Next Article in Journal
Magnetic Flux Concentration Technology Based on Soft Magnets and Superconductors
Previous Article in Journal
A New Type of Acidic OH-Groups in the LTL Zeolite
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Influence of Ni Doping on Oxygen Vacancy-Induced Changes in Structural and Chemical Properties of CeO2 Nanorods

1
School of Materials and Chemistry, University of Shanghai for Science and Technology, Shanghai 200093, China
2
Anhui Province Quartz Sand Purification and Photovoltaic Glass Engineering Research Center, Chuzhou 233100, China
3
Department of Environmental & Energy Engineering, The University of Suwon, 17, Wauan-gil, Bongdam-eup, Hwaseong-si 18323, Republic of Korea
*
Authors to whom correspondence should be addressed.
Crystals 2024, 14(8), 746; https://doi.org/10.3390/cryst14080746 (registering DOI)
Submission received: 19 July 2024 / Revised: 14 August 2024 / Accepted: 19 August 2024 / Published: 22 August 2024
(This article belongs to the Section Hybrid and Composite Crystalline Materials)

Abstract

:
In recent years, cerium dioxide (CeO2) has attracted considerable attention owing to its remarkable performance in various applications, including photocatalysis, fuel cells, and catalysis. This study explores the effect of nickel (Ni) doping on the structural, thermal, and chemical properties of CeO2 nanorods, particularly focusing on oxygen vacancy-related phenomena. Utilizing X-ray powder diffraction (XRD), alterations in crystal structure and peak shifts were observed, indicating successful Ni doping and the formation of Ni2O3 at higher doping levels, likely due to non-equilibrium reactions. Thermal gravimetric analysis (TGA) revealed changes in oxygen release mechanisms, with increasing Ni doping resulting in the release of lattice oxygen at lower temperatures. Raman spectroscopy corroborated these findings by identifying characteristic peaks associated with oxygen vacancies, facilitating the assessment of Ni doping levels. Ni-doped CeO2 can catalyze the ultrasonic degradation of methylene blue, which has good application prospects for catalytic ultrasonic degradation of organic pollutants. Overall, this study underscores the substantial impact of Ni doping on CeO2 nanorods, shedding light on tailored catalytic applications through the modulation of oxygen vacancies while preserving the nanorod morphology.

1. Introduction

In recent years, cerium dioxide (CeO2), as a typical rare earth metal oxide, has received widespread attention due to its excellent performance in photocatalysis [1,2,3], fuel cells [4,5,6,7], sensors [8,9,10,11], CO oxidation [12,13,14,15,16], water–gas shift reaction [17,18,19,20], and other fields [21,22,23]. This is mainly attributed to its two important characteristics. Firstly, in the CeO2 lattice, rapid conversion of two valence states (Ce4+/Ce3+) is achieved through the formation/loss of oxygen vacancies, thus CeO2 has excellent redox ability [24]. Secondly, CeO2 has a cubic fluorite structure containing many oxygen vacancies, which are beneficial for improving oxygen mobility [25,26].
CeO2 is an n-type semiconductor material with oxygen vacancy sites. The Kroger–Vink formula is expressed as 2CeO2 = 2Ce’Ce + Vö + 3OOx + 1/2O2↑, where Ce’Ce represents the presence of one-unit negative charge at the Ce4+ position, Vö represents an oxygen vacancy with a two-unit positive charge, and OOx represents the oxygen atom on the CeO2 lattice site. The presence of oxygen vacancies generates Ce3+, therefore CeO2 has a high lattice ion mobility and excellent oxygen storage and release ability and is used as a catalyst in various fields [27].
CeO2 nanocrystals typically expose low-index crystal planes (111), (110), and (100) [28]. Theoretical calculations indicate that the (110) plane has the lowest vacancy formation energy of 1.99 eV, the (100) plane has a vacancy formation energy of 2.27 eV, and the most stable (111) plane has a maximum vacancy formation energy of 2.60 eV. Consequently, the order of the formation energy of oxygen vacancies on different crystal planes of CeO2 is: (110) < (100) < (111) [29]. Therefore, the formation of oxygen vacancies on the CeO2 (110) crystal plane is easier. CeO2 has different exposed crystal planes based on its morphology. Polyhedral CeO2 mainly exposes (111) crystal planes, while cubic CeO2 mainly exposes (100) crystal planes and rod-shaped CeO2 mainly simultaneously exposes (110) and (100) crystal planes [28]. It is possible to regulate the morphology of CeO2 to alter oxygen vacancies. Yuan investigated the effect of CeO2 morphology on the catalytic activity of nitrobenzene hydrogen transfer reduction reaction. Oxygen vacancies and basic sites can selectively activate ethanol molecules to reduce nitro groups. The catalytic activity is sorted in the order: CeO2 nanorods, CeO2 nanopolyhedrons, and CeO2 nanocubes [30].
Doping different metals, such as precious metals, transition metals, alkali metals, and rare earth metals, into the CeO2 lattice can improve its catalytic activity and stability [31,32]. Through metal doping, lattice distortion can be induced, resulting in abundant oxygen vacancies and Ce3+, which are widely used in catalysis [33]. Transition metal catalysts have received considerable attention due to their low cost and excellent activity. Researchers have found that the substitution of Ce4+ with transition metals can significantly alter the geometric and electronic structures of CeO2 systems, leading to the reappearance of enriched electronic regions in CeO2 and the weakening of Ce-O or M-O bonds [34,35]. The results indicate that the atomic radius has a significant impact on the structure of doped CeO2, and radii larger or smaller than those of Ce4+ ions usually produce significant geometric distortions. La-doped CeO2 nanorod shows significantly higher H2 production compared to CeO2 in photocatalytic reaction [36]. We selected the fourth-period transition metal element Ni, which has the advantage of low cost, as the doping element to investigate its effect on oxygen vacancies in CeO2.
Methylene blue is a common dye, and untreated methylene blue can cause serious pollution to the water environment [37]. Organic dye wastewater has a deep chromaticity, and ultrasonic degradation of organic pollutants has been proven to be a useful method [38]. Ultrasound has been found to be an attractive and advanced method for degrading harmful organic compounds in water [39,40]. These local hotspots have a temperature of 5000 K, a pressure of 500 atm, and a lifespan of a few microseconds [41]. The water vapor inside the collapsed bubble undergoes pyrolysis to form hydroxyl and hydrogen radicals [42,43]. Ultrasonic degradation of organic compounds such as dyes is considered a highly effective method. Ni-doped CeO2 catalysts can generate a large number of oxygen vacancies, and combined with ultrasound methods, may contribute to the degradation of dyes.
In this article, we control the concentration of oxygen vacancies in CeO2 by combining morphology control and metal ion doping, thereby affecting its catalytic activity. This study prepared a series of Ni-doped CeO2 nanorods using the hydrothermal method. In addition, the structural, thermal stability, and changes in oxygen vacancy concentration between CeO2 and the doped CeO2 nanorods were thoroughly analyzed using a combination of X-ray power diffusion (XRD), transmission electron microscopy (TEM), thermogravimetry analysis (TGA), and Raman spectroscopy. Finally, the degradation reaction of methylene blue was used as a model, the effect of Ni-doped CeO2 catalytic ultrasonic degradation of methylene blue was studied, and the degradation mechanism was explored.

