Next Article in Journal
Diffusion-Limited Processes in Hydrogels with Chosen Applications from Drug Delivery to Electronic Components
Next Article in Special Issue
Stable Loading of TiO2 Catalysts on the Surface of Metal Substrate for Enhanced Photocatalytic Toluene Oxidation
Previous Article in Journal
Optimum Reaction Conditions for the Synthesis of Selenized Ornithogalum caudatum Ait. (Liliaceae) Polysaccharides and Measurement of Their Antioxidant Activity In Vivo
Previous Article in Special Issue
Dark–Light Tandem Catalytic Oxidation of Formaldehyde over SrBi2Ta2O9 Nanosheets
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enabling High Activity Catalyst Co3O4@CeO2 for Propane Catalytic Oxidation via Inverse Loading

1
SINOPEC Research Institute of Safety Engineering Co., Ltd., 339th Songling Road, Qingdao 266071, China
2
School of Physical Science and Technology, Guangxi University, Nanning 530004, China
3
Department of Chemical Engineering, Tsinghua University, Beijing 100084, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(15), 5930; https://doi.org/10.3390/molecules28155930
Submission received: 30 May 2023 / Revised: 30 July 2023 / Accepted: 3 August 2023 / Published: 7 August 2023
(This article belongs to the Special Issue Functional Nanomaterials in Green Chemistry)

Abstract

:
Propane catalytic oxidation is an important industrial chemical process. However, poor activity is frequently observed for stable C–H bonds, especially for non-noble catalysts in low temperature. Herein, we reported a controlled synthesis of catalyst Co3O4@CeO2–IE via inverse loading and proposed a strategy of oxygen vacancy for its high catalytic oxidation activity, achieving better performance than traditional supported catalyst Co3O4/CeO2–IM, i.e., the T50 (temperature at 50% propane conversion) of 217 °C vs. 235 °C and T90 (temperature at 90% propane conversion) of 268 °C vs. 348 °C at the propane space velocity of 60,000 mL g−1 h−1. Further investigations indicate that there are more enriched oxygen vacancies in Co3O4@CeO2–IE due to the unique preparation method. This work provides an element doping strategy to effectively boost the propane catalytic oxidation performance as well as a bright outlook for efficient environmental catalysts.

Graphical Abstract

1. Introduction

Environmental pollution, especially atmospheric environment pollution, is becoming an increasingly serious problem. Volatile organic compounds (VOCs) emission gained widespread attention in very recent years [1,2,3]. The term VOCs means any compound of carbon, excluding carbon monoxide, carbon dioxide, carbonic acid, metallic carbides or carbonates, and ammonium carbonate, which has become a main pollutant in the whole world. VOCs come from manifold sources, such as the petrochemical industry, construction materials, the printing industry and the electronics industry. In addition to the direct health effects, such as mutagenesis, teratogenesis and carcinogenesis, VOCs form 2.5 micrometer particulate matter (PM2.5) and damage the ozone layer through a series of photochemical reactions, thus seriously affecting the air quality around our living environment [4,5,6].
Propane, representing typical VOCs, is widely released from the energy activities involving liquefied natural gas (LNG), compressed natural gas (CNG) and liquefied petroleum gas (LPG), as well as other important industrial processes [7,8]. The effective treatment process of propane is catalytic oxidation for it provides high performance at a relatively low reaction temperature. However, there is still a problem in the catalytic oxidation of propane at low temperatures due to the stable C–H bonds in propane. Generally, precious metal (e.g., Au, Pd, Pt, Ru) catalysts show high activity in the propane catalytic oxidation [9,10,11,12]. Although noble metal catalysts, such as Pd and Pt, are supposed to be highly active and stable, their high expense, sintering rates, volatility, and the possibility of being poisoned by water or sulfur compounds limit their wide practical application. With gradually increasing awareness of environmental protection and strict emission control regulations, highly efficient cleaning technologies of dilute propane are in great demand. More attention has been paid to transition metal oxides (such as manganese, cobalt, and copper oxides) in recent years [13,14,15,16,17,18]. Co3O4 and its composite oxides emerge to be competitive alternative catalysts for noble metal catalysts due to their advantages, such as varied Co valence (Co2+ and Co3+), unfilled d orbitals, redox property, affluent active oxygen, rich resources and low prices. And it is one of the most effective active species in the cleavage of C–C and C–H bonds [19,20,21]. In the past decades, numerous efforts have been taken to improve the catalytic activity of Co3O4 based materials, for example, by controlling the size, dispersity and shape of the nanoparticles, the exposed crystal faces and the defect concentration of the surface [22,23,24]. New active sites can be constructed by surface engineering on the defects of Co3O4 based materials to regulate catalytic activity.
Oxygen vacancies (OV), one type of common defects, were first proposed in the 1960s [25]. OV can capture oxygen from the surrounding atmosphere and convert gaseous oxygen molecules into more reactive oxygen species. The mobility and activity of lattice oxygen species can be improved via the transmission effect of bulk vacancies. Later, researchers discovered that OV can be used as reaction sites to change the structure of the material and the electronic and chemical properties of the surface [26,27,28]. CeO2 acts as an attractive support for the catalyst to generate OV when Ce4+ is reduced to Ce3+ that corresponds to Ce2O3 and CeO2. As the most abundant rare earth element, its content in the crust is 0.0046%. It is an acidic substance with plentiful Lewis acid sites and a few Brønsted acid sites. Acidic sites can improve the mobility of reactive oxygen species, enhance the adsorption of propane, promote the breaking of C–H bonds and thus contribute to catalytic oxidation of propane [29,30,31,32]. Moreover, OV in the support can improve the dispersion and stability of catalysts by anchoring the active metal. However, as far as we know, there are few reports about the simple method for OV of the Co3O4−based catalyst for the propane catalytic oxidation. Thus, it is very necessary to develop a concise route to construct the OV of the catalyst for high catalytic activity and low temperatures of propane catalytic oxidation.
In this study, we designed an inverse loading method, called ion−exchange method, and accordingly synthesized a catalyst Co3O4@CeO2–IE (Figure 1) possessing unique structure different from the traditional supported catalyst Co3O4/CeO2–IM synthesized by impregnation method. For traditional impregnation (IM) methods, an inorganic salt precursor, Co(NO)3∙6H2O is used as the starting precursor to deposit the active species onto the outer surface of CeO2. These loading methods normally render limited contact area/density and relatively weak interfacial interaction between the active species and the support, resulting in easy aggregation of active species, simultaneously sharp reduction in the interfacial area/density during the pretreatment or reaction process and further impact on the catalytic performance (Figure 1a). In contrast, for the IE reaction, the Co(OH)2 nanosheet precursor was added into the solution of support CeO2 precursor Ce3+ solution and the ion–exchange (IE) reaction took place driven by the difference of solubility product (Ksp), as shown in Figure 1b. Co2+ was replaced with Ce3+ to form Co(OH)2@Ce(OH)4 mixed metal hydroxide (MMH) structure. The Co3O4@CeO2–IE catalyst was obtained by the subsequent calcination of Co(OH)2@Ce(OH)4 MMH. Co3O4@CeO2–IE exhibits a T50 of 217 °C and T90 of 268 °C at propane space velocity of 60,000 mL⋅g−1⋅h−1, comparing to T50 of 235 °C and T90 of 348 °C for the Co3O4/CeO2–IM. The investigations indicate that the enriched oxygen vacancies resulting from unique synthetic method, i.e., ion–exchange method, increase the propane catalytic oxidation efficiency.

