Next Article in Journal
Lanthanide Oxides in Ammonia Synthesis Catalysts: A Comprehensive Review
Next Article in Special Issue
New Insights into the Effect of Ce Doping on the Catalytic Performance and Hydrothermal Stability of Cu-USY Zeolite Catalysts for the Selective Catalytic Reduction of NO with NH3
Previous Article in Journal
A Fluidizable Catalyst for N-Butane Oxidative Dehydrogenation under Oxygen-Free Reaction Conditions
Previous Article in Special Issue
Application of Unconventional External-Field Treatments in Air Pollutants Removal over Zeolite-Based Adsorbents/Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Soot Oxidation Activity of a CuO-Doped CeO2 Catalyst via Acid Etching

1
Key Laboratory of Advanced Materials of Ministry of Education, School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
2
National Engineering Laboratory for Mobile Source Emission Control Technology, China Automotive Technology & Research Center, Tianjin 300300, China
3
State Key Laboratory of New Ceramics and Fine Processing, School of Materials Science and Engineering, Tsinghua University, Beijing 100084, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(12), 1463; https://doi.org/10.3390/catal13121463
Submission received: 26 October 2023 / Revised: 18 November 2023 / Accepted: 21 November 2023 / Published: 23 November 2023

Abstract

:
Copper oxides tend to agglomerate on the surface of CeO2, with a high amount of Cu. In this study, a CeO2 catalyst with a high CuO doping amount was treated with nitric acid to improve its catalytic performance for soot oxidation. The effect of acid etching on the structural properties of the CuO-doped CeO2 catalyst were elucidated. The characterization results indicated that aggregated CuO particles formed over CuCe. The acid etching resulted in a remarkable increase in the surface area of CuCe. Additionally, acid etching promoted the formation of surface-adsorbed oxygen species and oxygen vacancy, and reduced the content of CuOx species with weak interaction with CeO2. The soot temperature-programmed oxidation results show the acid etching of CuCe catalyst could reduce the T50 from 443 to 383 °C. The isothermal reaction results also suggest that acid etching of CuCe leads to an increase in reaction rate from 16.2 to 46.0 μmol min−1 g−1.

1. Introduction

The soot particles emitted by diesel engines can trigger respiratory and cardiovascular diseases, leading to severe health problems [1,2]. The most common after-processing technology for hindering soot discharge is the use of diesel particulate filters (DPFs) to trap particulates. The diesel exhaust temperature (200–500 °C) cannot meet the ignition temperature of soot (>600 °C) [3]. Hence, oxidation catalysts must be applied for catalyzing soot oxidation at lower temperatures. In the past decades, a good diversity of catalysts such as precious metal catalysts [4], CeO2-based oxides [5], and transition metal oxides [6] have been applied to soot removal for achieving DPFs regeneration. Because of the scarcity of precious metals, there is always an increasing impetus to exploit cost-effective and active alternatives to noble metal catalysts. In comparison with noble metals, inexpensive CeO2 and transition metal oxides are promising catalysts for soot combustion [7,8].
It was reported that modification of transition metal oxides via wet chemistry or impregnation methods is an efficient way of improving the redox ability of CeO2 catalysts [9]. Replacing Ce4+ with Fex+, Cox+, and Mnx+ can promote the formation of oxygen vacancies, thereby accelerating the diffusion of surface and bulk oxygen species and achieving improved soot oxidation activities [8,10]. Among transition metals, Cu mainly exists in three valence states of Cu0, Cu+ and Cu2+, which can promote the reduction/oxidation process of ceria via the Mars–van Krevelen (Mv-K) mechanism [11,12]. Liang et al. [13] suggested that Cu can improve the oxidation reaction by activating the O2 molecules on the surface and regulating the oxygen spillover to the ceria support. Due to the unique catalytic performance of Cu, it has gained tremendous attention from researchers. For example, the binary CuO-CeO2 system has been widely investigated in many catalytic reactions, such as soot oxidation [14], CO oxidation [15], and volatile organic compound oxidation [16].
Liang et al. [13] prepared CuO-doped CeO2 catalysts for soot oxidation using the sol-gel method. The results showed that CuOx nanoclusters highly dispersed on the surface of CeO2 are responsible for the activation of lattice oxygen, resulting in the enhancement of the soot oxidation activity of CuO-CeO2 (Tm = 356 and 522 °C in 10% O2 under tight and loose contact conditions). Wang et al. [17] suggested that highly dispersed copper species favor the enhanced redox capability of three-dimensionally ordered macroporous CuO–CeO2 catalysts at low temperatures, thus resulting in an enhanced soot oxidation (Tm = 551 and 427 °C in 5.0% O2 and 2500 ppm NO/5.0% O2 under loose contact conditions). Andana et al. [18] found that CuO/CeO2 catalysts are active for soot oxidation at low temperatures under an NO atmosphere due to the early NO evolution (T50 = 552 and 534 °C in 10% O2 and 550 ppm NO/10% O2 under loose contact conditions). Piumetti et al. [19] fabricated a set of Ce1-xCuxO2 (x ranging from 0.05 to 0.95) oxide catalysts, and evaluated the soot oxidation activity. The results showed that the Ce0.95Cu0.05O2 binary oxide catalyst exhibits the best catalytic activity (Tm = 384 °C in 50% air/50% N2 under tight contact conditions). Cui et al. [20] synthesized CuO-doped CeO2 nanosheets and evaluated their soot oxidation performance. The results demonstrated that CuO doping promoted the formation of oxygen species on the surface of CeO2, and enhanced the soot oxidation activity (T50 = 312 °C in 5% O2 under tight contact conditions; T50 = 360 °C in 1000 ppm/5% O2 under loose contact conditions). However, excessive CuO doping would suppress the catalytic activity of CeO2. Acid etching of catalytic materials can improve structural properties, thus influencing catalytic activity. For example, Zhao et al. [21] synthesized high-surface-area perovskite oxides LaMnO3+δ via an acid-etching method, and favorable NO catalytic oxidation activity was observed on the etched catalysts. Liang et al. [22] also found that nitric acid etching of Ag@CeO2 significantly increased the defect sites of CeO2 and improved the electrocatalytic performance.
In this study, a CuO-doped CeO2 catalyst was prepared via a hydrothermal method, and was then subjected to an acid-etching process. The intrinsic reactivity of the catalysts for soot oxidation was evaluated via temperature programmed oxidation (TPO) and an isothermal reaction model under tight contact conditions. The results showed that acid etching significantly improved the soot oxidation activity of the CuCe catalyst, and the corresponding mechanism of acid etching was discussed.

