Next Article in Journal
Hyperthermophilic Composting Technology for Organic Solid Waste Treatment: Recent Research Advances and Trends
Next Article in Special Issue
Effects of Ignition Timing on Combustion Characteristics of a Gasoline Direct Injection Engine with Added Compressed Natural Gas under Partial Load Conditions
Previous Article in Journal
Simultaneous Removal of Trivalent Arsenic and Nitrate Using Microbial Fuel Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Evaluation on the Catalytic Activity of a Novel CeZrK/rGO Nanocomposite for Soot Oxidation in Catalyzed Diesel Particulate Filter

1
College of Automotive and Mechanical Engineering, Changsha University of Science and Technology, Changsha 410114, China
2
School of Energy Science and Engineering, Central South University, Changsha 410083, China
*
Author to whom correspondence should be addressed.
Processes 2021, 9(4), 674; https://doi.org/10.3390/pr9040674
Submission received: 15 March 2021 / Revised: 1 April 2021 / Accepted: 8 April 2021 / Published: 12 April 2021
(This article belongs to the Special Issue Clean Combustion and Emission in Vehicle Power System)

Abstract

:
A nanostructured solid solution catalyst CeZrK/rGO for soot oxidation in catalyzed diesel particulate filter was synthesized using the dipping method. The reduced graphene oxide (rGO) was used as the catalyst carrier, and CeO2, ZrO2, and K2O were mixed with the molar ratio of 5:1:1, 5:2:2 and 5:3:3, which were referred to as Ce5Zr1K1/rGO, Ce5Zr2K2/rGO, and Ce5Zr3K3/rGO, respectively. The structure, morphology and catalytic activity of the CeZrK/rGO nanocomposites were thoroughly investigated and the results show that the CeZrK/rGO nanocomposites have nanoscale pore structure (36.1–36.9 nm), high-dispersion quality, large specific surface area (117.2–152.4 m2/g), small crystallite size (6.7–8.3 nm), abundant oxygen vacancies and superior redox capacity. The 50% soot conversion temperatures of Ce5Zr1K1/rGO, Ce5Zr2K2/rGO, and Ce5Zr3K3/rGO under tight contact condition were decreased to 352 °C, 339 °C and 358 °C respectively. The high catalytic activity of CeZrK/rGO nanocomposites can be ascribed to the following factors: the doping of Zr and K ions causes the nanocrystalline phase formation in CeZrK solid solutions, reduces the crystallite size, generates abundant oxygen vacancies and improves redox capacity; the rGO as a carrier provides a large specific surface area, thereby improving the contact between soot and catalyst.

1. Introduction

The diesel engine has been widely used in transportation and various non-road construction machinery due to a high thermal efficiency [1]. However, its exhausted emissions can cause serious environmental pollution and endanger human health at the same time, especially for soot which can enter the blood, trigger genetic mutations and create health risks [2,3]. In recent years, with the increasingly stringent emission regulations, the control technologies of the soot emission in diesel engines have become the focus of governments and scholars of various countries [4].
In the current scenario, soot is mainly caused by the incomplete combustion of hydrocarbon compounds inside the engine under high temperature and oxygen deficiency. In the post-treatment system, it gets into contact with the catalyst and reacts at high temperatures to form CO2. Diesel particulate filter (DPF) is an effective tool to control soot from diesel engine. Generally, the exhaust temperature ranges from 150 °C to 450 °C, while the oxidation temperature of soot exceeds 550 °C. It is, therefore, necessary to raise the temperature for soot combustion, but excessive soot combustion temperature will damage the filter material [5]. Therefore, a superior solution is to reduce the oxidation reaction temperature of soot using catalysts. Catalyzed diesel particulate filter (CDPF) does not need an extra heat source and complex control system, and thus it greatly reduces fuel consumption and engine cost. As the catalytic oxidation of soot is a reaction between O2, soot and the catalyst, the contact between soot and the catalyst and the activity of the catalyst are two key factors affecting the soot oxidation efficiency [6], which is a critical issue in CDPF regeneration. Accordingly, the design and preparation of catalysts with large contact area and high intrinsic activity are significant for the CDPF.
The noble metal catalysts have a superior intrinsic catalytic activity, but the price is too expensive, limiting their applications in CDPF [7]. It was established that the oxides of rare earth metal, like Ce0.5Zr0.5O2 [8], LaFeO3 [9], LaMn0.8Pt0.2O3 [10], can catalyze the oxidation of soot. Among them, CeO2 presents superior oxygen storage capacity (OSC) and catalytic performance over other traditional metal oxide catalysts in oxidizing soot at low temperature [11]. The introduction of external cations can cause the lattice distortion of CeO2 and increase oxygen vacancies, improving the redox capacity and catalytic activity of CeO2 [12]. It was reported by Atribak et al. [13] and Gross et al. [14] that the addition of Zr and K ions can enhance high-temperature stability and oxygen storage capacity of the catalysts, respectively. Thus, it can be inferred that the doping of these two elements in the catalyst can improve soot oxidation.
It is important to select a suitable substrate to increase the surface area, improve adsorption and provide more active sites on the catalyst. Graphene exists as a 2D nanomaterial with a honeycomb hexagonal structure composed of Sp2 hybridized carbon atoms (Olabi et al., 2020). Due to excellent physical and chemical properties, graphene has attracted extensive attention. The super-large specific surface area and special pore structure of graphene enable the metal nanoparticles loaded onto graphene to get a good dispersion, thus ensuring sufficient contact area between the active components of the catalyst and soot. Tan et al. [15] found that the maximum adsorption capacity of NiAl-LDH/rGO composite with a sandwich structure toward uranium is up to 277.8 mg/g. The high adsorption capacity has to do with the specific surface area and special pore structure characteristics of rGO. Due to the negatively charged nature of graphene and positively charged nature of the metal cation, the combination between rGO and metal oxide particles can be easily achieved by electrostatic interaction, which is conducive to the uniform dispersion of metal oxide particles on rGO [16]. Some researchers have pointed out that graphene-based materials can stabilize active sites and inhibit surface migration of metal nanoparticles, guaranteeing high catalytic stability even under severe operating conditions [17,18,19]. Therefore, the usage of graphene as the matrix of catalysts is a feasible strategy.
Still, the weak thermal-oxidative stability of bare graphene makes it easy to decompose during combustion. This impact is negative on the catalytic activity. Consequently, intellectual capital and resources are being committed to upturn the thermal stability of graphene by surface modification. Generally speaking, modified graphene is to compound other substances with active components through covalent or non-covalent interactions so that graphene has other special properties while maintaining its original properties [20]. Li et al. [21] portrayed that zirconium phosphate nanoparticles immobilized on the surface of graphene oxide (GO) yielded superior thermal stability and flame retardancy to the phenolic foams, and this is essentially ascribed to the harmonious impacts of zirconium phosphate nanoparticles and GO. Wu et al. [22] prepared graphene sheets using the hydrogen arc discharge exfoliation method. The materials exhibit a remarkable thermal stability from GO with an oxidization resistance temperature of 601 °C. The research performed by Bora et al. [23], Yuan et al. [24] and Zong et al. [25] found that rGO exhibited a higher thermal stability than GO due to the removal of thermally labile oxygen functional groups during the reduction process. Salavagione et al. [4] report a strategy for the synthesis of vinylalcohol/rGO nanocomposites and found the thermal stability was improved by more than 100 °C.
This study aims to develop a catalyst with high catalytic activity, i.e., CeO2-ZrO2-K2O/rGO (referred to as CeZrK/rGO), for soot oxidation in the CDPF. The structure and morphology of the catalyst were thoroughly characterized using field emission scanning electron microscopy (FESEM), transmission electron microscopy (TEM), N2 adsorption-desorption tests, X-ray powder diffraction (XRD), X-ray photoelectron spectroscopy (XPS), Fourier transform infrared spectrometry (FTIR), Raman spectra and H2-temperature programmed reduction (TPR) methods. The catalytic performance of the catalyst was evaluated using thermogravimetric analysis.

