Next Article in Journal
Impact of Different π-Bridges on the Photovoltaic Performance of A-D-D′-D-A Small Molecule-Based Donors
Previous Article in Journal
Research on Synthesis, Structure, and Catalytic Performance of Tetranuclear Copper(I) Clusters Supported by 2-Mercaptobenz-zole-Type Ligands
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Ternary Catalyst PW4@MOF-808@SBA-15 for Deep Extraction Oxidation Desulfurization of Model Diesel

by
Yan Gao
1,*,
Shuai Huang
1,
Nina Han
1 and
Jianshe Zhao
2
1
Department of Chemistry, Xinzhou Normal University, Xinzhou 034000, China
2
College of Chemistry & Materials Science, Northwest University, Xi’an 710069, China
*
Author to whom correspondence should be addressed.
Molecules 2024, 29(17), 4230; https://doi.org/10.3390/molecules29174230
Submission received: 31 July 2024 / Revised: 22 August 2024 / Accepted: 4 September 2024 / Published: 6 September 2024

Abstract

:
In this work, a novel heterogeneous catalyst consisting of peroxophosphotungstate, microporous MOF-808, and mesoporous SBA-15 was synthesized, characterized, and used to remove sulfides from model fuel. The prepared material, PW4@MOF-808@SBA-15, exhibits excellent catalytic activity with a desulfurization efficiency of 99.8% in 60 min for multicomponent simulated fuel, and the desulfurization rate can reach more than 90% after ten consecutive cycles. The excellent catalytic activity and reusability are attributed to the hierarchically porous hybrid material MOF-808@SBA-15, which can effectively encapsulate PW4 and provide a site for the oxidation of sulfides.

1. Introduction

With the continuous growth in the global demand for energy, the environmental problems caused by the massive use of traditional petrochemical energy have gradually become prominent [1,2]. Among them, problems such as air pollution and acid rain caused by the sulfur oxides emitted from fuel combustion are particularly prominent [3,4]. In view of the environmental issues caused by the combustion of sulfur-containing compounds in liquid fuels, strict environmental laws have been enacted around the world to limit the sulfur content in fuels, such as below 10 ppm for diesel [5,6]. There are many kinds of sulfides in fuel, which can be divided into two categories: active sulfur species (elemental sulfur, hydrogen sulfide, mercaptans) and inactive sulfur species (aromatic sulfides and heterocyclic sulfides) [7,8]. Among them, the removal of refractory organic sulfur compounds such as dibenzothiophene and its derivatives is challenging. Therefore, it is necessary to find efficient and economical desulfurization methods. Traditional hydrodesulfurization technology (HDS) is a method of removing sulfides from fuel, which converts organic sulfur compounds to hydrogen sulfide and hydrocarbons by reacting with hydrogen under the action of a catalyst [9,10,11]. However, HDS can only effectively remove the active sulfides and part of the inactive sulfides, and reaction conditions are harsh, which led to a burst of research on alternative methods such as extraction oxidation desulfurization technology (EODS) [12,13,14]. Pure extractive desulfurization (EDS) is a method to extract sulfide from fuel oil by using the difference in the solubility of sulfide in specific solvents and fuel oil. Oxidative desulfurization (ODS) can convert the organic sulfide in fuel into a more polar substance, and then the more polar substance is further separated from the fuel by extraction or adsorption to achieve the purpose of reducing the sulfur content. EODS is a combination of extraction and oxidation processes, in which sulfides are first extracted into the extraction phase and then oxidized to sulfoxides and/or sulfones in the presence of an oxidant and a catalyst. The process continues until a low-sulfur or even sulfur-free fuel is obtained [15,16]. EODS can be carried out with an oxidant and without any catalyst but is generally less effective and requires extremely harsh reaction conditions such as high temperature and pressure. Therefore, highly active and environmentally friendly catalysts are needed for the efficient removal of refractory sulfides through EODS. Common oxidants include O2, O3, H2O2, and tert-butyl peroxide, among which H2O2 is the most widely used oxidant during the EODS process because it is cheap, easily available, and environmentally friendly, and has a high active oxygen content [17,18,19,20]. Several metal-containing substances, including metal oxide, polyoxometalates (POMs), metal complexes, and metal–organic frameworks, have been applied to EODS [21,22,23]. In particular, polyoxometalates have become attractive candidates for EODS in recent years due to their simple synthesis, stable structure, and adjustable structure and properties [24]. Polyoxometalates (POMs) with peroxy bonds exhibit higher catalytic activity in EODS than traditional Keggin-type polyoxometalates [25,26]. This is due to the higher electrophilicity of some of the oxygen atoms in peroxopolyoxometalate, which can effectively improve the EODS efficiency. Salete et al. [26] prepared Keggin-type phosphotungstates modified with different cations ([BPy]3PW12, [BMIM]3PW12, and [HDPy]3PW12), which effectively removed several refractory sulfides from fuel using [BMIM]PF6 as the extraction agent. Meanwhile, the formation of peroxy complexes (PWxOy), which play a key role in the catalytic reaction, was verified through 31P NMR. In view of the above conclusions, the authors prepared peroxophosphotungstate (nBu4N)3{PO4[WO(O2)2]4} for fuel desulfurization, and the results showed that the catalyst could achieve a high desulfurization rate under more economical and energy-saving conditions [25]. Shi and co-workers [27] also prepared heterogeneous catalyst SiO2-BisILs [(PW12O40)3−] assembled from peroxophosphotungstate and ionic liquid brush via EODS, and the prepared material showed excellent catalytic activity for dibenzothiophene (DBT). Although POMs show superior activity in EODS, they cannot be separated from fuel due to their high solubility and cannot be reused continuously. Encapsulating POMs in solid materials with a large specific surface area and suitable pore structure is an effective way to solve the above shortcomings.
Metal–organic frameworks (MOFs), a class of ordered crystalline compounds, have become promising contemporary materials due to their tunable chemical functionalities, high specific surface area, and pore volume [28]. The development of new hybrid systems, which can combine the advantages of each component, has been attracting a large number of researchers. Some researchers have tried to combine MOFs with different materials, such as materials with two-dimensional structures (carbon nanotubes, graphene, and boron nitride) or materials with three-dimensional pore structures (MCM-41 and SBA-15), and so on. These MOF-based composites not only combine the advantages of the two materials but also show unique properties. Tapas Kumar Maji and Anindita Chakraborty [29] prepared a hybrid Mg-MOF-74@SBA-15 material through immobilizing Mg-MOF-74 nanocrystals into the mesopores of SBA-15, which has remarkable CO2 adsorption capacity at room temperature due to its unique pore structure. Dirk E. De Vos and co-workers [30] obtained a hybrid catalyst, (Zr)UiO-66(NH2)/SBA-15, by means of “solid-state” crystallization, i.e., the selective growth of (Zr)UiO-66(NH2) nanocrystals in mesoporous SBA-15, and the authors noted that this hybrid material, (Zr)UiO-66(NH2)/SBA-15, has higher catalytic activity and mechanical stability than pure MOFs.
MOF-808, as one of the {Zr6O8}-cluster-based MOFs, has attracted extensive attention due to its high hydrophilic porous surface, good biocompatibility, large cavity, and pore volume. However, the poor stability of MOF-808 has become the main factor restricting its large-scale commercial application. Mesoporous molecular sieve SBA-15 has attracted much attention from researchers due to its highly ordered mesoporous structure, good thermal stability, and high mechanical strength [31]. The hierarchically porous composite constructed by MOF material and mesoporous molecular sieve SBA-15 has a rich chemical environment and stable mechanical properties, which can avoid the disadvantages of MOF materials and provide more possibilities for the application of MOFs [29,30,32].
In this work, a more robust heterogeneous catalyst was constructed by combining peroxophosphotungstate [(C4H9)4N]3{PO4[WO(O2)2]4}, marked as (PW4), and hierarchically porous composite MOF-808@SBA-15 through a simple hydrothermal strategy, which was used for the removal of sulfides from model fuel. The structure of the target product was analyzed by various characterization methods, and then the desulfurization performance was studied using the multicomponent model fuel.

