetragonal and monoclinic ZrO2 phases (Figure 2a and Figure 3a). The lattice spacing of 0.3 nm measured in Figure 5c was consistent with the d-spacing of the ( 101 ) and ( 011 ) planes of the tetragonal ZrO2. Furthermore, the fast Fourier transform (FFT) image carried out through the [ 11 1 ¯ ] zone axis, attested the existence of pure tetragonal ZrO2 nanocrystals.
After calcination, as expected, an expressive growth of the ZrO2 nanoparticles and the presence of uniform nanoparticles in terms of size and shape were observed (Figure 5), exhibiting a nearly spherical shape, in agreement with SEM results (Figure 4a), and displaying an average particle size of 45.7 ± 9.9 nm. Figure 5e,f reveal that the measured lattice spacings of 0.32 and 0.37 nm are in good agreement with the d-spacing of the ( 1 ¯ 11 ) and ( 011 ) planes of the monoclinic ZrO2 phase, respectively. The FFT image captured through the [ 01 1 ¯ ] zone axis (Figure 5f) also confirmed the existence of pure monoclinic ZrO2 nanocrystals. From TEM measurements, it was observed that the monoclinic ZrO2 phase prevailed for the 800 °C calcinated material; however, some minor contaminations of tetragonal nanoparticles were also observed in accordance with the in situ XRD results (Figure 3c and Figure S1).

3.1.4. Photoluminescence Measurements

Figure 6 depicts the results obtained for both materials before and after calcination, with clear differences in the PL spectra. In the case of the as-synthesized nanopowder without calcination (Figure 6a), an asymmetric band was observed in the visible spectral region, with a maximum at ~415 nm (~2.99 eV) when excited with different wavelengths in the UV range. After calcination in air (Figure 6b), the visible band experienced a broadening and an increase in its intensity. Under 240 nm photon excitation, its peak position was located at ~459 nm (~2.70 eV), whereas for longer wavelengths (260–290 nm), the PL maximum shifted to longer wavelengths, around ~475 nm (~2.61 eV). Indeed, the band observed at ~459 nm is likely to be composed of several recombination channels, including the ones which are preferentially excited with photons with wavelengths in the 260–290 nm range. The PLE monitored at the maximum of the observed PL bands revealed, in both cases, an excitation band at ~310 nm (~4.00 eV), a noticed shoulder at ~260 nm (~4.77 eV) and an excitation band maximum at ~240 nm (~5.17 eV), indicating that those are the preferential population paths for the observation of such emission.
Broad visible bands were previously reported for the different undoped ZrO2 polymorphs, namely the monoclinic one; however, the nature of the defects that originate from them is still not fully understood [86]. Among the suggested origins, oxygen vacancies and their complexes have been pointed out as the most probable defects, even though the presence of some impurities, such as Ti, have also been proposed [86,87,88,89,90]. For instance, Perevalov et al. [89] studied non-stoichiometric zirconia samples prepared at different oxygen pressures and inferred that the origin of the blue emission band peaked at 2.7 eV is likely associated with charged oxygen vacancy states since a decrease of the oxygen pressure during the grown leads to an increase in the intensity of the mentioned band. Moreover, slight shifts in the peak position (between 2.6 and 2.8 eV) were seen depending on the stoichiometry of the samples. In line with the present results, the reported PLE spectra also evidence a maximum at 5.1–5.2 eV, which they associated with the optical absorption of a defect created by oxygen vacancies. Nevertheless, according to the literature, the expected band gap energy for the ZrO2 monoclinic phase predicted by theoretical models is in the range of ~5–6 eV [21,91,92,93], which is also in line with some experimental results [94,95,96]. Therefore, a likely explanation for the observed high energy excitation band that occurs in the vicinity of 240 nm (~5.17 eV) is that it corresponds to the band gap of the monoclinic ZrO2 phase, in fair agreement with the values theoretically predicted for this crystalline phase. As mentioned, the produced nanoparticles are polyphasic in nature, exhibiting a contribution of the tetragonal ZrO2 crystalline phase, which is expected to have a higher band gap energy than the monoclinic one. Therefore, the excitation shoulders/bands detected around ~260 nm and ~310 nm should be related to defects/impurities in electronic states inside the material’s band gap.
Ashraf et al. [90] analyzed ZrO2 monoclinic nano- and submicron crystals annealed at different temperatures and also observed broad emission bands from ~400 nm to 600 nm, whose spectral shape and intensity slightly varied depending on the annealing conditions and crystallite sizes. The authors assigned the PL emission centered at ~482 nm (~2.57 eV) to a transition involving F+ centers, i.e., transitions from the valence band to local mid-gap states associated with singly occupied anion (oxygen) vacancies. At the shorter wavelength region, a shoulder at 422 nm (~2.94 eV) was also visible in the case of the nanoscale ZrO2 before annealing and associated with a transition to singly ionized associated oxygen vacancy defects. After annealing, this shoulder disappeared, likely related to the dissociation of those centers into F+ centers by the aid of atmospheric oxygen diffusion into the ZrO2 particles [90]. A similar phenomenon may be occurring in the present materials, explaining the redshift observed in the peak position of the band after calcination. In the case of Ashraf et al. [90], further annealing in air at higher temperatures (1500 °C) promoted additional diffusion of oxygen into the crystals, leading to a decrease in the broad band PL intensity, strengthening the assumption of the involvement of oxygen vacancy defects in the origin of this band. Their PLE spectra showed an excitation band between ~250 nm and 350 nm, which suffered a blueshift from ~300 nm (~4.13 eV) to ~280 nm (~4.43 eV) after annealing. In 2021, Lokesha et al. [88] also studied monoclinic ZrO2 samples, observing a broad band between 400 and 650 nm and peaked at 499 nm (~2.48 eV), which, in conjugation with other techniques, such as EPR, they attributed to the F2+ center related with aggregates of the singly occupied oxygen vacancies. These broad bands are frequently deconvoluted into different components (recombination channels), which can be associated with different charged states of the vacancy-related defect [88,90,97]. As a result, the existence of distinct charge states of the F-centers can account for the fact that three excitation maxima were observed in the present samples, subsequently leading to a shift in the band peak position depending on the excitation wavelength, particularly in the case of the calcinated material. Hence, the presence of oxygen vacancy-related defects seems to be a fair hypothesis for the nature of the defects involved in the PL emission identified in the present materials.

3.2. Structural Characterization of the ZrO2 Powder and ZrOx Thin Films Produced by Solution Combustion Synthesis

3.2.1. ZrO2 Powder

X-ray Diffraction

Solution combustion synthesis is known to produce materials in powder form but also as thin films [58,60,98]. The ZrO2 powder was obtained by conventional solution combustion synthesis, with zirconium oxynitrate and urea as precursors and 2-methoxyethanol (2-ME) as a solvent. The prepared solution was further annealed in air at 350 °C for 1 h to obtain the ZrO2 powder. The XRD diffractogram of the ZrO2 powder is shown in Figure 7. The ZrO2 tetragonal phase is clearly identified with the peak at 30.176°, associated with the (111) diffraction plane [99] (ICDD 00-050-1089), along with other diffraction maxima that were assigned to the monoclinic phase (ICDD 00-037-1484). No impurities were detected, which suggests the presence of a highly pure ZrO2 powder.

