Next Article in Journal
Effect of Chalcogenophenes on Chiroptical Activity of Twisted Tetracenes: Computational Analysis, Synthesis and Crystal Structure Thereof
Next Article in Special Issue
Relation between Conventional and Starch-Assisted ASP Injection and Impact of Crystallinity on Flood Formation
Previous Article in Journal
One-Pot Graphene Supported Pt3Cu Nanoparticles—From Theory towards an Effective Molecular Oxygen Reduction Reaction Catalyst
Previous Article in Special Issue
Modulating the Oxygen Reduction Reaction Performance via Precisely Tuned Reactive Sites in Porphyrin-Based Covalent Organic Frameworks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis of Flower-Like Cobalt–Molybdenum Mixed-Oxide Microspheres for Deep Aerobic Oxidative Desulfurization of Fuel

1
Laboratory for Development & Application of Cold Plasma Technology, College of Chemistry and Chemical Engineering, Luoyang Normal University, Luoyang 471022, China
2
School of Agriculture and Bioengineering, Heze University, Heze 274015, China
3
Petrochemical Research Institute, PetroChina Company Limited, Beijing 102206, China
4
School of Environmental Engineering and Chemistry, Luoyang Institute of Science and Technology, Luoyang 471023, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(13), 5073; https://doi.org/10.3390/molecules28135073
Submission received: 10 June 2023 / Revised: 24 June 2023 / Accepted: 26 June 2023 / Published: 28 June 2023

Abstract

:
Flower-like cobalt–molybdenum mixed-oxide microspheres (CoMo-FMs) with hierarchical architecture were successfully synthesized via a hydrothermal process and subsequent calcination step. The characterization results show that CoMo-FMs were assembled from ultrathin mesoporous nanosheets with thicknesses of around 4.0 nm, providing the composite with a large pore volume and a massive surface area. The synthesized CoMo-FMs were employed as catalysts for the aerobic oxidative desulfurization (AODS) of fuel, and the reaction results show that the optimal catalyst (CoMo-FM-2) demonstrated an outstanding catalytic performance. Over CoMo-FM-2, various thiophenic sulfides could be effective removed at 80–110 °C under an atmospheric pressure, and a complete conversion of sulfides could be achieved in at least six consecutive cycles without a detectable change in chemical compositions. Further, the catalytic mechanism was explored by conducting systemic radical trapping and transformation experiments, and the excellent catalytic performance for CoMo-FMs should be mainly due to the synergistic effect of Mo and Co elements.

1. Introduction

The emission of sulfur oxides from the large-scale combustion of sulfur-containing fuels represents one of the main contributors of air pollution that cause severe environmental issues [1]. At the present time, hydrodesulfurization technology (HDS) is primarily applied in chemical industry to eliminate the environmental hazards; however, it requires harsh reaction conditions, incurring a substantial cost of investment and operation [2,3]. The content of organic sulfur in different fractions of crude oil shows an increasing trend with the rise in their boiling points; as a result, the sulfur in straight-run diesel oil can reach up to more than 10,000 ppm. Among the sulfides in straight-run diesel, the sulfur contained in dibenzothiophene (DBT) and its derivatives could be in excess of 1000 ppm in most cases. Due to the synergy of the electron effect and steric hindrance effect, these refractory sulfides are extremely difficult to remove with the traditional HDS [4,5]. Given that, it is necessary to develop supplement technologies that show such advantages as a low energy consumption and high conversion efficiency regarding the refractory sulfides [6,7,8].
Among the emerging alternative desulfurization technologies, oxidative desulfurization (ODS) has been considered to be the preferred one for the effective removal of refractory thiophenic compounds under mild reaction conditions [9,10,11,12]. According to the principle of ODS, refractory sulfides can be converted to sulfones and/or sulfoxides and then removed by means of solvent extraction [13,14,15]. In the previous studies, H2O2 was frequently taken as the oxidizing agent [16,17,18] but, obviously, its usage on a large scale in industry will inevitably face serious safety concerns. Recently, substituting dioxygen (O2) for H2O2 to achieve aerobic oxidative desulfurization (AODS) has become a research hotspot [19,20,21,22,23]. As the most environmentally friendly oxidant, however, O2 is relatively chemically inert due to its triplet ground state (3O2) [24,25,26]. As a result, for AODS reaction systems, irreversible sacrificial agents had to be used [27,28].
In the past ten years, great progress has been made in the catalyst preparation for AODS. A wide variety of catalytic materials have been reported, such as metal oxides [29,30,31,32], boron nitride (BN) [33,34,35], reduced graphene oxide (rGO) [36], deep eutectic solvents (DESs) [37], metal–organic frameworks (MOFs) [38], and polyoxometalates (POMs) [37,39,40]. However, from the point of view of performance, cost, and stability, those existing catalysts still cannot meet the requirements of industrial applications [41]. Developing more active, reliable, and low-cost AODS catalysts is highly desirable [41]. Further, AODS is a kind of heterogeneous catalytic reaction occurring in complex gas–liquid–solid systems. The structure and morphology of catalysts greatly affect the mass transfer process between phase interfaces [42]. Therefore, for the development of high-performance catalysts, it is necessary to make an adjustment to the shape and size of catalysts from the macro-scale and improve the connectivity of pores at the meso scale, thus improving the mass transfer efficiency between gas–solid and liquid–solid phases [43]. Notably, mixed-metal oxides (MMOs) have achieved rapid advances in the catalysis science [44,45,46]. MMOs are highly diverse with a designable structure and composition [47]. For MMOs, under the scientific design, hierarchical porous structures can be developed and desired electronic structures for O2 activation can be achieved so as to meet the requirements of multi-phase reactions of AODS.
Herein, we report the successful synthesis of flower-like Co-Mo-O microspheres (denoted as CoMo-FMs) via a hydrothermal process followed by a calcination step. Surface morphologic structures, electronic structures, and crystalline phases as well as chemical compositions of the resulting samples with different Co/Mo dosages were systematically investigated. Further, the catalytic performance of CoMo-FMs for the aerobic oxidative desulfurization (AODS) of fuel was tested under mild conditions with dioxygen, O2, in air as the oxidant. In addition, catalyst stability tests were performed by six cycles of experiments. Furthermore, a catalytic mechanism of CoMo bimetallic oxides catalyzing the reaction of AODS was put forward.

