Next Article in Journal
Silver-Doped Zeolitic Imidazolate Framework (Ag@ZIF-8): An Efficient Electrocatalyst for CO2 Conversion to Syngas
Next Article in Special Issue
Research Progress on Metal Oxides for the Selective Catalytic Reduction of NOx with Ammonia
Previous Article in Journal
Boosting Catalytic Combustion of Ethanol by Tuning Morphologies and Exposed Crystal Facets of α-Mn2O3
Previous Article in Special Issue
Cobalt-/pH-Modified V2O5-MoO3/TiO2 Catalyst with Enhanced Activity for the Low-Temperature Selective Catalytic Reduction Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Engineering Surface Properties of CuO/Ce0.6Zr0.4O2 Catalysts for Efficient Low-Temperature Toluene Oxidation

1
Faculty of Environment Science and Engineering, Kunming University of Science and Technology, Kunming 650500, China
2
Sino-Platinum Metals Chemical (Yunnan) Co., Ltd., Kunming 650500, China
*
Author to whom correspondence should be addressed.
Catalysts 2023, 13(5), 866; https://doi.org/10.3390/catal13050866
Submission received: 12 April 2023 / Revised: 28 April 2023 / Accepted: 4 May 2023 / Published: 10 May 2023

Abstract

:
The development of superior low-temperature catalytic performance and inexpensive catalysts for the removal of volatile organic compounds (VOCs) is crucial for their industrial application. Herein, CuO/Ce0.6Zr0.4O2 catalysts calcinated at different temperatures (Cu/CZ-X, X represented calcination temperature) were prepared and used to eliminate toluene. It can be found that Cu/CZ-550 presented the highest low-temperature catalytic activity, with the lowest temperature (220 °C) 50% conversion of toluene, the highest normalized reaction rate (3.1 × 10−5 mol·g−1·s−1 at 180 °C) and the lowest apparent activation energy value (86.3 ± 4.7 kJ·mol−1). Systematically, the surface properties analysis results showed that the optimum redox property, abundant oxygen vacancies, and plentiful surface Ce3+ species over Cu/CZ-550 were associated with the strong interaction between Cu and support could significantly favor the adsorption and activation of toluene, thus resulting in its superior catalytic performance.

1. Introduction

Volatile organic compounds (VOCs), as representative air pollutants, are mainly emitted from industrial processes and transportation [1,2], including toluene, benzene and alkanes, which significantly harm human health and the environment [3,4]. Catalytic combustion has received increasing attention due to its high efficiency and low energy consumption compared to other approaches used for VOC degradation [5,6,7]. However, the low-temperature activity and stability of catalysts do not satisfy the requirements for practical industrial applications and need to be further improved.
Active components are considered to be the key factor in determining catalytic performance. Noble catalysts (Pt, Au, Pd) exhibit high activity and stability, but wide utilization is significantly limited due to their high price and small reserves [8,9,10]. Recently, transition metal oxides (MnO [6,11,12], CuO [13,14,15], CoO [8,16,17]) have received considerable attention worldwide owning to their low price, good reducibility, and strong oxygen storage [18,19], but their low-temperature catalytic efficiency means that this issue is an open question [4,20]. Therefore, it is crucial to optimize transition metal oxides in terms of their superior catalytic activity at low temperatures and with high efficiency. Cu oxides in Cu-based catalysts are the most promising catalysts in terms of VOC combustion [13,14,21]. The mechanism of oxidation is determined by the oxidation state of Cu, which plays a vital part. The diffusion of lattice oxygen in CuO to the surface limits the rate of oxidation [3]. Additionally, the nature of the support plays a critical role in the catalytic activity of supported metal catalysts. The electronic state of the active metal can be modified or determined by the support through the interaction between the active components and the support [22,23], and, in some cases, the support is able to participate in the reaction, facilitating the kinetic process [24,25,26]. Cerium dioxide (CeO2) has abundant oxygen vacancies and outstanding redox properties as the catalyst support, which was closely related to the reversible Ce4+ ↔ Ce3+ redox reaction [27,28,29]. It has been shown that during the oxidation of CeO2, the lattice oxygen atoms are directly consumed in the reaction, forming oxygen vacancies [30]. The oxygen vacancies can adsorb and activate gaseous O2, forming adsorbed oxygen species that favor oxygen mobility [28,30,31]. However, the modification of pure CeO2 with other substances is essential on account of the thermal instability of CeO2 [32,33]. ZrO2, which has superior stability, in combination with CeO2, can form CexZr1-xO2 solid solutions that extremely improve the thermal stability [34,35,36,37]. Sara et al. found that the Zr4+ cations added to the lattice of CeO2 exhibited significant improvement in their oxygen storage capacity (OSC) and thermal stability and enhanced its catalytic activity at a low temperature [38]. In addition, Zr4+ replaced Ce3+ generates distortion due to the difference of Zr4+ and Ce3+ atomic sizes in CexZr1-xO2 solid solution, increasing oxygen vacancies and oxygen mobility [30].
In recent years, Cu-based catalysts have been considered to be the most promising catalysts for VOC combustion, and their surface properties are modulated to enhance the catalytic performance [15,39]. Song et al. synthesized a series of different molar ratio Cu-Ce binary oxides via the co-precipitation method and applied them to catalytic toluene oxidation [40]. They found that the catalysts that had a strong interaction between Cu and Ce exhibited excellent catalytic performance, with a toluene complete conversion temperature (T100) of 230 °C. In order to strengthen the synergistic interaction between CuO and CeO2, Li et al. synthesized a series of hollow-structured CuCeOx via a hydrothermal approach [17]. The review illustrated that the hollow structure induced the electron transfer between CuO and CeO2, which improved the adsorption of toluene and activation capacity with a T100 of 280 °C. Lately, Zeng et al. further modified and synthesized the CuO-CeO2 catalysts via a novel double redox method, which greatly enhanced the concentration of Ce3+ and oxygen vacancies and then promoted the activation capacity of toluene catalytic oxidation with a T100 of 250 °C [15]. Wang et al. reported that toluene catalytic oxidation performance was closely related to the synergistic interaction between CuO and Ce-ZrOx in recent years [41]. By comparing CuO/ZrO2, CeO2/ZrO2, and CuO-CeO2/ZrO2, their surface properties and the interaction between Cu-CeOx and ZrO2 made the catalysts exhibit good toluene oxidation performance, with a T100 of 350 °C. Additionally, an accepted catalytic activity over CuO/CexZr1-xO2, with a T100 lowered to 270 °C was reported, which mainly resulted from the interfacial interaction between CuO and CexZr1-xO2, the highly dispersed CuO, and the nature of the support. However, the mechanism of the enhancement was still not very clear and needs to be further investigated.
In the present work, the surface properties of CuO/Ce0.6Zr0.4O2 were engineered by varying the calcination temperature (450 °C, 550 °C, 650 °C, and 750 °C). The surface properties of the catalysts were investigated systematically using a series of measurements, such as Raman, TEM, XPS, H2-TPR, and the relationship with the catalytic performance was also explored.

