Next Article in Journal
One-Pot Synthesis of TiO2/Hectorite Composite and Its Photocatalytic Degradation of Methylene Blue
Previous Article in Journal
β-Arsenene Monolayer: A Promising Electrocatalyst for Anodic Chlorine Evolution Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Cordierite Monolith Catalysts with the Coating of K-Modified Spinel MnCo2O4 Oxide and Their Catalytic Performances for Soot Combustion

1
State Key Laboratory of Heavy Oil Processing, China University of Petroleum-Beijing, Beijing 102249, China
2
Institute of Catalysis for Energy and Environment, Shenyang Normal University, Shenyang 110034, China
*
Authors to whom correspondence should be addressed.
Catalysts 2022, 12(3), 295; https://doi.org/10.3390/catal12030295
Submission received: 24 January 2022 / Revised: 27 February 2022 / Accepted: 3 March 2022 / Published: 5 March 2022
(This article belongs to the Section Environmental Catalysis)

Abstract

:
Diesel engines are important for heavy-duty vehicles. However, particulate matter (PM) released from diesel exhaust should be eliminated. Nowadays, catalytic diesel particulate filters (CDPF) are recognized as a promising technology. In this work, a series of monolith Mn1−nKnCo2O4 catalysts were prepared by the simple citric acid method. The as-prepared catalysts displayed good catalytic performance for soot combustion and the Mn0.7K0.3Co2O4 catalyst gave the best catalytic performance among all the prepared samples. The T10 and Tm of Mn0.7K0.3Co2O4-HC catalyst for soot combustion are 310 and 439 °C, respectively. The physical and chemical properties of catalysts were characterized by means of SEM, XPS, H2-TPR, Raman and other techniques. The characterization results indicate that K substitution is favorable for the formation of oxygen vacancies, enhancing the mobility of active oxygen species, and improving the redox properties and so on. In-situ Raman results prove that the strength of Co-O bonds in the catalysts became weak during the reaction at high temperatures. In addition, SEM and ultrasonic test results show that the peeling rate of the coat-layer is less than 5%. The as-prepared catalysts can be taken as one kind of candidate catalyst for promising application in soot combustion because of its facile synthesis, low cost and high catalytic activity.

Graphical Abstract

1. Introduction

Diesel engines with excellent fuel efficiency have become the first choice for heavy-duty vehicles [1,2]. However, particulate matter (PM) released from diesel exhaust always threatens human health and environmental safety [3]. Because carbon is the main component of PM in diesel exhaust and its initiation temperature is higher than 450 °C, PM cannot be eliminated in the temperature range of diesel exhaust (150–450 °C) [4]. As the national emission standards become more stringent, it is necessary to develop catalytic purification technology to resolve the PM pollution. Nowadays, catalytic diesel particulate filters (CDPF) are considered a promising technology [5]. However, the design and preparation of high efficiency catalysts is still a difficult problem for the practical application of CDPF.
The catalytic combustion of PM is a solid–gas–solid heterogeneous catalytic reaction, and the diameter of PM is greater than 25 nm. Therefore, the catalytic performance of catalysts is influenced by two factors, including the redox property of catalysts and the contact efficiency between catalysts and PM. At present, the design and preparation of porous catalysts, especially three-dimensionally ordered microporous (3DOM) catalysts, contribute to significantly improving the contact between the soot and catalyst [6,7,8,9,10].
However, for CDPF, there are also two crucial aspects to improve the conditions of soot–catalyst contact in the filters [11,12]. The first one is the high dispersion and deep penetration of catalyst into the DPF walls by using suitable techniques; the second is how to avoid the cake layer by the accumulation of soot on the surface of DPF. To alleviate the contact problem, the design and preparation of high efficiency catalysts with the ability to activate oxygen at low temperatures is very important. In order to enhance the intrinsic activity and the ability to activate oxygen at low temperatures, a series of noble metals catalysts were also studied and they exhibit excellent catalytic performance for soot combustion, such as: Ag/Ce [13], Au–CoFe2O4 [14], Ag/Co-Ce [15], Pt/H-ZSM5 [16] and Ag/Al2O3 [17].
Solid solution with excellent oxidation performance is also a kind of catalyst for soot oxidation, such as perovskite (ABO3) [18,19], spinel (AB2O4) [20,21,22,23] or mullite (AB2O5) [24,25]. In our previous work, a kind of novel catalyst, which combines the advantages of three-dimensional ordered macropores (3DOM) and noble metals, has been successfully synthesized and the as-prepared catalysts effectively reduce the ignition temperature of soot, such as 3DOM Au0.04/Ce0.8Zr0.2O2 [26], 3DOM Aux/LaFeO3 [27], 3DOM Pt@CeO2−x/ZrO2 [28]. In addition, previous researchers found that the contact efficiency could be improved when catalysts possess a low melting point or alkali metal components [29,30,31]. For example, K/La2O3, K/MgO, KNO3/ZrO2 catalysts have similar soot combustion temperatures under tight and loose contact conditions [32,33,34,35]. Moreover, the complex oxides of transition metal and alkali metals exhibit the ability for reduction of the loss of alkali elements due to the synergistic effect between transition metals and alkali metals [36,37].
At present, a variety of excellent powder catalysts have been successfully developed in the laboratory. However, it should not be neglected that conventional powder catalysts will sharply increase the back pressure of exhaust gas pipeline. Therefore, diesel particulate filter (DPF) or honeycomb ceramics (HC) with the ability to catalyze the oxidation of soot is very essential, that is also called monolith catalyst. For instance, 3DOM LaCoO3/HC, LaKCoO3/HC, Co3O4/OMS-2/HC and Ce0.6Cu0.4O2/HC, these monolith catalysts displayed considerable performance for soot combustion [38,39,40,41]. Hernandez-Garrido prepared aluminum sols mixed with active components and studied the microscopic morphology of the slurry-coated honeycomb ceramics, revealing the interaction between them [42]. To make sure of the wide application of monolith catalyst, three factors for the preparation of monolith catalysts should be considered. The first is that the coating precursor liquid must have good fluidity, so that the honeycomb ceramics surface can be fully and evenly covered by the coating layer. The second is that inexpensive metals were used to prepare catalysts with good redox properties. The third is that honeycomb ceramics are used to ensure the smoothness of pipeline airflow to approximate the actual application situation.
Some researchers have studied the catalytic performance of alkali metals and spinel for soot combustion from the aspects of redox properties and intermediate products. However, the research on the effects between the structure and reactive oxygen species releasing, when the spinel A-site elements is substituted, especially the monolith spinel catalysts for soot oxidation are very few. Herein, a Co ion with good oxidizing ability is trivalent at the B position. The easily adjustable valence of Mn is beneficial to promoting the interaction between Mn and Co ions. In this work, MnCo2O4-HC spinel catalyst was synthesized and coated on honeycomb ceramics by the citric acid sol-solution method. To improve the catalytic performance of the as-prepared catalyst, the capacities for the adsorption and activation of oxygen of the MnCo2O4-HC catalyst were enhanced by introducing potassium, changing the unit cell structure of spinel and adjusting the valence of Mn and Co ions. At the same time, the catalyst coating can be more closely attached on the surface of honeycomb ceramics by citrate sol-solution, which is favorable for reducing the peeling rate of coating. In addition, the physical and chemical properties of catalysts were characterized successfully by SEM, XPS, H2-TPR, Raman spectroscopy and other techniques. The catalytic performances of the as-prepared monolith catalysts for soot combustion were also evaluated.

2. Results

2.1. Catalytic Performance of As-Prepared Monolith Catalysts for Soot Combustion

In order to explore the influence of K on the catalytic activity of MnCo2O4-HC spinel, the catalytic performance of Mn1−nKnCo2O4-HC with different n values (n = 0.1, 0.2, 0.3) was evaluated and the results are shown in Figure 1. According to CO2 concentration curves, the activities of catalysts sequentially increased in the following order, MnCo2O4-HC < Mn0.9K0.1Co2O4-HC < Mn0.8K0.2Co2O4-HC < Mn0.7K0.3Co2O4-HC. As shown in Table 1, when the n value is 0.3, the Mn0.7K0.3Co2O4-HC catalyst has the lowest temperature of T10, T50 and T90 among the as-prepared catalysts and the value are 310, 415 and 504 °C, respectively. Meantime, the selectivity of CO2 for Mn0.7K0.3Co2O4-HC catalyst is also higher than 99%. The above results indicate that potassium exhibits an obviously positive influence on catalytic performance for soot combustion. Compared with the case without a catalyst, the catalytic combustion temperature of soot on Mn0.7K0.3Co2O4-HC is significantly reduced. Due to its simple preparation method, it is considered one of the candidate catalysts that has the potential for application.

2.2. Resistance versus Sulfur and Water and Stability Test on Monolith Catalysts

As one kind of monolith catalyst for practical application, the properties of stability, sulfur and water resistance are very important. As shown in Figure 2a, the similar CO2 concentration curve indicated that the catalytic activity of the Mn0.7K0.3Co2O4-HC catalyst remained stable after five-cycle times. From Table 1, it can be seen that the catalytic activity of Mn0.7K0.3Co2O4-HC has remained stable until four-cycle times. The temperature of T10 for cycle 3 is 318 °C and slightly higher than the T10 of fresh Mn0.7K0.3Co2O4-HC (310 °C), while the T10 for the cycle 4 is decreased to 311 °C again. The T50 and T90 for cycle 3 and cycle 4 are located at 421, 423 °C and 506, 516 °C respectively, and the changing of temperature is also not obvious. However, T10, T50 and T90 of cycle 5 are located at 327, 431 and 513 °C, respectively, and the offset degree of T10 and T50 to higher temperature 15 °C than that of fresh Mn0.7K0.3Co2O4-HC. Figure 2b shows the catalytic activities of Mn0.7K0.3Co2O4-HC catalyst after pretreatment with SO2 and water vapor. After treatment for 6 h by SO2, the CO2 concentration curve was sharper than that of the fresh catalyst and the peak temperature (Tm) was shifted from 423 °C to 439 °C. However, the values of T10, T50 and T90 stayed the same as that of the fresh catalyst in Figure S1b. It perhaps means that SO2 exhibits a weak influence on the catalytic activity of the Mn0.7K0.3Co2O4-HC catalyst. When the catalyst was treated with water, the catalytic activity was sharply decreased, and the values of T10, T50 and T90 were shifted to the high temperature range of 370, 481 and 585 °C, respectively. The possible reason is that water vapor forms a competitive adsorption with NO on the surface of the catalyst, which hinders the path of NO participation in the reaction [43].