2. Results and Discussion

2.1. Synthesis and Characterization of CoMo-FMs

CoMo-FMs were obtained by hydro-thermal synthesis with CoCl2·and (NH4)6Mo7O24 as precursors followed by calcination steps. While keeping the amount of CoCl2 constant, the effects of Mo/Co ratios on the chemical composition, morphologies, and catalytic performance of the resulting samples were systematically investigated (Figure 1a). We first conducted inductively coupled plasma–atomic emission spectroscopy (ICP-AES) to quantify the elemental contents within the samples (Table 1). As expected, the Mo/Co atomic ratio increased from 0.151 to 0.488 with an increase in its dosage ratio. The powder X-ray diffraction (XRD) spectra of CoMo-FMs (Figure 1b) mainly contain the diffraction peaks of Co3O4 and β-CoMoO4 [17,48]. In addition, the diffraction peak intensity of β-CoMoO4 increases with an increase in the Mo/Co ratio. Fourier-transform infrared (FT-IR) spectra collection was performed in the range of 400~1600 cm−1 to characterize the chemical structure of CoMo-FMs (Figure 1c). The characteristic peaks at 655 and 549 cm−1 are assigned to the stretching mode of Co(II)-O and Co(III)-O bonds in the oxide Co3O4 [49]. The peak present at 943 cm−1 should be attributed to the distortion of the MoO42− tetrahedron because of the distortion in CoMoO4. The peak at approximately 847 cm−1 is identified as a characteristic peak of Mo-O-Mo [50,51]. It can be observed that the intensity of peaks assigned to Mo species get enhanced with the increasing of the Mo/Co ratio, consistent with the XRD results.
To probe the morphologies of CoMo-FMs, scanning electron microscope (SEM) characterization was conducted. Figure 2a–d confirm that all the catalysts exhibit a flower-like microsphere structure. The particle size of CoMo-FM-1 and CoMo-FM-2 is approximately 1.5 μm, and it becomes larger for CoMo-FM-3. When further increasing the Mo/Co ratio, the materials do not form a regular microsphere. Figure 2e–h demonstrate that the micro-flowers are constructed by ultrathin nanosheets. It can be observed that the nanosheets in CoMo-FM-1 display a smooth surface but prove to be porous in CoMo-FM-2, whereas the nanosheets in CoMo-FM-3 and CoMo-FM-4 exhibit uneven thickness and a small amount of lumps.
The micro/nano-structure of CoMo-FM-2 was then observed by transmission electron microscopy (TEM). Figure 3a clearly demonstrates that the flower-like microsphere consists of nanosheets extending inside out. Furthermore, the nanosheets have an ultrathin thickness of approximately 4.0 nm (Figure 3c). As revealed by the high-resolution transmission electron microscopy (HRTEM) image in Figure 3b, two distinctive crystal forms with the lattice distance of 0.2505 nm and 0.3415 nm are observed, belonging to the (311) and (220) planes of Co3O4 and β-CoMoO4, respectively, conforming to the XRD results. Moreover, the HRTEM image of CoMo-FM-2 (Figure 3b) indicates the presence of mesoporous pores (highlighted by the red circles) on nanosheets. As shown in Figure 3d–g, the energy-dispersive spectroscopy (EDS) elemental mapping spectra of CoMo-MF-2 demonstrate that the elements, Co, Mo, and O uniformly distribute in the sample. It also can be inferred from the EDS results (Figure 4) that the contents of Co, O, and Mo account for 25.84%, 66.97%, and 7.19%, respectively, which are consistent with those determined by ICP-AES (Table 1).
A N2 physical adsorption–desorption test was conducted to investigate the pore structures of CoMo-FMs. Condensation adsorption/desorption isotherms of type V are designated for all four obtained samples (Figure 5a). Moreover, there is a hysteresis loop at relative pressure (P/P0) between 0.6 and 1.0 for each sample, indicating the existence of meso-pores or macro-pores. In addition, it can be seen in Figure 5b that all samples contain meso- and macro-pores with sizes distributed in the region of 5–60 nm. The mesopores in nanosheets and larger gaps between nanosheets are responsible for their hierarchical porous structure. It is noteworthy that CoMo-FM-2 possessed the largest pore size, pore volume, and surface area, covering up to 103.46 m2/g in all the samples (Table 2), which may significantly increase the number and accessibility of the active sites.
To investigate the surface chemical composition and chemical states of elements of CoMo-FMs, X-ray photoelectron spectroscopy (XPS) spectra were collected and are shown in Figure 6.
It can be discovered from the wide scan curves (Figure 6a) that CoMo-FMs contain three elements, which are Co, O, and Mo. As indicated by Figure 6b, doublet Mo 3d peaks are displayed at around 230.5 and 233.6 eV, assigned to levels of 3d3/2 and 3d5/2, which are different from the MoO3 reference (231.52 and 234.65 eV). The results demonstrate that Mo species in CoMo-FMs exist in a mixed-valence state of Mo(V) and Mo(VI), and Mo species gain electrons from Co species in the mixed-metal oxides. Consequently, with a decrease in the Mo/Co molar ratio, the Mo 3d peaks are observed to shift toward a lower binding energy. The high-revolution curves of Co 2p (Figure 6c) were smoothly fitted by setting the area ratio of 2p1/2 and 2p3/2 as 1:2. For the 2p3/2 component, the peaks at the point of 778.2 and 781.2 eV belong to Co(II) and Co(III) in Co3O4 [52], while the peak at 779.1 eV is assigned to the Co species in CoMoO4 [53]. It is also observed that the increase in the Mo/Co ratio reduced the proportion of Co(II), which further confirms the electrical effect of Co species in the oxides. Figure 6d presents the XPS spectra of O 1s, with three peaks at 531.4, 529.7, and 528.4 eV deconvoluted. The three peaks are assigned to adsorbed water, adsorbed oxygen, and lattice oxygen, respectively [54]. The proportions of adsorbed oxygen were then quantitively calculated and compared (Table 3). The CoMo-FM-2 and CoMo-FM-3 samples possessed a larger Oad/Olat ratio, which indicates their higher concentration of oxygen vacancies.