2. Results and Discussion

2.1. Catalytic Activities for Soot Oxidation

The intrinsic activities of the catalysts were evaluated by conducting a soot–TPO reaction under tight contact conditions. Soot conversion as a function of reaction temperature over the as-prepared catalysts is profiled in Figure 1a, and the corresponding T50 data are listed in Table 1. All the catalysts exhibit a high CO2 selectivity (CO2/COx > 99%). The T50 values of CeO2 and CuCe are 412 and 443 °C, respectively, revealing that a high Cu loading level suppresses soot oxidation over CeO2. However, the CuCe-AE catalyst exhibits a much lower T50 value of 383 °C, displaying remarkably enhanced soot oxidation activity. Such a catalytic performance is normal among the Ce-based catalysts reported in the literature, as compared in Table 2, in which some important reaction parameters are also listed.
To further characterize the intrinsic activities of the as-prepared catalysts, the isothermal reactions were performed under tight contact conditions at 310 °C. At a low conversion rate (<15%), soot oxidation can be considered a surface reaction, while soot oxidation at a high conversion rate is limited by diffusion. Figure 1b shows the soot conversion amount as a function of reaction time. The reaction rate can be obtained from the slope of the lines after linear fitting, and the results are listed in Table 1. Similar to the soot–TPO results, a high Cu loading level can inhibit the reaction rate of CeO2. After the CuCe was subjected to acid etching, the reaction rate increased from 16.2 to 46.0 μmol min−1 g−1, which exceeds that of bare CeO2. These data confirm that the acid etching prompts the intrinsic reactivity of the CuO-doped CeO2 catalyst. The Ea of catalysts was also calculated using the Arrhenius method. The Arrhenius plots of lnr vs. 1/T are shown in Figure 1c, and the corresponding parameters are listed in Table 1. The Ea of the catalysts follows a sequence of CuCe-AE (83 kJ/mol) < CeO2 (96 kJ/mol) < CuCe (110 kJ/mol), which is in accordance with the catalytic performance in Figure 1a,b. Nevertheless, the Ea differences are not large, indicating that they conform with an identical soot oxidation reaction mechanism. To determine the stability of CuCe-AE, four consecutive cycles under the same conditions were conducted, with the results shown in Figure 1d. An obvious increase of 18 °C in the T50 value is observed for the second cycle, which is believed to be related importantly to the decrease in the surface area and sintering of the oxides in the used catalyst. There is no significant deactivation after the third cycle, although a small increase in the T50 (approximately 1–4 °C) occurs, and the catalytic behavior of the cycled CuCe-AE catalyst is still better than those of the fresh CeO2 and CuCe. These observations indicate the potential reusability of CuCe-AE catalyst in soot combustion.