2. Results

2.1. Structure and Morphology of the Obtained Samples

Two sets of composite samples with different loading directions, i.e., the Co3O4@CeO2–IE and Co3O4/CeO2–IM, and similar chemical compositions were synthesized (Table S1 in the Supporting Information (SI)). The morphology and hierarchical structure of the composites were characterized by a high–resolution transmission electron microscope (HRTEM) and energy dispersive spectrometer (EDS) elemental mapping analyses. The transmission electron microscope (TEM) images of Co(OH)2 show ultrathin nanosheets (about 10 nm) with clear lattice spacing of 0.26 nm (Figure S2). And the typical HRTEM images (Figure 2a) of Co3O4@CeO2–IE shows the CeO2 and Co3O4 particles are uniformly distributed. As shown in Figure 2b, the typical HRTEM images clearly show the Co3O4 (311), (220) and CeO2 (111) with interlayer spacing of 0.243 nm, 0.288 nm and 0.310 nm, respectively. EDS elemental mapping further proves the homogeneous distribution, implying a sufficient contact between Ce and Co of the Co3O4@CeO2–IE catalyst. (Figure 2c). In comparison, Co3O4/CeO2–IM shows a heterogeneous distribution between Co and Ce, and the CeO2 is reunited, meaning that there is a weak interaction between Co3O4 and CeO2 (Figure 2d–f). These results suggest that Co3O4@CeO2–IE mainly exposes the CeO2 inserts into Co3O4 surface, while Co3O4/CeO2–IM mainly stands on the surface of Co3O4. And the particle size of Co3O4@CeO2–IE (10.8 nm) and Co3O4/CeO2–IM (41.8 nm) was calculated by Scherrer equation, indicating that the smaller particles could be fabricated by ion–exchange methods.
The phase composition and crystallographic structure of Co3O4@CeO2–IE samples and Co3O4/CeO2–IM were examined by powder X-ray diffraction (PXRD). Figure 3a shows PXRD patterns Co3O4@CeO2–IE and Co3O4/CeO2–IM, which can be indexed to the composite phases of Co3O4 (JCPDS:42–1467) and CeO2 (JCPDS:34–0394), further confirming the similarity in the effect of IE and IM methods on the preparation of the catalyst. However, there is a discrepancy in the intensity of the diffraction peaks. Co3O4/CeO2–IM has higher intensity of the diffraction peaks thanCo3O4@CeO2–IE, suggesting that the former has better crystallinity than the latter and the uniform distribution affects the crystallinity of Co3O4 and CeO2. Notably, the characteristic peaks of Co3O4@CeO2–IE shift to lower 2θ values due to the diffusion of Ce4+ and Co2+, Co3+ into the opposite crystal lattice at the interface of Co3O4 and CeO2 while smaller Co2+ (74.5 pm) and Co3+ (61 pm) than Ce4+ (97 pm) [33], thus forming a mixed phase. Co3O4/CeO2–IM has almost the same peak positions as Co3O4 and CeO2, indicating that interactions between the Co3O4 and CeO2 seldom occur at the interface of Co3O4/CeO2–IM, consistent with the results of HRTEM analysis. The analysis results of nitrogen adsorption isotherms of the catalysts demonstrate that Co3O4@CeO2–IE (94.7 m2 g−1) has a much higher BET surface area than Ni–CeO2–IM (24.3 m2 g−1), meaning that the IE method can expose more active sites for the following catalytic applications (Figure 3b).

2.2. Catalytic Oxidation Performance

The propane catalytic oxidation test was performed in the fixed bed reactor at WHSV of 60,000 mL⋅g−1⋅h−1. The oxidative reactivity was evaluated by T50 and T90. Figure 4a shows that the Co3O4@CeO2−IE has much higher activity than Co3O4/CeO2–IM. Co3O4@CeO2−IE as it participates in the propane catalytic oxidation reaction, with T50 of 217 °C and T90 of 268 °C, significantly superior to T50 of 235 °C and T90 of 348 °C for Co3O4/CeO2–IM catalyst. The activation energies (Ea) were evaluated according to the Arrhenius plots in Figure 4b, demonstrating a much lower Ea value of 63.7 kJ⋅mol−1 for Co3O4@CeO2−IE than Co3O4/CeO2−IM (89.5 kJ⋅mol−1). And Co3O4@CeO2–IE shows a superior activity toward the total oxidation of propane to Co3O4/CeO2–IM, and its reaction rate at 235 °C is 9.80 × 10−7 mol g−1 s−1, is almost 1.5 times that of Co3O4/CeO2–IM (6.91 × 10−7 mol g−1 s−1). These findings confirm that Co3O4@CeO2−IE exhibits higher catalytic activity (Table S2).

