Next Article in Journal
Mechanical and Structural Characterization of Laser-Cladded Medium-Entropy FeNiCr-B4C Coatings
Previous Article in Journal
The Effects of Zirconium and Yttrium Addition on the Microstructure and Hardness of AlCuMgMn Alloy when Applying In Situ Heating during the Laser Melting Process
Previous Article in Special Issue
Design of Pt-Sn-Zn Nanomaterials for Successful Methanol Electrooxidation Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Novel Mesoporous and Multilayered Yb/N-Co-Doped CeO2 with Enhanced Oxygen Storage Capacity

1
Laboratory for Functional Materials, School of New Energy Materials and Chemistry, Leshan Normal University, Leshan 614000, China
2
Leshan West Silicon Materials Photovoltaic and New Energy Industry Technology Research Institute, Leshan 614000, China
3
College of Materials Science and Engineering, Sichuan University, Chengdu 610065, China
4
Department of Chemical Engineering, Illinois Institute of Technology, Chicago, IL 60616, USA
5
College of Materials Science and Engineering, National Engineering Research Center for Magnesium Alloys, Chongqing University, Chongqing 400044, China
*
Author to whom correspondence should be addressed.
Materials 2023, 16(15), 5478; https://doi.org/10.3390/ma16155478
Submission received: 25 June 2023 / Revised: 29 July 2023 / Accepted: 2 August 2023 / Published: 4 August 2023

Abstract

:
A cubic fluorite-type CeO2 with mesoporous multilayered morphology was synthesized by the solvothermal method followed by calcination in air, and its oxygen storage capacity (OSC) was quantified by the amount of O2 consumption per gram of CeO2 based on hydrogen temperature programmed reduction (H2–TPR) measurements. Doping CeO2 with ytterbium (Yb) and nitrogen (N) ions proved to be an effective route to improving its OSC in this work. The OSC of undoped CeO2 was 0.115 mmol O2/g and reached as high as 0.222 mmol O2/g upon the addition of 5 mol.% Yb(NO3)3∙5H2O, further enhanced to 0.274 mmol O2/g with the introduction of 20 mol.% triethanolamine. Both the introductions of Yb cations and N anions into the CeO2 lattice were conducive to the formation of more non-stoichiometric oxygen vacancy (VO) defects and reducible–reoxidizable Cen+ ions. To determine the structure performance relationships, the partial least squares method was employed to construct two linear functions for the doping level vs. lattice parameter and [VO] vs. OSC/SBET.

1. Introduction

Cerium oxide (ceria, CeO2) and CeO2-based binary or multiple composites are typical oxygen storage materials, which have a wide range of applications in VOC abatement, partial alkyne hydrogenation, oxidative dehydrogenation, and water gas shift [1,2,3,4,5,6,7], even in biomedical applications [8]. CeO2 is appealing because of its unique structure and intrinsic oxygen vacancy (VO) defect, which can be rapidly formed and eliminated on the surface of CeO2, giving the redox cycle of Ce4+ ⇔ Ce3+. This is the source of the oxygen storage capacity (OSC), which could be described by Reaction (1) and could also be written using Kroger and Vink notations as Reaction (2) [9]:
Ce O 2 Oxygen   storage Oxygen   release C e 1 x 4 + C e x 3 + O 2 x / 2 V O x / 2 + x / 4 O 2
Ce 2 O 3   2 CeO 2   2 Ce Ce + 3 O O × + V O
Doping CeO2 with other metallic cations proved to be efficient in enhancing its OSC [10,11]. Inspired by La/N [12] and Sm/N [13] co-doped TiO2 for enhanced photocatalytic activities, Yb/N doping should be a feasible method for modifying the OSC of CeO2 in this work. However, no reports had been found on the OSC of cationic and anionic co-doping of CeO2. Based on the similarity–intermiscibility theory and the similar ionic radii of the ytterbium cation (Yb3+; 0.98 Å) and the cerium cation (Ce4+; 0.97 Å), the doping of Yb cations into the CeO2 lattice was feasible, which could be supported by previous reports [14,15]. Nitrogen (N) was a favorite and preferred non-metal dopant due to its relatively small ionization energy and similar ionic radii to oxygen (O). However, besides N-doped TiO2 [16,17], there have been very limited reports on doping with N anions into the CeO2 lattice. Since the O2− anion (1.38 Å) has a larger size than that of the Ce4+ cation (0.97 Å) in the ionic crystal of CeO2, it was more difficult to be substituted by larger N3− anions (1.46 Å) [18,19]. In our previous work, we successfully synthesized N-doped CeO2 using ammonium persulfate as an inorganic nitrogen source [20].
Hence, the modification of CeO2 by co-doping with Yb cation and N anion should be an effective method to enhance its OSC. In our strategy, Yb-/N-doped CeO2 was prepared by a solvothermal method followed by calcination in air with Yb(NO3)3∙5H2O as a Yb source and triethanolamine (TEA) as a N source. For Yb/N-co-doped CeO2 samples, the nominal content of Yb was 5 mol.%, namely, Yb/(Yb + Ce) (mol.%) = 5, while the molar ratio of N was TEA/Ce(NO3)3∙6H2O = 5, 10, 15, 20, 25, and 30 mol.%, and the as-obtained samples were denoted 5%N/5%Yb-, 10%N/5%Yb-, 15%N/5%Yb-, 20%N/5%Yb-, 25%N/5%Yb-, and 30%N/5%Yb-doped CeO2, respectively. Moreover, 1~7 mol.% Yb-doped CeO2 was synthesized in the absence of TEA, while pure or undoped CeO2 was synthesized in the absence of Yb(NO3)3∙5H2O and TEA.

