1 s2.0 S0263876221002872 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 70

Journal Pre-proof

Catalytic Deoxygenation of Palm Oil and its Residue in Green Diesel


Production: a Current Technological Review<!–<ForCover>Mahdi HI,
Bazargan A, McKay G, Azelee NIW, Meili L, Catalytic Deoxygenation of
Palm Oil and its Residue in Green Diesel Production: a Current
Technological Review, Chemical Engineering Research and Design, doi:
10.1016/j.cherd.2021.07.009</ForCover>–>

Hilman Ibnu Mahdi, Alireza Bazargan, Gordon McKay, Nur Izyan


Wan Azelee, Lucas Meili

PII: S0263-8762(21)00287-2
DOI: https://doi.org/10.1016/j.cherd.2021.07.009
Reference: CHERD 4545

To appear in: Chemical Engineering Research and Design

Received Date: 23 January 2021


Revised Date: 1 June 2021
Accepted Date: 9 July 2021

Please cite this article as: { doi: https://doi.org/

This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2020 Published by Elsevier.


Catalytic Deoxygenation of Palm Oil and its Residue in Green Diesel
Production: a Current Technological Review
Hilman Ibnu Mahdi1, Alireza Bazargan2, Gordon McKay3, Nur Izyan Wan Azelee4,5, Lucas
Meili6
1
PT Global Amines Indonesia, Kapten Darmo Sugondo No.56 Street, Indro Legi, Kec. Gresik,
Kabupaten Gresik, East Java, 61124, Indonesia
2
School of Environment, College of Engineering, University of Tehran, Tehran, Iran
3
Division of Sustainable Development, College of Science and Engineering, Hamad Bin Khalifa
University, Education City, Qatar Foundation, Doha, Qatar
4
School of Chemical and Energy Engineering, Faculty of Engineering, Universiti Teknologi

of
Malaysia, 81310, UTM Skudai, Johor, Malaysia.
5
Institute of Bioproduct Development (IBD), Universiti Teknologi Malaysia, 81310, UTM
Skudai, Johor, Malaysia.

ro
6
Laboratório de Processos, Centro de Tecnologia, Universidade Federal de Alagoas, Av.
Lourival de Melo Mota, Tabuleiro dos Martins, Maceió, Alagoas, Brazil

-p
ORCID: 0000-0002-0803-8359 (Hilman Ibnu Mahdi); 0000-0001-9956-2377 (Alireza
Bazargan); 0000-0001-9965-7659 (Gordon McKay); 0000-0001-9415-4382 (Nur Izyan Wan
Azelee); 0000-0002-0307-8204 (Lucas Meili)
re
Corresponding authors: hilmanibnu.mahdi@globalamines.com; alireza.bazargan@ut.ac.ir;
gmckay@hbku.edu.qa; nur.izyan@utm.my and lucas.meili@ctec.ufal.br
Highlights
lP

 Green diesel consumption is annually forecasted to increase to 3% in the coming 2030;


 Deoxygenation technology is widely applied in the green diesel production;
 Zeolites are promising and potential catalysts for DO;
 The reusability studies are important to confirm catalysts stability and performance;
na

 The costing analysis of green diesel production need to be thoroughly investigated.

Abstract: Worldwide consumption of energy produced from fossil fuel is forecasted to grow.
ur

This trend leads to both environmental pollution and the depletion of fossil fuel resources. Green
diesel is a suitable candidate to substitute petroleum based-diesel due to its plentiful raw
materials, non-polluting production process, and cost-effectiveness. Green diesel production is
Jo

stated as simple, efficient, and relatively clean process. Deoxygenation (DO) is crowned as the
best available technology to produce green diesel from palm fatty acid distillate (a side product
of palm oil production) and other oils using heterogeneous catalysts such as zeolites. The
capability of catalysts can be improved by optimizing operating parameters, treating and
modifying catalysts and with the use of nano-sized catalytic materials. These activities contribute
to stronger and more active Bronsted-Lewis acid-base sites and enlarged crystallite sizes, which
improve the DO efficiency, selectivity, and reusability, to produce high-grade green diesel with
less oxygen content.
Keywords: Deoxygenation technology; Palm oil; Palm fatty acid distillate; Zeolites; Fatty acid
methyl ester.

Contents
1. Introduction
2. Green diesel production by heterogeneous reaction methods
2.1 The effect of different reaction routes

of
2.2 The role of metal oxides
2.3 The role of nickel phosphides
3. Green diesel production using zeolite catalysts

ro
4. Green diesel producing technologies
5. Catalytic deoxygenation of palm oil and palm fatty acid distillate
5.1 The effect of operating conditions
5.2 Leverage of catalyst treatment
-p
re
5.3 The role of metal loadings
5.4 The role of phosphides contents
lP

5.5 Nanoparticle catalysts


6. Final considerations
References
na
ur

1. Introduction
Jo

Approximately, 80% of global energy consumption originates from fossil fuels [1] and it is
predicted to grow at an average annual rate of 0.9% by the year 2030 [2, 3] and global energy
consumption is forecasted to reach 780 x 1015 Btu by 2035. From this value, 280 x 1015 Btu and
500 x 1015 Btu is associated with economic co-operation and development (OECD) or non-
OECD countries respectively, both of which exhibit the same upward trend [4].
Fossil fuels are widely utilized in all sectors [5] and contribute to economic growth in every
country [6]. Nevertheless, the massive and uncontrolled use of fossil fuels has caused a reduction
in underground reserves, as well as negative environmental impacts [7, 8] incorporating major
increases in greenhouse gas emissions, GHGs [9]. As the largest global energy consuming sector,
40% of worldwide energy consumption is attributed to transportation, in the form of gasoline, jet
fuel, diesel fuels, namely, green diesel [10] and biodiesel [11, 12] with a number of reviews
comparing the benefits of the diesel types [13, 14].
Due to the high demand of fuels in the transportation sector and the concern regarding the

of
depletion of fossil fuels and global warming, renewable, sustainable, and alternative energy
sources such as biofuels are necessary to replace petroleum-based products [15]. Biofuels [16,
17] can be divided into two types, namely bioethanol and biodiesel. Bioethanol [18] is produced

ro
from the fermentation of simple sugars or agricultural raw materials containing sugar while
biodiesel is produced from vegetable oils and methanol [19, 20]. Biofuel is cleaner than fossil

-p
fuels, and can be produced from edible and non-edible oils as well as animal fats using
heterogeneous solid catalysts and enzymes [21]. A wide range of conversion processes of source
re
materials have been investigated to produce biofuels, for example, lignin based materials [22],
seaweed [23], spent coffee grounds [24], and microwave pyrolysis of biomass [25]. Due to its
lP

minimal emissions, it is recognized as the greenest and most cost-efficient fuel with the potential
to be applied in diesel engines based on an industrial production scale [26, 27].
On the other hand, the catalysts used for the required reactions to produce biodiesel can be
na

derived from wastes, as has been previously demonstrated by our group. This allows for double-
sided benefits, both from the perspective of using waste products to produce valuable catalysts,
and, the production of biofuels as the main objective [28, 29].
ur

Hence, many different resources have been exploited for the production of biofuels to
achieve the increasing demand and to mitigate environmental pollution. The total global
Jo

production of edible vegetable oils such as palm, soybean, rapeseed, sunflower, and kernel oils
harnessed for biofuel production has been reported at 143.15 million tons (MT) in the last
decade. The highest production was 48.10 MT of palm oil (33.6% of the total production), and
the lowest production was 5.44 MT for palm kernel oil (3.8% of the total production) as depicted
in Figure 1. Whilst approximately 8.5% of the total production (12.17 MT) was produced from
other non-edible vegetable oils including jatropha, castor, karanja, macauba acid, and rubber
seed oils [30].

of
ro
-p
re
lP

Figure 1. The global production of vegetable oils applied in biofuels [30].

Among edible and non-edible vegetable oils that are widely exploited to produce biofuels is
na

palm oil which has been increasing. For example, palm oil production was 20 MT in 2000 and
has grown more than two fold to 50 MT in 2012 and 54.5 MT in 2020 [31-33]. [32–34].
Moreover it is estimated to increase over 78 MT and 84 MT in 2045 and 2050, respectively [33].
ur

Notwithstanding, from the palm oil produced in Indonesia, the majority is mostly exported to
other countries (67%), and the rest is used in the food (20%) and biodiesel (13%) industries [13,
Jo

19, 20]. Consequently, the optimization of palm oil production to use in the biofuels sector
receives vast attention and provides many challenges for researchers [34-36].
Biodiesel, also known as FAME (fatty acid methyl ester), is one of the most commercially
promising biofuels that is typically produced from palm oil and used as a substitution for diesel
fuel [37-41]. However, regular biodiesel may have a high oxygen content, poor stability [42] and
different fatty acid ester compounds compared with petroleum-derived-diesel [43]. Despite being
already used at an industrial scale and commercially applied in the transportation sector, the use
of regular biodiesels for diesel engines still has some significant hurdles to overcome. Therefore
better biofuels, such as green diesel, are highly sought after [44].
Recently, green diesel has been recognized as the most favorable biofuel due to its higher
heating value, energy density, cetane number, and better quality than petroleum-based diesel and
biodiesels. The term “Green Diesel” is often associated with the trademark of Honewell
company, as produced using the Ecofining process. That is why the phrase “Renewable Diesel”
is sometimes used instead of “Green Diesel”. In addition, green diesel has high stability and a

of
low oxygen content, as shown in Table 1. The differences in physio-chemical properties
between green diesel, petroleum-diesel, and regular biodiesel are compared with each other.
Green diesel is generally produced by deoxygenation technology (DO tech) which is a

ro
suitable process using various vegetable oils and solid catalysts [45]. Zeolites are highly
recommended catalysts due to their reusability, cost-effectiveness, stability, catalytic activity,

-p
high conversion, high yield, selectivity, and abundance [46, 47]. Producing green diesel from
crude palm oil (CPO) forsakes palm fatty acid distillate (PFAD) as its residue in the final
re
process, which makes the overall production less clean [48-50]. This will make CPO-based
diesel production a green and high-value industry.
lP

This article reviews the green diesel production from palm oil and PFAD using the DO
technology, where heterogeneous solid catalysts and zeolites are exploited. Optimized operating
conditions, catalyst treatments, doping procedures, phosphide loadings and nano-sized catalysts
na

are thoroughly discussed.


ur
Jo
Table 1. Physio-chemical properties of green diesel, petroleum-diesel, and biodiesel.

of
Physio-chemical properties of fuels
Containing Ref
Petroleum diesel Regular biodiesel Green diesel

ro
Carbon (wt.%) 86.8 76.2 84.9 [42]
Hydrogen (wt.%) 13.2 12.6 15.1 [42]
Specific gravity 0.84 0.88 0.78

-p
[42, 51]
Oxygen (wt.%) 0.0 11.2 0.0 [42, 51]
[42, 51,
Cetane number (CN) 40 – 67 45 – 65 70 – 90
52]

re
[42, 51,
Lower Heating Value (LHV) MJ/kg 42.34 – 43.1 37.2 – 38 43.7 – 44.5
53]
Density at 15 ◦C (kg/m3) 796 – 841 880 770 – 790 [42, 53]

Aromatic content, %/.vol


Sulfur Content (ppm)
Flash Point (oC)
lP
Polycyclic Aromatic Hydrocarbons (wt%) 1.5 – 4.4
22.4 – 35
< 10
54 – 148
-
-
<1
100 – 180
< 0.1
0.6
<1
59 – 138
[42]
[42, 53]
[51, 53]
na
[42, 53]
Cloud Point (◦C) - 5 to 3 - 5 up to 15 - 20 – 21 [51, 53]
Ash Content (wt%) 0.01 - < 0.001 [42]
Water Content (mg/kg) - - < 200 [42]
ur

Carbon Residue on 10% Distillation (wt%) - - < 0.1 [42]


Total Contamination (mg/kg) - - < 10 [42]
Water and Sediment (vol%) - - < 0.02 [42]
Jo

Copper Strip Corrosion Rating No 3 No 1 No 1 [42]


Oxidation Stability - 2 – 15 h < 25 g/m3 [42, 54]
Lubricity, Wear Scar Diameter at 60 oC (µm) 226 – 354 - > 700 [42, 52]
Viscosity at 40 oC (mm2/s) 1.9 – 4.1 2.9 – 11 2–4 [42, 53]
Stability Good Medium Good [51]
2. Green diesel production by heterogeneous reaction methods
There are many available pathways to convert vegetable oils into liquid fuels that are utilized
to substitute petroleum-based diesel and regular biodiesel with more environmentally friendly
and cost-effective options such as green diesel. Green diesel is notably produced by the
hydrodeoxygenation (HDO) process from a wide range of oils [34, 55, 56], which includes,
sunflower oil [57], rapeseed oil [58], algae oil and lipids [59], soybean oil [60], palm kernel oil
[61], Jatropha oil [62, 63], anisole [64], and vegetable oils [65]. The vegetable oils contain
triglycerides and FFAs with high oxygen content which can be removed by various reaction

of
methods such as hydrogenolysis (HG), dehydration (DH), DO, decarboxylation (DCX), and the
decarbonylation (DCN) processes [66]. By undergoing catalytic cracking (CC) and catalytic
isomerization (CI), green diesel is produced as portrayed in Figure 2. The figure describes the

ro
possibilities of available pathways to produce green diesel from FFAs.

-p
re
lP
na
ur
Jo

Figure 2. Deoxygenation of FFAs to produce green diesel [30].