3.2.2. Electron Microscopy

Figure 8 shows the SEM images of the ZrO2 powder resultant from the solution combustion synthesis. The formation of micro-sized plate-like structures is clear, as seen in Figure 8a. Nevertheless, Figure 8b indicates the presence of nano-sized grains composing the micro-sized structure, and Figure S2 suggests the stacking of several nanolayers. Individual nanoparticles were also observed. Moreover, the void-like nature of the particle observed in Figure S2 can be associated with the escape of gaseous combustion products that were formed during combustion synthesis [58]. EDS measurements were also carried out, and, as observed for the ZrO2 nanoparticles produced under microwave irradiation (Figure 4), the presence of a homogeneous distribution of Zr (Figure 8c) and O (Figure 8d) was clear. No impurities were detected by this technique.
The atomic percentages of Zr and O were also estimated by EDS point analysis (Table 2). As observed for the nanopowders synthesized under microwave irradiation, the values were within the expected range for pure zirconium oxide (the Zr/O ratio was 1:2.1).
The ZrO2 powder produced by solution combustion synthesis was also observed by TEM, confirming the presence of individual ZrO2 nanoparticles. From Figure 9a, nano-sized particles with different shapes can be observed, including nanospheres, nano squares and irregular-shaped nanoparticles (Figure 9a). Larger particles without a specific shape were also observed; nevertheless, the average particle size was 10 ± 7 nm. The lattice spacing of 0.3 nm was measured on an individual ZrO2 nanoparticle (Figure 9b), which was consistent with the d-spacing of the ( 101 ) plane of the tetragonal ZrO2. Nonetheless, as observed by XRD results, the powder was a mixture of tetragonal and monoclinic ZrO2 phases.

3.2.3. ZrOx Thin Films

X-ray Diffraction

The ZrOx thin films were produced with the same solution used for producing the ZrO2 powder by solution combustion. Spin coating was employed to deposit ZrOx thin films on silicon substrates to be tested as capacitors. Crystalline dielectric films may provide, for instance, pathways for leakage current due to the grain boundaries, which results in a small on/off ratio, leading to poor capacitor performance. On the contrary, amorphous thin films have a smooth surface, high stability, low interface state density, as well as large-area uniformity [100,101].
Grazing angle mode was performed to analyze the crystalline structure of ZrOx thin films annealed at 350 °C in air for 30 min on a hot plate. An annealing treatment at 350 °C was carried out since it was reported that annealing treatments in solution-processed high-κ ZrOx thin films between 300–600 °C are expected to increase O−2 content in ZrOx films, leading to a reduction of oxygen defects and/or hydroxyl groups and consequently of conduction paths [69,102].
As observed in Figure 10, some diffraction maxima are visible at 2θ = 51.4° and 2θ = 53.4°, accompanied by a bump at around 55° (2θ), which correspond to forbidden diffraction planes of a (100)-oriented silicon wafer that appear when the grazing angle mode is used at certain phi angle [103]. Apart from that, no other diffraction maxima were identified, indicating the amorphous nature of the produced ZrOx thin films, which is ideal for thin film capacitor applications.

3.3. Electrical Characterization of the ZrOx Capacitors

ZrOx thin films were implemented in metal-insulator-semiconductor (MIS) devices to evaluate their dielectric performance. Figure 11a show the typical C–V curve of the ZrOx thin film capacitors with a small hysteresis. The devices present good stability with a small change of the capacitance over the frequency (1 kHz–100 kHz), as depicted in Figure 11b. To calculate the dielectric contact, a cross-section of the device to determine the ZrOx thickness was performed, as shown in the inset of Figure 11b. The ZrOx thin films present a capacitance per unit area of 672 ± 28 nF cm−2 and a dielectric constant of 11.0 ± 0.5 (both calculated at 1 kHz). These values are lower when compared with vacuum deposited crystalline films due to the lower film density and presence of pores as a consequence of the solution process [71,72,104,105,106]. Nevertheless, the devices present a good breakdown field of 2.8 ± 0.1 MV/cm and a low leakage current density of (3.9 ± 1.1) × 10−7 A/cm2 at 1 MV/cm.
Figure 11c shows a typical current density versus breakdown field dependence. The results are in accordance with solution-based ZrOx thin film capacitors [71]. Once low leakage current density was reached (<10−5 A/cm2), it opens the window for future applications such as gate insulators for solution-based thin film transistors.

4. Conclusions

This work demonstrated that both the hydrothermal method assisted by microwave irradiation and solution combustion synthesis are reliable and simple production routes for ZrO2 nanomaterials, including nanoparticles and thin films. Moreover, the low-cost character of the study is evidenced by the inexpensive apparatus required to produce and deposit ZrOx thin films and thus fabricate energy storage devices. In both synthesis routes, XRD analysis revealed that the as-synthesized powders had a mixture of ZrO2 monoclinic and tetragonal phases. A calcination treatment at 800 °C for 15 min was imposed on the microwave synthesized nanopowder to promote the conversion into single-phase materials. Nevertheless, in situ XRD and TEM analysis revealed a minor presence of the ZrO2 tetragonal phase after the calcination treatment. Different nanostructures were observed after synthesis in both production routes. In the case of the microwave synthesized material, after calcination, near-spherical nanoparticles were observed, while plate-like structures composed of nano-sized grains were identified for the combustion synthesis. The effect of calcination on the optical properties of the microwave synthesized nanopowders was evaluated by PL and PLE, showing a clear broadening and an increased intensity of the ZrO2 band observed in the visible spectral region after calcination. This band was tentatively associated with the presence of oxygen vacancy-related defects. Solution combustion synthesis also resulted in thin films, and XRD analysis suggested their amorphous nature. SEM and FIB measurements of the ZrOx capacitors revealed that the thin films present a thickness of 14.2 ± 0.1 nm, which leads to a dielectric constant of 11.0 ± 0.5 (calculated at 1 kHz). The devices also presented a good breakdown field of 2.8 ± 0.1 MV/cm and a low leakage current density of (3.9 ± 1.1) × 10−7 A/cm2 at 1 MV/cm. From this study, the production of ZrO2 nanomaterials with two simple, cost-effective and easily upscaling synthesis routes and their integration on capacitors can be highlighted.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/en15176452/s1, Figure S1: High-resolution TEM image of the ZrO2 nanocrystals calcinated at 800 °C for 15 min. The inset shows the FFT images of the area (black square) indicated as A; Figure S2: SEM image of a ZrO2 particle produced by solution combustion synthesis and annealed at 350 °C.