2. Experimental Procedure

2.1. Starting Materials

Ce(NO3)3∙6H2O (99.95%), Yb(NO3)3∙5H2O (99.9%), and triethanolamine (TEA, ≥99.0%) were supplied by Aladdin Co. Ltd. (Shanghai, China). Ethylene glycol (AR) and ethanol (≥99.7%) were supplied by Chengdu Kelong Chemical Co. Ltd. (Chengdu, China). Distilled water was used in all experiments, and all chemicals were used as received without further purification.

2.2. Synthesis of Yb/N-Co-Doped CeO2

Yb/N-co-doped CeO2 was prepared by a solvothermal method followed by calcination in air. Typically, 3.8 mmol Ce(NO3)3∙6H2O, 0.2 mmol Yb(NO3)3∙5H2O, and the desired amounts of TEA were dissolved in a 30 mL mixed solution of ethylene glycol and distilled water (10 vol.% H2O) in a 50 mL Teflon bottle. Then, the bottle was sealed in a stainless-steel autoclave, transferred to an electric oven, and maintained at 200 °C for 24 h. After the autoclave naturally cooled to room temperature, the resulting precipitate was washed with distilled water and ethanol and then dried at 60 °C for 24 h. Finally, the Yb and N co-doping CeO2 powders were obtained by following calcination at 500 °C for 2 h in air. For the Yb/N-co-doping CeO2, Yb/(Yb + Ce) (mol.%) = 5, TEA/Ce(NO3)3∙6H2O (mol.%) = 5, 10, 15, 20, 25, and 30 mol.%, the as-obtained samples designated as 5%N/5%Yb-, 10%N/5%Yb-, 15%N/5%Yb-, 20%N/5%Yb-, 25%N/5%Yb-, and 30%N/5%Yb-doped CeO2, respectively.

2.3. Synthesis of Yb-Doped CeO2

Yb-doped CeO2 with different molar concentrations of Yb cations was synthesized using the same procedure as controls, but in the absence of TEA. Typically, appropriate amounts of Ce(NO3)3∙6H2O and Yb(NO3)3∙5H2O with a total of 4.0 mmol were dissolved in a 30 mL mixed solution of ethylene glycol and distilled water (10 vol.% H2O) in a 50 mL Teflon bottle. Then, the bottle was sealed in a stainless-steel autoclave, transferred to an electric oven, and maintained at 200 °C for 24 h. After the autoclave naturally cooled to room temperature, the resulting precipitate was washed with distilled water and ethanol and then dried at 60 °C for 24 h. Finally, the Yb-doped CeO2 powders were obtained by following calcination at 500 °C for 2 h in air. The as-obtained Yb-doped CeO2 powders with different molar concentrations of Yb were labeled as 1%Yb-doped CeO2, 2%Yb-doped CeO2, 3%Yb-doped CeO2, 4%Yb-doped CeO2, 5%Yb-doped CeO2, 6%Yb-doped CeO2, and 7%Yb-doped CeO2.

2.4. Synthesis of Undoped CeO2

Pure or undoped CeO2 was synthesized using the same procedure as controls, but in the absence of Yb(NO3)3∙5H2O and TEA. Typically, 4.0 mmol Ce(NO3)3∙6H2O was dissolved in a 30 mL mixed solution of ethylene glycol and distilled water (10 vol.% H2O) in a 50 mL Teflon bottle. Then, the bottle was sealed in a stainless-steel autoclave, transferred to an electric oven, and maintained at 200 °C for 24 h. After the autoclave naturally cooled to room temperature, the resulting precipitate was washed with distilled water and ethanol and then dried at 60 °C for 24 h. Finally, the pure CeO2 powders were obtained by following calcination at 500 °C for 2 h in air.

2.5. Characterization

The crystallographic phases of samples were characterized by X-ray diffraction (XRD, D/MAX 2200 PC, Rigaku, Japan) with 40 kV tube voltages and 40 mA current. The morphologies of samples were evaluated by field-emission scanning electron microscopy (SEM; JEOL–7500F, Tokyo, Japan) with an acceleration voltage of 5 kV. N2 adsorption–desorption isotherms were measured using a QuadraSorb SI surface area analyzer (Quantachrome, Boynton Beach, FL, USA), and the BET-specific surface areas were determined using the Brunauer–Emmett–Teller method. The surface composition and binding energy of samples were determined by X-ray photoelectron spectroscopy (XPS, ESCALAB 250Xi, Thermo Scientific, Waltham, MA, USA). The oxygen vacancy defects of samples were characterized using a Raman spectrometer (LabRAM Aramis, Horiba Jobin–Yvon, Paris, France) with a He–Cd laser of 325 nm.