2.2. Physicochemical Characterization of Catalyst

Table 3 summarizes the structural and textural features of the catalysts. As determined by ICP-AES, the Cu content is 20.0 and 2.3 wt.% for CuCe and CuCe-AE, respectively. Obviously, the acid etching significantly reduces the Cu content in the CuO-doped mixed oxides.
The TEM images were collected to observe the morphologies of the catalysts. As shown in Figure 2, CeO2 prepared by alkaline hydrothermal method displays a typical nanorod-like morphology (Figure 2a). Upon the addition of Cu(NO3)2, the CuCe (Figure 2b) and especially CuCe-AE (Figure 2c) exhibit irregular shape. Aggregated CuO particles with a size larger than 15 nm can be clearly observed on CuCe. Figure 2d illustrates the XRD pattern of the as-prepared catalysts. The diffraction peaks at 2θ of 28.5°, 33.1°, 47.5°, 56.3°, 59.1°, 69.4°, 76.7° and 79.1° are indexed to the (111), (200), (220), (311), (222), (400), (331) and (420) planes of CeO2 (JCPDS 34-0394). The half widths of the diffraction peak for CuO-doped catalysts are broader than pure CeO2, indicative of a decrease in the crystallite size of the oxides. The crystallite sizes of the catalysts were calculated via the Scherrer equation, and the results are listed in Table 3. The crystallite sizes of CeO2, CuCe and CuCe-AE are 10.5, 5.8 and 5.5 nm, respectively. In addition, the typical diffraction peaks of CuO particles appear at 2θ of 35.6 and 38.7o for the CuCe catalyst [19], suggesting the agglomeration of CuO at a high doping level. However, the diffraction peaks of CuO are not identified for CuCe-AE, implying that acid etching can remove the agglomerated CuO particles, while the highly dispersed CuOx clusters with a size below the XRD detection limit still exist.
To further identify Cu distribution on the catalyst surface, EDS-mapping analyses of the mixed oxides before and after acid etching were carried out, and the received images are compiled in Figure 3. The markedly agglomerated CuO particles are observed on CuCe, consistent with the XRD results. In these Cu-rich regions, a lower concentration of cerium occurs, indicating the non-uniform distribution of metal oxides. For CuCe-AE, the highly even Cu distribution is consistent with that of Ce, implying the removal of aggregated CuO particles and the reservation of homogeneously distributed mixed oxides.
The N2 sorption isotherms and BJH pore size distribution of the catalysts are shown in Figure 4, and the BET specific surface area and pore volume results are given in Table 3. The N2 sorption isotherms of all the catalysts show a type IV isotherm with H3 hysteresis loop, which is mainly caused by the irregular pores generated via particle accumulation. The measured pore volume is predominantly generated by the gap of primary particles. Notably, the specific surface area of CuCe-AE is about 40 m2/g higher than that of CuCe, suggesting that acid etching can enlarge the specific surface area of CuO-doped CeO2. A similar phenomenon was also observed by Zhao and co-workers [21]. They found that the acid-etching method not only maintained the perovskite-type structure of the LaMnO3+δ catalyst, but also improved its surface area. It is generally accepted that a large specific area facilitates the contact between the soot and catalyst, being therefore beneficial for soot oxidation [23].
XPS analysis was used to disclose the chemical state of Ce, O, and Cu on the catalyst surface. As listed in Table 4, the atom ratios of Cu/Ce on the surface of CuCe and CuCe-AE catalysts are 0.95 and 0.31, respectively. The measured values are not only determined by the actual content of surface elements, but also affected by the dispersion of metal oxides. Figure 5 shows the XPS full spectra of the catalysts and the fine spectrum in the Ce 3d, O 1s and Cu 2p regions, and the deconvolution results are listed in Table 4. In the spectra of Ce 3d in Figure 5b, six peaks assigned to Ce4+ 3d (u1–6: 882.4, 888.6, 898.3, 900.8, 907.5, and 916.7 eV) and two peaks attributed to Ce3+ 3d (v1–2: 884.5 and 902.0 eV) are observed after deconvolution [24,25]. The ratio of Ce3+/Ce4+ decreases after CuO doping. This is due to the fact that the electronegativity of Cu (1.90) is stronger than that of Ce (1.12). After copper doping, electrons transfer from Ce to Cu (Cu2+ + Ce3+→Cu+ + Ce4+), resulting in a decrease in the ratio of Ce3+/Ce4+. A similar phenomenon was observed by Piumetti and co-workers [19]. It is generally accepted that surface oxygen vacancy is normally correlated to the content of Ce3+ [26]. The ratios of Ce3+/Ce4+ for CuCe and CuCe-AE catalysts are 19.9% and 23.8%, respectively, suggesting that acid etching could increase the amount of surface oxygen vacancy on the CuO-doped CeO2. The increased Ce3+/Ce4+ ratio for CuCe-AE leads to an expansion of the ceria crystal cell compared with CuCe.
Figure 5c shows the Cu 2p spectra of CuCe and CuCe-AE. The two main peaks of CuCe at the binding energies of 953 and 933 eV correspond to Cu 2p1/2 orbital and Cu 2p3/2 orbital, respectively, accompanied with two satellite peaks at the binding energies of 942 and 962 eV [27]. This result is similar to the spectrum of CuO [28], which indicates that Cu species mainly exist in the form of CuO. The CuCe-AE catalyst demonstrates weak signals in the Cu 2p region, meaning that the surface Cu content is relatively low.
Two oxygen species can be observed in the O 1s spectra in Figure 5d. The binding energy of 529–530 eV is characteristic of lattice oxygen (denoted as Oβ), and the binding energy in the 531–532 eV region is related to surface oxygen species (including -OH, O 2 and O 2 2 , denoted as Oα) [29,30]. The relative content of surface oxygen species can be roughly estimated by the ratio of Oα/Oβ. The CuO doping results in a decrease in the proportion of Oα/Oβ. This is because a great quantity of agglomerated CuO particles on the surface of CeO2 leads to the surface oxygen species being covered. Notably, acid etching removes these aggregated CuO particles and recovers surface oxygen species to a great extent within the amount of CuO-doped CeO2.
Raman spectroscopy of CeO2 and CuO-doped catalysts was carried out to estimate the relative concentration of oxygen vacancies in CeO2 lattice with an indicative at the wavelength of 532 nm [31]. As displayed in Figure 6a, the as-prepared catalysts exhibit a strong peak at ~456 cm−1 and two weak peaks at ~600 and 1175 cm−1, which can be attributed to the F2g, defect-induced (D), and second-order longitudinal optical (2LO) modes of the fluorite phase, respectively [32,33]. The ratios of ID/IF2g, as a characteristic of bulk oxygen vacancy concentration, are listed in Table 3. In comparison with pure CeO2, CuO-doped CeO2 shows a higher ID/IF2g ratio. The enhanced ratio of bulk oxygen vacancy in CuO-doped CeO2 can be ascribed to the incorporation of smaller copper ions into the CeO2 lattice (rCux+ < rCex+). As a result, CuO doping could promote the formation of bulk oxygen vacancy in CeO2. The ID/IF2g ratios of CeO2, CuCe and CuCe-AE are 0.32, 0.38 and 0.40, respectively. The CuO doping increases the bulk oxygen vacancy concentration in the mixed oxides due to distortion of the ceria crystal cell, and the acid etching creates more bulk oxygen vacancies by removing aggregated CuO particles and strengthening the interaction of highly dispersed CuOx clusters with ceria. As we know, surface oxygen vacancy is essential for soot oxidation reaction [34]; the conclusion that enhanced bulk oxygen vacancy leads to the boosted catalytic activity may be not strict. However, the increase in bulk oxygen vacancy can accelerate the utilization of lattice oxygen CeO2, which is responsible for improved soot oxidation activity of CuCe-AE. When the surface active oxygen is consumed by the soot, the lattice oxygen will diffuse from the bulk of the oxides onto the surface, which is closely associated with the bulk oxygen vacancy [4,35].
The reducibility of the catalysts was investigated using H2-TPR experiments, and the results are shown in Figure 6b. For pure CeO2, two reduction peaks are observed at 510 and 750 °C, corresponding to the reduction of surface oxygen and bulk oxygen of CeO2, respectively [36]. Two intense reduction peaks are observed at 100–200 °C for CuO-doped CeO2, which is attributed to the reduction of CuO [37] and the promoted reduction of CeO2 by spillover effect. The presence of multiple peaks implies the existence of different copper species with varied strength with CeO2. The reduction peak at a low temperature (peak I) corresponds to the reduction of finely dispersed CuOx clusters with strong interaction with CeO2 [38,39]. The reduction peak at a high temperature (peak II) is attributed to the reduction of aggregated CuO particles with interaction with CeO2 [39]. For the CuCe catalyst, an intense reduction peak of CuO interacting with CeO2 at 174 °C can be clearly observed, which means that the substantial agglomerated CuO particles exist on its surface. Upon acid etching, the reduction peaks shift towards higher temperatures, and the content of aggregated CuO particles obviously decreases in peak intensity. The Gaussian fitting of the H2-TPR curves was further conducted for CuO-doped catalysts. The area of peak I of CuCe-AE (7920) is almost identical to that of CuCe (8310), while the area of peak II decreases sharply from 44,800 to 8280 after acid etching. These results demonstrate that the acid etching almost does not affect the finely dispersed CuOx, but removes the aggregated CuO particles significantly.