Through the schematic diagram presented in Figure 2, it is seen that green diesel can be
produced from various reactions such as CC, CI, HDO, DO, DCN, and DCX [30]. Murti et al.
[67] claimed that these pathways are not only for green diesel production but also for biodiesel
production, because these products consist of methyl and ethyl ester molecules and are produced
from vegetable oils containing high FFAs [68, 69]. However, green diesel differs from regular
biodiesel due to the processes used, including the hydrogenation process [70], pyrolytic
deoxygenation [71], hydrogenative decarboxylation (HDCX) in microwave [72], hydrotreating
[73, 74], hydrocracking (HC) [75], hydroconversion [76], and hydroprocessing [77] using

of
heterogeneous catalysts as listed in Table 2.
Baek et al. [72] performed experiments using two types of biofuel production techniques,
namely green diesel and bio-jet fuel over DCX of fatty acids using silicon nanowire array-

ro
assisted nano-sized Rhodium (Rh) catalyst (SiNA-Rh) under microwave irradiation. The low
electric power consumption was attributed to the microwave systems [78, 79] and the SiNA-Rh

-p
catalyst was highly efficient and reusable up to 20 cycles [72]. The DCX process was also
implemented by Vardon et al. [80] using Pt-Re/Carbon, Mizugaki et al. [81] using Ru-CeO2,
re
Yamada et al. [82] using nano-sized SiNA-Pd, and Zhang et al. [83] using Ru-HAP catalysts.
On the other hand, Moreira et al. [84] studied DO of macauba acid oil to produce green
lP

diesel and biokerosene with an activated biochar-assisted Co/Carbon catalyst. DO degree of 98.4
wt.% was reached at 350 oC, 30 bar and 10 wt.% of catalyst within 4 h. Meanwhile, the DO
process of macauba oils, proposed by Silva et al. [85] to produce green diesel and biokerosene
na

using Pd/C catalyst, resulted in 85 wt.% hydrocarbon content (mostly green diesel) at 300 oC, 10
bar, and 700 rpm mixing for 5 h. Green diesel production from macauba acid oil has also been
investigated in other studies [85-87].
ur

Apart from macauba oils, oleic acid could also be used in green diesel production as
reported by Shim et al. [88] using a CoMo catalyst through the sol gel synthesis method. The
Jo

resulting conversion and selectivity were 88.9% and 48.1%, respectively, and the removed
oxygen content was 69.6%. The researchers claimed that the CoMo catalyst showed excellent
catalytic performance and was an effective and potential catalyst in green diesel production [88].
Table 2. Green diesel fuel produced from various reaction routes.

of
Operating conditions
Conversion Selectivity Yield
Pathways Feedstocks Temp Pressure Time Catalyst Catalysts Ref
(%) (%) (%)
(oC) (Bar) (h) content

ro
Methyl
Hydroprocessing 240 25 4 20 mg 38 – 77 11.3 – 61.9 - Ni/beta zeolite [89]
palmitate oil
Vegetable Pt/γ-Al2O3, Pt/TiO2, and Pt-
Deoxygenation (DO) 320 50 4 0.24 gr - - 67.0 [36]

-p
oils MoOx/TiO2
Crude palm 280 –
Hydrogenation 40 – 90 1 0.15 gr - - 68.2 NiMo/ᵞ-Al2O3 [70]
oil (CPO) 380
Ni/ᵞ-Al2O3

re
Hydrogenation Karanja oil 360 - 2 20 wt.% 100 70 - [90]
MWCNTs impregnated by
Chicken fat 75 – different metal oxides such as
Deoxygenation (DO) 350 - 2 3 wt.% - 81 – 85 [91]
oil (CFO) 84 Ni–Mg, Ni–Mn, Ni–Cu, and

Hydrodeoxygenation
(HDO)
Hydrodeoxygenation
Rubber Seed
Oil (RSO)
Rubber Seed
300 –
400
300 –
lP
30 – 80

30 – 80
- - 100 - 84.94
Ni–Ce
NiMo/ᵞ-Al2O3

NiMo/ᵞ-Al2O3
[66]
na
5 - 100 19.1 - [92]
(HDO) Oil (RSO) 400
Decarboxylation Rubber Seed 300 –
30 – 80 5 - 100 81.7 - NiMo/ᵞ-Al2O3 [92]
(DCX) Oil (RSO) 400
Pyrolytic- Jatropha Atmosph MWCNTs assisted by NiO–
350 1 3 wt.% - - 75 [71]
ur

deoxygenation curcas oil ere Fe2O3 and NiO–ZnO


Hydrogenative
Nano-sized rhodium (Rh)
decarboxylation 500 mol
Fatty acids 200 10 24 - - 99 doped by silicon nanowire [72]
Jo

(HDCX) in ppm
array (SiNA-Rh)
microwave
Palm fatty
acid Atmosph
Deoxygenation (DO) 350 2 5 wt.% - 86.0 83.4 NiO-ZnO [49]
distillate ere
(PFAD)
Palm fatty
acid 5 – 20 Activated carbon-assisted
Deoxygenation (DO) 350 - 1 - 89 92 [45]

of
distillate wt.% CoMo
(PFAD)
Palm fatty
85.8

ro
acid Ni, Co, and Ni-Co doped
Deoxygenation (DO) 350 - 3 10 wt.% - > 85 and [48]
distillate SBA-15
88.1
(PFAD)

-p
Methyl ester
Hydrotreating 300 30 6 5 wt.% > 90 - > 60 Co and Ni-doped Zr‐SBA‐15 [73]
s
24,
260 – 29, USY, H-ZSM-5 and Al-SBA-

re
Deoxygenation (DO) Oleic acid 50 6 1.5 gr - - [93]
300 and 15-supported Ni2P
42
Macauba
Deoxygenation (DO)

Hydrocracking (HC)

Hydroconversion
acid oil
Palm Oil
Waste
350

400
380 –
lP
30

60

40
4

-
10 wt.%

0.9%
1 and 5
-

2.9 – 21.5
-

-
98.4
50 –
62.20
39 –
Co-Carbon (activated biochar)

Pd/Al2O3 and Pd-Fe/Al2O3

Mn-Al2O3-assisted NiMo
[84]

[75]

[76]
na
soybean oil 410 mol% 82
3, 12,
Hydrodeoxygenation Rubber Seed
350 35 - and 15 > 99 10.4 - HY Zeolite [94]
(HDO) Oil (RSO)
wt.%
300 –
ur

Hydroprocessing Soybean oil 10 3–5 0 – 25% - - 40 Niobium phosphate (NbOPO4) [77]


350
300 –
Deoxygenation (DO) Soybean oil 10 3–5 0 – 25% - 37 15 Niobium phosphate (NbOPO4) [95]
350
Jo

300 –
Deoxygenation (DO) Soybean oil 10 3–5 0 – 25% - - 38 Pd-Carbon [95]
350
300 –
Deoxygenation (DO) Soybean oil 10 3–5 0 – 25% - - 27 Fluid catalytic cracking (FCC) [95]
350
52
260 –
Deoxygenation (DO) Oleic acid 50 6 - 100 - and CoP-USY and Ni2P-USY [96]
340

of
36
Atmosph
Deoxygenation (DO) Oleic acid 300 3 - 98.7 54.2 - Pt-Ni-assisted Ce0.6Zr0.4O2 [97]
ere

ro
Atmosph
Deoxygenation (DO) Oleic acid 300 3 - 98.3 33.9 - Ni-assisted Ce0.6Zr0.4O2 [97]
ere
Atmosph
Deoxygenation (DO) Oleic acid 300 3 - 90.8 19 - Pd-Ni-assisted Ce0.6Zr0.4O2 [97]

-p
ere
Atmosph
Deoxygenation (DO) Oleic acid 300 3 - 88.3 19.9 - Ru-Ni-assisted Ce0.6Zr0.4O2 [97]
ere
3 wt.%

re
325 –
Deoxygenation (DO) Canola oil 31 - and 15 100 - - NiMo/CoMo [98]
400
wt.%
Atmosph
Deoxygenation (DO)

Deoxygenation (DO)

Deoxygenation (DO)
Oleic acid

Oleic acid

Oleic acid
300

300

300
lP
ere
Atmosph
ere
Atmosph
3

3
0.69 gr

0.69 gr

0.69 gr
88.9

62

84.9
48.1

19.7

22.8
-

-
CoMo-SG

CoMo-IWI

CoMo-HT
[88]

[88]

[88]
na
ere
Atmosph
Deoxygenation (DO) Oleic acid 300 3 0.69 gr 87.2 31 - CoMo-CP [88]
ere
Coffee
Hydrotreating 400 40 2 5% > 90 - > 20 NiMo/ᵞ-Al2O3 and Pd-Carbon [74]
ur

ground oil
Hydrodeoxygenation Palm kernel
300 10 5 5% 96 - 98 Pd-Carbon [99]
(HDO) fat
Macauba
Jo

Deoxygenation (DO) 300 10 5 5 wt.% - 97 – 100 85 Pd-Carbon [85]


oils
310 – Atmosph
Deoxygenation (DO) Palmitic acid - 3 gr 99.6 - 42.9 Ni2P/H-ZSM-22 [100]
360 ere
Crude palm
Hydroprocessing 400 40 3 5% > 60 - 51 Pd-Carbon [101]
oil (CPO)
MWCNTs = multi-walled carbon nanotubes; SG = sol gel; IWI = incipient wet impregnation; HT = hydrothermal; CP = co-
precipitation; Ni2P = nickel phosphide.

of
ro
-p
re
lP
na
ur
Jo
The use of ᵞ-Al2O3/NiMo catalysts for DO of oleic acid was also observed by Sagi et al.
[102] at 320 – 380 oC, 40 – 70 bar with H2/oleic acid ratio of 600 Nm3/m3 and by Szarvas et al.
[103] using phosphorus impregnation at 300 – 400 oC and 1 bar. Sousa et al. [99] investigated
DO of palmist fat using Pd/C catalyst at 300 oC and 10 bar for 5 h, resulting in 96% conversion
where 5% is produced bio-jet fuel [99].
Phimsen et al. [74] comprehensively compared the ᵞ-Al2O3/NiMo and Pd/C catalysts in the
hydrocracking of oil extracted from coffee grounds at 400 oC for 2 h. The yield obtained by the
Pd/C catalyst (22.3 wt.%) was higher than the ᵞ-Al2O3/NiMo catalyst (4.8 wt.%). Importantly,

of
the cetane index of green diesel produced over these two catalysts was much higher than
commercial diesel fuel.
The effect of bimetallic catalysts (i.e. ᵞ-Al2O3/Co, ᵞ-Al2O3/Ni, ᵞ-Al2O3/Pd, and ᵞ-Al2O3/Pt)

ro
was explored by Srifa et al. [104] with various metallic content ranging from 2 – 10 wt.% in the
DO of palm oil. The best catalytic performance was the catalyst supported by Co followed by Pd,

-p
Pt, and Ni. A larger metal size increased the turnover frequency (TOF), and a higher metal
content displayed a positive effect on the obtained results. For example, a higher yield of olefins
re
(i.e. C17) was achieved at 10% Co content (23.5%) compared to 5% Co content (21.9%) with
complete conversion (100%).
lP

Other than solid acid-base catalysts, zeolite catalysts can also be applied in green diesel
production from edible and non-edible oils. Wang et al. [105] harnessed several zeolites such as
SAPO-11, ZSM-5, ZSM-22, ZSM-23 and beta supported by Ni. The best catalyst was found to
na

be SAPO-11 doped with 8 wt.% Ni, resulting in the maximum conversion of 100% and a 74.8%
yield of green diesel at 370 oC and 30 bar. Boyas et al. [106] performed hydrocracking of
rapeseed oil in a batch reactor at 300 – 400 oC and 50 – 110 bar using Pt/H-Y, Pt/ZSM-5, and ᵞ-
ur

Al2O3/NiMo, the latter of which had the optimum yield. Meanwhile, the catalytic DO, CC, and
CI of palm oils were simultaneously carried out by Sousa et al. [107] using beta zeolite at 350 oC
Jo

and 10 bar for 5 h. High conversion (96%) was achieved under mild process conditions, with
reusable beta zeolite without the need for metal oxides doping [107].
Based on the above reports, the quality and quantity of green diesel production are strongly
affected by several factors such as different reaction routes, metal oxides, and the use of nickel-
phosphides on the catalyst. All these factors need to be carefully considered for optimum results
and high catalyst reusability. The authors highlighted it was significant to enhance the catalyst
activity over the nickel-phosphide because it makes the catalyst porosity, Bronsted-lewis, and
accessibility higher, stronger with pore enlargement leading to optimum yield and good
performance.

2.1 The effect of different reaction routes


Heterogeneous methods employed to produce green diesel consist of two different
processes, namely the chemical routes and physical routes. The chemical routes consist of
catalytic hydrothermal liquefaction [108], cracking [109, 110], emulsification [111, 112],
esterification [113, 114], including DO processes such as the HDO, DCX and DCN reactions

of
[115], as well as hydroprocessing [116] and steam reforming [114] methods. Meanwhile, the
physical routes include hot vapor filtration (HVF), removal char, adding solvents and the

ro
extracted organic acids processes [117-119].
The employed routes for producing green diesel from vegetable oils have a significant

-p
impact on the achieved results and catalyst performance [108, 120]. For instance, Ameen et al.
[66] investigated HDO of rubber seed oil (RSO) to produce green diesel by using NiMo/γ-Al2O3
re
catalyst. A 100% conversion of the RSO and 84.94 wt.% yield were achieved at 400 °C, weight
hourly space velocity (WHSV) of 1 per hour, H2/oil ratio of 400 N and pressure of 80 bar.
lP

Whilst, Zikri and Aznury [70] employed the NiMo/γ-Al2O3 catalyst in the hydrogenation of
CPO. The maximum yield was, 68.2%, obtained at 315 °C. Based on these two studies, the HDO
route of vegetable oils had a better performance than the hydrogenation route because of the
na

higher resulting yield. The HDO of RSO is an attractive and promising process from an
economic perspective to produce green diesel as reported by Cheah et al. [119].
The effect of different reaction methods (i.e. HDO and DCX) has been reported by Ameen
ur

et al. [92] in converting RSO into green diesel using NiMo/γ-Al2O3 catalyst. Both HDO and
DCX successfully resulted in 100% conversion. Whereas, the selectivity of DCX (81.7 wt.%)
Jo

was markedly higher than HDO (19.1 wt.% selectivity) at 400 °C, 80 bar, WHSV of 1 per hour,
and H2/oil ratio of 400 N.
On the other hand, Scaldaferri et al. [77, 95] converted soybean oil into green diesel using
niobium phosphate (NbOPO4) catalyst under hydroprocessing and DO methods at 300 – 500 °C,
10 bar, with a catalyst content of 0 – 25% within 3 – 5 h. The hydroprocessing production
method of soybean oil produced a green diesel yield of 40% [95], whereas the DO of soybean oil
had a lower yield (15%). Hence, choosing an appropriate reaction method is very crucial to
achieve the best results.
According to these observations, the highest green diesel yield was produced by the DCX
process followed by HDO, hydrogenation, hydro-processing and DO processes in spite of using
different feedstocks. The authors claim that the yield produced under the DO process can be
upgraded using the nickel-phosphides-modified heterogeneous catalysts.

