Next Article in Journal
Simple Rate Expression for Catalyzed Ammonia Decomposition for Fuel Cells
Previous Article in Journal
Phenolic Compounds and Antioxidant and Anti-Enzymatic Activities of Selected Adaptogenic Plants from South America, Asia, and Africa
Previous Article in Special Issue
Nickel-Atom Doping as a Potential Means to Enhance the Photoluminescence Performance of Carbon Dots
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Oxygen Storage Capacity of Porous CeO2 by Rare Earth Doping

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
School of Automotive Engineering, Yancheng Institute of Technology, Yancheng 224051, China
5
Department of Chemical Engineering, Illinois Institute of Technology, Chicago, IL 60616, USA
6
College of Materials Science and Engineering, National Engineering Research Center for Magnesium Alloys, Chongqing University, Chongqing 400044, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(16), 6005; https://doi.org/10.3390/molecules28166005
Submission received: 7 July 2023 / Revised: 28 July 2023 / Accepted: 8 August 2023 / Published: 10 August 2023
(This article belongs to the Special Issue Synthesis and Applications of Semiconductor Nanomaterials)

Abstract

:
CeO2 is an important rare earth (RE) oxide and has served as a typical oxygen storage material in practical applications. In the present study, the oxygen storage capacity (OSC) of CeO2 was enhanced by doping with other rare earth ions (RE, RE = Yb, Y, Sm and La). A series of Undoped and RE–doped CeO2 with different doping levels were synthesized using a solvothermal method following a subsequent calcination process, in which just Ce(NO3)3∙6H2O, RE(NO3)3∙nH2O, ethylene glycol and water were used as raw materials. Surprisingly, the Undoped CeO2 was proved to be a porous material with a multilayered special morphology without any additional templates in this work. The lattice parameters of CeO2 were refined by the least–squares method with highly pure NaCl as the internal standard for peak position calibrations, and the solubility limits of RE ions into CeO2 were determined; the amounts of reducible–reoxidizable Cen+ ions were estimated by fitting the Ce 3d core–levels XPS spectra; the non–stoichiometric oxygen vacancy (VO) defects of CeO2 were analyzed qualitatively and quantitatively by O 1s XPS fitting and Raman scattering; and the OSC was quantified by the amount of H2 consumption per gram of CeO2 based on hydrogen temperature programmed reduction (H2–TPR) measurements. The maximum [OSC] of CeO2 appeared at 5 mol.% Yb–, 4 mol.% Y–, 4 mol.% Sm– and 7 mol.% La–doping with the values of 0.444, 0.387, 0.352 and 0.380 mmol H2/g by an increase of 93.04, 68.26, 53.04 and 65.22%. Moreover, the dominant factor for promoting the OSC of RE–doped CeO2 was analyzed.

1. Introduction

Rare earth (RE), known as “Industrial vitamin”, “Industrial monosodium glutamate” and “Mother of new material”, has irreplaceable excellent magnetic, optical, and electrical properties, playing a huge role in improving product performance, increasing product variety and improving production efficiency. Although the amount is small, it can greatly optimize the properties of materials. In view of its large effect and low dosage, RE has become an important national strategic resource in improving product structure, increasing technological content, and promoting industry technological progress, and is broadly utilized in many fields, such as metallurgy, military, petrochemical, glass ceramics, agriculture and new materials, and so on [1,2,3]. Cerium (Ce) is the most abundant RE element in the crust of Earth, which has good redox performance, so that its oxide (cerium oxide, CeO2) shows excellent oxygen transport capacity and oxygen storage/release capacity. Moreover, CeO2 has the advantages of low toxicity and reusability, so it has attracted great attention in the detection of food biological and chemical substances, catalysis and fuel cell fields [4,5].
Oxygen storage materials are binary or multicomponent composite oxides, in which a CeO2 and CeO2–based solid solution are the main components. CeO2 is a significant N–type semiconductor material with high electrical conductivity, an excellent oxygen storage/release capacity and strong redox activity. Moreover, CeO2–based oxygen storage materials are one of the key materials in a three–way catalyst for automobile exhaust purification [6,7], as well as in water–gas–shift [8,9,10], ethanol steam reforming [11,12] and hydrocarbon reforming [13,14,15]. In an oxygen–rich environment, CeO2 can capture ambient oxygen into its own lattice, and release these stored oxygen quickly when the oxygen content of the reaction system is reduced. Because of this, CeO2–based oxygen storage materials can even determine the performance and service life of a catalyst [16]. Especially in the heterogeneous catalytic reactions, they can regulate the fluctuation of the oxygen content in the reaction system through their own oxygen storage and oxygen release characteristics, which can always maintain the best catalytic effect. This ability of CeO2–based composite oxides to store and release oxygen is called its oxygen storage capacity (OSC). The oxygen evolution and absorption equilibrium reaction can be described by Reaction (1) [17,18]:
C e O 2 O x y g e n   s t o r a g e O x y g e n   r e l e a s e C e 1 x 4 + C e x 3 + O 2 x / 2 V x / 2 + x / 4 O 2
where “VO” represents the oxygen vacancy defects produced via the vacancy compensation mechanism. Interestingly, CeO2 can exhibit a large deviation from stoichiometry at low oxygen partial pressure, forming nonstoichiometric oxide CeO2−x. Even after the loss of oxygen from the lattice and the consequent formation of numerous VO, CeO2−x still retains a fluorite crystal structure [19,20] and captures oxygen by filling the VO upon exposure to oxygen, accompanied by the recovery of CeO2 [21]. Moreover, the doping of other metallic elements into the CeO2 lattice could control their structure and physical properties [22,23,24], such as rare–earth elements [25,26,27], transition elements [28,29,30] and alkaline earth elements [31,32,33]. In spite of the successful synthesis of CeO2–based composite oxides, most of the previous reports have focused on the investigation of catalytic performances [34,35], transport properties [36,37] and the origin of room–temperature ferromagnetism [38,39], the theoretical data about OSC were usually quite scattered, and only a few fundamental studies on the OSC of doped CeO2 have been reported. For example, Singh [40] et al. synthesized a series of Ce1−xMxO2−σ (M = Zr, Ti, Pr, Y and Fe) nanocrystallites using the hydrothermal method using melamine and diethylenetriamine as complexing agents; up to 50% Zr and Y, 40% Ti, 25% Pr and 15% Fe were substituted for Ce4+ in CeO2, and Ce0.85Fe0.15O1.85 showed a higher OSC and higher CO conversion at a lower temperature than Ce1−xZrxO2. Ansari et al. [41] reported the redox properties of Fe–doped CeO2 nanoparticles obtained by a polyol–assisted co–precipitation process, and the 10 mol.% Fe doped CeO2 nanoparticles exhibited excellent reduction performance. Si et al. [42] prepared Ce1−xZrxO2 (x = 0~0.8) powders via a mild urea hydrolysis based on the hydrothermal method, and validated a linear relationship between the lattice strain and the OSC value of CeO2–ZrO2 solid solutions. Therefore, the microstructure and OSC of doped CeO2 have to be understood at a fundamental level through a series of dopants to design advanced materials.
For that, four rare earth elements (RE = Yb, Y, Sm and La) were selected as dopants to improve the OSC of CeO2 based on the similarity–intermiscibility theory. In order to avoid the influence of other ions on the doping effect, we only used Ce(NO3)3∙6H2O, RE(NO3)3∙nH2O, ethylene glycol and water as raw materials. Moreover, all experimental conditions and the purity of raw materials were the same, so, the comparison of structure and properties of RE–doped CeO2 was reliable and effective. Based on this, the influence of the dopant elements and their amounts on the non–stoichiometric VO and OSC were investigated and discussed. Surprisingly, the undoped CeO2 was proved to be a porous material with a multilayered morphology without any additional templates, and the effect of RE–doping on morphology of CeO2 also was investigated.