2. Materials and Methods

2.1. Reagents

Cerium(III) nitrate hexahydrate (Ce(NO3)3·6H2O, 99.9%, AR) was purchased from Shanghai Civi Chemical Technology Co., Ltd. (Shanghai, China). Nickel nitrate hexahydrate (Ni(NO3)3·6H2O, 99%, AR) was obtained from J&K Scientific (Beijing, China), and sodium hydroxide (NaOH, 99%, AR) was sourced from Shanghai Titan Scientific Co., Ltd. (Shanghai, China). Ultrapure water was produced using a laboratory water purification system (Hitech Instruments Co., Ltd., Hetai, China). All chemicals were used as received without further purification.

2.2. Hydrothermal Synthesis of Ni-Doped CeO2 Nanorods

Ni-doped CeO2 nanorods with various Ni2+ contents (0, 1, 5, 15 at%) were synthesized using the modified hydrothermal method, as described in Mai’s work [28]. Ce(NO3)3·6H2O and Ni(NO3)3·6H2O with different molar ratios were dissolved in 5 mL of deionized water. Meanwhile, 12.6 g of NaOH was dissolved in 30 mL of deionized water. Afterward, the two solutions were uniformly mixed and stirred at room temperature for 30 min. The resulting mixture was transferred to a 50 mL autoclave, which was then sealed and maintained at 373 K for 24 h. Subsequently, the autoclave was allowed to cool down to room temperature, and the obtained product was centrifuged and washed with deionized water and ethanol three times. The product was then dried overnight at 333 K.

2.3. Catalytic Degradation of Methylene Blue

The degradation of methylene blue solution was conducted under the condition of 293 ± 0.5 K. In a 10 mg/L methylene blue solution, 0.10 g of Ni-doped CeO2 catalyst was added, and ultrasound (25 kHz, 500 W) was used. Catalysts were taken at different times and filtered using a 0.45 μm microporous filter membrane. The absorbance of the methylene blue (MB) solution was measured at its maximum absorbance wavelength of 664 nm using a 721N Visible Spectrophotometer (INESA Analytical Instrument Co., Ltd., Shanghai, China) [37]. The removal efficiency of methylene blue was calculated according to the following formula:
Removal efficiency = (A0At)/A0 × 100%
A0 is the absorbance of the original methylene blue solution before degradation. At is the absorbance of the degraded methylene blue solution at time t.
Methylene blue (MB) standard solutions of 3 mg/L, 4 mg/L, 5 mg/L, 7 mg/L, and 9 mg/L were prepared using ultrapure water as the reference solution. The absorbance of the methylene blue solutions at a wavelength of 664 nm was measured, and the calibration curve was established. The solution exhibits a good linear relationship within the range of 3–9 mg/L, and its standard curve fits well. The calculated correlation coefficient is R2 = 0.99864, and the relationship between absorbance, A and MB concentration, C is A = 0.1762C + 0.02492.

2.4. Characterization

The X-ray powder Diffraction (XRD) patterns were recorded using a Rigaku diffractometer with Cu Kα radiation (λ = 1.5418 nm). The X-rays were operated at 40 kV and 40 mA. Patterns were collected in the 2θ range from 20° to 80°, with a scanning step of 0.02 and a scanning speed of 2.5°/min. The morphology and microstructure of the samples were characterized using a HT7800 transmission electron microscope (TEM) operating at 200 kV. Thermal gravimetric analysis (TGA-50, Shimadzu Corp., Kyoto, Japan) was employed to measure oxygen release with temperature increase. The temperature was raised from room temperature to 800 °C at a rate of 10 °C/min under a nitrogen atmosphere (gas flow rate: 24 mL/min). Raman spectroscopy was utilized to measure the presence of surface Ce3+ or oxygen vacancies. This analysis was performed at the Shanghai branch of the Beijing High Voltage Science Research Center (Gaoke, Shanghai, China) using the Renishaw inVia micro- Raman spectrometer from the UK. The experiment employed a 532 nm laser light source with a power of 50 mW, 1800 gr/mm grating, and a 50 times Leica long focal length lens. The required Raman spectrum for this experiment covers a scanning range of 50–1200 cm−1, with a scanning time of 10 s, 10 scanning times, and a resolution of 3 cm−1.