2.2 The role of metal oxides

of
The improvement of catalyst performance is usually carried out by doping metal oxides (i.e.
Pd, Pt, Ni, Co, Cu, Mo, Mn, Mg etc.). Improved catalytic performance with doping has led to
investigation of such modifications by many authors [4, 85, 98]. For example, Aliana-

ro
Nasharuddin et al. [91] studied the role of metal oxides (Ni-Mg, Ni-Mn, Ni-Cu, and Ni-Ce)
impregnated onto multi-walled carbon nanotube (MWCNT) catalysts for green diesel production

-p
with DO of chicken fat oil. The Ni-Mg-assisted MWCNT catalyst produced a 75% yield,
whereas the Ni-Mn-doped MWCNT catalyst resulted in an 81% yield. Higher yields were seen
re
when more metal oxide was used. The DO capability of metal-oxide-supported catalysts is
caused by lower acidity and stronger basicity [91] and larger active surface sites on the catalysts
lP

[98, 121, 122].


The MWCNT catalyst, supported by metal oxides of Fe and Zn, was investigated by Asikin-
Mijan et al. [71] in the DO of jatropha curcas oil. The acidity and basicity were strongly
na

enhanced by increasing the Fe and Zn contents leading to an improved yield (up to 79%) [71].
Elsewhere, the influence of Mn-Al metal doping into a NiMO catalyst was studied by Garrido et
al. [76]. The superior performance of the Mn-Al-assisted NiMO catalyst was affected by
ur

stabilizing Mn oxide onto the Al2O3 to redisperse the MoS2 active sites assisted by Ni oxides
[76]. An Al2O3-assisted NiMO catalyst was compared with other catalysts (i.e. Pd, CoMo, Ni, Pt
Jo

and Ru) also impregnated with Al2O3 by Veriansyah et al. [69]. The best catalyst performance
was achieved by NiMo catalyst (92.9%) followed by Pd (91.9%), CoMo (78.9%), Ni (60.8%), Pt
(50.8%), and Ru (39.7%) [69].
Ochoa‐Hernández et al. [73] have used Zr-SBA-15 catalyst supported by Ni and Co oxides.
The conversion and yield achieved by Zr-SBA-15 catalyst impregnated by Co oxide were higher
than those obtained over the Zr-SBA-15 catalyst doped by Ni. The better performance of the Co-
doped Zr-SBA-15 catalyst was correlated with a more extensive interaction between Zr and Co
content either on the surface or in the pores of the catalyst [73]. The synergistic effect of Zr and
Co content has also been reported by Li et al. [123] and Yanyong et al. [124]. Whilst,
Kamaruzaman et al. [48] compared SBA-15 catalyst supported by 5 wt.% Ni, Co, and Ni-Co.
The yields obtained from the Ni-Co/SBA-15 and the Ni/SBA-15 catalysts were 85.8% and
88.1%, respectively. Whereas the Co/SBA-15 catalyst had an unsatisfactory performance due to
the bigger particle size and the formation of carbon and coke, leading to the closure of pore sites
and the deactivation of the surface active sites. Co oxide showed a positive effect on the Zr-

of
SBA-15 catalyst, but a negative effect on the SBA-15 catalyst.
Recently, Srihanun et al. [75] explored a Pd-Al2O3 catalyst supported by an Fe loading
where the Fe/Pd-Al2O3 catalyst could enhance the active sites, pore diameter, and pore sizes of

ro
the catalyst leading to an increase in the green diesel yield. The yield obtained from the Pd-Al2O3
catalyst was 86% whereas the yield achieved from the Fe/Pd-Al2O3 catalyst was 94%.

-p
Based on the study, metal oxides play a crucial role in the catalytic performance and stability
leading to an increase in conversion, selectivity, and yield. The increase is caused by larger
re
active surface sites of the catalysts and higher Lewis-Bronsted acidity and basicity when the
metal oxides are added to the catalysts. Among those metal oxides, Ni oxide was widely applied
lP

in the modified catalysts and crowned as the best metal that could provide a better synergistic
effect and enlarging the accessibility for the reactants in the surface area as well as triggering a
high yield.
na

2.3 The role of nickel phosphides


Nickel phosphides (Ni2P) have also been shown to improve catalytic performance and
ur

stability and thus have been widely explored. Excellent stability [4] and better activity caused by
larger surface area sites have been reported [125]. Camargo et al. [93] investigated three different
Jo

catalysts, namely Al-SBA-15, ZSM-5, and USY catalysts synthesized with Ni2P. The catalytic
conversion performance was rated as Ni2P/Al-SBA-15 ≥ Ni2P/ZSM-5 ≥ Ni2P/USY. Moreover,
the Ni2P-doped Al-SBA-15 activity in DO of vegetable oils was much better than the metal-
oxides-assisted Al-SBA-15 activity. Ni2P/USY catalyst was also applied in DO of oleic acid by
Kochaputi et al. [96] and a 100% conversion was successfully achieved. The HY catalyst
supported with Ni2P was comprehensively studied by Zarchin et al. [126] where the Ni2P/HY
catalyst revealed a good isomerization and hydrocracking performance.
On the other hand, Phimsen et al. [127] compared Ni2P with nickel sulfide (NiS) and nickel
carbide (NiC) for hydrotreating oils into green diesel at 375 – 425 oC, 20 – 40 bar within 3 h.
Even though the best catalyst activity was achieved by NiC followed by Ni2P and NiS, the
highest green diesel yield was achieved by Ni2P followed by NiS and NiC. The yields obtained
at 20 bar from Ni2P, NiS, and NiC were 36.5%, 33.7%, and 17.8%, respectively. At 40 bar the
yields were 33.8%, 33.4%, and 23.3%, respectively. Smith carried out a similar comparison

of
between Ni2P and NiC catalysts.
A Ni2P-supported activated carbon (AC) catalyst was carbonized inside a tube furnace
(Ni2P/AC-TF) and an Iwasaki kiln (Ni2P/AC-IW) as reported by Ruangudomsakul et al. [128].

ro
The Ni2P/AC-IW had a better capability than the Ni2P/AC-TF, because it possessed a broader
accessibility for oils to the surface active sites of the catalyst. Jeong et al. [129] compared

-p
between Ni2P/SiO2 and Pt/Al2O3 in the hydroprocessing of palm oil. The Ni2P/SiO2
outperformed the Pt/Al2O3 due to less formation of carbon sediments and more substantial
re
Brønsted acid sites. The Pt/Al2O3 yielded a lot of carbon deposits leading to lowered DCX and
DCN performances caused by the reduced Pt active sites.
lP

Elsewhere, Liu et al. [100] observed the role of Ni2P in DO of palmitic acid at 350o C and 10
bar using ZSM-22 and compared ZSM-22 (without doping Ni2P) and Ni2P/ZSM-22. The highest
yield was achieved by Ni2P/ZSM-22 (45.7%), whereas the yield obtained by ZSM-22 was only
na

11.6%. Ivars-Barceló et al. [130] made a comparison between several pure noble metal catalysts
such as Pt, Ru, Ir, and Rh and those doped by Ni2P in the HDO of oils at 300 oC and 3 bar. The
researchers reported that the noble metal catalysts synthesized by Ni2P had a better performance
ur

than the pure noble metal catalysts. The HDO percentages of the Pt, Rh, and Ir catalysts were
20%, 40%, and 45%, respectively, compared with 30%, 90%, and 90% respectively for the Ni2P-
Jo

assisted Pt, Rh, and Ir catalysts. The suitable performance of the Ni2P-assisted noble metal
catalysts was attributed to more spacious surface H-active sites of the catalysts [130].
In another study, Rupflin et al. [131] compared Rh2P/SiO2, Rh/SiO2 and Ni-Mo/Al2O3
catalysts in hydrodesulfurization (HDS) in a batch reactor at 275 oC, 30 bar and 3000 rpm.
Among the catalysts, the highest conversion was obtained by the Rh2P/SiO2 catalyst (99%)
followed by Ni-Mo/Al2O3 (84.2%) and Rh/SiO2 (70.1%). The catalyst using phosphides (P) was
better than the ones doped with nickel (Ni), presumably due to sintering of NiO into the surface
active sites of the catalysts [131]. Meanwhile, a study conducted by Xin et al. using AC catalyst
impregnated with Ni2P and Ni12P5 for DO of palmitic acid with various Ni:P molar ratios
exhibited highly superior catalytic capability, attributed to a synergistic effect of Ni2P and Ni12P5
contents on the catalyst leading to increased formation of the desired products [132]. It seems
that doping with Ni2P could be an important key factor for enhancing catalytic capability.
It can be concluded from the mentioned studies that suitable reaction routes, the use of metal
oxides, and Ni2P-doping have a positive influence on catalytic behavior leading to maximum

of
quality and quantity of products. The Ni2P-dopped zeolites are more superior over the Ni2P-
dopped synthetic catalysts.

ro
-p
re
3. Green diesel production using zeolite catalysts
Currently, zeolites are highly used in the conversion of edible and non-edible oils into
lP

biofuels due to their highly accessible pore sites for the reactions to take place [56]. To optimize
the zeolite activities in green diesel production, some supports such as Ni2P, Pt, Al, Fe, Cu etc.
have been widely applied so that the zeolites can be reused several times with maintained
na

performance [105, 119, 133-135]. By the year of 2021, the use of zeolites as catalyst gains a
great attention to produce green diesel from vegetable oils because they are more cost-effective,
green catalyst, and plenteous. For instance, Zikri et al [136] harnessed natural zeolite as a
ur

catalyst to produce green diesel from crude palm oil (CPO) under hydrogenation technology at
pressure of 0.7 – 2 atm. Different zeolite contents of 1 wt.%, 2 wt.%, 3 wt.%, and 4 wt.% are
Jo

further investigated at various reaction temperatures of 350, 375 and 400 oC. Approximately,
37.30% of green diesel yield was achieved at 400 oC, 2 bar, and 3 wt.% of zeolite content. The
physical and chemical properties of green diesel produced have a similar property with diesel
fuel including density of 782 – 808 kg/m3, kinematic viscosity of 2.24 2.53 cSt, and flash point
of 55.1 – 58.5 oC [136].
Based on the investigation, natural zeolites are potential and promising catalysts to use in
green diesel production. The low yield is a great opportunity for others to improve the operating
process thus the obtained yield can be significantly enhanced. As investigated by Suhendi et
al[137] who studied influence of activation time on performance of Bayah natural zeolite from
Banten – Indonesia in palm oil shell pyrolysis process. The good performance was reached over
the longer reflux time and higher catalyst content[137]. The performance of Bayah natural
zeolite was also studied by Mahdi et al [138] in biodiesel waste treatment. The smaller zeolite
diameter size (0.46 mm), 9 wt.% of zeolite-to-reactant ratio, 3 wt.% of water content, and
reaction temperature of 100 oC played a crucial impact in increasing the conversion [138].

of
Not only natural zeolites can be implemented, but also synthetic zeolites are widely applied
in green diesel production. There have been several types of zeolites employed in green diesel

ro
production, namely USY, HY, ZSM-5, ZSM-22, SAPO-11, beta and others. For instance,
Camargo et al. [93] used USY, ZSM-5 and SBA-15 impregnated with Ni2P in a batch reactor at

-p
260 – 300 oC and 50 bar with a residence time of 6 h. The optimum green diesel yield was
achieved by Ni2P/SBA-15-(42%) followed by Ni2P/ZSM-5-(29%) and Ni2P/USY-(24%). The
re
Ni2P/SBA-15 showed the best performance due to having a higher pore diameter (7.23 nm) than
the kinetic diameter of oleic acid (0.55 nm) [139], thereby increasing the accessibility of the
lP

reactants to the active sites of the catalyst. The Ni2P/ZSM-5 and Ni2P/USY had less micropores
which caused limitations for the reaction to take place on the external crystal surface leading to
lower yields [140].
na

Table 3 summarizes some recent green diesel production studies using zeolites as the
catalysts. The table does not include studies in which zeolites have been used in the production
of other products such as isobutylene [134] and C4 olefins [135].
ur
Jo
Table 3. Green diesel production from different zeolites.

of
Operating parameters
Process Results
Zeolites Oils T P t Remarks Ref
Methods Supporting (%)
(oC) (Bar) (h)

ro
Crude 350
0.7 – The higher catalyst content and higher reaction temperature
Natural Palm Oil Hydrogenation – 3 - Y = 37.30 [136]
2.1 contributed to the increasing green diesel yield
(CPO) 400

-p
The synergistic interaction between La-Si-Al, enlarger pore
La, Co, Fe X = 99
Oleic Hydrodeoxyge diameter, and stronger acid sites of the La-supported beta
Beta 350 40 2 Mg, Mn, and S = [141]
acid nation catalyst contributed to enhance the activity due to possessing
and Zn 83
the larger accessibility of the reactant

re
Palm Hydroconversi The Pt-assisted Ni/NH4-beta catalyst improved the
Beta 360 40 4 Pt-Ni/NH4 Y = 28.7 [142]
olein on performance leading to obtain the highest yield
Beta and Macauba Beta zeolite resulted in higher conversion than the one
ZSM-5
Y

HZSM-
oils
Oil
sludge
Oleic
Deoxygenation
Hydroconversi
on
350

160
10

50 lP 3

18
-

Ni
X = 100

Y = 90.6
obtained over ZSM-5
The Ni-modified Y zeolite had a good reusability during 5
batches
The comprehensive mesopores of HZSM-5 contributed to
[143]

[144]
na
Deoxygenation 360 - 1 - Y = 65 enhance larger acid site accessibility leading to increasing [145]
5 acid
the yield
Al- The P-supported Al-MCM-41 exhibited the best
Oleic La, Ni, P,
MCM- Deoxygenation - - - Y = 82.30 deoxygenation activity due to high Bronsted-Lewis acidity [146]
acid Zr, and Ce
ur

41 leading to increase of hydrocarbons content


USY 260
Oleic Y = 24 USY and H-ZSM-5 were modified by nickel phosphide
and H- Deoxygenation – 50 - Ni2P [93]
acid and 29 (Ni2P)
Jo

