Next Article in Journal
Enhanced Diclofenac Photomineralization under Solar Light Using Ce1−xZnxO2−x Solid Solution Catalysts: Synergistic Effect of Photoexcited Electrons and Oxygen Vacancies
Next Article in Special Issue
Hard Template-Assisted Trans-Crystallization Synthesis of Hierarchically Porous Cu-SSZ-13 with Enhanced NH3-SCR Performance
Previous Article in Journal
Synthesis of Tungsten-Modified Sn3O4 through the Cetyltrimethylammonium Bromide-Assisted Solvothermal Method for Dye Decolorization under Visible Light Irradiation
Previous Article in Special Issue
The Catalyst of Ruthenium Nanoparticles Decorated Silicalite-1 Zeolite for Boosting Catalytic Soot Oxidation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Revealing the Roles of Cu/Ba on Ce-Based Passive NOx Adsorbers

1
Sichuan Provincial Environmental Protection Environmental Catalytic Materials Engineering Technology Center, College of Chemistry, Sichuan University, Chengdu 610064, China
2
Institute of New Energy and Low-Carbon Technology, Sichuan University, Chengdu 610207, China
3
Institute of Urban Environment, Chinese Academy of Sciences, Jimei Road 1799, Xiamen 361000, China
4
State Key Laboratory of Pollution Control and Resource Reuse, School of Environment, Nanjing University, Nanjing 210023, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(8), 1180; https://doi.org/10.3390/catal13081180
Submission received: 27 June 2023 / Revised: 27 July 2023 / Accepted: 31 July 2023 / Published: 2 August 2023

Abstract

:
At present, passive NOx adsorbers (PNAs) represent one of the most effective technologies for addressing NOx emissions from diesel engines during cold-start periods. Conventional PNAs, which primarily consist of noble metals (such as Pt, Pd, and Ag) loaded on metal oxides or zeolites, share the common drawback of high production costs. Consequently, developing low-cost PNAs with outstanding NOx storage performance remains a significant challenge. In this study, a series of CuxBa5Ce adsorbents were synthesized using the impregnation method, and a monolithic adsorbent was employed to evaluate NOx storage and release performance. Techniques such as XRD, UV-Vis DRs, H2-TPR, XPS, and in situ DRIFTs confirmed the crucial roles of Cu and Ba in NOx storage and release. Specifically, the incorporation of Cu into CeO2 enhanced NOx storage performance. Moreover, in the Cu3Ba5Ce adsorbent, the addition of Ba not only introduced new storage sites and altered the stability of NOx adsorption species but also helped prevent the aggregation of CuO, thereby prolonging the complete NOx storage duration and satisfying desorption temperature requirements. The Cu3Ba5Ce adsorbent exhibited the most favorable NOx storage performance, including a complete NOx storage time of 135 s and a NOx storage efficiency exceeding 50% at 80 °C over a 10 min period. While PNAs loaded with noble metals, such as Pd/CeO2 and Pt/CeO2, exhibited NOx storage efficiencies below 50% after adsorbing for 5 min at 80 °C. Therefore, this research offered a crucial strategy for developing non-noble-metal-loaded, Ce-based PNAs.

1. Introduction

Nitrogen oxides (NOx), comprising NO and NO2 emitted from diesel engines, represent a significant source of air pollution. The selective catalytic reduction in NOx with ammonia (NH3-SCR) is among the most effective technologies for mitigating NOx emissions. The operational temperature range of NH3-SCR catalysts typically exceeds 180 °C; however, the warm-up time from a cold start is usually 100–120 s [1,2]. With increasingly stringent emission regulations, further reducing NOx emissions from diesel engines will face a greater challenge, especially during the cold-start period. A new concept of passive NOx adsorbers (PNAs) was first proposed to store NOx by the Ford Motor Company in 2001 and was reintroduced as a “Low Temperature NOx Adsorber (LTNA)” in 2015. The adsorbent would adsorb and store NOx as nitrate/nitrites species during the cold-start period; once the downstream SCR catalyst temperature increased above 180 °C, the stored NOx would be released and reduced to N2 by NH3-SCR [3]. Hence, the PNAs should meet the following requirements: NOx should be adsorbed during the cold-start period and have enough complete storage time (at least 100 s); and NOx is desorbed in the operational temperature range of NH3-SCR catalysts [4].
Two types of PNAs have recently garnered significant interest in both academia and industry: noble metal (Pt, Pd, and Ag)-supported metal oxides (Al2O3, TiO2, CeO2, or their solid solutions) and Pd-supported zeolites (Pd/SSZ-13, Pd/Beta, and Pd/ZSM-5) adsorbents [5,6,7,8,9,10,11]. Among these, CeO2 is a candidate for PNA support because of its excellent redox properties and abundant oxygen vacancies, which facilitate the adsorption and storage of NOx [12,13,14,15,16,17]. Additionally, noble metals supported on CeO2 will further improve the NOx adsorption capacity, and the desorption temperature is about 250–460 °C [18]. However, the NOx adsorption capacity of Pt/CeO2 is usually larger than that of Pd/CeO2 [19], and both of the adsorbents exhibit NOx storage efficiencies below 50% after adsorbing for 5 min at 80 °C. Additionally, the related characterizations revealed that NO was trapped primarily on the Pt and Pd sites in the intake air, and then the trapped NO was transferred to the surface of CeO2 and stored as nitrate or nitrite species. And, transition metals (e.g., Pr, Y, Zr, etc.) are doped into CeO2 to generate solid solution oxides [20,21], which are beneficial to the formation of surface oxygen vacancies (Ce3+-X) and active oxygen (Ce4+-O*) and then improve the NOx storage efficiency [22,23]. It has been demonstrated that CuO can increase the adsorption and oxidation performances of NO [24]. Zeng et al. discovered that CuO highly dispersed on TiO2 displayed excellent catalytic oxidation activity of NO rather than CuO particles [25]. In addition, loading alkali (Na, K) and alkaline earth metals (Ba) on the adsorbents can also enhance the basicity and thus improve NOx storage capacity, e.g., BaO + NO + O2→Ba(NO3)2/Ba(NO2)2. This stored nitrate species will decompose at elevated temperatures to release NOx. Importantly, the nitrate and nitrite species formed on Ba were more stable than those stored on CeO2, thus contributing to NOx desorption in the operational temperature range of NH3-SCR catalysts [26,27].
Although noble-metal-loaded PNAs exhibit high NOx storage capacities at low temperatures [5,6,7,8,9,10,11,28], developing cost-effective, non-precious metal PNAs with high NOx storage capacities is essential. In this study, a series of CuxBa5Ce adsorbents supported on CeO2 with improved NOx storage capacities and storage stabilities were prepared by the step-impregnation method. Various characterizations, including X-ray diffraction, UV-Vis diffuse reflection spectroscopy, X-ray photoelectron spectroscopy, H2 temperature-programmed reduction, and in situ diffuse reflectance infrared Fourier transform spectra, were employed to investigate the presence of Cu and Ba on CeO2 and elucidate their roles on NOx adsorption and storage over CeO2-based PNAs. This work aims to provide a novel strategy for designing cost-effective, non-noble PNAs.

