Next Article in Journal
Comprehensive Analysis for Electromagnetic Shielding Method Based on Mesh Aluminium Plate for Electric Vehicle Wireless Charging Systems
Next Article in Special Issue
Comprehensive Review on Potential Contamination in Fuel Ethanol Production with Proposed Specific Guideline Criteria
Previous Article in Journal
Development of Renewable Energy Sources in the European Union in the Context of Sustainable Development Policy
Previous Article in Special Issue
In Situ Binder-Free and Hydrothermal Growth of Nanostructured NiCo2S4/Ni Electrodes for Solid-State Hybrid Supercapacitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Catalytic Hydrotreating of Crude Pongamia pinnata Oil to Bio-Hydrogenated Diesel over Sulfided NiMo Catalyst

by
Yuwadee Plaola
1,
Wanwipa Leangsiri
2,
Kanokthip Pongsiriyakul
1,
Worapon Kiatkittipong
1,*,
Atthapon Srifa
3,
Jun Wei Lim
4,
Prasert Reubroycharoen
5,
Kunlanan Kiatkittipong
6,
Apiluck Eiad-ua
7 and
Suttichai Assabumrungrat
2,8
1
Department of Chemical Engineering, Faculty of Engineering and Industrial Technology, Silpakorn University, Nakhon Pathom 73000, Thailand
2
Center of Excellence in Catalysis and Catalytic Reaction Engineering, Department of Chemical Engineering, Faculty of Engineering, Chulalongkorn University, Bangkok 10330, Thailand
3
Department of Chemical Engineering, Faculty of Engineering, Mahidol University, Nakhon Pathom 73170, Thailand
4
Department of Fundamental and Applied Sciences, HICoE-Centre for Biofuel and Biochemical Research, Institute of Self-Sustainable Building, Universiti Teknologi PETRONAS, Seri Iskandar 32610, Malaysia
5
Center of Excellence in Catalysis for Bioenergy and Renewable Chemicals (CBRC), Department of Chemical Technology, Faculty of Science, Chulalongkorn University, Bangkok 10330, Thailand
6
Department of Chemical Engineering, Faculty of Engineering, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand
7
College of Materials Innovation and Technology, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand
8
Bio-Circular-Green-economy Technology and Engineering Center (BCGeTEC), Department of Chemical Engineering, Faculty of Engineering, Chulalongkorn University, Bangkok 10330, Thailand
*
Author to whom correspondence should be addressed.
Submission received: 20 January 2022 / Revised: 10 February 2022 / Accepted: 15 February 2022 / Published: 19 February 2022
(This article belongs to the Special Issue From Unidisciplinary to Multidisciplinary Energy Research)

Abstract

:
This work studied the catalytic activity and stability of Ni-MoS2 supported on γ-Al2O3, SiO2, and TiO2 toward deoxygenation of different feedstocks, i.e., crude Pongamia pinnata oil (PPO) and refined palm olein (RPO). PPO was used as a renewable feedstock for bio-hydrogenated diesel production via catalytic hydrotreating under a temperature of 330 °C, H2 pressure of 50 bar, WHSV of 1.5 h−1, and H2/oil (v/v) of 1000 cm3/cm3 under continuous operation. The oil yield from a Soxhlet extraction of PPO was up to 26 wt.% on a dry basis, mainly consisting of C18 fatty acids. The catalytic activity in terms of conversion and diesel yield was in the same trend as increasing in the order of NiMo/γ-Al2O3 > NiMo/TiO2 > NiMo/SiO2. The hydrodeoxygenation (HDO) activity was more favorable over the sulfided NiMo supported on γ-Al2O3 and TiO2, while a high DCO was observed over the sulfided NiMo/SiO2 catalyst, which related to the properties of the support material and the intensity of metal–support interaction. The deactivation of NiMo/SiO2 and NiMo/TiO2 occurred in a short period, due to the phosphorus and alkali impurities in PPO which were not found in the case of RPO. NiMo/γ-Al2O3 exhibited the high resistance of impure feedstock with excellent stability. This indicates that the catalytic performance is influenced by the purity of the feedstock as well as the characteristics of the catalysts.

1. Introduction

In recent years, the use of uncertain and insecure petroleum energy has been suggested as a major problem causing carbon dioxide emissions, which directly affect global climate change [1,2]. Many researchers have focused on the reduction of carbon dioxide emissions by shifting to renewable fuels. Plant biomass is an important bioresource that can reduce the environmental impact as CO2 from the atmosphere is used for plant growth. Triglycerides in vegetable oils and animal fats, mainly consisting of C16–C18 fatty acids, have been utilized as important feedstocks for the production of renewable fuels due to their similar carbon chain length to diesel fuel. Generally, edible vegetable oils such as soybean oil [3], rapeseed oil [4], sunflower oil [5], and palm oil [6] have been employed as sources of triglyceride-based feedstock for biofuel production. To avoid the competitive issue of food supplies, several non-food feedstocks, such as inedible byproducts, e.g., palm fatty acid distillate [7,8], non-edible oil (e.g., Jatropha oil, rubber seed oil, and tall oil) [9,10], waste oils (e.g., waste cooking oils and waste animal fat) [11,12], and novel feedstocks (e.g., algae [13]), have received increased attention for biofuel production.
Pongamia pinnata is indigenous to the Indian subcontinent and Southeast Asia. It is a fast-growing plant that is particularly resistant to drought and salt, thus it could be cultivated on marginal land in broad tropical and subtropical areas, potentially reducing the issue of land use. Additionally, it has a high annual oil yield, so oil extracted from Pongamia pinnata seeds has been recently promoted as an energy renewable feedstock in several countries such as India and Australia. In India, it is the second-largest oil crop compared with Jatropha oil [14,15]. The advantages of Pongamia pinnata are high oil content and resistance to the drought-prone environment. The Pongamia pinnata seeds consist of 20%–42% of oil content based on dry weight [16]. In particular, the Pongamia pinnata can produce extracted oil of approximately 3600–4800 L ha−1 year−1 which could be competitive with palm oil of 5950 L ha−1 year−1 and higher than jatropha oil of 1892 L ha−1 year−1 [17]. The dominant fatty acids of extracted oil from Pongamia pinnata seeds are palmitic acid, stearic acid, linoleic acid, and eicosenoic acid, making them promising for biodiesel production [18]. The first-generation of biodiesel is the esterification of fatty acids and transesterification of triglycerides with alcohols, which produce the fatty acid methyl esters (FAMEs), ester-based biodiesel. The second-generation technology is catalytic hydrotreating to produce bio-hydrogenated diesel or green diesel which provides better fuel properties compared with biodiesel, such as no oxygen content, high cetane number, and, in particular, high thermal and oxidation stability [19,20,21,22]. This alkane-based renewable diesel can be produced by the catalytic hydrotreating of triglycerides and fatty acids via three main reaction pathways, including decarbonylation (DCO), decarboxylation (DCO2), and hydrodeoxygenation (HDO), which lead to the elimination of oxygen in the form of CO and H2O, or CO2, or H2O, respectively, as shown in Equations (1)–(3) [23].
R−CH2−COOH → R−CH3 + CO2
R−CH2−COOH → R=CH2 + CO + H2O
R−CH2−COOH + 3H2 → R−CH2−CH3 + 2H2O
Additionally, the generated CO and CO2 can further react with water or H2 via water gas shift (Equation (4)) and methanation (Equations (5) and (6)) reactions.
CO + H2O ↔ CO2 + H2
CO + 3H2 ↔ CH4 + H2O
CO2 + 4H2 ↔ CH4 + 2H2O
The bimetallic sulfide catalysts such as NiMoS2, CoMoS2, and NiWS2 supported on zeolites, carbon, Al2O3, SiO2, and TiO2 are suitable for deoxygenation of triglycerides due to high selectivity toward hydrodeoxygenation at moderate temperatures [24,25]. Among them, supported NiMoS2 catalysts showed the highest catalytic activity for the reaction. Al2O3 was selected as a support material because it has moderate acidity and a high surface area [26]. SiO2 shows a high surface area and pores volume, assisting the diffusion of the triglyceride molecules to the catalyst active sites [27]. TiO2 can facilitate the transformation of MoO3 to MoS2 which acted as an active species of the catalyst [28]. The long-term stability test was previously reported using the supported NiMo catalysts. The effect of reaction time on hydrocarbon conversion was investigated in the range of 80–120 h. Gong et al. [10] investigated the lifetime of sulfided NiMoP/Al2O3 catalyst using Jatropha oil as a feedstock under a temperature of 350 °C and 30 bar of H2 pressure. It was found that the catalyst was deactivated after a reaction time of 120 h due to the transformation of sulfide species to oxide species of the catalyst. Toba et al. [11] studied the addition of sulfur in feed during the deoxygenating of waste cooking oil over NiMo/B2O3-Al2O3 catalysts. The ratio of n-C17/(n-C17 + n-C18) remained constant at 20% for 80 h on-stream. It was confirmed that the sulfur leaching in the liquid product was high at the initial stage of the reaction, meanwhile, the sulfur contamination was low and the catalyst activity remained constant after 20–80 h on-stream. However, to our knowledge, the direct comparison of the effect of support materials including Al2O3, SiO2, and TiO2 on catalytic activity and stability tests over typical NiMo bimetallic sulfide catalysts especially on unpurified crude has not been fully investigated in the literature. In addition, Pongamia pinnata is currently a feedstock of interest for ester-based biodiesel production but there are rather limited studies on its conversion to bio-hydrogenated diesel [29,30,31]. Moreover, there is a lack of study on the catalyst stability by using Pongamia pinnata oil as a feedstock.
Therefore, this research focuses on the hydrotreating of Pongamia pinnata oil to produce bio-hydrogenated diesel over sulfided NiMo supported on different industrially important support materials, i.e., γ-Al2O3, SiO2, and TiO2. Different behaviors of catalytic activity and catalyst stability among these supports are highlighted. Moreover, refined oil such as palm olein was also used as a feedstock to reveal the effect of feedstock on these supports. The physical and chemical properties of the synthesized catalysts were characterized by XRD, H2-TPR, and N2 sorption.