2.6. Evaluation of OSC

Hydrogen temperature-programmed reduction (H2–TPR) measurements were employed to evaluate the OSC of Yb/N-doped CeO2 samples, which were performed using a TP–5080 instrument with a thermal conductivity detector (TCD). Briefly, 0.05 g of sample was pre-treated in a 5 vol.%–O2/N2 flow (30 mL/min) at 500 °C for 1 h and was cooled down under this flowing O2/N2. Then, the sample was purged with high-purity N2 to remove the excess O2 on the surface. Finally, a 5 vol.%–H2/N2 flow (30 mL/min) was introduced into the reactor, which was heating to ~960 °C (10 °C/min), and the change in H2 concentration of the outlet gases was monitored online by TCD.

3. Results and Discussion

Figure 1a shows the XRD patterns of undoped, 5%Yb-doped, 25%N/5%Yb-doped, and 30%N/5%Yb-doped CeO2. All identified peaks matched well with the standard CeO2 (JCPDS no. 34–0349) pattern, indicating that the CeO2 phase with a fluorite structure was obtained, and no other phases (such as CeN, YbN, and Yb2O3) were detected. Moreover, the corresponding lattice parameters were estimated using Bragg’s equation; the results and the appropriate linear fitting are shown in Figure 1b. The lattice parameters of Yb-doped CeO2 and Yb/N-co-doped CeO2 were greater than those of undoped CeO2, which revealed that the possibly partial substitutions of Ce4+ (0.97 Å) and O2− (1.38 Å) ions by the larger Yb3+ (0.98 Å) and N3− (1.46 Å) ions happened in the CeO2 lattice, and the local lattice distortion (such as lattice expansion) occurred as a result. For Yb-doped CeO2, the lattice parameters increased linearly with increasing Yb content until 5%, suggesting that the concentration of Yb in CeO2 reached the solid solubility limit with the addition of 5% Yb. Upon the introduction of TEA, the lattice parameters of 5%Yb-doped CeO2 continued to increase almost linearly until 20%N, implying that N anions were saturated in the 5%Yb-doped CeO2 lattice with the addition of 20%N. In addition, when Yb and N ions reached the solid solubility limit in CeO2, and the amount of Yb(NO3)3∙5H2O or TEA continued to increase, their lattice parameters decreased, which indicated that the solid solubility limits of Yb and N in CeO2 were supersaturated concentrations. This was attributed to the fact that CeO2 itself had a large number of defect structures, leaving CeO2 in a metastable state.
SEM was used to study the effect of Yb/N doping on the morphologies of CeO2. As seen in Figure 1c, the morphology of undoped CeO2 was a multilayered structure consisting of flakes, and these flakes intertwined to form an open porous structure. After the addition of Yb and N ions, the morphologies of 5%Yb-doped CeO2 and 25%N/5%Yb-doped CeO2 still maintained the multilayered structure, as shown in Figure 1d,e, respectively. Surprisingly, 30%N/5%Yb-doped CeO2 in Figure 1f has a completely different morphology from the original multilayered structure. It could be concluded that the addition of TEA had a certain effect on the morphology of CeO2. Due to TEA’s alkalescence, excessive TEA addition would destroy the original equilibrium of the self-assembled multilayer morphology in the solvothermal system, which promoted the formation of spheroid particle aggregates.
To further clarify the porous structures of the as-obtained CeO2, N2 adsorption–desorption experiments were conducted. Figure 1g shows the N2 adsorption–desorption isotherms of undoped, 5%Yb-doped, 25%N/5%Yb-doped, and 30%N/5%Yb-doped CeO2. All isotherms were consistent with type IV hysteresis loops, confirming their mesoporous structure [21]. The specific surface area (m2/g) was an objective, reliable, and physically meaningful size metric for porous materials, and the well-known Brunauer–Emmett–Teller (BET) equation was usually used to determine the specific surface area from the physical adsorption of a gas on a solid surface (labeled as SBET) [22]. The SBET of undoped, 5%Yb-doped, 25%N/5%Yb-doped, and 30%N/5%Yb-doped CeO2 powders was estimated and is shown in Figure 1h. The SBET of 5%Yb-doped CeO2 was almost constant with a value of 93.1 m2/g, while the SBET value of 25%N/5%Yb co-doped CeO2 was 107.3 m2/g, slightly higher than that of the undoped sample (96.0 m2/g), but the SBET value of 30%N/5%Yb-doped CeO2 was only 52.5 m2/g. It could be concluded that the introduction of TEA had a certain effect on the SBET, especially when the TEA addition exceeded a certain value. The morphology not only changed significantly, but the SBET also decreased sharply.
XPS analysis was performed to determine the chemical composition of Yb/N doped CeO2 and identify the influence of Yb/N doping on VO defects and Cen+ ions in the CeO2 crystal. Figure 2a shows the XPS survey spectrum of 20%N/5%Yb-doped CeO2. All wide-scan spectra showed clear CeO2 features by the signals ascribed to Ce 3d and 4d and O KLL and 1s. Fortunately, the Yb signal could also be detected by the presence of the Yb 4d peak. In addition, the weak peak of the N element could also be clearly observed from the high-resolution XPS spectrum of N 1s in the Figure 2a inset. Figure 2b shows the Ce 3d XPS core-level spectrum of 20%N/5%Yb-doped CeO2. The curve of the Ce 3d spectrum was fitted by eight peaks; the bands u4, u3, and u1 (and those for vi) were attributed to the Ce4+ state, while the u2 and v2 bands were due to the Ce3+ state [23]. Moreover, the O 1s XPS peak of 20%N/5%Yb-doped CeO2 is shown in Figure 2c, which was divided into four separate peaks by Gaussian distributions, indicative of the presence of four kinds of oxygen species on CeO2. The peaks labeled β, γ, and δ could be assigned to lattice oxygen (β for O–Ce4+ species, γ for O–Ce3+ species, and δ for O–Yb species), whereas the peak labeled α could be assigned to the chemisorption of oxygen or/and weekly bonded oxygen species related to VO defects [24].