2. Results and Discussion

2.1. Characterization

The crystal structures of the prepared materials were analyzed by powder XRD, as depicted in Figure 1. The obtained pattern for PW4 shows several characteristic peaks at 7–30°. For pure MOF-808, many strong peaks are shown at 2θ values of 4.3, 8.3, and 8.7°, which were assigned to the characteristic planes (111), (311), and (222). Tri-peaks were observed at low angles in the SBA-15 pattern, which were labeled as p6mm hexagonally symmetrical (100), (110), and (200) reflections, confirming the successful preparation of SBA-15. After incorporating PW4, no significant change was found in the characteristic peak, indicating that PW4 was uniformly dispersed in the SBA-15 channel. For PW4@MOF-808@SBA-15, the presence of the same signals from MOF-808 confirmed the successful incorporation of MOF-808 into SBA-15, and the absence of signals from PW4 indicated uniform encapsulation of PW4. Furthermore, the low-angle peaks in the two composites (PW4@SBA-15 and PW4@MOF-808@SBA-15) were similar to those of SBA-15, indicating that the ordered mesoporous pores corresponding to SBA-15 were not destroyed.
The IR spectra of the raw materials (PW4, MOF-808, and SBA-15) and two composites (PW4@SBA-15 and PW4@MOF-808@SBA-15) were recorded and are displayed in Figure 2. For PW4, several intense bands were observed at the range of 849–1083 cm−1, which is attributed to the Venturello structure anion {PO4[WO(O2)2]4}3. For MOF-808, the strong peak at 652 cm−1 is attributed to the stretching vibration of Zr-O, and two absorption bands are shown at 1618 and 1571 cm−1, corresponding to COO- asymmetric stretching. In the case of SBA-15, the strong absorption peaks at 1031, 813, and 460 cm−1 are attributed to the asymmetric and symmetrical tensile vibration and bending patterns of the Si-O-Si frame. The FT-IR spectrum of PW4@SBA-15 shows not only the characteristic peaks belonging to SBA-15 but also some characteristic peaks belonging to PW4. In the FT-IR spectrum of PW4@MOF-808@SBA-15, the same absorptions as those of SBA-15 and MOF-808 were observed, with a slight shift, indicating the successful immobilization of MOF-808, and the peaks attributed to PW4 were not witnessed, indicating the homogeneous dispersion of PW4 within the pore structure.
The trends in the thermogravimetric curves of SBA-15 and PW4@SBA-15 were similar except for the weight loss rate, which verified the presence of the active component PW4 in PW4@SBA-15 (Figure 3). The thermogravimetric curve for PW4@MOF-808@SBA-15 shows a similar profile to that of MOF-808. After the weight loss of water and solvent, two significant weight losses occurred between 258 °C and 590 °C, which correspond to the removal of BTC and Zr clusters.
The porous properties of pure SBA-15, PW4@SBA-15, and PW4@MOF-808@SBA-15 were determined through N2 adsorption–desorption experiments, and the test results are shown in Figure 4 and Table 1. The isotherm of the SBA-15 sample was a typical type IV with an H1-type hysteresis loop, indicating the presence of mesoporous pores. And the pore size distribution of the original SBA-15 was narrow, mainly in the range of 4.1–4.6 nm. The isotherm type of composite PW4@SBA-15 is was same as that of the original SBA-15, indicating that the original pore structure was preserved after loading PW4. The isotherm of PW4@MOF-808@SBA-15 exhibited a combination of type I (characteristic of microporous materials) and type IV (characteristic of mesoporous materials). That is, composite PW4@MOF-808@SBA-15 showed the presence of additional micropores compared to the mesoporous pores of the original SBA-15. Compared with pure SBA-15, the specific surface area and empty volume of PW4@SBA-15 decreased, caused by the occupation of PW4 in the channels. And the specific surface area and empty volume of PW4@MOF-808@SBA-15 had more obvious decreases, indicating that MOF-808 nanocrystals were partially filled in the channels. Although the SEM images (Figure 5) show the presence of few MOF-808 nanocrystals on the surface of SBA-15, the reduced adsorption amount and partial filling of the mesopores of SBA-15 confirmed that the channels of SBA-15 could serve as sites for the growth of MOF-808 nanoparticles [29]. In addition, it was possible to confirm that the channels of SBA-15 were occupied by comparing the size distribution plot.
The morphologies SBA-15, PW4@SBA-15, and PW4@MOF-808@SBA-15 were observed from SEM images (Figure 5). The SEM images show that SBA-15 was an irregular mass, and the morphology of PW4@SBA-15 had no significant change after the introduction of PW4. Although the appearance of PW4@MOF-808@SBA-15 was similar to that of pure SBA-15, the formation of MOF-808 on the surface of SBA-15 could be clearly seen. The chemical composition of PW4@MOF-808@SBA-15 was further evaluated using the EDS technique. The detection of characteristic elements P and W indicated the successful introduction of the active component PW4.
The TEM image of PW4@SBA-15 shows clear areas and darker areas corresponding to empty holes and hole walls, which are similar to those of pure SBA-15 (Figure 6). At the same time, scattered black spots were observed in the pores, indicating that PW4 was evenly distributed in the pores of SBA-15. The TEM image of PW4@MOF-808@SBA-15 shows that the black spots were denser and larger, which may have been because the pore channels were occupied by MOF-808, and the octahedral MOF-808 could be seen on the edge of the sample, which is also consistent with the SEM results. Elemental mapping was used to confirm the homogeneous distribution of guests in the SBA-15 sample. For PW4@SBA-15, the uniform distribution of P and W elements attributed to PW4 indicated that the active components were evenly distributed in the pores (Figure 7). For PW4@MOF-808@SBA-15, the P and W elements attributed to the PW4 and Zr elements attributed to MOF-808 were uniformly distributed, which further indicated that the active component and MOF-808 existed in the SBA-15 channels (Figure 8).
The composition and chemical environment of PW4@SBA-15 and PW4@MOF-808@SBA-15 were further elucidated by XPS (Figure 9). The survey scan of two composites, PW4@SBA-15 and PW4@MOF-808@SBA-15, revealed the presence of W 4f, C 1s, O 1s, N 1s, and P 2p elements, while Zr 3d was detected in PW4@MOF-808@SBA-15. The high-resolution spectra of W 4f for PW4@SBA-15 and PW4@MOF-808@SBA-15 are shown in Figure 9b. For PW4@SBA-15, the binding energies of 37.5 (W 4f5/2) and 35.4 eV (W 4f7/2) were assigned to the W(VI) oxidation state. For PW4@MOF-808@SBA-15, the negative chemical shift (ca. 0.16 eV) indicated the chemical interaction between PW4 and MOF-808 [33]. In addition, a peak of W(CO)x appeared at 31.4 eV, which further proved that the active ingredient PW4 may have been covalently linked to MOF-808 via the -COOH functional group on the organic linker [34]. Additionally, the high resolution of the Zr 3d spectrum of PW4@MOF-808@SBA-15 showed two peaks at 185.1 and 182.7 eV, which were assigned to Zr 3d5/2 and Zr 3d3/2.

