Next Article in Journal
Influence of the Salt Addition during the Fermentation Process on the Physical and Chemical Properties of Dried Yellow Beetroot
Previous Article in Journal
A Novel Method to Identify Key Nodes in Complex Networks Based on Degree and Neighborhood Information
Previous Article in Special Issue
Luminescent Materials: Synthesis, Characterization and Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancing the Photocatalytic Activity and Luminescent Properties of Rare-Earth-Doped CeO2 Nanoparticles

1
National Institute for Research and Development of Isotopic and Molecular Technologies, 67-103 Donat Street, 400293 Cluj-Napoca, Romania
2
Physics and Chemistry Department, Technical University of Cluj-Napoca, 28 Memorandumului Street, 400114 Cluj-Napoca, Romania
3
Materials Science and Engineering Department, Technical University of Cluj-Napoca, 103-105 Muncii Avenue, 400641 Cluj-Napoca, Romania
4
Raluca Ripan Institute for Research in Chemistry, Babes-Bolyai University, 30 Fântânele Street, 400294 Cluj-Napoca, Romania
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2024, 14(2), 522; https://doi.org/10.3390/app14020522
Submission received: 30 November 2023 / Revised: 3 January 2024 / Accepted: 5 January 2024 / Published: 7 January 2024
(This article belongs to the Special Issue Luminescent Materials: Synthesis, Characterization and Application)

Abstract

:
Our study is focused on the structural and morphological characteristics, optical behaviour and photocatalytic properties of undoped and 5 at% Eu3+-, Gd3+- and Y3+-doped CeO2 nanoparticles prepared by a green hybrid sol-gel combustion method. Several techniques such as X-ray diffraction powder (XRD), Transmission Electron Microscopy (TEM), UV-Vis spectroscopy, Photoluminescence spectroscopy (PL) and Electron Paramagnetic Resonance (EPR) have been used to investigate the obtained samples. Moreover, the correlation between the characteristics and properties has been studied. The nanoparticles observed by TEM exhibit a pseudo-spherical shape, except for Y3+-doped CeO2, which presents an acicular shape. The average size of undoped and rare-earth-doped CeO2 nanoparticles is below 10 nm, in good agreement with the calculations performed based on XRD analyses. From UV-Vis analyses it has been deduced that with doping the band gap energy decreases, which shows that additional levels are introduced by doping into the CeO2 band gap. The EPR spectra evidence similar behaviour for all doped samples. The photocatalytic activity was evaluated by the degradation of rhodamine B (RhB) under UV light irradiation. The photodegradation mechanism has been studied in depth based on the formation of electron-hole pairs, and to evidence the reactive oxygen species, ESR coupled with spin-trapping experiments was performed. In the case of Y-doped CeO2 nanoparticles, the generation of both •OOH and • O 2 radicals involved in RhB photodegradation was highlighted.

1. Introduction

Industrial progress has improved living standards, driving the development of human society. Therefore, some industrial pollutants are generated in wastewater effluents, such as heavy metals, radioactive nuclides, and organics [1]. Unfortunately, great amounts of wastewater are released into the environment, causing significant environmental pollution. Organic pollutants such as phenols, organic dyes and antibiotics are very toxic. These pollutants can be accumulated into aquatic organisms and from there along the food chain to the human body [2]. Thus, the discarding of organic pollutants represents one of the most important environmental problems. In recent years, many technologies have been employed in environmental remediation, including adsorption [3], biodegradation [4] and chemical oxidation [5]. The disadvantages of the above methods consist of the use of solvents and the generation of secondary metabolites which are very difficult to remove. One of the cost-effective methods that had been developed for the removal of organic contaminants from wastewater is Advanced Oxidation Processes (AOPs). AOPs using semiconductor photocatalysts have been proven to be a promising technique for water disinfection due to its high efficiency, acceptable price, high stability and the capability to degrade organic pollutants and mineralization to CO2 and H2O as the final products [6].
The efficiency of the photocatalytic process depends on the operational parameters that dominate the adsorption and photodegradation of dye molecules. Among these parameters, the most important are the initial pH of the dye solution, the concentration of the photocatalyst and dye, the reaction temperature and the dopants content [6,7]. Tirumala et al. show that controlling the incident light intensity and wavelength is a mode for controlling and tuning the activity and selectivity of the catalyst [8]. Also, in a doped semiconductor photocatalyst, the positively or negatively charged microenvironment around trapped holes or electrons influences the ability of photocatalyst to generate reactive oxygen species (ROS) involved in reactions of photo-oxidation and photo-reduction [9,10].
A great number of semiconductor metal oxides, such as TiO2 [11], ZnO [12], SnO2 [13], etc., have been used as photocatalytic materials. As one of the most popular metal oxide semiconductors, the rare earth oxide ceria (CeO2) is extensively used in many fields due to its ability to extract, store and release oxygen. Its applicative potential based on this special feature includes domains like catalysis, photocatalysis, solid oxide fuel cells, ceramics, sunscreens, solar cells, gas sensors and biomedical areas such as biosensing, bioimaging or targeted drug therapy, etc. [14,15,16,17,18,19]. Moreover, ceria has proved to be an ideal candidate as an artificial oxygen modulator, and this is vital for many biological processes [20]. Also, nanoceria is well-known for its enhanced catalytic reactivity due to easy conversion between cerium’s tetravalent and trivalent valence surface states [21]. In ceria, the tetravalent cerium ion, Ce4+, transforms into a trivalent state, Ce3+, and this transformation produces a good yield of oxygen vacancies without changes in the chemical environment [22]. Additionally, it is nontoxic, abundant and absorbs light in the near UV and also marginally in the visible region, properties which can be used in heterogeneous photocatalysis [23]. Thus, the application of CeO2 is obstructed. Many strategies have been used to enhance the catalytic/photocatalytic activity by tuning the structural and morphological characteristics [24]. One strategy to extend the utilization of visible light over CeO2 consists of suitable doping with various metal ions: isovalent cations (Ti4+, Zr4+, Hf4+, Sn4+, etc.) or aliovalent cations (Mn2+, Zn2+, Ca2+, La3+, Eu3+, Sm3+, Y3+, Eu3+, Mn3+, Sc3+, Gd3+, Sm3+, etc.) [25]. It has been demonstrated theoretically and experimentally that the substitution of cerium ions with isovalent dopant cations induces a decrease in oxygen vacancy formation energy due to the structural distortion, whereas in the case of aliovalent dopants the generation of extra oxygen vacancies into the ceria lattice has been evidenced due to the synergism between the structural distortion and electron modifications [25,26,27]. On the other hand, the trivalent cations of rare earth metals drastically modify the reactivity structure and optical emission of ceria [21,28,29,30]. The luminescent properties of undoped or rare earth (Eu3+, Sm3+, or Tb3+)-doped ceria have been reported and the emission mechanism proposed by considering the transfer of the charge carrier between the Ce 4f conduction band and the O 2p valence band. Also, the broad photoluminescence emission in the visible domain is attributed to defects, including oxygen vacancies in the ceria lattice with electronic energy levels below the 4f band [31,32,33].
To date, a lot of chemical strategies have been developed to synthesize undoped or doped ceria nanoparticles, such as solid-state reactions [34], co-precipitation [35], hydrothermal techniques [36], sol-gel [37], microwave-assisted synthesis [38], ultrasonic spray pyrolysis [39], sonochemical synthesis [40], combustion method [41] and so on, proving that the physical and chemical properties are strongly dependent on the synthetic route used. In this regard, Amoresi et al. studied the correlation between a wide range of CeO2 morphologies, such as nanorods, nanocubes, nanobeans and nanohexagons and the photocatalytic activity [42]. A similar study has been published by Ortega et al. with emphasis on the Eu-doped ceria nanostructures [43]. It has been demonstrated that enhanced photocatalytic performance is obtained by increasing the charge carrier concentration and light absorption capacity by defects induced by the dopants into CeO2 nanostructures. Nevertheless, studies are still ongoing for a deeper understanding of the correlations between the experimental parameters, structure, morphology, luminescence properties and photocatalytic activity.
Considering the previous research works reported in the literature, this paper focuses on the effects induced by the incorporation into the ceria lattice of Eu3+, Gd3+ and Y3+ aliovalent cations by studying the correlations between the structural and morphological characteristics and luminescence properties and photocatalytic activity. These materials have been synthesized by a green hybrid sol-gel combustion method and characterized by using various techniques, namely X-ray Diffraction, Transmission Electron Microscopy, UV-Vis spectroscopy, Photoluminescence (PL) and Electron Paramagnetic Resonance (EPR). Finally, we attempt to provide an overview of how the selected rare earth ions influence the ceria’s structure, morphology and photocatalytic activity.