2.3. Surface Chemistry Analysis of the Catalysts

To gain deep insight into the origin of differences in catalytic behaviors between Co3O4@CeO2–IE and Co3O4/CeO2–IM, X-ray photoelectron spectroscopy (XPS) analysis was used to access the electronic state of the catalyst. The Ce 3d XPS spectra for the two catalysts were performed. And the spectra were deconvoluted into ten sub−peaks. The peaks denoted as U and V are individually attributed to Ce 3d5/2 and Ce3d3/2, respectively. Among those ten peaks, U0, UI, V0 and VI are assigned to Ce3+ (red line) and the rest of them are related to Ce4+ (blue line) [34]. As depicted in Figure 5a, the XPS results reveal that the Ce3+ fraction (49.6%) on the surface of Co3O4@CeO2–IE is nearly twice more than that of Co3O4/CeO2–IM (28.6%). It is proposed that there are more oxygen defects on the surface of Co3O4@CeO2−IE (Ce4+ + OL → Ce3+ + OV) [35,36]. The relative oxygen vacancy (Ov) proportion on the surface of Co3O4@CeO2–IE (40.7%) is much higher than that of Co3O4/CeO2–IM (25.0%). The peaks at 780.5 eV and 795.8 eV are attributed to the Co 2p1/2 and Co 2p3/2 core line of Co2+, respectively, whereas those at 779.2 eV and 794.2 eV are related to Co3+ [37]. The ratio of Co2+/Co3+ (2.64) in Co3O4@CeO2–IE is much higher than that (1.96) in Co3O4/CeO2−IM due to the fact that the oxygen defects of the catalyst can promote the reduction in metal ions (Figure 5c and Table S3).
The structures of oxygen vacancies were further investigated by electron paramagnetic resonance (EPR). Figure 6a,b show that the two catalysts have the g value of 2.004, which can be attributed to the unpaired electrons trapped in the OV in the Co3O4/CeO2 materials. Co3O4@CeO2−IE has higher intensity of the peaks than Co3O4/CeO2–IM. The intensity of the peaks, which is associated with the number/density of the Ov, implies that Co3O4@CeO2−IE have more Ov.
The surface defects were further examined by the Raman spectroscopy (Figure 6c). Co3O4/CeO2−IM exhibits the signals of 191 (F2g1), 474 (Eg), 516 (F2g2), 615 (F2g3) and 680 cm−1 (A1g), which correspond to pure Co3O4 [38], indicating that Ce seldom interacts with Co3O4 in Co3O4@CeO2−IE (in accordance with the results of HRTEM and XRD analysis). All of the bands should be mainly assigned to the vibration mode of Co3O4. The sharp band at 461 cm−1 (F2g band) of CeO2 can be assigned to the vibration mode of CeO2 fluorspar structure (Figure S3) [39,40]. The F2g band of Co3O4@CeO2−IE shows a red−shift of 16 cm−1 to 445 cm−1 with a sharp peak, which can be attributed to the Co–O–Ce bond induced by residual stress or lattice distortion in CeO2 structure, further suggesting a strong interaction between Co3O4 and CeO2 in Co3O4@CeO2−IE [41].
O2−TPD profiles were determined to recognize the oxygen species desorbed from the surface as a function of temperature. Figure 6d shows more obvious adsorption (225 °C–600 °C) and (725 °C–950 °C) of oxygen species in Co3O4@CeO2−IE than in Co3O4@CeO2−IM, indicating more OV. The quantitative analysis shows that Co3O4@CeO2−IE has nearly three times higher desorptive capacity of O (0.0838 mmol g−1) than Co3O4/CeO2−IM (0.0284 mmol g−1). The defects OV caused by Ce doping can strengthen the adsorption, activation ability of oxygen molecules and surface oxygen species migration ability, resulting in the generation of abundant active oxygen species at low temperatures.
The above analysis results indicate that the Ce3+ ions are partly exchanged with the Co2+ ions in Co(OH)2, leading to a special interface between Co(OH)2 and Ce(OH)4 with the diffusion of Ce4+ and Co2+ into opposite crystal phases and thus more Ov of the interface in Co3O4@CeO2−IE via inverse loading. There existed the lattice distortion due to the radius of Co2+ or Co3+ being smaller than Ce4+ and the charge imbalance for Co2+ or Co3+ and Ce4+ in Co3O4@CeO2−IE, that may induce the more active oxygen species of Co–O–Ce in the interface than in the pure CeO2 which induced the formation of more interfacial oxygen vacancies, while the Co3O4 and CeO2 in Co3O4/CeO2−IM exhibited poor interaction.