2.2. Catalytic Performance

2.2.1. EODS Performance of Different Catalysts

The desulfurization of the simulated diesel containing dibenzothiophene (DBT), 4-methyldibenzothiophene (4-MDBT), and 4,6-dimethyldibenzothiophene (4,6-DMDBT), which are the most common refractory sulfur compounds in liquid fuels, was carried out using MeCN as the extraction agent, H2O2 as the oxidant, and different materials as the catalyst. In the typical desulfurization process, a biphasic system consisting of 0.75 mL of simulated diesel and equal volumes MeCN and 3 umol PW4 or composite containing 3 umol PW4 were strongly stirred for 10 min under 70 ℃ to activate the initial extraction of sulfur compounds from the model diesel to the extraction solvent. Then, H2O2 (0.21 mmol, H2O2/S molar ratio = 6) was poured into the system to start the oxidation stage. Periodically, the upper fuel phase was sampled to analyze the variation in sulfide concentration.
Pure PW4, as a homogeneous catalyst, demonstrated superior desulfurization performance, removing 99.2% of the sulfur in the simulated fuel within 40 min (Figure 10a). Although PW4 showed excellent catalytic activity in the catalytic oxidation of sulfides, it was difficult to recover from the reaction medium, which restricts its development in industrial applications. The heterogeneous catalyst prepared by filling PW4 in the SBA-15 channel for fuel desulfurization effectively improved the recyclability of the catalyst. The kinetic profile of the heterogeneous catalyst PW4@SBA-15 was similar to that of the homogeneous catalyst PW4, but only 98.9% of the sulfides was removed within an extended 100 min. Pure MOF-808 displayed moderate desulfurization activity with a sulfur removal rate of 77.1% in 100 min. Compared with pure MOF-808 and PW4@SBA-15, the ternary catalyst PW4@MOF-808@SBA-15 could remove sulfur compounds from fuel with higher efficiency, which was attributed to the uniform dispersion of PW4 molecules as a single active site and the synergistic action between PW4 and MOF-808. To verify that the active component PW4 was confined to the pore material, a leaching experiment was performed over the PW4@MOF-808@SBA-15 catalyst (Figure 10b). The results showed that the removal rate of sulfide remained basically unchanged after the catalyst filtration, indicating that the active substance was stably confined.
Considering the excellent catalytic activity of PW4@MOF-808@SBA-15 for model diesel, the ternary catalyst PW4@MOF-808@SBA-15 was used to desulfurize commercial diesel with a primary sulfur concentration of 1271 ppm. After treatment under the same conditions, the sulfur concentration of the commercial diesel was reduced to 105 ppm, corresponding to a 91.7% desulfurization rate. The desulfurization efficiency for commercial diesel was lower than that for simulated diesel, which may have been due to the more complex types of sulfides in real diesel.