ZSM-5 300
X = 80; The nanosized NaY assisted by NiO nanoparticle was
NaY Triolein Deoxygenation 380 - 1 NiO [46]
S = 93.7 synthesized over deposition–precipitation and impregnation
Ni, Cu, Co, X= HY zeolite was supported by transition metals (i.e. Ni, Cu,
HY Triolein Deoxygenation 380 - 2 Zn, and 76.21; Co, Zn, and Mn-Y) and the best catalyst performance was [47]
Mn-Y S = 92.61 the Ni-Y zeolite
Green diesel production from palmitic acid was obtained
Palmitic Hydrodeoxyge Ni and
ZSM-5 300 35 4 X = 99 over hydrodeoxygenation and hydroisomerization using Ni- [147]
acid nation Ni/Mo

of
Mo(-R)/ZSM-5
SAPO- Stearic Ni/SAPO-11 was synthesized under a dry gel supported by
Deoxygenation 290 40 3 Ni X = > 95 [148]
11 acid microwave method

ro
Methyl Hydroprocessi X = 77; Beta zeolite doped with nanosized Ni was synthesized over a
Beta 240 25 4 Ni [89]
palmitate ng S = 61.9 sol-immobilization route
> Different heating rates (2, 6, 10, 14, and 18 oC/min) were
ZSM-5 HDPE Cracking 525 - - Y = 61.05 [149]

-p
1 implemented in a pyrolysis reactor
X=
Various nanosized HY zeolites (i.e. 20 nm, 65 nm, 380 nm
HY Triolein Deoxygenation 380 0.01 4 - 71.16; [150]
and 2.75 μm) were employed
S = 72.15

re
X = > 99;
Rubber Hydrodeoxyge Y = 30;
HY 350 35 0 Ni The increasing yield was enhanced over higher Ni content [94]
seed oil nation S = 6.8 –

Beta
Palm
kernel oil
and palm
Deoxygenation
, cracking and
isomerization
350 10
lP 0 -
14.5

X = 96; Beta zeolite was applied without using metal oxides [107]
na
olein
Hydrodeoxyge X = 58
INZ Palm oil 375 12 2 Fe INZ was supported with 3 wt.% of Fe content [151]
nation and 89
Soybean X = 99.5; Ion exchange method with ammonia was explored to
ZSM-5 Hydrocracking 450 1.27 2 - [152]
ur

oil S = 42.11 enhance the catalyst acidity


n- 260 50
Hydroisomeriz X = 36.5; Different catalyst contents were employed to study the pour
Beta hexadeca – and 0 Pt/Al [153]
ation Y =35.5 point of the desired products
ne 320 100
Jo

HY zeolite doped with Ni content was synthesized over


Stearic Hydrodeoxyge X = 100;
HY 300 30 2 Ni impregnation method so that the Ni-Y zeolite could be [154]
acid nation S = > 20
reused several times
Note : T = Temperature; P = Pressure; t = time; X = Conversion; S = Selectivity; Y = Yield; Ni 2P = nickel phosphide; HDPE = high density polyethylene; INZ =
Indonesian natural zeolites (i.e. modernite and clinoptilolite).
In order to improve ZSM-5 performance in green diesel production, several methods
have been implemented as reported by others. For instance, Nugraha et al [155] used structure
directing agents (SDA) on mesopore/micropore structure of aluminosilicates (ALS) in green
diesel production under deoxygenation technology. Indonesia kaolin was firstly transformed into
the ALS under two-step hydrothermal, which used tetrapropylammonium hydroxide (TPAOH)
as SDA in the first step and utilized cetyltrimethylammonium bromide (CTAB) in the second
one to induce mesopores. Thereby, jacking up the conversion and green diesel selectivity was
attributed to higher surface active sites, enlarged mesopore volume, and higher acidity [155].

of
ZSM-5 was also observed by Freitas et al [143] in 2021 in green diesel production using
macauba oil under deoxygenation technology. The reaction was carried out at 350 oC, 10 bar,
and 200 rpm in a batch reactor within 5 h. The highest deoxygenation efficiency achieved was

ro
80%. By dint of the upper sustainability and economic feasibility, the green diesel production
using macauba oil as the feedstock and ZSM-5 as the catalyst is totally worthed to apply in

-p
industrial production scale in order to reduce the production cost and to bring down the
environmental pollution. Beside ZSM-5, this study also observed beta as the catalyst and
re
reavaled that the beta genarated higher deoxygenation efficiency (100%) than the one obtained
over ZSM-5 [143].
lP

Because beta had much more performance than ZSM-5 ability in green diesel production
using macauba oil under deoxygenation technology as reported by Freitas et al [143]. Hence,
Azreena et al [141] utilized the beta as the catalyst to produce green diesel using oleic acid under
na

hydrodeoxygenation process in which the beta was modified using metals such as La, Co, Fe,
Mg, Mn, and Zn in order to jack up the beta ability. Among these metal supports, the La-
supported beta was the best one and resulted in the highest conversion (99%) and selectivity
ur

(83%) obtained at 350 oC and 40 bar within 2 h. The La-modified beta indicated a superior
performance related to the better synergistic interaction between La-Si-Al, higher acid sites and
Jo

enlarger pore diameter. The enlarged porosity can affect higher diffusion of bulky molecules
leading to the more comprehensive accessibility of the reactant and perfect catalyst capability
[141].
Instead of metals of La, Co, Fe, Mg, Mn, and Zn, Chintakanan et al [142] employed Ni
metal supported on beta zeolite to produce biofuels such as bio-gasoline, bio-jet fuel and green
diesel using palm olein under hydroconversion technology at 360 oC and 40 bar within 4 h. The
Ni-assisted beta was very effective and efficient catalyst in biofuel production related to
improving conversion and biofuels yield [142]. Meanwhile, Silva et al [156] draw upon Aspen
Plus-8.8 software to produce green diesel using batch reactor at 300 oC, H2-to-oil molar ratio of
2.5:1, and 40 bar under deoxygenation technology. This Aspen Plus-8.8 software outcome 96%
of conversion of vegetable oils and 78% of green diesel yield and further assisted economic
anaylysis in green diesel production [156].
High silica-modified zeolite beta (HSZB) was successfully employed by Fereidooni et al
[157] using waste sunflower oil under trans-esterification technology. Tetrahydrofuran (THF)-20

of
wt.% was applied in as a co-solvent for mass transfer intensification. The highest conversion
produced over HSZB was 96.5% reached at methanol-to-oil ratio of 9:1, 2 wt.% of catalyst, and
300 rpm within 4 h in which the HSZB was reused during three cycles with maintained

ro
performance [157]. The 24 mol%-optimum yield was achieved using Zr-Al-modified beta
attributed to higher Bronsted-Lewis acidity as studied by Paniagua et al [158].

-p
On the other hand, mesoporous Y zeolite (MPYZ) harnessed to convert oil sludge into
green diesel was supported over 10% of Ni as observed by Kang et al [144]. The Ni-assisted Y
re
zeolite had good activity thus it resulted in high yield (90.6%) achieved at 160 oC and 50 bar
[144]. The higher stability and stronger acidity were caused by possessing three dimensional
lP

channel types [159]. While, Kochaputi et al. [96] compared USY assisted with Ni2P and CoP for
DO of oleic acid in a batch reactor at 260 – 340 oC, 50 bar, and 6 h. CoP/USY gave higher yield
than Ni2P/USY presumably due to larger active sites and higher acidity. Liu et al. [100] used
na

Ni2P/ZSM-22 for DO of palmitic acid at 310 – 360 oC and 10 bar with a reaction time of 2.5 h
and claimed that the Ni2P facilitated the DCX and DCN reactions. Brønsted and Lewis acid sites
on ZSM-22 had a positive impact on the isomerization and cracking processes. Thus, the
ur

synergistic role of Ni2P/ZSM-22 revealed excellent catalytic activity.


Lee et al. [147] used ZSM-5 doped with Ni and Ni-Mo(R), observing that the Ni-Mo(R)-
Jo

ZSM-5 resulted in 99% conversion and 70% yield at 300 oC for 4 h. The catalyst simultaneously
facilitated hydrodeoxygenation and hydroisomerization due to carbonyl oxygen adsorbed into the
Mo(R), which possesses a higher affinity and a good absorption capacity of hydrogen atoms into
Ni. ZSM-5, has been applied in the CC of HDPE where various heating rates and reaction times
impart an impact on the obtained yield as reported by Ghaffar et al [149]. The ZSM-5 lifetime
has been shown to increase by employing an ion exchange method with ammonia leading to
longer reaction times and increasing the yield as implemented by Zandonai et al. [152].
Apart from those catalysts, HY activity has been shown to increase with the addition of
transition metals (i.e. Ni, Cu, Co, Zn, and Mn-Y) as investigated and reported by Choo et al.
[47]. Among the transition metals studied, Ni-Y was the most effective and efficient catalyst in
producing green diesel with 76.21% conversion and a selectivity of 92.61%. The favorable
performance of Ni-Y is attributed to a synergistic effect between the lowered Bronsted/Lewis
sites and the strong hydrogenolysis capability of Ni to facilitate DCX reactions by minimizing

of
cracking and polymerization reactions. This finding was also confirmed by Ameen et al. [94]
who achieved higher yields by increasing the Ni content due to strong acid sites. The Ni-Y
catalyst could be reused several times as a result of the larger active sites and higher

ro
accessibility caused by the larger surface area and larger pore size as proposed by Hachemi et al.
[154].

-p
Beside the Ni content, nano-sized HY has shown outstanding capability as reported by Choo
et al. [150] who studied the effect of nano-sized HY with crystal sizes of 20, 65, and 380 nm as
re
well as larger 2.75 µm sized crystals. The smaller crystal size increased the conversion of oils to
green diesel due to stronger Brønsted and Lewis acidity, higher porosity, preferable accessibility
lP

and larger channel structures [150, 160]. These factors immensely affect catalyst performance
and stability [161]. Hence, the role of Ni content and nanosized HY play a positive impact on
catalyst ability. Choo et al. [46] showed that a smaller particle size led to better accessibility and
na

higher Brönsted -Lewis acid sites leading to high conversion (80%) and green diesel selectivity
(93.7%).
The Ni content resulted in favorable performance not only for HY but also for SAPO-11 as
ur

observed by Liu et al. [148], as for Ni-beta as investigated by Papanikolaou et al. [89]. Liu et al.
[148] utilized a dry gel synthesis method with microwave radiation at 120 – 180 oC for 2 h for
Jo

SAPO-11 preparation. The Ni-SAPO-11 synthesized with the dry gel-based microwave method
resulted in the conversion of more than 95%. Meanwhile, Papanikolaou et al [89]. harnessed the
sol-immobilization method for beta synthesis resulting in a conversion of 77% and selectivity of
61.9%. The noticeable ability of the nano-sized Ni-beta was due to the preferable accessibility of
the acidity caused by dealumination [89]. Approximately, 96% conversion using the beta at 350
C, 10 bar, and 5 h was achieved without using metal oxides [107]. Thus, the influential
parameters for zeolite performance in green diesel production are nickel content, metal oxides,
catalyst synthesis, and nano sized particles. The best zeolite producing the highest green diesel
yield and widely harnessed these days was the modified-beta zeolite related to the stronger
Bronsted and Lewis acid-base, enlarged porosity, easier accessibility and larger channel
structures.

of
4. Green diesel producing technologies
Green diesel technology using heterogeneous solid synthetic catalysts and natural catalysts

ro
like zeolites has been extensively investigated by many researchers both at laboratory scale and
at industrial scale for decades. Green diesel fuel has been produced with many processes

-p
including supercritical treatment [162, 163], pyrolysis [164], gasification, hydrogenation [162],
Fischer-Tropsch synthesis [165], deoxygenation, alkali catalyzed process [162], liquefaction, and
re
hydrotreating [166] technologies. Table 4 compares different operating parameters for green
diesel production under four main processes, namely supercritical alcoholysis (SCA),
lP

hydrogenation of waste vegetable oils (WVO), homogeneous alkali catalyzed alcoholysis


(HACA), and deoxygenation (DO). Among these technologies, the deoxygenation technology is
widely applied in the conversion of vegetable oils and animal fats into green diesel and is well
na

known as the most effective and efficient technology [66] for the time being.
Also, either the transesterification reaction under SCA or the WVO hydrogenation reaction
are nowadays readily implemented in green diesel production using WVO with high FFA
ur

content. On the other hand, HACA is extensively applied in green diesel production using oils
with low FFA content [162]. To lower the oxygen content in triglycerides, DO technology is
Jo

suitably implemented. This is because a high oxygen content can interrupt engine combustion
efficiency and lead to corrosion [42, 66].
Technologies of SCA, hydrogenation of WVO and HACA are industrially employed to
produce biodiesel. Nevertheless, they could also be implemented in green diesel production as
reported by Glisic et al. [162]. Meanwhile, DO technology is particularly applied for green diesel
production using different vegetable oils and catalysts [167-169] due to less energy consumption
and higher conversion, selectivity, and yield. The catalysts can also be reused, which leads to
reduced production costs.

of
ro
-p
re
lP
na
ur
Jo
Table 4. Comparison of operating parameters for green diesel production using different technologies.

f
oo
Parameters Hydrogenation of WVO Supercritical process, SCA Alkali catalyzed process, HACA Deoxygenation process, DO
Capacity (tonnes/year) 100,000 100,000 100,000 90,000 [170]
Temperatures (oC) 390 300 60 350
Pressure (Bar) 138 200 4 40

pr
Reaction time (h) 1.1 1.1 1.8 >1
Conversion 96 97 95 100
Yield - - - > 80

e-
Selectivity - - - > 80
Catalysts CoMo No NaOH Heterogeneous catalysts
Energy consumption 73,608 (MW/year) 63,320 (MW/year) 93,040 (MW/year) 44 Mj/kg [13, 21]

Pr
1. Less energy consumption
2. Simple technology
3. Higher conversion,
1. Lower operating conditions selectivity, and yield
l
1. Less environmental
1. Higher conversion
(safer) 4. Highly reusable catalysts
na
pollutions 2. Cost-efficiently 5. Lower production cost
Advantages 2. Suitable technology for high
2. Suitable technology for high 3. Cost-environmentally 6. Less environmental pollutions
FFAs content
FFAs content 4. Suitable technology for high 7. Suitable technology for the
FFAs content oils containing high FFAs and
ur

oxygen contents
8. Producing other biofuels like
biogasoline and biojet fuel
1. Higher operating conditions 1. Separation process is required to
Jo

1. Purification and separation


(higher explosive risk) remove the remaining catalyst
1. Higher energy consumption processes are required to
2. The formation of gas and coke after the reaction
2. Unsuitable technology for separate between the desired
Disadvantages can lower the green diesel yield 2. The process produces substantial
the oils containing high products and by-products
3. Unsuitable technology for number of wastewater
oxygen content 2. H2 make up is required
the oils containing high oxygen 3. Purification and separation
3. Wastewater is produced
content processes are required to separate
between the desired products and 4. Pretreatment of oils is

f
by-products required to remove the

oo
4. Higher energy consumption impurities
5. Unsuitable technology for the
oils containing high oxygen
content

pr
Ref [162, 171] [162, 172, 173] [162, 174, 175] [2, 14, 167-169]
WVO = waste vegetable oil; SCA= supercritical alcoholysis; HACA = homogeneous alkali catalyzed alcoholysis; DO = deoxygenation;
MW = Megawatts; Mj= Megajoules.

e-
l Pr
na
ur
Jo
In a typical WVO hydrogenation flowchart (Figure 3), there are four main pieces of
equipment, namely, a heat exchanger (HE), furnace, reactor, and distillation column. WVO and
hydrogen are heated by the HE followed by the furnace up to desired temperatures. Then, the
mixture reacts inside the reactor in the presence of catalysts such as CoMo at 390 oC and at a
pressure of 138 bar [162]. The products coming out from the hydrogenation reactor are separated
in a separator resulting in water, gas (i.e. CO, CO2, propane and light hydrocarbons), and liquid
hydrocarbons [2]. The liquid hydrocarbons are then separated in a distillation column producing
light hydrocarbons (i.e. C3 and C4), bio-gasoline, green diesel, and residue at 1 bar [162]. The

of
WVO hydrogenation technology requires an approximate energy consumption of 73,608
MW/year for a plant capacity of 100,000 tons/year [162, 171].

ro
-p
re
lP
na
ur
Jo

Figure 3. Simple WVO hydrogenation process for green diesel production adapted
from Glisic et al [162].

