Next Article in Journal
One Bicopper Complex with Good Affinity to Nitrate for Highly Selective Electrocatalytic Nitrate Reduction to Ammonia
Next Article in Special Issue
Synthesis and Application of Domestic Glassy Carbon TiO2 Nanocomposite for Electrocatalytic Triclosan Detection
Previous Article in Journal
Interface Engineering of SRu-mC3N4 Heterostructures for Enhanced Electrochemical Hydrazine Oxidation Reactions
Previous Article in Special Issue
Carbon Quantum Dot-Titanium Doped Strontium Ferrite Nanocomposite: Visible Light Active Photocatalyst to Degrade Nitroaromatics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

On Stability of High-Surface-Area Al2O3, TiO2, SiO2-Al2O3, and Activated Carbon Supports during Preparation of NiMo Sulfide Catalysts for Parallel Deoxygenation of Octanoic Acid and Hydrodesulfurization of 1-Benzothiophene

1
Institute of Chemical Process Fundamentals, Czech Academy of Sciences, Rozvojová 135, 165 02 Prague, Czech Republic
2
Institute of Catalysis, Bulgarian Academy of Sciences, Akad. G. Bonchev Str., Bl. 11, 1113 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(12), 1559; https://doi.org/10.3390/catal12121559
Submission received: 7 November 2022 / Revised: 24 November 2022 / Accepted: 28 November 2022 / Published: 2 December 2022
(This article belongs to the Special Issue Synthesis and Catalytic Application of Porous Carbon Materials)

Abstract

:
NiMo sulfide catalysts were prepared by the impregnation of high surface area supports with an aqueous solution made of NiCO3·2Ni(OH)2, MoO3 and citric acid, followed by freeze drying and sulfidation in H2S/H2 mixture. N2 physisorption and X-ray diffraction were selected to investigate the amphoteric oxides Al2O3 and TiO2, acidic SiO2-Al2O3 and activated carbon supports, fresh prepared sulfide NiMo catalysts and spent catalysts after model parallel reaction of octanoic acid deoxygenation and 1-benzothiophene hydrodesulfurization. The studied mesoporous amphoteric oxides Al2O3 and TiO2 did not lead to highly active NiMo catalysts due to the low hydrothermal stability of these supports during the preparation of the active sulfide phase and deoxygenation reaction. The most active catalyst based on oxidic support was the NiMo sulfide supported on acidic mesoporous SiO2-Al2O3, which was explained by the increased stability of this support to the water and CO/CO2 mixture during the activation of the sulfidic phase and deoxygenation reaction. The extraordinarily high stability of the activated carbon support led to outstanding activities of the sulfidic NiMo/C catalyst.

Graphical Abstract

1. Introduction

Modern refining calls for the hydrotreating of synthetic crude coming from waste plastic or renewables to reduce the environmental concerns of using fossil petroleum. These novel feeds bring into the hydrocarbon pool a high oxygen content of up to about 38 wt.% [1,2,3], mainly in the form of –OH or –COOH groups and their derivatives. During co-hydrotreating of synthetic or renewable feeds with fossil-based feeds, sulfide hydrotreating catalysts face extensive an hydrodeoxygenation (HDO) reaction in addition to an hydrodesulfurization (HDS) reaction. For this reason, the co-hydrotreating of Co(Ni)Mo [4,5,6,7,8,9,10,11,12,13,14,15,16], Ru [17], Re [18], or Pt [19,20] catalysts has been studied using feeds containing pyrolyzed biooils [4,5,6,7,8,9,10,11,12,17,18] or vegetable oils [13,14,15,16,19] together with thiophene derivatives [4,5,6,7,8,18] or petroleum fractions [9,10,11,12,13,14,15,16,19].
We have found in our previous work comparing MgO, C, ZrO2, TiO2, Al2O3, and MCM-41 supports [21] that the sulfidic NiMo phase on each support exhibited systematically higher HDO and HDS activity than the sulfidic CoMo and Mo phase. Nevertheless, this NiMo phase did not lead to highly active catalysts in HDO if it was supported on the amphoteric and basic oxides Al2O3, ZrO2 and MgO, respectively. The inhibition effect of water or CO/CO2 during HDO that had been reported previously [11,22,23,24,25,26,27] could explain this observation.
This inhibition principally occurs by two manners, either by destruction of the support or by poisoning of the active NiMoS phase.
First, the support structure and texture are destroyed by water. This phenomenon appeared obvious for high-surface-area MgO support [28,29] but it was also described for γ-Al2O3-supported NiMoS [23]. Gamma-alumina recrystallized to hydrous alpha-boehmite phase after treatment with water in the temperature region 140–380 °C. These structural changes were recorded by X-ray diffraction (XRD) and were accompanied with about a 26% decrease in specific surface area (SBET) [23].
Secondly, the active NiMoS phase suffers from partial oxidation. This oxidation mostly happens on the Ni that decorates the edges of the MoS2 slabs during HDO treatment. This Ni is considered as the most labile but the most active in HDO/HDS. Formation of nickel sulfates, oxides or migration of this labile nickel to form inactive nickel aluminates explains the activity losses [23]. Other inhibition can be due to the adsorption of CO/CO2 on NiMoS active sites [11]. CO/CO2 is formed by the decarbonylation and hydrodecarboxylation reactions during HDO. Moreover, CO/CO2 undergoes water–gas shift and methanation reactions at typical operation conditions of sulfidic NiMo catalysts, which slows down the intrinsic HDO/HDS [26]. Thus, avoiding decarbonylation and hydrodecarboxylation reactions during HDO is recommended [26,30].
The aim of the present work is to investigate the first means of HDO/HDS catalysts deactivation, i.e., the structural and textural changes of the support. The high-surface-area and mesoporous Al2O3, TiO2, SiO2-Al2O3 supports have been selected to deposit NiMo sulfides for parallel reactions of octanoic acid HDO and 1-benzothiophene HDS at 1.6 MPa and 330 °C. The high surface area of the mesopores should be able to disperse the high loading of the NiMo phase. The thin walls of these mesoporous supports should be more sensitive to destruction by H2O and/or an acidic CO2/H2O mixture during reduction sulfidation of catalysts’ precursors and/or the HDO/HDS. Activated carbon has been selected due to its inertness and high HDO/HDS activity [21] for the matter of comparison. We therefore elucidated the structural and textural properties of the supports, sulfide catalysts and sulfide catalysts after the HDO/HDS using XRD and nitrogen physisorption, respectively.