2. Experimental Methods

2.1. Catalyst Preparation

The schematic diagram of the catalyst’s preparation process is shown in Figure 1. All materials were of analytical grade and used without further purification. Graphene oxide was prepared by the Hummer method in the oxidation-reduction method [26]. Under low temperature stirring conditions, some amount of concentrated sulfuric acid and strong oxidant were sequentially added to the graphite powder, diluted with water and hydrogen peroxide, heated and dried, and then subjected to ultrasound exfoliation to obtain graphene oxide. Ce(NO3)3·6H2O, ZrO(NO3)2·6H2O, and KNO3 were dissolved in deionized water by magnetic stirring to generate the salt solution containing the metallic ions of Ce, Zr, and K. Then, 60 mg of GO was put into 30 mL of deionized water to generate the GO solution under the ultrasonic treatment for 0.5 h. The prepared solutions were mixed and then processed by the ultrasonic treatment for 0.5 h. The mixed solution was heated in a water bath at 80 °C for 1 h. Ammonia hydroxide was added into the heated solution and gently stirred until pH 10 ± 0.1 was reached. The obtained solution was hydrothermally treated at 160 °C for 12 h. After that, the solution was washed with deionized water by repeated centrifugations until the pH was adjusted to neutral. The solution was further freeze-dried for 18 h in a FD-1A-80 vacuum freeze drier to obtain the precursor. It was calcined at 200 °C for 2 h in a tube furnace under argon atmosphere to realize CeZrK/rGO.

2.2. Experimental Apparatus

The morphology and composition of catalysts were investigated by FESEM (Czech Nova NanoSEM 230, FEI Company, Hillsboro, OR, USA) equipped with energy dispersive X-ray (EDX). The TEM images were acquired by an FEI Tecnai G2 F20 at an accelerating voltage of 200 kV. The crystal structure of the catalysts was observed using a Rigaku D/Max 2550 XRD (Tokyo, Japan). The static adsorption and desorption tests were carried out on a Micromeritics ASAP2020 to obtain N2 adsorption/desorption isotherms. The specific surface area and the pore size distribution were calculated from the adsorption branch and desorption branch of the isomers based on Brunauer-Emmett-Teller (BET) and Barrett-Joyner-Halenda (BJH) methods, respectively. The XPS spectra was recorded on an ESCALAB 250Xi XPS in the analysis chamber of 2 × 10−9 Pa using Al Kα radiation (1486.6 eV), and the obtained binding energies were referenced to C 1s line set at 284.9 eV. The FT-IR spectra was obtained on a Nicolet iS50 FT-IR (Thermo Fisher Scientific, Waltham, MA, USA) using potassium bromide (KBr) beamsplitter. The Raman spectra were measured using a Horiba Jobin-Yvon HR800 Raman spectrometer (Kyoto, Japan) with a 532 nm line of Ar+ ion laser as the excitation source at room temperature. The TG 209 F3 Tarsus (Netzsch, Selb, Germany) was used to carry out the thermogravimetric analysis, under the temperature range between room temperature and 900 °C with a heating rate of 10 °C·min−1. The TPR measurements were carried out on an Auto Chem II 2920 instrument equipped with a thermal conductivity detector (TCD). Each time, 40 mg of the sample was heated from room temperature to 800 °C at a heating rate of 10 °C·min−1. A mixed gas consisting of 10% H2 in Ar at a fixed flow rate of 20 mL·min−1 was used as reductant.
In the experiment, the soot and the cerium-based catalyst were mixed in loose contact and close contact, respectively. In close contact conditions, the soot particles and the catalyst powder were mixed and ground in an agate mortar for 30 min. In loose contact conditions, a stainless-steel spatula is used to uniformly mix soot particles with the same mass ratio and catalyst powder for 3–4 min. Under the two working conditions, the masses of soot and cerium-based catalyst in the crucible were 4 mg and 40 mg. A thermogravimetry technique using Printex-U soot (Degussa, Frankfurt, Germany) as the model reactant was utilized to assess the catalytic soot activity of the catalysts. The catalytic activity of the catalysts for diesel soot combustion were realized using simulated air (21% O2 + 79% N2) or pure N2 under different contact modes. The thermogravimetric analysis was performed in a TG 209 F3 Tarsus under the temperature range between room temperature and 650 °C with a heating rate of 10 °C·min−1.