2.3. Morphologies of As-Prepared Monolith Catalysts

The photographs of the as-prepared catalysts are shown in Figure 3. From Figure 3a, it can be seen that the blank honeycomb ceramics monolith is white, and the channels are evenly and orderly distributed. When the Mn1−nKnCo2O4-HC active components were coated on the surface of blank honeycomb ceramics, the color of Mn1−nKnCo2O4-HC catalysts changed from white to black and was equally distributed (Figure 3b–e). Meanwhile, the channels were also well maintained. The above phenomena indicate that the Mn1−nKnCo2O4 active components are evenly coated on the honeycomb ceramics.
To more deeply investigate the morphologies of Mn1−nKnCo2O4-HC catalysts, the representative Mn0.7K0.3Co2O4-HC catalyst was characterized by SEM and the results are shown in the Figure 4. The views for the plane surface of blank honeycomb ceramics are shown in Figure 4a,b. The surface of the blank honeycomb ceramics is very rough. There are many irregular macropores on the surface under high resolution. Figure 4c,d shows that the surface of honeycomb ceramics was covered by the Mn0.7K0.3Co2O4-P active component based on the difference of plane and section. The image in Figure 4d indicates that the catalyst coat-layer exhibits lamellar appearance on the surface of the honeycomb ceramics with a large number of irregular macropores. Compared with the blank honeycomb ceramics, Figure 4d exhibits that the interface between the catalyst coat-layer and honeycomb ceramics was very clear. The thickness of the coating is 25.7 μ m. Additionally, the adhesion stability of the Mn0.7K0.3Co2O4-P active component on the surface of the honeycomb ceramics was also tested with the ultrasonic vibration test and the result is shown in Figure 4e. Based on the previously reported calculation method, the shedding rate of the Mn0.7K0.3Co2O4-HC catalyst is lower than 3.5%, indicating that the Mn0.7K0.3Co2O4-P active component can be well coated on the honeycomb ceramics [39]. Mn0.7K0.3Co2O4-HC displayed good cycle stability and adhesion, which meant that the catalyst possessed a longer lifetime.

2.4. EDS Mapping and X-ray Fluorescence of As-Prepared Monolith Catalysts

To obtain the distribution and content of elements for the prepared monolith catalysts, EDS mapping and X-Ray Fluorescence were carried out and the results are shown in Figure 5 and Table 2. The elemental mapping of Mg, Al, Si and active components on the ceramics and monolith catalysts are displayed in Figure 5. Mg, Al and Si are the main elements in ceramics and are distributed evenly. The elements of Mn and Co in MnCo2O4 have good dispersion on the surface of MnCo2O4-HC. For the Mn0.7K0.3Co2O4-HC catalyst, there is no obvious aggregation of K, indicating that K is homogeneously doped into the Mn0.7K0.3Co2O4 spinel.
As shown in Table 2, the XRF results indicate that the elements of Si, Al, and Mg are the main components for the blank ceramics, and their contents are 45.71%, 35.24% and 16.82%, respectively. Meanwhile, the blank ceramics also have little K element (0.21%) and others (2.02%). Because of the low loading amount of MnCo2O4, Si, Al, Mg and O elements of ceramics are still the main components in the MnCo2O4-HC catalyst. Due to a small amount of K in blank ceramics, the ratio of Mn to K is about 2.7 when the content of K in blank ceramics is deducted. The result is close to the theoretical stoichiometric ratio of Mn0.7K0.3Co2O4.

2.5. XRD Patterns of As-Prepared Monolith Catalysts

To deeply study the crystal structure of as-prepared catalysts, the monolith and corresponding powder Mn1−nKnCo2O4-HC catalysts were evaluated by the XRD technique and the results are shown in Figure 6. Figure 6a shows that the blank honeycomb ceramics, which are constituted by alumina, silica and magnesia, has very high crystallinity and very strong diffraction peaks. When Mn0.7K0.3Co2O4-P active components are coated on the surface of honeycomb ceramics, a weak characteristic peak, which belongs to the Mn1−nKnCo2O4-P spinel and is located at 2θ = 37o, can be observed. However, the characteristic peak of Mn1−nKnCo2O4-P in Figure 6a is so week that the crystal structure of the Mn1−nKnCo2O4-P spinel cannot be well identified. To more clearly explain the characteristic peaks, the XRD diffraction peaks of Mn1−nKnCo2O4-P spinel catalysts were also characterized (Figure 6b). The characteristic peaks of the samples at 2θ = 18.8°, 30.8°, 36.5°, 44.3°, 58.9° and 64.9° correspond to the crystal planes of (111), (220), (311), (400), (511), (440) for the spinel [44]. Figure 6b also indicates that the substitution of K metal can enhance the crystallinity of sample. As shown in the insert of Figure 6b, it can be seen that the diffraction peaks are shifted to high angle with increasing substitution amount of K in the 2θ range of 30–40°. This phenomenon indicates that K ions are successfully entered into the interior of the MnCo2O4-P spinel and cause lattice distortion [45].

2.6. Raman Spectra of As-Prepared Monolith Catalysts

Figure 7a,b shows the Raman spectra of Mn1−nKnCo2O4-HC catalysts with 532 nm and 325 nm laser wavelengths, respectively. As shown in Figure 7a, the catalysts exhibit obvious Raman vibration peaks in the two regions of 434–543 cm−1 and 625–709 cm−1. The peaks at low wavenumber are related to the spinel Eg mode, and the peaks at high wavenumber are belonged to the overlap peaks of the two vibration modes of F2g and A1g [46]. The Raman vibration peaks show significant shift under 532 nm laser wavelength when Mn ion is substituted by K in MnCo2O4-HC. The A1g vibration peak is shifted from 665 cm−1 to 673 cm−1 for Mn0.9K0.1Co2O4-HC catalyst. While the corresponding vibration peak of Mn0.7K0.3Co2O4-HC is continually shifted to 679 cm−1. The introduction of potassium element into the spinel structure makes Mn0.7K0.3Co2O4-HC’s vibration peak intensity higher than MnCo2O4-HC but lower than Co3O4 at the octahedral position of 679 cm−1, indicating that both the Co3+ cation-anion bond length and the coordination environment of the octahedral position have changed in the spinel lattice [47]. However, there is no corresponding vibration peak of manganese species in the Raman spectra [48].
Figure S2a shows the Raman spectrum of Co3O4 under 532 nm laser wavelength. There were five vibration peaks appeared in the detection range, which located at 191, 469, 512, 607 and 673 cm−1. These vibration peaks correspond to different vibration modes of Co3O4, such as, the vibration peak at 191 cm−1 belonging to the F2g1 vibration mode of the cobalt tetrahedron (CoO4). The vibration peaks at 469 and 512 cm−1 are attributed to the vibrations of Eg and F2g2 modes, respectively. The weak peak at 607 cm−1 is related to the F2g2 symmetrical vibration mode. The strong vibration peak at 673 cm−1 ascribed to A1g symmetrical vibration mode, which typically correlated with the octahedral point (CoO6) in the cobalt oxide spinel structure [49]. Different with cobalt oxide, manganese oxide only shows three Raman vibration peaks under laser wavelength of 532 nm. Raman band at 202 cm−1 can be attributed to external vibration, originating from the translational motion of MnO6 octahedron. The weak peak at 305 cm−1 may be related to the local crystal defect. The peak at 630 cm−1 can be assigned to symmetric stretch of Mn-O, which is perpendicular to the MnO6 octahedral [50].
The Raman spectra under 325 nm of Mn1−nKnCo2O4-HC catalysts are also provided in Figure 7b and Figure S2b,d. As shown in Figure S2b, the cobalt oxide calcined under the same conditions has five vibration peaks at 184, 468, 497, 606 and 661 cm−1, which belong to the F2g1 vibration mode, Eg, F2g2, F2g2 symmetrical vibration mode, and A1g symmetrical vibration mode of the surface cobalt species, respectively. Generally, manganese oxide is mainly composed of MnO2 and Mn2O3. MnO2 exhibits Raman vibration peaks at 512, 645 and 739 cm−1, while Mn2O3 has Raman peaks at 302 and 690 cm−1 [51]. The crystal structure of metal formed during calcination and intrinsic electronic properties can significantly affect the Raman signal [52]. The curve of pure manganese oxide displays weak vibration peaks located at 501, 635 and 701 cm−1 in Figure S2d [53]. As shown in Figure 7b, the Raman peaks at 463(468) and 497 cm−1 can be recognized as the Eg and F2g2 vibration modes and 606 cm−1 is the F2g2 symmetric vibration of Co-O bands. Compared with pure cobalt oxide and manganese oxide, no detection of vibration peak of Mn-O bands was observed and the vibration peaks at 184 and 661 cm−1 for Co-O bands disappeared in the Raman spectra of Mn1−nKnCo2O4-HC catalysts.

2.7. Soot-TPR

To deeply investigate the type of oxygen species in the Mn1−nKnCo2O4-P catalysts, soot-TPR for as-prepared powder catalysts were carried out and the results are shown in Figure 8a. During the soot-TPR process no gaseous oxygen was present in the reactant gas atmosphere. The oxygen species reacted with soot comes from the catalyst, including the adsorbed oxygen and lattice oxygen from the catalyst. The weak peak of CO2 at 200–300 °C is corresponded to the physical adsorption oxygen on the catalyst surface, namely α-Oads (superoxide, O2). The amount of such oxygen species is little; as a result, the amount of CO2 produced is also very little. The peak in the range of 490–670 °C for MnCo2O4-P is corresponded to the chemical adsorption state oxygen species β-Oads (per-oxygen, O22−), namely superoxide or peroxide. The peaks are the range of 379–670 °C for Mn0.9K0.1Co2O4-P are also attributed to the chemisorbed oxygen on the catalyst surface. From the Figure 8, it can be concluded that the temperature range of chemisorbed oxygen peak is shifted to low temperature with increasing of K substitution. This phenomenon is well agreed with the rule of catalytic activity. In addition, the peak at 700–800 °C corresponded to the surface lattice oxygen Olat of MnCo2O4-P [54,55]. Due to the K substitution, Mn0.8K0.2Co2O4-P and Mn0.7K0.3Co2O4-P exhibit the ability to release more active oxygen species at relatively low temperatures. Deconvolution of soot-TPR curves obtained by Gaussian Fitting are shown in the Figure 8b, based on the integrating results for peaks, the integral areas of Mn1−nKnCo2O4-P catalysts are about twice as long as that of MnCo2O4-P. The corresponding amounts of reactive oxygen species [O*], which are involved in the catalytic oxidation of soot, are listed in Table 3. The results indicate that the K substitution in the MnCo2O4-P spinel has changed the surface structure of catalysts and increased the number of surface oxygen species. These surface oxygen species are mainly derived from the bulk phase lattice oxygen of the catalyst subsurface. In addition, the increasing content of K substitution also enhances the mobility of these oxygen species.