3. Experimental

3.1. Catalyst Preparation

CuCe mixed oxide catalyst with a Cu/Ce mole ratio of 1:1 was prepared using a hydrothermal method. Briefly, Ce(NO3)3·6H2O (>99.9%, Nanjing Chemical Reagent, Nanjing, China) and Cu(NO3)3·3H2O (>99%, Nanjing Chemical Regent) were added into deionized water. NaOH solution (>98%, Nanjing Chemical Regent) was dropped into the mixed solution slowly and magnetically stirred for 30 min. The pure CeO2 was synthesized by a similar process but without Cu(NO3)3·3H2O addition. The mixed solution was transferred into a Teflon-lined stainless-steel autoclave, which was subsequently heated at 100 °C for 24 h. After cooling to room temperature, the obtained material was washed with deionized water and alcohol for three times and dried at 80 °C overnight. Finally, the powders were calcined at 500 °C in a muffle oven for 2 h.
The CuCe sample was then washed with nitric acid to reduce the Cu content. Briefly, 1 g of CuCe was dispersed in 2 M HNO3 solution and magnetically stirred for 12 h at room temperature. The mixture was filtered and washed with deionized water repeatedly until the filtrate was neutral. Then, the sample was dried at 80 °C in the oven overnight, and the obtained catalyst was denoted CuCe-AE.

3.2. Catalytic Activity Evaluation

Printex-U soot, which was consistently used as the model soot, was purchased from Degussa Company (Germany). In activity tests, the catalyst (50 mg) and soot (5 mg) were ground in a mortar for 5 min to achieve tight contact conditions. Tight contact maximizes the number of contact points between catalyst and soot, allows a rapid screening of catalysts, and facilitates the study of structure–performance relationships, although it is less representative of the real contact conditions that occur in a catalytic trap for diesel particulate removal [40]. To minimize the effect of hot spots caused by the exothermal nature of the reaction, the soot–catalyst mixture was diluted with silica pellets (150 mg), sandwiched by quartz wool, and placed in the vertical quartz reactor. Before each test, the mixture was pretreated in N2 (50 mL/min) at 200 °C for 30 min to eliminate surface adsorbed species. A gas mixture of 10% O2/N2 (80 mL/min) was then introduced, and the mixture was heated at a ramping rate (5 °C/min) to 600 °C. The concentrations of COx (CO2 and CO) in the outlet gas were measured using an infrared spectrometer (Thermo Nicolet iS10) [3,20,24]. The soot combustion efficiency was evaluated in terms of T50, corresponding to the temperature at which 50% of soot conversion was obtained. The selectivity to CO2 was calculated using the following equation.
SCO2 (%) = ACO2/(ACO2 + ACO) × 100%
ACO2 and ACO represent the total peak areas of produced CO2 and CO, respectively.
To determine the apparent activation energy (Ea) of the catalysts, the isothermal reactions were conducted under tight contact with the 10% O2/N2 flow of 150 mL/min. The soot conversion was kept below 15% to ensure the reaction in the kinetic regime, and the apparent activation energy was calculated according to the Arrhenius equation.
ln r = A E a R · 1 T
where A is the pre-exponential factor, R is the ideal gas constant, r is the specific soot oxidation rate (mol min−1 m−2), and T is temperature (K).

