Next Article in Journal
Facade-Based Bicelles as a New Tool for Production of Active Membrane Proteins in a Cell-Free System
Next Article in Special Issue
Graphitic Carbon Nitride/Zinc Oxide-Based Z-Scheme and S-Scheme Heterojunction Photocatalysts for the Photodegradation of Organic Pollutants
Previous Article in Journal
RNA-Based Strategies for Cancer Therapy: In Silico Design and Evaluation of ASOs for Targeted Exon Skipping
Previous Article in Special Issue
Amine Functionalization Leads to Enhanced Performance for Nickel- and Cobalt-Ferrite-Supported Palladium Catalysts in Nitrobenzene Hydrogenation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Bifunctional MoS2/Al2O3-Zeolite Catalysts in the Hydroprocessing of Methyl Palmitate

1
Boreskov Institute of Catalysis SB RAS, 630090 Novosibirsk, Russia
2
Faculty of Natural Sciences, Novosibirsk National Research University, 630090 Novosibirsk, Russia
3
Faculty of Chemistry, Lomonosov Moscow State University, 119991 Moscow, Russia
4
Topchiev Institute of Petrochemical Synthesis RAS, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(19), 14863; https://doi.org/10.3390/ijms241914863
Submission received: 15 September 2023 / Revised: 27 September 2023 / Accepted: 28 September 2023 / Published: 3 October 2023
(This article belongs to the Special Issue Catalysts: Design, Synthesis, and Molecular Applications)

Abstract

:
A series of bifunctional catalysts, MoS2/Al2O3 (70 wt.%), zeolite (30 wt.%) (zeolite—ZSM-5, ZSM-12, and ZSM-22), and silica aluminophosphate SAPO-11, were synthesized for hydroconversion of methyl palmitate (10 wt.% in dodecane) in a trickle-bed reactor. Mo loading was about 7 wt.%. Catalysts and supports were characterized by different physical-chemical methods (HRTEM-EDX, SEM-EDX, XRD, N2 physisorption, and FTIR spectroscopy). Hydroprocessing was performed at a temperature of 250–350 °C, hydrogen pressure of 3.0–5.0 MPa, liquid hourly space velocity (LHSV) of 36 h−1, and an H2/feed ratio of 600 Nm3/m3. Complete conversion of oxygen-containing compounds was achieved at 310 °C in the presence of MoS2/Al2O3-zeolite catalysts; the selectivity for the conversion of methyl palmitate via the ‘direct’ hydrodeoxygenation (HDO) route was over 85%. The yield of iso-alkanes gradually increases in order: MoS2/Al2O3 < MoS2/Al2O3-ZSM-12 < MoS2/Al2O3-ZSM-5 < MoS2/Al2O3-SAPO-11 < MoS2/Al2O3-ZSM-22. The sample MoS2/Al2O3-ZSM-22 demonstrated the highest yield of iso-alkanes (40%). The hydroisomerization activity of the catalysts was in good correlation with the concentration of Brønsted acid sites in the synthesized supports.

1. Introduction

Negative climate change caused by greenhouse gas emissions stimulates widespread expansion of the concept of “decarbonization” of the economy by reducing CO2 emissions. One of the approaches to solving this problem in the transport sector is the use of renewable bio-resources to obtain biofuel components that are not inferior in quality to products of petroleum origin [1,2,3]. The promising raw materials for the production of biofuel components are triglycerides of fatty acids (non-edible oils, substandard animal fats, food production waste), fatty acid esters, and free fatty acids [2,3,4,5,6]. HDO of natural lipids gives normal saturated C15-C18 hydrocarbons, which cannot be used as a drop-in paraffinic biofuel (equivalent functionally to petroleum-derived one and compatible with existing infrastructure) without further upgrading because of pour cold flow properties [2,3,4,5,6,7].
Therefore, hydroisomerization/hydrocracking of HDO products (alkanes C15-C18) is required, which gives [4,5,6,7]. Hydroprocessed ester and fatty acid (HEFA) is also known as renewable or green diesel, hydrotreated/hydrogenating vegetable oil (HVO), and hydrotreated biodiesel. Currently, processes with two catalytic stages are used in industry to obtain paraffinic biofuel components in diesel, kerosene, and gasoline ranges from lipid-based feedstocks, among them NExBTL, UOP/Eni EcofiningTM, UPM BioVerno, etc. [2,4,6,7]. At the first stage, hydrodeoxygenation of the feedstock is carried out over sulfide catalysts (Ni(Co)Mo/Al2O3); at the second stage (after thorough purification from sulfur-containing compounds, CO and CO2), hydroisomerization/hydrocracking of the obtained alkanes is carried out in the presence of catalysts based on noble metals (Pt on composite carriers, which include zeolites ZSM-22, ZSM-23, and SAPO-11).
The development of a one-stage process to obtain paraffinic biofuel components with the appropriate properties for hydroprocessing lipid-based feedstocks is an urgent task that has gotten a lot of attention in recent years [6,8]. The fatty acid methyl esters (FAME) are often used as a model compound to study the mechanism of reactions and catalytic properties of different materials in the hydroconversion of triglycerides. However, there is an opinion that it is more profitable to use the product of triglyceride transesterification, fatty acid methyl esters, which makes it possible to obtain value-added glycerol in the same process, save hydrogen, and reduce the carbon footprint [9,10,11]. In addition, the hydroconversion of FAME needs a lower reaction temperature and pressure in comparison with triglycerides [6].
It is generally accepted that the unavoidable condition for successive one-step hydroconversion of triglycerides and esters is the use of a polyfunctional catalyst that ensures the occurrence of several reactions (HDO and hydroisomerization/hydrocracking). According to a well-known concept, hydrodeoxygenation of esters first occurs to form normal alkanes with an even (via the direct hydrodeoxygenation route, with the removal of water) or odd (by decarbonylation/decarboxylation) number of carbon atoms [6,8,12]. Hydroisomerization of obtained n-alkanes proceeds via dehydrogenation to alkenes over metallic sites that are protonated by Brønsted acidic sites with carbenium ion formation and subsequent isomerization and hydrogenation [12,13]. Several articles dealing with the hydroprocessing of esters in a mixture of normal and iso-alkane have been reported over bifunctional catalysts, differing in zeolite component and metal function [6,8,14,15,16,17,18,19,20,21].
Pt/SAPO-11 bifunctional catalysts are widely studied in the hydroconversion of vegetable oils [6,8,14]. The use of Pt/Fe3SAPO-11 showed 100% conversion of FAME with 99.6% selectivity to C15-C18 alkanes and 34.8% selectivity of iso-C15-C18 alkanes with 34.8% in one-step hydrotreatment at 320, 4 MPa, run time 6 h in batch reactor [14]. Adding Sn increased the selectivity of iso-C15-C18 alkane formation over the Pt1Sn1/Fe3SAPO-11 catalyst from 34.8% to 62.7%. Despite high activity and selectivity, Pt-based catalysts are not preferred in industrial applications because of their high cost, low abundance, and sensitivity to poison. Hence, bifunctional catalysts containing transition metals (e.g., Ni, Co) have been tested in FAME hydroprocessing in recent years [15,16,17,18,19,20,21].
Ni/HZSM-5 catalysts with differing Si/Al ratios and Ni loadings were compared in the hydroprocessing of long-chain unsaturated fatty acid methyl esters [15]. Selectivity of 88.2% for C5-C18 liquid alkanes was obtained over 10 wt.% Ni/HZSM-5 (Si/Al = 25) at 280 °C, 0.8 MPa, LHSV of 4 h−1, and H2/oil molar ratio of 15, with isomerization selectivity of 27.0%. However, the conversion of FAME was only 85.1%, which decreased to 30.1% after operation for 80 h due to carbonaceous deposits. Hydroprocessing of microalgae biodiesel was performed over a 10% Ni/meso-Y zeolite catalyst [16]. A high isomerization ratio (46.4%) and selectivity to jet fuel range hydrocarbons (56.2%) were achieved, but conversion was 91.5% and Ni crystallite size was increased from 25 to 54 nm during hydroprocessing at 275 °C and 2.0 MPa. After the addition of 4% HPW to a 10% Ni/meso-Y catalyst, the production of jet fuel-range alkanes and iso-alkane selectivity increased along with an increase in strong acid density [17,18]. It was shown that Ni/meso-Y can produce 4,47% of aromatics, while Ni-based catalysts supported by Meso-ZSM-5, Meso-Hbeta, and SAPO-34 tend to produce more aromatics in the hydroconversion of microalgae oil in batch reactors at 370–410 °C for 8 h [19]. The Ni-based catalysts are prone to deactivation by coke deposition and agglomeration during the HDO process [15,16,20,21].
The comparison of Ni/SAPO-11, Co/SAPO-11, and NiCo/SAPO-11 in the hydroconversion of FAME was performed at 360–440 °C, 1.5 MPa, and a WHSV of 2.6 h−1 [20]. The catalyst Ni-Co/SAPO-11, containing 3% Ni and 6% Co, exhibited optimal catalytic properties, providing 100.0% conversion of FAMEs, 93.0% selectivity to C15-C18 hydrocarbons, and 36.1% of the isomerization ratio at 400 °C.
Ni/SAPO-11 and Ni2P/SAPO-11 catalysts were compared in hydroconversion of methyl laurate (ML) at 320–380 °C, 1.0–5.0 MPa, WHSV of 2–8 h−1, and H2/ML molar ratio of 25 [21]. Ni2P/SAPO-11 exhibited higher stability in comparison to Ni/SAPO-11 in HDO of ML, but both catalysts lost hydroisomerization activity. The ML conversion was close to 100% at 360 °C, 3.0 MPa, and WHSV of 2 h−1, while selectivity to iso-undecane and iso-dodecane decreased from 36.9% to 28.6% on Ni2P/SAPO-11 for 100 h. It was shown that sintering of Ni particles and formation of carbonaceous deposit were observed on spent Ni/SAPO-11, while no obvious increase of Ni2P particles took place, and carbonaceous deposit was a reason for the deactivation hydroisomerization activity of Ni2P/SAPO-11 [21].
Sulfided NiMo/SAPO-11 and NiMo/AlSBA-15 catalysts were studied in the hydroconversion of methyl stearate at 300–375 °C, 3 MPa, LHSV of 10 h−1, and volume H2/feed ratio of 600 [12]. Both NiMo catalysts provided high HDO conversion (above 99%) and isomerization activities, but NiMo/SAPO-11 exhibited a higher yield of iso-alkanes, while NiMo/AlSBA-15 catalysts additionally promoted the formation of cracked products. The authors conclude that moderate acidity and a suitable pore size of SAPO-11 provide for the formation of mono-branched isomers. MCM-41-supported sulfide Mo, CoMo, and NiMo catalysts yielded C4, C5, and C6 hydrocarbons and branched C7 and C8 hydrocarbons in the hydrodeoxygenation of octanoic acid at 330 °C and 1.6 MPa. The authors concluded that the higher activity of the sulfided NiMo/MCM-41 catalyst in isomerization/cracking reactions is associated with the promoting effect of MCM-41 on the acidity of the sulfided phase [22].
The efficiency of unpromoted sulfide MoS2 was demonstrated recently in MP hydrodeoxygenation [23], wherein alumina-supported MoS2 demonstrated high selectivity for the conversion of aliphatic esters through the direct HDO route without the formation of carbon oxides [24,25,26]. This property allows for the avoidance of the effect of carbon oxides on the catalyst lifetime and the additional purification of the recycle gas from COx [27]. To the best of our knowledge, the activity of MoS2 nanoparticles dispersed on zeolite-containing supports has not been studied yet in the hydroprocessing of aliphatic esters.
The purpose of this work is the comparative study of sulfide Mo-containing catalysts supported on granulated composite supports differing in the nature of the zeolite (Al2O3-ZSM-5, Al2O3-ZSM-12, Al2O3-SAPO-11, and Al2O3-ZSM-22) in the hydroprocessing of methyl palmitate. The MoS2/Al2O3-Z catalysts have been prepared with the use of organic additives, ensuring the high dispersion of MoS2 nanoparticles after proper sulfidation with DMDS/dodecane solution. Characterization has been performed using a wide set of different techniques to compare the MoS2 size/location and the acidity of the support depending on zeolite and to elucidate the possible correlations between physico-chemical and catalytic properties.