2. Results and Discussion

2.1. Catalytic Activity and Kinetic Study

The catalytic performance of toluene oxidation is presented in Figure 1a. The temperature required for a 50% conversion of toluene (T50) is usually used to compare the low-temperature performance of catalysts. It could be clearly seen from Figure 1a that the T50 of Cu/CZ-450, Cu/CZ-550, Cu/CZ-650, and Cu/CZ-750 was 230 °C, 220 °C, 237 °C and 272 °C, respectively. Obviously, with the increasing calcination temperature from 450 °C to 750 °C, T50 initially decreased but gradually raised. Typically, Cu/CZ-550, with the lowest value of T50, presented the highest low-temperature activity among these samples. The comparison of the toluene catalytic activity on a series of catalysts in the summarized literature, is listed in Table 1. The range of gas hourly space velocity (GHSV) values was 15,000 h−1–75,000 h−1 and 20,000 mL·g−1·h−1–66,000 mL·g−1·h−1. The T50 range present in the mentioned literature in Table 2 was 200–300 °C, and the T50 value of Cu/CZ-550 was 220 °C, which was lower than most of the catalysts in the summarized literatures. Therefore, the catalytic performance over Cu/CZ-550 may be superior to most of those in the literature. The preparation methods used in terms of catalysts in the summarized literature were not the same. Typically, in terms of two different types of catalysts, CuCeOx was prepared via a co-precipitation approach, and Cu/CZ was synthesized using an impregnation method. It was obvious that the calcination temperature had a certain effect on the catalytic activity. The optimization of the calcination temperature to engineer the surface was essential for obtaining superior catalytic performance.
In order to further compare the catalytic performance, the normalized reaction rate was calculated, and it is listed in Figure 1b. It could be found that the reaction rate of toluene oxidation at 180 °C over Cu/CZ-550 was up to 3.1 × 10−5 mol·g−1·s−1, which was almost 1.7 times, 3.1 times, and 6.7 times higher than those of Cu/CZ-450 (1.8 × 10−5 mol·g−1·s−1), Cu/CZ-650 (9.8 × 10−6 mol·g−1·s−1) and Cu/CZ-750 (4.6 × 10−6 mol·g−1·s−1), respectively. A similar trend was observed at 220 °C and 240 °C. The results confirmed the superior catalytic activity of Cu/CZ-550.
To further explore the difference in catalytic activity, the kinetic analysis was conducted by plotting the Arrhenius curves. The toluene conversions were kept below 20% at a high GHSV of 80,000 mL·g−1·h−1 to minimize the effect of internal and external diffusion [42]. As shown in Figure 1c, the coefficients (R2) for all catalysts were at least 0.99. The apparent activation energy (Ea) was calculated from the slope of the Arrhenius plot [43]. A relatively lower Ea value was observed for Cu/CZ-550 (86.3 ± 4.7 kJ·mol−1) compared to Cu/CZ-450 (105.5 ± 6.0 kJ·mol−1), Cu/CZ-650 (112.3 ± 4.5 kJ·mol−1), and Cu/CZ-750 (146.5 ± 5.4 kJ·mol−1), respectively. Generally, the catalysts with a low value could activate the reactants more easily, which resulted in excellent catalytic performance over Cu/CZ-550, which is in line with the results of the catalytic activity test.
Table 1. A comparison on catalytic activity of literature data on catalytic oxidation of toluene.
Table 1. A comparison on catalytic activity of literature data on catalytic oxidation of toluene.
SamplesCalcination
Temperature, °C
Reaction MixtureT50,
°C
T90,
°C
Reference
Cu0.4Ce0.6Oy500500 ppm toluene
20 vol.% O2/N2
GHSV: 50,000 h−1
230245[15]
5 wt% CuO/CeO2400500 ppm toluene
20vol.% O2/N2
GHSV: 75,000 h−1
210235[19]
Cu1Ce3Oy5001000 ppm toluene
21 vol.% O2/N2
GHSV: 60,000 h−1
200210[40]
CuO/CeO2-ZrO2 (nCe/nCu mole ratio = 5 and nCe/nZr
mole ratio = 9)
4001500 ppm toluene
GHSV: 24,000 h−1
300350[44]
Cu0.15Ce0.85Oy550600 ppm toluene
20 vol.% O2/He
GHSV: 50,000 h−1
240250[45]
Cu6Ce4Ox
(Cu/Ce atomic ratio = 6:4)
5501000 ppm toluene
20 vol.% O2/N2
GHSV: 36,000 mL·g−1·h−1
240260[39]
UiO-66-CeCu
(nCe/nCu mole ratio = 3)
500 ppm toluene
21 vol.% O2/N2
GHSV: 30,000 mL·g−1·h−1
220[13]
6% CuO/CeO2-ZrO27001000 ppm toluene
30 vol.% O2/Ar
277310[41]
8 wt% CuO/Ce0.8Zr0.2Oy4004400 ppm toluene
GHSV: 33,000 mL·g−1
210230[46]
Cu1Ce3Oy
(nCe/nCu mole ratio = 3)
5001.0 vol.% toluene
GHSV: 66,000 mL·g−1·h−1
223225[47]
Cu0.13Ce0.87Oy6001000 ppm toluene
GHSV: 15,000 h−1
470490[48]
8 wt% CuO/Ce0.6Zr0.4Oy550600 ppm toluene
21 vol.% O2/N2
GHSV: 20,000 mL·g−1·h−1
220240In this study
Table 2. Physical parameters, the ratio of ID/IF2g, and the value of Ea over all catalysts.
Table 2. Physical parameters, the ratio of ID/IF2g, and the value of Ea over all catalysts.
SamplesSurface Area (m2·g−1)Pore Volume (cm3·g−1)Average Pore Size
(nm)
Average Ce0.6Zr0.4O2 Crystal Sizes a (nm) Average CuO Crystal Sizes b (nm) ID/IF2g cEa d (kJ·mol−1)
Cu/CZ-45057.60.158.44.9 - e0.50105.5 ± 6.0
Cu/CZ-55050.80.1610.45.3 - e0.6188.3 ± 4.7
Cu/CZ-65042.00.1812.85.7 - e0.55112.3 ± 4.5
Cu/CZ-7509.50.0613.810.1 22.3 0.30146.5 ± 5.4
a From Ce0.6Zr0.4O2 (111) facet of the XRD results. b From the CuO (111) facet of XRD results. c From the Raman results. d From the kinetic analysis results. e Represents not detected.

2.2. Textural Characteristics

The textural properties of these catalysts were measured using N2 adsorption–desorption measurements. As shown in Figure 2a, all the Cu/CZ-X displayed a typical type-IV adsorption and desorption curve and H3-type hysteresis loop [44,47]. This was attributed to the capillary condensation of mesoporous materials, indicating the presence of irregular porosity [2]. Figure 2b shows the distribution of the pore size over the samples. The pore sizes of Cu/CZ-450, Cu/CZ-550, and Cu/CZ-650 were mainly concentrated at 5, 6, and 7 nm, respectively. However, the pore size of Cu/CZ-750 is located mainly around 30 nm, indicating that its texture was significantly impacted by the calcination temperature. The specific surface area, pore volume, and average pore size ae listed in Table 2. It was noticeable that the specific surface area of the catalysts progressively decreased from 57.6 m2·g−1 to 9.5 m2·g−1 when the calcination temperature increased from 450 °C to 750 °C. It confirmed the pore collapse over Cu/CZ-750, which is in line with its extremely low pore volume (Table 2). As the surface area increased, more surface-active centers were exposed to reactants, usually promoting catalytic activity [49]. Therefore, the Cu/CZ-750 exhibited inferior activity, which was closely related to the small specific surface area.