2.2. Catalytic Performances of CoMo-FMs

Taking DBT as a model compound, the performances of CoMo-FMs for catalysis in deep aerobic oxidative desulfurization were evaluated. As shown in Figure 7a, all the catalysts display a noticeable sulfur removal effect. CoMo-FM-2 possesses the largest specific surface area and a relatively higher concentration of Mo(V) species and oxygen vacancies, performing the best in this work, achieving total transformation of DBT in 5 h. The hierarchical porous structure of CoMo-FM-2 is beneficial for reducing the mass transfer resistance of oxygen from the liquid–gas interface to the active sites of catalysts, while Mo(V) species and oxygen vacancies are supposed to be related to the catalytic behavior.
The catalytic mechanism of CoMo bimetallic oxides in AODS was then explored. Co3O4 has been used as an active substance in the aerobic oxidation of alcohols. Nevertheless, sole Co3O4 as well as MoO3 manifested little activity upon the aerobic oxidation of DBT, as is indicated in Figure 7b. Control experiments were also carried out. Firstly, p-benzoquinone (PB) and dimethyl sulfoxide (DMSO) were charged into and employed as a radical scavenger of O2•− and OH, respectively, to identify the catalytic active substance [55,56]. The catalytic performance of CoMo-FM-2 deteriorated critically with the adding of PB, but was almost unaffected when DMSO was added. The results demonstrate that O2•− instead of OH was an active intermediate substance during the catalytic reaction, whereas Co3O4 was also reported to be capable of activate O2 by forming O2•−. Then, Co3O4 and benzaldehyde (BA) were used as catalysts and showed obvious catalytic effect by achieving 49.6% of DBT conversion. In this reaction system, the generated O2•− could be made further active by BA, forming active peroxyl species [27], and could then react with the thiophenic compounds. The above results reveal that both neat Co3O4 and CoMo bimetallic oxides can form active species with O2. However, the generated O2•− is rich in electrons, and its direct reaction with thiophenic compounds is thus prohibited, while the Mo(V) species with oxygen vacancy is ready to interact with O2•− to generate electrophilic species and hence realize the successful aerobic oxidation of sulfides (Scheme 1). Pure MoO3 does not have enough oxygen vacancies to form active species with O2•−, which also results in a low catalytic activity of the mixture of Co3O4 and MoO3. As shown in Figure 7c, as the temperature of the reaction rises, the conversions of DBT are clearly enhanced. It can be observed that complete DBT conversion is realized within 4 h at 100 °C.
Moreover, compared with those previously reported catalysts for the aerobic oxidation of DBT (Table 4), CoMo-FM-2 is evidenced to achieve a better catalytic effect under a similar reaction condition. In addition, it is observed that white needle-like crystals gradually precipitate in the oil phase during the cooling of the reaction system.
We then collected the solid by filtration, dissolved them in ethanol, and analyzed the composition with GC-MS. It was evidenced that DBT was converted into dibenzothiophene sulfone (DBTO2). In addition, the first-order reaction kinetics were analyzed, and apparent reaction rate constants at different measurement temperatures were measured and then calculated (Figure 8a). Moreover, the apparent activation energy Ea of aerobic DBT oxidation was found to be approximately 75.69 kJ/mol (Figure 8b).
In addition, the effect of catalyst dosage on DBT conversion efficiency was also investigated. It was found that DBT conversion did not improve significantly with the increase in the amount of catalyst as revealed by Figure 9. Even when the catalyst dosage increased from 20 mg to 40 mg and then to 60 mg, the DBT conversion was only slightly improved. This is because the catalyst has a very low density, and excessive addition of the catalyst may lead to poor dispersion.

2.3. Catalytic Activities on Various Sulfides

As is known, the sulfur compounds in actual diesel oil are varied, among which, benzothiophene (BT), DBT, and their derivatives account for the most. Therefore, the performance of CoMo-FM-2 for the oxidation of other two sulfides, BT and 4,6-dimethyl dibenzothiophene (4,6-DMDBT), was also investigated in this study, and the results are shown in Figure 7d. It can be seen that the conversion of 4,6-DMDBT shows similarity with DBT, and complete conversions are reached within 5 h, while the performance of CoMo-FM-2 for BT removal was much lower. The difference in oxidative reactivity for the three thiophene compounds should be correlated with the electron density of sulfur atoms in their molecular structures [61]. In addition, gas chromatography–mass spectrometry (GC-MS) was employed to detect the products from the aerobic oxidation of DBT and 4,6-DMDBT, and the analysis results show that the two thiophene compounds were converted into corresponding sulfones.

2.4. Reusability of the Catalyst

For a new catalyst, stability is a key performance index, and determines the reliability, economy, and ultimately the possibility of industrial applications of the catalyst. Given that, the catalyst CoMo-FM-2 was used to catalyze aerobic DBT oxidation for up to six cycles to test its stability. Briefly, CoMo-FM-2 was thoroughly mixed with model oil and stirring was maintained to catalyze the AODS reaction under the condition of 100 °C, 20 mL/min air, 20 mg catalyst, 20 g oil, and 500 ppm DBT for 4 h. After that, the slurry was centrifuged to collect the catalyst in the solid phase. Subsequently, the collected catalyst was washed by ethanol followed by vacuum drying for the next cycle. It is worth noting that the conversion of DBT remains 100% for each cycle experiment (Figure 10a), suggesting an excellent stability of the sample CoMo-FM-2. FTIR spectra were then used to compare the catalyst before and after recycle use and are shown in Figure 10b. After cycle usage for six times, CoMo-FM-2 still has clear FTIR characteristic absorption bands with no shift observed, while the new bands at 1250–1100 cm−1 are due to the generated sulfones adsorbed by the catalysts.

3. Materials and Methods

3.1. Chemicals

Benzothiophene (BT), dibenzothiophene (DBT), 4,6-dimethyl dibenzothiophene (4,6-DMDBT), cobalt (II) chloride hexahydrate (CoCl2·6H2O), ammonium heptamolybdate tetrahydrate ((NH4)6Mo7O24·4H2O), and hexamethylenetetramine (HMTA) were supplied by Aladdin Reagent Co., Ltd. (Shanghai, China). Ethanol (EtOH), benzaldehyde (BA), p-benzoquinone (PB), dimethyl sulfoxide (DMSO), and decahydronaphthalene were sourced from Tianjin Kermel Chemical Reagent Co., Ltd. (Tianjin, China). The purity of solid chemicals above was above 98%, the solvents were analytical grade, and all of them were used as received.

3.2. Preparation of CoMo-FMs

CoCl2·6H2O (1.5 mmol), (NH4)6Mo7O24·6H2O, and HMTA (10 mmol) were dissolved into 36.0 mL pure water and stirred to obtain a lavender solution, which was then homogenized and charged into a lining of a stainless-steel autoclave. After sealing, it was placed into an oven at 110 °C for 15 h followed by naturally cooling down to room temperature. The resulting slurry was centrifuged for collection of the solid product, which was then washed with pure water six times and followed by ethanol washing for one time. After drying at 60 °C for 8 h, the sample was subjected to calcination at 450 °C for 2 h to obtain CoMo bimetal oxides (denoted as CoMo-FM-x, x =1, 2, 3, and 4, with the dosage of (NH4)6Mo7O24·6H2O as 0.05, 0.10, 0.15, and 0.20 mmol, respectively).

3.3. Test of Catalytic Activity

Briefly, a sulfide (BT, DBT, or 4,6-DMDBT) was firstly dissolved in decahydronaphthalene to prepare a model oil with a sulfide concentration of 500 ppm. For the process of desulfurization experiments, the model oil (20 mL) and a certain amount of catalyst were placed into a 100 mL three-necked flask that was connected with a reflux condensing tube, a gas injector, and a port for the subsequent liquid sampling. Subsequently, the mixed solution was stirred continuously at 400 rpm at a desired temperature. Air was passed through a drying column to remove trace water prior to being bubbled into the liquid phase under atmospheric pressure for the reaction to be initiated. Sampling from the three-necked flask was performed at intervals for analyzing and determining the concentration of the residual sulfide by a GC-FID (HP-5 7820, Zhongke Huifen Instrument Co., Ltd., Beijing, China). For each AODS reaction, the experiment was repeated at least three times.