2. Materials and Methods

2.1. Materials

The γ-Al2O3, TiO2, and SiO2 supports were purchased from Alfa Aesar Chemical Co., Ltd. (Ward Hill, MA, USA). The catalyst precursors, Ni (NO3)2·6H2O (purity = 97.0%) and (NH4)6Mo7O24·4H2O (purity = 81.0%–83.0%) were purchased from Ajax Finechem Pty Ltd. (New South Walesm, Australia). The standard mixture for ASTM-2887 was supplied by Restek Co., Ltd. (Bellefonte, PA, USA). Hydrogen gas (99.99 vol%) was purchased from Linde Public Co., Ltd. (Bangkok, Thailand). Refined palm olein was obtained from Patum Vegetable Oil Co., Ltd. (Bangkok, Thailand).

2.2. Oil Extraction

The Pongamia pinnata seeds were crushed and sieved to a particle size of less than 1 mm and further dried in an oven overnight at 110 °C in order to remove moisture content. Subsequently, the extraction was employed using a Soxhlet extractor with n-hexane as solvent. Briefly, 400 g of dried Pongamia pinnata seeds and 3 L of n-hexane were loaded into the extraction thimble. An extraction time of 12 h was required per batch. Finally, the obtained solutions were further evaporated in order to recover n-hexane using a rotary evaporator. The percentage of oil extraction was calculated using the following equation.
% Oil   extraction = weight   of   the   extracted   oil dry   weight   of   Pongamia   pinnata   seeds ×   100

2.3. Catalysts Preparation

The γ-Al2O3, SiO2, and TiO2 supports were crushed and sieved to 0.3–1.0 mm diameter. The catalysts were prepared by an impregnation method using nickel (II) nitrate hexahydrate and ammonium heptamolybdate tetrahydrate as the corresponding metal salt precursors. After impregnation, the resultant samples were dried at 110 °C overnight and then calcined at 500 °C for 5 h with a ramping rate of 5 °C min−1 under air to obtain NiMo supported catalysts with Mo and Ni loading of 9.4 wt.% and 2.4 wt.%, respectively. The amount of metal was loaded according to our previous work [7].

2.4. Catalysts Characterization

The X-ray diffraction (XRD) patterns of prepared samples were obtained using an X-ray diffractometer (Bruker D8 Advance, Karlsruhe, Germany). The experiments were performed with CuKα radiation and over the 2θ ranges from 20° to 80°.
The specific surface area, total pore volume, and mean pore size diameter were determined from the N2 physisorption using a BELSORP mini II device (BELJapan Inc., Osaka, Japan). Prior to the experiments, the moisture on the catalyst surface was removed by a pretreatment system at 220 °C for 3 h under a helium flow rate of 50 cm3 min−1. The average pore size of samples was calculated from the desorption branch of isotherm by the BJH method.
The reduction behavior and reducibility of catalysts were measured by the H2-temperature-programmed reduction (TPR) using Autochem 2910 (Micromeritics, Norcross, GA, USA) with 10% H2 in N2. Prior to the experiments, the samples were pretreated at 150 °C for 2 h with a ramping rate of 10 °C min−1 under 30 cm3 min−1 of a N2 flow. The hydrogen consumption was analyzed using a thermal conductivity detector (TCD).

2.5. Catalytic Hydrotreating of Extracted Oil from Pongamia pinnata Seeds

The catalytic hydrotreating of extracted oil was carried out in a continuous fixed-bed reactor with an internal diameter of 0.9 cm. The catalyst (3.5 g) was loaded into the reactor and presulfided using a mixture of 3 wt.% carbon disulfide in n-heptane. The presulfidation was conducted under a H2 flow rate of 100 cm3 min−1 with a pressure of 40 bar. The temperature was increased from 30 to 400 °C with a ramp rate of 10 °C min−1. Subsequently, the heptane/carbon disulfide solution was supplied to the reactor with a flow rate of 0.17 cm3 min−1 for 3 h. In the catalytic reaction test, the reactor was heated to 330 °C and pressurized with H2 to 50 bar controlled by a back-pressure regulator. A high-pressure pump was used to introduce the extracted oil feed, while the H2 feed was controlled by a mass flow controller. After sulfidation, the catalysts were stabilized by flowing the feed through the reactor for 9 h before actual experiments were implemented. The conditions of the hydrotreating were temperature of 330 °C, H2 pressure of 50 bar, WHSV of 1.5 h−1, and H2/oil ratio of 1000 cm3/cm3, which were chosen based on our previous research [32]. The liquid product was collected for analysis. The reaction test was conducted by using fresh catalysts to study the effect of catalyst deactivation during the experiments, and no extra sulfur was supplied during the reaction for all experiments.

2.6. Analysis of the Extracted Oil from Pongamia pinnata Seeds and Hydrotreating Products

The percentage of fatty acid content in extracted oil was determined following the ASTM D664. The fatty acid composition of the extracted oil and the product identifications were confirmed by gas chromatography-mass spectroscopy analysis equipped with a capillary column (TR-FAME column, 30 m × 0.25 m × 0.25 µm). The elemental analysis including carbon, hydrogen, nitrogen, and sulfur in the extracted oil was characterized by CHNS/O analysis using a Thermo Scientific Flash 2000 elemental analyzer (Thermo Fisher Scientific, Waltham, MA, USA).
The gaseous products were analyzed by gas chromatography with a thermal conductivity detector (TCD) with a Porapak Q column (30 m × 0.25 mm × 0.25 µm). The injection and detection temperatures were maintained at 150 °C. The oven temperature was increased from 40 to 100 °C with a ramping rate of 20 °C min−1 and held at 100 °C for 11 min. The liquid products were analyzed by using gas chromatography (Shimadzu GC-14B, Kyoto, Japan) equipped with a flame ionization detector (FID) and a DB-2887 column (10 m × 0.53 mm × 3.00 µm). Simulated distillation (according to ASTM D2887) [33,34] based on the relationship between retention times and boiling ranges was employed to determine the boiling range distribution of the liquid products. The injector and detector temperatures were maintained at 350 °C and 320 °C, respectively. The temperature program was increased from 40 to 340 °C with a ramping rate of 15 °C min−1 and held at 340 °C for 20 min. The fatty acid and triglycerides conversion, gasoline selectivity, and diesel selectivity were calculated from simulated distillation data using the following equations:
Conversion   ( % ) = Feed FA + TG     Product FA + TG Feed FA + TG   ×   100
where
Feed FA + TG   is the weight percent of fatty acid and triglyceride in the feed.
Product FA + TG is the weight percent of fatty acid and triglyceride in the product.
Gasoline   selectivity   ( % ) =   Product C 5 C 11       Feed C 5 C 11   Feed FA + TG     Product FA + TG
Diesel   selectivity   ( % ) = Product C 12 C 22       Feed C 12 C 22 Feed FA + TG     Product FA + TG
where
Feed C 5 C 11 is the weight percent of hydrocarbons at a boiling point in the range of C5–C11 in the feed.
Product C 5 C 11   is the weight percent of hydrocarbons at a boiling point in the range of C5–C11 in the product.
Feed C 12 C 22 is the weight percent of hydrocarbons at a boiling point in the range of C12–C22 in the feed.
Product C 12 C 22   is the weight percent of hydrocarbons at a boiling point in the range of C12–C22 in the product.
Oil   phase   fraction = weight   of   oil   phase   product   weight   of   feed
Gasoline and diesel yield can be calculated by:
Gasoline yield = oil phase fraction × conversion × gasoline selectivity
Diesel yield = oil phase fraction × conversion × diesel selectivity
Yield of liquid fuel = diesel yield + gasoline yield
The functional group of the Pongamia pinnata oil and liquid products was identified by the Fourier transform infrared (FTIR) spectroscopy (Perkin Elmer spectrum GXFT-IR).

3. Results

3.1. Extraction of Pongamia pinnata Oil and Characterization of Oil Feedstock

From the extraction of oil using hexane as a solvent, the yield of the extracted oil from Pongamia pinnata seed was approximately 26% based on the dry weight of Pongamia pinnata seeds. Free fatty acids contained in the Pongamia pinnata oil were 5.7 wt.%. The fatty acid composition of Pongamia pinnata oil and refined palm olein is summarized in Table 1. The major compositions of the fatty acids (>85 wt.%) are oleic acid, palmitic acid, and linoleic acid, making them promising for diesel application.
The elements contained in the extracted oil were characterized by CHNS-O analysis as shown in Table 2. It should be noted that the extracted oil before the hydrotreating process contained high oxygen content due to the major triglyceride and free fatty acid composition of Pongamia pinnata oil.