Author Contributions

M.L.M. was responsible for producing and characterizing microwave synthesized materials, as well as writing the manuscript. H.d.V. was responsible for the production and characterization of the ZrOx films, and J.M. was responsible for the combustion powder synthesis. M.M. and A.P. were responsible for the XRD measurements. D.N., R.B. and E.C. were responsible for the overall scientific orientation and revising the manuscript. D.N. performed the TEM observations and analysis. T.M. and J.R. were responsible for the optical measurements, respective analyses and revising of the manuscript. E.F. and R.M. were responsible for supervising all the processes and obtaining funding. Part of the work resulted from the Master of Science thesis of H.d.V., titled: Effect of eco-friendly solvents in solution-based ZrOx dielectrics, developed at the Nova School of Science and Technology. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financed by national funds from FCT-Fundação para a Ciência e a Tecnologia, I.P., in the scope of the projects UI/BD/151292/2021 (Ph.D. research scholarship), LA/P/0037/2020, UIDP/50025/2020 and UIDB/50025/2020 of the Associate Laboratory Institute of Nanostructures, Nanomodelling and Nanofabrication-i3N, but also the 2021.03825.CEECIND. The work was also partially funded by the Nanomark collaborative project between INCM (Imprensa Nacional-Casa da Moeda) and CENIMAT/i3N. Acknowledgments also go to the EC project SYNERGY H2020-WIDESPREAD-2020-5, CSA, proposal n° 952169, EMERGE-2020-INFRAIA-2020-1, proposal n° 101008701, and to the European Community’s H2020 program under grant agreement No. 787410 (ERC-2018-AdG DIGISMART).