2. Experimental Procedure

2.1. Starting Materials

Ce(NO3)3∙6H2O (99.95%), Yb(NO3)3∙5H2O (99.9%), Y(NO3)3∙6H2O (99.9%), Sm(NO3)3∙6H2O (99.9%) and La(NO3)3∙6H2O (99.9%) were supplied by Aladdin Co., Ltd. (Shanghai, China). Ethylene glycol (99.5%) and ethanol (99.7%) were obtained from Chengdu Kelong Chemical Co., Ltd. (Chengdu, China). Distilled water was used in all experiments.

2.2. Synthesis of Undoped and RE–Doped CeO2

Firstly, the desired amounts of Ce(NO3)3∙6H2O and RE(NO3)3∙nH2O (RE = Yb, Y, Sm and La) with different RE/(RE + Ce) (mol.%) were dissolved in a mixed solution of 25 mL ethylene glycol and 5 mL distilled water, the total amount of Ce3+ and RE3+ ions was 4.0 mmol. Then, the mixed solution was decanted into a 50 mL Teflon–lined stainless steel autoclave and sealed. Subsequently, the solvothermal process lasted for 24 h at 200 °C. After the reaction, the resulting precipitates were collected by centrifugation, and washed thrice alternately with distilled water and ethanol. At this point, the precursors synthesized by the hydrothermal process were obtained after drying in air at 80 °C for 12 h. Finally, a series of RE–doped CeO2 powders were obtained by following calcination in air at 500 °C for 2 h. For comparison, the Undoped CeO2 was synthesized using the same procedure, albeit in the absence of dopants RE(NO3)3∙nH2O.

2.3. Characterization

The actual doping amounts of RE elements in CeO2 were determined using an inductively coupled plasma–atomic emission spectrometer (ICP–AES, SPECTRO ARCOS EOP, Kleve, Germany). The crystallographic phases of samples were characterized by X-ray diffraction (XRD, Rigaku D/MAX 2200 PC, Rigaku, Japan) analysis using graphite monochromatized Cu Karadiation with 40 kV tube voltages and a 40 mA current. The morphologies of CeO2 were observed by field–emission scanning electron microscopy (SEM, JEOL–7500F, Tokyo, Japan). The surface composition and binding energy of CeO2 were determined by X-ray photoelectron spectroscopy (XPS, ESCALAB 250, Thermo Scientific, Waltham, MA, USA). The natures of surface VO defects were identified using Raman spectroscopy (LabRAM Aramis, Horiba Jobin Yvon, Paris, France) with a He–Cd laser of 325 nm. N2 adsorption–desorption isotherms were measured on a QuadraSorb SI (Quantachrome, Boynton Beach, FL, USA), and the specific surface areas were determined using the Brunauer–Emmett–Teller method.

2.4. Evaluation of OSC

For the Undoped and RE–doped CeO2 samples synthesized using the hydrothermal process at 200 °C for 24 h, followed by calcination in air at 500 °C for 2 h, the hydrogen temperature programmed reduction (H2–TPR) measurements were employed to evaluate their OSC, which was carried out on a TP–5080 instrument with a thermal conductivity detector of gas chromatography. Typically, 50 mg CeO2 powder was pre–treated in a 5% O2/N2 stream at 500 °C for 1 h. After cooling down, the sample was purged with N2 to remove the excess O2. Then, a flow of 5% H2/N2 was introduced into the reactor with a flow rate of 30 mL/min, and the temperature was raised to ~650 °C with a heating rate of 10 °C/min.