2. Results and Discussion

2.1. Catalyst Characterization

The textural properties of the synthesized supports and Mo content in the catalysts are listed in Table 1. All supports had similar textural properties: surface area—about 170 m2/g, pore volume—about 0.5 cm3/g, and pore diameter—above 20 nm. The prepared catalysts contained about 7.0 wt.% molybdenum. Such a Mo concentration was chosen to get a monolayer on the support surface (4.0 at Mo/nm2) [28]. Wherein it was considered that MoS2 localized predominantly on the alumina surface.
According to the XRD data of the synthesized supports, the alumina and corresponding zeolite diffraction lines were clearly observed (Figure 1). The supports contained a nanocrystalline alumina phase of γ-Al2O3 (PDF № 00-029-0063, the cubic cell parameter was a = 7.915 Å, the determined average size of coherently scattering domain was 7.5 nm) and corresponding crystalline phase of zeolite ZSM-5 (PDF# 00-044-0003, the determined average size of coherently scattering domain was 80 nm), ZSM-12 (PDF# 00-086-2634, a = 24.863 Å, b = 5.012 Å, c = 24.372 Å β = 107.7⁰, the determined average size of coherently scattering domain was 45 nm), ZSM-22 (PDF# 00-038-0197, a = 13.83 Å, b = 17.41 Å, c = 5.042 Å, the determined average size of coherently scattering domain was 45 nm), and silicoaluminophosphate SAPO-11 Al2Si0.35P1.74 O8.05 (PDF# 00-047-0614, the determined average size of coherently scattering domain was 70 nm).
SEM pictures of zeolite materials and the final supports of Al2O3-zeolite are shown in Figure 2. Zeolite fragments presented on SEM images of composites support evidence of the preservation of zeolite structure in synthesized supports of Al2O3-zeolite. Moreover, EDX mapping of Al2O3-zeolite supports demonstrates a uniform distribution of zeolite in support granules. Zeolites in the synthesized supports of Al2O3-zeolite display different average particle sizes (930, 1010, 300, and 220 nm for ZSM-5, ZSM-12, SAPO-11, and ZSM-22), and their histograms of particle size distribution are given in Figure 3 (the scale was chosen so that the difference was visually seen).
The hydroxyl cover of Al2O3 and Al2O3-zeolite supports was studied by FTIR spectroscopy (Figure 4). The spectrum of pure alumina shows the vibration bands at ca. 3790, 3775, 3727, 3700–3685, and 3660 cm−1, which are typical for the FTIR spectrum of surface OH groups of γ-Al2O3 [29] and characterize the different types of terminal Al-OH and bridged Al−O(H)-Al groups. The spectra of alumina-zeolite composites present two groups of signals in the region of O-H stretching vibrations assigned to the hydroxyl groups of the zeolites and the alumina binder. The intensity of bands at 3790, 3770, 3727, and 3685–3700 cm−1 in the spectra of composites (except for the Al2O3-SAPO-11) is proportional to the binder content. In the spectrum of the Al2O3-SAPO-11 sample, a decrease in the intensity of the bands of binder hydroxyl is observed, possibly caused by the interaction of phosphate ions from SAPO-11 both with Al-OH and Al-O(H)-Al groups of alumina. The signal at 3676 cm−1 in the spectrum of this composite characterizes P-OH groups either in the structure of the PO4 tetrahedron at the external surface of silica aluminophosphates [30] or at the surface of PO4-doped alumina [31]. The framework Si-O(H)-Al groups of SAPO-11, corresponding to strong Brønsted acid sites (BAS), appear at 3628 cm−1 for Al2O3-SAPO-11 composite in accordance with [30] at 3602 cm−1 for Al2O3-ZSM-22 [32], at 3612 cm−1 for Al2O3-ZSM-5 [33] and at 3612 and 3575 cm−1 for Al2O3-ZSM-12 composites [34]. The intensity of the bands of bridged hydroxyls in zeolite channels for Al2O3-ZSM-5 and Al2O3-ZSM-22 composites is significantly higher than for Al2O3-ZSM-12 and Al2O3-SAPO-11 ones. The bands of hydroxyl groups attached to partially extra-framework Al-OH species of zeolites overlap with the peaks of bridged Al−O(H)-Al groups of Al2O3. The bands at 3745 and 3738–3740 cm−1 in the spectra of Al2O3-zeolite extrudates are assigned to terminal silanols and defect Si-OH groups located in close vicinity to the lattice imperfection or Lewis acid sites at the external surfaces of zeolite crystals [35], respectively.
The acid properties of the Al2O3-zeolite supports were studied by FTIR spectroscopy with progressive CO adsorption at liquid nitrogen temperature. Adsorption of CO on pure Al2O3 at low pressures (spectra not shown) leads to the appearance of bands at 2241, 2235, 2220–2218, and 2208–2206 cm−1, assigned to the coordinately bonded CO complexes with strong and moderate Lewis acid sites (LAS) [36]. An increase in CO pressure leads to the appearance of a band at about 2200 cm−1, red shifted to 2184–2186 cm−1 at increasing coverage, which is attributed to the CO complex with weak LAS of alumina. The signals at 2164 and 2158–2156 cm−1 indicate CO complexes with different types of Al−OH groups. The spectra of CO adsorbed on Al2O3-zeolite supports present bands related to CO adsorption both on pure alumina and on zeolites (Figure 5). The bands at 2225–2230 cm−1, which are attributed to the complexes of CO with strong LAS of zeolites, overlap with the same bands of CO complex with Al2O3 species. The concentration of strong and moderate LAS in Al2O3-zeolite composites varies insignificantly (except Al2O3-SAPO-11); the amount of weak LAS is the same and proportional to the alumina content in the composites. An increase in the concentration of moderate LAS with the band at 2206 cm−1, apparently related to Al3+ species modified by PO42- groups [31], is observed for the Al2O3-SAPO-11 support (Figure S1 in the Supplementary Materials).
The spectra of Al2O3-zeolite samples demonstrate additional signals at 2178–2170 and 2137–2138 cm−1 at the CO stretching region compared to the spectra of pure alumina. The first group of bands refers to CO complexes with BAS; the second peak characterizes physically or liquid-like adsorbed CO molecules in zeolite channels [33]. The spectra of Al2O3-ZSM-5 and Al2O3-ZSM-22 supports exhibit one signal for CO complexes with strong BAS at 2176 cm−1, which corresponds to the spectra of pure zeolites [37,38,39]. The spectra of Al2O3-SAPO-11 composite present one band for CO complexes with BAS at 2173 cm−1, red shifted to 2170 cm−1 at increasing coverage, which corresponds to moderate BAS in accordance with the value of CO-induced blue shift relative to the CO gas phase (ΔνCO = 30–27 cm−1). The band at 2178 cm−1 in the spectra of Al2O3-ZSM-12 support is assigned to the CO complex with strong BAS, while the signal at 2171 cm−1 belongs to the CO complex with moderate BAS. The two types of BASs, framework Si-O(H)-Al groups and extra-framework Al-OH groups, are also observed in the spectra of the original ZSM-12 zeolite [34]. The CO complex with Na+ impurities in the ZSM-12 zeolite additionally increases the intensity of the band at 2171–2170 cm−1 in the case of the Al2O3-ZSM-12 composite.
The weakly basic CO molecule is known to be a good probe molecule for testing the strength of BAS in zeolites and related materials [40]. During low-temperature CO adsorption on Al2O3-zeolite samples, the bands of acidic OH groups fully disappeared, and a new band appeared (Figure 6). The red frequency shift of OH stretching vibration at hydrogen bonding with carbon monoxide is traditionally used to estimate the acidity of hydroxyl groups. A new positive peak at about 3285 and 3300 cm−1 appears in the spectra of Al2O3-ZSM-22 and Al2O3-ZSM-5 samples, respectively, at low CO pressure. The value of the red frequency shift of the bands from the framework Si-O(H)-Al groups at hydrogen bonding with CO (ΔνOH…CO) is 317–320 cm−1 and similar to the magnitude for initial zeolites [37,38,39]. The corresponding blue frequency shift of the CO stretching bands for these composites is also the same (ΔνCO = 33 cm−1), which indicates a similar high acidity of the bridged hydroxyls. Quantitative data on BAS concentration and acid strength are given in Table 2. A large concentration of strong BAS for the Al2O3-ZSM-22 composite is obviously associated with a lower Si/Al ratio in the structure of the zeolite used. The shoulder at about 3400 cm−1 in the spectra of Al2O3-ZSM-5 and Al2O3-ZSM-22 supports changing in synchrony with the band at 3285–3300 cm−1 due to Fermi resonance [41]. Other positive bands appear at 3470–3480 cm−1 in the spectra of these composites and are related to hydrogen-bonded CO complexes with defect silanol groups (Si-O(H)…Al3+). The shift values (ΔνOH…CO = 270–260 cm−1) are slightly lower than the magnitude typical for bridged Si-O(H)-Al groups in the zeolite channel and correspond to moderately strong BAS. The concentration of these sites is negligible.
After CO adsorption on the Al2O3-ZSM-12 sample, the appearance of the new positive signal at 3285 cm−1 is observed. The values of the red frequency shift in the hydroxyl region (ΔνOH…CO = 333 cm−1) and the corresponding blue frequency shift in the carbonyl region (ΔνCO = 35 cm−1) in the spectra of this support are assigned to BAS with enhanced acidity that bridged hydroxyl in the zeolite channel for Al2O3-ZSM-22 and Al2O3-ZSM-5 composites. The magnitude of shifts is slightly higher than in pure zeolite [34]. The low concentration of strong BAS for the Al2O3-ZSM-12 composite compared to the Al2O3-ZSM-5 composite may be due to their partial exchange with Na+ impurities. The other band in the O–H stretching region after CO adsorption on the Al2O3-ZSM-12 composite is detected at ~3460–3490 cm−1 and is related to perturbation both of the extra-framework Al-OH groups of zeolites with the band at about 3670–3675 cm−1 and the defect silanols with the band at 3738–3740 cm−1. Apparently, ZSM-12 zeolite is partially dealuminated. According to the values of the red frequency shift, the extra-framework Al-OH groups in this zeolite are BAS with medium strength. Progressive CO adsorption on the Al2O3-SAPO-11 composite leads to the appearance of a strong positive band at 3378 cm−1, with the shoulder at 3470 cm−1, due to perturbation of bridged Si-O(H)-Al groups in the zeolite channels. The shift value (ΔνOH…CO = 258 cm−1) is significantly lower than the magnitude typical for bridged Si-O(H)-Al groups in the pure SAPO-11 channel (ΔνOH…CO = 310 cm−1) [42] and corresponds to moderately strong BAS. The change in the acidity of bridged Si-O(H)-Al groups in the zeolite channels can be probably caused by disruption of the SAPO-11 structure by the partial removal of phosphate groups during the molding of extrudates. The red shift of P-OH groups in zeolites after CO adsorption (ΔνOH…CO = 202 ÷ 198 cm−1) corresponds to somewhat weaker Brønsted acid sites.
Thus, the strength of framework BAS (bridged Si-O(H)-Al groups in zeolite channels) decreases in the series of Al2O3-zeolite supports as Al2O3-ZSM-12 > Al2O3-ZSM-22 ~ Al2O3-ZSM-5 >> Al2O3-SAPO-11, while the concentration of strong and moderate BAS of zeolites decreases in the following order: Al2O3-ZSM-22 > Al2O3-SAPO-11 > Al2O3-ZSM-5 >> Al2O3-ZSM-12.
In FTIR difference spectra during adsorption of CO on pure alumina, there are no bands in the region of 3200–3500 cm−1 [43], which are characteristic of CO complexes with BAS of zeolites. The terminal Al-OH groups of alumina are traditionally assigned basic properties, while the bridging hydroxyls have been shown to have weak acidic properties (ΔνOH…CO = 130 ÷ 100 cm−1). The formation of CO complexes with the Al-O(H)-Al groups of alumina during CO adsorption on Al2O3-zeolite supports occurs after saturation of the zeolite BAS (Figure S2 in the Supplementary Materials).
According to HRTEM images (Figure 7), a dispersed sulfide phase is presented on the surfaces of the sulfided catalysts, which is visualized as a black line (edges of MoS2 particles). The average size of nanoparticles varied from 4 to 6 nm; the stacking number was 1.5–1.7 for all catalysts. It should be noted that MoS2 nanoparticles were predominantly located on the alumina surface, and only a single species was presented on the surface of zeolite. This statement is illustrated for MoS2/Al2O3-SAPO-11 and MoS2/Al2O3-ZSM-22 catalysts in Figure 7. EDX mapping confirms this statement: sulfide species (Figure 8, green color) are more prevalent on alumina surfaces in comparison with zeolite surfaces (Figure 8, red color), where sulfide particles are far less prevalent.

