Next Article in Journal
Impact of the Framework Type on the Regeneration of Coked Zeolites by Non-Thermal Plasma in a Fixed Bed Dielectric Barrier Reactor
Next Article in Special Issue
Mesoporous Silica Nanosheets with Tunable Pore Lengths Supporting Metal Nanoparticles for Enhanced Hydrogenation Reactions
Previous Article in Journal
Synthesis of Plasmonic Photocatalysts for Water Splitting
Previous Article in Special Issue
SBA-15 as a Support for Effective Olefin Metathesis Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

SBA-15 Anchored Metal Containing Catalysts in the Oxidative Desulfurization Process

by
Marcello Crucianelli
1,*,
Bruno Mattia Bizzarri
2 and
Raffaele Saladino
2,*
1
Department of Physical and Chemical Sciences, University of L’Aquila, Via Vetoio, I-67100 Coppito (AQ), Italy
2
Department of Ecology and Biology, University of Tuscia, Largo dell’Università, 01100 Viterbo (VT), Italy
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(12), 984; https://doi.org/10.3390/catal9120984
Submission received: 30 October 2019 / Revised: 14 November 2019 / Accepted: 18 November 2019 / Published: 23 November 2019
(This article belongs to the Special Issue SBA-15 and Catalysis)

Abstract

:
Recalcitrant sulfur compounds are common impurities in crude oil. During combustion they produce SOx derivatives that are able to affect the atmospheric ozone layer, increasing the formation of acid rains, and reducing the life of the engine due to corrosion. In the last twenty years, many efforts have been devoted to develop conventional hydrodesulfurization (HDS) procedures, as well as alternative methods, such as selective adsorption, bio-desulfurization, oxidative desulfurization (ODS) under extractive conditions (ECODS), and others. Among them, the oxidative procedures have been usually accomplished by the use of toxic stoichiometric oxidants, namely potassium permanganate, sodium bromate and carboxylic and sulfonic peracids. As an alternative, increasing interest is devoted to selective and economical procedures based upon catalytic methods. Heterogeneous catalysis is of relevance in industrial ODS processes, since it reduces the leaching of active species and favors the recovery and reuse of the catalyst for successive transformations. The heterogenization of different types of high-valent metal transition-based organometallic complexes, able to promote the activation of stoichiometric benign oxidants like peroxides, can be achieved using various solid supports. Many successful cases have been frequently associated with the use of mesoporous silicas that have the advantage of easy surface modification by reaction with organosilanes, facilitating the immobilization of homogeneous catalysts. In this manuscript the application of SBA-15 as efficient support for different active metal species, able to promote the catalytic ODS of either model or real fuels is reviewed, highlighting its beneficial properties such as high surface area, narrow pore size distribution and tunable pore diameter dimensions. Related to this topic, the most relevant advances recently published, will be discussed and critically described.

1. Introduction

Santa Barbara Amorphous (SBA) materials are mesoporous silicate-aluminosilicates characterized by uniform pore size (4.6–30 nm), well-defined pore structure and size-distribution, high surface area, high thermal stability and the capability to support a large panel of active species [1,2]. A wide variety of SBA materials has been reported in the literature like SBA-1 (Pm3n, cubic), SBA-15 (P6mm, hexagonal) and SBA-16 (Im3m, cubic) [3]. SBA-15 is one of the most promising components of this family, due to its several applications, including selective adsorption processes, heterogeneous catalysis, different chemical transformations and gas storage [4]. SBA-15 is usually prepared by supramolecular non-ionic self-assembly between ethylene oxide/propylene oxide copolymer-template (Pluronic P123) and appropriate silica precursors (tetraethoxysilane TEOS and tetramethoxysilane TMOS) in a range of temperature from 30 °C to 120 °C, followed by calcination at high temperature to degrade the copolymer-template (Figure 1) [5,6].
In this process, the value of the ratio of ethylene oxide to propylene oxide can control the distribution function of the pores inside the basic hexagonally-arrayed channel framework. At low values of this ratio, the p6mm hexagonal morphology is favored instead of the cubic mesoporous arrangement prevailing at higher ratio values [7,8]. Alternative reaction pathways for the preparation of SBA-15 derivatives have been also reported [9,10,11]. SBA-15 materials have been well recognized as a promising support template for the synthesis of catalytic materials due to its uniform, hexagonally-arrayed channels with a narrow pore size distribution. These features, together with high surface area and hydrothermal stability, make it as an ideal support for the incorporation of various active molecules on its surface, provided that its mesoporous surface is previously and appropriately functionalized with various functional groups (i.e., amines, thiol, nitriles, halides etc.), grafting them either by co-condensation or by post synthesis grafting procedures [12].
SBA-15 has been applied in a large variety of industrial processes after functionalization with active species on both the bulky silica framework or on the surface of the material [13,14,15]. These processes include: (i) oxidative transformation of hydrocarbons, such as alkanes [16,17], alkenes [18,19] and aromatic derivatives [20,21]; (ii) oxidative transformation of alcohols [22]; (iii) CO oxidation [23]; reduction processes [24] Knoevenagel reactions [25]; waste water treatments [26,27]; and carrier applications in drug-delivery [28,29].
Environmental concerns associated with novel stringent European regulations for the tolerance limits of pollutants in the atmosphere [30] highlighted the interest for the application of SBA-15 materials in the removal of sulfur-containing compounds from petroleum, a process that is indicated as “deep desulfurization” [31]. Sulfur-containing compounds, such as thiols, sulfides and disulfides, produce sulfur oxides (SOx) upon combustion, that act as pollution agents at the global level, inducing acid rain and the formation of fine particulate matter of metal sulfates [32,33,34]. To avoid this problem, reductive hydrodesulfurization (HDS) is a well-stablished technology applied at the industrial level to purify crude fuel oils [35]. As a general procedure, this reaction is performed by using hydrogen in the presence of organometallic transition metal species as heterogeneous catalysts at both high temperature (up to 400 °C) and pressure (up to 100 atm) [36,37,38]. The application of SBA-15 in HDS processes has been reported and reviewed, encompassing the incorporation of Al3+, Ti4+ and Zr4+ cations into the original material to improve the catalyst activity. In the functionalization procedure, the post-synthesis grafting was more effective than the direct synthesis method, leading to a larger pore diameter and the limited diffusion of reactants [39]. However, HDS can yield only limited desulfurization, since refractory sulfur derivatives, such as alkyl-substituted benzothio-phenes (BTs) and dibenzothiophene (DBT), are resistant to reductive treatment [40,41] due to their high alky chain steric hindrance that makes the approach to the catalyst surface difficult [42]. Moreover, the reactivity of thiophene derivatives further decreases in the presence of polyaromatics and nitrogen-containing compounds, which are often components of low quality diesel fuel compounds [43,44].
For this reason, high temperature and pressure, as well as costly hydrogen consumption, are necessary to achieve ultra-low deep desulfurization (that is, to reach a sulfur concentration lower than 10 ppm) [45]. The oxidative desulfurization (ODS) is considered the most promising technology alternative to HDS [46]. In this latter case, the recalcitrant sulfur compounds are oxidized to corresponding sulfoxides and sulfones, followed by distillation, liquid–liquid extraction, or solid–liquid adsorption removal procedures [47]. ODS is accomplished using t-butyl-hydroperoxide (TBHP) [48], superoxides [49], sodium bromate (NaBrO3) [50] and hydrogen peroxide (H2O2), along with mixtures of formic acid with H2O2 [51,52,53,54] as primary oxidants. Hydrogen peroxide is the oxidant of choice for ODS, since it is a low cost, mild and environmentally benign reagent [55]. In this latter case, highly active transition metal derivatives have been selected for the activation of H2O2, often after immobilization on appropriate supports [56,57,58,59], and in alternative reaction solvents [60] [61,62]. Examples on the application of microporous and mesoporous material in ODS procedures have been previously reported [63], focusing on the application of mesoporous TiO2 [64], titanium-containing mesoporous molecular sieves [65] and MCM-41 [66], mesoporous phosphotungstic acid/TiO2 nanocomposites [67], mesoporous TS-1 using a hybrid SiO2-TiO2 xerogel procedure [68,69], vanadium polyoxometalate supported on mesoporous MCM-41[70] and phosphotungstic acid containing ionic liquid immobilized on magnetic mesoporous silica rod systems [71].
In the following sections, we will focus on the application of SBA-15 based catalysts in ODS, classifying them in terms of different families depending on the properties of the prevalent catalytic species as follows: (i) titanium oxides; (ii) vanadium oxides; (iii) molybdenum oxides; (iv) iron oxides; (v) tungsten oxides; (vi) silver oxides; (vii) polyoxometalates and (viii) miscellanea. This choice helps the reader to focus upon the role played by the metal (or metals) species in the oxidation system, highlighting the occurrence of specific synergy events with the SBA-15 support.

2. Titanium Oxides

Titanium dioxide TiO2/SBA-15 materials have been applied as catalytic adsorptive desulfurization systems (CADS) [72], by coupling between the oxidation of recalcitrant sulfur compounds and the selective adsorption of polar products on the SBA phase [73]. As an example, a large panel of TiO2/SBA-15 catalysts, differing in the loading of the active species, was prepared by an application of the facile wetness impregnation technology [74]. In this procedure, SBA-15 was treated with tetrabutyl titanate in ethanol under assisted ultrasound conditions [75]. Irrespective of the loading value, the titanium anatase phase largely prevailed in the samples, the active species being well dispersed on SBA-15 as a consequence of the strong electrostatic adsorption of the TiO2 precursor on SBA-15 with a low point of zero charge [76]. The efficacy of novel TiO2/SBA-15 was evaluated in the removal of DBT, MDBT and DMDBT as selected substrates, by comparing the simple adsorptive desulfurization capacity (ADS) [77] with the properties of the CADS approach (Figure 2). In this latter case, cumene hydroperoxide was used as the primary oxidant in acetonitrile at 35 °C. As a general trend, these CADS were two orders of magnitude more efficient than simple ADS, showing a desulfurization capacity higher than the previously reported UV-photocatalytic CDAS-TiO2/SiO2 procedure [78]. Under optimal experimental conditions, high desulfurization uptakes of 19.3, 12.7 and 7.2 mg-S/g-sorb were obtained. Different interaction sites were operative during these ADS and CADS processes. For CADS, TiO2 was considered as the main adsorption site for recalcitrant sulfur compounds [79], while for ADS, the recognition site was localized in the correspondence of the silanol groups of SBA-15 [80], probably involved in the formation of hydrogen bonds [81] with sulfur oxidized products. Additionally, TiO2/SBA-15 catalysts were used for five consecutive adsorption-regeneration cycle tests by acetone washing of the system followed with oxidative air treatment.
Mesoporous Ti-silica (pore size > 2 nm) showed higher catalytic activity and longer lifetimes than its microporous counterpart (pore size < 2 nm), such as titanium silicalite TS-1, thanks to larger access to the active site and the presence of coordinatively unsaturated tetrahedral titanium able to interact with the primary oxidant [82,83]. In accordance with these data, large cylindrical mesopores Ti-SBA-15 catalysts (pore size > 7 nm) have been prepared by post-grafting insertion of titanium alkoxide/acetylacetone (acac) chelates favoring the dispersion of tetrahedral Ti4+ sites on the mesopore surface of the material, followed by the generation of hydrophobic surface sites by treatment with tetramethyldisilazane (TMDS) to avoid poisoning effects of the polar sulfoxide and sulfone products [84,85]. TEOS was used as the silicon source and Pluronic P123 as the structure-directing agent, working at different temperatures (35 °C, 100 °C and 140 °C) in order to control the size of the mesopores [86,87]. This procedure allows the deposition of highly dispersed Ti4+ sites on the surface of the parent material without any appreciable modification of the original morphology, making easy the accessibility to active species [88]. The titanium contents [Ti/Si molar (atomic) ratio] in the novel Ti-SBA-15 catalysts were included in the range between 0.7% to 4.7%, and the pore size and pore volume increased by increasing the aging temperatures [89,90]. The novel catalysts have been applied in the oxidation of benzothiophene (BT), dibenzothiophene (DBT), 4-methyldibenzo--thiophene (4-MDBT) and 4,6-dimethyldibenzothiophene (4,6-DMDBT) as representative sulfur-containing compounds present in middle distillates [91]. The reaction was performed in an n-heptane/toluene mixture (8:2 wt %) at 80 °C under atmospheric pressure using an excess of cumene hydroperoxide (CHP) (CHP: Sulfur compounds 2:1) as its primary oxidant, to yield the corresponding sulfone derivatives (Figure 3).
All of the novel Ti-SBA-15 catalysts showed largely superior ODS activity compared to Ti-SBA-15 prepared by the conventional impregnation method, as a consequence of the presence of the desired active tetrahedral Ti4+ sites. Irrespective of the nature of the sulfur-containing substrates, quantitative conversion was obtained within half an hour at 80 °C. Similar rate constants were observed in the oxidation of DBT, 4-MDBT and 4,6-DMDBT for each catalyst at a given Ti loading, suggesting that the efficacy of Ti-SBA-15 materials was independent from the actual electron density and steric hindrance of the sulfur substrates. The loading of titanium, as well as the porosity degree of the catalyst, were found to be critical parameters for the activity of the system. The novel Ti-SBA-15 catalysts were also effective in the ODS of a residue hydrodesulfurization (RHDS) diesel sample containing 200 ppmw of refractory sulfur compounds, under fixed bed reactor conditions at 80 °C. The highest reactivity was again obtained for catalysts with the largest titanium loading and pore size.
With the aim to evaluate the effect of nitrogen compounds in the efficacy of the Ti-SBA-15 based ODS procedure, a catalyst produced by tetrabutyl orthotitanate (TBOT) grafting procedure [92] was applied for the oxidation of recalcitrant sulfur compounds in the presence of indole, carbazole and quinoline, as selected compounds commonly found in feed oils [93]. Reactions were performed by treating model sulfur and nitrogen compounds with an excess of tert-butyl hydroperoxide (TBHP) (TBHP/S-compound ratio = 2.5) for 1 h at 80 °C [94]. With respect to the simple ODS treatment, the Ti-SBA-15 activity was found to be drastically decreased in the presence of nitrogen compounds, in the following order: indole > quinoline > carbazole. These data were in accordance with results previously reported for the inhibitory effect of nitrogen compounds on MoO3/Al2O3 [95] and V2O5 catalysts [96] in ODS processes. The mechanism of inhibition may depend on the basic or non-basic properties of the nitrogen compound, the ODS of thiophene being inhibited only by basic conditions, while that of BT being affected by both basic and non-basic nitrogen compounds [97]. Remarkably, the overall ODS activity of Ti-SBA-15 increased in the presence of aromatic and aprotic solvents, such as tetralin, 1-methylnaphthalene and acetonitrile, respectively, due to the detected high solubility of both oxidized sulfur and nitrogen compounds in these reaction media.
As an alternative, trimethylammonium-functionalized SBA-15 polyoxometalate ((PW11Ti)2OH@TMS-SBA-15) has been applied in ODS procedures [98]. Ti(IV) easily substitutes the W(VI) metal center in the Keggin-type POMs, creating multicenter catalytic active sites with corner- or edge-sharing TiO6 octahedra, associated with oligomeric species by the formation of Ti–O–Ti bonds [99,100]. The preparation of the catalyst was performed through an impregnation technology specifically designed for the immobilization of polyoxometalates in SBA-15 materials [101,102], starting from freshly functionalized SBA-15 with N-trimethoxysilypropyl-N,N,N-trimethyl-ammonium chloride (TMS-SBA-15) and (PW11Ti)2OH. Under these experimental conditions a loading of 0.045 mmol/g of (PW11Ti)2OH was obtained, the absence of peaks characteristic for (PW11Ti)2OH in the XRD analysis strongly suggesting that the inorganic anions have been successfully incorporated inside the TMS-SBA-15 channels. The catalyst (PW11Ti)2OH@TMS-SBA-15 was then applied in the oxidation of a mixture of BT, DBT, 4-MDBT and 4,6-DMDBT in n-octane/acetonitrile (total sulfur concentration of 2350 ppm), using an excess of H2O2 as the primary oxidant (H2O2/S-compounds = 1:9) at 70 °C (Figure 4). The reaction proceeded by the transformation of the µ-hydroxo dimeric (PW11Ti)2OH into the active peroxo compound [HPW11O39TiO2]4− [103]. After 2 h, complete desulfurization was achieved. The lower activity observed for (PW11Ti)2OH@TMS-SBA-15 with respect to the homogeneous counterpart (PW11Ti)2OH was probably due to the presence of kinetic barriers for the access of substrate at the heterogeneous active site. The oxidative reactivity order DBT > 4-MDBT ≥ 4,6-DMDBT > BT was in accordance with electronic density effect and with some steric hindrance [104,105]. Moreover, (PW11Ti)2OH@TM-SBA-15 retained its structure and morphology after five consecutive ODS cycles.