3. Results and Discussion

3.1. Characterization Studies

The SEM and TEM techniques were employed to characterize the morphology and size of the catalysts as shown in Figure 2a,c,e and Figure 2b,d,f, respectively. It can be seen from Figure 2a that the substrate material of graphene exists in the form of a translucent sheet with wrinkles and folds. Looking at Figure 2c,e, it is apparent that metal oxide particles are uniformly and highly dispersed on graphene. It can be found from TEM images that Ce5Zr2K2/rGO presents the most evenly dispersed metal oxide particles compared to Ce5Zr1K1/rGO and Ce5Zr3K3/rGO. By using Nano measurer software [27], the average particle size of Ce5Zr1K1/rGO, Ce5Zr2K2/rGO, and Ce5Zr3K3/rGO were 9.65 nm, 7.42 nm, and 9.39 nm, respectively, which were smaller than that of CeZrOx catalysts (15–20 nm) [28]. It has been shown that graphene inhibits the surface migration of metal oxide nanoparticles and reduces the size of metal oxide nanoparticles [29].
To obtain the distributed information of Ce, Zr, and K elements in the Ce5Zr2K2/rGO catalysts, the EDX elemental mapping is shown in Figure 3. The catalysts presented a uniform and orderly element distribution, and elements Ce, Zr, K, C, and O were highly dispersed on the catalysts [19]. It can be further verified in TEM images that the rGO carrier was loaded with well-dispersed spherical metal oxide nanoparticles. The catalytic oxidation of soot belonged to the heterogeneous catalytic reaction, and thus the contact between soot and catalyst was one of the most important influencing factors on the catalytic activity and efficiency. Therefore, the well-dispersed metal oxide nanoparticles promoted the catalytic oxidation of soot.
The XRD patterns of CeO2 and Ce5Zr2K2/rGO catalysts are depicted in Figure 4. It was observed that all the catalysts show a typical cubic structured CeO2 peak. The XRD of CeZrK/rGO catalysts showed a series of reflections at 29° (111), 33° (200), 47° (220), 56° (331), 59° (222) and 69° (400) indexed to typical cubic spinel CeO2 (PDF#89-8436). The doping with Zr4+ and K+ irons caused the shifting of diffraction peak towards lower angle side since Ce4+ was partially replaced by Zr4+ and K+ ions, which led to the slight changes in cell parameter and crystallite size [30,31]. The peaks of CeZrK/rGO catalysts were wider than those of CeO2. Meanwhile, the half-peak width of Ce5Zr2K2/rGO was relatively wide among the CeZrK/rGO catalysts, indicating the crystallite size of Ce5Zr2K2/rGO was smaller than that of Ce5Zr1K1/rGO and Ce5Zr3K3/rGO. There were no obvious characteristic diffraction peaks observed for graphene-related phases. The complete exfoliation of graphene in the hybrid was achieved as a result of efficiently avoiding the restacking of the graphene [19]. Meanwhile, no other obvious diffractions were found in Zr-related and K-related phases, indicating that the Zr-related and K-related phases were non-crystalline states or presented as a part of CeO2-ZrO2-K2O solid solution.
As adsorption and diffusion are the critical processes in a heterogeneous catalytic reaction, the pore structure of the catalyst has a great impact on catalytic performance [32]. Figure 5 displays the N2 adsorption isotherms and pore size distribution of the as-prepared catalysts. A similar IV type isotherm with H3-type hysteresis loops on IUPAC classification suggested that the CeZrK/rGO catalysts had mesoporous distribution characteristic [33]. From Figure 5b,d,f, it can be seen that the most probable pore size of Ce5Zr1K1/rGO, Ce5Zr2K2/rGO, and Ce5Zr3K3/rGO were 36.9 nm, 36.8 nm, and 36.1 nm, respectively, which were larger than that of soot (>20 nm). At this time, most of the soot particles stayed on the external surface, and some entered the pores, which increased the contact area between the catalyst and the PM to a certain extent, thereby improving the catalytic activity.
Table 1 lists the average crystallite size and the specific surface area (SBET) of the catalysts. The average crystallite sizes was computed from XRD based on the Scherrer equation. The average crystallite size of Ce5Zr1K1/rGO, Ce5Zr2K2/rGO, and Ce5Zr3K3/rGO were 8.3 nm, 6.9 nm, and 6.7 nm, respectively, which were smaller than that of CeO2 because the lattice distortion caused by the doping of Zr and K ions inhibited the crystal growth of CeO2 phase [34]. The SBET can be obtained from N2 adsorption isotherms based on the multi-point BET. The as-prepared catalysts had a higher specific surface area compared to that of CeZrK catalysts (17 m2/g) in Ref. [35] and other Ce-based catalysts (55–82 m2/g) in Ref. [36]. Meanwhile, among the CeZrK/rGO catalysts, the specific surface area of Ce5Zr2K2/rGO (151.5 m2/g) and Ce5Zr3K3/rGO (152.4 m2/g) were slightly larger than that of Ce5Zr1K1/rGO (117.2 m2/g). It has proved that graphene-based catalysts have a large surface area and superior pore structure, which effectively guarantees the high catalytic activity of the CeZrK/rGO catalysts [37,38].
Figure 6 presents the Raman profiles of CeZrK/rGO catalysts and GO. The typical Raman peaks of D band and G band appeared around 1350 cm−1 and 1594 cm−1, respectively. The D band represents the structural defect caused by oxygen-containing functional groups on the C base, and the G band represents the E2g symmetric mode of sp2 hybrid C [39]. For the Raman profiles of the CeZrK/rGO catalysts, no bands occurred at 400~500 cm−1 assigned to CeO2 [29], showing good dispersion of the CeO2 particle on the graphene layers. The intensity ratio of D band to G band (ID/IG) is an indicator of disorder degree and the average size of the in-plane sp2 regions in graphite [39]. The ID/IG for Ce5Zr1K1/rGO, Ce5Zr2K2/rGO, Ce5Zr3K3/rGO, and GO were 1.10, 1.28, 1.21, and 0.99, respectively. The higher ID/IG means that there may be smaller sp2 regions and unrepaired defect sites in Ce5Zr2K2/rGO. The defect sites on the surface of rGO can efficiently inhibit the agglomeration of metal oxide nanoparticles to reduce the particle size and improve the dispersion characteristics [40], which is consistent with the above results observed in SEM and TEM.
The FTIR spectra of CeZrK/rGO catalysts and GO shown in Figure 7 reveals numerous O-containing functional groups such as (O-H) (3423 cm−1) and (C=O) (1735 cm−1) of COOH, (O-H) of tertiary C-OH (1384 cm−1), and (H2O) (1637 cm−1) [41]. Just on these O-containing groups of GO, during the synthesis of the catalyst, the mixed metal ions (Ce4+, Zr4+, K+) are first adsorbed uniformly on the surface of negatively charged GO sheets via electrostatic attractions. Compared to GO, the FTIR spectra of CeZrK/rGO catalysts was greatly reduced or even disappeared at 1735 cm−1 and 1384 cm−1, implying that GO was deoxidized and reduced to rGO, which was consistent with the phenomenon reported in Ref. [42]. Moreover, the infrared peak changed from 1637 cm−1 to 1654 cm−1 and 1560 cm−1 due to the stretching vibration of C=C. The band at 1228 cm−1 emerged from C-O-C stretch epoxy bands on rGO surface.
The oxidation states of the main elements Ce, Zr, O and C, including an estimate of the atomic ratio in the CeZrK/rGO catalysts, were achieved using the XPS measurement technique and the corresponding result is depicted in Figure 8. The atomic surface concentrations, the relative percentages, and the binding energy of Ce, Zr, O and C elements in the CeZrK/rGO catalysts are summarized in Table 2 and Table 3, respectively. The u and v in the XPS spectra of Ce 3d refer to the 3d3/2 and 3d5/2 spin-orbit components, respectively [43]. The peaks of u, u″, u‴ and v, v″, v‴ are attributed to Ce4+ ions, whereas u′ and v′ are attributed to Ce3+ ions. Therefore, it can be inferred that the CeZrK/rGO catalysts had both Ce4+ and Ce3+ ions. The Ce3+/Ce4+ ratios of the CeZrK/rGO catalysts were calculated by the ratio of the sum of the integrated areas of Ce3+ (u′, v′) to the sum of the integrated area of Ce4+ (u, v, u″, v″, u‴, v‴) as shown in Table 4. It was found that Ce5Zr2K2/rGO exhibits a higher concentration of Ce3+ ions, which indicates that Ce5Zr2K2/rGO contained more oxygen vacancies [29]. The XPS spectra of Zr 3d in Figure 8b shows that there is no difference in the binding energies of Zr element in the CeZrK/rGO catalysts. The binding energies of Zr 3d5/2 and Zr 3d3/2 were 182.6 eV and 185 eV, respectively, which can prove the existence of Zr+4 in the CeZrK/rGO catalysts [44]. The XPS spectra of O 1s in Figure 8c presents two types of surface oxygen in the CeZrK/rGO catalysts. The OI (529.8–530.0 eV) is the characteristic of the surface lattice oxygen (O2−) in cerium oxide, and the OII (531.5-531.8 eV) indicates the presence of chemisorbed oxygen on the surface of oxygen (O, O2−, O22−) belonging to the defect-oxygen or hydroxyl-like group [19]. In general, the surface-adsorbed oxygen was derived from the adsorption of gaseous O2 on the oxygen vacancies of the oxide catalyst. It is clear that Ce5Zr2K2/rGO had a much higher OII/OI peak area ratio than Ce5Zr1K1/rGO and Ce5Zr3K3/rGO, which effectively increased the concentration of surface-adsorbed oxygen. Compared to the lattice oxygen, chemical adsorbed oxygen (O) had higher mobility, and thus it played a more important role in the soot oxidation reaction. Figure 8d displays the deconvolution of C 1s peaks in the CeZrK/rGO catalysts. The main peaks are centred in 284.5–284.6 eV, 285.5 eV, and 288.4 eV, which represents graphite structure, C-O, and -O-C=O, respectively. The C-O and -O-C=O bonds in graphene indicate that graphene can provide abundant active sites for the directional connection of CeO2-ZrO2 particles. Additionally, the XPS failed to detect the presence of K element. It can be explained that the binding energy of K and C were mainly in the range of 291–296 eV and of 277–295 eV, respectively, and the content of C in the as-prepared catalysts was much larger than that of K, which may have interfered with the detection of K element. Generally, among the three CeZrK/rGO catalysts, the Ce5Zr2K2/rGO catalyst contained more oxygen vacancies which is beneficial to improve the OSC of CeO2, and the Ce5Zr2K2/rGO catalyst had a higher chemically adsorbed oxygen concentration.
Figure 9 shows the result of the H2-TPR analyses administered on GO and CeZrK/rGO catalysts to examine their redox behavior. It was shown that the reduction peak of GO appears at 663 °C, while no obvious reduction peak was observed for the CeZrK/rGO catalysts at 550–650 °C, indicating that the oxygen-containing functional groups on the GO surface are deoxidized during the preparation [45], which agrees with the FTIR results shown in Figure 7. Three obvious reduction peaks ranging from 250 °C to 600 °C were observed for the CeZrK/rGO catalysts. The shoulder peak at 360 °C corresponds to adsorbent oxygen, and the reduction peak at 465 °C corresponds to the reduction of the outermost layer of Ce4+ to Ce3+. The reduction peak ranging from 700 °C to 850 °C can be attributed to the reduction of CeO2 (Ce4+ inner layer) and the reduction of lattice oxygen (bulk reduction) [46]. Curve b has a higher peak intensity ratio (surface/bulk reduction) than that of other curves, which indicates the enhanced oxygen mobility within the lattice of Ce5Zr2K2/rGO. Due to the synergistic effect of Ce and Zr, a shoulder peak occurs around 301 °C. Moreover, the peak at around 465 °C in curve b moves to low temperature, indicating that the reducibility of Ce5Zr2K2/rGO improved. The reduction peak in curve c becomes flat due to the aggregation of excess metal oxides, which could impact negatively on the catalytic performance of Ce5Zr3K3/rGO. In general, the doping of Zr and K ions into CeO2 with graphene as the carrier can effectively promoted the formation of a solid solution. The graphene can enhance the redox capacity of cerium oxide at low temperatures [47]. All the CeZrK/rGO catalysts showed a superior redox capacity, and this reflects on the improved catalytic oxidation activity of the soot. Meanwhile, the redox capacity of the CeZrK/rGO catalysts was affected by the metal doping ratio. Among the CeZrK/rGO catalysts, Ce5Zr2K2/rGO presents the lowest reduction temperature in the H2-TPR profiles.