2. Materials and Methods

2.1. Materials

Commercial ammonium cerium (IV) nitrate, (NH4)2Ce(NO3)6, (Merck, 98.5%), gadolinium nitrate hexahydrate, Gd(NO3)3·6H2O (99.9%, Merck), europium acetate hydrate Eu(CH3-COO)3·xH2O (99.9%, Merck), yttrium acetate hydrate Y(CH3-COO)3·xH2O (99.9%, Merck), sucrose, C12H22O11 (99.5%, Sigma-Aldrich), pectin from citrus peel, (Sigma-Aldrich), rhodamine B (RhB), C28H31ClN2O3 (analytical standard, Merck), acetonitrile, C2H3N (ACN, HPLC-grade, Merck) and ortho-phosphoric acid (H3PO4, 85%, Merck) were used without any further purification.

2.2. Nanoparticle Synthesis

The undoped and Eu3+-, Gd3+- and Y3+-doped CeO2 nanoparticles were synthetized via a green hybrid sol-gel combustion method developed by Goga et al. [44] using sucrose as a poly-condensation agent and pectin as a chelation agent. Briefly, the corresponding amounts of metal salts—for the doped systems 5 at% of Eu3+, Gd3+ and Y3+ ions relative to CeO2—were dissolved in MiliQ water (0.05 M) at 60 °C for an hour under vigorous magnetic stirring. Similarly, a sucrose solution (0.1 M) was prepared using milli-Q water, at 60 °C, under magnetic stirring. The two solutions were mixed in the presence of the pectin (the sucrose to pectin weight ratio was 1:5). The pH value was monitored during the synthesis and no correction was necessary to maintain an acidic medium in the solution. The pH value of the final solutions was determined to be in the range of 1.5–2. The as-obtained mixture was dried in air, on the sand bath. At 215 °C, a spontaneous combustion of the synthesis mixture was observed and a dispersed yellow powder was obtained. The evolved gases were captured and absorbed in a soda lime aqueous solution. The following chemical reaction can be considered in the formation of ceria nanoparticles:
(1 − x)Ce3+ + xRE3+ + 2NH4+ + 6NO3 + 2C12H22O11 + 26O2→Ce(1−x)RExO2 + 24CO2↑ + 26H2O + 8NO2
The use of non-toxic reagents such as sucrose and pectin give the process a green synthesis approach, but the reaction is not an environmentally friendly reaction due to the production of nitrogen dioxide.

2.3. Characterization Techniques

The structural characterisation of the obtained nanoparticles was undertaken through X-ray diffraction (XRD). The equipment used was an INEL Equinox 3000 diffractometer, which works in reflection mode, and it is equipped with a curved multidetector. The wavelength used was the one characteristic of Co, Kα radiation (λ = 1.7903 Å). The mean crystallite size was calculated with the Scherrer method [45]. The lattice parameter was calculated using Celref version 3 refinement software [46]. TEM images were obtained by scanning transmission electron microscopy (STEM) using a Hitachi HD2700 Electron Microscope (Hitachi, Tokyo, Japan). Absorption spectra were recorded using a JASCO V570 UV–VIS-NIR Spectrophotometer equipped with an absolute reflectivity measurement JASCO ARN-475 accessory. Luminescent characteristics of samples were evaluated based on emission spectra registered with a Wavell JASCO FP-6500 spectrofluorimeter, equipped with a PMT R928 photomultiplier (glass filter UG1—Reichmann Feinoptik, GER). Electron paramagnetic resonance (EPR) spectroscopy measurements of powder samples were carried out using a Bruker E-500 ELEXSYS spectrometer (Karlsruhe, Germany) at room temperature under identical conditions: The measurements were conducted at Xband (9.52 GHz) microwave power 2 mW, with a modulation frequency of 100 kHz. To determine the reactive oxygen species (ROS) generation, the EPR spectrometer coupled with the spin trapping probe technique was used. As the spin trapping reagent, DMPO was used. The nanoparticles (10 mg) were dispersed in DMSO (1 mL) and homogenized in an ultrasound bath (30 min) before use. DMPO of 0.2 mol/L concentration was added to the solution. The obtained solution was subjected to visible irradiation for 25 min and immediately before measurement transferred into the quartz flat cell, optimized for liquid measurements. The photodegradation of Rhodamine B (RhB) was carried out in a Laboratory-Visible-Reactor system using a 400 W halogen lamp (Osram) [47]. The catalyst (5 mg) was suspended in an aqueous solution of RhB (1.0 × 10−5 mol·L−1, 10 mL) and then the mixture was stirred in the dark to achieve the adsorption equilibrium on the catalyst surface. Each degradation experiment was continuously conducted for 6 h. The mixture (3.5 mL) was withdrawn for analysis every 60 min. After separating the catalyst from the suspensions by centrifugation, the solution was analysed by the UV–Vis spectrophotometer by recording the maximum absorbance of RhB at 554 nm.
RhB removal rate is calculated using the following equation:
R e m o v a l   r a t e = A 0   A t A 0 × 100
where A0 and At represent the RhB absorbance initially and at time t, respectively.
The HPLC analyses were performed on a Jasco HPLC chromatograph (Jasco International Co., Ltd., Tokyo, Japan) that was equipped with an intelligent pump, PU-980, a ternary gradient unit, LG-980-02, an intelligent column thermostat, CO-2060 Plus, an intelligent detector, UV-975, an intelligent fluorescence detector, FP-2020 Plus, and an injection valve that was equipped with a 20 µL sample loop (Rheodyne, Thermo Fischer Scientific, Waltham, MA, USA). The system was controlled, and the experimental data processed with the ChromPass software (version v1.7, Jasco International Co., Ltd., Tokyo, Japan). Separation was performed on a Gemini chromatographic column (5μ C18, 110 Å, 250 × 4.6 mm, Phenomenex, Torrance, CA, USA) at 21 °C. Isocratic elution was performed with 45% of acetonitrile with 55% of 0.1% H3PO4 buffer. The flow rate was 1 mL/min and the injection volume was set to 20 μL. UV detection was performed by fluorescence at the excitation and emission wavelengths set at 535 nm and 580 nm. All the experiments were performed in triplicate. The determination of rhodamine B was performed according to the adapted method described by Chiang et al. [48].