2.4. Catalytic Mechanism and Density−Functional Theory (DFT) Calculations

Marse−van Krevelen (MvK) mechanism is suitable for most non−metal catalysts to oxidize VOCs. MvK mechanism is based on redox reaction. Its essence is that the lattice oxygen in the catalyst oxidizes VOCs. In this reaction process, firstly, when VOCs react with lattice oxygen and generate oxygen vacancies, the metal oxides are reduced consequently. Secondly, oxygen vacancies are filled by oxygen in the air. The adsorption capacity and oxygen transfer capacity are the key contributing factors of the MvK mechanism. It is widely accepted that the dissociative adsorption of C3H8 on the catalyst surfaces triggers the catalytic oxidation process [42,43]. We calculated DFT in order to further elucidate the adsorption mechanism of Co3O4@CeO2−IE. To simulate the sample, a Co3O4−(311) surface model was created to expose the most crystal faces in the experiments. The Co3O4−(311) surface, which contains OV (Co–OV) adjacent to Co atoms, was simulated first, and then Co3O4@CeO2−IE was simulated by replacing Co atoms with Ce atoms in Co3O4. In the simulations, VO is adjacent to both Co and Ce atoms (Co–OV–Ce). The reaction is initiated by the adsorption of propane to the surface, followed by the dehydrogenation of propane to produce free radicals, which are finally oxidized to CO2 and H2O, where the desorption of generated propane radicals from the surface of the catalyst is the rate−limiting step for the entire reaction. Hence, the adsorption energies of propane adsorption are studied. Two substrate structures for adsorption are shown in Figure 7a,b. The calculated adsorption free energies (ΔEads) of propane on the two substrates are 0.17 eV and −0.85 eV, respectively. That indicates that propane is not favorable for adsorption on the Co–OV substrate, but is favorable for adsorption on the Co–OV–Ce surface. This is mainly due to the fact that Ce atoms have a greater number of outer electrons than Co atoms, and Ce doping in the system not only induces a large number of O vacancy defects, but also increases the free electrons of the system. The electronic density of Ce in the Co3O4@CeO2−IE system near the Fermi level can be analyzed according to the density of states in Figure 7c,d, conductive to electron exchange with propane. The density of states of the Co–OV–Ce structure suggests that the additional electrons provided by the Ce atoms are mainly populated above the Fermi energy level 0.3–0.7 eV above the Fermi energy level. In order to better investigate the adsorption property of propane by the extra electrons from Ce atoms, we calculated the crystal orbital Hamilton population (COHP) between Ce and H after adsorption of propane (Figure S4). We found that the extra electrons provided in the Ce atoms coupled with the H atoms favor the adsorption and dehydrogenation of propane [44]. As shown in the charge density difference in Figure 7e,f, the presence of Ce atoms leads to a greater exchange of electron density between the substrate and propane. When Ce fully invades the surface, Co3O4@CeO2−IE has more Co–OV–Ce structures than Co3O4/CeO2−IM [45], and is more conducive to the decomposition of propane.

3. Discussion

In summary, we successfully fabricated a new kind of Co3O4@CeO2–IE catalyst by ion−exchange method. The obtained Co3O4@CeO2–IE presents superior catalytic activity in the propane catalytic oxidation. The results of XPS, EPR, Raman spectra and DFT calculations etc., indicate that Co3O4@CeO2–IE catalyst possesses abundant oxygen vacancies on the surface of Co3O4–CeO2, which adsorb the propane. This study not only presents a new kind of non–noble metal catalyst for efficient catalytic oxidation of propane, but also highlights a strategy for the design of the oxygen vacancy for an advanced catalyst.

4. Materials and Methods

4.1. Catalyst Preparation

4.1.1. Preparation of Co(OH)2

All chemicals were purchased from Aladdin, China, and used as received.
Synthesis of Co(OH)2 nanosheets. According to our previously reported hybridization route [46], the synthetic process for the Co(OH)2 was as follows: Commercial MgCO3 was calcined at 750 °C for two hours to obtain MgO. Then, the resultant MgO was put into distilled water with a MgO−to−H2O mass ratio of 1:10 under stirring for 24 h. Finally, the white Mg(OH)2 product was separated by filtration, washed with deionized water and absolute ethanol three times and dried at room temperature. The resultant Mg(OH)2 (0.73 g equivalent to 0.0125 mol) was added to 25 mL of an aqueous solution containing 0.0125 mol of Co(NO3)2⋅6H2O. After stirring vigorously for two hours at room temperature, the green Co(OH)2 product was separated by filtration, washed with deionized water and ethyl alcohol three times and dried at room temperature overnight.

4.1.2. Preparation of Co3O4@CeO2–IE Catalyst

The Co3O4@CeO2–IE surrounded catalyst was synthesized by ion−change method. The fresh 1.86 g Co(OH)2 was added into the 40 mL solution including of 0.22 g Ce(NO3)3∙6H2O and then stirred under room temperature for 30 min. And thus put the mixture into Teflon−lined stainless steel autoclave, sealed, and maintained at 120 °C for 12 h. When cooled to the room temperature, the yellow–greenish products of Co(OH)2/Ce(OH)4 were separated by filtration, washed with deionized water and ethanol six times, and dried at 80 °C overnight. The hydroxide products were calcinated at 300 °C for 2 h in muffle furnace denoted as Co3O4@CeO2–IE. The Co loading is 45.3 wt % and the Ce loading is 8.36 wt % which was determined by ICP.

4.1.3. Preparation of Co3O4/CeO2–IM Catalyst

The Co3O4/CeO2−IM catalyst was synthesized by wet impregnation method. CeO2 support was prepared by reference method. 5.82 g Co(NO3)2 ∙6H2O was added into deionized 5 mL water solution and then added 0.09 g CeO2 into the mixture. The slurry was evaporated under stirring at 110 °C until dried thoroughly. And then followed the calcination denoted as Co3O4/CeO2−IM. The Co loading was 45.4 wt % and the Ce loading was 8.38 wt % which was determined by ICP and the similar content as Co3O4@CeO2–IE.

4.2. Characterizations

The morphology was characterized using a Hitachi S−4800 scanning electron microscope (SEM). The chemical composition of the solids was determined by an inductively coupled plasma−atomic core line spectroscopy (ICP–AES) (Thermo Fisher iCAP PRO (OES)). XRD patterns were performed on a Rigaku Miniflex 600 using Ni−filtered Cu Ka radiation (k = 0.15408 nm) at 40 Kv and 40 mA, and the scope of data collection was 2θ = 10–80°. The HRTEM images were obtained using a FEI Talos F200X G2 electron microscope operated at 200 kV. Results of element mapping were obtained on a super–x equipped with an Energy Dispersive Spectrometer (EDS). Brunauer–Emmett–Teller (BET) surface area measurement was conducted on a ASAP 2460 instrument. Before the BET surface area measurement, the samples were dried at 300 °C for 4 h under vacuum. Temperature−programmed desorption of O2 (O2−TPD) was performed on a Tianjin XQ TP−5080B chemisorption instrument with a thermal conductivity detector (TCD). The sample (100 mg) was first at 300 °C for 1 h to remove moisture under the steam (30 mL/min). After cooling to room temperature, 5 vol% O2/N2 mixture was switched on with a flow rate of 25 mL·min−1 at 50 °C for 1 h, and then cooled down to room temperature in the oxidizing atmosphere. Then continuously purging in a He flow for 30 min, the measurement started from room temperature to 950 °C at a heating rate 10 °C·min−1 after the baseline of single was stable. Electron paramagnetic resonance (EPR, Bruker A300, Bremen, Germany) were tested by a FA−200 (JES) electron paramagnetic resonance spectrometer. Raman spectra were carried out on Renishaw inVia Qontor with 532 nm of incident light. X−ray photoelectron spectra (XPS) were recorded on Escalab 250Xi using Al Kα radiation (1486.6 eV, 150 W) with binding energies (BEs) calibrated against the C1s peak of adventitious carbon at 284.8 eV.