3.2. Evaluation on Catalytic Activity for Soot Oxidation

The catalytic activity of the CeZrK/rGO catalysts for soot oxidation were evaluated under the tight contact and loose contact modes in simulated air (21% O2 + 79% N2) by thermogravimetric analysis. Figure 10 shows the normalized soot conversion under the tight contact mode as the function of temperature over the as-prepared catalysts. To provide a better comparison for the catalytic activity, T50 was defined as the temperature at which 50% soot conversion is achieved and Tm was defined as the temperature at which the maximum rate of soot combustion occurs [48]. The results show that the T50 of Ce5Zr1K1/rGO, Ce5Zr2K2/rGO, Ce5Zr3K3/rGO, and CeO2 were 352 °C, 339 °C, 358 °C, and 407 °C, respectively. It can be inferred that the CeZrK/rGO catalysts had higher catalytic activity than CeO2. The difference of T50 among the CeZrK/rGO catalysts was small due to the tight contact between soot and the catalyst. Under this condition, the difference in the structure and morphology of the CeZrK/rGO catalysts had a small impact on the catalytic activity.
The loose contact mode is the most realistic approach because soot is mixed with the catalyst without any force exertion [49]. Figure 11 shows the normalized soot conversion under the loose contact mode as a function of temperature with and without as-prepared catalysts. The T50 for Ce5Zr1K1/rGO, Ce5Zr2K2/rGO, and Ce5Zr3K3/rGO were 390 °C, 383 °C, and 432 °C, respectively, which were far less than those of CeO2 (523 °C), rGO (532 °C) and catalyst-free (604 °C). The catalytic activity of rGO on soot oxidation was presented due to the fact that rGO contains residual structural defects, which improve oxygen reduction reactions. The higher catalytic activity of Ce5Zr2K2/rGO was closely related to its superior dispersion quality, small particle size (Figure 2d), large specific surface area (Table 1), and abundant oxygen vacancies (Figure 8 and Table 3) and redox capacity (Figure 9). It should be noted that the slope of the curve before 300 °C was arranged in the order of Ce5Zr3K3/rGO > Ce5Zr2K2/rGO > Ce5Zr1K1/rGO, which is consistent with the order of the specific surface area (Table 1) of the three CeZrK/rGO catalysts. The slope of the curve after 300 °C was arranged in the order of Ce5Zr2K2/rGO > Ce5Zr1K1/rGO > Ce5Zr3K3/rGO, which is consistent with the order of oxygen vacancy concentration (Ce3+/Ce4+) and the concentration of chemically adsorbed oxygen (O/O) (Table 4). This may be due to the fact that the specific surface area had a greater influence on the activity in the low temperature region, while the oxygen vacancy concentration had a greater influence on the activity in the medium and high temperature region. In order to avoid the influence of graphene on experimental data, the weight loss of modified graphene under air atmosphere was investigated. As shown in Figure 12, the weight loss was 2.3 (wt)% and mainly occurred at 520 °C, which had little impact on the TG experimental data.
The research on the surface oxygen species is necessary for the catalytic oxidation reaction. Therefore, the catalytic performance of the CeZrK/rGO catalysts in pure N2 was performed as shown in Figure 13. Due to the absence of gaseous O2, soot can only be oxidized by active adsorbed oxygen (O, O2-) and lattice oxygen (O2−) from the catalyst surface. It can be seen from DTG profiles that there was a trough for the CeZrK/rGO catalysts at 318~322 °C, which is attributed to the soot oxidation by adsorbed oxygen. The troughs for Ce5Zr2K2/rGO and Ce5Zr3K3/rGO at 513 °C and 521 °C, respectively, correspond to the soot oxidation by lattice oxygen. In comparison with the peak strength and the peak temperature, Ce5Zr2K2/rGO showed a better catalytic performance among the CeZrK/rGO catalysts, which can be derived from its abundant reactive oxygen species and reasonable metal doping concentration, improving the mobility of adsorbed oxygen and lattice oxygen.
Table 5 and Table 6 compare the catalytic performance of various catalysts with high catalytic activity, including perovskite catalysts, noble metal catalysts, and CeZrK/rGO catalysts for soot oxidation [9,19,30,50,51,52,53,54,55,56,57,58,59]. It is clear that the CeZrK/rGO catalysts by doping metal oxide particles onto graphene have higher catalytic activity than other catalysts, especially for Ce5Zr2K2/rGO. Generally, Ce5Zr2K2/rGO is a promising catalyst for the catalytic oxidation of soot in the CDPF owing to the advantages of easy synthesis, low cost, and high catalytic activity.

4. Conclusions

Today the environmental pollution [62,63,64,65] and energy crisis [66,67,68] have to learn another new problem. In this study, a new nanostructured solid solution catalyst CeZrK/rGO was prepared for the catalytic oxidation of soot by loading CeO2-ZrO2-K2O nanoparticles on the surface of rGO. Some major conclusions can be summarized as follows:
  • The use of rGO as the carrier of CeZrK/rGO catalysts provides a large specific surface area and special pore structure, increasing the contact between soot and catalyst compared to the Ce-based catalysts without rGO.
  • The doping of Zr and K ions in the CeZrK/rGO catalysts leads to the lattice distortion of CeO2, which can increase oxygen vacancies and active oxygen species, control the crystallite growth of CeO2, and ensure the dispersion quality of active sites.
  • The characterization studies show that the CeZrK/rGO catalysts have nanoscale pore structure, superior dispersion quality, small particle size, large specific surface area, as well as sufficient oxygen vacancies.
  • The T50 related to soot conversion of Ce5Zr1K1/rGO, Ce5Zr2K2/rGO, and Ce5Zr3K3/rGO under tight contact and loose contact conditions were 352 °C, 339 °C, 358 °C and 390 °C, 383 °C, 432 °C, respectively, which are smaller than those of some typical rare earth, perovskite, and precious metal catalysts, such as CeO2, 3DOM La0.8Ce0.2FeO3, Pt/Al2O3, etc. It is indicated that the CeZrK/rGO catalysts have high catalytic activity for soot oxidation, especially for Ce5Zr2K2/rGO.

