Next Article in Journal
Tetrachlorocobaltate-Catalyzed Methane Oxidation to Methyl Trifluoroacetate
Next Article in Special Issue
Carbon-Protected BiVO4—Cu2O Thin Film Tandem Cell for Solar Water Splitting Applications
Previous Article in Journal
Co(OH)2 Nanoflowers Decorated α-NiMoO4 Nanowires as a Bifunctional Electrocatalyst for Efficient Overall Water Splitting
Previous Article in Special Issue
Single-Atom Transition Metal Photocatalysts for Hydrogen Evolution Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

ZrO2-Based Photocatalysts for Wastewater Treatment: From Novel Modification Strategies to Mechanistic Insights

1
School of Advanced Chemical Sciences, Faculty of Basic Sciences, Shoolini University, Solan 173229, HP, India
2
Department of Organic Chemistry, Bioorganic Chemistry and Biotechnology, Silesian University of Technology, B. Krzywoustego 4, 44-100 Gliwice, Poland
3
Advanced Materials and Processes Research Institute, Hoshangabad Road, Bhopal 462026, MP, India
4
Department of Materials Science and Engineering, Korea University, 145, Anam-ro Seongbuk Gu, Seoul 02841, Korea
5
Faculty of Allied Health Sciences, Chettinad Academy of Research and Education (CARE), Kelambakkam, Chennai 630102, TN, India
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(11), 1418; https://doi.org/10.3390/catal12111418
Submission received: 22 August 2022 / Revised: 30 October 2022 / Accepted: 9 November 2022 / Published: 11 November 2022

Abstract

:
Zirconium dioxide (ZrO2) has garnered substantial research interest in the field of photocatalytic water treatment due to its appealing properties, such as thermal stability, considerable physical strength, and strong chemical resistance. However, the wide bandgap energy endorses less photoabsorption and rapid charge carrier recombination kinetics, thus restricting the photoactivity of ZrO2. Previously, vast research efforts have been made to improve the photoefficacy of ZrO2, and hence it is worth exploring the potential strategic modifications responsible for incremented photocatalytic efficiency. In this regard, the present review article emphasizes the optical, structural, and electronic features of ZrO2, which makes it an interesting photocatalytic material. The exceptional modification strategies that help to modulate the crystal structure, morphology, bandgap energy, and charge carrier kinetics are primarily discussed. The potential synthetic routes involving bottom-up and top-down methods are also outlined for understanding the rationale for incorporating these techniques. Moreover, the photocatalytic performance evaluation was done by investigating the photodegradation kinetics of various organic and inorganic pollutants degradation by ZrO2. Conclusively, in light of research advances involving ZrO2 photocatalyst, this review article may expedite further investigation for enhancing the large-scale photocatalytic applications for environmental and energy concerns.

1. Introduction

The past several decades have witnessed a tremendous rise in industrialization, agrochemical processes, and excessive exploitation of non-renewable resources, resulting in extreme deterioration of water resources [1,2]. Owing to the exclusion of these multifarious exactor pollutants from contaminated water and to abate their dangerous consequences on the ecosystem, several conventional methods for eliminating organic and inorganic pollutants have been used, including biodegradation, adsorption, coagulation, flocculation, chlorination, reverse osmosis, chemical precipitation, ion exchange, etc. [3,4]. However, these are not proven to be effective enough to eradicate these recalcitrant compounds because of their unreasonable resistance and less biodegradability [5,6,7]. Further, these approaches have several drawbacks, including high maintenance costs, significant sludge formation, limited scale application, excessive energy consumption, long processing period, and incomplete degradation of persistent pollutants [8,9,10]. Currently, numerous research efforts have addressed advanced oxidation processes (AOPs) because of their ability to generate reactive oxidative species for the effective and rapid degradation of these toxic industrial effluents to safe biodegradable mineralized salts, CO2, and H2O [11,12,13]. This special class of oxidation methods is environmentally benign, sustainable, and operates at feasible temperature and atmospheric pressure [14,15,16].
Photocatalysis is an interesting AOP that has been given substantial attention due to its features like effective degradation of pollutants without generating harmful by-products, cost-effectiveness, high efficiency, and the ability to operate under the illumination of solar light [17,18]. Heterogeneous photocatalysis is a photo redox reaction that proceeds with the harnessing of light by a semiconductor, specifically including the generation of e/h+ pairs, capturing of charges at the surface, and inter-interface charge transfer [19,20]. Typically, in photocatalysis, the initiating event is the photoexcitation of electron-hole pairs (EHP) via absorption of the radiation within the bandgap energy. The photoinduced electrons and holes reduce O2 into the superoxide radical •O2 (−0.33 V vs. SHE at pH = 7) and oxidize H2O and OH into •OH (2.8 V vs. SCE at pH = 0) radicals for photo-oxidation reactions, respectively. [14]. The hydroxyl radicals have a short life span due to a single unpaired electron which readily attacks a series of organic wastes (e.g., dyes, insecticides, pesticides, antibiotics, etc.) that are often hard to degrade, showing rate constants typically in the order of 106–109 M−1s−1 [17,21].
Water treatment by photocatalysis using inorganic semiconductors like TiO2, SnO2, CuO, and ZnO, has been vastly explored owing to its non-toxic nature, optical properties, and reasonable cost [22]. For instance, Fujishima and Honda, in 1972, reported the photoelectrolysis of water using TiO2, because of its cost-effectiveness, chemical and photochemical stability, and non-toxic nature [23]. Currently, ZrO2 as broad bandgap (3.25–5.1 eV) semiconductor has been widely explored due to the presence of a large number of oxygen vacancies on the surface, significant ion exchange ability and physiochemical stability [23]. However, ZrO2 possesses a high value of the dielectric constant, tensile strength, melting and boiling point, however its wide band gap and low separation rate of photogenerated electron and hole hinder the wide-scale applications [24]. To overcome this, various enhancement strategies like doping, heterojunction formation incorporation of cocatalysts, etc., have been potentially explored. For instance, Co doped ZrO2 synthesized by facile-chemical precipitation showed greater degrading activity against MB dye when exposed to visible light due to the improved dielectric properties [25]. However, coupling ZrO2 with narrow band semiconductors like MoS2, TiO2, and g-C3N4 also retards the photocarrier reassembly and augments its photocatalytic performance [26]. The optical properties of ZrO2 can be enhanced by introducing new electronic states by doping with rare earth metals. For example, the introduction of Ce (IV) ions in ZrO2 enhanced the photoresponse through the formation of midgap states and the Ce 4f empty states acted as a electron transfer bridge between VB and CB of ZrO2 by low energy photons [27].
Recent research has manifested that zirconia-based materials show high redox abilities, large surface area, and low thermal conductivity in photocatalytic processes. In this regard, Aldeen et al. comprehensively explored the altered ZrO2 based photocatalysts for organic pollutant degradation [24]. However, the typical insights of inorganic pollutant degradation, specifically the heavy metal ions present in waste water, are missing. Additionally, the importance of generating oxygen vacancies (OVs) in ZrO2 for the synergistic improvement in opto-electronic properties still needs further comprehension. Hence, the present review article highlights the features of ZrO2 semiconducting material for the photocatalytic removal of organic as well as inorganic pollutants present in wastewater. Starting from the optimal crystal structure of ZrO2, various phases and the subsequent photocatalytic features are comprehensively reviewed. Significant methods to circumvent the inherent drawbacks of the bare ZrO2 photocatalyst have been inclusively illustrated using different methods such as doping, heterojunction formation, morphology, and structural modulation along with the formation of OVs. After this, the review introduces facile synthesis routes involving bottom-up (solvothermal, Morphology control, and doping) and top-down methods (ultrasonication and chemical reduction) to design ZrO2-based photocatalytic materials. Then, in order to highlight the advancement of ZrO2 as a photocatalyst, the mechanistic insights of wastewater treatment via photocatalysis of organic and inorganic pollutants have been broadly reviewed. Lastly, the review summarizes various challenges in the field with a broad perspective to scale up the photocatalytic applications of ZrO2 with the purpose of solving the burgeoning environmental crisis in the world.

2. ZrO2 as Photocatalyst

ZrO2 is an n-type semiconducting material that possesses a bandgap energy of 5.0 eV, which facilitates its application in the heterogeneous catalysis [28,29]. Moreover, the tremendous properties of ZrO2, like thermal stability, high strength, strong chemical resistance, ductility, and low thermal conductivity makes it an alluring photocatalytic material [25]. Consequently, ZrO2 has been explored for different applications, such as oxygen sensors, power device electrolytes, colors for sharpened sunlight-based cells, the photocatalytic parting of water for hydrogen creation, the particular amalgamation of natural mixtures, air refinement, and photocatalytic abatement of organic and inorganic pollutants present in wastewater [30]. Certain other features of ZrO2, which endorse its usage as photocatalytic material, are described as follows:
  • ZrO2 is a transition metal oxide and a prominent candidate for photocatalysis owing to its high refractive index, high corrosion resistance, and good photochemical stability, as well as optical absorption ability [31,32].
  • ZrO2 is rich in OVs, which exist on the surface or in the lattice’s interstitial sites, which further promotes the optoelectronic properties of ZrO2 leading to boosted light harvesting ability and superior isolation of photocarriers [22].
  • The optical characteristics of ZrO2 are modified by the introduction of midgap states in the elemental doping, resulting in optimal visible light responsive photocatalyst [22]. By doping with ZrO2, ions diffuse throughout the lattice at interstitial sites, resulting in a new valance band maximum (VBM) or conduction band minimum (CBM), which narrows the optical bandgap of ZrO2 [33].