4.3. Catalyst Evaluation

Propane oxidation test was carried out in a fixed–bed reactor using reactant of 0.5 vol% C3H8 and 21 vol% O2, balanced with N2. The Weight–Hourly–Space–Velocity (WHSV) was set at 60,000 mL·g−1·h−1 with reaction temperature increased from room temperature to 500 °C (5 °C/min). The products were in situ tested by a gas chromatograph (Agilent, GC–7890B, Santa Clara, CA, USA). C3H8 conversion was acquired from the following formulas:
C 3 H 8   conversion   ( % ) = C 3 H 8 i n C 3 H 8 o u t C 3 H 8 i n × 100 ( % )
where in, [C3H8]in and [C3H8]out represent the inlet and outlet C3H8 concentrations, separately.
To obtain the information of the apparent activation energy (Ea), the reaction was controlled at the kinetic regime (C3H8 conversion is under 10%). The equation below was used for the calculation of Ea by acquiring the slope:
ln k = E a R T + C
where k represents the reaction rate constant, T is the absolute temperature and R stands for the gas constant.

4.4. Computational Details

First−principles calculations were carried out using the projector−augmented wave method implemented in the Quantum ESPRESSO based on density functional theory (DFT) [47,48]. The Perdew–Burke–Ernzerhof (PBE) functional of the generalized gradient approximation (GGA) was adopted for electron exchange and correlation interaction [49]. The van der Waals interactions between layers were corrected using the DFT−D3 functional [50]. The ions relaxation was achieved until the force for per atom was less than 0.02 eV/Å and the total energy converged to 10−5 eV. A vacuum spacing of 15 Å was used to prevent interaction between adjacent slabs. The change in free energy for the adsorption (ΔEads) of the target by the catalyst substrate is defined by Eq: ΔEads = Etotal – Esubstrate – Etarget, where Etotal is the total energy of the catalyst substrate and propane, Esubstrate and Etarget are the energies of the substrate and free target molecule, respectively.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28155930/s1, Figure S1: PXRD patterns of Co(OH)2; Figure S2: Typical (a) TEM and (b) HRTEM images of Co(OH)2 nanosheets; Figure S3: Raman spectrum of CeO2; Figure S4: The crystal orbital Hamilton population between Ce and H atoms in the Co3O4−OV−Ce adsorbed propane system. Table S1: Element compositions of the two catalysts analyzed by ICP technique; Table S2: The ration of Co, Ce and O ions from XPS; Table S3: Catalytic performances of Co3O4@CeO2−IE and Co3O4/CeO2−IM catalysts.

