Next Article in Journal
Prenatal Exposure to Mercury, Manganese, and Lead and Adverse Birth Outcomes in Suriname: A Population-Based Birth Cohort Study
Previous Article in Journal
Biochar Is Not Durable for Remediation of Heavy Metal-Contaminated Soils Affected by Acid-Mine Drainage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ce3+/Ce4+-Doped ZrO2/CuO Nanocomposite for Enhanced Photocatalytic Degradation of Methylene Blue under Visible Light

1
Faculty of Chemistry, Thai Nguyen University of Education, Thai Nguyen City 24000, Vietnam
2
Faculty of Physics, Thai Nguyen University of Education, Thai Nguyen City 24000, Vietnam
3
Institute of Applied Technology and Sustainable Development, Nguyen Tat Thanh University, Ho Chi Minh City 700000, Vietnam
4
Advanced Institute of Science and Technology, Hanoi University of Science and Technology, Hanoi City 100000, Vietnam
*
Authors to whom correspondence should be addressed.
Toxics 2022, 10(8), 463; https://doi.org/10.3390/toxics10080463
Submission received: 20 June 2022 / Revised: 17 July 2022 / Accepted: 18 July 2022 / Published: 10 August 2022
(This article belongs to the Section Toxicity Reduction and Environmental Remediation)

Abstract

:
In recent years, photocatalysis has been used as an environmentally friendly method for the degradation of organic pigments in water. In this study, Ce3+/Ce4+-doped ZrO2/CuO as a mixed semiconductor oxide was successfully prepared by a one-step hydrothermal method. The Ce3+/Ce4+-doped ZrO2/CuO has shown high degradation efficiency of methylene blue (MB), and the maximum degradation percentage was observed to be 94.5% at 180 min under irradiation visible light. The photocatalytic activity increases significantly by doping Ce3+/Ce4+ in ZrO2/CuO for MB degradation. Ce3+/Ce4+ doping is shown to reduce the (e-/h+) recombination rate and improve the charge transfer, leading to enhanced photocatalytic activity of materials. The materials were characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), transmission electron microscopy (TEM), FTIR, EDS, BET and diffuse reflectance spectroscopy (DRS).

1. Introduction

Water pollution is a global environmental problem as it can lead to the decomposition of aquatic ecosystems and can affect human health. Contaminants can include organic or inorganic compounds, metal ions, dyes, phenols, pesticides and detergents. In which the pollution of natural and synthetic organic dyes poses challenges because of their high carcinogenic nature due to the fact they contain azo functional groups that, during decomposition, can form amines and benzidine [1]. Organic dyes are persistent organic pollutants that are resistant to decomposition through chemical and biological processes. The dye molecules are practically non-biodegradable, so they persist longer in the environment, creating potential hazards [2].
Today, water pollution is treated by applying technologies such as adsorption, coagulation, filtration and photocatalysis. In which the heterogeneous photocatalysis method has gained wide scientific attention with high potential for applications [3]. Zirconia (ZrO2)-based photocatalysis for wastewater treatment is of great interest because ZrO2 is a semiconductor oxide with high mechanical strength, is non-toxic, and has high chemical stability and biocompatibility. However, ZrO2 has a large band gap; the Eg of ZrO2 ranges from 3.25 to 5.1 eV, depending on the fabrication method [4]. Therefore, ZrO2 needs to be irradiated with UV radiation to have a photocatalytic activity. UV radiation constitutes about 5% of the total electromagnetic radiation output from sunlight, limiting the practical applications of ZrO2.
To improve the photocatalytic performance of ZrO2, coupling with other semiconductor oxides has been shown to be an effective approach. The coupling produces mixed oxide semiconductors (MOS) that are capable of efficient charge separation due to the transition of electrons and holes generated from one semiconductor to another. CuO is often the semiconductor oxide of choice because it is a p-type semiconductor, has a small band gap of about 1.2 eV, is cheap and is environmentally friendly [5]. Renuka, L., et al., reported the synthesis of ZrO2/CuO materials with excellent photocatalytic activity under visible light by a simple combustion method. The photocatalytic activity of ZrO2/CuO obtained was 1.5 times higher than that of commercial P25 [6]. Nanda, B., et al., synthesized mesoporous CuO/ZrO2–MCM–41 nanocomposites with high photocatalytic activity, the photoreduction of Cr(VI) to Cr(III) had a 99% degradation efficiency in 30 min [7]. The synergistic effect between ZrO2, CuO and TiO2 oxides was also studied by Guerrero-Araque and Diana, et al. Co-catalysts are widely used to promote photocatalytic hydrogen production, especially CuO, which has shown a significant improvement in reaction rates [8].
Another use of low-photocatalytic activity is the electron-hole recombination process that occurs in the semiconductor oxide. In photocatalysis, the photocatalytic activity depends on the ability of the catalyst to create electron-hole pairs. Such an electron-hole pair is called an exciton. The excited electron and hole can recombine and release heat. Such exciton recombination leads to a decrease in photocatalytic activity. To reduce the recombination of electrons and holes, it is common to dope transition metal elements, especially rare earth, to prolongation of the exciton lifetime [9].
Recently, Piña-Pérez and Yanet, et al., doped Ce3+/Ce4+ into Al2O3, showing significantly improved electron-hole pair separation efficiency. This catalyst was chosen for the photodegradation of other phenolic derivatives such as 4–chlorophenol, p–cresol and 4–nitrophenol. Ce3+/Ce4+-doped Al2O3 synthesized by the sol-gel method showed better photocatalytic activity (TiO2, P25 Degussa) [10]. The redox pair of Ce4+/Ce3+-doped α-β phase of Bi2O3 was reported by Akshatha, S., et al. The Ce4+/Ce3+-doped Bi2O3 created a synergistic effect to enhance the photocatalytic activity to degrade Alizarin red S dye [11].
To the best of our knowledge, there are many studies on the doping of Ce3+/Ce4+ ions into oxide semiconductors for photocatalysis. However, the doping of Ce3+/Ce4+ ions into mixed semiconductor oxide is a topic that has not been fully investigated. In this paper, we use the hydrothermal method to prepare Ce3+/Ce4+-doped ZrO2/CuO. The Ce3+/Ce4+-doped ZrO2/CuO mixed oxide semiconductor was used for the photocatalytic degradation of methylene blue (MB). X-ray diffraction (XRD), field emission scanning electron microscopy (FE-SEM), transmission electron microscopy (TEM), infrared spectroscopy (FT-IR), energy dispersive X-ray spectroscopy, spectrophotometry UV-Vis and Brunauer–Emmett–Teller (BET) were studied.