3. Results and Discussion

XRD analyses were employed to identify the phase composition and crystallographic structure of the as–obtained precursors and samples. Figure 1a showed the XRD patterns of the precursors synthesized using the solvothermal process at 200 °C for 24 h before calcination. For the precursor synthesized without RE, its major phase component was CeCO3OH (JCPDS no. 52–0352), and similar profiles were observed for these precursors synthesized with the introduction of 10 mol.% RE in the solvothermal process. Figure 1b showed the XRD patterns of Undoped and 10 mol.% RE–doped samples synthesized at 200 °C for 24 h after calcination in air at 500 °C for 2 h. All identified peaks had a good match with the standard CeO2 pattern (cubic fluorite structure, JCPDS no. 34–0349), and the intensities of the corresponding diffraction peaks were comparable. Moreover, no impurity phases were detected, such as Yb3O4, Y3O4, Sm3O4 and La3O4. The absence of RE impurity phases could be explained as follows. The RE impurity phases in the sample might exist as highly dispersed amorphous species. Another possibility was that the RE impurity ions partially substituted the host Ce ions to form a solid solution. Compared with Undoped CeO2, the relative diffraction intensities of 10 mol.% RE–doped CeO2 showed no clear differences, suggesting that there was no preferential orientation or preferential crystal growth upon the incorporation of RE. In addition, compared with Undoped CeO2, a recognizable peak shift towards lower diffraction angles for 10 mol.% RE–doped CeO2 was observed. These findings indicate that the larger RE impurity ions partially substituted the host Ce ions to form the RE–based solid solution based on Bragg’s equation, and the cubic fluorite crystal structure of CeO2 was maintained.
When the impurity ions were introduced into the lattice of the matrix, its lattice parameter (a) would change. So, the change in the a value could be used to determine the solubility limit of these dopants in the matrix. In this work, the a value of CeO2 was refined by the least–squares method, in which the highly pure NaCl (≥99.999%) was selected as an internal standard to calibrate the peak position of CeO2. Figure 2a showed the XRD patterns of Undoped CeO2 and 10 mol.% RE–doped CeO2 with the internal standard of NaCl. It could be found that the diffraction intensities of (111) peak from CeO2 and (200) peak from NaCl were comparable, suggesting the feasibility of this internal standard method. Moreover, the a values of Undoped and 1~10 mol.% RE–doped CeO2 were calculated, and the calculated a as a function of RE contents in CeO2 were summarized in Figure 2b. From Figure 2b, the a values of all RE–doped CeO2 were greater than that of the Undoped one (5.4117 Å). Under the same doping concentration, the variation trend of a values was as follows: aYb < aY < aSm < aLa, which was consistent with the sequence of their ionic radii for CN8: RCe (0.97 Å) < RYb (0.98 Å) < RY (1.02 Å) < RSm (1.08 Å) < RLa (1.16 Å) according to Shannon’s compilation [43]. The increased a values after the introduction of RE indicated that the partial host Ce4+ (0.97 Å) ions substituted by the larger RE ions and the local lattice expansion of CeO2 crystal occurred as a result. Moreover, the a values linearly increased with increasing RE contents, reached a maximum at 5, 4, 4 and 7 mol.% for Yb, Y, Sm and La, before decreasing and maintaining a certain level for higher RE contents. This would indicated that the solubility limits of Yb, Y, Sm and La ions in CeO2 were 5, 4, 4 and 7 mol.%.
In order to further confirm the incorporation of RE ions and their effect on the CeO2 lattice, high–resolution electron microscopy (HR–TEM) was performed and the corresponding HR–TEM images of Undoped and 10 mol.% RE–doped CeO2 were synthesized using the hydrothermal process at 200 °C for 24 h, followed by calcination in air at 500 °C for 2 h, as shown in Figure 3. From the HR–TEM image of Undoped CeO2 in Figure 3a, the interplanar spacing was measured with a value of 0.3110 nm, which fitted well with the (111) plane of cubic CeO2, proving the generation of the CeO2 phase. After the incorporation of 10 mol.% RE (RE = Yb, Y, Sm and La), the interplanar spacings of CeO2 in Figure 3b–e had increased to 0.3178, 0.3202, 0.3209 and 0.3231 nm, respectively. Combined with XRD analysis results in Figure 2, both the local lattice expansion and the increased interplanar spacing indicated that these large RE (RYb = 0.98 Å; RY = 1.02 Å; RSm = 1.08 Å; RLa = 1.16 Å) impurity ions partially substituted the host Ce ions (RCe = 0.97 Å), and a solid solution was formed. Importantly, the size of the RE impurity ions was consistent with the trends of interplanar spacing. In other words, the larger the size of the doped RE ion, the greater the interplanar spacing of the as–obtained RE–doped CeO2. In addition, the practical RE contents in CeO2 were measured by ICP–AES, and the results are shown in Table 1. As observed in Table 1, it could be found that the practical RE contents in CeO2 were close to the corresponding nominal doped one.
XPS analysis was employed to probe the surface chemical composition and various oxidation states before and after RE–doping. Figure 4a–e shows the wide–scan XPS spectra of Undoped and 4 mol.% RE–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h, respectively. As observed, all wide–scan XPS spectra showed the clear CeO2 features by the signals of Ce 3d, Ce 4d and O 1s, in good agreement with those XPS patterns of Gd– [44], Y– [45] and Dy– [46] doped CeO2. It is worth noting that the obvious C 1s peaks located at ~284.8 eV were derived from adventitious carbon to calibrate the tested samples. Moreover, the faint RE 3d or RE 4d signals can be seen in the red dotted box in Figure 4b–e, and the corresponding Yb 4d, Y 3d, Sm 3d and La 3d XPS regions were recorded, as shown in Figure 4f–i, respectively. The characteristic peaks in Figure 4f–i implied that the Yb, Y, Sm and La elements were in +3 states. It indicated that the Yb, Y, Sm and La elements had been successfully incorporated into the CeO2 lattice with positive trivalent states (RE3+).
In order to understand the effect of RE–doping on Ce ions in the CeO2 crystal, the Ce 3d XPS regions of Undoped and 4 mol.% RE–doped CeO2 were recorded and fitted, as shown in Figure 5a–e. The Ce 3d XPS core–levels of all CeO2 samples were fitted into eight peaks, corresponding to four pairs of spin–orbit doublets (u1–4 and v1–4) of Ce ions, in which the ui and vi bands corresponded to the contributions of Ce 3d3/2 and Ce 3d5/2. Moreover, the bands of u4, u3 and u1 (and those for v4, v3, v1) were attributed to the Ce4+ state, while u2 and v2 were due to the Ce3+ state [47]. Meanwhile, the relative concentration of Ce3+ ions in CeO2, labeled as [Ce3+]XPS, could be calculated by the ratio of integrated peak areas of the peak related to the Ce3+ species (u2 and v2 peaks) to that of all peaks (u1–4 and v1–4 peaks) in Figure 5, and the results were summarized in Table 2. As observed, the [Ce3+]XPS values of 4 mol.% Yb, Y, Sm and La–doped CeO2 were 13.78, 12.60, 10.94 and 9.78%, respectively, higher than that of Undoped CeO2 (6.54%), which indicates that Undoped CeO2 itself contained a certain number of Ce3+ ions, and RE–doping could promote the formation of Ce3+ species. In other words, the amount of reducible–reoxidizable Cen+ (namely, Ce3+ ⇔ Ce4+) ions increased with the introduction of RE ions into CeO2 lattice, indicating that RE–doping was conducive to improving the redox capacity of CeO2.
To investigate the chemical states of O in CeO2, the O 1s core–level XPS spectra of Undoped and 4 mol.% RE–doped CeO2 were recorded and fitted, as shown in Figure 6a–e. For Undoped CeO2 in Figure 6a, its O 1s XPS spectrum could be curve–fitted into three peaks, indicating the presence of three kinds of oxygen species in CeO2. The peaks with a binding energy of ~529.8 and ~528.4 eV could be assigned to lattice oxygen of O–Ce(IV) species and O–Ce(III) species, respectively, whereas that of ~531.6 eV (yellow region peak) could be assigned to the chemisorption of oxygen or/and weakly bonded oxygen species related to VO defects. For the O 1s spectra of RE–doped CeO2 in Figure 6b–e, besides the above three peaks, a new curve fitting could be observed, which might be attributed to the corresponding O–RE species, namely, the O–Yb species at ~527.6 eV, O–Y species at ~528.2 eV, O–Sm species at ~528.2 eV and O–La species at ~532.9 eV. Furthermore, the relative VO content could be estimated by the ratio of the integrated area of the peak related to the VO defect (yellow region peak in Figure 6a–e) to that of all peaks, labeled as [VO]XPS, and the results were summarized in Table 2. As observed in Table 2, the calculated [VO]XPS values of 4 mol.% Yb, Y, Sm and La–doped CeO2 were 30.00, 26.82, 26.81 and 17.28%, respectively, higher than that of the Undoped one (13.42%). This result indicated that RE–doping was beneficial for the VO creation in CeO2.
From the results of XPS analyses in Figure 4, Figure 5 and Figure 6 and Table 2, it could be concluded that RE elements were successfully incorporated into the CeO2 lattice with positive trivalent states, and RE–doping could increase the amount of redox Cen+ (Ce3+/Ce4+) of CeO2, as well as the VO defects.
Due to its sensitivity to the VO defect, Raman scattering was employed to investigate the structure of Undoped and RE–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h [48,49]. For the Undoped CeO2 in Figure 7a, the peak at ~458 cm−1 was attributed to the triply degenerate F2g mode from the symmetric O–Ce–O stretching mode [50], while the weak peak at ~592 cm−1 was assigned to the optical LO mode related to VO defects [51,52,53]. Upon the incorporation of RE3+ ions into the CeO2 lattice, the band intensity of the F2g mode decreased, while that of the LO mode related to the VO defect increased (Figure 7b–e). It indicated that Undoped CeO2 itself had a certain number of VO defects and RE–doping could favor the presence of substoichiometric CeO2−x underscoring an increase in VO defects, as consistent with the analysis results of O 1s core–level XPS spectra in Figure 6.
