Next Article in Journal
Recent Progresses in the Catalytic Stereoselective Dearomatization of Pyridines
Previous Article in Journal
Two-Photon Fluorescent Probes for Amyloid-β Plaques Imaging In Vivo
Previous Article in Special Issue
Assessment of the Effects of Chitosan, Chitooligosaccharides and Their Derivatives on Lemna minor
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Catalytic Activity of Rare Earth Elements (REEs) in Advanced Oxidation Processes of Wastewater Pollutants: A Review

by
Lorenzo Saviano
1,†,
Antonios Apostolos Brouziotis
1,2,†,
Edith Guadalupe Padilla Suarez
1,
Antonietta Siciliano
1,*,
Marisa Spampinato
1,3,
Marco Guida
1,3,
Marco Trifuoggi
2,4,
Donatella Del Bianco
1,
Maurizio Carotenuto
5,
Vincenzo Romano Spica
6,
Giusy Lofrano
6 and
Giovanni Libralato
1
1
Department of Biology, University of Naples Federico II, 80126 Naples, Italy
2
Department of Chemical Sciences, University of Naples Federico II, 80126 Naples, Italy
3
NBFC, National Biodiversity Future Center, 90133 Palermo, Italy
4
CeSMA Advanced Metrological and Technological Service Center, University of Naples Federico II, 80126 Naples, Italy
5
Department of Chemistry and Biology “Adolfo Zambelli”, University of Salerno, 84084 Fisciano, Italy
6
Department of Movement, Human and Health Sciences, University of Rome “Foro Italico”, 00135 Rome, Italy
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2023, 28(17), 6185; https://doi.org/10.3390/molecules28176185
Submission received: 21 June 2023 / Revised: 12 August 2023 / Accepted: 18 August 2023 / Published: 22 August 2023
(This article belongs to the Special Issue Degradation of Drugs and Organic Pollutants in Wastewater)

Abstract

:
In recent years, sewage treatment plants did not effectively remove emerging water pollutants, leaving potential threats to human health and the environment. Advanced oxidation processes (AOPs) have emerged as a promising technology for the treatment of contaminated wastewater, and the addition of catalysts such as heavy metals has been shown to enhance their effectiveness. This review focuses on the use of rare earth elements (REEs) as catalysts in the AOP process for the degradation of organic pollutants. Cerium and La are the most studied REEs, and their mechanism of action is based on the oxygen vacancies and REE ion concentration in the catalysts. Metal oxide surfaces improve the decomposition of hydrogen peroxide to form hydroxide species, which degrade the organics. The review discusses the targets of AOPs, including pharmaceuticals, dyes, and other molecules such as alkaloids, herbicides, and phenols. The current state-of-the-art advances of REEs-based AOPs, including Fenton-like oxidation and photocatalytic oxidation, are also discussed, with an emphasis on their catalytic performance and mechanism. Additionally, factors affecting water chemistry, such as pH, temperature, dissolved oxygen, inorganic species, and natural organic matter, are analyzed. REEs have great potential for enhancing the removal of dangerous organics from aqueous solutions, and further research is needed to explore the photoFenton-like activity of REEs and their ideal implementation for wastewater treatment.

Graphical Abstract

1. Introduction

Over the past few decades, conventional wastewater treatment plants (WWTPs) showed low efficiencies in removing contaminants of emerging concern (CECs) [1,2,3]. CECs encompass a wide range of substances, including pharmaceuticals, personal care products, and dyes [1,4].
The chronic release of these substances in water bodies represents a serious threat to human health and ecosystems and requires urgent advancements in wastewater treatment technologies [5,6,7] focusing also on the perspective of treated wastewater reuse.
Advanced oxidation processes (AOPs) evidenced potential revolutionary solutions in the development of integrated systems for the treatment of contaminated wastewater [1,8]. The mechanism of action of AOPs relies on the generation of reactive oxygen species (ROS) with strong oxidation capabilities, such as hydroxyl (OH), sulfate (SO42−⦁), and superoxide ion radicals (O2−⦁), which quickly degrade a wide range of organic compounds in CO2, H2O, and inorganic acids [9,10,11]. The addition of catalysts, such as heavy metals, can promote the development of such radicals and improve the process efficiency [12,13,14].
AOPs, particularly Fenton oxidation, are commonly carried out under acidic conditions, leading to the consumption of significant amounts of acid and base reagents [15,16]. The effectiveness of AOPs relies on the proper dosage of reactive hydroxyl (HO) radicals, ensuring the desired level of treatment is achieved [17,18]. A major limitation of AOPs is their relatively high cost, primarily due to the use of expensive chemicals and increased energy consumption [17]. Additionally, AOPs can result in the formation of unknown recalcitrant by-products, which may be more nephrotoxic than the original compounds, posing unresolved challenges [1,7]. While AOPs are effective in scavenging reactive radicals from non-target substances, they may not be suitable for certain categories of toxic compounds that resist degradation by HO radicals.
Recently, the application of rare earth elements (REEs) in the advanced oxidation process (AOP) has gained growing interest [19]. The REEs are a group of seventeen chemical elements in the periodic table, the 15 lanthanides from lanthanum (La) to lutetium (Lu), to which scandium (Sc) and yttrium (Y) are added since the latter tend to occur in the same mineral deposits with lanthanides and exhibit similar chemical properties. REE ions have a particular electronic structure compared to transition metal ions and are gaining great importance as Lewis acids in Green Chemistry because they can be used in aqueous solvents or water without any deactivation; they can be easily recovered from the reaction medium by extraction and can be used several times without loss of activity [19,20]. The redox properties of REEs depend on the arrangement of the surfaces and the shape and size of their crystals, and, for this reason, they can be improved by checking the nanostructure [21].
REEs exhibit unique physicochemical properties, such as high surface area, reactivity, and redox potential [22]. These properties make REEs suitable to be used in AOPs, where they can enhance the performance of the processes [23,24]. REEs can be used as catalysts to promote the generation of reactive oxygen species (ROS) when exposed to UV radiation in photocatalysis processes [25], to catalyze Fenton-like processes [26], and to enhance the efficiency of ozonation by increasing the production of hydroxyl radicals [27].
To date, most applications are still at lab scale and an increasing number of studies are being performed to fill the existing gaps in the usage of REEs for polluted water treatment. The main limitations to full-scale applications deal with high energy and capital cost.
This review focuses on the use of REEs as catalysts for the degradation of CECs from wastewater, since their use, which is currently in place for wastewater treatment, lacks processing and economic sustainability.

2. Methodology of Data Collection

The REE-assisted catalytic removal of CECs has substantially grown in recent decades, as retrieved in a Scopus search (keywords: rare earth elements, catalyst, and wastewater) updated on 20 December 2022.
We have collected 60 papers dealing with the degradation of different CECs from wastewater using REEs in AOPs. As shown in Figure 1a where the number of papers issued each year on catalytic removal of contaminants in wastewater treatment using REEs have been reported, the first publication was made in 1995 [28], which means that the field is quite new. Since 2015, the number of publications has been increasing each year. Figure 1b shows the cumulative number of publications for each REE over the period of 1995–2022. Most processes were carried out by using Ce and La. Cerium has received by far most of the attention (148 publications), followed by La (44), Y (10), and Nd (10). Five REEs (Sc, Ho, Tm, Lu, and Pm) and their possible effective role in this field have never been explored.
Figure 2 presents the distribution of research papers that focused on the utilization of REE catalysts for the degradation of various types of contaminants. According to the data, pharmaceutical compounds and dyes accounted for the highest percentage of research papers, with approximately 33% and 42%, respectively. Alkaloids, herbicides, phenols, organic acids, cyanides, and other contaminants were also investigated to a lesser extent, each representing 3–8% of the papers. This distribution highlighted the varying levels of research emphasis on different categories of contaminants in relation to REE catalysts in the past.
The chemical, metallurgical, and physical properties of the REEs are governed by their electron configuration. In contrast to element Y, the lanthanides are characterized by the progressive filling of the 4f orbital [29]. Due to the low energy of these electrons, they do not directly participate in the bonding when forming compounds with other elements. As a consequence, this leads to a significant similarity in the chemical reactivity of the lanthanides. This uniform reactivity pattern is evident consistently throughout geochemical processes, rather than being unique to each individual element [29,30].
In Table 1, the essential details encompassing the atomic numbers and intricate valence shell electronic configurations pertaining to the distinctive set of REEs are provided.