Production of green diesel using the WVO hydrogenation process is a highly innovative
technology to mitigate environmental pollution caused by massive waste cooking oil production
in many countries such as the UK, producing 200,000 tons/year, and the EU countries, producing
up to 1 million tons/year [172]. Thus, sustainable and renewable diesel fuel, namely biodiesel
and green diesel produced from waste cooking oil have become a topic of focus [176].
The transesterification process carried out in SCA is non-catalytically operated at 240 – 270
C, and 200 – 280 bar [162, 175] where the supercritical gas phase forms at 270 oC and 200 bar
o

[172]. The SCA process has been extensively studied often showing first order reaction kinetics.
Kusdiana and Saka [177] transesterificated rapeseed oil under subcritical (at 200 oC) and
supercritical (at 500 oC) processes and concluded that the conversion was dramatically enhanced
under the supercritical process at the optimum temperature of 350 oC [177]. In other studies,

of
subcritical and supercritical processes studied by He at al [178] were conducted at 87 bar and
360 bar for the transesterification of soybean oil. Not only the conversion, but also the activation
energies under the supercritical process were higher than the subcritical process with 56 kJ/mol

ro
and 11.2 kJ/mol, respectively. The kinetic designs for such systems can be accurately optimized
with the Genetic Algorithm method [179].

-p
Song et al. [180] experimented with non-catalytic transesterification of palm oil with SCA at
200 – 400 oC and found that the desired product was strongly increased when the critical point
re
was reached with a methanol to oil ratio of 30. Moreover, Silva et al. [181] revealed that the
conversion was reduced when the temperature and pressure were kept below the critical point.
lP

The better results achieved under the supercritical condition were principally influenced by the
substantial alteration of the physical characteristics of the alcohols (i.e. methanol and ethanol)
[178]. Also, Madras et al. [182] compared the transesterification of sunflower oil using SCA at
na

200 – 400 oC and 200 bar using both the thermal process and an enzyme catalyst. The
experiment successfully resulted in 100% conversion using the thermal process, while using the
enzyme catalyst only managed to achieve 30% conversion.
ur

Apart from SCA, supercritical water (SCW) has also been used to produce renewable diesel
fuel, for example in the hydrotreatment of microalga at 400 oC and 34 bar using Pd/C catalyst as
Jo

reported by Duan and Savage [183]. A low yield was achieved in this study, that was ascribed to
the formation of gas and coke due to longer reaction times and higher catalyst content.
Nonetheless, Dickinson et al. [163] used Pt/C catalyst in the DO of benzofuran under SCW and
reported that the product selectivity was enhanced by higher water loadings.
Green diesel produced from HACA technology using KOH or NaOH catalysts at low
pressure, temperature, and a methanol:oil ratio of 6:1 has been industrially applied for years.
Mittelbach and Trathnigg [184] studied the HACA process for producing green diesel from
sunflower oil using KOH catalyst at 25 – 60 oC with methanol:oil ratios of 3:1 and 3.3:1 and a
catalyst content of 0.5 – 1.5%. The yield increases with longer reaction times. The HACA
process is usually carried out at relatively low pressures (4 – 10 bar) and temperatures (60 – 80
o
C) using homogeneous catalysts (i.e. alkali and acid) [171] where a particular process is
required to remove the remaining catalyst after the reaction [172].
Drawbacks of using homogeneous catalysis that the catalyst is often hard to separate from
the products in the final process step. Therefore, the use of heterogeneous catalysts to replace

of
homogeneous catalysts in HACA technology is desirable, and which can improve the reusability
during several cycles, and decrease corrosion. Homogeneous catalysts also produce more
wastewater. Heterogeneous catalysts are more secure, cost-efficient and environmentally-

ro
friendly [172] and are thus have been examined by many researchers [185-187].
Gopinath et al. [185] utilized a heterogeneous solid acid catalyst (SO42-/Zr-KIT-6) in green

-p
diesel production from oleic acid and Jatropha oil at 120 oC, methanol:oil ratio of 20:1, and
catalyst content of 4 wt.% for 6 h. The reaction resulted in a 96% conversion of oleic acid and
re
85% conversion of Jatropha oil yielding 95% and 80% of green diesel for oleic acid and Jatropha
oil, respectively. Moreover, the catalyst could be reused three times with good performance and
lP

stability. Elsewhere, Hongloi et al. [187] employed other heterogeneous catalysts, namely Ni-
ZrO2, Ni-ZSM-5, and Ni-AC in the reaction undertaken at low pressure. The Ni-ZrO2 catalyst
gave the highest conversion (98.33%) and the optimum yield (76%) due to larger acid surface
na

sites on the catalyst.


Even though the WVO hydrogenation, SCA, and HACA technologies exhibit excellent
performance when FFA content is high, these technologies are not suitable for application for
ur

triglycerides with high oxygen content. Hence, DO technology can be implemented in the early
stage for feedstock with too much oxygen [188]. The DO technology is designed to lower the
Jo

high oxygen content that can be problematic to the engine combustion and cause corrosion in the
equipment. Some have gone as far as designating the DO process as the best available
technology for converting vegetable oils and animal fats into green diesel [42, 66].
Being a highly promising option, numerous researchers have been motivated to study DO
technology. Nasharuddin et al. [91] deoxygenated chicken fat oil using multi-walled carbon
nanotube (MWCNT) catalysts doped by metal oxides to increase DO activity. The Ni-Mn-
supported catalyst resulted in 84% yield and 85% selectivity. Whilst, Mijan et al. [188]
harnessed MWCNT catalysts impregnated by Ni-Fe and Ni-Zn in the deoxygenation of Jatropha
Curcas oil. The catalyst could be reused up to 6 cycles producing green diesel with properties
similar to commercial petroleum-diesel [71]. On the other hand, Mijan et al. [188] claimed that
the efficiency of the DO process using the NiCo-doped MWCNT catalyst was 80% resulting in a
green diesel yield of more than 76% and selectivity of over 60% with four cycles.
Other catalysts can also be employed in DO, such as Ag-La-doped AC catalyst as exploited
by Alsultan et al. [189] in the DO of waste cooking oil. The larger basic and acid surface active

of
sites on the catalyst boosted the DCX and DCN reactions and reinforced the DO performance
leading to increasing selectivity of green diesel (93%), good stability, and high reusability (up to
six cycles). DCX and DCN reactions promoted by supported-catalysts were also observed by

ro
Silva et al. [85] using a Pd/C catalyst in the DO of macauba. The highest green diesel yield was
85% achieved at 300 oC, 10 bar, and 700 rpm within 5 h. Meanwhile, using the PFAD byproduct

-p
of palm oil production to re-produce green diesel was carried out by Gamal et al. [45] using
CoMo-assisted AC catalyst and resulted in 92% hydrocarbon yield and 89% of C15+C17
re
selectivity. The CoMo-AC catalyst was reused up to six runs without catalyst deactivation and
maintained a selectivity above 80%.
lP

In addition, Alsultan et al. [190] conducted the DO of waste cooking oil using a Co-La-
doped AC catalyst in a batch reactor. The different catalyst contents played a significant role in
increasing green diesel selectivity. For instance, the maximum selectivity obtained with 3 wt.%
na

catalyst content was 63% while the best selectivity achieved with 1 wt.% catalyst content was
93%. Sousa et al. [107] used beta zeolite in the DO of palm kernel oil and palm olein in a batch
reactor at 350 oC, 10 bar, and 800 rpm for 5 h. The reported DO technology efficiency was 83%
ur

in the first reaction and 100% afterwards. The beta zeolite was reused in four cycles and resulted
in green diesel yield of 80% for palm kernel oil and 96% for palm olein.
Jo

According to these reports, the DO process is seen as having a markedly high performance
and capability for green diesel production because not only can it be implemented for high FFA-
content feedstock to reduce oxygen content, but it has also shown a good efficiency leading to
favorable conversion of oils, selectivity, and high yield of green diesel. Importantly, the process
uses less energy and the highly reusable catalysts reduce production costs. Furthermore, DO
technology is a simple process and produces other biofuels like biogasoline and biojet fuel as
well as resulting in less environmental pollution. Considering these advantages, DO technology
can be crowned as the best available technology.

5. Catalytic deoxygenation of palm oil and palm fatty acid distillate


As mentioned, DO is a process used to remove oxygen content of vegetable oils [191]
conducted over three different routes namely HDO, DCX and DCN reactions under H2-free
conditions [43, 115]. It is an important cracking and isomerization steps to produce green diesel
and bio-jet fuel [95, 192] possessing similar properties with petroleum-biofuels including low

of
frosting point, high heat of incineration [107, 193] and less carbon content [60]. The catalytic
DO process can be carried out with different catalysts including noble metals (i.e. Pt, Pd and Ru)
[4, 194], sulfide metals, Ni-Co-assisted zeolites, alumina, and activated carbon [31, 85, 99, 104,

ro
105, 195, 196]. For instance, noble and non-noble metals consisting of Pd, Pt, Co, and Ni
impregnated γ-Al2O3 have been utilized in the DO reaction of crude palm oil (CPO) and oleic

-p
acid [104], while the Co and Ni metals doped with silica (Si) have been used to produce green
diesel from methyl esters [195].
re
In the DO process of palm oil and PFAD to produce green diesel, other products such as
biokerosene and biogasoline are produced as well [197-199]. Biokerosene can be used in stoves
lP

to replace petroleum-based kerosene, while biogasoline can be applied in vehicles such as cars
and motorcycles as a substitution for fossil-based gasoline. The green diesel is widely utilized in
diesel engines to substitute petroleum-diesel fuel and biodiesel. These biofuel products are
na

produced from vegetable oils including palm oil and PFAD with the DO and CC processes [200].
Aromatic compounds such as benzene, toluene, and xylene (BTX) can be generated from
aromatization, isomerization and alkylation processes while by-products such as CO2 and CO
ur

gasses can be harnessed to produce methanol (CH3OH) with a syngas process. The process cycle
in green diesel production as depicted in Figure 4 shows a continuous route, where the main
Jo

products can be utilized for the transportation sector and the side-products are harnessed to
produce other intermediate products like methanol to reduce environmental pollution .
of
ro
-p
Figure 4. Simple flowchart of green diesel production from palm oil and PFAD [201,
202].
re
Makertihartha et al. [201] deoxygenated palm oil in a batch reactor at 500 oC with WHSV of
lP

2.5 per hour in 2 h using ZSM-5 zeolite. The authors reported 100% conversion of palm oil and
catalyst reusability of 3 cycles, ascribed to lower activation energy and a lower Gibbs energy.
The effect of activation energy and Gibbs energy in thermodynamics theory for higher
na

conversion and catalyst stability was exhibited by Taufiqurrahman and Bhatia [202] where a
faster reaction rate of green diesel production from palm oil was correlated with the presence of
triglycerides and oxygen, which very effectively formed carbenium molecules on the catalysts at
ur

a lower activation energy as investigated by Hamayel [203].


Nowadays, the use of palm oil in green diesel production as a substitute for commercial
Jo

petroleum-diesel and biodiesel receives a great deal of attention in the world, particularly in
Indonesia, Malaysia, and Brazil, because the palm oil production in these countries has been
significantly increasing for more than 30 years [204]. Brazil designated around 180,000 hectares
of land for palm oil plantations from 2010 to 2013 [205].The demand rate of Indonesia’s palm
oil production increased up to 28 Mha in 2020 [204] and is predicted to increase up to 84.2
MT/year in the year 2050 [33]. Hence, green diesel produced from the DO process of palm oil
positively contributes towards sustainability and renewability of biofuel production in the
transportation sector. The route for green diesel production from palm oil using DO technology
is illustrated in Figure 5.

of
ro
-p
re
Figure 5. Typical DO technology for green diesel production from palm oil [2, 14, 42].
lP

Kalnes et al. [42] reported that DO technology results in complete conversion (100%), high
selectivity, and high green diesel yield. However, it also produces side-products such as light
na

hydrocarbons (propane), water, and carbon dioxide (CO2). Therefore, a separator is required to
separate between the main products and by-products, where the by-products are recycled into the
feed [13]. The main products are then mixed with hydrogen, and are further reacted in catalytic
ur

hydro-isomerization yielding hydrocarbons in the range of green diesel. Afterwards, the desired
products are eventually produced in a distillation column producing four different products
Jo

namely light gases, naphtha, bio-jet fuel and green diesel [14, 42].
The produced green diesel has been evaluated as better than petroleum-diesel in terms of
saving fossil fuels (66 – 84%) and reducing pollution (41 – 85%). It also requires fewer
vegetable oils to produce one ton of green diesel [13]. DO technology generates around 85 wt.%
of green diesel, 7 wt.% of naphtha and 3 – 9 wt.% of light hydrocarbons whereas regular
biodiesel technology produces above 90 wt.% of biodiesel and 10-20 of the total volume of
biodiesel produced is made up of glycerol [13, 206]. After which the glycerol can be re-utilized
to produce glycerol carbonate as reported by Hilman et al. [138]. However, DO technology is
simple, uses less energy [167], is less polluting, and produces not only green diesel but also
biogasoline and biojet fuel [14].
The DO process of palm oil has been extensively reviewed by many studies and its residue
namely palm fatty acid distillate (PFAD) oil can be recycled to re-produce green diesel to
enhance the productivity using multiform catalysts [48-50]. A feasibility evaluation of green
diesel production from palm oil was reported by Nitipat et al. [170] in a techno-economic study

of
with a rate of return of 13.58 % and payback period of about 15 months.
Fangkoch et al. [207] employed MoO3 catalyst synthesized with 0.5 – 3% Pt metal content
for DO of palm oil. The initial acid site on the MoO3 catalyst was 24 μmol NH3/gcat while the

ro
acid active site possessed by Pt/MoO3 catalyst was in the range of 50 – 152 μmol NH3/gcat.
Supporting Pt onto the MoO3 catalyst contributed to higher acid active sites leading to an

-p
increase in catalytic performance as a consequence of a synergistic effect between the MoO3 and
Pt metal. Meanwhile, Kaewtrakulchai et al. [55] used KOH-activated CoP/PS (porous carbon)
re
catalyst synthesized with the microwave method. The highest particle spread sites and the largest
active sites on the catalyst pyrolyzed at 600 oC yielded the highest quantity of hydrocarbons.
lP

Meanwhile, the residue of palm oil production, namely PFAD oil, has been re-utilized by Gamal
et al. [45] to re-produce green diesel using CoMo-AC catalysts. The resulting selectivity and
yield of green diesel were 89% and 92%, respectively, and the CoMo-AC catalyst can be
na

recycled up to six runs.