3.2. Catalyst Characterization

The N2 adsorption–desorption measurements were conducted to examine the textural properties of the catalysts. Table 3 summarizes the BET specific surface area, pore volume, and pore diameter of the bare supports and synthesized samples.
The surface area decreased in the following order: NiMo/SiO2 > NiMo/γ-Al2O3 > NiMo/TiO2 which was in a similar trend corresponding with their bare supports. The surface area, total pore volume, and mean pore diameter slightly decreased after loading with NiMo for all supports. The adsorption–desorption isotherms of the prepared samples are shown in Figure 1. It was found that the calcined NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 exhibited type IV isotherm, indicating the characteristics of a mesoporous structure. This result is consistent with the pore sizes of all the catalysts observed by the BET, which were in the range of 5–13 nm. The hysteresis loop consists of adsorption and desorption branches depending on the pore shape of the catalyst.
The phase identification and crystallinity of the synthesized catalyst were analyzed through XRD. The XRD patterns of the calcined, sulfided, and spent NiMo on γ-Al2O3, SiO2, and TiO2 catalysts are represented in Figure 2a–c. The peaks at 2θ = 32.60°, 38.49°, 56.7°, 61.37°, and 67.94° corresponding to the (220), (222), (400), (511), and (440) phase assigned to γ-Al2O3 were observed for the calcined and sulfided samples as shown in Figure 2a [10,11,35]. Besides, the XRD patterns of NiMo supported on TiO2 samples exhibited the rutile phase peak at 36° (101) and 55° (211) and anatase phase peak at 25° (101) and 48° (200), suggesting the mixed-phase of TiO2 in this present experiment [8,36].
Figure 2b shows the XRD patterns of the calcined samples. It was confirmed that the MoO2 and Mo4O11 phases were observed in the NiMo/SiO2 and NiMo/TiO2 samples [37]. The XRD pattern of NiO [38,39] was found in NiMo/SiO2. However, the MoO3 peak could not be clearly detected for all the catalysts [40]. After the presulfidation treatment, the characteristic peaks of MoS2 were obviously observed over NiMo/SiO2 (Figure 3) at 33.5°and 58° corresponding to the (100) and (110) phase, respectively [41,42,43]. On the other hand, the peaks of MoS2 could not be observed in NiMo/γ-Al2O3 catalyst because this peak overlaps with the support. In the case of the NiMo/TiO2 catalyst, the XRD pattern showed highly dispersed MoS2 on TiO2 support.
The H2-TPR experiments were employed to investigate the reducibility and metal-support interaction of the prepared samples. The H2-TPR profiles of the samples are shown in Figure 3. The main reduction peak was located between 300 and 450 °C which might correspond to the simultaneous reduction of NiO and NiMoO4 [44,45]. The TiO2-supported catalyst exhibited a weak metal–support interaction which was the most facile to reduce (reduction temperature of 350 °C). Meanwhile, the reduction peaks of γ-Al2O3 and SiO2-supported catalysts were shifted to higher reduction temperature probably due to the highly dispersed smaller-size metals on the γ-Al2O3 and SiO2 support [46], leading to the strong interaction between metal and support compared to the TiO2. The appearance of shoulder reduction peaks between 500 and 600 °C was attributed to the several reduction steps of MoO3, which could be involved in the reduction of MoO3 to MoO2 and finally MoO2 to Mo [44,45,47]. The high reduction temperatures above 650 °C were assigned to the reduction of all highly dispersed Mo species. As a result, NiMo/γ-Al2O3 was the most difficult one to reduce which could reach full reduction at higher 800 °C, probably due to the NiAl2O4 and Al2(MoO4)3 formation [44].

3.3. Catalytic Hydrotreating of Pongamia pinnata Oil

The triglycerides and fatty acids conversion, diesel selectivity, and gasoline selectivity over the sulfide NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts after stabilizing the reaction for 9 h are shown in Figure 4. The conversions of Pongamia pinnata oil of the sulfide NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts were 97.2%, 52.5%, and 84.4%, respectively. Moreover, the diesel selectivity was in a range of 95–99% for all the catalysts. The higher diesel selectivity was obtained for the sulfided NiMo/γ-Al2O3 and NiMo/TiO2 catalysts compared with the sulfided NiMo/SiO2 catalyst. Besides, the gasoline selectivity was slightly higher for the sulfided NiMo/SiO2 catalyst (approximately 5%), suggesting the cracking reaction to lighter hydrocarbons (mixture of n-C5 to n-C11) occurred over the catalyst.
Generally, the catalytic hydrotreating process could convert the triglycerides and fatty acids into a mixture of straight-chain alkanes. The oxygen atoms in the triglycerides and fatty acids were eliminated in the forms of CO, CO2, and H2O via decarbonylation, decarboxylation, and hydrodeoxygenation, respectively. The loss of carbon atoms by CO and CO2 generation during deoxygenation can cause more mass loss of the yield of the liquid product than that of O removal via hydrodeoxygenation.
As seen in Figure 5, the maximum theoretical yields of deoxygenated products were calculated in order to investigate the catalyst activity by assuming no side reaction occurrs. The values would be different depending on the type of starting materials and the reaction pathways (DCO/DCO2 and HDO). In this study, the maximum theoretical yields were calculated based on the compositions of 94.3 wt.% of triglyceride and 5.7 wt.% of fatty acid contained in the extracted Pongamia pinnata oil, which were 80.5 wt.% for DCO/DCO2 and 85.2 wt.% for HDO. The desired product yield (diesel + gasoline) of the sulfided NiMo/γ-Al2O3 catalyst almost reached the theoretical value comparable to the other two catalysts, due to its high deoxygenating performance and low cracking activity.
The hydrocarbon distribution in the liquid product of sulfided NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts is summarized in Table 4. n-C17 and n-C18 hydrocarbons are predominant hydrocarbon compositions corresponding to oleic acid, the main feed composition. Higher n-C18 (>50 wt.%) was observed over the γ-Al2O3 and TiO2 support catalysts, suggesting that HDO reaction is preferred for these two catalysts. On the other hand, DCO and DCO2 pathways were enhanced over the sulfided NiMo/SiO2 catalyst as evidenced by the high amount of n-C17 (>60 wt.%) in the liquid product. This result can also explain the highest gas yield produced over the SiO2-support catalyst due to the elimination of oxygen atoms in triglycerides and fatty acids in the forms of CO and CO2.
Considering the hydrocarbon distribution above, the DCO/DCO2 and HDO reactions occurred simultaneously over these three catalysts but with different selectivities. As obviously seen in Figure 6, the ratio of n-Cn-1/n-Cn in the liquid product was examined to represent the selectivity of DCO/DCO2 and HDO. The value of n-Cn-1/n-Cn decreased in the following order: NiMo/SiO2 > NiMo/γ-Al2O3 > NiMo/TiO2. For the γ-Al2O3- and TiO2-supported catalysts, low ratio values (less than 0.5) were achieved, while the highest value of almost 6 was obtained over the SiO2-supported catalyst. This can be described by the influence of Ni and Mo contents. Typically, Ni catalyst is favorable for DCO/DCO2 pathway and also enhances the cracking reaction, higher Ni offers higher DCO/DCO2 activity, which leads to the formation of Cn-1 hydrocarbon corresponding to the carbon contained in the feedstock. Meanwhile, the activity of the HDO reaction would be enhanced by introducing Mo species [44,48,49]. Imai et al. [49] investigated the effect of Ni, Mo, and NiMo supported on alumina for the hydrotreating of methyl laurate. They found that the product distributions were different over Ni, Mo, and NiMo catalysts. However, a sulfided NiMo/γ-Al2O3 catalyst exhibited the highest catalytic activity due to the synergistic effects between Ni and Mo species. In addition, the NiMo/γ-Al2O3 with high Mo content could enhance the conversion and the HDO reaction, as well as suppress the hydrocracking and methanation reaction.
In our investigation, the reaction was carried out with the same amount of Ni (2.4 wt.%) and Mo (9.4 wt.%) loadings; however, the product distributions were rather different. This phenomenon is probably caused by the various interactions between the supports and the metal-oxide/sulfide species. As previously stated, HDO selectivity is preferable over the MoS2 active site; however, as seen from the TPR (Figure 3), there was the formation of different types of metal oxide phase which could affect the MoS2 slab formation during the sulfidation treatment, resulting in distinctions in HDO and DCO/DCO2 selectivities, which have been extensively studied [50,51,52,53]. According to the literature, the most active phase, Type II, which is highly stacked and totally sulfided, can be formed at a weak metal–support interaction, while Type I, which is less stacked and partially sulfided, can be formed by strong interaction with the support. Based on our results, the TiO2-supported catalyst exhibited less metal–support interaction, resulting in more MoS2 active phase formation which leads to relatively high HDO activity compared to the other two catalysts. Meanwhile, the DCO/DCO2 selectivity was highly active over the SiO2-supported catalyst which could be ascribed to the MoS2 suppression by the strong metal–support interaction.
Interestingly, the SiO2-supported catalyst contains a relatively high surface area (see Table 1) which could be beneficial to catalytic activity by increasing the exposure of the active sites [46]. However, the conversion of triglyceride and fatty acid over the SiO2-supported catalyst is relatively low compared to the value obtained from the γ-Al2O3 and TiO2-supported catalysts. This result may be caused by the specific properties of the supporting materials. The deoxygenation activity also involves an acidity property of the catalyst. Mild acidity is required for the high deoxygenation activity, due to its favorable adsorption of triglycerides, fatty acid, and other oxygen-containing molecules [54,55,56], whereas the strong acidity promotes unwanted hydrocracking and increases coke formation, leading to catalyst deactivation [57]. There have been several reports indicating that high acidity support has a beneficial effect on hydrodeoxygenation activity. Kumar et al. [38] studied the impact of Ni supported on γ-Al2O3, SiO2, and HSZM-5 zeolite on stearic acid deoxygenation. They observed that utilizing the HSZM-5 catalyst resulted in the maximum conversion, which was attributed to the relatively strong acidity of the HSZM-5 catalyst compared to that of γ-Al2O3 and SiO2. This consequence was also observed by Peng et al. [58]. They studied the deoxygenation of palmitic acid with Ni supported over γ-Al2O3, SiO2, and ZrO2 and determined that conversion proceeded in the following order: ZrO2 > γ-Al2O3 > SiO2 due to the greater acidity of the ZrO2 support. According to the literature, infrared spectroscopy of adsorbed CO [59] and NH3-temperature-programmed desorption [60] were used to study the acidity of SiO2. There were no substantial numbers of acid sites for the SiO2 support. While the three varieties of Lewis acid sites (weak, medium, and strong) were found on γ-Al2O3 support [57,59,61], the medium-weak forms of Lewis acid were found on anatase TiO2 support [54,55]. Taking into account all of these remarks, the XRD results revealed that the TiO2 employed in this work had a mixed-phase (anatase + rutile); consequently, the acidity of the mixed-phase TiO2 was apparently lower than that stated in the literature. It is worth noting that bare Al2O3 and TiO2 have been reported to exhibit conversion and hydrodeoxygenation activity due to their acid-based characteristics; however, their activity is much higher once loaded with active metal [62,63,64,65]. While bare SiO2 poses very low acidity (inertness) and no significant activity has been observed when employing bare SiO2.
As a result, it might be speculated that the acidity of the catalyst increased in the following order: NiMo/SiO2 < NiMo/TiO2 < NiMo/γ-Al2O3, which corresponds to our conversion (NiMo/SiO2 < NiMo/TiO2 < NiMo/γ-Al2O3). Consequently, we hypothesize that the observed disparities in reaction pathways are not only driven by the type of active phase but also the acidity of the supporting materials.
Figure 7 shows the gas product compositions which mainly consist of CO, CO2, and CH4. The CO and CO2 could be generated via DCO and DCO2, whereas CH4 could be produced from the cracking of products in the liquid phase and also a methanation reaction in the gas phase. Based on the equilibrium constant for reverse water gas shift (RWGS) reaction, the calculated value at 330 °C is approximately 0.04, suggesting that the RWGS can be neglected to convert CO2 to CO. Hence, CO would be a primary gas product which is mainly generated from the oxygen removal of triglycerides and fatty acid via DCO. Therefore, the level of CO and CO2 content can elucidate the activity of DCO and DCO2 reactions of the three catalysts [8]. By comparison, the CO level was higher than the CO2, indicating that DCO has higher activity than DCO2 for all the catalysts. Due to the low methane percentage in the gaseous product, the methanation process might be minimal. However, a small amount of methane was generated when the NiMo/γ-Al2O3 catalyst was in use. This might be attributed to the high acidity of γ-Al2O3 support, which enhances the adsorption of triglyceride and fatty acid molecules in feedstock, potentially accelerating methane generation. This finding might support our previous assumption concerning the acidity of the catalysts. It is worth noting that the amount of methane from NiMoS of all these supports is much less pronounced than single Ni [66] or single Mo [67] in various active forms.
In order to confirm the elimination of oxygen atoms in triglycerides and fatty acids, the functional groups of the liquid products were characterized by Fourier transform infrared (FTIR) spectroscopy (see Figure 8). The absorption peak at 1704 cm−1 in the FTIR spectrum corresponds to the carboxylic functional groups of fatty acids. On the other hand, the absorption peaks at 1746 and 1164 cm−1 in the FTIR spectrum are assigned to the carbonyl groups (–C=O stretch) and ester groups (–C–O– stretch), respectively [41]. In the hydrotreating process, triglyceride was firstly hydrogenolysed to fatty acids and further deoxygenated to n-alkane. The FTIR spectrum of Pongamia pinnata oil exhibited the absorption peaks corresponding to triglyceride and fatty acids. Considering the liquid product after the hydrotreating process, the peaks intensity of triglyceride and fatty acid in the FTIR spectrum significantly decreased over sulfided NiMo supported on γ-Al2O3 and TiO2 catalysts. This result suggested that the triglyceride and fatty acid were almost completely converted to n-alkanes. On the other hand, the peak intensity of fatty acid was detected in the liquid product when the reaction was catalyzed by the sulfided NiMo/SiO2 catalyst.
The results of the elemental composition (CHNS/O) of extracted Pongamia pinnata oil and liquid products are summarized in Table 2. Thereafter the deoxygenation activity, and percent of carbon atom content in liquid product increased with decreasing oxygen concentration. The deoxygenation degree calculated based on the oxygen content of feedstock and liquid product is increased as follows: NiMo/SiO2 (37.7%) < NiMo/γ-Al2O3 (94.6%) < NiMo/TiO2 (98.3%). These results are attributed to the various intensities of metal-support interaction, as mentioned earlier. However, the relatively high catalytic performance of the NiMo/γ-Al2O3 compared to NiMo/TiO2 may be due to its higher surface area and greater mesopore size.
The H/C and O/C ratios were shown on the Van Krevelen diagram (Figure 9) to compare the elemental compositions of hydrotreated liquid products and extracted oil with petroleum fuel and hydrotreated vegetable oil (HVO) from Neste Oil. Pongamia pinnata oil had the highest O/C ratio of 0.08, which is consistent with the high amount of oxygen-containing molecules (triglycerides and fatty acid) content. In comparison to the starting feedstock, the hydrotreated products of the three catalysts displayed greater H/C and lower O/C values, particularly for the γ-Al2O3 and TiO2-supported catalysts, which were close to the value of petroleum-based oil (H/C = 1.6–1.8, O/C almost nil) [68,69]. Furthermore, the H/C value of NiMo/γ-Al2O3 (H/C = 2.17) is comparable to the value of HVO generated from Neste oil (H/C = 2.15) [70]. The SiO2-supported catalyst had a much higher O/C value than the other catalysts, which could be attributed to poor deoxygenation activity and strong cracking activity, as evidenced by the high gas and short hydrocarbon fraction generated. While a high H/C value ratio suggested the production of a liquid product with a high hydrogen content, this might be attributed to high hydrogenation and low cyclization/aromatization activities.
Surprisingly, slightly higher amounts of sulfur content were detected in the hydrotreated liquid products. They were 2.65%, 2.07%, and 2.56% for the NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts, respectively (Table 2). It is even higher than the value detected in the extracted Pongamia pinnata oil (2.16%). This result could demonstrate that there was some sulfur leaching from the NiMoS catalysts during the reaction, as the amount of sulfur content decreased after the reaction process (Table 5). In addition, the loss of sulfur in the spent catalysts, on the other hand, could be accounted for by the oxidation of the sulfide phase, which would be incorporated by oxygen atoms at the edge sulfur atoms in the MoS2 during the reaction.