3. AOP-Based REEs for CECs Removal

3.1. Pharmaceutical Degradation

Pharmaceutical toxic compounds can end up in wastewater via many routes, such as via their consumption by humans or their use in agriculture. Therefore, it is essential to discover effective ways for decomposing these hazardous compounds in wastewater. The use of REEs has been shown by several studies to be potentially useful. Numerous researchers utilized REEs in various forms as heterogeneous catalysts in AOPs (Table 1).
Nie et al. [31] carried out a Fenton-like reaction, at 30 °C, to take complete degradation of the antibiotic sulfamethoxazole (SMX), in the simultaneous presence of LaFeO3 and H2O2 at pH values ranging between 5.5 and 7.14. It was found that the SMX removal efficiency increased with the increase of the catalyst load and then reached a constant value when the LaFeO3 concentration was above 1.4 g/L. Furthermore, it was investigated that the LaFeO3 could be reused for at least 10 cycles and the reused catalyst kept the catalytic activity nearly as efficient as the fresh one. Ref. [32] investigated the photocatalytic activity for degradation of carbamazepine and caffeine using MIL-125-NH2, LaFeO3, and LaFeO3/MIL-125-NH2 photocatalysts under solar simulator at ambient temperatures. LaFeO3/MIL-125-NH2 composite exhibited higher carbamazepine and caffeine degradation than MIL-125-NH2 MOF and LaFeO3 perovskite and this is highlighted through the calculation of pseudo-first-order kinetics equation. LaFeO3/MIL-125-NH2 composite photocatalysts prepared by the self-assembly method have shown excellent and promising results in the degradation of pharmaceutical compounds.
Chen et al. (2022) [33] used LaFeO3/lignin-biochar (LFO/LG) catalysts prepared by a sol–gel pyrolysis to evaluate the degradation of ofloxacin (OFX) under visible light irradiation. The pure LaFeO3 (LFO) and the pure lignin-biochar (LG) were tested and prepared under the same conditions to compare which one had the best performance. The whole process was operated at a constant temperature in the presence of H2O2 (30 wt%) and all experiments were carried out under different initial pH, revealing later an optimal photodegradation efficiency at pH values above 6. In addition, the morphology, microstructure, energy band structure, and photoelectrochemical behavior of the composite catalysts were characterized and the relationship with catalytic performance was explained. Pure LFO has almost no adsorption of OFX when the system was left in the dark to reach a saturation adsorption state and the degradation efficiency for LFO is only 53.4% after a 75-min degradation reaction. Instead, LFO-LG samples showed a significant adsorption effect on OFX, and the degradation efficiency was also improved up to 95.6%.
In the study of [34], the electrochemical degradation of dipyrone using cerium catalysts was not effective during a two-hour electrolysis. However, the use of a CeO2/C gas diffusion electrode degraded all of the dipyrone in 20 min with 26% mineralization at −1.3 V, and after only 5 min with 57% mineralization at −1.1 V when the Fenton process was employed. The authors declared that ceria acts as an oxygen buffer leading to an increase in the local oxygen concentration, facilitating H2O2 formation and consequently improving the dipyrone degradation. FeCeOx can remove 83% of diclofenac in a heterogenous Fenton process and can be reused due to its chemical stability [35].
A heterogeneous Fenton catalyst, Ce0-Fe0-reduced graphene oxide (Ce–Fe–RGO), showed good catalytic performance and adsorption of sulfamethazine [13]. Ferrum-doped CeO2 nanosheets with various Fe/Ce ratios were evaluated for the Fenton-like degradation of salicylic acid (SA). The 2% Fe-doped CeO2 nanosheets showed the highest concentration of surface oxygen vacancies and catalytic activity under the optimum reaction conditions due to the complex adsorption of SA and H2O2 on the surface Ce3+. The catalytic activity of the nanosheets could be recovered, and so they could be reused in new cycles [36]. A surface of graphite felt loaded with CeOx accelerated the mineralization of carbamazepine via an E-peroxone process [37].
The Fe0/CeO2 nanocomposite enhanced the removal of tetracycline (TC) in the Fenton oxidation process and could be reused for further cycles. The authors suggested a synergistic effect between nanoscale Fe0 and CeO2 [19].
Hu et al. [38] tried out for the first time the use of cobalt (Co) and REE gadolinium (Gd) to investigate the removal of two antibiotics from wastewater, ciprofloxacin (CIP) and tetracycline. In particular, the sponge-like structure of Co- and Gd-modified biochar (MBC) was conducive to studying its adsorption capacity due to the presence of metal oxides and functional groups. Furthermore, the influence of pH on the absorption of antibiotics was studied by varying the pH between 3 and 11. It was observed that at pH 9, the antibiotic adsorption capacity of MBC was greatly increased. These data suggested that MBC is an innovative and effective adsorbent for the removal of antibiotics from complex contaminated water.
Luan et al. [39] synthesized photocatalysts Er2FeSbO7, BiTiSbO6, or N-TO by solid-state method to degrade enrofloxacin (ENR), a common antibiotic able to effectively kill Gram-positive and Gram-negative bacteria and mycoplasma under visible light irradiation. A new photocatalyst provided by Er2FeSbO7/BiTiSbO6 heterojunction (EBH) was prepared for the first time by the thermal solvent method and tested. It was observed that the photocatalytic activity among the four photocatalysts was as follows: EBHP > Er2FeSbO7 > BiTiSbO6 > N-TO. Therefore, it can be concluded that the EBHP process can be a potent method for treating pharmaceutical wastewater that is polluted by ENR.
In general, the studies highlighted the use of diverse catalysts to improve the efficiency of removing pharmaceutical compounds. Each study focused on targeting specific pharmaceutical compounds; this targeted approach allowed for a more accurate evaluation of the catalyst’s performance against particular contaminants. Among the studies conducted on CIP removal, the highest removal percentage (>99%) within a timeframe of 180 min was achieved by the study utilizing the MBC catalyst [38].
The pH can play a significant role in the reactivity and adsorption behavior of both the catalyst and the target compound. The studies reported a range of pH values, including pH 4, 5, 6.48, and 9, indicating that different pH levels can be effective depending on the specific catalyst and target compound.
Furthermore, the temperature of the reaction can influence the reaction kinetics and the stability of the catalyst. It was generally observed that moderate temperatures (20–25 °C) were preferred to strike a balance between reaction rates and energy consumption.
The pharmaceutical degradation studies highlighted the potential of Fenton-based processes for wastewater treatment, showcasing the reusability of FeCeOx [33], the catalytic performance and adsorption properties of Ce-Fe-RGO [13], and the degradation capabilities of ferrum-doped CeO2 nanosheets [19]. The results emphasize the versatility and effectiveness of Fenton processes in removing pharmaceutical compounds, offering promising solutions for the treatment of wastewater contaminated with these contaminants.