3.4. Characterization of Materials

The element content of Co and Mo in CoMo-FMs was quantitatively determined using the method of inductively coupled plasma–atomic emission spectroscopy (ICP-AES) (ICPE-9000, Shimadzu Co., Kyoto, Japan) Morphology of all the resulting catalysts was observed by a field emission scanning electron microscope (FE-SEM, XFlash 6100, Bruker, Karlsruhe, Germany). Powder X-ray diffraction (XRD) patterns of CoMo-FMs were identified by a D/max-2500V X-ray diffractometer (Rigaku Corp., Tokyo, Japan) using a Cu Kα (λ = 0.15418 nm) radiation source. Transmission electron microscopy (TEM), high-angle annular dark-field scanning TEM (HAADF-STEM), and energy-dispersive X-ray spectrometry (EDS) observations were conducted by using a transmission electron microscope (Talos F200X, Thermo Fisher Scientific, Waltham, MA, USA) equipped with an EDS equipment and a digitally acquired STEM imaging system. The Fourier-transform infrared (FT-IR) spectra of samples were obtained using a Nicolet is50 FT-IR spectrometer (Thermo Fisher Scientific, Waltham, MA, USA). The chemical states of several typical elements in CoMo-FMs were evaluated by X-ray photoelectron spectroscopy (XPS), performed on an X-ray photoelectron spectrometer (EscaLab Xi, Thermo Fisher Scientific, Madison, WI, USA). N2 physical adsorption/ desorption test was performed by a fully automated apparatus of physical adsorption (ASAP 2020 Plus HD88, Micromeritics Co., Norcross, GA, USA).

4. Conclusions

In summary, flower-like Co–Mo–O microspheres with hierarchical architecture were synthesized by a hydrothermal process followed by a calcination step. The micro spherical catalysts comprise Co3O4/CoMoO4 ultrathin porous nanosheets. It is shown that the diameter of the microspheres, the thickness of the nanosheets, and the diameter of channels on nanosheets are around 1.5 μm, 4 nm, and 10 nm, respectively, which makes the catalyst show an interconnected porous structure as well as a large pore size. The catalyst CoMo-FM-2 with a Mo/Co ratio of 0.265 shows the best catalytic performance for the AODS reaction; in particular, it could achieve the complete conversion of DBT and 4,6-DMDBT at 100 °C using atmospheric O2 as the oxidant source. It is worth noting that CoMo-FM-2 maintains almost constant catalytic activity for the AODS reaction during a six-cycle stability test. Radical trapping and transformation experiments reveal the synergistic effect of cobalt and molybdenum and the catalytic mechanism of the Co–Mo–O microspheres for the AODS reaction. Accordingly, the flower-like Co–Mo–O microspheres can be applied as robust and durable catalysts for AODS, and our results also reveal the significance of a rational design of a hierarchical structure for heterogeneous catalysis.

Author Contributions

Conceptualization, X.C. and J.W.; methodology, L.Z.; validation, L.Z.; formal analysis, R.T.; investigation, R.T., Y.W., Y.L., and X.W.; resources, J.W. and R.T.; data curation, Y.W., Y.L., and X.W.; writing—original draft preparation, X.C.; writing—review and editing, R.T., J.W., and L.Z.; supervision, R.T. and J.W.; project administration, X.C.; funding acquisition, L.Z. and X.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Key R&D and Promotion Special Projects (Science and Technology) of Henan Province, China, grant number 212102210214 and 232102240068.

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.