2. Results and Discussion

2.1. Structural Characteristics

X-ray diffraction patterns of the studied supports are provided in Figure 1. High-surface-area TiO2 was notably anatase (PDF 00-021-1272) while other supports were practically X-ray amorphous. Al2O3 exhibited broad peaks typical for its gamma phase.
After catalyst preparation, the X-ray patterns showed typical broad and low intensity peaks at about 33.5 (with a shoulder up to about 38) and 58.5 degrees that pointed on the presence of an X-ray amorphous phase of NiMo sulfides in all catalysts. Quite recently [31], the combination of Debye function analysis (DFA) applied to XRD data and high-resolution transmission electron microscopy (HRTEM) explained how these broad and low intensity peaks can be interpreted by the presence of nanocrystallites of MoS2. The larger shoulder at about 38 degrees was assigned to the presence of multilayers and therefore larger nanocrystallites of MoS2 [31]. This shoulder can be observed in our Al2O3- and TiO2-supported catalysts. However, it should be noted that the phase of anatase could also explain this shoulder in the TiO2-supported catalyst. For other catalysts, this shoulder is less pronounced compared to the intensity of the first peak at about 33.5 degrees.
Other recognized crystalline phases were observed only in the TiO2-supported cata-lyst. The phase of anatase (PDF 00-021-1272) after NiMo deposition was less pronounced in comparison with the original TiO2 support. Additionally, the reflections of orthorhombic TiO2 (PDF 01-082-1123) and brookite (PDF 00-029-1360) were identified. The low intensity but sharp peaks at 22.3 and 30.9 degrees can be explained by the presence of crystalline molybdena species such as hydrous molybdena (PDF 00-026-1449) or Mo9O26 (PDF 03-065-1292). This shows in the incomplete sulfidation of NiMo phase in the collapsed pores of the TiO2. This Mo species thus could not contribute to HDO/HDS reactions.
XRD documented that the high-surface-area TiO2 underwent deep structural changes during the catalyst preparation. The original structure of anatase deteriorated and the molybdena species recrystallized.

2.2. Textural Characteristics

2.2.1. Supports

The inner surface areas SBET of the supports and catalysts are shown in Table 1. Adsorption—desorption isotherms of N2 and pore-size distribution are summarized in Figure 2.
All supports exhibited SBET higher than 400 m2 g−1. Nevertheless, the oxidic supports possessed adsorption—desorption isotherms of type IV typical for materials containing mesopores while activated carbon was highly microporous with the isotherm of type I [32]. In order to quantify the inner surface area of the mesopores, SM, and the volume of micropores, VM, a t-plot analysis was done. The results are shown in Table 2. It was ascertained that practically all of the inner surface area of the oxidic supports was located in the mesopores (SBETSM). In contrast, the surface area of the mesopores of the activated carbon SM represented only about 24% of total SBET. The rest of this rather overestimated value of SBET originated from N2 sorption in the micropores with pore radii lower than 1 nm.
The following order of increasing mesopore surface area SM was found within the support studied: C < Al2O3 = TiO2 < SiO2-Al2O3.

2.2.2. NiMo Sulfide Catalysts

Deposition of the NiMo phase using an aqueous solution of citric acid followed by freeze-drying and sulfidation at 400 °C decreased the surface area SBET of the catalysts in comparison with the SBET of the parent supports. The surface area SBET of the sulfided catalysts were also normalized per unit weight of the support in Table 1, assuming the composition of the sulfidic phase NiS-MoS2, its content in the sulfided catalyst equal to 40 wt.%, and its inherent surface area as negligible. The SBET normalized in this way decreased by about 33, 71, 20, and 6% for the Al2O3-, TiO2-, SiO2-Al2O3-, and C-supported catalyst, respectively. If the inherent surface area of NiS-MoS2 nanocrystallites is not negligible, then this decrease in SBET would be higher. Clearly, this decrease in the SBET was significant for the Al2O3 and TiO2 supports while it was low for the SiO2-Al2O3 and C. In addition to the significant decrease in SBET and SM in Table 2, the studied TiO2 exhibited complex changes in pore-size distribution in Figure 2. Possible changes in the dispersion of NiMo sulfides (Figure 1) necessarily connected with changes in SBET and other textural parameters hardly explain these high differences between the two groups of supports. We concluded that the SiO2-Al2O3 represented the most stable support within the studied high-surface-area and mesoporous oxides. The activated carbon then exhibited extraordinary stability despite being microporous.
The explanation of this high stability of the SiO2-Al2O3 surface lays in the fact that SiO2-Al2O3 endures the reaction with water or CO/CO2 that is formed during reduction–sulfidation of the deposited complex of Ni, Mo and CA. In contrast, the Al2O3 and TiO2 surface is prone to hydrothermal deterioration. The complex of Ni, Mo and CA does not cover the entire support surface as was the case with the MoO3 monolayer in our previous works [33,34,35]. This surface (not covered by Mo) can react with the reaction water from the reduction–sulfidation or CO/CO2 originated from the citric acid, which causes the partial destruction of pore walls and a decrease in the SBET. The activated carbon is inert towards this type of chemical deterioration.