2.8. H2-TPR of As-Prepared Monolith Catalysts

Figure 9 shows the H2-TPR results of as-prepared catalysts. The MnCo2O4-HC catalyst exhibits two distinct reduction peaks. The reduction peak in the temperature range of 250–400 °C is attributed to the reduction of the Mn and Co metal cations to the intermediate state. The high temperature of 400–650 °C belongs to the intermediate state to the low valence state [56]. Because the reduction of K+ ions cannot be achieved at 800 °C, the reduction peaks of Mn0.9K0.1Co2O4-HC are still attributed to the reduction of Mn and Co species [57]. However, compared with the MnCo2O4-HC catalyst, the intensity of the reduction peak at 350 °C for k-substituted samples has a significant improvement and the reduction peak at 491 °C obviously shifts to a low temperature. In addition, the Mn0.9K0.1Co2O4-HC catalyst has a new weak reduction peak at 428 °C. The changes indicate that the surface and internal oxygen species of the Mn0.9K0.1Co2O4-HC catalyst were activated by K substitution, which is beneficial to soot combustion. When the amount of the K substitution was increased, Mn0.8K0.2Co2O4-HC shows a weak reduction peak at 200~300 °C, which is belonged to the weak adsorption of oxygen on the catalyst surface [58]. Compared with Mn0.9K0.1Co2O4-HC, the temperatures for reduction peaks of Mn0.8K0.2Co2O4-HC are shifted to low temperatures of 331 and 483 °C. For the Mn0.7K0.3Co2O4-HC catalyst, the reduction peaks are obviously different than with other catalysts. It has three reduction peaks, which are located at 215, 318 and 401 °C, respectively. Compared with other catalysts, the lowest reduction temperature indicates that the metal cations in Mn0.7K0.3Co2O4-HC can be easily reduced demonstrating that oxygen species were easy to move and participate in the soot combustion. To deeply illustrate the results of H2-TPR, the H2 consumption amounts of as-prepared catalysts were calculated by the integral peak area. As shown in Table 3, the total H2 consumption amounts for as-prepared catalysts are similar to each other. However, due to K substitution, the reduction temperature ranges and H2 consumption for each reduction peak are different from each other.

2.9. XPS Results of As-Prepared Monolith Catalysts

To more clearly study and obtain high accuracy data for the composition and valence state of each element, the XPS of Mn1−nKnCo2O4-HC were characterized and the results are exhibited in Figure 10. As shown in Figure 10a, the survey spectra clearly exhibit the difference in surface elements among Mn1−nKnCo2O4-HC. The potassium element doped inside the spinel lattice shows a distinct XPS peak, and the C, O, Mn, and Co elements contained on the catalyst’s surface also show clear peaks in the spectrum.
Figure 10b exhibits the results of Co2p XPS. There are two main peaks located at 780.0 and 795.1 eV, respectively, which can be assigned to Co 2p3/2 and Co 2p1/2 and the electron spin splitting energy is 15.2 eV [59]. Meanwhile, the satellite peak located at 784~792 eV also can be observed. Generally speaking, Mn1−nKnCo2O4 spinel oxides contain two valence states of Co3+ and Co2+, and the binding energies of Co 2p3/2 and Co 2p1/2 in mixed valence states are in the range of 779.5–780.3 eV and 795.0–795.5 eV, respectively. Because the binding energy difference defining the spin-orbit splitting of Co 2p3/2 and Co 2p1/2 is 15.2 eV, Co2p can be divided into two sets of doublet peaks by deconvolution. The peaks located at 780.1 and 795.2 eV belong to Co3+, while the peaks located at 781.6 and 796.2 eV belong to Co2+, respectively.
Figure 10c shows the XPS spectra of Mn 2p. It can be seen that Mn 2p3/2 and Mn 2p1/2 are located in the two intervals of 638~647 eV and 650~659 eV, respectively. According to the previous literature and standard oxides, the binding energy difference between the two spin peaks is fixed at 11.3 eV [60]. The valence states of Mn in spinel are mainly divalent and trivalent [60,61,62]. The binding energies of the two groups of sub-peaks are 642.0, 653.3 eV and 643.8, 655.1 eV, which are attributed to Mn2+ and Mn3+, respectively. Their ratios are listed in Table 4. The proportion of Mn2+ rose slightly with the increasing of the amount of potassium substitution.
The XPS spectra of O1s are shown in Figure 10d. The O1s XPS peaks centered at 530, 531.9 and 533.1 eV can be assigned to lattice oxygen (O2−), adsorbed oxygen species (O22−, O2) and hydroxyl, and surface adsorbed water of catalysts, respectively [62]. The ratio of adsorbed oxygen species in the surface oxygen species have an important impact on catalytic performance. The amount of oxygen species adsorbed by the oxygen vacancies surface is fixed under static conditions. Therefore, analyzing the ratio of adsorbed oxygen can infer the state of oxygen vacancies. The ratio of adsorbed oxygen is listed in Table 4. As the amount of potassium substituted for manganese increased, the proportion of adsorbed oxygen species also showed an upward trend, and this trend was also consistent with the gradual increase in catalytic performance, which meant that more oxygen vacancies were generated on the surface after potassium substitution.
Table 4 listed the content of each element on the surface of Mn1−nKnCo2O4-HC catalysts. With the increasing of substitution amount of K, the K species content on the surface of the catalysts also gradually increased. Based on the data of atomic ratios in Table 4, the ratio of manganese to cobalt species on the surface is close to 1:1, indicating that manganese species were more easily enriched on the catalyst surface. Meanwhile, the proportion of Mn2+ species is slightly raised. Interestingly, the content of Co3+ species rapidly increased correspondingly. The results of Table 4 illustrate that the spinel structure of as-prepared catalysts is not changed when K is partially replaced by manganese. However, the K substitution indeed causes lattice distortion and imbalance of the charge of the spinel oxide. To maintain the overall charge balance of the catalyst, there may occur two cases, including loss of electrons to form Mn2+δ or Co3+ increases the valence states of cations at A or B sites, reducing the oxygen coordination number in the spinel lattice and forming Mn1−nKnCo2O4−δ. In this work, the K substitution can easily result in the transformation of Co2+ to Co3+ and reduce the amount of coordinated oxygen and form oxygen vacancies. In addition, the H2-TPR results also prove that the distorted crystal lattices can create more oxygen vacancies on the catalyst’s surface and the ability for the adsorption of oxygen species is also improved. The effect of oxygen deficiency caused by potassium substitution and the change in the valence of metal cations on activity will be further described in following Section 3.

2.10. In-Situ Raman Results of As-Prepared Monolith Catalysts

The in-situ Raman results mainly reflect the changes of the metal-oxygen bonds between the surface and sub-surface layers in different reaction environments during the reaction, which should be mainly attributed to the change of coordination oxygen in the catalyst. Therefore, the in-situ Raman experiments for TPO of the mixture of Mn0.7K0.3Co2O4-HC catalysts and soot particles under a laser wavelength of 532 nm were carried out. Figure 11 exhibits the Raman spectra for a mixture of Mn0.7K0.3Co2O4-HC catalyst and soot under a 532 nm wavelength laser. Compared with the Raman spectrum of pure soot (Figure S3), the Raman peaks for soot particles do not overlap with the peaks of in-situ Raman for Mn0.7K0.3Co2O4-HC catalyst. Therefore, the vibrational peak of a cobalt octahedron A1g at 697 cm−1 is still obvious at room temperature. With the increasing temperature, the A1g vibration peak shifted to low wavenumber and peak intensity decreased. A1g vibration peak shifted to 667cm−1 when the temperature was increased to 450 °C and the peak intensity of the vibration peak reached the lowest value at 350 °C. At high temperature, the shift of the main peak to the lower wave number and the widening of the spectral peak are generally considered to be the increase of the lower lattice disorder, while the intensity decrease is attributed to the degradation of CoO6 octahedron [63].
Compared with visible laser Raman spectra (532 nm), ultraviolet Raman spectra (UV) laser (325 nm) can reflect better surface sensitivity [64]. Therefore, the in-situ Raman for mixture of Mn0.7K0.3Co2O4-HC catalyst and soot particles under the soot-TPO reaction atmosphere were tested and the results are shown in Figure 12. From the spectra, it can be seen that soot particles also exhibit obvious Raman peaks and its peak positions are located at about 1590 cm−1. With the increasing of reaction temperature, the Raman peak intensity for soot gradually decreased and the peaks basically disappeared when the temperature rose to 400 °C. It is worth noting that the intensity of the F2g2 vibration mode for Co-O at 606 cm−1 gradually decreased with increasing temperature. This may be because that the cobalt oxide species on the outer surface participated in the catalytic reaction and was partially reduced by the soot, but it was oxidized by the reaction atmosphere after the soot was reacted, which indicated that the absence of alkali metal potassium made the cobalt-oxygen bond in MnCo2O4-HC difficult to participate in the oxidation reaction of soot.