Data Availability Statement

The authors confirm that the data supporting the findings of this study are available within the article and its Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shah, N.A.; Gul, M.; Abbas, M.; Amin, M. Synthesis of Metal Oxide Semiconductor Nanostructures for Gas Sensors. In Gas Sensors; IntechOpen: London, UK, 2019; ISBN 978-1-78985-160-1. [Google Scholar]
  2. Roco, M.C.; Mirkin, C.A.; Hersam, M.C. Nanotechnology research directions for societal needs in 2020: Summary of international study. J. Nanoparticle Res. 2011, 13, 897–919. [Google Scholar] [CrossRef]
  3. Chavali, M.S.; Nikolova, M.P. Metal oxide nanoparticles and their applications in nanotechnology. SN Appl. Sci. 2019, 1, 1–30. [Google Scholar] [CrossRef]
  4. Ahmad, T.; Shahazad, M.; Phul, R. Hydrothermal synthesis, characterization and dielectric properties of zirconia nanoparticles. Mater. Sci. Eng. Int. J. 2017, 1, 100–104. [Google Scholar] [CrossRef]
  5. Nunes, D.; Pimentel, A.; Gonçalves, A.; Pereira, S.; Branquinho, R.; Barquinha, P.; Martins, R. Metal Oxide Nanostructures for Sensor Applications. Semicond. Sci. Technol. 2019, 34, 1–178. [Google Scholar] [CrossRef]
  6. Nunes, D.; Pimentel, A.; Santos, L.; Barquinha, P.; Pereira, L.; Fortunato, E.; Martins, R.; Pimentel, A.; Barquinha, P.; Pereira, L.; et al. Metal Oxide Nanostructures: Synthesis, Properties and Applications, 1st ed.; Korotcenkov, G., Ed.; Elsevier: Amsterdam, The Netherlands, 2018; ISBN 9780128115121. [Google Scholar]
  7. Nunes, D.; Fortunato, E.; Martins, R. Flexible nanostructured TiO2-based gas and UV sensors: A review. Discov. Mater. 2022, 2, 1–23. [Google Scholar] [CrossRef]
  8. Nunes, D.; Pimentel, A.; Branquinho, R.; Fortunato, E.; Martins, R. Metal oxide-based photocatalytic paper: A green alternative for environmental remediation. Catalysts 2021, 11, 504. [Google Scholar] [CrossRef]
  9. Vaidya, S.; Ahmad, T.; Agarwal, S.; Ganguli, A.K. Nanocrystalline Oxalate/Carbonate Precursors of Ce and Zr and Their Decompositions to CeO2 and ZrO2 Nanoparticles. J. Am. Ceram. Soc. 2007, 90, 863–869. [Google Scholar] [CrossRef]
  10. Behbahani, A.; Rowshanzamir, S.; Esmaeilifar, A. Hydrothermal Synthesis of Zirconia Nanoparticles from Commercial Zirconia. Procedia Eng. 2012, 42, 908–917. [Google Scholar] [CrossRef]
  11. Van Tran, T.; Nguyen, D.T.C.; Kumar, P.S.; Din, A.T.M.; Jalil, A.A.; Vo, D.V.N. Green synthesis of ZrO2 nanoparticles and nanocomposites for biomedical and environmental applications: A review. Environ. Chem. Lett. 2022, 20, 1309–1331. [Google Scholar] [CrossRef]
  12. Hassan, N.S.; Jalil, A.A. A review on self-modification of zirconium dioxide nanocatalysts with enhanced visible-light-driven photodegradation of organic pollutants. J. Hazard. Mater. 2022, 423 (Pt A), 126996. [Google Scholar] [CrossRef]
  13. Liu, B.; Wang, C.; Chen, Y.; -Structural, A.; Properties of, L.Y.; Zirconia Nanopowders Korsunska, C.-C.N.; Baran, M.; Polishchuk, Y.; Horti, N.C.; Kamatagi, M.D.; et al. Structural and optical properties of zirconium oxide (ZrO2) nanoparticles: Effect of calcination temperature. Nano Express 2020, 1, 1–10. [Google Scholar] [CrossRef]
  14. Boratto, M.H.; Lima, J.V.M.; Scalvi, L.V.A.; Graeff, C.F.O. Low-temperature ZrO2 thin films obtained by polymeric route for electronic applications. J. Mater. Sci. Mater. Electron. 2020, 31, 16065–16072. [Google Scholar] [CrossRef]
  15. Namavar, F.; Wang, G.; Cheung, C.L.; Sabirianov, R.F.; Zeng, X.C.; Mei, W.N.; Bai, J.; Brewer, J.R.; Haider, H.; Garvin, K.L. Thermal stability of nanostructurally stabilized zirconium oxide. Nanotechnology 2007, 18, 1–6. [Google Scholar] [CrossRef]
  16. Abd El-Lateef, H.M.; Khalaf, M.M. Corrosion resistance of ZrO2–TiO2 nanocomposite multilayer thin films coated on carbon steel in hydrochloric acid solution. Mater. Charact. 2015, 108, 29–41. [Google Scholar] [CrossRef]
  17. Septawendar, R.; Sutardi, S.; Karsono, U.; Sofiyaningsih, N. A Low-Cost, Facile Method on Production of Nano Zirconia and Silica from Local Zircon in a Large Scale Using a Sodium Carbonate Sintering Technology. J. Aust. Ceram. Soc. 2016, 52, 92–102. [Google Scholar]
  18. Xifeng, L.; Enlong, X.; Jianhua, Z. Low-temperature solution-processed zirconium oxide gate insulators for thin-film transistors. IEEE Trans. Electron Devices 2013, 60, 3413–3416. [Google Scholar] [CrossRef]
  19. Chang, J.P.; Lin, Y.S. Dielectric property and conduction mechanism of ultrathin zirconium oxide films. Appl. Phys. Lett. 2001, 79, 3666. [Google Scholar] [CrossRef]
  20. Zandiehnadem, F.; Murray, R.A.; Ching, W.Y. Electronic structures of three phases of zirconium oxide. Physica B+C 1988, 150, 19–24. [Google Scholar] [CrossRef]
  21. Králik, B.; Chang, E.K.; Louie, S.G. Structural properties and quasiparticle band structure of zirconia. Phys. Rev. B Condens. Matter 1998, 57, 7027–7036. [Google Scholar] [CrossRef]
  22. Poungchan, G.; Ksapabutr, B.; Panapoy, M. One-step synthesis of flower-like carbon-doped ZrO2 for visible-light-responsive photocatalyst. Mater. Des. 2016, 89, 137–145. [Google Scholar] [CrossRef]
  23. Soares, M.R.N. Development of Zirconia Based Phosphors for Application in Lighting and as Luminescent Bioprobes. Ph.D. Thesis, University of Aveiro, Aveiro, Portugal, 2016. Available online: https://ria.ua.pt/handle/10773/15884 (accessed on 21 July 2022).
  24. Nova, C.V.; Reis, K.A.; Pinheiro, A.L.; Dalmaschio, C.J.; Chiquito, A.J.; Teodoro, M.D.; Rodrigues, A.D.; Longo, E.; Pontes, F.M. Synthesis, characterization, photocatalytic, and antimicrobial activity of ZrO2 nanoparticles and Ag@ZrO2 nanocomposite prepared by the advanced oxidative process/hydrothermal route. J. Sol-Gel Sci. Technol. 2021, 98, 113–126. [Google Scholar] [CrossRef]
  25. Stolzenburg, P.; Freytag, A.; Bigall, N.C.; Garnweitner, G. Fractal growth of ZrO2 nanoparticles induced by synthesis conditions. CrystEngComm 2016, 18, 8396–8405. [Google Scholar] [CrossRef]
  26. Rafiq Hussain Siddiqui, M.; Ibrahim Al-Wassil, A.; Mohmmed Al-Otaibi, A.; Mohamad Mahfouz, R. Effects of Precursor on the Morphology and Size of ZrO2 Nanoparticles, Synthesized by Sol-gel Method in Non-aqueous Medium. Mater. Res. 2012, 15, 986–989. [Google Scholar] [CrossRef]
  27. Pang, Q.; Chen, H.; Wang, X.; Wang, T.; Wang, D.; Feng, S.; Lu, H.; Li, Q. Field Effect Transistors Based on In Situ Fabricated Graphene Scaffold–ZrO2 Nanofilms. Adv. Electron. Mater. 2018, 4, 1–6. [Google Scholar] [CrossRef]
  28. Sani, E.; Sciti, D.; Capiani, C.; Silvestroni, L. Colored zirconia with high absorbance and solar selectivity. Scr. Mater. 2020, 186, 147–151. [Google Scholar] [CrossRef]
  29. Sani, E.; Mercatelli, L.; Sans, J.L.; Sciti, D. Optical properties of black and white ZrO2 for solar receiver applications. Sol. Energy Mater. Sol. Cells 2015, 140, 477–482. [Google Scholar] [CrossRef]