A description of catalysts, operation condition, results and details of catalytic deoxygenation
of palm oil and palm fatty acid distillate for green diesel production are presented in Table 5.
ur
Jo
Table 5. Catalytic deoxygenation of palm oil and palm fatty acid distillate for green diesel production.

of
Operating conditions Results
Oils Catalysts Temp Pressure LHSV Catalyst X Y S Description Ref
R-cat
(oC) (Bar) (h-1) content (%) (%) (%)

ro
Microwave-supported activation is effectively applied
Palm 340 – 0.5 – to reduce reaction time. The CoP/PC effectively
CoP/PC 50 2.5 gram 100 77 58 - [55]
oil 420 1.5 enhanced the yield and selectivity optimized under

-p
operating parameters.
Higher yield was achieved by Ni-P/NaMOR catalyst
Palm Ni- 89. than that obtained from pure Ni2P and Ni2P/Ni12P5.
425 50 1 2.5 gram 100 - - [208]
oil P/NaMOR 5 The activity of the catalyst increased over the nickel

re
phosphides leading to higher yield.
Palm 98. The AC supported by Ni2P was synthesized in a tube
Ni-P/AC 350 50 1 4 gram 100 - - [128]
oil 3 furnace (AC_TF) and Iwasaki kiln (AC_IW).

Palm
oil
Ni/γ-
Αl2O3,
ZrO2 and
SiO2
250 –
400
20 – 30
1.2 –
4.8 lP
8 wt.% > 90 90 - -
Higher conversion was achieved by increasing
pressures. Catalyst deactivation occurred after 6 h of
reaction time due to blocking of some pores of the
catalysts.
[167]
na
Pt loadings significantly upgraded the catalyst activity
Palm 56. causing higher yield produced at 0.5 – 1% of Pt
Pt-MoO2 400 40 - 20 wt.% 100 - - [207]
oil 4 contents. The yield was reduced from 55.2% to 54.5%
at 2 – 3% of Pt loadings.
ur

0.06 – Pt/N-AC catalyst was applied in the reaction with the


Palm
Pt/N-AC 300 30 1.5 0.12 > 90 76 - - presence of nitrogen atom altering the electronic [209]
oil
gram density and assisting activity of electrophilic groups.
Jo

The hydrodeoxygenation process with methanol


synthesis yields a more profitable process compared
Palm Ni, Mo, or with other processes in an economic aspect with
300 50 - - 100 - - - [170]
oil NiMoS2 hydrogen recovery of 46.27% and reducing CO2 and
CO of 39.42% with rate of return (ROT) of 13.58%
and a payback period of 1.24 year.
The process was simultaneously carried out by
Al-
hydrocracking and deoxygenation including HDO,
Palm Ti/NiMo 270 – 0.8 – 92.

of
40 - 99 - - DCX and DCN processes in a batch reactor using [53]
oil and Al- 360 3.0 7
NiMo and CoMo catalysts with the presence of
Ti/CoMo
hydrogen.

ro
Palm NiMo/Al2 280 – 0.15 68. The best yield was produced at 315 oC which
40 – 90 1 - - - [70]
oil O3 380 gram 2 temperature over 315 oC lowered the yield.
Indonesia 58
Palm Fe content (3 wt.%) doped into zeolites played a

-p
natural 375 12 2 1 wt.% and - - - [151]
oil significant role towards the conversion.
zeolites 89
The catalyst was highly stable within a day due to
Palm 94. good-dispersed metallic Ni content of spinel structure

re
Ni-Al2O4 300 50 2 5.5 gram 100 - - [210]
oil 3 and the reduced temperatures gave a good
contribution towards the yield.
80 Beta zeolite without modifying with metal loading
Palm
oil

Palm
H-Beta

Ni/HBEA
350

260
10

40
-

-
lP
1.5 gram

0.2 gram
96

76
and
96

80
-

-
4

4
was a highly effective and efficient catalyst for the
deoxygenation of palm kernel and palm olein oils.
Broader accessibility for heavy particles was provided
by nano-sized Ni-HBEA causing better activity and
[107]

[211]
na
oil stability of the catalyst. The catalyst was reusable up
to 4 cycles.
92.
Ni/γ-
2 Ni/γ-Al2O3 had a better performance than the Co/γ-
Palm Al2O3 and
ur

300 50 1 5.5 gram 100 and - - Al2O3 due to the slower carbon deposition rate [212]
oil Co/γ-
88. causing a slight reduction in green diesel yield.
Al2O3
6
Co,Ni,Pd, 30 The γ-Al2O3 supported Co metal had better catalytic
Jo

Palm
and Pt/γ- 330 50 1 2.8 gram 100 – - - performance than the catalyst modified by other [104]
oil
Al2O3 60 metals. The TOF was escalated over larger metal size.
Larger surface area and higher mesoporous volumes
Palm Ni-SAPO- 79. provide the nano-sized Ni-SAPO-11 supplied enough
280 40 2.6 1 gram - - - [213]
oil 11 5 space and robust interactions in the dispersion of Ni
content.
Better performance on the catalyst was strongly
Atmosph caused by the synergistic effect between NiO and ZnO
PFAD NiO-ZnO 350 - 5 wt.% - 83 86 - [49]

of
ere due to the Brønsted and Lewis acidic sites on NiO-
ZnO.
Co/SBA-

ro
15, 85.
The lowered catalytic performance was induced by the
Ni/SBA- 8
Atmosph > deactivation of catalyst active sites and the blocking
PFAD 15, and 350 - 10 wt.% - and 5 [48]
ere 85 of the catalyst pores caused by the presence of cobalt

-p
Ni– 88.
due to formation of coke.
Co/SBA- 1
15
Atmosph 5 – 20 CoMo-AC showed perfect stability and resulted in

re
PFAD CoMo-AC 350 - - 92 89 6 [45]
ere wt.% yield and selectivity of over 80% within 6 cycles.
The reaction was carried out without solvent, and the
NiO/Al- Atmosph
PFAD 350 - 5% - 86 91 - selectivity of green diesel was slightly reduced at [50]

PFAD
SBA-15
NiO-
CaO/SiO
Al2O3
350
ere
Vacuum
pressure
(0.01)
- lP7% - 81 82 -
lower Si/Al ratio.
The acidity and basicity of the catalysts played a
crucial role to escalate the catalytic DO performance
as well as selectivity of green diesel.
[214]
na
PFAD = Palm fatty acid distillate; X = Conversion; Y = Yield; S = Selectivity; R-cat = Reusability of catalysts; LHSV = liquid
hourly space velocity; CoP/PC = cobalt phosphide/Porous carbon; NiP/NaMOR = nickel phosphide-mordenite in sodium
form; AC = activated carbon; TOF = turnover frequency.
ur
Jo
On the other hand, Baharudin et al. [49] deoxygenated PFAD using nano-sized Ni-Zn
catalysts (14 – 22 nm) with different Ni contents. Higher Ni content enhanced the crystallite size
of the catalyst leading to improvement in the Brønsted and Lewis acid active sites of the catalyst.
This is due to the synergistic impact between the active surface site and the Ni content. It was
claimed that PFAD was successfully converted into green diesel with a 83.4% hydrocarbons
yield and green diesel selectivity of 86%. In addition, Kamaruzaman et al. [48] used 10 wt.% of
SBA-15 catalyst supported by 5 wt.% of Co, Ni, and Ni-Co to re-produce green diesel from
PFAD to improve green diesel productivity. The hydrocarbon yields achieved by using Ni-SBA-

of
15 and Ni-Co-SBA-15 catalysts were 85.8% and 88.1%, respectively, as well as a green diesel
selectivity of more than 85%. The Co-SBA-15 catalyst produced smaller yields than the Ni-
SBA-15 and Ni-Co-SBA-15 catalysts due to larger size of the cobalt which caused the blocking

ro
of the catalyst pores and deactivation of the catalyst active surface sites by triggering the
formation of carbon and coke. Thus, the process significantly decreased the catalytic

-p
performance as well as the yield and selectivity obtained.
In addition, Baharudin et al. [50] also used Ni-SBA-15 catalyst to deoxygenate PFAD into
re
green diesel and obtained a 86% hydrocarbon yield and 91% green diesel selectivity using 9 – 10
nm Ni particles. Whilst, Mijan et al. [214] employed NiO-CaO/SiO2-Al2O3 and NiO/SiO2-Al2O3
lP

catalysts in an inert nitrogen flow that could preclude the catalyst active sites from poisoning.
The acidity and basicity of a catalyst can be hampered by poisoning on the active site thus
reducing the DO performance and lowering the yield and selectivity. Therefore, the utilization of
na

inert nitrogen flow and nano-sized catalysts gave a positive effect on the DO activity and the
results obtained.
So both palm oil and PFAD residue can be maximally harnessed to produce green diesel via
ur

DO technology on an industrial scale. DO technology improves productivity and is highly


attractive from a techno-economical perspective, promising, an effective and cleaner technology.
Jo

Nevertheless, several developments are still ongoing such as optimizing operating conditions,
catalyst treatments, metal oxide doping, phosphide loadings and nano-sized catalysts aiming to
improve the overall results and increasing reusability [167].
5.1 The effect of operating conditions
Different operating conditions such as temperatures, pressures, liquid hourly space velocities
(LHSV), reaction times, and catalyst content have been intensely studied to optimize catalytic
activity and stability [42, 43, 167, 194]. As a case in point, the effect of operating conditions in
the DO process of palm oil using heterogeneous catalysts, such as Al2O3, ZrO2 and SiO2
impregnated with Ni was thoroughly investigated by Papagerdis et al. [167]. In the investigation
carried out, higher pressure led to a positive impact on the increase of conversion for all types of
catalysts, whereas increasing reaction temperature also had favorable effects for each of the
catalysts at specific points (i.e. at 375 oC for Ni/Al2O3, at 300 oC for Ni/ZrO2, and at 350 oC for
Ni/SiO2). Catalyst deactivation occurred after 6 h of reaction time. Whilst, Pimenta et al. [215]
claimed that the reaction temperatures strongly affected the range of obtained aromatic

of
compounds. Also, the FFA concentration was shown neither to affect the DO efficiency nor the
reaction rate of catalytic activity.
Catalyst content plays an essential role in the reaction by speeding up the reaction, and

ro
improving the catalytic performance, activity, and stability [216, 217]. For instance, Mijan et al.
[188] maximized the reaction through applying different catalyst loadings (1–9%). Increasing

-p
DO performance was successfully achieved when certain catalyst contents, ranging from 1% to
5%, were used due to the larger active surface sites of the catalyst [218] with the highest yield
re
(80%) achieved at 5% catalyst loading [188]. However, more than 5% catalyst content had a
negative effect on the DO activity leading to lower yield caused by the probability of
lP

polymerization reactions [218, 219]. Moreover, Mijan et al. [220] agreed that the maximum
activity of the reaction occurred at 5 wt.% catalyst content using Ag-Ni/AC catalyst and with
Ag-Ni contents of 5 – 15 wt.%. The best operating condition conducted using 5 wt.% catalyst
na

content generated hydrocarbons yield of 78 – 95% and a green diesel selectivity of 82 – 83%.
The catalyst was able to be reused for up to five runs. More than 5 wt.% catalyst content did not
positively further influence the process due to cracking.
ur

Longer reaction times have been shown to have both positive and negative effects on yield
and selectivity. Mijan et al. [188] conducted catalytic DO under various reaction times (0.5 – 2
Jo

h) and observed a maximum of 80% at 1 h, which was slowly reduced as longer reaction times
were used. The reduced yield is caused by cracking of deoxygenated hydrocarbons producing
lighter molecules such as propane, butane and other fractions [188]. Pimenta et al. [215] claimed
that a longer reaction time would decrease the reaction rate (r), thus, reducing the yield according
to Equation 1 [215].
𝑚 𝑑𝑦𝐻𝐶
𝑟𝐻𝐶 = 𝑊 …………………………………………………………………….. (1)
𝑑𝑡
Where 𝑟𝐻𝐶 = reaction rate, 𝑚= mass of oil, 𝑊= mass of catalyst, 𝑑𝑦𝐻𝐶 = hydrocarbon
fractions obtained, and 𝑑𝑡 = reaction time.
However, Moreira et al. [84] showed an opposite result in which the conversion of Macauba
acid oil and the DO yield improved over longer reaction times. The conversions obtained in 2, 3,
and 4 h were 97.38%, 98.29%, and 98.80%, respectively, while the yields were 96.68%, 97.86%,
and 98.40%. Hence, the effect of longer reaction times is not straightforward, and it strongly
relies on the particular situation and the kind of catalyst used.
The contribution of reaction temperature is also not universal [67]. Increasing temperature

of
has been shown to improve the activation energy of the catalyst at certain temperatures, thereby
improving its capabilities. However, temperatures above the desired optimum point will overheat
the reaction and ruin the catalyst framework, thereby causing a decrease in the catalyst

ro
performance [134, 135]. In a study by Mijan et al. [188], an increasing yield occurred as the
temperature increased to 350 oC. However, as the temperature was further increased, the yield

-p
dropped at 400 oC. The decrease in yield and selectivity caused by an overheated reaction might
be due to the promotion of a second reaction leading to the production of light hydrocarbons as
re
reported by Arend et al. [221].
Mijan et al. [222] also studied the effect temperature in the DO process using CaO catalyst
lP

and reported that the yield was enhanced two fold from 27% to 54% at higher reaction
temperatures and conversions as high as 100% were obtained. Nevertheless, the yield and
selectivity significantly dropped to 45% at higher reaction temperatures (over 400 oC) caused by
na

cracking reactions during the DO process [55].