3. Discussion

3.1. Influencing Factors of Structure for Monolith Catalytic Activity

Catalytic combustion of soot particles in the diesel vehicles exhausts gas occurs on the surface of monolith catalyst, which is a typical solid–gas–solid multi-phase catalytic reaction. The oxidation property and ability to adsorb and activate the gas phase reactants determine the activity of catalysts. Based on the characterization results, K-substitution has a significant impact on the structure, coordination bond and oxygen vacancy of the MnCo2O4-HC catalysts. It is known that the radius of K ion is 0.133 nm, which is much larger than that of Mn2+ (r = 0.080 nm) and Co3+ (r = 0.063 nm). The diffraction peaks of MnCo2O4-HC in the XRD shifted to high angle with the increasing of K content, and the peak intensity also increased. The Raman spectra 532 and 325 nm for Mn0.7K0.3Co2O4-HC and MnCo2O4-HC (Figure 7) indicate that the coordinated oxygen environment of Mn0.7K0.3Co2O4-HC is significantly different from that of MnCo2O4-HC. The Co-O bond of CoO6 in the bulk phase is affected by K+ substitution and the Raman band moves to high wavenumber. The above results indicate that K has successfully entered the unit cell structure of MnCo2O4 spinel. Due to this influence, the Co-O bonds in CoO6 were weakened, and more structural defects and lattice distortions were generated in the bulk phase of catalyst [46]. That is, Mn1−nKnCo2O4-HC will generate more oxygen vacancies than MnCo2O4-HC, indicates that more adsorbed oxygen species can be generated on Mn0.7K0.3Co2O4-HC surface. Moreover, some amounts of Co2+ transformed into Co3+. The relationship of these factors has been displayed in Figure 13.
The in-situ Raman spectra under 532 and 325 nm clearly show the different changes in the bulk and surface phases of Mn1−nKnCo2O4-HC catalyst under the reaction conditions. On the one hand, the in-situ Raman under 532 nm is sensitive to bulk (subsurface layer) [60]. On the other hand, the catalytic oxidation of soot is closely related to the oxygen species in the catalysts, which can affect the lattice structure of catalysts when they migrate in different environments. The structure changing regular of MnCo2O4-HC, Mn0.9K0.1Co2O4-HC and Mn0.7K0.3Co2O4-HC under different atmospheres (NO, NO+O2 and O2) were studied and the results are shown in Figure S4. It can be seen that Raman peak of CoO6 at 679 cm−1 shifts to low wavenumber and its intensity decreases, which indicates that the adsorption and activation of NO and O2 in the gas phase is related to the symmetric stretching vibration of CoO6, and O-Co3+-O is the main active site. Figure S4 also exhibits that the peaks at 679 cm−1 of MnCo2O4-HC and Mn0.9K0.1Co2O4-HC shift to low wavenumbers more obviously than that of Mn0.7K0.3Co2O4-HC, which means that the A1g vibrational mode of the Co-O bonds was changed in the MnCo2O4-HC and Mn0.9K0.1Co2O4-HC catalysts. When K is enriched on the surface, this negative impact can be effectively alleviated. It may speculate that the presence of K is beneficial to maintaining the kind of O-Co3+-O bonds in CoO6, and promotes the activation of oxygen species during the process of soot combustion [65]. From the in-situ Raman spectra of Figure S4, the vibration peak intensity ratio of (CoO6) at 25 °C and 350 °C, which can be abbreviated as   I 350 / 25 , changes regularly with the increase of K content. Based on the results above, it can be concluded that the regular is related to the proportion of reactive oxygen species. The results are shown in Figure 13. As the K content increased, I 350 / 25 gradually decreased while the content of reactive oxygen species gradually increased over the monolith catalysts.
The symmetrical vibration peak of catalyst Mn0.7K0.3Co2O4-HC at 606 cm−1 gradually decreases during the reaction in the 325 Raman result, while the peak for MnCo2O4-HC has almost no change, as shown in Figure 12, which further illustrates the positive effect of surface potassium in promoting soot combustion.

3.2. The Influence of K Substitution on Monolith Catalytic Performance and Physicochemical Properties

The physicochemical properties and oxygen species are important for the enhancement of catalytic performance. To more deeply study the nature of high activity, the TPR temperature, active oxygen, H2 consumption, K content and catalytic performance for as-prepared monolith catalysts are summarized in Figure 14 and Figure 15. As shown in the Figure 14, it can be obtained that the peak temperature of H2-TPR over monolith catalysts decreases with increasing K content. Similarly, the peak temperature of soot-TPR also has the same tendency (Figure 15). The temperatures for T10 and T50 of as-prepared monolith catalysts are decreased with increasing K content, which means the catalytic activity is enhanced.
Based on the XPS results for Co3+ content in Table 4, because of the K substitution, the charge imbalance caused by the low charge of K can promote the production of more Co3+ and improve the oxidation ability of catalysts. The mobility of oxygen species in the lattice increased and the formation of active oxygen species easily occurred. The reduction peaks shifted to low temperature in soot-TPR and H2-TPR verifying this viewpoint. From Figure 14 and Figure 15 and Table S1, it can be observed that the amounts of active oxygen for as-prepared monolith catalysts obtained by soot-TPR follow the order: MnCo2O4-HC < Mn0.9K0.1Co2O4-HC ≈ Mn0.8K0.2Co2O4-HC ≈ Mn0.7K0.3Co2O4-HC. However, the amounts of active oxygen obtained by H2-TPR follow the order: MnCo2O4-HC < Mn0.8K0.2Co2O4-HC < Mn0.9K0.1Co2O4-HC ≈ Mn0.7K0.3Co2O4-HC.
The essence of soot-TPR and H2-TPR reaction is as follows:
C + 2[O]→CO2
2H + [O]→H2O.
As we all know, hydrogen can react with subsurface or bulk lattice oxygen. However, due to a lack of K, oxygen species in MnCo2O4-HC is difficult to release from the (CoO6) octahedral unit and react with soot. The K substitution can change the coordination environment in the Mn1−nKnCo2O4 catalysts and part of the lattice oxygen can transform into active oxygen and move to the surface. Moreover, the metal-oxygen bond interaction is further weakened with increasing of K content and the mobility of active lattice oxygen also gradually increased. Therefore, the active oxygen species in the subsurface layer can be easily desorbed because of K substitution and the reduction peaks gradually shifted to low temperature.
As a result, T10 and T50 of Mn0.7K0.3Co2O4-HC are significantly lower than other catalysts. In summary, the replacement of potassium for Mn leads to the enhancements of Co3+ and defect sites, the mobility of oxygen species and the increase of active lattice oxygen which can participate in the reaction, and thus the temperature for catalytic combustion of soot decreases.

4. Materials and Methods

4.1. Catalyst Preparation

4.1.1. Materials

The chemical reagents used in this article are all standard reagents. The name and purity information of the reagents involved in the preparation are as follows: manganese nitrate Mn(NO3)2 (50%), AR), cobalt nitrate ((NO3)2·6H2O, AR), potassium nitrate (KNO3, AR), anhydrous ethanol(CH3CH2OH, AR), citric acid (C6H8O7·H2O, AR). The above reagents are all from Sinopharm Chemical Reagent Co., Ltd. The honeycomb ceramics for preparing monolith catalyst was a conventional commercial cylindrical ceramic with 400 mesh. The diameter and height of ceramics were 50 × 50 mm.

4.1.2. Synthesis of Honeycomb Ceramics Monolith Catalysts with the Coating of Mn1−nKnCo2O4

The honeycomb ceramics monolith catalysts with the coating of Mn1−nKnCo2O4 were prepared by the citric acid complexation method. The potassium nitrate, manganese nitrate and cobalt nitrate were weighed according to the nominal composition of the spinels of MnCo2O4, Mn0.9K0.1Co2O4, Mn0.8K0.2Co2O4, Mn0.7K0.3Co2O4, respectively. The molar ratio of citric acid to total metal cation was 1:1. The weighed metal salts and citric acid were dissolved in ethanol and the concentration of metal cations was 2 mol/L in the solution. Then, the solution was slowly stirred for 3 h to form a homogenous purple solution. After that, the pretreated cylindrical honeycomb ceramics monolith (high × diameter = 50 × 50 mm, calcined at 600 °C) was repeatedly immersed in the above purple sol to ensure the sol can be coated on the inner wall of the honeycomb ceramics. The excess solution in the channel was removed by purge gas, and then dried in an oven at 80 °C for 24 h. The dried samples were heated to 550 °C at 2 °C/min in a muffle furnace for 6 h. Finally, the honeycomb ceramics monolith catalysts with the coating of K-modified MnCo2O4 spinel were obtained. The as-prepared monolith catalysts were named as MnCo2O4-HC, Mn0.9K0.1Co2O4-HC, Mn0.8K0.2Co2O4-HC and Mn0.7K0.3Co2O4-HC for the sake of clarity. In addition, to more clearly study the physicochemical properties of as-prepared catalysts, the corresponding Mn1−nKnCo2O4 powder samples were also prepared by the drying and calcination of the remaining impregnation purple solution and the powder were named MnCo2O4-P, Mn0.9K0.1Co2O4-P, Mn0.8K0.2Co2O4-P, Mn0.7K0.3Co2O4-P, respectively. To more clearly express the loading amounts, the weights of blank honeycomb ceramics (HC) and Mn1−nKnCo2O4 -HC are listed in Table 5. Compared with the blank honeycomb ceramics (HC), the Mn1−nKnCo2O4 active components are successfully loaded on the honeycomb ceramics and the loading weight ranges are about 2.5 g, which proves that the loading method is effective for the synthesis of monolith catalysts.

4.2. Physical and Chemical Characterizations

The crystal phase structures of the catalysts were measured with an X-ray diffractometer (Shimadzu XRD 6000 with Cu Kα radiation, λ = 0.15406 nm). The step size of X-ray diffraction is 0.02 degrees. The scan rate is 4 degrees per minute in the 2θ range of 5~90 degrees. The morphologies of catalysts were studied by scanning electron microscopy (ZEISS Gemini SEM 300, accelerating voltage 5 kV). The valence information and binding energy results of each element were characterized by X-ray photoelectron spectroscopy (XPS, PerkinElmer PHI-1600 ESCA spectrometer).
H2-TPR measurements were performed using a Micromeritics AutoChem II 2920 (USA) dynamic adsorption analyzer. The mass of the catalyst for H2-TPR is 200 mg. To remove moisture and impurities on the surface, the catalyst was pretreated in helium at 400 °C for 120 min. The catalyst was conducted the reducing test in the temperature range of 50–750 °C, and the reducing gas contained 5% hydrogen and 95% helium. Heating rate was 10 °C/min. The signal was detected and recorded by the thermal conductivity detector (TCD). Soot-TPR were tested on gas chromatography GC-4320. Firstly, 0.1 g powder catalyst and 0.01 g soot with the mode of loose contact were treated at 300 °C for 30 min in Ar (50 mL/min) to remove other impurities. The temperature was raised from 150 to 800 °C with heating rate of 2 °C/min. The composition and concentration of exhaust gas were detected by gas chromatography GC-4320 with an FID detector.
The Raman and in-situ Raman spectra were conducted on inVia Reflex-Renishaw spectrometer. The wavelengths of the laser used were 325 and 532 nm. The atmosphere composition of in-situ Raman was the same as the activity measurements.
In-situ Raman spectra were also obtained on inVia Reflex-Renishaw spectrometer. The reaction gas compositions were NO (2000 ppm), O2 (10 %) and He (equilibrium gas), and the total flow rate was 50 mL/min. For different atmospheres of in-situ Raman spectra, the gas compositions of NO (2000 ppm)/He (equilibrium gas), O2 (10 %)/He (equilibrium gas) and NO (2000 ppm), O2 (10 %)/He (equilibrium gas) were selected according to experimental conditions. The contact mode for monolith catalyst and soot was loose contact.
The characterization in this work was all obtained by research of the monolith catalysts, except for the results of the soot-TPR and partial XRD measurements.