3. Results and Discussion

The structural characteristics of undoped or RE-doped CeO2 samples were determined by X-ray diffraction measurements and the corresponding diffraction patterns are shown in Figure 1. All diffraction lines are well indexed to the face-centred cubic (fcc) fluorite structure of CeO2 (JCDD PDF 034-0394). No peaks belonging to any other phase have been observed in the XRD patterns, indicating the high structural purity of the samples. The effects of the dopant on the lattice parameters and crystallite sizes—calculated using the Scherrer method—are presented in Table 1.
It has been demonstrated that the substitution of Ce4+ ions with a larger ion radius dopant cation alters the local chemistry and induces an enlargement of the crystal lattice due to steric effects [49,50,51]. The substitution into the ceria lattice of octahedral Ce4+ ions (Ce4+ = 0.97 Å) with a larger sized dopant (Gd3+ = 1.053 Å; Eu3+ = 1.066 Å and Y3+ = 1.019 A [52]) induced a slight increase in the lattice parameter values. Regarding the crystallite sizes it can be noted that the small amount of dopant ions stimulates the nucleation and the growth processes of the nanocrystals. This tendency can be noticed for the crystallite sizes in the cases of Gd3+ and Eu3+. The introduction of Y3+ into the ceria lattice led to a diminishing of the crystallite size.
The morphological characterization of the samples has been achieved via TEM investigations—Figure 2. TEM images show the formation of pseudo-spherical nanoparticles with a high tendency to agglomerate. Analysis of these images disclosed that the particle sizes are below 10 nm, which is in good concurrence with the XRD results. This means that each particle is composed of only one crystallite. In the case of the Y-doped ceria sample, the presence of some aggregates with an acicular shape should be noticed. The changing of nanoparticles’ shapes by introducing different types of dopants indicate that the crystallisation process can be tailored.
The absorption spectrum of nanoparticles in the UV-Vis-NIR domain is affected by various factors such as optical band gap, oxygen vacancies, the presence of an impurity centre, etc. Figure 3a shows the absorbance spectra of pure and RE-doped ceria nanostructures and the spectra exhibit a strong absorbance in the UV region. In general, the absorption band below 400 nm is assigned to the transitions of charge transfer between O 2p and Ce 4f orbitals in cerium oxide [53]. Doping with RE caused a red shift in the absorbance band of ceria, indicating that 4f or 4d transition metals ions induce intermediate electronic energy levels between the O 2p and Ce 4f orbitals of cerium oxide [54]. To evaluate the energy band gap of the samples, Tauc’s relation was applied to the absorption spectra. Figure 3b displays the Tauc plots. The energy band gaps of RE-doped CeO2 nanoparticles are smaller than those corresponding to ceria (2.98 eV)—Table 1. The band gap narrowing of CeO2 by doping with RE ions can be explain by the formation of 4f for Gd and Eu or 4d for Y states below the conduction band edge, which extend the absorption of ceria into the visible region.
Photoluminescence spectroscopy is considered an important technique for understanding the role of the recombination of electron pairs and defects in modifying the photocatalytic activity of a material. The registered spectra for the ceria-based samples are presented in Figure 4. The literature data reported that the energy of emitted photons due to the electron transitions from the 4f band to the valence band in CeO2 lies in the range of 3.0–3.8 eV [21]. The band centred at 364 nm, presented in all spectra, should be attributed to this transition. Also, it is reported that the emission bands located from 400 to 500 nm are attributed to some defects, such as vacancies or dislocations in CeO2 [21,55]. Therefore, the conclusive transitions from Ce 4f level to O 2p level due to abundant defects like O vacancies and dislocations are responsible for emission peaks from the 430 nm to 494 nm domains observed in all spectra. The replacement of Ce4+ cations by RE3+ cations in the lattice of CeO2 generates the formation of defects and oxygen vacancies in order to preserve the charge balance.
The photoluminescence spectrum of Eu-doped ceria is significantly different from the other registered spectra, showing well-defined emission bands in the 540–620 region. These bands result directly from the incorporation of the Eu3+ ions into the CeO2 lattice. Depending on the excitation wavelength, possible f-f transitions of Eu3+ cations (5D07FJ) take place at 550 nm (J = 0), 593 nm (J = 1), 610–630 nm (J = 2), 650 nm (J = 3), and 710 nm (J = 4) [56]. It is already known that the Ce site in the fluorite CeO2 lattice presents Oh symmetry with eightfold oxygen coordination. Therefore, the Ce site possesses inversion symmetry and the electric dipole emission 5D07F2 could be forbidden and could not be observed only if the Eu3+ ion occupied the Ce site without changing the original inversion symmetry. When Eu3+ replaces Ce4+, it needs charge compensation in order to obtain a charge equilibrium. Two kinds of defects could be formed during charge compensation: * interstitial Eu3+ (four Eu3+ replacing three Ce4+) and **O2− vacancies (one Eu2O3 replacing two CeO2). These defects induce the symmetry distortion of the local environment around the Eu3+ ions. Moreover, due to the difference in ionic radius between Ce4+ and Eu3+, the replacement of Ce4+ ions with Eu3+ ions also generates the symmetry distortion of the local environment around the Eu3+ ions [56].
The presence in the spectrum (λexc = 325 nm) of 5D07F0 transition, around 549 nm, 5D07F1 transition at 592 nm and 5D07F2 at 606 and 612 nm, respectively, suggests that Eu3+ cations replace Ce4+ sites in the CeO2 lattice and moreover is indicative of the formation of oxygen vacancies [57,58,59,60].
To evidence the presence of paramagnetic species, the EPR technique was used. The spectra were similar, and as an example the experimental spectrum of CeO2-Y nanoparticles obtained at room temperature is illustrated in Figure 5a. The spectrum shows a signal at a higher value than the free spin value (ge = 2.0023), and two signals at lower g values. The EPR signal at g = 2.00935 was assigned to physisorbed oxygen molecules or to paramagnetic species such as • O 2 generated from these oxygen molecules [61], and that centred g = 1.964 has been attributed both to quasi-free electrons and to Ce3+ ions [62]. Both paramagnetic impurities are due to the non-complete stoichiometry obtained during the synthesis process.
The effect of the charge separation in semiconducting oxides under irradiation can be probed by EPR. Therefore, the CeO2-Y sample was irradiated with UV to check the modification of the spectrum (Figure 5a). Under irradiation, only the resonance signal centred at g = 2.00935 changes its intensity. This behaviour can be explained by the following: when CeO2 is irradiated with photons owing energy higher than that of the band gap, electron–hole pairs are generated. If these pairs do not recombine, the electrons are stabilized on cations, while the holes are stabilized by oxygen anions [63]. Thus, the photoexcited electrons are stabilized by Ce4+ ions, being reduced to Ce3+; the increases of the resonance signal attributed to this signal are not visible because of the high value of spin orbit coupling and therefore the very fast relaxation time [64].
To illustrate the charge generation and recombination, the intensity of the EPR signal centred at g = 2.00935 was monitored under UV irradiation in three on–off cycles. The obtained kinetic is exhibited in Figure 5b. It can be observed that when the UV is turned on, the EPR signal increased faster and charge carriers stayed at the excited energy level; after turning off the light, the signal decreased, but a little slowly, which means that the recombination process is longer. The on–off cycle experiments show that the process is reversible.
The photocatalytic activity of the samples was investigated against Rhodamine B (RhB) solution under visible irradiation. Prior to the photocatalytic experiments, the samples dispersed in solution were kept for 60 min in the dark to achieve an adsorption–desorption equilibrium. The RhB removal rate was monitored by the temporal evolution of RhB absorbance. As an example, Figure 6 shows the evolution in time after dark adsorption and the visible irradiation of RhB absorbance in the presence of the CeO2-Gd sample. It could be observed that the RhB absorbance decreases with irradiation time. The decrease in the absorption peak is due to the cleavage of the RhB chromophore ring structure, and the shift of the absorption maximum results in the N-deethylation of RhB molecules [65].
The RhB removal rate in the presence of all the samples is depicted in Figure 7. All doped samples possess enhanced adsorption capacity relative to ceria nanoparticles. Figure S2 shows the adsorption capacity of Y dopedCeO2 sample for 420 min which proves that after 60 min, the adsorption equilibrium was reached. The enhanced dark adsorption capacity of doped samples could be due to much more active sites for the dye molecules to react. The removal rate of the CeO2-Gd sample is almost three times higher than that corresponding to ceria.
The photocatalytic process could be described by the first order kinetic model following the equation
l n A t A 0 * = k i · t
where At—absorbance of RhB at t time; A 0 * —absorbance of RhB after dark adsorption; t—irradiation time; and ki—apparent kinetic constant. The obtained plots show a linear relationship with the irradiation time (Figure 8). In Table 2 are shown the calculated rate constants and correlation coefficients corresponding to Figure 8. From the obtained k values, it can be seen that the best performance was obtained using Y-CeO2, followed by Eu-CeO2 > Gd-CeO2 > CeO2.
The photocatalytic mechanism can be explained based on the formation of electron–hole pairs. Under visible irradiation, electrons absorb energy and move from the valence band to the conduction band, resulting in the appearance of electron–hole pairs. The electrons interact with oxygen and produce superoxide radicals and the holes interact with H2O, resulting in hydroxyl radicals. These radicals are known as radical oxygen species and are capable of degrading organic pollutants. The photocatalytic activity depends on the capacity of the photocatalyst to separate the electron–hole pairs. The most important factor is represented by the rate of electron–hole pair separation and recombination. The delay in recombination drives enhanced photocatalytic activity [24].
To evidence the reactive oxygen species, ESR coupled with spin-trapping experiments was performed. The experimental spectrum is shown in Figure 9. The obtained spectrum is a complex one, consisting of many resonance lines which are due to several species. To identify the species, a simulation of the spectrum was carried out. The simulated spectrum is also shown in Figure 9 and was obtained through a linear combination of the following spin adducts: ●DMPO–OCH3 (aN = 13.3 G, a H β = 8.3 G, a H γ = 1.9 G, relative concentration 60%), ●DMPO– O 2 (aN = 12.9 G, a H β = 11.4 G, a H γ = 2.6 G, relative concentration 31%) and nitroxide-like radicals (aN = 13.9 G, relative concentration 9%).
The proof of the ●OH radical existence in the system is due to the presence of ●DMPO–OCH3 spin adducts, since these spin adducts are generated by the reaction between ●OH radicals and DMSO solvent [66]. Nitroxide-like radicals were generated by the cleavage of the NAC bond followed by the ring opening of DMPO [67].
The roles of scavengers in elucidating the active radical species involved in photodegradation are to capture and suppress the action of the species under investigation during the photocatalytic process. The scavenger experiment was performed to evaluate the active radical species that participated in the photocatalytic process. Ethylenediaminetetraacetic acid (EDTA) and vitamin C (Vit C) were used as h+ and • O 2 radical scavengers, respectively [68]. The scavenger experimental conditions were the same as those used for the RhB degradation and 1.0 mM scavenger was added to the solution. In Figure 10, all samples showed lower RhB degradation than the control experiment (with no scavenger), which attested that these radical species are involved in the photocatalytic process. The dominant active species showed the weakest photocatalytic activity. Therefore, the lowest RhB degradation was obtained with Vit C, indicating that • O 2 radicals are the main species involved in RhB photodegradation.
Based on the above studies, the photocatalytic mechanisms of Y-, Gd- and Eu-doped CeO2 nanoparticles are illustrated in Figure 11. The doping of CeO2 results in a higher concentration of oxygen vacancies and additional doping transition levels. When the doped CeO2 is irradiated with visible light, the photogenerated electrons will move in the conduction band and the holes will be generated in the valence band. Part of the electrons will be trapped on the doping levels, limiting the recombination of the photoinduced electron and hole pairs [69]. Therefore, the electrons from the conduction band will interact with oxygen-producing superoxide radicals and the holes will interact with H2O, generating hydroxyl radicals. Both generated radical species will participate in a redox reaction, being capable of degrading the RhB molecule. From scavenger experiments it can be found that • O 2 radicals are the main species required to degrade RhB dye; indeed, • O 2 radicals are generated by the sample, but hydroxyl radicals are the majority, which explains the not-very-high photocatalytic performances of the samples.
The HPLC method has been used as a complementary method to determine the concentration in RhB of the solutions formed after the UV photodegradation process. The obtained concentration values for all samples are summarized in Table 3 and the chromatograms are presented in Figure S1—Supplementary Material.
The obtained results suggest a decreasing tendency for RhB concentration after the photodegradation of the pollutant solution with CeO2-doped catalysts (CeO2 > CeO2:Eu > CeO2:Y > CeO2:Gd), sustaining the previous results obtained using UV–Vis spectroscopy.