The relative concentration of Ce3+ ions in CeO2, labeled as [Ce3+], could be calculated by comparing the integrated peak areas of the peak related to Ce3+ ions to those of all peaks in Figure 2b. Meanwhile, the relative oxygen vacancy content in CeO2 (labeled as [VO]) could also be estimated from O 1s XPS in Figure 2c by the ratio of the integrated area of the α peak to that of all peaks. Figure 2d shows the calculated [Ce3+] and [VO] values of undoped, 5%Yb-doped, and 20%N/5%Yb-doped CeO2. Both [Ce3+] and [VO] values of 5%Yb-doped CeO2 were higher than those of undoped CeO2, which could be explained as follows. Yb3+ cations were introduced into the CeO2 lattice by the substitution of Ce4+ ions, and a substoichiometric CeO2–x unit was formed with increasing VO defects based on the vacancy compensation mechanism, accompanied by an increase in [Ce3+] due to the activation effect of doping. The substitution reaction of Ce4+ by Yb3+ cations could be written using Kroger and Vink notations as Reaction (3). Moreover, for 20%N/5%Yb-doped CeO2, the [Ce3+] and [VO] values were increased further compared with those of 5%Yb-doped CeO2, indicating that N anions had been incorporated into the CeO2 lattice to form solid solutions, as shown in Reaction (4) using Kroger and Vink notations:
Yb 2 O 3   2 CeO 2   2 Yb Ce + 3 O O × + V O
CeN   CeO 2   Ce Ce + N O + V O
Raman scattering is a very powerful tool for identifying the nature of surface VO defects [25]. Figure 3a compares the Raman spectra of undoped, 5%Yb-doped, and 20%N/5%Yb-doped CeO2. For undoped CeO2, the visible Raman spectrum was dominated by the strong F2g mode of the CeO2 fluorite phase at ~462 cm−1, as well as a weak band at ~592 cm−1 assigned to the defect-induced mode, related to VO defects. This finding indicated that a certain number of intrinsic VO defect sites existed in undoped CeO2. Both the band intensities at ~462 and ~596 cm−1 had changed with the introduction of Yb and N ions in the CeO2 lattice. Typically, compared with the defect-induced band of undoped CeO2 at ~592 cm−1, that of 5%Yb-doped CeO2 had comparable intensity as the F2g mode at ~462 cm−1. In the case of 20%N/5%Yb-doped CeO2, the band intensity of the defect-induced band was even stronger than that of the F2g band. Moreover, an alternative approach was provided to estimate the relative concentration of VO defects in CeO2, namely, the relative number of VO defects in CeO2 could be determined by the intensity ratio of the bands at ~592 and ~462 cm−1, labeled as I592/I462. As observed in Figure 3b, with the increasing amount of N in 5%Yb-doped CeO2, the relative concentration of VO defects, that is, the value of I592/I462, increased almost in a straight line and reached a maximum in 20%N/5%Yb-doped CeO2. This proved that the additive amount of N anions in fluorite CeO2 had an important influence on the formation of VO defects in Yb/N-co-doped CeO2.
Figure 4a displays the H2–TPR profiles of 5%Yb-doped and 20%N/5%Yb-doped CeO2, as well as undoped CeO2 for comparison. For the H2–TPR spectrum of undoped CeO2, the reduction occurred at 200 °C and reached two maximum H2 consumptions at ~505 and ~776 °C, implying the existence of at least two kinds of oxygen species at various coordination environments. In other words, the reduction behavior of pure CeO2 was divided into two steps: The reduction band at 200~600 °C corresponded to the reduction of surface oxygen species, while the reduction band above 600 °C was assigned to the reduction of bulk oxygen, which could migrate to the CeO2 surface through VO defects and react with H2. Compared to undoped CeO2, the 5%Yb-doped and 20%N/5%Yb-doped CeO2 could release a certain amount of oxygen below 200 °C, and there appeared to be a visible shoulder from 400 °C in the H2–TPR profiles. Moreover, the reduction bands at ~600 °C were far higher than the baseline. These findings indicated that Yb/N doping could effectively improve the OSC of CeO2 by increasing the number of VO defects. In addition, 5%Yb-doped and 20%N/5%Yb-doped CeO2 exhibited a slightly higher redox temperature than that of undoped CeO2, which was attributed to the substitutions of Ce4+ and O2− with the larger Yb3+ and N3− and improved the stability of the CeO2 lattice.
OSC is a fundamental characteristic and an important indicator of oxygen storage materials. In CeO2-supported catalysts, the porous structure of CeO2 is conducive to the loading of other active components, while the excellent OSC favors regulating the oxygen content of the catalytic system. Therefore, the quantification of OSC was the key to comparing their oxygen storage/release properties. For that, the OSC was quantified by the amount of H2 consumption per gram of CeO2 (mmol H2/g CeO2) based on H2–TPR curves. Finally, the quantified OSC was obtained by calculating the amount of O2 released per gram of sample (mmol O2/g CeO2) according to the H2 consumptions, and the quantified OSC at low temperatures is shown in Figure 4b. Compared with the OSC of undoped CeO2 (0.115 mmol O2/g), that of 5%Yb-doped CeO2 reached as high as 0.222 mmol O2/g with a 93.04% increase. Upon the introduction of N anions, the OSC of Yb/N-co-doped CeO2 continued to increase until 20%N, reached a maximum of 0.274 mmol O2/g with a 138.26% increase, decreasing at higher N content. In our previous work, the OSC of CeO2 increased by 119.02% and 40.22% upon 5% Hf-doping and 3% Sn-doping, respectively [26]. In our previous study, the OSC of N-doped CeO2 microspheres was 0.73 mmol O2/g [20]. Among the existing reported OSC data [27,28,29,30,31,32,33], our OSC in this work was above the average level. Even so, the idea of enhancing the OSC of CeO2 by cation/anion co-doping has been proven to be feasible in this work.
SBET is an important factor affecting the OSC of CeO2, so we developed a function for the OSC/SBET vs. lattice parameter as well as the OSC/SBET vs. VO concentration. These data were linearly fitted based on the partial least squares method, as shown in Figure 5a,b, and the calculated parameters are given as tables in Figure 5a,b. Both Figure 5a,b showed a linear dependence; however, the fitting of the OSC/SBET vs. lattice parameter showed a higher correlation coefficient (R2 = 0.98257) than that of OSC/SBET vs. VO concentration (R2 = 0.92349), implying that a larger lattice distortion produced a higher OSC per SBET for Yb/N-co-doped CeO2. This lattice distortion includes not only the VO defects, but also the lattice expansion, as well as the reducible–reoxidizable Cen+ ions, and so on. Thus, the OSC/SBET could be predicted as a function of the lattice parameter for Yb/N co-doping CeO2.