2.1. Crystal Structure

The three major polymorphs of ZrO2, namely monoclinic, tetragonal, and cubic crystal structures are depicted in Figure 1. At ambient temperature, the monoclinic stage remains constant, but when heated to 1170 °C, it transforms to the tetragonal ZrO2 and further to cubic ZrO2 by increasing the temperature up to 2370 °C. The transformation is reversible when the temperature is lowered; the shift from tetragonal to monoclinic occurs at a lower temperature (around 950 °C) [34]. Tetragonal ZrO2 is formed by the elongation of any of the three equal axes of cubic fluorite-type structure and the relocation of oxygen anions from their characteristic places in the same direction [35].
The main features of monoclinic ZrO2 involve sevenfold coordination of the Zr-atom and alteration of fluorite layers possessing OII ions in the tetrahedral coordination with triangular coordinated layers containing OI. Additionally, the nearest Zr-O distance is observed to be 3.77 Å, which signifies the coordination number to be seven (Figure 1a). The Zr-OII bond length in the OII coordinated tetrahedron varies from 2.16 to 2.26 Å, while the central bond angles at OI range from 100 to 132°. The Zr-OI bond length in the OI coordinated triangle is 2.04 Å, 2.10 Å, and 2.15 Å with 2.10 Å as an average angle, while the central angles at OI are 106°, 109°, and 145° [36]. The surface site of m-ZrO2 consists of oxygen vacancies, unsaturated surface hydroxyl groups (Zr-O pairs), and Lewis’s acid sites (Zr3+, Zr4+). In comparison to other phases, monoclinic-ZrO2 has a less symmetrical lattice structure, making it a more adaptable surface for catalytic activity [37].
In the case of tetragonal ZrO2, the Zr-atoms are surrounded by oxygen atoms (eight), four of which are in a flattened tetrahedron at a distance of 2.065 Å and the remaining four at a distance of 2.455 Å in an extended tetrahedron (Figure 1b). Each oxygen atom has two oxygen neighbors at 2.635 Å and four at 2.655 Å, which are connected to two Zr-atoms at 2.065 Å, and two more Zr-atoms at 2.455 Å [38]. The tetragonal phase is more active than the other two phases due to the presence of defects and its stable nature. t-ZrO2 is excellent for photocatalytic reactions because of its antioxidant activity for free radical trapping, high redox ability, chemical and optical stability, and environmental affinity [25]. The cubic phase is thermally stable above 2300 °C, so it is rarely produced at low temperatures (Figure 1c). Yttrium doping was used to successfully construct cubic ZrO2 structures with abundant OVs, resulting in cubic yttria-stabilized zirconia [39,40].

2.2. Bandgap Formation

Compared to the normal hydrogen electrode (NHE), ZrO2 possesses a strong reduction ability due to its comparatively high CBM, which is roughly 2.1 V less than the O2/•O2 redox potential, yet its VBM is more positive (approximately 0.7 V) than the H2O/•OH = 2.72 V (vs. NHE) potential [28,41]. The VB edge of ZrO2 fundamentally arises from the O 2p orbitals, whereas the CB edge is overwhelmed by the 4d orbitals of Zr. The VB potential is 2.77 V higher than the required potential for water oxidation, whereas the CB edge is 1.0 V lower in potential than the hydrogen reduction level [42]. The lowest CB potential is ca. 1.0 V (vs. NHE, pH 0), which is considerably lower than the TiO2 anatase (0.1 V) potential, and the highest VB potential is ca. +4.0 V, which is superior to TiO2 (+3.1 V), suggesting that the band potentials of ZrO2 are highly appropriate for various photocatalytic applications [27].

2.3. Opto-Electronic Properties

Density Functional Theory (DFT) studies can anticipate the characteristics of a material with less inconsistency and estimate bandgap due to the capability to detect quantum mechanical features of atoms, sub-atomic constituents, and inexact depiction of electronic interactions through generalized density approximation (GGA) and local density approximation (LDA), respectively. In the DFT analysis, the unit-cell structures of the oxide are optimized and determined by the lattice parameters for understanding the structural characteristics. Through DFT, it is found that ZrO2 exhibits twelve ionic structures containing four formula units in ZrO2, with seven O atoms surrounding the cation and two oxygen locations OI and OII in Wyckoff points. Each unit cell is thought to be made up of three distinct ions due to the varying atomic locations of the oxygen atom. An average bond length of 2.86 Å and 3.58 Å for Zr-O and Zr-Zr, respectively, was observed [43]. The structure of pure ZrO2 is found to be cubic, which belongs to the space group “Fm3m” (N 225) with zirconium occupied position at (0, 0, 0), and oxygen atoms at (1/4, 1/4, 1/4) and (3/4, 3/4, 3/4), respectively (Figure 2a) [32].
Interestingly, the DFT calculation studies are further helpful to explore the intricate electronic structures of ZrO2. The electrical levels of a crystal structure are presented by the bandgap positioning estimated through the Bloch vector (k) and the band index (n). When Fermi energy is set to the zero level, ZrO2 showed an indirect band gap, with the placement of CBM and VBM at M and N, respectively (Figure 2b). Furthermore, the density of states (DOS) indicates the number of unoccupied levels accessible for the electron to be occupied at a certain energy level. The total density of state (TDOS) and partial density of state (PDOS) charts can be used to show the contribution of distinct atoms. In the case of ZrO2, four different bands are present with the appearance of a lowest band around 27 eV. The next energy band starts from 19 to 16.5 eV with the upper VB positioning in between 5.3–0 eV and the CB at 4.14 eV. In the lowest VB states, Zr 3p states contribute the most while O 2 s state participates in the lower VB formation. In the VB below the Fermi level, O 2p states are the dominant states, with minor assistance from Zr 4d and 5 s hybridized orbitals. The Zr 4d and O 2p orbitals dominate the formation of CB as shown in Figure 2c [43]. The partial DOS (Figure 2d) further represents the characteristics of ionic bonding between the O 2p and Zr 4d orbitals placed below and above the Fermi energy level, respectively.

3. Modification Strategies

ZrO2 is a potential photocatalyst due to its advantages of appropriate redox potentials, exceptional oxygen storage capacity, high thermal and chemical strength, etc. However, the limitations of smaller surface area, restricted visible light captivation, and speedy charge carrier recombination rate reduced the efficiency of the ZrO2 photocatalyst. To overcome these drawbacks, various modification strategies have been incorporated so far, such as doping, heterojunction formation, morphology and structural modulation, and formation of OVs. These strategies are reported to increment the photocatalytic activity of the ZrO2 photocatalyst, as discussed below.

3.1. Morphology and Structural Modulation

Modulation in the morphology and structure is an effective means to enhance the photocatalytic performance of photocatalysts as it can increase the specific area, amend the surface-active sites for photocatalytic reactions, facilitate the mass transportation and light harvesting along with accelerated charge migration and suppressed charge recombination [44]. Generally, factors like pH, temperature, reducing agents, etc., dominantly influence the morphology of photocatalysts and in turn affect the photocatalytic performance. However, the morphology of ZrO2 photocatalyst can also be modified by varying the dopant concentration. In particular, the porosity of ZrO2 nanoparticles can be enhanced by varying the dopant concentration. For instance, the variation in the morphology of ZrO2 photocatalyst was observed by varying the concentration of Mg2+ ion dopant from 0.1 to 5 mol% [45]. From scanning electron microscope (SEM) results (Figure 3a), the hollow sphere structure of the ZrO2 photocatalyst was observed with a large number of cracks at the surface indicating a highly porous nature. The unique interaction between Zr4+ and Mg2+ ions resulted in a complex polysaccharides structure with strong bindings. Furthermore, the increase in the Mg2+ ion concentration in ZrO2 photocatalyst leading to increased light absorption capability due to the development of mid-gap states. Thus, Mg doping in ZrO2 could facilitate the enhancement in light harvesting, efficient inhibition in the reassembly of charge carriers, and incremented specific surface area corresponding to improved photocatalytic performance.
Ahmed et al. also studied the variation in morphology of the ZrO2 photocatalyst after doping with cobalt ions [25]. The SEM results (Figure 3b) reveal that both the un-doped and Co-doped ZrO2 photocatalyst presented spherical structures having identical size and showed ultrafine morphology. Thus, the parameters like shape and size intensities were influenced by the presence of defects that leads to the lowering of electron-hole pairs reassembly which contributed to boosted photocatalytic activity. In a different study, the cube-like or star-like morphology of the ZrO2 colloidal nanocrystal clusters (CNCs) was also observed by varying the reaction conditions [47]. The occurrence of Cl ions in the hydrothermal process resulted in the establishment of cube-like ZrO2 CNCs with a rough surface and increased particle size from 74 to 170 nm by incrementing the reaction time (six to ten hours). Furthermore, the morphology of ZrO2 CNCs changed to star-like structures by the combined action of both Cl and Na+ ions in the hydrothermal process due to its highly crystalline nature and systematic aggregation of nanocrystals. Additionally, the clustering of ZrO2 nanocrystals takes place at high temperatures (150–240 °C), which leads to the formation of a cube-like morphology.

