Next Article in Journal
The Application of the Intolerance of Uncertainty Model to Gambling Urge and Involvement
Next Article in Special Issue
The Effect of Iron Content on the Ammonia Selective Catalytic Reduction Reaction (NH3-SCR) Catalytic Performance of FeOx/SAPO-34
Previous Article in Journal
How Do the Young Perceive Urban Parks? A Study on Young Adults’ Landscape Preferences and Health Benefits in Urban Parks Based on the Landscape Perception Model
Previous Article in Special Issue
Low-Temperature Catalytic Ozonation of Multitype VOCs over Zeolite-Supported Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Soot Combustion over Cu–Co Spinel Catalysts: The Intrinsic Effects of Precursors on Catalytic Activity

1
Faculty of Maritime and Transportation, Ningbo University, Ningbo 315211, China
2
State Key Laboratory of Coal Combustion, School of Energy and Power Engineering, Huazhong University of Science and Technology, Wuhan 430074, China
*
Authors to whom correspondence should be addressed.
Int. J. Environ. Res. Public Health 2022, 19(22), 14737; https://doi.org/10.3390/ijerph192214737
Submission received: 27 August 2022 / Revised: 23 October 2022 / Accepted: 28 October 2022 / Published: 9 November 2022

Abstract

:
In this work, a series of CuCo2O4-x (x = N, A and C) catalysts were synthesized using different metal salt precursors by urea hydrothermal method for catalytic soot combustion. The effect of CuCo2O4-x catalysts on soot conversion and CO2 selectivity in both loose and tight contact mode was investigated. The CuCo2O4-N catalyst exhibited outstanding catalytic activity with the characteristic temperatures (T10, T50 and T90) of 451 °C, 520 °C and 558 °C, respectively, while the CO2 selectivity reached 98.8% during the reaction. With the addition of NO, the soot combustion was further accelerated over all catalysts. Compared with the loose contact mode, the soot conversion was improved in the tight contact mode. The CuCo2O4-N catalysts showed better textural properties compared to the CuCo2O4-A and CuCo2O4-C, such as higher specific surface areas and pore volumes. The XRD results confirmed that the formation of a CuCo2O4 crystal phase in all catalysts. However, the CuO crystal phase only presented in CuCo2O4-N and CuCo2O4-A. The relative contents of Cu2+, Co3+ and Oads on the surface of CuCo2O4-x (x = N, A and C) catalysts were analyzed by XPS. The CuCo2O4-N catalyst displayed the highest relative content of Cu2+, Co3+ and Oads. The activity of catalytic soot combustion showed a good correlation with the order of the relative contents of Cu2+, Co3+ and Oads. Additionally, the CuCo2O4-N catalyst exhibited lower reduction temperature compared to the CuCo2O4-A and CuCo2O4-C. The cycle tests clarified that the copper–cobalt spinel catalyst obtained good stability. In addition, based on the Mars–van Krevelen mechanism, the process of catalytic soot combustion was described combined with the electron transfer process and the role of oxygen species over CuCo2O4 spinel catalysts.

1. Introduction

Soot particles are emitted from incomplete combustion of hydrocarbon fuels in diesel engines [1]. Mitigating soot particle emissions from diesel engines has raised much attention due to the negative impact of soot particles on human health and environmental protection [2,3]. A diesel particulate filter (DPF) is the most effective after-treatment technology for soot particles from diesel engines. However, a common issue of pore blockage over DPFs after a period of use should be addressed [4]. One of the most effective measures to improve soot oxidation is to coat the DPF with catalysts and form a catalyzed DPF (CDPF) [5]. CDPFs with highly active catalysts can accelerate soot oxidation and lower the ignition temperatures.
Various catalysts have been proposed and optimized to accelerate the catalytic soot combustion, while Pt-based catalysts have shown superb soot oxidation performance under practical conditions [6,7]. However, due to the high cost of noble metals, transition metal-based catalysts have been intensively studied for catalytic soot oxidation more recently [8]. Copper–cobalt catalysts play an important role as substitutes for noble metals in the field of catalytic soot combustion due to their outstanding redox properties and thermal stability [9]. Zhang et al. reported that a spinel-type CuCo2O4 catalyst showed a lower T50 temperature (574 °C) of catalytic soot combustion than that of Co3O4 catalysts (580 °C) due to its higher relative content of Co3+ species [10]. Jampaiah et al. found that a CuCo-MnO2 catalyst exhibited a lower T90 temperature (485 °C) of catalytic soot combustion than that of NiCo-MnO2 catalysts (513 °C) under the loose contact mode due to more oxygen vacancies and active sites on the catalyst surfaces [11]. Zhang et al. tested the stability of CuCo2O4 spinel catalysts in four consecutive soot combustion cycles and found that the values of T50 for catalytic soot combustion increased from 539 °C to 544 °C, 547 °C and 549 °C, respectively [12].
Recently, the effect of metal salt precursor on the textural and redox properties of catalysts was found to play a key role in catalytic activity and product selectivity [13]. Yang et al. found that the catalytic activity of selective acetylene hydrogenation at 50 °C over the Pd/Al2O3 catalyst using acetate as the Pd precursor reached 78.6%, which was 31.1% higher than that using chloride precursors since Cl residuals existed in the form of P d C l 4 2 , resulting in the decrease of the electron density of Pd atoms [14]. Wang et al. reported that the n-hexadecane conversion over Pt/ZSM-22 catalysts with Pt(NH3)4Cl2 precursors was ~9.1% higher compared to that over catalysts using (Pt(NO3)2 and H2PtCl6 as the precursors due to the higher platinum dispersion [15]. Yun et al. found that the Ni/Al2O3 catalyst using acetate salt precursors achieved the highest activity (85.6%) in the steam reforming of acetic acid at 450 °C, while the catalytic activity of the catalyst prepared from chloride salts was only 38.7% [16]. These results indicated that the role of precursors should be preferentially considered during the designing and preparation of heterogeneous catalysts.
In this work, copper–cobalt spinel oxides were prepared with a urea hydrothermal method using different copper and cobalt precursors (nitrate, chloride and acetate). The physicochemical properties of these catalysts were characterized by N2 adsorption–desorption, X-ray diffraction (XRD), scanning electron microscopy (SEM), X-ray photoelectron spectroscopy (XPS), temperature-programed reduction of H2 (H2-TPR) and temperature-programed desorption of O2 (O2-TPD). The catalytic activity of soot combustion was investigated by temperature-programed oxidation (TPO) experiments under various reaction conditions.

2. Experimental Study

2.1. Catalyst Preparation

In this work, a series of CuCo2O4-x catalysts was synthesized by urea hydrothermal methods with different precursors, where x represented the types of copper and cobalt precursors, e.g., the nitrates (N), acetate (A) and chloride (C), respectively. All chemicals used were of analytical reagent grade and purchased from Macklin Co., Ltd. (Shanghai, China). Taking CuCo2O4-N as an example, 0.02 mol Cu(NO3)2·3H2O, 0.04 mol Co(NO3)3·6H2O and 0.06 mol CH4N2O were dissolved in 100 mL of deionized water and stirred for 1 h at room temperature. Afterwards, the mixed solution was transferred to a polytetrafluoroethylene-lined high pressure reactor placed in a thermostat at 150 °C for hydrothermal reaction. After 12 h, the resulting solution was filtered to obtain a precipitate. The precipitate was then washed three times with deionized water and dried in an oven at 110 °C overnight. Finally, the precipitate was calcined in a muffle furnace at 700 °C for 6 h and sieved to 40–60 meshes. The prepared catalysts were denoted as CuCo2O4-N, CuCo2O4-A and CuCo2O4-C, respectively.