2. Materials and Methods

2.1. Preparation of the Ce3+/Ce4+-Doped ZrO2/CuO Nanocomposites

The Ce3+/Ce4+-doped ZrO2/CuO nanocomposites were synthesized by the hydrothermal method. All chemicals were purchased from Merck (purity > 99%) and were used without further treatment: ZrCl4, Cu(NO3)2.6H2O, Ce(SO4)2.4H2O, NaOH, C2H5OH, Methylene blue.
In a typical experiment, the molar ratio of ZrO2/CuO was taken as 1:1, 10 mmol of ZrCl4 and 10 mmol of Cu(NO3)2 were dissolved in 50 mL of deionized water, stirred in a magnetic stirrer at room temperature for 30 min. Next, 0–0.8 mmol Ce(SO4)2 was added to the above solution. The mixture is stirred vigorously to obtain a clear and homogeneous mixture. After stirring these mixtures for 30 min, they were added to 10 mL of 2 M NH3 aqueous solution to gel.
The gel was then transferred into a Teflon-lined 100 mL stainless steel autoclave, which was heated in an oven at 200 °C for 12 h (heated with a Nabertherm furnace, heating rate 5 °C/min). For natural cooling, the entire solution was centrifuged to collect the precipitate and washed with deionized water, and then dried at 80 °C. Pure ZrO2 and CuO were prepared with the same procedure. Finally, all the prepared powders were calcined at 600 °C for 2 h and then allowed to cool to room temperature for further experiments.

2.2. Characterization

The morphology of the Ce3+/Ce4+-doped ZrO2/CuO nanocomposites was evaluated by using a field emission scanning electronic microscope (JEOL, JSM–7600F, JEOL Techniques, Tokyo, Japan). The surface morphology and microstructure of the nanocomposites were characterized by transmission electron microscopy (TEM) made with a (JEOL, JEM 1010, JEOL Techniques, Tokyo, Japan) operating at 200 kV. For the TEM analyses, the powders were dispersed in ethanol by sonication for 5 min.
FT-IR was performed in the range of 4000–400 cm−1 with the help of (FT-IR using a Perkin–Elmer Spectrum BX spectrometer (PerkinElmer Inc., Wellesley, MA, USA) using KBr pellets to identify the functional groups present in the sample. EDS (Hitachi TM4000Plus Tabletop Microscope(Hitachi High-Tech Corporation, Tokyo, Japan)) was the determining element in the composites. The UV-Vis absorption spectrum was recorded using a (UV–1700 PharmaSpec, Shimadzu, Kyoto, Japan) UV-Visible spectrophotometer. The X-ray powder diffraction patterns were characterized using (XRD, D8 Advance, Bruker, Germany) a diffractometer at room temperature (Cu-Kα radiation) with a nickel filter at a scan rate of 2°/min.

2.3. Dye Photodegradation

The photocatalytic activity of Ce3+/Ce4+-doped ZrO2/CuO nanocomposites was evaluated by investigating the degradation of methylene blue (MB). Using 30 mL of the MB solution with a dye concentration of 10 mgL−1, which was illuminated under a light-emitting 300W Xenon lamp (sunlight simulation) at irradiance (1 Wcm−2). Light filters were used to remove wavelengths λ < 420 nm. The distance from the lamp to the MB solution was 10 cm. To limit the influence of external light, the entire experimental system was placed in a dark chamber. The nanocomposite powder loaded into the MB solution was 20 mg. Then the suspension was stirred for 60 min in the dark to obtain balance adsorption and desorption equilibration of the system. About 3 mL was taken from the suspension at different irradiation times, it was then centrifuged at 4000 rpm for 10 min, and degradation measurements were performed by a UV-Vis spectrophotometer. The photodegradation efficiency was calculated using Equation (1):
% H = C 0 C t C 0 .100
where %H is the photodegradation efficiency, C0 and Ct are the concentrations of MB at time 0 and t (min), respectively.