2.3. Catalytic Activity

Table 1 summarizes the activities kHDO and kHDS of the prepared catalysts in the HDO/HDS reaction. The fitting of kHDO and kHDS by Equations (1) and (2) is shown in Figure 3.
The high-surface-area Al2O3 resulted in the catalyst with the lowest HDO and HDS activity. The TiO2-supported catalyst exhibited increased HDO activity while HDS activity remained low. It should be noted that these activities are lower than those achieved previously [21] over conventional γ-Al2O3 (SBET = 262 m2 g−1) and TiO2 (anatase, SBET = 140 m2 g−1) carriers. In our previous study [21], the NiMo sulfidic catalysts possessed kHDO and kHDS 105 and 128 mmol g−1 h−1 for the γ-Al2O3 and 275 and 90 mmol g−1 h−1 for the anatase, respectively. The unprecedently low activities reached currently over the high-surface-area and mesoporous Al2O3 and TiO2 can be explained by the complete textural collapse of these supports during catalyst preparation.
In contrast, the SiO2-Al2O3-supported NiMo catalyst exhibited the highest activities in HDO and HDS comparing the studied mesoporous oxidic supports, which well correlates with the highest SBET given in Table 1 and other textural parameters summarized in Table 2 and Figure 2. Nevertheless, the activated carbon had the most inert surface to hydrothermal deterioration during the preparation of NiMo sulfides, which resulted in the most active catalyst within the studied series. The presented high kHDO and kHDS corroborated previously published results [21] that were achieved over similar activated carbon (product no. GA05) and the NiMo phase. The activated carbon thus remains the most promising support resulting in extraordinary high HDO/HDS activities.

2.4. Textural Stability during HDO/HDS Reactions

Table 1 and Table 2 and Figure 2 also show how the studied textural parameters of the catalysts differed after 3-h time on stream. We hardly found significant changes in textural parameters or HDO/HDS activities after this simplified deactivation test. Clearly, the studied catalysts were stable during the HDO/HDS experiment. The most significant changes in textural properties thus took place during the preparation of the catalysts including the reduction–sulfidation procedure as was discussed above. The water, CO, or CO2 that are produced during the HDO did not continue the deterioration of the textural properties, which could be explained by the low partial pressure of these inhibitors.

2.5. Selectivity in HDO/HDS Reactions

2.5.1. HDO/HDS Selectivity

Table 1 shows that NiMo/Al2O3 (the least active catalyst) and NiMo/C (the most active catalyst) were less selective to HDO than to HDS while NiMo/TiO2 and NiMo/SiO2-Al2O3 were about two times more selective to HDO than to HDS. The explanation of this HDO/HDS selectivity behavior has been previously found in empiric correlation with the point of zero charge, PZC, of the support [21]. The amphoteric supports with PZC about 7 result in NiMo catalysts with low selectivity to HDO. This is the case of the studied Al2O3 and activated carbon. In contrast, slightly more acidic TiO2 with PZC about 5 and the SiO2-Al2O3 with PZC about 4.5 are more selective to HDS. In summary, the hydrothermal inertness of the surface of the activated carbon leading to preservation of the catalyst texture and low metal-support interactions resulted in extraordinarily high HDO and HDS activities in the NiMo/C compared with the NiMo/SiO2-Al2O3.

2.5.2. Selectivity to Reaction Intermediates

In contrast to HDO/HDS selectivity, the selectivities to reaction intermediates during HDO/HDS reactions were similar over all of the prepared catalysts as shown in Figure 3.
No octanol was observed during the HDO of octanoic acid. The formation of olefins, the intermediates of the HDO reaction, was pronounced at low space time while hydrogenation to n-octane and n-heptane predominated at higher space time. We observed the same in our previous work over NiMo sulfides supported on commercial carriers [21].
The Cn−1 pathway (decarboxylation/decarbonylation) yielding C7 olefins and finally n-heptane predominated over all catalysts. The Cn pathway (intrinsic hydrodeoxygenation) yielding C8 olefins and finally n-octane was less pronounced. The Al2O3-, TiO2-, and C-supported NiMo catalysts yielded about 67% of C7 hydrocarbons (from total yield of hydrocarbons) while the SiO2-Al2O3-supported NiMo catalyst yielded about 75% of C7 hydrocarbons. We did not observe any cracking (Cn−x pathway). The acidity of SiO2-Al2O3 containing 25 wt.% of Al2O3 was thus lower than that of the MCM-41 in our previous paper [21]. The formation of branched hydrocarbons such as 2-methyl-hexane, 3-methyl-hexane, 2-(or 4-)methyl-heptane, and 3-methyl-heptane was observed only over the SiO2-Al2O3-supported NiMo catalyst, but within amounts lower than 2% of C8 + C7 yields, and so it was not shown in Figure 3.
The formation of dihydrobenzothiophene (DH), the reaction intermediate of 1-benzotiophene HDS, was negligible over all catalysts.
Figure 3 clearly shows that the hydrogenation function of the NiMo/C catalyst is predominant because the presence of linear olefins is the least pronounced within the studied catalysts.

2.6. Recapitulative Discussion and Outlook

The present contribution touched on important aspects for the forthcoming research of HDO/HDS catalysts:
  • The method of catalyst preparation should be tailored to the specific character of each individual support.
  • The properly used chelating agent leads to highly active catalysts containing a variety of supports including Al2O3 [31,36,37] but it appeared detrimental for the studied high surface area Al2O3 and TiO2.
  • The supports on the base of SiO2-Al2O3 represent a promising alternative to single oxides because they could be tuned in a wide range of acid-base properties [38,39,40,41,42,43,44] and because they increase dispersion of MoS2 compared to SiO2 [37]. Acidity seems crucial for the HDO of fatty acid containing feeds because it influences the formation of linear or branched hydrocarbons [21,41].
  • The active phase should hydrogenate olefins, the reaction intermediates of the HDO of fatty-acid-containing feeds. In this respect, the most promising support of the NiMo phase was found to be the activated carbon. Activated carbon as well as the use of citric acid contributes to low metal-support interaction and stability of Type II NiMoS phase containing a high number of corner sites [36,37].
  • The active phase and mesoporous support should provide stability to water and CO/CO2 that has not been needed to such an extent in a single HDS reaction [28,34,35,45,46,47,48,49].