2.2. Catalyst Characterizations

N2 adsorption–desorption analysis was conducted at 77 K to determine the specific surface area, pore size distribution and average pore diameter of the catalysts using TriStar II 3020 (Micromeritics, Norcross, GA, USA). The specific surface areas were calculated using the Brunauer–Emmett–Teller (BET) method. Total pore volume and pore size distribution were calculated via the Barrett–Joyne–Halenda (BJH) method at the relative pressure of p/p0 = 0.99. The crystal structure of catalysts was observed by X-ray diffraction (XRD) using a diffractometer system (D-Max 2000, Rigaku, Tokyo, Japan) with Cu-Kα radiation operating at 40 kV and 30 mA. All catalysts were scanned in the range of 10° to 80° with a step size of 0.02°. A scanning electron microscopy (SEM) instrument (JSM-7001F, JEOL, Tokyo, Japan) was used to observe the surface morphology of catalysts at an accelerating voltage of 10 kV. The X-ray photoelectron spectra (XPS) were measured with Thermo Escalab 250Xi equipment with monochromatic Al-Ka X-ray radiation at 150 W. All binding energies were calibrated using the C 1s photoelectron peak at 284.8 eV. The redox properties of all catalysts were analyzed using the temperature-programed reduction of the H2 (H2-TPR) apparatus (Autochem II 2920, Micromeritics, Norcorss, GA, USA). To remove impurities, 16 mg catalysts were pretreated at 250 °C for 1 h and cooled down to room temperature before each test. Then, the catalysts were heated from room temperature to 800 °C at a heating rate of 10 °C·min−1 in a 30 mL·min−1 feeding gas flow (10 vol.% H2/Ar). The amount of H2 consumption was calculated based on the H2-TPR profiles. The profiles of the temperature-programed desorption of O2 (O2-TPD) were performed using the same apparatus as that for H2-TPR. Before the measurement, 200 mg of catalysts were pretreated in an He stream at 200 °C for 1 h and cooled down to room temperature. Then, the adsorption of O2 was conducted at 70 °C for 1 h in a gas mixture of 3 vol.% O2/He (30 mL·min−1). Subsequently, the sample was purged by a flowing pure He stream to remove excessive and weakly adsorbed O2. Finally, the sample was heated to 800 °C with a heating rate of 10 °C·min−1 in a pure He flow (30 mL·min−1), and the desorption profile was recorded.

2.3. Experimental System

Figure 1 shows a schematic diagram of the experimental setup. The catalytic activity of the CuCo2O4-x (x = N, C and A) catalysts was evaluated by temperature-programed oxidation (TPO) experiments. Printex-U (Degussa, with the size of 20–30 nm) was used as the model soot in this study. For each test, 180 mg of catalyst powder was mixed with 20 mg of soot in loose (mixing with a spatula for 5 min) or tight (grinding in an agate mortar for 5 min) contact mode. Then, the resulting soot–catalyst mixtures (200 mg) were mixed with 400 mg of inert silica (40–60 meshes) for another 5 min to avoid the formation of hot spots during the reaction. The experimental procedure was as follows: Firstly, the mixture of catalyst and inert silica after reaction was collected from the quartz tube and then placed in an agate mortar. Secondly, 20 mg of soot was mixed with the obtained mixture of the catalyst and inert silica after reaction in loose contact mode. Finally, the above obtained mixture of soot, catalyst and inert silica was filled back into the quartz tube for next catalytic soot combustion reaction cycle. The reaction temperature was increased from 50 °C to 700 °C at a heating rate of 5 °C·min−1.
All gas streams from the gas cylinders were regulated by mass flow controllers (Sevenstars D07-B, Beijing, China). The mixed gases (10 vol.% O2 with balanced N2) were fed into the reactor at the flow rate of 200 mL·min−1 during the TPO experiment. In addition, 1000 ppm NO was added into the feeding gas to investigate the effect of NO on catalytic soot combustion when necessary. The outlet concentrations of CO and CO2 were monitored online using an infrared (IR) gas analyzer (GXH-3010/3011AE, Huayun, Beijing, China) with the accuracy of ±3%. The temperatures at which 10%, 50% and 90% of the soot was oxidized (denoted as T10, T50 and T90, respectively) were recorded as indicators of the catalytic activity. Soot conversion (denoted as α) and CO2 selectivity (denoted as SCO2) were calculated by integrating the CO and CO2 concentration curves with time as follows:
α   ( % ) = 0 t ( [ C O 2 ] o u t + [ C O ] o u t ) d t M × 100 %
S C O 2 ( % ) = 0 t [ C O ] o u t d t 0 t ( [ C O 2 ] o u t + [ C O ] o u t ) d t × 100 %
where [CO2]out and [CO]out are the real-time concentrations of CO and CO2 at the reactor outlet, respectively, and M is the weight of the initially packed soot.

3. Results and Discussion

3.1. Textural Properties of the Catalysts

The specific surface area (SBET), pore volume and pore size of CuCo2O4-x (x = N, A and C) catalysts are obtained with a N2 adsorption–desorption experiment (Table 1). The CuCo2O4-N catalyst shows the largest specific surface area of 2.1 m2·g−1, followed by CuCo2O4-A (2.0 m2·g−1) and CuCo2O4-C (1.2 m2·g−1). The pore volumes of CuCo2O4-N, CuCo2O4-A and CuCo2O4-C catalysts are 4.5 mm3·g−1, 4.3 mm3·g−1 and 2.5 mm3·g−1, respectively. The specific surface area and pore volume of the CuCo2O4 catalysts prepared with nitrate and acetate metal salt show no significant differences, while the specific surface area and pore volume of CuCo2O4-C catalyst dramatically decreases. Yu et al. also reported the formation of hydrochloric acid from chloride salts during calcination, which may inhibit the formation of developed pore systems and negatively affect the specific surface area [16]. Compared with conventional metal oxides and supported catalysts, the specific surface area, pore volume and average pore size of the prepared Cu–Co spinel catalysts are rather low. The results could be ascribed to the formation of the well-crystallized spinel structure under high calcination temperature (700 °C) [12,17].
Figure 2 shows the XRD patterns of the CuCo2O4-x (x = N, A and C) catalysts. Sharp and intense diffraction peaks of copper–cobalt spinel phases are observed for all samples, suggesting the formation of well-crystallized structures at high calcination temperature of 700 °C. The diffraction peaks observed at 19.1°, 31.4°, 36.9°, 45.1°, 56.0°, 59.6°, 68.9° and 77.5° are attributed to tetragonal spinel crystalline of CuCo2O4 (JCPDS No. 01-1155) [18]. Meanwhile, metal oxide CuO phase (JCPDS No. 80-0076) is observed at 35.6°, 48.6° and 61.7° over CuCo2O4-N and CuCo2O4-A catalysts [19]. No distinct CuO phase is found on the CuCo2O4-C catalyst. Based on the characteristic peak of CuCo2O4 (3 1 1) crystal face, the crystal size of the CuCo2O4-x catalysts were calculated using the Scherrer equation (Table 1). The crystal size of the CuCo2O4-N catalyst (38.0 nm) is slightly smaller than that of CuCo2O4-A and CuCo2O4-C catalysts. Wen et al. also found that the Cl- coordination anion was an important contributor to the formation of larger clusters of CuCo2O4 [20].
Figure 3 shows the representative SEM images of the CuCo2O4-x (x = N, A and C) catalysts. The CuCo2O4-N catalyst exhibits a sheetlike morphology with tiny spherical particles on its surface. The stacking of spherical particles and its sheetlike morphologies promote the formation of porous structures (Figure 3a). The phenomenon of particle agglomeration is exhibited over the CuCo2O4-A and CuCo2O4-C catalyst (Figure 3c,e). As shown in Figure 3b, the spherical particles are dispersed on the sheetlike morphology and formed a dendritic structure. However, tiny spherical particles are hardly generated on the surfaces of the CuCo2O4-A and CuCo2O4-C catalysts (Figure 3d,f), while the agglomeration of bulk particles might have resulted in the blockage of the pore systems. These tiny spherical particles can improve the contact between the catalyst and soot particles, which can facilitate the utilization of the catalyst active sites for catalytic soot combustion. The particle sizes of CuCo2O4-A and CuCo2O4-C catalysts increased obviously compared to the CuCo2O4-N catalysts.