3. Results

3.1. X-ray Diffraction

The XRD pattern of the Ce3+/Ce4+-doped ZrO2/CuO nanocomposites is shown in Figure 1. All the peaks in the XRD pattern are well indexed and have all crystallized as multiphase at Ce3+/Ce4+ different doping concentrations. ZrO2/CuO and Ce3+/Ce4+-doped ZrO2/CuO exists in three phases, namely tetragonal ZrO2, monoclinic CuO and orthorhombic CuZrO3.
The diffraction peaks originating at 2θ = ~ 30.48°, 34.35°, 50.30°, 50.65° and 60.09° correspond to (101), (200), (112), (220) and (202) planes of the tetragonal phase of ZrO2 (JCPDS No. 00–050–1089), respectively. The diffraction peaks observed at 2θ = ~ 35.68°, 38.89°, 48.93°, 53.62°, 61.70°, 65.97° and 68.20° are associated with (002), (111), (−202), (020), (−113), (022) and (113) planes, signifying the monoclinic phase of CuO (JCPDS No. 48–1548).
We also observed diffraction peaks at 2θ = ~ 24.33°, 28.29°, 34.34°, 55.56° and 61.72° corresponded to (012), (112), (013), (313) and (402) planes of CuZrO3 (PDF Card No. 00–043–0953), respectively. CuZrO3 can be formed in a hydrothermal process, where the reaction between ZrO2 and CuO occurs according to the chemical equation:
ZrO 2   + CuO   t 0 CuZrO 3
It is interesting that when the Ce3+/Ce4+ doping concentration changes, the ratio between the phases changes. In the absence of Ce3+/Ce4+ doping, the material exists in all three phases, including tetragonal ZrO2, monoclinic CuO and orthorhombic CuZrO3. However, when increasing the doping concentration of Ce3+/Ce4+, the diffraction intensity of the CuZrO3 phase decreased. According to the authors Dean, James, et al., reported that CuZrO3 has a perovskite structure; this structure is less stable due to its low energy, which has a thermodynamic preference to decompose into CuO and ZrO2 [12]. We believe that increasing the concentration of Ce3+/Ce4+ doping increases the ability of phase segregation to CuO and ZrO2 is preferred.

3.2. Microstructure and Morphology of ZrO2/CuO-Doped Ce3+/Ce4+ Nanocomposites

The microstructure of the Ce3+/Ce4+-doped ZrO2/CuO synthesized by the hydrothermal method was analyzed by FE-SEM. Figure 2a–d show that the Ce3+/Ce4+-doped ZrO2/CuO catalysts are spheres with non-uniform size distribution. When the Ce3+/Ce4+ doping concentration changed from 2 mol% to 8 mol%, the morphology of the catalysts was not significantly affected.
The size of the catalyst, as well as the surface area, greatly influence the photocatalytic efficiency. The small size of the material shows that the ability to adsorb the MB on the surface will be better, creating conditions for the oxidizing radicals to easily react with the colourant.
To better understand the shape, size and crystal formation of the Ce3+/Ce4+-doped ZrO2/CuO catalyst, we studied the transmission electron microscopy (TEM) images.
Figure 3 shows that the particle size of the Ce3+/Ce4+-doped ZrO2/CuO catalyst is larger than that of ZrO2/CuO. According to the particle size distribution curves in Figure 3c–d, ZrO2/CuO concentrates around 20 nm, while Ce3+/Ce4+-doped ZrO2/CuO concentrates around 30 nm.

3.3. Fourier Transform Infrared Spectroscopy and Energy Dispersive X-ray Spectroscopy of the Ce3+/Ce4+-Doped ZrO2/CuO Nanocomposites

Figure 4 shows the FT-IR spectra of ZrO2/CuO and Ce3+/Ce4+-doped ZrO2/CuO. The Ce3+/Ce4+-doped ZrO2/CuO materials all have absorption maximums around 3113–3153 cm−1, which is typical for the –OH group of wet water. The absorption peaks at 1710–1712 cm−1 and 1556–1566 cm−1 are characteristic of the vibrations of the Zr–O–H group in the material. In addition, the materials have an absorption maximum of 601–607 cm−1, which can be attributed to the vibrational stretching mode of the Cu-O bond. Bands at 424–524 cm−1 can be assigned to the Zr–O vibration mode in composites.
Figure 4b shows the presence of elements Ce, Zr, Cu and O, confirming the presence of Ce doping in ZrO2/CuO nanocomposites. We performed EDS mapping, which is equipped with SEM. Figure 4c clearly show the uniform spread of Zr, Cu, O and Ce elements.