The band at ~590 cm−1 in Raman spectra was known to be associated with the VO defect and has been widely observed in substoichiometric CeO2−x [54]. From Figure 7a, the band intensity of both the F2g and LO modes obviously changed upon the incorporation of RE3+ ions into the CeO2 lattice, which was attributed to the increased lattice distortion caused by RE–doping and hence interfered with the vibrations of CeO2−x. It made the quantitative analysis of VO defects difficult. For this, an alternative approach to quantitatively estimate the relative contents of VO defects was adopted by the ratio of the integrated area of the LO mode to that of the F2g mode from the Raman spectra. Figure 8 showed the calculated relative VO concentrations of Undoped and 1~9 mol.% RE–doped CeO2 synthesized by the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h. As observed, there existed a certain amount of VO defects in Undoped CeO2, and the calculated value was 0.67, consistent with the analysis results of the O 1s core–level XPS spectra in Figure 6. These intrinsic VO defects might have evolved from the redox cycle of Cen+ in CeO2 (Ce3+ ⇔ Ce4+). The relative VO concentrations increased almost linearly with increasing RE contents, and reached maximum when the RE contents were 5, 4, 4 and 7 mol.% for Yb, Y, Sm and La–doped CeO2, and gradually decreased above this doping level. Before this turning point, the variation trend of relative VO concentration under the same doping concentration was as follows: Yb > Y > Sm > La, which was consistent with their electronegativity: χYb (1.26) > χY (1.22) > χSm (1.17) > χCe (1.12) > χLa (1.11). After the RE3+ ions substituted the host Ce ions into the CeO2 lattice, the bigger its electronegativity, the stronger its ability to attract the surrounding electrons to itself, and the surrounding O2− anions lost electrons more easily, thus resulting in extrinsic VO defects.
H2–TPR measurements were employed to evaluate the OSC of CeO2. Figure 9a–e illustrated the H2–TPR profiles of Undoped and 4 mol.% RE–doped CeO2 (RE = Yb, Y, Sm and La) synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h. For all CeO2 samples in Figure 9, one can clearly find a distinct H2 reduction band from 30 to 610 °C, with the strongest H2 reduction peak at ~510 °C; the maximum H2 consumption occurred at 510 °C and then decreased until ~600 °C, and after that it tended to rise. The reduction band from 30 °C to ~600 °C could be attributed to the reduction in surface/subsurface lattice oxygen, which was consistent with these reported results [55,56]. Before 200 °C, the RE–doped CeO2 in Figure 9b–e exhibited more H2 consumption than that of the Undoped CeO2; especially for 4 mol.% Y, Sm and La–doped CeO2, a minima at 170 °C occurred. This indicated that the specific surface area of CeO2 played a dominant role in its OSC at low temperatures. To prove this conjecture, we tested the specific surface areas of 4 mol.% Yb, Y, Sm and La–doped CeO2, and the results were summarized in Table 2. The specific surface areas of 4 mol.% Y, Sm and La–doped CeO2 were 98.1, 112.6 and 104.6 m2/g, respectively, higher than that of Undoped CeO2 (96.0 m2/g); however, these decreased after 4 mol.% Yb–doping (89.7 m2/g). Moreover, compared to Undoped CeO2 in Figure 9a, there appeared to be a visible shoulder from ~350 °C in the H2–TPR profiles of RE–doped CeO2 in Figure 9b–e, and the reduction bands of RE–doped CeO2 at ~600 °C were far higher than the baseline. These phenomena suggested that RE–doping optimized the surface states of CeO2, thereby enhancing its OSC.
OSC was the fundamental performance of CeO2 and CeO2–based oxygen storage materials; so, the quantification of OSC was the key to evaluate their oxygen storage/release property. For that, the OSC was quantified using the amount of H2 consumption per gram of CeO2 powders by measuring the corresponding peak areas of H2–TPR profiles in this work. The quantified OSC (labeled as [OSC], mmol H2/g CeO2) from 30 °C to ~600 °C, which was the value of H2 consumption per gram of CeO2 powders, is shown in Figure 10. The [OSC] of Undoped CeO2 was 0.23 mmol H2/g, indicating that pure CeO2 itself possessed a certain OSC, which was attributed to the unique structure of its intrinsic VO defect or the redox cycle of Ce3+ ⇔ Ce4+, supported by the XPS analyses in Figure 5 and Figure 6 and Raman analyses in Figure 7 and Figure 8. For Yb–doped CeO2, the [OSC] value reached a maximum with a doping level of 5 mol.% and decreased at a higher Yb content. Interestingly, Y–, Sm– and La–doped CeO2 also showed similar trends, reaching the maximum H2 consumptions with doping contents of 4, 4 and 7 mol.%, respectively. The [OSC] values of 5 mol.% Yb–, 4 mol.% Y–, 4 mol.% Sm– and 7 mol.% La–doped CeO2 were 0.444, 0.387, 0.352 and 0.380 mmol H2/g, with an increase of 93.04, 68.26, 53.04 and 65.22% compared with that of the Undoped one (0.230 mmol H2/g). These findings indicate that RE–doping could effectively improve the OSC of CeO2, combined with the H2–TPR curves. This enhanced OSC of RE–doped CeO2 could be explained as follows. When RE3+ ions were doped into the CeO2 lattice to substitute host Ce4+ ions, more VO defects would be generated to keep the electric neutrality of the fluorite structure, and a substoichiometric solid solution Ce1–xRExO2−σ (RE = Yb, Y, Sm and La) was formed based on RE–doping. During the H2 reduction of H2–TPR, H2 reacted with a chemisorbed oxygen from the CeO2 surface, which was fixed by intrinsic and extrinsic VO defects on the CeO2 surface. As the surface chemisorbed oxygen was gradually consumed, the intrinsic and extrinsic VO defects were exposed, and the bulk lattice oxygen began to move to the CeO2 surface for replenishment by the VO defects. The oxygen in the bulk RE–doped CeO2 diffused more easily to the surface to fill the VO defects than that in the Undoped CeO2 due to the activation effect of RE3+ dopants which induced oxygen mobility [57].
In order to investigate the effect of RE–doping on the morphology of CeO2, SEM was employed. Figure 11a–e showed the SEM images of Undoped and 10 mol.% Yb, Y, Sm and La–doped CeO2 particles synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h, respectively. From Figure 11a, it could be seen that the morphology of the Undoped CeO2 particle was a multilayered structure consisting of flakes, and these flakes intertwined to form an open porous structure. After the incorporation of 10 mol.% RE (RE = Yb, Y, Sm and La) into CeO2, the multilayered morphology was still maintained, as seen in Figure 11b–e. This finding indicates that the low concentration of RE–doping had little effect on the morphology of CeO2. Generally, CeO2 with a porous structure or special morphology was usually synthesized by a template–based method, in which either surfactants as soft templates or other porous inorganic material as hard templates were used. Surprisingly, the porous CeO2 with a multilayered morphology was obtained without any additional templates in this work. The abundant porous structure and highly specific surface area would undoubtedly enhance the OSC of CeO2. Further analysis of the porous structures was conducted using an N2 adsorption–desorption isotherm, as discussed later.
In order to further demonstrate the porous structure of CeO2, an N2 adsorption–desorption experiment was performed, and the N2 adsorption–desorption isotherm of Undoped CeO2 is shown in Figure 12a. As observed in Figure 12a, the isotherm was similar to the Langmuir IV(a) type according to the IUPAC classification, and an obvious hysteresis loop was observed in the relative pressure range of 0.4~1.0, attributable to the type H3. It suggests that Undoped CeO2 was a mesoporous material with a disordered mesoporous structures [58], and the isotherm was consistent with that of other reported porous CeO2 [59,60,61]. Moreover, the specific surface areas of Undoped CeO2 and RE–doped CeO2 with solubility limits were estimated based on the N2 adsorption–desorption experiment using a Brunauer–Emmett–Teller method, and the results are shown in Figure 12b as a histogram. Combined with the specific surface areas of 4 mol.% RE–doped CeO2 in Table 2, it can be found that RE–doping had a certain influence on the specific surface area of CeO2. However, the specific surface area was not the dominant factor for promoting the OSC of RE–doped CeO2. Among the CeO2 samples with 4 mol.% RE–doping, 4 mol.% Sm–doped CeO2 displayed the minimum [OSC] value of 0.352 mmol H2/g in Figure 10; however, it possessed the maximum specific surface area of 112.6 m2/g in Table 2. Among RE–doped CeO2 with saturation doping concentration, 5 mol.% Yb–doped CeO2 exhibited the minimum specific surface area of 93.1 m2/g in Figure 12b; however, it possessed the maximum [OSC] value of 0.444 mmol H2/g. Alternatively, the morphology was also not a major factor influencing the OSC of RE–doped CeO2, which is supported by the similar multilayered morphology in Figure 11a–e. Combined with the analyses of morphology and specific surface area of Undoped and RE–doped CeO2, one conclusion could be drawn that the enhanced OSC might be attributed to the incorporation of positive trivalent RE3+ ions into the CeO2 lattice, and partially substituted the host Ce4+ ions, promoting the formation of more VO defects and the oxidation/reduction cycle of Ce3+ ⇔ Ce4+. This result could be supported by the lattice parameter analysis in Figure 2, the O 1s XPS analysis in Figure 6 and the Raman spectra analysis in Figure 7.