3.2. Bandgap Tuning via Doping

The band gap of semiconductors can be modulated by elemental doping, which triggers the creation of the allowed defective electronic levels within the bandgap energy and promotes the electronic excitation with lesser energy. To date, doping of ZrO2 photocatalyst has been performed with noble metals [48], non-metals [49], and transition elements to adjust its band gap, amend the photoactivity, and inhibit the electron-hole pairs recombination rate [20]. For instance, Co2+/3+ ions with radii 0.74 Å comparable to Zr4+ (0.80 Å) are reported to be an appropriate dopant as it can modify the optical properties and dielectric characteristics of ZrO2 photocatalyst along with a reinforced charge separation rate leading to higher photocatalytic activity [25]. With the varied dopant concentration of Co from 3%, 5%, 7%, to 10%, the band gap energy of ZrO2 photocatalyst changed from 3.7 eV to 2.74, 2.40, 2.23, and 2.12 eV, respectively (Figure 3c), due to the presence of OVs and dislocation defects. However, the optimal Co-doping concentration was found to be 10% in ZrO2 photocatalyst for the highest degradation of 96.5% of methylene blue (MB). Other than the transition metal dopant, boron doping into oxygen vacancy-rich ZrO2 (B@ZrO2-OV) reportedly resulted in more impurity electronic levels, which significantly reduced the band gap value and boosted the capability of ZrO2 to harvest light in the visible region [28]. Consequently, the B@ZrO2-OV photocatalyst exhibited prompted photocatalytic performance by degrading 87% of RhB and 71% of the colorless phenol owing to enhanced photon harvesting capability.
The DFT study was used to analyze the decrement in bandgap energy of N doped ZrO2 photocatalyst that further facilitated the separation of photocarriers. The DFT calculation was helpful in analyzing the effect of N-doping in the ZrO2 photocatalyst with its subsequent impact on the band structure and electronic density of states in N-doped ZrO2 (ZrO2−xNx, x = 0.125) [46]. The results revealed that the band gap of ZrO2 photocatalyst narrowed down from 5.0 to 2.860 eV by doping with N atoms, as shown in Figure 3d,e. Furthermore, the VBM increased from 0 to 0.473 V, while the CBM decreased from 3.398 to 3.333 V. Consequently, the introduction of N-atoms in a substitutional mode of ZrO2 photocatalyst resulted in the narrowed bandgap and extension of visible light absorptivity giving rise to enhanced photocatalytic degradation of formaldehyde (HCHO) into CO2 under the exposure of visible light for 50 min.
Doping with Au has also facilitated the reduction in bandgap energy of ZrO2 reduced from 5.72 to 5.40 eV [50]. With the increase in Au concentration, interactions occur between charge carriers and ionized impurities leading to a decreased band gap and photoluminescence (PL) intensity of ZrO2 photocatalyst. Hence, doping is an effective way to boost the photoactivity of the ZrO2 photocatalyst with noble metals, non-metals, and transition metals by tuning the bandgap energy to some extent by varying the concentration of impurity atoms.

3.3. Heterojunction Formation

Heterojunction formation is one of the most important ways to achieve better photocatalytic activity by reducing the high charge recombination rate, less visible-light absorption, and low electron utilization ratio [51]. ZrO2 has reportedly formed type-I heterostructure systems for water treatment applications. For instance, the formation of a type-I heterojunction for modulating the photocatalytic efficiency of ZrO2/ZnO heterojunction was efficiently reported [52]. With the formation of optimized ZrO2/ZnO (1:2) heterojunction, the bandgap energy of the nanocomposite was decreased from 5.3 to 3.2 eV. Additionally, the ZrO2/ZnO (1:2) heterojunction exhibited 93% photocatalytic degradation of Acid orange 8 (AO 8) under UV light illumination. In a different study, the photo-absorption ability of ZrO2 photocatalyst was reportedly enhanced when coupled with Polypyrrole (PPy) in the type-II mechanistic mode [53]. The decreased eh+ recombination rate and enhanced isolation of charge carriers were accredited to the type-II system with reduced bandgap energy from 3.26 to 2.45 eV. Moreover, the photocatalytic performance of the ZrO2/PPy system was assessed by degrading Reactive Black 31 (RB 31; 77%) and 4-chlorophenol (4-CP; 98%) under the exposure of visible light for 60 min. The synergic effect between the ZrO2 and PPy photocatalysts after heterostructure formation improved the charge carrier kinetics and provided sufficient carriers at the surface for photoredox reactions. While ZrO2-based type-I and type-II heterojunctions have shown considerable photocatalytic performance, the accumulation of photocarriers on one photocatalyst only and the electrostatic forces of repulsion between charge carriers along with the decreased redox potential in type-II system hampers the overall photocatalytic performance [54]. Thus, the formation of the Z-scheme photocatalytic system is considered to be a prominent approach that has been vastly opted to expedite the photocatalysis process.
For example, the photoactivity of g-C3N4/ZrO2−x Z-scheme heterojunction was analyzed for wastewater treatment [55]. The report illustrated that, due to the involvement of the Z-scheme charge transference, the g-C3N4/ZrO2−x system exhibited extended visible light irradiation and enhanced separation of charge carriers. Additionally, the presence of abundant, spatially separated e- facilitated the formation of •O2 species, which primarily participated in the degradation process. Consequently, g-C3N4/ZrO2−x system showed 90.6% tetracycline hydrochloride (TC-H) degradation under 1 h exposure to visible light (Figure 4a). Similarly, the photodegradation of Congo Red (CR) dye was reported by Hao et al. by utilizing the Ag/CeO2-ZrO2 nanocomposite [56]. The formation of three-dimensionally ordered macroporous Ag/CeO2-ZrO2 composite enabled the absorption of reactive molecules on the surface of photocatalysts, which reduced the recombination probability of charge carriers leading to enhanced photocatalytic performance.
Interestingly, the arrangement of the S-scheme heterojunction is quite similar to Type-II but it follows a completely different charge transfer mechanism [57]. Typically, the photoexcited e and h+ are accumulated on the CB of reduction photocatalyst (RP) and VB of oxidation photocatalyst (OP), respectively, which corresponds to a strong redox potential. The remaining (unutilized) photocarriers are eliminated through recombination, leading to better photocatalytic reactions. For instance, the formation of the S-scheme g-C3N4/ZrO2 heterojunction was achieved via a simple calcination method for improved photocatalytic performance [26]. The g-C3N4/ZrO2 S-scheme heterojunction inhibited the recombination rate of photocarriers due to a sufficiently exposed large specific surface area with abundant active sites. Moreover, the optimized 15% g-C3N4/ZrO2 heterojunction exhibited excellent photocatalytic performance by degrading 82% RhB, 50% methyl orange (MO), and 98% acid orange II (AO II) under visible light irradiation for 150 min (Figure 4b).

3.4. Generation of Oxygen Vacancies

The introduction of OVs in the ZrO2 photocatalyst is a useful strategy to increase the absorption range in the visible light region and to promote the adsorption affinity leading to enhanced photocatalytic degradation efficacy. The formation of vacancies in ZrO2 crystal acts as a trapping site for one or more electrons, which gives rise to F-centers with the local states in the crystal energy gap and produces the new energy band near the Fermi level [58]. The determination of OVs and surface defects can be done with the help of PL intensity, X-ray photoelectron spectroscopy (XPS), and X-ray diffraction (XRD) pattern. Over the past few years, various works have reported the successful generation and identification of OVs in the ZrO2 photocatalyst.
For instance, OVs in ZrO2 were created by depositing Co on its surface [25]. The XRD pattern identified the presence of OVs due to the distinct ionic radii of Zr and Co ions. The slight shift in the peak position intensity of Raman spectra (Figure 4c) and decreased the band gap energy of Co-ZrO2 occurred by the presence of OVs and surface defects. Furthermore, the OVs trapped the photogenerated e- leading to the formation of F-centers in the crystal structure of ZrO2, which inhibit the charge carrier’s reassembly and amend the photocatalytic performance. Wahba et al. studied the formation of OVs in the Zr-doped ZrO2/ZnO heterojunction through experimental and simulation studies [22]. The XRD pattern revealed that, with the increasing concentration of Zr from 30 to 50 wt%, two diffraction peaks corresponding to monoclinic ZrO2 and wurtzite hexagonal ZnO were obtained as delineated in Figure 4d. Additionally, the Zr4+ ions were replaced by Zn2+ ions, which induced the OVs in the crystal structure for trapping charge carriers and increasing their separation. XPS analysis can also be employed to evaluate the presence of OVs in ZrO2 photocatalysts. For instance, Reddy et al. reported the formation of OVs in Ni-doped ZrO2 by analyzing the XPS spectra of O 1 s [59]. In detail, the XPS study showed three types of OVs on the ZrO2 surface that is neutral, singly charged and doubly charged oxygen vacancy with different binding energies. The binding energy 530.7 eV, and 532.1 eV corresponded to singly and doubly charged OVs in ZrO2 photocatalyst, respectively.
From the above study, it is evident that the modification strategies are significantly effective in incrementing the photocatalytic performance of the ZrO2 photocatalyst. The change in morphology and optoelectronic structure of ZrO2 leads to enhanced visible light absorption and superior charge isolation. Doping with suitable dopants and the formation of ZrO2-based heterojunctions resulted in narrowed bandgap energy and improved charge carrier kinetics enabling their effectiveness in photocatalytic performance. Additionally, the presence of OVs not only acts as trapping sites for photoinduced charge carriers, but also renders defective states in between the bandgap energy which synergistically enhances the photoactivity of ZrO2 nanomaterials.

4. Synthesis Methods

ZrO2 photocatalysts having different morphologies, sizes, and compositions have been designed by various bottom-up and top-down strategies for efficient photocatalytic activity. Different advantages and challenges in the fabrication of ZrO2 using distinct synthesis methods are discussed as follows.

4.1. Bottom-Up Strategies

4.1.1. Hydrothermal/Solvothermal Method

The hydrothermal/solvothermal strategy is a synthesis method for the preparation of the ZrO2 photocatalyst, which involves a sealed vessel with water or any organic solvent as a reaction medium [60]. The crystallization of catalyst occurs at high temperatures and pressure maintained in the vessel. This method leads to the formation of powders with fine size or narrow size distribution with good chemical homogeneity and is dependent on some reaction parameters such as reaction time, pH value, temperature, solvent, etc. Surfactants also inhibit the growth of crystals by forming an ultrathin structure. Parveen et al. showed the pH dependence of hydrothermal reaction involving coupling of PPy with ZrO2 photocatalyst [53]. Typically, the ZrO2/PPy composite was fabricated by maintaining the hydrothermal reaction condition (180 °C/24 h) with the adjustment of pH value. The individual PPy and ZrO2 system showed the planer-structured PPy nanosheets with spherical ZrO2 particles. TEM results also revealed the wrinkled lamer structure nanosheets of PPy and ZrO2 samples (Figure 5a). Besides pH, the hydrothermal temperature also affects the morphology of the particles. As Rashid et al. prepared the ZrO2 doped ZnCo2O4 nanocomposites by the hydrothermal process with sulfate as a precursor, ethylene glycol (EG) and water as solvents at 250 to 650 °C calcination temperature [61]. The maximum increase in the surface area of ZrO2-ZnCo2O4 was achieved at 350 °C calcined temperature. After calcination, the ZrO2-ZnCo2O4 showed the presence of granular particles. Furthermore, the SEM results revealed the appearance of uniform-size nanoparticles of ZrO2-ZnCo2O4 composites at higher temperatures (Figure 5b).