2.3. Structural Characteristics

XRD measurements of Cu/CZ-X were implemented to confirm the crystal phase of the samples, and these are shown in Figure 3. For all samples (Figure 3a), the main diffraction peaks at 29.2°, 33.8°, 48.5°, and 57.6° can be attributed to the (111), (200), (220), and (311) planes of Ce0.6Zr0.4O2 solid solution (PDF#38-1439), respectively, which is in good agreement with the ratio of the experimental preparation. It was clearly seen that the diffraction peaks at 29.8°, 49.7°, and 59.2° were attributed to the (003), (104), and (113) planes of ZrO2 (PDF#37-0031), respectively, when the calcination temperature was 450 °C, 650 °C, 750 °C. That was due to the separation of the Ce0.6Zr0.4O2 phase to produce ZrO2. However, Cu/CZ-550 did not have contributions from ZrO2 probably because of a slight difference in synthesis. The diffraction peak centered at 35.6° and 38.7° could be assigned to CuO (PDF#48-1548). It appeared to suggest that the intensity of the peak towards CuO increased with increasing calcination temperature (Figure 3b). However, almost no peak corresponding to CuO was observed in terms of Cu/CZ-550, indicating its high dispersion of copper species on the surface and the presence of a strong interaction between these species and Ce0.6Zr0.4O2 [2]. Ce0.6Zr0.4O2 and CuO crystal sizes were estimated using Scherrer’s formula from the Ce0.6Zr0.4O2 (111) and CuO (111) facets of XRD, respectively (Table 2). The support crystal sizes of Cu/CZ-450, Cu/CZ-550, Cu/CZ-650, and Cu/CZ-750 were, respectively, 4.9, 5.3, 5.7, and 10.1 nm, which indicated that the agglomeration at the high calcination temperature led to larger sizes. Indeed, the CuO crystal size of Cu/CZ-750 was 22.3 nm as a result of agglomeration.
The structure of the catalysts was further investigated using Raman measurements. As displayed in Figure 4a, the two peaks at 462 and 598 cm−1 were observed over all of the samples [28], which were attributed to the vibration model of octahedral local symmetry from the CeO2 lattice (F2g) and the defect-induced mode (D mode), respectively [27]. A blue shift of the feature (472 to 460 cm−1) toward the F2g mode was seen as increasing the calcination temperature, indicating the different structures present in these samples (Figure 4b). The D mode was attributed to the formation of oxygen vacancies, which were derived from structural defects [28]. The oxygen vacancy concentrations on the samples were assessed by estimating the ratio of the intensities in the D band and F2g band (ID/IF2g) [50]. Notably, as presented in Table 2, the ratio of ID/IF2g was remarkably higher in Cu/CZ-550 (0.61) than in Cu/CZ-450 (0.50), Cu/CZ-650 (0.55), and Cu/CZ-750 (0.30), respectively. This illustrated the highest oxygen vacancy concentration in Cu/CZ-550 than in the other samples. The high oxygen vacancy concentration was conducive to the absorption and activation of gaseous O2 and promoted the diffusion of reactant molecules, which is in agreement with the catalytic activity.
The structures of all of the samples were investigated using TEM. As displayed at Figure 5, it was easily observed that all samples presented clear lattice fringe distances of 0.31 nm and 0.27 nm, corresponding to CeO2 (111) and CuO (11-1), respectively. In addition, Cu/CZ-750 had a lattice fringe distance of 0.19 nm which was attributed to ZrO2 (104). The mean grain sizes and the distributions of grain size over the Cu/CZ-X samples were calculated using the TEM images (Figure 6), and they are listed in Table 3. The average grain sizes (d) of Cu/CZ-450, Cu/CZ-550, Cu/CZ-650, and Cu/CZ-750 were 4.9 nm, 7.3 nm, 9.2 nm, and 10.8 nm, respectively. Noteworthily, the grain sizes increased with the increasing calcination temperature.

2.4. Chemical State

The surface composition and chemical state of the samples were examined using XPS measurements. In Figure 7a, the peak centered at 933.1–933.8 eV corresponded to Cu+, while the peak at 935.4–935.9 eV and the two satellite peaks belonged to Cu2+ species [51,52,53]. As displayed in Table 3, the binding energy (BE) of Cu2+ over Cu/CZ-550 (934.8 eV) was quite different from those in Cu/CZ-450 (935.2 eV), Cu/CZ-550 (935.23 eV), and Cu/CZ-750 (935.4 eV), respectively. The results indicated that the chemical environments of Cu2+ over these samples were different. The ratio of the surface Cu2+/(Cu2+ + Cu+) is listed in Table 3. It is worth noting that the ratio of Cu2+/(Cu2+ + Cu+) gradually increased when increasing the calcination temperature to 550 °C. Nevertheless, the values of Cu2+/(Cu2+ + Cu+) decreased rapidly from 0.58 to 0.35 when further increasing the calcination temperature from 550 to 750 °C. Thus, Cu/CZ-550 possessed the highest concentration of surface Cu2+ species. Generally, the Cu2+ component was considered to be the catalytically active site involved in toluene oxidation [40]. Thus, Cu/CZ-550 would show accepted catalytic activity, which is in line with the results of the performance tests.
For the O 1s XPS spectra (Figure 7b), the binding energy of 529.6 eV, 531.4–531.5 eV, and 532.9 eV corresponded to the surface lattice oxygen (Olatt), surface oxygen (Osur), and other weakly bound oxygen species (Oabs), such as carbonate (CO32−), adsorbed molecular water, and hydroxyl (OH), over the samples, respectively [54,55]. The concentration of surface oxygen species could be estimated, and the results are summarized in Table 3. The contents of these surface oxygen over all samples were almost the same. The peaks (Figure 7c) at 182.1 eV and 184.5 eV were indexed to Zr 3d3/2 and Zr 3d5/2, which were ascribed to Zr4+ [56,57,58]. For all of the samples, there were no obvious differences in terms of Zr binding energy.
As described in Figure 7d, Ce 3d XPS spectra 3d3/2 and 3d5/2 were marked as u and v, respectively. The peaks at 917.3 eV (u4), 908.5 eV (u3), 900.8 eV (u1), 898.3 eV (v4), 889.6 eV (v3), and 882.7 eV (v1) were assigned to surface Ce4+ species, and the peaks at 903.5 eV (u2) and 885.2 eV (v2) corresponded to surface Ce3+ species. A higher ratio of Ce3+/(Ce3+ + Ce4+) (0.22) was found over Cu/CZ-550 in comparison to those over Cu/CZ-450 (0.19), Cu/CZ-650 (0.18), and Cu/CZ-750 (0.17), respectively. Ce3+ species were normally associated with the formation of oxygen vacancies, which could favor oxygen mobility [27,28]. Thus, Cu/CZ- might have more oxygen vacancies, which is in good agreement with Raman spectrum results, showing excellent catalytic activity.