4. Conclusions

In summary, a series of RE–substituted CeO2 was synthesized just using Ce(NO3)3∙6H2O, RE(NO3)3∙nH2O (RE = Yb, Y, Sm and La), ethylene glycol and water as raw materials. The Undoped CeO2 was proved to be a mesoporous material with a multilayered morphology; both its multilayered morphology and cubic fluorite structure could be maintained even after 10 mol.% RE introduction. The RE elements were successfully incorporated into the CeO2 lattice with positive trivalent states. RE–doping was beneficial for the oxidation/reduction cycle of Ce3+ ⇔ Ce4+, as well as the creation of extrinsic VO defects. The solubility limits of Yb, Y, Sm and La ions in CeO2 were determined as 5, 4, 4 and 7 mol.%. After the incorporation of larger RE3+, the lattice expansion of the CeO2 crystal occurred, and more VO defects appeared, which could induce the oxygen mobility from bulk to surface, and promote its OSC. The [OSC] values were 0.444, 0.387, 0.352 and 0.380 mmol/g, much higher than that of the Undoped one (0.230 mmol/g), with an increase of 93.04, 68.26, 53.04 and 65.22%, respectively. The enhanced OSC of RE–doped CeO2 should be attributed to the impurity–induced defects by the substitution of host Ce4+ with RE3+ into CeO2, rather than the effects of its specific surface area and morphology.

Author Contributions

Conceptualization, Y.X.; validation, P.W., Q.H., Y.Z. and Z.D.; investigation, Y.X., Y.Z. and Z.D.; resources, Y.X.; data curation, L.G., Y.Z. and Z.D.; writing—original draft, Y.X.; writing—review and editing, Y.X., L.G., Y.Z. and Z.D.; supervision, Z.D., P.W., Y.Z. and Q.H.; project administration, P.W., Q.H., L.G., Y.Z. 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), Leshan Normal University Research Program, China (KYPY2023–0001) and Fundamental Research Funds for the Central Universities, China (2023CDJXY–019).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Not available.