Author Contributions

Conceptualization, Y.L. (Yuqiang Li); methodology, G.W.; investigation, K.T. and D.W.; data curation, K.T. and D.W.; writing—original draft preparation, G.W.; writing—review and ed-iting, Y.L. (Yuelin Li); supervision, Y.L. (Yuqiang Li); project administration, G.W.; funding acquisition, G.W. and Y.L. (Yuqiang Li). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China, grant number 51976016, Research Foundation of Education Bureau of Hunan Province, grant number 18B149, Open Research Subject of Key Laboratory of Advanced Manufacture Technology for Automobile Parts, grant number 2017KLMT02.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the main text of the article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Ratcliff, M.A.; Dane, A.J.; Williams, A.; Ireland, J.; Luecke, J.; McCormick, R.L.; Voorhees, K.J. Diesel Particle Filter and Fuel Effects on Heavy-Duty Diesel Engine Emissions. Environ. Sci. Technol. 2010, 44, 8343. [Google Scholar] [CrossRef]
  2. Wu, G.; Wang, X.; Li, Y.; Liu, Z. A Realistic Skeletal Mechanism for the Oxidation of Biodiesel Surrogate Composed of Long Carbon Chain and Polyunsaturated Compounds. Fuel 2021, 289, 119934. [Google Scholar] [CrossRef]
  3. Wu, G.; Wu, D.; Li, Y.; Meng, L. Effect of Acetone-n-butanol-ethanol (ABE) as an Oxygenate on Combustion, Performance, and Emission Characteristics of a Spark Ignition Engine. J. Chem. 2020, 2020, 7468651. [Google Scholar] [CrossRef] [Green Version]
  4. Johnson, T.V. Review of Diesel Emissions and Control. Int. J. Engine Res. 2009, 10, 275. [Google Scholar] [CrossRef]
  5. Peck, R.; Becker, C. Experimental Investigations and Dynamic Simulation of Diesel Particulate Filter Systems. Chem. Eng. Technol. 2009, 32, 1411. [Google Scholar] [CrossRef]
  6. Cauda, E.; Fino, D.; Saracco, G.; Specchia, V. Preparation and Regeneration of a Catalytic Diesel Particulate filter. Chem. Eng. Sci. 2007, 62, 5182. [Google Scholar] [CrossRef]
  7. Shukla, P.C.; Gupta, T.; Labhasetwar, N.K.; Khobaragade, R.; Gupta, N.K.; Agarwal, A.K. Effectiveness of Non-Noble Metal Based Diesel Oxidation Catalysts on Particle Number Emissions from Diesel and Biodiesel Exhaust. Sci. Total Environ. 2017, 574, 1512. [Google Scholar] [CrossRef] [PubMed]
  8. Mamontov, E.; Brezny, R.; Koranne, M.; Egami, T. Nanoscale Heterogeneities and Oxygen Storage Capacity of Ce0.5Zr0.5O2. J. Phys. Chem. B. 2003, 107, 13007. [Google Scholar] [CrossRef]
  9. Wei, Y.; Liu, J.; Zhao, Z.; Chen, Y.; Xu, C.; Duan, A.; Jiang, G.; He, H. Highly Active Catalysts of Gold Nanoparticles Supported on Three-Dimensionally Ordered Macroporous LaFeO3 for Soot Oxidation. Angew. Chem. Int. Ed. 2011, 50, 2326. [Google Scholar] [CrossRef]
  10. Da, Y.; Zeng, L.; Wang, C.; Mao, T.; Chen, R.; Gong, C.; Fan, G. Catalytic oxidation of diesel soot particulates over Pt substituted LaMn1-xPtxO3 perovskite oxides. Catal. Today 2019, 327, 73. [Google Scholar] [CrossRef]
  11. Sun, J.; Zhao, Z.; Li, Y.; Yu, X.; Zhao, L.; Li, J.; Wei, Y.; Liu, J. Synthesis and Catalytic Performance of Macroporous La1-xCexCoO3 Perovskite Oxide Catalysts with High Oxygen Mobility for Catalytic Combustion of Soot. J. Rare Earths 2020, 38, 584. [Google Scholar] [CrossRef]
  12. Mukherjee, D.; Reddy, B.M. Noble Metal-Free CeO2-Based Mixed Oxides for CO and Soot Oxidation. Catal. Today 2018, 309, 227. [Google Scholar] [CrossRef]
  13. Atribak, I.; Bueno-López, A.; García-García, A. Combined Removal of Diesel Soot Particulates and NOx Over CeO2–ZrO2 Mixed Oxides. J. Catal. 2008, 259, 123. [Google Scholar] [CrossRef] [Green Version]
  14. Gross, M.S.; Ulla, M.A.; Querini, C.A. Catalytic Oxidation of Diesel Soot: New Characterization and Kinetic Evidence Related to the Reaction Mechanism on K/CeO2 catalyst. Appl. Catal. A 2009, 360, 81. [Google Scholar] [CrossRef]
  15. Tan, L.; Wang, Y.; Liu, Q.; Wang, J.; Jing, X.; Liu, L.; Liu, J.; Song, D. Enhanced Adsorption of Uranium (VI) Using a Three-Dimensional Layered Double Hydroxide/Graphene Hybrid Material. Chem. Eng. J. 2015, 259, 752. [Google Scholar] [CrossRef]
  16. Mei, X.; Yan, Q.; Lu, P.; Wang, J.; Cui, Y.; Nie, Y.; Umar, A.; Wang, Q. Synthesis of Pt/K2CO3/MgAlOx–reduced Graphene Oxide Hybrids as Promising NOx Storage–Reduction Catalysts with Superior Catalytic Performance. Sci. Rep. 2017, 7, 42862. [Google Scholar] [CrossRef] [Green Version]
  17. Pan, M.; Qian, W.; Wei, H.; Feng, D.; Pan, J. Effects on Performance and Emissions of Gasoline Compression Ignition Engine over a Wide Range of Internal Exhaust Gas Recirculation Rates under Lean Conditions. Fuel 2020, 265, 116881. [Google Scholar] [CrossRef]
  18. Wang, P.; Yi, J.; Sun, C.; Luo, P.; Lei, L. Evaluation of H2 on the Evolution Mechanism of NOx Storage and Reduction over Pt-Ba-Ce/γ-Al2O3 Catalysts. Engineering 2019, 5, 568. [Google Scholar] [CrossRef]
  19. Pan, M.; Huang, R.; Liao, J.; Ouyang, T.; Lv, D.; Huang, H. Effect of EGR dilution on Combustion, Performance and Emission Characteristics of a Diesel Engine Fueled with n-Pentanol and 2-Ethylhexyl Nitrate Additive. Energy. Conv. Manag. 2018, 176, 246. [Google Scholar] [CrossRef]
  20. Zhang, P.; Yang, B.; Ma, H.; Wu, Z. Graphene Modified Porous Organic Polymer Supported Phosphotungstic Acid Catalyst for Alkylation Desulfurization. Fuel 2021, 293, 120438. [Google Scholar] [CrossRef]
  21. Li, X.; Wang, Z.; Wu, L.; Tsai, T. One-step in Situ Synthesis of a Novel α-zirconium Phosphate/Graphene Oxide Hybrid and its Application in Phenolic Foam with Enhanced Mechanical Strength, Flame Retardancy and Thermal Stability. RSC Adv. 2016, 6, 74903. [Google Scholar] [CrossRef]
  22. Wu, Z.S.; Ren, W.; Gao, L.; Zhao, J.; Chen, Z.; Liu, B.; Tang, D.; Yu, B.; Jiang, C.; Cheng, H.M. Synthesis of Graphene Sheets with High Electrical Conductivity and Good Thermal Stability by Hydrogen Arc Discharge Exfoliation. ACS Nano 2009, 3, 411. [Google Scholar] [CrossRef]