References

  1. Gungor, A. Prediction of SO2 and NOx emissions for low-grade Turkish lignites in CFB combustors. Chem. Eng. J. 2009, 146, 388–400. [Google Scholar] [CrossRef]
  2. Mahmoudabadi, Z.S.; Rashidi, A.; Tavasoli, A. Synthesis of MoS2 quantum dots as a nanocatalyst for hydrodesulfurization of naphtha: Experimental and DFT study. J. Environ. Chem. Eng. 2020, 8, 103736. [Google Scholar] [CrossRef]
  3. Jamali, M.A.; Arvani, A.; Amini, M.M. Vanadium containing metal-organic frameworks as highly efficient catalysts for the oxidation of refractory aromatic sulfur compounds. ChemCatChem 2020, 13, 293–303. [Google Scholar] [CrossRef]
  4. Safa, M.A.; Bouresli, R.; Al-Majren, R.; Al-Shamary, T.; Ma, X.L. Oxidative desulfurization kinetics of refractory sulfur compounds in hydrotreated middle distillates. Fuel 2019, 239, 24–31. [Google Scholar] [CrossRef]
  5. Zhang, M.T.; Yan, T.T.; Dai, W.L.; Guan, N.J.; Li, L.D. Zeolite stabilized isolated molybdenum species for catalytic oxidative desulfurization. Acta Chim. Sin. 2020, 78, 1404–1410. [Google Scholar] [CrossRef]
  6. Xiong, J.; Yang, L.; Chao, Y.H.; Pang, J.Y.; Wu, P.W.; Zhang, M.; Zhu, W.S.; Li, H.M. A large number of low coordinated atoms in boron nitride for outstanding adsorptive desulfurization performance. Green Chem. 2016, 18, 3040–3047. [Google Scholar] [CrossRef]
  7. Caero, L.C.; Hernández, E.; Pedraza, F.; Murrieta, F. Oxidative desulfurization of synthetic diesel using supported catalysts. Catal. Today 2005, 107–108, 564–569. [Google Scholar] [CrossRef]
  8. Mhemed, H.A.; Gallego, M.M.; Largeau, J.F.; Kordoghli, S.; Zagrouba, F.; Tazerout, M. Gas adsorptive desulfurization of thiophene by spent coffee grounds-derived carbon optimized by response surface methodology: Isotherms and kinetics evaluation. J. Environ. Chem. Eng. 2020, 8, 104036. [Google Scholar] [CrossRef]
  9. Smolders, S.; Willhammar, T.; Krajnc, A.; Sentosun, K.; Wharmby, M.T.; Lomachenko, K.A.; Bals, S.; Mali, G.; Roeffaers, M.B.J.; De Vos, D.E.; et al. A titanium(iv)-based metal-organic framework featuring defect-rich Ti–O sheets as an oxidative desulfurization catalyst. Angew. Chem. Int. Ed. 2019, 58, 9160–9165. [Google Scholar] [CrossRef]
  10. Zhang, X.M.; Zhang, Z.H.; Zhang, B.H.; Yang, X.F.; Chang, X.; Zhou, Z.; Wang, D.H.; Zhang, M.H.; Bu, X.H. Synergistic effect of Zr-MOF on phosphomolybdic acid promotes efficient oxidative desulfurization. Appl. Catal. B-Environ. 2019, 256, 117804. [Google Scholar] [CrossRef]
  11. Baradaran, S.; Sadeghi, M.T. Desulfurization of non-hydrotreated kerosene using hydrodynamic cavitation assisted oxidative desulfurization (HCAOD) process. J. Environ. Chem. Eng. 2020, 8, 103832. [Google Scholar] [CrossRef]
  12. Qi, Z.Y.; Huang, Z.X.; Wang, H.X.; Li, L.; Ye, C.S.; Qiu, T. In situ bridging encapsulation of a carboxyl-functionalized phosphotungstic acid ionic liquid in UiO-66: A remarkable catalyst for oxidative desulfurization. Chem. Eng. Sci. 2020, 225, 115818. [Google Scholar] [CrossRef]
  13. Zuo, M.G.; Huang, X.Q.; Li, J.X.; Chang, Q.; Duan, Y.S.; Yan, L.J.; Xiao, Z.Z.; Mei, S.D.; Lu, S.X.; Yao, Y. Oxidative desulfurization in diesel via a titanium dioxide triggered thermocatalytic mechanism. Catal. Sci. Technol. 2019, 9, 2923–2930. [Google Scholar] [CrossRef]
  14. Dizaji, A.K.; Mokhtarani, B.; Mortaheb, H.R. Deep and fast oxidative desulfurization of fuels using graphene oxide-based phosphotungstic acid catalysts. Fuel 2019, 236, 717–729. [Google Scholar] [CrossRef]
  15. Kang, M.; Wang, X.; Zhang, J.; Lu, Y.; Chen, X.; Yang, L.; Wang, F. Boosting the photocatalytic oxidative desulfurization of dibenzothiophene by decoration of MWO4 (M=Cu, Zn, Ni) on WO3. J. Environ. Chem. Eng. 2019, 7, 102809. [Google Scholar] [CrossRef]
  16. Zhu, J.; Wu, P.W.; Chen, L.L.; He, J.; Wu, Y.C.; Wang, C.; Chao, Y.H.; Lu, L.J.; He, M.Q.; Zhu, W.; et al. 3D-printing of integrated spheres as a superior support of phosphotungstic acid for deep oxidative desulfurization of fuel. J. Energy. Chem. 2020, 45, 91–97. [Google Scholar] [CrossRef] [Green Version]
  17. Liu, Y.; Han, L.; Zhang, J.; Yao, R.; Zhan, H.; Yang, H.; Bai, L.; Yang, L.; Wei, D.; Wang, W.; et al. Morphology-controlled construction and aerobic oxidative desulfurization of hierarchical hollow Co–Ni–Mo–O mixed metal-oxide nanotubes. Ind. Eng. Chem. Res. 2020, 59, 6488–6496. [Google Scholar] [CrossRef]
  18. Ribeiro, S.O.; Juliao, D.; Cunha-Silva, L.; Domingues, V.F.; Valenca, R.; Ribeiro, J.C.; de Castro, B.; Balula, S.S. Catalytic oxidative/extractive desulfurization of model and untreated diesel using hybrid based zinc-substituted polyoxometalates. Fuel 2016, 166, 268–275. [Google Scholar] [CrossRef]
  19. Zeng, X.Y.; Xiao, X.Y.; Li, Y.; Chen, J.Y.; Wang, H.L. Deep desulfurization of liquid fuels with molecular oxygen through graphene photocatalytic oxidation. Appl. Catal. B-Environ. 2017, 209, 98–109. [Google Scholar] [CrossRef]
  20. Zhu, W.S.; Wang, C.; Li, H.P.; Wu, P.W.; Xun, S.H.; Jiang, W.; Chen, Z.G.; Zhao, Z.; Li, H.M. One-pot extraction combined with metal-free photochemical aerobic oxidative desulfurization in deep eutectic solvent. Green Chem. 2015, 17, 2464–2472. [Google Scholar] [CrossRef]
  21. Zhong, W.Z.; Liu, M.Q.; Dai, J.; Yang, J.; Mao, L.Q.; Yin, D.L. Synergistic hollow CoMo oxide dual catalysis for tandem oxygen transfer: Preferred aerobic epoxidation of cyclohexene to 1,2-epoxycyclohexane. Appl. Catal. B-Environ. 2018, 225, 180–196. [Google Scholar] [CrossRef]