2.2.2. Reusability

The reusability of both heterogeneous catalysts PW4@SBA-15 and PW4@MOF-808@SBA-15 was measured in order to assess their practical application potential (Figure 11). The solid catalyst remained in the MeCN extraction layer during the reaction process. After a single test, the reused catalyst was centrifuged, washed, and dried, and then added to a new desulfurization system containing fresh simulated diesel, extractant, and oxidizer for the next ECODS cycle. The results showed that the desulfurization activity of PW4@SBA-15 decreased significantly after six consecutive cycles with a sulfide removal rate of 89.5%, which may have been due to the loss of PW4 adsorbed in the SBA-15 channel with the progression of the test. The reusability of the ternary catalyst PW4@MOF-808@SBA-15 was significantly better than that of composite PW4@SBA-15, which may have been due to the presence of MOF-808 inhibiting the loss of active component PW4.

2.3. Comparison with Other Catalysts

A comparison of the removal of organic sulfur substrates from model fuel catalyzed by the ternary catalyst PW4@MOF-808@SBA-15 prepared in this work and POM-based catalysts reported in the literature is given in Table 2 and Figure 12. Compared with similar catalysts (PMo12@MOF-808@SBA-15, SRL-POM@MOF-199@MCM-41, and POM-MOF@Fibercloth), the catalyst PW4@MOF-808@SBA-15 prepared in this work can remove a variety of sulfides from fuel in a shorter time, which is mainly due to the high activity of the active component PW4. Compared with simple composites, the catalyst PW4@MOF-808@SBA-15 has the advantage of high catalytic activity against a variety of sulfides under mild conditions.

2.4. Possible Mechanism

On the basis of the reported studies, a possible mechanism for the removal of organic sulfides from model fuel by EODS technology is proposed. The EODS system consists of an upper fuel phase and a lower extraction phase, and the catalyst PW4@MOF-808@SBA-15 exists in the lower extraction phase. At the initial stage of the reaction, the sulfides in the fuel phase were first extracted into the lower extractant, and the initial extraction rate was in the range of 53.6–60.8% (as shown in Figure 10). Then, the addition of the oxidizing agent H2O2 initiated the catalytic oxidation reaction of the sulfides in the extraction phase. While the sulfides were oxidized to large polar sulfones, more sulfides in the upper fuel phase were continuously transferred to the extraction layer until a low-sulfur or sulfur-free fuel was obtained. During the catalytic oxidation stage, PW4 in the PW4@MOF-808@SBA-15 catalyst combined with H2O2 to form an active species, which is a key intermediate in the oxidation reaction of POM-based catalysts with H2O2 as an oxidizing agent [42]. Under the action of the active species, the sulfides were oxidized to sulfoxides, then further oxidized to sulfones (Figure 13) [43].

3. Experimental Section

3.1. Chemicals and Reagents

All reagents used in this work were purchased from commercial suppliers without further purification. For the synthesis of materials, phosphotungstic acid (H3PMo12O40·nH2O, Bide, Shanghai, China, AR), H2O2 (Sigma-Aldrich, St. Louis, MO, USA, 30 wt%), tetrabutylammonium chloride (Bu4NCl, 99%, Bide), pluronic P123 (Aldrich, St. Louis, MO, USA, AR), tetraethyl orthosilicate (TEOS, Sigma-Aldrich, 98%), zirconium tetrachloride (ZrCl4, Aldrich, 99.5%), benzene tricarboxylic acid (H3BTC, Aldrich, 95%) were used. Dibenzothiophene (DBT, 99%, Aladdin. Bay City, MI, USA), 4-methyldibenzothiophene (4-MDBT, 96%, Aladdin), 4,6-dimethyldibenzothiophene (4,6-DMDBT, 97%, Aladdin), N-octane (99.9%, Aladdin), tetradecane (99%, Aladdin) were used to carry out desulfurization tests.

3.2. Catalyst Preparation

3.2.1. Synthesis of PW4@SBA-15

Peroxophosphotungstate (Bu4N)3{PO4[WO(O2)2]4}, abbreviated as PW4, was obtained according to our previous work [44]. SBA-15 was obtained by the following experimental method: 1.5 g of pluronic P123 was poured into 56.3 mL of 2.4 M HCl aqueous under stirring at 40 °C, and then 3.3 g of TEOS was added to the above solution. The mixed solution was stirred for 24 h at 40 °C and then heated at 100 °C for 24 h. After cooling, the SBA-15 sample was obtained by calcining the isolated white solid for 5 h at 550 °C. PW4@SBA-15 was prepared by the impregnation method. Then, 1.0 g of SBA-15 was added to 5 mL of acetonitrile solution dissolved with 0.50 g of PW4 under magnetic stirring. After 24 h, the solid product, PW4@SBA-15, was collected and dried.

3.2.2. Synthesis of PW4@MOF-808@SBA-15

PW4@MOF-808@SBA-15 was obtained through the solvothermal method [35,36], as follows: 0.50 g of SBA-15, 0.50 g of PW4, 0.51 g of ZrCl4, and 0.16 g of H3BTC were poured into 25 mL of DMF and treated with ultrasound for 30 min. Then, the mixture was transferred to Teflon autoclaves and heated at 130 °C for 24 h. Finally, the solid was collected after cooling, washed with DMF and EtOH three times, and dried at 60 °C overnight; the yield was 63.5%.
Pure MOF-808 was synthesized by the same solvothermal method as PW4@MOF-808@SBA-15, without adding SBA-15 and PW4.