2. Results and Discussion

2.1. Results of NOx Adsorption/Desorption

The NOx adsorption and NOx storage efficiencies (NSE) of CeO2, CuxCe, Ba5Ce, and CuxBa5Ce adsorbents are shown in Figure 1. Among the investigated adsorbents, pure CeO2 exhibited a relatively poor NOx storage efficiency of 10.1% and no complete storage time. NOx storage performance was significantly improved after impregnated with 1 wt.% Cu on CeO2, and the NSE of Cu1Ce was increased to 20.8% from 10.1% in CeO2, but there was no change on the complete storage time. Increasing the loading of Cu, the NSE was increased to 43.4% in Cu3Ce and 40.7% in Cu5Ce, where the complete storage time of NOx was extended to 68 s in Cu3Ce and 74 s in Cu5Ce. The above results indicated that the NSE and complete storage time of NOx were both effectively improved via increased loading of Cu on CuxCe. To further improve the PNA performance of the adsorbents, 5 wt.% Ba was impregnated on CeO2. The NOx storage performance of Ba5Ce was not enhanced after loading Ba on CeO2; however, the NSE of the CuxBa5Ce adsorbents were evidently improved after introducing Cu into Ba5Ce. The NSEs of Cu3Ba5Ce and Cu5Ba5Ce increased to 52.2% and 49.7%, respectively, surpassing the NOx storage efficiencies of Ce-based PNAs loaded with Pd/Pt. And, their complete storage times were both extended to 135 s. According to the above phenomena, the NSEs and complete storage times of CuxCe adsorbents were mainly affected by the content of Cu, whereas the complete storage time would be further extended by introducing Ba into CeO2.
Figure 2 shows the desorption curves and the desorbed amount of NOx in different temperature ranges. The two desorption peaks located at ca. 140 °C and 340 °C were discovered on CuxCe, respectively. Cu3Ce possessed the maximum NOx desorption amount among the investigated CuxCe adsorbents, but the desorption amount of NOx would not be increased by further enhancing the loading of Cu to 5 wt.%, which was consistent with the NSE results. The above phenomena illustrated that the NSE and desorption amount of CuxCe were positively correlated. The desorption amounts of CuxBa5Ce were not changed by further loading Cu on Ba5Ce, but the desorption peak temperature of CuxBa5Ce shifted to higher temperatures, where the desorption temperature range was shifted from 150–340 °C of Cu3Ce to 200–460 °C of Cu3Ba5Ce, respectively. Although the desorption temperature of Cu3Ba5Ce was lower than those of Pt/Pd-loaded Ce-based PNAs (250–460 °C), it still met the practical requirements. The higher desorption temperature of CuxBa5Ce would be attributed to increase the storage sites of NOx and/or changing the adsorption species of NOx and its stability by introducing Ba into CeO2 [26,27].

2.2. Textural and Structural Properties

The XRD results showed the effect of the addition of Cu and Ba on the structural properties of CeO2. The main diffraction peaks were assigned to the fluorite cubic structure of CeO2 (JCPDS 34-0394) for all samples, as displayed in Figure 3. In the CuxCe adsorbents, no diffraction peaks attributed to CuO were observed on Cu1Ce, whereas the intensity increased with increasing the content of Cu from 3 wt.% to 5 wt.%. A new diffraction peak attributed to BaCO3 (2θ = 23.9°) was observed after further loading 5 wt.% Ba onto CeO2. According to the report [29], the “low temperature LT-BaCO3” was formed when the loading of Ba was 5 wt.%–6 wt.% via the reaction between BaO generated by the thermal decomposition of the precursor Ba(OAc)2 and CO2 in air, which displayed excellent NOx storage performance but was easily decomposed to BaO and CO2 between 400–800 °C. However, it was noteworthy that the intensity of the diffraction peak assigned to CuO was decreased after loading Cu onto Ba5Ce, and no CuO diffraction peak was observed on Cu5Ba5Ce. This result indicated that the loading of Ba onto CeO2 would effectively inhibit the aggregation of CuO.
It has been reported that a decrease in specific surface area leads to a decrease in NOx storage performance when CeO2 is used as a NOx storage site [30]. According to the textural and structural data in Table 1, the crystalline size of CeO2 almost did not change, but its specific surface area decreased after loading Cu and Ba. The specific surface area of Cu3Ce was 136.9 m2/g and declined by 23.5% to 104.7 m2/g of Cu3Ba5Ce, but the NSE of Cu3Ba5Ce was 7% higher than that of Cu3Ce, and the complete storage time was extended by 67 s, which indicated that the loading of Ba provided new storage sites and effectively compensated for the loss of CeO2 storage sites.
The UV-Vis diffuse reflectance spectrum was employed to investigate the coordination and oxidation states of Cu. Three bands were observed on all samples, as shown in Figure 4; the strong band around 275 nm would be attributed to O2−→Cu2+ ligand to the metal charge transfer (LCMT) transition. The band at 340–450 nm could be assigned to the charge transfer involving (Cu-O-Cu)2+ cluster-like species, and the broad band (black dashed box) at 600–800 nm could be ascribed to the 2Eg2T2g spin-allowed transition of Cu2+ situated in a distorted octahedral symmetry. The spectra of CuxCe and CuxBa5Ce revealed that all the adsorbents showed the 340–450 nm band, indicating that loading Cu on CeO2 would result in the formation of (Cu-O-Cu)2+ clusters, and similar results were reported by Shi et al. [31]. It has been demonstrated that the adsorption and storage of NOx were improved by the formation of (Cu-O-Cu)2+ clusters [32,33]. A red shift was observed on the absorption edge at 340–450 nm, which may have been due to an increase in the number of (Cu-O-Cu)2+ clusters caused by an increase in Cu loading [34,35]. Additionally, the larger CuO particles were formed on the Cu5Ce due to a clear edge adsorption at 750–800 nm, but the adsorption intensities of Cu5Ba5Ce were weaker than those of Cu5Ce, which indicated that the loading of Ba inhibited the aggregation of CuO, in agreement with the result of XRD [36].

