Next Article in Journal
Conversion of Glucose to 5-Hydroxymethylfurfural Using Consortium Catalyst in a Biphasic System and Mechanistic Insights
Previous Article in Journal
Performance of Eversa Transform 2.0 Lipase in Ester Production Using Babassu Oil (Orbignya sp.) and Tucuman Oil (Astrocaryum vulgar): A Comparative Study between Liquid and Immobilized Forms in Fe3O4 Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Comparative Studies on Effects of Metal Cation (La) and Non-Metal Anion (N) Doping on CeO2 Nanoparticles for Regenerative Scavenging of Reactive Oxygen Radicals

1
Department of Chemical Engineering, The University of Seoul, Siripdae-gil 13, Jeonnong-dong, Seoul 02504, Republic of Korea
2
JNTG Co., Ltd., 240-11 Naehyangan-gil, Jeongnam-myeon, Hwaseong-si 18523, Republic of Korea
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(3), 572; https://doi.org/10.3390/catal13030572
Submission received: 16 February 2023 / Revised: 6 March 2023 / Accepted: 7 March 2023 / Published: 11 March 2023
(This article belongs to the Section Catalytic Materials)

Abstract

:
The intrinsic effects of metal cation (La) and non-metallic anion (N) doping of CeO2 nanoparticles (NPs) for regenerative scavenging of reactive oxygen radicals were studied. La-doped CeO2 NPs were prepared by the conventional impregnation method at various La doping levels. N-doped CeO2 NPs were prepared by urea thermolysis with two different methods: (i) direct thermolysis of urea after physical mixing with CeO2 NPs and (ii) wet impregnation of CeO2 NPs with urea followed by thermolysis under inert N2 atmosphere. Physicochemical properties of samples were characterized by X-ray diffraction, X-ray photoelectron spectroscopy, Raman spectroscopy, and N2 sorption measurement. Radical scavenging properties of the samples were characterized by applying Fenton’s reaction. Results indicated that atomic N doping on CeO2 NPs significantly enhanced radical scavenging properties of CeO2 NPs, resulting in an activity of N-doped CeO2 about 3.6 times greater than the pristine CeO2 NPs and 1.6 times higher than the La-doped CeO2 NPs. This result suggests that anionic N doping of CeO2 NPs is highly effective in enhancing radical scavenging properties of CeO2 NPs, whereas such modifications have been typically practiced by hetero-metal doping with rare earth metal elements. A collective structure–property correlation analysis suggested that enhancement of radical scavenging properties of heteroatom-doped CeO2 NPs was largely attributed to an increase in surface oxygen vacancies on CeO2 NPs due to heteroatom doping.

1. Introduction

Cerium oxides are known to have regenerative scavenging properties for reactive oxygen radicals. Extensive studies have been conducted on their applications in various areas such as polymer electrolyte membrane fuel cells (PEMFCs) and biomedical antioxidants. It has been reported that adding cerium oxide nanoparticles (NPs) into the perfluorosulfonated acid (PFSA) membrane of PEMFCs can mitigate degradation of the polymer membrane caused by reactive hydroxyl and hydroperoxyl radicals generated during electrochemical cell reactions [1,2,3,4,5,6,7,8,9]. In this context, various metal oxides such as ZrO2, YSZ, TiO2, MnO2, and silica-supported Mn, Cr, and Co oxides [5,9,10,11,12] have been explored for radical scavenging applications, with cerium oxides being explored mainly due to their superior properties [13]. In parallel, there have been significant efforts to enhance the radical scavenging properties of cerium oxides. Previous studies have reported that regenerative radical scavenging properties of cerium oxides are attributed to redox cycles between Ce3+ and Ce4+ species involving surface lattice oxygen vacancies [14,15]. Effects of the size and exposed crystal facets of CeO2 NP have been emphasized, in which a decrease in its particle size can lead to an increase in surface oxygen vacancies [13].
The chemical compositions of cerium oxides can also vary by hetero-atom doping, which can modify the intrinsic properties of metal oxides. Various studies have reported promotional effects of hetero-metal doping as metal cations, such as with La [16,17,18], Eu [17,19], Nd [17], Gd [20], Pr [17,20], W [21], and Zr [20,22,23,24]. In these studies, enhancements of radical scavenging activities of heterometal-doped CeO2 are generally rationalized by increases in surface oxygen vacancies and Ce3+ contents. In contrast, only a few recent studies have reported the enhancement of radical scavenging properties of CeO2 NPs by non-metallic anion doping, such as nitrogen doping on the CeO2 surface [25]. Atomic N doping has been applied previously for band gap modification of TiO2 and CeO2 for photocatalytic applications [26,27,28]. Enhancement of catalytic activity of CeO2 NPs by N doping has also been reported for catalytic reduction of nitrogen oxides in automotive applications [29].
The aim of this work is to compare the intrinsic effects of metallic and non-metallic heteroatom doping of CeO2 NPs for radical scavenging properties. Various previous studies have reported enhancements in the radical scavenging properties of CeO2 by heteroatom doping. However, the materials were prepared and their properties were characterized under very different conditions. In addition, comparative studies on the effects of metal cation and non-metal anion dopants on the intrinsic radical scavenging activities of CeO2 NPs have not been reported yet. Herein, La-doped CeO2 and N-doped CeO2 NPs were prepared with various methods at controlled heteroatom doping levels and their intrinsic radical scavenging properties were studied considering doping species, concentration, and surface oxygen vacancy concentrations.