3. Vanadium Oxides

The use of vanadosilicates in ODS/ H2O2 processes has been reported to achieve a desulfurization yield higher than 80% [106]. In these reactions, mesoporous vanadosilicates showed higher activity than materials having different pore-size distribution, confirming the beneficial role of large-surface-area, wider pore size and thicker pore walls on the interaction between the active site and the substrate [107,108]. On the basis of these results, vanadium oxide (V2O5) was incorporated into the SBA-15 framework by a controlled grafting process, affording highly dispersed vanadium sites with distorted tetrahedral VO4 units on the silica surface [109]. These catalysts showed significative CODS activity when compared with other V2O5 supports (alumina, titania, ceria and niobia); the highest activity being obtained in the presence of tetravalent cations [110]. As an improvement of this study, vanadium oxide species were incorporated into SBA-15, and modified aluminum Al-SBA-15 and gallium Ga-SBA-15 materials, by both direct-synthesis and impregnation methodology [111].
In the first case, VCl3 was added to the pluronic 123/TEOS mixture (V/Si ratio of 1/30 and 1/54) to yield V-SBA-15, while in the second case, VOx-SBA-15, VOx-Al-SBA-15 and VOx-Ga-SBA-15 were obtained by treating the appropriate SBA-15 starting material with an aqueous solution of VCl3 followed by calcination at 500 °C. As a general trend, the incorporation of Al and Ga improved the dispersion of vanadium species, probably due to a better anchorage of the active species on more acidic supports. Irrespective from the applied procedure, the small angle X-ray diffraction analysis (XRD) showed that SBA-15 retained its mesoporous structure after the incorporation of aluminum and gallium [112], with small clusters of vanadium oxide species being highly dispersed on the surface of the support. However, N2 adsorption–desorption isotherm analysis showed a partial loss of the structure of the catalysts from the impregnation method in the correspondence of the highest value of vanadium loading, as a consequence of the incorporation of vanadium species into the mesoporous channels. In contrast, pore diameter had a marked increase when vanadium was introduced by direct synthesis, suggesting that vanadium atoms in part replaced Si atoms [113]. VOx-SBA-15 efficiently converted DBT to corresponding sulfone in acetonitrile using H2O2 as a primary oxidant, the highest conversion (100%) being obtained with the ratio V/Si = 1/30, corresponding to the prevalence of pseudo-tetrahedral VO43− species on the surface of the catalyst. Samples characterized by the higher loading value of vanadium showed lower activity as a consequence of the formation of less active VOx microcrystal chains. The modification of SBA-15 with Al and Ga further increased the desulfuration activity of VOx-Al-SBA-15 and VOx-Ga-SBA-15 catalysts, probably due to the formation of a larger number of highly dispersed VO43− species, associated with synergic effect of Lewis and Bronsted acid sites, derived from Ga or Al incorporation [114]. On the other hand, lower activity was found when vanadium was incorporated via direct synthesis in V-SBA-15, due to the presence of a kinetic barrier for the interaction between the active sites and DBT. The experimental design optimization of the ODS of DBT using the VOx-Ga-SBA-15 prepared by the impregnation method from Ga(NO3)3 has been more recently reported [115]. The higher levels of the oxidation of DBT were obtained employing the catalyst with 4 wt % of gallium and 6 wt % of vanadium, the optimal ratio in weight between DBT and VOx-Ga-SBA-15 being 4 at the H2O2/DBT molar ratio value of 5.
Examples of the application of VOx-SBA-15, prepared by the impregnation method from ammonium metavanadate (NH4VO3) and SBA-15, using molecular oxygen (O2) as oxidant and the ionic liquid 1-butyl-3-methylimidazolium tetrafluoroborate [Bmim]BF4 as a selective extraction phase, are reported [116]. In this catalytic system, the recalcitrant sulfur compounds (DBT, 4-MDTB and 4,6-DMDTB) are firstly extracted from the oil phase to the ionic liquid in which the VOx-SBA-15 catalyst is suspended. After the extraction step, the oxidation proceeds by vanadium activation of ionic liquid adsorbed molecular oxygen (O2) with formation of the corresponding radical anion derivative and successive oxygen atom transfer to sulfur [117]. The sulfones produced during the oxidation are successively removed from the surface of the catalyst thanks to the high solubility property of the ionic liquid, thus reducing the inhibitory effect that these products normally exert after their adsorption on SBA-15 (Figure 5) [118]. Under the optimal experimental conditions, the removal of DBT, as one of the most stable residues in the hydro-desulfurization (HDS) process, reached up to 99.3% (<3.5 ppm) after 7 h at 120 °C.

4. Molybdenum Oxides

The use of high oxidation states molybdenum Mo(VI) derivatives in the ODS procedure has been reported [119]. For example, MoO3 supported on amorphous silica afforded 82% reduction of the residual sulfur compounds in pre-hydrotreated diesel samples after activation with cumene hydroperoxide at 70 °C [120]. The activity of MoO3/SiO2 further increased after the introduction of different alkaline earth metals (Ca, Ba, Sr and Mg) in the silica phase [121], highlighting the key role played by the Lewis and Bronsted acid sites in the behavior of the oxidation [122]. In order to prepare molybdenum ODS catalysts with larger surface area and larger pore diameter, CoMo/SBA-15 compounds were obtained by impregnating the SBA-15 phase with alcoholic solutions at different concentrations of Co(NO3)2·6H2O and (NH4)6Mo7O24·4H2O, respectively [123]. The XRD analysis confirmed the retention of the mesoporous structure after the impregnation procedure with a slight alteration of the loop shape and pore distribution with respect to original SBA-15, due to percolation of Co and Mo oxides within mesopores, associated with a high dispersion of the metal oxides on the surface of the catalyst. Only Lewis acid sites were found on the surface of the novel catalysts, corresponding to a mixture of different molybdenum species. In particular, the structural analyses showed that the amount of MoO3 microcrystals increased with respect to the β-CoMoMoO4 phase by increasing the overall CoMo concentration [124,125]. CoMo/SBA-15 catalysts were studied in the oxidation of DBT in n-hexadecane, using H2O2 as a primary oxidant under a large panel of reaction temperature (from 60 °C to 80 °C) for 1h at atmospheric pressure. Irrespective of the loading of the catalyst, the activity decreased at higher temperature as a consequence of both H2O2 decomposition and lowering of the amount of the Lewis acid sites. In the optimal catalyst condition (that is, 20% in weight of (Co+Mo) metal oxide loading), a 90% conversion value of DBT was obtained at 60 °C. As evidenced by Raman spectra, the cobalt modified molybdate MoO3 microcrystals were the main active species during the oxidation of DBT, while β-CoMoO4 substantially inhibited the oxidative process.
The role that the defective structure of dispersed MoO3 active sites can play in the ODS procedure was successively evaluated by a deep analysis of the oxidation behavior of 4,6-DMDBT with the MoO3/SBA-15 catalyst [126]. MoO3/SBA-15 was prepared via impregnation of SBA-15 with different amounts of ammonium heptamolybdate (from 5 wt/% to 25 wt/%) as molybdenum precursor. The integrity of the mesoporous structure was found to be dependent from the loading factor, the regular mesoporous feature being observed within to 15 wt/% of Mo, after which the collapse of the structure gradually occurred [127]. XRD analysis of catalysts with 5, 10 and 15 wt/% of MoO3 showed the presence of highly dispersed molybdenum species with an average size smaller than 4 nm. At a higher loading value, orthorombic α-MoO3 crystals, with a certain oxygen deficiency, were formed [128]; the crystallite size being increased by increasing the amount of molybdenum. The oxygen defects are electron-deficient Lewis acid sites and may serve as surface active centers for surface adsorption and reaction. The presence of α-MoO3 crystals in 20 and 25 wt % MoO3/SBA-15 catalysts was further confirmed by Raman and XPS analyses [129]. The novel MoO3/SBA-15 catalysts were evaluated in the oxidation of 4,6-DMDBT with H2O2 in n-hexadecane at different reaction temperatures (from 50 °C to 70 °C) for 1 h, in the presence or in the absence of formic acid as co-oxidant species. Performic acid (HC(O)OOH) is easily produced by reaction between H2O2 and formic acid [130]. As a general trend, the conversion of 4,6-DMDBT was found to be proportional to the number of Lewis acid sites present on the surface of the catalyst during the oxidation performed with H2O2 alone, the highest conversion value being obtained in the correspondence of the higher loading factor (15–20 wt/%). In the presence of formic acid, 4,6–DMDBT oxidation was significantly affected by the formation of surface peroxometallic complex and Lewis acidity. Under the optimal reaction condition using 15 and 20 wt % MoO3/SBA-15 catalysts and the H2O2/formic acid mixture, more than 99% 4,6–DMDBT could be removed at 70 °C within 30 min. Figure 6 describes the suggested mechanism for the oxidation performed with the H2O2/formic acid mixture. In this reaction, scheme performic acid is coordinated by Mo6+ ions to generate a reactive peroxometallic complex [131] able to transfer the oxygen atom to the substrate. In this process, oxygen defective α-MoO3 sites can perform as molecular recognition sites for 4,6-DMDBT. As suggested by the authors, when the amount of MoO3 was greater than 20 wt % (i.e., 25 wt % MoO3), the presence of the largest MoO3 particles lead to lower dispersion and a lower number of exposed Mo ions, thus disfavoring the 4,6-DMDBT adsorption and oxidation.

5. Iron Oxides

Iron-based ionic liquids with redox properties have been applied in ODS processes, thanks to their high reactivity and capability to extract polar sulfone and sulfoxide derivatives from the crude oil [132]. Examples of desulfuration processes mediated by iron-based ionic liquids are reported, including the use of 1-n-butyl-3-methylimidazolium metal chloride derivatives [C4mim] Cl/MCl2 (M = Zn, Fe, Cu, Mg, Sn, Co) [133], dialkylpyridinium tetrachloroferrates ionic liquids [C4MPy] FeCl4 [134] and Fenton-like ionic liquids [135,136]. The activity and recyclability of iron-based ionic liquids was significantly improved by the immobilization of appropriate precursors on SBA-15, as in the case of the preparation of 1-methyl-3-(trimethoxysilylpropyl)-imidazolium [pmim] tetrachloro-ferrate/SBA-15 catalyst, namely [pmim]FeCl4-SBA-15. The synthetic procedure for the synthesis of [pmim]FeCl4-SBA-15 required two successive steps (Figure 7): (i) the grafting of [pmim]Cl on SBA-15 to yield the [pmim]Cl-SBA-15 intermediate and (ii) the ion exchange of chorine with tetrachloroferrate by treatment of [pmim]Cl-SBA-15 with FeCl3 in acetonitrile [137].
The XRD analysis of [pmim] FeCl4-SBA-15 showed a long-range, ordered structure with a well-defined hexagonal lattice (p6mm) motif, associated with the expected functionalization of the mesopore channels, as highlighted by the expected decrease in the intensity of the peaks [138]. The textural properties of SBA-15 were substantially maintained after immobilization of [pmim] Cl and on subsequent anchoring of FeCl4. The [pmim] FeCl4-SBA-15 was applied in the ODS of a model fuel sample obtained by dissolving DBT, BT and DT in n-octane, using H2O2 as the primary oxidant. The concentration of sulfur compounds in the model fuel decreased in the order of DBT > BT > DT, with an overall 94.3% removal of DBT in the optimal conditions, consisting in the presence of [Omim] BF4 as a co-solvent. As suggested by the authors, the reaction proceeded with the initial extraction of sulfur compounds from the oil, followed by a Fenton-like oxidation process.
As an alternative, the Fe-SBA-15 catalyst was prepared by an impregnation procedure, using Fe(NO3)3·9H2O as source of iron [139]. Fe-SBA-15 showed well-resolved XRD diffraction peaks, confirming the p6mm hexagonal symmetry of mesostructured SBA-15. The decrease of the peaks’ intensity suggested a slight distortion in the material, with a partial diminution of the pore’s diameter, due to the lower ionic radio of Fe with respect to Si [140,141,142]. In particular, Fe-SBA-15 contained a well-dispersed hematite phase (α-Fe2O3) without any presence of the magnetite phase (Fe3O4).
To evaluate the ODS capacity, Fe-SBA-15 was applied in the oxidation of BT, DBT and 4,6-DMDBT, in a three-phase system (n-dodecane, acetonitrile and catalyst), with H2O2 as primary oxidant at 60 °C. Under these experimental conditions, more than 90% of sulfur was removed in the first 15 min of reaction time, DBT showing the higher reactivity as a consequence of its recognized high electron density on the sulfur atom [143]. Note that Fe-SBA-15 was less reactive than Fe-MCM-48 under similar experimental conditions, probably due to the difference in surface area, pore volume and dispersion of the active sites. Ultrasound-assisted technology improved the efficacy of iron-based SBA-15 catalysts. For example, Fe-SBA-15 and Fe/Zr-SBA-15 catalysts with high ODS capacity have been prepared from the corresponding mineral supports by an ultrasound-assisted procedure, using Fe(NO3)3·9H2O as the iron source [144]. Irrespective of experimental conditions, spectroscopic data indicated the presence of a well dispersed hematite phase (in both α-Fe2O3 and γ-Fe2O3 forms) as an active species. Both catalysts showed the presence of Lewis acid sites, which amount increased by increasing the iron loading. The catalytic activity of Fe-SBA-15 and Fe/Zr-SBA-15 was evaluated in the oxidation of DBT in n-hexadecane and H2O2 as the primary oxidant. As a general trend, the higher activity was observed in the presence of the highest value of loading of Fe (30 wt/%), Fe/Zr-SBA-15 being more active than Fe-SBA-15, probably as a consequence of the role played by the zirconium-increased surface acidity in the promotion of γ-Fe2O3 with respect to the formation of amorphous, and less reactive, iron oxide.

6. Tungsten Oxides

Nitrogen functionalized active materials have been successfully applied in heterogeneous catalysis [145,146], including the oxidative transformations of sulfur derivatives [147,148,149], thanks to the high coordinative capability of the nitrogen-containing framework towards metal species [150,151], associated with tunable acid-base properties. With the aim of obtaining a solid catalyst, to be employed in the ODS of gasoline, with large specific surface and high catalytic activity, WO3 was introduced, for the first time, into the SBA-15 mesoporous molecular sieve, under conventional hydrothermal conditions in strong acidic solution, using H2WO4 as a tungsten source in the presence of P123 triblock copolymer and cetyltrimethylammonium bromide as template. This system showed a promising oxidation activity, being able to reduce the sulfur content of gasoline (540 ppm) up to 91% in 80 min [152]. Examples of the preparation and application of tungstate nitrogen-containing catalysts are reported [153], nitride derivative being used in the oxidation of model sulfur compounds under visible light [154]. With the aim of further improving green applications of tungstate nitrogen-containing catalysts, WOx/N-SBA-15 materials have been prepared and used in the ODS procedure [155]. WOx-SBA-15 was prepared as starting material by direct synthesis in the presence of TEOS as the silica source and P123 as our structure-directing agent, using NaWO4·2H2O as tungstate source. Nitrogen rich WOx/N-SBA-15 was successively obtained by treating WOx-SBA-15 with a mixture of ethylene diamine NH2C2H4NH2/CCl4 at reflux, followed by calcination at high temperature (600 and 800 °C). The novel catalyst showed the presence of well-dispersed WOx species in a nanocluster form, and was characterized by a channel-like pore arrangement of the 2D hexagonal (honey comb) type. The presence of nitrogen-containing CNx-like frameworks was confirmed by both Raman and X-ray photoelectron (XP) spectra, in association with graphite-like carbon originated by the ethylene diamine degradation. The CNx framework was characterized by the presence of pyridine-nitrogen-like atoms (N-py), nitrile moieties and a C–N–C graphitic-like ring motif, suggesting a complex rearrangement of nitrogen and carbon atoms during the preparation of the catalyst. WOx/N-SBA-15 was a catalyst more efficient than WOx-SBA-15 in the ODS procedure, suggesting a beneficial role of the CNx framework in the oxidation. Under optimal experimental conditions, the quantitative conversion of DBT was obtained with WOx/N-SBA-15 after 2 h at 100 °C, using H2O2 as primary the oxidant in acetonitrile. Moreover, the activity of the catalyst increased with the increase of the loading of the WOx species. Most probably, the reaction proceeded by the formation of highly reactive tungstate-peroxo complexes [156], even if the authors have not investigated the possible role of the CNx framework in this process.