4. Conclusions

In summary, CeO2 nanoparticles doped with rare earth ions (Eu3+, Gd3+ and Y3+) were synthetized via a green hybrid sol-gel combustion method. The nanoparticles have a face-centred cubic (fcc) fluorite structure, and the slight increase in the cell parameters sustains the replacement of Ce4+ with dopant ions. Undoped CeO2 has a small crystallite size (around 8 nm) and by doping a slight increase was noticed, except for Y ions. The particle size evaluated by TEM is in accordance with the crystallite size evaluated by XRD, sustaining the formation of monocrystalline particles. By doping the band gap, the energy decreases, which shows that additional levels are introduced by doping into the CeO2 band gap. The EPR spectra evidence similar behaviour for all doped samples. The intense signal at g~2.0093 was assigned to physisorbed oxygen molecules or to paramagnetic species such as • O 2 generated from these oxygen molecules, and a lower one at g~1.964 was due to quasi-free electrons or Ce3+ ions. Under UV light irradiation, the signal at g~2.0093 increases in intensity, which suggests an efficient charge separation by capturing the photogenerated holes with oxygen anions. The charge generation and recombination processes are reversible. The RhB removal rate was tested under UV light irradiation. The removal of pollutant molecules is due to two processes, adsorption and photocatalysis. The best performance was obtained using Gd-CeO2, with the adsorption playing an important role. But the best photocatalytic activity was obtained for Y-doped CeO2. The spin-trapping experiment revealed the generation by Y-CeO2 nanoparticles of both •OOH and • O 2 radicals involved in RhB photodegradation. In our future works, studies of the photocatalytic degradation of other organic pollutants will be considered.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app14020522/s1, Figure S1: HPLC chromatograms of RhB for the initial solution—a; undoped CeO2—b; Eu-doped CeO2—c; Y-doped CeO2—d; Gd-doped CeO2—e; Figure S2: The dark adsorption capacity of CeO2-Y sample for 420 min.