2. Results and Discussion

2.1. Physicochemical Properties of La-Doped CeO2 NPs

Figure 1a shows photographs of pristine CeO2 and La-doped CeO2 NPs obtained at various La loading amounts (2, 5, and 10 wt.%). Pristine CeO2 NPs showed a typical bright yellow color. The La-doped CeO2 NPs also displayed a yellow color. However, the color became darker when La loading amount increased. Figure 1b displays XRD patterns of pristine CeO2 and La-doped CeO2 NPs at various La loading levels. All these samples exhibited a typical diffraction pattern of a face-centered cubic (FCC) crystal structure of CeO2 (JCPDS #82-0792) without any additional diffraction peaks. Formation of any hetero-crystalline phases by La doping treatment was not found. Table 1 shows the cubic lattice parameter (a, Å) of samples calculated from the (111) diffraction peak of the spectra. The pristine CeO2 showed a lattice parameter value of 5.408 Å, which was close to the average value of a stoichiometric CeO2 (5.4112 Å). The La-doped CeO2 samples showed cubic lattice parameter values (5.405–5.408 Å) similar to the pristine CeO2. BET surface areas of the pristine CeO2 and La-doped CeO2 samples are summarized in Table 1. The pristine CeO2 NPs showed a moderate surface area of 31.8 m2 g−1. La-doped CeO2 samples exhibited a gradual decrease in their surface areas with an increase in La loading amount. The average surface area of 5 wt.% La-CeO2 NPs was 19.8 m2 g−1, which was about 34% lower than the initial surface area of the pristine CeO2 (28.5 m2 g−1).
Figure 2a shows XPS spectra of the pristine CeO2 and La-doped CeO2 NPs at different La loading levels. La-doped CeO2 samples showed typical peaks of lanthanum oxides at 3d binding energy levels [30]. Table 2 summarizes surface atomic contents of La on samples measured by XPS analysis. The results showed incorporation of La on the sample surface at various loading levels as prepared accordingly. Figure 2b shows deconvoluted XPS spectra of the samples at Ce 3d binding energy level characterized for quantifying Ce3+ and Ce4+ concentrations on the pristine CeO2 and La−CeO2 NPs. The relative Ce3+ concentration was defined by the Ce3+/(Ce4+ + Ce3+) atomic ratio on the sample surface. As shown in Table 2, the relative Ce3+ concentration slightly decreased as the La loading on the CeO2 NPs increased.
Figure 3a shows Raman spectra of the pristine CeO2 and La-doped CeO2 NPs at various La loading levels. Samples exhibited the typical characteristic Raman peak of CeO2 at 460 cm−1, corresponding to the symmetrical stretching mode of the Ce–O8 vibrational unit [23,31]. Oxygen vacancy concentrations of samples were estimated from broadening of the peak at 460 cm−1 [25,32]. As shown in Figure 3b), oxygen vacancy concentrations of the pristine CeO2 increased by approximately 143–271% with the La doping treatment.

2.2. Radical Scavenging Properties of Pristine CeO2 and La-Doped CeO2 NPs

Radical scavenging properties of the pristine CeO2 and La-doped CeO2 NPs were characterized by conducting Fenton’s reaction with Rhodamine-B (RhB) in a solution. Briefly, reactive hydroxyl radicals generated by Fenton’s reactions (Equation (1)) degraded RhB molecules (Equation (3)) in the solution:
H 2 O 2 + Fe 2 + HO + HO + Fe 3 +
Fe 3 + + H 2 O 2 Fe 2 + + HOO + H +
HO + RhB Degradation   of   RhB  
Therefore, scavenging of reactive radicals by a radical scavenger could reduce the decay of RhB concentration in the solution, which could be used to compare radical scavenging properties of materials under the same reaction condition [33].
Figure 4a shows typical UV–Vis spectra of Fenton’s solution taken every 10 min for 1 h, with the addition of pristine CeO2 or La-doped CeO2 NPs (25 mg of the radical scavenger sample per 1 mL Fenton’s solution). All spectra displayed a peak at 550 nm attributed to light absorption by RhB in the solution. However, the peak intensity attenuated over time, indicating a decrease in RhB concentration due to degradation caused by hydroxyl radicals generated by Fenton’s reaction. Figure 4b shows relative RhB concentration (Ct/Ci) in the Fenton’s solution over time, where Ct is the RhB concentration at the given time and Ci is the initial RhB concentration in the solution. The results showed a significant and rapid decrease of RhB concentration in the blank solution. The relative RhB concentration in the solution was less than 10% after 60 min. In comparison, the decay of RhB concentration in the presence of pristine CeO2 NPs was much less, retaining about 40% of the initial RhB concentration after completion of the Fenton’s reaction. Radical scavenging properties of La-doped CeO2 NPs depended significantly on the La loading amount. The 5 wt.% La-CeO2 NPs exhibited a higher radical scavenging property than the pristine CeO2, retaining about 50% of the initial RhB concentration after completion of the reaction. The 2 wt.% La-CeO2 NPs showed similar radical scavenging properties to the pristine CeO2 NPs. The 10 wt.% La-CeO2 NPs displayed much smaller radical scavenging properties than the pristine CeO2, suggesting detrimental effects of excessive La loading on the radical scavenging properties of CeO2 NPs. Figure 4c displays photographs of the solution taken before and after completion of the Fenton’s reaction. The results clearly showed that the color of the blank solution significantly changed from a thick red to a pale pink due to degradation of RhB in the solution. Comparatively, the solution containing the 5 wt.% La-CeO2 NPs exhibited a thick red color similar to that of the initial solution, demonstrating the significant radical scavenging properties of La-CeO2 NPs.