4.3. Activity Measurements

Catalyst activity was tested by a temperature-programmed oxidation (TPO) reaction. Soot particles (Printex-U diameter ~25 nm) were used to simulate PM2.5 in diesel exhaust, which contained C (92.0%), H (0.7%), O (3.5%), N. (0.1%), S (0.2%), and others (3.5%). The active components of the monolith catalysts were spinel catalyst coatings. Therefore, the mass ratio of soot to active components was 1:10 in the activity test. The heating rate of TPO was 2 °C/min. The reaction atmosphere contained NO (2000 ppm), O2 (10 %), argon gas as equilibrium gas, and the total flow rate was 50 ml/min. The soot particles were mixed in an appropriate amount of deionized and thoroughly stirred to form a suspension. At this time, the honeycomb ceramics monolith was immersed in the suspension and the parallel channel would be filled by it. Then the honeycomb ceramics monolith catalyst was dried in the vacuum oven at 80 °C for 48 hours to achieve the purpose of removing moisture. In this way, the soot was in a state of loose contact with the catalyst coating of honeycomb ceramics.
The cycle test followed these steps. Firstly, an activity test was performed on the monolith catalyst. Then, the reacted monolith catalyst was immersed in the same soot suspension and the surface of monolith catalyst was filled by soot suspension. After that, the impregnated monolith catalyst was dried at 80 °C in a vacuum oven for 48 h until the solvent evaporated. Finally, the dried monolith catalyst can be tested for the second cycle under the same reaction conditions for the activity test. In the following cycles the above steps were repeated.
The method for SO2 resistance test was carried out in the atmosphere containing 200 ppm SO2 with the equilibrium gas Ar and poisoning treatment at 300 °C for 6 h. Gas flow was 50 mL/min. The water resistance test was similar with the reaction conditions for activity test except for 2% water vapor in the reaction gases. The water vapor was obtained by continuous bubbling method under 25 °C in gas washing bottle.
To achieve the loose contact mode for soot-TPR, the powdered catalyst and soot particles were mixed in a burette with a length of 10 cm and a diameter of 2 cm, and then the burette was vibrated up and down for a total of 10 min. In the process of vibration, the mixture was stirred every 2 min.
The activities of catalysts were reflected by the combustion temperatures of soot particles. The gas chromatography GC-4320 was used to detect the concentration of CO and CO2 produced in the soot combustion. Total amount of CO2 produced by soot was obtained by integration of concentration-temperature curve, and then the proportion of CO2 at every temperature point was calculated to draw soot conversion-temperature relationship curve. Soot conversion rate reached 10% and the corresponding temperature was considered, as T10. T50 and T90 were achieved by the same principle. In addition, Tm, the temperature at which the CO2 concentration in the reaction reached its maximum, was also one of the indicators for reflecting catalyst activity. During the catalytic combustion of soot, the generation of CO should be minimized. Therefore, the selectivity of CO2 was also an important indicator for catalysts. The selectivity and maximum selectivity of CO2 in the product were calculated according to formulas (1) and (2), respectively
S CO 2 = [ CO 2 ] out [ CO 2 ] out + [ CO ] out × 100 %
S CO 2 m = [ CO 2 ] out max [ CO 2 ] out max + [ CO ] out max × 100 % .
The Mn0.7K0.3Co2O4-HC catalyst was put in the water and treated in ultrasonic for 5~30 min. The frequency of the ultrasonic instrument (KQ-250E) is 40 KHz with the power of 250 W. The peeling rate for the coat-layer was calculated as follows formula.
Peeling   rate ( % ) = m f m t m f m HC .
mf and mt are the weights of fresh and treated monolith catalysts, respectively. mHC is the weight of blank honeycomb ceramics.

5. Conclusions

In this work, a series of monolith Mn1−nKnCo2O4 catalysts with different contents of K substitution were prepared by the simple citric acid method and they exhibit good catalytic performance for soot combustion. Among the as-prepared catalysts, the Mn0.7K0.3Co2O4-HC catalyst exhibited the best catalytic performance, and the values of T10 and Tm for soot combustion are 310 and 439 °C, respectively. The results of XRD and Raman confirm that the K ions entered the spinel lattice. The XPS results indicate that the amounts of Co3+ and the oxygen vacancy increased with the increasing of K content. More importantly, H2-TPR and soot-TPR also demonstrate that the mobility of active oxygen is enhanced and many more active oxygen species can participate in soot combustion than that of the MnCo2O4-HC catalyst due to the K substitution for Mn. Meanwhile, the catalyst coating is more closely attached to the honeycomb ceramics surface and the coat-layer peeling rate is less than 5%. However, the as-prepared catalysts can be taken as one kind of candidate catalyst for in-depth research because of their facile synthesis, low cost and high catalytic activity. The results of this work highlight inexpensive metals and facile coating methods for the efficient preparation of monolith catalysts. Despite the systematic investigation of monolith catalysts for soot catalytic oxidation based on the MnCo2O4 spinel developed in the laboratory, the main challenge remains to improve the redox properties and contact efficiency of the coatings. The as-prepared catalysts can be taken as one kind of candidate catalyst for in-depth research for practical application because of facile synthesis, low cost and high catalytic activity.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal12030295/s1, Figure S1: The soot conversion curves of cycle test (a) and resistance of sulfur and water (b) over Mn0.7K0.3Co2O4-HC monolith catalyst; Figure S2: Raman spectra of cobalt oxide and manganese oxide under different laser wavelength of 532 nm (a,c) and 325 nm (b,d); Figure S3: Raman spectrum of soot under laser wavelength of 532 nm; Figure S4: In-situ Raman spectra of MnCo2O4-HC (a,d and g), Mn0.9K0.1Co2O4-HC (b,e and h) and Mn0.7K0.3Co2O4-HC (c,f and i) monolith catalysts under laser wavelength of 532 nm for different atmosphere. Reaction conditions: heating rate was 2 °C/min. The reaction atmosphere contained NO (2000 ppm), O2 (10%) and He as equilibrium gas, and the total flow rate was 50 ml/min Table S1: The summary of catalytic performance, TPR temperature, cationic valences and active oxygen content.

Author Contributions

Conceptualization, Z.Z.; Investigation, Y.W. and K.Z.; Visualization, K.Z.; Writing original draft, K.Z.; Resources, J.L. (Jianmei Li) and D.L.; Validation, L.W. and R.L.; Methodology, K.Z. and B.L.; Project administration, X.Y. and J.L. (Jian Liu); Funding Acquisition, Z.Z.; Supervision: Z.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Key Research and Development Program of MOST (2017YFE0131200) for collaboration between China and Poland; NSFC (22072095, U1908204, 21761162016); University Joint Education Project for China-Central and Eastern European Countries (2021097); National Engineering Laboratory for Mobile Source Emission Control Technology (NELMS2018A04); University level innovation team of Shenyang Normal University; Major Incubation Program of Shenyang Normal University (ZD201901)

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time as the data also form part of an ongoing study.