3.2. Redox Properties of the Catalysts

The Co 2p spectra of the CuCo2O4-x (x = N, A and C) catalysts are shown in Figure 4a. The peaks of Co 2p3/2 and Co 2p1/2 are observed in the range of 776.0–784.0 eV and 792.0–800.0 eV, respectively [21]. The peaks between 784.0–792.0 eV and 800.0–808.0 eV are attributed to the satellite peaks of Co2+ [22,23]. The Co3+ and Co2+ signals are obtained after the deconvolution of the Co 2p3/2 and Co 2p1/2 spectra. The peaks centered at 779.7 eV and 794.8 eV correspond to the Co3+ species, while the peaks located at 780.8 eV and 796.4 eV belong to the Co2+ species [24]. The Cu 2p spectra of all catalysts are also given in Figure 4b. The peaks observed at 932.6 eV belong to the reduced Cu species (Cu+ or Cu0), while the prominent signals at 934.6 eV are ascribed to the Cu2+ species [25]. Additionally, the satellite peaks between 937.0 eV to 946.0 eV also confirm the existence of divalent Cu species [19]. Figure 4c shows the Cu LMM Auger spectra of all CuCo2O4 catalysts. The peaks at the kinetic energy of 912.6 eV correspond to the Cu+ species, while the peaks at 917.8 eV are attributed to the Cu2+ species [26]. However, Cu0 species are not observed on the Cu LMM Auger spectrum. These results suggest that the reduced copper species on the surfaces of CuCo2O4-x (x = N, A and C) catalysts mainly exists as Cu+. The de-convoluted XPS signals of O 1s are shown in Figure 4d. Two types of oxygen species present on the surface of the CuCo2O4 catalysts. The peaks around 529.8 eV correspond to the lattice oxygen species (Olatt), while the peaks around 531.4 eV are attributed to the adsorbed oxygen species (Oads) [27].
The relative contents of Cu2+ and Co3+ on the surface of CuCo2O4-x catalysts are given in Table 2. The highest relative content of Co3+/Cototal (38.4%) is obtained over the CuCo2O4-N catalyst, followed by CuCo2O4-A (35.1%) and CuCo2O4-C (33.8%). Moreover, the highest relative content of Cu2+/Cutotal is also achieved over the CuCo2O4-N catalyst (51.7%), followed by CuCo2O4-A (48.2%) and CuCo2O4-C (47.8%). The highest relative content of Oads/(Oads + Olatt) (46.8%) is found over the CuCo2O4-N catalyst, followed by CuCo2O4-A (37.2%) and CuCo2O4-C (36.5%) (Table 2). The adsorbed oxygen species were more chemically active than lattice oxygen in catalytic soot combustion, indicating a better soot conversion performance over the CuCo2O4-N catalyst with more Oads species [28].
For the CuCo2O4-N catalyst, two reduction peaks are observed (Figure 5). The first reduction peak in the range of 150–200 °C is attributed to the reduction of aggregated CuO, while the second peak between 200 °C and 250 °C can be ascribed to the reduction of Cu–Co mixed oxides [29]. The CuCo2O4-A catalyst exhibits shoulder peaks at 200–300 °C, the first reduction peak (244 °C) corresponds to the reduction of Cu–Co mixed oxides, and the latter reduction peak at 278 °C represents the reduction of Co3+ to Co2+ [30]. The CuCo2O4-C catalyst also shows two major peaks between 250 °C and 350 °C. The weak reduction peak at 299 °C is ascribed to the reduction of Co3+ to Co2+, while the reduction peak at 331 °C represents the reduction of Co2+ to Co0+ [31]. These results suggest that CuCo2O4-N catalyst has better reducibility at relatively low temperatures. The amount of H2 consumption was in the order of CuCo2O4-C > CuCo2O4-A > CuCo2O4-N. Although the CuCo2O4-C catalysts show the higher H2 consumption (14.5 mmol·g−1), the H2 consumption of the CuCo2O4-N catalyst is likely mainly concentrated in the low temperature range (150–250 °C). Therefore, the distribution of oxygen species was further investigated.
Figure 6a shows the O2-TPD profiles of the CuCo2O4-x (x = N, A and C) catalysts. The oxygen desorption peaks below 500 °C belong to adsorbed oxygen species (e.g., O2, O 2 and O-, labeled as α-O2) [32], while the oxygen desorption peaks between 600 °C and 850 °C belong to the lattice oxygen species (O2−, labeled as β-O2) [33]. Figure 6b shows the enlarged O2-TPD profiles in the temperature range of 50 °C to 550 °C. The oxygen desorption peaks within 50–300 °C and 300–500 °C correspond to the physically adsorbed oxygen (α1-O2) and chemically adsorbed oxygen (α2-O2) species, respectively [34,35]. Generally, the adsorption of oxygen followed the procedure of O2 O 2 →O-→O2− [27]. The CuCo2O4-N catalyst shows an extra oxygen desorption peak in the temperature range of 200 °C to 250 °C compared to the CuCo2O4-A and CuCo2O4-C. The extra oxygen desorption peak could be attributed to physically adsorbed oxygen (α1-O2), indicating that the catalysts had better oxygen mobility [36]. Similarly, Li et al. also reported that the CuCo2O4 and NiCo2O4 catalysts showed an extra oxygen desorption peak in the temperature range of 200 °C to 250 °C compared to the ZnCo2O4. The catalytic performance of CuCo2O4 and NiCo2O4 catalysts were found to be far superior to that of ZnCo2O4 in toluene combustion [7]. The desorption amount of α1-O2 and α2-O2 species were calculated according to the desorption peaks in the O2-TPD profiles. As shown in Table 2, the amount of O2 desorption was in the order of CuCo2O4-N (41.2 μmol·g−1) > CuCo2O4-A (27.1 μmol·g−1) > CuCo2O4-C (25.6 μmol·g−1), which confirms the existence of more adsorbed oxygen species over the CuCo2O4-N catalyst.