3.3. Catalyst Characterization

The Cu content of catalysts was determined via inductively coupled plasma atomic emission spectrometry (ICP-AES, J-A1100, Jarrell-Ash, Waltham, MA, USA). The micromorphology of catalysts was characterized using a transmission electron microscope (TEM) (JEM-2100, JEOL, Tokyo, Japan). X-ray diffraction (XRD) spectra of samples were identified with an Ultima (Rigaku, Tokyo, Japan) machine operated at 40 kV and 40 mA, with a monochromator using Cu Kα radiation (λ = 0.15406 nm). The XRD diffractograms were collected in the 2θ range between 10° and 80°, with a step size of 10°/min. To check the distribution of Cu element in samples, an energy-dispersive spectrometry (EDS) analysis was carried out. The specific surface area of catalysts was measured via N2 sorption at −196 °C using the Brunauer–Emmett–Teller (BET) method on an ASAP 2020 apparatus (Micromeritics, Unterschleissheim, Germany). The sample was vacuum-pretreated at 300 °C for 3 h prior to measurement. The H2 temperature-programmed reduction (H2-TPR) was carried out on a Finesorb-3010 instrument (Fantai, Hangzhou, China) equipped with a thermal conductivity detector (TCD). About 10 mg of the sample was placed in a U-shaped quartz tube and pretreated in N2 (50 mL/min) at 200 °C for 1 h. After cooling to room temperature, 7.0 vol% H2/Ar (50 mL/min) was introduced, and the sample was then heated from 25 to 800 °C with a ramp rate of 10 °C/min. Raman spectroscopy of samples was performed on a LabRAM Aramis (Horiba, Tokyo, Japan) with a 532 nm Ar+ laser beam. The valence states of surface O, Ce and Cu were acquired via X-ray photoelectron spectroscopy (XPS) (ULVAC-PHI5000, Kanagawa, Japan). The binding energy was calibrated with a C1s peak at 284.6 eV.

4. Conclusions

In this work, acid etching was applied to tune the loading amount and form of copper oxide species in ceria-based mixed oxides. An abnormal decrease in the soot oxidation activity of a CuO-doped CeO2 catalyst was observed, with a T50 increase of 31 °C. This was ascribed to the extra high Cu loading (20 wt.%), resulting in the formation of aggregated CuO particles in addition to highly dispersed CuOx clusters. The strong interaction of CuOx clusters with CeO2 promoted the formation of bulk oxygen vacancy in CeO2, but the amount of surface oxygen species decreased obviously due to coverage by agglomerated CuO particles. Correspondingly, the intrinsic reactivity of CuCe was reduced for soot oxidation.
Importantly, these aggregated CuO particles can be removed via acid etching, resulting in a remarkable increase in the catalyst surface area and restoration of surface oxygen vacancies. Meanwhile, highly dispersed CuOx clusters facilitating the generation of bulk oxygen vacancy are mostly reserved in the etched catalyst. As a result, the etched mixed oxides show the lowest light-off temperature and apparent activation energy for soot oxidation via an enhanced Mv-K reaction mechanism.

Author Contributions

Conceptualization, C.Z. and X.W.; Data curation, C.Z. and X.W.; Formal analysis, C.Z. and Z.L.; Funding acquisition, X.W. and D.W.; Investigation, C.Z. and X.W.; Methodology, R.R. and X.W.; Project administration, D.W.; Resources, Z.L., R.R. and D.W.; Supervision, X.W.; Visualization, Z.L. and R.R.; Validation, Z.L.; Writing—original draft, C.Z.; Writing—review and editing, C.Z., X.W., Z.L. and R.R. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by projects of the National Key Research and Development Program of China (No. 2023YFC3707202), Mobile Source Emission Control Technology (NELMS2020A08) and Key Laboratory of Advanced Materials of Ministry of Education of China (ADV22-18).

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors would like to thank the help from Letong Yang and Letao Li.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work that was reported in this paper.