3.4. Catalyst Stability and Deactivation

The catalytic stability of the NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts was evaluated through 15 h time-on-stream using extracted Pongamia pinnata oil at a temperature of 330 °C, H2 pressure of 50 bar, WHSV of 1.5 h−1, and H2/oil ratio of 1000 cm3/cm3. As can be seen in Figure 10a, The NiMo/γ-Al2O3 exhibited the highest conversion and stability for 15 h, meanwhile, the NiMo/SiO2 and NiMo/TiO2 catalysts gradually decreased in conversion. On the other hand, unexpectedly, the NiMo/γ-Al2O3 and NiMo/SiO2 catalysts showed good catalytic activity (100% conversion) and remained stable for 42 h when using refined palm olein as feedstock (Figure 10b). This may be caused by the impurities in the Pongamia pinnata oil (Table 6) which were quantified using the inductively coupled plasma (ICP) technique. According to Kubicka and Horacek [71], an abundance of alkali content in feedstock may have a deleterious influence on the catalyst. They discovered that the presence of alkalis (Ca, K, Na, and Mg) in waste rapeseed oil increased the rate of CoMoS/γ-Al2O3 catalyst deactivation due to their deposition on the catalyst surface, resulting in active site blockage/poisoning. Similarly, Arora et al. [72] studied the effect of potassium-containing feedstock and discovered that when the quantity of potassium is high, the catalyst activity is inhibited. This is due to potassium atoms preferentially adsorbed on the edge vacancy sites of the MoS2 slabs.
The effect of potassium doping on MoS2 for CO hydrogenation has also been studied using DFT calculation by Andersen et al. [73]. Their findings revealed that doping K on MoS2 enhances surface basicity owing to increased electron charge, and also blocks the Mo and S edge, inhibiting CO adsorption and limiting hydrogen accessing the MoS2 surface. In agreement with May et al. [74,75], they found that modifying the CoMoS/Al2O3 commercial catalyst with potassium resulted in decreased hydrogenation activity for a synthetic FCC gasoline, due to changing electronic properties of sulfide phase and also decreasing the acid site of support. Furthermore, calcium has been correlated to a reduction in catalyst activity, particularly HDS and HDO activities, as well as a minor decrease in hydrogenation activity [76]. Comparing these results between using Pongamia pinnata oil and refined palm olein, it is presumed that the suppression on catalytic activity is considerably more pronounced with NiMo/SiO2 than with NiMo/γ-Al2O3 and NiMo/TiO2. These results were probably caused by the partially facilitated by the γ-Al2O3 and TiO2 and catalyzed by their acid sites. Furthermore, the prolonged activity of NiMo/γ-Al2O3 for Pongamia pinnate oil was tested, which revealed that the catalyst activity dropped after 75 h, while the stability of the catalyst for refined palm olein was still longer, over 75 h.
Taking into account the presence of phosphorus (P) in feedstock, it was reported to promote oligomerization/polymerization reactions, resulting in increased formation of high molecular weight compounds, eventually leading to increased carbonaceous deposition on the surface of the catalyst and rapid catalyst deactivation [71,78]. As can be seen in Figure 11, significantly greater carbonaceous deposition was discovered during the hydrotreating of Pongamia pinnata oil than in the hydrotreating of refined palm olein in our study. TGA was performed in atmospheric air to quantify the formation of carbonaceous species on the surface of catalysts. The first weight loss at temperatures below 200 °C can be ascribed to evaporation of moisture adsorbed, whereas temperatures between 300 and 450 °C can be attributed to volatile substance decomposition. At higher temperatures ranging from 450 to 800 °C, it is ascribed to the decomposition of various carbonaceous species including heavy hydrocarbons accumulated inside the catalyst pores or more stable amorphous coke (450–500 °C), as well as the oxidation of graphite and graphene carbon (600–800 °C) [79,80]. The decomposition of a substrate at a temperature above 800 °C could be related to the oxidation of inorganic carbon (i.e., CaCO3 [80] and K2CO3 [81]). Evidently, larger weight losses were detected in the catalyst employed for Pongamia pinnata oil, especially when compared to the SiO2- and TiO2-supported catalysts, showing that the presence of phosphorus content might accelerate the rate of coke formation.
Ultimately, it is possible to conclude that differences in catalytic performance are caused not only by different characteristics of metals and supporting materials but also by the purity of feedstock, which appears to have a significant impact on declining catalytic activity and catalyst deactivation. Catalyst deactivation during 42 h of time-on-stream when using refined feed (refined palm olein) could be negligible for all three catalysts. However, the difference in deactivation was more pronounced when being tested under unrefined with high metal content, e.g., P. pintnata oil. Though tested under unrefined is found to be an aggressive condition for coking, it is likely that catalyst poisoning and coking is not the main cause of the deactivation.