Author Contributions

Conceptualization: X.W., W.L. and N.S.; methodology: X.W. and W.L.; software: C.L.; validation: X.W., W.L., T.Z. and J.Z.; formal analysis: X.W.; investigation: X.W. and W.L.; resources: B.S. and W.X.; data curation: X.W. and C.L.; writing—original drat preparation: X.W.; writing—review and editing: Z.S.; visualization: X.W.; supervision: W.X., N.S. and Z.S.; project administration: J.J., B.S. and W.X. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by SINOPEC Research Institute of Safety Engineering Co., Ltd.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Mellouki, A.; Wallington, T.J.; Chen, J. Atmospheric chemistry of oxygenated volatile organic compounds: Impacts on air quality and climate. Chem. Rev. 2015, 115, 3984–4014. [Google Scholar] [CrossRef]
  2. Ravi Kumar, Y.; Deshmukh, K.; Kovářík, T.; Khadheer Pasha, S.K. A systematic review on 2D materials for volatile organic compound sensing. Coordin. Chem. Rev. 2022, 461, 214502. [Google Scholar] [CrossRef]
  3. He, C.; Cheng, J.; Zhang, X.; Douthwaite, M.; Pattisson, S.; Hao, Z. Recent Advances in the Catalytic Oxidation of Volatile Organic Compounds: A Review Based on Pollutant Sorts and Sources. Chem. Rev. 2019, 119, 4471–4568. [Google Scholar] [CrossRef]
  4. Li, P.; Kim, S.; Jin, J.; Do, H.C.; Park, J.H. Efficient photodegradation of volatile organic compounds by iron-based metal-organic frameworks with high adsorption capacity. Appl. Catal. B Environ. 2020, 263, 118284. [Google Scholar] [CrossRef]
  5. Yin, F.; Yue, W.; Li, Y.; Gao, S.; Zhang, C.; Kan, H.; Niu, H.; Wang, W.; Guo, Y. Carbon–based nanomaterials for the detection of volatile organic compounds: A review. Carbon 2021, 180, 274–297. [Google Scholar] [CrossRef]
  6. Khatib, M.; Haick, H. Sensors for Volatile Organic Compounds. ACS Nano 2022, 16, 7080–7115. [Google Scholar] [CrossRef]
  7. Nadjafi, M.; Abdala, P.M.; Verel, R.; Hosseini, D.; Safonova, O.V.; Fedorov, A.; Müller, C.R. Reducibility and Dispersion Influence the Activity in Silica-Supported Vanadium-Based Catalysts for the Oxidative Dehydrogenation of Propane: The Case of Sodium Decavanadate. ACS Catal. 2020, 10, 2314–2321. [Google Scholar] [CrossRef]
  8. Zhang, M.; Sui, X.; Zhang, X.; Niu, M.; Li, C.; Wan, H.; Qiao, Z.-A.; Xie, H.; Li, X. Multi-defects engineering of NiCo2O4 for catalytic propane oxidation. Appl. Surf. Sci. 2022, 600, 154040. [Google Scholar] [CrossRef]
  9. Liotta, L.F. Catalytic oxidation of volatile organic compounds on supported noble metals. Appl. Catal. B Environ. 2010, 100, 403–412. [Google Scholar] [CrossRef]
  10. Scirè, S.; Liotta, L.F. Supported gold catalysts for the total oxidation of volatile organic compounds. Appl. Catal. B Environ. 2012, 125, 222–246. [Google Scholar] [CrossRef]
  11. 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 2012, 356, 100–105. [Google Scholar] [CrossRef]
  12. Chen, J.; Lv, X.; Xu, W.; Li, X.; Chen, J.; Jia, H. Utilizing Cl coordination to facilitate Ru–Ag self–assembling into alloy and recover thermally-inactivated catalyst for propane combustion. Appl. Catal. B Environ. 2021, 290, 119989. [Google Scholar] [CrossRef]
  13. Chen, K.; Li, W.; Zhu, C.; Yuan, L. Construction of Hollow Cobalt Tetroxide Nanocages through the Metal Salt Bifunctional Etching Strategy for Catalytic Oxidation of Propane at Ultrahigh Space Velocity. ACS Appl. Nano Mater. 2022, 5, 6575–6584. [Google Scholar] [CrossRef]
  14. Li, W.; Wang, J.; Gong, H. Catalytic Combustion of VOCs on Non-noble Metal Catalysts. Catal. Today 2009, 148, 81–87. [Google Scholar] [CrossRef]
  15. Feng, C.; Jiang, F.; Xiong, G.; Chen, C.; Wang, Z.; Pan, Y.; Fei, Z.; Lu, Y.; Li, X.; Zhang, R.; et al. Revelation of Mn4+–Osur–Mn3+ active site and combined Langmuir-Hinshelwood mechanism in propane total oxidation at low temperature over MnO2. Chem. Eng. J. 2023, 451, 138868. [Google Scholar] [CrossRef]
  16. Sun, L.; Liang, X.; Liu, H.; Cao, H.; Liu, X.; Jin, Y.; Li, X.; Chen, S.; Wu, X. Activation of Co–O bond in (110) facet exposed Co3O4 by Cu doping for the boost of propane catalytic oxidation. J. Hazard Mater. 2023, 452, 131319. [Google Scholar] [CrossRef]
  17. Chen, K.; Li, W.; Zhou, Z.; Huang, Q.; Liu, Y.; Duan, Q. Hydroxyl groups attached to Co2+ on the surface of Co3O4: A promising structure for propane catalytic oxidation. Catal. Sci. Technol. 2020, 10, 2573–2582. [Google Scholar] [CrossRef]
  18. Wu, S.; Liu, H.; Huang, Z.; Xu, H.; Shen, W. O-vacancy-rich porous MnO2 nanosheets as highly efficient catalysts for propane catalytic oxidation. Appl. Catal. B Environ. 2022, 312, 121387. [Google Scholar] [CrossRef]
  19. Xiao, Z.; Huang, Y.C.; Dong, C.L.; Xie, C.; Liu, Z.; Du, S.; Chen, W.; Yan, D.; Tao, L.; Shu, Z.; et al. Operando Identification of the Dynamic Behavior of Oxygen Vacancy-Rich Co3O4 for Oxygen Evolution Reaction. J. Am. Chem. Soc. 2020, 142, 12087–12095. [Google Scholar] [CrossRef]
  20. Li, Z.; Zhang, Y.; Feng, Y.; Cheng, C.Q.; Qiu, W.K.; Dong, C.K.; Liu, H.; Du, X.W. Co3O4 Nanoparticles with Ultrasmall Size and Abundant Oxygen Vacancies for Boosting Oxygen Involved Reactions. Adv. Fun. Mater. 2019, 29, 1903444. [Google Scholar] [CrossRef]
  21. Hu, J.; Zeng, X.; Wang, G.; Qian, B.; Liu, Y.; Hu, X.; He, B.; Zhang, L.; Zhang, X. Modulating mesoporous Co3O4 hollow nanospheres with oxygen vacancies for highly efficient peroxymonosulfate activation. Chem. Eng. J. 2020, 400, 125869. [Google Scholar] [CrossRef]