4. Conclusions

In summary, Yb/N-doped mesoporous CeO2 with a multilayered structure was synthesized by a solvothermal method using Yb(NO3)3∙5H2O and triethanolamine as cationic and anionic dopants, respectively. Both the original morphologies and fluorite crystal structure of undoped CeO2 could be maintained even after the incorporation of 25%N/5%Yb, but the incorporation of Yb/N into CeO2 could slightly affect their SBET. Both the introduction of Yb cations and N anions could cause lattice expansion, accompanied by the formation of more VO defects and reducible–reoxidizable Cen+ ions. Compared with the undoped CeO2 (0.115 mmol O2/g), the OSC of 5%Yb-doped CeO2 increased to 0.222 mmol O2/g with a 93.04% increase, while that of 20%N/5%Yb-doped CeO2 enhanced to 0.274 mmol O2/g with a 138.26% increase.

Author Contributions

Conceptualization, Y.X.; validation, P.W.; formal analysis, P.W.; Investigation, Y.X. and P.W.; Resources, Y.X.; data curation, P.W.; writing—original draft, Y.X.; writing—review & editing, Y.X., L.G. and Z.D.; supervision, Z.D.; project administration, L.G. and Z.D.; funding acquisition, Y.X. and Z.D. All authors have read and agreed to the published version of the manuscript.