3. Results and Discussion

The phase structure of the products was characterized by X-ray powder diffraction (XRD), as shown in Figure 1, and the related data are summarized in Table 1. The relationship between grain size and half peak width can be obtained through the Scherrer equation, D = ()/(βcos θ), where D is the grain size, K is a constant, and β is the half peak width. The half peak width is inversely proportional to the crystallite size.
The characteristic diffraction peaks of pure CeO2 samples appear at 28.5°, 33.1°, 47.5°, 56.3°, 59.1°, 69.4°, 76.7°, and 79.1°, corresponding to the cubic fluorite crystal planes of CeO2 for 111, 200, 220, 311, 222, 400, 331, and 420, respectively. As the Ni doping amount increases, the width of the diffraction peaks also increases, indicating a decrease in crystallite size with increasing Ni doping amount (Table 1).
Additionally, upon comparison with the (111) peak of pure CeO2, the peaks of the doped samples are observed to shift to higher angles. This phenomenon occurs due to the replacement of cerium sites in the cerium oxide lattice by Ni2+, with a radius of 0.63 Å, resulting in the observed shift of the cerium oxide peaks [44,45]. This shift indicates the successful doping of metallic nickel into the cerium oxide lattice. According to the Bragg equation, 2dsinθ = , after doping with nickel metal, the increase in θ value leads to a slight decrease in the interplanar spacing of the crystal cell, causing slight lattice shrinking [46].
When the doping amount reaches 15%, a new diffraction peak appears near 38°. The peak in question may correspond to Ni2O3, indicating that some of the excess Ni species did not successfully dope into the CeO2 structure and instead underwent a secondary reaction. Therefore, it can be confirmed that there was hardly any peak shift. Furthermore, it has been observed that as the nickel doping level increases, there is a tendency for the crystallite size to decrease. However, at 15% doping, it has been confirmed that this trend deviates significantly.
Therefore, at doping levels of 1% and 5%, only diffraction peaks corresponding to CeO2 are observed, confirming a trend of high-angle shifts and decreasing crystallite size. This suggests that nickel can be fully doped into CeO2 at these levels.
Figure 2a–d presents TEM images of CeO2 and Ni-doped CeO2. From Figure 2a, it is evident that rod-shaped CeO2 nanomaterials have been successfully synthesized. As depicted in Figure 2b, even with a 1% doping amount, the rod-shaped morphology of CeO2 nanomaterials remains unchanged. With a doping amount of 5% (Figure 2c), the rod-shaped structure becomes wider. In Figure 2d, with a doping amount of 15%, although the structure remains rod-shaped, the length of the rods significantly decreases while the width increases.
Figure 3 shows the thermogravimetric analysis of undoped CeO2 and Ni-doped CeO2 nanorods. All the samples were measured without calcination. Up to about 150 °C, the weight loss due to the loss of surface moisture can be ignored. The weight loss rate of undoped CeO2 is about 3.08%, the weight loss rate of 1% Ni-CeO2 is about 7.95%, the weight loss rate of 5% Ni-CeO2 is about 5.88%, and the weight loss rate of 15% Ni-CeO2 is about 3.33%; therefore, the weight loss above 150 °C can be attributed to the loss of surface or lattice oxygen. It can be observed that the pattern of weight loss changes around 320 °C, indicating different energy release mechanisms between surface and lattice oxygen [47,48]. By analyzing the weight loss in each temperature range (150–320 °C), the weight loss due to surface oxygen release was approximately 2.35 wt% for pure CeO2, 3.79 wt% for 1% Ni-CeO2, 5.04 wt% for 5% Ni-CeO2, and 4.34 wt% for 15% Ni-CeO2. On the other hand (320–800 °C), the weight loss due to lattice oxygen release was approximately 1.23 wt% for pure CeO2, 2.62 wt% for 1% Ni-CeO2, 2.63 wt% for 5% Ni-CeO2, and 2.08 wt% for 15% Ni-CeO2. The weight loss rate after Ni doping was higher than that without doping, and generally decreased with increasing doping level (Table 2). This may be attributed to the non-equilibrium reactions leading to the formation of NiOx when Ni is added at high concentrations [49]. Furthermore, as the doping effect increases, the activation of surface oxygen also increases, leading to a greater impact on weight loss at lower temperatures. In other words, in the temperature range below 800 °C, the influence of surface oxygen release becomes more significant with increased doping. Calculating the ratio of surface oxygen to lattice oxygen, it increased to 1.45 for 1% Ni-CeO2, 1.92 for 5% Ni-CeO2, and 2.23 for 15% Ni-CeO2.
As shown in Figure 4, the Raman spectra provide the following insights: both undoped and Ni-doped CeO2 samples exhibit characteristic peaks around 462 cm–1, originating from the F2g vibration of CeO2 due to the symmetric arrangement of oxygen atoms in the CeO2 lattice [50]. Additionally, a weak absorption peak is observed near 595 cm–1, corresponding to Frenkel-type oxygen vacancies [51]. Notably, the characteristic peak of undoped CeO2 at 462 cm–1 appears strongest among the four images. As the doping level of Ni increases, the intensity of absorption peaks at this location gradually diminishes due to the substitutional reaction C e 4 + + N i 2 + N i 3 + + C e 3 + , leading to the generation of oxygen vacancies and thus symmetry degradation [52,53]. Furthermore, an increase in Ni doping may lead to a decrease in particle size, which in turn affects the linewidth of the main Raman peak at 462 cm−1 [54].
To gauge the relative concentration of surface oxygen vacancies in the samples, the intensity of the absorption peak at 595 cm−1 (A595) was compared to that at 462 cm−1 (A462) [55]. A higher ratio indicates a greater presence of oxygen vacancies on the sample surface [56,57]. Specific ratios are provided in Table 3. As evident from these findings, increasing Ni doping correlates with a rise in the concentration of oxygen vacancies on the surface of cerium dioxide, peaking at a 5% doping level. However, at a 15% doping level, the peak intensity slightly decreases, attributed to incomplete Ni doping and the formation of Ni2O3, consistent with previous discussions.
To evaluate the chemical properties of the synthesized nanoparticles, an acoustic cavitation test was conducted using methylene blue, a surrogate for organic pollutants, as shown in Figure 5 (Refer to the Section 2 for the calibration curve in Figure 5a). Acoustic cavitation occurs when ultrasound of sufficient intensity passes through a liquid, causing small bubbles (cavitation nuclei) present in the liquid to rapidly increase under negative pressure. During the subsequent positive pressure cycle, the bubbles undergo adiabatic compression and collapse, creating extremely short and intense pressure pulses at the moment of collapse. High temperatures and high pressures are generated inside the bubbles, making this method very effective for the degradation of organic materials.
As shown in Figure 5b, the results after 20 min of reaction indicate that 1% and 5% Ni-CeO2 show higher degradation rates than undoped CeO2. Initially, both samples exhibit similar degradation rates, but as the reaction time increases, the 5% Ni-CeO2 shows a higher degradation rate, confirming that the higher the surface oxygen concentration, the higher the degradation rate (Table 2 and Table 3). In contrast, 15% Ni-CeO2 initially shows a degradation rate similar to that of pure CeO2, as the nickel doping is minimal at the start. However, as the reaction progresses, the degradation rate gradually increases, and after 40 min, it increases even more, reaching a degradation rate of 45.9% at 80 min. This is thought to be due to the additional catalytic action of nickel oxide in forming ·OH radicals, rather than the effect of doping itself [58].