2.3. TPR

H2-TPR was performed to investigate the reducibility of the different adsorbents, as shown in Figure 5. There were two or three reduction peaks between 100 °C and 250 °C after loading Cu onto CeO2, but the reduction peak at 500–600 °C could have been attributed to the reduction in the surface active oxygen on CeO2 and Ba5Ce. The low-temperature peak (α) was attributed to the reduction peak of highly dispersed Cu species, and the medium-temperature peak (β) was attributed to the reduction in (Cu-O-Cu)2+ clusters, which would have been consistent with the results of the UV-Vis DRs [37]. Two reduction peaks were observed on Cu1Ce, whereas a new (γ) peak that appeared on Cu3Ce and Cu5Ce would be assigned to the reduction in CuO particles with increasing Cu loading, which could be accordance with the results of XRD [31]. On the contrary, both (α) and (β) peaks could be identified, and no reduction peak assigned to the CuO particles was observed on CuxBa5Ce with an increase in Cu loading, which further demonstrated that the loading of Ba inhibited the formation of CuO particles, in agreement with the results of UV-Vis DRs and XRD.

2.4. XPS

XPS was performed to identify the chemical states of elements on adsorbents. The XPS spectra of Cu 2p, O 1s, and Ce 3d are shown in Figure 6, and the concentrations of surface Cu2+, Ce3+, chemically adsorbed oxygen, and the atomic ratios are listed in Table 2.
As shown in Figure 6a, the spectra of Cu 2p were centered at 932.28, 933.78, 951.88, and 953.38 eV and represented the photoelectrons emitted from Cu+/Cu0 2p3/2, Cu2+ 2p3/2, Cu+/Cu0 2p1/2, and Cu2+ 2p1/2, respectively. The shift to higher binding energy with the increase in Cu loading amount indicated that the interaction between Cu and Ce was enhanced. The concentration of Cu2+ on the adsorbents was calculated as (Cu2+(%) = A(Cu2+)/Σ[A(Cu2+) + A(Cu+)] × 100 (A represents the peak area of each element)). As listed in Table 2, the concentrations of surface Cu2+ increased in the following order: Cu1Ce (71.6) < Cu3Ce (76.2%) < Cu5Ce (81.7%) and Cu1Ba5Ce (61.9%) < Cu3Ba5Ce (66.2%) < Cu5Ba5Ce (71.2%), and the ratios of Cu/Ce, Cu/Ba, and O/Ce gradually increased with the increase in Cu loading and the decrease in Ce content. The results illustrated that Cu2+ was the main chemical valence of Cu on all adsorbents, and its concentration increased with the increase in Cu content. However, it is noteworthy that the concentration of Cu2+ was approximately 10% lower after loading 5 wt.% Ba onto CeO2. The possible explanation was that the electron density around Cu was affected by the strong electron donation from doped Ba species [38], thus reducing the concentration of Cu2+. Additionally, the ratio of Ba/Ce decreased with increasing Cu loading, which may have been due to the fact that excessive Cu loading would cover the Ba on the surface, leading to a decrease in Ba concentration, as evidenced by the decreased surface content of Ba, listed in Table 2.
The O 1s XPS spectra of the adsorbents are shown in Figure 6b. The band at the higher binding energy (531.48 eV) is ascribed to the surface oxygen O22− and O (Oα; yellow color); the peak at the lower binding energy (529.23 eV) is assigned to the lattice oxygen O2− (Oβ; green color) [39]. The concentration of Oα as (Oα(%) = A(Oα)/Σ[A(Oα) + A(Oβ)] × 100 (A represents the peak area of each element)). For CuxCe and CuxBa5Ce adsorbents, there was a slight decrease in surface oxygen (Oα) concentration with the increase in Cu content; however, it is noteworthy that the concentration of Oα on the CuxBa5Ce adsorbents is 1.5 times that of the CuxCe adsorbents. For CuxBa5Ce, more surface oxygen is related to the maintenance of charge balance after doping Ba [40]. Surface oxygen, acting as active oxygen species, is favorable to boost the formation of NO2, and then the NOx storage is advanced.
The Ce 3d XPS spectra of the adsorbents are shown in Figure 6c, and the corresponding assignments are defined [41]. Two groups of spin orbital multiplets attributed to 3d3/2 and 3d5/2 are denoted as “u” and “v” in the binding energy range of 880–920 eV. It is widely reported that the u1 and v1 belong to the orbital peaks of the 3d104f1 electron state Ce3+, and u0, u2, u3, v0, v2, and v3 belong to the orbital peaks of 3d104f0 electron state Ce4+. The concentration of Ce3+ on the adsorbents is calculated as (Ce3+(%) = A(v1) + A(u1)/Σ[A(v) + A(u)] × 100 (A represents the peak area of each element)). The concentration of Ce3+ decreases in the following order: CeO2 (19.1%) > Cu1Ce (16.8%) > Cu3Ce (15.4%) > Cu5Ce (14.2%) and Ba5Ce (19.5%) > Cu1Ba5Ce (15.8%) > Cu3Ba5Ce (14.4%) > Cu5Ba5Ce (13.4%). The results indicate that the concentration of Ce4+ is increased with the increase in Cu content. According to the literature [42], the increased amounts of Cu2+ and Ce4+ are beneficial to generate more crystal lattice oxygen (Oβ) species, providing a material basis for the formation of highly reactive oxygen species. This may have been the reason for the slight decrease in the concentration of surface oxygen (Oα) species.
In short, the XPS results show that Cu2+and Ce4+ are the main chemical valences of Cu and Ce for all adsorbents, which can provide a material basis for reactive oxygen species and the formation of (Cu-O-Cu)2+ clusters. Additionally, the increase in surface oxygen species is favorable for the improvement of oxidation activity after the addition of Ba to CeO2 and then improving the NSE of CeO2.