2.3. Physicochemical Properties of N-Doped CeO2 NPs

Atomic N doping of the pristine CeO2 was conducted with the urea thermolysis method. Scheme 1 briefly shows the procedure for atomic N doping on the pristine CeO2 NPs. As described in the experimental section, N doping was conducted by urea thermolysis via two different approaches: (i) mixing pristine CeO2 NPs with urea powder (the dry method); and (ii) impregnating pristine CeO2 NPs with aqueous solution of urea (the wet method). These samples were subsequently treated at 873 K under two different conditions: inert N2 or air atmosphere. Figure 5a shows photographs of the pristine CeO2, N-CeO2-D-I, N-CeO2-W-I, and N-CeO2-W-A samples. The pristine CeO2 NPs showed a typical bright yellow color. The N-CeO2-D-I and N-CeO2-W-I samples prepared by the heat treatment in an inert N2 atmosphere exhibited dark brown colors. However, samples prepared in an air atmosphere (N-CeO2-W-A) showed a bright yellow color similar to the pristine CeO2. Figure 5b shows the XRD patterns of the pristine CeO2 and N-doped CeO2 samples. As discussed previously, the pristine CeO2 NPs exhibited the typical diffraction pattern of FCC crystal structure of CeO2. The N-doped CeO2 samples also exhibited the typical diffraction pattern of CeO2, retaining the bulk FCC crystal structure of the pristine CeO2. Table 3 shows cubic lattice parameter values and BET surface areas of the pristine CeO2 and N-doped CeO2 NPs prepared by various methods. The N-doped CeO2 samples exhibited similar cubic lattice parameter values (5.406–5.408 Å) to the pristine CeO2 NPs (5.4112 Å). Surface areas of the N-doped CeO2 samples showed considerable differences depending on the N doping method. The N-CeO2-W-I NPs prepared by the urea impregnation followed by thermolysis in N2 atmosphere showed a moderate surface area (22.3 m2 g−1), which was about 21% smaller than that of the pristine CeO2 (28.5 m2 g−1). The surface area of the N-CeO2-D-I prepared by the dry method was far smaller (12.2 m2 g−1), which was about 57% lower than that of the pristine CeO2. Differently, the surface area of the N-CeO2-W-A NPs (30.0 m2 g−1) prepared in the air atmosphere was similar to that of the pristine CeO2 NPs.
Figure 6a shows the deconvoluted XPS spectra of the pristine CeO2 and N-doped CeO2 samples at the N 1s binding energy level. It was notable that N doping did not occur on the N-CeO2-W-A NPs prepared by heat treatment in an air atmosphere. Differently, N-CeO2-D-I and the N-CeO2-W-I samples prepared in an inert N2 atmosphere exhibited two deconvoluted peaks centered at binding energies of 396 and 399 eV, respectively, that could be attributed to Ce-N and C-O bonding configurations. Table 4 shows surface N atomic contents and relative bonding configuration obtained from XPS results. The N-CeO2-D-I NPs exhibited a higher N atomic content (7.51%) than the CeO2-W-I NPs (4.45%), with a high Ce-N bonding configuration (51.6%). Figure 6b shows the deconvoluted XPS spectra of samples at Ce 3d binding energy level for quantifying Ce3+ and Ce4+ concentrations. The results indicated that Ce3+ concentrations on N-doped CeO2 samples (24.7–28%) were similar to or slightly higher than that on the pristine CeO2 (25.4%).
Figure 7a shows Raman spectra of the pristine CeO2 and N-doped CeO2 samples. Oxygen vacancy concentrations on samples were estimated from peak broadening [25,32]. The results are summarized in Figure 7b. Surface oxygen vacancy concentrations on the N-CeO2-D-I and the N-CeO2-W-I samples prepared by urea thermolysis in an inert N2 atmosphere were about 4-fold and 8-fold higher than that on the pristine CeO2. However, the N-CeO2-W-A sample prepared in an air atmosphere showed practically the same oxygen vacancy concentration as the pristine CeO2. Similar oxygen vacancy concentrations of the pristine CeO2 and the N-CeO2-W-A were reasonable considering that N doping did not occur on the N-CeO2-W-A NPs as shown in XPS results. Collectively, these results clearly indicate that N doping of CeO2 NPs could lead to a significant increase in surface oxygen vacancy concentration on the surface.

2.4. Radical Scavenging Properties of N-Doped CeO2 NPs

Radical scavenging properties of the N-doped CeO2 NPs are characterized by applying the Fenton’s reaction as previously described. Figure 8a compares the UV–Vis spectra of Fenton’s solution taken every 10 min for 1 h after addition of a radical scavenger (pristine CeO2, N-CeO2-D-I, N-CeO2-W-I, or N-CeO2-W-A NPs, 25 mg of radical scavenger per 1 mL Fenton’s solution). Figure 8b shows changes of RhB concentration (Ct) normalized by its initial concentration (Ci) in the solution measured over time. The results showed that normalized RhB concentration (Ct/Ci) decreased over time due to degradation of RhB caused by hydroxyl radicals generated by Fenton’s reaction. These results clearly showed that RhB was retained at much higher concentrations in solutions in the presence of N-doped CeO2 samples (N-CeO2-D-I and the N-CeO2-W-I) than the pristine CeO2 NPs, reflecting significant enhancements to the radical scavenging properties of CeO2 NPs by N doping on the surface. In comparison, N-CeO2-W-A NPs prepared by heat treatment in air atmosphere showed practically the same radical scavenging property as pristine CeO2 NPs. As discussed previously, N doping did not occur on the N-CeO2-W-A sample and its surface area was also practically the same as the pristine CeO2. These results support that N doping of the CeO2 NPs is the primary reason for the enhancement of the radical scavenging properties of N-doped CeO2 NPs. These results collectively suggest that the doping method has significant effects on radical scavenging properties of N-doped CeO2 NPs. The wet method conducted with urea impregnation and subsequent heat treatment in an inert atmosphere gave rise to higher radical scavenging properties than the dry method conducted by heat treatment of a mixture of urea powder and CeO2 NPs. Figure 8c compares photographs of the initial solution and the solutions after completion of Fenton’s reaction. The results clearly show that the solutions containing N-doped CeO2 samples maintained the initial thick red color, confirming the high radical scavenging properties of the N-doped CeO2 NPs.