3.4. Nitrogen Adsorption-Desorption Isotherms of the Ce3+/Ce4+-Doped ZrO2/CuO Nanocomposites

The specific surface areas of the Ce3+/Ce4+-doped ZrO2/CuO and ZrO2/CuO nanocomposites were determined by measuring the nitrogen adsorption-desorption isotherms, as shown in Figure 5. All of them are all type III adsorption isotherms [13]. The BET surface areas of ZrO2/CuO and Ce3+/Ce4+-doped ZrO2/CuO are 15.59 and 13.52 m2g−1. The large surface area of the catalyst facilitates the adsorption of the pigment to the surface of the material, thereby improving the photocatalyst efficiency.

3.5. UV-Vis Absorption Spectrum of Ce3+/Ce4+-Doped ZrO2/CuO Nanocomposites

Figure 6 shows the UV-Vis absorption spectrum and the bandgap energy of the Ce3+/Ce4+-doped ZrO2/CuO. The results show that the Ce3+/Ce4+-doped ZrO2/CuO oxide composite material has an absorption region from UV to the visible region (200–1000 nm). To calculate the band gap, the Tauc relation is used:
αhν = A(hν − Eg)n
where α is the absorption coefficient, h is the Planck constant, ν is the frequency of the incident photon, A is a constant that depends on the transition probability, Eg is the band gap energy and n is an index that depends on the transition probability of the nature of the electronic transition.
The bandgap energy of Ce3+/Ce4+-doped ZrO2/CuO was found to be in the range of 1.445–1.475 eV. When increasing the Ce3+/Ce4+ doping concentration from 0 to 8 mol%, the band gap increased slightly from 1.445 to 1.475 eV. When Ce was doped into ZrO2/CuO, the 5d and 6s orbitals of Ce overlapping with the d orbitals of Cu and Zr can increase the size of the valence band and also lower the position of its maximum, thereby increasing the band gap of mixed semiconductor oxide. The increase in the optical band gap upon Ce doping is similar to the increase in the band gap of Ag2O upon Zn doping, as reported by De, Arup Kumar et al. [14]. The decrease in the band gap of the ZrO2/CuO nanocomposites material compared with that of pure ZrO2 is mainly due to the introduction of CuO into the ZrO2 host matrix. It was found that when two metals combine, the band gap decreases and thus a shift to the visible region is observed [15].
The narrowing of the band gap can be attributed to the appearance of an impurity region (CuO) formed by overlapping impurity states and local states formed by the combination of Cu 2p and Zr 3p [7]. This result is similar to previous studies [16]. The photocatalytic reaction is initiated by the absorption of light with an energy equal to or greater than the band gap of the semiconductor. The band gap is in the range of 1.445–1.475 eV, so it is expected that the catalyst material can absorb visible light.

3.6. Photocatalytic Activity of Ce3+/Ce4+-Doped ZrO2/CuO Nanocomposites

To investigate the adsorption balance, 20 mg of 8 mol% Ce3+/Ce4+-doped ZrO2/CuO was added into 30 mL of 10 mgL−1 MB solution. The suspension was stirred in the dark; samples were measured via UV-Vis spectroscopy. Figure 7 shows that after the first 30 min, the concentration of MB strongly decreased, then at 60 min slightly decreased, and at 90 min, the concentration of MB increased slightly (desorption). It is shown that after 60 min of adsorption, equilibration has occurred.
Figure 8 shows the UV-Vis absorption spectrum of MB under visible light at different times in the presence of Ce3+/Ce4+-doped ZrO2/CuO. The evolution of MB photodegradation was monitored by measuring the absorbance at the wavelength of about λ = 662 nm at different times. The maximum absorption peaks of MB gradually decreased and almost disappeared in 180 min of illumination.
The reactions that occur during photocatalysis are described below. The reaction begins by generating excitons on the metal oxide surface (MeO) [17]:
MeO + hν → MeO (h+ + e)
Oxidation reactions due to the photocatalytic effect:
h + + H2O → H+ + •OH
2h+ + 2 H2O → 2H+ + H2O2
H2O2 → 2 •OH
Reduction reaction due to the photocatalytic effect:
e + O2 → •O2
•O2 + H2O + H+ → H2O2 + O2
H2O2 → 2•OH
•OH + Dye → Degradation
•O2 + Dye → Degradation
Figure 9a shows the ZrO2/CuO catalyst, the photocatalytic efficiency of MB decomposition by Ce3+/Ce4+-doped ZrO2/CuO increased significantly. The Ce3+/Ce4+ doping concentration increased, and the photocatalytic efficiency under visible light also increased. Although the Eg values of ZrO2/CuO and ZrO2/CuO-doped Ce3+/Ce4+ were not significantly different, the results showed that the photocatalytic degradation efficiency was enhanced by Ce3+/Ce4+ doping. This could be due to the fact that as the Ce3+/Ce4+ ion concentration increases, an increase in oxygen vacancy defects in the crystal lattice occurs. This leads to the redox pair Ce3+/Ce4+ being capable of shifting between Ce2O3 and CeO2, which reduces electron-hole recombination [18].
The MB degradation processes all show first-order kinetics by plotting ln(C/C0) versus irradiation time, t. The apparent response rate constant (kapp) was calculated from the slope of the curve, as illustrated in Figure 9b.
The kapp value continuously increased as the Ce3+/Ce4+ doping concentration increased, with x = 0, kapp = 0.0074 min−1; x = 2, kapp = 0.0093 min−1; x = 4, kapp = 0.0122 min−1; x = 6, kapp = 0.0123 min−1; and x = 8, kapp = 0.0138 min−1. It was shown that 8 mol% Ce3+/Ce4+-doped ZrO2/CuO exhibited the highest kapp = 0.0138 min−1, much higher than that of undoped ZrO2/CuO with Ce3+/Ce4+.
Figure 9c shows the stability of the Ce3+/Ce4+-doped ZrO2/CuO nanocomposites. The stability of the MOS material was evaluated by performing photocatalyst recycling experiments for a duration of 180 min. The activity was found to be roughly the same in the two repeated runs, and then there was a slight decrease in the third and fourth runs.
We assume that there is metal-to-metal charge transfer (MMCT). The MMCT effect is assumed to be: Zr4+–O–Ce4+ to Zr4+–O–Ce3+ and Zr4+–O–Cu2+ to Zr4+–O–Cu+. The effect is when two different metals in the mixture form an oxygen bridge, which helps electrons and holes move efficiently and avoid recombination, increasing the photocatalytic efficiency of the Ce3+/Ce4+-doped ZrO2/CuO catalyst [19]. Furthermore, Ce3+/Ce4+ ions act as electron-trapping sites reducing charge-pair recombination and thus enhancing photocatalytic activity [20,21].
Table 1 shows the photocatalytic degradation of MB using the various heterojunction photocatalyst. It shows that the combination of semiconductor oxides, or metal ion doping, shows an improved rate constant compared to single semiconductor oxides.
The ZrO2/CuO-doped Ce3+/Ce4+ photocatalyst can be reproducible and without destruction. This reproducible process can be confirmed by FT-IR, as shown in Figure 10. It is clear that the characteristic bands of MB, which are at 1383, 1324, 1245, 884 and 670 cm−1, appear after adsorbing [31], while they disappear after 180 min of irradiation. Combined with the results of photocatalytic reuse experiments, it is proven that the Ce3+/Ce4+-doped ZrO2/CuO nanocomposites have stability.