2.5. Reducibility

The reducibility of these samples was analyzed using H2-TPR measurements, and the profiles of all of the samples are depicted in Figure 8. The reduction peaks located at 50–275 °C could be divided into three peaks α, β, and γ. Peak α, β, and γ were ascribed to the reduction of well dispersed copper species [19,59], the reduction in copper incorporated into CeO2 [15,60], and the reduced interaction of CuO with the support [40,59], respectively. Compared with other catalysts, the reduction peak position of peak α and β in Cu/CZ-550 shifted to a lower reduction temperature. It revealed that Cu/CZ-550 had a superior redox property. In order to further compare the reducibility of these samples, the H2 consumptions were calculated from the peak area and calibrated using CuO as an internal standard. As listed in Table 4, the hydrogen consumption of Cu/CZ-550 was up to 120.6 μmol·gcat−1, which was higher than those of Cu/CZ-450 (103.7 μmol·gcat−1), Cu/CZ-650 (111.9 μmol·gcat−1), and Cu/CZ-750 (86.1 μmol·gcat−1). The result indicated the highest concentration of surface-reducible oxygen over Cu/CZ-550 and proved that Cu/CZ-550 had outstanding reducibility. This considerable low-temperature reducibility could contribute to enhancing catalytic performance [61]. Thus, Cu/CZ-550 presented appropriate catalytic efficiency.

3. Materials and Methods

3.1. Chemicals

Cerium nitrate hexahydrate (Ce(NO3)3·6H2O), zirconium nitrate pentahydrate (Zr(NO3)4·5H2O), and copper nitrate trihydrate (Cu(NO3)2·3H2O) were obtained from Sinopharm Chemical Reagent Co., Ltd. (Shanghai, China). Ammonia water (NH3·H2O) and ammonium carbonate ((NH4)2CO3) were purchased from Shanghai Macklin Biochemical Co., Ltd. (Shanghai, China) and Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China), respectively.

3.2. Preparations of Catalysts

3.2.1. Synthesis of Ce0.6Zr0.4O2 Supports

A ceria–zirconia support was prepared via the co-precipitation approach, and the molar ratio of Ce and Zr was 1.5. 7.351 g Ce(NO3)3·6H2O and 7.268 g Zr(NO3)4·5H2O, which were, respectively, dissolved into 50 mL of deionized water to form transparent solutions and were then mixed by stirring for 10 min at room temperature. The saturated (NH4)2CO3 (120 mL) solution was added dropwise to the liquid mixture while stirring, which was followed by adding the right amount of NH3·H2O solution steadily under constant stirring at room temperature until the pH of the solution reached about 9–10. Then, the resulting turbid solution was aged at 80 °C for 6 h, filtrated, and washed with deionized water until reaching pH = 7. The resulting solid was dried at 80 °C for 12 h to obtain powders and was transferred into a muffle furnace to calcinate at 550 °C for 3 h. The obtained solid was the Ce0.6Zr0.4O2 support.

3.2.2. Synthesis of CuO/Ce0.6Zr0.4O2 Catalysts

CuO/Ce0.6Zr0.4O2 were synthesized using an impregnation method. Concretely, excessive NH3·H2O solution was dropped into 0.486 g Cu(NO3)2·3H2O under stirring until Cu(NO3)2·3H2O was dissolved completely at room temperature. Subsequently, 1.840 g of the ceria–zirconia support was added to the resulting mixture and stirred vigorously. The product obtained was dried at 80 °C overnight and divided into four parts to calcinate at varied temperatures of 450 °C, 550 °C, 650 °C, and 750 °C for 3 hm with a heating rate of 2 °C·min−1 in air. The final obtained catalysts are named as follows: Cu/CZ-X, where X represents the calcination temperature. For example, the catalyst calcinated at 450 °C is named Cu/CZ-450. The normalized amount of CuO was 8%.

3.3. Characterization of Catalysts

Nitrogen adsorption–desorption isotherms were performed at liquid N2 temperature (-196 °C) on a Micromeritics TriStar II 3020 instrument and were applied to determine the specific surface areas (SSAs) using the Brunauer–Emmett–Teller (BET) method. Before the analysis, the samples were degassed at 300 °C for 4 h to clean the surfaces. The pore volume and pore size distribution were discovered through the desorption branch of the N2 adsorption isotherm based on the Barrett–Joyner–Halenda (BJH) method. Wide-angle X-ray diffraction (XRD) experiments were carried out using a Rigaku Ultima IV at 40 kV and 40 mA by means of Cu-Kα (λ = 0.15406 nm) radiation. The diffraction was recorded from 5° to 85°, with a rate of 8°/min. Raman spectroscopy was obtained via a Renishaw Via Reflex 2000 microscopic Raman spectrometer equipped with a Leica microscopy system and a microscope (50×) using laser excitation (λ = 532 nm). Additionally, the Raman spectra were calibrated using a silicon wafer before the experiment at 520.5 cm−1. Transmission electron microscopy (TEM) images were taken using the Tecnai F30 device. The specimens were pretreated by ultrasonically suspending the samples in ethanol, and, subsequently, the suspension droplets were cast on a carbon-coated copper grid. The corresponding copper species sizes and ceria–zirconia solution sizes of the samples were estimated using Nano Measurer software. X-ray photoelectron spectroscopy (XPS) analyses were collected on a ThermoFischer ESCALAB 250Xi spectrometer with an Al Kα radiation source (hv = 1486.6 eV). The C 1s photoemission line was used for the binding energy (BE) calibration based on the peak position at 284.8 eV. H2 temperature programmed reduction (H2-TPR) experiment was undertaken on a TP-5080-D fitted with TCD. Prior to the experiments, 50 mg Cu/CZ-X samples were pretreated under N2 flow (30 mL/min) to remove surface impurities at 400 °C for 30 min. After cooling to room temperature, the TPR test changed the gas flow to a H2/Ar stream (30 mL·min−1), which was introduced in the feed stream instead of the N2. The catalysts were measured continuously with a range from 50 °C to 850 °C at a heating rate of 8 °C/min. The H2 consumption amount was quantified through calibration with the standard sample of CuO (99.99%).

3.4. Catalytic Measurements

The activity tests of different catalysts were carried out on the fixed bed reaction (i.d. = 6 mm) and filled with 300 mg of samples on a sieve fraction of a 40–60 mesh. The feed-stock gas mixture was made up of 600 ppm toluene and 20% O2, balanced by N2 with a continuous gas flow maintained at 100 mL·min−1, corresponding to a GHSV of 20,000 mL·g−1·h−1. Before the reaction, the catalysts in the fixed bed reaction were activated by air, which was passed for 30 min. The catalytic activity was tested at a temperature range of 180–340 °C, with an interval of 20 °C, and each test temperature was held for at least 30 min. During the tests, the inlet and outlet toluene gas concentrations were detected using a gas chromatograph (Fuli 9790) equipped with a flame ionization detector (FID) in real time. Toluene conversion was calculated using Equation (1).
X toluene ( % ) = [ Toluene ] in [ Toluene ] out [ Toluene ] out × 100 %
where [Toluene]in and [Toluene]out represent the toluene concentration measured in the inlet gas and the outlet gas, respectively.
The reaction rates (r, mol·g−1·s−1) were obtained by following Equation (2) [28].
r   mol · g 1 · s 1 = C inlet ×   F m cat · ln ( 1 X toluene )
where Cinlet (ppm) refers to the toluene concentration in the inlet gas, F (mol·s−1) is the flow rate, mcat (g) is the mass of catalyst.