Author Contributions

Conceptualization, D.T. and A.M.; methodology, D.T., A.P., R.B. and A.M.; investigation, D.T., A.P., R.B.S., R.B., T.F.M., I.P. and M.F.; writing—original draft preparation, D.T. and A.M.; writing—review and editing, D.T., A.P., R.B.S., T.F.M., I.P., M.F. and A.M.; visualization, D.T., A.P. and A.M.; supervision, M.F. and A.M.; project administration, A.M. All authors have read and agreed to the published version of the manuscript.

Funding

Financial support was provided by a grant from the Ministry of Research, Innovation and Digitalization, CNCS—UEFISCDI, project number PN-III-P4-PCE-2021-1561, and from the MCID through the “Nucleu” Programe within the National Plan for Research, Development and Innovation 2022–2027, project PN 23 24 01 03.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Shah, M.P. Chemistry in the Environment. In Biological Treatment of Industrial Wastewater; The Royal Society of Chemistry: London, UK, 2021; pp. 1–11. ISBN 978-1-83916-539-9. [Google Scholar]
  2. Wang, C.C.; Li, J.R.; Lv, X.L.; Zhang, Y.Q.; Guo, G. Photocatalytic organic pollutants degradation in metal–organic frameworks. Energy Environ. Sci. 2014, 7, 2831–2867. [Google Scholar] [CrossRef]
  3. Ali, I.; Gupta, V.K. Advances in water treatment by adsorption technology. Nat. Protoc. 2006, 1, 2661–2667. [Google Scholar] [CrossRef] [PubMed]
  4. Kagle, J.; Porter, A.W.; Murdoch, R.W.; Rivera-Cancel, G.; Hay, A.G. Biodegradation of pharmaceutical and personal care products. Adv. Appl. Microbiol. 2009, 67, 65–108. [Google Scholar] [CrossRef] [PubMed]
  5. Wenk, J.; Aeschbacher, M.; Salhi, E.; Canonica, S.; von Gunten, U.; Sander, M. Chemical Oxidation of Dissolved Organic Matter by Chlorine Dioxide, Chlorine, And Ozone: Effects on its optical and antioxidant properties. Environ. Sci. Technol. 2013, 47, 11147–11156. [Google Scholar] [CrossRef] [PubMed]
  6. Hejazi, R.; Mahjoub, A.R.; Khavar, A.H.C.; Khazaee, Z. Novel visible-light-responsive rGO-ZnO@Bi2MoO6 nanocomposite with enhanced light harvesting and Z-scheme charge transfer for photodegradation and detoxification of RhB. Solid State Sci. 2019, 95, 105934. [Google Scholar] [CrossRef]
  7. Rafiq, A.; Ikram, M.; Ali, S.; Niaz, F.; Khan, M.; Khan, Q.; Maqbool, M. Photocatalytic degradation of dyes using semiconductor photocatalysts to clean industrial water pollution. J. Ind. Eng. Chem. 2021, 97, 111–128. [Google Scholar] [CrossRef]
  8. Vasiljevic, Z.Z.; Dojcinovic, M.P.; Jankovic-Castvan, I.; Ognjanovic, M.; Tadic, N.B.; Stojadinovic, S.; Brankovic, G.O.; Nokoloc, M.V. Photocatalytic degradation of methylene blue under natural sunlight using iron titanate nanoparticles prepared by a modified sol–gel method. R. Soc. Open Sci. 2020, 7, 200708. [Google Scholar] [CrossRef]
  9. Tirumala, R.T.A.; Khatri, N.; Ramakrishnan, S.B.; Mohammadparast, F.; Khan, M.T.; Tan, S.; Wagle, P.; Puri, S.; McIloray, D.N.; Kalkan, A.K.; et al. Tuning Catalytic Activity and Selectivity in Photocatalysis on Mie-Resonant Cuprous Oxide Particles: Distinguishing Electromagnetic Field Enhancement Effect from the Heating Effect. ACS Sustain. Chem. Eng. 2023, 11, 15931–15940. [Google Scholar] [CrossRef]
  10. Xu, W.; Jain, P.K.; Beberwyck, B.J.; Alivisatos, A.P. Probing Redox Photocatalysis of Trapped Electrons and Holes on Single Sb-doped Titania Nanorod Surfaces. J. Am. Chem. Soc. 2012, 134, 3946–3949. [Google Scholar] [CrossRef]
  11. Toloman, D.; Popa, A.; Stefan, M.; Pana, O.; Silipas, T.D.; Macavei, S.; Barbu-Tudoran, L. Impact of Gd ions from the lattice of TiO2 nanoparticles on the formation of reactive oxygen species during the degradation of RhB under visible light irradiation. Mater. Sci. Semicond. Process. 2017, 71, 61–68. [Google Scholar] [CrossRef]
  12. Toloman, D.; Mesaros, A.; Popa, A.; Silipas, T.D.; Neamtu, S.; Katona, G. V-doped ZnO particles: Synthesis, structural, optical and photocatalytic properties. J. Mater. Sci. Mater. Electron. 2016, 27, 5691–5698. [Google Scholar] [CrossRef]
  13. Toloman, D.; Popa, A.; Stefan, M.; Silipas, T.D.; Suciu, R.C.; Barbu-Tudoran, L.; Pana, O. Enhanced photocatalytic activity of Co doped SnO2 nanoparticles by controlling the oxygen vacancy states. Opt. Mater. 2020, 110, 110472. [Google Scholar] [CrossRef]
  14. Ma, R.; Zhang, S.; Wen, T.; Gu, P.; Li, L.; Zhao, G.; Niu, F.; Huang, Q.; Tang, Z.; Wang, X. A critical review on visible-light-response CeO2-based photocatalysts with enhanced photooxidation of organic pollutants. Catal. Today 2019, 335, 20–30. [Google Scholar] [CrossRef]
  15. Banavar, S.; Deshpande, A.; Sur, S.; Andreescu, S. Ceria nanoparticle theranostics: Harnessing antioxidant properties in biomedicine and beyond. J. Phys. Mater. 2021, 4, 042003. [Google Scholar] [CrossRef]
  16. Rashid, S.; Zaman, Z.; Nasir, M.; Ahmed, A.; Andreescu, S.; Liaquat, M.; Hayat, A. Nanoceria surface: The most sensitive redox-triggered one step nano-amplifier for fluorescence signal of ochratoxin A. J. Nanostructure Chem. 2022, 12, 223–233. [Google Scholar] [CrossRef]
  17. Othman, A.; Hayat, A.; Andreescu, S. Eu-doped ceria nanocrystals as nanoenzyme fluorescent probes for biosensing. Appl. Nano Mater. 2018, 1, 5722–5735. [Google Scholar] [CrossRef]
  18. Gubernatorova, E.O.; Liu, X.; Othman, A.; Muraoka, W.T.; Koroleva, E.; Andreescu, S.; Tumanov, A.V. Europium-doped cerium oxide nanoparticles limit reactive oxygen species formation and ameliorate intestinal ischemia–reperfusion injury. Adv. Healthc. Mater. 2017, 6, 1700176. [Google Scholar] [CrossRef]
  19. Pansambal, S.; Oza, R.; Borgave, S.; Chauhan, A.; Bardapurkar, P.; Vyas, S.; Ghotekar, S. Bioengineered cerium oxide (CeO2) nanoparticles and their diverse applications: A review. Appl. Nanosci. 2023, 13, 6067–6092. [Google Scholar] [CrossRef]