4. Conclusions

The Pongamia pinnata oil was extracted using Soxhlet extraction with hexane as the solvent. On a dry basis, the oil production was approximately 26 wt.%. Oleic acid, palmitic acid, and linoleic acid were the key fatty acid compositions (>85 wt.%), which made them attractive for biodiesel production. The bio-hydrotreated diesel was successfully produced through the hydrotreating of Pongamia pinnata oil in a fixed-bed reactor. To investigate the effect of different supporting materials, γ-Al2O3, SiO2, and TiO2 were all loaded with the same quantity of Ni-Mo. In terms of conversion and diesel yield, the hydrotreating activity increased in the following order: NiMo/γ-Al2O3 > NiMo/TiO2 > NiMo/SiO2. The metal–support interaction of each catalyst results in the activity and selectivity pathway due to the difference of MoS2 active phase formation. The SiO2-supported catalyst shows low activity with a high DCO selective pathway, leading to the formation of n-C17 hydrocarbon as the main product. The HDO pathway is suppressed as a result of the strong metal–support interaction, which has a deleterious influence on the formation of the MoS2 active phase. While the γ-Al2O3- and TiO2-supported catalysts had a higher selectivity for the HDO pathway, which produces n-C18 hydrocarbon as the main product. This was probably owing to the greater number of MoS2 active sites in weak supporting interaction. The best performance of NiMo/γ-Al2O3 may be attributed to the synergistic effect of pore size, acidity, and support interaction. The long-term stability of the three catalysts in Pongamia pinnata oil was investigated compared to that in refined palm olein. In refined palm olein, 100% conversion was obtained using NiMo/γ-Al2O3 and NiMo/SiO2. This is due to the presence of phosphorus and alkali impurities in Pongamia pinnata oil, which could accelerate the formation of coke and poisoning or blocking of the active site. As a result, excellent catalytic performance depends not only on the properties of catalyst materials but also on the quality of the feedstock.

Author Contributions

Conceptualization, W.K.; methodology, validation, formal analysis, investigation, data curation, writing—original draft preparation, Y.P., W.L. and K.P.; writing—review and editing, K.P., W.K., A.S., J.W.L., P.R., K.K., A.E.-u. and S.A.; supervision, W.K. and S.A.; funding acquisition, W.K. and S.A. All authors have read and agreed to the published version of the manuscript.