3.3. Catalyst Characterization

The functional group information of different samples was acquired through a Fourier-transform infrared (FT-IR) spectrometer (EQUINOX 55) using the KBr pellet method in the wavenumber range of 4000–400 cm−1 (Bruker, Ettlingen, Germany). The crystal structures of the prepared materials were acquired by powder X-ray diffraction (XRD) on a Bruker D8 Advance (Bruker, Ettlingen, Germany) and Shimadzu XRD-7000S (Shimadzu, Tokyo, Japan) at room temperature. The thermal stability of the prepared materials was analyzed on a NETZSCH STA 449A thermal analyzer (Netzsch-GeräTebau GmbH, Selb, Germany) from 30 to 1000 °C with a heating rate of 5 °C/min under a nitrogen atmosphere. The physical structure parameters of the porous materials were collected by a N2 adsorption–desorption apparatus on a JW-BK 200 at 77 K (CASIO, Tokyo, Japan), and all samples were degassed at 120 °C for 6 h before the test. The morphology and element composition of the samples were examined by a scanning electron microscope (SEM, SU8010, Hitachi, Tokyo, Japan) with electron energy-dispersive spectroscopy (EDX) at 3 kV. The element distribution of the samples was further examined by transmission electron microscopy (TEM) on a Talos F200X instrument (Thermo Fisher Scientific, Waltham, MA, USA). Before the test, about 5 mg of the powder sample was fully dispersed in 1 mL of ethanol under ultrasound and then dropped on the surface of the copper mesh. The surface chemical composition and electronic states of the materials were analyzed by X-ray photoelectron spectroscopy (XPS) on PHI-Veroprobe 5000 Ⅲ spectrometer (ULVAC-PHI, Maozaki, Japan) with an Al Kα X-ray source. The sulfide concentration during the reaction was periodically measured by a gas chromatograph (Agilent 7890, Agilent Technologies, Santa Clara, CA, USA) equipped with an HP-5 type (30 m × 0.25 mm) column; the temperatures of the inlet, detector, and oven were 150 °C, 250 °C, and 150 °C, respectively.

3.4. Evaluation of EODS Efficiency

A given amount of DBT, 4-MDBT, and 4,6-DMDBT was completely dissolved in n-octane to obtain a simulated diesel with a total sulfur content of 1500 ppm. EODS tests were executed in a closed glass container filled with simulated oil, acetonitrile as the extractant, and catalyst under continuous stirring. More precisely, 3 µmol PW4 or composite containing 3 µmol PW4, 0.75 mL of model diesel, and an equal volume of acetonitrile were added to the reactor in sequence and stirred for 10 min under 70 °C to achieve extraction equilibrium. Then, the oxidation process was initiated by adding 0.21 mmol of hydrogen peroxide to the mixture. During the whole process, the upper diesel phase was periodically sampled, and the sulfide content was quantitatively analyzed by the internal standard method for gas chromatography.

3.5. Catalyst Recovery

After the reaction, the mixture was centrifuged for 10 min to recover the solid catalyst. Then, the catalyst was fully washed with octane and acetonitrile and dried overnight at 60 °C. The recovered catalyst was added to the fresh EODS system to re-evaluate its desulfurization performance. Several consecutive tests were carried out to assess its reusability.

4. Conclusions

A novel ternary catalyst, PW4@MOF-808@SBA-15, was successfully prepared through the facile solvothermal method in the presence of mesoporous SBA-15 and active-center PW4 for the deep desulfurization of fuel. The characterization test results showed that the small MOF-808 nanocrystals grew in the pores and surfaces of SBA-15 to form a complex, MOF-808@SBA-15, and PW4 was encapsulated within the pores of MOF-808@SBA-15. The ternary catalyst PW4@MOF-808@SBA-15 showed high desulfurization activity and stability in multicomponent model fuel. This is mainly due to three reasons: the high dispersion of the active-center PW4 unit, the presence of mesoporous channels promoting the diffusion of reactants and products, and the mesoporous silica carrier improving the stability of the material. Although this ternary catalyst, PW4@MOF-808@SBA-15, had an excellent desulfurization effect with a simulated fuel, its application in real fuel needs to be further improved. This new type of ternary composite will open new directions in materials science, as the hybridization of different components will yield more new properties.

Author Contributions

Y.G.: Conceptualization, Methodology, Investigation, Writing—Original Draft, Funding Acquisition. S.H.: Formal Analysis. N.H.: Formal Analysis. J.Z.: Conceptualization. All authors have read and agreed to the published version of the manuscript.