4.1.2. Crystal Structure Engineering

The introduction of an impure atom can distort the crystal structure and change the balance of the ZrO2 photocatalyst by creating the OVs on its surface. Doping with suitable dopants further leads to a change in the morphology of the nanoparticles. For example, Wahba et al. analyzed the variation in morphology of Zr doped ZrO2/ZnO nanocomposite by preparing it with EG as a complexing agent [22]. The 2.5 wt% of Zr concentration changed the morphology of nanoparticles to elongated non-uniform distribution with large size. By increasing the concentration up to 5 wt%, the particle exhibited uniform distribution with less agglomeration. Moreover, the SEM results also showed the presence of homogeneous particle size and uniform distribution with 30, 40, and 50 wt% of Zr concentration in ZrO2/ZnO nanocomposite. In a different study, Navio et al. studied the distinction in the morphology of ZrO2 photocatalysts by doping with Fe [62]. The Fe/ZrO2 was prepared by maintaining the appropriate room temperature, and pH (9–10) of the sample. The SEM results revealed that the particles have a large shape and dimension distribution. With increasing the Fe content in the sample, the particles exhibited a high degree of wrinkled texture due to surface deposition. TEM descriptions indicated no substantial change in the morphology of the sample. However, in some cases, enlarged and different round-size grain particles were also observed.
Various studies have been reported to date to study the morphology of ZrO2 photocatalysts. Agorku et al. prepared the multi-element doped ZrO2 with Gd, C, N, and S elements to study the morphology of the photocatalyst [63]. The TEM results revealed the presence of a small-size multi-element doped ZrO2 with spherical shape and distinct grain boundaries. However, the SEM monographs showed uniform morphology with a crystalline-shaped photocatalyst. Furthermore, 0.6% Gd doped C, N, and S-ZrO2 exhibited the highest photocatalytic activity due to reduced bandgap energy and improved surface-active sites of the photocatalyst as a result of multi-element doping. Chen et al. also studied the morphology of g-C3N4 loaded ZrO2−x nanotubes [55]. ZrO2 nanotubes are tailored through the anodic oxidation of ZrO2 in a two-electrode system with a 20 mm diameter for each electrode. The prepared ZrO2−x nanotubes were further annealed at 550 °C for 3 h in the presence of Argon. The SEM results showed the clean surface morphology of ZrO2−x nanotubes with a 50 nm diameter. By increasing the g-C3N4 concentration, the optimal 0.09CN/ZrO2−x sample exhibited completely covered structure nanotubes. Furthermore, TEM results revealed the regular and clean tubular morphology of ZrO2−x nanotubes whereas the tubular morphology of 0.06 CN/ZrO2−x nanotubes with apparent attachments was further observed by increasing the g-C3N4 concentration. The presence of g-C3N4 in the composite was confirmed with the help of Fourier transform infrared (FTIR) spectroscopy as delineated in Figure 5c. The broad band of g-C3N4 in the NH2 group stretching mode was observed at 3178 cm−1, while for ZrO2−x, the NTs showed a broad band of OH stretching at 3420 cm−1. The Zr-O bond in ZrO2 corresponds to 490, 580, and 780 cm−1 bands vibrations and the composite showed a peak at 2046 cm−1. The as-synthesized composite exhibited 90.6% degradation of tetracycline hydrochloride after 1 h of visible light exposure.

4.2. Top-Down Strategies

4.2.1. Thermal Treatment

Thermal treatment is a significant method utilized for the introduction of OVs in ZrO2 at high temperatures, which leads to the cleavage of bonds and further boosts the escape of atoms from the lattice. At definite temperatures, the metals become decomposed to form nanoparticles by involving a chemical reaction. For instance, Zhang et al. prepared an S-scheme g-C3N4/ZrO2 heterostructure via a simple calcination method [26]. Experimental results explained the morphology changes in g-C3N4/ZrO2 heterojunction after variation in calcination temperature. The presence of irregularly arranged nanoparticles on the surface of ZrO2 with the variation in calcination temperature was confirmed by SEM results. With variation in the g-C3N4 concentration, the surface of ZrO2 photocatalyst became dimmer by the existence of spindle-shaped nanoparticles. It has been observed that the mixing of two metal oxide phases can also help to modulate the photocatalytic activity by inhibiting the crystal growth during the heterostructure formation via calcination. For instance, the heterojunction formation between ZrO2 and Bi2O3 via post-thermal treatment rendered a facile morphology control approach by inhibiting the crystal growth of Bi2O3 [64]. From the TEM images, an optimal particle size of 20 nm was observed and was in good agreement with the XRD results. The UV-Visible diffuse reflectance spectroscopy (UV-vis-DRS) spectra of Bi2O3-ZrO2, Bi2O3, and ZrO2 were depicted in Figure 5d where a red shift in the composite was observed due to the electron–electron transition in between bare ZrO2 and Bi2O3. The spectra also explained the absorbance intensity shift of Bi2O3-ZrO2 towards the longer wavelength than Bi2O3 and ZrO2. Hence, thermal treatment can be successfully employed to modulate the surface morphology of ZrO2 and ZrO2-based photocatalysts.

4.2.2. Ultrasonication

Sonication involves the dispersion of the precursor in a solvent followed by further exfoliation of materials. Furthermore, shock waves and high-speed inter-particle collisions also help to improve the surface structure and may sometimes induce the formation of OVs. Zarei et al. synthesized ZrO2/g-C3N4 heterojunction via mixing of g-C3N4 powder into ZrO2 colloidal dispersion through ultrasonic treatment at 30 °C for 2 h [65]. The integration of ZrO2 nanoparticles into g-C3N4 increased their surface area to 51.8 m2 g−1 due to its distinct morphology. The TEM results indicated that g-C3N4 showed planer and sheet-like structures while ZrO2 nanoparticles were spherical in shape with a 15 nm particle size. Additionally, the XRD pattern revealed the existence of strong and sharp diffraction peaks, which confirmed a high degree of crystallinity in the sample (Figure 5e). Furthermore, the incorporation of ultrasonic treatment with various dopant concentrations can also disturb the morphology of the ZrO2 photocatalyst. For example, Yu et al. synthesized the TiO2-N/ZrO2−xNx heterojunction by subjecting the precursors such as ZrOCl2·8H2O and Ti (OC4H9)4 under ultrasonic treatment [46]. With the variation in the concentration of N, the transformation of tetragonal ZrO2 to monoclinic was strictly inhibited with resultantly improved surface-active sites. The XRD patterns helped to understand the interactions between N atoms with O atoms in ZrO2 which led to the substitution of O by nitrogen dopants due to high-energy ultrasonic waves.

4.2.3. Chemical Reduction

Chemical reduction is a method used for the reduction of ionic salt in the presence of surfactants by using reducing agents such as sodium borohydride, lithium aluminum hydride, etc. Interestingly, OVs can also be introduced with the help of reducing agents and by varying the concentration of reductants, the content of vacancies can be controlled. For instance, Qin et al. produced a B@ZrO2-OV photocatalyst through the introduction of NaBH4 as reducing agent [28]. In this process, ZrO2 and NaBH4 were taken in different weight ratios and heated at 650 °C for 2 h followed by the annealing of the sample. More importantly, the weight ratios of ZrO2:NaBH4 was optimized to 1:0.4 because, beyond this, a further increase in the amount of NaBH4 meant that the crystal phase of ZrO2 was transformed into ZrB2. The electron paramagnetic resonance (EPR) spectroscopy confirmed the existence of OVs in the B@ZrO2 photocatalyst at 2.003 g-value. Moreover, the XPS analysis showed three deconvoluted peaks at 529.9 eV, 531.4 eV, and 532.8 eV ascribed to the lattice oxygen, oxygen deficient, and chemically adsorbed oxygen species, respectively.

5. Photodegradation and Mechanism

Environmental contamination by organic dyes, pharmaceuticals, and insecticides is posing a serious threat worldwide owing to their toxic, cancerous nature and difficulty in total decomposition causing a shortage in drinking water. Thus far, various traditional procedures, including biological, physical, and chemical methods have been used to remove these pollutants, however the applicability of these technologies is limited due to their high cost, risk of hazardous byproducts, and poor efficiency. As a result, a more direct and effective way to deal with these contaminants is highly essential [55,66]. Water treatment employing a simple method like photocatalysis is in the limelight owing to its low energy requirement, cost-effectiveness, capacity to adsorb a wide range of natural and inorganic substances on the photocatalyst’s surface, and complete degradation without hazardous by-products [67]. Even photocatalysis seemed to be a promising method due to the easy availability of a large amount of solar energy [23]. Semiconductor photocatalytic activity has potential applications in the degradation of environmental contaminants and pollutant transformation. Thus, it has aroused a lot of attention [68].