2.5. NOx-DRIFTs

NOx adsorption behaviors of CuxCe samples were investigated by in situ DRIFTs measurements, as shown in Figure 7. The DRIFTs bands would be attributed to bridging bidentate nitrate (1577, 1562 cm–1), monodentate nitrite (1457 cm−1), monodentate nitrate (1290–1243 cm–1), and bridging bidentate nitrite (1303, 1186 cm–1) species found on pure CeO2 [43,44,45]. The band of nitrite species decreased, whereas the band of nitrate species increased after injecting NO + O2 for 10 min, which meant that the nitrite species were oxidized to nitrate species. In addition, the intensity of the adsorption bands would not be further strengthened by extend the injection time, indicating that the pure CeO2 was saturated after NOx adsorption for 10 min. For Cu1Ce, the main adsorption bands would be attributed to the bidentate nitrite species (1186 cm−1), and the corresponding intensities were gradually increased after injecting NO + O2 for 5 min. However, the adsorption intensities of the bidentate nitrite species (1211 cm−1) on the Cu3Ce adsorbent were significantly higher than those of Cu1Ce within 10 min. For Cu5Ce, the adsorption intensity of the bidentate nitrite species (1211 cm−1) was not further enhanced within 10 min, and the main species were changed to bridged bidentate nitrate (1567 cm−1), monodentate nitrite (1457 cm−1), and chelated bidentate nitrate (1278 cm−1) species after 10 min, which may have been that the oxidation activity of NO was enhanced after loading excess Cu to facilitate the formation of various nitrate species [32]. The intensities of these nitrate and nitrite species on all CuxCe adsorbents were significantly higher than those of pure CeO2. The above results suggested that the NOx adsorption species were mainly in the presence of nitrite species on all CuxCe adsorbents within 10 min, and increasing the loading of Cu was advantageous to promote NOx adsorption.
Figure 8 shows the NOx adsorption behaviors of CuxBa5Ce samples. For pure Ba5Ce, the DRIFT bands at 1565, 1196, and 1126 cm–1 were detected after injecting NO + O2 for 30 min, which were attributed to bridging bidentate nitrate species on Ce sites, monodentate nitrites, and monodentate nitrates on Ba sites, respectively [46]. Loading 1 wt.% Cu on Ba5Ce, the monodentate nitrite species (1196 cm–1) were found in the initial exposure stage and obtained their maximum intensities. Then, the monodentate nitrite species (1196 cm–1) decreased slowly after 10 min, with plenty of bridging bidentate nitrites (1234 cm–1) forming on Ba sites, and bridging bidentate nitrates (1565, 1604 cm–1) forming on Ce sites. For Cu3Ba5Ce, the adsorption intensity of monodentate nitrite species (1196 cm–1) formed on Ba sites was higher than those of Cu1Ba5Ce within 10 min. Moreover, the nitrate species would not be changed with further extended the adsorption time, and a weak shift to high frequency was found. After further increasing the loading of Cu to 5 wt.%, the adsorption intensities of bridging bidentate nitrates (1565, 1604 cm–1) and monodentate nitrites (1024 cm–1) formed on Ce sites were significantly higher than Cu1B5Ce and Cu3B5Ce. The monodentate nitrite species (1196 cm–1) showed a significant red shift with increased the adsorption time of NOx and transformed into bridging bidentate nitrites (1234 cm–1) after 30 min, indicating that 5 wt.% Cu loading favored the formation of bidentate nitrite on Ba sites and monodentate nitrites on Ce sites.
The adsorbed nitrate/nitrite species on the Ce sites displayed poor thermal stability than those adsorbed on the Ba sites, which was probably attributed to the covalent or ionic bonds between Ba and nitrogenous species [47]. And, it was known that there were significant differences among the thermal stabilities of different nitrate/nitrite species (nitrite < bidentate nitrates < monodentate nitrates < ionic nitrates) [47]. Therefore, the thermal stabilities of the adsorbed nitrate/nitrite species increased in the following order: bidentate nitrite species (1211, 1186 cm−1) formed on Ce sites < bidentate nitrites (1234 cm–1) formed on Ba sites < monodentate nitrites (1196 cm–1) formed on Ba sites. The above results demonstrated that the loading of Cu was beneficial to improve the amount of NOx adsorption without the change in NOx adsorbed species, whereas the loading of Ba provided new adsorption sites and changed the NOx adsorption species to strengthen thermal stability.

3. Materials and Methods

3.1. Preparation of Adsorbents

A series of CuO/CeO2 adsorbents were prepared by the conventional incipient wetness impregnation method from the precursors Cu(NO3)2·3H2O (AR grade, Keshi, China) and the commercial CeO2 powder (Sinocat, Chengdu, China) with different mass percentages (x = 1, 3, and 5) of CuO, labeled as CuxCe. Next, the as-prepared powders were achieved after drying at 105 °C for 12 h and calcining at 550 °C for 5 h.
In total, 5 wt.% BaO (denoted as Ba) from the precursor Ba(OAc)2 (AR grade, Keshi, China) was supported on the commercial CeO2 powder to prepare Ba5Ce as the same prepared process of CuxCe. Then, different mass percentages (x = 1, 3, and 5) of CuO from the precursor Cu(NO3)2·3H2O were impregnated on the as-prepared Ba5Ce powder to prepare CuxBa5Ce (x = 1, 3, and 5) and followed the same dried and calcined process as CuxCe and Ba5Ce.
The resulting seven adsorbent powders were subsequently coated on honeycomb cordierites (cylinder, diameter 11 mm, length 26 mm, bulk 2.5 cm3, 62 cell/cm2, Corning Ltd., Corning, NY, USA) to achieve the monolithic adsorbents. The detailed preparation process of the monolithic adsorbents referred to our previous works [48].