  23. Bora, C.; Bharali, P.; Baglari, S.; Dolui, S.K.; Konwar, B.K. Strong and Conductive Reduced Graphene Oxide/Polyester Resin Composite Films with Improved Mechanical Strength, Thermal Stability and its Antibacterial Activity. Compos. Sci. Technol. 2013, 87, 1. [Google Scholar] [CrossRef]
  24. Yuan, B.; Song, L.; Liew, K.M.; Hu, Y. Solid Acid-Reduced Graphene Oxide Nanohybrid for Enhancing Thermal Stability, Mechanical Property and Flame Retardancy of Polypropylene. RSC Adv. 2015, 5, 41307. [Google Scholar] [CrossRef]
  25. Zong, M.; Huang, Y.; Zhao, Y.; Sun, X.; Qu, C.; Luo, D.; Zheng, J. Facile Preparation, High Microwave Absorption and Microwave Absorbing Mechanism of RGO–Fe3O4 Composites. RSC Adv. 2013, 3, 23638. [Google Scholar] [CrossRef]
  26. Khira, Z.; Nasr, S.; Dorra, J.; Karima, H. Investigations on electrical conductivity and dielectric properties ofgraphene oxide nanosheets synthetized from modified Hummer’s method. J. Mol. Struct. 2020, 1216, 128304. [Google Scholar]
  27. Xiao, Q.; Sun, Y.; Zhang, J.; Li, Q. Size-dependent of Chromium (VI) Adsorption on Nano α-Fe2O3 Surface. Appl. Surf. Sci. 2015, 356, 18. [Google Scholar] [CrossRef]
  28. Sajeevan, A.C.; Sajith, V. Synthesis of Stable Cerium Zirconium Oxide Nanoparticle—Diesel Suspension and Investigation of its Effects on Diesel Properties and Smoke. Fuel 2016, 183, 155. [Google Scholar] [CrossRef]
  29. Lian, P.; Zhu, X.; Liang, S.; Li, Z.; Yang, W.; Wang, H. High Reversible Capacity of SnO2/graphene Nanocomposite as an Anode Material for Lithium-Ion Batteries. Electrochim. Acta 2011, 56, 4532. [Google Scholar] [CrossRef]
  30. Rangaswamy, A.; Sudarsanam, P.; Reddy, B.M. Rare Earth Metal Doped CeO2-based Catalytic Materials for Diesel Soot Oxidation at Lower Temperatures. J. Rare Earths 2015, 33, 1162. [Google Scholar] [CrossRef]
  31. You, X.; Sheng, Z.; Yu, D.; Yang, L.; Xiao, X.; Wang, S. Influence of Mn/Ce ratio on the Physicochemical Properties and Catalytic Performance of Graphene Supported MnOx-CeO2 Oxides for NH3-SCR at Low Temperature. Appl. Surf. Sci. 2017, 423, 845. [Google Scholar] [CrossRef]
  32. Julkapli, N.M.; Bagheri, S. Graphene Supported Heterogeneous Catalysts: An Overview. Int. J. Hydrog. Energy 2015, 40, 948–979. [Google Scholar] [CrossRef]
  33. Thommes, M.; Kaneko, K.; Neimark, A.V.; Olivier, J.P.; Rodriguez-Reinoso, F.; Rouquerol, J.; Sing, K.S. Physisorption of Gases, with Special Reference to the Evaluation of Surface Area and Pore Size Distribution (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1051. [Google Scholar] [CrossRef] [Green Version]
  34. Shi, Z.M.; Jin, L.N. Influence of La3+/Ce3+-doping on Phase Transformation and Crystal Growth in TiO2-15wt% ZnO Gels. J. Non-Cryst. Solids 2009, 355, 213. [Google Scholar] [CrossRef]
  35. Weng, D.; Li, J.; Wu, X.; Si, Z. Modification of CeO2-ZrO2 Catalyst by Potassium for NOx-Assisted Soot Oxidation. J. Environ. Sci. 2011, 23, 145. [Google Scholar] [CrossRef]
  36. Wei, Y.; Zhao, Z.; Jiao, J.; Liu, J.; Duan, A.; Jiang, G. Preparation of Ultrafine Ce-based Oxide Nanoparticles and Their Catalytic Performances for Diesel Soot Combustion. J. Rare Earths 2014, 32, 124. [Google Scholar] [CrossRef]
  37. Zheng, C.; Zhou, X.; Cao, H.; Wang, G.; Liu, Z. Synthesis of Porous Graphene/Activated Carbon Composite with High Packing Density and Large Specific Surface Area for Supercapacitor Electrode Material. J. Power Sources 2014, 258, 290. [Google Scholar] [CrossRef]
  38. Zhang, L.; Zhang, F.; Yang, X.; Long, G.; Wu, Y.; Zhang, T.; Leng, K.; Huang, Y.; Ma, Y.; Yu, A.; et al. Porous 3D Graphene-Based Bulk Materials with Exceptional High Surface Area and Excellent Conductivity for Supercapacitors. Sci. Rep. 2013, 3, 1408. [Google Scholar] [CrossRef] [Green Version]
  39. Vallés, C.; Núñez, J.D.; Benito, A.M.; Maser, W.K. Flexible Conductive Graphene Paper Obtained by Direct and Gentle Annealing of Graphene Oxide Paper. Carbon 2012, 50, 835. [Google Scholar] [CrossRef] [Green Version]
  40. Xie, R.; Fan, G.; Ma, Q.; Yang, L.; Li, F. Facile Synthesis and Enhanced Catalytic Performance of Graphene-Supported Ni Nanocatalyst from a Layered Double Hydroxide-Based Composite Precursor. J. Mater. Chem. A. 2014, 2, 7880. [Google Scholar] [CrossRef]
  41. Strankowski, M.; Włodarczyk, D.; Piszczyk, Ł.; Strankowska, J. Polyurethane Nanocomposites Containing Reduced Graphene Oxide, FTIR, Raman, and XRD Studies. J. Spectrosc. 2016, 2016, 7520741. [Google Scholar] [CrossRef] [Green Version]
  42. Fan, X.; Peng, W.; Li, Y.; Li, X.; Wang, S.; Zhang, G.; Zhang, F. Deoxygenation of Exfoliated Graphite Oxide under Alkaline Conditions: A Green Route to Graphene Preparation. Adv. Mater. 2008, 20, 4490. [Google Scholar] [CrossRef]
  43. Larachi, F.; Pierre, J.; Adnot, A.; Bernis, A. Ce 3d XPS Study of Composite CexMn1−xO2−y Wet Oxidation Catalysts. Appl. Surf. Sci. 2002, 195, 236. [Google Scholar] [CrossRef]
  44. Atiqullah, M.; Akhtar, M.N.; Faiz, M.; Moman, A.; Abu-Raqabah, A.H.; Khan, J.H.; Wazeer, M.I. Surface Chemistry of Selected Supported Metallocene Catalysts Studied by DR-FTIR, CPMAS NMR, and XPS Techniques. Surf. Interface Anal. 2006, 38, 1319. [Google Scholar] [CrossRef]
  45. Hajjar, Z.; Kazemeini, M.; Rashidi, A.; Bazmi, M. In Situ and Simultaneous Synthesis of a Novel Graphene-Based Catalyst for Deep Hydrodesulfurization of Naphtha. Catal. Lett. 2015, 145, 1660. [Google Scholar] [CrossRef]
  46. Li, H.; Lu, G.; Dai, Q.; Wang, Y.; Guo, Y.; Guo, Y. Efficient Low-Temperature Catalytic Combustion of Trichloroethylene over Flower-Like Mesoporous Mn-doped CeO2 Microspheres. Appl. Catal. B 2011, 102, 475. [Google Scholar] [CrossRef]
  47. Jiao, J.Z.; Li, S.H.; Huang, B.C. Preparation of Manganese Oxides Supported on Graphene Catalysts and their Activity in Low-Temperature NH3-SCR. Acta Phys. Chim. Sin. 2015, 31, 1383. [Google Scholar]