  22. Iablokov, V.; Barbosa, R.; Pollefeyt, G.; Van Driessche, I.; Chenakin, S.; Kruse, N. Catalytic CO Oxidation over Well-Defined Cobalt Oxide Nanoparticles: Size-Reactivity Correlation. ACS Catal. 2015, 5, 5714–5718. [Google Scholar] [CrossRef]
  23. Sun, H.; Ang, H.M.; Tadé, M.O.; Wang, S. Co3O4 nanocrystals with predominantly exposed facets: Synthesis, environmental and energy applications. J. Mater. Chem. A 2013, 1, 14427–14442. [Google Scholar] [CrossRef] [Green Version]
  24. Xie, X.; Li, Y.; Liu, Z.Q.; Haruta, M.; Shen, W.-J. Low-temperature oxidation of CO catalysed by Co3O4 nanorods. Nature 2009, 458, 746–749. [Google Scholar] [CrossRef] [PubMed]
  25. Tompkins, F.C. Superficial Chemistry and Solid Imperfections. Nature 1960, 186, 3–6. [Google Scholar] [CrossRef]
  26. Li, H.; Xiao, Z.; Liu, P.; Wang, H.; Geng, J.; Lei, H.; Zhuo, O. Interfaces and Oxygen Vacancies-Enriched Catalysts Derived from Cu-Mn-Al Hydrotalcite towards High-Efficient Water–Gas Shift Reaction. Molecules 2023, 28, 1522. [Google Scholar] [CrossRef] [PubMed]
  27. Jiang, W.; Loh, H.; Low, B.Q.L.; Zhu, H.; Low, J.; Heng, J.Z.X.; Tang, K.Y.; Li, Z.; Loh, X.J.; Ye, E.; et al. Role of oxygen vacancy in metal oxides for photocatalytic CO2 reduction. Appl. Catal. B Environ. 2023, 321, 122079. [Google Scholar] [CrossRef]
  28. Zhu, K.; Shi, F.; Zhu, X.; Yang, W. The roles of oxygen vacancies in electrocatalytic oxygen evolution reaction. Nano Energy 2020, 73, 104761. [Google Scholar] [CrossRef]
  29. Zhang, G.; Zhou, Y.; Yang, Y.; Kong, T.; Song, Y.; Zhang, S.; Zheng, H. Elucidating the Role of Surface Ce4+ and Oxygen Vacancies of CeO2 in the Direct Synthesis of Dimethyl Carbonate from CO2 and Methanol. Molecules 2023, 28, 3785. [Google Scholar] [CrossRef]
  30. 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, 54, 12684–12692. [Google Scholar] [CrossRef]
  31. Yang, M.; Shen, G.; Wang, Q.; Deng, K.; Liu, M.; Chen, Y.; Gong, Y.; Wang, Z. Roles of Oxygen Vacancies of CeO2 and Mn-Doped CeO2 with the Same Morphology in Benzene Catalytic Oxidation. Molecules 2021, 26, 6363. [Google Scholar] [CrossRef]
  32. 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]
  33. Singha, R.K.; Shukla, A.; Yadav, A.; Sivakumar Konathala, L.N.; Bal, R. Effect of Metal-Support Interaction on Activity and Stability of Ni-CeO2 Catalyst for Partial Oxidation of Methane. Appl. Catal. B Environ. 2017, 202, 473–488. [Google Scholar] [CrossRef]
  34. Wei, X.; Wang, X.; Pu, Y.; Liu, A.; Chen, C.; Zou, W.; Zheng, Y.; Huang, J.; Zhang, Y.; Yang, Y.; et al. Facile Ball-Milling Synthesis of CeO2/g-C3N4 Z-Scheme Heterojunction for Synergistic Adsorption and Photodegradation of Methylene Blue: Characteristics, Kinetics, Models, and Mechanisms. Chem. Eng. J. 2021, 420, 127719. [Google Scholar] [CrossRef]
  35. Jiang, F.; Wang, S.; Liu, B.; Liu, J.; Wang, L.; Xiao, Y.; Xu, Y.; Liu, X. Insights into the Influence of CeO2 Crystal Facet on CO2 Hydrogenation to Methanol over Pd/CeO2 Catalysts. ACS Catal. 2020, 10, 11493–11509. [Google Scholar] [CrossRef]
  36. Chang, S.; Li, M.; Hua, Q.; Zhang, L.; Ma, Y.; Ye, B.; Huang, W. Shape-Dependent Interplay Between Oxygen Vacancies and Ag–CeO2 Interaction in Ag/CeO2 Catalysts and Their Influence on the Catalytic Activity. J. Catal. 2012, 293, 195–204. [Google Scholar] [CrossRef]
  37. Han, X.; Li, J.; Lu, J.; Luo, S.; Wan, J.; Li, B.; Hu, C.; Cheng, X. High mass-loading NiCoLDH nanosheet arrays grown on carbon cloth by electrodeposition for excellent electrochemical energy storage. Nano Energy 2021, 86, 106079. [Google Scholar] [CrossRef]
  38. Mo, S.; Zhang, Q.; Li, S.; Ren, Q.; Zhang, M.; Xue, Y.; Peng, R.; Xiao, H.; Chen, Y.; Ye, D. Integrated Cobalt Oxide Based Nanoarray Catalysts with Hierarchical Architectures: In Situ Raman Spectroscopy Investigation on the Carbon Monoxide Reaction Mechanism. ChemCatChem 2018, 10, 3012–3026. [Google Scholar] [CrossRef]
  39. Spanier, J.E.; Robinson, R.D.; Zhang, F.; Chan, S.-W.; Herman, I.P. Size-Dependent Properties of CeO2−y Nanoparticles as Studied by Raman Scattering. Phys. Rev. B 2001, 64, 245407–245414. [Google Scholar] [CrossRef] [Green Version]
  40. Lin, J.; Li, L.; Huang, Y.; Zhang, W.; Wang, X.; Wang, A.; Zhang, T. In Situ Calorimetric Study: Structural Effects on Adsorption and Catalytic Performances for CO Oxidation over Ir-in-CeO2 and Ir-on-CeO2 Catalysts. J. Phys. Chem. C 2011, 115, 16509–16517. [Google Scholar] [CrossRef]
  41. Chen, L.; Liu, F.; Li, X.; Tao, Q.; Huang, Z.; Zuo, Q.; Chen, Y.; Li, T.; Fu, M.; Ye, D. Surface adsorbed and lattice oxygen activated by the CeO2/Co3O4 interface for enhancive catalytic soot combustion: Experimental and theoretical investigations. J. Colloid Interface Sci. 2023, 638, 109–122. [Google Scholar] [CrossRef]
  42. Wu, S.; Liu, H.; Huang, Z.; Xu, H.; Shen, W. Mn1ZrxOy mixed oxides with abundant oxygen vacancies for propane catalytic oxidation: Insights into the contribution of Zr doping. Chem. Eng. J. 2023, 452, 139341. [Google Scholar] [CrossRef]
  43. Zhu, W.; Chen, X.; Jin, J.; Di, X.; Liang, C.; Liu, Z. Insight into catalytic properties of Co3O4-CeO2 binary oxides for propane total oxidation. Chin. J. Catal. 2020, 41, 679–690. [Google Scholar] [CrossRef]