4. Conclusions

Ce3+/Ce4+-doped ZrO2/CuO nanocomposites have been synthesized by the simple hydrothermal method. The obtained material is a mixed semiconductor oxide ZrO2 and CuO. The coupling between CuO and ZrO2 oxides has shown that the light absorption capacity is extended from UV to the visible region. Photocatalytic activity is enhanced when doping Ce3+/Ce4+ ions because Ce3+/Ce4+ ions have higher separation efficiency of charge carriers, which reduces electron-hole recombination. The Ce3+/Ce4+-doped ZrO2/CuO nanocomposites showed significantly improved photocatalytic activity. High photocatalysis activity with 94.5% degradation efficiency of MB was achieved after 180 min under visible light irradiation. It is believed that these materials will have promising applications in the wastewater treatment field.

Author Contributions

Writing—original draft preparation, M.N.C., L.T.H.N. and V.H.P.; software, M.X.T. and T.H.D.; data curation, T.T.A.D.; conceptualization, M.A.P., T.C.Q.N. and L.T.T.N.; writing—review and editing M.N.C., T.K.N.T. and V.H.P.; Supervision: M.N.C.; L.T.H.N. and M.X.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data are available within the manuscript.

Acknowledgments

The authors thank Thai Nguyen University of Education (TNUE, Vietnam) for providing all the necessary facilities to complete this study.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Maheshwari, K.; Agrawal, M.; Gupta, A.B. Dye Pollution in Water and Wastewater. In Novel Materials for Dye-Containing Wastewater Treatment; Springer: Berlin/Heidelberg, Germany, 2021; pp. 1–25. [Google Scholar]
  2. Alharbi, O.M.L.; Khattab, R.A.; Ali, I. Health and environmental effects of persistent organic pollutants. J. Mol. Liq. 2018, 263, 442–453. [Google Scholar] [CrossRef]
  3. Anwer, H.; Mahmood, A.; Lee, J.; Kim, K.-H.; Park, J.-W.; Yip, A.C.K. Photocatalysts for degradation of dyes in industrial effluents: Opportunities and challenges. Nano Res. 2019, 12, 955–972. [Google Scholar] [CrossRef]
  4. Reddy, B.M.; Khan, A. Recent advances on TiO2-ZrO2 mixed oxides as catalysts and catalyst supports. Catal. Rev. 2005, 47, 257–296. [Google Scholar] [CrossRef]
  5. Fazl, F.; Gholivand, M.B. High performance electrochemical method for simultaneous determination dopamine, serotonin, and tryptophan by ZrO2–CuO co-doped CeO2 modified carbon paste electrode. Talanta 2022, 239, 122982. [Google Scholar] [CrossRef]
  6. Renuka, L.; Anantharaju, K.S.; Vidya, Y.S.; Nagaswarupa, H.P.; Prashantha, S.C.; Sharma, S.C.; Nagabhushana, H.; Darshan, G.P. A simple combustion method for the synthesis of multi-functional ZrO2/CuO nanocomposites: Excellent performance as Sunlight photocatalysts and enhanced latent fingerprint detection. Appl. Catal. B Environ. 2017, 210, 97–115. [Google Scholar] [CrossRef]
  7. Nanda, B.; Pradhan, A.C.; Parida, K.M. Fabrication of mesoporous CuO/ZrO2-MCM-41 nanocomposites for photocatalytic reduction of Cr (VI). Chem. Eng. J. 2017, 316, 1122–1135. [Google Scholar] [CrossRef] [Green Version]
  8. Guerrero-Araque, D.; Acevedo-Peña, P.; Ramírez-Ortega, D.; Calderon, H.A.; Gómez, R. Charge transfer processes involved in photocatalytic hydrogen production over CuO/ZrO2–TiO2 materials. Int. J. Hydrogen Energy 2017, 42, 9744–9753. [Google Scholar] [CrossRef]