3. Materials and Methods

The Al2O3 and SiO2-Al2O3 were synthesized by previously reported methods [33] and [50], respectively. SiO2-Al2O3 contained 25 wt.% of Al2O3, which was found optimal to achieve high HDS activity [51]. TiO2 was a commercial support (NanoScale Materials Inc., Manhattan, KA, USA, Product No. AC312-0100). Activated carbon, C, was a commercial support (Slovenské lučobné závody a.s., Hnúšťa, Slovakia, Product No. GA1). All supports were uniformized to particle size fraction 0.16-0.32 mm by grinding and sieving.
The impregnation solution was prepared by dissolution of NiCO3·2Ni(OH)2 (Sigma-Aldrich Chemie GmbH, Steinheim, Germany, Product No. 544183), MoO3 (Sigma-Aldrich Chemie GmbH, Steinheim, Germany, Product No. 69850) and citric acid, CA, (Lach-Ner s.r.o., Neratovice, Czechia, Product No. 10019-AP0) in water. The molar ratio of Ni:Mo:CA was 1.00:2.33:3.33. The support with impregnation solution was stirred for 15 min at laboratory temperature, cooled in liquid nitrogen for 1 min and dried in a freeze-dryer for 3 days. Freeze drying was chosen to lessen the interaction of Mo species with the support to form undesirable three dimensional heteropolymolybdates [52]. The relatively long 3-day freeze-drying time is necessary to remove moisture from the pores of the supports. We verified on part of the samples that an additional drying in a conventional oven at 110 °C did not lead to any weight loss. The samples removed from the freeze dryer were immediately sulfided in a gaseous H2S:H2 mixture (1:10) at 400 °C for 1 h. The catalysts were stored in gas tight ampules under N2. Chemical analysis by atomic absorption spectroscopy, AAS, confirmed the same content of Mo and Ni in all catalysts (19.0 ± 0.5 wt.% Mo and 5.0 ± 0.3 wt.% Ni).
X-ray diffraction measurements, XRD, were done on Bruker D8 Advanced Eco Powder X-ray analyzer (Bruker Corporation, Karlsruhe, Germany) with filtered CuKα radiation at 40 kV acceleration and 25 mA current of the X-ray tube, scan step 0.02° and 4 s accumulation time at each step. Physisorption of N2 was performed over the supports and catalysts on an ASAP 2020 Micromeritics instrument (Micromeritics Instrument Corporation, Norcross, GE, USA) after degassing at 350 °C and 1 Pa vacuum for 24 h. The adsorption–desorption isotherms of nitrogen at 77 K were evaluated by the standard Brunauer–Emmett–Teller procedure [53] for the p/p0 range = 0.05–0.25 to calculate the specific surface area SBET. The pore-size distribution (pore radius 1–10 nm) was evaluated from both the adsorption and desorption branch of N2 isotherm by the Barrett–Joyner–Halenda (BJH) method [54] assuming a cylindrical pore geometry. The Lecloux–Pirard standard isotherm [55] was used for the t–plot and for the pore–size distribution evaluation. Volume of micropores, VM, and surface area of mesopores, SM, were also determined [56]. The determination of SBET, SM, and VM were made with an accuracy of ±5%.
The activity of catalyst in parallel HDO/HDS of octanoic acid (OA)/1-benzothiophene (BT) was determined by the previously reported method [21] in laboratory-scale tubular flow reactor at 330 °C and 1.6 MPa. The catalyst weight, W, was 0.1, 0.2 and 0.4 g. The NiMo/C was additionally measured at weight 0.03 g due to its high activity; as reported therein. The reaction was run at three feed rates, F, of OA and BT: 7.7 mmol h−1, 10.3 mmol h−1, and 15.5 mmol h−1.The pseudo-first order rate constant, kHDO, of C8 + C7 hydrocarbons formation, yC8+C7, during the HDO of OA and the pseudo-first order rate constant, kHDS, of ethylbenzene (EB) formation, yEB, during the HDS of BT was fitted by non-linear regression and used as the index of catalyst activity:
yC8+C7 = 1 − exp (−kHDO × W/FOA)
yEB = 1 − exp (−kHDS × W/FBT)
In addition to this activity test [21], the catalysts were also compared at relative yields of C8 + C7 hydrocarbons yC8+C7 = 0.5 in terms of stability. The yield yC8+C7 = 0.5 was achieved at 50% conversion of OA. The catalysts were taken from the reactor after 3-h time on stream and their textural properties were characterized by N2 physisorption. The determination of kHDO and kHDS were made with an accuracy of ±10%.

4. Conclusions

The co-impregnation of the high-surface-area mesoporous Al2O3 and TiO2 supports with aqueous solution of NiCO3·2Ni(OH)2, MoO3 and citric acid, followed by freeze drying and reduction–sulfidation resulted in catalysts of low surface area and low activities in hydrodeoxygenation of octanoic acid and hydrodesulfurization of 1-benzothiophene. In contrast, this method of preparation was successful for the high-surface-area SiO2-Al2O3 support containing 25 wt.% of Al2O3 and the activated carbon. The BET area of the sulfidic NiMo/SiO2-Al2O3 catalyst with Ni and Mo contents of 5 and 20 wt.%, respectively, was 240 m2 g−1. This surface area practically represents the surface area of the mesopores and it remained the same after the HDO/HDS reaction. This catalyst mostly yielded linear hydrocarbons by HDO convenient for diesel fuel and it also exhibited high activity in HDS. Nevertheless, the most active catalyst in both HDO and HDS was NiMo/C. At the same loading of NiMo, the mesopore surface area was only about 140 m2 g−1 due to high microporosity of the activated carbon. The textural stability of the activated carbon towards the hydrothermal deterioration during the reduction–sulfidation step of catalyst preparation was the highest among the studied supports.