  48. Krishna, K.; Bueno-López, A.; Makkee, M.; Moulijn, J.A. Potential Rare Earth Modified CeO2 Catalysts for Soot Oxidation: I. Characterisation and Catalytic Activity with O2. Appl. Catal. B 2007, 75, 189. [Google Scholar] [CrossRef] [Green Version]
  49. Sui, L.; Yu, L. Diesel Soot Oxidation Catalyzed by Co-Ba-K Catalysts: Evaluation of the Performance of the Catalysts. Chem. Eng. J. 2008, 142, 327. [Google Scholar] [CrossRef]
  50. Liu, S.; Wu, X.; Weng, D.; Li, M.; Fan, J. Sulfation of Pt/Al2O3 Catalyst for Soot Oxidation: High Utilization of NO2 and Oxidation of Surface Oxygenated Complexes. Appl. Catal. B 2013, 138, 199. [Google Scholar] [CrossRef]
  51. Guilhaume, N.; Bassou, B.; Bergeret, G.; Bianchi, D.; Bosselet, F.; Desmartin-Chomel, A.; Jouguet, B.; Mirodatos, C. In situ Investigation of Diesel Soot Combustion Over an AgMnOx Catalyst. Appl. Catal. B 2012, 119, 287. [Google Scholar] [CrossRef]
  52. Zeng, L.; Cui, L.; Wang, C.; Guo, W.; Gong, C. In-situ Modified the Surface of Pt-Doped Perovskite Catalyst for Soot Oxidation. J. Hazard. Mater. 2020, 383, 121210. [Google Scholar] [CrossRef] [PubMed]
  53. Feng, N.; Chen, C.; Meng, J.; Liu, G.; Fang, F.; Wang, L.; Wan, H.; Guan, G. K–Mn Supported on Three-Dimensionally Ordered Macroporous La0.8Ce0.2FeO3 Catalysts for the Catalytic Combustion of Soot. Appl. Surf. Sci. 2017, 399, 114. [Google Scholar] [CrossRef]
  54. Xu, J.; Liu, J.; Zhao, Z.; Zheng, J.; Zhang, G.; Duan, A.; Jiang, G. Three-Dimensionally Ordered Macroporous LaCoxFe1−xO3 Perovskite-Type Complex Oxide Catalysts for Diesel Soot Combustion. Catal. Today 2010, 153, 136. [Google Scholar] [CrossRef]
  55. Zhang, G.; Zhao, Z.; Liu, J.; Jiang, G.; Duan, A.; Zheng, J.; Chen, S.; Zhou, R. Three Dimensionally Ordered Macroporous Ce1−xZrxO2 Solid Solutions for Diesel Soot Combustion. Chem. Commun. 2010, 46, 457. [Google Scholar] [CrossRef]
  56. Zhao, P.; Feng, N.; Fang, F.; Liu, G.; Chen, L.; Meng, J.; Chen, C.; Wang, L.; Wan, H.; Guan, G. Facile Synthesis of Three-Dimensional Ordered Macroporous Sr1−xKxTiO3 Perovskites with Enhanced Catalytic Activity for Soot Combustion. Catal. Sci. Technol. 2018, 8, 5462. [Google Scholar] [CrossRef]
  57. Mei, X.; Xiong, J.; Wei, Y.; Wang, C.; Wu, Q.; Zhao, Z.; Liu, J. Three-Dimensional Ordered Macroporous Perovskite-Type La1–xKxNiO3 Catalysts with Enhanced Catalytic Activity for Soot Combustion: The Effect of K-Substitution. Chin. J. Catal. 2019, 40, 722. [Google Scholar] [CrossRef]
  58. Alinezhadchamazketi, A.; Khodadadi, A.A.; Mortazavi, Y.; Nemati, A. Catalytic Evaluation of Promoted CeO2-ZrO2 by Transition, Alkali, and Alkaline-Earth Metal Oxides for Diesel Soot Oxidation. J. Environ. Sci. 2013, 25, 2498. [Google Scholar] [CrossRef]
  59. Matarrese, R.; Morandi, S.; Castoldi, L.; Villa, P.; Lietti, L. Removal of NOx and Soot Over Ce/Zr/K/Me (Me = Fe, Pt, Ru, Au) Oxide Catalysts. Appl. Catal. B 2017, 201, 318. [Google Scholar] [CrossRef]
  60. Liu, S.; Wu, X.; Lin, Y.; Li, M.; Weng, D. Active Oxygen-Assisted NO-NO2 Recycling and Decomposition of Surface Oxygenated Species on Diesel Soot with Pt/Ce0.6Zr0.4O2 Catalyst. Chin. J. Catal. 2014, 35, 407. [Google Scholar] [CrossRef]
  61. Wang, H.; Zhao, Z.; Xu, C.; Duan, A.; Liu, J.; Chi, Y. Simultaneous Removal of Soot Particles and NO From Diesel Engines Over LaBO3 Perovskite-Type Oxides. J. Phys. Chem. 2008, 29, 649. [Google Scholar]
  62. Han, D.; E, J.; Deng, Y.; Chen, J.; Leng, E.; Liao, G.; Zhao, X.; Feng, C.; Zhang, F. A review of studies using hydrocarbon adsorption material for reducing hydrocarbon emissions from cold start of gasoline engine. Renew. Sustain. Energy Rev. 2021, 135, 110079. [Google Scholar] [CrossRef]
  63. Zhao, X.; E, J.; Liao, G.; Zhang, F.; Chen, J.; Deng, Y. Numerical simulation study on soot continuous regeneration combustion model of diesel particulate filter under exhaust gas heavy load. Fuel 2021, 287, 119795. [Google Scholar] [CrossRef]
  64. E, J.; Zhao, M.; Zuo, Q.; Zhang, B.; Zhang, Z.; Peng, Q.; Han, D.; Zhao, X.; Deng, Y. Effects analysis on diesel soot continuous regeneration performance of a rotary microwave-assisted regeneration diesel particulate filter. Fuel 2020, 260, 116353. [Google Scholar] [CrossRef]
  65. Zhang, Z.; Ye, J.; Tan, D.; Feng, Z.; Luo, J.; Tan, Y.; Huang, Y. The effects of Fe2O3 based DOC and SCR catalyst on the combustion and emission characteristics of a diesel engine fueled with biodiesel. Fuel 2021, 290, 120039. [Google Scholar] [CrossRef]
  66. Wei, K.; Yang, Y.; Hong, Y.; Zhong, D. A review on ice detection technology and ice elimination technology for wind turbine. Wind Energy 2019, 23, 433–457. [Google Scholar] [CrossRef]
  67. Zuo, H.; Liu, G.; E, J.; Zuo, W.; Wei, K.; Hu, W.; Tan, J.; Zhong, D. Catastrophic analysis on the stability of a large dish solar thermal power generation system with wind-induced vibration. Sol. Energy 2019, 183, 40–49. [Google Scholar] [CrossRef]
  68. Zuo, H.; Tan, J.; Wei, K.; Huang, Z.; Zhong, D.; Xie, F. Effects of different poses and wind speeds on wind-induced vibration characteristics of a dish solar concentrator system. Renew. Energy 2021, 168, 1308–1326. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of CeZrK/rGO catalysts preparation.
Figure 1. Schematic diagram of CeZrK/rGO catalysts preparation.
Processes 09 00674 g001
Figure 2. Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images of CeZrK/rGO catalysts.
Figure 2. Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images of CeZrK/rGO catalysts.
Processes 09 00674 g002
Figure 3. Energy dispersive X-ray (EDX) elemental mapping of Ce5Zr2K2/rGO catalyst.
Figure 3. Energy dispersive X-ray (EDX) elemental mapping of Ce5Zr2K2/rGO catalyst.
Processes 09 00674 g003