  20. Saule, T.X.T.; Molinari, M.; Das, S.; Bhatta, U.M.; Mobus, G.; Parker, S.C.; Seal, S.; Sayle, D.C. Environment-mediated structure, surface redox activity and reactivity of ceria nanoparticles. Nanoscale 2013, 5, 6063–6073. [Google Scholar] [CrossRef]
  21. Leel, N.S.; Kiran, M.; Kumawat, M.K.; Alvi, P.A.; Vats, V.S.; Patidar, D.; Dalela, B.; Kumar, S.; Dalela, S. Oxygen vacancy driven luminescence, ferromagnetic and electronic structure properties of Eu doped CeO2 nanoparticles. J. Lumin. 2023, 263, 119981. [Google Scholar] [CrossRef]
  22. Yang, H.; Xu, B.; Zhang, Q.; Yuan, S.; Zhang, Z.; Liu, Y.; Nan, Z.; Zhang, M.; Ohno, T. Boosting visible-light-driven photocatalytic performance of waxberry-like CeO2 by samarium doping and silver QDs anchoring. Appl. Catal. B 2021, 286, 119845. [Google Scholar] [CrossRef]
  23. Kumar, M.P.; Suganya Josephine, G.A.; Sivasamy, A. Oxidation of organic dye using nanocrystalline rare earth metal ion doped CeO2 under UV and Visible light irradiations. J. Mol. Liq. 2017, 242, 789–797. [Google Scholar] [CrossRef]
  24. Singh, K.; Kumar, K.; Srivastava, S.; Chowdhury, A. Effect of rare-earth doping in CeO2 matrix: Correlations with structure, catalytic and visible light photocatalytic properties. Ceram. Int. 2017, 43, 17041–17047. [Google Scholar] [CrossRef]
  25. Vinodkumar, T.; Rao, B.G.; Reddy, B.M. Influence of isovalent and aliovalent dopants on the reactivity of cerium oxide for catalytic applications. Catal. Today 2015, 253, 57–64. [Google Scholar] [CrossRef]
  26. Trovarelli, A. Structural and oxygen storage/release properties of CeO2-based solid solutions. Comments Inorg. Chem. 1999, 20, 263–284. [Google Scholar] [CrossRef]
  27. Tang, Y.; Zhang, H.; Cui, L.; Ouyang, C.; Shi, S.; Tang, W.; Li, T.; Lee, J.S.; Chen, L. First-principles investigation on redox properties of M-doped CeO2 (M = Mn, Pr, Sn, Zr). Phys. Rev. B 2010, 82, 125104. [Google Scholar] [CrossRef]
  28. Reddy, B.M.; Katta, L.; Thrimurthulu, G. Novel nanocrystalline Ce1−xLaxO2−δ (x = 0.2) solid solutions: Structural characteristics and catalytic performance. Chem. Mater. 2010, 22, 467–475. [Google Scholar] [CrossRef]
  29. Reddy, B.M.; Thrimurthulu, G.; Katta, L.; Yamada, Y.; Park, S.E. Structural characteristics, and catalytic activity of nanocrystalline ceria-praseodymia solid solutions. J. Phys. Chem. C 2009, 113, 15882–15890. [Google Scholar] [CrossRef]
  30. Wang, Z.; Quan, Z.; Lin, J. Remarkable changes in the optical properties of CeO2 nanocrystals induced by lanthanide ions doping. Inorg. Chem. 2007, 46, 5237–5242. [Google Scholar] [CrossRef]
  31. Ramjeyanthi, N.; Alagar, M.; Muthuraman, D. Synthesis, structural and optical behavior of cerium oxide nanoparticles by co-precipitation method. Int. J. Sci. Res. Sci. Tech. 2018, 4, 44–51. [Google Scholar] [CrossRef]
  32. Phokha, S.; Pinitsoontorn, S.; Chirawatkul, P.; Poo-arporn, Y.; Maensiri, S. Synthesis, characterization, and magnetic properties of monodisperse CeO2 nanospheres prepared by PVP-assisted hydrothermal method. Nanoscale Res. Lett. 2012, 7, 425. [Google Scholar] [CrossRef] [PubMed]
  33. Marabelli, F.; Wachter, P. Covalent insulator CeO2: Optical reflectivity measurements. Phys. Rev. B 1987, 36, 1238–1243. [Google Scholar] [CrossRef] [PubMed]
  34. Li, F.; Yu, X.H.; Pan, H.J.; Wang, M.L.; Xin, X.Q. Syntheses of MO2 (M = Si, Ce, Sn) nanoparticles by solid-state reactions at ambient temperature. Solid State Sci. 2000, 2, 767–772. [Google Scholar] [CrossRef]
  35. Derevyannikova, E.A.; Kardash, T.Y.; Kibis, L.S.; Slavinskaya, E.M.; Svetlichnyi, V.A.; Stonkus, O.A.; Ivanova, S.; Boronin, A.I. The structure and catalytic properties of Rh-doped CeO2 catalysts. Phys. Chem. Chem. Phys. 2017, 19, 31883–31897. [Google Scholar] [CrossRef] [PubMed]
  36. Zheng, H.; Hong, Y.; Xu, J.; Xue, B.; Li, Y.-X. Transesterification of ethylene carbonate to dimethyl carbonate catalyzed by CeO2 materials with various morphologies. Catal. Commun. 2018, 106, 6–10. [Google Scholar] [CrossRef]
  37. Si, R.; Zhang, Y.W.; You, L.P.; Yan, C.H. Rare-Earth Oxide Nanopolyhedra, Nanoplates, and Nanodisks. Angew. Chem. 2005, 44, 3256. [Google Scholar] [CrossRef]
  38. Phuruangrat, A.; Thongtem, S.; Thongtem, T. Microwave-assisted hydrothermal synthesis and characterization of CeO2 nanowires for using as a photocatalytic material. Mater. Lett. 2017, 196, 61–63. [Google Scholar] [CrossRef]
  39. Santiago, A.A.G.; Andrade Neto, N.F.; Longo, E.; Paskocimas, C.A.; Motta, F.V.; Bomio, M.R.D. Fast and continuous obtaining of Eu3+ doped CeO2 microspheres by ultrasonic spray pyrolysis: Characterization and photocatalytic activity. J. Mater. Sci. Mater. Electron. 2019, 30, 11508–11519. [Google Scholar] [CrossRef]
  40. Yin, L.X.; Wang, Y.Q.; Pang, G.S.; Koltypin, Y.; Gedanken, A.J. Sonochemical syntesis of cerium oxide nanoparticles—Effect of additives and quantum size effect. Colloid Interface Sci. 2002, 246, 78–84. [Google Scholar] [CrossRef]
  41. Bakkiyaraj, R.; Bharath, G.; Hasini Ramsait, K.; Abdel-Wahab, A.; Alsharaeh, E.H.; Chen, S.M.; Balakrishnan, M. Solution combustion synthesis and physico-chemical properties of ultrafine CeO2 nanoparticles and their photocatalytic activity. RSC Adv. 2016, 6, 51238–51245. [Google Scholar] [CrossRef]
  42. Amoresi, R.A.C.; Oliveira, R.C.; Marana, N.L.; De Almeida, P.B.; Prata, P.S.; Zaghete, M.A.; Longo, E.; Sambrano, J.R.; Simoes, A.Z. CeO2 nanoparticle morphologies and their corresponding crystalline planes for the photocatalytic degradation of organic pollutants. ACS Appl. Nano Mater. 2019, 2, 6513–6526. [Google Scholar] [CrossRef]