7. Silver Oxides

Different supported silver adsorbents have been reported in the selective removal of sulfur compounds from fuels via a strong complexation mechanism, including Ag2O/titania materials in jet fuel desulfurization [157,158], Ag+/SBA-15 and Ag+/SiO2 for the selective adsorption of dibenzothiophene (DBT) [159], and Ag+ supported zeolite and activated carbon for the adsorption of both BT and DBT [160,161]. Moreover, supported silver oxides or silver can also act as oxidative catalysts in the epoxidation of alkenes [162,163,164], aldehydes [165], and various organic compounds [166] and carbon oxide [167]. The capability of silver oxides to contemporarily perform as active catalytic species and adsorption phases has been applied in the ODS procedure by preparation of novel AgxO-SBA-15 materials and their use in the adsorption/oxidation of BT, DBT, 4-MDBT and 4,6-DMDBT, as model fuel compounds [168]. AgxO-SBA-15 catalysts were prepared by an ultrasound-assisted impregnation process of SBA-15 and AgNO3 water solutions at different concentration. Irrespective from the loading factor, the catalysts retained the original mesoporosity of SBA-15 showing AgxO species uniformly dispersed in the hexagonally-ordered network (average diameter value of the particle around 5–6 nm), the presence of larger AgxO particles being detected only at the highest loading factor (25 wt/%). In the XRD analysis, AgO particles largely prevailed in the fresh sample, while Ag2O and Ag become the main silver species in the aged sample, as a consequence of degradation processes occurring during the time for unstable high valence AgO oxide at the nano-size scale [169,170]. AgxO-SBA-15-aged samples showed a high desulfurization activity in the presence of air under ambient conditions. The addition of air dramatically enhanced the activity with respect to the simple adsorption processes. On the other hand, freshly prepared AgxO-SBA-15 catalysts showed a remarkable desulfurization capacity even in the absence of air. In this latter case the desulfurization capacity followed the order of 4,6-DMDBT > 4-MDBT > DBT > BT, consistent with the order of sulfur electron density on the substrate. Irrespective of the origin of the catalysts, sulfones were detected as reaction products in the presence of air. Noteworthy, sulfones were also detected with freshly prepared AgxO-SBA-15 catalyst in the absence of air, suggesting that this catalyst may also perform as an endogenous source of oxygen for the oxidation. The general reaction mechanism for fresh and aged AgxO-SBA-15 catalysts is described in Figure 8.
Surface-distributed, nano-size [O]-retaining AgO active sites are responsible for the oxygen atom transfer from freshly prepared AgxO-SBA-15 catalyst to substrate in the absence of air (Panel A). A different mechanism occurs with aged AgxO-SBA-15 in the presence of air (Panel B), in which case Ag2O and metallic Ag become the main active species for the direct activation of dioxygen [171]. During the aging AgO is progressively transformed into Ag2O.

8. Polyoxometalates

Polyoxometalates (POMs) are formed by cations and polyanion clusters having structural diversity (Figure 9), where the basic construction units are based upon the MOx (x = 5, 6) oxometal polyhedral [172,173,174]. The M species are early transition metals (e.g., V, Nb, Ta, Mo, W, and so on) at their highest oxidation states. Some polyanions are centered by heteroatoms (X) that significantly affect their properties. These heteroatoms are usually main-group elements, including Si, P, S, Ge, As, Se, B, Al and Ga, but are not limited to these. The bulky polyanions have a highly negative charge, and their surfaces are rich of oxygen atoms able to donate electrons; consequently, they can be considered as soft bases. At the same time, the metal ions on the skeleton of polyanions possess unoccupied orbitals, able to accept electrons. By this way, polyanions can also act as Lewis acids. Hence, POMs may play the roles of Lewis acid and Lewis base, under different conditions. Generally speaking, the protonation of an oxometallate ion under particular conditions (e.g., pH, concentration, temperature, solvent) gives rise to the polycondensation of the tetrahedral MO42 units and the formation of more complex structures called polyanions:
nMO42 + 2mH+ ⇌ MnO(4n−m)(2n−m) + mH2O
If the condensation occurs between similar species, the reaction yields an octahedral isopolyanion (IPA) having the general formula MnO(4n−m)(2n−m) (usually n 6). Conversely, if the condensation of the oxoanions occurs around either, a central heteroatom (e.g., X = Si, P, As, Ge, etc.) or another metal atom, the reaction leads to the formation of a heteropolyanion (HPA) with the general formula of XsMnOmy. Depending on experimental assembling conditions, POMs are formed by IPAs and HPAs, having two main structures, namely (i) Keggin (X/M = 1/12) and (ii) Dawson (X/M = 2/18) type structures. The many different elements which can act as heteroatoms in HPA complexes with various coordination numbers leads to the formation of several other complex polyhedra structures known as Anderson, Lindqvist, Waugh and Silverton (Figure 9) [175].
The Keggin type HPAs containing tungsten or molybdenum addenda atoms have received great attention for the very variable chemical properties, which may be easily tuned through changes in their composition and structure. The replacement of their protons by large radium cations (i.e., Cs+) makes insoluble the Keggin HPAs and increases their surface area and thermal stability [176]. On the other hand, the removal of tungsten or molybdenum atoms from the parent structure generates vacancies (Lacunar Keggin HPA), which can give them alternative catalytic activity [177].
In spite of an overgrowing interest toward the preparation of a huge number of inorganic–organic hybrid architectures based on POMs, due to their extreme versatility of compositions, high structural stability and special properties like strong acidity and rich redox chemistry, at the same time these compounds frequently show several drawbacks. They have, indeed, relatively small specific surface areas that hinder accessibility to their active sites, thus reducing their catalytic activity; in addition, their high solubility in polar solvents prevents their simple recovery from complex mixtures. To avoid such disadvantages and to meet the requirement of sustainable chemistry, great effort has been focused on immobilization of the active POMs on different solid supports, such as silica with high-surface areas, activated carbon, molecular sieves and graphene or graphene oxide, to improve their dispersion and heterogeneity [178,179,180]. For all the above reasons, it is reasonable that the POMs compounds have received considerable interest for what concerns their use for demanding catalytic ODS processes [181], and in particular, as immobilized catalysts on mesoporous SBA-15 materials.
Hereafter, the main results obtained in the last years in the oxidative desulfurization of authentic or model fuels promoted by heterogeneous catalysts based on POMs derivatives, will be reviewed grouping the results into two main sub-categories formed, respectively, by phosphomolybdic or phosphotungstic containing POMs.

8.1. Phosphomolybdic Containing POMs

Starting about fifteen years ago, a novel phosphomolybdic (HPMo)/SiO2 mesoporous composite was prepared by the sol–gel method, working with 12-phosphomolybdic acid, TEOS and the triblock copolymer EO20PO70EO20 (Pluronic P123) as a template [182]. Authors declared that in this compound, the HPMo was highly dispersed inside the silica framework, while maintaining its original Keggin structure. The heterogeneous catalytic system was active and stable toward the ODS of a model fuel only formed by DBT, with H2O2 as oxidant in a large excess (O/S molar ratio = 12). Some of the same authors published another paper where the phosphomolybdic (HPMo) and phosphotungstic (HPW) heteropolyacids, were chemically anchored onto the amino functionalized SBA-15 channels, with the aim to study their catalytic activity for the ODS of model fuels [183]. The SBA-15 functionalization with aminosilane groups, through aminopropyltriethoxysilane (APTES) [184], is aimed to increase the electrostatic interactions between the amino moiety of the support and the heteropolyacid groups, thus increasing the chemical stability of the heterogeneous catalyst. Interestingly, by comparing the two catalytic systems, it was evident that the catalyst HPW/H2N-SBA-15 was slightly more active in comparison with HPMo/H2N-SBA-15, justifying these experimental data with the conversion of the original Keggin structure of starting HPA (with its original redox potential) into a new polyoxoperoxo species, after the oxidation with H2O2. On the other hand, it is known that the ability of tungsten atoms to combine with TBHP is higher than molybdenum atoms. Thus with increasing the tungsten loading, the ability of HPAs to produce more active peroxometalate intermediates becomes stronger, also in the presence of TBHP as main oxidant.
An interesting method for the deep desulfurization of model fuels and exploitable as an alternative to severe HDS, indicated that the condensation reaction with formaldehyde catalyzed by phosphomolybdic acid inside selected supports such as SBA-15, acting both as support and adsorbent of the reaction products, was effective for the specific removal of thiophenic and benzothiophenic compounds, while preserving the fuel quality [185]. Indeed, as assessed by authors, the phosphomolybdic acid promoting the condensation reaction with formaldehyde, affording polymerization products, can occur only with thiophenic or benzothiophenic compounds, thus preserving olefins or the other aromatic components of fuels. The coupling with an oxidant as peracetic acid allowed a strong reduction in the total sulfur content lower than 15 ppm.
Frequently, the phosphomolybdic-based HPA has been modified with the addition of other metals in order to tune the catalytic activity of the corresponding POMs, according to specific requirements. This is the case of an heterogeneous catalyst based on a molybdovanadophosphoric acid framework supported on both pure SBA-15 or H2N-SBA-15, after the silica functionalization with APTES [186].
It was observed that, in spite of a comparable high activity in the oxidation of sulfur compounds of a model oil (DBT in n-hexane), the catalyst PMoV2/SBA-15-NH2 formed through grafting of molybdovanadophosphoric acid onto H2N-SBA-15, showed good recyclability in comparison of the catalyst PMoV2/SBA-15 impregnated on unmodified SBA-15 silica. The higher stability against leaching of the grafted catalyst, has been attributed to the strong electrostatic binding between cation and anion due to the in situ formation of the salt ≡ Si(CH2)NH3·PMoV2, not possible in the unmodified silica, where only weak interactions between silanol groups and high soluble PMoV2 moiety, can occur. In another case, the stability of immobilized molybdovanadophosphoric acid was increased by using zirconium modified mesoporous SBA-15 able to form insoluble salts of HPAs on the silica surface. The 11-Molybdo-vanadophosphoric acid supported on Zr-modified silica (MoV/Zr/SBA-15) worked as an active and stable catalyst for the TBHP promoted ODS of model oil [187]. It is known that, in particular cases, the length and pore size of mesoporous silica channels imply both crucial factors and effectiveness in fluid phase reactions [188]. As a consequence, channels with the larger pore size and shorter length can overcome limitations like mass transfer, accessibility and diffusion in the liquid phase reactions. For this reason platelet SBA-15, characterized by shorter channels parallel to its thickness, has been considered more promising to serve as a support for catalytic systems. Accordingly, cesium salts of tungsten-substituted molybdophosphoric acid, CsxH3−x [PMo12−yWyO40], (x = 1–3, y = 2–10), supported on platelet SBA-15 have been synthesized, via a two-step impregnation method, and used as oxidative desulfurization catalysts, using TBHP as oxidant [189]. It is interesting to note that for the selected H3PMo8W4O40 HPA, loaded with different amounts of cesium ion, the observed initial catalytic performances followed the order Cs2Mo8W4/SBA > Cs1Mo8W4/SBA > Cs3Mo8W4/SBA. As known, HPA supported on mesoporous silica shows strong Bronsted acidity (higher Bronsted acid sites vs Lewis acid sites), being the Lewis acidity increased by replacing of H+ with Cs+ cations. It has been clarified [190] that both kinds of acid sites may be present in Cs1Mo8W4/SBA and Cs2Mo8W4/SBA species, even if the number of Lewis acid sites of the former catalyst is lower than that of the latter, and being the Bronsted acidity irrelevant for Cs3Mo8W4/SBA, due to the absence of H+ ions. As well accepted, the oxidation of sulfur substrates activated by molybdenum- or tungsten-polyoxometalate catalyst implies the formation of active peroxo species at the metal site, followed by transferring the oxygen atom to the sulfur acceptor (Figure 10).
From literature, it is known that the formation of peroxo-metallate species, is energetically favored by the combined presence of Lewis and Bronsted acid sites [191]. As a consequence, it has been supposed that the presence of both Lewis and Bronsted sites on the Cs2Mo8W4/SBA catalyst, facilitates the formation of peroxo-HPA intermediates, thus increasing the oxidation reaction efficiency.
Among several advantages on the use of mesoporous SBA-15 as versatile support for heterogeneous catalysis, one of the main drawbacks may be found on its hydrophilic nature which could be a detrimental aspect if we are dealing with the catalysis of reactions containing both aqueous phase and hydrophobic starting materials, as in the case of ODS of authentic fuels. In some cases, it has been reported that the hydrophobization of mesoporous supports could markedly enhance the stability and performance of the catalyst system. To achieve this goal, the most explored approach was based on the functionalization of SBA-15 with imidazole-based ionic liquids (ILs). Noteworthy, due to the introduction of IL, the heterogeneous catalyst exhibited good wettability for the model oil, which could provide easier access to catalytically active sites, for reagents. As a well explored general method, the doping of SBA-15 by IL was based firstly on the preparation of the appropriate precursor 1-Methyl-3-(triethoxysilylpropyl)-imidazolium chloride ([pmim]Cl) ionic liquid, by using N-methyl imidazolium and 3-chloropropyltriethoxysilan, and then by grafting the latter IL on SBA-15, under usual conditions. At this point, several different HPAs may be immobilized affording several different types of hybrid functional materials (Figure 11) depending on the nature of metals (M, M’) and heteroatoms (X).
By this way, several authors published the preparation of heterogeneous PMOs catalysts, ranging from HPMoV2 [192], to HPMo [193] based materials, all following the same immobilization procedure over the functionalized IL-SBA-15 silica. These hybrid catalysts displayed high catalytic performances for removing sulfur compounds from model oil (especially in the case of HPMo system, where the H2O2 oxidant was used in the lowest amount) and high stability and recyclability properties, acting both as a catalyst and as an adsorbent toward the sulfur oxidized products. Interestingly, XRD analyses confirmed that the ordered structure of SBA-15 is maintained after the IL grafting and the successive HPA loading. Concerning the latter point, an optimal value should be carefully chosen in order to avoid that IL-SBA-15 channels can be blocked by the HPA overloading, thus reducing the surface area and consequently the catalytic activity.