The effect of increasing pressure is also dependent on the particular process although it has
more often than not been positive [223]. Rakmae et al. [208] observed different pressures (i.e. 1
ur

– 50 bar) in the DO of palm oil using Ni2P/Na-MOR catalyst and concluded that green diesel
yield increased with pressure. Furthermore, Srifa et al. [133] studied the effect of different
Jo

pressures (15, 30, 50, and 80 bar) in the DO of palm oil using a NiMo/ᵞ-Al2O3 catalyst. Complete
conversion (100%) was stable at higher pressures and the product yield was enhanced as the
pressure increased. The obtained yield also increased to 95.2% at 80 bar although it experiences
a slight stall from 30 to 50 bar. The HDO yield was also improved over higher reaction pressures
from 67.8% to 78.1%[167]. It is speculated that the absorption of hydrogen onto surface active
sites of the catalyst at higher pressures leads to increased hydrogen solubility inside the oils and
thus a higher yield [224].
Meanwhile, the LHSV which controls the contact time between the reactants and catalysts is
one of the most crucial parameters [133, 223] particularly in the cracking and isomerization
reactions in the DO process [126]. Kaewtrakulchai et al. [55] varied the LHSV from 0.5 to 1.5
per hour in the DO of palm oil using CoP/PC catalyst. The conversion obtained at 0.5 and 1.0/h
was 100% whereas at 1.5/h the conversion was slightly reduced to 85.9%. The highest
hydrocarbon yield was 75.9% consisting of 50.4% green diesel yield at LHSV of 1.0/h while it

of
was lessened to less than 70% at LHSV of 1.5/h. The same trend was found for the green diesel
selectivity which dramatically decreased from over 90% at LHSV of 0.5/h to less than 50% at
LHSV of 1.5/h.

ro
In another study, the conversion of palm oil was not affected at higher LHSV, but the yield
was reduced from 95% to 84.3% as the LHSV increased from 0.25 to 5/h [133]. The reduction of

-p
LHSV favored the DO process, thus upgrading the yield and selectivity of green diesel as
claimed by Chen et al. [225]. A lower LHSV has been shown to escalate the conversion of oils,
re
yield, and selectivity of green diesel. Higher LHSV causes quick cracking and isomerization
reactions which contribute to plugging in the reactor caused by the formation of free fatty acids
lP

and esters [226-229].


In our view, the best result and best operating condition are overall significantly affected by
catalyst content among others because the proper catalyst content in the reaction provides the
na

enlarged surface active sites, thus gives a stronger Bronsted-Lewis acidity. In addition, it
facilitates the reactants to react caused by a larger accessibility of the reactants to produce green
diesel leading to the better capability and higher yield.
ur

5.2 Leverage of catalyst treatment


Jo

The method by which a catalyst is synthesized or the post-synthesis processes such as the
removal of impurities can have considerable effects on the DO process [66, 92, 188]. For
instance, Choo et al. [46] investigated 65 nm Y zeolite (Y65) synthesized by two different
methods namely the impregnation (IM-Y65) method and deposition precipitation (DP-Y65) in
the DO of triolein. The Y zeolite had a surface area of 62 m2/g while the Y zeolite synthesized
over IM and DP had a surface area of 43 and 124 m2/g, respectively. The conversion obtained
from the DP-Y65 (74.0%) was understandably higher than that of IM-Y65 (46.2%) with the Y65
conversion standing at only 20.8% at 380 oC in 30 minutes. The DP-Y65 exhibited a higher
green diesel selectivity (88.9%) than the IM-Y65 (78.2%) when Ni content was added [46]. The
best DO capability was achieved by the DP-Y65 and resulted in the highest conversion of 80%
and green diesel selectivity of 93.7% at 380 oC in 1 h [46]. Very small NiO particles (3.57 ± 0.40
nm), increased the accessibility (H.F value of 0.084), and a higher Brönsted-Lewis acidity ratio
(0.29) was also an important factor in increasing the accessibility of reactants and the catalytic
deoxygenation activity thus aiding the diffusion of triolein particles (~4.4 nm) into the active

of
surface sites to convert them to the desired products [230]. On the contrary, the poor DO ability
shown by the IM-Y65 was attributed to its smaller external surface area (43 m2/g) and the lower
strength of its acid sites (1.14 mmol/g) [46].

ro
Impregnation consists of two general types known as dry IM (DIM) and wet IM (WIM).
Hensley et al. [231] synthesized Pd-Fe2O3 catalyst using DIM at 300 oC and produced a yield of

-p
56% and selectivity of 92%. Zheng et al. [232] used the DIM method to synthesize Cu/SiO2
catalyst at 290 oC and resulted in a yield of 70% and selectivity of 14.2%. Whilst, Nikul'shin et
re
al. [233] applied DIM for the synthesis of MoS/Al2O3 catalyst and obtained a yield of 70% and
selectivity of 7.9%. Conversely, WIM was conducted by Tran et al. [234] for Co/MCM-41
lP

zeolite synthesis, and by Alharbi et al. [235] for Ru/CsPW-1 catalyst. Yields as high as 100%
[234] were obtained by the Co/MCM-41 zeolite at 400 oC and a selectivity of 100% was
achieved by the Ru/CsPW-1 catalyst at 100 oC [235].
na

As the evidence reviewed shows, the synthesis of the catalysts plays a crucial role in their
ability for catalytic DO. The synthesis route can change the physical and chemical properties of
the catalysts leading to increase or decrease in DO performance. Therefore, catalyst treatments
ur

are almost always applied to remove impurities and to enlarge the active surface sites. For the
time being, the impregnation method for catalyst treatment has been widely used because of the
Jo

simplicity of the process.

5.3 Role of metal loadings


Metal oxides are often used to improve the catalyst activity and can be divided in the two
main groups of noble metals (i.e. ruthenium (Ru), rhodium (Rh), palladium (Pd), silver (Ag),
osmium (Os), iridium (Ir), platinum (Pt), and gold (Ag) etc.) and transition metals (i.e. zinc (Zn),
copper (Cu), scandium (Sc), titanium (Ti), vanadium (V), chromium (Cr), manganese (Mn), iron
(Fe), cobalt (Co), molybdenum (Mo), and nickel (Ni) etc.) [236, 237]. Catalysts impregnated by
these metal oxides have shown good activity and stability and are able to produce high green
diesel yields [222, 238].
For instance, Srihanun et al. [75] used two different catalysts namely Pd/Al2O3 and Fe-

of
doped Pd/Al2O3 in green diesel production from palm oil. The increasing active sites, porosity,
and pore size of the Fe-Pd/Al2O3 catalyst were caused by Fe loading into the catalyst. Doping Fe
into the Pd/Al2O3 catalyst could produce a higher yield of biofuels (94%) compared to Pd/Al2O3

ro
catalyst alone (86%) with a green diesel proportion of 50 – 62.02%.
Using a combination of a noble metal and a transition metal was also studied by Fangkoch et

-p
al. [207] using Pt-doped MoO2 catalyst with a Pt content of 0.5 – 3% in the DO process of palm
oil. The acidity of the catalyst was greatly improved, showing excellent DCX and DCN
re
processes and resulting in high conversion and yield [121, 239]. However, a further increase in
Pt loading resulted in unfavorable activity because of the formation of clotted Pt particles leading
lP

to the blocking of some pores on the catalyst active sites [207]. The highest green diesel yield
was, 56.4%, achieved by MoO2 catalyst doped by 1% Pt which decreased to 55.2% at 2% Pt and
further decreased to 54.5% at 3% Pt content [207].
na

Although catalysts incorporating noble metals often give great results, the high price of
noble metals becomes the main problem for the application of such catalysts on a large scale. On
the contrary, transition metals are much cheaper, hence if comparable performance is obtained,
ur

their use becomes attractive [240]. This is why transition metals have been widely explored as
catalysts, catalyst supports, or for improving catalysts which also incorporate noble metals. For
Jo

instance, Wang [241] used ZSM-5 supported by Ru and NiMo catalyst in a study.
Approximately, a yield of 16% for bio-jet fuel and 20 – 29% for green diesel was obtained by the
Ru-ZSM-5 which was similar to the NiMo catalyst. Therefore, transition metal catalysts are
continuously investigated for cost-efficiency and comparable results. Moreover, the effect of
doping transition metals into catalysts was observed by Choo et al. [47] who utilized Y zeolite
synthesized over Ni, Cu, Co, Zn and Mn. The best DO activity was shown by Ni-Y followed by
Co-Y, Cu-Y, Mn-Y, and Zn-Y. The Ni-Y zeolite successfully produced the maximum
conversion (76.21%) and green diesel selectivity of 92.61% was achieved at 380 oC for 2 h. The
excellent catalytic performance of the Ni-Y zeolite was the result of the synergistic influence
between the high hydrogenolysis activity of Ni doping, the decreased Bronsted-Lewis active acid
sites which facilitate the DCX reaction, pressurizing the CC and polymerization of heavy
hydrocarbon compounds into lighter hydrocarbons [47].
The excellent performance of Ni-doped catalysts was also proven by Kamaruzaman et al.
[48] who employed 10 wt.% SBA-15 catalyst assisted by 5 wt.% Ni, Co, and Ni-Co in the DO
reaction of PFAD to evaluate green diesel production from CPO at 350 oC in 3 h. The catalytic

of
performance shown by the Ni-assisted SBA-15 catalyst was much better than the Co-assisted
SBA-15 catalyst due to the larger size of Co particles (5.45 nm) and formation of carbon and

ro
coke leading to the blockage of some pores on the active surface sites as well as deactivation of
the SBA-15 catalyst. The Ni-SBA-15 catalyst resulted in the highest hydrocarbon yield of 85.8%

-p
and green diesel selectivity of 97.2% and was reused up to 5 cycles resulting in a hydrocarbon
yield of 16.3% [48]. The deactivation effect of catalysts supported by Co metal was also
re
witnessed by Srifa et al. [212] using Co-Al2O3 catalyst in the DO of palm oil.
Elsewhere, Liang et al. [242] demonstrated the powerful catalytic performance of
lP

MoO2/Mo2CTx catalyst doped with Ni metal which led to 100% conversion and high selectivity
caused by the synergistic effects between Ni metal and the active catalyst sites. More active sites,
larger crystallite size, smaller particle size, lower reduction, and less coke formation as observed
na

by Ameen et al. [243] are attributed to Ni doping, and that is why the application of this metal
has been widely studied.
Apart from Ni, Garrido et al. [76] used Mn added into Al2O3 to support NiMo catalysts for
ur

converting vegetable oil into green diesel at 380 oC and 40 bar. The Mn-Al2O3-assisted NiMo
catalysts showed great activity and stability, as a consequence of synergistic stabilization
Jo

between Mn and Al2O3 in redispersion of Ni-driven MoS2 active surface sites [76].
Based on the reviewed studies, it is safe to say that catalysts using transition metals are
highly promising compared to noble metal compound groups, because they are cheaper, and yet,
exhibit comparable results. If the metal oxide has a larger size than the kinetic diameter of the
oils, higher accessibility of the reactants into the pores of the catalysts becomes possible. Ni
metal was crowned as the best metal that has been often utilized in many biofuel processes
including green diesel production.

5.4 Role of phosphides contents


Furthermore, the performance of metal-oxides-doped catalysts can be significantly improved
by doping them with phosphides. This is because the phosphides provide highly spacious H-
active surface sites on the catalysts and larger accessibility of the reactants into the catalyst pores
[130]. Phosphides have been widely tested in the form of Ni2P and Co2P in DO of vegetable oils

of
[126, 233, 234, 237].
The utilization of Ni2P was first introduced more than 60 years ago by Sweeny et al. [244] in
1958 by reducing nitrobenzene and hydrogen into aniline. In the years that followed, it was

ro
further implemented for other experiments [245, 246] specifically in the production of biofuels
and green diesel [247-249]. Moreover, the catalysts incorporating Ni2P content have shown

-p
better performance than the ones which only had metal oxides. Camargo et al. [93] studied the
effect of adding Ni2P on USY, ZSM-5, and SBA-15 catalysts in the DO reaction of oleic acid in
re
a batch reactor at 260, 280, and 300 oC and 50 bar in 6 h. The Ni2P/SBA-15 revealed a favorable
DO capability when compared with the SBA-15 doped by noble metals. The highest yield was
lP

achieved by Ni2P/SBA-15 (42%) followed by Ni2P/ZSM-5 (29%) and Ni2P/USY (24%) because
the Ni2P/SBA-15 had a larger micropore volume (0.764 cm3/g) than Ni2P/ZSM-5 (0.368 cm3/g)
and Ni2P/USY (0.168 cm3/g) and thus facilitated entrance into the catalysts pores [93].
na

Elsewhere, Rakmae et al. [208] concluded that the conversion of palm oil and the yield of
green diesel as well as the HDO yield could be significantly improved when using Na-MOR
zeolite assisted by Ni2P due to the higher Lewis-Bronsted acidity and larger active surface sites
ur

on the Ni2P-aided catalysts [245]. The highest conversion achieved was 100% and the green
diesel yield and HDO yield were both over 80% [208]. Meanwhile, Xin et al. [132] employed
Jo