  22. Lu, S.Y.; Lu, Y.; Jin, M.; Bao, S.J.; Li, W.Y.; Yu, L. Design and fabrication of highly sensitive and stable biochip for glucose biosensing. Appl. Surf. Sci. 2017, 422, 900–904. [Google Scholar] [CrossRef]
  23. Yang, H.; Zhang, Q.; Zhang, J.; Yang, L.; Ma, Z.; Wang, L.; Li, H.; Bai, L.; Wei, D.; Wang, W.; et al. Cellulose nanocrystal shelled with poly(ionic liquid)/polyoxometalate hybrid as efficient catalyst for aerobic oxidative desulfurization. J. Colloid. Interface Sci. 2019, 554, 572–579. [Google Scholar] [CrossRef] [PubMed]
  24. Zeng, X.Y.; Xiao, X.Y.; Chen, J.Y.; Wang, H.L. Electron-hole interactions in choline-phosphotungstic acid boosting molecular oxygen activation for fuel desulfurization. Appl. Catal. B-Environ. 2019, 248, 573–586. [Google Scholar] [CrossRef]
  25. Abednatanzi, S.; Gohari Derakhshandeh, P.; Leus, K.; Vrielinck, H.; Callens, F.; Schmidt, J.; Savateev, A.; Van Der Voort, P. Metal-free activation of molecular oxygen by covalent triazine frameworks for selective aerobic oxidation. Sci. Adv. 2020, 6, eaaz2310. [Google Scholar] [CrossRef] [Green Version]
  26. Lin, F.; Wang, D.G.; Jiang, Z.X.; Ma, Y.; Li, J.; Li, R.G.; Li, C. Photocatalytic oxidation of thiophene on BiVO4 with dual co-catalysts Pt and RuO2 under visible light irradiation using molecular oxygen as oxidant. Energy Environ. Sci. 2012, 5, 6400–6406. [Google Scholar] [CrossRef]
  27. Murata, S.; Murata, K.; Kidena, K.; Nomura, M. A novel oxidative desulfurization system for diesel fuels with molecular oxygen in the presence of cobalt catalysts and aldehydes. Energy Fuels 2004, 18, 116–121. [Google Scholar] [CrossRef]
  28. Lu, Y.; Wang, Y.; Gao, L.; Chen, J.; Mao, J.; Xue, Q.; Liu, Y.; Wu, H.; Gao, G.; He, M. Aerobic oxidative desulfurization: A promising approach for sulfur removal from fuels. ChemSusChem 2008, 1, 302–306. [Google Scholar] [CrossRef]
  29. Shi, Y.; Liu, G.; Zhang, B.; Zhang, X. Oxidation of refractory sulfur compounds with molecular oxygen over a Ce–Mo–O catalyst. Green Chem. 2016, 18, 5273–5279. [Google Scholar] [CrossRef]
  30. Wang, C.; Qiu, Y.; Wu, H.Y.; Yang, W.S.; Zhu, Q.; Chen, Z.G.; Xun, S.H.; Zhu, W.S.; Li, H.M. Construction of 2D-2D V2O5/BNNS nanocomposites for improved aerobic oxidative desulfurization performance. Fuel 2020, 270, 117498. [Google Scholar] [CrossRef]
  31. Lu, L.J.; He, J.; Wu, P.W.; Wu, Y.C.; Chao, Y.H.; Li, H.P.; Tao, D.J.; Fan, L.; Li, H.M.; Zhu, W.S. Taming electronic properties of boron nitride nanosheets as metal-free catalysts for aerobic oxidative desulfurization of fuels. Green Chem. 2018, 20, 4453–4460. [Google Scholar] [CrossRef]
  32. Wu, P.; Yang, S.; Zhu, W.; Li, H.; Chao, Y.; Zhu, H.; Li, H.; Dai, S. Tailoring N-terminated defective edges of porous boron nitride for enhanced aerobic catalysis. Small 2017, 13, 1701857. [Google Scholar] [CrossRef]
  33. Li, H.; Fu, W.; Yin, J.; Zhang, J.; Li, Y.; Jiang, D.; Lv, N.; Zhu, W. Rational design of the carbon doping of hexagonal boron nitride for oxygen activation and oxidative desulfurization. Phys. Chem. Chem. Phys. 2020, 22, 24310–24319. [Google Scholar] [CrossRef] [PubMed]
  34. Gu, Q.Q.; Wen, G.D.; Ding, Y.X.; Wu, K.H.; Chen, C.M.; Su, D.S. Reduced graphene oxide: A metal-free catalyst for aerobic oxidative desulfurization. Green Chem. 2017, 19, 1175–1181. [Google Scholar] [CrossRef]
  35. Gomez-Paricio, A.; Santiago-Portillo, A.; Navalon, S.; Concepcion, P.; Alvaro, M.; Garcia, H. MIL-101 promotes the efficient aerobic oxidative desulfurization of dibenzothiophenes. Green Chem. 2016, 18, 508–515. [Google Scholar] [CrossRef]
  36. Zhang, Q.; Zhang, J.; Yang, H.; Dong, Y.; Liu, Y.; Yang, L.; Wei, D.; Wang, W.; Bai, L.; Chen, H. Efficient aerobic oxidative desulfurization over Co–Mo–O bimetallic oxide catalysts. Catal. Sci. Technol. 2019, 9, 2915–2922. [Google Scholar] [CrossRef]
  37. Sun, L.L.; Su, T.; Xu, J.J.; Hao, D.M.; Liao, W.P.; Zhao, Y.C.; Ren, W.Z.; Deng, C.L.; Lu, H.Y. Aerobic oxidative desulfurization coupling of Co polyanion catalysts and p-TsOH-based deep eutectic solvents through a biomimetic approach. Green Chem. 2019, 21, 2629–2634. [Google Scholar] [CrossRef]
  38. Jiang, W.; Xiao, J.; Dong, L.; Wang, C.; Li, H.P.; Luo, Y.P.; Zhu, W.S.; Li, H.M. Polyoxometalate-based poly(ionic liquid) as a precursor for superhydrophobic magnetic carbon composite catalysts toward aerobic oxidative desulfurization. ACS Sustain. Chem. Eng. 2019, 7, 15755–15761. [Google Scholar] [CrossRef]
  39. Tang, N.F.; Jiang, Z.X.; Li, C. Oxidation of refractory sulfur-containing compounds with molecular oxygen catalyzed by vanadoperiodate. Green Chem. 2015, 17, 817–820. [Google Scholar] [CrossRef]
  40. Lü, H.; Ren, W.; Liao, W.; Chen, W.; Li, Y.; Suo, Z. Aerobic oxidative desulfurization of model diesel using a B-type anderson catalyst [(C18H37)2N(CH3)2]3Co(OH)6Mo6O18·3H2O. Appl. Catal. B-Environ. 2013, 138–139, 79–83. [Google Scholar] [CrossRef]
  41. Shafiq, I.; Shafique, S.; Akhter, P.; Ishaq, M.; Yang, W.; Hussain, M. Recent breakthroughs in deep aerobic oxidative desulfurization of petroleum refinery products. J. Clean. Prod. 2021, 294, 125731. [Google Scholar] [CrossRef]
  42. Zhang, M.; Liao, W.; Wei, Y.; Wang, C.; Fu, Y.; Gao, Y.; Zhu, L.; Zhu, W.; Li, H. Aerobic oxidative desulfurization by nanoporous tungsten oxide with oxygen defects. ACS Appl. Nano. Mater. 2021, 4, 1085–1093. [Google Scholar] [CrossRef]
  43. Rajendran, A.; Cui, T.Y.; Fan, H.X.; Yang, Z.F.; Feng, J.; Li, W.Y. A comprehensive review on oxidative desulfurization catalysts targeting clean energy and environment. J. Mater. Chem. A 2020, 8, 2246–2285. [Google Scholar] [CrossRef]