3.2. Dye Degradation

Wastewater from the textile industry contains residual dyes which are hardly biodegradable. Advanced Oxidation Processes (AOPs) are alternative techniques for the destruction of dyes and many other organic products in wastewater and effluents. In Table 2, the degradation efficiency of various dyes using REE catalysts is highlighted, including the type of contaminant, initial concentration, reaction conditions, and the corresponding degradation percentage achieved by the REE catalysts.
The presence of La in the La–Fe montmorillonite (La–Fe MMT) composite enhanced the degradation efficiency of the organic dyes Rhodamine B (RhB) and methylene blue (MB) and resulted in a little iron leaching, consequently increasing the lifetime of the system. Due to its low cost, efficient reactivity, high stability, and low metal leaching, it seems a great potential as a green catalyst for the heterogeneous Fenton-like degradation of hazardous dyes [40]. FeOCl doped with Y or La showed efficient Fenton catalytic activity for ibuprofen degradation under simulated solar light [41]. Tungsten oxide composites (WO3) doped with La, Gd, or Er showed an enhanced photocatalytic degradation of organic dyes compared to the pure composites [42].
Huang et al. (2014) [43] doped hydroxyl FeAl intercalated montmorillonite with La or Ce and saw an improvement in the photo-Fenton catalytic oxidation of Reactive Blue 19 under natural sunlight irradiation. The photo-Fenton activity of pristine CoFe2O4 on the oxidation of five organic dyes was enhanced by inserting a very small quantity of La or Ce cations into its spinel structure in the presence of two different oxidants [44].
Divya and Renuka (2015) [45] doped nanoceria with five different elements (Cu, Fe, Zr, Dy, and La) and investigated the heterogenous Fenton-like oxidation of MB. CuO showed the highest catalytic activity, which is attributed to the synergistic effect between copper and ceria. The Cu–Mn/CeO2/SBA-15 catalysts can efficiently degrade (>99%) high-concentration dye pollutants. The excellent catalytic performance could be attributed to the well-dispersed catalytic nanoparticles inside the mesopores and the catalytic synergic effect of CeO2 [46]. Xie, Yibing, and Chunwei Yuan [47] showed that Ce4+-TiO2 sol and nanocrystallites can photodegrade the reactive brilliant red dye (X-3B) with the dye/Ce4+-TiO2/visible-light reaction system.
Singh et al. (2016) [48] prepared several Cu/zeolite Y catalysts to test the decolorization, degradation, and mineralization of recalcitrant diazo dye (Congo red) in a heterogeneous Fenton-like process. The catalyst with 7.5 wt% Cu showed the maximum degradation, decolorization, and mineralization of 93.58%, 95.34%, and 79.52% after 2.5, 2, and 4 h, respectively, along with reproducible activity for three cycles. The addition of Y on the surface of BiW2O6 composites enhanced the degradation of RhB compared to the individual composites under visible light illumination, by producing holes (h+) and superoxide radicals [49].
Seroglazova et al. (2022) [50] showed that Praseodymium orthoferrite (PrFeO3) nanopowders can achieve 100% efficiency of photo-Fenton-like degradation of methyl violet. The PrFeO3/CeO2-based nanocomposites can degrade methyl violet (MV) under the presence of visible light via a photocatalytic Fenton-like activity. These nanoparticles can be reused due to their stability and be separated from the reaction solution [51]. In the study of [52], Pr-CdWO4 nanoparticles could degrade the toxic synthetic azo dye Remazol Black B by 93.9% at optimal conditions.
Nd-doped ZnO nanoneedles can degrade MB under UV light. Especially 1% Nd can degrade 2.5 times more with respect to undoped ZnO [53]. The LaFeO3 perovskites exhibit better photocatalytic activity towards the oxidative degradation of dye molecules when doped with Nd, Eu, Gd, or Dy [54].
Keerthana et al. (2021) [55] showed that zinc ferrite (ZnFe2O4) nanoparticles doped with 2% Samarium (Sm) can remove 65% of MB dye with the help of visible light irradiation within 1 h and, due to their stability, could be reused for more than three cycles. In the study of [56], Sm-doped CeO2 nanoparticles demonstrated a very efficient photodegradation of UV-irradiated rose bengal dye. Wang et al. (2017) [57] showed that Sm-doped Bi2MoO6 photocatalysts could degrade more effectively RhB with respect to the individual particles.
The europium-nitrogen (EueN) co-doped TiO2/Sepiolite nanocomposites, prepared through microwave-hydrothermal treatment, can be used to effectively eliminate dyestuff in real wastewater treatment. In particular, in order to evaluate the degradation of an anionic dye, Orange G, the photocatalytic activities were conducted under solar irradiation. The Eu-N co-doped TiO2/Sep NCs, with various amounts of EuIII, photocatalysts exhibited improved photoactivity as compared to mono-doped and undoped samples [58]. In other studies, europium, in the +2 and +3 oxidation state, has been used as a dopant in VOC degradation over metal oxide catalysts such as TiO2 for photocatalytic oxidation of RhB and MB [59,60] and BiVO4 for photocatalytic oxidation of methyl orange [61]. EuFeO3 nanoparticles have photo-Fenton-like catalytic activity, as indicated by the efficient decolorization of RhB aqueous solution under visible-light irradiation, and they can be recovered and reused due to their excellent photocatalytic stability [62]. Xu et al. (2014) [63] showed that the addition of Eu3+ ions on the surface of Bi2WO6 composites enhances the photocatalytic degradation of RhB under visible light irradiation by producing hydroxyl radicals.
Sharmin et al. (2022) [64] synthesized 10% Gd-doped BiFeO3 nanoparticles (BGFO) via a simple and cost-effective hydrothermal technique at a lower reaction temperature of 160 °C. Following this approach, the synthesized nanoparticles exhibited photocatalytic activity towards the degradation of hazardous industrial pollutants, such as RhB and MB pharmaceutical pollutants such as ciprofloxacin (CIP) and levofloxacin (LFX) under simulated solar irradiation. The photocatalyst BGFO demonstrated 96% and 97% degradation of RhB and MB within 240 and 180 min of solar irradiation, respectively, and about 80% and 79% degradation of CIP and LFX were obtained within 240 min in the same conditions. Similarly, through the hydrothermal method, Sn-Gd2O3 nanomaterial was prepared and mixed with a hot aqueous solution (T > 60 °C) of gelatin polymer, followed by cross-linking, for the efficient reduction of water pollutants [65]. The catalytic activity of Sn-Gd2O3 NPs was analyzed against different dyes, such as 4-nitrophenol (4-NP), 2,6-dinitrophenol (2,6-DNP), 2-nitrophenol (2-NP), MB, methyl orange (MO) and congo red, the presence of a strong reducing agent, NaBH4. These latter two anionic azo dyes are used as indicators as well as in the food, textile, paper and pulp, cotton, silk, plastic, cosmetic, pharmaceutical, rubber, printing, and wool industries [66]. Therefore, this scientific approach has led to the fabrication of a new hydrogel nanocomposite for the removal of water pollutants, especially in the reduction of the activity of industrial dyes. Composites of Bi2MoO6 showed an enhanced photodegradation of RhB and TC when doped with Gd and Pt ions, by producing hydroxyl radicals, holes in the valence band (h+), and superoxide radicals [67]. Yu et al. (2016) showed that Bi2MoO6 nanoplate crystals perform a better photocatalytic degradation of RhB when doped with Gd3+ ions, with the production of hydroxyl radicals as a proposed mechanism of action [68]. Wu et al. (2019) [69] showed that the TiO2 NPs demonstrated an enhanced photocatalytic degradation of MB and RhB when doped with Gd3+ ion or Gd2O3.
Vinosha et al. (2022) [70] showed that zinc ferrite (ZnFe2O4) nanocrystals doped with dysprosium (Dy) exhibit a photo-Fenton activity against MB with an efficiency of 97.3% within 45 min. This ferrite can act as a magnetic recyclable catalyst even after five cycles with an insignificant lessening of elements.
Liang et al. (2006) [71] showed that Er3+-TiO2 catalysts had higher adsorption equilibrium constants and better adsorption capacity than pure TiO2, and the degradation and mineralization of orange I under both UV radiation and visible light were more efficient with Er3+-TiO2 catalyst than with pure TiO2.
Tikhanova et al. (2021) [72] showed that the heterojunction h-YbFeO3/o-YbFeO3 photocatalyst demonstrates an enhanced photo-Fenton decolorization of MV under visible light compared to pure o-YbFeO3. A novel developed 2.0%Yb/2.0%Er/2.0%Pr-BiW2O6 catalyst could degrade the Congo red dye, TC, RhB, and MB by 90.2%, 52.3%, 95.5%, and 64.6%, respectively, under visible light illumination. The proposed mechanism of action was the promoted reactive oxide species production and photocarrier separation [73].
The addition of Ln2O3 (Ln = Sm, Eu or Tb) on the surface of BiVO4 composites enhanced the degradation of RhB with respect to the individual composites, by producing superoxide radicals, holes (h+), and hydroxyl radicals [74]. In the study of [75], the Bi2WO6 photocatalyst showed enhanced performance in the degradation of RhD when doped with the combination of Tb/Eu, Dy/Sm, or Er/Nd. The proposed mechanism of action was the production of hydroxyl radicals, holes (h+), and superoxide radicals.
Among the catalysts discussed, the most promising results for the removal of dyes were obtained with nanoceria doped with Cu, Fe, Zr, Dy, or La, which achieved a remarkable removal efficiency of over 99% MB [45]. These catalysts demonstrated a high level of effectiveness in degrading and eliminating MB from wastewater, showcasing their potential for applications in water treatment. However, it is important to consider that the performance of catalysts can be influenced by various factors, such as reaction conditions and the specific dye compound being targeted. Optimal results may vary depending on these factors and further exploration and optimization are necessary to discover even more efficient catalysts for MB removal. By improving catalyst design, fine-tuning reaction conditions, and exploring novel materials, it may be possible to achieve even higher removal efficiencies and enhance the overall sustainability and efficiency of water treatment processes.

3.3. Degradation of Other Target Pollutants

With the rapid development of industrialization, wastewater has increasingly enriched itself with various substances which often include, in addition to those already extensively discussed above, alkaloids, herbicides, phenols, etc. Unconventional wastewater treatment methods are presented below based on photocatalysis and Fenton-like reaction using REEs for the abatement of these molecules. In the first-ever study using REEs for removing pollutants from aqueous solutions, [33] studied the use of solid compounds of Y(III), La(III), Ce(III), Nd(III), Sm(III), Gd(III), and Ce(IV) in the form of oxides, hydrous oxides and basic carbonates for the removal of fluoride ions, and only Ce(IV) was capable of removing them effectively without a significant dissolution even at pH 2.
Fouad et al. (2021) [76] tried to modify the LaFeO3 catalyst with Ti (Ti-LaFeO3) by developing photocatalysis and photo-Fenton processes under UV and visible light for the degradation of carbofuran (CBF). It was observed that about 20% of CBF was degraded by H2O2 without catalyst loading under UV irradiation. In the presence of UV light and the absence of H2O2, the CBF degradation was 59%, which was mainly due to the photocatalysis process induced by the illumination of Ti-LaFeO3. The degradation of CBF was 73% under metal-halide lamp irradiation in the system of Ti-LaFeO3/H2O2, which confirms the activity of the catalyst under visible light. Moreover, the degradation was improved to 89.5% when this lamp was replaced by UV light. Heterogeneous catalytic wet peroxide oxidation (CWPO) was proposed in this work, aiming to obtain degradation rates and faster mineralization of recalcitrant pollutants like CBF [77]. The results showed that the efficiency of the CWPO process by Ti-LaFeO3 is higher than other photocatalytic processes used in the literature for the degradation of CBF. Furthermore, the catalyst could maintain its catalytic activity for five consecutive cycles. In the study of [78], the La-doped TiO2 photocatalyst could degrade acetone and NO under visible light by 38% and 98%, respectively.
Gogoi et al. (2017) [79] used Fe3O4-CeO2 nanocomposite materials as a Fenton-like heterogeneous catalyst for the degradation of catechol. Fe3O4-CeO2 (15 wt%) showed the maximum degradation of catechol compared to the other prepared catalysts. The catalyst could also be recovered and used for five consecutive cycles with excellent catalytic activity. Cerium chloride exhibited radical production in the presence of hydrogen peroxide, as assessed by relaxation of supercoiled plasmid DNA and the production of radical cation of 2,2′-azinobis-(3-ethylbenzthiazoline-6-sulfonic acid), and so it is proposed that cerium is capable of redox-cycling with peroxide to generate damaging oxygen radicals [80]. Orge et al. (2012) [81] tested ceria and Ce-based mixed oxides (Ce-Sm, Ce-La, and Ce-Zr with different compositions) as catalysts in the ozonation of oxalic acid, aniline, and the textile dye C. I. Reactive Blue 5. CeO2-ZrO2 catalyst with higher than 25 wt.% of ZrO2 presented the best catalytic oxidation of oxalic acid, which may be related to the larger percentage of the Ce (III) species on their surfaces. On the other hand, the ozonation catalyzed by 25 wt.% of Zr, Sm, or La oxides with Ce oxide or by Ce oxide did not present significant differences. In addition, the association of O3 and all mentioned catalysts leads to an improved TOC removal from the dye and almost complete mineralization in all cases after about two hours. The highest mineralization rate of aniline was achieved by CeO2 and Ce0.75Zr0.25O2 with the latter being the sample with the largest amount of oxygen vacancies number in its structure. Cerium nanotubes catalysts (CeNTs) doped with different amounts of Nb2O5 (0–10 wt.%) were used for the catalytic reduction of NOx with NH3 (NH3-SCR) in the presence of CH2Cl2. The 10 wt.% Nb-CeNTs catalyst presented the best oxidation of NH3-SCR and degradation efficiency of CH2Cl2 among the prepared catalysts, due to its abundant surface oxygen species and acid sites, superior redox capability, the interaction between Nb and Ce, higher ratio of Nb4+/(Nb5++ Nb4+) and Ce3+/(Ce3++ Ce4+), and the special tubular structure of the Ce nanotube [82].
A magnetic nanoscaled Fe3O4/CeO2 composite can degrade 4-chlorophenol and be reused for six successive cycles [83].
A ZnO nanoparticle doped with 2% Ce can catalyze the oxidation of cyanide and subsequently of cyanate as well under UV-A light or natural sunlight [84].
Sharma et al. (2016) [44] doped cobalt ferrite (CoFe2O4) with La or Ce and observed an enhancement in its photo-Fenton degradation properties of various organic pollutants using two different inorganic oxidants: hydrogen peroxide (H2O2) and potassium peroxymonosulphate (KHSO5), with respect to pure cobalt ferrite. The doped catalyst could also be reused in wastewater treatment without a decline in its catalytic activity due to its chemical stability.
Samarium-doped ZnO nanorods (Sm/ZNRs) can exhibit a higher photocatalytic degradation of phenol aqueous solution under visible light irradiation than that of the pure ZNRs due to their high charge separation efficiency and OH generation ability. These nanorods could also be reused for more cycles [85].
Cymes et al. (2020) [86] doped cryptomelane with europium via either co-precipitation or ion exchange in order to promote the catalytic oxidation of ethanol. Eu-doping by co-precipitation generally improves catalytically advantageous physical and chemical properties whereas Eu-doping by ion exchange shows deleterious effects.
Composites of Bi2MoO6 showed an enhanced photodegradation of 4-Chlorophenol when doped with Gd and Pt ions, by producing hydroxyl radicals, holes in the valence band (h+), and superoxide radicals [49]. In the study of [63], the Bi2WO6 photocatalyst showed enhanced performance in the degradation of phenol when doped with the combination of Tb/Eu, Dy/Sm, or Er/Nd. The proposed mechanism of action was the production of hydroxyl radicals, holes (h+), and superoxide radicals.
Table 2 shows which REEs have been studied for photocatalysis and which for Fenton-like catalysis.
Table 3, Table 4 and Table 5 below summarize all the catalyst-doped REEs divided and ordered according to publication date.