Funding

We appreciate the financial support of this work by the Fundamental Research Program of Shanxi Province (No. 202103021223359), Research Project Supported by Shanxi Scholarship Council of China (No. 2023-164), Scientific and Technological Innovation Programs (STIP) of Higher Education Institutions in Shanxi (No. 2021L453), Xinzhou Normal University Fund (No. 2021KY06), and Xinzhou Normal University PhD startup fund (No. 00001045).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article, further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bian, J.; Zhou, Z.; Yu, J.; Shi, X.; Huang, P.; Du, W. A comprehensive energy flow analysis method based on hydrogen energy to electric energy conversion. J. Phys. Conf. Ser. 2024, 2788, 012004. [Google Scholar] [CrossRef]
  2. Yildiz, I.; Caliskan, H.; Mori, K. Energy, exergy and environmental assessments of biodiesel and diesel fuels for an internal combustion engine using silicon carbide particulate filter. J. Therm. Anal. Calorim. 2021, 145, 739–750. [Google Scholar] [CrossRef]
  3. Li, L.; Li, H.; Peng, L.; Li, Y.; Zhou, Y.; Chai, F.; Mo, Z.; Chen, Z.; Mao, J.; Wang, W. Characterization of precipitation in the background of atmospheric pollutants reduction in Guilin: Temporal variation and source apportionment. J. Environ. Sci. 2020, 98, 1–13. [Google Scholar] [CrossRef] [PubMed]
  4. Lu, C.-W.; Hung, H.-Y.; Sung, H.-C.; Sheu, Y.-S.; Lin, W.-L.; Wu, S.-P. Total sulfur determination in petroleum fuels for routine quality control by sector field inductively coupled plasma mass spectrometry after dilution treatment. J. Anal. At. Spectrom. 2019, 34, 570–576. [Google Scholar] [CrossRef]
  5. Petrova, D.; Lyubimenko, V.; Ivanov, E.; Gushchin, P.; Kolesnikov, I. Energy Basics of Catalytic Hydrodesulfurization of Diesel Fuels. Catalysts 2022, 12, 1301. [Google Scholar] [CrossRef]
  6. de Azambuja, A.O.; Bücker, F.; de Quadros, P.D.; Zhalnina, K.; Dias, R.; Vacaro, B.B.; Correa, C.; Ferrão, M.F.; de Oliveira Camargo, F.A.; Triplett, E.; et al. Microbial community composition in Brazilian stored diesel fuel of varying sulfur content, using high-throughput sequencing. Fuel 2017, 189, 340–349. [Google Scholar] [CrossRef]
  7. Zhang, H.; Chen, L.; Chen, Y.; Wang, Z. Removal of sulfide from fuels by ionic liquids: Prospects for the future. Braz. J. Chem. Eng. 2023, 40, 929–963. [Google Scholar] [CrossRef]
  8. Duan, Z.; Bian, H.; Gao, Z.; Zhu, L.; Xia, D. Green Fuel Desulfurization with β-Cyclodextrin Aqueous Solution for Thiophenic Sulfides by Molecular Inclusion. Energy Fuels 2019, 33, 9690–9701. [Google Scholar] [CrossRef]
  9. Wang, E.; Yang, F.; Song, M.; Chen, G.; Zhang, Q.; Wang, F.; Bing, L.; Wang, G.; Han, D. Recent advances in the unsupported catalysts for the hydrodesulfurization of fuel. Fuel Process. Technol. 2022, 235, 107386. [Google Scholar] [CrossRef]
  10. Zhang, C.; Liu, X.; Liu, T.; Jiang, Z.; Li, C. Optimizing both the CoMo/Al2O3 catalyst and the technology for selectivity enhancement in the hydrodesulfurization of FCC gasoline. Appl. Catal. A: Gen. 2019, 575, 187–197. [Google Scholar] [CrossRef]
  11. Srour, H.; Devers, E.; Mekki-Berrada, A.; Toufaily, J.; Hamieh, T.; Batiot-Dupeyrat, C.; Pinard, L. Regeneration of an aged hydrodesulfurization catalyst: Conventional thermal vs non-thermal plasma technology. Fuel 2021, 306, 121674. [Google Scholar] [CrossRef]
  12. Zaidi, Z.; Gupta, Y.; Gayatri, S.L.; Singh, A. A comprehensive discussion on fuel combustion and desulfurization technologies. Inorg. Chem. Commun. 2023, 154, 110964. [Google Scholar] [CrossRef]
  13. Adhami, M.; Movahedirad, S.; Sobati, M.A. Novel Method for Desulfurization of Mixed Fuel via Microbubble Oxidation Followed by Microtube Extraction. Energy Fuels 2024, 38, 2153–2166. [Google Scholar] [CrossRef]
  14. Zhang, X.; Zhang, J.; Yin, J.; Liu, X.; Qiu, W.; He, J.; Jiang, W.; Zhu, L.; Li, H.; Li, H. Mo-MOF-based ionanofluids for highly efficient extraction coupled catalytic oxidative desulfurization. Sep. Purif. Technol. 2025, 353, 128289. [Google Scholar] [CrossRef]
  15. Sun, S.; Yu, X.; Guo, Y.; Chen, L.; Wang, X.; Jiang, Z. Temperature-Responsive Polyoxometalate Catalysts for DBT Desulfurization in One-Pot Oxidation Combined with Extraction. Catal. Surv. Asia 2016, 20, 98–108. [Google Scholar] [CrossRef]
  16. Mohammed, W.T.; Almilly, R.F.K.; Al-Ali, S.B.A. Desulfurization of Diesel Fuel by Oxidation and Solvent Extraction. J. Eng. 2015, 21, 87–102. [Google Scholar] [CrossRef]
  17. Dou, S.-Y.; Wang, R. High-efficient utilization of gaseous oxidants in oxidative desulfurization based on sulfonated carbon materials. Fuel Process. Technol. 2022, 235, 107372. [Google Scholar] [CrossRef]
  18. Chen, L.; Ren, J.-T.; Wang, H.-Y.; Sun, M.-L.; Yuan, Z.-Y. Engineering a Local Hydrophilic Environment in Fuel Oil for Efficient Oxidative Desulfurization with Minimum H2O2 Oxidant. ACS Catal. 2023, 13, 12125–12133. [Google Scholar] [CrossRef]
  19. Adhami, M.; Movahedirad, S.; Sobati, M.A. Oxidative desulfurization of fuels using gaseous oxidants: A review. J. Sulfur Chem. 2022, 43, 685–712. [Google Scholar] [CrossRef]