Author Contributions

Conceptualization, writing and funding acquisition, L.K.; methodology K.S.; investigation, M.K.; data curation, J.K.; resources, R.P., formal analysis, M.L.; visualization, D.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Czech Science Foundation grant no. 17-22490S, the Technology Agency of the Czech Republic grant No. TN01000048 and mobility project CAS-BAS grant no. BAS-20-01.

Data Availability Statement

Not applicable.

Acknowledgments

The senior scientists M. Zdražil and O. Šolcová are acknowledged for their help with reaction progress kinetic analysis (RPKA) and textural analysis (BET, BJH), respectively.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Furimsky, E. Catalytic hydrodeoxygenation. Appl. Catal. A 2000, 199, 147–190. [Google Scholar] [CrossRef]
  2. Mortensen, P.M.; Grunwaldt, J.D.; Jensen, P.A.; Knudsen, K.G.; Jensen, A.D. A review of catalytic upgrading of bio-oil to engine fuels. Appl. Catal. A 2011, 407, 1–19. [Google Scholar] [CrossRef]
  3. Furimsky, E. Hydroprocessing challenges in biofuels production. Catal. Today 2013, 217, 13–56. [Google Scholar] [CrossRef]
  4. Platanitis, P.; Panagiotou, G.D.; Bourikas, K.; Kodulis, C.; Lycourghiotis, A. Hydrodeoxygenation of phenol over hydrotreatment catalysts in their reduced and sulfided states. Open Catal. J. 2014, 7, 18–25. [Google Scholar] [CrossRef] [Green Version]
  5. Xia, L.; Xia, Z.; Tang, W.; Wang, H.; Fang, M. Hydrogenation of Model Compounds Catalyzed by MCM-41-Supported Nickel Phosphide. Adv. Mater. Res. 2014, 864–867, 366–372. [Google Scholar] [CrossRef]
  6. Wang, W.; Zhu, G.; Li, L.; Tan, S.; Wu, K.; Zhang, X.; Yang, Y. Facile hydrothermal synthesis of flower-like Co–Mo–S catalysts and their high activities in the hydrodeoxygenation of p-cresol and hydrodesulfurization of benzothiophene. Fuel 2016, 174, 1–8. [Google Scholar] [CrossRef]
  7. Dhandapani, B.; Clair, T.; Oyama, S.T. Simultaneous hydrodesulfurization, hydrodeoxygenation, and hydrogenation with molybdenum carbide. Appl. Catal. A 1998, 168, 219–228. [Google Scholar] [CrossRef]
  8. Odebunmi, E.O.; Ollis, D.F. Catalytic hydrodeoxygenation: II. Interactions between catalytic hydrodeoxygenation of m-cresol and hydrodesulfurization of benzothiophene and dibenzothiophene. J. Catal. 1983, 80, 65–75. [Google Scholar] [CrossRef]
  9. Pstrowska, K.; Walendziewski, J.; Łuzny, R.; Stolarski, M. Hydroprocessing of rapeseed pyrolysis bio-oil over NiMo/Al2O3 catalyst. Catal. Today 2014, 223, 54–65. [Google Scholar] [CrossRef]
  10. Pstrowska, K.; Walendziewski, J.; Stolarski, M. Hydrorefining of oil from rapeseed cake pyrolysis over NiMo/Al2O3 catalyst. Fuel Process. Technol. 2014, 128, 191–198. [Google Scholar] [CrossRef]
  11. Pinheiro, A.; Hudebine, D.; Dupassieux, N.; Geantet, C. Impact of oxygenated compounds from lignocellulosic biomass pyrolysis oils on gas oil hydrotreatment. Eng. Fuel. 2009, 23, 1007–1014. [Google Scholar] [CrossRef]
  12. Bui, V.N.; Toussaint, G.; Laurenti, D.; Mirodatos, C.; Geantet, C. Co-processing of pyrolisis bio oils and gas oil for new generation of bio-fuels: Hydrodeoxygenation of guaïacol and SRGO mixed feed. Catal. Today 2009, 143, 172–178. [Google Scholar] [CrossRef]
  13. Vonortas, A.; Kubicka, D.; Papayannakos, N. Catalytic co-hydroprocessing of gasoil–palm oil/AVO mixtures over a NiMo/γ-Al2O3 catalyst. Fuel 2014, 116, 49–55. [Google Scholar] [CrossRef]
  14. Varakin, A.N.; Salnikov, V.A.; Nikulshina, M.S.; Maslakov, K.I.; Mozhaev, A.V.; Nikulshin, P.A. Beneficial role of carbon in Co(Ni)MoS catalysts supported on carbon-coated alumina for co-hydrotreating of sunflower oil with straight-run gas oil. Catal. Today 2017, 292, 110–120. [Google Scholar] [CrossRef]
  15. Nikulshin, P.A.; Salnikov, V.A.; Pimerzin, A.A.; Eremina, Y.V.; Koklyukhin, A.S.; Tsvetkov, V.S.; Pimerzin, A.A. Co-hydrotreating of straight-run diesel fraction and vegetable oil on Co(Ni)-PMo/Al2O3 catalysts. Petrol. Chem. 2016, 56, 56–61. [Google Scholar] [CrossRef]
  16. Eller, Z.; Varga, Z.; Hancsók, J. Renewable Jet Fuel from Kerosene/Coconut Oil Mixtures with Catalytic Hydrogenation. Energ. Fuel. 2019, 33, 6444–6453. [Google Scholar] [CrossRef]
  17. Mercader, F.M.; Groeneveld, M.J.; Kersten, S.R.A.; Geantet, C.; Toussaint, G.; Way, N.W.J.; Schaverien, C.J.; Hogendoorn, K.J.A. Hydrodeoxygenation of pyrolysis oil fractions: Process understanding and quality assessment through co-processing in refinery units. Eng. Environ. Sci. 2011, 4, 985–997. [Google Scholar]