4. Potential Environmental Impacts of REE Catalysts

As for the use of REEs in AOPs, different perspectives are to be considered. Further improvement needs to be implemented in water treatments; as of now, several methods have been adopted that increase efficiency and provide better handling of waste, such as Fenton-like oxidation processes [87]. Studies focused on the use of lanthanides have demonstrated an improvement in wastewater treatments, as their use can produce on average 30% less sludge compared to other alternatives such as iron and aluminum [88]. The use of REEs could represent an alternative for reducing wastelands generated by sludge [88] and improving the removal of organic pollutants by enhancing photocatalytic activity [23].
Moreover, care should be taken to avoid the loss and transfer of particles into the environment, contributing to the anthropogenic input of REEs. The authors of [89] found that the mixture of lanthanides used in wastewater treatment had an antagonistic behavior, which could reduce their environmental impact. However, up to now, the knowledge of the toxicity of REEs is very limited, as the environmental fate is highly dependent on the system, such as the presence of organic matter, pH, and presence of cations [90] which influence the speciation, which, by extension influences the bioaccumulation and bioavailability. Because of the lack of knowledge regarding the potential effects of REEs, the regulation for the discharge of REEs is still lacking. In contrast, many restrictions regarding the presence of organic pollutants have been applied globally [91].

5. Conclusions

The study of REEs for their potential use in removing toxic organic compounds from aqueous solutions is a relatively recent field, with the first publication on this topic occurring only 28 years ago. Among the REEs, cerium has been extensively studied for its photo Fenton-like catalytic activity against organic pollutants, followed by lanthanum. However, the remaining REEs have received little attention, with five of them having never been studied in this context.
REEs have been shown to be promising for removing hazardous organic compounds from aqueous solutions, such as wastewater. The effectiveness of REEs in this regard depends mainly on the presence of oxygen vacancies and the concentration of REE ions in the catalysts. Specifically, the metal oxide surfaces promote the decomposition of hydrogen peroxide, which leads to the formation of hydroxide species responsible for the degradation of organic compounds.
Moreover, AOPs, which include REE-combined catalysts and other parameters, should be chosen based on the nature of the contaminants and the acid-base properties of the wastewater to be treated.
Due to their exceptional properties, REEs have the potential to revolutionize wastewater treatment. However, achieving the production of clean water without generating traces of persistent toxins remains a challenging goal.

Author Contributions

L.S.: Conceptualization, Data Curation, Writing—original draft. A.A.B.: Conceptualization, Data Curation, Writing—original draft. E.G.P.S.: Writing—original draft. A.S.: Data Curation, Writing—original draft, Supervision. M.S.: Data Curation. M.G.: Funding acquisition, data curation, and writing—review & editing. M.T.: Funding acquisition, data curation, and writing—review & editing. D.D.B.: Data curation. M.C.: Data Curation—original draft revision. V.R.S.: Conceptualization—data curation. G.L. (Giusy Lofrano): Supervision, Writing—original draft. G.L. (Giovanni Libralato): Writing—original draft, Supervision, Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

Funding for this project has been provided through multiple sources including the European Union’s Horizon 2020 research and innovation program under the Marie Skłodowska-Curie grant agreement No. 857989. Additionally, support has been received from NBFC to the University of Palermo Institute, which is funded by the Italian Ministry of University and Research, PNRR, Missione 4 Componente 2, “Dalla ricerca all’impresa”, Investimento 1.4, Project CN00000033.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors state that the research was carried out without any potential conflicts of interest arising from commercial or financial relationships.