Funding

This study was financially supported by the Opening Project of Crystalline Silicon Photovoltaic New Energy Research Institute, China (2022CHXK002), the Leshan Normal University Research Program, China (KYPY2023–0001), and the Fundamental Research Funds for the Central Universities, China (2023CDJXY–019).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gaálová, J.; Topka, P. Gold and Ceria as Catalysts for VOC Abatement: A Review. Catalysts 2021, 11, 789. [Google Scholar] [CrossRef]
  2. Grabchenko, M.V.; Mikheeva, N.N.; Mamontov, G.V.; Salaev, M.A.; Liotta, L.F.; Vodyankina, O.V. Ag/CeO2 Composites for Catalytic Abatement of CO, Soot and VOCs. Catalysts 2018, 8, 285. [Google Scholar] [CrossRef] [Green Version]
  3. Ganduglia-Pirovano, M.V. The non–innocent role of cerium oxide in heterogeneous catalysis: A theoretical perspective. Catal. Today 2015, 253, 20–32. [Google Scholar] [CrossRef]
  4. Stoian, M.; Rogé, V.; Lazar, L.; Maurer, T.; Védrine, J.C.; Marcu, I.C.; Fechete, I. Total Oxidation of Methane on Oxide and Mixed Oxide Ceria-Containing Catalysts. Catalysts 2021, 11, 427. [Google Scholar] [CrossRef]
  5. Fornasiero, P.; Trovarelli, A. Catalysis by ceria. Catal. Today 2015, 253, 1–2. [Google Scholar] [CrossRef]
  6. Wang, L.; Deo, S.; Dooley, K.; Janik, M.J.; Rioux, R.M. Influence of metal nuclearity and physicochemical properties of ceria on the oxidation of carbon monoxide. Chin. J. Catal. 2020, 41, 951–962. [Google Scholar] [CrossRef]
  7. Reyna-Alvarado, J.; López-Galán, O.A.; Ramos, M.; Rodríguez, J.; Pérez-Hernández, R. A theoretical catalytic mechanism for methanol reforming in CeO2 vs Ni/CeO2 by energy transition states profiles. Catal. Today 2022, 392–393, 146–153. [Google Scholar] [CrossRef]
  8. Shcherbakov, A.B.; Ivanov, V.K.; Zholobak, N.M.; Ivanova, O.S.; Krysanov, E.Y.; Baranchikov, A.E.; Spivak, N.Y.; Tretyakov, Y.D. Nanocrystalline ceria based materials—Perspectives for biomedical application. Biophysics 2011, 56, 987–1004. [Google Scholar] [CrossRef]
  9. Lu, J.; Asahina, S.; Takami, S.; Yoko, A.; Seong, G.; Tomai, T.; Adschiri, T. Interconnected 3D Framework of CeO2 with High Oxygen Storage Capacity: High-Resolution Scanning Electron Microscopic Observation. ACS Appl. Nano Mater. 2020, 3, 2346–2353. [Google Scholar] [CrossRef]
  10. Zhu, Y.; Chen, C.; Cheng, P.; Ma, J.; Yang, W.; Yang, W.; Peng, Y.; Huang, Y.; Zhang, S.; Seong, G. Recent advances in hydrothermal synthesis of facet-controlled CeO2-based nanomaterials. Dalton Trans. 2022, 51, 6506–6518. [Google Scholar] [CrossRef]
  11. Ray, K.; Sinhamahapatra, A.; Sengupta, S.; Prasad, M. Ni/CexZr1–xO2 catalyst prepared via one-step co-precipitation for CO2 reforming of CH4 to produce syngas: Role of oxygen storage capacity (OSC) and oxygen vacancy formation energy (OVFE). J. Mater. Sci. 2022, 57, 2839–2856. [Google Scholar] [CrossRef]
  12. Zhang, X.; Zhou, G.; Zhang, H.; Wu, C.; Song, H. Characterization and activity of visible light-driven TiO2 photocatalysts co-doped with nitrogen and lanthanum. Transit. Met. Chem. 2011, 36, 217–222. [Google Scholar] [CrossRef]
  13. Ma, Y.; Zhang, J.; Tian, B.; Chen, F.; Wang, L. Synthesis and characterization of thermally stable Sm,N co-doped TiO2 with highly visible light activity. J. Hazard. Mater. 2010, 182, 386–393. [Google Scholar] [CrossRef]
  14. Sheng, Y.; Yang, L.; Luan, H.; Liu, Z.; Yu, Y.; Li, J.; Dai, N. Improvement of radiation resistance by introducing CeO2 in Yb-doped silicate glasses. J. Nucl. Mater. 2012, 427, 58–61. [Google Scholar] [CrossRef]
  15. Singh, A.K.; Kumar, K.; Rai, S.B.; Kumar, D. Upconversion studies on Yb3+/Er3+ doped CeO2 and CeF3 phosphors: Enhanced near infrared emission. Solid State Commun. 2013, 169, 1–5. [Google Scholar] [CrossRef]