  9. Rajendran, S.; Khan, M.M.; Gracia, F.; Qin, J.; Gupta, V.K.; Arumainathan, S. Ce3+-ion-induced visible-light photocatalytic degradation and electrochemical activity of ZnO/CeO2 nanocomposite. Sci. Rep. 2016, 6, 1–11. [Google Scholar] [CrossRef] [Green Version]
  10. Piña-Pérez, Y.; Tzompantzi-Morales, F.; Pérez-Hernández, R.; Arroyo-Murillo, R.; Acevedo-Peña, P.; Gómez-Romero, R. Photocatalytic activity of Al2O3 improved by the addition of Ce3+/Ce4+ synthesized by the sol-gel method. Photodegradation of phenolic compounds using UV light. Fuel 2017, 198, 11–21. [Google Scholar] [CrossRef]
  11. Akshatha, S.; Sreenivasa, S.; Parashuram, L.; Kumar, V.U.; Sharma, S.C.; Nagabhushana, H.; Kumar, S.; Maiyalagan, T. Synergistic effect of hybrid Ce3+/Ce4+ doped Bi2O3 nano-sphere photocatalyst for enhanced photocatalytic degradation of alizarin red S dye and its NUV excited photoluminescence studies. J. Environ. Chem. Eng. 2019, 7, 103053. [Google Scholar] [CrossRef]
  12. Dean, J.; Yang, Y.; Veser, G.; Mpourmpakis, G. CuZrO3: If it exists it should be a sandwich. Phys. Chem. Chem. Phys. 2021, 23, 23748–23757. [Google Scholar] [CrossRef] [PubMed]
  13. Donohue, M.D.; Aranovich, G.L. Classification of Gibbs adsorption isotherms. Adv. Colloid Interface Sci. 1998, 76, 137–152. [Google Scholar] [CrossRef]
  14. De, A.K.; Majumdar, S.; Pal, S.; Kumar, S.; Sinha, I. Zn doping induced band gap widening of Ag2O nanoparticles. J. Alloys Compd. 2020, 832, 154127. [Google Scholar] [CrossRef]
  15. Pelaez, M.; Nolan, N.T.; Pillai, S.C.; Seery, M.K.; Falaras, P.; Kontos, A.G.; Dunlop, P.S.M.; Hamilton, J.W.J.; Byrne, J.A.; O’Shea, K. A review on the visible light active titanium dioxide photocatalysts for environmental applications. Appl. Catal. B Environ. 2012, 125, 331–349. [Google Scholar] [CrossRef] [Green Version]
  16. Sapawe, N. Hybridization of zirconia, zinc and iron supported on HY zeolite as a solar-based catalyst for the rapid decolorization of various dyes. New J. Chem. 2015, 39, 4526–4533. [Google Scholar] [CrossRef]
  17. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 photocatalysis: Mechanisms and materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef]
  18. Wang, X.; Zhai, B.; Yang, M.; Han, W.; Shao, X. ZrO2/CeO2 nanocomposite: Two step synthesis, microstructure, and visible-light photocatalytic activity. Mater. Lett. 2013, 112, 90–93. [Google Scholar] [CrossRef]
  19. Raees, A.; Jamal, M.A.; Ahmed, I.; Silanpaa, M.; Saad Algarni, T. Synthesis and characterization of CeO2/CuO nanocomposites for photocatalytic degradation of methylene blue in visible light. Coatings 2021, 11, 305. [Google Scholar] [CrossRef]
  20. Ansari, S.; Khan, M.; Omaish, M.; Kalathil, S.; Lee, J.; Cho, M. Band gap engineering of CeO2 Nanostructure by electrochemically active biofilm for visible light applications. RSC Adv. 2014, 4, 16782–16791. [Google Scholar] [CrossRef]
  21. Khan, M.M.; Ansari, S.A.; Pradhan, D.; Han, D.H.; Lee, J.; Cho, M.H. Defect-induced band gap narrowed CeO2 nanostructures for visible light activities. Ind. Eng. Chem. Res. 2014, 53, 9754–9763. [Google Scholar] [CrossRef]
  22. Nuengmatcha, P.; Porrawatkul, P.; Chanthai, S.; Sricharoen, P.; Limchoowong, N. Enhanced photocatalytic degradation of methylene blue using Fe2O3/graphene/CuO nanocomposites under visible light. J. Environ. Chem. Eng. 2019, 7, 103438. [Google Scholar] [CrossRef]