  44. Kim, S.O.; Sastri, C.V.; Seo, M.S.; Kim, J.; Nam, W. Dioxygen activation and catalytic aerobic oxidation by a mononuclear nonheme iron(ii) complex. J. Am. Chem. Soc. 2005, 127, 4178–4179. [Google Scholar] [CrossRef] [Green Version]
  45. Wachs, I.E. Recent conceptual advances in the catalysis science of mixed metal oxide catalytic materials. Catal. Today 2005, 100, 79–94. [Google Scholar] [CrossRef]
  46. Lu, S.Y.; Li, S.W.; Jin, M.; Gao, J.C.; Zhang, Y. Greatly boosting electrochemical hydrogen evolution reaction over Ni3S2 nanosheets rationally decorated by Ni3Sn2S2 quantum dots. Appl. Catal. B-Environ. 2020, 267, 118675. [Google Scholar] [CrossRef]
  47. Tian, Y.; Zhou, M.; Meng, X.; Miao, Y.; Zhang, D.J.M.C. Physics Needle-like comoo with multi-modal porosity for pseudocapacitors. Mater. Chem. Phys. 2017, 198, 258–265. [Google Scholar] [CrossRef]
  48. Zhang, X.; Li, J.; Yang, Y.; Zhang, S.; Zhu, H.; Zhu, X.; Xing, H.; Zhang, Y.; Huang, B.; Guo, S.; et al. Co3O4/Fe0.33Co0.66P interface nanowire for enhancing water oxidation catalysis at high current density. Adv. Mater. 2018, 30, e1803551. [Google Scholar] [CrossRef]
  49. Dong, Y.; Zhang, J.; Ma, Z.; Xu, H.; Yang, H.; Yang, L.; Bai, L.; Wei, D.; Wang, W.; Chen, H. Preparation of co-mo-o ultrathin nanosheets with outstanding catalytic performance in aerobic oxidative desulfurization. Chem. Commun. 2019, 55, 13995–13998. [Google Scholar] [CrossRef]
  50. Yang, H.W.; Bai, L.J.; Wei, D.L.; Yang, L.X.; Wang, W.X.; Chen, H.; Niu, Y.Z.; Xue, Z.X. Ionic self-assembly of poly(ionic liquid)-polyoxometalate hybrids for selective adsorption of anionic dyes. Chem. Eng. J. 2019, 358, 850–859. [Google Scholar] [CrossRef]
  51. Li, J.; Zhang, Q.; Zhang, J.; Wu, N.; Liu, G.; Chen, H.; Yuan, C.; Liu, X. Optimizing electronic structure of porous Ni/MoO2 heterostructure to boost alkaline hydrogen evolution reaction. J. Colloid Interface Sci. 2022, 627, 862–871. [Google Scholar] [CrossRef] [PubMed]
  52. Cai, Y.F.; Xu, J.; Guo, Y.; Liu, J.Y. Ultrathin, polycrystalline, two-dimensional Co3O4 for low-temperature CO oxidation. ACS Catal. 2019, 9, 2558–2567. [Google Scholar] [CrossRef]
  53. Yu, M.Q.; Jiang, L.X.; Yang, H.G. Ultrathin nanosheets constructed CoMoO4 porous flowers with high activity for electrocatalytic oxygen evolution. Chem. Commun. 2015, 51, 14361–14364. [Google Scholar] [CrossRef] [PubMed]
  54. Santos, V.P.; Pereira, M.F.R.; Orfao, J.J.M.; Figueiredo, J.L. The role of lattice oxygen on the activity of manganese oxides towards the oxidation of volatile organic compounds. Appl. Catal. B-Environ. 2010, 99, 353–363. [Google Scholar] [CrossRef]
  55. Rodriguez, E.M.; Marquez, G.; Tena, M.; Alvarez, P.M.; Beltran, F.J. Determination of main species involved in the first steps of TiO2 photocatalytic degradation of organics with the use of scavengers: The case of ofloxacin. Appl. Catal. B-Environ. 2015, 178, 44–53. [Google Scholar] [CrossRef]
  56. Panganamala, R.V.; Sharma, H.M.; Heikkila, R.E.; Geer, J.C.; Cornwell, D.G. Role of hydroxyl radical scavengers dimethyl sulfoxide, alcohols and methional in the inhibition of prostaglandin biosynthesis. Prostaglandins 1976, 11, 599–607. [Google Scholar] [CrossRef]
  57. Shi, M.; Zhang, D.; Yu, X.; Li, Y.; Wang, X.; Yang, W. Deep oxidative desulfurization catalyzed by (NH4)5H6PV8Mo4O40 using molecular oxygen as an oxidant. Fuel Process. Technol. 2017, 160, 136–142. [Google Scholar] [CrossRef]
  58. Yu, J.; Zhu, Z.; Ding, Q.; Zhang, Y.; Wu, X.; Sun, L.; Du, J. Oxidative desulfurization of dibenzothiophene with molecular oxygen using cobalt and copper salen complexes encapsulated in nay zeolite. Catal. Today 2019, 339, 105–112. [Google Scholar] [CrossRef]
  59. Li, X.; Gu, Y.; Chu, H.; Ye, G.; Zhou, W.; Xu, W.; Sun, Y. MFM-300(V) as an active heterogeneous catalyst for deep desulfurization of fuel oil by aerobic oxidation. Appl. Catal. A–Gen. 2019, 584, 117152. [Google Scholar] [CrossRef]
  60. Yu, X.; Shi, M.; Yan, S.; Wang, H.; Wang, X.; Yang, W. Designation of choline functionalized polyoxometalates as highly active catalysts in aerobic desulfurization on a combined oxidation and extraction procedure. Fuel 2017, 207, 13–21. [Google Scholar] [CrossRef]
  61. Qiu, J.; Wang, G.; Zeng, D.; Tang, Y.; Wang, M.; Li, Y. Oxidative desulfurization of diesel fuel using amphiphilic quaternary ammonium phosphomolybdate catalysts. Fuel Process. Technol. 2009, 90, 1538–1542. [Google Scholar] [CrossRef]
Figure 1. (a) Schematic illustration of AODS process over CoMo-FMs. (b) XRD patterns and (c) FT-IR spectra of CoMo-FMs.
Figure 1. (a) Schematic illustration of AODS process over CoMo-FMs. (b) XRD patterns and (c) FT-IR spectra of CoMo-FMs.
Molecules 28 05073 g001
Figure 2. SEM characterization of the CoMo flower-like microsphere catalysts. (a,e) CoMo-FM-1, (b,f) CoMo-FM-2, (c,g) CoMo-FM-3, and (d,h) CoMo-FM-4.
Figure 2. SEM characterization of the CoMo flower-like microsphere catalysts. (a,e) CoMo-FM-1, (b,f) CoMo-FM-2, (c,g) CoMo-FM-3, and (d,h) CoMo-FM-4.
Molecules 28 05073 g002
Figure 3. (a) TEM, (b,c) HRTEM, and (d) high-angle annular dark-field scanning TEM (HAADF-STEM) images of CoMo-FM-2 and corresponding EDS elemental mappings of (e) Co, (f) Mo, and (g) O.
Figure 3. (a) TEM, (b,c) HRTEM, and (d) high-angle annular dark-field scanning TEM (HAADF-STEM) images of CoMo-FM-2 and corresponding EDS elemental mappings of (e) Co, (f) Mo, and (g) O.
Molecules 28 05073 g003
Figure 4. EDS results of CoMo-FM-2 performed by TEM.
Figure 4. EDS results of CoMo-FM-2 performed by TEM.
Molecules 28 05073 g004
Figure 5. (a) N2 condensation adsorption/desorption isotherms and (b) the pore diameter distributions of CoMo-FMs.