2.5. Comparison of Radical Scavenging Properties of La- and N-Doped CeO2 NPs

Radical scavenging properties of pristine CeO2, La-doped CeO2, and N-doped CeO2 NPs were compared based on the RhB concentration remained in the solution after Fenton’s reaction for 60 min. Figure 9a shows relative RhB concentrations in the blank and the radical scavenger containing solutions after completion of Fenton’s reaction. For comparison, the 5 wt.% La-CeO2 and the N-CeO2-W-I NPs were chosen among the La-doped and the N-doped CeO2 samples because these two samples exhibited the best radical scavenging properties in each group of samples. The results showed that RhB concentration fell rapidly to 10% in the blank solution after 60 min. In the presence of pristine CeO2 NPs, about 38% of the initial RhB concentration was retained after the test. Comparatively, the RhB concentration was retained at significantly higher concentrations with La-CeO2 NPs (50%) and N-CeO2 NPs (80%) than with pristine CeO2 NPs, indicating a considerable enhancement to radical scavenging activities by La and the N atomic doping on CeO2 NPs.
Intrinsic radical scavenging properties of pristine CeO2, La-CeO2, and N-CeO2 NPs were estimated by normalization of radical scavenging properties with their BET surface areas. Figure 9b shows the surface area-normalized radical scavenging properties of samples. The results revealed that intrinsic radical scavenging properties of samples had the following decreasing order: N-CeO2 > La-CeO2 > pristine CeO2 NPs. The intrinsic radical scavenging property of the N-CeO2 NPs was about 3.6 times greater than that of the pristine CeO2 and 1.6 times greater than that of the La-CeO2 NPs. These results suggest that atomic N doping is highly effective for enhancing the radical scavenging properties of CeO2 NPs.
The radical scavenging properties of pristine CeO2, La-doped CeO2, and N-doped CeO2 NPs were correlated with surface oxygen vacancy concentrations on these samples. Figure 9c shows the RhB concentrations retained after 60 min of Fenton’s reaction in the solution with respect to surface oxygen vacancy concentrations of the radical scavenger samples. The collective structure-property correlation analysis showed that RhB concentration retained in the solution increased with surface oxygen vacancy concentration of the radical scavengers. These results suggest that oxygen vacancies play a primary role in regenerative radical scavenging activities of these samples. Comparative studies revealed that atomic N doping of CeO2 NPs was highly effective for increasing surface oxygen vacancies on CeO2 NPs, which resulted in large enhancements in the intrinsic radical scavenging properties of CeO2 NPs.

3. Experimental

3.1. Atomic Hetero-Metal (La) Doping on CeO2 NPs

Scheme 1 describes the atomic doping of metal cation (La) and non-mental anion (N) on pristine CeO2 NPs prepared in this work. La doping of CeO2 NPs was conducted with the incipient wetness impregnation method. Briefly, an aqueous solution of La(NO3)3 was prepared by dissolving 2.6 g of La(NO3)3 (Alfa Aesar, 99%) in 5 mL of D.I. H2O. This 1.58 mL La(NO3)3 solution was then mixed dropwise with 5 g of pristine CeO2 powder (Sigma-Aldrich, 99.9%) for uniform impregnation (5 wt.% La loading). The resultant was then dried at 353 K in a convection oven for 12 h. The sample was placed in a ceramic boat and placed inside a horizontal tubular quartz reactor (I.D. = 50 mm, L = 1,200 mm) equipped with a temperature-controlled electric furnace. Both ends of the reactor were capped gas tight, and air (99.999%, Deokyang, Republic of Korea) was fed to the reactor (100 mL min−1). While maintaining the air flow through the reactor, the temperature was raised (ramp = 10 K min−1) and held at 873 K for 4 h. The reactor was then cooled to room temperature under the air flow. The loading amount of La on the sample was varied (2, 5, and 10 wt.%) by changing the concentration of La(NO3)3 in the impregnation solution.

3.2. Atomic Non-Metallic Heteroatom (N) Doping on CeO2 NPs

Atomic N doping on CeO2 NPs was conducted by urea thermolysis with two different methods: (i) dry urea thermolysis and ii) wet urea impregnation followed by thermolysis. In the dry method, CeO2 NPs were physically mixed with solid urea powder (98%, Alfa Aesar) at a fixed weight ratio (CeO2/urea = 4) and placed in a ceramic boat. The sample boat was placed inside a horizontal tubular quartz reactor. Both ends of the reactor were capped gas tight. The inside reactor was purged by flowing an inert N2 gas (100 mL min−1, 99.999%, Deokyang, Republic of Korea) at 298 K for 1 h. The reactor temperature was raised (ramp = 10 K min−1) and held at 873 K for 2 h, flowing N2 gas through the reactor. Finally, the reactor was cooled to room temperature under the N2 gas flow. The resulting sample is denoted as N−CeO2-D-I, where D and I indicate the dry method and inert atmosphere, respectively.
In the wet impregnation method, CeO2 NPs were impregnated with an aqueous solution of urea. The impregnation amount of urea on the CeO2 was fixed at 3 wt.%. Urea solution was prepared by dissolving 2.1 g of urea powder in 5 mL of D.I. H2O. Then, 1.58 mL of urea solution was dropwise mixed with 5 g CeO2 NPs for uniform impregnation. The urea impregnated CeO2 was then dried in a convection oven at 353 K for 12 h. The sample was placed in a ceramic boat and loaded in the tubular reactor. Subsequent heat treatment procedures were the same as those described above in the dry method except that it was conducted in two different atmospheres: inert N2 and air. The sample was treated flowing N2 gas (99.999%, Deokyang Korea) or air (99.999%, Deokyang Korea) at 873 K for 4 h. Resulting samples are denoted as N-CeO2-W-I and N-CeO2-W-A, respectively, where W and A indicated wet method and air atmosphere, respectively.

3.3. Characterizations

Surface atomic compositions of the pristine CeO2, La-doped CeO2, and N-doped CeO2 NPs were characterized by X-ray photoelectron spectroscopy (XPS, Theta Probe AR-XPS, Thermo Fisher Scientific, Waltham, MA, U.S.A.). The atomic contents of Ce3+ and Ce4+ on the samples were obtained by deconvolution of corresponding characteristic peaks at the Ce 3d binding energy level according to previously reported methods in the literature [25,34]. Doping amount and structural configuration of La and the N species on La-doped and N-doped CeO2 NPs were obtained by deconvolution of the La 3d and the N 1s peak of the XPS spectra. The crystal structure of each sample was characterized by X-ray diffraction (XRD, SmartLab, Rigaku, Tokyo, Japan) with monochromic Cu–Kα radiation operated at 3 kW (scan rate = 0.05° min−1). Defects on the sample surfaces were characterized by Raman spectroscopy (Vertex 80, Bruker, Billerica, MA, U.S.A.) conducted at 532 nm laser wavelength. The concentration of surface oxygen vacancies was calculated from the full width at half maximum (FWHM) of the characteristic Raman peak according to the method reported in the literature [25,32]. The specific surface area of each sample was obtained with the Brunauer–Emmett–Teller (BET) method using an N2 adsorption–desorption isotherm obtained in a volumetric unit (Tristar, Micromeritics, Norcross, GA, USA).

3.4. Measurement of Radical Scavenging Properties

Radical scavenging properties of the pristine CeO2, La-doped CeO2, and N-doped CeO2 NPs were characterized by conducting Fenton’s reaction after adding Rhodamine B (RhB) and radical scavenger into the solution. Added RhB was used as a marker to estimate radical scavenging properties of the scavenger materials added into the solution. First, an aqueous solution was prepared by dissolving 0.6 mg RhB (Alfa Aesar) and 3 mg FeSO4 (98.5%, Junsei Chem., Tokyo, Japan) in 100 mL D.I. H2O. Then, 50 mg of the radical scavenger was added to 2 mL of the solution with stirring. Separately, an aqueous solution of H2O2 (0.08 vol.%) was prepared by mixing H2O2 (30 wt.%, DaeJung Chem., Siheung, Republic of Korea) and D.I. H2O. Finally, a fixed amount (3 μL) of the aqueous H2O2 solution was added dropwise to the first solution containing FeSO4 and stirred for 10 min. Subsequently, light adsorption spectra of RhB in the mixed solution were obtained using a UV–Vis spectrometer (Mega-800, Sinco, Daejeon, Republic of Korea) at a wavelength of 300 to 700 nm. Dropwise H2O2 addition, stirring, and UV–Vis measurements were repeated at 10 min intervals for 1 h to measure changes of RhB concentration in the solution over time. Radical scavenging properties of materials added into the solution were estimated from the extent of RhB concentration decay in the solution in comparison with that in the blank Fenton’s solution measured without adding radical scavengers.

4. Conclusions

The effects of metal cation and non-metal anion doping on CeO2 NPs for radical scavenging properties were studied. La-doped CeO2 NPs (8 atomic % La) showed about 2.0 times greater intrinsic radical scavenging activities than the pristine CeO2 NPs. In comparison, N-doped CeO2 NPs (4.5 atomic-% N) prepared by a wet urea impregnation with thermolysis method exhibited about 2.8 times higher intrinsic radical scavenging activities than the pristine CeO2, suggesting the high effectiveness of atomic N doping for enhancing radical scavenging properties. A collective structure–property correlation analysis conducted with pristine CeO2, La-doped CeO2, and N-doped CeO2 NPs suggested that surface oxygen vacancies provide relevant active sites for regenerative radical scavenging. Surface oxygen vacancy concentrations could be effectively introduced on CeO2 NPs by heteroatom doping, which gave rise to significant enhancements of radical scavenging properties of CeO2 NPs.

Author Contributions

Conceptualization, D.L. and E.-S.L.; investigation, J.P., S.H. and J.-Y.J.; formal analysis, J.P. and S.H.; writing-original draft preparation, J.P.; supervision, D.L. All authors have read and agreed to the published version of the manuscript.

Funding

E.-S. Lee and J.-Y. Jyoung acknowledge financial support through a grant (#P0017349) from the Institute for Advancement of Technology funded by The Ministry of Trade, Industry and Energy of Republic of Korea.

Data Availability Statement

Data is contained within the article.

Acknowledgments

D. Lee acknowledges the support from the University of Seoul by the 2021 Research Fund.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Rui, Z.; Liu, J. Understanding of free radical scavengers used in highly durable proton exchange membranes. Prog. Nat. Sci. 2020, 30, 732–742. [Google Scholar] [CrossRef]
  2. Prabhakaran, V.; Arges, C.G.; Ramani, V. Investigation of polymer electrolyte membrane chemical degradation and degradation mitigation using in-situ fluorescence spectroscopy. Proc. Natl. Acad. Sci. USA 2012, 109, 1029–1034. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Wang, L.; Advani, S.G.; Prasad, A.K. Degradation reduction of polymer electrolyte membranes using CeO2 as a free-radical scavenger in catalyst layer. Electrochim. Acta 2013, 109, 775–780. [Google Scholar] [CrossRef]
  4. Pearman, B.P.; Mohajeri, N.; Slattery, D.K.; Hampton, M.D.; Seal, S.; Cullen, D.A. The chemical behavior and degradation mitigation effect of cerium oxide nanoparticles in perfluorosulfonic acid polymer electrolyte membranes. Polym. Degrad. Stab. 2013, 98, 1766–1772. [Google Scholar] [CrossRef]
  5. Weissbach, T.; Peckham, T.J.; Holdcroft, S. CeO2, ZrO2 and YSZ as mitigating additives against degradation of proton exchange membranes by free radicals. J. Membr. Sci. 2016, 498, 94–104. [Google Scholar] [CrossRef]
  6. Pearman, B.P.; Mohajeri, N.M.; Brooker, R.P.; Rodgers, M.P.; Slattery, D.K.; Hampton, M.D.; Cullen, M.D.; Seal, S. The degradation mitigation effect of cerium oxide in polymer electrolyte membranes in extended fuel cell durability tests. J. Power Sources 2013, 225, 75–83. [Google Scholar] [CrossRef]
  7. Breitwieser, M.; Klose, C.; Hartmann, A.; Buchler, A.; Klingele, M.; Vierrath, S.; Zengerle, R.; Thiele, S. Cerium Oxide Decorated Polymer Nanofibers as Effective Membrane Reinforcement for Durable, High-Performance Fuel Cells. Adv. Energy Mater. 2017, 7, 1602100. [Google Scholar] [CrossRef]
  8. Zhiyan, R.; Qingbing, L.; Youxiu, H.; Rui, D.; Jia, L.; Jia, L.; Jianguo, L. Ceria nanorods as highly stable free radical scavengers for highly durable proton exchange membranes. RSC Adv. 2021, 11, 32012–32021. [Google Scholar] [CrossRef] [PubMed]
  9. Gubler, L.; Koppenol, W.H. Kinetic Simulation of the Chemical Stabilization Mechanism in Fuel Cell Membranes Using Cerium and Manganese Redox Couples. J. Electrochem. Soc. 2012, 159, B211–B218. [Google Scholar] [CrossRef]
  10. Taghizadeh, M.T.; Vatanparast, M. Ultrasonic-assisted synthesis of ZrO2 nanoparticles and their application to improve the chemical stability of Nafion membrane in proton exchange membrane (PEM) fuel cells. J. Colloid Interface Sci. 2016, 483, 1–10. [Google Scholar] [CrossRef] [PubMed]
  11. D’Urso, C.; Oldani, C.; Baglio, V.; Merlo, L.; Arico, A.S. Immobilized transition metal-based radical scavengers and their effect on durability of Aquivion® perfluorosulfonic acid membranes. J. Power Sources 2016, 301, 317–325. [Google Scholar] [CrossRef]
  12. Hiroki, A.; LaVerne, J.A. Decomposition of Hydrogen Peroxide at Water−Ceramic Oxide Interfaces. J. Phys. Chem. B 2015, 109, 3364–3370. [Google Scholar] [CrossRef] [PubMed]
  13. Kwon, H.J.; Shin, K.; Soh, M.; Chang, H.; Kim, J.; Lee, J.; Ko, G.; Kim, B.H.; Kim, D.; Hyeon, T. Large-Scale Synthesis and Medical Applications of Uniform-Sized Metal Oxide Nanoparticles. Adv. Mater. 2018, 30, 1704290. [Google Scholar] [CrossRef]
  14. Lawler, R.; Cho, J.; Ham, H.C.; Ju, H.; Lee, S.W.; Kim, J.Y.; Choi, J.I.; Jang, S.S. CeO2(111) Surface with Oxygen Vacancy for Radical Scavenging: A Density Functional Theory Approach. J. Phys. Chem. C 2020, 124, 20950–20959. [Google Scholar] [CrossRef]
  15. Bedar, A.; Singh, B.G.; Tewari, P.K.; Bindal, R.C.; Kar, S. Kinetics studies on free radical scavenging property of ceria in polysulfone–ceria radiation resistant mixed-matrix membrane. Int. J. Chem. React. Eng. 2021, 19, 779–785. [Google Scholar] [CrossRef]
  16. Fernandez-Garcia, S.; Jiang, L.; Tinoco, M.; Hungria, A.B.; Han, J.; Blanco, G.; Calvino, J.J.; Chen, X. Enhanced Hydroxyl Radical Scavenging Activity by Doping Lanthanum in Ceria Nanocubes. J. Phys. Chem. C 2016, 120, 1891–1901. [Google Scholar] [CrossRef]
  17. Vinothkumar, G.; Rengaraj, S.; Arunkumar, P.; Cha, S.W.; Babu, K.S. Ionic Radii and Concentration Dependency of RE3+ (Eu3+, Nd3+, Pr3+, and La3+)-Doped Cerium Oxide Nanoparticles for Enhanced Multienzyme-Mimetic and Hydroxyl Radical Scavenging Activity. J. Phys. Chem. C 2019, 123, 541–553. [Google Scholar] [CrossRef]
  18. Loche, D.; Morgan, L.M.; Casu, A.; Mountjoy, G.; O’Regan, C.; Corrias, A.; Falqui, A. Determining the maximum lanthanum incorporation in the fluorite structure of La-dopedceria nanocubes for enhanced redox ability. RSC Adv. 2019, 9, 6745–6751. [Google Scholar] [CrossRef] [Green Version]
  19. Hernandez-Castillo, Y.; Garcia-Hernandez, M.; Lopez-Marure, A.; Luna-Dominguz, J.H.; Lopez-Camacho, P.Y.; de Jesus Morales-amirez, A. Antioxidant activity of cerium oxide as a function of europium doped content. Ceram. Int. 2018, 45, 2303–2308. [Google Scholar] [CrossRef]
  20. Baker, A.M.; Stewart, S.M.; Ramaiyan, K.P.; Banham, D.; Ye, S.; Garzon, F.; Mukundan, R.; Borup, R.L. Doped Ceria Nanoparticles with Reduced Solubility and Improved Peroxide Decomposition Activity for PEM Fuel Cells. J. Electrochem. Soc. 2021, 168, 024507. [Google Scholar] [CrossRef]
  21. Li, Z.; Yang, W.; Xie, L.; Li, Y.; Liu, Y.; Sun, Y.; Bu, Y.; Mi, X.; Zhan, S.; Hu, W. Prominent role of oxygen vacancy for superoxide radical and hydroxyl radical formation to promote electro-Fenton like reaction by W-doped CeO2 composites. Appl. Surf. Sci. 2021, 549, 149262. [Google Scholar] [CrossRef]
  22. Soh, M.; Kang, D.-W.; Jeong, H.-G.; Kim, D.; Kim, D.Y.; Yang, W.; Song, C.; Baik, S.; Choi, I.-Y.; Ki, S.-K.; et al. Ceria–Zirconia Nanoparticles as an Enhanced Multi-Antioxidant for Sepsis Treatment. Angew. Chem. Int. Ed. 2017, 129, 11557–11561. [Google Scholar] [CrossRef]
  23. Trogadas, P.; Parrondo, J.; Ramani, V. CeO2 Surface Oxygen Vacancy Concentration Governs in Situ Free Radical Scavenging Efficacy in Polymer Electrolytes. ACS Appl. Mater. Interfaces 2012, 4, 5098–5102. [Google Scholar] [CrossRef] [PubMed]
  24. Baker, A.M.; Williams, S.T.D.; Mukundan, R.; Spernjak, D.; Advani, S.G.; Prasad, A.K.; Borup, R.L. Zr-doped ceria additives for enhanced PEM fuel cell durability and radical scavenger stability. J. Mater. Chem. A 2017, 5, 15073–15079. [Google Scholar] [CrossRef]
  25. Prabhakaran, V.; Ramani, V. Structurally-Tuned Nitrogen-Doped Cerium Oxide Exhibits. Exceptional Regenerative Free Radical Scavenging Activity in Polymer Electrolytes. J. Electrochem. Soc. 2014, 161, F1–F9. [Google Scholar] [CrossRef]
  26. Asahi, R.; Morikawa, T.; Ohwaki, T.; Aoki, K.; Taga, Y. Visible-Light Photocatalysis in Nitrogen-Doped Titanium Oxides. Science 2001, 293, 269–271. [Google Scholar] [CrossRef]
  27. Jorge, A.B.; Sakatani, Y.; Boissiere, 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]
  28. Sun, D.; Gu, M.; Li, R.; Yin, S.; Song, X.; Zhao, B.; Li, C.; Li, J.; Feng, Z.; Sato, T. Effects of nitrogen content in monocrystalline nano-CeO2 on the degradation of dye in indoor lighting. Appl. Surf. Sci. 2013, 280, 693–697. [Google Scholar] [CrossRef]
  29. Cong, Q.; Chen, L.; Wang, X.; Ma, H.; Zhao, J.; Li, S.; Hou, Y.; Li, W. Promotional effect of nitrogen-doping on a ceria unary oxide catalyst with rich oxygen vacancies for selective catalytic reduction of NO with NH3. Chem. Eng. J. 2020, 379, 122302. [Google Scholar] [CrossRef]
  30. Sundingm, M.F.; Hadidi, K.; Diplas, S.; Løvvik, O.M.; Nordy, T.E.; Gunnæs, A.E. XPS characterisation of in situ treated lanthanum oxide and hydroxide using tailored charge referencing and peak fitting procedures. J. Electron Spectros. Relat. Phenomena 2011, 184, 399–409. [Google Scholar] [CrossRef]
  31. Wang, X.; Jiang, Z.; Zheng, B.; Xie, Z.; Zheng, L. Synthesis and shape-dependent catalytic properties of CeO2 nanocubes and truncated octahedra. CrystEngComm 2012, 14, 7579–7582. [Google Scholar] [CrossRef]
  32. Kosacki, I.; Suzuki, T.; Anderson, H.U.; Colomban, P. Raman scattering and lattice defects in nanocrystalline CeO2 thin films. Solid State Ion 2002, 149, 99–105. [Google Scholar] [CrossRef]
  33. Yu, F.; Xu, D.; Lei, R.; Li, N.; Li, K. Free-radical scavenging capacity using the Fenton reaction with rhodamine B as the spectrophotometric indicator. J. Agric. Food Chem. 2008, 56, 730–735. [Google Scholar] [CrossRef] [PubMed]
  34. Zhang, F.; Wang, P.; Koberstein, J.; Khalid, S.; Chan, S.-W. Cerium oxidation state in ceria nanoparticles studied with X-ray photoelectron spectroscopy and absorption near edge spectroscopy. Surf. Sci. 2004, 563, 74–82. [Google Scholar] [CrossRef]
Figure 1. (a) photographs of the pristine CeO2 and La-doped CeO2 NPs at different La loading amount, (b) XRD pattern of the pristine CeO2 and the La-doped CeO2 NPs.
Figure 1. (a) photographs of the pristine CeO2 and La-doped CeO2 NPs at different La loading amount, (b) XRD pattern of the pristine CeO2 and the La-doped CeO2 NPs.
Catalysts 13 00572 g001
Figure 2. XPS spectra of the pristine CeO2 and the La-doped CeO2 NPs: (a) La 3d, (b) Ce 3d binding energy level.
Figure 2. XPS spectra of the pristine CeO2 and the La-doped CeO2 NPs: (a) La 3d, (b) Ce 3d binding energy level.
Catalysts 13 00572 g002
Figure 3. (a) Raman spectra of the pristine CeO2 and the La-doped CeO2 NPs. (b) Oxygen vacancy concentration on the samples estimated by the Raman spectroscopy.
Figure 3. (a) Raman spectra of the pristine CeO2 and the La-doped CeO2 NPs. (b) Oxygen vacancy concentration on the samples estimated by the Raman spectroscopy.
Catalysts 13 00572 g003
Figure 4. (a) UV spectra of the Fenton’s solutions with RhB and the radical scavengers added. (b) Relative RhB concentration in the solution over time after adding radical scavenger in the solution. (c) Photograph of the solution before and after the radical scavenging test.
Figure 4. (a) UV spectra of the Fenton’s solutions with RhB and the radical scavengers added. (b) Relative RhB concentration in the solution over time after adding radical scavenger in the solution. (c) Photograph of the solution before and after the radical scavenging test.
Catalysts 13 00572 g004
Scheme 1. Heteroatom doping on CeO2 nanoparticles: (a) cationic La doping, (b) anionic N doping.
Scheme 1. Heteroatom doping on CeO2 nanoparticles: (a) cationic La doping, (b) anionic N doping.
Catalysts 13 00572 sch001
Figure 5. (a) Photographs of the pristine CeO2 and the N-doped CeO2 NPs. (b) XRD spectra of the pristine CeO2 and the N-doped CeO2 prepared by various methods.
Figure 5. (a) Photographs of the pristine CeO2 and the N-doped CeO2 NPs. (b) XRD spectra of the pristine CeO2 and the N-doped CeO2 prepared by various methods.
Catalysts 13 00572 g005
Figure 6. XPS spectra of the pristine CeO2 and the N-doped CeO2 NPs: (a) N 1s, (b) Ce 3d binding energy level.
Figure 6. XPS spectra of the pristine CeO2 and the N-doped CeO2 NPs: (a) N 1s, (b) Ce 3d binding energy level.
Catalysts 13 00572 g006
Figure 7. (a) Raman spectra of the pristine CeO2 and the La-doped CeO2 NPs. (b) Oxygen vacancy concentration on the samples estimated by the Raman spectroscopy.
Figure 7. (a) Raman spectra of the pristine CeO2 and the La-doped CeO2 NPs. (b) Oxygen vacancy concentration on the samples estimated by the Raman spectroscopy.
Catalysts 13 00572 g007
Figure 8. (a) UV–Vis spectra of the Fenton’s solutions with RhB and the radical scavengers added. (b) The relative RhB concentration in the solution over time after adding the radical scavenger in the solution. (c) Photograph of the solution before and after the radical scavenging test.
Figure 8. (a) UV–Vis spectra of the Fenton’s solutions with RhB and the radical scavengers added. (b) The relative RhB concentration in the solution over time after adding the radical scavenger in the solution. (c) Photograph of the solution before and after the radical scavenging test.
Catalysts 13 00572 g008
Figure 9. (a) Comparison of radical scavenging properties of the pristine CeO2, La-doped CeO2, and the N-doped CeO2 NPs. (b) Surface area-normalized radical scavenging properties of the samples. (c) Correlation of radical scavenging properties of the pristine CeO2, La-CeO2, and N-CeO2 NPs with their surface oxygen vacancy concentrations.
Figure 9. (a) Comparison of radical scavenging properties of the pristine CeO2, La-doped CeO2, and the N-doped CeO2 NPs. (b) Surface area-normalized radical scavenging properties of the samples. (c) Correlation of radical scavenging properties of the pristine CeO2, La-CeO2, and N-CeO2 NPs with their surface oxygen vacancy concentrations.
Catalysts 13 00572 g009
Table 1. Cubic lattice parameter and BET surface area of the pristine CeO2 and the La-doped CeO2 NPs.
Table 1. Cubic lattice parameter and BET surface area of the pristine CeO2 and the La-doped CeO2 NPs.
SampleCubic Lattice Parameter,
a (Å)
BET Surface Area (m2 g−1)
Pristine CeO25.40828.5
2 wt% La-CeO25.40825.9
5 wt% La-CeO25.40919.8
10 wt% La-CeO25.40513.5
Table 2. Atomic content and the relative Ce concentration on the pristine CeO2 and the La-doped CeO2 NPs.
Table 2. Atomic content and the relative Ce concentration on the pristine CeO2 and the La-doped CeO2 NPs.
SampleAbsolute Atomic % Relative Ce
Concentration (%)
CeCOLaCe3+Ce4+
Pristine CeO218.2330.0451.73-25.474.6
2 wt.%
La-CeO2
19.2517.0759.254.4221.678.4
5 wt.%
La-CeO2
15.1715.8360.998.0022.078.0
10 wt.%
La-CeO2
13.6114.7462.489.1820.579.5
Table 3. Cubic lattice parameter and BET surface area of the pristine CeO2 and the N-doped CeO2 NPs.
Table 3. Cubic lattice parameter and BET surface area of the pristine CeO2 and the N-doped CeO2 NPs.
SampleCubic Lattice Parameter,
a (Å)
BET Surface Area (m2 g−1)
Pristine CeO25.40828.5
N-CeO2-D-I5.40612.2
N-CeO2-W-I5.40822.3
N-CeO2-W-A5.40830.0
Table 4. Atomic content and the relative Ce concentration on the pristine CeO2 and the N-doped CeO2 NPs.
Table 4. Atomic content and the relative Ce concentration on the pristine CeO2 and the N-doped CeO2 NPs.
SampleAbsolute Atomic % Relative Ce Concentration (%)
CeCONCe3+Ce4+
Pristine CeO218.2330.0451.73-25.474.6
N-CeO2-D-I22.0416.7353.727.5128.072.0
N-CeO2-W-A22.3320.4457.22-25.274.8
N-CeO2-W-I24.710.8360.154.4524.775.3
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

Paick, J.; Hong, S.; Jyoung, J.-Y.; Lee, E.-S.; Lee, D. Comparative Studies on Effects of Metal Cation (La) and Non-Metal Anion (N) Doping on CeO2 Nanoparticles for Regenerative Scavenging of Reactive Oxygen Radicals. Catalysts 2023, 13, 572. https://doi.org/10.3390/catal13030572

AMA Style

Paick J, Hong S, Jyoung J-Y, Lee E-S, Lee D. Comparative Studies on Effects of Metal Cation (La) and Non-Metal Anion (N) Doping on CeO2 Nanoparticles for Regenerative Scavenging of Reactive Oxygen Radicals. Catalysts. 2023; 13(3):572. https://doi.org/10.3390/catal13030572

Chicago/Turabian Style

Paick, Jihun, Seunghee Hong, Jy-Young Jyoung, Eun-Sook Lee, and Doohwan Lee. 2023. "Comparative Studies on Effects of Metal Cation (La) and Non-Metal Anion (N) Doping on CeO2 Nanoparticles for Regenerative Scavenging of Reactive Oxygen Radicals" Catalysts 13, no. 3: 572. https://doi.org/10.3390/catal13030572

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