3.2. Catalytic Activity and Characterization

The PNA performances of adsorbents were tested using a continuous flow fixed-bed quartz reactor. The testing process procedures are shown in Figure 9. The NO, O2, and N2 flows were controlled by a 50 mL/min, 100 mL/min, and 1 L/min mass flowmeter before entering the reactor, respectively. The simulated reaction conditions were as follows: 500 ppm NO, 8 vol.% O2, N2 as balance gas, and a gas concentration error of about 10 ppm. The total space velocity of the gas was 30,000 h−1, corresponding to a total flow rate of 1250 mL/min. The outlet concentrations of NO, NO2, N2O, and NH3 were monitored by a Gas Fourier transform infrared spectrometer (Nicolet Antaris IGS-6700, Thermo Fisher Scientific, Waltham, MA, USA). NOx was provided by the China National Testing Institute (Chengdu, China), and O2 was provided by the Dongfeng Industrial Gas Company (Chengdu, China).
The NOx storage efficiency (hereafter denoted as NSE) and NOx desorption amounts (μmol) were calculated according to the following formula:
NSE = ( 1 0 t ( [ NO x ] out ) dt 0 t ( [ NO ] in ) dt ) × 100 %
NO x   desorption   amounts = V × t ( T 0 ) t ( T ) ( [ NO x ] out ) dt 22.4 × 1000
where t is the NOx storage time; V is the total flow rate; [NO]in is the inlet NO concentration during NOx storage; [NOx]out is the outlet NOx concentration during either NOx storage or the subsequent NOx desorption period; t(To) is the start time of NOx-TPD corresponding to the NOx storage temperature before the temperature is raised; and t(T) is the end time of NOx-TPD corresponding to the desired NOx desorption temperature.
The N2 adsorption–desorption was measured on a Quadrosorb SI automated surface area and a pore size analyzer (Quantachrome, Boynton Beach, FL, USA) at 77 K, and the sample was pre-treated at 300 °C for 3 h prior to the measurement. The surface areas of the adsorbents were calculated by the Brunauer–Emmett–Teller theory.
Powder X-ray diffraction (XRD) experiments were performed on a Rigaku DX-2500 diffractometer (Rigaku, Tokyo, Japan) using Cu Ka (l = 0.15406 nm) radiation. The tube voltage and current were 40 kV and 25 mA, respectively. XRD powder diffractograms were recorded at 0.031°/s intervals in the 2θ range of 10–90°.
The X-ray photoelectron spectra (XPS) data were recorded by the Thermo Scientific K-Alpha spectrometer with 12 kV high pressure and 6 mA operating current using Al Kα radiation. C 1 s (284.8 eV) was used for the internal standard to calibrate the binding energy.
UV-Vis diffuse reflectance spectra (UV-Vis DRS) were recorded in the range of 200–800 nm on a Shimatsu Uv-3600i Plus (Shimadzu, Tokyo, Japan).
Temperature programmed reduction with H2 (H2-TPR) experiment was carried out on a TP-5076 TPD/TPR dynamic adsorption (Xianquan, China) with a thermal conductivity detector. The samples approximately of 100 mg were pre-treated in a quartz tubular microreactor in a flow of pure N2 at 450 °C for 1 h and then cooled to 30 °C. The reduction was carried out in a flow of 5 vol.% H2-95 vol.% N2 from 30 °C to 900 °C with a heating rate of 10 °C/min.
In situ DRIFT Spectra were performed using a Nicolet Nexus 6700 FTIR spectrometer (Nicolet, Waltham, MA, USA) with a diffuse reflectance infrared Fourier transform spectroscopy (Drifts) equipped with a reaction cell and a KBr window. The reaction temperature was controlled by an Omega program temperature controller (Omega, Norwalk, CT, USA) and a heating rate of 10 °C/min. The powder sample was activated in N2 at 450 °C for 30 min. The background spectra of the sample were recorded in pure N2 from 450 to 80 °C. The adsorbents were exposed to 2 vol.% NO-98 vol.% N2/5 vol.% O2-95 vol.% N2 at 80 °C for 30 min, and the NOx adsorption spectra at 80 °C of adsorbents were collected.

4. Conclusions

The transition metal Cu and alkaline earth metal Ba were utilized to enhance the NOx storage efficiency and desorption performance of CeO2. The specific roles of Cu and Ba in this process were investigated thoroughly. Cu3Ba5Ce with 3 wt.% of Cu and 5 wt.% of Ba showed the highest NOx storage efficiency of 52.2% at 80 °C for 10 min, surpassing the NOx storage efficiency of Ce-based PNAs loaded with Pd/Pt. The desorption temperature range was 200–460 °C. The optimized functions of Cu and Ba could be summarized as follows. First, the loading of Cu on CeO2 was beneficial to improve the amount of NOx adsorption and storage, where Cu was present as Cu2+ in the form of (Cu-O-Cu)2+ clusters when the loading of Cu was ≤3 wt.% and agglomerated into CuO particles with the further increasing of the content of Cu to 5 wt.%, thus reducing the storage of NOx. Then, the formations of CuO particles were inhibited, and more surface oxygen participated in the storage of NOx after the addition of Ba to CeO2. Also, CuxBa5Ce provided new storage sites, and nitrate/nitrite species adsorbed on the Ba sites were more thermally stable than those adsorbed on the Ce sites, which resulted in a shift of NOx desorption temperature in the NH3-SCR reaction temperature window. Consequently, the NOx storage efficiency, complete storage time, and desorption performance of CeO2 were effectively enhanced by the introduction of the non-precious metals Cu and Ba, along with further optimizations. The economy of Ce-based PNAs could be significantly improved.

Author Contributions

Conceptualization, H.X. and J.W.; methodology, M.P. and Y.F.; validation, H.X., J.W. and M.P.; formal analysis, M.P.; investigation, M.P.; resources, H.X., J.W., and Y.C.; data curation, M.P. and Y.F.; writing—original draft preparation, M.P.; writing—review and editing, H.X., W.T. and Z.L.; supervision, H.X., J.W. and Y.C.; project administration, H.X.; funding acquisition, H.X. and J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (No. 22072098) and Sichuan Science and Technology Program (No. 2022ZHCG0125).