Figure 5. (a) N2 condensation adsorption/desorption isotherms and (b) the pore diameter distributions of CoMo-FMs.
Molecules 28 05073 g005
Figure 6. (a) XPS survey spectrums of CoMo-FMs and high-revolution XPS spectrums of (b) Mo3d, (c) Co2p, and (d) O1s.
Figure 6. (a) XPS survey spectrums of CoMo-FMs and high-revolution XPS spectrums of (b) Mo3d, (c) Co2p, and (d) O1s.
Molecules 28 05073 g006
Figure 7. (a) Aerobic catalytic performance of CoMo-FMs. Reaction conditions: 20 mg catalyst, 20 g oil, 500 ppm DBT, and 100 °C. (b) Comparison of the effect of aerobic DBT oxidation in different catalytic systems. Reaction condition: 20 mg catalyst, 500 ppm DBT, 100 °C, and 4 h. (c) Catalytic performance of CoMo-FMs at different reaction temperatures in aerobic DBT oxidation. Reaction conditions: 20 mg catalyst, 20 g oil, and 500 ppm DBT (d) Exploring the catalytic effect of CoMo-FM-2 on different sulfides. Reaction condition: 20 mg catalyst, 20 g oil, 500 ppm initial sulfur content, and 100 °C.
Figure 7. (a) Aerobic catalytic performance of CoMo-FMs. Reaction conditions: 20 mg catalyst, 20 g oil, 500 ppm DBT, and 100 °C. (b) Comparison of the effect of aerobic DBT oxidation in different catalytic systems. Reaction condition: 20 mg catalyst, 500 ppm DBT, 100 °C, and 4 h. (c) Catalytic performance of CoMo-FMs at different reaction temperatures in aerobic DBT oxidation. Reaction conditions: 20 mg catalyst, 20 g oil, and 500 ppm DBT (d) Exploring the catalytic effect of CoMo-FM-2 on different sulfides. Reaction condition: 20 mg catalyst, 20 g oil, 500 ppm initial sulfur content, and 100 °C.
Molecules 28 05073 g007
Scheme 1. A proposed mechanism of thiophene oxidation over CoMo-FM.
Scheme 1. A proposed mechanism of thiophene oxidation over CoMo-FM.
Molecules 28 05073 sch001
Figure 8. (a) Fitting of the kinetic data with pseudo-first-order model, and (b) the calculation of activation energy with Arrhenius equation for aerobic DBT oxidation over CoMo-FM-2.
Figure 8. (a) Fitting of the kinetic data with pseudo-first-order model, and (b) the calculation of activation energy with Arrhenius equation for aerobic DBT oxidation over CoMo-FM-2.
Molecules 28 05073 g008
Figure 9. Aerobic catalytic performance of CoMo-FM-2 with different catalyst dosages for the oxidation of DBT. Reaction condition: initial DBT content = 500 ppm, moil = 20 g, T = 100 °C.
Figure 9. Aerobic catalytic performance of CoMo-FM-2 with different catalyst dosages for the oxidation of DBT. Reaction condition: initial DBT content = 500 ppm, moil = 20 g, T = 100 °C.
Molecules 28 05073 g009
Figure 10. (a) Recycling performance of CoMo-FM-2 in aerobic oxidation of DBT (air flow rate of 20 mL/min) (b) FT-IR characterization of fresh and recycled CoMo-FM-2.
Figure 10. (a) Recycling performance of CoMo-FM-2 in aerobic oxidation of DBT (air flow rate of 20 mL/min) (b) FT-IR characterization of fresh and recycled CoMo-FM-2.
Molecules 28 05073 g010
Table 1. Elemental percentage of CoMo-FM determined by ICP-AES.
Table 1. Elemental percentage of CoMo-FM determined by ICP-AES.
SampleElemental Percentage (atm.%)Mo/Co Ratio
CoMo
CoMo-MF-129.22%4.42%0.151
CoMo-MF-225.93%6.87%0.265
CoMo-MF-322.88%8.51%0.372
CoMo-MF-421.36%10.43%0.488
Table 2. Porous texture of CoMo-MFs.
Table 2. Porous texture of CoMo-MFs.
SampleSpecific Pore Volume
/cm3 g−1
Specific Surface Area
/m2 g−1
Mean Pore Size
/nm
CoMo-FM-10.5597.4817.48
CoMo-FM-20.59103.4617.55
CoMo-FM-30.482.9515.82
CoMo-FM-40.3272.8114.01
Table 3. Summarization of chemical states of Mo and O determined by XPS.
Table 3. Summarization of chemical states of Mo and O determined by XPS.
SampleMo 3d5/2O1s
OlatOadWaterad
CoMo-FM-1230.44528.68 (51.34%)529.55 (34.59%)531.01 (14.07%)
CoMo-FM-2230.48528.71 (44.54%)529.57 (33.44%)531.03 (22.02%)
CoMo-FM-3230.61528.89 (48.94%)529.73 (43.67%)531.19 (7.39%)
CoMo-FM-4230.78528.94 (50.43%)529.79 (27.65%)531.25 (21.92%)
Table 4. Comparation of catalytic performances with different catalysts.
Table 4. Comparation of catalytic performances with different catalysts.
EntryCatalystSubstrateOxidantReaction Conditions aConversionRef.
1CoMo-FM-3DBTAir100 °C, 20 mg/20 g, 4 h100%This work
2(NH4)5H6PV8Mo4O40DBTO2100 °C, 20 mg, 6 h88.4%[57]
3r-GODBTO2140 °C, 5 mg/25 mL, 4 h100%[34]
4MoOx/MC-600DBTO2120 °C, 10 mg, 8 h100%[38]
5Ce–Mo–ODBTO2100 °C, 100 mg, 6 h100%[29]
6(Cu-Co)(salen)YDBTO2100 °C, 200 mg, 4 h97.6%[58]
7MFM-300(V)DBTO2120 °C, 0.75 g/L, 5 h99.6%[59]
8ChxNa5-xIMo6O24DBTO2100 °C, 0.1 mmol/6 mL, 5 h100%[60]
9MIL-101DBTO2120 °C, 500 mg/L, 15 h100%[35]
a Reaction conditions: reaction temperature, catalytic dosage/model oil volume, reaction time.
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

Cao, X.; Tong, R.; Wang, J.; Zhang, L.; Wang, Y.; Lou, Y.; Wang, X. Synthesis of Flower-Like Cobalt–Molybdenum Mixed-Oxide Microspheres for Deep Aerobic Oxidative Desulfurization of Fuel. Molecules 2023, 28, 5073. https://doi.org/10.3390/molecules28135073

AMA Style

Cao X, Tong R, Wang J, Zhang L, Wang Y, Lou Y, Wang X. Synthesis of Flower-Like Cobalt–Molybdenum Mixed-Oxide Microspheres for Deep Aerobic Oxidative Desulfurization of Fuel. Molecules. 2023; 28(13):5073. https://doi.org/10.3390/molecules28135073

Chicago/Turabian Style

Cao, Xinxiang, Ruijian Tong, Jingyuan Wang, Lan Zhang, Yulan Wang, Yan Lou, and Xiaomeng Wang. 2023. "Synthesis of Flower-Like Cobalt–Molybdenum Mixed-Oxide Microspheres for Deep Aerobic Oxidative Desulfurization of Fuel" Molecules 28, no. 13: 5073. https://doi.org/10.3390/molecules28135073

Article Metrics

Back to TopTop