8.2. Phosphotungstic Containing POMs

Since the last fifteen years, several articles reported the preparation of heterogeneous catalysts based on the simple wet impregnation of variable amounts of phosphotungstic acid over mesoporous SBA-15 or mixed HY-SBA-15 composite zeolite [194], with the aim to study their catalytic activity in the ODS process applied either to model fuels [195] or real feeds [196] with H2O2, TBHP o air [197] as main oxidants. In general, these catalytic systems showed high activity and stability, being possible to reuse the heterogeneous catalysts after the regeneration step. Supports based on commercial silica showed worst results in comparison with mesoporous SBA-15, with the same tungsten loading, thus evidencing the positive role on the use of a mesoporous support allowing on the one hand the presence of well-dispersed tungsten species and on the other reducing the possible catalyst deactivation caused by sulfone deposition during oxidation. Nevertheless, some problems remain in SBA-15-supported catalysts being possible that phosphotungstic acids easily leached into a solution because of the weak interaction between them and SBA-15. Therefore, the search for catalysts with high activity and good recovery remains essential. From this point of view, the creativity of the researchers to find innovative solutions, ranged from the preparation of micro-mesoporous composite molecular sieves (ZSM-5/SBA-15) doped with different metal ions as Zr, Ag, Ce, thus balancing advantages/disadvantages of both types of supports [198] to the immobilization of phosphotungstic acids (HPW) onto composite supports as pentaethylenehexaamine (PEHA)-preloaded acidic Zr/SBA-15 with short pore channels [199]. In the latter, the synergic role of the zirconium-ion-modified acidic SBA-15 and the auxiliary PEHA molecules toward the high dispersion and stabilization of HPW clusters is believed to endow the resulting nanocomposites with high reusability in comparison to the zirconium free materials. A deep investigation on the potential effects of the preparation methods of mesoporous HPW/SBA-15 catalysts on their catalytic performances on the ODS of a model oil, has been reported [200]. In this work, authors focused their study on two kind of catalysts prepared through silica aminofunctionalization (AF) or evaporation-induced self-assembly (EISA) methods, respectively (Figure 12). The study revealed that the catalyst prepared by EISA method gave better results, in terms of higher catalytic activity and stability in comparison with that obtained by AF procedure. The reasons may be found on the fact that, according to structural and spectroscopic analyses, in the first case the catalyst maintain the Keggin structure of original HPW molecules and has a high surface area; in the second strategy, the mesostructure of SBA-15 resulted partially blocked after the introduction of HPW molecules, and the Keggin structure of HPW molecules was damaged.
Anyway, the easy modification of native SBA-15 surface through a post-grafting procedure, in order to introduce more robust functional group (i.e., APTES to give H2N-SBA-15 or N-(3-trimethoxysilylpropyl) tributylammonium to give tba-SBA-15) for the anchoring of different types of HPAs, justifies the large number of published articles appeared in literature, in the last decade. Among them, H2N-SBA-15 and tba-SBA-15 supports were used to heterogenize, through an impregnation method, Keggin-type derivatives like lacunar polyanions [PW11O39]7− (PW11), resulting by the removal of one WO4+ unit, or sandwich-type [Eu(PW11O39)2]11− anion affording robust and active catalysts for the ODS of model oil and real diesel, either under biphasic (diesel/acetonitrile 1:1) or solvent-free conditions [201,202]. Noteworthy, the PW11/H2N-SBA-15 composite system showed to maintain its activity for eight consecutive cycles. In another case, the amino functionalization of SBA-15 has been obtained either by one-step synthesis (co-condensation method, that is introducing the APTES into mesoporous silica in the presence of copolymer Pluronic P123) [203] or by the classical post-grafting procedure [204], having in mind the heterogenization, in both cases, of the Keggin-type 12-tungstophosphoric [PW12O40]3 HPA. In this case, the heterogeneous catalysts have been only used for the study of model oils with H2O2.
Composite catalysts based on tricopper(II)-substituted sandwich-type polyoxotungstates, having the general formula (Cu3X2W18, X = BiIII, SbIII) supported over H2N-SBA-15, have been prepared to be used in the H2O2 promoted ODS of thiophene [205]. Authors suggested that the best detected loading amount of Cu3X2W18 species over the mesoporous silica was up to 4.7%, while worst activities were observed for higher levels of active species, as a result of the decreasing of surface area and pore diameters, thus hindering the flow of the reaction substrates into the pore channel of H2N-SBA-15. No relevant differences, on the catalytic activity, have been detected for the two different polyoxotungstates containing BiIII or SbIII, respectively. In a recent paper, a contribution based on the use of a composite material formed by a phosphorous-containing tetranuclear peroxotungstate {PO4[WO2(O2)2]4}3 Venturello anion symbolized as PW4, immobilized on trimethylammonium functionalized SBA-15 (PW4/tma-SBA-15) through an impregnation method, employed as catalyst for the ODS of model and real fuels, has been published [206]. Impressive data concerning activity and stability of heterogenized POM, have been reported by authors; moreover, carefully investigations on the fresh heterogeneous catalyst and on the material recovered after the successive recycling steps, showed that during the heterogenization and oxidation experiments an equilibrium between the two species PW4 and the α-Keggin heteropolyanion [PW12O40]3 (PW12), cannot be ruled out.
Accordingly to what above reported for the immobilization of phosphomolybdic based POMs, several authors published examples describing the preparation of phosphotungstic acid immobilized over ionic liquid-modified SBA-15 support too, having in mind to enhance the stability and the performances of the catalyst system. An interesting paper described the preparation of a series of hybrid polyoxometalates formed by the same anion α-Keggin phosphotungstate PW12 and three different ionic liquid cations, namely 1-butyl-3-methylimidazolium (BMIM), 1-butylpyridinium (BPy) and hexadecylpyridinium (HDPy), having the general formula [IL]3PW12 [207]. The activity and stability of these catalysts toward the ODS of model diesel under extractive conditions (ECODS), that is using [BMIM]PF6 or acetonitrile as extraction solvent, have been studied in comparison with the phosphotungstate PW12 anion immobilized over the functionalized mesoporous tma-SBA-15 (see above). The latter (PW12/tma-SBA-15) showed to be the most stable composite system having similar catalytic performances with the three homogeneous [IL]3PW12 catalysts and comparable activity both in acetonitrile and [BMIM]PF6, even if with high recycling capacity. On the other hand, the classical ionic liquid derivatization of SBA-15, based on the preparation of the appropriate precursor 1-Methyl-3-(triethoxysilylpropyl)-imidazolium chloride ([pmim]Cl) ionic liquid and on the next grafting of the latter on SBA-15, allowed the anchoring of different HPA, ranging from the silicotungstic acid H4[SiO4(W3O9)4] [208] to the classical phosphotungstic PW12 anion [209]. In both cases, irrespective to the specific type of polyoxometalate composition, the anchoring on the IL functionalized support gave a system with higher catalytic activity for the H2O2 promoted ODS of model oils in comparison to the system obtained by unmodified SBA-15. As outlined by the authors, the reasons may be found on the enhanced wettability for the model oil promoted by the IL functionalization of the SBA-15 (changing its nature from hydrophilic to hydrophobic support), thus affording a catalytic system with high surface area, high accessibility of substrate and oxidant toward the active sites and substantially no leaching of the active species in the reaction mixture. Finally, the loading amount of HPA need to be carefully evaluated because beyond a certain value that varies for each species, the sulfur removal efficiency decrease due to the occurrence of mass transfer problems caused by the decrease of surface area and pore size of the support, thus limiting the successful contact between the substrates and the active species.

9. Miscellanea

In close correlation with SBA-15, the SBA-16 material is characterized by a more ordered three-dimensional (3-D) cage-like mesoporous structure, representing a further major achievement in the synthesis of mesoporous materials [210]. Pluronic F127 is used as a template to synthesize SBA-16 instead of P123 [211], providing larger pore diameter and wall thickness, as well as, higher thermal and hydrothermal stability [212]. Unfortunately, the preparation of SBA-16 requires acid conditions which are not compatible with the incorporation of a large panel of metal derivatives by traditional procedures, due to the occurrence of leaching processes. To avoid this drawback, Ti-SBA-16 mesoporous catalysts have been prepared (with various Ti loadings of 5, 10, and 15 wt%) by the evaporation-induced self-assembly method, and successively applied in the ODS procedure [213]. In this procedure, F127 in ethanol was treated with the appropriate amount of tetrabutyl titanate and TEOS under acid conditions (pH 2), followed by removal of the organic solvent and calcination at high temperature. Ti-SBA-16 showed the typical XRD signals of mesostructured material characterized by a slightly disordered structure. The presence of the silica-titanium framework was unambiguously confirmed by FTIR analysis [214] and XPS analysis, this latter highlighting that all Ti ions were incorporated onto the SiO2 structural motif [215,216,217]. Irrespective from the value of the Ti loading factor, all catalysts demonstrated a high activity in the oxidation of DBT as a model compound in n-octane/MeOH at 60 °C, using H2O2 as primary oxidant. A slight decrease in the activity of the catalysts was observed only at the highest 15 wt/% of Ti loading, probably due to the presence of major distortion effects on the mesoporous framework. Under optimal experimental conditions (that is 10 wt/% Ti-SBA-16, 60 °C for 3 h) 99% of sulfur removal was obtained.

10. Conclusions

The best catalysts and experimental conditions for the desulfurization process are summarized in Table 1, where the selection criteria have been focused upon the nature of the catalyst modification, the operating conditions and the percentage of the sulfur removal. As a general trend, the wetness impregnation and the grafting insertion technologies showed a similar efficacy in the activity of SBA-15 based metal oxide catalysts, with the only exception of TiO2, in which case the grafting procedure yielded the most reactive system. The low expensive, mild and environmentally friendly H2O2 was applied as main oxidant in most of the reported procedure, highlighting the high catalytic performance of the metal oxides. The use of ionic liquids as selective extraction solvent or, in alternative, the employment of carboxylic acid as co-oxidant, does not appear to significantly improve the efficiency of the desulfurization process, with respect to simpler procedures. Among the metal oxides, V2O5/SBA-15 was the only catalyst able to activate dioxygen (even if at relatively high temperature, 120 °C), while [pmim] FeCl4/SBA-15 was operative at the lowest reaction temperature (30 °C). With respect to the recyclability of the systems, Fe/SBA-15, TiO2/SiO2 and V2O5/SBA-15, were effective systems ranging from 4 to 5 and 6 successive runs, respectively. In the case of SBA-15 based polyoxometalates, wet impregnation or self-assembly strategies were the mainly used strategies for their heterogenization, varying from phosphomolybdic to phosphotungstic containing systems. Excellent results in terms of sulfur removal percentage and recyclability efficiencies have been observed after the POMs heterogenization on previously functionalized SBA-15, through the introduction of amino groups, trialkylammonium moieties or selected ionic liquids, even if excellent results were observed after simple doping of micro-mesoporous composite molecular sieves (ZSM-5/SBA-15) with Zr, Ag and Ce metal ions.

Author Contributions

Conceptualization, writing—review and editing, M.C. and R.S.; graphics editing and literature review, B.M.B.