  18. Sepúlveda, C.; Escalona, N.; García, R.; Laurenti, D.; Vrinat, M. Hydrodeoxygenation and hydrodesulfurization co-processing over ReS2 supported catalysts. Catal. Today 2012, 195, 101–105. [Google Scholar] [CrossRef]
  19. Vonortas, A.; Papayannakos, N. Hydrodesulphurization and hydrodeoxygenation of gasoil-vegetable oil mixtures over a Pt/γ-Al2O3 catalyst. Fuel Process. Technol. 2016, 150, 126–131. [Google Scholar] [CrossRef]
  20. Morales-Hernández, G.; Pacheco-Sosa, J.G.; Escobar-Aguilar, J.; Torres, J.G.T.; Pérez-Vidal, H.; Lunagómez-Rocha, M.A.; De La Cruz-Romero, D.; Del Ángel-Vicente, P. Improving platinum dispersion on sba-15 by titania addition. Rev. Mex. Ing. Quim. 2020, 19, 997–1010. [Google Scholar] [CrossRef] [Green Version]
  21. 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 Reac. Kinet. Mech. Cat. 2019, 127, 887–902. [Google Scholar] [CrossRef]
  22. Yoshimura, Y.; Sato, T.; Shimada, H.; Matsubayashi, N.; Nishijima, A. Influences of oxygen-containing substances on deactivation of sulfided molybdate catalysts. Appl. Catal. 1991, 73, 55–63. [Google Scholar] [CrossRef]
  23. Laurent, E.; Delmon, B. Influence of water in the deactivation of a sulfided NiMoγ-Al2O3 catalyst during hydrodeoxygenation. J. Catal. 1994, 146, 281–291. [Google Scholar] [CrossRef]
  24. Li, Y.; Wang, R.; Chang, L. Study of reactions over sulfide catalysts in CO–CO2–H2–H2O system. Catal. Today 1999, 51, 25–38. [Google Scholar] [CrossRef]
  25. Senol, O.I.; Viljava, T.R.; Krause, A.O.I. Hydrodeoxygenation of aliphatic esters on sulphided NiMo/γ-Al2O3 and CoMo/γ-Al2O3 catalyst: The effect of water. Catal. Today 2005, 106, 186–189. [Google Scholar]
  26. 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. Eng. Fuel. 2011, 25, 804–812. [Google Scholar] [CrossRef]
  27. 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]
  28. Mejias, J.A.; Berry, A.J.; Refson, K.; Fraser, D.G. The kinetics and mechanism of MgO dissolution Chem. Phys. Lett. 1999, 314, 558–563. [Google Scholar]
  29. Klicpera, T.; Zdražil, M. Preparation of high-activity MgO-supported Co–Mo and Ni–Mo sulfide hydrodesulfurization catalysts. J. Catal. 2002, 206, 314–320. [Google Scholar] [CrossRef]
  30. 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]
  31. Pakharukova, V.P.; Yatsenko, D.A.; Gerasimov, E.Y.; Vlasova, E.; Bukhtiyarova, G.A.; Tsybulya, S.V. Total Scattering Debye Function Analysis: Effective Approach for Structural Studies of Supported MoS2-Based Hydrotreating Catalysts. Ind. Eng. Chem. Res. 2020, 59, 10914–10922. [Google Scholar] [CrossRef]
  32. Gregg, S.J.; Sing, K.S.W. Adsorption, Surface Area and Porosity, 2nd ed.; Academic Press: London, UK, 1982; pp. 1–303. [Google Scholar]
  33. Kaluža, L.; Gulková, D.; Šolcová, O.; Žilková, N.; Čejka, J. Hydrotreating catalysts supported on organized mesoporous alumina: Optimization of Mo deposition and promotional effects of Co and Ni. Appl. Catal. A Gen. 2008, 351, 93–101. [Google Scholar] [CrossRef]
  34. Kaluža, L.; Zdražil, M.; Žilková, N.; Čejka, J. High activity of highly loaded MoS2 hydrodesulfurization catalysts supported on organised mesoporous alumina. Catal. Commun. 2002, 3, 151–157. [Google Scholar] [CrossRef]
  35. Čejka, J.; Žilková, N.; Kaluža, L.; Zdražil, M. Mesoporous alumina as a support for hydrodesulfurization catalysts. Stud. Surf. Sci. Catal. 2002, 141, 243–250. [Google Scholar]
  36. Escobar, J.; Barrera, M.C.; Gutiérrez, A.W.; Terrazas, J.E. Benzothiophene hydrodesulfurization over NiMo/alumina catalysts modified by citric acid. Effect of addition stage of organic modifier. Fuel Process. Technol. 2017, 156, 33–42. [Google Scholar] [CrossRef]
  37. Hensen, E.J.M.; Kooyman, P.J.; Van der Meer, Y.; Van der Kraan, A.M.; De Beer, V.H.J.; Van Veen, J.A.R.; Van Santen, R.A. The Relation between Morphology and Hydrotreating Activity for Supported MoS2 Particles. J. Catal. 2001, 199, 224–235. [Google Scholar] [CrossRef]
  38. Nguyen, M.T.; Tayakout-Fayolle, M.; Chainet, F.; Pirngruber, G.D.; Geantet, C. Use of kinetic modeling for investigating support acidity effects of NiMo sulfide catalysts on quinoline hydrodenitrogenation. C. Appl. Catal. A Gen. 2017, 530, 132–144. [Google Scholar] [CrossRef]
  39. Coumans, A.E.; Hensen, E.J.M. A real support effect on the hydrodeoxygenation of methyl oleate by sulfided NiMo catalysts. Catal. Today 2017, 298, 181–189. [Google Scholar] [CrossRef]
  40. Escalona, G.; Rai, A.; Betancourt, P.; Sinha, A.K. Selective poly-aromatics saturation and ring opening during hydroprocessing of light cycle oil over sulfided Ni-Mo/SiO2-Al2O3 catalyst. Fuel 2018, 219, 270–278. [Google Scholar] [CrossRef]