4. Conclusions

Summarizing the research findings, it can be concluded that doping Ni into CeO2 nanorods significantly impacts their structural, thermal, and chemical properties. XRD analysis revealed changes in crystal structure and peak shifts, indicating successful Ni doping and the formation of Ni2O3 at high doping levels due to non-equilibrium reactions. TGA results demonstrated variations in oxygen release mechanisms, with increasing Ni doping leading to the emission of some lattice oxygen at lower temperatures, as confirmed by changes in weight ratio across temperature ranges. Additionally, the presence of characteristic Raman peaks associated with oxygen vacancies allowed for assessing the extent of Ni doping based on peak intensity changes. Catalysts with rich oxygen vacancies under ultrasound help generate more ·OH radicals, promoting the degradation of methylene blue. Therefore, through this study, successful adjustment of properties for catalytic applications was achieved by controlling the Ni doping level while maintaining the CeO2 nanorod type.

Author Contributions

Conceptualization, Y.Z. (Yuanzheng Zhu); methodology, Y.Z. (Yuanzheng Zhu); investigation, W.W., H.L., P.C. and C.L.; data curation, H.L., G.C. and H.W.; formal analysis, Y.Z. (Yuedie Zhang); writing—original draft preparation, Y.Z. (Yuanzheng Zhu); writing—review and editing, G.S.; supervision, C.C. and G.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the ‘College Student Innovation Training Program’ of the University of Shanghai for Science and Technology (XJ2023484). It was also funded by research grants from The University of Suwon in 2023 (U2023-0169) and the Open Foundation of Anhui Province Quartz Sand Purification and Photovoltaic Glass Engineering Research Center (2023KF06).