Figure 4. X-ray powder diffraction (XRD) patterns of CeO2 and Ce5Zr2K2/rGO catalysts.
Figure 4. X-ray powder diffraction (XRD) patterns of CeO2 and Ce5Zr2K2/rGO catalysts.
Processes 09 00674 g004
Figure 5. N2 adsorption isotherms and pore size distributions of CeZrK/rGO catalysts.
Figure 5. N2 adsorption isotherms and pore size distributions of CeZrK/rGO catalysts.
Processes 09 00674 g005
Figure 6. Raman profiles of CeZrK/rGO catalysts and graphene oxide (GO).
Figure 6. Raman profiles of CeZrK/rGO catalysts and graphene oxide (GO).
Processes 09 00674 g006
Figure 7. Fourier transform infrared spectrometry (FTIR) spectra of CeZrK/rGO catalysts and GO.
Figure 7. Fourier transform infrared spectrometry (FTIR) spectra of CeZrK/rGO catalysts and GO.
Processes 09 00674 g007
Figure 8. X-ray photoelectron spectroscopy (XPS) spectra for (a) Ce 3d, (b) Zr 3d, (c) O 1s and (d) C 1s of the CeZrK/rGO catalysts.
Figure 8. X-ray photoelectron spectroscopy (XPS) spectra for (a) Ce 3d, (b) Zr 3d, (c) O 1s and (d) C 1s of the CeZrK/rGO catalysts.
Processes 09 00674 g008
Figure 9. H2-temperature programmed reduction (H2-TPR) profiles of GO and CeZrK/rGO catalysts.
Figure 9. H2-temperature programmed reduction (H2-TPR) profiles of GO and CeZrK/rGO catalysts.
Processes 09 00674 g009
Figure 10. Normalized soot conversion versus temperature of CeZrK/rGO catalysts under tight contact mode in simulated air (21% O2 + 79% N2).
Figure 10. Normalized soot conversion versus temperature of CeZrK/rGO catalysts under tight contact mode in simulated air (21% O2 + 79% N2).
Processes 09 00674 g010
Figure 11. Normalized soot conversion versus temperature of CeZrK/rGO catalysts and free-catalyst under loose contact mode in simulated air (21% O2 + 79% N2).
Figure 11. Normalized soot conversion versus temperature of CeZrK/rGO catalysts and free-catalyst under loose contact mode in simulated air (21% O2 + 79% N2).
Processes 09 00674 g011
Figure 12. TG/DTG (thermo gravimetric analysis/derivative thermo gravimetric analysis) profiles of Ce5Zr2K2/rGO in simulated air (21% O2 + 79% N2).
Figure 12. TG/DTG (thermo gravimetric analysis/derivative thermo gravimetric analysis) profiles of Ce5Zr2K2/rGO in simulated air (21% O2 + 79% N2).
Processes 09 00674 g012
Figure 13. DTG plots of soot combustion in tight contact mode over CeZrK/rGO catalysts in a pure N2 atmosphere.
Figure 13. DTG plots of soot combustion in tight contact mode over CeZrK/rGO catalysts in a pure N2 atmosphere.
Processes 09 00674 g013
Table 1. Average crystallite size and specific surface area of the CeZrK/rGO catalysts.
Table 1. Average crystallite size and specific surface area of the CeZrK/rGO catalysts.
CatalystsAverage Crystallite Size/nmSBET/m2·g−1
Ce5Zr1K18.3117.2
Ce5Zr2K26.9151.5
Ce5Zr3K36.7152.4
CeO28.9-
Table 2. Atomic surface composition of CeZrK/rGO catalysts determined from XPS.
Table 2. Atomic surface composition of CeZrK/rGO catalysts determined from XPS.
CatalystsAtomic Surface Composition (%)
CeZrOCCeOC
Ce3+Ce4+OOC-CC-O-O-C=O
Ce5Zr1K1/rGO7.31.3525.9665.3961.2638.7441.4558.5557.3526.8915.76
Ce5Zr2K2/rGO4.682.1127.0066.2161.8338.1740.8459.1659.0725.3015.62
Ce5Zr3K3/rGO4.473.3424.5067.6860.1439.8643.5156.4959.9924.7815.23
Table 3. Binding energies of Ce 3d3/2 (u‴), Zr 3d5/2, O 1s, and O 1s of CeZrK/rGO catalysts.
Table 3. Binding energies of Ce 3d3/2 (u‴), Zr 3d5/2, O 1s, and O 1s of CeZrK/rGO catalysts.
CatalystsCe 3d3/2 (u‴)/eVZr 3d5/2/eVO 1s/eVO 1s/eV
Ce5Zr1K1/rGO917.3182.6529.8531.6
Ce5Zr2K2/rGO916.9182.5530.0531.8
Ce5Zr3K3/rGO917.0182.7529.9531.5
Table 4. Ce3+/Ce4+, O/O, and Ce/Zr ratios of CeZrK/rGO catalysts.
Table 4. Ce3+/Ce4+, O/O, and Ce/Zr ratios of CeZrK/rGO catalysts.
CatalystsCe3+/Ce4+O/O/Ce/Zr
Ce5Zr1K1/rGO1.581.415.4
Ce5Zr2K2/rGO1.621.452.2
Ce5Zr3K3/rGO1.511.301.3
Table 5. T50 of various catalysts for soot oxidation under tight contact and loose contact conditions.
Table 5. T50 of various catalysts for soot oxidation under tight contact and loose contact conditions.
Catalyst NameT50 (°C)Refs.
TightLoose
Catalyst-free-604This work
CeO2-Sm2O3417517[30]
Pt/Al2O3-464[50]
Ag/MnOx-498[51]
LaCo0.94Pt0.06O3-437[52]
3DOM La0.8Ce0.2FeO3-434[53]
3DOM LaCo0.5 Fe0.5O3-397[54]
3DOM Ce0.8Zr0.2O2-396[55]
3DOM Sr0.8K0.2TiO3-382[56]
CeO2 a419523This work
Ce5Zr1K1/rGO a352390This work
Ce5Zr2K2/rGO a339383This work
Ce5Zr3K3/rGO a358432This work
a Based on TGA methods at the heating rate of 10 °C/min.
Table 6. Tm of various catalysts for soot oxidation under tight contact and loose contact conditions.
Table 6. Tm of various catalysts for soot oxidation under tight contact and loose contact conditions.
Catalyst NameTm (°C)Refs.
TightLoose
Au0.08/LaFeO3-359[9]
Pt0.1Ce0.68Zr0.31-430[60]
Pt0.1Al0.99-464[60]
3DOM La0.95K0.05NiO3-341[57]
LaNiO3-431[61]
CoMgAl-LDH/rGO433477[19]
Ce0.68Zr0.32O2440620[58]
K0.2Ce0.544Zr0.256O2440490[58]
CeZrK mixed oxides-445[35]
CeZrKRu mixed oxides-391[59]
CeO2416588This work
Ce5Zr1K1/rGO342347This work
Ce5Zr2K2/rGO329344This work
Ce5Zr3K3/rGO339330This work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, G.; Tang, K.; Wu, D.; Li, Y.; Li, Y. Experimental Evaluation on the Catalytic Activity of a Novel CeZrK/rGO Nanocomposite for Soot Oxidation in Catalyzed Diesel Particulate Filter. Processes 2021, 9, 674. https://doi.org/10.3390/pr9040674

AMA Style

Wu G, Tang K, Wu D, Li Y, Li Y. Experimental Evaluation on the Catalytic Activity of a Novel CeZrK/rGO Nanocomposite for Soot Oxidation in Catalyzed Diesel Particulate Filter. Processes. 2021; 9(4):674. https://doi.org/10.3390/pr9040674

Chicago/Turabian Style

Wu, Gang, Ke Tang, Deng Wu, Yuelin Li, and Yuqiang Li. 2021. "Experimental Evaluation on the Catalytic Activity of a Novel CeZrK/rGO Nanocomposite for Soot Oxidation in Catalyzed Diesel Particulate Filter" Processes 9, no. 4: 674. https://doi.org/10.3390/pr9040674

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