  30. Silvestroni, L.; Capiani, C.; Sciti, D.; Sani, E. Coloring zirconium oxide for novel energy saving industrial applications. Renew. Energy 2022, 190, 223–231. [Google Scholar] [CrossRef]
  31. Silva, J.P.B.; Silva, J.M.B.; Sekhar, K.C.; Palneedi, H.; Istrate, M.C.; Negrea, R.F.; Ghica, C.; Chahboun, A.; Pereira, M.; Gomes, M.J.M. Energy storage performance of ferroelectric ZrO2 film capacitors: Effect of HfO2:Al2O3 dielectric insert layer. J. Mater. Chem. A 2020, 8, 14171–14177. [Google Scholar] [CrossRef]
  32. Pešić, M.; Hoffmann, M.; Richter, C.; Mikolajick, T.; Schroeder, U. Nonvolatile Random Access Memory and Energy Storage Based on Antiferroelectric Like Hysteresis in ZrO2. Adv. Funct. Mater. 2016, 26, 7486–7494. [Google Scholar] [CrossRef]
  33. Wang, Y.; Wang, Y.; Zeng, H.; Wei, X. Ultra-high energy storage density of transparent capacitors based on linear dielectric ZrO2 thin films with the thickness scaled up to hundreds nanometers. Appl. Phys. Lett. 2022, 120, 023904. [Google Scholar] [CrossRef]
  34. Nunes, D.; Vilarigues, M.; Correia, J.B.; Carvalho, P.A. Nickel–carbon nanocomposites: Synthesis, structural changes and strengthening mechanisms. Acta Mater. 2012, 60, 737–747. [Google Scholar] [CrossRef] [Green Version]
  35. Vidya, Y.S.; Anantharaju, K.S.; Nagabhushana, H.; Sharma, S.C.; Nagaswarupa, H.P.; Prashantha, S.C.; Shivakumara, C. Danithkumar Combustion synthesized tetragonal ZrO2: Eu3+ nanophosphors: Structural and photoluminescence studies. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 135, 241–251. [Google Scholar] [CrossRef] [PubMed]
  36. Bumajdad, A.; Nazeer, A.A.; Al Sagheer, F.; Nahar, S.; Zaki, M.I. Controlled Synthesis of ZrO2 Nanoparticles with Tailored Size, Morphology and Crystal Phases via Organic/Inorganic Hybrid Films. Sci. Rep. 2018, 8, 3695. [Google Scholar] [CrossRef] [PubMed]
  37. Jiang, C.; Wang, F.; Wu, N.; Liu, X. Up- and down-conversion cubic zirconia and hafnia nanobelts. Adv. Mater. 2008, 20, 4826–4829. [Google Scholar] [CrossRef]
  38. Cao, H.; Qiu, X.; Luo, B.; Liang, Y.; Zhang, Y.; Tan, R.; Zhao, M.; Zhu, Q. Synthesis and Room-Temperature Ultraviolet Photoluminescence Properties of Zirconia Nanowires. Adv. Funct. Mater. 2004, 14, 1–4. [Google Scholar] [CrossRef]
  39. Stępień, M.; Handzlik, P.; Fitzner, K. Synthesis of ZrO2 nanotubes in inorganic and organic electrolytes by anodic oxidation of zirconium. J. Solid State Electrochem. 2014, 18, 3081–3090. [Google Scholar] [CrossRef]
  40. Mangla, O.; Roy, S. Monoclinic Zirconium Oxide Nanostructures Having Tunable Band Gap Synthesized under Extremely Non-Equilibrium Plasma Conditions. Proceedings 2019, 3, 10. [Google Scholar] [CrossRef]
  41. Yue, M.; Cui, M.; Zhang, N.; Long, Z.; Huang, X. Characterization of CeO2-ZrO2 mixed oxides prepared by two different co-precipitation methods. J. Rare Earths 2013, 31, 251–256. [Google Scholar] [CrossRef]
  42. Lim, H.S.; Ahmad, A.; Hamzah, H. Synthesis of zirconium oxide nanoparticle by sol-gel technique. AIP Conf. Proc. 2013, 1571, 812. [Google Scholar] [CrossRef]
  43. Zinatloo-Ajabshir, S.; Salavati-Niasari, M. Synthesis of pure nanocrystalline ZrO2 via a simple sonochemical-assisted route. J. Ind. Eng. Chem. 2014, 20, 3313–3319. [Google Scholar] [CrossRef]
  44. Liang, J.; Jiang, X.; Liu, G.; Deng, Z.; Zhuang, J.; Li, F.; Li, Y. Characterization and synthesis of pure ZrO2 nanopowders via sonochemical method. Mater. Res. Bull. 2003, 38, 161–168. [Google Scholar] [CrossRef]
  45. Beer, S.M.J.; Samelor, D.; Abdel Aal, A.; Etzkorn, J.; Rogalla, D.; Turgambaeva, A.E.; Esvan, J.; Kostka, A.; Vahlas, C.; Devi, A. Direct liquid injection chemical vapor deposition of ZrO2 films from a heteroleptic Zr precursor: Interplay between film characteristics and corrosion protection of stainless steel. J. Mater. Res. Technol. 2021, 13, 1599–1614. [Google Scholar] [CrossRef]
  46. Chang, Y.; Li, X. Bin Preparation of ZrO2 spherical nanometer powders by emulsion processing route. Trans. Nonferrous Met. Soc. China 2006, 16, s332–s336. [Google Scholar] [CrossRef]
  47. Lee, M.H.; Tai, C.Y.; Lu, C.H. Synthesis of spherical zirconia by reverse emulsion precipitation. Korean J. Chem. Eng. 1999, 16, 818–822. [Google Scholar] [CrossRef]
  48. Zhao, N.; Pan, D.; Nie, W.; Ji, X. Two-phase synthesis of shape-controlled colloidal zirconia nanocrystals and their characterization. J. Am. Chem. Soc. 2006, 128, 10118–10124. [Google Scholar] [CrossRef]
  49. Guo, M.; Wang, G.; Zhao, Y.; Li, H.; Tang, K.; Zhao, Y.; Burgess, K. Preparation of Nano-ZrO2 powder via a microwave-assisted hydrothermal method. Ceram. Int. 2021, 47, 12425–12432. [Google Scholar] [CrossRef]
  50. Gonell, F.; Portehault, D.; Julián-López, B.; Vallé, K.; Sanchez, C.; Corma, A. One step microwave-assisted synthesis of nanocrystalline WOx–ZrO2 acid catalysts. Catal. Sci. Technol. 2016, 6, 8257–8267. [Google Scholar] [CrossRef]
  51. Bondioli, F.; Bonamartini Corradi, A.; Ferrari, A.M.; Leonelli, C. Synthesis of Zirconia Nanoparticles in a Continuous-Flow Microwave Reactor. J. Am. Ceram. Soc. 2008, 91, 3746–3748. [Google Scholar] [CrossRef]
  52. Kim, J.R.; Myeong, W.J.; Ihm, S.K. Characteristics of CeO2–ZrO2 mixed oxide prepared by continuous hydrothermal synthesis in supercritical water as support of Rh catalyst for catalytic reduction of NO by CO. J. Catal. 2009, 263, 123–133. [Google Scholar] [CrossRef]
  53. Li, Q.; Liu, L.; Wang, Z.; Wang, X. Continuous Hydrothermal Flow Synthesis and Characterization of ZrO2 Nanoparticles Doped with CeO2 in Supercritical Water. Nanomaterials 2022, 12, 668. [Google Scholar] [CrossRef]
  54. Hobbs, H.; Briddon, S.; Lester, E. The synthesis and fluorescent properties of nanoparticulate ZrO2 doped with Eu using continuous hydrothermal synthesis. Green Chem. 2009, 11, 484–491. [Google Scholar] [CrossRef]
  55. Dudnik, E.V. Modern methods for hydrothermal synthesis of ZrO2-based nanocrystalline powders. Powder Metall. Met. Ceram. 2009, 48, 238–248. [Google Scholar] [CrossRef]
  56. Liu, L.; Wang, S.; Zhang, B.; Jiang, G.; Yang, J. Supercritical hydrothermal synthesis of nano-ZrO2: Influence of technological parameters and mechanism. J. Alloy. Compd. 2022, 898, 16287. [Google Scholar] [CrossRef]
  57. Darr, J.A.; Zhang, J.; Makwana, N.M.; Weng, X. Continuous Hydrothermal Synthesis of Inorganic Nanoparticles: Applications and Future Directions. Chem. Rev. 2017, 117, 11125–11238. [Google Scholar] [CrossRef]
  58. Ponkumar, S.; Janaki, K.; Prakashbabu, D.; B Ramalingam, H.; Munirathnam, K.; Hari Krishna, R. Solution Combustion Synthesis of ZrO2:Tb3+ Nanophosphors Viable for WLEDs. Mater. Today Proc. 2018, 5, 10717–10721. [Google Scholar] [CrossRef]