Funding

The financial support from Reinventing University System Program by the Ministry of Higher Education, Science, Research and Innovation, Thailand (Fiscal Year 2021); the Thailand 4.0 Innovation Hub—Bioenergy for building an innovation-based economy, Council of University Presidents of Thailand which is allocated to Chulalongkorn University and Silpakorn University is gratefully acknowledged. The authors also would like to acknowledge the Research Chair Grant supported by the National Science and Technology Development Agency (NSTDA).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Soytas, U.; Sari, R.; Ewing, B.T. Energy consumption, income, and carbon emissions in the United States. Ecol. Econ. 2007, 62, 482–489. [Google Scholar] [CrossRef]
  2. Lau, L.C.; Lee, K.T.; Mohamed, A.R. Global warming mitigation and renewable energy policy development from the Kyoto Protocol to the Accord—A comment. Renew. Sustain. Energy Rev. 2012, 16, 5280–5284. [Google Scholar] [CrossRef]
  3. Kim, S.K.; Brand, S.; Lee, H.S.; Kim, Y.; Kim, J. Production of renewable diesel by hydrotreatment of soybean oil: Effect of reaction parameters. Chem. Eng. J. 2013, 228, 114–123. [Google Scholar] [CrossRef]
  4. Priecel, P.; Kubička, D.; Čapek, L.; Bastl, Z.; Ryšánek, P. The role of Ni species in the deoxygenation of rapeseed oil over NiMo-alumina catalysts. Appl. Catal. A Gen. 2011, 397, 127–137. [Google Scholar] [CrossRef]
  5. Onyestyák, G.; Harnos, S.; Szegedi, Á.; Kalló, D. Sunflower oil to green diesel over Raney-type Ni-catalyst. Fuel 2012, 102, 282–288. [Google Scholar] [CrossRef]
  6. Guzman, A.; Torres, J.E.; Prada, L.P.; Nuñez, M.L. Hydroprocessing of crude palm oil at pilot plant scale. Catal. Today 2010, 156, 38–43. [Google Scholar] [CrossRef]
  7. Kiatkittipong, W.; Phimsen, S.; Kiatkittipong, K.; Wongsakulphasatch, S.; Laosiripojana, N.; Assabumrungrat, S. Diesel-like hydrocarbon production from hydroprocessing of relevant refining palm oil. Fuel Process. Technol. 2013, 116, 16–26. [Google Scholar] [CrossRef]
  8. Pongsiriyakul, K.; Kiatkittipong, W.; Kiatkittipong, K.; Laosiripojana, N.; Faungnawakij, K.; Adhikari, S.; Assabumrungrat, S. Alternative hydrocarbon biofuel production via hydrotreating under a synthesis gas atmosphere. Energy Fuels 2017, 31, 12256–12262. [Google Scholar] [CrossRef]
  9. Yenumala, S.R.; Maity, S.K.; Shee, D. Hydrodeoxygenation of karanja oil over supported nickel catalysts: Influence of support and nickel loading. Catal. Sci. Technol. 2016, 6, 3156–3165. [Google Scholar] [CrossRef]
  10. Gong, S.; Shinozaki, A.; Shi, M.; Qian, E.W. Hydrotreating of jatropha oil over alumina based catalysts. Energy Fuels 2012, 26, 2394–2399. [Google Scholar] [CrossRef]
  11. Toba, M.; Abe, Y.; Kuramochi, H.; Osako, M.; Mochizuki, T.; Yoshimura, Y. Hydrodeoxygenation of waste vegetable oil over sulfide catalysts. Catal. Today 2011, 164, 533–537. [Google Scholar] [CrossRef]
  12. Bezergianni, S.; Dimitriadis, A.; Kalogianni, A.; Knudsen, K.G. Toward hydrotreating of waste cooking oil for biodiesel production. Effect of pressure, H2/oil ratio, and liquid hourly space velocity. Ind. Eng. Chem. Res. 2011, 50, 3874–3879. [Google Scholar] [CrossRef]
  13. Loe, R.; Santillan-jimenez, E.; Morgan, T.; Sewell, L.; Ji, Y.; Jones, S.; Isaacs, M.A.; Lee, A.F.; Crocker, M. Effect of Cu and Sn promotion on the catalytic deoxygenation of model and algal lipids to fuel-like hydrocarbons over supported Ni catalysts. Appl. Catal. B Environ. 2016, 191, 147–156. [Google Scholar] [CrossRef]
  14. Mohibbeazam, M.; Waris, A.; Nahar, N. Prospects and potential of fatty acid methyl esters of some non-traditional seed oils for use as biodiesel in India. Biomass Bioenergy 2005, 29, 293–302. [Google Scholar] [CrossRef]
  15. Demirbas, A. Potential resources of non-edible oils for biodiesel. Energy Sources Part B Econ. Plan. Policy 2009, 4, 310–314. [Google Scholar] [CrossRef]
  16. Sarin, A.; Arora, R.; Singh, N.P.; Sarin, R.; Sharma, M.; Malhotra, R.K. Effect of metal contaminants and antioxidants on the oxidation stability of the methyl ester of Pongamia. J. Am. Oil Chem. Soc. 2010, 87, 567–572. [Google Scholar] [CrossRef]
  17. Islam, A.K.M.A.; Chakrabarty, S.; Yaakob, Z.; Ahiduzzaman, M.; Islam, A.K.M.M. Koroch (Pongamia pinnata): A promising unexploited resources for the tropics and subtropics. In Forest Biomass – From Trees to Energy; Gonçalves, A.C., Sousa, A., Malico, I., Eds.; IntechOpen: Rijeka, Hrvatska, 2021; pp. 1–19. ISBN 978-1-83962-971-6. [Google Scholar]
  18. Bobade, S.; Khyade, V. Preparation of methyl ester (biodiesel) from karanja (Pongamia pinnata) oil. Res. J. Chem. Sci. 2012, 2, 43–50. [Google Scholar]
  19. Hari, T.K.; Yaakob, Z. Production of diesel fuel by the hydrotreatment of jatropha oil derived fatty acid methyl esters over γ-Al2O3 and SiO2 supported NiCo bimetallic catalysts. React. Kinet. Mech. Catal. 2015, 116, 131–145. [Google Scholar] [CrossRef]
  20. Taromi, A.A.; Kaliaguine, S.; Afshar Taromi, A.; Kaliaguine, S. Green diesel production via continuous hydrotreatment of triglycerides over mesostructured γ-alumina supported NiMo/CoMo catalysts. Fuel Process. Technol. 2018, 171, 20–30. [Google Scholar] [CrossRef]
  21. Poddar, M.K.; Anand, M.; Farooqui, S.A.; Martin, G.J.O.; Maurya, M.R.; Sinha, A.K. Hydroprocessing of lipids extracted from marine microalgae Nannochloropsis sp. over sulfided CoMoP/Al2O3 catalyst. Biomass Bioenergy 2018, 119, 31–36. [Google Scholar] [CrossRef]
  22. Bezergianni, S.; Dimitriadis, A. Comparison between different types of renewable diesel. Renew. Sustain. Energy Rev. 2013, 21, 110–116. [Google Scholar] [CrossRef]
  23. Phimsen, S.; Kiatkittipong, W.; Yamada, H.; Tagawa, T.; Kiatkittipong, K.; Laosiripojana, N.; Assabumrungrat, S. Oil extracted from spent coffee grounds for bio-hydrotreated diesel production. Energy Convers. Manag. 2016, 126, 1028–1036. [Google Scholar] [CrossRef]
  24. Veriansyah, B.; Han, J.Y.; Kim, S.K.; Hong, S.A.; Kim, Y.J.; Lim, J.S.; Shu, Y.W.; Oh, S.G.; Kim, J. Production of renewable diesel by hydroprocessing of soybean oil: Effect of catalysts. Fuel 2012, 94, 578–585. [Google Scholar] [CrossRef]
  25. Liu, J.; Fan, K.; Tian, W.; Liu, C.; Rong, L. Hydroprocessing of Jatropha oil over NiMoCe/Al2O3 catalyst. Int. J. Hydrogen Energy 2012, 37, 17731–17737. [Google Scholar] [CrossRef]
  26. Krokidis, X.; Raybaud, P.; Gobichon, A.-E.; Rebours, B.; Euzen, P.; Toulhoat, H. Theoretical study of the dehydration process of boehmite to γ-Alumina. J. Phys. Chem. B 2001, 105, 5121–5130. [Google Scholar] [CrossRef]
  27. Gong, S.; Shinozaki, A.; Qian, E.W. Role of support in hydrotreatment of jatropha oil over sulfided NiMo catalysts. Ind. Eng. Chem. Res. 2012, 51, 13953–13960. [Google Scholar] [CrossRef]
  28. Wei, Z.B.; Yan, W.; Zhang, H.; Ren, T.; Xin, Q.; Li, Z. Hydrodesulfurization activity of NiMo/TiO2Al2O3 catalysts. Appl. Catal. A Gen. 1998, 167, 39–48. [Google Scholar] [CrossRef]
  29. Scott, P.T.; Pregelj, L.; Chen, N.; Hadler, J.S.; Djordjevic, M.A.; Gresshoff, P.M. Pongamia pinnata: An untapped resource for the biofuels industry of the future. BioEnergy Res. 2008, 1, 2–11. [Google Scholar] [CrossRef]
  30. Kaushik, N.; Mann, S.; Kumar, K. Pongamia pinnata: A candidate tree for biodiesel feedstock. Energy Sources Part A Recover. Util. Environ. Eff. 2015, 37, 1526–1533. [Google Scholar] [CrossRef]
  31. Panpraneecharoen, S.; Punsuvon, V. Biodiesel from crude Pongamia pinnata oil under subcritical methanol conditions with calcium methoxide catalyst. Energy Sources Part A Recover. Util. Environ. Eff. 2016, 38, 2225–2230. [Google Scholar] [CrossRef]
  32. Srifa, A.; Faungnawakij, K.; Itthibenchapong, V.; Viriya-empikul, N.; Charinpanitkul, T.; Assabumrungrat, S. Production of bio-hydrogenated diesel by catalytic hydrotreating of palm oil over NiMoS2/γ-Al2O3 catalyst. Bioresour. Technol. 2014, 158, 81–90. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, C.; Firor, R. Simulated Distillation System for ASTM D2887 Based on the Agilent 6890N GC; Agilent Technologies: Wilmington, DE, USA, 2005; pp. 1–10. [Google Scholar]
  34. ASTM D2887-97a; Standard Test Method for Boiling Range Distribution of Petroleum Fractions by Gas Chromatography. Annual Book of Standards Volume 05.02. Annual Book of Standards Volume 05.02; ASTM International: West Conshohocken, PA, USA, 2017.
  35. Palcheva, R.; Kaluza, L.; Spojakina, A.; Jiratova, K.; Tyuliev, G. NiMo/γ-Al2O3 catalysts from Ni heteropolyoxomolybdate and effect of alumina modification by B, Co, or Ni. Chin. J. Catal. 2012, 33, 952–961. [Google Scholar] [CrossRef]
  36. Thamaphat, K.; Limsuwan, P.; Ngotawornchai, B. Phase characterization of TiO2 powder by XRD and TEM. Kasetsart J. (Nat. Sci.) 2008, 42, 357–361. [Google Scholar]
  37. Wu, D.; Shen, R.; Yang, R.; Ji, W.; Jiang, M.; Ding, W.; Peng, L. Mixed molybdenum oxides with superior performances as an advanced anode material for lithium-ion batteries. Sci. Rep. 2017, 7, 44697. [Google Scholar] [CrossRef]
  38. Kumar, P.; Yenumala, S.R.; Maity, S.K.; Shee, D. Kinetics of hydrodeoxygenation of stearic acid using supported nickel catalysts: Effects of supports. Appl. Catal. A Gen. 2014, 471, 28–38. [Google Scholar] [CrossRef]
  39. Silvester, L.; Ipsakis, D.; Antzara, A.; Heracleous, E.; Lemonidou, A.A.; Bukur, D.B. Development of NiO-based oxygen carrier materials: Effect of support on redox behavior and carbon deposition in methane. Energy Fuels 2016, 30, 8597–8612. [Google Scholar] [CrossRef]