5.1. Organic Pollutant Degradation

Organic pollutants are identified as harmful for human health as they are toxic, carcinogenic, and teratogenic, so the use of heterogeneous photocatalysis to decontaminate wastewater containing organic compounds is an attractive solution to the effluent treatment problem since it brings about chemical degradation into non-toxic molecules like carbon dioxide and water [29,66]. Charge transfer and electron transfer are involved in the degradation of organic pollutants, and these processes are impacted by the presence of charged species [69]. The alluring features of ZrO2 involving physicochemical stability and distinctive band structures render optimal photodegradation ability, however the broad bandgap and low separation rate of photocarriers limit its wide-scale practical uses [55]. However, the synergistic coupling of ferrites with zirconia boosts the overall photoactivity by the combined action of inhibited charge separation and integration of photo and Fenton activity. In detail, Figure 6a depicts the suggested photocatalytic process on the ZrO2/ZnFe2O4 nanocomposite (ZZFO 12 NC) for the photodegradation of indigo carmine dye (98%) [31]. After the excitation and isolation of photocarriers, the reaction of photoinduced electrons oxygen yielded superoxide radicals (•O2), which further reacted with H2O to produce H2O2. This generated H2O2 reaction with α-Fe2O3 in ZZFO 12 further produced •OH radicals and participated in dye degradation, as shown in Equations (1)–(7)
ZZFO + h ν ZrO 2 h + + ZnFe 2 O 4   e _  
ZnFe 2 O 4   e _ + O 2   O 2
O 2 + H 2 O H 2 O 2 + OH  
H 2 O 2 + α Fe 2 O 3 2   OH  
h + + OH 2   OH  
OH + indigo   carmine Degradation
O 2 + indigo   carmine Degradation
Other than dyes, phenolic compounds have also been treated effectively with ZrO2-based nanocomposites. For example, ZrO2/g-C3N4 nanocomposite synthesized by ultrasonication method for the efficient degradation of 4-chlorophenol (4-CP) (Figure 6b) in water with the highest degradation rate of 0.0173 min−1 [65]. The synergistic action of higher surface area with exposed active sites facilitates the adsorption of 4-CP and the surface-induced charges accelerated the formation of •O2 radicals which boosted the degradation efficiency. Moreover, from the radical scavenging results, it was confirmed that •O2 radicals and h+ primarily participated in the degradation process.

5.2. Inorganic Pollutant Degradation

The higher concentration of inorganic pollutants in the water due to their discharge from various industries like fertilizers, refineries, and pharmaceuticals is causing a major threat to human as well as aquatic lives. Thus the photocatalytic treatment of such toxic effluents using ZrO2 as a photocatalyst can be a potential solution. In this regard, enhancement in Cr (VI) photoreduction under visible light irradiation is achieved by developing a Bi2O3-ZrO2 nanocomposite [64]. Bi2O3-ZrO2 has a greater photocatalytic activity (92.30%) after an irradiation period of 180 min. The mechanism for enhanced photocatalytic activity of Bi2O3-ZrO2 is explained according to previously published literature on the photocatalytic activity of nanocomposites. Typically, Cr (VI) is photo-reduced by CB electrons captured by ZrO2, and the VB holes further oxidized H2O to create O2 (Figure 6c). The detailed photocatalytic processes that occur on the surface of Bi2O3-ZrO2 are summarized in Equations (8)–(11):
Bi 2 O 3 + hv Bi 2 O 3 h VB + +   e CB
Bi 2 O 3 e CB + ZrO 2 Bi 2 O 3 + ZrO 2 e CB
ZrO 2 6 e CB + Cr 2 O 7 2 + 14 H + 2 Cr 3 + + 7 H 2 O + ZrO 2
2 H 2 O + 4 h + O 2 + 4 H +
Various ZrO2-based semiconductor photocatalysts with incremented photodegradation activity are summarized in Table 1 [59,70,71,72,73,74,75,76,77,78,79].

6. Concluding Remarks

Among various metal-oxide photocatalysts, ZrO2 is an n-type wide bandgap semiconducting material with substantial features like physicochemical stability, facile synthesis, cost-effectiveness, etc. In spite of various fascinating properties of ZrO2, there are various factors that limit its practical applications. Therefore, to enhance its photocatalytic activity, suitable modification techniques such as doping, heterojunction formation, and vacancy creations have been opted. Considering that, the present study highlights the recent advances in the photocatalytic activity of ZrO2 and its nanocomposites. The structural features, formation of band edge potentials, and the characteristics of optical as well as electronic properties are comprehensively discussed by integrating experimental and computational studies. Various strategic modulations, such as morphology and structural control, bandgap engineering, heterostructure construction, and vacancy engineering are the major approaches for synergistically improving the charge carrier’s dynamics and light harvesting ability of ZrO2. Moreover, various synthetic routes involving bottom-up and top-down methods can be utilized to modulate the physicochemical features of ZrO2 and attain effective photocatalytic degradation performance. Thus, it is envisaged that, with the aid of the research advances discussed in this review, future research can be focused to explore the photocatalytic performance of modified ZrO2 for large-scale photocatalytic applications.