References

  1. Lofrano, G.; Libralato, G.; Casaburi, A.; Siciliano, A.; Iannece, P.; Guida, M.; Pucci, L.; Dentice, E.; Carotenuto, M. Municipal wastewater spiramycin removal by conventional treatments and heterogeneous photocatalysis. Sci. Total Environ. 2018, 624, 461–469. [Google Scholar] [CrossRef]
  2. Priya, A.K.; Gnanasekaran, L.; Rajendran, S.; Qin, J.; Vasseghian, Y. Occurrences and removal of pharmaceutical and personal care products from aquatic systems using advanced treatment-A review. Environ. Res. 2022, 204, 112298. [Google Scholar] [CrossRef] [PubMed]
  3. Rout, P.R.; Zhang, T.C.; Bhunia, P.; Surampalli, R.Y. Treatment technologies for emerging contaminants in wastewater treatment plants: A review. Sci. Total Environ. 2021, 753, 141990. [Google Scholar] [CrossRef] [PubMed]
  4. Lofrano, G.; Libralato, G.; Carotenuto, M.; Guida, M.; Inglese, M.; Siciliano, A.; Meriç, S. Emerging concern from short-term textile leaching: A preliminary ecotoxicological survey. Bull. Environ. Contam. Toxicol. 2016, 97, 646–652. [Google Scholar] [CrossRef] [PubMed]
  5. Ugwu, E.I.; Karri, R.R.; Nnaji, C.C.; John, J.; Padmanaban, V.; Othmani, A.; Ikechukwu, E.L.; Helal, W.M. Application of green nanocomposites in removal of toxic chemicals, heavy metals, radioactive materials, and pesticides from aquatic water bodies. In Sustainable Nanotechnology for Environmental Remediation; Elsevier: Amsterdam, The Netherlands, 2022; pp. 321–346. [Google Scholar]
  6. Karri, R.R.; Ravindran, G.; Dehghani, M.H. Wastewater—Sources, toxicity, and their consequences to human health. In Soft Computing Techniques in Solid Waste and Wastewater Management; Elsevier: Amsterdam, The Netherlands, 2021; pp. 3–33. [Google Scholar]
  7. Pedrazzani, R.; Bertanza, G.; Brnardić, I.; Cetecioglu, Z.; Dries, J.; Dvarionienė, J.; García-Fernández, A.J.; Langenhoff, A.; Libralato, G.; Lofrano, G. Opinion paper about organic trace pollutants in wastewater: Toxicity assessment in a European perspective. Sci. Total Environ. 2019, 651, 3202–3221. [Google Scholar] [CrossRef] [PubMed]
  8. Werkneh, A.A.; Rene, E.R. Applications of nanotechnology and biotechnology for sustainable water and wastewater treatment. In Water and Wastewater Treatment Technologies; Springer: Berlin/Heidelberg, Germany, 2019; pp. 405–430. [Google Scholar]
  9. 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]
  10. Spasiano, D.; Siciliano, A.; Race, M.; Marotta, R.; Guida, M.; Andreozzi, R.; Pirozzi, F. Biodegradation, ecotoxicity and UV254/H2O2 treatment of imidazole, 1-methyl-imidazole and N, N’-alkyl-imidazolium chlorides in water. Water Res. 2016, 106, 450–460. [Google Scholar] [CrossRef] [PubMed]
  11. Lofrano, G.; Rizzo, L.; Grassi, M.; Belgiorno, V. Advanced oxidation of catechol: A comparison among photocatalysis, Fenton and photo-Fenton processes. Desalination 2009, 249, 878–883. [Google Scholar] [CrossRef]
  12. Wang, S. A comparative study of Fenton and Fenton-like reaction kinetics in decolourisation of wastewater. Dye Pigment 2008, 76, 714–720. [Google Scholar] [CrossRef]
  13. Wang, N.; Zheng, T.; Zhang, G.; Wang, P. A review on Fenton-like processes for organic wastewater treatment. J. Environ. Chem. Eng. 2016, 4, 762–787. [Google Scholar] [CrossRef]
  14. Kuang, J.C. Characteristics of Electron Structure and 4f Orbital of Rare Earths and Their Reinforcement of Wastewater Degradation. Adv. Mater. Res. 2014, 1022, 72–75. [Google Scholar] [CrossRef]
  15. Atalay, S.; Ersöz, G. Novel Catalysts in Advanced Oxidation of Organic Pollutants; Springer: Berlin/Heidelberg, Germany, 2016. [Google Scholar]
  16. Acharya, B.K.; Pathak, H.; Mohana, S.; Shouche, Y.; Singh, V.; Madamwar, D. Kinetic modelling and microbial community assessment of anaerobic biphasic fixed film bioreactor treating distillery spent wash. Water Res. 2011, 45, 4248–4259. [Google Scholar] [CrossRef] [PubMed]
  17. Kumar, V.; Shah, M.P. Advanced oxidation processes for complex wastewater treatment. In Advanced Oxidation Processes for Effluent Treatment Plants; Elsevier: Amsterdam, The Netherlands, 2021; pp. 1–31. [Google Scholar]
  18. Oturan, M.A.; Aaron, J.-J. Advanced oxidation processes in water/wastewater treatment: Principles and applications. A review. Crit. Rev. Environ. Sci. Technol. 2014, 44, 2577–2641. [Google Scholar] [CrossRef]
  19. Zhang, N.; Chen, J.; Fang, Z.; Tsang, E.P. Ceria accelerated nanoscale zerovalent iron assisted heterogenous Fenton oxidation of tetracycline. Chem. Eng. J. 2019, 369, 588–599. [Google Scholar] [CrossRef]
  20. Oliverio, M.; Nardi, M.; Costanzo, P.; Di Gioia, M.L.; Procopio, A. Erbium salts as non-toxic catalysts compatible with alternative reaction media. Sustainability 2018, 10, 721. [Google Scholar] [CrossRef]
  21. Zinatloo-Ajabshir, S.; Emsaki, M.; Hosseinzadeh, G. Innovative construction of a novel lanthanide cerate nanostructured photocatalyst for efficient treatment of contaminated water under sunlight. J. Colloid Interface Sci. 2022, 619, 1–13. [Google Scholar] [CrossRef] [PubMed]
  22. Ambaye, T.G.; Vaccari, M.; Castro, F.D.; Prasad, S.; Rtimi, S. Emerging technologies for the recovery of rare earth elements (REEs) from the end-of-life electronic wastes: A review on progress, challenges, and perspectives. Environ. Sci. Pollut. Res. 2020, 27, 36052–36074. [Google Scholar] [CrossRef] [PubMed]
  23. Chan, S.H.S.; Yeong Wu, T.; Juan, J.C.; Teh, C.Y. Recent developments of metal oxide semiconductors as photocatalysts in advanced oxidation processes (AOPs) for treatment of dye waste-water. J. Chem. Technol. Biotechnol. 2011, 86, 1130–1158. [Google Scholar] [CrossRef]
  24. Bokare, A.D.; Choi, W. Review of iron-free Fenton-like systems for activating H2O2 in advanced oxidation processes. J. Hazard. Mater. 2014, 275, 121–135. [Google Scholar] [CrossRef]
  25. Onozuka, K.; Kawakami, Y.; Imai, H.; Yokoi, T.; Tatsumi, T.; Kondo, J. Perovskite-type La2Ti2O7 mesoporous photocatalyst. J. Solid State Chem. 2012, 192, 87–92. [Google Scholar] [CrossRef]
  26. Assila, O.; Barros, Ó.; Fonseca, A.M.; Parpot, P.; Soares, O.S.; Pereira, M.F.; Zerrouq, F.; Kherbeche, A.; Rombi, E.; Tavares, T. Degradation of pollutants in water by Fenton-like oxidation over LaFe-catalysts: Optimization by experimental design. Microporous Mesoporous Mater. 2023, 349, 112422. [Google Scholar] [CrossRef]
  27. Song, J.; Ma, N.; Chen, W.; Chen, J.; Dai, Q. Insights into mechanism of catalytic ozonation of cinnamyl alcohol over core–shell Fe3O4@SiO2@La2O3 catalyst. Sep. Purif. Technol. 2022, 282, 119969. [Google Scholar] [CrossRef]
  28. Tokunaga, S.; Haron, M.J.; Wasay, S.A.; Wong, K.F.; Laosangthum, K.; Uchiumi, A. Removal of fluoride ions from aqueous solutions by multivalent metal compounds. Int. J. Environ. Stud. 1995, 48, 17–28. [Google Scholar] [CrossRef]
  29. Liu, S. Electronic structure of rare earth metals. Handb. Phys. Chem. Rare Earths 1978, 1, 233–335. [Google Scholar]
  30. Been, E.; Lee, W.-S.; Hwang, H.Y.; Cui, Y.; Zaanen, J.; Devereaux, T.; Moritz, B.; Jia, C. Electronic structure trends across the rare-earth series in superconducting infinite-layer nickelates. Phys. Rev. X 2021, 11, 011050. [Google Scholar] [CrossRef]
  31. Nie, Y.; Zhang, L.; Li, Y.-Y.; Hu, C. Enhanced Fenton-like degradation of refractory organic compounds by surface complex formation of LaFeO3 and H2O2. J. Hazard. Mater. 2015, 294, 195–200. [Google Scholar] [CrossRef] [PubMed]
  32. Younes, H.A.; Taha, M.; Khaled, R.; Mahmoud, H.M.; Abdelhameed, R.M. Perovskite/metal-organic framework photocatalyst: A novel nominee for eco-friendly uptake of pharmaceuticals from wastewater. J. Alloys Compd. 2023, 930, 167322. [Google Scholar] [CrossRef]
  33. Chen, X.; Zhang, M.; Qin, H.; Zhou, J.; Shen, Q.; Wang, K.; Chen, W.; Liu, M.; Li, N. Synergy effect between adsorption and heterogeneous photo-Fenton-like catalysis on LaFeO3/lignin-biochar composites for high efficiency degradation of ofloxacin under visible light. Sep. Purif. Technol. 2022, 280, 119751. [Google Scholar] [CrossRef]
  34. Assumpção, M.H.M.T.; Moraes, A.; De Souza, R.; Reis, R.M.; Rocha, R.S.; Gaubeur, I.; Calegaro, M.L.; Hammer, P.; Lanza, M.R.d.V.; Santos, M.C.d. Degradation of dipyrone via advanced oxidation processes using a cerium nanostructured electrocatalyst material. Appl. Catal. A Gen. 2013, 462, 256–261. [Google Scholar] [CrossRef]