3.3. Activity of CuCo2O4-x for Soot Conversion

3.3.1. Soot Conversion in O2/N2

The catalytic activity of CuCo2O4-x (x = N, A and C) catalysts for soot combustion were studied using a TPO experiment under the loose contact mode. Firstly, the soot conversion was investigated under the carrier gases of 10 vol.% O2 balanced with N2. The characteristic temperatures (T10, T50 and T90) of soot conversion are 530 °C, 586 °C and 614 °C, respectively in the absence of a catalyst, which is much higher than those in the presence of the CuCo2O4 catalysts (Figure 7 and Table 3). The catalysts prepared with different metal salt precursors exhibit different catalytic activity in the soot conversion. Among them, the CuCo2O4-N catalyst achieves the highest soot conversion, while the values of T10, T50 and T90 are 451 °C, 520 °C and 558 °C, respectively. The T50 values for the CuCo2O4-x (x = N, A and C) catalysts followed the order of CuCo2O4-N (520 °C) < CuCo2O4-A (539 °C) < CuCo2O4-C (550 °C) < no catalysts (586 °C). The CO2 selectivity is also improved remarkably from 81.8% over the CuCo2O4-C to almost 100% over the CuCo2O4-N catalysts.
The CuCo2O4-N and CuCo2O4-A catalysts exhibit larger specific surface area and pore volume compared to the CuCo2O4-C catalyst. It is widely recognized that a larger specific surface area could facilitate the contact between gaseous reactants (e.g., O2, NO and NO2) and the catalyst [37,38]. SEM images further confirmed the generation of more developed pore structure systems of the CuCo2O4-N catalyst compared to the CuCo2O4-A and CuCo2O4-C catalysts. Fang et al. reported that a perovskite-type macro/mesoporous La1–xKxFeO3–δ catalysts with large specific surface area and pore volume could improve the utilization of catalytic sites in the soot combustion reaction [39]. Furthermore, the presence of an appropriate amount of single metal oxide CuO on the CuCo2O4 surface could contribute to the crystal lattice distortion of the spinel phase and promote the formation of a defect structure, thereby improving the catalytic activity of the soot combustion reaction [40].
The redox properties of CuCo2O4-x catalysts played a crucial role in catalytic soot combustion. The higher relative content of Co3+/Cototal (38.4%) and Cu2+/Cutotal (51.7%) obtained over CuCo2O4-N catalysts were much higher compared to the CuCo2O4-A (35.1% and 48.2%) and CuCo2O4-C (33.8% and 47.8%), respectively. Spinel-type CuCo2O4 catalysts possessed outstanding redox properties since the synergistic effects between Co3+ and Cu2+ species could enhance the adsorption-activation properties of oxygen species for soot conversion [41]. Abundant Co3+ species would increase the anionic defects on catalyst surfaces, leading to the formation of more oxygen vacancies [42]. Zhang et al. reported that the presence of Cu2+ species induced structural defects on cobalt oxide and weakened the Co-O bonds, which could facilitate the activation of oxygen and improve the reducibility of CuCo2O4 [43]. In addition, the CuCo2O4-N catalysts showed the highest relative content of Oads/(Oads + Olatt) (46.8%) compared to the CuCo2O4-A (37.2%) and CuCo2O4-C (36.5%). The Oads species possessed better mobility than Olatt species and could participate in catalytic soot combustion via the contact points between catalyst pellets and soot particulates [35].
The lower reduction peaks temperature proved that the CuCo2O4-N catalysts (173 °C) has excellent low temperature reduction performances compared to the CuCo2O4-A (184 °C) and CuCo2O4-C (247 °C). The total H2 consumption amount of the CuCo2O4-N catalysts (13.4 mmol·g−1) was slightly lower than that of CuCo2O4-A (13.6 mmol·g−1) and CuCo2O4-C (14.5 mmol·g−1). The CuCo2O4-N catalyst possessed more adsorbed oxygen species (41.2 μmol·g−1) compared to the CuCo2O4-A (27.1 μmol·g−1) and CuCo2O4-C (25.6 μmol·g−1). It was suggested that the CuCo2O4-N catalysts could release more adsorbed oxygen species below 500 °C [44]. Compared with lattice oxygen species, adsorbed oxygen species were more important in soot conversion due to their better oxygen mobility [45]. He et al. also reported that the adsorbed oxygen species released at low temperatures were more important than lattice oxygen species since the lattice oxygen species could be activated and released only at high temperatures [35]. Lower reduction temperature and abundant adsorbed oxygen species may be crucial factors for soot conversion over the CuCo2O4-N catalyst at lower temperature.

3.3.2. Soot Conversion in NO/O2/N2

NO is one of the main components in diesel engine exhaust and has a great effect on soot conversion. Therefore, soot conversion was investigated over the CuCo2O4-x (x = N, A and C) catalysts in the presence of 1000 ppm NO (Figure 8 and Table 4). The addition of NO resulted in lower T10, T50 and T90 values regardless of the presence of a catalyst. The T50 values for the CuCo2O4-x (x = N, A and C) catalysts also follow the order of CuCo2O4-N (414 °C) < CuCo2O4-A (458 °C) < CuCo2O4-C (485 °C) < no catalysts (575 °C).
Using CuCo2O4-N as an example, the T50 value decreases from 520 °C to 414 °C with NO addition, while the CO2 selectivity also decreases from 98.8% to 96.6%. NO could be oxidized with O2 to form NO2 in gas phase (Reaction (3)) [46]. NO could be also adsorbed on the surface of CuCo2O4-N catalyst and then oxidized by the active oxygen to form NO2 species, while could promote the formation of surface oxygen vacancies (Reaction (4)). NO2 is a stronger oxidant and participate in catalytic soot combustion as a mobile gaseous oxidant [46]. Possible pathways for the reaction of NO2 with soot are shown in Reaction (5) [47]. NO2 could attack the soot on the surface, resulting in the generation of CO and NO, while CO was eventually oxidized by O2 to CO2, with NO again participating in the NOx-assisted catalytic soot combustion [37]. However, the slight decrease in CO2 selectivity could be attributed to the rapid oxidation of soot by NO2, resulting in the incomplete oxidation of soot to CO [48].
2 N O + O 2 2 N O 2
N O + O * N O 2 + O v
N O 2 + C ( S oo t ) C O + N O
where O* represents the active oxygen species, Ov represents the surface oxygen vacancy and C(Soot) represents the soot particulate.

3.3.3. Effect of the Contact Mode

The contact mode between the catalyst and soot could greatly affect the activity of soot conversion [41]. The soot conversion and CO2 selectivity were investigated under both loose and tight contact modes over the CuCo2O4-N catalyst (Figure 9 and Table 5). Under the tight contact mode, soot conversion is accelerated at lower temperatures regardless of the presence of NO compared to the loose contact mode, while the CO2 selectivity reaches up to 100%. The increase of catalyst–soot contact points in the tight contact mode could contribute to better utilization of active sites on the catalyst surface compared to the loose contact mode [4,27]. Machida et al. found that the utilization/transfer of lattice oxygen species was more efficient under the tight contact mode. Besides, the case of “O2 slip” may be decreased in the tight contact mode, which increases the utilization of the released oxygen species to the soot combustion [49].

3.3.4. Stability Test

The stability test of CuCo2O4-N catalysts was conducted for four consecutive cycles of soot combustion. As shown in Figure 10, the characteristic temperatures (T10, T50 and T90) increased slightly with the increase in the number of cycles of soot combustion. For examples, the values of T50 for catalytic soot combustion for each cycle are 520 °C, 524 °C, 526 °C and 528 °C, respectively. Figure 11 shows the deconvoluted XPS spectra of O 1s for the CuCo2O4-N catalyst after four consecutive cycles of soot combustion. The relative content of Oads/(Oads + Olatt) of CuCo2O4-N catalyst decreases from 46.8% to 40.6% (Table 6). The slight decrease in soot catalytic activity may be due to the minor attenuation of the relative content of Oads/(Oads + Olatt). Chen et al. also reported the catalytic soot performance of Ag/Co3O4 and found that the decrease in adsorbed oxygen species was an important reason for the decrease of catalytic soot combustion [50].