Acknowledgments

Thanks to all reviewers and editors for suggestions and checking of the manuscript. Also thanks to colleagues at China University of Petroleum (Beijing, China) and Shenyang Normal University for their contributions to the experiments and manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Twigg, M.V. Progress and future challenges in controlling automotive exhaust gas emissions. Appl. Catal. B Environ. 2007, 7, 2–15. [Google Scholar] [CrossRef]
  2. Liu, S.; Wu, X.; Weng, D.; Ran, R. Ceria-Based Catalysts for Soot Oxidation: A Review. J. Rare Earths 2015, 33, 567–590. [Google Scholar] [CrossRef]
  3. Duvvuri, P.P.; Sukumaran, S.; Shrivastava, R.K.; Sreedhara, S. Modeling soot particle size distribution in diesel engines. Fuel 2019, 243, 70–78. [Google Scholar] [CrossRef]
  4. Liu, J.; Zhao, Z.; Xu, C.; Duan, A.; Jiang, G. Comparative Study on Physicochemical Properties and Combustion Behaviors of Diesel Particulates and Model Soot. Energ. Fuels 2010, 24, 3778–3783. [Google Scholar] [CrossRef]
  5. Johnson, T.V. Review of diesel emissions and control. Int. J. Engine. Res. 2009, 10, 275–285. [Google Scholar] [CrossRef]
  6. Sadakane, M.; Asanuma, T.; Kubo, J.; Ueda, W. Facile Procedure to Prepare Three-Dimensionally Ordered Macroporous (3DOM) Perovskite-type Mixed Metal Oxides by Colloidal Crystal Templating Method. Chem. Mater. 2005, 17, 3546–3551. [Google Scholar] [CrossRef]
  7. Sadakane, M.; Takahashi, C.; Kato, N.; Asanuma, T.; Ihara, H.; Ueda, W. Three-dimensionally ordered macroporous mixed iron oxide; Preparation and structural characterization of inverse opals with skeleton structure. Chem. Lett. 2006, 25, 480–481. [Google Scholar] [CrossRef]
  8. Sadakane, M.; Takahashi, C.; Kato, N.; Ogihara, H.; Nodasaka, Y.; Doi, Y.; Hinatsu, Y.; Ueda, W. Three-dimensionally ordered macroporous (3DOM) materials of spinel-type mixed iron oxides. Synthesis, structural characterization, and formation mechanism of inverse opals with a skeleton structure. Bull. Chem. Soc. Jpn. 2007, 80, 677–685. [Google Scholar] [CrossRef] [Green Version]
  9. Zhai, G.; Wang, J.; Chen, Z.; Yang, S.; Men, Y. Highly enhanced soot oxidation activity over 3DOM Co3O4-CeO2 catalysts by synergistic promoting effect. J. Hazard. Mater. 2019, 363, 214–226. [Google Scholar] [CrossRef]
  10. 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, 457–459. [Google Scholar] [CrossRef]
  11. Sarli, D.V.; Landi, G.; Lisi, L.; Benedetto, A.D. Ceria-Coated Diesel Particulate Filters for Continuous Regeneration. AIChE J. 2017, 63, 3442–3449. [Google Scholar] [CrossRef]
  12. Sarli, V.D.; Landi, G.; Lisi, L.; Saliva, A.; Benedetto, A.D. Catalytic diesel particulate filters with highly dispersed ceria: Effect of the soot-catalyst contact on the regeneration performance the soot-catalyst contact on the regeneration performance. Appl. Catal. B Environ. 2016, 197, 116–124. [Google Scholar] [CrossRef]
  13. Sarli, V.D.; Landi, G.; Benedetto, A.D.; Lisi, L. Synergy Between Ceria and Metals (Ag or Cu) in Catalytic Diesel Particulate Filters: Efect of the Metal Content and of the Preparation Method on the Regeneration Performance. Top. Catal. 2021, 64, 256–269. [Google Scholar] [CrossRef]
  14. Khan, A.U.; Ullah, S.; Yuan, Q.; Ali, S.; Ahmad, A.; Khan, Z.U.H.; Rahman, A.U. In Situ Fabrication of Au–CoFe2O4: An Efficient Catalyst for Soot Oxidation. Appl. Nanosci. 2020, 10, 3901–3910. [Google Scholar] [CrossRef]
  15. Zou, G.; Fan, Z.; Yao, X.; Zhang, Y.; Zhang, Z.; Chen, M.; Shangguan, W. Catalytic Performance of Ag/Co-Ce Composite Oxides During Soot Combustion in O2 and NOx: Insights into the Effects of Silver. Chin. J. Catal. 2017, 38, 564–572. [Google Scholar] [CrossRef]
  16. Gao, Y.; Yang, W.; Wu, X.; Liu, S.; Weng, D.; Ran, R. Controllable Synthesis of Supported Platinum Catalysts: Acidic Support Effect and Soot Oxidation Catalysis. Catal. Sci. Technol. 2017, 7, 3268–3274. [Google Scholar] [CrossRef]
  17. Gao, Y.; Wu, X.; Liu, S.; Ogura, M.; Liu, M.; Weng, D. Aggregation and Redispersion of Silver Species on Alumina and Sulphated Alumina Supports for Soot Oxidation. Catal. Sci. Technol. 2017, 7, 3524–3530. [Google Scholar] [CrossRef]
  18. Milt, V.G.; Ulla, M.A.; Miró, E.E. NOx Trapping and Soot Combustion on BaCoO3−y Perovskite: LRS and FTIR Characterization. Appl. Catal. B Environ. 2005, 57, 13–21. [Google Scholar] [CrossRef]
  19. Feng, N.J.; Meng, J.; Wu, Y.; Chen, C.; Wang, L.; Gao, L.; Wan, H.; Guan, G.F. KNO3 Supported on Three-Dimensionally Ordered Macroporous La0.8Ce0.2Mn1−xFexO3 for Soot Removal. Catal. Sci. Technol. 2016, 6, 2930–2941. [Google Scholar] [CrossRef]
  20. Shangguan, W.F.; Teraoka, Y.; Kagawa, S. Promotion Effect of Potassium on the Catalytic Property of CuFe2O4 for the Simultaneous Removal of NOx and Diesel Soot Particulate. Appl. Catal. B Environ. 1998, 16, 149–154. [Google Scholar] [CrossRef]
  21. Shangguan, W.F.; Teraoka, Y.; Kagawa, S. Kinetics of Soot-O2, Soot-No and Soot-O2-NO Reactions over Spinel-Type CuFe2O4 Catalyst. Appl. Catal. B Environ. 1997, 12, 237–247. [Google Scholar] [CrossRef]
  22. Zhao, Y.; Wen, Z.; Huang, Y.; Duan, X.; Cao, Y.; Ye, L.; Jiang, L.; Yuan, Y. Low-Temperature Soot Combustion over Ceria Modified MgAl2O4-Supported Ag Nanoparticles. Catal. Commun. 2018, 111, 26–30. [Google Scholar] [CrossRef]
  23. Zhao, H.; Zhou, X.X.; Pan, L.Y.; Wang, M.; Chen, H.R.; Shi, J.L. Facile Synthesis of Spinel Cu1.5Mn1.5O4 Microspheres with High Activity for the Catalytic Combustion of Diesel Soot. RSC Adv. 2017, 7, 20451–20459. [Google Scholar] [CrossRef] [Green Version]
  24. Chen, Y.; Shen, G.; Lang, Y.; Chen, R.; Jia, L.; Yue, J.; Shen, M.; Du, C.; Shan, B. Promoting Soot Combustion Efficiency by Strengthening the Adsorption of NOx on the 3DOM Mullite Catalyst. J. Catal. 2020, 384, 96–105. [Google Scholar] [CrossRef]
  25. Feng, Z.; Liu, Q.; Chen, Y.; Zhao, P.; Peng, Q.; Cao, K.; Chen, R.; Shen, M.; Shan, B. Macroporous SmMn2O5 Mullite for NOx-Assisted Soot Combustion. Catal. Sci. Technol. 2017, 7, 838–847. [Google Scholar] [CrossRef]
  26. Wei, Y.; Liu, J.; Zhao, Z.; Duan, A.; Jiang, G.; Xu, C.; Gao, J.; He, H.; Wang, X. Three-Dimensionally Ordered Macroporous Ce0.8Zr0.2O2-Supported Gold Nanoparticles: Synthesis with Controllable Size and Super-Catalytic Performance for Soot Oxidation. Energy Environ. Sci. 2011, 4, 2959–2970. [Google Scholar] [CrossRef]
  27. Wei, Y.C.; Liu, J.A.; Zhao, Z.; Chen, Y.S.; Xu, C.M.; Duan, A.J.; Jiang, G.Y.; He, H. Highly Active Catalysts of Gold Nanoparticles Supported on Three-Dimensionally Ordered Macroporous LaFeO3 for Soot Oxidation. Angew. Chem.-Int. 2011, 50, 2326–2329. [Google Scholar] [CrossRef]
  28. Li, Y.Z.; Du, Y.H.; Wei, Y.C.; Zhao, Z.; Jin, B.F.; Zhang, X.D.; Liu, J. Catalysts of 3d Ordered Macroporous ZrO2-Supported Core-Shell Pt@CeO2−x Nanoparticles: Effect of the Optimized Pt-CeO2 Interface on Improving the Catalytic Activity and Stability of Soot Oxidation. Catal. Sci. Technol. 2017, 7, 968–981. [Google Scholar] [CrossRef]
  29. Liu, J.; Zhao, Z.; Xu, C.-m.; Duan, A.J.; Meng, T.; Bao, X.J. Simultaneous Removal of NOx and Diesel Soot Particulates over Nanometric La2−xKxCuO4 Complex Oxide Catalysts. Catal. Today 2007, 119, 267–272. [Google Scholar] [CrossRef]
  30. Li, Q.; Wang, X.; Chen, H.; Xin, Y.; Tian, G.; Lu, C.; Zhang, Z.; Zheng, L.; Zheng, L. K-Supported Catalysts for Diesel Soot Combustion: Making a Balance between Activity and Stability. Catal. Today 2016, 264, 171–179. [Google Scholar] [CrossRef]
  31. Zhao, H.; Zhou, X.; Huang, W.; Pan, L.; Wang, M.; Li, Q.; Shi, J.; Chen, H. Effect of Potassium Nitrate Modification on the Performance of Copper-Manganese Oxide Catalyst for Enhanced Soot Combustion. ChemCatChem 2018, 10, 1455–1463. [Google Scholar] [CrossRef]
  32. Milt, V.G.; Pissarello, M.L.; Miró, E.E.; Querini, C.A. Abatement of Diesel-Exhaust Pollutants: NOx Storage and Soot Combustion on K/La2O3 Catalysts. Appl. Catal. B Environ. 2003, 41, 397–414. [Google Scholar] [CrossRef]
  33. Pisarello, M.L.; Milt, V.; Peralta, M.A.; Querini, C.A.; Miró, E.E. Simultaneous Removal of Soot and Nitrogen Oxides from Diesel Engine Exhausts. Catal. Today 2002, 75, 465–470. [Google Scholar] [CrossRef]
  34. Jiménez, R.; García, X.; Cellier, C.; Ruiz, P.; Gordon, A.L. Soot Combustion with K/MgO as Catalyst. Appl. Catal. A Gen. 2006, 297, 125–134. [Google Scholar] [CrossRef]
  35. Carrascull, A.; Lick, I.D.; Ponzi, E.N.; Ponzi, M.I. Catalytic Combustion of Soot with a O2/NO Mixture. KNO3/ZrO2 Catalysts. Catal. Commun. 2003, 4, 124–128. [Google Scholar] [CrossRef]
  36. An, H.; Kilroy, C.; McGinn, P.J. Combinatorial Synthesis and Characterization of Alkali Metal Doped Oxides for Diesel Soot Combustion. Catal. Today 2004, 98, 423–429. [Google Scholar] [CrossRef]