  35. Chong, S.; Zhang, G.; Zhang, N.; Liu, Y.; Zhu, J.; Huang, T.; Fang, S. Preparation of FeCeOx by ultrasonic impregnation method for heterogeneous Fenton degradation of diclofenac. Ultrason. Sonochem. 2016, 32, 231–240. [Google Scholar] [CrossRef]
  36. Wang, W.; Zhu, Q.; Qin, F.; Dai, Q.; Wang, X. Fe doped CeO2 nanosheets as Fenton-like heterogeneous catalysts for degradation of salicylic acid. Chem. Eng. J. 2018, 333, 226–239. [Google Scholar] [CrossRef]
  37. Wang, X.; Jin, Y.; Chen, W.; Zou, R.; Xie, J.; Tang, Y.; Li, X.; Li, L. Electro-catalytic activity of CeOx modified graphite felt for carbamazepine degradation via E-peroxone process. Front. Environ. Sci. Eng. 2021, 15, 122. [Google Scholar] [CrossRef]
  38. Hu, B.; Tang, Y.; Wang, X.; Wu, L.; Nong, J.; Yang, X.; Guo, J. Cobalt-gadolinium modified biochar as an adsorbent for antibiotics in single and binary systems. Microchem. J. 2021, 166, 106235. [Google Scholar] [CrossRef]
  39. Luan, J.; Liu, W.; Yao, Y.; Ma, B.; Niu, B.; Yang, G.; Wei, Z. Synthesis and Property Examination of Er2FeSbO7/BiTiSbO6 Heterojunction Composite Catalyst and Light-Catalyzed Retrogradation of Enrofloxacin in Pharmaceutical Waste Water under Visible Light Irradiation. Materials 2022, 15, 5906. [Google Scholar] [CrossRef] [PubMed]
  40. Fida, H.; Zhang, G.; Guo, S.; Naeem, A. Heterogeneous Fenton degradation of organic dyes in batch and fixed bed using La-Fe montmorillonite as catalyst. J. Colloid Interface Sci. 2017, 490, 859–868. [Google Scholar] [CrossRef] [PubMed]
  41. Shi, X.; Cui, C.; Zhang, L.; Zhang, J.; Liu, G. FeOCl/Ln (Ln = La or Y): Efficient photo-Fenton catalysts for ibuprofen degradation. New J. Chem. 2019, 43, 16273–16280. [Google Scholar] [CrossRef]
  42. Tahir, M.B.; Sagir, M. Carbon nanodots and rare metals (RM = La, Gd, Er) doped tungsten oxide nanostructures for photocatalytic dyes degradation and hydrogen production. Sep. Purif. Technol. 2019, 209, 94–102. [Google Scholar] [CrossRef]
  43. Huang, Z.; Wu, P.; Li, H.; Li, W.; Zhu, Y.; Zhu, N. Synthesis and catalytic properties of La or Ce doped hydroxy-FeAl intercalated montmorillonite used as heterogeneous photo Fenton catalysts under sunlight irradiation. RSC Adv. 2014, 4, 6500–6507. [Google Scholar] [CrossRef]
  44. Sharma, R.; Bansal, S.; Singhal, S. Augmenting the catalytic activity of CoFe2O4 by substituting rare earth cations into the spinel structure. RSC Adv. 2016, 6, 71676–71691. [Google Scholar] [CrossRef]
  45. Divya, T.; Renuka, N. Modulated heterogeneous Fenton-like activity of ‘M’doped nanoceria systems (M = Cu, Fe, Zr, Dy, La): Influence of reduction potential of doped cations. J. Mol. Catal. A Chem. 2015, 408, 41–47. [Google Scholar] [CrossRef]
  46. Fan, J.; Jiang, X.; Min, H.; Li, D.; Ran, X.; Zou, L.; Sun, Y.; Li, W.; Yang, J.; Teng, W. Facile preparation of Cu–Mn/CeO2/SBA-15 catalysts using ceria as an auxiliary for advanced oxidation processes. J. Mater. Chem. A 2014, 2, 10654–10661. [Google Scholar] [CrossRef]
  47. Xie, Y.; Yuan, C. Visible-light responsive cerium ion modified titania sol and nanocrystallites for X-3B dye photodegradation. Appl. Catal. B Environ. 2003, 46, 251–259. [Google Scholar] [CrossRef]
  48. Singh, L.; Rekha, P.; Chand, S. Cu-impregnated zeolite Y as highly active and stable heterogeneous Fenton-like catalyst for degradation of Congo red dye. Sep. Purif. Technol. 2016, 170, 321–336. [Google Scholar] [CrossRef]
  49. Cao, R.; Huang, H.; Tian, N.; Zhang, Y.; Guo, Y.; Zhang, T. Novel Y doped Bi2WO6 photocatalyst: Hydrothermal fabrication, characterization and enhanced visible-light-driven photocatalytic activity for Rhodamine B degradation and photocurrent generation. Mater. Charact. 2015, 101, 166–172. [Google Scholar] [CrossRef]
  50. Seroglazova, A.S.; Popkov, V.I. Synthesis of highly active and visible-light-driven PrFeO3 photocatalyst using solution combustion approach and succinic acid as fuel. Nanosyst. Phys. Chem. Math. 2022, 13, 649–654. [Google Scholar] [CrossRef]
  51. Seroglazova, A.S.; Chebanenko, M.I.; Nevedomskyi, V.N.; Popkov, V.I. Solution combustion synthesis of novel PrFeO3/CeO2 nanocomposite with enhanced photo-Fenton activity under visible light. Ceram. Int. 2023, 49, 15468–15479. [Google Scholar] [CrossRef]
  52. Ahmadi, S.; Rahdar, A.; Igwegbe, C.A.; Mortazavi-Derazkola, S.; Banach, A.M.; Rahdar, S.; Singh, A.K.; Rodriguez-Couto, S.; Kyzas, G.Z. Praseodymium-doped cadmium tungstate (CdWO4) nanoparticles for dye degradation with sonocatalytic process. Polyhedron 2020, 190, 114792. [Google Scholar] [CrossRef]
  53. Yayapao, O.; Thongtem, T.; Phuruangrat, A.; Thongtem, S. Ultrasonic-assisted synthesis of Nd-doped ZnO for photocatalysis. Mater. Lett. 2013, 90, 83–86. [Google Scholar] [CrossRef]
  54. Dhiman, M.; Singhal, S. Effect of doping of different rare earth (europium, gadolinium, dysprosium and neodymium) metal ions on structural, optical and photocatalytic properties of LaFeO3 perovskites. J. Rare Earths 2019, 37, 1279–1287. [Google Scholar] [CrossRef]
  55. Keerthana, S.; Yuvakkumar, R.; Kumar, P.S.; Ravi, G.; Velauthapillai, D. Rare earth metal (Sm) doped zinc ferrite (ZnFe2O4) for improved photocatalytic elimination of toxic dye from aquatic system. Environ. Res. 2021, 197, 111047. [Google Scholar] [CrossRef]
  56. Chahal, S.; Rani, N.; Kumar, A.; Kumar, P. Electronic structure and photocatalytic activity of samarium doped cerium oxide nanoparticles for hazardous rose bengal dye degradation. Vacuum 2020, 172, 109075. [Google Scholar] [CrossRef]
  57. Wang, M.; You, M.; Guo, P.; Tang, H.; Lv, C.; Zhang, Y.; Zhu, T.; Han, J. Hydrothermal synthesis of Sm-doped Bi2MoO6 and its high photocatalytic performance for the degradation of Rhodamine B. J. Alloys Compd. 2017, 728, 739–746. [Google Scholar] [CrossRef]
  58. Zhou, F.; Wang, H.; Zhou, S.; Liu, Y.; Yan, C. Fabrication of europium-nitrogen co-doped TiO2/Sepiolite nanocomposites and its improved photocatalytic activity in real wastewater treatment. Appl. Clay Sci. 2020, 197, 105791. [Google Scholar] [CrossRef]
  59. Zhang, Y.; Zhang, H.; Xu, Y.; Wang, Y. Europium doped nanocrystalline titanium dioxide: Preparation, phase transformation and photocatalytic properties. J. Mater. Chem. 2003, 13, 2261–2265. [Google Scholar] [CrossRef]
  60. Hernández, J.V.; Coste, S.; Murillo, A.G.; Romo, F.C.; Kassiba, A. Effects of metal doping (Cu, Ag, Eu) on the electronic and optical behavior of nanostructured TiO2. J. Alloys Compd. 2017, 710, 355–363. [Google Scholar] [CrossRef]
  61. Zhang, A.; Zhang, J. Effects of europium doping on the photocatalytic behavior of BiVO4. J. Hazard. Mater. 2010, 173, 265–272. [Google Scholar] [CrossRef] [PubMed]
  62. Ju, L.; Chen, Z.; Fang, L.; Dong, W.; Zheng, F.; Shen, M. Sol–gel synthesis and photo-Fenton-like catalytic activity of EuFeO3 nanoparticles. J. Am. Ceram. Soc. 2011, 94, 3418–3424. [Google Scholar] [CrossRef]
  63. Xu, X.; Ge, Y.; Li, B.; Fan, F.; Wang, F. Shape evolution of Eu-doped Bi2WO6 and their photocatalytic properties. Mater. Res. Bull. 2014, 59, 329–336. [Google Scholar] [CrossRef]
  64. Sharmin, F.; Basith, M. Highly efficient photocatalytic degradation of hazardous industrial and pharmaceutical pollutants using gadolinium doped BiFeO3 nanoparticles. J. Alloys Compd. 2022, 901, 163604. [Google Scholar] [CrossRef]
  65. Marwani, H.M.; Ahmad, S.; Rahman, M.M. Catalytic reduction of environmental pollutants with biopolymer hydrogel cross-linked gelatin conjugated tin-doped gadolinium oxide nanocomposites. Gels 2022, 8, 86. [Google Scholar] [CrossRef]
  66. Ismail, M.; Akhtar, K.; Khan, M.; Kamal, T.; Khan, M.A.; M Asiri, A.; Seo, J.; Khan, S.B. Pollution, toxicity and carcinogenicity of organic dyes and their catalytic bio-remediation. Curr. Pharm. Des. 2019, 25, 3645–3663. [Google Scholar] [CrossRef] [PubMed]
  67. Li, H.; Li, W.; Wang, F.; Liu, X.; Ren, C.; Miao, X. Fabrication of Pt nanoparticles decorated Gd-doped Bi2MoO6 nanosheets: Design, radicals regulating and mechanism of Gd/Pt-Bi2MoO6 photocatalyst. Appl. Surf. Sci. 2018, 427, 1046–1053. [Google Scholar] [CrossRef]
  68. Yu, C.; Wu, Z.; Liu, R.; He, H.; Fan, W.; Xue, S. The effects of Gd3+ doping on the physical structure and photocatalytic performance of Bi2MoO6 nanoplate crystals. J. Phys. Chem. Solids 2016, 93, 7–13. [Google Scholar] [CrossRef]
  69. Wu, D.; Li, C.; Zhang, D.; Wang, L.; Zhang, X.; Shi, Z.; Lin, Q. Enhanced photocatalytic activity of Gd3+ doped TiO2 and Gd2O3 modified TiO2 prepared via ball milling method. J. Rare Earths 2019, 37, 845–852. [Google Scholar] [CrossRef]