References

  1. Arnold, D. Composition-driven structural phase transitions in rare-earth-doped BiFeO3 ceramics: A review. IEEE Trans. Ultrason. Ferroelectr. Freq. Control. 2015, 62, 62–82. [Google Scholar] [CrossRef] [PubMed]
  2. Pan, H.; Ren, Y.; Fu, H.; Zhao, H.; Wang, L.; Meng, X.; Qin, G. Recent developments in rare-earth free wrought magnesium alloys having high strength: A review. J. Alloys Compd. 2016, 663, 321–331. [Google Scholar] [CrossRef]
  3. Ma, Z.; Liu, H.; Luo, Z. A General Method for the Large-Scale Preparation of Multifunctional Magnetic Sm-Fe/Co Nanomaterials. Adv. Funct. Mater. 2023, 33, 2301350. [Google Scholar] [CrossRef]
  4. Fornasiero, P.; Trovarelli, A. Catalysis by ceria. Catal. Today 2015, 253, 1–2. [Google Scholar] [CrossRef]
  5. 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]
  6. Liu, Y.; Yang, J.; Yang, J.; Wang, L.; Wang, Y.; Zhan, W.; Guo, Y.; Zhao, Y.; Guo, Y. Understanding the three-way catalytic reaction on Pd/CeO2 by tuning the chemical state of Pd. Appl. Surf. Sci. 2021, 556, 149766. [Google Scholar] [CrossRef]
  7. Yang, X.; Yang, L.; Lin, S.; Zhou, R. Investigation on properties of Pd/CeO2-ZrO2-Pr2O3 catalysts with different Ce/Zr molar ratios and its application for automotive emission control. J. Hazard. Mater. 2015, 285, 182–189. [Google Scholar] [CrossRef]
  8. Li, Y.; Kottwitz, M.; Vincent, J.L.; Enright, M.J.; Liu, Z.; Zhang, L.; Huang, J.; Senanayake, S.D.; Yang, W.C.D.; Crozier, P.A.; et al. Dynamic structure of active sites in ceria-supported Pt catalysts for the water gas shift reaction. Nat. Commun. 2021, 12, 914. [Google Scholar] [CrossRef]
  9. Dai, H.; Zhang, A.; Xiong, S.; Xiao, X.; Zhou, C.; Pan, Y. The catalytic performance of Ga2O3-CeO2 composite oxides over reverse water gas shift reaction. ChemCatChem. 2022, 14, e202200049. [Google Scholar] [CrossRef]
  10. Si, R.; Flytzani-Stephanopoulos, M. Shape and Crystal-Plane Effects of Nanoscale Ceria on the Activity of Au-CeO2 Catalysts for the Water-Gas Shift Reaction. Angew. Chem. 2008, 120, 2926–2929. [Google Scholar] [CrossRef]
  11. Xiao, Z.; Wu, C.; Wang, L.; Xu, J.; Zheng, Q.; Pan, L.; Zou, J.; Zhang, X.; Li, G. Boosting hydrogen production from steam reforming of ethanol on nickel by lanthanum doped ceria. Appl. Catal. B-Environ. 2021, 286, 119884. [Google Scholar] [CrossRef]
  12. Moraes, T.S.; Neto, R.C.R.; Ribeiro, M.C.; Mattos, L.V.; Kourtelesis, M.; Verykios, X.; Noronha, F.B. Effects of Ceria Morphology on Catalytic Performance of Ni/CeO2 Catalysts for Low Temperature Steam Reforming of Ethanol. Top. Catal. 2015, 58, 281–294. [Google Scholar] [CrossRef]
  13. Kim, J.; Ryou, Y.; Kim, T.H.; Hwang, G.; Bang, J.; Jung, J.; Bang, Y.; Kim, D.H. Highly selective production of syngas (>99%) in the partial oxidation of methane at 480 °C over Pd/CeO2 catalyst promoted by HCl. Appl. Surf. Sci. 2021, 560, 150043. [Google Scholar] [CrossRef]
  14. Liu, H.; Lei, Q.; Miao, R.; Sun, M.; Qin, C.; Zhang, L.; Ye, G.; Yao, Y.; Huang, B.; Ma, Z. Asymmetric Coordination of Single-atom Co Sites Achieves Efficient Dehydrogenation Catalysis. Adv. Funct. Mater. 2022, 32, 2207408. [Google Scholar] [CrossRef]
  15. Prasad, D.H.; Ji, H.I.; Kim, H.R.; Son, J.W.; Kim, B.K.; Lee, H.W.; Lee, J.H. Effect of nickel nano-particle sintering on methane reforming activity of Ni-CGO cermet anodes for internal steam reforming SOFCs. Appl. Catal. B-Environ. 2011, 101, 531–539. [Google Scholar] [CrossRef]
  16. Liu, T.; Yang, R.; Zhang, G.; Wu, W.; Yang, Z.; Lin, R.; Wang, X.; Jiang, Y. Mechanism of selective catalytic reduction of NO with NH3 over CeO2-TiO2: Insight from in-situ DRIFTS and DFT calculations. Appl. Surf. Sci. 2021, 568, 150764. [Google Scholar] [CrossRef]
  17. 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]
  18. Li, R.; Yabe, S.; Yamashita, M.; Momose, S.; Yoshida, S.; Yin, S.; Sato, T. UV-shielding properties of zinc oxide-doped ceria fine powders derived via soft solution chemical routes. Mater. Chem. Phys. 2022, 75, 39–44. [Google Scholar] [CrossRef]
  19. Chen, X.; Zhan, S.; Chen, D.; He, C.; Tian, S.; Xiong, Y. Grey Fe-CeO2-σ for boosting photocatalytic ozonation of refractory pollutants: Roles of surface and bulk oxygen vacancies. Appl. Catal. B-Environ. 2021, 286, 119928. [Google Scholar] [CrossRef]
  20. Trovarelli, A. Structural and oxygen storage/release properties of CeO2-based solid solutions. Comment. Inorg. Chem. 1999, 20, 263–284. [Google Scholar] [CrossRef]
  21. Herz, R.K. Dynamic behavior of automotive catalysts. 1. Catalyst oxidation and reduction. Ind. Eng. Chem. Res. 1981, 20, 451–457. [Google Scholar] [CrossRef]
  22. McFarland, E.W.; Metiu, H. Catalysis by doped oxides. Chem. Rev. 2013, 113, 4391–4427. [Google Scholar] [CrossRef] [PubMed]
  23. Xiao, Y.; Tan, S.; Wang, D.; Wu, J.; Jia, T.; Liu, Q.; Qi, Y.; Qi, X.; He, P.; Zhou, M. CeO2/BiOIO3 heterojunction with oxygen vacancies and Ce4+/Ce3+ redox centers synergistically enhanced photocatalytic removal heavy metal. Appl. Surf. Sci. 2020, 530, 147116. [Google Scholar] [CrossRef]
  24. Gayen, A.; Priolkar, K.R.; Sarode, P.R.; Jayaram, V.; Hegde, M.S.; Subbanna, G.N.; Emura, S. Ce1−xRhxO2−δ Solid Solution Formation in Combustion-Synthesized Rh/CeO2 Catalyst Studied by XRD, TEM, XPS, and EXAFS. Chem. Mater. 2004, 16, 2317–2328. [Google Scholar] [CrossRef]
  25. Swathi, S.; Yuvakkumar, R.; Kumar, P.S.; Ravi, G.; Thambidurai, M.; Dang, C.; Velauthapillai, D. Gadolinium doped CeO2 for efficient oxygen and hydrogen evolution reaction. Fuel 2022, 310, 122319. [Google Scholar] [CrossRef]
  26. Sawka, A.; Kwatera, A. Low temperature synthesis of Y2O3-doped CeO2 layers using MOCVD. Mat. Sci. Eng. B 2022, 276, 115580. [Google Scholar] [CrossRef]
  27. Balaguer, M.; Solís, C.; Serra, J.M. Structural-Transport Properties Relationships on Ce1−xLnxO2−δ System (Ln = Gd, La, Tb, Pr, Eu, Er, Yb, Nd) and Effect of Cobalt Addition. J. Phys. Chem. C 2012, 116, 7975–7982. [Google Scholar] [CrossRef]
  28. Zhang, L.; Spezzati, G.; Muravev, V.; Verheijen, M.A.; Hensen, E. Improved Pd/CeO2 catalysts for low-temperature no reduction: Activation of CeO2 lattice oxygen by Fe doping. ACS Catal. 2021, 11, 5614–5627. [Google Scholar] [CrossRef]
  29. Li, S.; Zhang, H.; Wu, L.; Zhao, H.; Li, L.; Sun, C.; An, B. Vacancy-engineered CeO2/Co heterostructure anchored on the nitrogen-doped porous carbon nanosheet arrays vertically grown on carbon cloth as an integrated cathode for the oxygen reduction reaction of rechargeable Zn-air battery. J. Mater. Chem. A 2022, 10, 9858–9868. [Google Scholar] [CrossRef]
  30. Zhang, J.; Guo, J.; Liu, W.; Wang, S.; Xie, A.; Liu, X.; Wang, J.; Yang, Y. Facile Preparation of Mn+-Doped (M = Cu, Co, Ni, Mn) Hierarchically Mesoporous CeO2 Nanoparticles with Enhanced Catalytic Activity for CO Oxidation. Eur. J. Inorg. Chem. 2015, 6, 969–976. [Google Scholar] [CrossRef]