Data Availability Statement

The data presented in this study are available within this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yang, J.; Xie, N.; Zhang, J.; Fan, W.; Huang, Y.; Tong, Y. Defect Engineering Enhances the Charge Separation of CeO2 Nanorods toward Photocatalytic Methyl Blue Oxidation. Nanomaterials 2020, 10, 2307. [Google Scholar] [CrossRef]
  2. Murray, S.; Wei, W.; Hart, R.; Fan, J.; Chen, W.; Lu, P. Solar Degradation of Toxic Colorants in Polluted Water by Thermally Tuned Ceria Nanocrystal-Based Nanofibers. ACS Appl. Nano Mater. 2020, 3, 11194–11202. [Google Scholar] [CrossRef]
  3. Yu, D.; Jia, Y.; Yang, Z.; Zhang, H.; Zhao, J.; Zhao, Y.; Weng, B.; Dai, W.; Li, Z.; Wang, P.; et al. Solar Photocatalytic Oxidation of Methane to Methanol with Water over RuOx/ZnO/CeO2 Nanorods. ACS Sustain. Chem. Eng. 2022, 10, 16–22. [Google Scholar] [CrossRef]
  4. Chen, T.; Zheng, G.; Liu, K.; Zhang, G.; Huang, Z.; Liu, M.; Zhou, J.; Wang, S. Application of CuNi–CeO2 Fuel Electrode in Oxygen Electrode Supported Reversible Solid Oxide Cell. Int. J. Hydrogen Energy 2023, 48, 9565–9573. [Google Scholar] [CrossRef]
  5. Zhao, M.; Geng, S.; Chen, G.; Wang, F. Efficient FeCoNi/CeO2 Coatings for Solid Oxide Fuel Cell Steel Interconnect Applications. Int. J. Hydrogen Energy 2024, 50, 1087–1094. [Google Scholar] [CrossRef]
  6. Li, J.; Xie, J.; Li, D.; Yu, L.; Xu, C.; Yan, S.; Lu, Y. An Interface Heterostructure of NiO and CeO2 for Using Electrolytes of Low-Temperature Solid Oxide Fuel Cells. Nanomaterials 2021, 11, 2004. [Google Scholar] [CrossRef]
  7. Zhang, Y.; Zhu, D.; Jia, X.; Liu, J.; Li, X.; Ouyang, Y.; Li, Z.; Gao, X.; Zhu, C. Novel n–i CeO2/a-Al2O3 Heterostructure Electrolyte Derived from the Insulator a-Al2O3 for Fuel Cells. ACS Appl. Mater. Interfaces 2023, 15, 2419–2428. [Google Scholar] [CrossRef]
  8. Ema, T.; Choi, P.G.; Takami, S.; Masuda, Y. Facet-Controlled Synthesis of CeO2 Nanoparticles for High-Performance CeO2 Nanoparticle/SnO2 Nanosheet Hybrid Gas Sensors. ACS Appl. Mater. Interfaces 2022, 14, 56998–57007. [Google Scholar] [CrossRef]
  9. Hu, Q.; Huang, B.; Li, Y.; Zhang, S.; Zhang, Y.; Hua, X.; Liu, G.; Li, B.; Zhou, J.; Xie, E.; et al. Methanol Gas Detection of Electrospun CeO2 Nanofibers by Regulating Ce3+/Ce4+ Mole Ratio via Pd Doping. Sens. Actuators B Chem. 2020, 307, 127638. [Google Scholar] [CrossRef]
  10. Hu, Q.; Jiang, H.; Zhang, W.; Wang, X.; Wang, X.; Zhang, Z. Unveiling the Synergistic Effects of Hydrogen Annealing on CeO2 Nanofibers for Highly Sensitive Acetone Gas Detection: Role of Ce3+ Ions and Oxygen Vacancies. Appl. Surf. Sci. 2023, 640, 158411. [Google Scholar] [CrossRef]
  11. Yuan, K.; Wang, C.-Y.; Zhu, L.-Y.; Cao, Q.; Yang, J.-H.; Li, X.-X.; Huang, W.; Wang, Y.-Y.; Lu, H.-L.; Zhang, D.W. Fabrication of a Micro-Electromechanical System-Based Acetone Gas Sensor Using CeO2 Nanodot-Decorated WO3 Nanowires. ACS Appl. Mater. Interfaces 2020, 12, 14095–14104. [Google Scholar] [CrossRef]
  12. Basina, G.; Polychronopoulou, K.; Zedan, A.F.; Dimos, K.; Katsiotis, M.S.; Fotopoulos, A.P.; Ismail, I.; Tzitzios, V. Ultrasmall Metal-Doped CeO2 Nanoparticles for Low-Temperature CO Oxidation. ACS Appl. Nano Mater. 2020, 3, 10805–10813. [Google Scholar] [CrossRef]
  13. Ahasan, M.R.; Wang, R. CeO2 Nanorods Supported CuOx-RuOx Bimetallic Catalysts for Low Temperature CO Oxidation. J. Colloid Interface Sci. 2024, 654, 1378–1392. [Google Scholar] [CrossRef]
  14. Fan, J.; Hu, S.; Li, C.; Wang, Y.; Chen, G. Effect of Loading Method on Catalytic Performance of Pt/CeO2 System for CO Oxidation. Mol. Catal. 2024, 558, 114013. [Google Scholar] [CrossRef]
  15. Kim, H.Y.; Lee, H.M.; Henkelman, G. CO Oxidation Mechanism on CeO2-Supported Au Nanoparticles. J. Am. Chem. Soc. 2012, 134, 1560–1570. [Google Scholar] [CrossRef]
  16. Zhou, Z.; Zhang, J.; Liu, Y. Promoted Low-Temperature CO Oxidation Activity of CeO2 by Cu Doping: The Important Role of Oxygen Vacancies. Fuel 2024, 359, 130434. [Google Scholar] [CrossRef]
  17. Reina, T.R.; Gonzalez-Castaño, M.; Lopez-Flores, V.; Martínez, T.L.M.; Zitolo, A.; Ivanova, S.; Xu, W.; Centeno, M.A.; Rodriguez, J.A.; Odriozola, J.A. Au and Pt Remain Unoxidized on a CeO2-Based Catalyst during the Water–Gas Shift Reaction. J. Am. Chem. Soc. 2022, 144, 446–453. [Google Scholar] [CrossRef]