2.2. The Effect of Zeolite Type on Hydrodeoxygenation of Methyl Palmitate

The conversion of fatty acid esters can follow two routes: ‘direct’ hydrodeoxygenation (‘direct’ HDO) and hydrodecarboxylation/hydrodecarbonylation (DeCOx). In the presence of a MoS2 catalyst, the conversion of fatty acid esters proceeded mainly via a ‘direct’ hydrodeoxygenation pathway to form hexadecane (C16H34) and water, with the formation of carbon oxides only in trace amounts [25,44,45].
Hydrodeoxygenation (HDO) of methyl palmitate (MP) was performed at a temperature range of 250–350 °C, at H2 pressure of 3.0 MPa, H2/feed ratio of 600 Nm3/m3, and LHSV of 36 h−1. Methyl palmitate conversion is increased with the temperature rising from 250 to 310 °C (Figure 9). Hexadecanol, hexadecanal, palmitic acid, palmityl palmitate, and methyl hexadecyl were detected as oxygen intermediate products over MoS2/Al2O3-zeolite catalysts in MP hydrodeoxygenation, in consistency with the previous results [23,24,46]. At a temperature range of 250–290 °C, normal and unsaturated C15-C16 alkanes were also observed.
Figure 9 shows that the addition of zeolite to alumina has a slight influence on MP conversion. Conversions of all-oxygen-containing compounds, including both intermediates and methyl palmitate, were calculated using the contents of oxygen in the reaction mixture before and after the reaction by means of elemental analysis (Equations (1) and (2)), and the results are presented in Figure 10. According to these results, the addition of zeolite to the support leads to an increase in the conversion of oxygen-containing compounds. Taking into account that the conversion of methyl palmitate weakly depends on the composition of the carrier, we can conclude that the addition of zeolite leads to an acceleration of the HDO reactions of intermediate oxygen-containing compounds [12].
Complete MP and oxygen conversion were achieved at 310 °C in the presence of all catalysts (Figure 9 and Figure 10). Normal and iso-alkanes (C15 and C16) were detected under conditions where complete oxygen conversion was achieved (at temperatures above 310 °C). Cracked products were detected in negligible amounts over MoS2/Al2O3-SAPO-11 and MoS2/Al2O3-ZSM-22 catalysts: 2 and 4% at 350 °C, respectively. The maximum yield of cracked products was observed for MoS2/Al2O3-ZSM-5 (18%) and MoS2/Al2O3-ZSM-12 (12%) catalysts.
The selectivity for the conversion of methyl palmitate via the ‘direct’ hydrodeoxygenation route in the presence of MoS2/Al2O3-zeolite catalysts was over 85% (Figure 11). Temperature increase leads to a decrease in the selectivity of the C16H34 formation via the ‘direct’ HDO route over all catalysts due to occurring DeCOx reactions (Figure 11) [47]. It can be seen that the addition of zeolite to alumina resulted in an enhancement of the DeCOx route in the hydroprocessing of MP over sulfide catalysts (Figure 11). The lowest HDO selectivity was observed over the MoS2/Al2O3-ZSM-22 catalyst. It can be explained by the highest concentration of strong BAS on the Al2O3-ZSM-22 support surface (Table 2) that could favor hydrodecarboxylation/hydrodecarbonylation reactions of methyl palmitate. Methane and negligible amounts of carbon monoxide were detected in the gas phase.
The catalyst stability in the hydrodeoxygenation of methyl palmitate was checked after 40 h at a temperature of 290 °C. Oxygen conversion was changed slightly: from 81.5 to 80.0% for MoS2/Al2O3-ZSM-5, from 81.2 to 77.0% for MoS2/Al2O3-ZSM-12, from 85.5 to 82.9 for MoS2/Al2O3-ZSM-22, and from 72.7 to 68.0% for MoS2/Al2O3-SAPO-11. Thus, the change in catalyst activity during the experiment can be neglected.

2.3. The Effect of Zeolite Type on Hydroisomerization of Methyl Palmitate

The isomerization process over MoS2/Al2O3-zeolite catalysts was studied under the conditions of complete conversion of oxygenates, i.e., temperature above 310 °C and pressure between 3.0 and 5.0 MPa. The catalytic activity of the sulfide samples during the hydroisomerization of methyl palmitate was compared by the yield of isomeric C16H34 and C15H32 alkanes in the reaction products.
According to the obtained results, the yield of iso-alkanes gradually increases in the following order: MoS2/Al2O3 < MoS2/Al2O3-ZSM-12 < MoS2/Al2O3-ZSM-5 < MoS2/Al2O3-SAPO-11< MoS2/Al2O3-ZSM-22: yield of iso-alkanes did not exceed 5% over MoS2/Al2O3, 13.5% and 7.4% for MoS2/Al2O3-ZSM-5 and MoS2/Al2O3-ZSM-12 samples, accordingly at 310 °C, 3.0 MPa, 600 Nm3/m3, 36 h−1. In the presence of the MoS2/Al2O3-SAPO-11 catalyst, the yield of iso-alkanes increases to 24%. The most active catalyst in MP hydroisomerization was MoS2/Al2O3-ZSM-22, with a yield of isomerized C16H34 and C15H32 alkanes of 40% (Figure 12). The observed sequence coincides with the increase in the BAS concentration order of zeolite-containing supports: Al2O3-ZSM-12 << Al2O3-ZSM-5 < Al2O3-SAPO-11 Al2O3-ZSM-22. The hydroisomerization activity of sulfide catalysts is proportional to the number of BAS [12,48].
It was observed that catalytic properties depend not only on the acidity of samples but also on the pore structure and framework topology of zeolites in the catalyst’s composition [49]. The MP molecule has a length of 22 Å and a width of 2.2 Å (Figure 13). According to the literature, data to isomerize MP molecules should be available to diffuse into pores and channels of zeolite [12]. Catalytic experiments showed that catalysts prepared with ZSM-22 and SAPO-11 demonstrated better performance in the hydroisomerization of methyl palmitate. It is correlated with the BAS concentration of synthesized zeolite-containing supports. Moreover, the better performance of ZSM-22- and SAPO-11-containing catalysts could probably be explained by the smaller average crystallite size of zeolite in comparison with catalysts prepared with ZSM-5 and ZSM-12 (Figure 3). We can propose that zeolite with a smaller crystallite size gives a more uniform (homogeneous) distribution in the support, which in turn provides closer proximity between zeolite and sulfide entities. There is no consensus in the literature about the influence of zeolite particle size on the efficiency of zeolite-containing catalysts in hydroprocessing [50,51,52,53,54]. Acidity is probably a more significant factor than the pore structure and framework topology of zeolites.
A temperature increase from 310 to 350 °C resulted in a decrease in iso-alkane yield over all MoS2/Al2O3-zeolite catalysts: from 40% to 26% over MoS2/Al2O3-ZSM-22; from 24% to 14% in the presence of MoS2/Al2O3-SAPO-11 catalyst (Figure 12). Catalytic experiments showed a decrease in iso-alkane yield with a temperature rise accompanied by an increase in normal C16 and C15 alkanes, while the content of cracked products changed slightly under the reaction conditions. Currently, we do not have a reasonable explanation for the observed dependence; a thorough study of the mechanism of ester and HDO intermediate transformation may help to elucidate this issue in the future.
In addition, the effect of pressure (3.0 and 5.0 MPa) on the MP hydroisomerization over catalysts containing ZSM-22 zeolite and SAPO-11 was also investigated. The reaction was carried out at a temperature of 350 °C, an LHSV of 36 h−1, and a H2/feed ratio of 600 Nm3/m3. A pressure increase from 3.0 to 5.0 promoted MP conversion via the ‘direct’ HDO route: HDO selectivity increased from 88.4 to 90.7% over MoS2/Al2O3-SAPO-11 and from 85.4 to 88.9% over MoS2/Al2O3-ZSM-22 catalyst, in agreement with previous results [15,25,47]. The yield of iso-alkanes decreases with pressure increase from 26% to 14.5% over MoS2/Al2O3-ZSM-22 catalyst and from 15% to 10% over MoS2/Al2O3-SAPO-11 sample (Figure 14). The reason for this could be the acceleration of hydrogenation of olefins, which, according to the generally accepted mechanism, are intermediate products in hydroisomerization and hydrocracking reactions [12].
In the literature, pressure and temperature increases resulted in an increase in the iso-alkane yield [57]. The authors performed MP hydrotreating at a high temperature of 350–410 °C and a pressure of 6.0–12.0 MPa over sulfided MoO3/ZrPOx in a batch reactor. High temperatures activated the stable alkanes, and the yield of iso-alkanes increased. Our catalytic tests were performed in a lower temperature and pressure range. A decrease in the activity of sulfide catalysts in the hydroisomerization of methyl palmitate was observed with increasing pressure and temperature, which is related to the reaction mechanism. The conversion of methyl palmitate over sulfide catalysts is quite complex, including hydrodeoxygenation and hydroisomerization reactions. Presumably, alkane isomers are formed not from the final product of hydrodeoxygenation (n-hexadecane), but from intermediate products of methyl palmitate conversion (alcohol and olefins).

3. Materials and Methods

3.1. Support Preparation

Four high-silica zeolite powders with different framework types were used to prepare catalysts (Table 3). All samples (except ZSM-12) were purchased from Zeolyst Corp.
The synthesis of zeolite ZSM-12 was carried out using the following reagents: a colloidal solution of silicon dioxide LUDOX HS-40 (40 wt.%, Sigma-Aldrich), aluminum sulfate octadecahydrate (Al2(SO4)3·18H2O, Sigma-Aldrich, 99%), methyltriethylammonium chloride ([CH3N(C2H5)3]Cl, Sigma-Aldrich, 97%, abbreviated [MTEA]Cl), sodium hydroxide (NaOH, Komponent-Reaktiv, 98%), and ammonium nitrate (NH4NO3, Khimmed, 98%). Solution A, consisting of 12.6 g of distilled water, 0.4 g of Al2(SO4)3·18H2O, 1 g of NaOH, and 3.3 g [MTEA]Cl used as a template, was stirred until all of the components became completely dissolved. Solution B, consisting of 25.2 g of a 40% (wt.) colloidal solution of silicon dioxide of the brand LUDOX HS-40 and 10.1 g of distilled water, was stirred until the reaction mixture was homogeneous. Solution A was dropped into solution B and stirred gently. The gel was poured into a Teflon liner and placed into the autoclave, which was heated at 155 °C for 120 h. The product was filtered off, washed with distilled water, dried at 110 °C for 12 h, and calcined at 550 °C for 10 h (heating rate 1 deg∙min−1). To obtain the H-form of zeolite, the synthesized material was treated 3 times with a 1 M aqueous solution of NH4NO3 at 80 °C for 17 h. The solid product was filtered off, washed with distilled water, dried at 110 °C for 12 h, and calcined at 550 °C for 8 h (heating rate 1 deg∙min−1).
Alumina support was prepared by HNO3 peptization of pseudoboehmite (Disperal 20, Sasol Germany GmbH, Hamburg, Germany). Zeolite-containing granular supports were prepared by mixing pseudoboehmite (Disperal 20, Sasol Germany GmbH, Hamburg, Germany) and zeolite powders, followed by peptization with nitric acid, and then piston extrusion through a trefoil-shaped die. After extruding, support granules were dried at 110 °C for 12 h and then calcined at 550 °C in air flow for 6 h. Zeolite content was 30 wt.% in all calcined composite supports. Synthesized supports were denoted as Al2O3-ZSM-5, Al2O3-ZSM-12, Al2O3-ZSM-22, and Al2O3-SAPO-11.

3.2. Catalyst Preparation

Mo catalysts were prepared by incipient wetness impregnation of synthesized alumina and zeolite-containing extrudates by aqua solution containing ammonium heptamolybdate ((NH4)6Mo7O24 4H2O from Vekton, Saint Petersburg, Russia) and citric acid monohydrate (C6H8O7⋅H2O from Vekton, Saint Petersburg, Russia). Mo content was about 7.0 wt.% after calcination of the catalysts at 550 °C for 4 h.

3.3. Support and Catalyst Characterization

The textural properties of the synthesized supports were determined using nitrogen physisorption at 77 K with an Autosorb-6B-Kr instrument (“Quantachrome Instruments”, New York, NY, USA).
The elemental analysis was performed using inductively coupled plasma atomic emission spectroscopy (ICPAES) on Optima 4300 DV (“Perkin Elmer”, Norwalk, CT, USA). The Mo content was determined after calcination of the catalysts at 550 °C for 4 h.
X-ray powder diffraction (XRD) patterns of supports and catalysts were obtained with an instrument STOE STADI MP (“STOE”, Darmstadt, Germany) with a detector MYTHEN2 1K using MoKα radiation (wavelength λ = 0.7093Å). The measurements were carried out in a range of 2θ from 2 to 40°, with a scanning step of 0.015°.
The acidity of Al2O3-zeolite supports and pure Al2O3 was characterized by FTIR spectroscopy of adsorbed carbon monoxide. FTIR spectra were recorded on a Shimadzu FTIR-8300 spectrometer (Shimadzu, Tokyo, Japan) within the spectral range of 700–6000 cm−1, resolution of 4 cm−1, and 300 scans for signal accumulation. The powder samples were pressed into thin, self-supporting wafers of 0.010–0.012 g × cm−2 density and pretreated in a home-made IR cell at 500 °C for 2 h under a dynamic vacuum of less than 10−3 Pa. In the presented spectra, the absorbance was normalized to sample wafer density. CO was introduced at liquid nitrogen temperature in doses from a low pressure of 0.1 mbar up to an equilibrium pressure of 10 mbar. The concentration of Brønsted acid sites (BAS) was determined from the integral intensity of the bands assigned to hydrogen-bonded complexes of CO molecules with the OH groups using the following molar integral absorption coefficient values: A0 = 54 cm/μmol for the complexes with νOH...CO ~ 3280–3380 cm−1 and A0 = 27 cm/μmol for the complexes with νOH...CO 3500 cm−1 [61].
The morphology of supports was studied using a Hitachi Regulus SU8230 FESEM scanning electron microscope (Hitachi, Tokyo, Japan) with an accelerating voltage of 2 and 5 kV in the modes of secondary (SE) and backscattered (BSE) electrons using an upper (U) detector, which makes it possible to obtain microscopic images in phase and topographic contrasts. The study of the chemical composition was also carried out on a Hitachi Regulus SU8230 FESEM scanning electron microscope (Hitachi, Tokyo, Japan) with an accelerating voltage of 20 kV. The device is equipped with an AztecLive (Oxford Instruments, Oxford, UK) energy-dispersive X-ray characteristic spectrometer (EDX) with a semiconductor Si detector with an energy resolution of 128 eV.
The morphology of the sulfide phase of the catalysts after hydroprocessing was studied by high-resolution transmission electron microscopy (HRTEM) using a ThemisZ electron microscope (“Thermo Fisher Scientific”, Waltham, MA, USA) with an accelerating voltage of 200 kV and a limiting resolution of 0.07 nm. Images were recorded using a Ceta 16 CCD array (“Thermo Fisher Scientific”, Waltham, MA, USA). The instrument is equipped with a SuperX (“Thermo Fisher Scientific”, Waltham, MA, USA) energy-dispersive characteristic X-ray spectrometer (EDX) with a semiconductor Si detector with an energy resolution of 128 eV. To obtain statistical information, the structural parameters of ca. 500 particles were measured.

3.4. Catalytic Experiments

The catalytic experiments were performed using an experimental setup with a trickle-bed reactor with an inner diameter of 12 mm and a length of 370 mm. In each experiment, 0.5 mL of catalyst (0.25–0.50 mm size fraction) was diluted with inert material, carborundum (0.1–0.25 mm size fraction), in a 1:8 volume ratio. Prior to the catalytic experiments, the catalysts were activated by in-situ sulfidation with dimethyl disulfide in dodecane (0.6 wt.% sulfur) at H2 pressure—3.5 MPa, H2/feed ratio—300 Nm3/m3, and LHSV—20 h−1. Sulfidation was performed at a temperature of 340 °C for 4 h with a heating rate of 25 °C/h.
Hydroprocessing of methyl palmitate was carried out at a temperature range of 250–350 °C, H2 pressure of 3.0 and 5.0 MPa, H2/feed ratio of 600 Nm3/m3, and LHSV of 36 h−1. The feed was 10 wt.% of methyl palmitate (1.17 wt.% O) with dimethyl disulfide (0.6 wt.% sulfur) in dodecane. Dimethyl disulfide was added to the feedstock to maintain the sulfide form of the active component (MoS2). The duration of each step was 6 h.
To check catalyst stability, oxygen conversion was compared in the first and last stages carried out in the same conditions in each experiment (290 °C, 3.0 MPa, H2/feed ratio of 600 Nm3/m3, and LHSV of 36 h−1).

3.5. Product Analysis

The products of methyl palmitate (MP) conversion were analyzed using an Agilent 6890N gas chromatograph (“Agilent Technologies”, USA, Wilmington) equipped with a flame ionization detector and an HP-1MS quartz capillary column (30 m × 0.32 mm × 1 μm). Methyl palmitate conversion was calculated as follows (1):
X M P = C M P 0 C M P C M P 0 × 100 % ,
where C M P 0 is the chromatogram peak area of MP in the feed and C O is the chromatogram peak area of MP in the final product.
The total oxygen content in liquid samples was determined using a Vario EL Cube elemental CHNSO analyzer (“Elementar Analysensysteme GmbH”, Langenselbold, Germany).
Oxygen conversion was calculated as (2):
X O = C O 0 C O C O 0 × 100 % ,        
where C O 0 is the total oxygen content in the feed and C O is the total oxygen content in the final product.
The gas phase during the MP hydroprocessing was analyzed online using a gas chromatograph Chromos 1000 (“Chromos”, Omsk, Russia) equipped with a methanator and a flame ionization detector.
The selectivity of the ‘direct’ HDO route (HDO selectivity) was calculated as follows (3):
S = C 16 C 16 + C 15 × 100 % ,
where C16 is the content of C16 alkanes (normal + iso) in the final product, and C15 is the content of C15 alkanes (normal + iso) in the final product.
The yield of iso-alkanes was calculated as follows (4):
Y i s o = i C 16 + i C 15 ( C 16 + C 15 ) 100 % ,
where i-C16 and i-C15 are the contents of iso-C16 and iso-C15 alkanes in the final product at complete oxygen conversion; ( C 16 + C 15 ) represents the sum of normal and iso-alkanes at complete oxygen conversion.

4. Conclusions

Synthesized composite zeolite-containing supports (30 wt.% zeolite and 70% Al2O3) and corresponding sulfide Mo-containing catalysts were characterized by XRD, HRTEM, and SEM. According to XRD data, the structure of zeolites was preserved in synthesized supports and catalysts. The uniform distribution of zeolite crystallites in composite materials (Al2O3-ZSM-5, Al2O3-ZSM-12, Al2O3-ZSM-22, and Al2O3-SAPO-11) was confirmed by SEM-EDX. 100% conversion of oxygen was observed at 310 °C over sulfided Mo/Al2O3-zeolite catalysts in the hydroprocessing of methyl palmitate. A temperature rise from 310 to 350 °C resulted in a decrease in HDO selectivity. It was found that the addition of zeolite to alumina has a slight influence on MP conversion, but the effect on the conversion of oxygen-containing compounds is greater. The activity of MoS2/Al2O3-zeolite catalysts in the production of isomerized alkanes in MP hydroconversion is in good correlation with the concentration of Brønsted acid sites. The yield of iso-alkanes in the hydroisomerization of MP increased in the following order: Al2O3 < Al2O3-ZSM-12 < Al2O3-ZSM-5 < Al2O3-SAPO-11 < Al2O3-ZSM-22. The yield of iso-alkanes was affected by temperature and hydrogen pressure. An increase in temperature and pressure resulted in a decrease in iso-alkane yield. This observation can probably be explained by the reaction mechanism under the given reaction conditions. Iso-alkanes are formed from intermediate products of methyl palmitate HDO, not from alkanes.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241914863/s1.

Author Contributions

Data curation and writing—original draft preparation, E.V. and A.N.; investigation, Y.Z., I.S. and P.A.; characterization by FTIR spectroscopy, I.D.; characterization by XRD, V.P.; characterization by SEM, E.S.; preparation of ZSM-12, D.T.; supervision, A.M.; supervision and writing—review and editing, G.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Russian Science Foundation, grant number 22-13-00371.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data is available upon request from the corresponding authors.

Acknowledgments

The authors are grateful to E. Yu. Gerasimov for HRTEM analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Panoutsou, C.; Germer, S.; Karka, P.; Papadokostantakis, S.; Kroyan, Y.; Wojcieszyk, M.; Maniatis, K.; Marchand, P.; Landalv, I. Advanced biofuels to decarbonise European transport by 2030: Markets, challenges, and policies that impact their successful market uptake. Energy Strategy Rev. 2021, 34, 100633. [Google Scholar] [CrossRef]
  2. Douvartzides, S.L.; Charisiou, N.D.; Papageridis, K.N.; Goula, M.A. Green Diesel: Biomass Feedstocks, Production Technologies, Catalytic Research, Fuel Properties and Performance in Compression Ignition Internal Combustion Engines. Energies 2019, 12, 809. [Google Scholar] [CrossRef]
  3. Mittelbach, M. Fuels from oils and fats: Recent developments and perspectives. Eur. J. Lipid Sci. Technol. 2015, 117, 1832–1846. [Google Scholar] [CrossRef]
  4. Vasquez, M.C.; Silva, E.E.; Castillo, E.F. Hydrotreatment of vegetable oils: A review of the technologies and its developments for jet biofuel production. Biomass Bioenergy 2017, 105, 197–206. [Google Scholar] [CrossRef]
  5. Goh, B.H.H.; Chong, C.T.; Ge, Y.; Ong, H.C.; Ng, J.-H.; Tian, B.; Ashokkumar, V.; Lim, S.; Seljak, T.; Józsa, V. Progress in utilisation of waste cooking oil for sustainable biodiesel and biojet fuel production. Energy Convers. Manag. 2020, 223, 113296. [Google Scholar] [CrossRef]
  6. Maki-Arvela, P.; Martinez-Klimov, M.; Murzin, D.Y. Hydroconversion of fatty acids and vegetable oils for production of jet fuels. Fuel 2021, 306, 121673. [Google Scholar] [CrossRef]
  7. Long, F.; Liu, W.G.; Jiang, X.; Zhai, Q.L.; Cao, X.C.; Jiang, J.C.; Xu, J.M. State-of-the-art technologies for biofuel production from triglycerides: A review. Renew. Sustain. Energy Rev. 2021, 148, 111269. [Google Scholar] [CrossRef]
  8. Yeletsky, P.M.; Kukushkin, R.G.; Yakovlev, V.A.; Chen, B.H. Recent advances in one-stage conversion of lipid-based biomass-derived oils into fuel components—Aromatics and isomerized alkanes. Fuel 2020, 278, 118255. [Google Scholar] [CrossRef]
  9. Hari, T.K.; Yaakob, Z. Production of diesel fuel by the hydrotreatment of jatropha oil derived fatty acid methyl esters over gamma-Al2O3 and SiO2 supported NiCo bimetallic catalysts. React. Kinet. Mech. Catal. 2015, 116, 131–145. [Google Scholar] [CrossRef]
  10. Bezergianni, S.; Dimitriadis, A.; Chrysikou, L.P. Quality and sustainability comparison of one- vs. two-step catalytic hydroprocessing of waste cooking oil. Fuel 2014, 118, 300–307. [Google Scholar] [CrossRef]
  11. Hsu, H.W.; Chang, Y.H.; Wang, W.C. Techno-economic analysis of used cooking oil to jet fuel production under uncertainty through three-, two-, and one-step conversion processes. J. Clean. Prod. 2021, 289, 125778. [Google Scholar] [CrossRef]
  12. Qian, E.W.; Chen, N.; Gong, S. Role of support in deoxygenation and isomerization of methyl stearate over nickel-molybdenum catalysts. J. Mol. Catal. A-Chem. 2014, 387, 76–85. [Google Scholar] [CrossRef]
  13. Akhmedov, V.M.; Al-Khowaiter, S.H. Recent advances and future aspects in the selective isomerization of high n-alkanes. Catal. Rev.-Sci. Eng. 2007, 49, 33–139. [Google Scholar] [CrossRef]
  14. Li, X.; Fan, Q.; Wu, Y.; Lin, X.; Ma, S.; Li, S.; Ye, Y.; Wang, D.; Cheng, J.; Zheng, Z. Enhancing hydrodeoxygenation-isomerization of FAME over M-SAPO-11 in one-step process: Effect of in-situ isomorphic substitution of transition metals and synergy of PtxSny alloy. Chem. Eng. J. 2023, 452, 139528. [Google Scholar] [CrossRef]
  15. Chen, L.G.; Li, H.W.; Fu, J.Y.; Miao, C.L.; Lv, P.M.; Yuan, Z.H. Catalytic hydroprocessing of fatty acid methyl esters to renewable alkane fuels over Ni/HZSM-5 catalyst. Catal. Today 2016, 259, 266–276. [Google Scholar] [CrossRef]
  16. Cheng, J.; Zhang, Z.; Zhang, X.; Fan, Z.T.; Liu, J.F.; Zhou, J.H. Continuous hydroprocessing of microalgae biodiesel to jet fuel range hydrocarbons promoted by Ni/hierarchical mesoporous Y zeolite catalyst. Int. J. Hydrogen Energy 2019, 44, 11765–11773. [Google Scholar] [CrossRef]
  17. Cheng, J.; Zhang, Z.; Zhang, X.; Liu, J.F.; Zhou, J.H.; Cen, K.F. Hydrodeoxygenation and hydrocracking of microalgae biodiesel to produce jet biofuel over H3PW12O40-Ni/hierarchical mesoporous zeolite Y catalyst. Fuel 2019, 245, 384–391. [Google Scholar] [CrossRef]
  18. Zhang, Z.; Cheng, J.; Zhu, Y.X.; Guo, H.; Yang, W.J. Jet fuel range hydrocarbons production through competitive pathways of hydrocracking and isomerization over HPW-Ni/MCM-41 catalyst. Fuel 2020, 269, 117465. [Google Scholar] [CrossRef]
  19. Li, T.; Cheng, J.; Zhang, X.; Liu, J.F.; Huang, R.; Zhou, J.H. Jet range hydrocarbons converted from microalgal biodiesel over mesoporous zeolite-based catalysts. Int. J. Hydrogen Energy 2018, 43, 9988–9993. [Google Scholar] [CrossRef]
  20. Xiao, Y.; Shang, J.; Zhai, M.; Qiao, C. Hydrodeoxygenation of fatty acid methyl esters and simultaneous products isomerization over bimetallic Ni-Co/SAPO-11 catalysts. Int. J. Energy Res. 2021, 45, 9648–9656. [Google Scholar] [CrossRef]
  21. Zhao, S.; Li, M.; Chu, Y.; Chen, J. Hydroconversion of Methyl Laurate as a Model Compound to Hydrocarbons on Bifunctional Ni2P/SAPO-11: Simultaneous Comparison with the Performance of NUSAPO-11. Energy Fuels 2014, 28, 7122–7132. [Google Scholar] [CrossRef]
  22. Kaluža, L.; Karban, J.; Gulková, D. Activity and selectivity of Co (Ni) Mo sulfides supported on MgO, Al2O3, ZrO2, TiO2, MCM-41 and activated carbon in parallel hydrodeoxygenation of octanoic acid and hydrodesulfurization of 1-benzothiophene. React. Kinet. Mech. Catal. 2019, 127, 887–902. [Google Scholar] [CrossRef]
  23. Cai, Z.; Ding, Y.; Zhang, J.; Yu, P.; Ma, Y.; Cao, Y.; Zheng, Y.; Huang, K.; Jiang, L. In situ generation of dispersed MoS2 catalysts from oil-soluble Mo-based ionic liquids for highly effective biolipids hydrodeoxygenation. J. Catal. 2023, 423, 50–61. [Google Scholar] [CrossRef]
  24. Deliy, I.V.; Vlasova, E.N.; Nuzhdin, A.L.; Gerasimov, E.Y.; Bukhtiyarova, G.A. Hydrodeoxygenation of methyl palmitate over sulfided Mo/Al2O3, CoMo/Al2O3 and NiMo/Al2O3 catalysts. Rsc Adv. 2014, 4, 2242–2250. [Google Scholar] [CrossRef]
  25. Kaluza, L.; Kubicka, D. The comparison of Co, Ni, Mo, CoMo and NiMo sulfided catalysts in rapeseed oil hydrodeoxygenation. React. Kinet. Mech. Catal. 2017, 122, 333–341. [Google Scholar] [CrossRef]
  26. Dupont, C.; Lemeur, R.; Daudin, A.; Raybaud, P. Hydrodeoxygenation pathways catalyzed by MoS2 and NiMoS active phases: A DFT study. J. Catal. 2011, 279, 276–286. [Google Scholar] [CrossRef]
  27. Kubickova, I.; Kubicka, D. Utilization of Triglycerides and Related Feedstocks for Production of Clean Hydrocarbon Fuels and Petrochemicals: A Review. Waste Biomass Valorization 2010, 1, 293–308. [Google Scholar] [CrossRef]
  28. Bergwerff, J.A.; Jansen, M.; Visser, T.; de Jong, K.P.; Weckhuysen, B.M. Influence of the preparation method on the hydrotreating activity of MoS2/Al2O3 extrudates: A Raman microspectroscopy study on the genesis of the active phase. J. Catal. 2006, 243, 292–302. [Google Scholar] [CrossRef]
  29. Busca, G. Structural, surface, and catalytic properties of aluminas. In Advances in Catalysis; Elsevier: Amsterdam, The Netherlands, 2014; Volume 57, pp. 319–404. [Google Scholar]
  30. Meriaudeau, P.; Tuan, V.; Nghiem, V.T.; Lai, S.; Hung, L.; Naccache, C. SAPO-11, SAPO-31, and SAPO-41 molecular sieves: Synthesis, characterization, and catalytic properties inn-octane hydroisomerization. J. Catal. 1997, 169, 55–66. [Google Scholar] [CrossRef]
  31. Morterra, C.; Magnacca, G.; Demaestri, P. Surface characterization of modified aluminas: III. Surface-features of PO4-doped Al2O3. J. Catal. 1995, 152, 384–395. [Google Scholar] [CrossRef]
  32. Verboekend, D.; Chabaneix, A.M.; Thomas, K.; Gilson, J.-P.; Pérez-Ramírez, J. Mesoporous ZSM-22 zeolite obtained by desilication: Peculiarities associated with crystal morphology and aluminium distribution. CrystEngComm 2011, 13, 3408–3416. [Google Scholar] [CrossRef]
  33. Kustov, L.; Kazanskii, V.; Beran, S.; Kubelkova, L.; Jiru, P. Adsorption of carbon monoxide on ZSM-5 zeolites: Infrared spectroscopic study and quantum-chemical calculations. J. Phys. Chem. 1987, 91, 5247–5251. [Google Scholar] [CrossRef]
  34. Dimitrov, L.; Mihaylov, M.; Hadjiivanov, K.; Mavrodinova, V. Catalytic properties and acidity of ZSM-12 zeolite with different textures. Microporous Mesoporous Mater. 2011, 143, 291–301. [Google Scholar] [CrossRef]
  35. Gabrienko, A.A.; Danilova, I.G.; Arzumanov, S.S.; Toktarev, A.V.; Freude, D.; Stepanov, A.G. Strong acidity of silanol groups of zeolite beta: Evidence from the studies by IR spectroscopy of adsorbed CO and 1H MAS NMR. Microporous Mesoporous Mater. 2010, 131, 210–216. [Google Scholar] [CrossRef]
  36. Morterra, C.; Bolis, V.; Magnacca, G. IR spectroscopic and microcalorimetric characterization of Lewis acid sites on (transition phase) Al2O3 using adsorbed CO. Langmuir 1994, 10, 1812–1824. [Google Scholar] [CrossRef]
  37. Zecchina, A.; Bordiga, S.; Spoto, G.; Scarano, D.; Petrini, G. Leofanti, G. Padovan, M. Otero Areàn, C. Low-temperature Fourier-transform Infrared Investigation of the Interaction of CO with Nanosized ZSM5 and Silicalite. J. Chem. Soc. Faraday Trans. 1992, 88, 2959–2969. [Google Scholar] [CrossRef]
  38. Gabrienko, A.A.; Danilova, I.G.; Arzumanov, S.S.; Freude, D.; Stepanov, A.G. Does the Zn2+ Species Introduced into H-ZSM-5 Zeolite Affect the Strength of Brønsted Acid Sites? Chemcatchem 2020, 12, 478–487. [Google Scholar] [CrossRef]
  39. Kwak, S.H. Spectroscopic Characterization of the Surface Hydroxyls of Zeolitic Catalysts. Master’s Thesis, University of Oslo, Oslo, Norway, 2014. [Google Scholar]
  40. Hadjiivanov, K. Identification and characterization of surface hydroxyl groups by infrared spectroscopy. In Advances in Catalysis; Elsevier: Amsterdam, The Netherlands, 2014; Volume 57, pp. 99–318. [Google Scholar]
  41. Chakarova, K.; Hadjiivanov, K. Problems in the IR measuring the acidity of zeolite bridging hydroxyls by low-temperature CO adsorption. Chem. Commun. 2011, 47, 1878–1880. [Google Scholar] [CrossRef]
  42. Höchtl, M.; Jentys, A.; Vinek, H. Acidity of SAPO and CoAPO molecular sieves and their activity in the hydroisomerization of n-heptane. Microporous Mesoporous Mater. 1999, 31, 271–285. [Google Scholar] [CrossRef]
  43. Nazimov, D.; Klimov, O.; Danilova, I.; Trukhan, S.; Saiko, A.; Cherepanova, S.; Chesalov, Y.A.; Martyanov, O.; Noskov, A. Effect of alumina polymorph on the dehydrogenation activity of supported chromia/alumina catalysts. J. Catal. 2020, 391, 35–47. [Google Scholar] [CrossRef]
  44. de Brimont, M.R.; Dupont, C.; Daudin, A.; Geantet, C.; Raybaud, P. Deoxygenation mechanisms on Ni-promoted MoS2 bulk catalysts: A combined experimental and theoretical study. J. Catal. 2012, 286, 153–164. [Google Scholar] [CrossRef]
  45. Vlasova, E.N.; Bukhtiyarova, G.A.; Deliy, I.V.; Aleksandrov, P.V.; Porsin, A.A.; Panafidin, M.A.; Gerasimov, E.Y.; Bukhtiyarov, V.I. The effect of rapeseed oil and carbon monoxide on SRGO hydrotreating over sulfide CoMo/Al2O3 and NiMo/Al2O3 catalysts. Catal. Today 2020, 357, 526–533. [Google Scholar] [CrossRef]
  46. Coumans, A.E.; Hensen, E.J.M. A model compound (methyl oleate, oleic acid, triolein) study of triglycerides hydrodeoxygenation over alumina-supported NiMo sulfide. Appl. Catal. B-Environ. 2017, 201, 290–301. [Google Scholar] [CrossRef]
  47. Yoosuk, B.; Sanggam, P.; Wiengket, S.; Prasassarakich, P. Hydrodeoxygenation of oleic acid and palmitic acid to hydrocarbon-like biofuel over unsupported Ni-Mo and Co-Mo sulfide catalysts. Renew. Energy 2019, 139, 1391–1399. [Google Scholar] [CrossRef]
  48. Zhang, M.; Chen, Y.J.; Wang, L.; Zhang, Q.M.; Tsang, C.W.; Liang, C.H. Shape Selectivity in Hydroisomerization of Hexadecane over Pt Supported on 10-Ring Zeolites: ZSM-22, ZSM-23, ZSM-35, and ZSM-48. Ind. Eng. Chem. Res. 2016, 55, 6069–6078. [Google Scholar] [CrossRef]
  49. Lanzafame, P.; Perathoner, S.; Centi, G.; Heracleous, E.; Iliopoulou, E.F.; Triantafyllidis, K.S.; Lappas, A.A. Effect of the Structure and Mesoporosity in Ni/Zeolite Catalysts for n-Hexadecane Hydroisomerisation and Hydrocracking. Chemcatchem 2017, 9, 1632–1640. [Google Scholar] [CrossRef]
  50. Valtchev, V.; Majano, G.; Mintova, S.; Pérez-Ramírez, J. Tailored crystalline microporous materials by post-synthesis modification. Chem. Soc. Rev. 2013, 42, 263–290. [Google Scholar] [CrossRef] [PubMed]
  51. Pérez-Ramírez, J.; Christensen, C.H.; Egeblad, K.; Christensen, C.H.; Groen, J.C. Hierarchical zeolites: Enhanced utilisation of microporous crystals in catalysis by advances in materials design. Chem. Soc. Rev. 2008, 37, 2530–2542. [Google Scholar] [CrossRef]
  52. Schwieger, W.; Machoke, A.G.; Weissenberger, T.; Inayat, A.; Selvam, T.; Klumpp, M.; Inayat, A. Hierarchy concepts: Classification and preparation strategies for zeolite containing materials with hierarchical porosity. Chem. Soc. Rev. 2016, 45, 3353–3376. [Google Scholar] [CrossRef]
  53. Roth, W.J.; Nachtigall, P.; Morris, R.E.; Cejka, J. Two-dimensional zeolites: Current status and perspectives. Chem. Rev. 2014, 114, 4807–4837. [Google Scholar] [CrossRef]
  54. Jang, H.-G.; Min, H.-K.; Lee, J.K.; Hong, S.B.; Seo, G. SAPO-34 and ZSM-5 nanocrystals’ size effects on their catalysis of methanol-to-olefin reactions. Appl. Catal. A Gen. 2012, 437, 120–130. [Google Scholar] [CrossRef]
  55. Available online: http://www.chemspider.com/Chemical-Structure.7889 (accessed on 4 September 2023).
  56. Available online: https://avogadro.cc/ (accessed on 4 September 2023).
  57. Cai, Z.; Wang, Y.; Cao, Y.; Yu, P.; Ding, Y.; Ma, Y.; Zheng, Y.; Huang, K.; Jiang, L. Direct production of isomerized biodiesel over MoS2/ZrPOx under solvent-free conditions. Fuel 2023, 337, 127175. [Google Scholar] [CrossRef]
  58. Available online: https://asia.iza-structure.org (accessed on 4 September 2023).
  59. Fyfe, C.; Gies, H.; Kokotailo, G.; Marler, B.; Cox, D. Crystal structure of silica-ZSM-12 by the combined use of hgh-resolution solid-state MAS NMR spectroscopy and synchrotron X-ray powder diffraction. J. Phys. Chem. 1990, 94, 3718–3721. [Google Scholar] [CrossRef]
  60. Borade, R.B.; Clearfield, A. A comparative study of acidic properties of SAPO-5, -11, -34 and -37 molecular sieves. J. Mol. Catal. 1994, 88, 249–265. [Google Scholar] [CrossRef]
  61. Dik, P.; Klimov, O.; Danilova, I.; Leonova, K.; Pereyma, V.Y.; Budukva, S.; Uvarkina, D.; Kazakov, M.; Noskov, A. Hydroprocessing of hydrocracker bottom on Pd containing bifunctional catalysts. Catal. Today 2016, 271, 154–162. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of synthesized Al2O3-zeolite composite supports ((a)—Al2O3-ZSM-5, (b)—Al2O3-ZSM-12, (c)—Al2O3-SAPO-11, and (d)—Al2O3-ZSM-22).
Figure 1. XRD patterns of synthesized Al2O3-zeolite composite supports ((a)—Al2O3-ZSM-5, (b)—Al2O3-ZSM-12, (c)—Al2O3-SAPO-11, and (d)—Al2O3-ZSM-22).
Ijms 24 14863 g001
Figure 2. SEM images of zeolite material (left) and EDX maps of composite supports of Al2O3-zeolite (right).
Figure 2. SEM images of zeolite material (left) and EDX maps of composite supports of Al2O3-zeolite (right).
Ijms 24 14863 g002aIjms 24 14863 g002b
Figure 3. Histograms of zeolite particle size distribution for supports Al2O3-zeolite (a) Al2O3-ZSM-5 (average particle size of 930 nm), (b) Al2O3-ZSM-12 (average particle size of 1010 nm), (c) Al2O3-SAPO-11 (average particle size of 300 nm), and (d) Al2O3-ZSM-22 (average particle size of 220 nm).
Figure 3. Histograms of zeolite particle size distribution for supports Al2O3-zeolite (a) Al2O3-ZSM-5 (average particle size of 930 nm), (b) Al2O3-ZSM-12 (average particle size of 1010 nm), (c) Al2O3-SAPO-11 (average particle size of 300 nm), and (d) Al2O3-ZSM-22 (average particle size of 220 nm).
Ijms 24 14863 g003
Figure 4. FTIR spectra of Al2O3 and Al2O3-zeolite supports after outgassing at 500 °C in the O-H stretching region.
Figure 4. FTIR spectra of Al2O3 and Al2O3-zeolite supports after outgassing at 500 °C in the O-H stretching region.
Ijms 24 14863 g004
Figure 5. FTIR spectra of CO absorbed at liquid nitrogen temperature on Al2O3-zeolite supports: Al2O3-ZSM-5 (A), Al2O3-ZSM-22 (B), Al2O3-ZSM-12 (C), and Al2O3-SAPO-11 (D). Equilibrium CO pressures used ranged from 0.1 (bottom curve) to 5 mbar (top curve). The inset shows enlarged spectra at low CO pressure. All spectra are background-corrected.
Figure 5. FTIR spectra of CO absorbed at liquid nitrogen temperature on Al2O3-zeolite supports: Al2O3-ZSM-5 (A), Al2O3-ZSM-22 (B), Al2O3-ZSM-12 (C), and Al2O3-SAPO-11 (D). Equilibrium CO pressures used ranged from 0.1 (bottom curve) to 5 mbar (top curve). The inset shows enlarged spectra at low CO pressure. All spectra are background-corrected.
Ijms 24 14863 g005
Figure 6. FTIR difference spectra of the OH stretching region during adsorption of CO at liquid nitrogen temperature and low equilibrium CO pressure of 0.3, 0.5, and 1 mbar on Al2O3-zeolite supports: (a) Al2O3-ZSM-22 and Al2O3-ZSM-5; (b) Al2O3-ZSM-12 and Al2O3-SAPO-11.
Figure 6. FTIR difference spectra of the OH stretching region during adsorption of CO at liquid nitrogen temperature and low equilibrium CO pressure of 0.3, 0.5, and 1 mbar on Al2O3-zeolite supports: (a) Al2O3-ZSM-22 and Al2O3-ZSM-5; (b) Al2O3-ZSM-12 and Al2O3-SAPO-11.
Ijms 24 14863 g006
Figure 7. HRTEM images of MoS2/Al2O3-SAPO-11 (a) and MoS2/Al2O3-ZSM-22 (b) catalysts (circle—sulfide nanoparticles on the alumina surface).
Figure 7. HRTEM images of MoS2/Al2O3-SAPO-11 (a) and MoS2/Al2O3-ZSM-22 (b) catalysts (circle—sulfide nanoparticles on the alumina surface).
Ijms 24 14863 g007
Figure 8. HRTEM-EDX mapping of MoS2/Al2O3-SAPO-11(a) and MoS2/Al2O3-ZSM-22 (b) catalysts.
Figure 8. HRTEM-EDX mapping of MoS2/Al2O3-SAPO-11(a) and MoS2/Al2O3-ZSM-22 (b) catalysts.
Ijms 24 14863 g008
Figure 9. MP conversion in dependence on temperature over MoS2/Al2O3-zeolite catalysts (reaction conditions: 250–350 °C, 3.0 MPa, 600 Nm3/m3, 36 h−1).
Figure 9. MP conversion in dependence on temperature over MoS2/Al2O3-zeolite catalysts (reaction conditions: 250–350 °C, 3.0 MPa, 600 Nm3/m3, 36 h−1).
Ijms 24 14863 g009
Figure 10. The conversion of oxygen-containing compounds in dependence on temperature over MoS2/Al2O3-zeolite catalysts (reaction conditions: 250–350 °C, 3.0 MPa, 600 Nm3/m3, 36 h−1).
Figure 10. The conversion of oxygen-containing compounds in dependence on temperature over MoS2/Al2O3-zeolite catalysts (reaction conditions: 250–350 °C, 3.0 MPa, 600 Nm3/m3, 36 h−1).
Ijms 24 14863 g010
Figure 11. Temperature effect on HDO selectivity over MoS2/Al2O3-zeolite catalysts (reaction conditions: 310–350 °C, 3.0 MPa, 600 Nm3/m3, 36 h−1).
Figure 11. Temperature effect on HDO selectivity over MoS2/Al2O3-zeolite catalysts (reaction conditions: 310–350 °C, 3.0 MPa, 600 Nm3/m3, 36 h−1).
Ijms 24 14863 g011
Figure 12. Temperature effect on yield of isomeric C16H34 and C15H32 alkanes over MoS2/Al2O3-SAPO-11 and MoS2/Al2O3-ZSM-22 catalysts (reaction conditions: 310–350 °C, 3.0 MPa, 600 Nm3/m3, 36 h−1).
Figure 12. Temperature effect on yield of isomeric C16H34 and C15H32 alkanes over MoS2/Al2O3-SAPO-11 and MoS2/Al2O3-ZSM-22 catalysts (reaction conditions: 310–350 °C, 3.0 MPa, 600 Nm3/m3, 36 h−1).
Ijms 24 14863 g012
Figure 13. Dimensions of methyl palmitate. Gray, white, and red balls are carbon, hydrogen, and oxygen atoms, respectively [55,56].
Figure 13. Dimensions of methyl palmitate. Gray, white, and red balls are carbon, hydrogen, and oxygen atoms, respectively [55,56].
Ijms 24 14863 g013
Figure 14. Pressure effect on yield of isomeric C16H34 and C15H32 alkanes over MoS2/Al2O3-SAPO-11 and MoS2/Al2O3-ZSM-22 catalysts (reaction conditions: 350 °C, 3.0 and 5.0 MPa, 600 Nm3/m3, 36 h−1).
Figure 14. Pressure effect on yield of isomeric C16H34 and C15H32 alkanes over MoS2/Al2O3-SAPO-11 and MoS2/Al2O3-ZSM-22 catalysts (reaction conditions: 350 °C, 3.0 and 5.0 MPa, 600 Nm3/m3, 36 h−1).
Ijms 24 14863 g014
Table 1. The properties of the prepared catalysts.
Table 1. The properties of the prepared catalysts.
CatalystsMo, wt.%SupportTextural Properties of the Support
Surface Area, m2/gPore Volume, cm3/gPore Diameter, nm
Mo/Al2O36.95Al2O31330.6625.1
Mo/Al2O3-ZSM-56.90Al2O3-ZSM-52020.4825.6
Mo/Al2O3-ZSM-126.96Al2O3-ZSM-121650.4922.8
Mo/Al2O3-ZSM-226.90Al2O3-ZSM-221750.5325.5
Mo/Al2O3-SAPO-116.97Al2O3-SAPO-111770.4222.6
Table 2. Type, acid strength, and concentration of Brønsted acid sites of Al2O3-zeolite composite supports.
Table 2. Type, acid strength, and concentration of Brønsted acid sites of Al2O3-zeolite composite supports.
Al2O3-Zeolite CompositesType of Zeolite SitesIR Frequency Shift/cm−1BAS Concentration
(μmol g−1)
ΔνOH…CO aΔνCO b
Al2O3-ZSM-5Framework
Si-O(H)-Al groups
–317 8.1
Extra-framework
Si-O(H)…Al3+ groups
–(260 ÷ 270)+331.8
Al2O3-ZSM-22Framework
Si-O(H)-Al groups
–320 11.4
Extra-framework
Si-O(H)…Al3+ groups
–(260 ÷ 270)+331.0
Al2O3-ZSM-12Framework
Si-O(H)-Al groups
–333+352.7
Extra-framework
Al-OH groups
–(200÷196)+282.6
Al2O3-SAPO-11Framework
Si-O(H)-Al groups
–258+307.4
P-OH groups–(202÷198)+27~4 c
a Red frequency shift of the bands of O-H groups at hydrogen bonding with CO. b Blue frequency shift of the CO stretching bands at hydrogen bonding of CO with OH- groups relative to the gas phase CO. c The contribution of the P-OH groups for the SAPO-11 is approximate (molar integral absorption coefficients for the P-OH group are unknown).
Table 3. List of zeolites relevant to this work with details.
Table 3. List of zeolites relevant to this work with details.
MaterialSiO2/Al2O3
Mole Ratio
Framework TypeChannelsSize of Channels
ZSM-5280MFI3D, 10 MR5.3 × 5.6 Å [010]
5.1 × 5.5 Å [100] [58]
ZSM-12280MTW1D, 12 MR5.6 × 7.7 Å [010] [59]
ZSM-2297TON1D, 10 MR4.6 × 5.7 Å [001] [58]
SAPO-11SiO2/Al2O3/P2O5 =
0.25/1.0/0.8
AEL1D, 10 MR3.9 × 6.3 Å [001] [60]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Vlasova, E.; Zhao, Y.; Danilova, I.; Aleksandrov, P.; Shamanaev, I.; Nuzhdin, A.; Suprun, E.; Pakharukova, V.; Tsaplin, D.; Maksimov, A.; et al. Bifunctional MoS2/Al2O3-Zeolite Catalysts in the Hydroprocessing of Methyl Palmitate. Int. J. Mol. Sci. 2023, 24, 14863. https://doi.org/10.3390/ijms241914863

AMA Style

Vlasova E, Zhao Y, Danilova I, Aleksandrov P, Shamanaev I, Nuzhdin A, Suprun E, Pakharukova V, Tsaplin D, Maksimov A, et al. Bifunctional MoS2/Al2O3-Zeolite Catalysts in the Hydroprocessing of Methyl Palmitate. International Journal of Molecular Sciences. 2023; 24(19):14863. https://doi.org/10.3390/ijms241914863

Chicago/Turabian Style

Vlasova, Evgeniya, Yiheng Zhao, Irina Danilova, Pavel Aleksandrov, Ivan Shamanaev, Alexey Nuzhdin, Evgeniy Suprun, Vera Pakharukova, Dmitriy Tsaplin, Anton Maksimov, and et al. 2023. "Bifunctional MoS2/Al2O3-Zeolite Catalysts in the Hydroprocessing of Methyl Palmitate" International Journal of Molecular Sciences 24, no. 19: 14863. https://doi.org/10.3390/ijms241914863

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