  31. Mordekovitz, Y.; Shelly, L.; Rosen, B.A.; Hayun, S. Surface properties of Ca, Ti-doped CeO2 and their influence on the reverse water-gas shift reaction. J. Am. Ceram. Soc. 2021, 104, 2237–2247. [Google Scholar] [CrossRef]
  32. Carey, J.J.; Nolan, M. Cation doping size effect for methane activation on alkaline earth metal doping of the CeO2 (111) surface. Catal. Sci. Technol. 2016, 6, 3544–3558. [Google Scholar] [CrossRef]
  33. Murakami, K.; Mizutani, Y.; Sampei, H.; Ishikawa, A.; Sekine, Y. Manipulation of co adsorption over Me1/CeO2 by heterocation doping: Key roles of single-atom adsorption energy. J. Chem. Phys. 2021, 154, 164705. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, Y.; Zhang, Q.; Lin, Y.; Huang, W.; Ding, D.; Zheng, Y.; Chen, M.; Wan, H. Insight into the high efficiency of Cu/CeO2 (110) catalysts for preferential oxidation of CO from hydrogen rich fuel. Appl. Surf. Sci. 2021, 566, 150707. [Google Scholar] [CrossRef]
  35. Liu, H.; Chen, J.; Wang, Y.; Yin, R.; Yang, W.; Wang, G.; Si, W.; Peng, Y.; Li, J. Interaction mechanism for simultaneous elimination of nitrogen oxides and toluene over the bifunctional CeO2-TiO2 mixed oxide catalyst. Environ. Sci. Technol. 2022, 56, 4467–4476. [Google Scholar] [CrossRef]
  36. Chen, L.A.; Lu, Y.S.; Lin, Y.T.; Lee, Y.L. Preparation and characterization of cerium-based conversion coating on a Fe50Mn30Co10Cr10 dual-phase high-entropy alloy. Appl. Surf. Sci. 2021, 562, 150200. [Google Scholar] [CrossRef]
  37. Yang, S.M.; Lee, S.; Jian, J.; Zhang, W.; Lu, P.; Jia, Q.; Wang, H.; Noh, T.T.; Kalinin, S.V.; MacManus-Driscoll, J.L. Strongly enhanced oxygen ion transport through samarium-doped CeO2 nanopillars in nanocomposite films. Nat. Commun. 2015, 6, 8588. [Google Scholar] [CrossRef] [Green Version]
  38. Tarui, K.; Oomori, T.; Ito, Y.; Yamamoto, T. Origin of room-temperature ferromagnetism in Co-doped CeO2. Phys. B Condens. Matter 2021, 619, 413158. [Google Scholar] [CrossRef]
  39. Motaung, D.E.; Tshabalala, Z.P.; Makgwane, P.R.; Mahmoud, F.A.; Oosthuizen, D.N.; Cummings, F.R.; Leshabane, N.; Hintsho-Mbita, N.; Li, X.; Ray, S.S.; et al. Multi-functioning of CeO2-SnO2 heterostructure as room temperature ferromagnetism and chemiresistive sensors. J. Alloy. Compd. 2022, 906, 164317. [Google Scholar] [CrossRef]
  40. Singh, P.; Hegde, M.S. Controlled synthesis of nanocrystalline CeO2 and Ce1−xMxO2−δ (M = Zr, Y, Ti, Pr and Fe) solid solutions by the hydrothermal method: Structure and oxygen storage capacity. J. Solid State Chem. 2008, 181, 3248–3256. [Google Scholar] [CrossRef]
  41. Ansari, A.A.; Labis, J.; Alam, M.; Ramay, S.M.; Ahmad, N.; Mahmood, A. Physicochemical and Redox Characteristics of Fe Ion-doped CeO2 Nanoparticles. J. Chin. Chem. Soc. 2015, 62, 925–932. [Google Scholar] [CrossRef]
  42. Si, R.; Zhang, Y.W.; Li, S.J.; Lin, B.X.; Yan, C.H. Urea-Based Hydrothermally Derived Homogeneous Nanostructured Ce1−xZrxO2 (x = 0–0.8) Solid Solutions: A Strong Correlation between Oxygen Storage Capacity and Lattice Strain. J. Phys. Chem. B 2004, 108, 12481–12488. [Google Scholar] [CrossRef]
  43. Shannon, R.T. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  44. Soni, B.; Makkar, S.; Biswas, S. Effects of surface structure and defect behavior on the magnetic, electrical, and photocatalytic properties of Gd-doped CeO2 nanoparticles synthesized by a simple chemical process. Mater. Charact. 2021, 174, 110990. [Google Scholar] [CrossRef]
  45. Xu, B.; Zhang, Q.; Yuan, S.; Liu, S.; Zhang, M.; Ohno, T. Synthesis and photocatalytic performance of yttrium-doped CeO2 with a hollow sphere structure. Catal. Today 2017, 281, 135–143. [Google Scholar] [CrossRef] [Green Version]
  46. 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]
  47. Burroughs, P.; Hamnett, A.; Orchard, A.F.; Thornton, G. Satellite structure in the X-ray photoelectron spectra of some binary and mixed oxides of lanthanum and cerium. J. Chem. Soc. 1976, 17, 1686–1698. [Google Scholar] [CrossRef]
  48. Fernández-García, M.; Martínez-Arias, A.; Guerrero-Ruiz, A.; Conesa, J.C.; Soria, J. Ce-Zr-Ca Ternary Mixed Oxides: Structural Characteristics and Oxygen Handling Properties. J. Catal. 2002, 211, 326–334. [Google Scholar] [CrossRef]
  49. Dohčević-Mitrović, Z.D.; Grujić-Brojčin, M.; Šćepanović, M.; Popović, Z.V.; Bošković, S.; Matović, B.; Zinkevich, M.; Aldinger, F. Ce1−xY(Nd)xO2−δ nanopowders: Potential materials for intermediate temperature solid oxide fuel cells. J. Phys. Condens. Matter 2006, 18, S2061–S2068. [Google Scholar] [CrossRef]
  50. Lin, X.M.; Li, L.P.; Li, G.S.; Su, W.H. Transport property and Raman spectra of nanocrystalline solid solutions Ce0.8Nd0.2O2−δ with different particle size. Mater. Chem. Phys. 2001, 69, 236–240. [Google Scholar] [CrossRef]
  51. Fernández-García, M.; Wang, X.Q.; Belver, C.; Iglesias-Juez, A.; Hanson, J.C.; Rodriguez, J. Ca Doping of Nanosize Ce-Zr and Ce-Tb Solid Solutions: Structural and Electronic Effects. Chem. Mater. 2005, 17, 4181–4193. [Google Scholar] [CrossRef]
  52. McBride, J.R.; Hass, K.C.; Poindexter, B.D.; Weber, W.H. Raman and x-ray studies of Ce1−xRExO2−y, where RE= La, Pr, Nd, Eu, Gd, and Tb. J. Appl. Phys. 1994, 76, 2435–2441. [Google Scholar] [CrossRef]
  53. Spanier, J.E.; Robinson, R.D.; Zhang, F.; Chan, S.W.; Herman, I.P. Size-dependent properties of CeO2−y nanoparticles as studied by Raman scattering. Phys. Rev. B 2001, 64, 245407. [Google Scholar] [CrossRef] [Green Version]
  54. Pu, Z.Y.; Lu, J.Q.; Luo, M.F.; Xie, Y.L. Study Of Oxygen Vacancies In Ce0.9Pr0.1O2−δ Solid Solution By In Situ X-Ray Diffraction And In Situ Raman Spectroscopy. J. Phys. Chem. C 2007, 111, 18695–18702. [Google Scholar] [CrossRef]
  55. Tang, X.L.; Zhang, B.C.; Li, Y.; Xu, Y.D.; Xin, Q.; Shen, W.J. Carbon monoxide oxidation over CuO/CeO2 catalysts. Catal. Today 2004, 93, 191–198. [Google Scholar] [CrossRef]
  56. 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]
  57. Katta, L.; Sudarsanam, P.; Thrimurthulu, G.; Reddy, B.M. Doped nanosized ceria solid solutions for low temperature soot oxidation: Zirconium versus lanthanum promoters. Appl. Catal. B-Environ. 2010, 101, 101–108. [Google Scholar] [CrossRef]
  58. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S.W. Physisorption of gases, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051–1069. [Google Scholar] [CrossRef] [Green Version]
  59. Zhang, Y.; Shi, R.; Yang, P.; Song, X.; Zhu, Y.; Ma, Q. Fabrication of electronspun porous CeO2 nanofibers with large surface area for pollutants removal. Ceram. Int. 2016, 42, 14028–14035. [Google Scholar] [CrossRef]
  60. Wang, Y.; Bai, X.; Wang, F.; Kang, S.; Yin, C.; Li, X. Nanocasting synthesis of chromium doped mesoporous CeO2 with enhanced visible-light photocatalytic CO2 reduction performance. J. Hazard. Mater. 2017, 372, 69–76. [Google Scholar] [CrossRef]