  41. Ramesh, A.; Tamizhdurai, P.; Krishnan, P.S.; Ponnusamy, V.K.; Sakthinathan, S.; Shanthi, K. Catalytic transformation of non-edible oils to biofuels through hydrodeoxygenation using Mo-Ni/mesoporous alumina-silica catalysts. Fuel 2020, 262, 116494. [Google Scholar] [CrossRef]
  42. Ramesh, A.; Tamizhdurai, P.; Mangesh, V.L.; Palanichamy, K.; Gopinath, S.; Sureshkumar, K.; Shanthi, K. Mg/SiO2–Al2O3 supported nickel catalysts for the production of naphthenic hydrocarbon fuel by hydro-de-oxygenation of eugenol. Int. J. Hydrogen. Energ. 2019, 44, 25607–25620. [Google Scholar] [CrossRef]
  43. Agliullin, M.R.; Danilova, I.G.; Faizullin, A.V.; Amarantov, S.V.; Bubennov, S.V.; Prosochkina, T.R.; Grigoreva, N.G.; Paukshtis, E.A.; Kutepov, B.I. Sol-gel synthesis of mesoporous aluminosilicates with a narrow pore size distribution and catalytic activity thereof in the oligomerization of dec-1-ene. Micropor. Mesopor. Mat. 2016, 230, 118–127. [Google Scholar] [CrossRef]
  44. Serrano, D.P.; Escola, J.M.; Briones, L.; Arroyo, M. Selective hydrodecarboxylation of fatty acids into long-chain hydrocarbons catalyzed by Pd/Al-SBA-15. Micropor. Mesopor. Mat. 2019, 280, 88–96. [Google Scholar] [CrossRef]
  45. Dzwigaj, S.; Louis, C.; Breysse, M.; Cattenot, M.; Bellière, V.; Geantet, C.; Vrinat, M.; Blanchard, P.; Payen, E.; Inoue, S.; et al. New generation of titanium dioxide support for hydrodesulfurization. Appl. Catal. B: Environ. 2003, 41, 181–191. [Google Scholar] [CrossRef]
  46. Wang, H.; Xiao, B.; Cheng, X.; Wang, C.; Zhao, L.; Zhu, Y.; Zhu, J.; Lu, X. NiMo catalysts supported on graphene-modified mesoporous TiO2 toward highly efficient hydrodesulfurization of dibenzothiophene. Appl. Catal. A Gen. 2015, 502, 157–165. [Google Scholar] [CrossRef]
  47. Platanitis, P.; Panagiotou, G.D.; Bourikas, K.; Kordulis, C.; Fierro, J.L.G.; Lycourghiotis, A. Preparation of un-promoted molybdenum HDS catalysts supported on titania by equilibrium deposition filtration: Optimization of the preparative parameters and investigation of the promoting action of titania. J. Mol. Catal. A Chem. 2016, 412, 1–12. [Google Scholar] [CrossRef]
  48. Naboulsi, I.; Lebeau, B.; Aponte, C.F.L.; Brunet, S.; Mallet, M.; Michelin, L.; Bonne, M.; Carteret, C.; Blin, J.L. Selective direct desulfurization way (DDS) with CoMoS supported over mesostructured titania for the deep hydrodesulfurization of 4, 6-dimethydibenzothiophene. Appl. Catal. A Gen. 2018, 563, 91–97. [Google Scholar] [CrossRef]
  49. Li, M.; Song, J.; Yue, F.; Pan, F.; Yan, W.; Hua, Z.; Li, L.; Yang, Z.; Li, L.; Wen, G.; et al. Complete Hydrodesulfurization of Dibenzothiophene via Direct Desulfurization Pathway over Mesoporous TiO2-Supported NiMo Catalyst Incorporated with Potassium. Catalysts 2019, 9, 448. [Google Scholar] [CrossRef] [Green Version]
  50. Vít, Z.; Šolcová, O. Synthesis and properties of mesoporous silica–alumina with narrow pore size distribution obtained without use of pore-regulating agents. Micropor. Mesopor. Mater. 2006, 96, 197–204. [Google Scholar] [CrossRef]
  51. Kaluža, L.; Gulková, D.; Vít, Z.; Zdražil, M. Water-assisted spreading of MoO3 onto SiO2–Al2O3 supports for preparation of sulfide CoMo hydrodesulfurization catalysts. Fuel 2013, 112, 272–276. [Google Scholar] [CrossRef]
  52. Carrier, X.; Lambert, J.F.; Kuba, S.; Knözinger, H.; Che, M. Influence of ageing on MoO3 formation in the preparation of alumina-supported Mo catalysts. J. Mol. Struct. 2003, 656, 231–238. [Google Scholar] [CrossRef]
  53. Brunauer, S.; Emmett, P.H.; Teller, E. Adsorption of gases in multimolecular layers. J. Am. Chem. Soc. 1938, 60, 309–319. [Google Scholar] [CrossRef]
  54. Barrett, E.P.; Joyner, L.G.; Halenda, P.P. The determination of pore volume and area distributions in porous substances. I. Computations from nitrogen isotherms J. Am. Chem. Soc. 1951, 73, 373–380. [Google Scholar]
  55. Lecloux, A.; Pirard, J.P. The importance of standard isotherms in the analysis of adsorption isotherms for determining the porous texture of solids. J. Colloid Interface Sci. 1979, 70, 265–281. [Google Scholar] [CrossRef]
  56. Schneider, P. Adsorption isotherms of microporous-mesoporous solids revisited. Appl. Catal. A Gen. 1995, 129, 157–165. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns of supports and sulfide NiMo catalysts. Characteristic peaks of TiO2: (+) anatase, PDF 00-021-1272; (o) orthorhombic TiO2, PDF 01-082-1123; and (x) possibly brookite, PDF 00-029-1360. Characteristic peaks of Mo oxidic species (*): MoO3.H2O, PDF 00-026-1449 or Mo9O26, PDF 03-065-1292. Characteristics low intensity and broad peaks of gamma Al2O3 (□) and nanocrystallites of MoS2 (•).