3.5. Kinetic Analysis

The Kinetic study was carried out at a high GHSV of 80,000 mL·g−1·h−1 and 75 mg samples. Additionally, the activation energy (Ea) of toluene oxidation was calculated according to the Arrhenius equation. The toluene concentration of the inlet gas was 600 ppm and was controlled to below 20%, where the effect of internal and external diffusion was negligible. The activation energy was determined according to the Arrhenius equation, as in Equation (3). [42]
r   μ mol · L 1 · s 1 = A   exp [ E a RT ] C
where r represents the reaction rate (μmol·L−1·s−1), A is the pre-exponential factor (s−1), R is the molar gas constant, and T refers to the reaction temperature (K). In this study, the temperature was between 488 K and 558 K.

4. Conclusions

A series of Cu/CZ--X (X = 450 °C, 550 °C, 650 °C, 750 °C) prepared using an impregnation method were used to explore the effect of calcination temperature on the toluene removal catalytic activity. Cu/CZ-550 exhibited excellent low-temperature catalytic activity, with a 50% toluene conversion temperature of 220 °C. In addition, Cu/CZ-550 presented a high normalized reaction rate (3.1 × 105 mol·g1·s1 at 180 °C) and a relatively low apparent activation energy value (86.3 ± 4.7 kJ·mol1). Systematically, the characterization results revealed that the optimum redox property, abundant oxygen vacancies, and a large number of surface Ce3+ species were positively correlated with its superior catalytic performance. On the contrary, low specific surface area caused by structure collapse and large grain sizes were responsible for the inferior catalytic activity of Cu/CZ-750. This work will potentially guide the design of highly efficient and outstanding low-temperature activity catalysts for VOC degradation.

Author Contributions

Methodology, M.W.; formal analysis, M.W. and D.Z.; investigation, M.Z., J.W. (Jingge Wang), D.Z., J.L. and J.W. (Junwei Wang); resources, Q.Z.; data curation, M.W. and D.Z.; writing—original draft preparation, M.W.; writing—review and editing, M.W., J.C. and Q.Z.; supervision, Q.Z., Y.Z., J.C. and P.N.; funding acquisition, J.C. and Q.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China, grant number 22206066 and the Yunnan Fundamental Research Projects, grant number 202201AU070084.

Data Availability Statement

The data that support the findings of this study are available on request from the corresponding author [J.C.], upon reasonable request.