  23. Acedo-Mendoza, A.G.; Infantes-Molina, A.; Vargas-Hernández, D.; Chávez-Sánchez, C.A.; Rodríguez-Castellón, E.; Tánori-Córdova, J.C. Photodegradation of methylene blue and methyl orange with CuO supported on ZnO photocatalysts: The effect of copper loading and reaction temperature. Mater. Sci. Semicond. Process. 2020, 119, 105257. [Google Scholar] [CrossRef]
  24. Van Huan, P.; Tam, P.D.; Pham, V.H. Visible-Light Photocatalysts of ZrO2/AgCl: Eu3+ Nanoparticles. J. Electron. Mater. 2019, 48, 5294–5300. [Google Scholar] [CrossRef]
  25. De Moraes, N.P.; de Azeredo, C.A.S.H.; Bacetto, L.A.; da Silva, M.L.C.P.; Rodrigues, L.A. The effect of C-doping on the properties and photocatalytic activity of ZrO2 prepared via sol-gel route. Optik 2018, 165, 302–309. [Google Scholar] [CrossRef]
  26. Zheng, J.; Sun, L.; Jiao, C.; Shao, Q.; Lin, J.; Pan, D.; Naik, N.; Guo, Z. Hydrothermally synthesized Ti/Zr bimetallic MOFs derived N self-doped TiO2/ZrO2 composite catalysts with enhanced photocatalytic degradation of methylene blue. Colloids Surf. A: Physicochem. Eng. Asp. 2021, 623, 126629. [Google Scholar] [CrossRef]
  27. Das, R.S.; Warkhade, S.K.; Kumar, A.; Wankhade, A.V. Graphene oxide-based zirconium oxide nanocomposite for enhanced visible light-driven photocatalytic activity. Res. Chem. Intermed. 2019, 45, 1689–1705. [Google Scholar] [CrossRef]
  28. Majumder, D.; Chakraborty, I.; Mandal, K.; Roy, S. Facet-dependent photodegradation of methylene blue using pristine CeO2 nanostructures. ACS Omega 2019, 4, 4243–4251. [Google Scholar] [CrossRef] [Green Version]
  29. Zeleke, M.A.; Kuo, D.H. Synthesis and application of V2O5-CeO2 nanocomposite catalyst for enhanced degradation of methylene blue under visible light illumination. Chemosphere 2019, 235, 935–944. [Google Scholar] [CrossRef]
  30. Quang, D.A.; Toan, T.T.T.; Tung, T.Q.; Hoa, T.T.; Mau, T.X.; Khieu, D.Q. Synthesis of CeO2/TiO2 nanotubes and heterogeneous photocatalytic degradation of methylene blue. J. Environ. Chem. Eng. 2018, 6, 5999–6011. [Google Scholar]
  31. Yu, Z.; Chuang, S.S. Probing methylene blue photocatalytic degradation by adsorbed ethanol with in situ IR. J. Phys. Chem. C 2007, 111, 13813–13820. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the ZrO2/CuO nanocomposite powder with doping of Ce3+/Ce4+ (0–8 mol%).
Figure 1. XRD patterns of the ZrO2/CuO nanocomposite powder with doping of Ce3+/Ce4+ (0–8 mol%).
Toxics 10 00463 g001
Figure 2. FE-SEM of x mol% Ce3+/Ce4+-doped ZrO2/CuO: (a) x = 2; (b) x = 4; (c) x = 6; (d) x = 8.
Figure 2. FE-SEM of x mol% Ce3+/Ce4+-doped ZrO2/CuO: (a) x = 2; (b) x = 4; (c) x = 6; (d) x = 8.
Toxics 10 00463 g002
Figure 3. TEM and the particle size distribution curve of ZrO2/CuO (a,c) and 2 mol% Ce3+/Ce4+-doped ZrO2/CuO (b,d).
Figure 3. TEM and the particle size distribution curve of ZrO2/CuO (a,c) and 2 mol% Ce3+/Ce4+-doped ZrO2/CuO (b,d).
Toxics 10 00463 g003
Figure 4. (a) FT-IR of x mol% Ce3+/Ce4+-doped ZrO2/CuO (x = 0; 2; 4; 6), (b) EDS, (c) mapping imaging of 8 mol% Ce3+/Ce4+-doped ZrO2/CuO.
Figure 4. (a) FT-IR of x mol% Ce3+/Ce4+-doped ZrO2/CuO (x = 0; 2; 4; 6), (b) EDS, (c) mapping imaging of 8 mol% Ce3+/Ce4+-doped ZrO2/CuO.
Toxics 10 00463 g004
Figure 5. Nitrogen adsorption-desorption isotherms of (a) ZrO2/CuO, and (b) 2 mol% Ce3+/Ce4+-doped ZrO2/CuO.
Figure 5. Nitrogen adsorption-desorption isotherms of (a) ZrO2/CuO, and (b) 2 mol% Ce3+/Ce4+-doped ZrO2/CuO.
Toxics 10 00463 g005
Figure 6. (a) The UV-Vis DRS and (b) The Kubelka-Munk energy curve versus the band gap (eV) of the x mol% Ce3+/Ce4+-doped ZrO2/CuO (x = 0; 2; 4; 6; 8).