4. Reaction Mechanisms for Catalytic Soot Combustion

The potential reaction pathways of catalytic soot combustion over CuCo2O4 were discussed. The excellent catalytic soot combustion of the copper–cobalt spinel catalyst performance depended on the redox properties of the catalysts [41]. The interactions between Cu and Co species in the CuCo2O4 catalyst played a crucial role in the redox reactions. Previous studies showed that the redox pairs of Cu2+/Cu+ and Co3+/Co2+ were involved in the electron transfer process from Co3+ to Cu2+ within the Cu2+–O–Co3+ connections in the copper–cobalt spinel catalysts [31,51]. The Cu2+–O–Co3+ connections could bridge the oxygen transfer within the structure and reduce the redox potential of the Cu species, which ensures the improvement of reducibility for both Cu and Co oxides in the Cu–Co catalysts (Reaction (6)), facilitating the NOx-assisted mechanism in the soot combustion reaction [46].
C u 2 + / C u + O 2 / O 2 e C o 3 + / C o 2 +
Catalytic soot combustion over the copper–cobalt spinel catalyst followed the Mars–van Krevelen (MvK) mechanism, in which the abundance of active oxygen species directly determined the performance of catalytic soot combustion [5]. At the beginning of soot conversion, plenty of active oxygen species on the catalyst surface would come into contact with the soot, resulting in soot combustion at a relatively low temperature and the subsequent formation of oxygen vacancies (Ov) (Reaction (7)) [49]. As the reaction proceeded, the consumed reactive oxygen species could be replenished in two ways. Firstly, the reduction of high-valence Cu2+ and Co3+ to low-valence Cu+ and Co2+ released the oxygen species and formed oxygen vacancies, making the surface of the catalyst ready for the oxygen species adsorption from the gas phase (O2(gas)) and accelerating the conversion of the gas phase oxygen species to the adsorbed oxygen species (Reactions (8) and (9)) [7]. The relative content of adsorbed oxygen species of the used CuCo2O4-N catalysts (40.6%) was significantly decreased compared to the fresh CuCo2O4-N catalysts (46.8%), and the activity of soot conversion was also decreased. The adsorbed oxygen species played a crucial role in catalytic soot combustion, which could be transformed to active oxygen species such as O2, O 2 and O (Reactions (10)–(12)) [52]. Moreover, some lattice oxygen species took over the oxygen vacancies and were transferred into adsorbed oxygen species. The adsorbed oxygen species were spilled over to the soot particle surface and further contributed to the soot combustion reaction [35]. All these active oxygen species could directly participate in catalytic soot combustion reaction via the contact points between catalysts and soot particulates to form CO and CO2 [37]. In addition, the released active oxygen species for the conversion of NO oxidation of NO2 [42]. Moreover, the Cu+ and Co2+ species were reoxidized to Cu2+ and Co3+ due to the replenishment of oxygen vacancies and participation in catalytic soot combustion (Reaction (13)), and a portion of the gaseous NO molecules were also converted to NO2 (Reaction (4)) [53]. Hence, the facilitated electron transfer process and abundant adsorbed oxygen species endowed the copper–cobalt spinel catalysts with outstanding soot oxidation activity.
C ( Soot ) + O * CO x + O v
C u 2 + + C o 3 + C u + + C o + + O ads + O v
O 2 ( g a s ) + O v O a d s
O a d s + e O 2
O 2 + e O
O + e O 2
C u + + C o 2 + + O v C u 2 + + C o 3 + + 2 e

5. Conclusions

To obtain a fundamental understanding on the catalytic soot combustion performance of the CuCo2O4 catalysts prepared with different metal salt precursors, the relationships between the textural and redox properties of the catalysts and soot conversion were investigated.
(1)
The tetragonal spinel crystals of CuCo2O4 were formed for all catalysts, while the single-metal oxide CuO species was formed only on the CuCo2O4-N and CuCo2O4-A catalysts. The CuCo2O4-N catalysts exhibited a higher specific surface area and well-developed pore structure.
(2)
The type of metal precursors could also profoundly affect the redox properties on CuCo2O4 spinel catalysts. The higher relative content of Co3+ (38.4%), Cu2+ (51.7%) and Oads (46.8%) species were obtained over the CuCo2O4-N catalysts. Meanwhile, CuCo2O4-N catalysts (173 °C) showed a lower reduction temperature over CuCo2O4-A (184 °C) and CuCo2O4-C (247 °C). The highest amount of surface adsorbed oxygen species (41.2 μmol·g−1) was achieved over the CuCo2O4-N catalysts.
(3)
The highest soot conversion activity and CO2 selectivity were obtained over the CuCo2O4-N catalysts regardless of the soot combustion conditions. The effects of the contact mode (loose and tight) and NO addition on soot conversion were also investigated. A good correlation between soot conversion and the textural and redox properties of the catalysts were observed. The reaction mechanisms and pathways of the CuCo2O4 for catalytic soot combustion were also established.