  59. Samantaray, S.; Mishra, B.G.; Pradhan, D.K.; Hota, G. Solution combustion synthesis and physicochemical characterization of ZrO2-MoO3 nanocomposite oxides prepared using different fuels. Ceram. Int. 2011, 37, 3101–3108. [Google Scholar] [CrossRef]
  60. Purohit, R.D.; Saha, S.; Tyagi, A.K. Combustion synthesis of nanocrystalline ZrO2 powder: XRD, Raman spectroscopy and TEM studies. Mater. Sci. Eng. B Solid-State Mater. Adv. Technol. 2006, 130, 57–60. [Google Scholar] [CrossRef]
  61. Carlos, E.; Martins, R.; Fortunato, E.; Branquinho, R. Solution Combustion Synthesis: Towards a Sustainable Approach for Metal Oxides. Chem.—A Eur. J. 2020, 26, 9099–9125. [Google Scholar] [CrossRef]
  62. Branquinho, R.; Santa, A.; Carlos, E.; Salgueiro, D.; Barquinha, P.; Martins, R.; Fortunato, E. Solution Combustion Synthesis: Applications in Oxide Electronics. In Developments in Combustion Technology; IntechOpen: London, UK, 2016; ISBN 978-953-51-2669-0. [Google Scholar]
  63. Moreira, M.; Carlos, E.; Dias, C.; Deuermeier, J.; Pereira, M.; Barquinha, P.; Branquinho, R.; Martins, R.; Fortunato, E. Tailoring IGZO Composition for Enhanced Fully Solution-Based Thin Film Transistors. Nanomaterials 2019, 9, 1273. [Google Scholar] [CrossRef]
  64. Merten, D.; Broekaert, J.A.C.; Brandt, R.; Jakubowski, N. Analysis of ZrO2 powders by microwave assisted digestion at high pressure and ICP atomic spectrometry. J. Anal. At. Spectrom. 1999, 14, 1093–1098. [Google Scholar] [CrossRef]
  65. Iqbal, Z.; Sadiq, S.; Sadiq, M.; Khan, I.; Saeed, K. Effect of Microwave Irradiation on the Catalytic Activity of Tetragonal Zirconia: Selective Hydrogenation of Aldehyde. Arab. J. Sci. Eng. 2021, 47, 5841–5848. [Google Scholar] [CrossRef]
  66. Srikalyani, C.H.; Sultana, S. Microwave Synthesis of ZrO2 Nanoparticles and its Reinforcement in Geo-Polymer gel. Int. J. ChemTech Res. 2017, 10, 1–6. [Google Scholar]
  67. Goharshadi, E.K.; Hadadian, M. Effect of calcination temperature on structural, vibrational, optical, and rheological properties of zirconia nanoparticles. Ceram. Int. 2012, 38, 1771–1777. [Google Scholar] [CrossRef]
  68. Padovini, D.S.S.; Pontes, D.S.L.; Dalmaschio, C.J.; Pontes, F.M.; Longo, E. Facile synthesis and characterization of ZrO2 nanoparticles prepared by the AOP/hydrothermal route. RSC Adv. 2014, 4, 38484–38490. [Google Scholar] [CrossRef]
  69. Liu, G.X.; Liu, A.; Meng, Y.; Shan, F.K.; Shin, B.C.; Lee, W.J.; Cho, C.R. Annealing dependence of solution-processed ultra-thin ZrOx films for gate dielectric applications. J. Nanosci. Nanotechnol. 2015, 15, 2185–2191. [Google Scholar] [CrossRef]
  70. Valle, H.R.L. Effect of Eco-Friendly Solvents in Solution-Based ZrOx Dielectrics. Master’s Thesis, FCT-UNL, Caparica, Portugal, 2019. Available online: https://run.unl.pt/handle/10362/91651 (accessed on 21 July 2022).
  71. Seon, J.B.; Cho, N.K.; Yoo, G.; Kim, Y.S.; Char, K. Solution-processed amorphous ZrO2 gate dielectric films synthesized by a non-hydrolytic sol–gel route. RSC Adv. 2018, 8, 39115–39119. [Google Scholar] [CrossRef]
  72. Gong, Y.; Zhao, K.; He, H.; Cai, W.; Tang, N.; Ning, H.; Wu, S.; Gao, J.; Zhou, G.; Lu, X.; et al. Solution processable high quality ZrO2 dielectric films for low operation voltage and flexible organic thin film transistor applications. J. Phys. D Appl. Phys. 2018, 51, 115105. [Google Scholar] [CrossRef]
  73. Wang, S.; Xia, G. A facile low-cost preparation of high-k ZrO2 dielectric films for superior thin-film transistors. Ceram. Int. 2019, 45, 23666–23672. [Google Scholar] [CrossRef]
  74. Jung, S.H.; Han, H.S.; Kim, Y.B.; Kim, D.S.; Deshpande, N.G.; Oh, S.J.; Choi, J.H.; Cho, H.K. Toward ultraviolet solution processed ZrOx/IZO transistors with top-gate and dual-gate operation: Selection of solvents, precursors, stabilizers, and additive elements. J. Alloy. Compd. 2020, 847, 156431. [Google Scholar] [CrossRef]
  75. Luo, C.; Huang, T.; Li, C.; Zhang, Y.; Zou, Z.; Li, Y.; Tao, R.; Gao, J.; Zhou, G.; Lu, X.; et al. Enhancement of electrical properties of solution-processed oxide thin film transistors using ZrO2 gate dielectrics deposited by an oxygen-doped solution. J. Phys. D Appl. Phys. 2021, 54, 125101. [Google Scholar] [CrossRef]
  76. Aminipoya, H.; Ghomi, A.B.; Niazi, A. Comparative synthesis of ZrO2 nanoparticles by green and co-precipitation methods: The effect of template on structure. Int. J. Nano Dimens 2021, 12, 59–66. [Google Scholar]
  77. Ji, P.; Wang, Z.; Shang, X.; Zhang, Y.; Liu, Y.; Mao, Z.; Shi, X. Direct observation of enhanced Raman scattering on nano-sized ZrO2 substrate: Charge-transfer contribution. Front. Chem. 2019, 7, 245. [Google Scholar] [CrossRef] [PubMed]
  78. Soares, M.R.N.; Nico, C.; Peres, M.; Ferreira, N.; Fernandes, A.J.S.; Monteiro, T.; Costa, F.M. Structural and optical properties of europium doped zirconia single crystals fibers grown by laser floating zone. J. Appl. Phys. 2011, 109, 013516. [Google Scholar] [CrossRef] [Green Version]
  79. Merle, T.; Guinebretiere, R.; Mirgorodsky, A.; Quintard, P. Polarized Raman spectra of tetragonal pure ZrO2 measured on epitaxial films. Phys. Rev. B—Condens. Matter Mater. Phys. 2002, 65, 1443021–1443026. [Google Scholar] [CrossRef]
  80. Fernandez Lopez, E.; Sanchez Escribano, V.; Panizza, M.; Carnasciali, M.M.; Busca, G. Vibrational and electronic spectroscopic properties of zirconia powders. J. Mater. Chem. 2001, 11, 1891–1897. [Google Scholar] [CrossRef]
  81. Ding, S.; Zhao, J.; Yu, Q. Effect of zirconia polymorph on vapor-phase ketonization of propionic acid. Catalysts 2019, 9, 768. [Google Scholar] [CrossRef]
  82. Gopal, R.; Jain, J.; Goyal, A.; Gupta, D.K.; Nagar, M. Formation of nano-sized cubic zirconia by aqueous sol–gel route. J. Aust. Ceram. Soc. 2018, 54, 691–700. [Google Scholar] [CrossRef]
  83. Basahel, S.N.; Ali, T.T.; Mokhtar, M.; Narasimharao, K. Influence of crystal structure of nanosized ZrO2 on photocatalytic degradation of methyl orange. Nanoscale Res. Lett. 2015, 10, 73. [Google Scholar] [CrossRef]
  84. Lopes, D. Low-Cost Upconversion Nanomaterials for Anti-Counterfeiting Solutions. Master’s Thesis, FCT-UNL, Caparica, Portugal, 2020. Available online: https://run.unl.pt/handle/10362/113701 (accessed on 22 July 2022).
  85. Prasad, K.; Pinjari, D.V.; Pandit, A.B.; Mhaske, S.T. Synthesis of zirconium dioxide by ultrasound assisted precipitation: Effect of calcination temperature. Ultrason. Sonochemistry 2011, 18, 1128–1137. [Google Scholar] [CrossRef]
  86. Carvalho, J.M.; Rodrigues, L.C.V.; Felinto, M.C.F.C.; Nunes, L.A.O.; Hölsä, J.; Brito, H.F. Structure–property relationship of luminescent zirconia nanomaterials obtained by sol–gel method. J. Mater. Sci. 2015, 50, 873–881. [Google Scholar] [CrossRef]
  87. Kiisk, V.; Puust, L.; Utt, K.; Maaroos, A.; Mändar, H.; Viviani, E.; Piccinelli, F.; Saar, R.; Joost, U.; Sildos, I. Photo-, thermo- and optically stimulated luminescence of monoclinic zirconia. J. Lumin. 2016, 174, 49–55. [Google Scholar] [CrossRef]