References

  1. Li, Y.; Guo, H.; Xiong, J.; Ma, Y.; Li, X.; Zhang, P.; Zhang, S.; Wei, Y. The Catalyst of Ruthenium Nanoparticles Decorated Silicalite-1 Zeolite for Boosting Catalytic Soot Oxidation. Catalysts 2022, 13, 1167. [Google Scholar] [CrossRef]
  2. Leonardi, S.A.; Miró, E.E.; Milt, V.G. Activity of catalytic ceramic papers to remove soot particles-A study of different types of Soot. Catalysts 2022, 12, 855. [Google Scholar] [CrossRef]
  3. Zhu, S.; Shi, S.; Zheng, X.; Wang, X.; Yu, G.; Jiang, Y.; Feng, J.; Zhu, L.; Zhang, G. Enhanced oxygen vacancies in Ce-Doped SnO2 nanofibers for highly efficient soot catalytic combustion. Catalysts 2022, 12, 596. [Google Scholar] [CrossRef]
  4. Zheng, C.; Mao, D.; Xu, Z.; Zheng, S. Strong Ru-CeO2 interaction boosts catalytic activity and stability of Ru supported on CeO2 nanocube for soot oxidation. J. Catal. 2022, 411, 122–134. [Google Scholar] [CrossRef]
  5. Zhang, H.; Hu, W.; Zhou, C.; Liu, H.; Yao, P.; Wang, J.; Chen, Y. A new understanding of CeO2-ZrO2 catalysts calcinated at different temperatures: Reduction property and soot-O2 reaction. Appl. Catal. A 2018, 563, 204–215. [Google Scholar] [CrossRef]
  6. Ji, F.; Men, Y.; Wang, J.; Sun, Y.; Wang, Z.; Zhao, B.; Tao, X.; Xu, G. Promoting diesel soot combustion efficiency by tailoring the shapes and crystal facets of nanoscale Mn3O4. Appl. Catal. B 2019, 242, 227–237. [Google Scholar] [CrossRef]
  7. Xu, J.; Lu, G.; Guo, Y.; Guo, Y.; Gong, X. A highly effective catalyst of Co-CeO2 for the oxidation of diesel soot: The excellent NO oxidation activity and NOx storage capacity. Appl. Catal. A 2017, 535, 1–8. [Google Scholar] [CrossRef]
  8. Zhang, Z.; Han, D.; Wei, S.; Zhang, Y. Determination of active site densities and mechanisms for soot combustion with O2 on Fe-doped CeO2 mixed oxides. J. Catal. 2010, 276, 16–23. [Google Scholar] [CrossRef]
  9. Lin, X.; Li, S.; He, H.; Wu, Z.; Wu, J.; Chen, L.; Ye, D.; Fu, M. Evolution of oxygen vacancies in MnOx-CeO2 mixed oxides for soot oxidation. Appl. Catal. B 2018, 223, 91–102. [Google Scholar] [CrossRef]
  10. Zhang, H.; Zhou, C.; Galvez, M.; Costa, P.; Chen, Y. MnOx-CeO2 mixed oxides as the catalyst for NO-assisted soot oxidation: The key role of NO adsorption/desorption on catalytic activity. Appl. Surf. Sci. 2018, 462, 678–684. [Google Scholar] [CrossRef]
  11. Jia, A.; Jiang, S.; Lu, J.; Luo, M. Study of catalytic activity at the CuO-CeO2 interface for CO oxidation. J. Phys. Chem. C 2010, 114, 21605–21610. [Google Scholar] [CrossRef]
  12. Davó-Quinonero, A.; Navlani-García, M.; Lozano-Castelló, D.; Bueno-López, A.; Anderson, J.A. Role of hydroxyl groups in the preferential oxidation of CO over copper oxide-cerium oxide catalysts. ACS Catal. 2016, 6, 1723–1731. [Google Scholar] [CrossRef]
  13. Liang, Q.; Wu, X.; Weng, D.; Xu, H. Oxygen activation on Cu/Mn–Ce mixed oxides and the role in diesel soot oxidation. Catal. Today 2008, 139, 113–118. [Google Scholar] [CrossRef]
  14. Nakagawa, K.; Ohshima, T.; Tezuka, Y.; Katayama, M.; Katoh, M.; Sugiyama, S. Morphological effects of CeO2 nanostructures for catalytic soot combustion of CuO/CeO2. Catal. Today 2015, 246, 67–71. [Google Scholar] [CrossRef]
  15. Avgouropoulos, G.; Ioannides, T. Selective CO oxidation over CuO-CeO2 catalysts prepared via the urea–nitrate combustion method. Appl. Catal. A 2003, 244, 155–167. [Google Scholar] [CrossRef]
  16. Heynderickx, P.M.; Thybaut, J.W.; Poelman, H.; Poelman, D.; Marin, G.B. The total oxidation of propane over supported Cu and Ce oxides: A comparison of single and binary metal oxides. J. Catal. 2010, 272, 109–120. [Google Scholar] [CrossRef]
  17. Wang, J.; Cheng, L.; An, W.; Xu, J.; Men, Y. Boosting soot combustion efficiencies over CuO–CeO2 catalysts with a 3DOM structure. Catal. Sci. Technol. 2016, 6, 7342–7350. [Google Scholar] [CrossRef]
  18. Andana, T.; Piumetti, M.; Bensaid, S.; Veyre, L.; Thieuleux, C.; Russo, N.; Fino, D.; Quadrelli, E.A.; Pirone, R. CuO nanoparticles supported by ceria for NOx-assisted soot oxidation: Insight into catalytic activity and sintering. Appl. Catal. B 2017, 216, 41–58. [Google Scholar] [CrossRef]
  19. Piumetti, M.; Bensaid, S.; Andana, T.; Russo, N.; Pirone, R.; Fino, D. Cerium-copper oxides prepared by solution combustion synthesis for total oxidation reactions: From powder catalysts to structured reactors. Appl. Catal. B 2017, 205, 455–468. [Google Scholar] [CrossRef]
  20. Cui, B.; Yan, S.; Xia, Y.; Li, K.; Li, S.; Wang, D.; Ye, Y.; Liu, Y. CuxCe1-xO2 nanoflakes with improved catalytic activity and thermal stability for diesel soot combustion. Appl. Catal. A 2019, 578, 20–29. [Google Scholar] [CrossRef]