Funding

This work was supported by MIUR (Ministero dell’Istruzione, dell’Università della Ricerca Italiano), PRIN 2017 project “ORIGINALE CHEMIAE in Antiviral Strategy—Origin and Modernization of Multi-Component Chemistry as a Source of Innovative Broad Spectrum Antiviral Strategy”, cod. 2017BMK8JR_002.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dos Santos, S.M.L.; Nogueira, K.A.B.; De Souza Gama, M.; Lima, J.D.F.; Da Silva Júnior, I.J.; De Azevedo, D.C.S. Synthesis and characterization of ordered mesoporous silica (SBA-15 and SBA-16) for adsorption of biomolecules. Microporous Mesoporous Mater. 2013, 180, 284–292. [Google Scholar] [CrossRef]
  2. Gaffar, I.; Rajamanickam, M. Mn-incorporated SBA-1 cubic mesoporous silicates: Synthesis and characterization. Mater. Chem. Phys. 2015, 161, 237–242. [Google Scholar]
  3. Chaudhary, V.; Sharma, S. An overview of ordered mesoporous material SBA-15: Synthesis, functionalization and application in oxidation reactions. J. Porous Mater. 2017, 24, 741–749. [Google Scholar] [CrossRef]
  4. Zhao, D.; Sun, J.; Li, Q.; Stucky, G.D. Morphological control of highly ordered mesoporous silica SBA-15. Chem. Mater. 2000, 12, 275–279. [Google Scholar] [CrossRef]
  5. Shen, S.; Chen, F.; Chow, P.S.; Phanapavudhikul, P.; Zhu, K.; Tan, R.B.H. Synthesis of SBA-15 mesoporous silica via dry-gel conversion route. Microporous Mesoporous Mater. 2006, 92, 300–308. [Google Scholar] [CrossRef]
  6. Zhao, D.; Feng, J.; Huo, Q.; Melosh, N.; Fredrickson, G.H.; Chmelka, B.F.; Stucky, G.D. Triblock copolymer syntheses of mesoporous silica with periodic 50 to 300 angstrom pores. Science 1998, 279, 548–552. [Google Scholar] [CrossRef]
  7. Zhang, F.; Yan, Y.; Yang, H.; Meng, Y.; Yu, C.; Tu, B.; Zhao, D. Understanding effect of wall structure on the hydrothermal stability of mesostructured silica SBA-15. J. Phys. Chem. B 2005, 109, 8723–8732. [Google Scholar] [CrossRef]
  8. Miyazawa, K.; Inagaki, S. Control of the microporosity within the pore walls of ordered mesoporous silica SBA-15. Chem. Commun. 2000, 21, 2121–2122. [Google Scholar] [CrossRef]
  9. Yang, L.; Qi, Y.; Yuan, X.; Shen, J.; Kim, J. Direct synthesis, characterization and catalytic application of SBA-15 containing heteropolyacid H3PW12O40. J. Mol. Catal. A 2005, 229, 199–205. [Google Scholar] [CrossRef]
  10. Mohammadi Ziarani, G.; Lashgari, N.; Badiei, A. Application of Organoamine-functionalized Mesoporous Silica (SBA-Pr-NH2) as a Nano Base Catalyst in Organic Reactions. Curr. Org. Chem. 2017, 21, 674–687. [Google Scholar] [CrossRef]
  11. Wu, S.H.; Lin, H.P. Synthesis of mesoporous silica nanoparticles. Chem. Soc. Rev. 2013, 42, 3862–3875. [Google Scholar] [CrossRef] [PubMed]
  12. Hoffmann, F.; Cornelius, M.; Morell, J.; Fröba, M. Silica-based mesoporous organic-inorganic hybrid materials. Angew. Chem. Int. Ed. 2006, 45, 3216–3251. [Google Scholar] [CrossRef] [PubMed]
  13. Martinez, F.; Han, Y.J.; Stucky, G.; Sotelo, J.L.; Ovejerob, G.; Melero, J.A. Synthesis and characterisation of iron-containing SBA-15 mesoporous silica. Stud. Surf. Sci. Catal. 2002, 142, 1109–1116. [Google Scholar]
  14. García-Martínez, J.C.; González Uribe, H.A.; González-Brambila, M.M.; Colín-Luna, J.A.; Escobedo-García, Y.E.; López-Gaona, A.; Alvarado-Perea, L. Selective adsorption of nitrogen compounds using silica-based mesoporous materials as a pretreatment for deep hydrodesulfurization. Catal. Today 2018, 305, 40–48. [Google Scholar] [CrossRef]
  15. Luan, Z. Alumination and ion exchange of mesoporous SBA-15 molecular sieves. Chem. Mater. 1999, 11, 1621–1627. [Google Scholar] [CrossRef]
  16. Venezia, A.M.; Di Carlo, G.; Liotta, L.F.; Pantaleo, G.; Kantcheva, M. Effect of Ti(IV) loading on CH4 oxidation activity and SO2 tolerance of Pd catalysts supported on silica SBA-15 and HMS. Appl. Catal. B 2011, 6, 529–539. [Google Scholar] [CrossRef]
  17. Mishra, G.S.; Machado, K.; Kumar, A. Highly selective n-alkanes oxidation to ketones with molecular oxygen catalyzed by SBA-15 supported rhenium catalysts. J. Ind. Eng. Chem. 2014, 20, 2228–2235. [Google Scholar] [CrossRef]
  18. Hu, J.; Li, K.; Li, W.; Ma, F.; Guo, Y. Selective oxidation of styrene to benzaldehyde catalyzed by Schiff base-modified ordered mesoporous silica materials impregnated with the transition metal-monosubstituted Keggin-type polyoxometalates. Appl. Catal. A 2009, 364, 211–220. [Google Scholar] [CrossRef]
  19. Yang, Y.; Zhang, Y.; Hao, S.; Guan, J.; Ding, H.; Shang, F.; Qiu, P.; Kan, Q. Heterogenization of functionalized Cu(II) and VO(IV) Schiff base complexes by direct immobilization onto amino-modified SBA-15: Styrene oxidation catalysts with enhanced reactivity. Appl. Catal. A 2010, 381, 274–281. [Google Scholar] [CrossRef]
  20. Jin, M.; Kim, J.H.; Kim, J.M.; Jeon, J.K.; Jurng, J.; Bae, G.N.; Park, Y.K. Benzene oxidation with ozone over MnOx/SBA-15 catalysts. Catal. Today 2013, 204, 108–113. [Google Scholar] [CrossRef]
  21. Neeli, C.K.P.; Narani, A.; Marella, R.K.; Rao, K.S.R.; Burri, D.R. Selective benzylic oxidation of alkylaromatics over Cu/SBA-15 catalysts under solvent-free conditions. Catal. Commun. 2013, 39, 5–9. [Google Scholar] [CrossRef]
  22. Cruz, P.; Perez, Y.; Hierro, I.D.; Fajardo, M. Copper, copper oxide nanoparticles and copper complexes supported on mesoporous SBA-15 as catalyst in the selective oxidation of benzyl alcohol in acqueous phase. Microporous Mesoporous Mater. 2016, 220, 136. [Google Scholar] [CrossRef]
  23. Liotta, L.F.; Pantaleo, G.; Puleo, F.; Venezia, A.M. Au/CeO2-SBA-15 catalysts for CO oxidation: Effect of ceria loading on physic-chemical properties and catalytic performances. Catal. Today 2012, 187, 10–19. [Google Scholar] [CrossRef]
  24. Zhang, L.X.; Shi, J.L.; Yu, J.; Hua, Z.L.; Zhao, X.G.; Ruan, M.L. A new in-situ reduction route for the synthesis of Pt nanoclusters in the channels of mesoporous silica SBA-15. Adv. Mater. 2002, 14, 1510–1513. [Google Scholar] [CrossRef]
  25. Xueguang, W.; Kyle, S.K.L.; Jerry, C.C.; Soofin, C. Direct Synthesis and Catalytic Applications of Ordered Large Pore Aminopropyl-Functionalized SBA-15 Mesoporous Materials. J. Phys. Chem. 2005, 109, 1763–1769. [Google Scholar]
  26. Haribandhu, C.; Subhajit, D.; Ashis, S. Synthesis and use of SBA-15 adsorbent for dye-loaded wastewater treatment. J. Environ. Chem. Eng. 2015, 3, 2866–2874. [Google Scholar]
  27. Tummino, M.L.; Testa, M.L.; Malandrino, M.; Gamberini, R.; Prevot, A.B.; Magnacca, G.; Laurenti, E. Green Waste-Derived Substances Immobilized on SBA-15 Silica: Surface Properties, Adsorbing and Photosensitizing Activities towards Organic and Inorganic Substrates. Nanomaterials 2019, 9, 162. [Google Scholar] [CrossRef] [PubMed]
  28. Moritz, M.; Łaniecki, M. Application of SBA-15 mesoporous material as the carrier for drug formulation systems. Papaverine hydrochloride adsorption and release study. Powder Technol. 2012, 230, 106–111. [Google Scholar] [CrossRef]
  29. Szewczyk, A.; Prokopowicz, M. Amino-modified mesoporous silica SBA-15 as bifunctional drug delivery system for cefazolin: Release profile and mineralization potential. Mater. Lett. 2018, 227, 136–140. [Google Scholar] [CrossRef]
  30. EU Regulations. Available online: http://www.dieselnet.com/standards/eu/fuel.php (accessed on 3 June 2013).
  31. Sharipov, A.K.; Nigmatullin, V.R. Removal of Sulfur from Hydrotreated Diesel Fuel. Chem. Technol. Fuels Oils 2005, 41, 225–229. [Google Scholar] [CrossRef]
  32. González-García, O.; Cedeño-Caero, L. V-Mo based catalysts for oxidative desulfurization of diesel fuel. Catal. Today 2009, 148, 42–48. [Google Scholar] [CrossRef]
  33. Rezvani, M.A.; Shaterian, M.; Aghbolagh, Z.S.; Babaei, R. Oxidative desulfurization of gasoline catalyzed by IMID@PMA@CS nanocomposite as a high-performance amphiphilic nano catalyst. Environ. Prog. Sustain. Energy 2018, 37, 1891–1900. [Google Scholar] [CrossRef]
  34. Rezvani, M.A.; Shaterian, M.; Akbarzadeh, F.; Khandan, S. Deep oxidative desulfurization of gasoline induced by PMoCu@MgCu2O4-PVA composite as a high-performance heterogeneous nanocatalyst. Chem. Eng. J. 2018, 333, 537–544. [Google Scholar] [CrossRef]
  35. Eßer, J.; Wasserscheid, P.; Jess, A. Deep desulfurization of oil refinery streams by extraction with ionic liquids. Green Chem. 2004, 6, 316–322. [Google Scholar] [CrossRef]
  36. Zeelani, G.G.; Lal Pal, S. A Review on Desulfurization Techniques of Liquid Fuels. Int. J. Sci. Res. 2016, 5, 2413–2419. [Google Scholar]
  37. Vsudevan, P.T.; Fierro, J.L.G. A Review of Deep Hydrodesulfurization Catalysis. Catal. Rev. Sci. Eng. 1996, 38, 2. [Google Scholar] [CrossRef]
  38. Tanimu, A.; Alhooshani, K. Advanced Hydrodesulfurization Catalysts: A Review of Design and Synthesis. Energy Fuels 2019, 33, 2810–2838. [Google Scholar] [CrossRef]
  39. Huirache-Acuña, R.; Rufino, N.; Peza-Ledesma, C.L.; Javier, L.R.; Alonso-Núñe, G.; Pawelec, B.; Rivera-Muñoz, E.M. SBA-15 Mesoporous Silica as Catalytic Support for Hydrodesulfurization Catalysts-Review. Materials 2013, 6, 4139–4167. [Google Scholar] [CrossRef] [Green Version]
  40. Oyama, S.T.; Zhao, H.; Freund, H.J.; Asakura, K.; Włodarczyk, R.; Sierka, M. Unprecedented selectivity to the direct desulfurization (DDS) pathway in a highly active FeNi bimetallic phosphide catalyst. J. Catal. 2012, 285, 1–5. [Google Scholar] [CrossRef]
  41. Jiang, Z.X.; Liu, Y.; Sun, X.P.; Tian, F.P.; Sun, F.X.; Liang, C.H.; You, W.S.; Han, C.R.; Li, C. Activated Carbons Chemically Modified by Concentrated H2SO4 for the Adsorption of the Pollutants from Wastewater and the Dibenzothiophene from Fuel Oils. Langmuir 2003, 19, 731–736. [Google Scholar] [CrossRef]
  42. Al-Shahrani, F.; Xiao, T.; Llewellyn, S.A.; Barri, S.; Jiang, Z.; Shi, H.; Martinie, G.; Green, M.L.H. Desulfurization of diesel via the H2O2 oxidation of aromatic sulfides to sulfones using a tungstate catalyst. Appl. Catal. B. 2007, 73, 311–316. [Google Scholar] [CrossRef]
  43. Babich, I.V.; Moulijn, J.A. Science and technology of novel processes for deep desulfurization of oil refinery streams: A review. Fuel 2003, 82, 607–631. [Google Scholar] [CrossRef]
  44. Schulz, H.; Böhringer, W.; Waller, P.; Ousmanov, F. Gas oil deep hydrodesulfurization: Refractory compounds and retarded kinetics. Catal. Today 1999, 49, 87–97. [Google Scholar] [CrossRef]
  45. Dou, S.-Y.; Wang, R. Recent Advances in New Catalysts for Fuel Oil Desulfurization. Curr. Org. Chem. 2017, 21, 1019–1036. [Google Scholar] [CrossRef]
  46. Zhu, W.; Li, H.; Jiang, X.; Yan, Y.; Lu, J.; Xia, J. Oxidative Desulfurization of Fuels Catalyzed by Peroxotungsten and Peroxomolybdenum Complexes in Ionic Liquids. Energy Fuels 2007, 21, 2514–2516. [Google Scholar] [CrossRef]
  47. Saladino, R.; Botta, G.; Crucianelli, M. Advances in Nanostrostructured Transition Metal Catalysis Applied to Oxidative Desulfurization. In Applying Nanotechnology to the Desulfurization Process in Petroleum Engineering; Saleh, T.A., Ed.; IGI Global, 2016; pp. 1–555. [Google Scholar]
  48. Liu, Y.-Y.; Leus, K.; Sun, Z.; Li, X.; Depauw, H.; Wang, A.; Zhang, J.; Van Der Voort, P. Catalytic oxidative desulfurization of model and real diesel over a molybdenum anchored metal-organic framework. Microporous Mesoporous Mat. 2019, 277, 245–252. [Google Scholar] [CrossRef]
  49. Chan, N.Y.; Lin, T.-Y.; Yen, T.F. Superoxides: Alternative Oxidants for the Oxidative Desulfurization Process. Energy Fuels 2008, 22, 3326–3328. [Google Scholar] [CrossRef]
  50. Shaabani, A.; Behnam, M.; Rezayan, A.H. Tungstophosphoric Acid (H3PW12O40) Catalyzed Oxidation of Organic Compounds with NaBrO3. Catal. Commun. 2009, 10, 1074–1078. [Google Scholar] [CrossRef]
  51. Campos-Martin, J.M.; Capel-Sanchez, M.C.; Fierro, J.L.G. Highly efficient deep desulfurization of fuels by chemical oxidation. Green Chem. 2004, 6, 557–562. [Google Scholar] [CrossRef]
  52. Garcìa-Gutiérrez, J.L.; Fuentes, G.A.; Hernàndez-Teràn, M.E.; Murrieta, F.; Navarrete, J.; Jiménez-Cruz, F. Ultra-deep oxidative desulfurization of diesel fuel with H2O2 catalyzed under mild conditions by polymolybdates supported on Al2O3. Appl. Catal. 2006, 305, 15–20. [Google Scholar] [CrossRef]
  53. Ciclosi, M.; Dinoi, C.; Gonsalvi, L.; Peruzzini, M.; Manoury, E.; Poli, R. Oxidation of Thiophene Derivatives with H2O2 in Acetonitrile Catalyzed by [Cp*2M2O5] (M = Mo, W): A Kinetic Study. Organometallics 2008, 27, 2281–2286. [Google Scholar] [CrossRef]
  54. Liu, G.; Cao, Y.; Jiang, R.; Wang, L.; Zhang, X.; Mi, Z. Oxidative desulfurization of jet fuels and its impact on thermal-oxidative stability. Energy Fuel 2009, 23, 5978–5985. [Google Scholar] [CrossRef]
  55. Noyori, R.; Aoki, M.; Sato, K. Green oxidation with aqueous hydrogen peroxide. Chem. Commun. 2003, 1977–1986. [Google Scholar] [CrossRef] [PubMed]
  56. Maksimchuk, N.; Lee, J.S.; Ayupov, A.; Chang, J.-S.; Kholdeeva, O. Cyclohexene Oxidation with H2O2 over Metal-Organic Framework MIL-125(Ti): The Effect of Protons on Reactivity. Catalysts 2019, 9, 324. [Google Scholar] [CrossRef] [Green Version]
  57. Di Giuseppe, A.; Crucianelli, M.; De Angelis, F.; Crestini, C.; Saladino, R. Efficient Oxidation of Thiophene Derivatives with Homogeneous and Heterogeneous MTO/H2O2 systems: A novel approach for Oxidative Desulfurization (ODS) of Diesel Fuel. Appl. Catal. B 2009, 89, 239–245. [Google Scholar] [CrossRef]
  58. Saladino, R.; Crucianelli, M.; De Angelis, F. Methyltrioxorhenium catalysis in nonconventional solvents: A great catalyst in safe reaction medium. ChemSusChem 2010, 3, 524–540. [Google Scholar]
  59. Meman, N.M.; Pourkhalil, M.; Rashidi, A.; Zarenezhad, B. Synthesis, characterization and operation of a functionalized multi-walled CNT supported MnOx nanocatalyst for deep oxidative desulfurization of four petroleum fractions. J. Ind. Eng. Chem. 2014, 20, 4054–4058. [Google Scholar] [CrossRef]
  60. Balzani, V.; Credi, A.; Venture, M. Photochemical conversion of solar energy. ChemSusChem 2008, 1, 26–58. [Google Scholar] [CrossRef]
  61. Ramirez-Verduzco, L.F.; De los Reyes, J.A.; Torres-Garcia, E. Solvent Effect in Homogeneous and Heterogeneous Reactions to Remove Dibenzothiophene by an Oxidation-Extraction Scheme. Ind. Eng. Chem. Res. 2008, 47, 5353–5361. [Google Scholar] [CrossRef]
  62. Rodríguez-Cabo, B.; Rodríguez, H.; Rodil, E.; Arce, A.; Soto, A. Extractive and oxidative-extractive desulfurization of fuels with ionic liquids. Fuel 2014, 117, 882–889. [Google Scholar] [CrossRef]
