Next Article in Journal
Review of Experimental Setups for Plasmonic Photocatalytic Reactions
Next Article in Special Issue
Catalytic Pyrolysis of Aliphatic Carboxylic Acids into Symmetric Ketones over Ceria-Based Catalysts: Kinetics, Isotope Effect and Mechanism
Previous Article in Journal
Sequestration and Oxidation of Cr(III) by Fungal Mn Oxides with Mn(II) Oxidizing Activity
Previous Article in Special Issue
Catalytic Decomposition of Oleic Acid to Fuels and Chemicals: Roles of Catalyst Acidity and Basicity on Product Distribution and Reaction Pathways
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhancement of HDO Activity of MoP/SiO2 Catalyst in Physical Mixture with Alumina or Zeolites

by
Ivan V. Shamanaev
*,
Irina V. Deliy
,
Evgeny Yu. Gerasimov
,
Vera P. Pakharukova
and
Galina A. Bukhtiyarova
Boreskov Institute of Catalysis, Lavrentiev Ave. 5, 630090 Novosibirsk, Russia
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(1), 45; https://doi.org/10.3390/catal10010045
Submission received: 21 November 2019 / Revised: 16 December 2019 / Accepted: 24 December 2019 / Published: 31 December 2019
(This article belongs to the Special Issue Catalysis for the Production of Sustainable Fuels and Chemicals)

Abstract

:
Catalytic properties of physical mixture of MoP/SiO2 catalyst with SiC, γ-Al2O3, SAPO-11 and zeolite β have been compared in hydrodeoxygenation of methyl palmitate (MP). MoP/SiO2 catalyst (11.5 wt% of Mo, Mo/P = 1) was synthesized using TPR method and characterized with N2 physisorption, elemental analysis, H2-TPR, XRD and TEM. Trickle-bed reactor was used for catalytic properties investigation at hydrogen pressure of 3 MPa, and 290 °C. The conversions of MP and overall oxygen-containing compounds have been increased significantly (from 59 to about 100%) when γ-Al2O3 or zeolite materials were used instead of inert SiC. MP can be converted to palmitic acid through acid-catalyzed hydrolysis along with metal-catalyzed hydrogenolysis, and as a consequence the addition of material possessing acid sites to MoP/SiO2 catalyst could lead to acceleration of MP hydrodeoxygenation through acid-catalyzed reactions. Isomerization and cracking of alkane were observed over the physical mixture of MoP/SiO2 with zeolites, but the selectivity of MP conversion trough the HDO reaction route is remained on the high level exceeding 90%.

Graphical Abstract

1. Introduction

In the last decade, the research efforts in the transportation fuels production from renewable sources have grown due to environmental issues and limitation of oil reserves [1,2,3,4]. One of the versatile and already commercial routes to produce diesel and aviation fuels is hydrodeoxygenation of triglyceride-containing feedstock, including animal fats, waste frying oils, non-edible vegetable oils and so on [5,6]. Oxygen removal from triglyceride molecules includes several alternative routes [7,8,9]: hydrodeoxygenation (HDO) or decarboxylation/decarbonylation (HDeCOx). The HDO pathway results in alkanes with the same numbers of carbon atoms in the chain and molecules of water, while HDeCOx reactions give COx molecules and alkanes with the reduced carbon chains [7]. The HDO reaction route is more attractive in terms of atom economy, greenhouse gases production and hydrogen recycle [10,11]. Thus, the current research issue is the design of the catalysts with high activity and high selectivity towards HDO pathway.
Sulfided Ni(Co)Mo/Al2O3 catalysts are widely used and studied in triglycerides and esters hydrodeoxygenation [5,6,7,12,13,14,15], but the necessity of the sulfiding agent feeding to save the catalyst in sulfide state stimulate the development of new catalytic systems containing supported metal Ni [16,17,18,19,20,21], base metal carbides [22,23,24], nitrides [25] or phosphides [26,27,28,29,30,31,32,33,34,35]. Among them, transition metal phosphide catalysts demonstrate promising catalytic properties in hydrodeoxygenation of fatty acids or alkyl esters [26,28,30,32,36], vegetable oils [33], co-processing of renewable oils with petroleum products [37]. The activity of silica-supported catalysts in the HDO of methyl laurate is decreased in the following order [26]: Ni2P > MoP > CoP–Co2P > WP > Fe2P–FeP.
Nevertheless, the high selectivity of MoP-based catalysts in the conversion of aliphatic oxygenates through the HDO route [26] makes MoP/SiO2 the system of choice from the consideration of carbon atom economy and effluent gases composition, because carbon oxides make difficult purification of circulating hydrogen.
It was reported that acid sites can affect appreciably the HDO of aliphatic esters [26,29,38,39,40,41,42,43,44,45,46]. The authors of [26,29,39,40] have proposed, that Brønsted acidic sites of silica-supported nickel phosphide catalysts participate in the reaction of carboxylic ester to carboxylic acid. Also, the rate of ethyl stearate HDO over Ru/TiO2 catalyst was increased when the catalyst was mixed with the grains of γ-Al2O3 [41]. The synergetic effect of Ni2P/SiO2 and γ-Al2O3 physical mixture was observed in methyl palmitate HDO resulting in the enhancement of methyl palmitate conversion over Ni2P/SiO2 catalyst after mixing it with alumina grains [42]. It was observed that Ni2P catalyst supported on γ-Al2O3 exhibits higher conversion of methyl palmitate HDO in comparison with silica-supported one, though both catalysts contain Ni2P nanoparticles with close mean particle size and nearly the same nickel content [43]. The authors [42,43] proposed that the acid sites of the alumina increase the reaction rate of methyl palmitate hydrolysis, that enhance the methyl palmitate conversion. The assumption was made earlier [44,45], that the Lewis acid sites of the support enhance the rate of hydrolysis of aliphatic esters over alumina-supported sulfide catalysts.
The above examples of activity enhancement in the presence of acidic material allowed us to suggest that the replacement of silica support with the alumina may increase the catalytic activity of MoP catalyst in methyl palmitate HDO without reducing selectivity. The synthesis of supported MoP/γ-Al2O3 catalysts is sophisticated task due to interaction between alumina surface and phosphate groups that makes the reduction difficult and prevents the formation of MoP nanoparticles on alumina support [47,48,49]. Up to now, the silica-supported MoP catalysts have been studied in the HDO reactions, due to relatively easy reduction of phosphate groups that makes possible the formation of highly dispersed MoP phase on silica support [26,50].
The goal of the current study is the investigation of the catalytic properties of mechanical mixtures of MoP/SiO2 catalyst and materials with different acidity (SiC, γ-Al2O3, SAPO-11 and zeolite β) in methyl palmitate hydrodeoxygenation as the representative model component of triglyceride-based feedstock.