Author Contributions

Conceptualization, X.Z., F.Z. and X.L.; Data curation, C.Z., X.Z. and G.C.; Formal analysis, C.Z., F.Z., X.L., G.C. and G.Y.; Funding acquisition, X.Z.; Investigation, F.Z.; Resources, G.Y.; Supervision, X.Z. and Z.Z.; Validation, Z.Z.; Writing—original draft, C.Z.; Writing—review and editing, X.Z. 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 No. 51976093).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful for Wentai Wang for the technical support of catalyst characterization and for the experimental platform provided by Ningbo University.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Choi, M.; Park, S. Optimization of multiple-stage fuel injection and optical analysis of the combustion process in a heavy-duty diesel engine. Fuel Process. Technol. 2022, 228, 107137. [Google Scholar] [CrossRef]
  2. Jeong, S.; Lee, J.H.; Ha, J.H.; Kim, J.; Kim, I.; Bae, S. An exploratory study of the relationships between diesel engine exhaust particle inhalation, pulmonary inflammation and anxious behavior. Int. J. Environ. Res. Public Health 2021, 18, 1166. [Google Scholar] [CrossRef]
  3. Jiaqiang, E.; Zheng, P.; Han, D.; Zhao, X.; Deng, Y. Effects analysis on soot combustion performance enhancement in a rotary diesel particulate filter unit during continuous microwave heating. Fuel 2020, 276, 118043. [Google Scholar] [CrossRef]
  4. Schobing, J.; Tschamber, V.; Brillard, A.; Leyssens, G. Biodiesel soot combustion: Analysis of the soot-catalyst contact in different experimental conditions. Fuel 2022, 312, 122854. [Google Scholar] [CrossRef]
  5. Wang, X.; Jin, B.; Feng, R.; Liu, W.; Weng, D.; Wu, X.; Liu, S. A robust core-shell silver soot oxidation catalyst driven by Co3O4: Effect of tandem oxygen delivery and Co3O4-CeO2 synergy. Appl. Catal. B Environ. 2019, 250, 132–142. [Google Scholar] [CrossRef]
  6. Shen, J.; Rao, C.; Fu, Z.; Feng, X.; Liu, J.; Fan, X.; Peng, H.; Xu, X.; Tan, C.; Wang, X. The influence on the structural and redox property of CuO by using different precursors and precipitants for catalytic soot combustion. Appl. Surf. Sci. 2018, 453, 204–213. [Google Scholar] [CrossRef]
  7. Li, K.; Li, T.; Dai, Y.; Quan, Y.; Zhao, J.; Ren, J. Highly active urchin-like MCo2O4 (M=Co, Cu, Ni or Zn) spinel for toluene catalytic combustion. Fuel 2022, 318, 123648. [Google Scholar] [CrossRef]
  8. Lin, C.S.; Liu, H.Y.; Guo, M.; Zhao, Y.H.; Su, X.; Zhang, P.Y.; Zhang, Y.F. Plasmon-induced broad spectrum photocatalytic overall water splitting: Through non-noble bimetal nanoparticles hybrid with reduced graphene oxide. Colloids Surf. A Physicochem. Eng. Asp. 2022, 646, 128962. [Google Scholar] [CrossRef]
  9. Fino, D.; Bensaid, S.; Piumetti, M.; Russo, N. A review on the catalytic combustion of soot in Diesel particulate filters for automotive applications: From powder catalysts to structured reactors. Appl. Catal. A Gen. 2016, 509, 75–96. [Google Scholar] [CrossRef]
  10. Zhang, S.; Zhu, X.; Zheng, C.; Hu, D.; Zhang, J.; Gao, X. Study on catalytic soot oxidation over spinel type ACo2O4 (A=Co, Ni, Cu, Zn) catalysts. Aerosol Air Qual. Res. 2017, 17, 2317–2327. [Google Scholar] [CrossRef]
  11. Jampaiah, D.; Velisoju, V.K.; Venkataswamy, P.; Coyle, V.E.; Nafady, A.; Reddy, B.M.; Bhargava, S.K. Nanowire morphology of mono- and bidoped α-MnO2 catalysts for remarkable enhancement in soot oxidation. ACS Appl. Mater. Interfaces 2017, 9, 32652–32666. [Google Scholar] [CrossRef]
  12. Zhang, F.; Zhu, X.; Wu, H.; Wu, X.; Zhou, Z.; Chen, G.; Yang, G. Activity and stability of Cu-based spinel-type complex oxides for diesel soot combustion. ChemistrySelect 2021, 6, 14019–14026. [Google Scholar] [CrossRef]
  13. Okal, J.; Zawadzki, M.; Kraszkiewicz, P.; Adamska, K. Ru/CeO2 catalysts for combustion of mixture of light hydrocarbons: Effect of preparation method and metal salt precursors. Appl. Catal. A Gen. 2018, 549, 161–169. [Google Scholar] [CrossRef]
  14. Yang, T.; Zhao, M.; Wang, X.; Ma, R.; Liu, Y.; He, Y.; Li, D. Influence of active metal precursors on the structure and catalytic behavior of Pd/Al2O3 catalysts for selective acetylene hydrogenation. Catal. Lett. 2021, 152, 227–238. [Google Scholar] [CrossRef]
  15. Wang, Y.; Tao, Z.; Wu, B.; Xu, J.; Huo, C.; Li, K.; Chen, H.; Yang, Y.; Li, Y. Effect of metal precursors on the performance of Pt/ZSM-22 catalysts for n-hexadecane hydroisomerization. J. Catal. 2015, 322, 1–13. [Google Scholar] [CrossRef]
  16. Yu, Z.; Hu, X.; Jia, P.; Zhang, Z.; Dong, D.; Hu, G.; Hu, S.; Wang, Y.; Xiang, J. Steam reforming of acetic acid over nickel-based catalysts: The intrinsic effects of nickel precursors on behaviors of nickel catalysts. Appl. Catal. B Environ. 2018, 237, 538–553. [Google Scholar] [CrossRef]
  17. Liu, H.; Zhuang, M.; Zhang, Z.; Dai, X.; Qian, L.; Yan, Z. Catalytic removal of soot particles over MnCo2O4 catalysts prepared by the auto-combustion method. Chem. Pap. 2018, 72, 1973–1979. [Google Scholar] [CrossRef]
  18. Jin, C.; Cui, Y.; Zhang, G.; Luo, W.; Liu, Y.; Sun, Y.; Tian, Z.; Zheng, W. Synthesis of copper-cobalt hybrid oxide microflowers as electrode material for supercapacitors. Chem. Eng. J. 2018, 343, 331–339. [Google Scholar] [CrossRef]
  19. Lu, J.; Zhong, J.; Ren, Q.; Li, J.; Song, L.; Mo, S.; Zhang, M.; Chen, P.; Fu, M.; Ye, D. Construction of Cu-Ce interface for boosting toluene oxidation: Study of Cu-Ce interaction and intermediates identified by in situ DRIFTS. Chin. Chem. Lett. 2021, 32, 3435–3439. [Google Scholar] [CrossRef]
  20. Wen, X.; Xu, L.; Chen, M.; Shi, Y.; Lv, C.; Cui, Y.; Wu, X.; Cheng, G.; Wu, C.-E.; Miao, Z.; et al. Exploring the influence of nickel precursors on constructing efficient Ni-based CO2 methanation catalysts assisted with in-situ technologies. Appl. Catal. B Environ. 2021, 297, 120486. [Google Scholar] [CrossRef]
  21. Kang, T.; Kim, J. Optimal cobalt-based catalyst containing high-ratio of oxygen vacancy synthesized from metal-organic-framework (MOF) for oxygen evolution reaction (OER) enhancement. Appl. Surf. Sci. 2021, 560, 150035. [Google Scholar] [CrossRef]
  22. Li, Y.; Han, W.; Wang, R.; Weng, L.-T.; Serrano-Lotina, A.; Bañares, M.A.; Wang, Q.; Yeung, K.L. Performance of an aliovalent-substituted CoCeOx catalyst from bimetallic MOF for VOC oxidation in air. Appl. Catal. B Environ. 2020, 275, 119121. [Google Scholar] [CrossRef]