  63. Taguchi, A.; Schuth, F. Ordered mesoporous materials in catalysis. Microporous Mesoporous Mater. 2005, 77, 1–45. [Google Scholar] [CrossRef]
  64. Huang, D.; Wang, Y.J.; Cui, Y.C.; Luo, G.S. Direct synthesis of mesoporous TiO2 and its catalytic performance in DBT oxidative desulfurization. Microporous Mesoporous Mater. 2008, 116, 378–385. [Google Scholar] [CrossRef]
  65. Tanev, P.T.; Chibwe, M.; Pinnavaia, T.J. Titanium-containing mesoporous molecular sieves for catalytic oxidation of aromatic compounds. Nature 1994, 368, 321–323. [Google Scholar] [CrossRef]
  66. Chica, A.; Corma, A.; Dómine, M.E. Catalytic oxidative desulfurization (ODS) of diesel fuel on a continuous fixed-bed reactor. J. Catal. 2006, 242, 299–308. [Google Scholar] [CrossRef]
  67. Yan, X.M.; Mei, P.; Lei, J.; Mi, Y.; Xiong, L.; Guo, L. Synthesis and characterization of mesoporous phosphotungstic acid/TiO2 nanocompositive as a novel oxidative desulfurization catalyst. J. Mol. Chem. Catal. A Chem. 2009, 304, 52–57. [Google Scholar] [CrossRef]
  68. Seung-Tae, Y.; Kwang-Eun, J.; Soon-Yong, J.; Wha-Seung, A. Synthesis of mesoporous TS-1 using a hybrid SiO2-TiO2 xerogel for catalytic oxidative desulfurization. Mater. Res. Bull. 2012, 47, 4398–4402. [Google Scholar]
  69. Sengupta, A.; Kamble, P.D.; Basu, J.K.; Sengupta, S. Kinetic study and optimization of oxidative desulfurization of benzothiophene using mesoporous titanium silicate-1 catalyst. Ind. Eng. Chem. Res. 2012, 51, 147–157. [Google Scholar] [CrossRef]
  70. Tangestaninejad, S.; Mirkhani, V.; Moghadam, M.; Mohammadpoor-Baltork, I.; Shams, E.; Salavati, H. Hydrocarbon oxidation catalyzed by vanadium polyoxometalate supported on mesoporous MCM-41 under ultrasonic irradiation. Ultrason. Sonochem. 2008, 15, 438–447. [Google Scholar] [CrossRef]
  71. Chen, Y.; Zhang, F.; Fang, Y.; Zhu, X.; Zhen, W.; Wang, R.; Ma, J. Phosphotungstic acid containing ionic liquid immobilized on magnetic mesoporous silica rod catalyst for the oxidation of dibenzothiophene with H2O2. Catal. Commun. 2013, 38, 54–58. [Google Scholar] [CrossRef]
  72. Wang, X.; Zhang, W.; Wu, L.; Ye, F.; Xiao, J.; Li, Z. One-pot photocatalysis-assisted adsorptive desulfurization of diesel over doped-TiO2 under ambient conditions. RSC Adv. 2014, 4, 56567–56570. [Google Scholar] [CrossRef]
  73. Shen, C.; Wang, Y.J.; Xu, J.H.; Luo, G.S. Synthesis of TS-1 on porous glass beads for catalytic oxidative desulfurization. Chem Eng. J. 2015, 259, 552–561. [Google Scholar] [CrossRef]
  74. Sentorun-Shalaby, C.; Saha, S.K.; Ma, X.L.; Song, C.S. Mesoporous-molecular-sievesupported nickel sorbents for adsorptive desulfurization of commercial ultralow-sulfur diesel fuel. Appl. Catal. B. 2011, 101, 718–726. [Google Scholar] [CrossRef]
  75. Ren, X.; Miao, G.; Xiao, Z.; Ye, F.; Li, Z.; Wang, H.; Xiao, J. Catalytic adsorptive desulfurization of model diesel fuel using TiO2/SBA-15 under mild conditions. Fuel 2016, 174, 118–125. [Google Scholar] [CrossRef]
  76. Hao, X.; Quach, L.; Korah, J.; Spieker, W.A.; Regalbuto, J.R. The control of platinum impregnation by PZC alteration of oxides and carbon. J. Mol. Catal. A Chem. 2004, 219, 97–107. [Google Scholar] [CrossRef]
  77. Lee, K.X.; Valla, J.A. Adsorptive desulfurization of liquid hydrocarbons using zeolite-based sorbents: A comprehensive review. React. Chem. Eng. 2019, 4, 1357–1386. [Google Scholar] [CrossRef]
  78. Miao, G.; Ye, F.; Wu, L.; Ren, X.; Xiao, J.; Li, Z.; Wang, H. Selective adsorption of thiophenic compounds from fuel over TiO2/SiO2 under UV-irradiation. J. Hazard. Mater. 2015, 300, 426–432. [Google Scholar] [CrossRef]
  79. Rourke, C.O.; Bowler, D.R. Adsorption of thiophene-conjugated sensitizers on TiO2 anatase (101). J. Phys. Chem. 2010, 114, 20240–20248. [Google Scholar]
  80. Zhuravlev, L.T. Concentration of hydroxyl groups on the surface of amorphous silicas. Langmuir 1987, 3, 316–318. [Google Scholar] [CrossRef]
  81. Collins, F.M.; Lucy, A.R.; Sharp, C. Oxidative desulphurization of oils via hydrogen peroxide and heteropolyanion catalysis. J. Mol. Catal. A Chem. 1997, 117, 397–403. [Google Scholar] [CrossRef]
  82. Gao, G.; Cheng, S.; An, Y.; Si, X.; Fu, X.; Liu, Y.; Zhang, H.; Wu, P.; He, M.Y. Oxidative Desulfurization of Aromatic Sulfur Compounds over Titanosilicates. ChemCatChem. 2010, 2, 459–466. [Google Scholar] [CrossRef]
  83. Capel-Sanchez, M.C.; Campos-Martin, J.M.; Fierro, J.L.G. Removal of refractory organosulfur compounds via oxidation with hydrogen peroxide on amorphous Ti/SiO2 catalysts. Energy Environ. Sci. 2010, 3, 328–333. [Google Scholar] [CrossRef] [Green Version]
  84. Bérubé, F.; Khadhraoui, A.; Janicke, M.T.; Kleitz, F.; Kaliaguine, S. Optimizing Silica Synthesis for the Preparation of Mesoporous Ti-SBA-15 Epoxidation Catalysts. Ind. Eng. Chem. Res. 2010, 49, 6977–6985. [Google Scholar] [CrossRef]
  85. Berube’, F.; Nohair, B.; Kleitz, F.; Kaliaguine, S. Controlled Postgrafting of Titanium Chelates for Improved Synthesis of Ti-SBA-15 Epoxidation Catalysts. Chem. Mater. 2010, 22, 1988–2000. [Google Scholar] [CrossRef]
  86. Choi, M.; Heo, W.; Kleitz, F.; Ryoo, R. Facile synthesis of high quality mesoporous SBA-15 with enhanced control of the porous network connectivity and wall thickness. Chem. Commun. 2003, 1340–1341. [Google Scholar] [CrossRef] [PubMed]
  87. Kleitz, F.; Bérubé, F.; Guillet-Nicolas, R.; Yang, C.M.; Thommes, M. Probing Adsorption, Pore Condensation, and Hysteresis Behavior of Pure Fluids in Three-Dimensional Cubic Mesoporous KIT-6 Silica. J. Phys. Chem. C 2010, 114, 9344–9355. [Google Scholar] [CrossRef]
  88. Cedeňo-Caero, L.; Ramos-Luna, M.; Méndez-Cruz, M.; Ramírez-Solís, J. Oxidative desulfurization of dibenzothiophene compounds with titania based catalysts. Catal. Today 2011, 172, 189–194. [Google Scholar] [CrossRef]
  89. Zhang, W.-H.; Lu, J.; Han, B.; Li, M.; Xiu, J.; Ying, P.; Can, L. Direct Synthesis and Characterization of Titanium-Substituted Mesoporous Molecular Sieve SBA-15. Chem. Mater. 2002, 14, 3413–3421. [Google Scholar] [CrossRef]
  90. Kruk, M.; Jaroniec, M.; Ko, C.H.; Ryoo, R. Characterization of the Porous Structure of SBA-15. Chem. Mater. 2000, 12, 1961–1968. [Google Scholar] [CrossRef]
  91. Kim, T.-W.; Kim, M.-J.; Kleitz, F.; Nair, M.M.; Guillet Nicolas, R.; Jeong, K.-E.; Chae, H.-J.; Kim, C.-U.; Jeong, S.-Y. Tailor-Made Mesoporous Ti-SBA-15 Catalysts for Oxidative Desulfurization of Refractory Aromatic Sulfur Compounds in Transport Fuel. ChemCatChem 2012, 4, 687–697. [Google Scholar] [CrossRef]
  92. Bianconi, A.; Fritsch, E.; Calas, G.; Petiau, J. X-ray-absorption near-edge structure of 3d transition elements in tetrahedral coordination: The effect of bond-length variation. Phys. Rev. B Condens. Matter. 1985, 32, 4292–4295. [Google Scholar] [CrossRef]
  93. Laredo, G.C.; Leyva, S.; Alvarez, R.; Mares, M.T.; Castillo, J.; Cano, J.L. Nitrogen compounds characterization in atmospheric gas oil and light cycle oil from a blend of Mexican crudes. Fuel 2002, 81, 1341–1350. [Google Scholar] [CrossRef]
  94. Cho, K.-S.; Le, Y.-K. Effects of nitrogen compounds, aromatics, and aprotic solvents on the oxidative desulfurization (ODS) of light cycle oil over Ti-SBA-15 catalyst. Appl. Catal. B 2014, 147, 35–42. [Google Scholar] [CrossRef] [Green Version]
  95. Ishihara, A.; Wang, D.; Dumeignil, F.; Amano, H.; Qian, E.W.; Kabe, T. Oxidative desulfurization and denitrogenation of a light gas oil using an oxidation/adsorption continuous flow process. Appl. Catal. A 2005, 279, 279–287. [Google Scholar] [CrossRef]
  96. Caero, L.C.; Navarro, J.F.; Gutierrez-Alejandre, A. Oxidative desulfurization of synthetic diesel using supported catalysts: Part II. Effect of oxidant and nitrogen-compounds on extraction–oxidation process. Catal. Today 2006, 116, 562–568. [Google Scholar] [CrossRef]
  97. Jia, Y.; Li, G.; Ning, G.; Jin, C. The effect of N-containing compounds on oxidative desulphurization of liquid fuel. Catal. Today 2009, 140, 192–196. [Google Scholar] [CrossRef]
  98. Mirante, F.; Ribeiro, S.O.; De Castro, B.; Granadeiro, C.M.; Balula, S.S. Sustainable Desulfurization Processes Catalyzed by TitaniumPolyoxometalate@TM-SBA-15. Top. Catal. 2017, 60, 1140–1150. [Google Scholar] [CrossRef]
  99. Kholdeeva, O.A.; Kovaleva, L.A.; Maksimovskaya, R.I.; Maksimov, G.M. Kinetics and mechanism of thioether oxidation with H2O2 in the presence of Ti (IV)-substituted heteropolytungstates. J. Mol. Catal. A Chem. 2000, 158, 223–229. [Google Scholar] [CrossRef]
  100. Maksimov, G.M.; Maksimovskaya, R.I.; Kholdeeva, O.A.; Fedotov, M.A.; Zaikovskii, V.I.; Vasil’ev, V.G.; Arzumanov, S.S. Structure and properties of H8(PW11TiO39)2O heteropoly acid. J. Struct. Chem. 2009, 50, 618–627. [Google Scholar] [CrossRef]
  101. Balula, S.S.; Cunha-Silva, L.; Santos, I.C.M.S.; Estrada, A.C.; Fernandes, A.C.; Cavaleiro, J.A.S.; Pires, J.; Freire, C.; Cavaleiro, A.M.V. Mono-substituted silicotungstates as active catalysts for sustainable oxidations: Homo- and heterogeneous performance. New J. Chem. 2013, 37, 2341–2350. [Google Scholar] [CrossRef]
  102. Balula, S.S.; Santos, I.C.M.S.; Cunha-Silva, L.; Carvalho, A.P.; Pires, J.; Freire, C.; Cavaleiro, J.A.S.; de Castro, B.; Cavaleiro, A.M.V. Phosphotungstates as catalysts for monoterpenes oxidation: Homo- and heterogeneous performance. Catal. Today 2013, 203, 95–102. [Google Scholar] [CrossRef]
  103. Kholdeeva, O.A.; Maksimovskaya, R.I. Titanium- and zirconium-monosubstituted polyoxometalates as molecular models for studying mechanisms of oxidation catalysis. J. Mol. Catal. A Chem. 2007, 262, 7–24. [Google Scholar] [CrossRef]
  104. Xu, J.; Zhao, S.; Chen, W.; Wang, M.; Song, Y.F. Highly efficient extraction and oxidative desulfurization system using Na7H2LaW10O36·32 H2O in [bmim]BF4 at room temperature. Chem. A Eur. J. 2012, 18, 4775–4781. [Google Scholar] [CrossRef] [PubMed]
  105. Zhu, W.; Huang, W.; Li, H.; Zhang, M.; Jiang, W.; Chen, G.; Han, C. Polyoxometalate-based ionic liquids as catalysts for deep desulfurization of fuels. Fuel Process. Technol. 2011, 92, 1842–1848. [Google Scholar] [CrossRef]
  106. Caero, L.C.; Hernández, E.; Pedraza, F.; Murrieta, F. Oxidative desulfurization of synthetic diesel using supported catalysts: Part, I. Study of the operation conditions with a vanadium oxide based catalyst. Catal. Today 2005, 107–108, 564–569. [Google Scholar] [CrossRef]
  107. Shiraishi, Y.; Naito, T.; Hirai, T. Vanadosilicate Molecular Sieve as a Catalyst for Oxidative Desulfurization of Light Oil. Ind. Eng. Chem. Res. 2003, 42, 6034–6039. [Google Scholar] [CrossRef]
  108. Anisimov, A.V.; Fedorova, E.V.; Lesnugin, A.Z.; Senyavin, V.M.; Aslanov, L.A.; Rybakov, V.B.; Tarakanova, A.V. Vanadium peroxocomplexes as oxidation catalysts of sulfur organic compounds by hydrogen peroxide in bi-phase systems. Catal. Today 2003, 78, 319–325. [Google Scholar] [CrossRef]
  109. Du, G.; Lim, S.; Pinault, M.; Wang, C.; Fang, F.; Pfefferle, L.; Haller, G.L. Synthesis, characterization, and catalytic performance of highly dispersed vanadium grafted SBA-15 catalyst. J. Catal. 2008, 253, 74–90. [Google Scholar] [CrossRef]
  110. Cedeño-Caero, L.; Gomez-Bernal, H.; Fraustro-Cuevas, A.; Guerra-Gomez, H.D.; Cuevas-Garcia, R. Oxidative desulfurization of synthetic diesel using supported catalysts. Part III. Support effect on vanadium-based catalysts. Catal. Today 2008, 133–135, 244–254. [Google Scholar] [CrossRef]
  111. Rivoira, L.; Martínez, M.L.; Anunziata, O.; Beltramone, A. Vanadium oxide supported on mesoporous SBA-15 modified with Al and Ga as a highly active catalyst in the ODS of DBT. Microporous Mesoporous Mater. 2017, 254, 96–113. [Google Scholar] [CrossRef]
  112. Yang, P.; Zhao, D.; Chmelka, B.F.; Stucky, G.D. Triblock-copolymer-directed syntheses of large-pore mesoporous silica fibers. Chem. Mater. 1998, 10, 2033–2036. [Google Scholar] [CrossRef]
  113. Liu, Y.M.; Cao, Y.; Yi, W.; Feng, W.L.; Dai, W.L.; Yan, S.R.; He, H.Y.; Fan, K.N. Vanadium oxide supported on mesoporous SBA-15 as highly selective catalysts in the oxidative dehydrogenation of propane. J. Catal. 2004, 224, 417–428. [Google Scholar] [CrossRef]
  114. Kumar, R.; Sithambaram, S.; Suib, S.L. Cyclohexane oxidation catalyzed by manganese oxide octahedral molecular sieves-Effect of acidity of the catalyst. J. Catal. 2009, 262, 304–313. [Google Scholar] [CrossRef]
  115. Rivoira, L.P.; Cussa, J.; Martínez, M.L.; Beltramone, A.R. Experimental design optimization of the ODS of DBT using vanadium oxide supported on mesoporous Ga-SBA-15. Catal. Today 2018. [Google Scholar] [CrossRef]
  116. Wang, C.; Chen, Z.; Yao, X.; Jiang, W.; Zhang, M.; Li, H.; Liu, H.; Zhu, W.; Li, H. One-pot extraction and aerobic oxidative desulfurization with highly dispersed V2O5/SBA-15 catalyst in ionic liquids. RSC Adv. 2017, 7, 39383. [Google Scholar] [CrossRef] [Green Version]
  117. Gao, X.; Xu, J. The oxygen activated by the active vanadium species for the selective oxidation of benzene to phenol. Catal. Lett. 2006, 111, 203. [Google Scholar] [CrossRef]
  118. Jia, Y.; Li, G.; Ning, G. Efficient oxidative desulfurization (ODS) of model fuel with H2O2 catalyzed by MoO3/γ-Al2O3 under mild and solvent free conditions. Fuel Process. Technol. 2011, 92, 106–111. [Google Scholar] [CrossRef]
  119. Hu, Z.; Lu, S.; Huang, X.; Li, J.; Duan, Y.; Yan, L.; Yao, Y.; Liao, X. Molybdenum anchored on NH2-modified spherical SiO2: A highly efficient and stable catalyst for oxidative desulfurization of fuel oil. Appl. Organomet. Chem. 2018, 32, e4521. [Google Scholar] [CrossRef]
  120. Han, X.; Wang, A.; Wang, X.; Li, X.; Wang, Y.; Hu, Y. Catalytic performance of P-modified MoO3/SiO2 in oxidative desulfurization by cumene hydroperoxide. Catal. Commun. 2013, 42, 6–9. [Google Scholar] [CrossRef]
  121. Chang, J.; Wang, A.; Liu, J.; Li, X.; Hu, Y. Oxidation of dibenzothiophene with cumene hydroperoxide on MoO3/SiO2 modified with alkaline earth metals. Catal. Today 2010, 149, 122–126. [Google Scholar] [CrossRef]
  122. Shao, X.; Zhang, X.; Yu, W.; Wu, Y.; Qin, Y.; Sun, Z.; Song, L. Effects of surface acidities of MCM-41 modified with MoO3 on adsorptive desulfurization of gasoline. Appl. Surf. Sci. 2012, 263, 1–7. [Google Scholar] [CrossRef]
  123. Ramos, J.M.; Wang, J.A.; Chen, L.F.; Arellano, U.; Ramírez, S.P.; Sotelo, R.; Schachat, P. Synthesis and catalytic evaluation of CoMo/SBA-15 catalysts for oxidative removal of dibenzothiophene from a model diesel. Catal. Commum. 2015, 72, 57–62. [Google Scholar] [CrossRef]