Author Contributions

Writing—original draft preparation, V.R. and A.S. (Amit Sharma); writing—review and editing, P.S. and A.K.; visualization and project administration, P.S., S.T. and A.S. (Archana Singh); conceptualization, validation and supervision, Q.V.L., V.H.N. and P.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hosseini-Bandegharaei, A. Adsorption and photocatalysis compiled toxic dyes mineralization using graphitic carbon nitride modified ZnFe2O4 and CoFe2O4 photocatalysts supported onto N-doped graphene. Desalination Water Treat 2020, 191, 381–399. [Google Scholar]
  2. Kumar, A.; Raizada, P.; Hosseini-Bandegharaei, A.; Thakur, V.K.; Nguyen, V.H.; Singh, P. C-, N-Vacancy defect engineered polymeric carbon nitride towards photocatalysis: Viewpoints and challenges. J. Mater. Chem. A 2021, 9, 111–153. [Google Scholar] [CrossRef]
  3. Pirhashemi, M.; Habibi-Yangjeh, A.; Pouran, S.R. Review on the criteria anticipated for the fabrication of highly efficient ZnO-based visible-light-driven photocatalysts. J. Ind. Eng. Chem. 2018, 62, 1–25. [Google Scholar] [CrossRef]
  4. Kumari, P.; Bahadur, N.; Kong, L.; O’Dell, L.A.; Merenda, A.; Dumee, L. Engineering Schottky-like and heterojunction material for enhanced photocatalysis performance-a review. Adv. Mater. 2022, 3, 2309–2323. [Google Scholar] [CrossRef]
  5. Almeida, F.; Grzebielucka, E.C.; Antunes, S.R.M.; Borges, C.P.F.; Andrade, A.V.C.; Souza, É.C.F. Visible light activated magnetic photocatalysts for water treatment. J. Environ. Manag. 2020, 273, 111143. [Google Scholar] [CrossRef] [PubMed]
  6. Alshorifi, F.T.; Ali, S.L.; Salama, R.S. Promotional Synergistic Effect of Cs–Au NPs on the Performance of Cs–Au/MgFe2O4 Catalysts in Catalysis 3, 4-Dihydropyrimidin-2 (1H)-Ones and Degradation of RhB Dye. J. Inorg. Organomet. Polym. Mater. 2022, 32, 3765–3776. [Google Scholar] [CrossRef]
  7. Ahamad, T.; Naushad, M.; Ubaidullah, M.; Alshehri, S. Fabrication of highly porous polymeric nanocomposite for the removal of radioactive U (VI) and Eu (III) ions from aqueous solution. Polym. J. 2020, 12, 2940. [Google Scholar] [CrossRef]
  8. Balakrishnan, T.; Kiran, K.U.V.; Kumar, S.M.; Raju, A.; Kumar, S.A.; Mayavan, S. Copper particles decorated boron carbon nitride, an efficient catalyst for methyl orange oxidation under sonication and sunlight irradiation. J. Environ. Chem. Eng. 2017, 5, 564–571. [Google Scholar] [CrossRef]
  9. Alshorifi, F.T.; Alswat, A.A.; Salama, R.S. Gold-selenide quantum dots supported onto cesium ferrite nanocomposites for the efficient degradation of rhodamine B. Heliyon 2022, 8, e09652. [Google Scholar] [CrossRef]
  10. Al-Enizi, A.M.; Ubaidullah, M.; Ahmed, J.; Ahamad, T.; Ahmad, T.; Shaikh, S.F.; Naushad, M. Synthesis of NiOx@ NPC composite for high-performance supercapacitor via waste PET plastic-derived Ni-MOF. Compos. B. Eng. 2020, 183, 107655. [Google Scholar] [CrossRef]
  11. Hasija, V.; Raizada, P.; Sudhaik, A.; Sharma, K.; Kumar, A.; Singh, P.; Jonnalagadda, S.B.; Thakur, V.K. Recent advances in noble metal free doped graphitic carbon nitride based nanohybrids for photocatalysis of organic contaminants in water: A review. Appl. Mater. Today 2019, 15, 494–524. [Google Scholar] [CrossRef]
  12. El-Hakam, S.A.; ALShorifi, F.T.; Salama, R.S.; Gamal, S.; El-Yazeed, W.A.; Ibrahim, A.A.; Ahmed, A.I. Application of nanostructured mesoporous silica/bismuth vanadate composite catalysts for the degradation of methylene blue and brilliant green. J. Mater. Res. Technol. 2022, 18, 1963–1976. [Google Scholar] [CrossRef]
  13. Thakur, N.; Kumar, K.; Thakur, V.K.; Soni, S.; Kumar, A.; Samant, S.S. Antibacterial and photocatalytic activity of undoped and (Ag, Fe) co-doped CuO nanoparticles via microwave-assisted method. Nanofabrication 2022, 7. [Google Scholar] [CrossRef]
  14. Andreozzi, R.; Caprio, V.; Insola, A.; Marotta, R. Advanced oxidation processes (AOP) for water purification and recovery. Catal. Today 1999, 53, 51–59. [Google Scholar] [CrossRef]
  15. Jangwan, J.; Kumar, S.S.; Kumar, V.; Kumar, A.; Kumar, D. A review on the capability of zinc oxide and iron oxides nanomaterials, as a water decontaminating agent: Adsorption and photocatalysis. Appl. Water Sci. 2022, 12, 1–14. [Google Scholar]
  16. Hussain, N.; Bilal, M.; Iqbal, H.M. Carbon-based nanomaterials with multipurpose attributes for water treatment: Greening the 21st-century nanostructure materials deployment. Biomater. Polym. Horiz. 2022, 1, 1–11. [Google Scholar] [CrossRef]
  17. Pournemati, K.; Habibi-Yangjeh, A.; Khataee, A. Rational design of TiO2/MnMoO4/MoO3 nanocomposites: Visible-light-promoted photocatalysts for decomposition of tetracycline with tandem nn heterojunctions. Colloids Surf. A Physicochem. Eng. Asp. 2022, 655, 130315. [Google Scholar] [CrossRef]
  18. Pelosato, R.; Bolognino, I.; Fontana, F.; Sora, I.N. Applications of Heterogeneous Photocatalysis to the Degradation of Oxytetracycline in Water: A Review. Molecules 2022, 27, 2743. [Google Scholar] [CrossRef]
  19. Murillo-Sierra, J.; Hernández-Ramírez, A.; Hinojosa-Reyes, L.; Guzmán-Mar, J. A review on the development of visible light-responsive WO3-based photocatalysts for environmental applications. Adv. Chem. Eng. 2021, 5, 100070. [Google Scholar] [CrossRef]
  20. Luo, S.; Zhang, C.; Almatrafi, E.; Yan, M.; Liu, Y.; Fu, Y.; Wang, Z.; Li, L.; Zhou, C.; Xu, P. Photocatalytic water purification with graphitic C3N4-based composites: Enhancement, mechanisms, and performance. Appl. Mater. Today 2021, 24, 101118. [Google Scholar] [CrossRef]
  21. Sang, Y.; Liu, H.; Umar, A. Photocatalysis from UV/Vis to near-infrared light: Towards full solar-light spectrum activity. ChemCatChem 2015, 7, 559–573. [Google Scholar] [CrossRef]
  22. Wahba, M.A.; Yakout, S.M.; Mohamed, W.A.; Galal, H.R. Remarkable photocatalytic activity of Zr doped ZnO and ZrO2/ZnO nanocomposites: Structural, morphological and photoluminescence properties. Mater. Chem. Phys. 2020, 256, 123754. [Google Scholar] [CrossRef]
  23. Pirzada, B.M.; Mir, N.A.; Qutub, N.; Mehraj, O.; Sabir, S.; Muneer, M. Synthesis, characterization and optimization of photocatalytic activity of TiO2/ZrO2 nanocomposite heterostructures. Mater. Sci. Eng. B 2015, 193, 137–145. [Google Scholar] [CrossRef]
  24. Aldeen, E.S.; Jalil, A.A.; Mim, R.S.; Alhebshi, A.; Hassan, N.S.; Saravanan, R. Altered zirconium dioxide based photocatalyst for enhancement of organic pollutants degradation: A review. Chemosphere 2022, 304, 135349. [Google Scholar] [CrossRef]
  25. Ahmed, W.; Iqbal, J. Co doped ZrO2 nanoparticles: An efficient visible light triggered photocatalyst with enhanced structural, optical and dielectric characteristics. Ceram. Int. 2020, 46, 25833–25844. [Google Scholar] [CrossRef]
  26. Zhang, K.; Zhou, M.; Yu, C.; Yang, K.; Li, X.; Dai, W.; Guan, J.; Shu, Q.; Huang, W. Construction of S-scheme g-C3N4/ZrO2 heterostructures for enhancing photocatalytic disposals of pollutants and electrocatalytic hydrogen evolution. Dyes Pigm. 2020, 180, 108525. [Google Scholar] [CrossRef]
  27. García-López, E.; Marcì, G.; Pomilla, F.; Paganini, M.; Gionco, C.; Giamello, E.; Palmisano, L. ZrO2 Based materials as photocatalysts for 2-propanol oxidation by using UV and solar light irradiation and tests for CO2 reduction. Catal. Today 2018, 313, 100–105. [Google Scholar] [CrossRef]
  28. Qin, Y.; Ding, Z.; Guo, W.; Guo, X.; Hou, C.; Jiang, B.-P.; Liu, C.-G.; Shen, X.-C. A full solar light spectrum responsive B@ ZrO2–OV photocatalyst: A synergistic strategy for visible-to-NIR photon harvesting. ACS Sustain. Chem. Eng. 2020, 8, 13039–13047. [Google Scholar] [CrossRef]
  29. Wang, Y.; Zhang, Y.; Lu, H.; Chen, Y.; Liu, Z.; Su, S.; Xue, Y.; Yao, J.; Zeng, H. Novel N-doped ZrO2 with enhanced visible-light photocatalytic activity for hydrogen production and degradation of organic dyes. RSC Adv. 2018, 8, 6752–6758. [Google Scholar] [CrossRef] [Green Version]
  30. Kaviyarasu, K.; Kotsedi, L.; Simo, A.; Fuku, X.; Mola, G.T.; Kennedy, J.; Maaza, M. Photocatalytic activity of ZrO2 doped lead dioxide nanocomposites: Investigation of structural and optical microscopy of RhB organic dye. Appl. Surf. Sci. 2017, 421, 234–239. [Google Scholar] [CrossRef]
  31. Renuka, L.; Anantharaju, K.; Vidya, Y.; Nagabhushana, H.; Uma, B.; Malini, S.; More, S.S.; Koppad, P. Porous network ZrO2/ZnFe2O4 nanocomposite with heterojunction towards industrial water purification under sunlight: Enhanced charge separation and elucidation of photo-mechanism. Ceram. Int. 2021, 47, 14845–14861. [Google Scholar] [CrossRef]
  32. Idrissi, S.; Ziti, S.; Labrim, H.; Bahmad, L. Sulfur doping effect on the electronic properties of zirconium dioxide ZrO2. Mater. Sci. Eng. B 2021, 270, 115200. [Google Scholar] [CrossRef]
  33. Faizan, M.; Siddique, M.N.; Ahmad, S.; Tripathi, P.; Riyajuddin, S. Tunable luminescence in Ce3+/Mn2+ co-doped ZrO2 nanophosphor integrated with theoretical studies on possible (ZrO2) n clusters using DFT method. J. Alloys Compd. 2021, 853, 157378. [Google Scholar] [CrossRef]
  34. Zhou, H.; Yao, P.; Gong, T.; Xiao, Y.; Zhang, Z.; Zhao, L.; Fan, K.; Deng, M. Effects of ZrO2 crystal structure on the tribological properties of copper metal matrix composites. Tribol. Int. 2019, 138, 380–391. [Google Scholar] [CrossRef]