  88. Lokesha, H.S.; Nagabhushana, K.R.; Singh, F.; Tatumi, S.H.; Prinsloo, A.R.E.; Sheppard, C.J. Unraveling the Charge State of Oxygen Vacancies in Monoclinic ZrO2 and Spectroscopic Properties of ZrO2:Sm 3+ Phosphor. J. Phys. Chem. C 2021, 125, 27106–27117. [Google Scholar] [CrossRef]
  89. Perevalov, T.V.; Gulyaev, D.V.; Aliev, V.S.; Zhuravlev, K.S.; Gritsenko, V.A.; Yelisseyev, A.P. The origin of 2.7 eV blue luminescence band in zirconium oxide. J. Appl. Phys. 2014, 116, 244109. [Google Scholar] [CrossRef]
  90. Ashraf, S.; Irfan, M.; Kim, D.; Jang, J.-H.; Han, W.-T.; Jho, Y.-D. Optical influence of annealing in nano- and submicron-scale ZrO2 powders. Ceram. Int. 2014, 40, 8513–8518. [Google Scholar] [CrossRef]
  91. Jiang, H.; Gomez-Abal, R.I.; Rinke, P.; Scheffler, M. Electronic band structure of zirconia and hafnia polymorphs from the GW perspective. Phys. Rev. B 2010, 81, 085119. [Google Scholar] [CrossRef]
  92. Gallino, F.; Di Valentin, C.; Pacchioni, G. Band gap engineering of bulk ZrO2 by Ti doping. Phys. Chem. Chem. Phys. 2011, 13, 17667–17675. [Google Scholar] [CrossRef]
  93. Li, J.; Meng, S.; Niu, J.; Lu, H. Electronic structures and optical properties of monoclinic ZrO2 studied by first-principles local density approximation + U approach. J. Adv. Ceram. 2017, 6, 43–49. [Google Scholar] [CrossRef]
  94. Singh, H.; Sunaina; Yadav, K.K.; Bajpai, V.K.; Jha, M. Tuning the bandgap of m-ZrO2 by incorporation of copper nanoparticles into visible region for the treatment of organic pollutants. Mater. Res. Bull. 2020, 123, 110698. [Google Scholar] [CrossRef]
  95. Teeparthi, S.R.; Awin, E.W.; Kumar, R. Dominating role of crystal structure over defect chemistry in black and white zirconia on visible light photocatalytic activity. Sci. Rep. 2018, 8, 5541. [Google Scholar] [CrossRef]
  96. Sakfali, J.; Ben Chaabene, S.; Akkari, R.; Dappozze, F.; Berhault, G.; Guillard, C.; Saïd Zina, M. High photocatalytic activity of aerogel tetragonal and monoclinic ZrO2 samples. J. Photochem. Photobiol. A Chem. 2022, 430, 113970. [Google Scholar] [CrossRef]
  97. Smits, K.; Grigorjeva, L.; Millers, D.; Sarakovskis, A.; Grabis, J.; Lojkowski, W. Intrinsic defect related luminescence in ZrO2. J. Lumin. 2011, 131, 2058–2062. [Google Scholar] [CrossRef]
  98. Sil, A.; Goldfine, E.A.; Huang, W.; Bedzyk, M.J.; Medvedeva, J.E.; Facchetti, A. Role of Fluoride Doping in Low-Temperature Combustion-Synthesized ZrOx Dielectric Films. ACS Appl. Mater. Interfaces 2022, 14, 12340–12349. [Google Scholar] [CrossRef]
  99. Pradhan, S.; Mishra, B.G. Catalytic application of SO42-/Fe-ZrO2 nanoparticles synthesized by a urea hydrolysis method for environmentally benign one pot synthesis of 1,8-dioxodecahydroacridines. RSC Adv. 2015, 5, 86179–86190. [Google Scholar] [CrossRef]
  100. Cai, W.; Zhu, Z.; Wei, J.; Fang, Z.; Ning, H.; Zheng, Z.; Zhou, S.; Yao, R.; Peng, J.; Lu, X. A simple method for high-performance, solution-processed, amorphous ZrO2 gate insulator TFT with a high concentration precursor. Materials 2017, 10, 972. [Google Scholar] [CrossRef] [PubMed]
  101. Meng, Y.; Liu, G.; Liu, A.; Song, H.; Hou, Y.; Shin, B.; Shan, F. Low-temperature fabrication of high performance indium oxide thin film transistors. RSC Adv. 2015, 5, 37807–37813. [Google Scholar] [CrossRef]
  102. Zhu, L.; He, G.; Lv, J.; Fortunato, E.; Martins, R. Fully solution-induced high performance indium oxide thin film transistors with ZrOx high-k gate dielectrics. RSC Adv. 2018, 8, 16788–16799. [Google Scholar] [CrossRef]
  103. Lee, H.; Zhang, X.; Kim, J.W.; Kim, E.J.; Park, J. Investigation of the Electrical Characteristics of Bilayer ZnO/In2O3 Thin-Film Transistors Fabricated by Solution Processing. Materials 2018, 11, 2103. [Google Scholar] [CrossRef]
  104. Oluwabi, A.T.; Katerski, A.; Carlos, E.; Branquinho, R.; Mere, A.; Krunks, M.; Fortunato, E.; Pereira, L.; Oja Acik, I. Application of ultrasonic sprayed zirconium oxide dielectric in zinc tin oxide-based thin film transistor. J. Mater. Chem. C 2020, 8, 3730–3739. [Google Scholar] [CrossRef]
  105. Zhu, C.; Liu, A.; Liu, G.; Jiang, G.; Meng, Y.; Fortunato, E.; Martins, R.; Shan, F. Low-temperature, nontoxic water-induced high-k zirconium oxide dielectrics for low-voltage, high-performance oxide thin-film transistors. J. Mater. Chem. C 2016, 4, 10715–10721. [Google Scholar] [CrossRef]
  106. Boratto, M.H.; Congiu, M.; dos Santos, S.B.O.; Scalvi, L.V.A. Annealing temperature influence on sol-gel processed zirconium oxide thin films for electronic applications. Ceram. Int. 2018, 44, 10790–10796. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Steps to fabricate the MIS capacitor after the cleaning procedure: (a) surface treatment on silicon (Si) substrate by UV-ozone for 15 min; (b) deposition of the zirconium oxide precursor solution by spin coating technique; (c) annealing treatment of the substrate for film densification; (d) scheme of the device after the deposition of top electrodes (80 nm Al) by thermal evaporation; (e) scheme of the device after the deposition of the bottom electrode (80 nm Al) by thermal evaporation.
Figure 1. Steps to fabricate the MIS capacitor after the cleaning procedure: (a) surface treatment on silicon (Si) substrate by UV-ozone for 15 min; (b) deposition of the zirconium oxide precursor solution by spin coating technique; (c) annealing treatment of the substrate for film densification; (d) scheme of the device after the deposition of top electrodes (80 nm Al) by thermal evaporation; (e) scheme of the device after the deposition of the bottom electrode (80 nm Al) by thermal evaporation.
Energies 15 06452 g001
Figure 2. Raman spectra of the synthesized ZrO2 nanopowders: (a) before and (b) after calcination at 800 °C for 15 min. The black and orange vertical dash lines represent the expected frequencies for the monoclinic and tetragonal ZrO2, respectively, and according to [77,79,81].
Figure 2. Raman spectra of the synthesized ZrO2 nanopowders: (a) before and (b) after calcination at 800 °C for 15 min. The black and orange vertical dash lines represent the expected frequencies for the monoclinic and tetragonal ZrO2, respectively, and according to [77,79,81].
Energies 15 06452 g002
Figure 3. (a) In situ XRD diffractograms of the ZrO2 nanopowder synthesized under microwave irradiation with different annealing temperatures ranging from RT to 800 °C with a plateau time of 15 min. For comparison, the simulated monoclinic, tetragonal and cubic zirconia structures are also presented. (b) XRD diffractogram of ZrO2 nanopowder synthesized under microwave irradiation and calcinated at 800 °C for 15 min. (c) In situ XRD diffractograms of the ZrO2 nanopowder at 800 °C for different annealing times. Green arrows represent an increase of monoclinic phase, while the red arrow indicates a decrease of tetragonal phase.