  124. La Parola, V.; Degnello, G.; Tewell, C.R.; Venrzial, A.M. Structural characterization of silica supported CoMo catalysts by UV, Raman spectroscopy, XPS and X-ray diffraction techniques. Appl. Catal. A Gen. 2002, 235, 171–180. [Google Scholar] [CrossRef]
  125. Briggs, D.; Seah, M.P. (Eds.) Practical Surface analysis, Auger and X-ray Photoelectron Spectroscopy, 2nd ed.; Wiley: New York, NY, USA, 1990; p. 607. [Google Scholar]
  126. González, J.; Wang, J.A.; Chen, L.; Manríquez, M.; Salmones, J.; Limas, R.; Arellano, U. Quantitative determination of oxygen defects, surface lewis acidity, and catalytic properties of mesoporous MoO3/SBA-15 catalysts. J. Solid State Chem. 2018, 263, 100–111. [Google Scholar] [CrossRef]
  127. Basahel, S.N.; Ali, T.T.; Narasimharao, K.; Bagabas, A.A.; Mokhtar, M. Effect of iron oxide loading on the phase transformation and physicochemical properties of nanosized mesoporous ZrO2. Mater. Res. Bull. 2012, 47, 3463–3472. [Google Scholar] [CrossRef]
  128. Epifani, M.; Imperatori, P.; Mirenghi, L.; Schioppa, M.; Siciliano, P. Synthesis and Characterization of MoO3 Thin Films and Powders from a Molybdenum Chloromethoxide. Chem. Mater. 2004, 16, 5495–5501. [Google Scholar] [CrossRef]
  129. He, T.; Yao, J. Photochromism of molybdenum oxide. J. Photochem. Photobiol. C Photochem. Rev. 2003, 4, 125–143. [Google Scholar] [CrossRef]
  130. De Filippis, P.; Scarsella, M.; Verdone, N. Oxidative Desulfurization I: Peroxyformic Acid Oxidation of Benzothiophene and Dibenzothiophene. Ind. Eng. Chem. Res. 2009, 48, 1372–1375. [Google Scholar] [CrossRef]
  131. Arends, I.W.C.E.; Sheldon, R.A. Activities and stabilities of heterogeneous catalysts in selectiveliquid phase oxidations: Recent developments. Appl. Catal. A 2001, 212, 175–187. [Google Scholar] [CrossRef]
  132. Wang, Q.L.; Lei, L.C.; Zhu, J.K.; Yang, B.; Li, Z.J. Deep desulfurization of fuels by extraction with 4-dimethylaminopyridinium-based ionic liquids. Energy Fuels 2013, 27, 4617–4623. [Google Scholar] [CrossRef]
  133. Chen, X.C.; Song, D.D.; Asumana, C.; Yu, G.R. Deep oxidative desulfurization of diesel fuels by Lewis acidic ionic liquids based on 1-n-butyl-3-methylimidazolium metal chloride. J. Mol. Catal. A 2012, 359, 8–13. [Google Scholar] [CrossRef]
  134. Nie, Y.; Dong, Y.X.; Bai, L.; Dong, H.F.; Zhang, X.P. Fast oxidative desulfurization of fuel oil using dialkylpyridinium tetrachloroferrates ionic liquids. Fuel 2013, 103, 997–1002. [Google Scholar] [CrossRef]
  135. Zhu, W.S.; Zhang, J.T.; Li, H.M.; Chao, Y.H.; Jiang, W.; Yin, S.; Liu, H. Fenton-like ionic liquids/H2O2 system: One-pot extraction combined with oxidation desulfurization of fuel. RSC Adv. 2012, 2, 658–664. [Google Scholar] [CrossRef]
  136. Zhang, J.T.; Zhu, W.S.; Li, H.M.; Jiang, W.; Jiang, Y.Q.; Huang, W.L.; Yan, Y.S. Deep oxidative desulfurization of fuels by Fenton-like reagent in ionic liquids. Green Chem. 2009, 11, 1801–1807. [Google Scholar] [CrossRef]
  137. Ding, W.; Zhu, W.; Xiong, J.; Yang, L.; Wei, A.; Zhang, M.; Li, H. Novel heterogeneous iron-based redox ionic liquid supported on SBA-15 for deep oxidative desulfurization of fuels. Chem. Eng. J. 2015, 266, 213–221. [Google Scholar] [CrossRef]
  138. Bordoloi, A.; Sahoo, S.; Lefebvre, F.; Halligudi, S. Heteropoly acid-based supported ionic liquid-phase catalyst for the selective oxidation of alcohols. J. Catal. 2008, 259, 232–239. [Google Scholar] [CrossRef]
  139. Rivoira, L.; Juárez, J.; Martínez, M.L.; Beltramone, A. Iron-modified mesoporous materials as catalysts for ODS of sulfur compounds. Catal. Today 2018. [Google Scholar] [CrossRef]
  140. Sanjini, N.S.; Velmathi, S. Iron impregnated SBA-15, a mild and efficient catalyst for the catalytic hydride transfer reduction of aromatic nitro compounds. RSC Adv. 2014, 4, 15381–15388. [Google Scholar] [CrossRef]
  141. Chaliha, S.; Bhattacharyya, K.G. Fe(III)-, Co(II)- and Ni(II)-impregnated MCM41 for wet oxidative destruction of 2,4-dichlorophenol in water. Catal. Today 2009, 141, 225–233. [Google Scholar] [CrossRef]
  142. Barpanda, P.; Recham, N.; Chotard, J.N.; Djellab, K.; Walker, W.; Armand, M.; Tarascon, J.M. Structure and electrochemical properties of novel mixed Li(Fe1−xMx)SO4F (M = Co, Ni, Mn) phases fabricated by low temperature ionothermal synthesis. J. Mater. Chem. 2010, 20, 1659–1668. [Google Scholar] [CrossRef]
  143. Otsuki, T.; Nonaka, N.; Takashima, W.; Qian, A.; Ishihara, T.; Imai, T.; Kabe, T. Oxidative Desulfurization of Light Gas Oil and Vacuum Gas Oil by Oxidation and Solvent Extraction. Energy Fuels 2000, 14, 1232–1239. [Google Scholar] [CrossRef]
  144. Ramos, J.M.; Wang, J.A.; Flores, S.O.; Chen, L.F.; Nava, N.; Navarrete, J.; Domínguez, J.M.; Szpunar, J.A. Ultrasound-assisted synthesis and catalytic activity of mesostructured FeOx/ SBA-15 and FeOx/Zr-SBA-15 catalysts for the oxidative desulfurization of model diesel. Catal. Today 2018. [Google Scholar] [CrossRef]
  145. Rodríguez-Reinoso, F. The role of carbon materials in heterogeneous catalysis. Carbon 1998, 36, 159–175. [Google Scholar] [CrossRef]
  146. Serp, P.; Figueiredo, J.L. Carbon Materials for Catalysis; John Wiley & Sons Inc.: Hoboken, NJ, USA, 2009. [Google Scholar]
  147. Tharamani, C.N.; Bordoloi, A.; Schuhmann, W.; Muhler, M. Mesoporous nitrogen-rich carbon materials as catalysts for the oxygen reduction reaction. ChemSusChem 2012, 5, 637–641. [Google Scholar]
  148. Su, F.C.; Mathew, S.; Lipner, G.; Fu, X.; Antonietti, M.; Blechert, S.; Wang, X. mpg-C3N4-Catalyzed Selective Oxidation of Alcohols Using O2 and Visible Light. J. Am. Chem. Soc. 2010, 132, 16299–16301. [Google Scholar] [CrossRef] [PubMed]
  149. Piccinino, D.; Abdalghani, I.; Botta, G.; Crucianelli, M.; Passacantando, M.; Di Vacri, M.L.; Saladino, R. Preparation of wrapped carbon nanotubes poly(4-vinylpyridine)/MTO based heterogeneous catalysts for the oxidative desulfurization (ODS) of model and synthetic diesel fuel. Appl. Catal. B Environ. 2017, 200, 392–401. [Google Scholar] [CrossRef]
  150. Jin, X.; Balasubramanian, V.V.; Selvan, S.T.; Sawant, D.P.; Chari, M.A.; Lu, G.Q.; Vinu, A. Highly ordered mesoporous carbon nitride nanoparticles with high nitrogen content: A metal-free basic catalyst. Angew. Chem. Int. Ed. 2009, 48, 7884–7887. [Google Scholar] [CrossRef]
  151. Goyal, R.; Sarkar, B.; Lucus, N.; Bordoloi, A. Acid–Base Cooperative Catalysis over Mesoporous Nitrogen-Rich Carbon. ChemCatChem 2014, 6, 3091–3095. [Google Scholar] [CrossRef]
  152. Li, X.; Huang, S.; Xu, Q.; Yang, Y. Preparation of WO3-SBA-15 mesoporous molecular sieve and its performance as an oxidative desulfurization catalyst. Transit. Met. Chem. 2009, 34, 943–947. [Google Scholar] [CrossRef]
  153. Wang, H.; Gong, Y.; Wang, Y. Cellulose-based hydrophobic carbon aerogels as versatile and superior adsorbents for sewage treatment. RSC Adv. 2014, 4, 45753. [Google Scholar] [CrossRef]
  154. Zhang, P.; Wang, Y.; Li, H.; Antonietti, M. Metal-free oxidation of sulfides by carbon nitride with visible light illumination at room temperature. Green Chem. 2012, 14, 1904–1908. [Google Scholar] [CrossRef]
  155. Goyal, R.; Dumbre, D.; Sivakumar Konathala, L.N.; Pandey, M.; Bordoloi, A. Oxidative coupling of aniline and desulfurization over nitrogen rich mesoporous carbon. Catal. Sci. Technol. 2015, 5, 3632–3638. [Google Scholar] [CrossRef]
  156. Wang, X.; Wu, J.; Zhao, M.; Lv, Y.; Li, G.; Hu, C.J. Partial Oxidation of Toluene in CH3COOH by H2O2 in the Presence of VO (acac)2 Catalyst. J. Phys. Chem. C 2009, 113, 14270–14278. [Google Scholar] [CrossRef]
  157. Nair, S.; Tatarchuk, B.J. Supported silver adsorbents for selective removal of sulfur species from hydrocarbon fuels. Fuel 2010, 89, 3218–3225. [Google Scholar] [CrossRef]
  158. Samokhvalov, A.; Duin, E.C.; Nair, S. Study ofthe surface chemical reactions of thiophene with Ag/titania by the complementary temperature-programmed electron spin resonance, temperature-programmed desorption, and X-ray photoelectron spectroscopy: Adsorption, desorption, and sorbent regeneration mechanisms. J. Phys. Chem. C 2010, 114, 4075–4085. [Google Scholar]
  159. Mckinley, S.G.; Angelici, R.J. Deep desulfurization by selective adsorption of dibenzothiophenes on Ag+/SBA-15 and Ag+/SiO2. Chem. Commun. 2003, 2620–2621. [Google Scholar] [CrossRef]
  160. Xiao, J.; Li, Z.; Liu, B. Adsorption of benzothiophene and dibenzothiophene on ion-impregnated activated carbons and ion-exchanged Y zeolites. Energy Fuels 2008, 22, 3858–3863. [Google Scholar] [CrossRef]
  161. Xiao, J.; Bian, G.A.; Zhang, W.; Li, Z. Adsorption of dibenzothiophene on Ag/Cu/ Fe-supported activated carbons prepared by ultrasonic-assisted impregnation. J. Chem. Eng. 2010, 55, 5818–5823. [Google Scholar] [CrossRef]
  162. Ghosh, S.; Acharyya, S.S.; Tiwari, R.; Sarkar, B.; Singha, R.K.; Pendem, C.; Sasaki, T.; Bal, R. Selective Oxidation of Propylene to Propylene Oxide over Silver-Supported Tungsten Oxide Nanostructure with Molecular Oxygen. ACS Catal. 2014, 4, 2169–2174. [Google Scholar] [CrossRef]
  163. Hosseini, S.M.; Hosseini-Monfared, H.; Abbasi, V.; Khoshroo, M.R. Selective oxidation of hydrocarbons under air using recoverable silver ferrite–graphene (AgFeO2–G) nanocomposite: A good catalyst for green chemistry. Inorg. Chem. Commun. 2016, 67, 72–79. [Google Scholar] [CrossRef]
  164. Fellah, M.F.; Santen, R.A.; Onal, I. Epoxidation of ethylene by silver oxide (Ag2O) cluster: A density functional theory study. Catal. Lett. 2011, 141, 762–771. [Google Scholar] [CrossRef]
  165. Zhang, J.; Li, Y.; Zhang, Y. Effect of support on the activity of Ag-based catalysts for formaldehyde oxidation. Sci. Rep. 2015, 5, 12950. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Li, Y.; Zhang, X.; He, H.; Yu, Y.; Yuan, T.; Tian, Z.; Wang, J.; Li, Y. Effect of the pressure on the catalytic oxidation of volatile organic compounds over Ag/Al2O3 catalyst. Appl. Catal. B Environ. 2009, 89, 659–664. [Google Scholar] [CrossRef]
  167. Zhang, X.; Qu, Z.; Li, X.; Wen, M. Studies of silver species for low-temperature CO oxidation on Ag/SiO2 catalysts. Sep. Purif. Technol. 2010, 72, 395–400. [Google Scholar] [CrossRef]
  168. Ye, F.; Miao, G.; Wu, L.; Wu, Y.; Li, Z.; Song, C.; Xiao, J. [O]-induced reactive adsorptive desulfurization of liquid fuel over Ag X O@SBA-15 under ambient conditions. Chem. Eng. Sci. 2017, 168, 225–234. [Google Scholar] [CrossRef]
  169. Dellasega, D.; Facibeni, A.; Fonzo, F.D. Nanostructured Ag4O4 films with enhanced antibacterial activity. Nanotechnology 2008, 19, 475602. [Google Scholar] [CrossRef] [PubMed]
  170. Waterhouse, G. Influence of catalyst morphology on the performance of electrolytic silver catalysts for the partial oxidation of methanol to formaldehyde. Appl. Catal. A Gen. 2004, 266, 257–273. [Google Scholar] [CrossRef]
  171. Derouin, J.; Farber, R.G.; Heslop, S.L.; Killelea, D.R. Formation of surface oxides and Ag2O thin films with atomic oxygen on Ag(111). Surf. Sci. 2015, 641, L1–L4. [Google Scholar] [CrossRef]
  172. Car, P.-E.; Patzke, G.R. The Fascination of Polyoxometalate Chemistry. Inorganics 2015, 3, 511–515. [Google Scholar] [CrossRef] [Green Version]
  173. Wang, S.-S.; Yang, G.-Y. Recent Advances in Polyoxometalate-Catalyzed Reactions. Chem. Rev. 2015, 115, 4893–4962. [Google Scholar] [CrossRef]
  174. Ammam, M. Polyoxometalates: Formation, structures, principal properties, main deposition methods and application in sensing. J. Mater. Chem. A 2013, 1, 6291–6312. [Google Scholar] [CrossRef]
  175. Omwoma, S.; Gore, C.T.; Ji, Y.; Hu, C.; Song, Y.-F. Environmentally benign polyoxometalate materials. Coord. Chem. Rev. 2015, 286, 17–29. [Google Scholar] [CrossRef]
  176. Narasimharao, K.; Brown, D.R.; Lee, A.F.; Newman, A.D.; Siril, P.F.; Tavener, S.J.; Wilson, K. Structure–activity relations in Cs-doped heteropoly acid catalysts for biodiesel production. J. Catal. 2007, 248, 226–234. [Google Scholar] [CrossRef]
  177. Contreras Coronel, N.; da Silva, M.J. Lacunar Keggin Heteropolyacid Salts: Soluble, Solid and Solid-Supported Catalysts. J. Clust. Sci. 2018, 29, 195–205. [Google Scholar] [CrossRef]
  178. Lotfian, N.; Heravi, M.M.; Mirzaei, M.; Heidari, B. Applications of inorganic-organic hybrid architectures based on polyoxometalates in catalyzed and photocatalyzed chemical transformations. Appl. Organomet. Chem. 2019, 33, 4808. [Google Scholar] [CrossRef]
  179. Ye, J.-J.; Wu, C.-D. Immobilization of polyoxometalates in crystalline solids for highly efficient heterogeneous catalysis. Dalton Trans. 2016, 45, 10101–10112. [Google Scholar] [CrossRef]
  180. Dufaud, V.; Lefebvre, F. Inorganic Hybrid Materials with Encapsulated Polyoxometalates. Materials 2010, 3, 682–703. [Google Scholar] [CrossRef] [Green Version]
  181. Hossain, M.N.; Park, H.C.; Choi, H.S. A Comprehensive Review on Catalytic Oxidative Desulfurization of Liquid Fuel Oil. Catalysts 2019, 9, 229. [Google Scholar] [CrossRef] [Green Version]
  182. Yan, X.; Yan, J.; Mei, P.; Lei, J. Synthesis, characterization and catalytic properties of mesoporous HPMo/SiO2 composite. J. Wuhan Univ. Technol. 2008, 23, 834–838. [Google Scholar] [CrossRef]
  183. Yan, X.; Lei, J.; Liu, D.; Guo, L.; Wu, Y. Oxidation reactivities of organic sulfur compounds in fuel oil using immobilized heteropoly acid as catalyst. J. Wuhan Univ. Technol. 2007, 22, 320–324. [Google Scholar] [CrossRef]
  184. Kaleta, W.; Nowinska, K. Immobilisation of Heteropoly Anions in Si-MCM-41 Channels by Means of Chemical Bonding to Aminosilane Groups. Chem. Commun. 2001, 535–536. [Google Scholar] [CrossRef]
  185. Wang, S.-Q.; Zhou, L.; Su, W.; Sun, Y.; Zhou, Y. Deep Desulfurization of Transportation Fuels by Characteristic Reaction Resided in Adsorbents. AIChE J. 2009, 55, 1872–1881. [Google Scholar] [CrossRef]
  186. Li, J.; Hu, B.; Tan, J.; Zhuang, J. Deep oxidative desulfurization of fuels catalyzed by molybdovanadophosphoric acid on amino-functionalized SBA-15 using hydrogen peroxide as oxidant. Transit. Met. Chem. 2013, 38, 495–501. [Google Scholar] [CrossRef]
  187. Chamack, M.; Mahjoub, A.R.; Aghayan, H. Catalytic performance of vanadium-substituted molybdophosphoric acid supported on zirconium modified mesoporous silica in oxidative desulfurization. Chem. Eng. Res. Des. 2015, 94, 565–572. [Google Scholar] [CrossRef]