2. Results and Discussion

2.1. Catalyst Characterization

Physicochemical properties of MoP/SiO2, prepared using phosphate and phosphite precursors by TPR are shown in Table 1. MoP_A was prepared using ammonium paramolybdate and ammonium phosphate, MoP_I was prepared using ammonium paramolybdate and phosphorous acid. MoP is a bulk sample, prepared for comparison and XRD analysis. Initial Mo/P molar ratio in impregnating solution was 1. This ratio is usually used to prepare MoP catalysts [26,51] and the samples with this ratio were shown to be more active in hydrotreating reactions [52]. After reduction Mo/P ratio is slightly decreased to 0.91 for MoP_A and MoP and to 0.83 for MoP_I (Table 1). This can be due to Mo losses at impregnation step. Mo/P ratio determined from EDX data differ slightly from the chemical analysis results, but still remains close to Mo/P = 1.
Initial SiO2 support has SBET = 300 m2/g and Dpore = 10.6 nm. MoP_A and MoP_I have decreased value of SBET and Dpore in comparison with the silica support. Both active component and unreduced phosphate groups can block pores and decrease SBET. Bulk MoP has quite high area due to presence of large quantities of citric acid in the precursor, which are transformed to carbon after reduction.
The precursors and POx/SiO2 samples were investigated by H2-TPR (Figure 1). There are two peaks for the bulk MoP precursor: at 490 and at 720 °C. The first corresponds to Mo+6 → Mo+4 reduction, the second corresponds to Mo+4 → Moδ+ and P+5 → Pδ− phosphide formation. Reduction of phosphates in POx/SiO2 sample occurs at higher temperatures (> 750 °C, Figure 1). H2-TPR of MoP_A and MoP_I precursors have the maximums at the temperatures similar to the reduction temperature of bulk MoP. H2-TPR shows that precursor nature does not influence the temperatures of reduction of silica-supported MoP precursors. It was shown that in the case of Ni-based phosphide catalysts during H2-TPR of phosphite precursors the additional amount of hydrogen is produced, which is detected in H2-TPR curves [40,43]. For the Mo-phosphide precursors there is only slight difference between H2-TPR curves of MoP_A and MoP_I precursors (Figure 1). This can be attributed to oxidation of HPO32− by paramolybdate after impregnation:
2Mo7O246− + 7HPO32− + H2O → 14MoO3 + 7PO43− + 9H+
Mo7O246− + 7HPO32− + 4H2O → 7MoO32− + 7PO43− + 15H+
These reactions are visible, because presence of Mo(IV/V) species gives the characteristic blue color of the catalysts. Moreover, the reduction peak of MoO3 in H2-TPR of MoP_I sample is almost invisible (Figure 1), which corresponds to the presence of Mo, preferably in Mo (IV/V) states.
The calculated consumption of H2 during MoP_A reduction is 1.1 mmol/g, while the MoP_I consumes 0.80 mmol/g. This difference can be explained by the different states of Mo species.
XRD analysis of unreduced bulk MoP precursor did not show presence of any crystal phases (Figure 2), only increased intensity at low 2θ angles, which corresponds to amorphous carbon residues. XRD pattern of reduced bulk MoP sample include well defied maximums, corresponding to MoP phase (PDF № 24-0771). This pattern also has some amorphous phase corresponding to an increased base line in a 2θ range of 35–40°. DXRD of MoP particles is 12 nm (Table 1). In supported MoP_A and MoP_I samples, there were broad signals of SiO2 at ~20°; low-intensive broadening is also observed in the range of 35–45° that can be due to highly dispersed MoP (<3 nm) nanoparticles.
Figure 3 shows images of TEM analysis of MoP_A, MoP_I and bulk MoP samples. Size distributions of MoP particles are shown for MoP_A and MoP_I samples. Supported MoP_A and MoP_I samples have quite small particles (1.4 and 2.0 nm correspondingly, Table 1) and narrow particle size distributions. MoP sample has particles of 10–50 nm; therefore, there is no particle size distribution on Figure 3 for this sample.
In preliminary MP HDO experiments (LHSV = 60 h−1, T = 290 °C, PH2 = 3 MPa, H2/feed = 600 cm3/cm3) 59 and 60% MP conversion were obtained over MoP_A and MoP_I, respectively. MoP_A sample was used for catalytic experiments with different diluters.
The diluters used for catalytic experiments in this work were SiC, γ-Al2O3, SAPO-11 and zeolite β (Table 2). XRD analysis proved corresponding crystal structures for these materials (Figure S1 in Supporting Information).
Unlike SiC, an inert material with extremely low SBET (1 m2/g), γ-Al2O3, SAPO-11 and zeolite β have developed textural structures (Table 2). The acidity of these materials, in accordance with NH3-TPD data, is decreased in the following order: zeolite β > SAPO-11 > γ-Al2O3 (Table 2).
Zeolite β is characterized by Brønsted acid sites and three-dimensional system of channels, with sizes of 0.56 × 0.56 and 0.66 × 0.67 nm [53]. SAPO-11 has one-dimensional system of channels with size of 0.40 × 0.65 nm. SAPO-11 is a mesoporous structure, which has Brønsted acid sites of middle strength [54].

2.2. MP HDO Results

Catalytic properties of physical mixtures of MoP/SiO2 with SiC, γ-Al2O3, SAPO-11 or zeolite β were investigated in MP HDO at T = 290 °C, PH2 = 3 MPa, H2/feed = 600 cm3/cm3, LHSV = 60 or 10 h−1, using 0.4 or 1.4 g of catalyst and volume ratio of Vcat:Vdiluter = 1:10.5 or 1:1.5. Stable conversion as a function of time on stream was obtained for all tested systems (Figure S2 in Supporting Information). The results of comparative study of these systems in the methyl palmitate HDO are presented on Figure 4, which displays the values of MP conversion (XMP) and oxygen conversion (XO). MoP/SiO2 sample diluted with inert SiC grains displays the lowest XMP (59%) and XO (55%) values, whereas XMP and XO data for systems containing γ-Al2O3, zeolite β and SAPO-11 are higher than 90%. It is well known that HDO of fatty acid esters includes consecutive and parallel reactions: hydrogenolysis of C–O (HDO) and C–C (HDeCOx) bonds, hydrolysis of ester, hydrogenation of double bounds, dehydration of alcohols and esterification of acids, which occur over metal and acid active sites of catalysts, supports and diluters [26,42,55]. In particular, MP can be transformed at metal centers through hydrogenolysis to acid or aldehyde, or via acid-catalyzed hydrolysis to acid [56]. Thus, the higher conversions of MP over MoP/SiO2 catalyst in the presence of Al2O3, SAPO-11 or zeolite β can be explained by the higher rate of methyl palmitate hydrolysis, which is favored by the acid sites of these materials.
The main HDO product over MoP/SiO2–SiC and MoP/SiO2–γ-Al2O3 is hexadecane at 100% conversion of MP; the selectivity of MP conversion through the HDO pathway exceeds 96% (Figure 5). Thus, the use of γ-Al2O3 instead of SiC in the mixture with MoP/SiO2 increases the MP conversion, but has minor impact on the product distribution. The dilution of MoP/SiO2 with the same amount of SAPO-11 or zeolite β instead of SiC decreases the selectivity of C16 alkane formation in the course of MP HDO and increases the quantity of iso-C16, iso-C15 and C6–C14 hydrocarbons among the liquid products. Especially increased formation of hydrocarbons with ≤14 carbon atoms in the hydrocarbon chains is observed over MoP/SiO2–zeolite β catalytic system. Apparently, Brønsted acid sites in MoP/SiO2–SAPO-11 and MoP/SiO2–zeolite β result in partial cracking and structural isomerization of linear C15 and C16 hydrocarbons, which are formed in the HDO reaction of the ester. In these cases, the secondary reactions of the alkanes that are formed in MP HDO make it unattainable to calculate selectivities of HDO and HDeCOx pathways using the results of liquid phase analysis. To estimate the effect of SAPO-11 and zeolite β on the selectivity of MP conversion the amounts of carbon oxides were compared for MoP/SiO2–SiC, MoP/SiO2–SAPO-11 and MoP/SiO2–zeolite β (Figure 6). The results obtained let us see that the use of MoP/SiO2–SAPO-11 and MoP/SiO2–zeolite β induces only negligible increase in the amounts of carbon oxides in the gas phase in comparison with MoP/SiO2–SiC system. Appreciable amounts of light hydrocarbons with the carbon number ≤5 were detected in the gas-phase products of MP HDO over MoP/SiO2–zeolite β that is the evidence of alkane overcracking in the presence of zeolite β.
The above results show that the use of the MoP/SiO2 catalyst in the mixtures with acidic materials, like γ-Al2O3, SAPO-11 and zeolite β improves the catalytic activity of MoP catalyst in the hydrodeoxygenation of MP without reducing selectivity through HDO pathway. MoP/SiO2–γ-Al2O3 system gives hexadecane with the selectivity up to 96%, while MoP/SiO2–SAPO-11 produces iso-alkane and some amount of C6–C14 hydrocarbons. Therefore, MoP catalysts on SAPO-11 are perspective in one-pot HDO/hydroisomerization of aliphatic esters.

3. Materials and Methods

3.1. Catalysts Preparation

Commercial silica “KSKG” (SBET = 300 m2/g, Vpore = 0.80 cm3/g, Dpore = 10.6 nm) was obtained from ChromAnalit (Moscow, Russia), it was used as support material. Commercial SiC (Chelyabinsk Plant of Abrasive Materials, Chelyabinsk, Russia), alumina “IKGO-1” (γ-Al2O3, Promkataliz, Ryazan, Russia), zeolite β in H form (Angarsk catalyst and organic synthesis plant, Angarsk, Russia), SAPO-11 (Zeolyst International, Conshohocken, PA, USA) were utilized as diluters. Before catalysts preparation and experiments silica and alumina were dried (at 110 °C, 7 h) and calcined (at 500 °C, 4 h). Afterwards the materials were crushed using mortar and pestle, and sieved to obtain granules with size of 0.25–0.50 mm.
Preparation method of MoP/SiO2 samples
Incipient wetness impregnation method was used to prepare precursors of MoP/SiO2 samples. Mo/P molar ratio in impregnation water solution was taken equal 1 (Mo content~11.5 wt%). The label MoP_A was chosen for the MoP/SiO2 sample, which was prepared from phosphate solution, the label MoP_I—for the MoP/SiO2 sample, which was prepared from phosphite solution.
Phosphate method (MoP_A). To synthesis samples using this technique 2.357 g of (NH4)2HPO4 (Alfa Aesar, Kandel, Germany) and 3.151 g of (NH4)6Mo7O24⋅4H2O (Reakhim, Moscow, Russia) were dissolved under stirring in distilled water. Subsequently the support (8.5 g SiO2) was impregnated by this solution. After drying on air overnight (at room temperature) the precursors were dried for 4 h at 110 °C. Calcination of the dried samples was carried out at 500 °C for 4 h, and reduction of the calcined samples (TPR) was performed in H2 flow (100 mL/(min·g)). TPR has the following program: heating to 370 °C (with the rate of 3 °C/min), heating to 650 °C (with the rate of 1 °C/min), keeping the samples at 650 °C (for 1 h). Passivation of the reduced samples (to prevent oxidation on air atmosphere) was carried out at room temperature in 1 vol% O2/He (at 40 mL/(min·g) for 2 h).
Phosphite method (MoP_I). First, silica support (8.5 g) was impregnated with aqueous solution containing 3.151 g of (NH4)6Mo7O24⋅4H2O and 1.464 g of H3PO3. Then, after drying overnight and subsequently at 80 °C (for 24 h), the samples were reduced (TPR) in H2 flow (100 mL/(min·g)). TPR was carried as follows: heating to 650 °C (with the rate of 1 °C/min), keeping at 650 (for 1 h). The same passivation technique was carried out for MoP_I samples as in case of MoP_A samples.
POx/SiO2 sample. SiO2 (5 g) was impregnated by the solution of 2.357 g of (NH4)2HPO4 in distilled water. Drying was performed at room temperature overnight and at 110 °C (for 4 h). After calcination at 500 °C (for 4 h), the precursor was reduced and passivated according the same TPR and passivation programs as for MoP_A and MoP_I precursors.
Bulk MoP reference sample preparation method. The bulk molybdenum phosphide sample (bulk MoP) was prepared as follows: to the water solution of (NH4)2HPO4 was added water solution of (NH4)6Mo7O24. The Mo/P initial molar ratio was equal to 1. After evaporation of the water, the obtained sample was dried at 87 °C (for 48 h) and then at 124 °C (for 24 h). Calcination of the sample was performed at 500 °C (for 5 h). Afterwards, TPR was carried out (at H2 flow of 150 mL/(min·g)): heating to 650 °C (with the rate of 1 °C/min), keeping at 650 °C (for 1 h). The obtained bulk sample was passivated in flow of 1 vol.% O2/He (80 mL/(min·g), for 3 h).

3.2. Catalysts Characterization

Nitrogen physisorption was used to determine textural characteristics of the MoP, MoP/SiO2 and diluters. It was carried out at 77 K on ASAP 2400 instrument (USA).
Inductively coupled plasma atomic emission spectroscopy (ICP-AES) was used to perform elemental analysis of the MoP and MoP/SiO2 samples. ICP-AES was carried out on Optima 4300 DV (Perkin Elmer, Villebon-sur-Yvette, France).
Bruker D8 Advance X-ray diffractometer (Bruker, Bremen, Germany) was used to carry out XRD analysis. CuKα radiation (with λ = 1.5418 Å) was used, and 2θ scanning range was from 10 to 70. To determine crystal phases JCPDS database was used [57]. Scherrer equation was used to calculate average crystallite size (DXRD).
Transmission electron microscope JEM-2010 (JEOL, Tokyo, Japan) was used to obtain high resolution transmission electron microscopy (TEM) images. The accelerating voltage and resolution were 200 kV and 0.14 nm correspondingly. EDX spectrometer was used to analyze elemental composition.
H2 temperature-programmed reduction (H2-TPR) was performed to determine reduction temperatures of the prepared samples. The sample mass was about 0.10 g, it was heated in 10 vol.% H2/Ar (60 mL/min, with rate of 10 °C/min) from 80 °C to 900 °C. Thermal conductivity detector (TCD) was used to determine H2 consumption.

3.3. Experimental Setup and Procedure

Methyl palmitate (hexadecanoic acid methyl ester, ≥97%), n-dodecane (≥99%) and n-octane (≥99.5%) were obtained from Sigma-Aldrich (St. Louis, MO, USA), Acros Organics (New Jersey, NJ, USA) and Kriokhrom (Saint Petersburg, Russia), correspondingly.
A trickle-bed reactor was used to evaluate catalytic activity of the prepared MoP catalysts in methyl palmitate HDO (Figure S3 in Supporting Information). The reaction conditions: T = 290 °C, PH2 = 3.0 MPa, volume ratio H2/feed = 600, LHSV = 60 or 10 h−1. The liquid feed consisted of: 10 wt% of methyl palmitate, 0.5 wt% of n-octane (internal standard), the rest is n-dodecane. The catalysts particles were diluted with SiC, Al2O3, SAPO-11 or zeolite β (0.2–0.3 mm) in a ratio 1:10.5 or 1:1.5, placed in the catalytic reactor between. Above and under the catalytic layer there were SiC layers. Before methyl palmitate HDO experiments reduction of the catalyst precursors was performed in situ by TPR technique described earlier in catalysts preparation section. Collecting of liquid product effluent was carried out hourly until the reaction reaches steady-state. For all experiments time on stream = 8 h (Figure S2 in Supporting Information).
XMP—methyl palmitate conversion was calculated as follows:
X M P = ( 1 C M P C M P o ) 100 %
coMP and cMP (mol/L)—methyl palmitate concentration in the initial feed and in the liquid product correspondingly.
XO—oxygen-containing compounds conversion was calculated as follows:
X O C C = ( 1 C O C O o ) 100 %
coO and cO (mol/L)—oxygen concentration in the initial feed and in the liquid product correspondingly.
Si—selectivity was calculated as follows:
S i = ( c i c M P o c M P ) 100 %
ci (mol/L)—concentration of the ith compound in the liquid product.

3.4. Analysis of the Reaction Products

7000 GC/MS Triple QQQ GC System 7890A (Agilent Technologies, Santa Clara, CA, USA) was used to identify HDO products. The GC was equipped with VF-5MS column (length 30 m, inner diameter 0.25 mm, film thickness 0.25 µm). Quantitative analysis was performed using GC with flame ionization detector (FID, Agilent 6890N, USA) and an HP-1MS column (length 30 m, inner diameter 0.32 mm, film thickness 1.0 µm). Carbon balance for all experiments was not lower than 97%.
Elemental analyzer Vario EL Cube (Elementar Analysensysteme GmbH, Langenselbold, Germany) was used to determine oxygen content in the liquid reaction mixture.
The concentration of gas-phase products was determined using GC Khromos-1000 with FID (Khromos, Dzerzhinsk, Russia). CO and CO2 were analyzed after methanation over Pd catalyst.

4. Conclusions

In this work catalytic properties of MoP/SiO2 catalyst in MP HDO were studied in mechanical mixture with materials having different acidity: SiC, γ-Al2O3, SAPO-11 and zeolite β.
A method was chosen for the preparation of SiO2-supported MoP catalyst. Two samples were prepared using (NH4)6Mo7O24 and different P precursors: (NH4)2HPO4 and H3PO3 (MoP_A and MoP_I samples, correspondingly). It was shown that H2-TPR curves for MoP_A and MoP_I samples are coinciding regardless of P precursor that can be explained by the oxidation of H3PO3 in the reaction with ammonium paramolybdate during impregnation and drying of support. The silica-supported MoP catalysts samples were synthesized by reduction (TPR) of the precursors at 650 °C, and according to TEM data, the particle sizes of MoP nanoparticles were 1.4 and 2.0 nm for MoP_A and MoP_I samples, respectively. Both catalysts (MoP_A and MoP_I) display similar activity in the MP hydrodeoxygenation; MoP_A sample (designated as MoP/SiO2) was chosen for the experiments with different diluters.
The comparative study of MoP/SiO2–SiC, MoP/SiO2–γ-Al2O3, MoP/SiO2–SAPO-11 and MoP/SiO2–zeolite β systems has demonstrated the enhancement of MoP/SiO2 catalyst activity in MP HDO in mechanical mixtures with alumina, SAPO-11 and zeolite β. The higher conversions of MP over MoP/SiO2 catalyst in the presence of Al2O3, SAPO-11 and zeolite β can be explained by the higher rate of methyl palmitate hydrolysis, which is favored by the acid sites of these materials. The use of γ-Al2O3, SAPO-11 and zeolite β instead of SiC in the mixture with MoP/SiO2 increases the MP conversion, but has minor impact on the HDO selectivity. It was shown that selectivities of methyl palmitate conversion through the HDO pathways are nearly the same for the different systems, because the amounts of CO and CO2 in the gas-phase products differ only slightly. The main HDO product observed in the liquid products of MP HDO over MoP/SiO2–SiC and MoP/SiO2–γ-Al2O3 is hexadecane; the selectivity of HDO pathway exceeds 96% at 100% of MP conversion. The use of MoP/SiO2 in the mixture with the SAPO-11 or zeolite β instead of SiC changes the product distribution due to partial cracking and structural isomerization of linear C15 and C16 hydrocarbons, which are formed in HDO reaction of ester. Thus, MoP/SiO2–SAPO-11 system produces iso-alkane and some amount of C6–C14 hydrocarbons, while increased formation of hydrocarbons with ≤ 14 carbon atoms in the hydrocarbon chains is observed over MoP/SiO2–zeolite β catalytic system.
The obtained results show, that the use of the MoP/SiO2 catalyst in the mixtures with γ-Al2O3, SAPO-11 and zeolite β enhances the catalytic activity of MoP catalyst in the hydrodeoxygenation of MP without reducing selectivity through HDO pathway. The product distribution in the liquid phase depends on the dilutor nature. Linear alkanes are produced in the presence of γ-Al2O3. SAPO-11 can be considered as the perspective material for the one-pot HDO/hydroisomerization of aliphatic esters.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/1/45/s1, Figure S1: XRD patterns of materials used as catalyst diluters, Figure S2: MP conversion vs time on stream for MoP/SiO2 catalysts diluted with different materials, Figure S3: The scheme of the experimental setup.

Author Contributions

Formal analysis, I.V.S. and I.V.D.; TEM analysis, E.Y.G.; XRD analysis, V.P.P.; writing and the draft preparation, I.V.D., I.V.S. and G.A.B.; reviewing and editing, I.V.D., G.A.B. and I.V.S.; supervision, G.A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by Ministry of Science and Higher Education of the Russian Federation (project AAAA-A17-117041710075-0). The authors of the paper are grateful to M.V. Shashkov and V.A. Utkin for analysis and identification of HDO reaction products by GC/MS.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Choudhary, T.V.; Phillips, C.B. Renewable fuels via catalytic hydrodeoxygenation. Appl. Catal. A Gen. 2011, 397, 1–12. [Google Scholar] [CrossRef]
  2. Serrano-Ruiz, J.C.; Ramos-Fernández, E.V.; Sepúlveda-Escribano, A. From biodiesel and bioethanol to liquid hydrocarbonfuels: New hydrotreating and advanced microbial technologies. Energy Environ. Sci. 2012, 5, 5638–5652. [Google Scholar] [CrossRef]
  3. Huber, G.W.; Iborra, S.; Corma, A. Synthesis of transportation fuels from biomass: Chemistry, catalysts, and engineering. Chem. Rev. 2006, 106, 4044–4098. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Melero, J.A.; Iglesias, J.; Garcia, A. Biomass as renewable feedstock in standard refinery units. Feasibility, opportunities and challenges. Energy Environ. Sci. 2012, 5, 7393–7420. [Google Scholar] [CrossRef]
  5. Khan, S.; Kay Lup, A.N.; Qureshi, K.M.; Abnisa, F.; Wan Daud, W.M.A.; Patah, M.F.A. A review on deoxygenation of triglycerides for jet fuel range hydrocarbons. J. Anal. Appl. Pyrolysis 2019, 140, 1–24. [Google Scholar] [CrossRef]
  6. Why, E.S.K.; Ong, H.C.; Lee, H.V.; Gan, Y.Y.; Chen, W.-H.; Chong, C.T. Renewable aviation fuel by advanced hydroprocessing of biomass: Challenges and perspective. Energy Convers. Manag. 2019, 199, 112015. [Google Scholar] [CrossRef]
  7. Kubička, D.; Tukač, V. Chemical Engineering for Renewables Conversion. Advances in Chemical Engineering. In Advances in Chemical Engineering; Murzin, D.Y., Ed.; Elsevier: Amsterdam, The Netherlands, 2013; Volume 42, pp. 141–194. ISBN 9780123865052. [Google Scholar]
  8. Zharova, P.A.; Chistyakov, A.V.; Shapovalov, S.S.; Pasynskii, A.A.; Tsodikov, M.V. Original Pt-Sn/Al2O3 catalyst for selective hydrodeoxygenation of vegetable oils. Energy 2019, 172, 18–25. [Google Scholar] [CrossRef]
  9. Kukushkin, R.G.; Reshetnikov, S.I.; Zavarukhin, S.G.; Eletskii, P.M.; Yakovlev, V.A. Kinetics of the Hydrodeoxygenation of Ethyl Ester of Decanoic Acid over the Ni–Cu–Mo/ Al2O3 Catalyst. Catal. Ind. 2019, 11, 191–197. [Google Scholar] [CrossRef]
  10. Janampelli, S.; Darbha, S. Selective and reusable Pt-WOx/ Al2O3 catalyst for deoxygenation of fatty acids and their esters to diesel-range hydrocarbons. Catal. Today 2018, 309, 219–226. [Google Scholar] [CrossRef]
  11. Gosselink, R.W.; Stellwagen, D.R.; Bitter, J.H. Tungsten-Based Catalysts for Selective Deoxygenation. Angew. Chem. Int. Ed. 2013, 52, 5089–5092. [Google Scholar] [CrossRef]
  12. Wang, H.; Rogers, K.; Zhang, H.; Li, G.; Pu, J.; Zheng, H.; Lin, H.; Zheng, Y.; Ng, S. The Effects of Catalyst Support and Temperature on the Hydrotreating of Waste Cooking Oil (WCO) over CoMo Sulfided Catalysts. Catalysts 2019, 9, 689. [Google Scholar] [CrossRef] [Green Version]
  13. Zhao, X.; Wei, L.; Cheng, S.; Julson, J. Review of Heterogeneous Catalysts for Catalytically Upgrading Vegetable Oils into Hydrocarbon Biofuels. Catalysts 2017, 7, 83. [Google Scholar] [CrossRef]
  14. Pinheiro, A.; Dupassieux, N.; Hudebine, D.; Geantet, C. Impact of the Presence of Carbon Monoxide and Carbon Dioxide on Gas Oil Hydrotreatment: Investigation on Liquids from Biomass Cotreatment with Petroleum Cuts. Energy Fuels 2011, 25, 804–812. [Google Scholar] [CrossRef]
  15. Philippe, M.; Richard, F.; Hudebine, D.; Brunet, S. Transformation of dibenzothiophenes model molecules over CoMoP/ Al2O3 catalyst in the presence of oxygenated compounds. Appl. Catal. B Environ. 2013, 132, 493–498. [Google Scholar] [CrossRef]
  16. Liu, S.; Zhu, Q.; Guan, Q.; He, L.; Li, W. Bio-aviation fuel production from hydroprocessing castor oil promoted by the nickel-based bifunctional catalysts. Bioresour. Technol. 2015, 183, 93–100. [Google Scholar] [CrossRef]
  17. Li, X.; Luo, X. Preparation of mesoporous activated carbon supported Ni catalyst for deoxygenation of stearic acid into hydrocarbons. Environ. Prog. Sustain. Energy 2015, 34, 607–612. [Google Scholar] [CrossRef]
  18. Kukushkin, R.G.; Bulavchenko, O.A.; Kaichev, V.V.; Yakovlev, V.A. Influence of Mo on catalytic activity of Ni-based catalysts in hydrodeoxygenation of esters. Appl. Catal. B Environ. 2015, 163, 531–538. [Google Scholar] [CrossRef]
  19. Yakovlev, V.A.; Khromova, S.A.; Sherstyuk, O.V.; Dundich, V.O.; Ermakov, D.Y.; Novopashina, V.M.; Lebedev, M.Y.; Bulavchenko, O.; Parmon, V.N. Development of new catalytic systems for upgraded bio-fuels production from bio-crude-oil and biodiesel. Catal. Today 2009, 144, 362–366. [Google Scholar] [CrossRef]
  20. Zuo, H.; Liu, Q.; Wang, T.; Ma, L.; Zhang, Q.; Zhang, Q. Hydrodeoxygenation of Methyl Palmitate over Supported Ni Catalysts for Diesel-like Fuel Production. Energy Fuels 2012, 26, 3747–3755. [Google Scholar] [CrossRef]
  21. Ochoa-Hernández, C.; Yang, Y.; Pizarro, P.; de la Peña O’Shea, V.A.; Coronado, J.M.; Serrano, D.P. Hydrocarbons production through hydrotreating of methyl esters over Ni and Co supported on SBA-15 and Al-SBA-15. Catal. Today 2013, 210, 81–88. [Google Scholar] [CrossRef]
  22. Ki Kim, S.; Yoon, D.; Lee, S.-C.; Kim, J. Mo2C/Graphene Nanocomposite as a Hydrodeoxygenation Catalyst for the Production of Diesel Range Hydrocarbons. ACS Catal. 2015, 5, 3292–3303. [Google Scholar] [CrossRef]
  23. Hollak, S.A.W.; Gosselink, R.W.; van Es, D.S.; Bitter, J.H. Comparison of tungsten and molybdenum carbide catalysts for the hydrodeoxygenation of oleic acid. ACS Catal. 2013, 3, 2837–2844. [Google Scholar] [CrossRef]
  24. Wang, H.; Yan, S.; Salley, S.O.; Simon Ng, K.Y. Support effects on hydrotreating of soybean oil over NiMo carbide catalyst. Fuel 2013, 111, 81–87. [Google Scholar] [CrossRef]
  25. Monnier, J.; Sulimma, H.; Dalai, A.; Caravaggio, G. Hydrodeoxygenation of oleic acid and canola oil over alumina-supported metal nitrides. Appl. Catal. A Gen. 2010, 382, 176–180. [Google Scholar] [CrossRef]
  26. Chen, J.; Shi, H.; Li, L.; Li, K. Deoxygenation of methyl laurate as a model compound to hydrocarbons on transition metal phosphide catalysts. Appl. Catal. B Environ. 2014, 144, 870–884. [Google Scholar] [CrossRef]
  27. Chen, J.; Yang, Y.; Shi, H.; Li, M.; Chu, Y.; Pan, Z.; Yu, X. Regulating product distribution in deoxygenation of methyl laurate on silica-supported Ni–Mo phosphides: Effect of Ni/Mo ratio. Fuel 2014, 129, 1–10. [Google Scholar] [CrossRef]
  28. Shi, H.; Chen, J.; Yang, Y.; Tian, S. Catalytic deoxygenation of methyl laurate as a model compound to hydrocarbons on nickel phosphide catalysts: Remarkable support effect. Fuel Process. Technol. 2013, 116, 161–170. [Google Scholar] [CrossRef]
  29. Yang, Y.; Chen, J.; Shi, H. Deoxygenation of Methyl Laurate as a Model Compound to Hydrocarbons on Ni2P/SiO2, Ni2P/MCM-41, and Ni2P /SBA-15 catalysts with different dispersions. Energy Fuels 2013, 27, 3400–3409. [Google Scholar] [CrossRef]
  30. Yang, Y.; Ochoa-Hernández, C.; de la Peña O’Shea, V.A.; Coronado, J.M.; Serrano, D.P. Ni2P/SBA-15 As a Hydrodeoxygenation Catalyst with Enhanced Selectivity for the Conversion of Methyl Oleate Into n-Octadecane. ACS Catal. 2012, 2, 592–598. [Google Scholar] [CrossRef]
  31. Yang, Y.; Ochoa-Hernández, C.; Pizarro, P.; de la Peña O’Shea, V.A.; Coronado, J.M.; Serrano, D.P. Synthesis of Nickel Phosphide Nanorods as Catalyst for the Hydrotreating of Methyl Oleate. Top. Catal. 2012, 55, 991–998. [Google Scholar] [CrossRef]
  32. Yang, Y.; Ochoa-Hernández, C.; Pizarro, P.; de la Peña O’Shea, V.A.; Coronado, J.M.; Serrano, D.P. Influence of the Ni/P ratio and metal loading on the performance of NixPy/SBA-15 catalysts for the hydrodeoxygenation of methyl oleate. Fuel 2015, 144, 60–70. [Google Scholar] [CrossRef]
  33. Zarchin, R.; Rabaev, M.; Vidruk-Nehemya, R.; Landau, M.V.; Herskowitz, M. Hydroprocessing of soybean oil on nickel-phosphide supported catalysts. Fuel 2015, 139, 684–691. [Google Scholar] [CrossRef]
  34. Pan, Z.; Wang, R.; Li, M.; Chu, Y.; Chen, J. Deoxygenation of methyl laurate to hydrocarbons on silica-supported Ni-Mo phosphides: Effect of calcination temperatures of precursor. J. Energy Chem. 2015, 24, 77–86. [Google Scholar] [CrossRef]
  35. Xue, Y.; Guan, Q.; Li, W. Synthesis of bulk and supported nickel phosphide using microwave radiation for hydrodeoxygenation of methyl palmitate. RSC Adv. 2015, 5, 53623–53628. [Google Scholar] [CrossRef]
  36. Zheng, Z.; Li, M.-F.; Chu, Y.; Chen, J.-X. Influence of CS2 on performance of Ni2P/SiO2 for deoxygenation of methyl laurate as a model compound to hydrocarbons: Simultaneous investigation on catalyst deactivation. Fuel Process. Technol. 2015, 134, 259–269. [Google Scholar] [CrossRef]
  37. Lee, S.I.; Kim, D.W.; Jeon, H.J.; Ju, S.J.; Ryu, J.W.; Oh, S.H.; Koh, J.H. Metal Phosphorus Compound for Preparing Biodiesel and Method Preparing Biodiesel Using the Same. U.S. Patent US2014/0150332A1, 5 June 2014. [Google Scholar]
  38. Alvarez-Galvan, M.; Campos-Martin, J.; Fierro, J. Transition Metal Phosphides for the Catalytic Hydrodeoxygenation of Waste Oils into Green Diesel. Catalysts 2019, 9, 293. [Google Scholar] [CrossRef] [Green Version]
  39. Deliy, I.; Shamanaev, I.; Gerasimov, E.; Pakharukova, V.; Yakovlev, I.; Lapina, O.; Aleksandrov, P.; Bukhtiyarova, G. HDO of Methyl Palmitate over Silica-Supported Ni Phosphides: Insight into Ni/P Effect. Catalysts 2017, 7, 298. [Google Scholar] [CrossRef]
  40. Shamanaev, I.V.; Deliy, I.V.; Aleksandrov, P.V.; Gerasimov, E.Y.; Pakharukova, V.P.; Kodenev, E.G.; Ayupov, A.B.; Andreev, A.S.; Lapina, O.B.; Bukhtiyarova, G.A. Effect of precursor on the catalytic properties of Ni2P/SiO2 in methyl palmitate hydrodeoxygenation. RSC Adv. 2016, 6, 30372–30383. [Google Scholar] [CrossRef]
  41. He, L.; Wu, C.; Cheng, H.; Yu, Y.; Zhao, F. Highly selective and efficient catalytic conversion of ethyl stearate into liquid hydrocarbons over a Ru/TiO2 catalyst under mild conditions. Catal. Sci. Technol. 2012, 2, 1328. [Google Scholar] [CrossRef]
  42. Shamanaev, I.; Aleksandrov, P.; Kodenev, E.; Pakharukova, V.; Gerasimov, E.; Deliy, I.; Bukhtiyarova, G. Synergetic Effect of Ni2P/SiO2 and γ-Al2O3 Physical Mixture in Hydrodeoxygenation of Methyl Palmitate. Catalysts 2017, 7, 329. [Google Scholar] [CrossRef] [Green Version]
  43. Deliy, I.; Shamanaev, I.; Aleksandrov, P.; Gerasimov, E.; Pakharukova, V.; Kodenev, E.; Yakovlev, I.; Lapina, O.; Bukhtiyarova, G. Support Effect on the Performance of Ni2P Catalysts in the Hydrodeoxygenation of Methyl Palmitate. Catalysts 2018, 8, 515. [Google Scholar] [CrossRef] [Green Version]
  44. Laurent, E.; Delmon, B. Influence of water in the deactivation of a sulfided NiMo?-Al2O3 catalyst during hydrodeoxygenation. J. Catal. 1994, 146, 281–291. [Google Scholar] [CrossRef]
  45. Şenol, O.İ.; Viljava, T.-R.; Krause, A.O.I. Hydrodeoxygenation of methyl esters on sulphided NiMo/γ- Al2O3 and CoMo/γ-Al2O3 catalysts. Catal. Today 2005, 100, 331–335. [Google Scholar] [CrossRef]
  46. Shamanaev, I.V.; Deliy, I.V.; Aleksandrov, P.V.; Reshetnikov, S.I.; Bukhtiyarova, G.A. Methyl palmitate hydrodeoxygenation over silica-supported nickel phosphide catalysts in flow reactor: Experimental and kinetic study. J. Chem. Technol. Biotechnol. 2019. [Google Scholar] [CrossRef]
  47. Prins, R.; Bussell, M.E. Metal Phosphides: Preparation, Characterization and Catalytic Reactivity. Catal. Lett. 2012, 142, 1413–1436. [Google Scholar] [CrossRef]
  48. Liu, X.; Xu, L.; Zhang, B. Essential elucidation for preparation of supported nickel phosphide upon nickel phosphate precursor. J. Solid State Chem. 2014, 212, 13–22. [Google Scholar] [CrossRef]
  49. Peroni, M.; Mancino, G.; Baráth, E.; Gutiérrez, O.Y.; Lercher, J.A. Bulk and γ-Al2O3 supported Ni2P and MoP for hydrodeoxygenation of palmitic acid. Appl. Catal. B Environ. 2016, 180, 301–311. [Google Scholar] [CrossRef]
  50. Li, K.; Wang, R.; Chen, J. Hydrodeoxygenation of Anisole over Silica-Supported Ni2P, MoP, and NiMoP Catalysts. Energy Fuels 2011, 25, 854–863. [Google Scholar] [CrossRef]
  51. Bui, P.; Cecilia, J.A.; Oyama, S.T.; Takagaki, A.; Infantes-Molina, A.; Zhao, H.; Li, D.; Rodríguez-Castellón, E.; Jiménez López, A. Studies of the synthesis of transition metal phosphides and their activity in the hydrodeoxygenation of a biofuel model compound. J. Catal. 2012, 294, 184–198. [Google Scholar] [CrossRef]
  52. Montesinos-Castellanos, A.; Zepeda, T.A.; Pawelec, B.; Fierro, J.L.G.; De Los Reyes, J.A. Preparation, characterization, and performance of alumina-supported nanostructured Mo-phosphide systems. Chem. Mater. 2007, 19, 5627–5636. [Google Scholar] [CrossRef]
  53. Newsam, J.M.; Treacy, M.M.J.; Koetsier, W.T.; Gruyter, C.B.D. Structural Characterization of Zeolite Beta. Proc. R. Soc. A Math. Phys. Eng. Sci. 1988, 420, 375–405. [Google Scholar] [CrossRef]
  54. Flanigen, E.M.; Lok, B.M.; Patton, R.L.; Wilson, S.T. Aluminophosphate molecular sieves and the periodic table. Pure Appl. Chem. 1986, 58, 1351–1358. [Google Scholar] [CrossRef] [Green Version]
  55. Peroni, M.; Lee, I.; Huang, X.; Baráth, E.; Gutiérrez, O.Y.; Lercher, J.A. Deoxygenation of Palmitic Acid on Unsupported Transition-Metal Phosphides. ACS Catal. 2017, 7, 6331–6341. [Google Scholar] [CrossRef]
  56. Gosselink, R.W.; Hollak, S.A.W.; Chang, S.W.; Van Haveren, J.; De Jong, K.P.; Bitter, J.H.; van Es, D.S. Reaction pathways for the deoxygenation of vegetable oils and related model compounds. ChemSusChem 2013, 6, 1576–1594. [Google Scholar] [CrossRef] [PubMed]
  57. File, P.D. JCPDS—International Centre for Diffraction Data; ICDD: Swarthmore, PA, USA, 1997. [Google Scholar]
Figure 1. H2-TPR curves of POx/SiO2, supported MoP_A, MoP_I and bulk MoP oxide precursors.
Figure 1. H2-TPR curves of POx/SiO2, supported MoP_A, MoP_I and bulk MoP oxide precursors.
Catalysts 10 00045 g001
Figure 2. XRD of reduced MoP_A, MoP_I and bulk MoP samples and unreduced MoP precursor.
Figure 2. XRD of reduced MoP_A, MoP_I and bulk MoP samples and unreduced MoP precursor.
Catalysts 10 00045 g002
Figure 3. Images of TEM MoP_A (a), MoP_I (b), bulk MoP (c,d).
Figure 3. Images of TEM MoP_A (a), MoP_I (b), bulk MoP (c,d).
Catalysts 10 00045 g003
Figure 4. MP and O conversion in MP HDO over MoP/SiO2 catalyst diluted with different materials. Vcat = 0.4 cm3, LHSV = 60 h−1, T = 290 °C, PH2 = 3 MPa, H2/feed = 600 cm3/cm3.
Figure 4. MP and O conversion in MP HDO over MoP/SiO2 catalyst diluted with different materials. Vcat = 0.4 cm3, LHSV = 60 h−1, T = 290 °C, PH2 = 3 MPa, H2/feed = 600 cm3/cm3.
Catalysts 10 00045 g004
Figure 5. Selectivities of MP HDO products over MoP/SiO2 catalyst diluted with different materials (XMP = 100%). C16O—hexadecanal, C16OH—hexadecanol, PA—palmitic acid, PP—palmityl palmitate. Vcat = 1.8 cm3, LHSV = 10 h−1, T = 290 °C, PH2 = 3 MPa, H2/feed = 600 cm3/cm3.
Figure 5. Selectivities of MP HDO products over MoP/SiO2 catalyst diluted with different materials (XMP = 100%). C16O—hexadecanal, C16OH—hexadecanol, PA—palmitic acid, PP—palmityl palmitate. Vcat = 1.8 cm3, LHSV = 10 h−1, T = 290 °C, PH2 = 3 MPa, H2/feed = 600 cm3/cm3.
Catalysts 10 00045 g005
Figure 6. Concentration of gas-phase products of MP HDO over MoP/SiO2 catalyst diluted with different materials (XMP = 100%). Vcat = 1.8 cm3, LHSV = 10 h−1, T = 290 °C, PH2 = 3 MPa, H2/feed = 600 cm3/cm3.
Figure 6. Concentration of gas-phase products of MP HDO over MoP/SiO2 catalyst diluted with different materials (XMP = 100%). Vcat = 1.8 cm3, LHSV = 10 h−1, T = 290 °C, PH2 = 3 MPa, H2/feed = 600 cm3/cm3.
Catalysts 10 00045 g006
Table 1. Physicochemical properties of SiO2, MoP/SiO2 and MoP samples.
Table 1. Physicochemical properties of SiO2, MoP/SiO2 and MoP samples.
SampleMo, wt%P, wt%Mo/P Molar Ratio in Reduced CatalystMo/P
EDX
SBET, m2/gDpore, nmDTEM, nmXRD PhaseDXRD, nm
SiO230010.6
MoP_A11.54.00.911.0217810.41.4amorphous phase
MoP_I11.54.30.830.972357.92.0amorphous phase
MoP51.718.40.911.172112.0MoP12
Table 2. Physicochemical properties of the diluters.
Table 2. Physicochemical properties of the diluters.
DiluterSBET, m2/gVpore, cm3/gDpore, nmNH3-TPD, 10−3 mol/g
SiC1
γ-Al2O32350.7913.40.421
SAPO-112950.263.51.11
zeolite β6090.493.21.92

Share and Cite

MDPI and ACS Style

Shamanaev, I.V.; Deliy, I.V.; Gerasimov, E.Y.; Pakharukova, V.P.; Bukhtiyarova, G.A. Enhancement of HDO Activity of MoP/SiO2 Catalyst in Physical Mixture with Alumina or Zeolites. Catalysts 2020, 10, 45. https://doi.org/10.3390/catal10010045

AMA Style

Shamanaev IV, Deliy IV, Gerasimov EY, Pakharukova VP, Bukhtiyarova GA. Enhancement of HDO Activity of MoP/SiO2 Catalyst in Physical Mixture with Alumina or Zeolites. Catalysts. 2020; 10(1):45. https://doi.org/10.3390/catal10010045

Chicago/Turabian Style

Shamanaev, Ivan V., Irina V. Deliy, Evgeny Yu. Gerasimov, Vera P. Pakharukova, and Galina A. Bukhtiyarova. 2020. "Enhancement of HDO Activity of MoP/SiO2 Catalyst in Physical Mixture with Alumina or Zeolites" Catalysts 10, no. 1: 45. https://doi.org/10.3390/catal10010045

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