  20. Ye, G.; Yang, Z.; Shi, G.; Huang, R.; Liu, X.; Zhang, Q.; Song, R. Synthesis of amorphous titanium isophthalic acid catalyst with hierarchical porosity for efficient oxidative desulfurization of model feed with minimum oxidant. Sep. Purif. Technol. 2025, 354, 128669. [Google Scholar] [CrossRef]
  21. Bertleff, B.; Claußnitzer, J.; Korth, W.; Wasserscheid, P.; Jess, A.; Albert, J. Extraction Coupled Oxidative Desulfurization of Fuels to Sulfate and Water-Soluble Sulfur Compounds Using Polyoxometalate Catalysts and Molecular Oxygen. ACS Sustain. Chem. Eng. 2017, 5, 4110–4118. [Google Scholar] [CrossRef]
  22. Zou, J.; Lin, Y.; Wu, S.; Wu, M.; Yang, C. Construction of bifunctional 3-D ordered mesoporous catalyst for oxidative desulfurization. Sep. Purif. Technol. 2021, 264, 118434. [Google Scholar] [CrossRef]
  23. 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]
  24. Taghizadeh, M.; Mehrvarz, E.; Taghipour, A. Polyoxometalate as an effective catalyst for the oxidative desulfurization of liquid fuels: A critical review. Rev. Chem. Eng. 2020, 36, 831–858. [Google Scholar] [CrossRef]
  25. Julião, D.; Mirante, F.; Ribeiro, S.O.; Gomes, A.C.; Valença, R.; Ribeiro, J.C.; Pillinger, M.; de Castro, B.; Gonçalves, I.S.; Balula, S.S. Deep oxidative desulfurization of diesel fuels using homogeneous and SBA-15-supported peroxophosphotungstate catalysts. Fuel 2019, 241, 616–624. [Google Scholar] [CrossRef]
  26. Ribeiro, S.O.; Duarte, B.; De Castro, B.; Granadeiro, C.M.; Balula, S.S. Improving the Catalytic Performance of Keggin [PW12O40]3− for Oxidative Desulfurization: Ionic Liquids versus SBA-15 Composite. Materials 2018, 11, 1196. [Google Scholar] [CrossRef] [PubMed]
  27. Shi, X.Y.; Sun, M.; Fan, J.; Wang, P.M.; Ma, W.J.; Wei, J.F. Deep oxidative desulfurization of benzothiophene and dibenzothiophene with a peroxophosphotungstate–ionic liquid brush assembly. Appl. Organomet. Chem. 2015, 29, 633–637. [Google Scholar] [CrossRef]
  28. Jayaramulu, K.; Geyer, F.; Schneemann, A.; Kment, Š.; Otyepka, M.; Zboril, R.; Vollmer, D.; Fischer, R.A. Hydrophobic Metal–Organic Frameworks. Adv. Mater. 2019, 31, 1900820. [Google Scholar] [CrossRef]
  29. Chakraborty, A.; Maji, T.K. Mg-MOF-74@SBA-15 hybrids: Synthesis; characterization, and adsorption properties. APL Mater. 2014, 2, 124107. [Google Scholar] [CrossRef]
  30. Cirujano, F.G.; Luz, I.; Soukri, M.; Van, C.; Vankelecom, I.F.J.; Lail, M.; De, D.E. Boosting the Catalytic Performance of Metal–Organic Frameworks for Steroid Transformations by Confinement within a Mesoporous Scaffold. Angew. Chem. Int. Ed. 2017, 56, 13302–13306. [Google Scholar] [CrossRef]
  31. Marcucci, S.M.P.; Zanin, G.M.; Arroyo, P.A. Synthesis of SBA-15 and pore-expanded SBA-15 and surface modification with tin for covalent lipase immobilization. Microporous Mesoporous Mater. 2022, 337, 111951. [Google Scholar] [CrossRef]
  32. Liu, H.; Ding, W.; Lei, S.; Tian, X.; Zhou, F. Selective Adsorption of CH4/N2 on Ni-based MOF/SBA-15 Composite Materials. Nanomaterials 2019, 9, 149. [Google Scholar] [CrossRef] [PubMed]
  33. Tian, P.; He, X.; Li, W.; Zhao, L.; Fang, W.; Chen, H.; Zhang, F.; Zhang, W.; Wang, W. Zr-MOFs based on Keggin-type polyoxometalates for photocatalytic hydrogen production. J. Mater. Sci. 2018, 53, 12016–12029. [Google Scholar] [CrossRef]
  34. Zhang, X.-M.; Zhang, Z.; Zhang, B.; Yang, X.; 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]
  35. Fernandes, S.C.; Leo, P.; Santos-Vieira, I.; de Castro, B.; Cunha-Silva, L.; Balula, S.S. Improving Oxidative Catalytic Efficiency for Fuels Desulfurization Using Hybrid Materials Based in MOF-808@SBA-15. ChemCatChem 2024, 16, e202400355. [Google Scholar] [CrossRef]
  36. Li, S.-W.; Yang, Z.; Gao, R.-M.; Zhang, G.; Zhao, J.-S. Direct synthesis of mesoporous SRL-POM@MOF-199@MCM-41 and its highly catalytic performance for the oxidesulfurization of DBT. Appl. Catal. B: Environ. 2018, 221, 574–583. [Google Scholar] [CrossRef]
  37. Li, S.-W.; Zhang, H.-Y.; Dong, S.-M.; Zhao, J.-S.; Li, R.-X. Highly efficient preformed heteropolyacid catalysts for the deep oxidative desulfurization: MOF as a bridge role under nucleation theory. J. Environ. Chem. Eng. 2022, 10, 107298. [Google Scholar] [CrossRef]
  38. Haruna, A.; Merican, Z.M.A.; Musa, S.G.; Rahman, M.B.A. MOF-808(Zr)-supported with Keggin Polyoxometalates as an efficient oxidative desulfurization catalyst. J. Taiwan Inst. Chem. Eng. 2023, 147, 104919. [Google Scholar] [CrossRef]
  39. Mirante, F.; Ribeiro, S.O.; de Castro, B.; Granadeiro, C.M.; Balula, S.S. Sustainable Desulfurization Processes Catalyzed by Titanium-Polyoxometalate@TM-SBA-15. Top. Catal. 2017, 60, 1140–1150. [Google Scholar] [CrossRef]
  40. Akopyan, A.; Polikarpova, P.; Gul, O.; Anisimov, A.; Karakhanov, E. Catalysts Based on Acidic SBA-15 for Deep Oxidative Desulfurization of Model Fuels. Energy Fuels 2020, 34, 14611–14619. [Google Scholar] [CrossRef]
  41. Zhu, W.; Wu, P.; Chao, Y.; Li, H.; Zou, F.; Xun, S.; Zhu, F.; Zhao, Z. A Novel Reaction-Controlled Foam-Type Polyoxometalate Catalyst for Deep Oxidative Desulfurization of Fuels. Ind. Eng. Chem. Res. 2013, 52, 17399–17406. [Google Scholar] [CrossRef]
  42. Zhu, B.; Yan, L.-K.; Guan, W.; Su, Z.-M. DFT characterization on the mechanism of sulfoxidation with H2O2 catalyzed by tetranuclear peroxotungstates [XO4{WO(O2)2}4]n− (X = SiIV, PV, SVI, AsV, and SeVI). Dalton Trans. 2015, 44, 9063–9070. [Google Scholar] [CrossRef] [PubMed]