Figure 3. (a) In situ XRD diffractograms of the ZrO2 nanopowder synthesized under microwave irradiation with different annealing temperatures ranging from RT to 800 °C with a plateau time of 15 min. For comparison, the simulated monoclinic, tetragonal and cubic zirconia structures are also presented. (b) XRD diffractogram of ZrO2 nanopowder synthesized under microwave irradiation and calcinated at 800 °C for 15 min. (c) In situ XRD diffractograms of the ZrO2 nanopowder at 800 °C for different annealing times. Green arrows represent an increase of monoclinic phase, while the red arrow indicates a decrease of tetragonal phase.
Energies 15 06452 g003
Figure 4. (a) SEM image of the ZrO2 nanoparticles synthesized by microwave irradiation with a further calcination step at 800 °C for 15 min, together with the corresponding EDS maps of Zr (b) and O (c).
Figure 4. (a) SEM image of the ZrO2 nanoparticles synthesized by microwave irradiation with a further calcination step at 800 °C for 15 min, together with the corresponding EDS maps of Zr (b) and O (c).
Energies 15 06452 g004
Figure 5. TEM images of the as-synthesized ZrO2 nanoparticles before and after calcination at 800 °C for 15 min. (a) Bright-field TEM image of the ZrO2 nanoparticles before calcination and (b) and (c) high-resolution TEM images of the ZrO2 nanocrystals. (d) Bright-field TEM image of the ZrO2 nanoparticles after calcination and (e,f) high-resolution TEM images of the ZrO2 nanocrystals. The insets in (c,f) show the FFT images of the areas (black squares) indicated as (A,B), respectively.
Figure 5. TEM images of the as-synthesized ZrO2 nanoparticles before and after calcination at 800 °C for 15 min. (a) Bright-field TEM image of the ZrO2 nanoparticles before calcination and (b) and (c) high-resolution TEM images of the ZrO2 nanocrystals. (d) Bright-field TEM image of the ZrO2 nanoparticles after calcination and (e,f) high-resolution TEM images of the ZrO2 nanocrystals. The insets in (c,f) show the FFT images of the areas (black squares) indicated as (A,B), respectively.
Energies 15 06452 g005
Figure 6. PLE/PL spectra acquired at RT on the ZrO2 nanopowders (a) before and (b) after a calcination treatment at 800 °C for 15 min.
Figure 6. PLE/PL spectra acquired at RT on the ZrO2 nanopowders (a) before and (b) after a calcination treatment at 800 °C for 15 min.
Energies 15 06452 g006
Figure 7. XRD diffractogram of ZrO2 powder produced by solution combustion synthesis and annealed in air at 350 °C for 1 h (represented in red). For comparison, the simulated monoclinic, tetragonal and cubic zirconia structures are also presented.
Figure 7. XRD diffractogram of ZrO2 powder produced by solution combustion synthesis and annealed in air at 350 °C for 1 h (represented in red). For comparison, the simulated monoclinic, tetragonal and cubic zirconia structures are also presented.
Energies 15 06452 g007
Figure 8. (a,b) SEM images of the ZrO2 particle produced by the combustion synthesis method and after annealing in air at 350 °C for 1 h, together with the corresponding EDS maps of Zr (c) and O (d).
Figure 8. (a,b) SEM images of the ZrO2 particle produced by the combustion synthesis method and after annealing in air at 350 °C for 1 h, together with the corresponding EDS maps of Zr (c) and O (d).
Energies 15 06452 g008
Figure 9. TEM images of the ZrO2 powder produced with solution combustion synthesis and annealed in air at 350 °C for 1 h. (a) Bright-field TEM image of the ZrO2 nanoparticles and (b) high-resolution TEM images of the ZrO2 nanocrystals. The inset in (b) shows the FFT images of the area (black square) indicated as C.
Figure 9. TEM images of the ZrO2 powder produced with solution combustion synthesis and annealed in air at 350 °C for 1 h. (a) Bright-field TEM image of the ZrO2 nanoparticles and (b) high-resolution TEM images of the ZrO2 nanocrystals. The inset in (b) shows the FFT images of the area (black square) indicated as C.
Energies 15 06452 g009
Figure 10. Grazing incidence XRD diffractogram of the spin-coated ZrOx thin films annealed at 350 °C in air for 30 min and deposited on a silicon substrate.
Figure 10. Grazing incidence XRD diffractogram of the spin-coated ZrOx thin films annealed at 350 °C in air for 30 min and deposited on a silicon substrate.
Energies 15 06452 g010
Figure 11. Electrical characteristics of solution-based ZrOx MIS capacitors produced at 350 °C: (a) typical capacitance-voltage characteristics at 100 kHz; (b) average of capacitance versus frequency characteristics, and the inset shows the cross-section SEM of the devices; (c) typical leakage current density versus the breakdown field.
Figure 11. Electrical characteristics of solution-based ZrOx MIS capacitors produced at 350 °C: (a) typical capacitance-voltage characteristics at 100 kHz; (b) average of capacitance versus frequency characteristics, and the inset shows the cross-section SEM of the devices; (c) typical leakage current density versus the breakdown field.
Energies 15 06452 g011
Table 1. EDS point analysis of the ZrO2 nanopowders before and after the calcination treatment at 800 °C for 15 min.
Table 1. EDS point analysis of the ZrO2 nanopowders before and after the calcination treatment at 800 °C for 15 min.
ZrO2 Nanopowder before the Calcination TreatmentZrO2 Nanopowder after the Calcination Treatment
ElementsZrOZrO
at.%33.366.726.973.1
Table 2. EDS point analysis of the ZrO2 nanopowder produced by the solution combustion synthesis.
Table 2. EDS point analysis of the ZrO2 nanopowder produced by the solution combustion synthesis.
ElementsZrO
at.%31.968.1
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Matias, M.L.; Carlos, E.; Branquinho, R.; do Valle, H.; Marcelino, J.; Morais, M.; Pimentel, A.; Rodrigues, J.; Monteiro, T.; Fortunato, E.; et al. A Comparison between Solution-Based Synthesis Methods of ZrO2 Nanomaterials for Energy Storage Applications. Energies 2022, 15, 6452. https://doi.org/10.3390/en15176452

AMA Style

Matias ML, Carlos E, Branquinho R, do Valle H, Marcelino J, Morais M, Pimentel A, Rodrigues J, Monteiro T, Fortunato E, et al. A Comparison between Solution-Based Synthesis Methods of ZrO2 Nanomaterials for Energy Storage Applications. Energies. 2022; 15(17):6452. https://doi.org/10.3390/en15176452

Chicago/Turabian Style

Matias, Maria Leonor, Emanuel Carlos, Rita Branquinho, Hadassa do Valle, João Marcelino, Maria Morais, Ana Pimentel, Joana Rodrigues, Teresa Monteiro, Elvira Fortunato, and et al. 2022. "A Comparison between Solution-Based Synthesis Methods of ZrO2 Nanomaterials for Energy Storage Applications" Energies 15, no. 17: 6452. https://doi.org/10.3390/en15176452

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