  23. Xiao, S.Y.; Liu, Y.; Wu, X.F.; Gan, L.T.; Lin, H.Y.; Zheng, L.R.; Dai, S.; Liu, P.F.; Yang, H.G. A low-valent cobalt oxide co-catalyst to boost photocatalytic water oxidation via enhanced hole-capturing ability. J. Mater. Chem. A 2021, 9, 14786–14792. [Google Scholar] [CrossRef]
  24. Huang, C.; Yu, Y.; Tang, X.; Liu, Z.; Zhang, J.; Ye, C.; Ye, Y.; Zhang, R. Hydrogen generation by ammonia decomposition over Co/CeO2 catalyst: Influence of support morphologies. Appl. Surf. Sci. 2020, 532, 147335. [Google Scholar] [CrossRef]
  25. Andana, T.; Piumetti, M.; Bensaid, S.; Veyre, L.; Thieuleux, C.; Russo, N.; Fino, D.; Quadrelli, E.A.; Pirone, R. CuO nanoparticles supported by ceria for NOx-assisted soot oxidation: Insight into catalytic activity and sintering. Appl. Catal. B Environ. 2017, 216, 41–58. [Google Scholar] [CrossRef]
  26. Du, H.; Ma, X.; Jiang, M.; Yan, P.; Conrad Zhang, Z. Highly efficient Cu/SiO2 catalyst derived from ethanolamine modification for furfural hydrogenation. Appl. Catal. A Gen. 2020, 598, 117598. [Google Scholar] [CrossRef]
  27. Liang, H.; Jin, B.; Li, M.; Yuan, X.; Wan, J.; Liu, W.; Wu, X.; Liu, S. Highly reactive and thermally stable Ag/YSZ catalysts with macroporous fiber-like morphology for soot combustion. Appl. Catal. B Environ. 2021, 294, 120271. [Google Scholar] [CrossRef]
  28. Xu, Y.; Qu, Z.P.; Ren, Y.W.; Dong, C. Enhancement of toluene oxidation performance over Cu-Mn composite oxides by regulating oxygen vacancy. Appl. Surf. Sci. 2021, 560, 149983. [Google Scholar] [CrossRef]
  29. Wang, J.; Chernavskii, P.A.; Khodakov, A.Y.; Wang, Y. Structure and catalytic performance of alumina-supported copper-cobalt catalysts for carbon monoxide hydrogenation. J. Catal. 2012, 286, 51–61. [Google Scholar] [CrossRef]
  30. Jia, A.-P.; Hu, G.-S.; Meng, L.; Xie, Y.-L.; Lu, J.-Q.; Luo, M.-F. CO oxidation over CuO/Ce1-xCuxO2−δ and Ce1−xCuxO2−δ catalysts: Synergetic effects and kinetic study. J. Catal. 2012, 289, 199–209. [Google Scholar] [CrossRef]
  31. Song, L.; Xiong, J.; Cheng, H.; Lu, J.; Liu, P.; Fu, M.; Wu, J.; Chen, L.; Huang, H.; Ye, D. In-Situ characterizations to investigate the nature of Co3+ coordination environment to activate surface adsorbed oxygen for methane oxidation. Appl. Surf. Sci. 2021, 556, 149713. [Google Scholar] [CrossRef]
  32. Lu, S.; Wang, F.; Chen, C.; Huang, F.; Li, K. Catalytic oxidation of formaldehyde over CeO2-Co3O4 catalysts. J. Rare Earths 2017, 35, 867–874. [Google Scholar] [CrossRef]
  33. 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 monolithic cordierite for diesel soot combustion. Appl. Catal. B Environ. 2021, 286, 119932. [Google Scholar] [CrossRef]
  34. Cui, B.; Yan, S.; Xia, Y.; Li, K.; Li, S.; Wang, D.; Ye, Y.; Liu, Y.-Q. Cu Ce1-O2 nanoflakes with improved catalytic activity and thermal stability for diesel soot combustion. Appl. Catal. A Gen. 2019, 578, 20–29. [Google Scholar] [CrossRef]
  35. He, J.; Yao, P.; Qiu, J.; Zhang, H.; Jiao, Y.; Wang, J.; Chen, Y. Enhancement effect of oxygen mobility over Ce0.5Zr0.5O2 catalysts doped by multivalent metal oxides for soot combustion. Fuel 2021, 286, 119359. [Google Scholar] [CrossRef]
  36. Zhang, Y.; Zeng, Z.; Li, Y.; Hou, Y.; Hu, J.; Huang, Z. Effect of the A-site cation over spinel AMn2O4 (A=Cu2+, Ni2+, Zn2+) for toluene combustion: Enhancement of the synergy and the oxygen activation ability. Fuel 2021, 288, 119700. [Google Scholar] [CrossRef]
  37. Zhao, M.; Deng, J.; Liu, J.; Li, Y.; Liu, J.; Duan, Z.; Xiong, J.; Zhao, Z.; Wei, Y.; Song, W.; et al. Roles of surface-active oxygen species on 3DOM cobalt-based spinel catalysts MxCo3-xO4 (M=Zn and Ni) for NOx-assisted soot oxidation. ACS Catal. 2019, 9, 7548–7567. [Google Scholar] [CrossRef]
  38. Geng, Q.; Wang, H.; Wang, J.; Hong, J.; Sun, W.; Wu, Y.; Wang, Y. Boosting the capacity of aqueous Li-ion capacitors via pinpoint surgery in nanocoral-like covalent organic frameworks. Small Methods 2022, 6, 2200314. [Google Scholar] [CrossRef]
  39. Feng, F.; Zheng, Y.; Shen, X.; Zheng, Q.; Dai, S.; Zhang, X.; Huang, Y.; Liu, Z.; Yan, K. Characteristics of back corona discharge in a honeycomb catalyst and its application for treatment of volatile organic compounds. Environ. Sci. Technol. 2015, 49, 6831–6837. [Google Scholar] [CrossRef]
  40. Roberts, C.A.; Paidi, V.K.; Shepit, M.; Peck, T.C.; Stamm Masias, K.L.; van Lierop, J.; Reddy, G.K. Effect of Cu substitution on the structure and reactivity of CuxCo3-xO4 spinel catalysts for direct NOx decomposition. Catal. Today 2021, 360, 204–212. [Google Scholar] [CrossRef]
  41. 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]
  42. Yu, D.; Peng, C.; Yu, X.; Wang, L.; Li, K.; Zhao, Z.; Li, Z. Facile preparation of amorphous CenMnOx catalysts and their good catalytic performance for soot combustion. Fuel 2022, 307, 121803. [Google Scholar] [CrossRef]
  43. Zhang, W.; Descorme, C.; Valverde, J.L.; Giroir-Fendler, A. Cu-Co mixed oxide catalysts for the total oxidation of toluene and propane. Catal. Today 2022, 384–386, 238–245. [Google Scholar] [CrossRef]
  44. Ambrogio, M.; Saracco, G.; Specchia, V. Combining filtration and catalytic combustion in particulate traps for diesel exhaust treatment. Chem. Eng. Sci. 2001, 56, 1613–1621. [Google Scholar] [CrossRef]
  45. Cao, C.; Xing, L.; Yang, Y.; Tian, Y.; Ding, T.; Zhang, J.; Hu, T.; Zheng, L.; Li, X. The monolithic transition metal oxide crossed nanosheets used for diesel soot combustion under gravitational contact mode. Appl. Surf. Sci. 2017, 406, 245–253. [Google Scholar] [CrossRef]
  46. Yeste, M.P.; Cauqui, M.Á.; Giménez-Mañogil, J.; Martínez-Munuera, J.C.; Muñoz, M.Á.; García-García, A. Catalytic activity of Cu and Co supported on ceria-yttria-zirconia oxides for the diesel soot combustion reaction in the presence of NOx. Chem. Eng. J. 2020, 380, 122370. [Google Scholar] [CrossRef] [Green Version]
  47. 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]
  48. Zhang, H.; Zhou, C.; Galvez, M.E.; Da Costa, P.; Chen, Y. MnOx-CeO2 mixed oxides as the catalyst for NO-assisted soot oxidation: The key role of NO adsorption/desorption on catalytic activity. Appl. Surf. Sci. 2018, 462, 678–684. [Google Scholar] [CrossRef]
  49. Martínez-Munuera, J.C.; Zoccoli, M.; Giménez-Mañogil, J.; García-García, A. Lattice oxygen activity in ceria-praseodymia mixed oxides for soot oxidation in catalysed gasoline particle filters. Appl. Catal. B Environ. 2019, 245, 706–720. [Google Scholar] [CrossRef] [Green Version]
  50. Chen, L.; Li, T.; Zhang, J.; Wang, J.; Chen, P.; Fu, M.; Wu, J.; Ye, D. Chemisorbed superoxide species enhanced the high catalytic performance of Ag/Co3O4 nanocubes for soot oxidation. ACS Appl. Mater. Interfaces 2021, 13, 21436–21449. [Google Scholar] [CrossRef]