  43. Ortega, P.P.; Amoresi, R.A.C.; Teodoro, M.D.; Longo, E.; Ponce, M.A.; Simoes, A.Z. Relationship among morphology, photoluminescence emission, and photocatalytic activity of Eu-doped ceria nanostructures: A surface-type effect. Ceram. Int. 2023, 49, 21411–21421. [Google Scholar] [CrossRef]
  44. Suciu, C.; Hoffmann, A.C.; Vik, A.; Goga, F. Effect of calcination conditions and precursor proportions on the properties of YSZ nanoparticles obtained by modified sol–gel route. Chem. Eng. J. 2008, 138, 608–615. [Google Scholar] [CrossRef]
  45. Scherrer, P. Estimation of the size and structure of colloidal particles by Röentgen rays. Nach. Ges. Wiss. Gottingen 1918, 98–100. [Google Scholar]
  46. Laugier, J.; Bochu, B. CELREF V3, Developed at the Laboratoire des Matériaux et du Génie Physique; Ecole Nationale Supérieure de Physique de Grenoble (INPG): Grenoble, France, 2003. [Google Scholar]
  47. Stefan, M.; Leostean, C.; Pana, O.; Toloman, D.; Popa, A.; Perhaita, I.; Senila, M.; Marincas, O.; Barbu-Tudoran, L. Magnetic recoverable Fe3O4-TiO2:Eu composite nanoparticles with enhanced photocatalytic activity. Appl. Surf. Sci. 2016, 390, 248–259. [Google Scholar] [CrossRef]
  48. Chiang, T.L.; Wang, Y.C.; Ding, W.H. Trace determination of rhodamine B and rhodamine 6G dyes in aqueous samples by solid-phase extraction and high-performance liquid chromatography coupled with fluorescence detection. J. Chin. Chem. Soc. 2012, 59, 515–519. [Google Scholar] [CrossRef]
  49. Yang, N.; Orgiani, P.; Di Bartolomeo, E.; Foglietti, V.; Torelli, P.; Ievlev, A.V.; Rossi, G.; Licoccia, S.; Balestrino, G.; Kalinin, S.V.; et al. Effects of dopant ionic radius on cerium reduction in epitaxial cerium oxide thin films. J. Phys. Chem. C 2017, 121, 8841–8849. [Google Scholar] [CrossRef]
  50. Hong, S.J.; Virkar, A.V. Lattice parameters and densities of rare-earth oxide doped ceria electrolytes. J. Am. Ceram. Soc. 1995, 78, 433–439. [Google Scholar] [CrossRef]
  51. Shannon, R.D.; Prewitt, C.T. Effective ionic radii in oxides and fluorides. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1969, 25, 925–946. [Google Scholar] [CrossRef]
  52. Shannon, R.D. Revised Effective Ionic Radii and Systematic Studies of Interatomic Distances in Halides and Chalcogenides. Acta Cryst. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  53. Goharshadia, E.K.; Samiee, S.; Nancarrow, P. Fabrication of cerium oxide nanoparticles: Characterization and optical properties. J. Colloid Interface Sci. 2011, 356, 473–480. [Google Scholar] [CrossRef] [PubMed]
  54. Deus, R.C.; Cortés, J.A.; Ramirez, M.A.; Ponce, M.A.; Andres, J.; Rocha, L.S.R.; Longo, E.; Simões, A.Z. Photoluminescence properties of cerium oxide nanoparticles as a function of lanthanum content. Mater. Res. Bull. 2015, 70, 416–423. [Google Scholar] [CrossRef]
  55. Sun, C.; Li, H.; Zhang, H.; Wang, Z.; Chen, L. Controlled synthesis of CeO2 nanorods by a solvothermal method. Nanotechnology 2005, 16, 1454–1463. [Google Scholar] [CrossRef]
  56. Liu, X.; Chen, S.; Wang, X. Synthesis and photoluminescence of CeO2:Eu3+ phosphor powders. J. Lumin. 2007, 127, 650–654. [Google Scholar] [CrossRef]
  57. Singh, M.; Goyal, M.; Devlal, K. Size and shape effects on the band gap of semiconductor compound nanomaterials. J. Taibah Univ. Sci. 2018, 12, 470–475. [Google Scholar] [CrossRef]
  58. Binnemans, K. Interpretation of europium(III) spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef]
  59. Li, L.; Yang, H.K.; Moon, B.K.; Fu, Z.; Guo, C.; Jeong, J.H.; Yi, S.S.; Jang, K.; Lee, H.S. Photoluminescence properties of CeO2:Eu3+ nanoparticles synthesized by a sol-gel method. J. Phys. Chem. C 2009, 113, 610–617. [Google Scholar] [CrossRef]
  60. Vimal, G.; Mani, K.P.; Biju, P.R.; Joseph, C.; Unnikrishnan, N.V.; Ittyachen, M.A. Structural studies and luminescence properties of CeO2:Eu3+ nanophosphors synthesized by oxalate precursor method. Appl. Nanosci. 2015, 5, 837–846. [Google Scholar] [CrossRef]
  61. Hernández-Alonso, M.D.; Hungria, A.B.; Martinez-Arias, A.; Fernández-Garcia, M.; Coronado, J.M.; Carlos Conesa, J.; Soria, J. EPR study of the photoassisted formation of radicals on CeO2 nanoparticles employed for toluene photooxidation. Appl. Catal. B Environ. 2004, 50, 167–175. [Google Scholar] [CrossRef]
  62. Oliva, C.; Termignone, G.; Vatti, F.P.; Forni, L.; Vishniakov, A.V. Electron paramagnetic resonance spectra of CeO2 catalyst for CO oxidation. J. Mater. Sci. 1996, 31, 6333–6338. [Google Scholar] [CrossRef]
  63. Soria, J.; Conesa, J.C.; Martinez-Arias, A. Characterization of surface defects in CeO2 modified by incorporation of precious metals from chloride salts precursors: An EPR study using oxygen as probe molecule. Colloids Surf. A Physicochem. Eng. Asp. 1999, 158, 67–74. [Google Scholar] [CrossRef]
  64. Cerrato, E.; Gionco, C.; Paganini, M.C.; Giamello, E.; Albanese, E.; Pacchioni, G. Origin of visible light photoactivity of the CeO2/ZnO heterojunction. ACS Appl. Energy Mater. 2018, 1, 4247–4260. [Google Scholar] [CrossRef]
  65. Li, Y.F.; Zhang, W.P.; Li, X.; Yu, Y. TiO2 nanoparticles with high ability for selective adsorption and photodegradation of textile dyes under visible light by feasible preparation. J. Phys. Chem. Solids. 2014, 75, 86–93. [Google Scholar] [CrossRef]
  66. Brezova, V.; Gabcova, S.; Dvoranova, D.; Stasko, A. Reactive oxygen species produced upon photoexcitation of sunscreens containing titanium dioxide (an EPR study). J. Photochem. Photobiol. B 2005, 79, 121–134. [Google Scholar] [CrossRef] [PubMed]
  67. Lázaro-Martínez, M.; Leal-Denis, M.F.; Piehl, L.L.; Rubín de Celis, E.; Buldain, G.Y.; Dall’-Orto, V.C. Studies on the activation of hydrogen peroxide for color removal in the presence of a new Cu(II)-polyampholyte heterogeneous catalyst. Appl. Catal. B Environ. 2008, 82, 273–283. [Google Scholar] [CrossRef]