  18. Yu, J.; Qin, X.; Yang, Y.; Lv, M.; Yin, P.; Wang, L.; Ren, Z.; Song, B.; Li, Q.; Zheng, L.; et al. Highly Stable Pt/CeO2 Catalyst with Embedding Structure toward Water–Gas Shift Reaction. J. Am. Chem. Soc. 2024, 146, 1071–1080. [Google Scholar] [CrossRef]
  19. Shen, X.; Li, Z.; Xu, J.; Li, W.; Tao, Y.; Ran, J.; Yang, Z.; Sun, K.; Yao, S.; Wu, Z.; et al. Upgrading the Low Temperature Water Gas Shift Reaction by Integrating Plasma with a CuOx/CeO2 Catalyst. J. Catal. 2023, 421, 324–331. [Google Scholar] [CrossRef]
  20. Zhang, K.; Guo, Q.; Wang, Y.; Cao, P.; Zhang, J.; Heggen, M.; Mayer, J.; Dunin-Borkowski, R.E.; Wang, F. Ethylene Carbonylation to 3-Pentanone with In Situ Hydrogen via a Water–Gas Shift Reaction on Rh/CeO2. ACS Catal. 2023, 13, 3164–3169. [Google Scholar] [CrossRef]
  21. Yu, Y.; He, J.; Wang, T.; Qiu, X.; Chen, K.; Wu, Q.; Shi, D.; Zhang, Y.; Li, H. One-Step Production of Pt–CeO2/N-CNTs Electrocatalysts with High Catalytic Performance toward Methanol Oxidation. Int. J. Hydrogen Energy 2023, 48, 29565–29582. [Google Scholar] [CrossRef]
  22. Jiang, F.; Wang, S.; Liu, B.; Liu, J.; Wang, L.; Xiao, Y.; Xu, Y.; Liu, X. Insights into the Influence of CeO2 Crystal Facet on CO2 Hydrogenation to Methanol over Pd/CeO2 Catalysts. ACS Catal. 2020, 10, 11493–11509. [Google Scholar] [CrossRef]
  23. Guo, Y.; Qin, Y.; Liu, H.; Wang, H.; Han, J.; Zhu, X.; Ge, Q. CeO2 Facet-Dependent Surface Reactive Intermediates and Activity during Ketonization of Propionic Acid. ACS Catal. 2022, 12, 2998–3012. [Google Scholar] [CrossRef]
  24. Reed, K.; Cormack, A.; Kulkarni, A.; Mayton, M.; Sayle, D.; Klaessig, F.; Stadler, B. Exploring the Properties and Applications of Nanoceria: Is There Still Plenty of Room at the Bottom? Environ. Sci. Nano 2014, 1, 390–405. [Google Scholar] [CrossRef]
  25. Chauhan, S.; Richards, G.J.; Mori, T.; Yan, P.; Hill, J.P.; Ariga, K.; Zou, J.; Drennan, J. Fabrication of a Nano-Structured Pt-Loaded Cerium Oxide Nanowire and Its Anode Performance in the Methanol Electro-Oxidation Reaction. J. Mater. Chem. A 2013, 1, 6262. [Google Scholar] [CrossRef]
  26. Li, G.; Lu, F.; Wei, X.; Song, X.; Sun, Z.; Yang, Z.; Yang, S. Nanoporous Ag–CeO2 Ribbons Prepared by Chemical Dealloying and Their Electrocatalytic Properties. J. Mater. Chem. A 2013, 1, 4974. [Google Scholar] [CrossRef]
  27. Wu, K.; Sun, L.; Yan, C. Recent Progress in Well-Controlled Synthesis of Ceria-Based Nanocatalysts towards Enhanced Catalytic Performance. Adv. Energy Mater. 2016, 6, 1600501. [Google Scholar] [CrossRef]
  28. Mai, H.-X.; Sun, L.-D.; Zhang, Y.-W.; Si, R.; Feng, W.; Zhang, H.-P.; Liu, H.-C.; Yan, C.-H. Shape-Selective Synthesis and Oxygen Storage Behavior of Ceria Nanopolyhedra, Nanorods, and Nanocubes. J. Phys. Chem. B 2005, 109, 24380–24385. [Google Scholar] [CrossRef] [PubMed]
  29. Nolan, M.; Fearon, J.; Watson, G. Oxygen Vacancy Formation and Migration in Ceria. Solid State Ion. 2006, 177, 3069–3074. [Google Scholar] [CrossRef]
  30. Yuan, Z.; Huang, L.; Liu, Y.; Sun, Y.; Wang, G.; Li, X.; Lercher, J.A.; Zhang, Z. Synergy of Oxygen Vacancies and Base Sites for Transfer Hydrogenation of Nitroarenes on Ceria Nanorods. Angew. Chem. 2024, 136, e202317339. [Google Scholar] [CrossRef]
  31. Abdulwahab, K.O.; Khan, M.M.; Jennings, J.R. Doped Ceria Nanomaterials: Preparation, Properties, and Uses. ACS Omega 2023, 8, 30802–30823. [Google Scholar] [CrossRef] [PubMed]
  32. Hajjiah, A.; Samir, E.; Shehata, N.; Salah, M. Lanthanide-Doped Ceria Nanoparticles as Backside Coaters to Improve Silicon Solar Cell Efficiency. Nanomaterials 2018, 8, 357. [Google Scholar] [CrossRef]
  33. Thormann, J.; Maier, L.; Pfeifer, P.; Kunz, U.; Deutschmann, O.; Schubert, K. Steam Reforming of Hexadecane over a Rh/CeO2 Catalyst in Microchannels: Experimental and Numerical Investigation. Int. J. Hydrogen Energy 2009, 34, 5108–5120. [Google Scholar] [CrossRef]
  34. Yildirim, H.; Pachter, R. Extrinsic Dopant Effects on Oxygen Vacancy Formation Energies in ZrO2 with Implication for Memristive Device Performance. ACS Appl. Electron. Mater. 2019, 1, 467–477. [Google Scholar] [CrossRef]