  37. An, H.; McGinn, P.J. Catalytic Behavior of Potassium Containing Compounds for Diesel Soot Combustion. Appl. Catal. B Environ. 2006, 62, 46–56. [Google Scholar] [CrossRef]
  38. Tang, L.; Zhao, Z.; Wei, Y.; Liu, J.; Peng, Y.; Li, K. Study on the Coating of Nano-Particle and 3DOM LaCoO3 Perovskite-Type Complex Oxide on Cordierite Monolith and the Catalytic Performances for Soot Oxidation: The Effect of Washcoat Materials of Alumina, Silica and Titania. Catal. Today 2017, 297, 131–142. [Google Scholar] [CrossRef]
  39. Tang, L.; Zhao, Z.; Li, K.; Yu, X.; Wei, Y.; Liu, J.; Peng, Y.; Li, Y.; Chen, Y. Highly Active Monolith Catalysts of LaKCoO3 Perovskite-Type Complex Oxide on Alumina-Washcoated Diesel Particulate Filter and the Catalytic Performances for the Combustion of Soot. Catal. Today 2020, 339, 159–173. [Google Scholar] [CrossRef]
  40. Yang, Y.; Zhao, D.; Gao, Z.; Tian, Y.; Ding, T.; Zhang, J.; Jiang, Z.; Li, X. Interface Interaction Induced Oxygen Activation of Cactus-Like Co3O4/Oms-2 Nanorod Catalysts in Situ Grown on Monolith Cordierite for Diesel Soot Combustion. Appl. Catal. B Environ. 2021, 286, 119932–119942. [Google Scholar] [CrossRef]
  41. Nascimento, L.F.; Serra, O.A. Washcoating of Cordierite Honeycomb with Ceria-Copper Mixed Oxides for Catalytic Diesel Soot Combustion. Process Saf. Environ. 2016, 101, 134–143. [Google Scholar] [CrossRef]
  42. Hernandez-Garrido Juan, C.; Gomez Diana, M.; Gaona, D.; Vidal, H.; Gatica José, M.; Sanz, O.; Rebled José, M.; Peiro, F.; Calvino José, J. Combined (S)TEM-FIB Insight into the Influence of the Preparation Method on the Final Surface Structure of a Co3O4/La-Modified-CeO2 Washcoated Monolith Catalyst. J. Phys. Chem. C. 2013, 117, 13028–13036. [Google Scholar] [CrossRef]
  43. Hernández-Giménez, A.M.; Lozano-Castelló, D.; Bueno-López, A. Effect of CO2, H2O and SO2 in the Ceria-Catalyzed Combustion of Soot under Simulated Diesel Exhaust Conditions. Appl. Catal. B Environ. 2014, 148-149, 406–414. [Google Scholar] [CrossRef] [Green Version]
  44. Zheng, J.; Lei, Z. Incorporation of CoO Nanoparticles in 3D Marigold Flower-Like Hierarchical Architecture MnCo2O4 for Highly Boosting Solar Light Photo-Oxidation and Reduction Ability. Appl. Catal. B Environ. 2018, 237, 1–8. [Google Scholar] [CrossRef]
  45. Liang, Q.; Qu, X.; Bia, N.; Chen, H.; Zou, X.; Li, G. Alkali metal-incorporated spinel oxide nanofibers enable high performance detection of formaldehyde at ppb level. J. Hazard. Mater. 2020, 400, 123301–123309. [Google Scholar] [CrossRef] [PubMed]
  46. Wang, X.; Liu, Y.; Zhang, T.; Luo, Y.; Lan, Z.; Zhang, K.; Zuo, J.; Jiang, L.; Wang, R. Geometrical-Site-Dependent Catalytic Activity of Ordered Mesoporous Co-Based Spinel for Benzene Oxidation: In Situ Drifts Study Coupled with Raman and XAFS Spectroscopy. ACS Catal. 2017, 7, 1626–1636. [Google Scholar] [CrossRef]
  47. Bai, B.; Li, J. Positive Effects of K+ Ions on Three-Dimensional Mesoporous Ag/Co3O4 Catalyst for HCHO Oxidation. ACS Catal. 2014, 4, 2753–2762. [Google Scholar] [CrossRef]
  48. Yuvaraj, S.; Vignesh, A.; Shanmugam, S.; Kalai Selvan, R. Nitrogen-Doped Multi-Walled Carbon Nanotubes-MnCo2O4 Microsphere as Electrocatalyst for Efficient Oxygen Reduction Reaction. Int. J. Hydrogen Energy 2016, 41, 15199–15207. [Google Scholar] [CrossRef]
  49. Xiao, B.; Zhao, K.; Zhang, L.; Cai, T.; Zhang, X.; Wang, Z.; Yuan, J.; Yang, L.; Gao, P.; He, D. A Green and Facile Synthesis of Co3O4 Monolith Catalyst with Enhanced Total Oxidation of Propane Performance. Catal. Commun. 2018, 116, 1–4. [Google Scholar] [CrossRef]
  50. Hastuti, E.; Subhan, A.; Amonpattaratkit, P.; Zainuri, M.; Suasmoro, S. The Effects of Fe-Doping on MnO2: Phase Transitions, Defect Structures and Its Influence on Electrical Properties. RSC Adv. 2021, 11, 7808–7823. [Google Scholar] [CrossRef]
  51. Fernandez, J.F.; Caballero, A.C.; Villegas, M.; Khatib, S.J.; Banares, M.A.; Fierro, J.L.G.; Costa-Kramer, J.L.; Lopez-Ponce, E.; Mart’ ın-Gonzalez, M.S.; Briones, F.; et al. Structure and Magnetism in the Zn–Mn–O System: A Candidate for Room Temperature Ferromagnetic Semiconductor. J. Eur. Ceram. Soc. 2006, 26, 3017–3025. [Google Scholar] [CrossRef]
  52. Julien, C.M.; Massot, M.; Poinsignon, C. Lattice Vibrations of Manganese Oxides: Part I. Periodic Structures. Spectrochim. Acta A 2004, 60, 689–700. [Google Scholar] [CrossRef]
  53. Fu, F.; Zhang, Y.; Zhang, Y.; Chen, Y. Synthesis of Mn-Doped and Anatase/Rutile Mixed-Phase TiO2 Nanofibers for High Photoactivity Performance. Catal. Sci. Technol. 2021, 11, 4181–4195. [Google Scholar] [CrossRef]
  54. Liu, J.; Zhao, Z.; Xu, C.-m.; Duan, A.J. Simultaneous Removal of NOx and Diesel Soot over Nanometer Ln-Na-Cu-O Perovskite-Like Complex Oxide Catalysts. Appl. Catal. B Environ. 2008, 78, 61–72. [Google Scholar] [CrossRef]
  55. Wang, X.; Zhang, Y.; Li, Q.; Wang, Z.; Zhang, Z. Identification of Active Oxygen Species for Soot Combustion on LaMnO3 Perovskite. Catal. Sci. Technol. 2012, 2, 1822–1824. [Google Scholar] [CrossRef]
  56. Wang, J.; Yang, G.; Cheng, L.; Shin, E.W.; Men, Y. Three-Dimensionally Ordered Macroporous Spinel-Type MCr2O4 (M = Co, Ni, Zn, Mn) Catalysts with Highly Enhanced Catalytic Performance for Soot Combustion. Catal. Sci. Technol. 2015, 5, 4594–4601. [Google Scholar] [CrossRef]
  57. Huang, X.; Zheng, H.; Lu, G.; Wang, P.; Xing, L.; Wang, J.; Wang, G. Enhanced Water Splitting Electrocatalysis over MnCo2O4 Via Introduction of Suitable Ce Content. ACS Sustain. Chem. Eng. 2019, 7, 1169–1177. [Google Scholar] [CrossRef]
  58. Qiu, M.; Zhan, S.; Yu, H.; Zhu, D.; Wang, S. Facile Preparation of Ordered Mesoporous MnCo2O4 for Low-Temperature Selective Catalytic Reduction of NO with NH3. Nanoscale 2015, 7, 2568–2577. [Google Scholar] [CrossRef]
  59. Xiong, J.; Wu, Q.; Mei, X.; Liu, J.; Wei, Y.; Zhao, Z.; Wu, D.; Li, J. Fabrication of Spinel-Type PdxCo3-xO4 Binary Active Sites on 3D Ordered Meso-Macroporous Ce-Zr-O2 with Enhanced Activity for Catalytic Soot Oxidation. ACS Catal. 2018, 8, 7915–7930. [Google Scholar] [CrossRef]
  60. Cai, N.; Fu, J.; Chan, V.; Liu, M.; Chen, W.; Wang, J.; Zeng, H.; Yu, F. MnCo2O4@Nitrogen-Doped Carbon Nanofiber Composites with Meso-Microporous Structure for High-Performance Symmetric Supercapacitors. J. Alloy Compd. 2019, 782, 251–262. [Google Scholar] [CrossRef]
  61. Zhu, F.; Zhang, J.; Yang, B.; Shi, X.; Lu, C.; Yin, J.; Yu, Y.; Hu, X. Peanut Shaped MnCo2O4 Winded by Multi-Walled Carbon Nanotubes as an Efficient Cathode Catalyst for Li-O2 Batteries. J. Alloy Compd. 2018, 749, 433–440. [Google Scholar] [CrossRef]
  62. Zou, G.; Xu, Y.; Wang, S.; Chen, M.; Shangguan, W. The synergistic effect in Co-Ce oxides for catalytic oxidation of diesel soot. Catal. Sci. Technol. 2015, 5, 1084–1092. [Google Scholar] [CrossRef]
  63. Lunsford, J.H.; Yang, X.; Haller, K.; Laane, J.; Mestl, G.; Knoezinger, H. In Situ Raman Spectroscopy of Peroxide Ions on Barium/Magnesium Oxide Catalysts. J. Phys. Chem. 1993, 97, 13810–13813. [Google Scholar] [CrossRef]
  64. Li, M.; Feng, Z.; Xiong, G.; Ying, P.; Xin, Q.; Li, C. Phase Transformation in the Surface Region of Zirconia Detected by UV Raman Spectroscopy. J. Phys. Chem. B 2001, 105, 8107–8111. [Google Scholar] [CrossRef]
  65. Cao, C.; Xing, L.; Yang, Y.; Tian, Y.; Ding, T.; Zhang, J.; Hu, T.; Zheng, L.; Li, X. Diesel Soot Elimination over Potassium-Promoted Co3O4 Nanowires Monolithic Catalysts under Gravitation Contact Mode. Appl. Catal. B Environ. 2017, 218, 32–45. [Google Scholar] [CrossRef]
Figure 1. The curves of CO2 concentration and selectivity to CO2 (a) and soot conversion curves (b) over as-prepared monolith catalysts.
Figure 1. The curves of CO2 concentration and selectivity to CO2 (a) and soot conversion curves (b) over as-prepared monolith catalysts.
Catalysts 12 00295 g001
Figure 2. The CO2 concentration curves of cycle test (a) and resistance versus sulfur and water (b) over Mn0.7K0.3Co2O4-HC monolith catalyst.
Figure 2. The CO2 concentration curves of cycle test (a) and resistance versus sulfur and water (b) over Mn0.7K0.3Co2O4-HC monolith catalyst.
Catalysts 12 00295 g002
Figure 3. The photographs of as-prepared monolith catalysts. Blank honeycomb ceramics monolith (a), MnCo2O4-HC (b), Mn0.9K0.1Co2O4-HC (c), Mn0.8K0.2Co2O4-HC (d), Mn0.7K0.3Co2O4-HC (e).
Figure 3. The photographs of as-prepared monolith catalysts. Blank honeycomb ceramics monolith (a), MnCo2O4-HC (b), Mn0.9K0.1Co2O4-HC (c), Mn0.8K0.2Co2O4-HC (d), Mn0.7K0.3Co2O4-HC (e).
Catalysts 12 00295 g003
Figure 4. SEM images of blank honeycomb ceramics monolith (a,b), Mn0.7K0.3Co2O4-HC (c,d) and ultrasonic vibration test (e).
Figure 4. SEM images of blank honeycomb ceramics monolith (a,b), Mn0.7K0.3Co2O4-HC (c,d) and ultrasonic vibration test (e).