  35. Lamas, D.; Rosso, A.; Anzorena, M.S.; Fernández, A.; Bellino, M.; Cabezas, M.; de Reca, N.W.; Craievich, A. Crystal structure of pure ZrO2 nanopowders. Scr. Mater. 2006, 55, 553–556. [Google Scholar] [CrossRef]
  36. McCullough, J.T.; Trueblood, K. The crystal structure of baddeleyite (monoclinic ZrO2). Acta Crystallogr. 1959, 12, 507–511. [Google Scholar] [CrossRef]
  37. Singh, H.; 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]
  38. Teufer, G. The crystal structure of tetragonal ZrO2. Acta Crystallogr. 1962, 15, 1187. [Google Scholar] [CrossRef]
  39. Lin, Y.-F.; Liang, F.-L. ZrO2/carbon aerogel composites: A study on the effect of the crystal ZrO2 structure on cationic dye adsorption. J. Taiwan Inst. Chem. Eng. 2016, 65, 78–82. [Google Scholar] [CrossRef]
  40. Morinaga, M.; Adachi, H.; Tsukada, M. Electronic structure and phase stability of ZrO2. J. Phys. Chem. Solids 1983, 44, 301–306. [Google Scholar] [CrossRef]
  41. Koe, W.S.; Lee, J.W.; Chong, W.C.; Pang, Y.L.; Sim, L.C. An overview of photocatalytic degradation: Photocatalysts, mechanisms, and development of photocatalytic membrane. Environ. Sci. Pollut. Res. 2020, 27, 2522–2565. [Google Scholar] [CrossRef] [PubMed]
  42. Wang, J.; Huang, J.; Meng, J.; Li, Q.; Yang, J. Double-hole codoped huge-gap semiconductor ZrO2 for visible-light photocatalysis. Phys. Chem. Chem. Phys. 2016, 18, 17517–17524. [Google Scholar] [CrossRef] [PubMed]
  43. Mazumder, J.T.; Mayengbam, R.; Tripathy, S. Theoretical investigation on structural, electronic, optical and elastic properties of TiO2, SnO2, ZrO2 and HfO2 using SCAN meta-GGA functional: A DFT study. Mater. Chem. Phys. 2020, 254, 123474. [Google Scholar] [CrossRef]
  44. Yang, R.; Zhang, Y.; Fan, Y.; Wang, R.; Zhu, R.; Tang, Y.; Yin, Z.; Zeng, Z. InVO4-based photocatalysts for energy and environmental applications. Chem. Eng. 2022, 428, 131145. [Google Scholar] [CrossRef]
  45. Renuka, L.; Anantharaju, K.; Sharma, S.; Nagaswarupa, H.; Prashantha, S.; Nagabhushana, H.; Vidya, Y. Hollow microspheres Mg-doped ZrO2 nanoparticles: Green assisted synthesis and applications in photocatalysis and photoluminescence. J. Alloys Compd. 2016, 672, 609–622. [Google Scholar] [CrossRef]
  46. Yu, Y.; Zhang, P.; Kuang, Y.; Ding, Y.; Yao, J.; Xu, J.; Cao, Y. Adjustment and Control of Energy Levels for TiO2–N/ZrO2–xNx with Enhanced Visible Light Photocatalytic Activity. J. Phys. Chem. C 2014, 118, 20982–20988. [Google Scholar] [CrossRef]
  47. Xia, Y.; Sun, Q.; Wang, D.; Zeng, X.-F.; Wang, J.-X.; Chen, J.-F. Surfactant-free aqueous dispersions of shape-and size-controlled zirconia colloidal nanocrystal clusters with enhanced photocatalytic activity. Langmuir 2019, 35, 11755–11763. [Google Scholar] [CrossRef]
  48. Fu, X.; Clark, L.A.; Yang, Q.; Anderson, M.A. Enhanced photocatalytic performance of titania-based binary metal oxides: TiO2/SiO2 and TiO2/ZrO2. Environ. Sci. Technol. 1996, 30, 647–653. [Google Scholar] [CrossRef]
  49. Linnik, O.; Shestopal, N.; Smirnova, N.; Eremenko, A.; Korduban, O.; Kandyba, V.; Kryshchuk, T.; Socol, G.; Stefan, N.; Popescu-Pelin, G. Correlation between electronic structure and photocatalytic properties of non-metal doped TiO2/ZrO2 thin films obtained by pulsed laser deposition method. Vacuum 2015, 114, 166–171. [Google Scholar] [CrossRef]
  50. Berlin, I.J.; Joy, K. Optical enhancement of Au doped ZrO2 thin films by sol–gel dip coating method. Physica B Condens. Matter 2015, 457, 182–187. [Google Scholar] [CrossRef]
  51. Anwer, H.; Park, J.-W. Synthesis and characterization of a heterojunction rGO/ZrO2/Ag3PO4 nanocomposite for degradation of organic contaminants. J. Hazard. Mater. 2018, 358, 416–426. [Google Scholar] [CrossRef] [PubMed]
  52. Gurushantha, K.; Renuka, L.; Anantharaju, K.; Vidya, Y.; Nagaswarupa, H.; Prashantha, S.; Nagabhushana, H. Photocatalytic and photoluminescence studies of ZrO2/ZnO nanocomposite for LED and waste water treatment applications. Mater. Today Proc. 2017, 4, 11747–11755. [Google Scholar] [CrossRef]
  53. Parveen, A.; Surumbarkuzhali, N. Spatial separation of photo-generated carriers and enhanced photocatalytic performance on ZrO2 catalysts via coupling with PPy. Inorg. Chem. Commun. 2020, 120, 108153. [Google Scholar] [CrossRef]
  54. Habibi-Yangjeh, A.; Basharnavaz, H. Ni/P, Pt/P and Pd/P-modified graphitic carbon nitride nanosheets for hydrogen storage application using a DFT investigation. Mol. Phy. 2022, e2124934. [Google Scholar] [CrossRef]
  55. Chen, Q.; Yang, W.; Zhu, J.; Fu, L.; Li, D.; Zhou, L. Enhanced visible light photocatalytic activity of g-C3N4 decorated ZrO2−x nanotubes heterostructure for degradation of tetracycline hydrochloride. J. Hazard. Mater. 2020, 384, 121275. [Google Scholar] [CrossRef]
  56. Hao, Y.; Li, L.; Liu, D.; Yu, H.; Zhou, Q. The synergy of SPR effect and Z-scheme of Ag on enhanced photocatalytic performance of 3DOM Ag/CeO2-ZrO2 composite. J. Mol. Catal. 2018, 447, 37–46. [Google Scholar] [CrossRef]
  57. Xu, Q.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. S-scheme heterojunction photocatalyst. Chem 2020, 6, 1543–1559. [Google Scholar] [CrossRef]
  58. Sobolev, A.; Varaksin, A.; Keda, O.; Khaimenov, A. Electronic structure and charge state of oxygen vacancies in perfect ZrO2 crystals. Phys. Status Solidi B Basic Res. 1990, 162, 165–171. [Google Scholar] [CrossRef]
  59. Reddy, C.V.; Reddy, I.N.; Ravindranadh, K.; Reddy, K.R.; Kim, D.; Shim, J. Ni-dopant concentration effect of ZrO2 photocatalyst on photoelectrochemical water splitting and efficient removal of toxic organic pollutants. Sep. Purif. Technol. 2020, 252, 117352. [Google Scholar] [CrossRef]
  60. Ma, H.; He, Y.; Chen, P.; Wang, H.; Sun, Y.; Li, J.; Dong, F.; Xie, G.; Sheng, J. Ultrathin Two-Dimensional Bi-Based photocatalysts: Synthetic strategies, surface defects, and reaction mechanisms. Chem. Eng. J. 2021, 417, 129305. [Google Scholar] [CrossRef]
  61. Rashid, J.; Barakat, M.; Mohamed, R.; Ibrahim, I. Enhancement of photocatalytic activity of zinc/cobalt spinel oxides by doping with ZrO2 for visible light photocatalytic degradation of 2-chlorophenol in wastewater. J. Photochem. Photobiol. 2014, 284, 1–7. [Google Scholar] [CrossRef]
  62. Navio, J.; Hidalgo, M.; Colon, G.; Botta, S.; Litter, M. Preparation and physicochemical properties of ZrO2 and Fe/ZrO2 prepared by a sol− gel technique. Langmuir 2001, 17, 202–210. [Google Scholar] [CrossRef]
  63. Agorku, E.; Pandey, A.; Mamba, B.; Mishra, A. Gd, C, N, S multi-doped ZrO2 for photocatalytic degradation of indigo carmine dye from synthetic water under simulated solar light. Mater. Today Proc. 2015, 2, 3909–3920. [Google Scholar] [CrossRef]
  64. Vignesh, K.; Priyanka, R.; Rajarajan, M.; Suganthi, A. Photoreduction of Cr (VI) in water using Bi2O3–ZrO2 nanocomposite under visible light irradiation. J. Mater. Sci. Eng. B 2013, 178, 149–157. [Google Scholar] [CrossRef]
  65. Zarei, M. Ultrasonic-assisted preparation of ZrO2/g-C3N4 nanocomposites with high visible-light photocatalytic activity for degradation of 4-chlorophenol in water. Water-Energy Nexus 2020, 3, 135–142. [Google Scholar] [CrossRef]
  66. Zhao, W.; Adeel, M.; Zhang, P.; Zhou, P.; Huang, L.; Zhao, Y.; Ahmad, M.A.; Shakoor, N.; Lou, B.; Jiang, Y.; et al. A critical review on surface modified nano-catalysts application for photocatalytic degradation of volatile organic compounds. Environ. Sci. Nano 2022, 9, 61–80. [Google Scholar] [CrossRef]
  67. Xu, F.; Lai, C.; Zhang, M.; Li, B.; Liu, S.; Chen, M.; Li, L.; Xu, Y.; Qin, L.; Fu, Y. Facile one-pot synthesis of carbon self-doped graphitic carbon nitride loaded with ultra-low ceric dioxide for high-efficiency environmental photocatalysis: Organic pollutants degradation and hexavalent chromium reduction. J. Colloid Interface Sci. 2021, 601, 196–208. [Google Scholar] [CrossRef]
  68. Zheng, H.; Liu, K.; Cao, H.; Zhang, X. L-Lysine-assisted synthesis of ZrO2 nanocrystals and their application in photocatalysis. J. Phys. Chem. C 2009, 113, 18259–18263. [Google Scholar] [CrossRef]
  69. Polisetti, S.; Deshpande, P.A.; Madras, G. Photocatalytic activity of combustion synthesized ZrO2 and ZrO2–TiO2 mixed oxides. Ind. Eng. Chem. Res. 2011, 50, 12915–12924. [Google Scholar] [CrossRef]
  70. de Moraes, N.P.; de Azeredo, C.A.S.H.; Bacetto, L.A.; da Silva, M.L.C.P.; Rodrigues, L.A. The effect of C-doping on the properties and photocatalytic activity of ZrO2 prepared via sol-gel route. Optik 2018, 165, 302–309. [Google Scholar] [CrossRef]
  71. Keerthana, S.P.; Yuvakkumar, R.; Kumar, P.S.; Ravi, G.; Velauthapillai, D. Nd doped ZrO2 photocatalyst for organic pollutants degradation in wastewater. Environ. Technol. Innov. 2022, 28, 102851. [Google Scholar] [CrossRef]
  72. Yaghoubi, A.; Ramazani, A.; Taghavi Fardood, S. Synthesis of Al2O3/ZrO2 nanocomposite and the study of its effects on photocatalytic degradation of reactive blue 222 and reactive yellow 145 dyes. ChemistrySelect 2020, 5, 9966–9973. [Google Scholar] [CrossRef]
  73. Du, W.; Zhu, Z.; Zhang, X.; Wang, D.; Liu, D.; Qian, X.; Du, J. RE/ZrO2 (RE= Sm, Eu) composite oxide nano-materials: Synthesis and applications in photocatalysis. Mater. Res. Bull. 2013, 48, 3735–3742. [Google Scholar] [CrossRef]