Figure 6. (a) The UV-Vis DRS and (b) The Kubelka-Munk energy curve versus the band gap (eV) of the x mol% Ce3+/Ce4+-doped ZrO2/CuO (x = 0; 2; 4; 6; 8).
Toxics 10 00463 g006
Figure 7. UV-Vis spectrum of MB after being adsorbed by ZrO2/CuO-doped 8 mol% Ce3+/Ce4+.
Figure 7. UV-Vis spectrum of MB after being adsorbed by ZrO2/CuO-doped 8 mol% Ce3+/Ce4+.
Toxics 10 00463 g007
Figure 8. UV-Vis absorbance of MB under visible light at different times in the presence of x mol% Ce3+/Ce4+-doped ZrO2/CuO: (a) x = 0; (b) x = 2; (c) x = 4; (d) x = 6; (e) x = 8.
Figure 8. UV-Vis absorbance of MB under visible light at different times in the presence of x mol% Ce3+/Ce4+-doped ZrO2/CuO: (a) x = 0; (b) x = 2; (c) x = 4; (d) x = 6; (e) x = 8.
Toxics 10 00463 g008
Figure 9. (a) Photodegradation of MB, (b) the reaction rate constant k of the MB degradation in the presence of x mol% Ce3+/Ce4+doped ZrO2/CuO (x = 0; 2; 4; 6; 8), (c) reuse of 8 mol% of Ce3+/Ce4+ doping ZrO2/CuO.
Figure 9. (a) Photodegradation of MB, (b) the reaction rate constant k of the MB degradation in the presence of x mol% Ce3+/Ce4+doped ZrO2/CuO (x = 0; 2; 4; 6; 8), (c) reuse of 8 mol% of Ce3+/Ce4+ doping ZrO2/CuO.
Toxics 10 00463 g009
Figure 10. FT-IR spectrum of (a) MB, (b) MB adsorbed onto Ce3+/Ce4+-doped ZrO2/CuO before irradiation, (c) MB adsorbed on Ce3+/Ce4+-doped ZrO2/CuO 180 min after irradiation.
Figure 10. FT-IR spectrum of (a) MB, (b) MB adsorbed onto Ce3+/Ce4+-doped ZrO2/CuO before irradiation, (c) MB adsorbed on Ce3+/Ce4+-doped ZrO2/CuO 180 min after irradiation.
Toxics 10 00463 g010
Table 1. Photocatalytic degradation of MB dye using the various heterojunction photocatalyst.
Table 1. Photocatalytic degradation of MB dye using the various heterojunction photocatalyst.
PhotocatalystSynthesis MethodLight SourceRate Constant (min−1)Reference
Fe2O3/graphene/CuOSolvothermalVisible72.5 × 10−3[22]
CuO/ZnOImpregnationUV lamp182 × 10−3[23]
ZrO2/AgCl:Eu3+Sol-gelVisible14 × 10−3[24]
C-doped ZrO2Sol-gelUVC lamp7.3 × 10−3[25]
N-TiO2/ZrO2HydrothermalUV light29 × 10−3[26]
ZrO2/CuOCombustionVisible11.19 × 10−3[6]
GO-ZrO2Co-precipitationVisible57.5 × 10−3[27]
CeO2PrecipitationUV light12.1 × 10−3[28]
V2O5-CeO2PrecipitationVisible108 × 10−3[29]
CeO2/TiO2PrecipitationVisible34 × 10−3[30]
Ce3+/Ce4+-doped ZrO2/CuOHydrothermalVisible13.8 × 10−3This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chu, M.N.; Nguyen, L.T.H.; Truong, M.X.; Do, T.H.; Duong, T.T.A.; Nguyen, L.T.T.; Pham, M.A.; Tran, T.K.N.; Ngo, T.C.Q.; Pham, V.H. Ce3+/Ce4+-Doped ZrO2/CuO Nanocomposite for Enhanced Photocatalytic Degradation of Methylene Blue under Visible Light. Toxics 2022, 10, 463. https://doi.org/10.3390/toxics10080463

AMA Style

Chu MN, Nguyen LTH, Truong MX, Do TH, Duong TTA, Nguyen LTT, Pham MA, Tran TKN, Ngo TCQ, Pham VH. Ce3+/Ce4+-Doped ZrO2/CuO Nanocomposite for Enhanced Photocatalytic Degradation of Methylene Blue under Visible Light. Toxics. 2022; 10(8):463. https://doi.org/10.3390/toxics10080463

Chicago/Turabian Style

Chu, Manh Nhuong, Lan T. H. Nguyen, Mai Xuan Truong, Tra Huong Do, Thi Tu Anh Duong, Loan T. T. Nguyen, Mai An Pham, Thi Kim Ngan Tran, Thi Cam Quyen Ngo, and Van Huan Pham. 2022. "Ce3+/Ce4+-Doped ZrO2/CuO Nanocomposite for Enhanced Photocatalytic Degradation of Methylene Blue under Visible Light" Toxics 10, no. 8: 463. https://doi.org/10.3390/toxics10080463

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