Catalysts 12 00295 g004
Figure 5. EDS result of elemental distribution on monolith catalysts and blank ceramics.
Figure 5. EDS result of elemental distribution on monolith catalysts and blank ceramics.
Catalysts 12 00295 g005aCatalysts 12 00295 g005b
Figure 6. XRD of Mn1−nKnCo2O4-HC monolith catalysts (a) and the corresponding powder catalysts (b).
Figure 6. XRD of Mn1−nKnCo2O4-HC monolith catalysts (a) and the corresponding powder catalysts (b).
Catalysts 12 00295 g006
Figure 7. Raman spectra of as-prepared monolith catalysts under different laser wavelength of 532 nm (a) and 325 nm (b).
Figure 7. Raman spectra of as-prepared monolith catalysts under different laser wavelength of 532 nm (a) and 325 nm (b).
Catalysts 12 00295 g007
Figure 8. Soot-TPR curves (a) and peak fitting (b) of as-prepared catalysts: MnCo2O4-P (1), Mn0.9K0.1Co2O4-P (2), Mn0.8K0.2Co2O4-P (3), Mn0.7K0.3Co2O4-P (4).
Figure 8. Soot-TPR curves (a) and peak fitting (b) of as-prepared catalysts: MnCo2O4-P (1), Mn0.9K0.1Co2O4-P (2), Mn0.8K0.2Co2O4-P (3), Mn0.7K0.3Co2O4-P (4).
Catalysts 12 00295 g008
Figure 9. H2-TPR curves of as-prepared catalysts: MnCo2O4-HC (a), Mn0.9K0.1Co2O4-HC (b), Mn0.8K0.2Co2O4-HC (c), Mn0.7K0.3Co2O4-HC (d).
Figure 9. H2-TPR curves of as-prepared catalysts: MnCo2O4-HC (a), Mn0.9K0.1Co2O4-HC (b), Mn0.8K0.2Co2O4-HC (c), Mn0.7K0.3Co2O4-HC (d).
Catalysts 12 00295 g009
Figure 10. XPS Spectra of Mn1−nKnCo2O4-HC catalysts: surface wide spectrum (a), Co2p (b), Mn2p (c) and O1s (d).
Figure 10. XPS Spectra of Mn1−nKnCo2O4-HC catalysts: surface wide spectrum (a), Co2p (b), Mn2p (c) and O1s (d).
Catalysts 12 00295 g010
Figure 11. In-situ Raman spectra of Mn0.7K0.3Co2O4-HC catalyst under laser wavelength of 532 nm. Reaction conditions: heating rate was 2 °C/min. The reaction atmosphere contained NO (2000 ppm), O2 (10%) and He as equilibrium gas, and the total flow rate was 50 mL/min. Monolith catalyst maintains loose contact with soot.
Figure 11. In-situ Raman spectra of Mn0.7K0.3Co2O4-HC catalyst under laser wavelength of 532 nm. Reaction conditions: heating rate was 2 °C/min. The reaction atmosphere contained NO (2000 ppm), O2 (10%) and He as equilibrium gas, and the total flow rate was 50 mL/min. Monolith catalyst maintains loose contact with soot.
Catalysts 12 00295 g011
Figure 12. In-situ Raman spectra of MnCo2O4-HC (a) and Mn0.7K0.3Co2O4-HC (b) catalysts under laser wavelength of 325 nm. Reaction conditions: heating rate was 2 °C/min. The reaction atmosphere contained NO (2000 ppm), O2 (10%) and He as equilibrium gas, and the total flow rate was 50 mL/min. Monolith catalyst maintains loose contact with soot.
Figure 12. In-situ Raman spectra of MnCo2O4-HC (a) and Mn0.7K0.3Co2O4-HC (b) catalysts under laser wavelength of 325 nm. Reaction conditions: heating rate was 2 °C/min. The reaction atmosphere contained NO (2000 ppm), O2 (10%) and He as equilibrium gas, and the total flow rate was 50 mL/min. Monolith catalyst maintains loose contact with soot.
Catalysts 12 00295 g012
Figure 13. The relationship between in-situ Raman vibration peak intensity and ratio of O2 and O22− for as-prepared catalysts under laser wavelength of 532 nm.
Figure 13. The relationship between in-situ Raman vibration peak intensity and ratio of O2 and O22− for as-prepared catalysts under laser wavelength of 532 nm.
Catalysts 12 00295 g013
Figure 14. The relationships among catalytic performance, K content, peak temperature of H2-TPR and active oxygen in as-prepared monolith catalysts.
Figure 14. The relationships among catalytic performance, K content, peak temperature of H2-TPR and active oxygen in as-prepared monolith catalysts.
Catalysts 12 00295 g014
Figure 15. The relationships among catalytic performance, K content, peak temperature of soot-TPR and active oxygen in as-prepared monolith catalysts.
Figure 15. The relationships among catalytic performance, K content, peak temperature of soot-TPR and active oxygen in as-prepared monolith catalysts.
Catalysts 12 00295 g015
Table 1. Catalytic performance, stability and tolerance of sulfur dioxide and water for soot combustion of as-prepared monolith catalysts.
Table 1. Catalytic performance, stability and tolerance of sulfur dioxide and water for soot combustion of as-prepared monolith catalysts.
Title 1T10/°CT50/°CT90/°CTm/°C S C O 2 m / %
Soot (without catalyst)45156260758838.9
MnCo2O4-HC33645153848998.2
Mn0.9K0.1Co2O4-HC32042851143697.3
Mn0.8K0.2Co2O4-HC31342450844299.4
Mn0.7K0.3Co2O4-HC31041550443999.1
Mn0.7K0.3Co2O4-HC-cycle 231141550342998.9
Mn0.7K0.3Co2O4-HC-cycle 331842150646999.2
Mn0.7K0.3Co2O4-HC-cycle 431142351645398.2
Mn0.7K0.3Co2O4-HC-cycle 532743151344598.5
Mn0.7K0.3Co2O4-HC (SO2)31241449643899.3
Mn0.7K0.3Co2O4-HC (H2O)37048057348698.9
Reaction conditions: heating rate was 2 °C/min. The reaction atmosphere contained NO (2000 ppm), O2 (10%) and argon gas as equilibrium gas, and the total flow rate was 50 ml/min.
Table 2. XRF results of oxide content on monolith catalysts and blank ceramics.
Table 2. XRF results of oxide content on monolith catalysts and blank ceramics.
ElementBlank Ceramics (wt %)MnCo2O4-HC (wt %)Mn0.7K0.3Co2O4-HC (wt %)
MnO--0.640.44
Co3O4--0.960.91
K2O0.210.230.42
SiO245.7148.2345.71
Al2O335.2434.7935.24
MgO16.8213.4416.83
other2.021.710.45
Table 3. The temperature of each peak in the Soot-TPR and H2-TPR curves over as-prepared monolith catalysts.
Table 3. The temperature of each peak in the Soot-TPR and H2-TPR curves over as-prepared monolith catalysts.
CatalystSoot-TPR aH2-TPR b
Peak 1 (°C)Peak 2 (°C)Peak 3 (°C)R1 (mol × 10−5)Peak 1 (°C)Peak 2 (°C)Peak 3 (°C)R2 (mmol/g)
MnCo2O4533687-8.81350509-0.165
Mn0.9K0.1Co2O442659473414.73504284910.204
Mn0.8K0.2Co2O4473--16.92403314830.180
Mn0.7K0.3Co2O4407435-15.12153184010.214
R1 is the amount of active [O*] consumed by Mn1−nKnCo2O4-P catalysts to generate CO2 in the soot-TPR reaction; R2 is the amount of hydrogen consumed by Mn1−nKnCo2O4-HC catalysts in H2-TPR reaction; Soot-TPR a: the reaction was carried out on a powder catalyst; H2-TPR b: The reaction was carried out on fragments of the monolithic catalyst.
Table 4. XPS results for transition metal valence and surface oxygen species of as-prepared monolith catalysts.
Table 4. XPS results for transition metal valence and surface oxygen species of as-prepared monolith catalysts.
CatalystsMn SpeciesCo SpeciesO SpeciesK Species
Atomic aMn2+bMn3+bAtomic aCo2+cCo3+cO2+ O22−dO2−dAtomic a
MnCo2O4-HC15.3%62.2%37.8%15.9%48.5 %51.5%17.1%82.9%
Mn0.9K0.1Co2O4-HC14.3%62.7%37.3%15.2%41.9%58.1%20.6%79.4%2.2%
Mn0.8K0.2Co2O4-HC17.3%63.7%36.3%17.0%34.6%65.4%21.3%78.7%3.1 %
Mn0.7K0.3Co2O4-HC14.7%64.1%35.9%15.4%30.9%69.1%25.8%74.2%4.3%
a: The data of atomic (%) came from XPS Survey; b: The sub-peaks were deconvoluted by XPS peak with deviation value less than 6 and the ratio was calculated according to formula Mn n + Mn 2 + + Mn 3 + ; c: The sub-peaks were deconvoluted by XPS peak with deviation value less than 6 and the ratio was calculated according to formula Co n + Co 2 + + Co 3 + ; d: The sub-peaks were deconvoluted by XPS peak with deviation value less than 6 and the ratio was calculated according to formula O 2 + O 2 2 O 2 + O 2 2 + O 2 .
Table 5. The weights of monolith Mn1−nKnCo2O4-HC catalysts and blank honeycomb ceramics (HC).
Table 5. The weights of monolith Mn1−nKnCo2O4-HC catalysts and blank honeycomb ceramics (HC).
CatalystmHC/gmfre/gΔm/g
MnCo2O4-HC55.6758.292.62
Mn0.9K0.1Co2O4-HC56.3158.782.47
Mn0.8K0.2Co2O4-HC57.3959.922.53
Mn0.7K0.3Co2O4-HC55.6558.242.59
mHC: the weight of blank ceramics; mfer: the weight of fresh monolith catalyst; Δm: the weight of coat-layer after calcined. Δm = mfer−mhc.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhao, K.; Li, J.; Wang, L.; Li, D.; Liu, B.; Li, R.; Yu, X.; Wei, Y.; Liu, J.; Zhao, Z. Preparation of Cordierite Monolith Catalysts with the Coating of K-Modified Spinel MnCo2O4 Oxide and Their Catalytic Performances for Soot Combustion. Catalysts 2022, 12, 295. https://doi.org/10.3390/catal12030295

AMA Style

Zhao K, Li J, Wang L, Li D, Liu B, Li R, Yu X, Wei Y, Liu J, Zhao Z. Preparation of Cordierite Monolith Catalysts with the Coating of K-Modified Spinel MnCo2O4 Oxide and Their Catalytic Performances for Soot Combustion. Catalysts. 2022; 12(3):295. https://doi.org/10.3390/catal12030295

Chicago/Turabian Style

Zhao, Kun, Jianmei Li, Lanyi Wang, Dong Li, Bonan Liu, Renjie Li, Xuehua Yu, Yuechang Wei, Jian Liu, and Zhen Zhao. 2022. "Preparation of Cordierite Monolith Catalysts with the Coating of K-Modified Spinel MnCo2O4 Oxide and Their Catalytic Performances for Soot Combustion" Catalysts 12, no. 3: 295. https://doi.org/10.3390/catal12030295

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