  16. Xiang, Q.; Yu, J.; Wang, W.; Jaroniec, M. Nitrogen self-doped nanosized TiO2 sheets with exposed {001} facets for enhanced visible-light photocatalytic activity. Chem. Commun. 2011, 47, 6906–6908. [Google Scholar] [CrossRef]
  17. Yin, Z.; Wang, Z.; Wang, J.; Wang, X.; Song, T.; Wang, Z.; Ma, Y. HF promoted increased nitrogen doping in TiO2(B) photocatalyst. Chem. Commun. 2020, 56, 5609–5612. [Google Scholar] [CrossRef]
  18. Jorge, A.B.; Sakatani, Y.; Boissière, C.; Laberty-Roberts, C.; Sauthier, G.; Fraxedas, J.; Sanchez, C.; Fuertes, A. Nanocrystalline N-doped ceria porous thin films as efficient visible-active photocatalysts. J. Mater. Chem. 2012, 22, 3220–3226. [Google Scholar] [CrossRef]
  19. Jorg, A.B.; Fraxedas, J.; Cantarero, A.; Williams, A.J.; Rodgers, J.; Attfield, J.P.; Fuertes, A. Nitrogen doping of ceria. Chem. Mater. 2008, 20, 1682–1684. [Google Scholar] [CrossRef]
  20. Xu, Y.; Li, R. Wet-chemical synthesis and characterization of nitrogen-doped CeO2 powders for oxygen storage capacity. Appl. Surf. Sci. 2018, 455, 997–1004. [Google Scholar] [CrossRef]
  21. Li, Y.; Zhang, X.; Chao, Z.; Gan, M.; Liao, J.; Ye, X.; You, W.; Fu, J.; Wen, H. A multidentate polymer microreactor route for green mass fabrication of mesoporous NaYF4 clusters. Chem. Commun. 2022, 58, 1764–1767. [Google Scholar] [CrossRef]
  22. Hill, T.L. Adsorption from a one-dimensional lattice gas and the Brunauer–Emmett–Teller equation. Proc. Natl. Acad. Sci. USA 1996, 93, 14328–14332. [Google Scholar] [CrossRef] [PubMed]
  23. Rui, Z.; Qin, Z.; Wei, Z. Enhanced catalytic performance of F-doped CeO2–TiO2 catalysts in selective catalytic reduction of NO with NH3 at low temperatures. Res. Chem. Intermed. 2015, 41, 3479–3490. [Google Scholar] [CrossRef]
  24. Zhang, Y.; Zhang, L.; Deng, J.; Dai, H.; He, H. Controlled Synthesis, Characterization, and Morphology-Dependent Reducibility of Ceria-Zirconia-Yttria Solid Solutions with Nanorod-like, Microspherical, Microbowknot-like, and Micro-octahedral Shapes. Inorg. Chem. 2009, 48, 2181–2192. [Google Scholar] [CrossRef] [PubMed]
  25. Wu, Z.; Li, M.; Howe, J.; Meyer, H.M.; Overbury, S.H. Probing Defect Sites on CeO2 Nanocrystals with Well-Defined Surface Planes by Raman Spectroscopy and O2 Adsorption. Langmuir 2010, 26, 16595–16606. [Google Scholar] [CrossRef]
  26. Min, P.; Zhang, S.; Xu, Y.; Li, R. Enhanced oxygen storage capacity of CeO2 with doping-induced unstable crystal structure. Appl. Surf. Sci. 2018, 448, 435–443. [Google Scholar] [CrossRef]
  27. Ke, J.; Xiao, J.W.; Zhu, W.; Liu, H.; Si, R.; Zhang, Y.W.; Yan, C.H. Dopant-Induced Modification of Active Site Structure and Surface Bonding Mode for High-Performance Nanocatalysts: CO Oxidation on Capping-free (110)-oriented CeO2:Ln (Ln = La–Lu) Nanowires. J. Am. Chem. Soc. 2013, 135, 15191–15200. [Google Scholar] [CrossRef]
  28. Li, H.; Lu, G.; Dai, Q.; Wang, Y.; Guo, Y.; Guo, Y. Hierarchical Organization and Catalytic Activity of High-Surface-Area Mesoporous Ceria Microspheres Prepared via Hydrothermal Routes. ACS Appl. Mater. Interfaces 2010, 2, 838–846. [Google Scholar] [CrossRef]
  29. Wang, F.; Xu, L.; Shi, W.; Zhang, J.; Wu, K.; Zhao, Y.; Li, H.; Li, H.; Xu, G.; Chen, W. Thermally stable Ir/Ce0.9La0.1O2 catalyst for high temperature methane dry reforming reaction. Nano Res. 2016, 10, 364–380. [Google Scholar] [CrossRef]
  30. Zhou, K.; Wang, X.; Sun, X.; Peng, Q.; Li, Y. Enhanced catalytic activity of ceria nanorods from well-defined reactive crystal planes. J. Catal. 2005, 229, 206–212. [Google Scholar] [CrossRef]