Acknowledgments

The author thanks the National Natural Science Foundation of China for their support in the development of this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liotta, L.F. Catalytic oxidation of volatile organic compounds on supported noble metals. Appl. Catal. B Environ. 2010, 1003–1004, 403–412. [Google Scholar] [CrossRef]
  2. Wei, G.; Zhang, Q.; Zhang, D.; Wang, J.; Tang, T.; Wang, H.; Liu, X.; Song, Z.; Ning, P. The influence of annealing temperature on copper-manganese catalyst towards the catalytic combustion of toluene: The mechanism study. Appl. Surf. Sci. 2019, 497, 143777. [Google Scholar] [CrossRef]
  3. Kamal, M.S.; Razzak, S.A.; Hossain, M.M. Catalytic oxidation of volatile organic compounds (VOCs)—A review. Atmos. Environ. 2016, 140, 117–134. [Google Scholar] [CrossRef]
  4. Yang, C.; Miao, G.; Pi, Y.; Xia, Q.; Wu, J.; Li, Z.; Xiao, J. Abatement of various types of VOCs by adsorption/catalytic oxidation: A review. Chem. Eng. J. 2019, 370, 1128–1153. [Google Scholar] [CrossRef]
  5. Cen, B.H.; Tang, C.; Lu, J.Q.; Chen, J.; Luo, M.F. Different roles of MoO3 and Nb2O5 promotion in short-chain alkane combustion over Pt/ZrO2 catalysts. Chin. J. Catal. 2021, 4212, 2287–2295. [Google Scholar] [CrossRef]
  6. Liao, Y.; Zhang, X.; Peng, R.; Zhao, M.; Ye, D. Catalytic properties of manganese oxide polyhedra with hollow and solid morphologies in toluene removal. Appl. Surf. Sci. 2017, 405, 20–28. [Google Scholar] [CrossRef]
  7. Sun, Y.; Fan, J.; Cheng, H.; Mo, S.; Ke, Y.; Ren, Q.; Fu, M.; Chen, P.; Wu, J.; Ye, D. Investigation into the roles of different oxygen species in toluene oxidation over manganese-supported platinum catalysts. Mol. Catal. 2021, 507, 111569. [Google Scholar] [CrossRef]
  8. Li, Y.; Chen, T.; Zhao, S.; Wu, P.; Chong, Y.; Li, A.; Zhao, Y.; Chen, G.; Jin, X.; Qiu, Y.; et al. Engineering Cobalt Oxide with Coexisting Cobalt Defects and Oxygen Vacancies for Enhanced Catalytic Oxidation of Toluene. ACS Catal. 2022, 129, 4906–4917. [Google Scholar] [CrossRef]
  9. Zhou, C.; Zhang, H.; Zhang, Z.; Li, L. Improved reactivity for toluene oxidation on MnOx/CeO2-ZrO2 catalyst by the synthesis of cubic-tetragonal interfaces. Appl. Surf. Sci. 2021, 539, 148188. [Google Scholar] [CrossRef]
  10. Xu, X.; Li, L.; Yu, F.; Peng, H.; Fang, X.; Wang, X. Mesoporous high surface area NiO synthesized with soft templates: Remarkable for catalytic CH4 deep oxidation. Mol. Catal. 2017, 441, 81–91. [Google Scholar] [CrossRef]
  11. Ye, Z.; Giraudon, J.M.; Nuns, N.; Simon, P.; De Geyter, N.; Morent, R.; Lamonier, J.F. Influence of the preparation method on the activity of copper-manganese oxides for toluene total oxidation. Appl. Catal. B Environ. 2018, 223, 154–166. [Google Scholar] [CrossRef]
  12. Zhao, L.; Zhang, Z.; Li, Y.; Leng, X.; Zhang, T.; Yuan, F.; Niu, X.; Zhu, Y. Synthesis of CeaMnOx hollow microsphere with hierarchical structure and its excellent catalytic performance for toluene combustion. Appl. Catal. B Environ. 2019, 245, 502–512. [Google Scholar] [CrossRef]
  13. Hou, X.; Bian, Y.; Yang, L. One-step synthesis of transition metal modified UiO-66-Ce metal-organic framework: Catalytic oxidation of toluene and investigation of the mechanism. Microporous Mesoporous Mater. 2022, 345, 112214. [Google Scholar] [CrossRef]
  14. Kondratowicz, T.; Drozdek, M.; Rokicińska, A.; Natkański, P.; Michalik, M.; Kuśtrowski, P. Novel CuO-containing catalysts based on ZrO2 hollow spheres for total oxidation of toluene. Microporous Mesoporous Mater. 2019, 279, 446–455. [Google Scholar] [CrossRef]
  15. Zeng, Y.; Haw, K.G.; Wang, Z.; Wang, Y.; Zhang, S.; Hongmanorom, P.; Zhong, Q.; Kawi, S. Double redox process to synthesize CuO-CeO2 catalysts with strong Cu-Ce interaction for efficient toluene oxidation. J. Hazard. Mater. 2021, 404 Pt A, 124088. [Google Scholar] [CrossRef]
  16. Hu, F.; Peng, Y.; Chen, J.; Liu, S.; Song, H.; Li, J. Low content of CoOx supported on nanocrystalline CeO2 for toluene combustion: The importance of interfaces between active sites and supports. Appl. Catal. B Environ. 2019, 240, 329–336. [Google Scholar] [CrossRef]
  17. Li, G.; Zhang, C.; Wang, Z.; Huang, H.; Peng, H.; Li, X. Fabrication of mesoporous Co3O4 oxides by acid treatment and their catalytic performances for toluene oxidation. Appl. Catal. A Gen. 2018, 550, 67–76. [Google Scholar] [CrossRef]
  18. Yun, J.; Wu, L.; Hao, Q.; Teng, Z.; Gao, X.; Dou, B.; Bin, F. Non-equilibrium plasma enhanced oxygen vacancies of CuO/CeO2 nanorod catalysts for toluene oxidation. J. Environ. Chem. Eng. 2022, 103, 107847. [Google Scholar] [CrossRef]
  19. Zeng, Y.; Wang, Y.; Song, F.; Zhang, S.; Zhong, Q. The effect of CuO loading on different method prepared CeO2 catalyst for toluene oxidation. Sci. Total Environ. 2020, 712, 135635. [Google Scholar] [CrossRef]
  20. Mo, S.; Zhang, Q.; Li, J.; Sun, Y.; Ren, Q.; Zou, S.; Zhang, Q.; Lu, J.; Fu, M.; Mo, D.; et al. Highly efficient mesoporous MnO2 catalysts for the total toluene oxidation: Oxygen-Vacancy defect engineering and involved intermediates using in situ DRIFTS. Appl. Catal. B Environ. 2020, 264, 118464. [Google Scholar] [CrossRef]
  21. Li, Y.; Peng, H.; Xu, X.; Peng, Y.; Wang, X. Facile preparation of mesoporous Cu–Sn solid solutions as active catalysts for CO oxidation. RSC Adv. 2015, 533, 25755–25764. [Google Scholar] [CrossRef]
  22. Tang, H.; Wei, J.; Liu, F.; Qiao, B.; Pan, X.; Li, L.; Liu, J.; Wang, J.; Zhang, T. Strong Metal-Support Interactions between Gold Nanoparticles and Nonoxides. J. Am. Chem. Soc. 2016, 1381, 56–59. [Google Scholar] [CrossRef]
  23. Wang, Z.; Huang, Z.; Brosnahan, J.T.; Zhang, S.; Guo, Y.; Guo, Y.; Wang, L.; Wang, Y.; Zhan, W. Ru/CeO2 Catalyst with Optimized CeO2 Support Morphology and Surface Facets for Propane Combustion. Environ. Sci. Technol. 2019, 539, 5349–5358. [Google Scholar] [CrossRef] [PubMed]
  24. Ma, C.; Wang, D.; Xue, W.; Dou, B.; Wang, H.; Hao, Z. Investigation of formaldehyde oxidation over Co3O4-CeO2 and Au/Co3O4-CeO2 catalysts at room temperature: Effective removal and determination of reaction mechanism. Environ. Sci. Technol. 2011, 458, 3628–3634. [Google Scholar] [CrossRef]
  25. Wang, B.; Wu, X.; Ran, R.; Si, Z.; Weng, D. IR characterization of propane oxidation on Pt/CeO2–ZrO2: The reaction mechanism and the role of Pt. J. Mol. Catal. A Chem. 2012, 356, 100–105. [Google Scholar] [CrossRef]