  61. Zhao, P.S.; Gao, X.M.; Zhu, F.X.; Hu, X.M.; Zhang, L.L. Ultrasonic-assisted Solution-Phase Synthesis and Property Studies of Hierarchical Layer-by-Layer Mesoporous CeO2. Bull. Korean Chem. Soc. 2018, 39, 375–380. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the Undoped and 10 mol.% RE–doped samples synthesized using a solvothermal process at 200 °C for 24 h (a) before and (b) after calcination in air at 500 °C for 2 h.
Figure 1. XRD patterns of the Undoped and 10 mol.% RE–doped samples synthesized using a solvothermal process at 200 °C for 24 h (a) before and (b) after calcination in air at 500 °C for 2 h.
Molecules 28 06005 g001
Figure 2. (a) XRD patterns of Undoped and 10 mol.% RE–doped CeO2 with the internal standard of NaCl, (b) lattice parameter and fitting curves of RE–doped CeO2 (RE = Yb, Y, Sm and La, [RE] ≤ 10 mol.%). The RE (mol.%) in CeO2 = 0 represented the Undoped CeO2 sample.
Figure 2. (a) XRD patterns of Undoped and 10 mol.% RE–doped CeO2 with the internal standard of NaCl, (b) lattice parameter and fitting curves of RE–doped CeO2 (RE = Yb, Y, Sm and La, [RE] ≤ 10 mol.%). The RE (mol.%) in CeO2 = 0 represented the Undoped CeO2 sample.
Molecules 28 06005 g002
Figure 3. HR–TEM images of (a) Undoped CeO2 and 10 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h.
Figure 3. HR–TEM images of (a) Undoped CeO2 and 10 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h.
Molecules 28 06005 g003
Figure 4. Full–range XPS spectra of (a) Undoped, 4 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h; corresponding XPS regions of (f) Yb 4d, (g) Y 3d, (h) Sm 3d and (i) La 3d.
Figure 4. Full–range XPS spectra of (a) Undoped, 4 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h; corresponding XPS regions of (f) Yb 4d, (g) Y 3d, (h) Sm 3d and (i) La 3d.
Molecules 28 06005 g004
Figure 5. Ce 3d core–levels XPS spectra of (a) Undoped, 4 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h.
Figure 5. Ce 3d core–levels XPS spectra of (a) Undoped, 4 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h.
Molecules 28 06005 g005
Figure 6. O 1s core–level XPS spectra of (a) Undoped, 4 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h.
Figure 6. O 1s core–level XPS spectra of (a) Undoped, 4 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h.
Molecules 28 06005 g006
Figure 7. Raman spectra of (a) Undoped, 4 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h.
Figure 7. Raman spectra of (a) Undoped, 4 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h.
Molecules 28 06005 g007
Figure 8. Relative VO concentrations of 0~9 mol.% RE–doped CeO2 calculated using integral area ratio from Raman spectra (RE = Yb, Y, Sm and La). The RE (mol.%) in CeO2 = 0 represented the Undoped CeO2 sample.
Figure 8. Relative VO concentrations of 0~9 mol.% RE–doped CeO2 calculated using integral area ratio from Raman spectra (RE = Yb, Y, Sm and La). The RE (mol.%) in CeO2 = 0 represented the Undoped CeO2 sample.
Molecules 28 06005 g008
Figure 9. H2–TPR profiles of (a) Undoped, 4 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h. (30 mL/min 5%–H2/N2 flow; Heating rate 10 °C/min).
Figure 9. H2–TPR profiles of (a) Undoped, 4 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h. (30 mL/min 5%–H2/N2 flow; Heating rate 10 °C/min).
Molecules 28 06005 g009
Figure 10. Relative [OSC] values of 0~9 mol.% RE–doped CeO2 calculated by measuring the corresponding peak areas of H2–TPR profiles (RE = Yb, Y, Sm and La). Note: [OSC] was the value of quantified OSC using the amount of H2 consumption per gram of CeO2 powders (mmol H2/g CeO2) by measuring the corresponding peak areas of H2–TPR profiles from 30 °C to ~600 °C.
Figure 10. Relative [OSC] values of 0~9 mol.% RE–doped CeO2 calculated by measuring the corresponding peak areas of H2–TPR profiles (RE = Yb, Y, Sm and La). Note: [OSC] was the value of quantified OSC using the amount of H2 consumption per gram of CeO2 powders (mmol H2/g CeO2) by measuring the corresponding peak areas of H2–TPR profiles from 30 °C to ~600 °C.
Molecules 28 06005 g010
Figure 11. SEM images of (a) Undoped, 10 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h.
Figure 11. SEM images of (a) Undoped, 10 mol.% (b) Yb, (c) Y, (d) Sm and (e) La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h.
Molecules 28 06005 g011
Figure 12. (a) N2 adsorption–desorption isotherm of Undoped CeO2, (b) specific surface areas of Undoped, 5 mol.% Yb–doped, 4 mol.% Y–doped, 5 mol.% Sm–doped, 7 mol.% La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h. Note: Specific surface areas were determined based on N2 sorption experiment using the Brunauer–Emmett–Teller method.
Figure 12. (a) N2 adsorption–desorption isotherm of Undoped CeO2, (b) specific surface areas of Undoped, 5 mol.% Yb–doped, 4 mol.% Y–doped, 5 mol.% Sm–doped, 7 mol.% La–doped CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h. Note: Specific surface areas were determined based on N2 sorption experiment using the Brunauer–Emmett–Teller method.
Molecules 28 06005 g012
Table 1. Practical contents and nominal contents of RE in CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h (RE = Yb, Y, Sm and La).
Table 1. Practical contents and nominal contents of RE in CeO2 synthesized using the hydrothermal process at 200 °C for 24 h and followed by calcination in air at 500 °C for 2 h (RE = Yb, Y, Sm and La).
RE in CeO2 (mol.%)YbYSmLa
* Nominal contents259249249279
* Practical RE contents2.585.269.621.924.248.672.414.279.382.196.799.25
* Nominal content (mol.%): w   ( mol . % ) = n R E n R E + n C e × 100 ,   ( n R E + n C e = 4   mmol ) . * Practical RE contents (mol.%): The actual RE doping amounts in CeO2 were determined using ICP–AES, where CeO2 was dissolved in a mixed solution of HNO3–H2O2.
Table 2. [Ce3+]XPS and [VO]XPS of Undoped CeO2 and 4 mol.% RE–doped CeO2 synthesized using solvothermal method at 200 °C for 24 h followed by calcination in air at 500 °C for 2 h (RE = Yb, Y, Sm and La).
Table 2. [Ce3+]XPS and [VO]XPS of Undoped CeO2 and 4 mol.% RE–doped CeO2 synthesized using solvothermal method at 200 °C for 24 h followed by calcination in air at 500 °C for 2 h (RE = Yb, Y, Sm and La).
SampleUndoped CeO24 mol.% RE–Doped CeO2
Parameter YbYSmLa
[Ce3+]XPS (%)6.5413.7812.6010.949.78
[VO]XPS (%)13.4230.0226.8226.8117.28
Specific surface area (m2/g)96.089.798.1112.6104.6
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.; Hou, Q.; Wu, P.; Zhou, Y.; Ding, Z. Enhanced Oxygen Storage Capacity of Porous CeO2 by Rare Earth Doping. Molecules 2023, 28, 6005. https://doi.org/10.3390/molecules28166005

AMA Style

Xu Y, Gao L, Hou Q, Wu P, Zhou Y, Ding Z. Enhanced Oxygen Storage Capacity of Porous CeO2 by Rare Earth Doping. Molecules. 2023; 28(16):6005. https://doi.org/10.3390/molecules28166005

Chicago/Turabian Style

Xu, Yaohui, Liangjuan Gao, Quanhui Hou, Pingkeng Wu, Yunxuan Zhou, and Zhao Ding. 2023. "Enhanced Oxygen Storage Capacity of Porous CeO2 by Rare Earth Doping" Molecules 28, no. 16: 6005. https://doi.org/10.3390/molecules28166005

Article Metrics

Back to TopTop