Data Availability Statement

Data will be made available upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chang, H.-L.; Chen, H.-Y.; Koo, K.; Rieck, J.; Blakeman, P. Gasoline Cold Start Concept (gCSC™) Technology for Low Temperature Emission Control. SAE Int. J. Fuels Lubr. 2014, 7, 480–488. [Google Scholar] [CrossRef]
  2. Chen, H.-Y.; Collier, J.E.; Liu, D.; Mantarosie, L.; Durán-Martín, D.; Novák, V.; Rajaram, R.R.; Thompsett, D. Low Temperature NO Storage of Zeolite Supported Pd for Low Temperature Diesel Engine Emission Control. Catal. Lett. 2016, 146, 1706–1711. [Google Scholar] [CrossRef]
  3. Ji, Y.; Xu, D.; Bai, S.; Graham, U.; Crocker, M.; Chen, B.; Shi, C.; Harris, D.; Scapens, D.; Darab, J. Pt- and Pd-promoted CeO2–ZrO2 for passive NOx adsorber applications. Ind. Eng. Chem. Res. 2017, 56, 111–125. [Google Scholar] [CrossRef]
  4. Lin, Q.J.; Pei, M.M.; Yao, P.; Xu, S.; Xu, S.H.; Liu, S.; Xu, H.D.; Dan, Y.; Chen, Y.Q. Determining hydrothermal deactivation mechanisms on Cu/SAPO-34 NH3-SCR catalysts at low-and high-reaction regions: Establishing roles of different reaction sites. Rare Met. 2022, 41, 1899–1910. [Google Scholar] [CrossRef]
  5. Johnson, T.V. Vehicular Emissions in Review. SAE Int. J. Engines 2012, 5, 216–234. [Google Scholar] [CrossRef] [Green Version]
  6. Ji, Y.; Bai, S.; Crocker, M. Al2O3-based passive NOx adsorbers for low temperature applications. Appl. Catal. B Environ. 2015, 170–171, 283–292. [Google Scholar] [CrossRef] [Green Version]
  7. Khivantsev, K.; Gao, F.; Kovarik, L.; Wang, Y.; Szanyi, J. Molecular Level Understanding of How Oxygen and Carbon Monoxide Improve NOx Storage in Palladium/SSZ-13 Passive NOx Adsorbers: The Role of NO+ and Pd(II)(CO)(NO) Species. J. Phys. Chem. C 2018, 122, 10820–10827. [Google Scholar] [CrossRef]
  8. Khivantsev, K.; Jaegers, N.R.; Kovarik, L.; Hanson, J.C.; Tao, F.F.; Tang, Y.; Zhang, X.; Koleva, I.Z.; Aleksandrov, H.A.; Vayssilov, G.N.; et al. Achieving Atomic Dispersion of Highly Loaded Transition Metals in Small-Pore Zeolite SSZ-13: High-Capacity and High-Efficiency Low-Temperature CO and Passive NO(x) Adsorbers. Angew. Chem. Int. Edtion 2018, 57, 16672–16677. [Google Scholar] [CrossRef]
  9. Vu, A.; Luo, J.; Li, J.; Epling, W.S. Effects of CO on Pd/BEA Passive NOx Adsorbers. Catal. Lett. 2017, 147, 745–750. [Google Scholar] [CrossRef]
  10. Zhao, H.; Hill, A.J.; Ma, L.; Bhat, A.; Jing, G.; Schwank, J.W. Progress and future challenges in passive NO adsorption over Pd/zeolite catalysts. Catal. Sci. Technol. 2021, 11, 5986–6000. [Google Scholar] [CrossRef]
  11. Tsukamoto, Y.; Nishioka, H.; Imai, D.; Sobue, Y.; Takagi, N.; Tanaka, T.; Hamaguchi, T. Development of New Concept Catalyst for Low CO2 Emission Diesel Engine Using NOx Adsorption at Low Temperatures. In Proceedings of the SAE World Congress & Exhibition, Detroit, MI, USA, 24–26 April 2012; SAE International: Warrendale, PA, USA, 2012. [Google Scholar] [CrossRef]
  12. Zheng, Y.; Kovarik, L.; Engelhard, M.H.; Wang, Y.; Wang, Y.; Gao, F.; Szanyi, J. Low-Temperature Pd/Zeolite Passive NOx Adsorbers: Structure, Performance, and Adsorption Chemistry. J. Phys. Chem. C 2017, 121, 15793–15803. [Google Scholar] [CrossRef]
  13. Le Phuc, N.; Corbos, E.C.; Courtois, X.; Can, F.; Marecot, P.; Duprez, D. NOx storage and reduction properties of Pt/CexZr1−xO2 mixed oxides: Sulfur resistance and regeneration, and ammonia formation. Appl. Catal. B Environ. 2009, 93, 12–21. [Google Scholar] [CrossRef]
  14. Levasseur, B.; Ebrahim, A.M.; Bandosz, T.J. Role of Zr4+ cations in NO2 adsorption on Ce1−xZrxO2 mixed oxides at ambient conditions. Langmuir 2011, 27, 9379–9386. [Google Scholar] [CrossRef]
  15. Yu, J.; Si, Z.; Chen, L.; Wu, X.; Weng, D. Selective catalytic reduction of NO by ammonia over phosphate-containing Ce0.75Zr0.25O2 solids. Appl. Catal. B Environ. 2015, 163, 223–232. [Google Scholar] [CrossRef]
  16. Hartridge, A.; Krishna, M.; Bhattacharya, A. Optical constants of nanocrystalline lanthanide-doped ceria thin films with the fluorite structure. J. Phys. Chem. Solids 1998, 59, 859–866. [Google Scholar] [CrossRef]
  17. Mamontov, E.; Egami, T.; Brezny, R.; Koranne, M.; Tyagi, S. Lattice defects and oxygen storage capacity of nanocrystalline ceria and ceria-zirconia. J. Phys. Chem. B 2000, 104, 11110–11116. [Google Scholar] [CrossRef]
  18. Skorodumova, N.V.; Simak, S.I.; Lundqvist, B.I.; Abrikosov, I.A.; Johansson, B. Quantum origin of the oxygen storage capability of ceria. Phys. Rev. Lett. 2002, 89, 166601. [Google Scholar] [CrossRef]
  19. Ryou, Y.; Lee, J.; Lee, H.; Kim, C.H.; Kim, D.H. Low temperature NO adsorption over hydrothermally aged Pd/CeO2 for cold start application. Catal. Today 2018, 307, 93–101. [Google Scholar] [CrossRef]
  20. Jones, S.; Ji, Y.; Crocker, M. Ceria-Based Catalysts for Low Temperature NOx Storage and Release. Catal. Lett. 2016, 146, 909–917. [Google Scholar] [CrossRef] [Green Version]
  21. Lamonier, J.; Kulyova, S.; Zhilinskaya, E.; Kostyuk, B.; Lunin, V.; Aboukaïs, A.J.K. Combustion of Carbon Black Catalyzed by Transition Metal-Promoted Y2O3–CeO2–ZrO2 Solid Solutions1. Kinet. Catal. 2004, 45, 429–435. [Google Scholar] [CrossRef]