  68. Yang, Z.; Xu, M.; Liu, Y.; He, F.; Gao, F.; Su, Y.; Wei, H.; Zhang, Y. Nitrogen-doped, carbon-rich, highly photoluminescent carbon dots from ammonium citrate. Nanoscale 2014, 6, 1890–1895. [Google Scholar] [CrossRef]
  69. Veedu, S.N.; Jose, S.; Narendranath, S.B.; Kurup, M.R.P.; Periyat, P. Visible light-driven photocatalytic degradation of methylene blue dye over bismuth-doped cerium oxide mesoporous nanoparticles. Environ. Sci. Pollut. Res. 2021, 28, 4147–4155. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of pure ceria, CeO2; Y-doped CeO2, CeO2:Y; Eu-doped ceria, CeO2:Eu; and Gd-doped ceria, CeO2:Gd. All reflections have been indexed according to the JCDD PDF 034-0394 for CeO2 face-centred cubic fluorite structure.
Figure 1. X-ray diffraction patterns of pure ceria, CeO2; Y-doped CeO2, CeO2:Y; Eu-doped ceria, CeO2:Eu; and Gd-doped ceria, CeO2:Gd. All reflections have been indexed according to the JCDD PDF 034-0394 for CeO2 face-centred cubic fluorite structure.
Applsci 14 00522 g001
Figure 2. TEM images of undoped ceria, CeO2; Y-doped CeO2, CeO2:Y; Eu-doped ceria, CeO2:Eu; and Gd-doped ceria, CeO2:Gd. The scale bar is 50 nm for all images, except the inset of CeO2:Y; here, the scale bar is 2 μm.
Figure 2. TEM images of undoped ceria, CeO2; Y-doped CeO2, CeO2:Y; Eu-doped ceria, CeO2:Eu; and Gd-doped ceria, CeO2:Gd. The scale bar is 50 nm for all images, except the inset of CeO2:Y; here, the scale bar is 2 μm.
Applsci 14 00522 g002
Figure 3. (a) Room temperature optical absorption spectra and (b) Tauc plots of undoped and RE doped ceria nanoparticles.
Figure 3. (a) Room temperature optical absorption spectra and (b) Tauc plots of undoped and RE doped ceria nanoparticles.
Applsci 14 00522 g003
Figure 4. The room temperature photoluminescence spectra of undoped and RE-doped CeO2exc = 325 nm).
Figure 4. The room temperature photoluminescence spectra of undoped and RE-doped CeO2exc = 325 nm).
Applsci 14 00522 g004
Figure 5. (a) EPR spectra of CeO2-Y nanoparticles obtained without and under UV irradiation. (b) Monitorization of • O 2 signal under UV irradiation in three on-off cycles.
Figure 5. (a) EPR spectra of CeO2-Y nanoparticles obtained without and under UV irradiation. (b) Monitorization of • O 2 signal under UV irradiation in three on-off cycles.
Applsci 14 00522 g005
Figure 6. Evolution in time of RhB absorbance after dark adsorption and visible irradiation in the presence of CeO2-Gd sample.
Figure 6. Evolution in time of RhB absorbance after dark adsorption and visible irradiation in the presence of CeO2-Gd sample.
Applsci 14 00522 g006
Figure 7. Removal rate of RhB solution in the presence of each sample.
Figure 7. Removal rate of RhB solution in the presence of each sample.
Applsci 14 00522 g007
Figure 8. Evaluation of photodegradation kinetic.
Figure 8. Evaluation of photodegradation kinetic.
Applsci 14 00522 g008
Figure 9. Experimental and simulated spectra of DMPO spin adducts generated by Y-CeO2 photocatalyst suspension in DMSO after 25 min visible irradiation.
Figure 9. Experimental and simulated spectra of DMPO spin adducts generated by Y-CeO2 photocatalyst suspension in DMSO after 25 min visible irradiation.
Applsci 14 00522 g009
Figure 10. Scavenger experiment results of Y-CeO2 nanoparticles for RhB degradation (control experiment = no scavenger).
Figure 10. Scavenger experiment results of Y-CeO2 nanoparticles for RhB degradation (control experiment = no scavenger).
Applsci 14 00522 g010
Figure 11. Degradation mechanism of Y-doped CeO2 nanoparticles.
Figure 11. Degradation mechanism of Y-doped CeO2 nanoparticles.
Applsci 14 00522 g011
Table 1. Average crystallite sizes (D) and lattice parameters (a) of undoped and doped ceria.
Table 1. Average crystallite sizes (D) and lattice parameters (a) of undoped and doped ceria.
SampleCrystallite Size
D (nm) a
Lattice Parameter
a (Å) a
Energy Band Gap
Eg (eV)
Undoped CeO28.165.41052.98
CeO2:Eu8.405.41302.93
CeO2:Gd8.405.41292.88
CeO2:Y7.795.41132.93
a Calculates from XRD.
Table 2. The apparent first-order rate constant k of photocatalytic degradation and correlation coefficient, R2.
Table 2. The apparent first-order rate constant k of photocatalytic degradation and correlation coefficient, R2.
SampleK × 10−3R2
CeO20.370.9860
CeO2:Eu1.130.9441
CeO2:Gd1.090.9244
CeO2:Y1.240.9594
Table 3. HPLC-FD concentrations of RhB before and after photodegradation with CeO2-doped catalysts (CeO2:Eu; CeO2:Gd; CeO2:Y).
Table 3. HPLC-FD concentrations of RhB before and after photodegradation with CeO2-doped catalysts (CeO2:Eu; CeO2:Gd; CeO2:Y).
CatalystNo CatalystCeO2CeO2:EuCeO2:YCeO2:Gd
Rhodamine B [mg/L]4.6402.4401.3860.7000.430
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

Toloman, D.; Popa, A.; Sonher, R.B.; Bortnic, R.; Marinca, T.F.; Perhaita, I.; Filip, M.; Mesaros, A. Enhancing the Photocatalytic Activity and Luminescent Properties of Rare-Earth-Doped CeO2 Nanoparticles. Appl. Sci. 2024, 14, 522. https://doi.org/10.3390/app14020522

AMA Style

Toloman D, Popa A, Sonher RB, Bortnic R, Marinca TF, Perhaita I, Filip M, Mesaros A. Enhancing the Photocatalytic Activity and Luminescent Properties of Rare-Earth-Doped CeO2 Nanoparticles. Applied Sciences. 2024; 14(2):522. https://doi.org/10.3390/app14020522

Chicago/Turabian Style

Toloman, Dana, Adriana Popa, Ramona Bianca Sonher, Rares Bortnic, Traian Florin Marinca, Ioana Perhaita, Miuța Filip, and Amalia Mesaros. 2024. "Enhancing the Photocatalytic Activity and Luminescent Properties of Rare-Earth-Doped CeO2 Nanoparticles" Applied Sciences 14, no. 2: 522. https://doi.org/10.3390/app14020522

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