  74. Zhou, Q.; Li, L.; Zhang, X.; Yang, H.; Cheng, Y.; Che, H.; Wang, L.; Cao, Y. Construction of heterojunction and homojunction to improve the photocatalytic performance of ZnO quantum dots sensitization three-dimensional ordered hollow sphere ZrO2–TiO2 arrays. Int. J. Hydrog. Energy 2020, 45, 31812–31824. [Google Scholar] [CrossRef]
  75. Du, W.; Wang, X.; Li, H.; Ma, D.; Hou, S.; Zhang, J.; Qian, X.; Pang, H. ZrO2/Dy2O3 Solid Solution Nano-Materials: Tunable Composition, Visible light–Responsive Photocatalytic Activities and Reaction Mechanism. J. Am. Ceram. Soc. 2013, 96, 2979–2986. [Google Scholar] [CrossRef]
  76. Carbuloni, C.F.; Savoia, J.E.; Santos, J.S.; Pereira, C.A.; Marques, R.G.; Ribeiro, V.A.; Ferrari, A.M. Degradation of metformin in water by TiO2–ZrO2 photocatalysis. J. Environ. Manag. 2020, 262, 110347. [Google Scholar] [CrossRef]
  77. Tanzifi, M.; Jahanshahi, M.; Peyravi, M.; Khalili, S. A morphological decoration of g-C3N4/ZrO2 heterojunctions as a visible light activated photocatalyst for degradation of various organic pollutants. J. Environ. Chem. Eng. 2022, 10, 108600. [Google Scholar] [CrossRef]
  78. Alotaibi, M.R.; Mahmoud, M.H.H. Promptness of tetracycline pollutant degradation via CuCo2O4@ZrO2 nanocomposites photocatalyst. Opt. Mater. 2022, 126, 112200. [Google Scholar] [CrossRef]
  79. Abdi, J.; Yahyanezhad, M.; Sakhaie, S.; Vossoughi, M.; Alemzadeh, I. Synthesis of porous TiO2/ZrO2 photocatalyst derived from zirconium metal organic framework for degradation of organic pollutants under visible light irradiation. J. Environ. Chem. Eng. 2019, 7, 103096. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of crystal structures of ZrO2 (a) monoclinic ZrO2, (b) tetragonal ZrO2, and (c) cubic ZrO2, reprinted with permission from Elsevier (License NO. 5332051438018) [34].
Figure 1. Schematic representation of crystal structures of ZrO2 (a) monoclinic ZrO2, (b) tetragonal ZrO2, and (c) cubic ZrO2, reprinted with permission from Elsevier (License NO. 5332051438018) [34].
Catalysts 12 01418 g001
Figure 2. (a) Systematic diagram of pure ZrO2 by Xcrysden package, reprinted with permission from Elsevier (License NO. 5332600596921) [32] (b) band structure and total density of state of ZrO2. (c) Partial density of states (PDOS) and total density of states (TDOS) of ZrO2, (d) PDOS and TDOS of pure ZrO2, reprinted with permission from Elsevier (License NO. 5332600898565) [43].
Figure 2. (a) Systematic diagram of pure ZrO2 by Xcrysden package, reprinted with permission from Elsevier (License NO. 5332600596921) [32] (b) band structure and total density of state of ZrO2. (c) Partial density of states (PDOS) and total density of states (TDOS) of ZrO2, (d) PDOS and TDOS of pure ZrO2, reprinted with permission from Elsevier (License NO. 5332600898565) [43].
Catalysts 12 01418 g002
Figure 3. SEM images of (a) ZrO2: Mg2+ 2 mol%, reprinted with permission from Elsevier (License NO. 5332611047818) [45], (b) Co-doped ZrO2, (c) photodegradation of MB of 10% by Co doped ZrO2 in visible light exposure, reprinted with permission from Elsevier (License NO. 5332971176132) [25], (d) band structure of pure ZrO2, and (e) ZrO2−xNx [46].
Figure 3. SEM images of (a) ZrO2: Mg2+ 2 mol%, reprinted with permission from Elsevier (License NO. 5332611047818) [45], (b) Co-doped ZrO2, (c) photodegradation of MB of 10% by Co doped ZrO2 in visible light exposure, reprinted with permission from Elsevier (License NO. 5332971176132) [25], (d) band structure of pure ZrO2, and (e) ZrO2−xNx [46].
Catalysts 12 01418 g003
Figure 4. (a) Systematic representation of TC-H degradation in g-C3N4/ZrO2−x NTs, reprinted with permission from Elsevier (License NO. 5332980412880) [55]. (b) The plausible photocatalytic mechanism of g-C3N4/ZrO2 heterostructure, reproduced with permission from Elsevier (License NO. 5332980598897) [26]. (c) Raman spectra of pure ZrO2 and Co doped ZrO2 NPs, reprinted with permission from Elsevier (License NO. 5332971176132) [25]. (d) XRD analysis of pure ZnO and Zr-doped ZnO, reprinted with permission from Elsevier (License NO. 5332980782041) [22].
Figure 4. (a) Systematic representation of TC-H degradation in g-C3N4/ZrO2−x NTs, reprinted with permission from Elsevier (License NO. 5332980412880) [55]. (b) The plausible photocatalytic mechanism of g-C3N4/ZrO2 heterostructure, reproduced with permission from Elsevier (License NO. 5332980598897) [26]. (c) Raman spectra of pure ZrO2 and Co doped ZrO2 NPs, reprinted with permission from Elsevier (License NO. 5332971176132) [25]. (d) XRD analysis of pure ZnO and Zr-doped ZnO, reprinted with permission from Elsevier (License NO. 5332980782041) [22].
Catalysts 12 01418 g004
Figure 5. (a) HRTEM images and SAED pattern of ZrO2/ PPy composite, reproduced with permission from Elsevier (License NO. 5333050908229) [53], (b) SEM images of calcined ZrO2-ZnCo2O4 at 650 °C, reprinted with permission from Elsevier (License NO. 5333051169811) [61], (c) FTIR spectrum of g-C3N4, ZrO2 NTs and g-C3N4/ZrO2−x NTs, reprinted with permission from Elsevier (License NO. 5332980412880) [55], (d) UV-vis-DRS spectrum of Bi2O3-ZrO2, Bi2O3, and ZrO2, reproduced with permission from Elsevier (License NO. 5333060789631) [64], and (e) XRD pattern of 30 wt% ZrO2/g-C3N4 [65].
Figure 5. (a) HRTEM images and SAED pattern of ZrO2/ PPy composite, reproduced with permission from Elsevier (License NO. 5333050908229) [53], (b) SEM images of calcined ZrO2-ZnCo2O4 at 650 °C, reprinted with permission from Elsevier (License NO. 5333051169811) [61], (c) FTIR spectrum of g-C3N4, ZrO2 NTs and g-C3N4/ZrO2−x NTs, reprinted with permission from Elsevier (License NO. 5332980412880) [55], (d) UV-vis-DRS spectrum of Bi2O3-ZrO2, Bi2O3, and ZrO2, reproduced with permission from Elsevier (License NO. 5333060789631) [64], and (e) XRD pattern of 30 wt% ZrO2/g-C3N4 [65].
Catalysts 12 01418 g005
Figure 6. (a) Photodegradation pathway of indigo carmine dye under sunlight exposure for ZrO2/ZnFe2O4 composite, reprinted with permission from Elsevier (License NO. 5333060926293) [31], (b) schematic representation of charge carriers’ separation on 30 wt% ZrO2/g-C3N4 composites under light illumination [65], and (c) systematic presentation of photocarriers transfer in Bi2O3-ZrO2 hybrid in presence of visible light, reproduced with permission from Elsevier (License NO. 5333060789631) [64].
Figure 6. (a) Photodegradation pathway of indigo carmine dye under sunlight exposure for ZrO2/ZnFe2O4 composite, reprinted with permission from Elsevier (License NO. 5333060926293) [31], (b) schematic representation of charge carriers’ separation on 30 wt% ZrO2/g-C3N4 composites under light illumination [65], and (c) systematic presentation of photocarriers transfer in Bi2O3-ZrO2 hybrid in presence of visible light, reproduced with permission from Elsevier (License NO. 5333060789631) [64].
Catalysts 12 01418 g006
Table 1. Summary of various ZrO2-based semiconductor photocatalysts for the incremented photocatalytic degradation of aqueous phase pollutants.
Table 1. Summary of various ZrO2-based semiconductor photocatalysts for the incremented photocatalytic degradation of aqueous phase pollutants.
PhotocatalystEnhancement StrategiesReaction ParameterTargeted PollutantPhotocatalyst EfficiencyRef.
Ni doped ZrO2DopingVisible light lamp (>400 nm), 15 mg photocatalystMethylene blue90.2% in 100 min.[59]
C-doped ZrO2DopingPL-L lamp, 0.2 g/L photocatalystMethylene blue75%[70]
Nd doped ZrO2DopingpH = 7Methylene blue, Rhodamine B, and acetophenone90%, 77%, and 60%[71]
Al2O3/ZrO2HeterojunctionVisible light, 0.04 g catalystReactive blue 222 and Reactive yellow 14591.4% and 94.6% in 60 min.[72]
RE/ZrO2 (RE = Sm, Eu)Heterojunction350 W Xenon lampMethylene blue and Rhodamine B100% in 30 min. and 96.3% in 90 min.[73]
ZnO QDs@ZrO2-TiO2HeterojunctionUltra violet lightCongo Red94.62%[74]
ZrO2/Dy2O3HeterojunctionXenon lamp with cut-off UV filterRhodamine B, and Methylene blue100% in 30 min. and 87.79%[75]
TiO2-ZrO2Heterojunction125 W Mercury lamp, pH = 7.6, 10 mg/L photocatalystMetformin92% in 150 min.[76]
g-C3N4/ZrO2Heterojunction50 W LED lamp, 30 mg photocatalystMethylene blue, Rhodamine B, Congo Red, and Tetracycline96%, 98%, 90%, and 83%[77]
CuCo2O4@ZrO2HeterojunctionVisible lightTetracycline95%[78]
TiO2/ZrO2Heterojunction100 W LED lampRhodamine B90%[79]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rani, V.; Sharma, A.; Kumar, A.; Singh, P.; Thakur, S.; Singh, A.; Le, Q.V.; Nguyen, V.H.; Raizada, P. ZrO2-Based Photocatalysts for Wastewater Treatment: From Novel Modification Strategies to Mechanistic Insights. Catalysts 2022, 12, 1418. https://doi.org/10.3390/catal12111418

AMA Style

Rani V, Sharma A, Kumar A, Singh P, Thakur S, Singh A, Le QV, Nguyen VH, Raizada P. ZrO2-Based Photocatalysts for Wastewater Treatment: From Novel Modification Strategies to Mechanistic Insights. Catalysts. 2022; 12(11):1418. https://doi.org/10.3390/catal12111418

Chicago/Turabian Style

Rani, Vandna, Amit Sharma, Abhinandan Kumar, Pardeep Singh, Sourbh Thakur, Archana Singh, Quyet Van Le, Van Huy Nguyen, and Pankaj Raizada. 2022. "ZrO2-Based Photocatalysts for Wastewater Treatment: From Novel Modification Strategies to Mechanistic Insights" Catalysts 12, no. 11: 1418. https://doi.org/10.3390/catal12111418

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