  22. Wang, Y.; Wang, J.; Chen, H.; Yao, M.; Li, Y. Preparation and NOx-assisted soot oxidation activity of a CuO–CeO2 mixed oxide catalyst. Chem. Eng. Sci. 2015, 135, 294–300. [Google Scholar] [CrossRef]
  23. Jones, S.; Ji, Y.; Bueno-Lopez, A.; Song, Y.; Crocker, M. CeO2-M2O3 Passive NOx Adsorbers for Cold Start Applications. Emiss. Control Sci. Technol. 2016, 3, 59–72. [Google Scholar] [CrossRef]
  24. Priya, N.S.; Somayaji, C.; Kanagaraj, S. Synthesis and characterization of Nd3+-doped Ce0.6Zr0.4O2 and its doping significance on oxygen storage capacity. Rare Met. 2021, 40, 231–236. [Google Scholar] [CrossRef]
  25. Zhang, Z.-S.; Crocker, M.; Chen, B.-B.; Bai, Z.-F.; Wang, X.-K.; Shi, C. Pt-free, non-thermal plasma-assisted NO storage and reduction over M/Ba/Al2O3 (M = Mn, Fe, Co, Ni, Cu) catalysts. Catal. Today 2015, 256, 115–123. [Google Scholar] [CrossRef] [Green Version]
  26. Zeng, Y.; Wang, T.; Zhang, S.; Wang, Y.; Zhong, Q. Sol–gel synthesis of CuO-TiO2 catalyst with high dispersion CuO species for selective catalytic oxidation of NO. Appl. Surf. Sci. 2017, 411, 227–234. [Google Scholar] [CrossRef]
  27. Wang, X.; Yu, Y.; He, H. Effects of temperature and reductant type on the process of NOx storage reduction over Pt/Ba/CeO2 catalysts. Appl. Catal. B Environ. 2011, 104, 151–160. [Google Scholar] [CrossRef]
  28. Cai, H.; Zhang, X.; Wang, P.; Li, K.; Shao, H.; Liu, G.; Qiao, G. Influence of Interaction Components on the NOx Adsorption Properties of Cu-Ba-Ce Composite Catalysts. J. Nanosci. Nanotechnol. 2019, 19, 4039–4045. [Google Scholar] [CrossRef]
  29. Piacentini, M.; Maciejewski, M.; Baiker, A. Pt-Ba/alumina NOx storage-reduction catalysts: Effect of Ba-loading on build-up, stability and reactivity of Ba-containing phases. Appl. Catal. B Environ. 2005, 59, 187–195. [Google Scholar] [CrossRef]
  30. Eberhardt, M.; Riedel, R.; Göbel, U.; Theis, J.; Lox, E.S. Fundamental investigations of thermal aging phenomena of model NOx storage systems. Top. Catal. 2004, 30–31, 135–142. [Google Scholar] [CrossRef]
  31. Shi, C.; Xu, L.; Zhu, A.; Zhang, Y.; Au, C.T. Copper Oxide Clusters Stabilized by Ceria for CO, C3H6, and NO Abatement. Chin. J. Catal. 2012, 33, 1455–1462. [Google Scholar] [CrossRef]
  32. Turco, M.; Bagnasco, G.; Costantino, U.; Marmottini, F.; Montanari, T.; Ramis, G.; Busca, G. Production of hydrogen from oxidative steam reforming of methanol. J. Catal. 2004, 228, 56–65. [Google Scholar] [CrossRef]
  33. Velu, S.; Suzuki, K.; Okazaki, M.; Kapoor, M.P.; Osaki, T.; Ohashi, F. Oxidative Steam Reforming of Methanol over CuZnAl(Zr)-Oxide Catalysts for the Selective Production of Hydrogen for Fuel Cells: Catalyst Characterization and Performance Evaluation. J. Catal. 2000, 194, 373–384. [Google Scholar] [CrossRef]
  34. Wang, Z.; Sun, X.; Liu, J.; Li, X. The NO oxidation performance over Cu/Ce0.8Zr0.2O2 catalyst. Surf. Interfaces 2017, 6, 103–109. [Google Scholar] [CrossRef]
  35. Yin, M.; Wu, C.-K.; Lou, Y.; Burda, C.; Koberstein, J.T.; Zhu, Y.; O’Brien, S. Copper Oxide Nanocrystals. J. Am. Chem. Soc. 2005, 127, 9506–9511. [Google Scholar] [CrossRef]
  36. Aguila, G.; Guerrero, S.; Araya, P. Effect of the preparation method and calcination temperature on the oxidation activity of CO at low temperature on CuO–CeO2/SiO2 catalysts. Appl. Catal. A Gen. 2013, 462–463, 56–63. [Google Scholar] [CrossRef]
  37. Yz, A.; Long, L.A.; Zc, A.; Jw, A.; Wen, Z.A.; Sz, A.; Mfa, B.; Lca, B.; Dya, B. Highly efficient Cu/CeO2-hollow nanospheres catalyst for the reverse water-gas shift reaction: Investigation on the role of oxygen vacancies through in situ UV-Raman and DRIFTS. Appl. Surf. Sci. 2020, 516, 146035. [Google Scholar] [CrossRef]
  38. Kobayashi, T.; Yamada, T.; Kayano, K. Effect of basic metal additives on NOx reduction property of Pd-based three-way catalyst. Appl. Catal. B Environ. 2001, 30, 287–292. [Google Scholar] [CrossRef]
  39. Liu, S.; Liu, J.; Lin, Q.; Xu, S.; Wang, J.; Xu, H.; Chen, Y. Solvent Effects on the Low-Temperature NH3–SCR Activity and Hydrothermal Stability of WO3/SiO2@CeZrOx Catalyst. ACS Sustain. Chem. Eng. 2020, 8, 13418–13429. [Google Scholar] [CrossRef]
  40. Li, F.; Lei, L.; Yi, J.; Dou, C.; Meng, Z.; Wang, P. Performance, Structure and Mechanisms of Pd Catalyst Supported on M-Doped (M=La, Ba and K) CeO2 for Methane Oxidation. Catal. Lett. 2022, 153, 1847–1858. [Google Scholar] [CrossRef]
  41. Liu, S.; Yao, P.; Pei, M.M.; Lin, Q.J.; Xu, S.H.; Wang, J.L.; Xu, H.D.; Chen, Y.Q. Significant differences of NH3-SCR performances between monoclinic and hexagonal WO3 on Ce-based catalysts. Environ. Sci. Nano 2021, 8, 2988–3000. [Google Scholar] [CrossRef]
  42. Lu, Y.; Duan, L.; Sun, Z.; Chen, J. Flame spray pyrolysis synthesized CuO–CeO2 composite for catalytic combustion of C3H6. Proc. Combust. Inst. 2021, 38, 6513–6520. [Google Scholar] [CrossRef]