Figure 1. X-ray diffraction (XRD) patterns of supports and sulfide NiMo catalysts. Characteristic peaks of TiO2: (+) anatase, PDF 00-021-1272; (o) orthorhombic TiO2, PDF 01-082-1123; and (x) possibly brookite, PDF 00-029-1360. Characteristic peaks of Mo oxidic species (*): MoO3.H2O, PDF 00-026-1449 or Mo9O26, PDF 03-065-1292. Characteristics low intensity and broad peaks of gamma Al2O3 (□) and nanocrystallites of MoS2 (•).
Catalysts 12 01559 g001
Figure 2. Textural properties of the support (blue lines), sulfided catalysts (black lines) and spent catalysts after 3 h time of stream at 50% conversion of octanoic acid (red lines). N2 adsorption desorption isotherms (a), pore-size distribution from the adsorption branch of the isotherm (b), and pore-size distribution from the desorption branch of the isotherm (c).
Figure 2. Textural properties of the support (blue lines), sulfided catalysts (black lines) and spent catalysts after 3 h time of stream at 50% conversion of octanoic acid (red lines). N2 adsorption desorption isotherms (a), pore-size distribution from the adsorption branch of the isotherm (b), and pore-size distribution from the desorption branch of the isotherm (c).
Catalysts 12 01559 g002
Figure 3. The formation of linear C8 and C7 hydrocarbons and ethylbenzene (EB) during parallel HDO/HDS of octanoic acid (OA) and 1-benzothiophene (BT) over the studied catalysts: solid lines—fitting of activity kHDO and kHDS by Equations (1) and (2), respectively.
Figure 3. The formation of linear C8 and C7 hydrocarbons and ethylbenzene (EB) during parallel HDO/HDS of octanoic acid (OA) and 1-benzothiophene (BT) over the studied catalysts: solid lines—fitting of activity kHDO and kHDS by Equations (1) and (2), respectively.
Catalysts 12 01559 g003
Table 1. The comparison of selected textural and catalytic properties.
Table 1. The comparison of selected textural and catalytic properties.
Surface Area SBET (m2 g−1)Catalyst Activity k
Support aCatalysts(mmol g−1 h−1)
Fresh bSpent ckHDOkHDS
Per Gram of SupportPer Gram of CatalystPer Gram of SupportPer Gram of CatalystPer Gram of Support
Al2O34131652751542571951
TiO2417721208614312666
SiO2-Al2O350023939825742819890
C919517862515858250420
aSBET of original support; b SBET of sulfided NiMo catalyst; c SBET of NiMo catalyst after 3-h time on stream at 50% conversion of octanoic acid.
Table 2. The comparison of specific surface area SBET with mesopore surface area SM and volume of micropores VM normalized per gram of sample (SM and VM calculated by t-plot method).
Table 2. The comparison of specific surface area SBET with mesopore surface area SM and volume of micropores VM normalized per gram of sample (SM and VM calculated by t-plot method).
SupportFresh CatalystSpent Catalyst a
SBET
m2 g−1
SM
m2 g−1
VM
mm3 g−1
SBET
m2 g−1
SM
m2 g−1
VM
mm3 g−1
SBET
m2 g−1
SM
m2 g−1
VM
mm3 g−1
Al2O3413399216516421541434
TiO2417415<17272<18686<1
SiO2-Al2O3500500<12392051525722911
C919224336517140194515139193
a after 3 h time on stream at 50% conversion of octanoic acid.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kaluža, L.; Soukup, K.; Koštejn, M.; Karban, J.; Palcheva, R.; Laube, M.; Gulková, D. On Stability of High-Surface-Area Al2O3, TiO2, SiO2-Al2O3, and Activated Carbon Supports during Preparation of NiMo Sulfide Catalysts for Parallel Deoxygenation of Octanoic Acid and Hydrodesulfurization of 1-Benzothiophene. Catalysts 2022, 12, 1559. https://doi.org/10.3390/catal12121559

AMA Style

Kaluža L, Soukup K, Koštejn M, Karban J, Palcheva R, Laube M, Gulková D. On Stability of High-Surface-Area Al2O3, TiO2, SiO2-Al2O3, and Activated Carbon Supports during Preparation of NiMo Sulfide Catalysts for Parallel Deoxygenation of Octanoic Acid and Hydrodesulfurization of 1-Benzothiophene. Catalysts. 2022; 12(12):1559. https://doi.org/10.3390/catal12121559

Chicago/Turabian Style

Kaluža, Luděk, Karel Soukup, Martin Koštejn, Jindřich Karban, Radostina Palcheva, Marek Laube, and Daniela Gulková. 2022. "On Stability of High-Surface-Area Al2O3, TiO2, SiO2-Al2O3, and Activated Carbon Supports during Preparation of NiMo Sulfide Catalysts for Parallel Deoxygenation of Octanoic Acid and Hydrodesulfurization of 1-Benzothiophene" Catalysts 12, no. 12: 1559. https://doi.org/10.3390/catal12121559

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