  40. Horiguchi, J.; Kobayashi, Y.; Kobayashi, S.; Yamazaki, Y.; Omata, K.; Nagao, D.; Konno, M.; Yamada, M. Mesoporous NiO–Al2O3 catalyst for high pressure partial oxidation of methane to syngas. Appl. Catal. A Gen. 2011, 392, 86–92. [Google Scholar] [CrossRef]
  41. Veeramalai, C.P.; Li, F.; Liu, Y.; Xu, Z.; Guo, T.; Kim, T.W. Enhanced field emission properties of molybdenum disulphide few layer nanosheets synthesized by hydrothermal method. Appl. Surf. Sci. 2016, 389, 1017–1022. [Google Scholar] [CrossRef]
  42. Yin, C.; Zhao, L.; Bai, Z.; Liu, H.; Liu, Y.; Liu, C. A novel porous ammonium nickel molybdate as the catalyst precursor towards deep hydrodesulfurization of gas oil. Fuel 2013, 107, 873–878. [Google Scholar] [CrossRef]
  43. Zhu, X.; Yang, C.; Xiao, F.; Wang, J.; Su, X. Synthesis of nano-TiO2-decorated MoS2 nanosheets for lithium ion batteries. New J. Chem. 2015, 39, 683–688. [Google Scholar] [CrossRef]
  44. Kumar, P.; Maity, S.K.; Shee, D. Role of NiMo alloy and Ni species in the performance of NiMo/Alumina catalysts for hydrodeoxygenation of stearic acid: A kinetic study. ACS Omega 2019, 4, 2833–2843. [Google Scholar] [CrossRef] [PubMed]
  45. Yeletsky, P.M.; Reina, T.R.; Bulavchenko, O.A.; Saraev, A.A.; Gerasimov, E.Y.; Zaikina, O.O.; Bermúdez, J.M.; Arcelus-Arrillaga, P.; Yakovlev, V.A.; Millan, M. Phenanthrene catalytic cracking in supercritical water: Effect of the reaction medium on NiMo/SiO2 catalysts. Catal. Today 2019, 329, 197–205. [Google Scholar] [CrossRef]
  46. Imai, H.; Abe, M.; Terasaka, K.; Suzuki, T.; Li, X.; Yokoi, T. Hydroconversion of methyl laurate over silica-supported Ni–Mo catalysts with different Ni sizes. Fuel Process. Technol. 2018, 180, 166–172. [Google Scholar] [CrossRef]
  47. Qu, L. MAS NMR, TPR, and TEM studies of the interaction of NiMo with alumina and silica–alumina supports. J. Catal. 2003, 215, 7–13. [Google Scholar] [CrossRef]
  48. Kubička, D.; Kaluža, L. Deoxygenation of vegetable oils over sulfided Ni, Mo and NiMo catalysts. Appl. Catal. A Gen. 2010, 372, 199–208. [Google Scholar] [CrossRef]
  49. Imai, H.; Kimura, T.; Terasaka, K.; Li, X.; Sakashita, K.; Asaoka, S.; Al-Khattaf, S.S. Hydroconversion of fatty acid derivative over supported Ni-Mo catalysts under low hydrogen pressure. Catal. Today 2018, 303, 185–190. [Google Scholar] [CrossRef]
  50. Liu, Z.; Han, W.; Hu, D.; Sun, S.; Hu, A.; Wang, Z.; Jia, Y.; Zhao, X.; Yang, Q. Effects of Ni–Al2O3 interaction on NiMo/Al2O3 hydrodesulfurization catalysts. J. Catal. 2020, 387, 62–72. [Google Scholar] [CrossRef]
  51. Ninh, T.K.T.; Massin, L.; Laurenti, D.; Vrinat, M. A new approach in the evaluation of the support effect for NiMo hydrodesulfurization catalysts. Appl. Catal. A Gen. 2011, 407, 29–39. [Google Scholar] [CrossRef]
  52. Candia, R.; Sorensen, O.; Topsøe, N.Y. Active phase in sulfided NiMo/γ-Al2O3 catalyst. Bull. Soc. Chim. Belg. 1984, 93, 763–773. [Google Scholar] [CrossRef]
  53. Stanislaus, A.; Marafi, A.; Rana, M.S. Recent advances in the science and technology of ultra low sulfur diesel (ULSD) production. Catal. Today 2010, 153, 1–68. [Google Scholar] [CrossRef]
  54. Lee, H.; Kim, H.; Yu, M.J.; Ko, C.H.; Jeon, J.-K.; Jae, J.; Park, S.H.; Jung, S.-C.; Park, Y.-K. Catalytic hydrodeoxygenation of bio-oil model compounds over Pt/HY catalyst. Sci. Rep. 2016, 6, 28765. [Google Scholar] [CrossRef] [PubMed]
  55. Benson, T.J.; Hernandez, R.; French, W.T.; Alley, E.G.; Holmes, W.E. Elucidation of the catalytic cracking pathway for unsaturated mono-, di-, and triacylglycerides on solid acid catalysts. J. Mol. Catal. A Chem. 2009, 303, 117–123. [Google Scholar] [CrossRef]
  56. Oi, L.E.; Choo, M.-Y.; Lee, H.V.; Taufiq-Yap, Y.H.; Cheng, C.K.; Juan, J.C. Catalytic deoxygenation of triolein to green fuel over mesoporous TiO2 aided by in situ hydrogen production. Int. J. Hydrogen Energy 2020, 45, 11605–11614. [Google Scholar] [CrossRef]
  57. Busca, G. Acid catalysts in industrial hydrocarbon chemistry. Chem. Rev. 2007, 107, 5366–5410. [Google Scholar] [CrossRef] [PubMed]
  58. Peng, B.; Zhao, C.; Kasakov, S.; Foraita, S.; Lercher, J.A. Manipulating catalytic pathways: Deoxygenation of palmitic acid on multifunctional catalysts. Chem. A Eur. J. 2013, 19, 4732–4741. [Google Scholar] [CrossRef]
  59. Hensen, E.J.M.; Poduval, D.G.; Degirmenci, V.; Ligthart, D.A.J.M.; Chen, W.; Maugé, F.; Rigutto, M.S.; van Veen, J.A.R. Acidity characterization of amorphous silica–alumina. J. Phys. Chem. C 2012, 116, 21416–21429. [Google Scholar] [CrossRef]
  60. Lomate, S.; Sultana, A.; Fujitani, T. Effect of SiO2 support properties on the performance of Cu–SiO2 catalysts for the hydrogenation of levulinic acid to gamma valerolactone using formic acid as a hydrogen source. Catal. Sci. Technol. 2017, 7, 3073–3083. [Google Scholar] [CrossRef]
  61. Chen, H.; Wang, Q.; Zhang, X.; Wang, L. Effect of support on the NiMo phase and its catalytic hydrodeoxygenation of triglycerides. Fuel 2015, 159, 430–435. [Google Scholar] [CrossRef]
  62. Vonghia, E.; Boocock, D.G.B.; Konar, S.K.; Leung, A. Pathways for the deoxygenation of triglycerides to aliphatic hydrocarbons over activated alumina. Energy Fuels 1995, 9, 1090–1096. [Google Scholar] [CrossRef]
  63. Bui, V.N.; Laurenti, D.; Delichère, P.; Geantet, C. Hydrodeoxygenation of guaiacol: Part II: Support effect for CoMoS catalysts on HDO activity and selectivity. Appl. Catal. B Environ. 2011, 101, 246–255. [Google Scholar] [CrossRef]
  64. Bai, X.; Duan, P.; Xu, Y.; Zhang, A.; Savage, P.E. Hydrothermal catalytic processing of pretreated algal oil: A catalyst screening study. Fuel 2014, 120, 141–149. [Google Scholar] [CrossRef]
  65. Laurent, E.; Delmon, B. Study of the hydrodeoxygenation of carbonyl, carboxylic and guaiacyl groups over sulfided CoMo/γ-Al2O3 and NiMo/γ-Al2O3 catalysts: I. Catalytic reaction schemes. Appl. Catal. A Gen. 1994, 109, 77–96. [Google Scholar] [CrossRef]
  66. Phimsen, S.; Kiatkittipong, W.; Yamada, H.; Tagawa, T.; Kiatkittipong, K.; Laosiripojana, N.; Assabumrungrat, S. Nickel sulfide, nickel phosphide and nickel carbide catalysts for bio-hydrotreated fuel production. Energy Convers. Manag. 2017, 151, 324–333. [Google Scholar] [CrossRef]
  67. Horáček, J.; Akhmetzyanova, U.; Skuhrovcová, L.; Tišler, Z.; de Paz Carmona, H. Alumina-supported MoNx, MoCx and MoPx catalysts for the hydrotreatment of rapeseed oil. Appl. Catal. B Environ. 2020, 263, 118328. [Google Scholar] [CrossRef]
  68. Alleman, T.L.; McCormick, R.L.; Christensen, E.D.; Fioroni, G.; Moriarty, K.; Yanowitz, J. Biodiesel Handling and Use Guide, 5th ed.; National Renewable Energy Lab (NREL): Golden, CO, USA, 2016; No. NREL/BK-5400-66521; DOE/GO-102016–4875. [Google Scholar]
  69. Rinaldi, R.; Schüth, F. Design of solid catalysts for the conversion of biomass. Energy Environ. Sci. 2009, 2, 610–626. [Google Scholar] [CrossRef]
  70. Aatola, H.; Larmi, M.; Sarjovaara, T.; Mikkonen, S. Hydrotreated vegetable Oil (HVO) as a renewable diesel fuel: Trade-off between NOx, particulate emission, and fuel consumption of a heavy duty engine. SAE Int. J. Engines 2009, 1, 1251–1262. [Google Scholar] [CrossRef]
  71. Kubička, D.; Horáček, J. Deactivation of HDS catalysts in deoxygenation of vegetable oils. Appl. Catal. A Gen. 2011, 394, 9–17. [Google Scholar] [CrossRef]
  72. Arora, P.; Abdolahi, H.; Cheah, Y.W.; Salam, M.A.; Grennfelt, E.L.; Rådberg, H.; Creaser, D.; Olsson, L. The role of catalyst poisons during hydrodeoxygenation of renewable oils. Catal. Today 2021, 367, 28–42. [Google Scholar] [CrossRef]
  73. Andersen, A.; Kathmann, S.M.; Lilga, M.A.; Albrecht, K.O.; Hallen, R.T.; Mei, D. Adsorption of Potassium on MoS2(100) Surface: A First-Principles Investigation. J. Phys. Chem. C 2011, 115, 9025–9040. [Google Scholar] [CrossRef]
  74. Mey, D.; Brunet, S.; Perot, G.; Diehl, F. Catalytic deep HDS of model FCC feed over a CoMo/Al2O3 catalyst modified by potassium. In Proceedings of the 2003 SPE/EPA/DOE Exploration Production Environmental Conference, San Antonio, TX, USA, 10–12 March 2003; pp. 44–45. [Google Scholar]
  75. Mey, D.; Brunet, S.; Canaff, C.; Maugé, F.; Bouchy, C.; Diehl, F. HDS of a model FCC gasoline over a sulfided CoMo/Al2O3 catalyst: Effect of the addition of potassium. J. Catal. 2004, 227, 436–447. [Google Scholar] [CrossRef]
  76. Cable, T.L.; Massoth, F.E.; Thomas, M.G. A basic study of catalyst aging in the H-coal process. Fuel Process. Technol. 1985, 10, 105–120. [Google Scholar] [CrossRef]
  77. Silva, S.M.; Sampaio, K.A.; Ceriani, R.; Verhé, R.; Stevens, C.; De Greyt, W.; Meirelles, A.J.A. Effect of type of bleaching earth on the final color of refined palm oil. LWT Food Sci. Technol. 2014, 59, 1258–1264. [Google Scholar] [CrossRef]
  78. de Klerk, A. Oligomerization of 1-Hexene and 1-Octene over solid acid catalysts. Ind. Eng. Chem. Res. 2005, 44, 3887–3893. [Google Scholar] [CrossRef]
  79. Charisiou, N.D.; Douvartzides, S.L.; Siakavelas, G.I.; Tzounis, L.; Sebastian, V.; Stolojan, V.; Hinder, S.J.; Baker, M.A.; Polychronopoulou, K.; Goula, M.A. The relationship between reaction temperature and carbon deposition on nickel catalysts based on Al2O3, ZrO2 or SiO2 supports during the biogas dry reforming reaction. Catalysts 2019, 9, 676. [Google Scholar] [CrossRef]