  51. Guo, X.; Li, J.; Zhou, R. Catalytic performance of manganese doped CuO-CeO2 catalysts for selective oxidation of CO in hydrogen-rich gas. Fuel 2016, 163, 56–64. [Google Scholar] [CrossRef]
  52. Lee, J.H.; Lee, B.J.; Lee, D.-W.; Choung, J.W.; Kim, C.H.; Lee, K.-Y. Synergistic effect of Cu on a Ag-loaded CeO2 catalyst for soot oxidation with improved generation of active oxygen species and reducibility. Fuel 2020, 275, 117930. [Google Scholar] [CrossRef]
  53. Yang, Q.; Gu, F.; Tang, Y.; Zhang, H.; Liu, Q.; Zhong, Z.; Su, F. A Co3O4-CeO2 functionalized SBA-15 monolith with a three-dimensional framework improves NOx-assisted soot combustion. RSC Adv. 2015, 5, 26815–26822. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the experimental setup.
Figure 1. Schematic diagram of the experimental setup.
Ijerph 19 14737 g001
Figure 2. XRD patterns of the CuCo2O4-x (x = N, A and C) catalysts.
Figure 2. XRD patterns of the CuCo2O4-x (x = N, A and C) catalysts.
Ijerph 19 14737 g002
Figure 3. SEM images of the CuCo2O4-N (a,b), CuCo2O4-A (c,d) and CuCo2O4-C (e,f) catalysts.
Figure 3. SEM images of the CuCo2O4-N (a,b), CuCo2O4-A (c,d) and CuCo2O4-C (e,f) catalysts.
Ijerph 19 14737 g003
Figure 4. XPS spectra of the CuCo2O4-x (x = N, A and C) catalysts: (a) Co 2p; (b) Cu 2p; (c) Cu LMM; (d) O 1s.
Figure 4. XPS spectra of the CuCo2O4-x (x = N, A and C) catalysts: (a) Co 2p; (b) Cu 2p; (c) Cu LMM; (d) O 1s.
Ijerph 19 14737 g004
Figure 5. H2-TPR profiles of the CuCo2O4-x (x = N, A and C) catalysts.
Figure 5. H2-TPR profiles of the CuCo2O4-x (x = N, A and C) catalysts.
Ijerph 19 14737 g005
Figure 6. O2-TPD profiles (a) and the enlarged O2-TPD profiles in the temperature range of 50 °C to 550 °C (b) of the CuCo2O4-x (x = N, A and C) catalysts.
Figure 6. O2-TPD profiles (a) and the enlarged O2-TPD profiles in the temperature range of 50 °C to 550 °C (b) of the CuCo2O4-x (x = N, A and C) catalysts.
Ijerph 19 14737 g006
Figure 7. Soot conversion over the CuCo2O4-x (x = N, A and C) catalysts under loose contact mode under the carrier gas of 10 vol.% O2 with balanced N2.
Figure 7. Soot conversion over the CuCo2O4-x (x = N, A and C) catalysts under loose contact mode under the carrier gas of 10 vol.% O2 with balanced N2.
Ijerph 19 14737 g007
Figure 8. Soot conversion of the CuCo2O4-x (x = N, A and C) catalysts under the loose contact mode with 1000 ppm NO.
Figure 8. Soot conversion of the CuCo2O4-x (x = N, A and C) catalysts under the loose contact mode with 1000 ppm NO.
Ijerph 19 14737 g008
Figure 9. Soot conversion over the CuCo2O4-N catalyst under the loose and tight contact modes.
Figure 9. Soot conversion over the CuCo2O4-N catalyst under the loose and tight contact modes.
Ijerph 19 14737 g009
Figure 10. Stability of CuCo2O4-N catalyst in cycle tests for soot combustion.
Figure 10. Stability of CuCo2O4-N catalyst in cycle tests for soot combustion.
Ijerph 19 14737 g010
Figure 11. XPS spectra of O 1 s of CuCo2O4-N catalyst before and after reaction.
Figure 11. XPS spectra of O 1 s of CuCo2O4-N catalyst before and after reaction.
Ijerph 19 14737 g011
Table 1. Textural properties of the CuCo2O4-x (x = N, A and C) catalysts.
Table 1. Textural properties of the CuCo2O4-x (x = N, A and C) catalysts.
CatalystsSBET
(m2·g−1)
Pore Volume
(mm3·g−1)
Average Pore Diameter
(nm)
Average Crystal Size (nm) *
CuCo2O4-N2.14.58.338.0
CuCo2O4-A2.04.28.638.3
CuCo2O4-C1.22.58.538.8
* Calculated by the Scherrer equation, based on the characteristic peak of CuCo2O4 (3 1 1) crystal face located at the 2θ of 36.8°.
Table 2. Redox properties of the CuCo2O4-x (x = N, A and C) catalysts.
Table 2. Redox properties of the CuCo2O4-x (x = N, A and C) catalysts.
CatalystsCo3+/Cototal
(%)
Cu2+/Cutotal (%)Oads/(Oads+ Olatt)
(%)
H2 Consumption (mmol·g−1)O2 Uptake (μmol·g−1)
CuCo2O4-N38.451.746.813.441.2
CuCo2O4-A35.148.237.213.627.1
CuCo2O4-C33.847.836.514.525.6
Table 3. Catalytic activity of the CuCo2O4-x (x = N, A and C) catalysts for soot combustion under loose contact mode in 10 vol.% O2 with balanced N2.
Table 3. Catalytic activity of the CuCo2O4-x (x = N, A and C) catalysts for soot combustion under loose contact mode in 10 vol.% O2 with balanced N2.
Catalysts10 vol.% O2/N2
T10 (°C)T50 (°C)T90 (°C)SCO2 (%)
CuCo2O4-N45152055898.8
CuCo2O4-A49653956995.0
CuCo2O4-C50455059481.8
No catalysts53058661558.7
Table 4. Catalytic activity of the CuCo2O4-x (x = N, A and C) catalysts for soot combustion under the loose contact mode with 1000 ppm NO.
Table 4. Catalytic activity of the CuCo2O4-x (x = N, A and C) catalysts for soot combustion under the loose contact mode with 1000 ppm NO.
CatalystsT10 (°C)T50 (°C)T90 (°C)SCO2 (%)
CuCo2O4-N349414 482 96.6
CuCo2O4-A381458 494 94.3
CuCo2O4-C392485 534 79.7
No catalysts496575 615 43.6
Table 5. Catalytic activity of the CuCo2O4-N catalysts for soot combustion under the loose and tight contact modes.
Table 5. Catalytic activity of the CuCo2O4-N catalysts for soot combustion under the loose and tight contact modes.
Contact ModeT10 (°C)T50 (°C)T90 (°C)SCO2 (%)
Loose contact45152055898.8
Loose contact + NO34941448296.6
Tight contact 385428499100
Tight contact + NO339394448100
Table 6. XPS parameters of CuCo2O4-N catalyst before and after reaction.
Table 6. XPS parameters of CuCo2O4-N catalyst before and after reaction.
CatalystsOads/(Oads + Olatt) (%)
CuCo2O4-N (fresh)46.8
CuCo2O4-N (used)40.6
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhou, C.; Zhu, X.; Zhang, F.; Li, X.; Chen, G.; Zhou, Z.; Yang, G. Soot Combustion over Cu–Co Spinel Catalysts: The Intrinsic Effects of Precursors on Catalytic Activity. Int. J. Environ. Res. Public Health 2022, 19, 14737. https://doi.org/10.3390/ijerph192214737

AMA Style

Zhou C, Zhu X, Zhang F, Li X, Chen G, Zhou Z, Yang G. Soot Combustion over Cu–Co Spinel Catalysts: The Intrinsic Effects of Precursors on Catalytic Activity. International Journal of Environmental Research and Public Health. 2022; 19(22):14737. https://doi.org/10.3390/ijerph192214737

Chicago/Turabian Style

Zhou, Chunlin, Xinbo Zhu, Fei Zhang, Xinbao Li, Geng Chen, Zijian Zhou, and Guohua Yang. 2022. "Soot Combustion over Cu–Co Spinel Catalysts: The Intrinsic Effects of Precursors on Catalytic Activity" International Journal of Environmental Research and Public Health 19, no. 22: 14737. https://doi.org/10.3390/ijerph192214737

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