  43. Mihaylov, M.Y.; Ivanova, E.Z.; Aleksandrov, H.A.; Petkov, P.S.; Vayssilov, G.N.; Hadjiivanov, K.I. Species formed during NO adsorption and NO+O2 co-adsorption on ceria: A combined FTIR and DFT study. Mol. Catal. 2018, 451, 114–124. [Google Scholar] [CrossRef]
  44. Mihaylov, M.; Zdravkova, V.; Ivanova, E.; Aleksandrov, H.; Petkov, P.; Vayssilov, G.; Hadjiivanov, K. Infrared spectra of surface nitrates: Revision of the current opinions based on the case study of ceria. J. Catal. 2021, 394, 245–258. [Google Scholar] [CrossRef]
  45. Mym, A.; Ezi, A.; Gnv, B.; Kiha, C. Revisiting ceria-NOx interaction: FTIR studies. Catal. Today 2020, 357, 613–620. [Google Scholar] [CrossRef]
  46. Symalla, M.O.; Drochner, A.; Vogel, H.; Philipp, S.; Gbel, U.; Müller, W. IR-study of formation of nitrite and nitrate during NOx-adsorption on NSR-catalysts-compounds CeO2 and BaO/CeO2. Top. Catal. 2007, 42–43, 199–202. [Google Scholar] [CrossRef]
  47. Wang, L.; Ran, R.; Wu, X.; Li, M.; Weng, D. In situ DRIFTS study of NOx adsorption behavior on Ba/CeO2 catalysts. J. Rare Earths 2013, 31, 1074–1080. [Google Scholar] [CrossRef]
  48. Lin, Q.; Xu, S.; Zhao, H.; Liu, S.; Xu, H.; Dan, Y.; Chen, Y. Highlights on Key Roles of Y on the Hydrothermal Stability at 900 °C of Cu/SSZ-39 for NH3-SCR. ACS Catal. 2022, 12, 14026–14039. [Google Scholar] [CrossRef]
Figure 1. (a) NOx adsorption curves; (b) NSE and complete storage time of different adsorbents at 80 °C for 10 min. Feed: 500 ppm NO, 8 vol.% O2 and N2 balance.
Figure 1. (a) NOx adsorption curves; (b) NSE and complete storage time of different adsorbents at 80 °C for 10 min. Feed: 500 ppm NO, 8 vol.% O2 and N2 balance.
Catalysts 13 01180 g001
Figure 2. (a) NOx desorption curves after NOx storage at 80 °C for 10 min; (b) NOx desorption amounts in two temperature ranges: 80–280 °C and 280–550 °C.
Figure 2. (a) NOx desorption curves after NOx storage at 80 °C for 10 min; (b) NOx desorption amounts in two temperature ranges: 80–280 °C and 280–550 °C.
Catalysts 13 01180 g002
Figure 3. XRD patterns of the adsorbents.
Figure 3. XRD patterns of the adsorbents.
Catalysts 13 01180 g003
Figure 4. UV-Vis DRS of Cu in CuxCe (a) and CuxBa5Ce (b) adsorbents.
Figure 4. UV-Vis DRS of Cu in CuxCe (a) and CuxBa5Ce (b) adsorbents.
Catalysts 13 01180 g004
Figure 5. H2-TPR curves over the samples.
Figure 5. H2-TPR curves over the samples.
Catalysts 13 01180 g005
Figure 6. XPS spectra of (a) Cu 2p, (b) O 1 s Ce 3d, and (c) Ce 3d for the adsorbents.
Figure 6. XPS spectra of (a) Cu 2p, (b) O 1 s Ce 3d, and (c) Ce 3d for the adsorbents.
Catalysts 13 01180 g006
Figure 7. In situ DRIFTs of NOx adsorption over CuxCe adsorbents. Adsorption conditions: 2 vol.% NO-98 vol.% N2/5 vol.% O2-95 vol.% N2 at 80 °C for 30 min.
Figure 7. In situ DRIFTs of NOx adsorption over CuxCe adsorbents. Adsorption conditions: 2 vol.% NO-98 vol.% N2/5 vol.% O2-95 vol.% N2 at 80 °C for 30 min.
Catalysts 13 01180 g007
Figure 8. In situ DRIFTs of NOx adsorption over CuxBa5Ce adsorbents. Adsorption conditions: 2 vol.% NO-98 vol.% N2/5 vol.% O2-95 vol.% N2 at 80 °C for 30 min.
Figure 8. In situ DRIFTs of NOx adsorption over CuxBa5Ce adsorbents. Adsorption conditions: 2 vol.% NO-98 vol.% N2/5 vol.% O2-95 vol.% N2 at 80 °C for 30 min.
Catalysts 13 01180 g008
Figure 9. Procedures for testing NO adsorption/desorption performance on adsorbents.
Figure 9. Procedures for testing NO adsorption/desorption performance on adsorbents.
Catalysts 13 01180 g009
Table 1. The main properties of studied samples.
Table 1. The main properties of studied samples.
SamplesSBET (m2/g)Total Pore Volume (mL/g)Average Pore Radius (nm)Crystalline Size of CeO2 a (nm)
CeO2152.30.253.37.6
Cu1Ce143.60.213.17.6
Cu3Ce136.90.223.27.6
Cu5Ce133.90.233.47.6
Ba5Ce119.80.223.37.6
Cu1Ba5Ce116.10.203.57.7
Cu3Ba5Ce104.70.193.77.8
Cu5Ba5Ce104.10.173.37.7
a Crystallite size of CeO2 determined from the XRD diffraction peak by Scherrer’s equation.
Table 2. Surface compositions of the adsorbents.
Table 2. Surface compositions of the adsorbents.
SamplesCu2+/CuOα/OCe3+/CeO/CeCu/CeBa/CeCu/BaBa 3d a
CeO2-26.519.54.82
Cu1Ce71.626.216.84.77 0.24
Cu3Ce76.225.815.45.27 0.36
Cu5Ce81.725.614.25.29 0.46
Ba5Ce-39.3 19.1 5.52 0.31 3.51
Cu1Ba5Ce61.938.9 15.85.78 0.27 0.29 0.86 3.09
Cu3Ba5Ce66.237.9 14.45.88 0.32 0.25 1.30 2.54
Cu5Ba5Ce71.237.9 13.46.11 0.39 0.25 1.53 2.49
a Surface composition (at.%).
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

Pei, M.; Fan, Y.; Xu, H.; Lian, Z.; Tan, W.; Wang, J.; Chen, Y. Revealing the Roles of Cu/Ba on Ce-Based Passive NOx Adsorbers. Catalysts 2023, 13, 1180. https://doi.org/10.3390/catal13081180

AMA Style

Pei M, Fan Y, Xu H, Lian Z, Tan W, Wang J, Chen Y. Revealing the Roles of Cu/Ba on Ce-Based Passive NOx Adsorbers. Catalysts. 2023; 13(8):1180. https://doi.org/10.3390/catal13081180

Chicago/Turabian Style

Pei, Mingming, Yuxin Fan, Haidi Xu, Zhihua Lian, Wei Tan, Jianli Wang, and Yaoqiang Chen. 2023. "Revealing the Roles of Cu/Ba on Ce-Based Passive NOx Adsorbers" Catalysts 13, no. 8: 1180. https://doi.org/10.3390/catal13081180

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