  31. Dutta, G.; Waghmare, U.V.; Baidya, T.; Hegde, M.S.; Priolkar, K.R.; Sarode, P.R. Origin of Enhanced Reducibility/Oxygen Storage Capacity of Ce1–xTixO2 Compared to CeO2 or TiO2. Chem. Mater. 2006, 18, 3249–3256. [Google Scholar] [CrossRef]
  32. Shen, Q.; Wu, M.; Wang, H.; He, C.; Hao, Z.; Wei, W.; Sun, Y. Facile synthesis of catalytically active CeO2 for soot combustion. Catal. Sci. Technol. 2015, 5, 1941–1952. [Google Scholar] [CrossRef]
  33. Ji, X.; Liu, X.; Tong, X.; Luo, T.; Wu, H.; Meng, X.; Zhan, Z. Enhanced activities of nano-CeO2–δ@430L composites by zirconium doping for hydrogen electro-oxidation in solid oxide fuel cells. Int. J. Hydrogen Energy 2016, 41, 11331–11339. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of undoped, 5%Yb-doped, 25%N/5%Yb-doped, and 30%N/5%Yb-doped CeO2 with standard CeO2 (JCPDS No. 34–0349). (b) Lattice parameters and fitting curves of Yb/N-co-doped CeO2. SEM images of (c) undoped, (d) 5%Yb-doped, (e) 25%N/5%Yb-doped, and (f) 30%N/5%Yb-doped CeO2. Corresponding (g) N2 adsorption–desorption isotherms and (h) specific surface areas.
Figure 1. (a) XRD patterns of undoped, 5%Yb-doped, 25%N/5%Yb-doped, and 30%N/5%Yb-doped CeO2 with standard CeO2 (JCPDS No. 34–0349). (b) Lattice parameters and fitting curves of Yb/N-co-doped CeO2. SEM images of (c) undoped, (d) 5%Yb-doped, (e) 25%N/5%Yb-doped, and (f) 30%N/5%Yb-doped CeO2. Corresponding (g) N2 adsorption–desorption isotherms and (h) specific surface areas.
Materials 16 05478 g001
Figure 2. (a) Full-range XPS spectrum, (b) Ce 3d, and (c) O 1s core-level spectra of 20%N/5%Yb-doped CeO2 (the inset in Figure 2a is the corresponding N 1s core-level XPS spectrum). (d) [Ce3+] and [VO] values of undoped, 5%Yb-doped, and 20%N/5%Yb-doped CeO2.
Figure 2. (a) Full-range XPS spectrum, (b) Ce 3d, and (c) O 1s core-level spectra of 20%N/5%Yb-doped CeO2 (the inset in Figure 2a is the corresponding N 1s core-level XPS spectrum). (d) [Ce3+] and [VO] values of undoped, 5%Yb-doped, and 20%N/5%Yb-doped CeO2.
Materials 16 05478 g002
Figure 3. (a) Raman spectra of undoped, 5%Yb-doped, and 20%N/5%Yb-doped CeO2. (b) Relative VO concentrations of Yb/N doped CeO2 calculated using I592/I462 from Raman spectra.
Figure 3. (a) Raman spectra of undoped, 5%Yb-doped, and 20%N/5%Yb-doped CeO2. (b) Relative VO concentrations of Yb/N doped CeO2 calculated using I592/I462 from Raman spectra.
Materials 16 05478 g003
Figure 4. (a) H2–TPR profiles of undoped, 5%Yb-doped, and 20%N/5%Yb-doped CeO2. (b) Relative low-temperature OSC of Yb/N-doped CeO2.
Figure 4. (a) H2–TPR profiles of undoped, 5%Yb-doped, and 20%N/5%Yb-doped CeO2. (b) Relative low-temperature OSC of Yb/N-doped CeO2.
Materials 16 05478 g004
Figure 5. OSC per SBET as a function of (a) OSC/SBET vs. lattice parameter and (b) OSC/SBET vs. VO concentration for Yb/N-co-doped CeO2.
Figure 5. OSC per SBET as a function of (a) OSC/SBET vs. lattice parameter and (b) OSC/SBET vs. VO concentration for Yb/N-co-doped CeO2.
Materials 16 05478 g005
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

Xu, Y.; Gao, L.; Wu, P.; Ding, Z. Novel Mesoporous and Multilayered Yb/N-Co-Doped CeO2 with Enhanced Oxygen Storage Capacity. Materials 2023, 16, 5478. https://doi.org/10.3390/ma16155478

AMA Style

Xu Y, Gao L, Wu P, Ding Z. Novel Mesoporous and Multilayered Yb/N-Co-Doped CeO2 with Enhanced Oxygen Storage Capacity. Materials. 2023; 16(15):5478. https://doi.org/10.3390/ma16155478

Chicago/Turabian Style

Xu, Yaohui, Liangjuan Gao, Pingkeng Wu, and Zhao Ding. 2023. "Novel Mesoporous and Multilayered Yb/N-Co-Doped CeO2 with Enhanced Oxygen Storage Capacity" Materials 16, no. 15: 5478. https://doi.org/10.3390/ma16155478

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