  43. Hidalgo, G.; Barozzino-Consiglio, G.; Robeyns, K.; Devillers, M.; Gaigneaux, E.M. Formation; Reactivity, and Catalytic Behavior of a Keggin Polyoxometalate/Bipyridine Hybrid in the Epoxidation of Cyclooctene with H2O2. Inorg. Chem. 2023, 62, 8576–8588. [Google Scholar] [CrossRef]
  44. Gao, Y.; Jia, S.; Zhou, H.; Chen, Z.; Yan, M.; Zhao, J. In situ encapsulation of peroxophosphotungstate in ZIF-8: A highly active, reusable and structurally stable catalyst for desulfurization. Fuel Process. Technol. 2024, 254, 108033. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the obtained materials in 2Θ ranges from 3 to 50° (a) and 0.5 to 3° (b).
Figure 1. XRD patterns of the obtained materials in 2Θ ranges from 3 to 50° (a) and 0.5 to 3° (b).
Molecules 29 04230 g001
Figure 2. FT-IR spectra of the obtained materials.
Figure 2. FT-IR spectra of the obtained materials.
Molecules 29 04230 g002
Figure 3. Thermogravimetric curves of the obtained materials.
Figure 3. Thermogravimetric curves of the obtained materials.
Molecules 29 04230 g003
Figure 4. N2 adsorption–desorption isotherms (a) and pore size distribution (b) of the obtained materials: SBA-15, PW4@SBA-15, and PW4@MOF-808@SBA-15.
Figure 4. N2 adsorption–desorption isotherms (a) and pore size distribution (b) of the obtained materials: SBA-15, PW4@SBA-15, and PW4@MOF-808@SBA-15.
Molecules 29 04230 g004
Figure 5. SEM images of the SBA-15 (a,b), PW4@SBA-15 (c,d), and PW4@MOF-808@SBA-15 (e,f) at different magnifications, and energy-dispersive X-ray spectroscopy (EDS) spectrum of PW4@MOF-808@SBA-15.
Figure 5. SEM images of the SBA-15 (a,b), PW4@SBA-15 (c,d), and PW4@MOF-808@SBA-15 (e,f) at different magnifications, and energy-dispersive X-ray spectroscopy (EDS) spectrum of PW4@MOF-808@SBA-15.
Molecules 29 04230 g005
Figure 6. TEM images of SBA-15 (a), PW4@SBA-15 (b), and PW4@MOF-808@SBA-15 (c).
Figure 6. TEM images of SBA-15 (a), PW4@SBA-15 (b), and PW4@MOF-808@SBA-15 (c).
Molecules 29 04230 g006
Figure 7. Elemental mapping of PW4@SBA-15.
Figure 7. Elemental mapping of PW4@SBA-15.
Molecules 29 04230 g007
Figure 8. Elemental mapping of PW4@MOF-808@SBA-15.
Figure 8. Elemental mapping of PW4@MOF-808@SBA-15.
Molecules 29 04230 g008
Figure 9. XPS spectra representing PW4@SBA-15 and PW4@MOF-808@SBA-15: (a) survey scan, (b) W4f, (c) Zr3d.
Figure 9. XPS spectra representing PW4@SBA-15 and PW4@MOF-808@SBA-15: (a) survey scan, (b) W4f, (c) Zr3d.
Molecules 29 04230 g009
Figure 10. Desulfurization of the multicomponent model diesel catalyzed by different materials (a) and leaching experiments of the catalyst PW4@MOF-808@SBA-15 (b).
Figure 10. Desulfurization of the multicomponent model diesel catalyzed by different materials (a) and leaching experiments of the catalyst PW4@MOF-808@SBA-15 (b).
Molecules 29 04230 g010
Figure 11. Reusability of heterogeneous catalysts PW4@SBA-15 and PW4@MOF-808@SBA-15.
Figure 11. Reusability of heterogeneous catalysts PW4@SBA-15 and PW4@MOF-808@SBA-15.
Molecules 29 04230 g011
Figure 12. Desulfurization activity comparison (A [35], B [36] C [37], D [38] E [39], F [40] G [41]).
Figure 12. Desulfurization activity comparison (A [35], B [36] C [37], D [38] E [39], F [40] G [41]).
Molecules 29 04230 g012
Figure 13. The proposed mechanism of EODS over the PW4@MOF-808@SBA-15 catalyst.
Figure 13. The proposed mechanism of EODS over the PW4@MOF-808@SBA-15 catalyst.
Molecules 29 04230 g013
Table 1. BET surface area, average pore size, and total pore volume of SBA-15, PW4@SBA-15, and PW4@MOF-808@SBA-15.
Table 1. BET surface area, average pore size, and total pore volume of SBA-15, PW4@SBA-15, and PW4@MOF-808@SBA-15.
SampleSBET (m²/g)V (cm3/g)D (nm)
SBA-15647.160.643.86
PW4@SBA-15416.290.474.36
PW4@MOF-808@SBA-15477.050.403.37
Table 2. Comparison of catalytic systems for oxidative desulfurization.
Table 2. Comparison of catalytic systems for oxidative desulfurization.
No.CatalystsOxidantExtractantTemperature (°C)Time (min)Substrate Removal EfficiencyRefs.
1PMo12@MOF-808@SBA-15H2O2MeOH70480DBT, 4-MDBT, 4,6-DMDBT *96.8[35]
2SRL-POM@MOF-199@MCM-41O2-60150DBT100[36]
3POM-MOF@FiberclothO2-9575DBT100[37]
4PW11@MOF-808H2O2MeCN6030DBT100[38]
5(PW11Ti)2OH@TM-SBA-15H2O2MeCN7060BT, DBT, 4-MDBT, 4,6-DMDBT *91%[39]
6HPMo/SBA-15H2O2-60330DBT100[40]
7HPWA-SBA-15t-BuOOH-50120DBT98.4[41]
8PW4@MOF-808@SBA-15H2O2MeCN7060DBT, 4-MDBT, 4,6-DMDBT *99.7This work
* The model fuel contained several sulfides at the same 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

Gao, Y.; Huang, S.; Han, N.; Zhao, J. A Novel Ternary Catalyst PW4@MOF-808@SBA-15 for Deep Extraction Oxidation Desulfurization of Model Diesel. Molecules 2024, 29, 4230. https://doi.org/10.3390/molecules29174230

AMA Style

Gao Y, Huang S, Han N, Zhao J. A Novel Ternary Catalyst PW4@MOF-808@SBA-15 for Deep Extraction Oxidation Desulfurization of Model Diesel. Molecules. 2024; 29(17):4230. https://doi.org/10.3390/molecules29174230

Chicago/Turabian Style

Gao, Yan, Shuai Huang, Nina Han, and Jianshe Zhao. 2024. "A Novel Ternary Catalyst PW4@MOF-808@SBA-15 for Deep Extraction Oxidation Desulfurization of Model Diesel" Molecules 29, no. 17: 4230. https://doi.org/10.3390/molecules29174230

Article Metrics

Back to TopTop