  70. Vinosha, P.A.; Vinsla, J.A.; Madhavan, J.; Devanesan, S.; AlSalhi, M.S.; Nicoletti, M.; Xavier, B. Impact of dysprosium doped (Dy) zinc ferrite (ZnFe2O4) nanocrystals in photo-fenton exclusion of recalcitrant organic pollutant. Environ. Res. 2022, 203, 111913. [Google Scholar] [CrossRef] [PubMed]
  71. Liang, C.-H.; Hou, M.-F.; Zhou, S.-G.; Li, F.-B.; Liu, C.-S.; Liu, T.-X.; Gao, Y.-X.; Wang, X.-G.; Lü, J.-L. The effect of erbium on the adsorption and photodegradation of orange I in aqueous Er3+-TiO2 suspension. J. Hazard. Mater. 2006, 138, 471–478. [Google Scholar] [CrossRef] [PubMed]
  72. Tikhanova, S.M.; Lebedev, L.A.; Martinson, K.D.; Chebanenko, M.I.; Buryanenko, I.V.; Semenov, V.G.; Nevedomskiy, V.N.; Popkov, V.I. The synthesis of novel heterojunction h-YbFeO3/o-YbFeO3 photocatalyst with enhanced Fenton-like activity under visible-light. New J. Chem. 2021, 45, 1541–1550. [Google Scholar] [CrossRef]
  73. Li, X.; Li, W.; Liu, X.; Geng, L.; Fan, H.; Ma, X.; Dong, M.; Qiu, H. The construction of Yb/Er/Pr triple-doped Bi2WO6 superior photocatalyst and the regulation of superoxide and hydroxyl radicals. Appl. Surf. Sci. 2022, 592, 153311. [Google Scholar] [CrossRef]
  74. Gu, S.; Li, W.; Bian, Y.; Wang, F.; Li, H.; Liu, X. Highly-visible-light photocatalytic performance derived from a lanthanide self-redox cycle in Ln2O3/BiVO4 (Ln: Sm, Eu, Tb) redox heterojunction. J. Phys. Chem. C 2016, 120, 19242–19251. [Google Scholar] [CrossRef]
  75. Li, H.; Li, W.; Wang, F.; Liu, X.; Ren, C. Fabrication of two lanthanides co-doped Bi2MoO6 photocatalyst: Selection, design and mechanism of Ln1/Ln2 redox couple for enhancing photocatalytic activity. Appl. Catal. B Environ. 2017, 217, 378–387. [Google Scholar] [CrossRef]
  76. Fouad, K.; Alalm, M.G.; Bassyouni, M.; Saleh, M.Y. Optimization of catalytic wet peroxide oxidation of carbofuran by Ti-LaFeO3 dual photocatalyst. Environ. Technol. Innov. 2021, 23, 101778. [Google Scholar] [CrossRef]
  77. Garcia-Muñoz, P.; Lefevre, C.; Robert, D.; Keller, N. Ti-substituted LaFeO3 perovskite as photoassisted CWPO catalyst for water treatment. Appl. Catal. B Environ. 2019, 248, 120–128. [Google Scholar] [CrossRef]
  78. Ho, C.-C.; Kang, F.; Chang, G.-M.; You, S.-J.; Wang, Y.-F. Application of recycled lanthanum-doped TiO2 immobilized on commercial air filter for visible-light photocatalytic degradation of acetone and NO. Appl. Surf. Sci. 2019, 465, 31–40. [Google Scholar] [CrossRef]
  79. Gogoi, A.; Navgire, M.; Sarma, K.C.; Gogoi, P. Fe3O4-CeO2 metal oxide nanocomposite as a Fenton-like heterogeneous catalyst for degradation of catechol. Chem. Eng. J. 2017, 311, 153–162. [Google Scholar] [CrossRef]
  80. Heckert, E.G.; Karakoti, A.S.; Seal, S.; Self, W.T. The role of cerium redox state in the SOD mimetic activity of nanoceria. Biomaterials 2008, 29, 2705–2709. [Google Scholar] [CrossRef] [PubMed]
  81. Orge, C.A.; Órfão, J.J.; Pereira, M.F.; de Farias, A.M.D.; Fraga, M.A. Ceria and cerium-based mixed oxides as ozonation catalysts. Chem. Eng. J. 2012, 200, 499–505. [Google Scholar] [CrossRef]
  82. Ouyang, W.; Zhou, Y.; Fei, X.; Bai, Y.; Wang, H.; Wu, Z. Simultaneous removal of NO and dichloromethane (CH2Cl2) over Nb-loaded cerium nanotubes catalyst. J. Environ. Sci. 2022, 111, 175–184. [Google Scholar] [CrossRef] [PubMed]
  83. Xu, L.; Wang, J. Magnetic nanoscaled Fe3O4/CeO2 composite as an efficient Fenton-like heterogeneous catalyst for degradation of 4-chlorophenol. Environ. Sci. Technol. 2012, 46, 10145–10153. [Google Scholar] [CrossRef]
  84. Karunakaran, C.; Gomathisankar, P.; Manikandan, G. Preparation and characterization of antimicrobial Ce-doped ZnO nanoparticles for photocatalytic detoxification of cyanide. Mater. Chem. Phys. 2010, 123, 585–594. [Google Scholar] [CrossRef]
  85. Sin, J.-C.; Lam, S.-M.; Lee, K.-T.; Mohamed, A.R. Preparation and photocatalytic properties of visible light-driven samarium-doped ZnO nanorods. Ceram. Int. 2013, 39, 5833–5843. [Google Scholar] [CrossRef]
  86. Cymes, B.A.; Almquist, C.B.; Krekeler, M.P. Europium-doped cryptomelane: Multi-pathway synthesis, characterization, and evaluation for the gas phase catalytic oxidation of ethanol. Appl. Catal. A Gen. 2020, 589, 117310. [Google Scholar] [CrossRef]
  87. Chong, M.N.; Jin, B.; Chow, C.W.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef] [PubMed]
  88. Kajjumba, G.W.; Marti, E.J. A review of the application of cerium and lanthanum in phosphorus removal during wastewater treatment: Characteristics, mechanism, and recovery. Chemosphere 2022, 309, 136462. [Google Scholar] [CrossRef] [PubMed]
  89. Kajjumba, G.W.; Attene-Ramos, M.; Marti, E.J. Toxicity of lanthanide coagulants assessed using four in vitro bioassays. Sci. Total Environ. 2021, 800, 149556. [Google Scholar] [CrossRef] [PubMed]
  90. Gonzalez, V.; Vignati, D.A.; Leyval, C.; Giamberini, L. Environmental fate and ecotoxicity of lanthanides: Are they a uniform group beyond chemistry? Environ. Int. 2014, 71, 148–157. [Google Scholar] [CrossRef] [PubMed]
  91. Ghadiri, M.; Ghasemi Darehnaei, M.; Sabbaghian, S.; Shirazian, S. Computational simulation for transport of priority organic pollutants through nanoporous membranes. Chem. Eng. Technol. 2013, 36, 507–512. [Google Scholar] [CrossRef]
Figure 1. (a) The number of papers issued each year on catalytic removal of contaminants in wastewater treatment using REEs over the period of 1995–2022; (b) a cumulative number of publications for each REE over the period of 1995–2022.
Figure 1. (a) The number of papers issued each year on catalytic removal of contaminants in wastewater treatment using REEs over the period of 1995–2022; (b) a cumulative number of publications for each REE over the period of 1995–2022.
Molecules 28 06185 g001
Figure 2. Percentage distribution of research papers on the utilization of REE catalysts for the degradation of different contaminant types.
Figure 2. Percentage distribution of research papers on the utilization of REE catalysts for the degradation of different contaminant types.
Molecules 28 06185 g002
Table 1. Atomic Number and Valence Shell Electronic Configuration of the Rare Earth Elements.
Table 1. Atomic Number and Valence Shell Electronic Configuration of the Rare Earth Elements.
ElementAtomic NumberValence Shell Electronic Configuration
Y39[Ar] 3d1 4s2
La57[Ar] 4d1 5s3
Ce58[Xe] 4f1 5d1 6s2
Pr59[Xe] 4f3 5d0 6s2
Nd60[Xe] 4f4 5d0 6s2
Pm61[Xe] 4f5 5d0 6s2
Sm62[Xe] 4f6 5d0 6s2
Eu63[Xe] 4f7 5d0 6s2
Gd64[Xe] 4f7 5d1 6s2
Tb65[Xe] 4f9 5d0 6s2
Dy66[Xe] 4f10 5d0 6s2
Ho67[Xe] 4f11 5d0 6s2
Er68[Xe] 4f12 5d0 6s2
Tm69[Xe] 4f13 5d0 6s2
Yb70[Xe] 4f14 5d0 6s2
Lu71[Xe] 4f14 5d1 6s2
Table 2. REEs studied for simple photocatalysis and Fenton-like photocatalysis.
Table 2. REEs studied for simple photocatalysis and Fenton-like photocatalysis.
Type of ReactionREEs Studied
PhotocatalysisY, La, Ce, Nd, Sm, Eu, Gd, Tb, Dy, Er, Yb
Fenton-like catalysisY, La, Ce, Pr, Nd, Eu, Gd, Dy, Yb
Table 3. REE catalysts used for pharmaceutical degradation.
Table 3. REE catalysts used for pharmaceutical degradation.
Catalyst
Doped REE
TargetDose of
Catalyst
Dose of Target Time (min)ConditionsProposed
Mechanism
Removal (%)Reference
LaFeO3Sulfamethoxazole 1.4 g/L3 mg/L120pH = 6.48, T = 30 °CProduction of •OH and superoxide radicals (O2/HOO)100%[31]
LaFeO3/MIL-125-NH2Carbamazepine (CBZ) and Caffeine (CAF)250 mg/L5 mg/L CBZ and 1 mg/L CAF60Visible lightProduction of •OH and O274% CBZ
87% CAF
[32]
LaFeO3/lignin-biochar (LFO/LG)Ofloxacin (OFX)250 mg/L30 mg/L75Visible light, addition of H2O2Production of •OH95.60%[33]
CeO2/C gas diffusion electrode (GDE)Dipyrone-100 mg/L20-Production of •OH100%[34]
FeCeOxDiclofenac0.5 g/L20 mg/L40pH = 5, addition of H2O2, ambient temperatureProduction of •OH83%[35]
CeO2 nanosheets doped with FeSalicylic acid250 mg/L50 mg/L120Addition of H2O2, pH = 4, T = 55 °CProduction of superoxide radicals (O2/HOO)96% by 2wt% Fe-CeO2[36]
CeOx modified graphite feltCarbamazepine-10 mg/L60pH = 5, T = 25 °CProduction of •OH69.40%[37]
Sponge-like structure of Co- and Gd-modified biochar (MBC)Ciprofloxacin (CIP)/Tetracycline (TC)1.1 g/L20 mg/L180pH = 9-99% of CIP or TC[38]
Er2FeSbO7/BiTiSbO6 heterojunction (EBH) catalystEnrofloxacin (ENR)0.75 g/L0.025 mM150Visible light, T = 20 °C, Production of •OH99,16%[39]
FeOCl doped with Y or LAIbuprofen0.5 g/L5 mg/L20Neutral pH, addition of H2O2, room temperature, darkProduction of •OH84–82% by 0.9wt% FeOCl/La and 1.2wt% FeOCl/Y, respectively[40]