  44. Maintz, S.; Deringer, V.; Tchougreeff, A.; Dronskowski, R. LOBSTER: A Tool to Extract Chemical Bonding from Plane-Wave Based DFT. J. Comput. Chem. 2016, 37, 1030–1035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Song, C.; Zhan, Q.; Liu, F.; Wang, C.; Li, H.; Wang, X.; Guo, X.; Cheng, Y.; Sun, W.; Wang, L.; et al. Overturned Loading of Inert CeO2 to Active Co3O4 for Unusually Improved Catalytic Activity in Fenton-Like Reactions. Angew. Chem. Int. Ed. 2022, 61, e202200406. [Google Scholar] [CrossRef]
  46. Wang, X.; Song, H.; Ma, S.; Li, M.; He, G.; Xie, M.; Guo, X. Template ion-exchange synthesis of Co-Ni composite hydroxides nanosheets for supercapacitor with unprecedented rate capability. Chem. Eng. J. 2021, 432, 134319. [Google Scholar] [CrossRef]
  47. Giannozzi, P.; Baroni, S.; Bonini, N.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Chiarotti, G.L.; Cococcioni, M.; Dabo, I.; et al. QUANTUM ESPRESSO: A modular and open-source software project for quantum simulations of materials. J. Phys. Condens. Matter 2009, 21, 395502. [Google Scholar] [CrossRef] [PubMed]
  48. Giannozzi, P.; Andreussi, O.; Brumme, T.; Bunau, O.; Buongiorno Nardelli, M.; Calandra, M.; Car, R.; Cavazzoni, C.; Ceresoli, D.; Cococcioni, M.; et al. Advanced capabilities for materials modelling with Quantum ESPRESSO. J. Phys. Condens. Matter 2017, 29, 465901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Van Setten, M.J.; Giantomassi, M.; Bousquet, E.; Verstraete, M.J.; Hamann, D.R.; Gonze, X.; Rignanese, G.M. The PseudoDojo: Training and grading a 85 element optimized norm-conserving pseudopotential table. Comput. Phys. Commun. 2018, 226, 39–54. [Google Scholar] [CrossRef] [Green Version]
  50. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Schematic illustration of the synthesis of (a) Co3O4/CeO2–IM prepared by impregnation method and (b) Co3O4@CeO2–IE catalyst via inverse loading.
Figure 1. Schematic illustration of the synthesis of (a) Co3O4/CeO2–IM prepared by impregnation method and (b) Co3O4@CeO2–IE catalyst via inverse loading.
Molecules 28 05930 g001
Figure 2. The typical HRTEM images of (a,b) Co3O4@CeO2–IE catalyst. (d,e) Co3O4/CeO2–IM catalyst, (c) EDS elemental mappings of Co3O4@CeO2–IE and (f) Co3O4/CeO2−IM.
Figure 2. The typical HRTEM images of (a,b) Co3O4@CeO2–IE catalyst. (d,e) Co3O4/CeO2–IM catalyst, (c) EDS elemental mappings of Co3O4@CeO2–IE and (f) Co3O4/CeO2−IM.
Molecules 28 05930 g002
Figure 3. (a) Powder XRD patterns and (b) N2 adsorption/desorption isotherms of Co3O4@CeO2−IE and Co3O4/CeO2−IM.
Figure 3. (a) Powder XRD patterns and (b) N2 adsorption/desorption isotherms of Co3O4@CeO2−IE and Co3O4/CeO2−IM.
Molecules 28 05930 g003
Figure 4. (a) Catalytic performance and (b) Arrhenius plots of Co3O4@CeO2−IE and Co3O4/CeO2−IM.
Figure 4. (a) Catalytic performance and (b) Arrhenius plots of Co3O4@CeO2−IE and Co3O4/CeO2−IM.
Molecules 28 05930 g004
Figure 5. XPS spectra of Ce3d (a) (Ce3+: V0, VI, U0, UI (red line); Ce4+: V, VII, VIII, U, UII, UIII (blue line)), (b) O1s and (c) Co 2p for Co3O4@CeO2–IE and Co3O4/CeO2–IM.
Figure 5. XPS spectra of Ce3d (a) (Ce3+: V0, VI, U0, UI (red line); Ce4+: V, VII, VIII, U, UII, UIII (blue line)), (b) O1s and (c) Co 2p for Co3O4@CeO2–IE and Co3O4/CeO2–IM.
Molecules 28 05930 g005
Figure 6. (a,b) EPR, (c) Raman spectra and (d) O2−TPD profiles of Co3O4@CeO2−IE and Co3O4/CeO2−IM.
Figure 6. (a,b) EPR, (c) Raman spectra and (d) O2−TPD profiles of Co3O4@CeO2−IE and Co3O4/CeO2−IM.
Molecules 28 05930 g006
Figure 7. (a,b) The structures and(c,d) the electronic density of states (DOS) of the Co3O4−OV and Co3O4−OV−Ce substrates, respectively. The charge density difference of the substrate after propane adsorption is shown. The green dashed line represents the Fermi energy level (e,f), where the blue region represents a decrease in charge density and the yellow region represents an increase in charge density. The isosurface level is set at 0.002 e/Å3.
Figure 7. (a,b) The structures and(c,d) the electronic density of states (DOS) of the Co3O4−OV and Co3O4−OV−Ce substrates, respectively. The charge density difference of the substrate after propane adsorption is shown. The green dashed line represents the Fermi energy level (e,f), where the blue region represents a decrease in charge density and the yellow region represents an increase in charge density. The isosurface level is set at 0.002 e/Å3.
Molecules 28 05930 g007
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, X.; Liang, W.; Lin, C.; Zhang, T.; Zhang, J.; Sheng, N.; Song, Z.; Jiang, J.; Sun, B.; Xu, W. Enabling High Activity Catalyst Co3O4@CeO2 for Propane Catalytic Oxidation via Inverse Loading. Molecules 2023, 28, 5930. https://doi.org/10.3390/molecules28155930

AMA Style

Wang X, Liang W, Lin C, Zhang T, Zhang J, Sheng N, Song Z, Jiang J, Sun B, Xu W. Enabling High Activity Catalyst Co3O4@CeO2 for Propane Catalytic Oxidation via Inverse Loading. Molecules. 2023; 28(15):5930. https://doi.org/10.3390/molecules28155930

Chicago/Turabian Style

Wang, Xuan, Wei Liang, Changqing Lin, Tie Zhang, Jing Zhang, Nan Sheng, Zhaoning Song, Jie Jiang, Bing Sun, and Wei Xu. 2023. "Enabling High Activity Catalyst Co3O4@CeO2 for Propane Catalytic Oxidation via Inverse Loading" Molecules 28, no. 15: 5930. https://doi.org/10.3390/molecules28155930

Article Metrics

Back to TopTop