  35. Sen, S.; Edwards, T.; Kim, S.K.; Kim, S. Investigation of the Potential Energy Landscape for Vacancy Dynamics in Sc-Doped CeO2. Chem. Mater. 2014, 26, 1918–1924. [Google Scholar] [CrossRef]
  36. Nabi, G.; Ali, W.; Majid, A.; Alharbi, T.; Saeed, S.; Albedah, M.A. Controlled Growth of Bi-Functional La Doped CeO2 Nanorods for Photocatalytic H2 Production and Supercapacitor Applications. Int. J. Hydrogen Energy 2022, 47, 15480–15490. [Google Scholar] [CrossRef]
  37. Yusuf, L.A.; Ertekin, Z.; Fletcher, S.; Symes, M.D. Enhanced ultrasonic degradation of methylene blue using a catalyst-free dual-frequency treatment. Ultrason. Sonochem. 2024, 103, 106792. [Google Scholar] [CrossRef] [PubMed]
  38. Adewuyi, Y.G. Sonochemistry: Environmental science and engineering applications. Ind. Eng. Chem. Res. 2021, 40, 4681–4715. [Google Scholar] [CrossRef]
  39. Chowdhury, P.; Viraraghavan, T. Sonochemical degradation of chlorinated organic compounds, phenolic compounds and organic dyes—A review. Sci. Total Environ. 2009, 407, 2474–2492. [Google Scholar] [CrossRef]
  40. Bi, Y.G.; Zheng, Y.H.; Tang, L.; Guo, J.; Zhou, S.Q. Optimization of process parameters for enhanced degradation of methylene blue by trough ultrasonic. J. Phys. Conf. Ser. 2022, 2152, 012027. [Google Scholar] [CrossRef]
  41. Suslick, K.S. Sonochemistry. Science 1990, 247, 1439–1445. [Google Scholar] [CrossRef]
  42. Makino, K.; Mossoba, M.M.; Riesz, P. Chemical effects of ultrasound on aqueous solutions. Evidence for ·OH and ·H by spin trapping. J. Am. Chem. Soc. 1982, 104, 3537–3539. [Google Scholar] [CrossRef]
  43. Makino, K.; Mossoba, M.M.; Riesz, P. Chemical effects of ultrasound on aqueous solutions. Formation of hydroxyl radicals and hydrogen atoms. J. Phys. Chem. 1983, 87, 1369–1377. [Google Scholar] [CrossRef]
  44. Xiao, M.; Han, D.; Yang, X.; Tsona Tchinda, N.; Du, L.; Guo, Y.; Wei, Y.; Yu, X.; Ge, M. Ni-Doping-Induced Oxygen Vacancy in Pt-CeO2 Catalyst for Toluene Oxidation: Enhanced Catalytic Activity, Water-Resistance, and SO2-Tolerance. Appl. Catal. B Environ. 2023, 323, 122173. [Google Scholar] [CrossRef]
  45. Yang, Z.; Zheng, D.; Yue, X.; Wang, K.; Hou, Y.; Dai, W.; Fu, X. The Synergy of Ni Doping and Oxygen Vacancies over CeO2 in Visible Light-Assisted Thermal Catalytic Methanation Reaction. Appl. Surf. Sci. 2023, 615, 156311. [Google Scholar] [CrossRef]
  46. Liu, Z.; Ma, H.; Sorrell, C.C.; Koshy, P.; Wang, B.; Hart, J.N. Enhancement of Light Absorption and Oxygen Vacancy Formation in CeO2 by Transition Metal Doping: A DFT Study. Appl. Catal. Gen. 2024, 670, 119544. [Google Scholar] [CrossRef]
  47. Lee, J.; Ryou, Y.; Chan, X.; Kim, T.J.; Kim, D.H. How Pt Interacts with CeO2 under the Reducing and Oxidizing Environments at Elevated Temperature: The Origin of Improved Thermal Stability of Pt/CeO2 Compared to CeO2. J. Phys. Chem. C 2016, 120, 25870–25879. [Google Scholar] [CrossRef]
  48. D’Angelo, A.M.; Wu, Z.; Overbury, S.H.; Chaffee, A.L. Cu-Enhanced Surface Defects and Lattice Mobility of Pr-CeO2 Mixed Oxides. J. Phys. Chem. C 2016, 120, 27996–28008. [Google Scholar] [CrossRef]
  49. Zhu, Y.; Seong, G.; Noguchi, T.; Yoko, A.; Tomai, T.; Takami, S.; Adschiri, T. Highly Cr-Substituted CeO2 Nanoparticles Synthesized Using a Non-Equilibrium Supercritical Hydrothermal Process: High Oxygen Storage Capacity Materials Designed for a Low-Temperature Bitumen Upgrading Process. ACS Appl. Energy Mater. 2020, 3, 4305–4319. [Google Scholar] [CrossRef]
  50. Shyu, J.Z.; Weber, W.H.; Gandhi, H.S. Surface Characterization of Alumina-Supported Ceria. J. Phys. Chem. 1988, 92, 4964–4970. [Google Scholar] [CrossRef]
  51. Popović, Z.V.; Dohčević-Mitrović, Z.; Konstantinović, M.J.; Šćepanović, M. Raman Scattering Characterization of Nanopowders and Nanowires (Rods). J. Raman Spectrosc. 2007, 38, 750–755. [Google Scholar] [CrossRef]
  52. Du, X.; Dai, Q.; Wei, Q.; Huang, Y. Nanosheets-Assembled Ni (Co) Doped CeO2 Microspheres toward NO + CO Reaction. Appl. Catal. Gen. 2020, 602, 117728. [Google Scholar] [CrossRef]