  26. Wu, H.; Wang, L.; Shen, Z.; Zhao, J. Catalytic oxidation of toluene and p-xylene using gold supported on Co3O4 catalyst prepared by colloidal precipitation method. J. Mol. Catal. A Chem. 2011, 351, 188–195. [Google Scholar] [CrossRef]
  27. Su, Z.; Si, W.; Liu, H.; Xiong, S.; Chu, X.; Yang, W.; Peng, Y.; Chen, J.; Cao, X.; Li, J. Boosting the Catalytic Performance of CeO2 in Toluene Combustion via the Ce-Ce Homogeneous Interface. Environ. Sci. Technol. 2021, 5518, 12630–12639. [Google Scholar] [CrossRef]
  28. Su, Z.; Yang, W.; Wang, C.; Xiong, S.; Cao, X.; Peng, Y.; Si, W.; Weng, Y.; Xue, M.; Li, J. Roles of Oxygen Vacancies in the Bulk and Surface of CeO2 for Toluene Catalytic Combustion. Environ. Sci. Technol. 2020, 5419, 12684–12692. [Google Scholar] [CrossRef]
  29. Liu, C.; Xian, H.; Jiang, Z.; Wang, L.; Zhang, J.; Zheng, L.; Tan, Y.; Li, X. Insight into the improvement effect of the Ce doping into the SnO2 catalyst for the catalytic combustion of methane. Appl. Catal. B Environ. 2015, 176–177, 542–552. [Google Scholar] [CrossRef]
  30. Lou, B.; Shakoor, N.; Adeel, M.; Zhang, P.; Huang, L.; Zhao, Y.; Zhao, W.; Jiang, Y.; Rui, Y. Catalytic oxidation of volatile organic compounds by non-noble metal catalyst: Current advancement and future prospectives. J. Clean. Prod. 2022, 363, 132523. [Google Scholar] [CrossRef]
  31. Song, W.; Ji, J.; Guo, K.; Wang, X.; Wei, X.; Cai, Y.; Tan, W.; Li, L.; Sun, J.; Tang, C.; et al. Solid-phase impregnation promotes Ce doping in TiO2 for boosted denitration of CeO2/TiO2 catalysts. Chin. Chem. Lett. 2022, 332, 935–938. [Google Scholar] [CrossRef]
  32. Chen, X.; Chen, X.; Yu, E.; Cai, S.; Jia, H.; Chen, J.; Liang, P. In situ pyrolysis of Ce-MOF to prepare CeO2 catalyst with obviously improved catalytic performance for toluene combustion. Chem. Eng. J. 2018, 344, 469–479. [Google Scholar] [CrossRef]
  33. Yu, G.; Zhu, L.; Zhang, G.; Qin, G.; Fu, H.; Ji, F.; Zhao, J. Preparation and characterization of the continuous titanium-doped ZrO2 mesoporous fibers with large surface area. J. Porous Mater. 2013, 211, 105–112. [Google Scholar] [CrossRef]
  34. Shukla, M.K.; Balyan, Y.; Kumar, A.; Bhaskar, T.; Dhar, A. Catalytic oxidation of soot by CeO2–ZrO2 catalysts: Role of Zr. Mater. Chem. Phys. 2022, 286, 126161. [Google Scholar] [CrossRef]
  35. Wen, J.; Xie, Y.; Ma, Y.; Sun, H.; Wang, H.; Liu, M.; Zhang, Q.; Chen, J. Engineering of surface properties of Ni-CeZrAl catalysts for dry reforming of methane. Fuel 2022, 308, 122008. [Google Scholar] [CrossRef]
  36. Guo, F.; Li, J.; Zhang, Y.; Yang, X. Enhanced Stability and Catalytic Performance of Active Rh Sites on Al2O3 Via Atomic Layer Deposited ZrO2. J. Phys. Chem. Lett. 2022, 1338, 8825–8832. [Google Scholar] [CrossRef]
  37. Xu, H.; Sun, M.; Liu, S.; Li, Y.; Wang, J.; Chen, Y. Effect of the calcination temperature of cerium–zirconium mixed oxides on the structure and catalytic performance of WO3/CeZrO2 monolithic catalyst for selective catalytic reduction of NOx with NH3. RSC Adv. 2017, 739, 24177–24187. [Google Scholar] [CrossRef]
  38. Abdollahzadeh-Ghom, S.; Zamani, C.; Andreu, T.; Epifani, M.; Morante, J.R. Improvement of oxygen storage capacity using mesoporous ceria–zirconia solid solutions. Appl. Catal. B Environ. 2011, 108–109, 32–38. [Google Scholar] [CrossRef]
  39. Li, L.; Zhang, C.; Chen, F.; Xiang, Y.; Yan, J.; Chu, W. Facile fabrication of hollow structured Cu-Ce binary oxides and their catalytic properties for toluene combustion. Catal. Today 2021, 376, 239–246. [Google Scholar] [CrossRef]
  40. Song, B.; Li, C.; Du, X.; Li, S.; Zhang, Y.; Lyu, Y.; Zhou, Q. Superior performance of Cu-Ce binary oxides for toluene catalytic oxidation: Cu-Ce synergistic effect and reaction pathways. Fuel 2021, 306, 121654. [Google Scholar] [CrossRef]
  41. Wang, L.; Yin, G.; Yang, Y.; Zhang, X. Enhanced CO oxidation and toluene oxidation on CuCeZr catalysts derived from UiO-66 metal organic frameworks. React. Kinet. Mech. Catal. 2019, 1281, 193–204. [Google Scholar] [CrossRef]
  42. Liu, M.; Yang, X.; Tian, Z.; Wang, H.; Yin, L.; Chen, J.; Guan, Q.; Yang, H.; Zhang, Q. Insights into the role of strontium in catalytic combustion of toluene over La1-xSrxCoO3 perovskite catalysts. Phys. Chem. Chem. Phys. 2022, 246, 3686–3694. [Google Scholar] [CrossRef] [PubMed]
  43. Tathod, A.P.; Hayek, N.; Shpasser, D.; Simakov, D.S.A.; Gazit, O.M. Mediating interaction strength between nickel and zirconia using a mixed oxide nanosheets interlayer for methane dry reforming. Appl. Catal. B Environ. 2019, 249, 106–115. [Google Scholar] [CrossRef]
  44. Dou, B.; Yang, D.; Kang, T.; Xu, Y.; Hao, Q.; Bin, F.; Xu, X. Morphology effects of CeO2-ZrO2 on the catalytic performance of CuO/CeO2-ZrO2 for toluene oxidation. Carbon Resour. Convers. 2021, 4, 55–60. [Google Scholar] [CrossRef]
  45. Delimaris, D.; Ioannides, T. VOC oxidation over CuO–CeO2 catalysts prepared by a combustion method. Appl. Catal. B Environ. 2009, 891–892, 295–302. [Google Scholar] [CrossRef]
  46. Deng, Q.-F.; Ren, T.-Z.; Agula, B.; Liu, Y.; Yuan, Z.-Y. Mesoporous CexZr1−xO2 solid solutions supported CuO nanocatalysts for toluene total oxidation. J. Ind. Eng. Chem. 2014, 205, 3303–3312. [Google Scholar] [CrossRef]
  47. Zhou, G.; Lan, H.; Yang, X.; Du, Q.; Xie, H.; Fu, M. Effects of the structure of Ce@Cu catalysts on the catalytic combustion of toluene in air. Ceram. Int. 2013, 394, 3677–3683. [Google Scholar] [CrossRef]
  48. Hu, C. Catalytic combustion kinetics of acetone and toluene over Cu0.13Ce0.87Oy catalyst. Chem. Eng. J. 2011, 1683, 1185–1192. [Google Scholar] [CrossRef]
  49. Liu, Z.-G.; Chai, S.-H.; Binder, A.; Li, Y.-Y.; Ji, L.-T.; Dai, S. Influence of calcination temperature on the structure and catalytic performance of CuOx-CoOy-CeO2 ternary mixed oxide for CO oxidation. Appl. Catal. A Gen. 2013, 451, 282–288. [Google Scholar] [CrossRef]
  50. Mi, R.; Li, D.; Hu, Z.; Yang, R.T. Morphology Effects of CeO2 Nanomaterials on the Catalytic Combustion of Toluene: A Combined Kinetics and Diffuse Reflectance Infrared Fourier Transform Spectroscopy Study. ACS Catal. 2021, 1113, 7876–7889. [Google Scholar] [CrossRef]
  51. Du, Y.; Xiao, G.; Guo, Z.; Lin, B.; Fu, M.; Ye, D.; Hu, Y. A high-performance and stable Cu/Beta for adsorption-catalytic oxidation in-situ destruction of low concentration toluene. Sci. Total Environ. 2022, 833, 155288. [Google Scholar] [CrossRef] [PubMed]