Gd doped BiFeO3 nanoparticles (BGFO)Ciprofloxacin (CIP)/Levofloxacin (LFX)--240Solar illuminationProduction of O280%/79%[64]
Table 4. REE catalysts used for dyes degradation.
Table 4. REE catalysts used for dyes degradation.
Catalyst
Doped REE
TargetDose of
Catalyst
Dose of Target Time (min)ConditionsProposed
Mechanism
Removal (%)Reference
La–Fe MMTRhodamine B (RhB)/Methylene Blue (MB)1 g/L100 mg/L60Neutral pH, addition of H2O2Production of •OH97% MB
96% RhB
[40]
Tungsten oxide composites (WO3) doped with La, Gd, or ErOrganic dyes300 mg/L-90Visible light, Production of O298% for MB and >90% for MT, MO, TC and CV by 2% Gd-doped WO3 [42]
FeAl/Ce-Mts and FeAl/La-MtsReactive Blue 190.5 g/L0.12 mM180Sunlight, addition of H2O2, pH = 3.0Production of •OH100% by FeAl/Ce1.0-Mt and 99.7% with FeAl/La1.0-Mt[43]
CoFe2O4 doped with La or CeRemazol black 5, remazol brilliant yellow, o-nitrophenol, m-nitrophenol (MNP) and p-nitrophenol0.5 g/L-ott-30Visible light, room temperature, pH = 2.5, addition of H2O2Production of •OH-[44]
Nanoceria doped with Cu, Fe, Zr, Dy or LaMB80 mg/L55.2 mg/L420pH = 9.6, T = 27 °C, UV lightProduction of •OH>99% by CuO/CeO2[45]
Cu-Mn/CeO2/SBA-15RhB200 mg/L2 g/L210pH = 7.0, addition of H2O2, atmospheric pressure and constant temperatureProduction of •OH0.99[46]
Ce4+-TiO2 sol and nanocrystallitesBrilliant red dye (X-3B)1 g/L100 mg/L120Visible light Production of •OH and O283.1% by Ce4+-TiO2 nanocrystallines and 99.9% by Ce4+-TiO2 sol[47]
Cu/zeolite YCongo red (CR)1 g/L0.143 mM150Addition of H2O2, pH = 7.0Production of •OH93.58% by 7.5wt% Cu[48]
PrFeO3Methyl violet (MV)250 mg/L23.2 mg/L60Addition of H2O2, visible lightProduction of •OH-[50]
PrFeO3/CeO2-based nanocompositesMV250 mg/L23.2 mg/L30Visible light, addition of H2O2Production of •OH80.1% by 9 wt% CeO2[51]
Pr-CdWO4 NPsRemazol Black B350 mg/L100 mg/L100pH = 3.0, T = 25 °C, addition of enhancersProduction of •OH93.90%[52]
ZnO nanoneedles doped with NdMB5 g/L10−5 M300UV lightProduction of O292% by 1% Nd-doped ZnO[53]
LaFeO3 perovskites doped with Eu, Gd, Dy, or NdSafranine-O and remazol brilliant yellow0.5 g/L15 mg/L SO and 60 mg/L RBY20, 35, 35, 20Visible light, pH = 2.0, adding H2O2Production of •OHMore than 90%[54]
ZnFe2O4 nanocrystals doped with SmMB0.1 g/L1 mg/L60Visible lightProduction of •OH65% by 2 wt% Sm[55]
Sm doped CeO2 nanoparticlesRose bengal dye1 g/L5 mg/L90UV lightProduction of •OH84% by 4 wt% Sm and 89% by 6 wt% Sm[56]
Eu-N co-doped TiO2/Sepiolite NCsOrange G0.8 g/L10 mg/L540Visible light, pH = 3.0Production of •OHMore than 98% by 0.6 wt% Eu[58]
Eu3+–TiO2RhB2 g/L10−5 M30UV lightFormation of a complex between the doped lanthanide ions and substrates96%[59]
Eu2+–TiO3MB0.125 M2 × 10−5 M60UV light, pH = 8.3Induced surface plasmon resonances-[60]
Eu2+,3+−BiVO4Methyl orange (MO)2 g/L10 mg/L180Visible lightTrapping of photogenerated electrons in the catalyst by Eu dopant93.6% by 1.46 wt% Eu[61]
EuFeO3 NPsRhB1 g/L5 mg/L180Visible light, room temperature, addition of H2O2Production of •OH71%[62]
Gd doped BiFeO3 NPs (BGFO)RhB/MB- 240/180Solar illuminationProduction of O296%/97%[64]
Sn-Gd2O3 NPs4-NP, 2,6-DNP, 2-NP, MB, MO and CR8 g/L0.07 mM18Addition of NaBH4Electron transfer from the catalyst to the dye-[65]
ZnFe2O4 nanocrystals doped with DyMB1 g/L20 mg/L45Visible light, addition of H2O2Production of •OH97.30%[66]
Er3+–TiO2Orange I1 g/L6 × 10−5 M60T = 25 °C, UV light or visible light, pH = 7.0Production of •OHAbout 100% under UV and about 80% under visible light by 1.5 wt% Er[70]
Heterojunction h-YbFeO3/o-YbFeO3MV125 mg/L0.25 mM180Visible light, addition of H2O2Production of •OH64%[72]
Yb/Er/Pr-Bi2WO6 catalystCR/TC/RhB/MB1 g/L-60Visible lightReactive oxide species production and photocarrier separation90.2%/52.3%/95.5%/64.6%[73]
BiVO4 composites doped with Ln2O3 (Ln = Sm, Eu, or Tb)RhB1 g/L10 mmol/L210Visible light, 25 °CProduction of •O2, holes in the valence band (h+), and •OH55.1%, 51.8%, and 57.3% by 2% Sm2O3/BiVO4, 2% Eu2O3/BiVO4, and 10% Tb2O3/BiVO4, respetively[74]
Y-Bi2WO6 photocatalystRhB1 g/L0.01 mM240Visible lightProduction of holes (h+) and O285% by 1% Y-Bi2WO6[49]
Gd/Pt-Bi2MoO6 compositesRhB/TCs1 g/L12 mg/L RhB/20 mg/L TCs80 for RhB and 90 for TCsUV visible lightProduction of •O2, holes in the valence band (h+), and •OH95% of RhB and 77.6% of TCs by 2%Gd/2%Pt-Bi2MoO6[67]
Eu-doped Bi2WO6 compositesRhB1 g/L10 mg/L60Visible light, 25 °C, addition of H2O2Production of OH radicals98%[63]
Ln1/Ln2 co-doped Bi2MoO6 photocatalysts (Ln1/Ln2 = Tb/Eu, Dy/Sm, Er/Nd)RhB1 g/L12 mg/L240UV visible lightProduction of OH, holes (h+) and superoxide radicals O295.9% by 3%Tb/3%Eu, 98.5% by 3%Dy/3%Sm and 91.6% by 2%Er/2%Nd [75]
Gd3+ doped Bi2MoO6 nanoplate crystalsRhB0.5 g/L20 mg/L10Visible light, 20 °CProduction of OH and holes (h+) 84% by 6%Gd/Bi2MoO6[68]
Gd3+/TiO2 and Gd2O3/TiO2 NPsMV/RhB0.2 g/L for MV and 0.1 mg/L for RhB25 mg/L/30 mg/L60UV visible lightProduction of •O2, holes in the valence band (h+), and •OH97.9% by 2.5% Gd3+/TiO2[69]
Sm-doped Bi2MoO6 photocatalytsRhB0.2 g/L5 mg/L50Visible lightProduction of •O2− and holes (h+)89% by 0.8% Sm-doped Bi2MoO6[57]
Table 5. REE catalysts used for herbicide, alkaloid, cyanide, organic acid, and phenol degradation.
Table 5. REE catalysts used for herbicide, alkaloid, cyanide, organic acid, and phenol degradation.
Catalyst
Doped REE
TargetDose of
Catalyst
Dose of Target Time (min)ConditionsProposed MechanismRemoval (%)Reference
Ti-LaFeO3Carbofuran700 mg/L7 mg/L180Adding H2O2, pH = 3.0Production of •OH91%[76]
Ti-substituted LaFeO34-Chlorophenol0.5 g/L25 mg/L-Circumneutral pH, ambient atmospheric pressure and temperature, UV-A light, addition of H2O2Production of •OH100% by 3.2% Ti[77]
Fe3O4-CeO2Catechol1 g/L10 mg/L60Adding H2O2, room temperature, pH = 2.4Production of ROS100% by Fe3O4-CeO2 (15 wt%)[79]
Ce-Sm, Ce-La, Ce-ZrOxalic acid/Aniline/C.I. Reactive Blue 5-C0,oxalic acid = C0,aniline = 1 mM, C0,dye = 50 mg L−1180/30/15pH0,oxalic acid ≈ 3.0, pH0,aniline ≈ 6.5, pH0,dye ≈ 5.5, constant gas (oxygen + ozone) flow rate and constant inlet ozone concentration, 25 °CProduction of •OH100% of oxalic acid with more than 25% of Sm, La or Zr/100% of aniline by Ce0.75Zr0.25O2/100% by Ce0.75Zr0.25O2[80]
Fe3O4/CeO2 composite4-Chlorophenol2 g/L0.78 mM90Addition of H2O2, T = 30 °C, pH = 3.0Production of •OH100%[81]
ZnO nanoparticles doped with CeCyanide4 g/L-60UV-A light or natural sunlight, pH = 12.5Production of •OH-[84]
Sm/ZNRsPhenol1 g/L20 mg/L480Visible light, room temperatureProduction of •OH radicals89.5% by 1 wt% Sm/ZNRs[85]
Cryptomelane doped with EuEthanol50 mg--Dark, T = 175–200 °CProduction of O2100%[86]
Cerium nanotubes catalysts (CeNTs)NOx-600 mg/L600Addition of CH2Cl, 200 °CProduction of •OH100% by 10% Nb-CeNTs[82]
Gd/Pt-Bi2MoO6 composites4-Chlorophenol1 g/L15 mg/L300UV visible lightProduction of •O2, holes in the valence band (h+), and •OH80% by 2%Gd/2%Pt-Bi2MoO6[49]
Ln1/Ln2 co-doped Bi2MoO6 photocatalysts (Ln1/Ln2 = Tb/Eu, Dy/Sm, Er/Nd)Phenol1 g/L15 mg/L300UV visible lightProduction of •O2, holes in the valence band (h+), and •OH76.2%, 79.1% and 70.7% by 3%Tb/3%Eu, 3%Dy/3%Sm and 2%Er/2%Nd, respectively[75]
La-doped TiO2 photocatalystAcetone/NO-1000 ppb/500 ppb30UV visible lightProduction of •O2 and •OH38% of acetone and 98% of NO by 0.5%La-TiO2[78]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Saviano, L.; Brouziotis, A.A.; Padilla Suarez, E.G.; Siciliano, A.; Spampinato, M.; Guida, M.; Trifuoggi, M.; Del Bianco, D.; Carotenuto, M.; Romano Spica, V.; et al. Catalytic Activity of Rare Earth Elements (REEs) in Advanced Oxidation Processes of Wastewater Pollutants: A Review. Molecules 2023, 28, 6185. https://doi.org/10.3390/molecules28176185

AMA Style

Saviano L, Brouziotis AA, Padilla Suarez EG, Siciliano A, Spampinato M, Guida M, Trifuoggi M, Del Bianco D, Carotenuto M, Romano Spica V, et al. Catalytic Activity of Rare Earth Elements (REEs) in Advanced Oxidation Processes of Wastewater Pollutants: A Review. Molecules. 2023; 28(17):6185. https://doi.org/10.3390/molecules28176185

Chicago/Turabian Style

Saviano, Lorenzo, Antonios Apostolos Brouziotis, Edith Guadalupe Padilla Suarez, Antonietta Siciliano, Marisa Spampinato, Marco Guida, Marco Trifuoggi, Donatella Del Bianco, Maurizio Carotenuto, Vincenzo Romano Spica, and et al. 2023. "Catalytic Activity of Rare Earth Elements (REEs) in Advanced Oxidation Processes of Wastewater Pollutants: A Review" Molecules 28, no. 17: 6185. https://doi.org/10.3390/molecules28176185

Article Metrics

Back to TopTop