  53. Tiwari, S.; Khatun, N.; Patra, N.; Yadav, A.K.; Bhattacharya, D.; Jha, S.N.; Tseng, C.M.; Liu, S.W.; Biring, S.; Sen, S. Role of Oxygen Vacancies in Co/Ni Substituted CeO2: A Comparative Study. Ceram. Int. 2019, 45, 3823–3832. [Google Scholar] [CrossRef]
  54. Weber, W.H.; Hass, K.C.; McBride, J.R. Raman study of CeO2: Second-order scattering, lattice dynamics, and particle-size effects. Phys. Rev. B Condens. Matter Mater. Phys. 1993, 48, 178–185. [Google Scholar] [CrossRef] [PubMed]
  55. Li, L.; Chen, F.; Lu, J.-Q.; Luo, M.-F. Study of Defect Sites in Ce1−xMxO2−δ (x = 0.2) Solid Solutions Using Raman Spectroscopy. J. Phys. Chem. A 2011, 115, 7972–7977. [Google Scholar] [CrossRef] [PubMed]
  56. Zhao, P.; Feng, N.; Tan, B.; Huo, Z.; Liu, G.; Wan, H.; Guan, G. Exposed Crystal Facet Tuning of CeO2 for Boosting Catalytic Soot Combustion: The Effect of La Dopant. J. Environ. Chem. Eng. 2022, 10, 108503. [Google Scholar] [CrossRef]
  57. Cao, Y.; Zhao, L.; Gutmann, T.; Xu, Y.; Dong, L.; Buntkowsky, G.; Gao, F. Getting Insights into the Influence of Crystal Plane Effect of Shaped Ceria on Its Catalytic Performances. J. Phys. Chem. C 2018, 122, 20402–20409. [Google Scholar] [CrossRef]
  58. Parameswaran, S.; Shanmugam, P.; Bakkiyaraj, R.; Venugopal, T. Fabrication and assessment of CuO and NiO-infused MnO2 nanocomposites: Characterization, methylene blue degradation, and antibacterial efficacy. J. Mol. Struct. 2024, 1303, 137560. [Google Scholar] [CrossRef]
Figure 1. XRD spectra of CeO2 and Ni-doped CeO2. The right figure is an enlarged view of a specific range of the left figure.
Figure 1. XRD spectra of CeO2 and Ni-doped CeO2. The right figure is an enlarged view of a specific range of the left figure.
Crystals 14 00746 g001
Figure 2. TEM images of CeO2 and Ni-doped CeO2: (a) CeO2; (b) 1% Ni-CeO2; (c) 5% Ni-CeO2; (d) 15% Ni-CeO2.
Figure 2. TEM images of CeO2 and Ni-doped CeO2: (a) CeO2; (b) 1% Ni-CeO2; (c) 5% Ni-CeO2; (d) 15% Ni-CeO2.
Crystals 14 00746 g002
Figure 3. The TG curve of CeO2 and Ni-doped CeO2.
Figure 3. The TG curve of CeO2 and Ni-doped CeO2.
Crystals 14 00746 g003
Figure 4. Raman spectra of CeO2 and Ni-doped CeO2. The dashed lines indicate 462 and 595 cm–1.
Figure 4. Raman spectra of CeO2 and Ni-doped CeO2. The dashed lines indicate 462 and 595 cm–1.
Crystals 14 00746 g004
Figure 5. Effect of Ni doping amount on CeO2 catalyzing capability: (a) Calibration curve; (b) Degradation depending on reaction time.
Figure 5. Effect of Ni doping amount on CeO2 catalyzing capability: (a) Calibration curve; (b) Degradation depending on reaction time.
Crystals 14 00746 g005
Table 1. Lattice constant and crystallite size of the samples.
Table 1. Lattice constant and crystallite size of the samples.
CeO21% Ni-CeO25% Ni-CeO215% Ni-CeO2
Lattice constant (Å)5.41105.40645.39825.4230
Crystallite size (nm)128656092
Table 2. The ratio of surface oxygen/lattice oxygen of samples.
Table 2. The ratio of surface oxygen/lattice oxygen of samples.
CeO21% Ni-CeO25% Ni-CeO215% Ni-CeO2
Surface oxygen2.35 wt%3.79 wt%5.04 wt%4.34 wt%
Lattice oxygen1.23 wt%2.62 wt%2.63 wt%2.08 wt%
Total oxygen3.58 wt%6.41 wt%7.67 wt%6.42 wt%
Table 3. The ratio of A595/A462 in Raman spectra.
Table 3. The ratio of A595/A462 in Raman spectra.
CeO21% Ni-CeO25% Ni-CeO215% Ni-CeO2
A595/A4629.79%23.9%30.5%26.7%
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

Zhu, Y.; Wang, W.; Chen, G.; Li, H.; Zhang, Y.; Liu, C.; Wang, H.; Cheng, P.; Chen, C.; Seong, G. Influence of Ni Doping on Oxygen Vacancy-Induced Changes in Structural and Chemical Properties of CeO2 Nanorods. Crystals 2024, 14, 746. https://doi.org/10.3390/cryst14080746

AMA Style

Zhu Y, Wang W, Chen G, Li H, Zhang Y, Liu C, Wang H, Cheng P, Chen C, Seong G. Influence of Ni Doping on Oxygen Vacancy-Induced Changes in Structural and Chemical Properties of CeO2 Nanorods. Crystals. 2024; 14(8):746. https://doi.org/10.3390/cryst14080746

Chicago/Turabian Style

Zhu, Yuanzheng, Weixia Wang, Gejunxiang Chen, Huyi Li, Yuedie Zhang, Chang Liu, Hao Wang, Ping Cheng, Chunguang Chen, and Gimyeong Seong. 2024. "Influence of Ni Doping on Oxygen Vacancy-Induced Changes in Structural and Chemical Properties of CeO2 Nanorods" Crystals 14, no. 8: 746. https://doi.org/10.3390/cryst14080746

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