  188. Wang, X.; Zhang, X.; Wang, Y.; Liu, H.; Qiu, J.; Wang, J.; Han, W.; Yeung, K.L. Investigating the role of zeolite nanocrystal seeds in the synthesis of mesoporous catalysts with zeolite wall structure. Chem. Mater. 2011, 23, 4469–4479. [Google Scholar] [CrossRef]
  189. Chamack, M.; Mahjoub, A.R.; Aghayan, H. Cesium salts of tungsten-substituted molybdophosphoric acid immobilized onto platelet mesoporous silica: Efficient catalysts for oxidative desulfurization of dibenzothiophene. Chem. Eng. J. 2014, 255, 686–694. [Google Scholar] [CrossRef]
  190. Trakarnpruk, W.; Jatupisarnpong, J. Acidic and cesium salts of polyoxometalates with and without vanadium supported on MCM-41 as catalysts for oxidation of cyclohexane with H2O2. Appl. Petrochem. Res. 2013, 3, 9–15. [Google Scholar] [CrossRef] [Green Version]
  191. Torres-Garcia, E.; Galano, A.; Rodriguez-Gattorno, G. Oxidative desulfurization (ODS) of organosulfur compounds catalyzed by peroxo-metallate complexes of WOx–ZrO2: Thermochemical, structural, and reactivity indexes analyses. J. Catal. 2011, 282, 201–208. [Google Scholar] [CrossRef]
  192. Zhuang, J.; Jin, X.; Shen, X.; Tan, J.; Nie, L.; Xiong, J.; Hu, B. Preparation of Ionic Liquid-modified SBA-15 Doped with Molybdovanadophosphoric Acid for Oxidative Desulfurization. Bull. Korean Chem. Soc. 2015, 36, 1784–1790. [Google Scholar] [CrossRef]
  193. Xiong, J.; Zhu, W.; Ding, W.; Yang, L.; Zhang, M.; Jiang, W.; Zhao, Z.; Li, H. Hydrophobic mesoporous silica-supported heteropolyacid induced by ionic liquid as a high efficiency catalyst for the oxidative desulfurization of fuel. RSC Adv. 2015, 5, 16847–16855. [Google Scholar] [CrossRef]
  194. Cai, T.-F.; Li, H.-P.; Zhao, H. A Study on the Oxidation-extraction Desulfurization of FCC Gasoline Over Supported Phosphotungstic Acid/H2O2/Dimethyl Sulfoxide (DMSO). Pet. Sci. Technol. 2014, 32, 1713–1719. [Google Scholar] [CrossRef]
  195. Yang, L.; Li, J.; Yuan, X.; Shen, J.; Qi, Y. One step non-hydrodesulfurization of fuel oil: Catalyzed oxidation adsorption desulfurization over HPWA-SBA-15. J. Mol. Catal. A 2007, 262, 114–118. [Google Scholar] [CrossRef]
  196. Estephane, G.; Lancelot, C.; Blanchard, P.; Toufaily, J.; Hamiye, T.; Lamonier, C. Sulfur compounds reactivity in the ODS of model and real feeds on W–SBA based catalysts. RSC Adv. 2018, 8, 13714–13721. [Google Scholar] [CrossRef] [Green Version]
  197. Dai, W.; Zhou, Y.; Wang, S.; Long, L.; Zhou, L. Catalytic Oxidation of Thiophenes with Air and PW/SBA-15. Ads. Sci. Technol. 2007, 25, 385–394. [Google Scholar] [CrossRef]
  198. Lu, C.; Fu, H.; Li, H.; Zhao, H.; Cai, T. Oxidation-extraction desulfurization of model oil over Zr-ZSM-5/SBA-15 and kinetic study. Front. Chem. Sci. Eng. 2014, 8, 203–211. [Google Scholar] [CrossRef]
  199. Qi, H.-X.; Zhai, S.-R.; Zhang, W.; Zhai, B.; An, Q.-D. Recyclable HPW/PEHA/ZrSBA-15 toward efficient oxidative desulfurization of DBT with hydrogen peroxide. Catal. Commun. 2015, 72, 53–56. [Google Scholar] [CrossRef]
  200. Lei, J.; Chen, L.; Yang, P.; Du, X.; Yan, X. Oxidative desulfurization of diesel fuel by mesoporous phosphotungstic acid/SiO2: The effect of preparation methods on catalytic performance. J. Porous Mater. 2013, 20, 1379–1385. [Google Scholar] [CrossRef]
  201. Ribeiro, S.O.; Granadeiro, C.M.; Almeida, P.L.; Pires, J.; Capel-Sanchez, M.C.; Campos-Martin, J.M.; Gago, S.; de Castro, B.; Balula, S.S. Oxidative desulfurization strategies using Keggin-type polyoxometalate catalysts: Biphasic versus solvent-free systems. Catal. Today 2019, 333, 226–236. [Google Scholar] [CrossRef]
  202. Ribeiro, S.O.; Nogueira, L.S.; Gago, S.; Almeida, P.L.; Corvo, M.C.; de Castro, B.; Granadeiro, C.M.; Balula, S.S. Desulfurization process conciliating heterogeneous oxidation and liquid extraction: Organic solvent or centrifugation/water? Appl. Catal. A 2017, 542, 359–367. [Google Scholar] [CrossRef]
  203. Pham, X.N.; Tran, D.L.; Pham, T.D.; Nguyen, Q.M.; van Thi, T.T.; Van, H.D. One-step synthesis, characterization and oxidative desulfurization of 12-tungstophosphoric heteropolyanions immobilized on amino functionalized SBA-15. Adv. Powder Technol. 2018, 29, 58–65. [Google Scholar] [CrossRef] [Green Version]
  204. Pham, X.N.; van Doan, H. Activity and stability of amino-functionalized SBA-15 immobilized 12-tungstophosphoric acid in the oxidative desulfurization of a diesel fuel model with H2O2. Chem. Eng. Commun. 2019, 206, 1139–1151. [Google Scholar] [CrossRef]
  205. Zhang, Y.; Wu, L.; Dong, X.; Wu, P.; Hu, H.; Xue, G. Copper(II)-Substituted Polyoxotungstates Immobilized on Amine-Functionalized SBA-15: Efficient Heterogeneous Catalysts for Liquid Phase Oxidative Reaction. Catal. Lett. 2016, 146, 2468–2477. [Google Scholar] [CrossRef]
  206. Julião, D.; Mirante, F.O.; 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]
  207. 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] [Green Version]
  208. Yuan, J.; Xiong, J.; Wang, J.; Ding, W.; Yang, L.; Zhang, M.; Zhu, W.; Li, H. Structure and catalytic oxidative desulfurization properties of SBA-15 supported silicotungstic acid ionic liquid. J. Porous Mater. 2016, 23, 823–831. [Google Scholar] [CrossRef]
  209. Xiong, J.; Zhu, W.; Ding, W.; Yang, L.; Chao, Y.; Li, H.; Zhu, F.; Li, H. Phosphotungstic Acid Immobilized on Ionic Liquid-Modified SBA-15: Efficient Hydrophobic Heterogeneous Catalyst for Oxidative Desulfurization in Fuel. Ind. Eng. Chem. Res. 2014, 53, 19895–19904. [Google Scholar] [CrossRef]
  210. Wang, L.; Fan, J.; Tian, B.; Yang, H.; Yu, C.; Tu, B.; Zhao, D. Synthesis and characterization of small pore thick-walled SBA-16 templated by oligomeric surfactant with ultra-long hydrophilic chains. Microporous Mesoporous Mater. 2004, 67, 135–141. [Google Scholar] [CrossRef]
  211. Zhao, D.; Huo, Q.; Feng, J.; Chmelka, B.F.; Stucky, G.D. Nonionic Triblock and Star Diblock Copolymer and Oligomeric Surfactant Syntheses of Highly Ordered, Hydrothermally Stable, Mesoporous Silica Structures. J. Am. Chem. Soc. 1998, 120, 6024–6036. [Google Scholar] [CrossRef]
  212. Gallo, J.M.R.; Bisio, C.; Marchese, L.; Pastore, H.O. Surface acidity of novel mesostructured silicas with framework aluminum obtained by SBA-16 related synthesis. Microporous Mesoporous Mater. 2008, 111, 632–635. [Google Scholar] [CrossRef]
  213. Shah, A.T.; Li, B.; Ali Abdalla, Z.E. Direct synthesis of Ti-containing SBA-16-type mesoporous material by the evaporation-induced self-assembly method and its catalytic performance for oxidative desulfurization. J. Colloid Interface Sci. 2009, 336, 707–711. [Google Scholar] [CrossRef]
  214. Boccuti, M.R.; Rao, K.M.; Zecchina, A.; Leofanti, G.; Petrini, G. Spectroscopic Characterization of Silicalite and Titanium-Silicalite. Stud. Surf. Sci. Catal. 1989, 48, 133–144. [Google Scholar]
  215. Araújo, R.S.; Azevedo, D.C.S.; Rodríguez-Castellón, E.; Jiménez-López, A.; Cavalcante, C.L. Al and Ti-containing mesoporous molecular sieves: Synthesis, characterization and redox activity in the anthracene oxidation. J. Mol. Catal. A Chem. 2008, 281, 154–163. [Google Scholar] [CrossRef]
  216. Eimer, G.A.; Chanquia, C.M.; Sapag, K.; Herrero, E.R. The role of different parameters of synthesis in the final structure of Ti-containing mesoporous materials. Microporous Mesoporous Mater. 2008, 116, 670–676. [Google Scholar] [CrossRef] [Green Version]
  217. Zepeda, T.A. Comparison and performance of different sulphided Ti-loaded mesostructured silica-supported CoMo catalysts in deep HDS. Appl. Catal. A 2008, 347, 148–161. [Google Scholar] [CrossRef]
Figure 1. Sketch of the general procedure for the Santa Barbara Amorphous (SBA)-15 preparation.
Figure 1. Sketch of the general procedure for the Santa Barbara Amorphous (SBA)-15 preparation.
Catalysts 09 00984 g001
Figure 2. Schematic representation of the catalytic adsorptive desulfurization systems (CADS) versus the simple adsorptive desulfurization capacity (ADS) mechanism operating during the removal of recalcitrant sulfur compounds by TiO2/SBA-15 catalysts.
Figure 2. Schematic representation of the catalytic adsorptive desulfurization systems (CADS) versus the simple adsorptive desulfurization capacity (ADS) mechanism operating during the removal of recalcitrant sulfur compounds by TiO2/SBA-15 catalysts.
Catalysts 09 00984 g002
Figure 3. Schematic representation of the ODS process by active tetrahedral Ti4+ sites in Ti-SBA-15 catalysts.
Figure 3. Schematic representation of the ODS process by active tetrahedral Ti4+ sites in Ti-SBA-15 catalysts.
Catalysts 09 00984 g003
Figure 4. Oxidation of sulfur-containing compounds by the (PW11Ti)2OH@TMS-SBA-15/H2O2 system.
Figure 4. Oxidation of sulfur-containing compounds by the (PW11Ti)2OH@TMS-SBA-15/H2O2 system.
Catalysts 09 00984 g004
Figure 5. The extractive-catalytic desulfurization oxidative cycle of VOx-SBA-15 in the presence of 1-butyl-3-methylimidazolium tetrafluoroborate.
Figure 5. The extractive-catalytic desulfurization oxidative cycle of VOx-SBA-15 in the presence of 1-butyl-3-methylimidazolium tetrafluoroborate.
Catalysts 09 00984 g005
Figure 6. Schematic representation of the oxidation of 4,6-DMDBT with MoO3/SBA-15 using the H2O2/formic acid mixture. Reactive Mo peroxometallic complex transfers the oxygen atom to the substrate, while oxygen defective α-MoO3 sites perform as specific molecular recognition sites.
Figure 6. Schematic representation of the oxidation of 4,6-DMDBT with MoO3/SBA-15 using the H2O2/formic acid mixture. Reactive Mo peroxometallic complex transfers the oxygen atom to the substrate, while oxygen defective α-MoO3 sites perform as specific molecular recognition sites.
Catalysts 09 00984 g006
Figure 7. Preparation of iron-based redox ionic liquid [pmim] FeCl4 supported on SBA-15.
Figure 7. Preparation of iron-based redox ionic liquid [pmim] FeCl4 supported on SBA-15.
Catalysts 09 00984 g007
Figure 8. Schematic representation of the oxidation of sulfur compounds in the presence of AgxO-SBA-15. Panel A: Freshly prepared catalysts in the absence of air. Panel B: Aged catalyst in the presence of air.
Figure 8. Schematic representation of the oxidation of sulfur compounds in the presence of AgxO-SBA-15. Panel A: Freshly prepared catalysts in the absence of air. Panel B: Aged catalyst in the presence of air.
Catalysts 09 00984 g008
Figure 9. Sketches of classical polyoxometalates (POMs) structures in polyhedral illustrations.
Figure 9. Sketches of classical polyoxometalates (POMs) structures in polyhedral illustrations.
Catalysts 09 00984 g009
Figure 10. Proposed mechanism for the oxidation of organosulfur compounds by ROOH as oxidant and heterogeneous HPA as catalyst.
Figure 10. Proposed mechanism for the oxidation of organosulfur compounds by ROOH as oxidant and heterogeneous HPA as catalyst.
Catalysts 09 00984 g010
Figure 11. Synthetic route for the preparation of HPA immobilized over IL-SBA-15.
Figure 11. Synthetic route for the preparation of HPA immobilized over IL-SBA-15.
Catalysts 09 00984 g011
Figure 12. HPA immobilization strategies: through amino post- synthetic derivatization of SBA-15 (above) or by surfactant mediated self-assembly (below).
Figure 12. HPA immobilization strategies: through amino post- synthetic derivatization of SBA-15 (above) or by surfactant mediated self-assembly (below).
Catalysts 09 00984 g012
Table 1. Selection of optimal catalyst modification and operating conditions for SBA-15 anchored metal containing catalysts in the oxidative desulfurization process.
Table 1. Selection of optimal catalyst modification and operating conditions for SBA-15 anchored metal containing catalysts in the oxidative desulfurization process.
Catalyst ModificationOperating Conditions aSulfur Removal (%) bCatalyst Recycle
(n° of Runs) c
Reference
TiO2/SiO2
sol-gel method
BT, DBT, MDBT, and DMDBT, UV-irrad., air, toluene/dodecane, 25 °C, 2 h.n.c.5[78]
TiO2/SBA-15
grafting insertion
BT, DBT, 4-MDBT, and 4,6-DMDBT, CHP, n-heptane/toluene, 80 °C.>99n.a.[91]
V2O5/SBA-15
wetness impregnation
Selective extraction phase DBT, 4-MDTB, 4,6-DMDTB, [Bmim]BF4, O2, 120 °C, 6 h.996[116]
CoMo/SBA-15
wetness impregnation
DBT, hexadecane, H2O2, 60 °C, 1 h.90n.a.[125]
MoO3/SBA-15
wetness impregnation
Formic acid as co-oxidant species, 4,6-DMDBT, H2O2, hexadecane, 70 °C, 1 h.99n.a.[130]
[pmim]FeCl4/SBA-15
grafting insertion
BT and DBT, H2O2, octane, 30 °C, 1 h.94.3n.a.[137]
Fe/SBA-15
wetness impregnation
BT, DBT and 4,6-DMDBT, H2O2, n-dodecane and CH3CN, 60 °C, 15 min.904[139]
WOx/SBA-15
grafting insertion
DBT, H2O2, n-dodecane, 100 °C, 2 h.>99n.a.[155]
MoV/Zr/SBA-15
wetness impregnation
DBT, TBHP, n-hexane, 60 °C, 75 min.98.54[187]
Cs2Mo8W4/SBA-15
wetness impregnation
DBT, TBHP, n-hexane, 60 °C, 80 min.>994[189]
HPMo/IL-SBA-15
wetness impregnation
DBT, H2O2, n-octane, 60 °C, 90 min.>905[193]
HPW/PEHA/Zr/SBA-15
layer by layer strategy
DBT, H2O2, n-octane, 60 °C, 60 min.956[199]
HPW/SiO2-EISA
self-assembly
BT, DBT and 4,6-DMDBT, H2O2, petroleum ether/CH3CN, 60 °C, 2 h>997[200]
PW11/H2N-SBA-15
wetness impregnation
BT, DBT, 4-MDBT and 4,6-DMDBT, H2O2, n-octane and CH3CN, 70 °C, 60 min.>998[201]
PW4/tma-SBA-15
wetness impregnation
BT, DBT and 4,6-DMDBT, H2O2, n-octane and [Bmim]PF6 or MeCN, 70 °C, 3 h>9910[206]
HSiW/IL-SBA-15
wetness impregnation
BT, DBT and 4,6-DMDBT, H2O2, n-octane, 60 °C, 2 h968[208]
TiO2/SBA-16
self-assembly
DBT, n-octane, 60 °C, 3 h>995[213]
a BT:benzothiophene; DBT: dibenzothiophene; 4-MDBT: 4-methyldibenzothiophene; 4,6-DMDBT: 4,6- dimethyl dibenzothiophene. CHP: cumene hydroperoxide. [Bmim]BF4: 1-butyl-3-methylimidazolium tetrafluoroborate. [pmim]: 1-methyl-3-(trimethoxysilylpropyl)-imidazolium. b n.c.: percentage of S removal, not calculated. c n.a.: not available.

Share and Cite

MDPI and ACS Style

Crucianelli, M.; Bizzarri, B.M.; Saladino, R. SBA-15 Anchored Metal Containing Catalysts in the Oxidative Desulfurization Process. Catalysts 2019, 9, 984. https://doi.org/10.3390/catal9120984

AMA Style

Crucianelli M, Bizzarri BM, Saladino R. SBA-15 Anchored Metal Containing Catalysts in the Oxidative Desulfurization Process. Catalysts. 2019; 9(12):984. https://doi.org/10.3390/catal9120984

Chicago/Turabian Style

Crucianelli, Marcello, Bruno Mattia Bizzarri, and Raffaele Saladino. 2019. "SBA-15 Anchored Metal Containing Catalysts in the Oxidative Desulfurization Process" Catalysts 9, no. 12: 984. https://doi.org/10.3390/catal9120984

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop