Next Article in Journal
New Insight into the Interplay of Method of Deposition, Chemical State of Pd, Oxygen Storage Capability and Catalytic Activity of Pd-Containing Perovskite Catalysts for Combustion of Methane
Next Article in Special Issue
Facile Construction of Magnetic Ionic Liquid Supported Silica for Aerobic Oxidative Desulfurization in Fuel
Previous Article in Journal
Manganese Molybdenum Oxide Micro Rods Adorned Porous Carbon Hybrid Electrocatalyst for Electrochemical Determination of Furazolidone in Environmental Fluids
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Adding Chelating Ligands on the Catalytic Performance of Rh-Promoted MoS2 in the Hydrodesulfurization of Dibenzothiophene

by
Siphumelele Majodina
*,
Zenixole R. Tshentu
and
Adeniyi S. Ogunlaja
*
Department of Chemistry, Nelson Mandela University, P.O. Box 77000, Gqeberha 6031, South Africa
*
Authors to whom correspondence should be addressed.
Submission received: 15 October 2021 / Revised: 11 November 2021 / Accepted: 14 November 2021 / Published: 18 November 2021
(This article belongs to the Special Issue Frontiers in Heterogeneous Catalysts for Desulfurization of Fuel Oil)

Abstract

:
Hydrodesulfurization (HDS) is a widely used process currently employed in petroleum refineries to eliminate organosulfur compounds in fuels. The current hydrotreating process struggles to remove organosulfur compounds with a steric hindrance due to the electronic nature of the current catalysts employed. In this work, the effects of adding chelating ligands such as ethylenediaminetetraacetic acid (EDTA), citric acid (CA) and acetic acid (AA) to rhodium (Rh) and active molybdenum (Mo) species for dibenzothiophene (DBT) HDS catalytic activity was evaluated. HDS activities followed the order of RhMo/ɣ-Al2O3 (88%) > RhMo-AA/ɣ-Al2O3 (73%) > RhMo-CA/ɣ-Al2O3 (72%) > RhMo-EDTA/ɣ-Al2O3 (68%). The observed trend was attributed to the different chelating ligands with varying electronic properties, thus influencing the metal–support interaction and the favorable reduction of the Mo species. RhMo/ɣ-Al2O3 offered the highest HDS activity due to its (i) lower metal–support interaction energy, as observed from the RhMo/ɣ-Al2O3 band gap of 3.779 eV and the slight shift toward the lower BE of Mo 3d, (ii) increased Mo-O-Mo species (NMo-O-Mo ~1.975) and (iii) better sulfidation of Rh and MoO in RhMo/ɣ-Al2O3 compared to the chelated catalysts. The obtained data provides that HDS catalytic activity was mainly driven by the structural nature of the RhMo-based catalyst, which influences the formation of more active sites that can enhance the HDS activity.

1. Introduction

Deleterious refractory organosulfur compounds in fuel oils have contributed to SOx emissions [1,2]. Therefore, it has become very important to remove these compounds due to the introduction of the strict environmental regulations of the Euro V limits of 10 ppmS in diesel fuels [1,2].
Hydrodesulfurization (HDS) is the most commonly used technology to produce clean fuels by employing hydrotreating catalysts, mainly from Co(Ni)/Mo oxides supported on alumina [3,4,5,6,7,8,9,10,11]. The production of ultra-low fuels using the HDS process at present requires extreme and expensive operating conditions, viz., high temperatures, hydrogen and highly active catalysts.
However, other methods of fuel desulfurization have been reported, and these techniques are bio-desulfurization (BDS), oxidative-desulfurization (ODS), adsorptive and extractive desulfurization (ADS and EDS) [1,2]. Bio-desulfurization (BDS), which involves the use of sulfur-consuming bacteria to desulfurize fuels, is nonetheless limited in meeting very deep desulfurization, as 50–200 ppmS have been reported [1]. The ODS process of recalcitrant sulfur-containing compounds such as dibenzothiophene have been reported to produce nonessential side products during oxidation [2]. Adsorptive and extractive desulfurization (ADS and EDS) is a process which employs the use of use of solid sorbents and extraction solvents for removing the recalcitrant sulfur-containing compounds, respectively. Adsorbents have been limited in organosulfur compound selectivity, as well as low adsorption capacity, while the organosulfur extraction solvents such as dimethylsulfoxide (DMSO) and N,N-dimethylformamide (DMF) have high boiling points, and hence the solvent recovery may be impossible [1,2]. The use of ODS, BDS, ADS and EDS desulfurization techniques are currently limited, as they may pose problems for larger scale applications [1,2]. As a result, more attention is required to improve the HDS system by redesigning the current HDS catalysts’ inadequacies and by introducing chelating ligands and precious metals (PGMs), which could drive HDS under mild conditions [3,4,5,6,7,8,9,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26].
Several transition metal sulfides have been reported as possible candidates for the HDS catalysts [15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30]. Rhodium-based catalysts have exhibited promising properties, hence showing great potential as an HDS catalyst. A few HDS studies of dibenzothiophene (DBT) over Rh-based catalysts have been reported. Lee et al. [27] reported that the RhCs/Al2O3 catalyst was more active than a conventional CoMo/Al2O3 catalyst for hydrotreating dibenzothiophene. RhCs/Al2O3-catalyzed DBT hydrodesulfurization was mainly controlled by the DDS (direct desulfurization) mechanism. Similarly, a synergetic effect was reported with the RhMo/ɣ-Al2O3 catalysts, suggesting that Rh and Mo interact when employed for DBT HDS [28]. While there are reports on the use of RhMo-based catalysts [27,28], studies with regards to the influence of chelating ligands, especially acetic acid (AA), on hydrotreating activity is lacking. Chelating ligands are molecules with two or more donor atoms available to bind a metal cation, and they have been reported to improve hydrotreating activity [15,17,29,30,31].
The scientific novelty of the research is to arrive at a fundamental understanding of the nature of the sulfur tolerance of the supported Rh-Mo catalysts (chelated and unchelated RhMoS/ɣ-Al2O3), and to offer clarification of the synergetic effect of chelating ligands such as ethylenediaminetetraacetic acid (EDTA), citric acid (CA) and acetic acid (AA) on the individual metal, the Rh and the Mo components in the catalytic hydrodesulfurization of DBT. The originality of the research was in the design and the application of the nanostructured RhMo catalysts in the presence of the chelating ligands so to ensure the uniform composition of the catalytic metal species for ease of physico-chemical characterization and the fundamental understanding of the structure-activity relationships. In this study, the HDS catalytic activity of the RhMo-based catalysts were carried out on dibenzothiophene (DBT). The as-synthesized catalysts were characterized using powder X-ray diffraction (PXRD), ultraviolet-visible diffuse reflectance spectroscopy (UV-vis DRS), Fourier transform infrared spectroscopy (FT-IR), scanning electron microscope (SEM), energy-dispersive X-ray spectroscopy (EDX), X-ray photoelectron spectroscopy (XPS), transmission electron microscopes (TEM) and thermogravimetric analysis (TGA)—differential scanning calorimetry (DSC) (TGA-DSC) to determine the catalysts’ bulk chemical compositions, morphologies and thermal stability.
The main findings are as follows:
(1) The prepared catalysts presented band gaps of 3.779 eV (RhMo/ɣ-Al2O3), 4.341 eV (RhMo-EDTA/ɣ-Al2O3), 4.394 eV (RhMo-AA/ɣ-Al2O3) and 4.478 eV (RhMo-CA/ɣ-Al2O3), respectively.
(2) The introduction of different chelating ligands increases the metal–support interaction, which prevents the formation of easily reduced Mo species.
(3) The HDS activity decreased in the following order: RhMo/ɣ-Al2O3 (88%) > RhMo-AA/ɣ-Al2O3 (73%) > RhMo-CA/ɣ-Al2O3 (72%) > RhMo-EDTA/ɣ-Al2O3 (68%).
(4) The observed catalytic results were ascribed to the introduction of different ligands, thus increasing the metal–support interaction and increasing the e-charge transfer from the valance band Rh 4d orbital to the conduction band of the Mo species. This led to the excessive weakening of the Mo-S bond by inhibiting the absorption of sulfur (S) compound (DBT) on the active sites, hence leading to a reduced activity.

2. Results and Discussion

2.1. UV-Vis Spectroscopy

UV-vis spectroscopy was applied to study RhMo/γ-Al2O3, RhMo-EDTA/γ-Al2O3 and RhMo-CA/γ-Al2O3, and are shown in Figure 1. As shown in Figure 1, the UV-vis spectra recorded for RhMo/γ-Al2O3 and RhMo(x)/γ-Al2O3 (x = EDTA, CA and AA) exhibited a broad absorption band at 210–290 nm and were assigned to the O2−→Mo6+ ligand–metal charge transfer transitions in an octahedral environment [32]. A weak band observed in the region of 340–360 nm with RhMo/γ-Al2O3 was associated with Rh(III) oxides, and another weak broad band at a visible region displayed at 452 nm was due to the presence of Rh(III) in oxide form [33]. A weak absorbance at 530 nm provided strong evidence of well-dispersed octahedral Rh oxide species [34,35], which are known to be easily reduced and sulfided [36,37,38]. A band around 550–680 nm on RhMo/γ-Al2O3 was due to Rh3+ interacting with γ-Al2O3 support to form a rhodium aluminate complex [32]. With the addition of the ligands, the absorbance of Rh3+ shifts, as do the weak broad bands around 430 nm, 410–470 nm and 420 nm associated with the metal to ligand transfer, and Rh-O species for RhMo-AA/γ-Al2O3, RhMo-EDTA/γ-Al2O3 and RhMo-CA/γ-Al2O3 were observed. The observed shifts to the lower wavelength suggested that the Rh–ligand complexes inhibit the formation of the Rh-γ-Al2O3 phase [32,39]. The shift to a lower wavelength also indicated a decreased agglomeration of the Mo species [32,40,41].

2.2. Band Gaps of RhMo Catalysts

The Eg value obtained from the Tauc and Davis–Mott Equation (1) demonstrates the dispersion of the Mo species. The band gaps for RhMo/γ-Al2O3, RhMo-EDTA/γ-Al2O3, RhMo-CA/γ-Al2O3 and RhMo-AA/Al2O3 were determined and are displayed in Figure S1. According to the literature, the higher the Eg value, the more improved the dispersion of the Mo species [42]. The obtained band gap for RhMo/Al2O3 was 3.779 eV (Figure S1), the Eg value obtained for RhMo-EDTA/γ-Al2O3 was 4.341 eV, the Eg value RhMo-AA/γ-Al2O3 was Eg = 4.394 eV and the Eg value for RhMo-CA/γ-Al2O3 was 4.478 eV. The RhMo-CA/γ-Al2O3 catalyst exhibited the highest Eg value, which implies a decrease in the average particle size and an increased charge transfer [41,42,43,44].
( α h ν ) 1 n = A   ( h ν E g )
where α is the absorption coefficient, hν is the incident photon energy, A is the proportionality constant, Eg is the optical band gap energy and n represents the nature of the electronic transition (n = 1/2 for direct transition).
The bridging Mo-O-Mo bonds, which determine the degree of polymerization/aggregation of Mo(VI), were determined from the Eg values by using the formula (NMo-O-Mo = 11.8 − 26Eg) presented by Tian et al. (2010) [42]. It was established that the higher Eg values of the catalysts corresponded to the lower average number (of covalent bridging of the central Mo6+ cation) nearest to the Mo6+ neighbours (Table 1), thus confirming the Mo(VI) cation structural variations in the catalysts. RhMo/γ-Al2O3 was reported to offer more polymeric/aggregated Mo species compared to the chelated catalysts (RhMo-AA/γ-Al2O3, RhMo-EDTA/γ-Al2O3 and RhMo-CA/γ-Al2O3).

2.3. Fourier Transform Infrared Spectroscopy (FT-IR)

FT-IR bands at 3600–2800 cm−1 correspond to the OH (from H2O) stretching. At the region between 550 cm−1, the corresponding band was assigned to the Mo-O-Mo bridge stretching, while the bands that were found, and the Mo=O stretching, were located at 950 cm−1 [45]. RhMo-EDTA/Al2O3 showed an absorption band at 978 cm−1, which could be assigned to the Mo-N band [35,45]. At 1588 cm−1, the absorption band could be assigned to -COO vibrations with H2O [46] (Figure S2).

2.4. Energy Dispersion Spectroscopy (EDX)

The EDS of the sulfided RhMo/Al2O3, RhMo-EDTA/Al2O3, RhMo-AA/Al2O3 and RhMo-CA/Al2O3 catalysts confirmed that the catalysts are made up of Rh, Mo, O, S, C and Al (Figure S3). The peaks at ~2.37, ~2.7, ~2.81 and ~3.28 KeV corresponded to the theoretical Lα, Kα and Kβ of Rh, respectively. The peaks at ~2.3 and 2.81 KeV corresponded to the theoretical Lα, Kα and Kβ of Mo, and the O peak was obtained at ~0.5 KeV and the S peak at ~2.3 KeV. The additional peak at ~1.5 KeV corresponded to Al from the support, and the presence of carbon (~0.28 KeV) was due to the carbon tape that was used for the sample analysis [47]. Table 2 illustrates the qualitative atomic percentage of the present elements for each HDS catalyst. The S/Mo atomic ratio for the catalysts are around (0.47–4.63) and the amounts of carbon (5.46 ≤ C/Mo ≤ 31.88) are found in all of the MoS2 catalysts. The main source of carbon is most probably the heptane solvent [48,49].

2.5. X-ray Diffraction (XRD)

The XRD analysis was performed to identify the diffraction phases and dispersion of the synthesized RhMo/γ-Al2O3 and RhMo-x/γ-Al2O3 (x = EDTA, acetic acid (AA), citric acid (CA)) catalysts. Figure 2 showed that all of the RhMo catalysts in the oxide phase had similar diffraction patterns at 2θ = 19.6°, 32.0°, 37.6°, 39.5°, 45.5°, 60.9° and 67.0°, and were assigned to (220), (311), (222), (400), (511) and (440), characteristic of the γ-Al2O3 face-centered cubic phase, respectively. For the RhMo/ɣ-Al2O3 catalyst, more patterns were observed at 2θ = 12.1°, 18.5° and 28.5°, which were due to the orthorhombic MoO3 crystalline phase [50], and the diffraction pattern at 49.0° could be ascribed to the monoclinic crystalline phase of MoO3. The pattern observed at 2θ = 34.1° could be attributed to Rh2O3 phase, and diffraction patterns at 56.0° and 57.8° were also observed. All of the chelated catalysts showed the characteristic reflections of alumina supports and very weak reflection peaks for RhMo-CA/ɣ-Al2O3, which indicated that the addition of the chelating agent could promote the redispersion of the bulk MoO3 [40,51,52]. The broadness and the amorphous nature of the diffraction pattern observed in the chelated catalysts indicated the absence of crystalline MoO3 (RhMo-CA/ɣ-Al2O3). Sulfided RhMo/ɣ-Al2O3 presented a hexagonal MoS2 phase at 14.8°, 29.5°, 33.3°, 38.5° and 60.4°. According to the obtained result, it was shown that the addition of the chelating ligands resulted in the better dispersion of molybdenum oxide [53]. Additional peaks attributed to the rhodium sulfide phase (Rh2S3 and/or Rh3S4), with characteristic peaks at 2Ɵ = 36–42, were detected on all the sulfided RhMo-x/ɣ-Al2O3, with RhMo/ɣ-Al2O3 exhibiting more characteristic peaks compared to RhMo-x/ɣ-Al2O3. The broad diffraction peaks of the sulfided catalysts (Figure 2b–d) compared to RhMo/ɣ-Al2O3 (Figure 2a) showed bulk and relatively smaller crystallite sizes. The RhMo catalysts’ crystallite sizes decreased in the order of RhMo/ɣ-Al2O3 (5.903 nm) > RhMo-CA/ɣ-Al2O3 (5.809 nm) > RhMo-EDTA/ɣ-Al2O3 (5.770 nm) > RhMo-AA/ɣ-Al2O3 (5.750 nm).

2.6. XPS Analysis

The XPS survey spectrum of sulfided RhMo/ɣ-Al2O3 and RhMo-EDTA/ɣ-Al2O3 catalysts with the detected species, viz., S 2p, Rh 3d and Mo 3d, are presented in Figure 3 and Figure 4, respectively. The survey scan spectrum, shown in Figure 3a and Figure 4a, demonstrated the presence of the key elements, O 1s, S 2p, C 1s, Al 2p and 2s, Rh 3d and Mo 3d in the catalysts. The highly resolved measurements of these individual elements of O 1s, C 1s, Al 2p, and Al 2s are demonstrated in Figures S3a–d and S4a–d for the RhMo/ɣ-Al2O3 and RhMo-EDTA/ɣ-Al2O3 catalysts. The binding energies of these elements are shown in Table 3, and these energies signify the presence of the elements on the catalysts. The XPS result of the Rh 3d peaks for RhMo/ɣ-Al2O3 showed doublets at the binding energies of 306.5 and 310.6 eV, and for RhMo-EDTA/ɣ-Al2O3, the Rh 3d binding energies were obtained at 305.0 and 310.1 eV, respectively [54,55]. Mo 3d showed three characteristic peaks observed at 226, 230.0, and 233.2 eV for RhMo/ɣ-Al2O3 and two visible peaks at 226 and 230.0 eV for RhMo-EDTA/ɣ-Al2O3, respectively.
To investigate the different phases within the samples for Rh 3d and Mo 3d for the sulfided RhMo/ɣ-Al2O3 and RhMo-EDTA/ɣ-Al2O3, the spectra were carefully deconvoluted and the obtained results are presented in Figure 5a–d. The rhodium oxide (Rh2O3) showed characteristic peaks between 307–310 for Rh 3d5/2, and for Rh 3d3/2 showed characteristic peaks at a region between 312–315 eV [52,56], and the corresponding results are presented in Figure 5a,c. The Rh2S3 phase showed a doublet at the binding energies of 307–309.2 and 312.7–314 eV in the sulfided RhMo/ɣ-Al2O3 and RhMo-EDTA/ɣ-Al2O3 samples (Table S2), corresponding to the Rh 3d5/2 and Rh 3d3/2 states for rhodium sulfide, respectively [54,55,56]. The Mo 3d was comprised of three main peaks with oxidation states of +V (oxide), +V (oxysulfide) and +IV (sulfide) [57,58]. Figure 5b,d shows the deconvolution of the Mo 3d spectra. In the case of Mo, its deconvolution consisted of Mo4+ (228.4–229.1 eV, sulfide MoS2), Mo5+ (229.7–230.5 eV, oxysulfide MoSxOy) and Mo6+ (232.1–232.7 eV, oxide MoO3) [59].
The doublet at a binding energy (BE) of (±) 229 eV and (±) 232.1 0.1 eV was attributed to Mo 3d5/2 and Mo 3d3/2 levels of MoS2 (Mo4+), and the two contributions observed at (±) 230.0 eV and (±) 235.3 0.1 eV were assigned to Mo 3d5/2 and Mo 3d3/2 of Mo oxysulfide (MoOxSy, Mo5+) [59], while the binding energy of the Mo 3d5/2 component located at (±) 232.2 0.1 eV was assigned to the Mo6+ (MoOx) species [60], and for the Mo 3d3/2 energy level, the binding energy was (±) 236.0 eV, respectively [61]. The peak presented at 226.3 and 226.1 eV was ascribed to the S 2s level of sulfur (Figure 5e,f). The analysis results, including the detailed binding energies and the sulfidation degree of the Mo species obtained by the deconvolution, are shown in Table S1, and the different phase compositions of the catalysts are calculated from the area of the deconvoluted peaks. Figure 5e,f displays a contribution at 162.1 eV at the lower binding energy of the S 2p peak originating from the S2− precursor [62], and it was clearly visible in both catalysts. The S 2p region suggests the existence of S2−, S22− and SO42− species [63,64]. The peaks at 162.3 and 163.5 eV are assigned to the S2− in the 2p3/2 and 2p1/2 levels, respectively, in MoS2 [40,51,63,65], and the characteristic peak in S 2p at 166.7 eV was attributed to SO42− [66]. The sulfidation degree of the Mo species was calculated by the following Formula (2):
[ Mo 4 + ] ( % ) = A Mo 4 + A Mo 4 + + A Mo 5 + + A Mo 6 + × 100 %
where [Mo4+] is the sulfidation degree of the samples, and AMo4+, AMo5+ and AMo6+ are the areas of the peaks which are assigned to the Mo4+, Mo5+ and Mo6+ species, respectively [67,68].
A catalyst with a higher sulfidation degree (MoS2) would suggest that there was a lower metal–support interaction with the active metal, bringing about easier catalyst reduction and sulfidation [67]. The slight shift toward the lower BE of Mo 3d in the RhMo/ɣ-Al2O3 catalyst (Table S1) could be attributed to a weaker metal–support interaction (caused by electron effects of the defects at the surface on the alumina support), which enhances the HDS catalytic activity [68,69].

2.7. Transmission Electron Microscopy (TEM)

Information on active MoS2 crystallite dispersion and sizes in the sulfided RhMo/ɣ-Al2O3 and RhMo-x/ɣ-Al2O3 (where x = AA, EDTA, CA) was obtained by means of TEM measurement. Figure 6a–d shows the distribution of MoS2 crystallites in the sulfided catalysts with and without the chelating agents, and their statistical distribution results for the length of the MoS2 slab. The addition of the chelating agents influenced the particle size distribution. The RhMo-CA/ɣ-Al2O3 had the lowest average diameter (1.86 nm) and the RhMo/ɣ-Al2O3 resulted in the highest average diameter (4.72 nm). The decrease of the MoS2 slabs due to the chelating ligand’s decomposition improved the dispersion of the active phase [70,71]. The average slab length in diameter observed for the catalysts are shown in Table 4, and the average slab length of the MoS2 slab with the highest frequency was distributed between 2.5–6.0 nm. Along with these crystallites, there were a few regions with big agglomerations of molybdenum sulfide, as confirmed by the strings of highly stacked crystallites.
The results for the MoS2 slab average diameter in Table 4 indicated that the MoS2 dispersion obtained for the RhMo catalysts decreased in the order of RhMo-CA/ɣ-Al2O3 > RhMo-AA/ɣ-Al2O3 > RhMo/ɣ-Al2O3 > RhMo-EDTA/ɣ-Al2O3. The higher dispersion for the chelated catalysts was due to the complexation of metal-chelating ligand, reducing the metal–support interaction and leading to the delay of the sulfidation of the metals. The high MoS2 dispersion could facilitate the generation of more active sites [72,73,74].

2.8. Scanning Electron Microscopy (SEM)

The SEM images of RhMoOx are shown in Figure 7a,c,e,g for RhMo/ɣ-Al2O3 and RhMo-x/ɣ-Al2O3 (x = EDTA, AA, CA). The images showed that particles are closely spherical in shape with an average uniform distribution. All the samples had an average particle size (82.5–102.6 μm). Figure 7b,d,f,h represents the sulfided RhMo/ɣ-Al2O3 and RhMo-x/ɣ-Al2O3 (x = EDTA, AA, CA) catalysts, and the zoomed images show that all the chelated catalysts highly agglomerated with spherical-like materials with fluffy-like particles, which could indicate that the catalysts are porous in nature. The particle distributions for the sulfided catalysts could not be measured due to the high agglomeration of the particles.

2.9. Stability of Catalysts—TGA and DSC Thermal Analyses

TGA-DSC is a technique used for thermal analysis to characterize materials by measuring their change in mass as a function of temperature. It is coupled with DSC to provide complementary information such as measuring the heat flow as a function of time and temperature at a controlled environment.
RhMo/ɣ-Al2O3: The first weight loss (2.5%) for RhMo/ɣ-Al2O3 occurred in the range of 50–150 °C due to the desorption of the physically adsorbed water from the surface of the catalyst, and this was accompanied by a broad exothermic peak in the range of 120–210 °C (Figure 8a). A second weight loss of 3% was observed between 200–580 °C, and it was reflected by a very weak endothermic peak between 360–610 °C, associated with the decomposition of nitrate radical, hexaammonium molybdate and dihydroxylation [75]. An endothermic peak at 810 °C was observed, and it was attributed to the formation of a stable MoO3 phase.
RhMo-EDTA/ɣ-Al2O3: The first weight loss (5.2%) in Figure 8b was below 100 °C and was mainly due to water desorption. From 160–800 °C, there was a gradual weight loss of 3% and not many events were happening in those stages. This weight loss was associated with the loss of the complex, and a partial dehydration–decomposition of the Rh and Mo species of RhMo-EDTA/ɣ-Al2O3. The first and second weight losses were accompanied by a broad exothermic peak at a maximum of 180 °C, corresponding to the decomposition of the complex. A broad exothermic peak between 280–670 °C was observed, and was due to the decomposition of EDTA and the further combustion of the residual organic matrix [76], and to the total transformation of the partially decomposed Rh and Mo precursor species into the catalyst oxidic precursor [77].
RhMo-AA/ɣ-Al2O3: The results obtained for RhMo-AA/ɣ-Al2O3 (Figure 8c) showed a weight loss (3.6%) taking place in the range of 50–150 °C followed by an exothermic peak 90 °C, which was mainly due to water desorption. The subsequent weight loss of 2.8% between 180–410 °C was attributed to the dehydration–decomposition of precursor species and the partial dehydroxylation of alumina [77]. A second exothermic peak occurred at a maximum of 280 °C, which corresponded to the decomposition of the complex (metal–AA) and the total decomposition and partial dehydration–decomposition of Rh and Mo precursor species. A third weight loss (1.8%) between 410–850 °C was due to the formation of the monometallic oxidic precursor. The DCS curve displayed an endothermic peak at a maximum of 580 °C.
RhMo-CA/ɣ-Al2O3: The results obtained for RhMo-CA/Al2O3 (Figure 8d) showed a weight loss of 5.2% below 150 °C, followed by an endothermic peak of 80 °C, was mainly due to H2O removal. The second weight loss of 2.8% between 180–400 °C was attributed to a decomposition and combustion of the precursor species, and the complete breakdown of citric acid [46]. A broad and weak exothermic peak occurred at a maximum of 400 °C, which corresponded to the decomposition of the remaining complex (metal–CA) and the total decomposition and partial dehydration–decomposition of Rh and Mo precursor species. Above 400 °C, not much loss of weight loss was observed, and this indicated the formation of stable metallic oxidic precursors.

2.10. Catalytic Activity

The conversion of dibenzothiophene (DBT) was used to estimate the catalytic activity in HDS (Equation (1)). RhMo/ɣ-Al2O3 (88%) had the highest catalytic activity, and the activity for the chelated catalysts followed this order: RhMo-AA/ɣ-Al2O3 (73%) > RhMo-CA/ɣ-Al2O3 (72%) > RhMo-EDTA/ɣ-Al2O3 (68%) (Table 5). The observed catalytic results were ascribed to the introduction of the different ligands, which increased the metal–support interaction and increased the e-charge transfer (energy band gap) from the valance band Rh 4d orbital to the conduction band of the Mo species. The absence led to a weaker Mo-S bond strength, a higher concentration of CUS and a higher HDS activity [60]. Crystallite sizes were also observed to influence the catalytic activity, as RhMo/ɣ-Al2O3 (with crystallite size of 5.903 nm) presented the highest activity, and this may be due to the formation of bigger MoS2 crystals when compared to others. A combined electron donating effect of the chelates, and the crystallite sizes of MoS2, may have influenced the chelated catalyst activity.
The chelated RhMo/ɣ-Al2O3 catalysts resulted in slightly lower catalytic activity due to the formation of rhodium-chelating ligands and a molybdenum–chelate complex. For all the catalysts, DBT converted mainly via the DDS pathway (Table 5). There was not much difference obtained in terms of selectivity when comparing the HYD/DDS selectivity ratio. RhMo/ɣ-Al2O3 showed a slightly higher HYD/DDS ratio of ~0.20 when compared to the chelated catalysts (see GC chromatogram, Figures S5–S7). The addition of the chelating ligands showed a slight difference, and therefore we can conclude that the addition of the chelating ligand on the catalysts did not have much influence on the selectivity (Table 5).
The HDS selectivity correlated linearly with the slab length of the MoS2 phase (TEM), the longer slab length indicated a high ratio of edge/corner and better HDS selectivity, with RhMo/ɣ-Al2O3 presenting the longest slab length [64,77], and the edge sites only catalysed the HDS reaction [78]. The values presented by the current RhMo catalysts (Table 6) exhibit certain benefits and compare well with the other catalysts reported in the literature [79,80,81,82,83,84]. We concluded that the HDS % conversion and desulfurization route was influenced by the catalyst composition, the electronic properties and the HDS reaction conditions.

2.11. Proposed Mechanism

The large energy gap (Eg) values of the chelated ligands RhMo-AA/ɣ-Al2O3 (4.341 eV), RhMo-EDTA/ɣ-Al2O3 (4.394 eV) and RhMo-CA/ɣ-Al2O3 (4.478 eV), supported the increased charge transfer of Rh, chelates and Mo species catalysts when compared to RhMo/ɣ-Al2O3 (3.779 eV). However, according to Figure 5a–d, the higher amount of Mo-S/RhMo-S phases are formed in the absence of a chelating ligand, owing to the electron transfer between the Rh and Mo-phase. Furthermore, the BE of O 1s increases upon chelation, confirming less neutralization of the surface O-H in ɣ-Al2O3 (O 1s 531.08 eV). The O 1s of RhMo/Al2O3 is 529.139 eV, and for RhMo-EDTA/Al2O3 is 529.280 eV (Figure 9).
Thus, this implies that ligand presence decreases the neutralization of the surface Brønsted acid site OHδ+ of ɣ-Al2O3 (an observed increase in the oxygen binding energies), and this prevents the formation of Mo-S/RhMo-S bonds, but more Rh-S bonds are formed due to the higher charge transfer between Rh and the chelating ligands (see Figure 5c,d). Generally, the free electrons promoted catalytic activities by donating electrons to the conduction band of the Mo species, thus promoting Mo-S/RhMo-S bond cleavage to form more coordinatively unsaturated sites (CUS) (Scheme 1). In a typical HDS reaction, the Rh2O3 phase is reduced to metallic rhodium (BE of Rh 3d5/2 at 307.0–307.1 eV, predominately observed in RhMo/ɣ-Al2O3 [55]) and Rh2O3-x(RhO), with unpaired free electrons in the Rh 4d orbital and oxygen vacancies transferred to the conduction band of Mo species, thereby promoting the cleavage of Mo-S/RhMo-S to form the CUS. The increased charge transfer (energy gap (Eg) values) in chelated RhMo/ɣ-Al2O3 between Rh and Mo species led to the excessive weakening of the Mo-S bond, preventing S-compound absorption on the active sites, thereby leading to reduced activity [84].

3. Experimental Section

3.1. Materials

All chemicals used were obtained from Merck/Sigma-Aldrich, South Africa. These include rhodium(III) chloride (98%), ammonium heptamolybdate (99%), ethylenediaminetetraacetic acid (EDTA, 97%), citric acid monohydrate (CA, 99.5%), acetic acid (AA, 99%), heptane, dibenzothiophene (98%) and gamma alumina support (ɣ-Al2O3).

3.2. Synthesis of RhMo Catalysts Prepared with Ethylenediaminetetraacetic Acid (EDTA), Citric Acid (CA) and Acetic Acid (AA)

Rh(x)Mo(y) catalysts were prepared by wet impregnations of the precursor salts [28]. Unchelated catalyst: Rh from RhCl3 (0.0421 g, 2 × 10−4 mol) and Mo from (NH4)6Mo7O24·4H2O (0.496 g, 4 × 10−4 mol) were added in 30 mL deionized water to obtain the desired metal content ratio (Rh/Rh + Mo) 0.3, and the pH adjusted to pH = 9. The solution was added to γ-alumina (1 g), and the resulting mixture was transferred to a 100 mL Teflon-lined stainless steel autoclave, and hydrothermally treated at 453 K for 4 h. The resulting solid was filtered, dried at 393 K for 12 h and calcined at 773 K for 4 h to obtain the RhMo oxide on γ-alumina.
Chelated catalysts: Generally, for the synthesis of chelated catalyst, molar ratios 1:2 Rh to chelates (EDTA, CA and AA) was employed. RhMo-AA/γ-Al2O3: A mixture of RhCl3 (0.0421 g, 2 × 10−4 mol), (NH4)6Mo7O24·4H2O (0.496 g, 4 × 10−4 mol) and AA (0.0241 g, 4 × 10−4 mol) was dissolved in 20 mL H2O solution. RhMo-CA/γ-Al2O3: A mixture of RhCl3 (0.0421 g, 2 × 10−4 mol), (NH4)6Mo7O24·4H2O (0.496 g, 4 × 10−4 mol) and AA (0.0770 g, 4 × 10−4 mol) was dissolved in 20 mL H2O solution. RhMo-EDTA/γ-Al2O3: A mixture of RhCl3 (0.0421 g, 2 × 10−4 mol), (NH4)6Mo7O24·4H2O (0.496 g, 4 × 10−4 mol) and EDTA (0.0745 g, 2 × 10−4 mol) was dissolved in 20 mL H2O solution. In all of the solution mixtures, the molar ratio with Rh/Rh + Mo molar ratio of 0.3 was added to γ-Al2O3 (1 g, calcined at 500 °C) and the mixture was stirred for 4 h and pH adjusted to 9. The resulting solid was dried at 120 °C overnight to obtain RhMo-AA/γ-Al2O3, RhMo-CA/γ-Al2O3 and RhMo-EDTA/γ-Al2O3. For the chelated catalysts, RhMo-AA/γ-Al2O3, RhMo-CA/γ-Al2O3 and RhMo-EDTA/γ-Al2O3 were only treated at 120 °C to preserve chelating ligands (CA, AA and EDTA) in catalyst until the activation stage [46,50].

3.3. Catalyst Characterization

Ultraviolet-visible diffuse reflectance spectroscopy (DRS) and band gap energies of the catalysts were processed from a Shimadzu UV-vis DRS spectrophotometer UV-3100 UV-vis spectrophotometer from a wavelength range from 200 to 800 nm.
FT-IR spectroscopy of the catalysts was acquired using a Bruker Tensor 27 platinum ATR-FTIR spectrometer (wavelength range from 4000 to 400 cm−1).
Thermogravimetric analysis (TGA-DSC) was measured using a Perkin Elmer STA 6000 at a heating range of 55 to 900 °C at 20 °C/min with N2 flow of 30 mL min−1.
X-ray powder diffraction (XRD) analysis was carried out on a Bruker D2 powder X-ray diffractometer using Cu radiation with a LynxEye detector with a scan range of 5 to 80° 2 theta.
Milled samples (RhMo/γ-Al2O3, RhMo-AA/γ-Al2O3, RhMo-CA/γ-Al2O3 and RhMo-EDTA/γ-Al2O3) were gold coated and imaged for morphological evaluation using a JOEL 7001f scanning electron microscope (SEM). JEOL JEM-2010 transmission electron microscope (TEM) operated at 200 kV was employed for TEM imaging.
X-ray photoelectron spectrometer (XPS) was performed on a Kratos Axis Ultra X-ray photoelectron spectrometer equipped with a monochromatic Al Kα source (1486.6 eV).

3.4. Catalyst Sulfidation and Hydrodesulfurization Measurements

The sulfidation and HDS tests were carried in a 2 L Parr pressure reactor 4842 (350 bar, max tem = 425 °C). For catalyst sulfidation: HDS catalysts (2 g, mol/mL ratio of Mo:DBT (~1:100)) were pressurized to 4.0 MPa (40 bar) in a 100 mL heptane solution containing 10 wt.% of CS2 (sulfiding agent) under hydrogen flow (40 mL/min) for 4 h and at 573 K to ensure complete sulfidation.
After sulfidation, the reactor was cooled down to room temperature for the dibenzothiophene HDS test. At this temperature, the liquid feed was switched to dibenzothiophene solution (0.22 g, 1.194 × 10−5 mol/mL). The temperature was adjusted to 573 K under H2 pressure of 4.0 MPa (40 bar) and maintained for 6 h. The dibenzothiophene content was measured with an Agilent 6890 gas chromatograph equipped with a FID detector and a 30 m × 0.25 mm × 0.25 μm capillary column (ZB-5MSi, 5% Phenyl column). HDS catalytic activity was estimated using (3):
X HDS ( % ) = C DBT 0 C DBT C DBT 0 × 100
where C DBT 0 is the DBT content in the feedstock (wt.%) and CDBT is the DBT content in the products (wt.%) [29]. Therefore, the catalytic selectivity ratio between the hydrogenation (HYD) and direct desulfurization (DDS) is estimated from Equation (4):
S HYD / S DDS   = C x C DBT 0 C DBT  
where Cx is the content of phenylcyclohexane (PhCh) or biphenyl (BP) [30].

4. Conclusions

In the present work, a series of new catalysts (RhMo/Al2O3, RhMo-x/Al2O3 (where x = EDTA, AA, CA) were successfully synthesized and characterized. The UV-vis analysis confirmed the presence of octahedral molybdate species between 320–360 nm for RhMo/Al2O3, and the shift to a lower wavelength in the visible part of 220 nm for the chelated catalysts was observed and indicated the formation of less polymerized molybdate species and heteropoplymolybdates. Band gaps of 3.779 eV (RhMo/ɣ-Al2O3), 4.341 eV (RhM-EDTA/ɣ-Al2O3), 4.394 eV (RhMo-AA/ɣ-Al2O3) and 4.478 eV (RhMo-CA/ɣ-Al2O3) were obtained. The TEM imaging confirmed that the materials had fringe-like morphologies, deemed as MoS2 slabs. The chelated catalysts showed a greater dispersion when compared with the unchelated catalysts, and this was confirmed by XRD analysis by the absence of crystalline peaks for the chelated catalysts. RhMo/Al2O3 resulted in higher catalytic activity when compared with the chelated catalysts, and this was confirmed by XPS showing more MoS2 phases of the RhMo/Al2O3 catalyst to be (63%), and it was also confirmed by HDS activity where RhMo/Al2O3 exhibited the highest DBT conversion of (88%). The addition of the chelating ligands (EDTA, AA and CA) resulted in lower HDS activity. Since this is a new catalyst, a lot of parameters must be investigated, such as the sulfidation temperature for Rh, the crystallite size effects, the molar ratio of Rh:chelating ligand to be used, and the type of the chelating ligand to be used. This would help to understand the catalyst and how the activity could be enhanced.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal11111398/s1, Figure S1: The Eg values for RhMo/Al2O3, RhMo-x/Al2O3 (x = EDTA, AA, CA) obtained from a UV-Vis spectra; Figure S2. FT-IR spectra of RhMo/ɣ-Al2O3, RhMo-EDTA/ɣ-Al2O3, RhMo-AA/ɣ-Al2O3, and RhMo-CA/ɣ-Al2O3 catalysts; Figure S3. EDX analysis for (a) RhMo/Al2O3, (b) RhMo-EDTA/Al2O3, (c) RhMo-AA/Al2O3, (d) RhMo-CA/Al2O3 catalysts.; Figure S4. XPS spectra for RhMo/Al2O3 different elemental contributions of (a) O 1s, (b) C 1s, (c) Al 2p, (d) Al 2s; Figure S5. XPS spectra for CoMo-EDTA/ɣ-Al2O3 different elemental contributions of (a) O 1s, (b) C 1s, (c) Al 2p, (d) Al 2s; Figure S6. GC chromatogram of DBT before HDS; Figure S7. GC chromatogram of DBT after HDS using (A) RhMo/Al2O3, (B) RhMo-EDTA/Al2O3, (C) RhMo-AA/Al2O3, (D) RhMo-CA/Al2O3, DBT = dibenzothiophene, BP = biphenyl, PhCH = biphenyl cyclohexane, BCH = bicyclohexyl, THDBT= tetrahydrodibenzothiophene; Table S1. XPS parameters of the different distributions (BE) of Mo 3d obtained for chelated and unchelated RhMo/Al2O3 catalysts; Table S2. XPS parameters of the contributions of Rh 3d obtained for unchelated and chelated RhMo/Al2O3 catalysts.

Author Contributions

Conceptualization, Z.R.T. and A.S.O.; methodology, S.M.; software, S.M. and A.S.O.; validation, S.M. and A.S.O.; formal analysis, S.M.; investigation, S.M.; resources, Z.R.T. and A.S.O.; data curation, S.M. and A.S.O.; writing—original draft preparation, S.M. and A.S.O.; writing—review and editing, S.M., Z.R.T. and A.S.O.; visualization, S.M. and A.S.O.; supervision, Z.R.T. and A.S.O.; project administration, Z.R.T. and A.S.O.; funding acquisition, A.S.O. All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the National Research Foundation (NRF) of South Africa (TTK170422228302). This research was also supported by Sasol South Africa (Pty) Ltd. and Nelson Mandela University.

Data Availability Statement

Data is contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no known conflicts of interests.

References

  1. Dembaremba, T.O.; Ogunlaja, A.S.; Tshentu, Z.R. Coordination polymers and polymer nanofibers for effective adsorptive desulfurization. In Nanocomposites Desulfurization Fuels; Saleh, T., Ed.; IGI Global: Hershey, PA, USA, 2020; pp. 168–234. Available online: https://0-www-igi--global-com.brum.beds.ac.uk/chapter/coordination-polymers-and-polymer-nanofibers-for-effective-adsorptive-desulfurization/246161 (accessed on 4 March 2021). [CrossRef]
  2. Tahira, S.; Qazib, U.Y.; Naseema, Z.; Tahira, N.; Zahida, M.; Javaidc, R.; Shahid, I. Deep eutectic solvents as alternative green solvents for the efficient desulfurization of liquid fuel: A comprehensive review. Fuel 2021, 305, 121502. [Google Scholar] [CrossRef]
  3. Nascimento, I.G.; Locatel, W.R.; Magalhães, B.C.; Travalloni, L.; Zotin, J.L.; da Silva, M.A.P. Kinetics of dibenzothiophene hydrodesulfurization reactions using CoMoP/Al2O3 and NiMoP/Al2O3. Catal. Today 2021, 381, 200–208. [Google Scholar] [CrossRef]
  4. Mundotiya, S.; Singh, R.; Saha, S.; Kakkar, R.; Pal, S.; Kunzru, D.; Pala, R.G.S.; Sivakumar, S. Effect of Sodium on Ni-Promoted MoS2 Catalyst for Hydrodesulfurization Reaction: Combined Experimental and Simulation Study. Energy Fuels 2021, 35, 2368–2378. [Google Scholar] [CrossRef]
  5. Huang, T.; Peng, Q.; Gai, H.; Fan, Y. Surfactant-Confined Synthesis of CoMoS Catalysts Using Polyoxometalate Precursors for Superior Fuel Hydrodesulfurization. Energy Fuels 2021, 35, 2402–2415. [Google Scholar] [CrossRef]
  6. Zhang, M.; Wang, C.; Han, Z.; Wang, K.; Zhang, Z.; Guan, G.; Abudula, A.; Xu, G. Catalyst Ni-Mo/Al2O3 promoted with infrared heating calcination for hydrodesulfurization of shale oil. Fuel 2021, 305, 121537. [Google Scholar] [CrossRef]
  7. Vinogradov, N.A.; Glotov, A.P.; Savinov, A.A.; Vutolkina, A.V.; Vinokurov, V.A.; Pimerzin, A.A. The mesoporous silicate-alumina composites application as supports for bifunctional sulfide catalysts for n-hexadecane hydroconversion. J. Porous Mater. 2021, 28, 1449–1458. [Google Scholar] [CrossRef]
  8. Ghosh, S.; Courthéoux, L.; Brunet, S.; Lacroix-Desmazes, P.; Pradel, A.; Girard, E.; Uzio, D. Hybrid CoMoS—Polyaniline nanowires catalysts for hydrodesulfurisation applications. Appl. Catal. A Gen. 2021, 623, 118264. [Google Scholar] [CrossRef]
  9. Soltanali, S.; Mashayekhi, M.; Mohaddecy, S.R.S. Comprehensive investigation of the effect of adding phosphorus and/or boron to NiMo/γ-Al2O3 catalyst in diesel fuel hydrotreating. Process Saf. Environ. Prot. 2020, 137, 273–281. [Google Scholar] [CrossRef]
  10. Hamidi, R.; Khoshbin, R.; Karimzadeh, R. Facile fabrication, characterization and catalytic activity of a NiMo/Al2O3 nanocatalyst via a solution combustion method used in a low temperature hydrodesulfurization process: The effect of fuel to oxidant ratio. RSC Adv. 2020, 10, 12439–12450. [Google Scholar] [CrossRef] [Green Version]
  11. Shafiq, I.; Shafique, S.; Akhter, P.; Yang, W.; Hussain, W. Recent developments in alumina supported hydrodesulfurization catalysts for the production of sulfur-free refinery products: A technical review. Catal. Rev. 2020, 1–86. [Google Scholar] [CrossRef]
  12. Vutolkina, A.V.; Glotov, A.P.; Zanina, A.V.; Mahmutov, D.F.; Maksimov, A.L.; Karakhanov, E.A. Mesoporous Al-HMS and Al-MCM-41 supported Ni-Mo sulfide catalysts for HYD and HDS via in situ hydrogen generation through a WGSR. Catal. Today 2019, 329, 156–166. [Google Scholar] [CrossRef]
  13. Rinaldi, N.; Kubota, T.; Okamoto, Y. Effect of citric acid addition on the hydrodesulfurization activity of MoO3/Al2O3 catalysts. Appl. Catal. A: Gen. 2010, 374, 228–236. [Google Scholar] [CrossRef]
  14. Ninh, T.K.T.; Massin, L.; Laurenti, D.; Vrinat, M. A new approach in the evaluation of the support effect for NiMo hydrodesulfurization catalysts. Appl. Catal. A: Gen. 2011, 407, 29–39. [Google Scholar] [CrossRef]
  15. Sun, M.; Nicosia, D.; Prins, R. The effects of fluorine, phosphate and chelating agents on hydrotreating catalysts and catalysis. Catal. Today 2003, 86, 173–189. [Google Scholar] [CrossRef]
  16. Zhang, X.; Wei, W.; Yin, X.; Li, G.; Li, L.; Guo, R.; Liu, X.; Guo, Q.; Shen, B. Migration and rearrangement of Fe–Zn species in the preparation of Fe–Zn bimetallic catalysts and their effects on the hydrodesulfurization reactivity. Energy Fuels 2021, 35, 16768–16777. [Google Scholar] [CrossRef]
  17. Ramírez, J.; Romualdo-Escobar, D.; Castillo-Villalón, P.; Gutiérrez-Alejandre, A. Improved NiMoSA catalysts: Analysis of EDTA post-treatment in the HDS of 4,6-DMDBT. Catal. Today 2020, 349, 168–177. [Google Scholar] [CrossRef]
  18. Hussain, M.; Song, S.-K.; Lee, J.-H.; Ihm, S.-K. Characteristics of CoMo catalysts supported on modified MCM-41 and MCM-48 materials for thiophene hydrodesulfurization. Ind. Eng. Chem. Res. 2006, 45, 536–543. [Google Scholar] [CrossRef]
  19. Hussain, M.; Ihm, S.K. Synthesis, characterization, and hydrodesulfurization activity of new mesoporous carbon supported transition metal sulfide catalysts. Ind. Eng. Chem. Res. 2009, 48, 698–707. [Google Scholar] [CrossRef]
  20. Hussain, M.; Yun, J.S.; Ihm, S.-K.; Russo, N.; Geobaldo, F. Synthesis, characterization, and thiophene hydrodesulfurization activity of novel macroporous and mesomacroporous carbon. Ind. Eng. Chem. Res. 2011, 50, 2530–2535. [Google Scholar] [CrossRef]
  21. Aray, Y.; Rodriguez, J.; Vega, D.; Rodriguez-Arias, E.N. Correlation of the topology of the electron density of pyrite-type transition metal sulfides with their catalytic activity in hydrodesulfurization. Angew. Chem. Int. Ed. 2000, 39, 3810–3813. [Google Scholar] [CrossRef]
  22. Vissers, J.P.R.; Groot, C.K.; Van Oers, E.M.; Beer, D.; Prins, R. Carbon-supported transition metal sulfides. Bull. Des Sociétés Chim. Belg. 1984, 93, 813–822. [Google Scholar] [CrossRef]
  23. Ledoux, M.J.; Michaux, O.; Agostini, G.; Panissod, P. The influence of sulfide structures on the hydrodesulfurization activity of carbon-supported catalysts. J. Catal. 1986, 102, 275–288. [Google Scholar] [CrossRef]
  24. Raybaud, P.; Hafner, J.; Kresse, G.; Toulhoat, H. Ab initio density functional studies of transition-metal sulphides: II. Electronic structure. J. Phys. Condens. Matter 1997, 9, 11107. [Google Scholar] [CrossRef]
  25. Toulhoat, H.; Raybaud, P.; Kasztelan, S.; Kresse, G.; Hafner, J. Transition metals to sulfur binding energies relationship to catalytic activities in HDS: Back to Sabatier with first principle calculations. Catal. Today 1999, 50, 629–636. [Google Scholar] [CrossRef]
  26. Bataille, F.; Lemberton, J.L.; Michaud, P.; Pérot, G.; Vrinat, M.; Lemaire, M.; Kasztelan, S. Alkyldibenzothiophenes hydrodesulfurization-promoter effect, reactivity, and reaction mechanism. J. Catal. 2000, 191, 409–422. [Google Scholar] [CrossRef]
  27. Lee, J.; Ishihara, A.; Dumeignil, F.; Miyazaki, K.; Oomori, Y.; Qian, E.W.; Kabe, T. Novel hydrodesulfurization catalysts derived from a rhodium carbonyl complex. J. Mol. Catal. A Chem. 2004, 209, 155–162. [Google Scholar] [CrossRef]
  28. Giraldo, S.A.; Pinzón, M.H.; Centeno, A. Behavior of catalysts with rhodium in simultaneous hydrodesulfurization and hydrogenation reactions. Catal. Today 2008, 133, 239–243. [Google Scholar] [CrossRef]
  29. Rana, M.S.; Ramírez, J.; Gutiérrez-Alejandre, A.; Ancheyta, J.; Cedeño, L.; Maity, S.K. Support effects in CoMo hydrodesulfurization catalysts prepared with EDTA as a chelating agent. J. Catal. 2007, 246, 100–108. [Google Scholar] [CrossRef]
  30. Stanislaus, A.; Marafi, A.; Rana, M.S. Recent advances in the science and technology of ultra-low sulfur diesel (ULSD) production. Catal. Today 2010, 153, 1–68. [Google Scholar] [CrossRef]
  31. Ríos-Caloch, G.; Santes, V.; Escobar, J.; Valle-Orta, M.; Barrera, M.C.; Hernández-Barrera, M. Effect of chitosan addition on NiMo/Al2O3 catalysts for Dibenzothiophene Hydrodesulfurization. Int. J. Chem. Reactor Eng. 2012, 10, 1–24. [Google Scholar] [CrossRef]
  32. Bui, N.Q.; Geantet, C.; Berhault, G. Activation of regenerated CoMo/Al2O3 hydrotreating catalysts by organic additives–The particular case of maleic acid. Appl. Catal. A Gen. 2019, 572, 185–196. [Google Scholar] [CrossRef]
  33. Mendes, F.M.; Schmal, M. The cyclohexanol dehydrogenation on Rh-CuAl2O3 catalysts Part 1. Characterization of the catalyst. Appl. Catal. A Gen. 1997, 151, 393–408. [Google Scholar] [CrossRef]
  34. Bergwerff, J.A.; Visser, T.; Weckhuysen, B.M. On the interaction between Co-and Mo-complexes in impregnation solutions used for the preparation of Al2O3-supported HDS catalysts: A combined Raman/UV–vis–NIR spectroscopy study. Catal. Today 2008, 130, 117–125. [Google Scholar] [CrossRef] [Green Version]
  35. Vatutina, Y.V.; Klimov, O.V.; Stolyarova, E.A.; Nadeina, K.A.; Danilova, I.G.; Chesalov, Y.A.; Noskov, A.S. Influence of the phosphorus addition ways on properties of CoMo-catalysts of hydrotreating. Catal. Today 2019, 329, 13–23. [Google Scholar] [CrossRef]
  36. Matralis, H.; Papadopoulou, C.; Lycourghiotis, A. Fluorinated hydrotreatment catalysts effect of the deposition order of F− ions on F-CoMo/γ-Al2O3 catalysts. Appl. Catal. A Gen. 1994, 116, 221–236. [Google Scholar] [CrossRef]
  37. Papadopoulou, C.; Vakros, J.; Matralis, H.K.; Voyiatzis, G.A.; Kordulis, C. Preparation, characterization, and catalytic activity of CoMo/γ-Al2O3 catalysts prepared by equilibrium deposition filtration and conventional impregnation techniques. J. Colloid Interface Sci. 2004, 274, 159–166. [Google Scholar] [CrossRef]
  38. Huirache-Acuña, R.; Zepeda, T.A.; Rivera-Muñoz, E.M.; Nava, R.; Loricera, C.V.; Pawelec, B. Characterization and HDS performance of sulfided CoMoW catalysts supported on mesoporous Al-SBA-16 substrates. Fuel 2015, 149, 149–161. [Google Scholar] [CrossRef]
  39. Rinaldi, N.; Al-Dalama, K.; Kubota, T.; Okamoto, Y. Preparation of Co–Mo/B2O3/Al2O3 catalysts for hydrodesulfurization: Effect of citric acid addition. Appl. Catal. A Gen. 2009, 360, 130–136. [Google Scholar] [CrossRef]
  40. Zhang, Y.; Han, W.; Long, X.; Nie, H. Redispersion effects of citric acid on CoMo/γ-Al2O3 hydrodesulfurization catalysts. Catal. Commun. 2016, 82, 20–23. [Google Scholar] [CrossRef]
  41. Weber, R. Effect of local structure on the UV-Visible absorption edges of molybdenum oxide clusters and supported molybdenum oxides. J. Catal. 1995, 151, 470–474. [Google Scholar] [CrossRef]
  42. Tian, H.; Roberts, C.A.; Wachs, I.E. Molecular structural determination of molybdena in different environments: Aqueous solutions, bulk mixed oxides, and supported MoO3 catalysts. J. Phys. Chem. C 2010, 114, 14110–14120. [Google Scholar] [CrossRef]
  43. Xu, J.; Huang, T.; Fan, Y. Highly efficient NiMo/SiO2-Al2O3 hydrodesulfurization catalyst prepared from gemini surfactant-dispersed Mo precursor. Appl. Catal. B Environ. 2017, 203, 839–850. [Google Scholar] [CrossRef]
  44. Escobar, J.; Barrera, M.C.; Gutiérrez, A.W.; Terrazas, J.E. Benzothiophene hydrodesulfurization over NiMo/alumina catalysts modified by citric acid. Effect of addition stage of organic modifier. Fuel Process. Technol. 2017, 156, 33–42. [Google Scholar] [CrossRef]
  45. Cabello, C.I.; Cabrerizo, F.M.; Alvarez, A.; Thomas, H.J. Decamolybdodicobaltate (III) heteropolyanion: Structural, spectroscopical, thermal and hydrotreating catalytic properties. J. Mol. Catal. A Chem. 2002, 186, 89–100. [Google Scholar] [CrossRef]
  46. Valencia, D.; Klimova, T. Citric acid loading for MoS2-based catalysts supported on SBA-15. New catalytic materials with high hydrogenolysis ability in hydrodesulfurization. Appl. Catal. B Environ. 2013, 129, 137–145. [Google Scholar] [CrossRef]
  47. Yurum, A.; Karakas, G. Synthesis of Na-, Fe-, and Co-promoted TiO2/multiwalled carbon nanotube composites and their use as a photocatalyst. Turk. J. Chem. 2017, 41, 440–454. [Google Scholar] [CrossRef]
  48. Trakarnpruk, W.; Seentrakoon, B. Hydrodesulfurization activity of MoS2 and bimetallic catalysts prepared by in situ decomposition of thiosalt. Ind. Eng. Chem. Res. 2007, 46, 1874–1882. [Google Scholar] [CrossRef]
  49. Alonso, G.; Petranovskii, V.; Del Valle, M.; Cruz-Reyes, J.; Licea-Claverie, A.; Fuentes, S. Preparation of WS2 catalysts by in situ decomposition of tetraalkylammonium thiotungstates. Appl. Catal. A Gen. 2000, 197, 87–97. [Google Scholar] [CrossRef]
  50. Badoga, S.; Mouli, K.C.; Soni, K.K.; Dalai, A.K.; Adjaye, J. Beneficial influence of EDTA on the structure and catalytic properties of sulfided NiMo/SBA-15 catalysts for hydrotreating of light gas oil. Appl. Catal. B Environ. 2012, 125, 67–84. [Google Scholar] [CrossRef]
  51. Calderón-Magdaleno, M.Á.; Mendoza-Nieto, J.A.; Klimova, T.E. Effect of the amount of citric acid used in the preparation of NiMo/SBA-15 catalysts on their performance in HDS of dibenzothiophene-type compounds. Catal. Today. 2014, 220, 78–88. [Google Scholar] [CrossRef]
  52. Chen, S.; Luo, L.; Cheng, X. Influence of preparation method on the performance of Pd/ZrO2-Al2O3 catalysts for HDS. Indian J. Chem. Technol. 2009, 16, 272–277. [Google Scholar]
  53. Gutiérrez, O.Y.; Pérez, F.; Fuentes, G.A.; Bokhimi, X.; Klimova, T. Deep HDS over NiMo/Zr-SBA-15 catalysts with varying MoO3 loading. Catal. Today 2008, 130, 292–301. [Google Scholar] [CrossRef]
  54. Hara, K.; Iwahashi, K.; Takakusagi, S.; Uosaki, K.; Sawamura, M. Construction of self-assembled monolayer terminated with N-heterocyclic carbene–rhodium(I) complex moiety. Surf. Sci. 2007, 601, 5127–5132. [Google Scholar] [CrossRef]
  55. Larichev, Y.V.; Netskina, O.V.; Komova, O.V.; Simagina, V.I. Comparative XPS study of Rh/Al2O3 and Rh/TiO2 as catalysts for NaBH4 hydrolysis. Int. J. Hydrog. Energy 2010, 35, 6501–6507. [Google Scholar] [CrossRef]
  56. Masud, J.; Van Nguyen, T.; Singh, N.; McFarland, E.; Ikenberry, M.; Hohn, K.; Hwang, B.J. A RhxSy/C catalyst for the hydrogen oxidation and hydrogen evolution reactions in HBr. J. Electrochem. Soc. 2015, 162, F455. [Google Scholar] [CrossRef] [Green Version]
  57. Galtayries, A.; Wisniewski, S.; Grimblot, J. Formation of thin oxide and sulphide films on polycrystalline molybdenum foils: Characterization by XPS and surface potential variations. J. Electron Spectrosc. Relat. Phenom. 1997, 87, 31–44. [Google Scholar] [CrossRef]
  58. Venezia, A.M. X-ray photoelectron spectroscopy (XPS) for catalysts characterization. Catal. Today 2003, 77, 359–370. [Google Scholar] [CrossRef]
  59. Budukva, S.V.; Klimov, O.V.; Uvarkina, D.D.; Chesalov, Y.A.; Prosvirin, I.P.; Larina, T.V.; Noskov, A.S. Effect of citric acid and triethylene glycol addition on the reactivation of CoMo/γ-Al2O3 hydrotreating catalysts. Catal. Today 2019, 329, 35–43. [Google Scholar] [CrossRef]
  60. Xu, J.; Guo, Y.; Huang, T.; Fan, Y. Hexamethonium bromide-assisted synthesis of CoMo/graphene catalysts for selective hydrodesulfurization. Appl. Catal. B Environ. 2019, 244, 385–395. [Google Scholar] [CrossRef]
  61. Huang, T.; Xu, J.; Fan, Y. Effects of concentration and microstructure of active phases on the selective hydrodesulfurization performance of sulfided CoMo/Al2O3 catalysts. Appl. Catal. B Environ. 2018, 220, 42–56. [Google Scholar] [CrossRef]
  62. Matsumoto, K.; Matsumoto, T.; Kawano, M.; Ohnuki, H.; Shichi, Y.; Nishide, T.; Sato, T. Syntheses and crystal structures of disulfide-bridged binuclear ruthenium compounds: The First UV—Vis, Raman, ESR, and XPS Spectroscopic Characterization of a Valence-Averaged Mixed-Valent RuIIISSRuII Core. J. Am. Chem. Soc. 1996, 118, 3597–3609. [Google Scholar] [CrossRef]
  63. Tang, M.; Zhou, L.; Du, M.; Lyu, Z.; Wen, X.D.; Li, X.; Ge, H. A novel reactive adsorption desulfurization Ni/MnO adsorbent and its hydrodesulfurization ability compared with Ni/ZnO. Catal. Commun. 2015, 61, 37–40. [Google Scholar] [CrossRef]
  64. Liu, B.; Chai, Y.; Li, Y.; Wang, A.; Liu, Y.; Liu, C. Effect of sulfidation atmosphere on the performance of the CoMo/γ-Al2O3 catalysts in hydrodesulfurization of FCC gasoline. Appl. Catal. A Gen. 2014, 471, 70–79. [Google Scholar] [CrossRef]
  65. Wang, H.; Liu, S.; Smith, K.J. Understanding selectivity changes during hydrodesulfurization of dibenzothiophene on Mo2C/carbon catalysts. J. Catal. 2019, 369, 427–439. [Google Scholar] [CrossRef]
  66. Rashidi, F.; Sasaki, T.; Rashidi, A.M.; Kharat, A.N.; Jozani, K.J. Ultradeep hydrodesulfurization of diesel fuels using highly efficient nanoalumina-supported catalysts: Impact of support, phosphorus, and/or boron on the structure and catalytic activity. J. Catal. 2013, 299, 321–335. [Google Scholar] [CrossRef]
  67. Wang, X.; Du, P.; Chi, K.; Duan, A.; Xu, C.; Zhao, Z.; Zhang, H. Synthesis of NiMo catalysts supported on mesoporous silica FDU-12 with different morphologies and their catalytic performance of DBT HDS. Catal. Today 2017, 291, 146–152. [Google Scholar] [CrossRef]
  68. Nikulshina, M.; Kokliukhin, A.; Mozhaev, A.; Nikulshin, P. CoMo/Al2O3 hydrotreating catalysts prepared from single Co2Mo10-heteropolyacid at extremely high metal loading. Catal. Commun. 2019, 127, 51–57. [Google Scholar] [CrossRef]
  69. Muhammad, Y.; Rahman, A.U.; Rashid, H.U.; Sahibzada, M.; Subhan, S.; Tong, Z. Hydrodesulfurization of dibenzothiophene using Pd-promoted Co–Mo/Al₂O₃ and Ni–Mo/Al₂O₃ catalysts coupled with ionic liquids at ambient operating conditions. RSC Adv. 2019, 9, 10371–10385. [Google Scholar] [CrossRef]
  70. Pereyma, V.Y.; Klimov, O.V.; Prosvirin, I.P.; Gerasimov, E.Y.; Yashnik, S.A.; Noskov, A.S. Effect of thermal treatment on morphology and catalytic performance of NiW/Al2O3 catalysts prepared using citric acid as chelating agent. Catal. Today 2018, 305, 162–170. [Google Scholar] [CrossRef]
  71. Zhang, M.H.; Fan, J.Y.; Chi, K.; Duan, A.J.; Zhao, Z.; Meng, X.L.; Zhang, H.L. Synthesis, characterization, and catalytic performance of NiMo catalysts supported on different crystal alumina materials in the hydrodesulfurization of diesel. Fuel Process. Technol. 2017, 156, 446–453. [Google Scholar] [CrossRef]
  72. Shimizu, T.; Hiroshima, K.; Honma, T.; Mochizuki, T.; Yamada, M. Highly active hydrotreatment catalysts prepared with chelating agents. Catal. Today 1998, 45, 271–276. [Google Scholar] [CrossRef]
  73. Okamoto, Y.; Ochiai, K.; Kawano, M.; Kubota, T. Evaluation of the maximum potential activity of Co–Mo/Al2O3 catalysts for hydrodesulfurization. J. Catal. 2004, 222, 143–151. [Google Scholar] [CrossRef]
  74. Yoosuk, B.; Kim, J.H.; Song, C.; Ngamcharussrivichai, C.; Prasassarakich, P. Highly active MoS2, CoMoS2 and NiMoS2 unsupported catalysts prepared by hydrothermal synthesis for hydrodesulfurization of 4, 6-dimethyldibenzothiophene. Catal. Today 2008, 130, 14–23. [Google Scholar] [CrossRef]
  75. Escobar, M.C.; Barrera, J.A.; De los Reyes, J.A.; Toledo, V.; Santes, J.A. Colín, Effect of chelating ligands on Ni-Mo impregnation over wide-pore ZrO2-TiO2. J. Mol. Catal. A Chem. 2008, 278, 33–40. [Google Scholar] [CrossRef]
  76. González-Cortés, S.L.; Xiao, T.C.; Costa, P.M.; Fontal, B.; Green, M.L. Urea–organic matrix method: An alternative approach to prepare Co-MoS2/γ-Al2O3 HDS catalyst. Appl. Catal. A Gen. 2004, 270, 209–222. [Google Scholar] [CrossRef]
  77. González-Cortés, S.L.; Xiao, T.C.; Rodulfo-Baechler, S.M.; Green, M.L. Impact of the urea–matrix combustion method on the HDS performance of Ni-MoS2/γ-Al2O3 catalysts. J. Mol. Catal. A Chem. 2005, 240, 214–225. [Google Scholar]
  78. Bin, L.I.U.; Lei, L.I.U.; Chai, Y.M.; Zhao, J.C.; Liu, C.G. Essential role of promoter Co on the MoS2 catalyst in selective hydrodesulfurization of FCC gasoline. J. Fuel Chem. Technol. 2018, 46, 441–450. [Google Scholar]
  79. Li, X.; Bai, J.; Wang, A.; Prins, R.; Wang, Y. Hydrodesulfurization of dibenzothiophene and its hydrogenated intermediates over bulk Ni2P. Top. Catal. 2011, 54, 290–298. [Google Scholar] [CrossRef]
  80. García-Martínez, J.C.; Dutta, A.; Chávez, G.; De los Reyes, J.A.; Castillo-Araiza, C.O. Hydrodesulfurization of dibenzothiophene in a micro trickle bed catalytic reactor under operating conditions from reactive distillation. Int. J. Chem. React. Eng. 2015, 14, 769–783. [Google Scholar] [CrossRef]
  81. Torres-García, N.L.; Huirache-Acuña, R.; Zepeda-Partida, T.A.; Pawelec, B.; Fierro, J.L.G.; Vázquez-Salas, P.J.; Maya-Yescas, R.; Rivera-Garnica, J.M. Trimetallic RuxMoNi catalysts supported on SBA-15 for the hydrodesulfurization of dibenzothiophene. Int. J. Chem. React. Eng. 2019, 17, 2194–5748. [Google Scholar] [CrossRef]
  82. Li, M.; Song, J.; Yue, F.; Pan, F.; Yan, W.; Hua, Z.; Li, L.; Yang, Z.; Li, L.; Wen, G.; et al. Complete Hydrodesulfurization of Dibenzothiophene via Direct Desulfurization Pathway over Mesoporous TiO2-Supported NiMo Catalyst Incorporated with Potassium. Catalysts 2019, 9, 448. [Google Scholar] [CrossRef] [Green Version]
  83. Li, X.; Wang, A.; Wang, Y.; Chen, Y.; Liu, Y.; Hu, Y. Hydrodesulfurization of dibenzothiophene over Ni-Mo sulfides supported by proton-exchanged siliceous MCM-41. Catal. Lett. 2002, 84, 107–113. [Google Scholar] [CrossRef]
  84. Liu, X.; Liu, J.; Li, L.; Guo, R.; Zhang, X.; Ren, S.; Guo, Q.; Wen, X.-D.; Shen, B. Hydrodesulfurization of dibenzothiophene on TiO2−x-modified Fe-based catalysts: Electron transfer behavior between TiO2−x and Fe species. ACS Catal. 2020, 10, 9019–9033. [Google Scholar] [CrossRef]
Figure 1. UV-vis spectra of RhMo/Al2O3 and RhMo-x/γ-Al2O3 (x = EDTA, AA, CA) catalysts.
Figure 1. UV-vis spectra of RhMo/Al2O3 and RhMo-x/γ-Al2O3 (x = EDTA, AA, CA) catalysts.
Catalysts 11 01398 g001
Figure 2. XRD diffraction pattern for oxide and sulfided (a) RhMo/ɣ-Al2O3, (b) RhMo-EDTA/ɣ-Al2O3, (c) RhMo-AA/ɣ-Al2O3, (d) RhMo-CA/ɣ-Al2O3 catalysts.
Figure 2. XRD diffraction pattern for oxide and sulfided (a) RhMo/ɣ-Al2O3, (b) RhMo-EDTA/ɣ-Al2O3, (c) RhMo-AA/ɣ-Al2O3, (d) RhMo-CA/ɣ-Al2O3 catalysts.
Catalysts 11 01398 g002
Figure 3. XPS spectra for (a) RhMo/ɣ-Al2O3 survey spectrum with different elemental contributions, (b) S 2p, (c) Rh 3d, (d) Mo 3d.
Figure 3. XPS spectra for (a) RhMo/ɣ-Al2O3 survey spectrum with different elemental contributions, (b) S 2p, (c) Rh 3d, (d) Mo 3d.
Catalysts 11 01398 g003
Figure 4. XPS spectra for (a) RhMo-EDTA/ɣ-Al2O3 survey spectrum with different elemental contributions, (b) S 2p, (c) Rh 3d, (d) Mo 3d.
Figure 4. XPS spectra for (a) RhMo-EDTA/ɣ-Al2O3 survey spectrum with different elemental contributions, (b) S 2p, (c) Rh 3d, (d) Mo 3d.
Catalysts 11 01398 g004
Figure 5. XPS deconvolution of RhMo/ɣ-Al2O3 and RhMo-EDTA/ɣ-Al2O3, where: (a) Rh 3d for RhMo/ɣ-Al2O3; (b) Mo 3d for RhMo/ɣ-Al2O3; (c) Rh 3d for RhMo-EDTA/ɣ-Al2O3; (d) Mo 3d for RhMo-EDTA/ɣ-Al2O3; (e) S 2p for RhMo/ɣ-Al2O3; (f) S 2p RhMo-EDTA/ɣ-Al2O3.
Figure 5. XPS deconvolution of RhMo/ɣ-Al2O3 and RhMo-EDTA/ɣ-Al2O3, where: (a) Rh 3d for RhMo/ɣ-Al2O3; (b) Mo 3d for RhMo/ɣ-Al2O3; (c) Rh 3d for RhMo-EDTA/ɣ-Al2O3; (d) Mo 3d for RhMo-EDTA/ɣ-Al2O3; (e) S 2p for RhMo/ɣ-Al2O3; (f) S 2p RhMo-EDTA/ɣ-Al2O3.
Catalysts 11 01398 g005
Figure 6. TEM images for (a) RhMo/ɣ-Al2O3, (b) RhMo-EDTA/ɣ-Al2O3, (c) RhMo-AA/ɣ-Al2O3, (d) RhMo-CA/ɣ-Al2O3.
Figure 6. TEM images for (a) RhMo/ɣ-Al2O3, (b) RhMo-EDTA/ɣ-Al2O3, (c) RhMo-AA/ɣ-Al2O3, (d) RhMo-CA/ɣ-Al2O3.
Catalysts 11 01398 g006
Figure 7. SEM images for RhMo/ɣ-Al2O3 (a) oxide, (b) sulfided; RhMo-EDTA/ɣ-Al2O3 (c) oxide, (d) sulfided; RhMo-AA/ɣ-Al2O3 (e) oxide, (f) sulfided; RhMo-CA/ɣ-Al2O3 (g) oxide, (h) sulfided catalysts.
Figure 7. SEM images for RhMo/ɣ-Al2O3 (a) oxide, (b) sulfided; RhMo-EDTA/ɣ-Al2O3 (c) oxide, (d) sulfided; RhMo-AA/ɣ-Al2O3 (e) oxide, (f) sulfided; RhMo-CA/ɣ-Al2O3 (g) oxide, (h) sulfided catalysts.
Catalysts 11 01398 g007
Figure 8. TGA-DSC curves of (a) RhMo/ɣ-Al2O3, (b) RhMo-EDTA/ɣ-Al2O3, (c) RhMo-AA/ɣ-Al2O3, (d) RhMo-CA/ɣ-Al2O3 catalysts.
Figure 8. TGA-DSC curves of (a) RhMo/ɣ-Al2O3, (b) RhMo-EDTA/ɣ-Al2O3, (c) RhMo-AA/ɣ-Al2O3, (d) RhMo-CA/ɣ-Al2O3 catalysts.
Catalysts 11 01398 g008
Figure 9. XPS spectra of O 1s of RhMo/ɣ-Al2O3 and RhMo-EDTA/ɣ-Al2O3.
Figure 9. XPS spectra of O 1s of RhMo/ɣ-Al2O3 and RhMo-EDTA/ɣ-Al2O3.
Catalysts 11 01398 g009
Scheme 1. Proposed mechanisms of DDS and HDS reaction on RhMo chelated catalysts in HDS of DBT.
Scheme 1. Proposed mechanisms of DDS and HDS reaction on RhMo chelated catalysts in HDS of DBT.
Catalysts 11 01398 sch001
Table 1. Band gap energy and average number of nearest Mo6+ neighbors (NMo-O-Mo) in deposited clusters, as determined from UV spectra of the oxide RhMo catalysts.
Table 1. Band gap energy and average number of nearest Mo6+ neighbors (NMo-O-Mo) in deposited clusters, as determined from UV spectra of the oxide RhMo catalysts.
CatalystEg ValuesNMo-O-Mo
RhMo/γ-Al2O33.7791.975
RhMo-AA/γ-Al2O34.3410.5134
RhMo-EDTA/γ-Al2O34.3940.3756
RhMo-CA/γ-Al2O34.4780.1572
NMo-O-Mo = 11.8–2.6Eg.
Table 2. Qualitative atomic percentage of Rh, C, O, S and Mo for sulfided HDS catalysts.
Table 2. Qualitative atomic percentage of Rh, C, O, S and Mo for sulfided HDS catalysts.
CatalystsAtomic Percentage (wt. %)S/MoC/Mo
C KO KAl KS KRh LMo L
RhMo/ɣ-Al2O37.9763.2426.470.680.171.460.475.46
RhMo-EDTA/ɣ-Al2O310.2057.5230.031.480.440.324.6331.88
RhMo-AA/ɣ-Al2O39.3862.6125.631.350.190.731.8512.85
RhMo-CA/ɣ-Al2O312.0757.5626.881.600.940.901.7813.41
Table 3. Binding energies determined in XPS experiments for RhMo/Al2O3 and RhMo-EDTA/ɣ-Al2O3.
Table 3. Binding energies determined in XPS experiments for RhMo/Al2O3 and RhMo-EDTA/ɣ-Al2O3.
Elements (eV)RhMo/ɣ-Al2O3RhMo-EDTA/ɣ-Al2O3
C 1s289.5285.5
O 1s530.0529.0
Mo 3d226.2; 230.0; 233.2226.0; 230.0
Rh 3d306.5; 310.6305.0; 310.1
S 2p160.1; 167.0160.5; 166.1
Al 2p72.472.0
Al 2s117.0117.0
Table 4. Average length of MoS2 crystallites in RhMo/Al2O3 and RhMo-x/ɣ-Al2O3 (x = EDTA, AA, CA).
Table 4. Average length of MoS2 crystallites in RhMo/Al2O3 and RhMo-x/ɣ-Al2O3 (x = EDTA, AA, CA).
CatalystsAverage Diameter ± SD (nm)
RhMo/ɣ-Al2O34.4 (±1.38)
RhMo-EDTA/ɣ-Al2O34.1 (±1.220)
RhMo-AA/ɣ-Al2O33.3 (±0.757)
RhMo-CA/ɣ-Al2O31.6 (±0.860)
Table 5. Catalytic performances of RhMo/ɣ-Al2O3 and RhMo-x/ɣ-Al2O3 (x = EDTA, AA, CA) in hydrotreating of DBT as simulated fuel.
Table 5. Catalytic performances of RhMo/ɣ-Al2O3 and RhMo-x/ɣ-Al2O3 (x = EDTA, AA, CA) in hydrotreating of DBT as simulated fuel.
CatalystsCrystallite Sizes (nm)Eg ValuesHDS (%)BP(%)PhCh(%)HYD/DDS RatioTOF (h−1) a
RhMo/ɣ-Al2O35.9033.7798865130.2051
RhMo-EDTA/ɣ-Al2O35.7704.394681610.0660
RhMo-AA/ɣ-Al2O35.7504.341736530.0579
RhMo-CA/ɣ-Al2O35.8094.478723620.06223
Catalyst (molybdenum content) employed = 0.1 g (4.119 × 10−5 moles). Hydrodesulfurization (HDS) time = 6 h; reaction temperature = 300 °C; reaction pressure = 40 bar. Phenylcyclohexane (PhCh) or biphenyl (BP) a TOF, h−1: (turnover frequency).
Table 6. Comparison of catalyst performance with literature reports in DBT hydrodesulfurization.
Table 6. Comparison of catalyst performance with literature reports in DBT hydrodesulfurization.
CatalystsModel CompoundReaction Temperature (°C)HDS (%)Reaction Pressure (Bar)Reference
RhMo/ɣ-Al2O3DBT3008840This work
RhMo-EDTA/ɣ-Al2O3DBT3006840This work
RhMo-AA/ɣ-Al2O3DBT3007340This work
RhMo-CA/ɣ-Al2O3DBT3007240This work
Ni2PDBT3403540[79]
Ni2PTH-DBT3405040[79]
NiMoP/γ-Al2O3DBT<32022–90<25[80]
RuxMoNiDBT32024–9254.5[81]
NiMoDBT3206254.5[81]
NiMo/TiO2-6DBT3009020[82]
NiMo/MCM-41-NaDBT300>9550[83]
Fe-Zn/TiO2-Al2O3DBT380>9840[84]
RhMo/ɣ-Al2O3DBT3108450[28]
Dibenzothiophene (DBT); TH-DBT = 1,2,3,4-tetrahydro-dibenzothiophene (TH-DBT).
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Majodina, S.; Tshentu, Z.R.; Ogunlaja, A.S. Effect of Adding Chelating Ligands on the Catalytic Performance of Rh-Promoted MoS2 in the Hydrodesulfurization of Dibenzothiophene. Catalysts 2021, 11, 1398. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11111398

AMA Style

Majodina S, Tshentu ZR, Ogunlaja AS. Effect of Adding Chelating Ligands on the Catalytic Performance of Rh-Promoted MoS2 in the Hydrodesulfurization of Dibenzothiophene. Catalysts. 2021; 11(11):1398. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11111398

Chicago/Turabian Style

Majodina, Siphumelele, Zenixole R. Tshentu, and Adeniyi S. Ogunlaja. 2021. "Effect of Adding Chelating Ligands on the Catalytic Performance of Rh-Promoted MoS2 in the Hydrodesulfurization of Dibenzothiophene" Catalysts 11, no. 11: 1398. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11111398

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