  52. Leng, X.; Wang, W.; He, L.; Li, D.; Chen, J.; Luo, M. Novel insights into diethylamine catalytic combustion over CuO catalysts supported by SSZ-13: Undesirable product NOx as a crucial intermediate for N2 generation. Mol. Catal. 2021, 516, 111952. [Google Scholar] [CrossRef]
  53. Lin, Q.; Liu, S.; Xu, S.; Liu, J.; Xu, H.; Chen, Y.; Dan, Y. Fabricate surface structure-stabilized Cu/BEA with hydrothermal-resistant via si-deposition for NOx abatement. Mol. Catal. 2020, 495, 111153. [Google Scholar] [CrossRef]
  54. Xie, Y.; Chen, J.; Wu, X.; Wen, J.; Zhao, R.; Li, Z.; Tian, G.; Zhang, Q.; Ning, P.; Hao, J. Frustrated Lewis Pairs Boosting Low-Temperature CO2 Methanation Performance over Ni/CeO2 Nanocatalysts. ACS Catal. 2022, 1217, 10587–10602. [Google Scholar] [CrossRef]
  55. Zhang, J.; Qin, X.; Chu, X.; Chen, M.; Chen, X.; Chen, J.; He, H.; Zhang, C. Tuning Metal-Support Interaction of Pt-CeO2 Catalysts for Enhanced Oxidation Reactivity. Environ. Sci. Technol. 2021, 5524, 16687–16698. [Google Scholar] [CrossRef] [PubMed]
  56. Chai, G.; Zhang, W.; Liotta, L.F.; Li, M.; Guo, Y.; Giroir-Fendler, A. Total oxidation of propane over Co3O4-based catalysts: Elucidating the influence of Zr dopant. Appl. Catal. B Environ. 2021, 298, 120606. [Google Scholar] [CrossRef]
  57. Zhang, X.; Li, M.; Cui, X.; Niu, X.; Zhu, Y. Enhancing catalytic activity for toluene and acetone oxidation over ZraCo1-aOx catalysts by doping Zr to improve the oxygen activation capacity due to formation of Zr-O-Co bonds. Chem. Eng. J. 2023, 465, 142857. [Google Scholar] [CrossRef]
  58. Chen, G.; Hong, D.; Xia, H.; Sun, W.; Shao, S.; Gong, B.; Wang, S.; Wu, J.; Wang, X.; Dai, Q. Amorphous and homogeneously Zr-doped MnOx with enhanced acid and redox properties for catalytic oxidation of 1,2-Dichloroethane. Chem. Eng. J. 2022, 428, 131067. [Google Scholar] [CrossRef]
  59. Cao, J.-L.; Wang, Y.; Zhang, T.-Y.; Wu, S.-H.; Yuan, Z.-Y. Preparation, characterization and catalytic behavior of nanostructured mesoporous CuO/Ce0.8Zr0.2O2 catalysts for low-temperature CO oxidation. Appl. Catal. B Environ. 2008, 781–782, 120–128. [Google Scholar] [CrossRef]
  60. Qi, L.; Yu, Q.; Dai, Y.; Tang, C.; Liu, L.; Zhang, H.; Gao, F.; Dong, L.; Chen, Y. Influence of cerium precursors on the structure and reducibility of mesoporous CuO-CeO2 catalysts for CO oxidation. Appl. Catal. B Environ. 2012, 119–120, 308–320. [Google Scholar] [CrossRef]
  61. Yang, X.; Ma, X.; Yu, X.; Ge, M. Exploration of strong metal-support interaction in zirconia supported catalysts for toluene oxidation. Appl. Catal. B Environ. 2020, 263, 118355. [Google Scholar] [CrossRef]
Figure 1. (a) Catalytic Performance of Cu/CZ-X samples, (b) normalized reaction rates of Cu/CZ-X samples for toluene catalytic oxidation, and (c) kinetic analysis.
Figure 1. (a) Catalytic Performance of Cu/CZ-X samples, (b) normalized reaction rates of Cu/CZ-X samples for toluene catalytic oxidation, and (c) kinetic analysis.
Catalysts 13 00866 g001
Figure 2. (a) N2 adsorption–desorption isotherms and (b) BJH pore size distributions of Cu/CZ-X catalysts.
Figure 2. (a) N2 adsorption–desorption isotherms and (b) BJH pore size distributions of Cu/CZ-X catalysts.
Catalysts 13 00866 g002
Figure 3. (a) XRD pattern of Cu/CZ-X catalysts and (b) partially magnified profiles.
Figure 3. (a) XRD pattern of Cu/CZ-X catalysts and (b) partially magnified profiles.
Catalysts 13 00866 g003
Figure 4. (a) Raman spectra of Cu/CZ-X catalysts and (b) partially magnified profiles.
Figure 4. (a) Raman spectra of Cu/CZ-X catalysts and (b) partially magnified profiles.
Catalysts 13 00866 g004
Figure 5. TEM images of (a) Cu/CZ-450, (b) Cu/CZ-550, (c) Cu/CZ-650, and (d) Cu/CZ-750 catalysts.
Figure 5. TEM images of (a) Cu/CZ-450, (b) Cu/CZ-550, (c) Cu/CZ-650, and (d) Cu/CZ-750 catalysts.
Catalysts 13 00866 g005
Figure 6. The distribution of grain sizes over (a) Cu/CZ-450, (b) Cu/CZ-550, (c) Cu/CZ-650, and (d) Cu/CZ-750 catalysts.
Figure 6. The distribution of grain sizes over (a) Cu/CZ-450, (b) Cu/CZ-550, (c) Cu/CZ-650, and (d) Cu/CZ-750 catalysts.
Catalysts 13 00866 g006aCatalysts 13 00866 g006b
Figure 7. XPS spectra of the Cu/CZ-X catalysts: (a) Cu 2p, (b) O 1s, (c) Zr 3d, and (d) Ce 3d.
Figure 7. XPS spectra of the Cu/CZ-X catalysts: (a) Cu 2p, (b) O 1s, (c) Zr 3d, and (d) Ce 3d.
Catalysts 13 00866 g007aCatalysts 13 00866 g007b
Figure 8. H2-TPR profiles of Cu/CZ-X catalysts.
Figure 8. H2-TPR profiles of Cu/CZ-X catalysts.
Catalysts 13 00866 g008
Table 3. Surface information and grain sizes of Cu/CZ-X catalysts.
Table 3. Surface information and grain sizes of Cu/CZ-X catalysts.
SamplesCu2+ BE
(eV)
Cu2+/(Cu+ + Cu2+)Osur/(Osur + Olatt)Ce3+/(Ce3+ + Ce4+) aGrain Sizes b (nm)
Cu/CZ-450935.20.540.270.195.3
Cu/CZ-550934.80.580.280.225.9
Cu/CZ-650935.20.520.260.186.3
Cu/CZ-750935.40.350.270.1710.8
a Calculated from (Sv2 + Su2)/(Sv1 + Sv2 + Sv3 + Sv4 + Su1 + Su2 + Su3 + Su4).b Obtained from the TEM results.
Table 4. H2 Consumption Quantification Results and kinetic features of toluene conversion on the as-prepared catalysts.
Table 4. H2 Consumption Quantification Results and kinetic features of toluene conversion on the as-prepared catalysts.
samplesPeak α H2
Consumption a (μmol·gcat−1)
Peak β H2
Consumption a
(μmol·gcat−1)
Peak γ H2
Consumption a
(μmol·gcat−1)
Total H2
Consumption a (μmol·gcat−1)
Cu/CZ-45027.153.423.3103.7
Cu/CZ-55014.0100.66.0120.6
Cu/CZ-6509.560.442.0111.9
Cu/CZ-7505.665.516.086.1
a Calculated from the H2-TPR results.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, M.; Zhang, Q.; Zou, M.; Wang, J.; Zhu, D.; Liu, J.; Wang, J.; Zuo, Y.; Chen, J.; Ning, P. Engineering Surface Properties of CuO/Ce0.6Zr0.4O2 Catalysts for Efficient Low-Temperature Toluene Oxidation. Catalysts 2023, 13, 866. https://doi.org/10.3390/catal13050866

AMA Style

Wang M, Zhang Q, Zou M, Wang J, Zhu D, Liu J, Wang J, Zuo Y, Chen J, Ning P. Engineering Surface Properties of CuO/Ce0.6Zr0.4O2 Catalysts for Efficient Low-Temperature Toluene Oxidation. Catalysts. 2023; 13(5):866. https://doi.org/10.3390/catal13050866

Chicago/Turabian Style

Wang, Mingyue, Qiulin Zhang, Meilin Zou, Jingge Wang, Danrui Zhu, Jiaying Liu, Junwei Wang, Yang Zuo, Jianjun Chen, and Ping Ning. 2023. "Engineering Surface Properties of CuO/Ce0.6Zr0.4O2 Catalysts for Efficient Low-Temperature Toluene Oxidation" Catalysts 13, no. 5: 866. https://doi.org/10.3390/catal13050866

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