  80. Zethof, J.H.T.; Leue, M.; Vogel, C.; Stoner, S.W.; Kalbitz, K. Identifying and quantifying geogenic organic carbon in soils—The case of graphite. SOIL 2019, 5, 383–398. [Google Scholar] [CrossRef]
  81. Arvelakis, S.; Jensen, P.A.; Dam-Johansen, K. Simultaneous thermal analysis (STA) on ash from high-alkali biomass. Energy Fuels 2004, 18, 1066–1076. [Google Scholar] [CrossRef]
Figure 1. The adsorption–desorption isotherms of the calcined NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts.
Figure 1. The adsorption–desorption isotherms of the calcined NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts.
Energies 15 01547 g001
Figure 2. The XRD patterns of the calcined (a), sulfide (b), and post-run (c) of NiMo/SiO2, NiMo/TiO2, and NiMo/γ-Al2O3 catalysts.
Figure 2. The XRD patterns of the calcined (a), sulfide (b), and post-run (c) of NiMo/SiO2, NiMo/TiO2, and NiMo/γ-Al2O3 catalysts.
Energies 15 01547 g002aEnergies 15 01547 g002b
Figure 3. H2-temperature-programmed reduction profiles of the calcined NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts.
Figure 3. H2-temperature-programmed reduction profiles of the calcined NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts.
Energies 15 01547 g003
Figure 4. The conversion, diesel selectivity, and gasoline selectivity of the NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts. All experiments were conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV = 1.5 h−1, and H2/oil ratio = 1000 cm3/cm3.
Figure 4. The conversion, diesel selectivity, and gasoline selectivity of the NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts. All experiments were conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV = 1.5 h−1, and H2/oil ratio = 1000 cm3/cm3.
Energies 15 01547 g004
Figure 5. The diesel yield, gasoline yield, and gas yield of NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts. All experiments were conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV = 1.5 h−1, and H2/oil ratio = 1000 cm3/cm3.
Figure 5. The diesel yield, gasoline yield, and gas yield of NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts. All experiments were conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV = 1.5 h−1, and H2/oil ratio = 1000 cm3/cm3.
Energies 15 01547 g005
Figure 6. Effect of catalyst on n-Cn-1 and n-Cn in diesel range of NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts. All experiments were conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV = 1.5 h−1, and H2/oil ratio = 1000 cm3/cm3.
Figure 6. Effect of catalyst on n-Cn-1 and n-Cn in diesel range of NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts. All experiments were conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV = 1.5 h−1, and H2/oil ratio = 1000 cm3/cm3.
Energies 15 01547 g006
Figure 7. Mole fraction of gaseous products from the reactions (excluding hydrogen) performed over NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts. All experiments were conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV of 1.5 h−1, and H2/oil ratio of 1000 cm3/cm3.
Figure 7. Mole fraction of gaseous products from the reactions (excluding hydrogen) performed over NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts. All experiments were conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV of 1.5 h−1, and H2/oil ratio of 1000 cm3/cm3.
Energies 15 01547 g007
Figure 8. The FTIR spectra of Pongamia pinnata seed extracted oil and deoxygenated liquid product over NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2. All experiments were conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV of 1.5 h−1, and H2/oil ratio of 1000 cm3/cm3. The product was collected after stabilizing the reaction for 9 h.
Figure 8. The FTIR spectra of Pongamia pinnata seed extracted oil and deoxygenated liquid product over NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2. All experiments were conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV of 1.5 h−1, and H2/oil ratio of 1000 cm3/cm3. The product was collected after stabilizing the reaction for 9 h.
Energies 15 01547 g008
Figure 9. Comparison of H/C and O/C ratio for the hydrotreated liquid product over sulfide NiMo/γ-Al2O3 (BHD-Al2O3), NiMo/SiO2 (BHD-SiO2), and NiMo/TiO2 (BHD-TiO2) catalysts and Pongamia pinnata oil (PPO) with petroleum fuel and hydrotreated vegetable oil (HVO) from Neste oil.
Figure 9. Comparison of H/C and O/C ratio for the hydrotreated liquid product over sulfide NiMo/γ-Al2O3 (BHD-Al2O3), NiMo/SiO2 (BHD-SiO2), and NiMo/TiO2 (BHD-TiO2) catalysts and Pongamia pinnata oil (PPO) with petroleum fuel and hydrotreated vegetable oil (HVO) from Neste oil.
Energies 15 01547 g009
Figure 10. Effect of reaction time on conversion of (a) Pongamia Pinnata oil and (b) refined palm olein, over the various supported NiMoS. The experiment was conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV of 1.5 h−1, and H2/oil ratio of 1000 cm3/cm3.
Figure 10. Effect of reaction time on conversion of (a) Pongamia Pinnata oil and (b) refined palm olein, over the various supported NiMoS. The experiment was conducted at a temperature of 330 °C, H2 pressure of 50 bar, WHSV of 1.5 h−1, and H2/oil ratio of 1000 cm3/cm3.
Energies 15 01547 g010
Figure 11. Thermogravimetric (TGA) as function of temperature of the spent NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts used in hydrotreating of (a) Pongamia pinnata oil and (b) refined palm olein (RPO).
Figure 11. Thermogravimetric (TGA) as function of temperature of the spent NiMo/γ-Al2O3, NiMo/SiO2, and NiMo/TiO2 catalysts used in hydrotreating of (a) Pongamia pinnata oil and (b) refined palm olein (RPO).
Energies 15 01547 g011
Table 1. The fatty acid distributions of feedstocks.
Table 1. The fatty acid distributions of feedstocks.
Fatty Acid (wt.%)Pongamia pinnata OilRefined Palm Olein
C12:0 (Lauric acid)0.00.4
C14:0 (Myristic acid)0.00.8
C16:0 (Palmitic acid)15.137.4
C16:1 (Palmitoleic acid)0.00.2
C18:0 (Stearic acid)4.73.6
C18:1 (n-9 cis, Oleic acid)54.445.8
C18:2 (n-6 cis, Linoleic acid)13.511.1
C18:3 (n-3, Linolenic acid)0.20.3
C20:0 (Arachidic acid)1.10.3
C20:1 (n-9, Eicosenoic acid)1.50.1
C22:0 (Behenic acid)8.30.0
Percent of fatty acid (wt.%)5.7<1
Table 2. The elemental analysis of the feedstocks and liquid products of supported sulfide NiMo over γ-Al2O3, SiO2, and TiO2 catalysts.
Table 2. The elemental analysis of the feedstocks and liquid products of supported sulfide NiMo over γ-Al2O3, SiO2, and TiO2 catalysts.
SampleC (%)H (%)N (%)S (%)O (%)
Extracted oil76.9912.050.282.168.52
Hydrotreated extracted oilNiMoS/γ-Al2O381.6614.750.482.650.46
NiMoS/SiO279.7312.780.112.075.31
NiMoS/TiO281.3115.970.022.560.14
Table 3. Physicochemical properties of the calcined catalysts.
Table 3. Physicochemical properties of the calcined catalysts.
CatalystsSurface Area (m2 g−1)Total Pore Volume (cm3 g−1)Mean Pore Diameter (nm)
γ-Al2O32400.509.2
NiMo/γ-Al2O32090.408.0
SiO22890.9413.9
NiMo/SiO22340.7912.7
TiO21550.386.2
NiMo/TiO21470.335.4
Table 4. Composition of the liquid product from the catalytic hydrotreating of Pongamia pinnata oil over the sulfided NiMo catalysts.
Table 4. Composition of the liquid product from the catalytic hydrotreating of Pongamia pinnata oil over the sulfided NiMo catalysts.
CatalystsLiquid Product Composition (%)
<n-C15n-C15n-C16n-C17n-C18n-C19n-C20n-C21n-C22>n-C22
NiMo/γ-Al2O30.123.548.4522.1050.671.132.242.695.643.43
NiMo/SiO20.6110.821.9561.2010.521.880.556.471.784.21
NiMo/TiO20.271.8010.5810.3361.840.552.581.556.743.76
Table 5. Sulfur content of sulfide catalysts initially and after 15 h of the reaction analyzed by energy dispersive X-ray (EDX).
Table 5. Sulfur content of sulfide catalysts initially and after 15 h of the reaction analyzed by energy dispersive X-ray (EDX).
CatalystsSulfur (%)
Sulfided NiMo/γ-Al2O312.922
Spent NiMo/γ-Al2O39.701
Sulfided NiMo/SiO215.130
Spent NiMo/SiO213.014
Sulfided NiMo/TiO211.965
Spent NiMo/TiO212.311
Table 6. The impurities in the feedstocks.
Table 6. The impurities in the feedstocks.
SampleConcentration (ppm)
NaKCaMgP
Pongamia pinnata oil113168231.5108.3509.5
Refined palm olein--<0.1<0.5<2.5 [77]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Plaola, Y.; Leangsiri, W.; Pongsiriyakul, K.; Kiatkittipong, W.; Srifa, A.; Lim, J.W.; Reubroycharoen, P.; Kiatkittipong, K.; Eiad-ua, A.; Assabumrungrat, S. Catalytic Hydrotreating of Crude Pongamia pinnata Oil to Bio-Hydrogenated Diesel over Sulfided NiMo Catalyst. Energies 2022, 15, 1547. https://0-doi-org.brum.beds.ac.uk/10.3390/en15041547

AMA Style

Plaola Y, Leangsiri W, Pongsiriyakul K, Kiatkittipong W, Srifa A, Lim JW, Reubroycharoen P, Kiatkittipong K, Eiad-ua A, Assabumrungrat S. Catalytic Hydrotreating of Crude Pongamia pinnata Oil to Bio-Hydrogenated Diesel over Sulfided NiMo Catalyst. Energies. 2022; 15(4):1547. https://0-doi-org.brum.beds.ac.uk/10.3390/en15041547

Chicago/Turabian Style

Plaola, Yuwadee, Wanwipa Leangsiri, Kanokthip Pongsiriyakul, Worapon Kiatkittipong, Atthapon Srifa, Jun Wei Lim, Prasert Reubroycharoen, Kunlanan Kiatkittipong, Apiluck Eiad-ua, and Suttichai Assabumrungrat. 2022. "Catalytic Hydrotreating of Crude Pongamia pinnata Oil to Bio-Hydrogenated Diesel over Sulfided NiMo Catalyst" Energies 15, no. 4: 1547. https://0-doi-org.brum.beds.ac.uk/10.3390/en15041547

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