  21. Zhao, B.; Ran, R.; Sun, L.; Guo, X.; Wu, X.; Weng, D. NO catalytic oxidation over an ultra-large surface area LaMnO3+δ perovskite synthesized by an acid-etching method. RSC Adv. 2016, 6, 69855–69860. [Google Scholar] [CrossRef]
  22. Liang, F.; Yu, Y.; Zhou, W.; Xu, X.; Zhu, Z. Highly defective CeO2 as a promoter for efficient and stable water oxidation. J. Mater. Chem. A 2015, 3, 634–640. [Google Scholar] [CrossRef]
  23. Wei, Y.; Liu, J.; Zhao, Z.; Duan, A.; Jiang, G.; Xu, C.; Gao, J.; He, H.; Wang, X. Three-dimensionally ordered macroporous Ce0.8Zr0.2O2-supported gold nanoparticles: Synthesis with controllable size and super-catalytic performance for soot oxidation. Energy Environ. Sci. 2011, 4, 2959–2970. [Google Scholar] [CrossRef]
  24. Ji, J.; Jing, M.; Wang, X.; Tan, W.; Guo, K.; Li, L.; Wang, X.; Song, W.; Cheng, L.; Sun, J.; et al. Activating low-temperature NH3-SCR catalyst by breaking the strong interface between acid and redox sites: A case of model Ce2(SO4)3-CeO2 study. J. Catal. 2021, 399, 212–223. [Google Scholar] [CrossRef]
  25. Li, J.; Liu, Z.; Cullen, D.; Hu, W.; Huang, J.; Yao, L.; Peng, Z.; Liao, P.; Wang, R. Distribution and valence state of Ru species on CeO2 supports: Support shape effect and its influence on CO oxidation. ACS Catal. 2019, 9, 11088–11103. [Google Scholar] [CrossRef]
  26. Deng, X.; Li, M.; Zhang, J.; Hu, X.; Zheng, J.; Zhang, N.; Chen, B. Constructing nano-structure on silver/ceria-zirconia towards highly active and stable catalyst for soot oxidation. Chem. Eng. J. 2017, 313, 544–555. [Google Scholar] [CrossRef]
  27. Zhang, C.; Wang, J.; Yang, S.; Liang, H.; Men, Y. Boosting total oxidation of acetone over spinel MCo2O4 (M = Co, Ni, Cu) hollow mesoporous spheres by cation-substituting effect. J. Colloid Interface Sci. 2019, 539, 65–75. [Google Scholar] [CrossRef] [PubMed]
  28. Yang, S.; Wang, J.; Chai, W.; Zhu, J.; Men, Y. Enhanced soot oxidation activity over CuO/CeO2 mesoporous nanosheets. Catal. Sci. Technol. 2019, 9, 1699–1709. [Google Scholar] [CrossRef]
  29. Othmane, A.; Essamlali, Y.; Fihri, A.; Larzek, M.; Zahouily, M. Effect of calcination temperature on the structure and catalytic performance of copper-ceria mixed oxide catalysts in phenol hydroxylation. RSC Adv. 2017, 7, 12586–12597. [Google Scholar]
  30. Xu, X.; Liu, L.; Tong, Y.; Fang, X.; Xu, J.; Jiang, D.; Wang, X. Facile Cr3+-doping strategy dramatically promoting Ru/CeO2 for low-temperature CO2 methanation: Unraveling the roles of surface oxygen vacancies and hydroxyl groups. ACS Catal. 2021, 11, 5762–5775. [Google Scholar] [CrossRef]
  31. Chang, S.; Li, M.; Hua, Q.; Zhang, L.; Ma, Y.; Ye, B.; Huang, W. Shape-dependent interplay between oxygen vacancies and Ag–CeO2 interaction in Ag/CeO2 catalysts and their influence on the catalytic activity. J. Catal. 2012, 293, 195–204. [Google Scholar] [CrossRef]
  32. Wu, Z.; Li, M.; Howe, J.; Meyer, H.M.; Overbury, S.H. Probing defect sites on CeO2 nanocrystals with well-defined surface planes by Raman spectroscopy and O2 adsorption. Langmuir 2010, 26, 6595–16606. [Google Scholar] [CrossRef] [PubMed]
  33. Li, Y.; Wei, Z.; Gao, F.; Kovarik, L.; Peden, C.H.F.; Wang, Y. Effects of CeO2 support facets on VOx/CeO2 catalysts in oxidative dehydrogenation of methanol. J. Catal. 2014, 315, 15–24. [Google Scholar] [CrossRef]
  34. Mao, X.; Liu, S.; Liu, W.; Wu, X.; Liu, S. A simple model catalyst study to distinguish the roles of different oxygen species in propane and soot combustion. Appl. Catal. B 2022, 310, 121331. [Google Scholar] [CrossRef]
  35. Gao, Y.; Duan, A.; Liu, S.; Wu, X.; Liu, W.; Li, M.; Chen, S.; Wang, X.; Weng, D. Study of Ag/CexNd1-xO2 nanocubes as soot oxidation catalysts for gasoline particulate filters: Balancing catalyst activity and stability by Nd doping. Appl. Catal. B 2017, 203, 116–126. [Google Scholar] [CrossRef]
  36. Cárdenas-Arenas, A.; Quindimil, A.; Davó-Quiñonero, Q.; Bailón-García, E.; Lozano-Castelló, D.; De-La-Torre, U.; Pereda-Ayo, B.; González-Marcos, J.A.; González-Velasco, J.R.; Bueno-López, A. Isotopic and In Situ DRIFTS study of the CO2 methanation mechanism using Ni/CeO2 and Ni/Al2O3 catalysts. Appl. Catal. B 2020, 265, 118538. [Google Scholar] [CrossRef]
  37. Zimmer, P.; Tschöpe, A.; Birringer, R. Temperature-programmed reaction spectroscopy of ceria- and Cu/ceria-supported oxide catalyst. J. Catal. 2002, 205, 339–345. [Google Scholar] [CrossRef]
  38. Sun, S.; Mao, D.; Yu, J.; Yang, Z.; Lu, G.; Ma, Z. Low-temperature CO oxidation on CuO/CeO2 catalysts: The significant effect of copper precursor and calcination temperature. Catal. Sci. Technol. 2015, 5, 3166–3181. [Google Scholar] [CrossRef]
  39. Gamarra, D.; Cámara, A.L.; Monte, M.; Rasmussen, S.B.; Chinchilla, L.E.; Hungría, A.B.; Munuera, G.; Gyorffy, N.; Schay, Z.; Corberán, V.C.; et al. Preferential oxidation of CO in excess H2 over CuO/CeO2 catalysts: Characterization and performance as a function of the exposed face present in the CeO2 support. Appl. Catal. B 2013, 130–131, 224–238. [Google Scholar] [CrossRef]
  40. Piumetti, M.; Bensaid, S.; Russo, N.; Fino, D. Nanostructured ceria-based catalysts for soot combustion: Investigations on the surface sensitivity. Appl. Catal. B 2015, 165, 742–751. [Google Scholar] [CrossRef]
Figure 1. (a) Temperature−programmed oxidation of soot, (b) isothermal oxidation of soot at 310 °C, (c) Arrhenius plots of the catalysts under tight contact modes, and (d) cycled test of CuCe−AE catalyst.
Figure 1. (a) Temperature−programmed oxidation of soot, (b) isothermal oxidation of soot at 310 °C, (c) Arrhenius plots of the catalysts under tight contact modes, and (d) cycled test of CuCe−AE catalyst.
Catalysts 13 01463 g001
Figure 2. TEM images of (a) CeO2, (b) CuCe and (c) CuCe-AE, and (d) XRD patterns of the catalysts.
Figure 2. TEM images of (a) CeO2, (b) CuCe and (c) CuCe-AE, and (d) XRD patterns of the catalysts.
Catalysts 13 01463 g002
Figure 3. HAADF-STEM images of (a) CuCe and (b) CuCe-AE, and corresponding EDS elemental mapping of (c,d) Cu, (e,f) Ce, and (g,h) O.
Figure 3. HAADF-STEM images of (a) CuCe and (b) CuCe-AE, and corresponding EDS elemental mapping of (c,d) Cu, (e,f) Ce, and (g,h) O.
Catalysts 13 01463 g003
Figure 4. (a) N2 adsorption−desorption isothermals and (b) BJH pore diameter distribution of the catalysts.
Figure 4. (a) N2 adsorption−desorption isothermals and (b) BJH pore diameter distribution of the catalysts.
Catalysts 13 01463 g004
Figure 5. (a) The full XPS spectra of the catalysts and fine spectra of (b) Ce 3d, (c) Cu 2p, and (d) O 1 s.
Figure 5. (a) The full XPS spectra of the catalysts and fine spectra of (b) Ce 3d, (c) Cu 2p, and (d) O 1 s.
Catalysts 13 01463 g005
Figure 6. (a) Raman spectra and (b) H2−TPR profiles of the catalysts.
Figure 6. (a) Raman spectra and (b) H2−TPR profiles of the catalysts.
Catalysts 13 01463 g006
Table 1. Reaction parameters of the catalysts for soot oxidation.
Table 1. Reaction parameters of the catalysts for soot oxidation.
CatalystT50 (°C) aSCO2 (%) aReaction Rate (μmol min−1 g−1) bEa (kJ/mol) c
CeO241299.830.896
CuCe44399.016.2110
CuCe-AE38399.546.083
a Obtained from soot-TPO. b Determined by isothermal reaction at 310 °C. c Determined by Arrhenius equation.
Table 2. Comparison of activity of CuCe-AE with Ce-based catalysts previously reported in the literature under tight contact conditions.
Table 2. Comparison of activity of CuCe-AE with Ce-based catalysts previously reported in the literature under tight contact conditions.
CatalystReactant GasFlow Rate (mL/min)T50 (°C)Reference
CuCe-AE10%O2/N280383This work
Sn0.7Ce0.3O2O2/N22684373
Ru/CeO210%O2/N2803864
Cu-Ce10%O2/N250035613
CuO/Ce-DDAAir-41314
Ce0.95Cu0.05O210%O2/N210038419
Cu0.1Ce0.9O2-NF5%O2/Ar5031020
Table 3. Structural and textural properties of the catalysts.
Table 3. Structural and textural properties of the catalysts.
CatalystCu Content (wt.%) aD (nm) bSBET (m2/g) cVpore(cm3/g) cID/IF2g d
CeO2/10.51040.300.32
CuCe20.05.8980.220.38
CuCe-AE2.35.41450.270.40
a Determined by ICP-AES. b Crystallite size estimated via the Scherrer equation using the (111) peak of CeO2. c Determined by N2 sorption. d Obtained from Raman spectra.
Table 4. XPS results of the catalysts.
Table 4. XPS results of the catalysts.
CatalystSurface
Cu/Ce
Ce 3dCe3+/Ce4+ (%)O 1sOα/Oβ (%)
Ce3+ (%)Ce4+ (%)Oα (%)Oβ (%)
CeO2-18.381.722.4287238.8
CuCe0.9516.683.419.9257533.3
CuCe-AE0.3119.280.823.8386261.3
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

Zheng, C.; Wu, X.; Li, Z.; Ran, R.; Weng, D. Enhanced Soot Oxidation Activity of a CuO-Doped CeO2 Catalyst via Acid Etching. Catalysts 2023, 13, 1463. https://doi.org/10.3390/catal13121463

AMA Style

Zheng C, Wu X, Li Z, Ran R, Weng D. Enhanced Soot Oxidation Activity of a CuO-Doped CeO2 Catalyst via Acid Etching. Catalysts. 2023; 13(12):1463. https://doi.org/10.3390/catal13121463

Chicago/Turabian Style

Zheng, Changlong, Xiaodong Wu, Zhenguo Li, Rui Ran, and Duan Weng. 2023. "Enhanced Soot Oxidation Activity of a CuO-Doped CeO2 Catalyst via Acid Etching" Catalysts 13, no. 12: 1463. https://doi.org/10.3390/catal13121463

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