Ni2P-synthesized AC catalyst at 350 oC and 1 bar and obtained 86.2 – 100% conversion with
41.4 – 56% hydrocarbon yield and green diesel selectivity of 40.1 – 74.9% [132].
In another study, Wagner et al. [247] reported that Y zeolite impregnated with Ni2P was
highly active and could be reused many times. Meanwhile, Zarchin et al. [126] have used a
Ni2P/SiO2 catalyst and a Ni2P-HY zeolite at 340 – 420 oC and 30 bar. The catalysts resulted in
100% conversion and the highest yields of 81.7% for Ni2P/HY zeolite and 82% for Ni2P/SiO2
catalyst was achieved. Furthermore, Liu et al. [100] used ZSM-22 augmented by Ni2P for the DO
process in a fixed bed reactor at 310 – 360 oC and 10 bar for 2.5 h and reached 99.6% conversion
and 42.9% yield. The researchers claimed that the Brønsted acidity of ZSM-22 crucially
contributed to the CI reaction, whereas the Lewis acidity played an important role for the CC
reaction.
Another metal phosphide, namely cobalt phosphide (Co2P), has also been used in green
diesel production via the DO of palm oil. Kaewtrakulchai et al. [55] employed porous carbon
(PC) catalyst impregnated by Co2P in the DO of palm oil at 340 – 420 oC and LHSV of 0.5 – 1.5

of
/h. Complete conversion of palm oil and the highest green diesel selectivity achieved by the
Co2P/PC catalyst at 340 oC were 100% and 67%, respectively [55].
Kochaputi et al. [96] used Ni2P and Co2P impregnated into USY zeolite for DO in a batch

ro
reactor at 260 – 340 oC and 50 bar with a residence time of 6 h. The two catalysts resulted in
complete conversion (100%) and the highest yield of 52% for Co2P/USY and 36% for Ni2P/USY

-p
[93]. It seems that the Co2P/USY outperformed the Ni2P/USY due to larger crystallite size (i.e.
58.4 nm for Co2P/USY and 39.2 nm for Ni2P/USY) and higher total pore volume (i.e. 0.28 cc/g
re
for Co2P/USY and 0.24 for Ni2P/USY) [96].
Niobium has also been used to synthesize catalysts for the DO of vegetable oils as well due
lP

to its attractive acid and redox characteristics [250-252]. Niobium phosphate (NbOPO4)-assisted
catalysts for chemical and biofuel production are a possibility [253, 254] to be considered for
green diesel production [77, 95]. Scaldaferri et al. [77] applied niobium phosphate catalysts in
na

the DO process at 300 – 350 oC, 10 bar, and catalyst contents of 0 – 25% for 3 – 5 h. The
optimization assays provided elevated hydrocarbon contents (97%) and upper yields of biofuels.
Bio-jet fuel was produced at a higher yield, 62%, and green diesel was acquired at a yield of
ur

40%. The performance of niobium-synthesized catalysts has been optimized under various
conditions [255, 256] including different calcination temperatures as reported by Rade et al.
Jo

[257], various alcohols used as studied by Bassan et al. [258], and different operating parameters
as investigated by Conceição et al. [259]. Researchers should be encouraged to use Ni2P for
supporting the catalyst because, either the synthetic catalysts or zeolite-typed catalysts have a
better performance when they are modified over Ni2P supports. This approach can be viewed
both in economic aspects and from the green environment aspects.
5.5 Nanoparticle catalysts
Nano-sized catalyst particles have exhibited excellent performance by providing higher
surface areas, increased accessibility and rapid diffusion of reactants and products over the
catalyst surface [260]. Nano-catalysts have been explored in DO reactions [213] such as in the
case of using multi-walled carbon nanotubes (MWCNT) covered with NiO–Fe2O3 and NiO–ZnO
catalysts [71]. The researchers observed that the high catalyst content and metal oxide content
did not affect the DO activity but the nano-sized carbon particles contributed to catalyst stability,
and selectivity (89% of hydrocarbon and 79% of n-(C15+C17)) caused by stronger Bronsted-

of
Lewis acid sites. The green diesel produced over the NiO–Fe2O3 catalyst possessed the highest
heating value and the lowest oxygen content, as well as better physicochemical properties
compared with commercial diesel fuels and regular biodiesel. The nanosize-assisted NiO–Fe2O3

ro
catalyst could be reused for up to 6 cycles, and provided the strongest acid and basic active sites
with 5,313 mol/g and 9,046 mol/g, respectively [261, 262].

-p
Nano-catalysts have also been applied in the treatment of PFAD to re-produce green diesel
as conducted by Baharudin et al. [49] using NiO–ZnO nano catalysts at 350 oC for 2 h. Large
re
crystalline meso-macro-sized ZnO particles were created by the co-precipitation method in
catalyst synthesis. The nanosized NiO content was then added. Stronger Bronsted-Lewis acidity
lP

of the NiO-ZnO catalyst was attributed to the synergistic influence between the active sites of the
catalyst and the support caused by higher crystallite sizes. Finally, an 83.4% hydrocarbon yield
with green diesel selectivity of 86% was achieved.
na

The co-precipitation method in nanosized Cu/Ni-Al2O3 catalyst synthesis was also


implemented by Gousi et al. [217] in the DO process at 310 oC and 40 bar where the catalyst size
was 3 – 5 nm with an active surface area of 192 – 285 m2/g. The nanosized Cu/Ni-Al2O3 catalyst
ur

revealed that the highest catalytic ability was caused by stronger acidity [45]. Elsewhere,
approximately, an 89% hydrocarbon yield and 93% green diesel selectivity were achieved by an
Jo

Ag-La/AC nano-catalyst even after 6 cycles as reported by Alsultan et al. [189].


The main advantages of nano-sized catalyst particles is that they give higher Bronsted-Lewis
acid-basic sites, enlarged crystallite sizes, and larger active sites; which in turn improve the DO
activity, catalyst reusability, green diesel selectivity and final quality of the green diesel.
6. Final considerations
Fossil fuels donate 80% of the world's energy consumption that is annually predicted to
increase to 0.9% from 725 x 1015 Btu in 2030 to 780 x 1015 Btu in 2035. Either economic
cooperation and development (ECOD) countries or non ECOD countries show the same trend,
which is an escalation in requirements to 280 x 1015 Btu and 500 x 1015 Btu, successively, by the
year 2035. It triggers a faster growth in economic progress in every country, where global energy
consumption is mostly (40%) accounted for by the transportation sectors, such as, gasoline, jet
fuel, and diesel fuels. However, the large-scale and uncontrolled use of fossil fuels contributes to

of
the depletion of fossil resources and global warming caused by CO2, CO, hydrocarbons, and
NOx produced.
Hence, a renewable, sustainable, and alternative biofuel namely green diesel is crucially

ro
required to displace petroleum-based diesel and biodiesel regarding effluent, no pollution, and
economically attractive/cheap, where the green diesel consumption is annually portended to

-p
increase to 3% in the coming 2030. Even, the green diesel is considered as the greenest, most
cost-efficient, and potential fuel to be used in the diesel engines and produced on an industrial
re
scale in the recent and future to reduce the environmental pollution and to fulfill the increasing
diesel fuel demand.
lP

The green diesel production is mostly carried out through HDO processing using vegetable
oils possessing high triglycerides or FFAs content as well as a high oxygen content which can be
further eliminated under HG, DH, DO, DCX, and DCN processes. Simultaneously occurring, the
na

second step reactions such as CC and CI facilitates cracking of heavy hydrocarbons into lighter
hydrocarbons and by-products (i.e. CO2, CO, and H2O). To minimize these undesired products,
several modifications are required during the process such as application of different reaction
ur

routes, doping metal oxides, and adding nickel phosphides. These developments play a crucial
role in the catalytic performance and the catalyst stability which can increase conversion,
Jo

selectivity, and yield.


Among those various reaction pathways, the DO process is widely applied in the green
diesel production using heterogeneous catalysts and zeolites where the zeolites are a promising
and potential catalyst these days as a substitution for solid acid and base catalysts. These newer
catalysts are proving attractive due to their regenerability, plentiful nature, inexpensive, and
highly stable framework structure. Different calcination temperatures and zeolite treatments are
being conducted to improve the zeolite ability, leading to increasing catalytic performance.
Catalyst treatments can remove impurities and change the physical and chemical properties
of the catalysts. Whilst, doping metal oxides (i.e. noble and transition metals) and phosphides
contents (i.e. Ni2P, Co2P and NbOPO4) are responsible for higher active surface sites, enlarged
crystallite size, and stronger Bronsted-Lewis acid-basic sites. Meanwhile, nano-sized catalysts
provide more exposed surface area, enlarged accessibility and rapid diffusion for reactant and
product over the catalyst surface. All of those modifications have a significant role for improving

of
the catalytic DO and catalyst stability thus increasing DO efficiency, catalyst reusability, green
diesel selectivity and high-grade quality of green diesel.
A very important point in the HDO processes is the consumption of hydrogen. Since the

ro
hydrogen requirement is significant, it should not be derived from unsustainable and
nonrenewable sources. In addition, since it is rare for refineries to have a hydrogen surplus, the

-p
hydrogen required for HDO will often need to be generated or provided from the outside [263].
However, a supply of hydrogen from external sources is cost prohibitive due to the considerable
re
expenses of transport and storage, and therefore, on-site generation is the more practical option.
There are several options for generating hydrogen on-site, such as gasification of biomass,
lP

shifting CO to H2 followed by scrubbing CO2, steam reforming of an aqueous solution, or, the
electrolysis of water. For example, the area of electrochemical HDO has come under
investigation in recent years because of its ability to operate at ambient conditions without
na

external supply of hydrogen gas [264].


Finally, a summary of the more promising as well as the less promising areas of research in
green diesel production is provided in Table 6 as per the authors’ opinions.
ur
Jo
Table 6. Promising as opposed to less promising areas of research as identified by the
authors for green diesel production.

More promising areas of research in Less promising areas of research in green


green diesel production diesel production
Environmentally friendly Limited use of the green diesel
Renewable and sustainable biofuel High production cost
Abundant and inexpensive feedstocks Government policy of unfriendly feedstocks

of
The efficient and clean combustion Engine must be firstly modified to use green
technologies diesel
Modern conversion technologies

ro
Competitive fuels, heat, and electricity
Less pollution caused by CO2, CO, NOx,
SOx, and hydrocarbons

-p
Low-cost biofuel
Non-oxygenated biofuel
re
7. Conclusion
This current review demonstrates the variety of processing technologies and catalysts that
lP

are suitable for green diesel production from a variety of feedstocks. Among all the various
reaction pathways discussed in this review, the DO process is widely applied in the green diesel
production. Zeolites and heterogeneous catalysts are promising and potential catalysts to
na

substitute the older conventional solid acid and base catalysts due to their regenerability
potential, abundant, and plentiful availability, low and inexpensive cost, and highly stable
framework structure. Catalyst treatments can remove the impurities and change its physical and
ur

chemical properties to improve the catalytic performance. Doping metal oxides and phosphides
contents are responsible for higher active surface sites, enlarged crystallite size, and stronger
Jo

Bronsted-Lewis acid-basic sites. Meanwhile, nano-sized catalysts provide more exposed surface
area, enlarged accessibility, and rapid diffusion for reactant and product over the catalyst surface.
All the modifications discussed have a significant role in improving the catalyst stability, thus
increasing DO efficiency, reusability, green diesel selectivity, and high-grade quality of green
diesel.
However, despite the huge benefits that can be obtained from green diesel production, one
important aspect that needs to be seriously taken into consideration, which is the way to cope
with the limited amount of feedstock available to satisfy the potentially high demand for the fuel.
The reusability studies for the treated zeolites also require an intensive investigation to confirm
its stability and performance. Furthermore, the cost analysis of the overall green diesel
production for large scale and high capacity units needs to be thoroughly investigated to confirm
the profit margin from the business point of view.

Declaration of interests

of
The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.

ro
Acknowledgement
N.I.W.A. greatly appreciate the Universiti Teknologi Malaysia (UTM) and the Ministry of

-p
Education (MOE) Malaysia through Fundamental Research Grant Scheme
(FRGS/1/2018/STG01/UTM/01/4) for the financial and facilities support. L.M. thanks to the
re
Coordination of Superior Level Staff Improvement (CAPES/Brazil), the National Council for
Scientific and Technological Development (CNPq/Brazil) and the Alagoas State Research
lP

Support Foundation (FAPEAL/Brazil). The findings achieved herein are solely the responsibility
of the authors. H.I.M greatly appreciates PT Global Amine Indonesia and PT Wilmar Nabati
Indonesia for the supports and some of good advices in order to make the paper more perfect.
na
ur

Refferences
Jo

[1] Hancsók, J., S. Magyar, and A. Holló. Importance of isoparaffins in the crude oil refining
industry. in The 8th International Conference on Chemical and Process Engineering, Italy,
Naples, Ischia. 2007.
[2] Wibowo, A.A., et al. Green diesel production from waste vegetable oil: A simulation study. in
AIP Conference Proceedings. 2020. AIP Publishing LLC.
[3] Conti, J., et al., International energy outlook 2016 with projections to 2040. 2016, USDOE
Energy Information Administration (EIA), Washington, DC (United States ….
[4] Arun, N., R.V. Sharma, and A.K. Dalai, Green diesel synthesis by hydrodeoxygenation of bio-
based feedstocks: Strategies for catalyst design and development. Renewable and
Sustainable Energy Reviews, 2015. 48: p. 240-255.
[5] IEA, G.E., CO2 Status Report 2018. International Energy Agency, Paris, 2019.
[6] Capellán-Pérez, I., et al., Fossil fuel depletion and socio-economic scenarios: An integrated
approach. Energy, 2014. 77: p. 641-666.
[7] Hughes, L. and J. Rudolph, Future world oil production: growth, plateau, or peak? Current
Opinion in Environmental Sustainability, 2011. 3(4): p. 225-234.
[8] Gregg, J.S., R.J. Andres, and G. Marland, China: Emissions pattern of the world leader in CO2
emissions from fossil fuel consumption and cement production. Geophysical Research
Letters, 2008. 35(8).
[9] Gurney, K.R., et al., Toward Accurate, Policy-Relevant Fossil Fuel CO2 Emission Landscapes.

of
Environmental Science & Technology, 2020. 54(16): p. 9896-9907.
[10] Kalnes, T., T. Marker, and D.R. Shonnard, Green diesel: a second generation biofuel.
International Journal of chemical reactor engineering, 2007. 5(1).

ro
[11] Sheehan, J., et al., An overview of biodiesel and petroleum diesel life cycles. 1998.
[12] Van Gerpen, J.H., C.L. Peterson, and C.E. Goering, Biodiesel: an alternative fuel for
compression ignition engines. 2007: American Society of Agricultural and Biological
Engineers Kentucky, USA.

-p
[13] Kalnes, T.N., et al., A technoeconomic and environmental life cycle comparison of green
diesel to biodiesel and syndiesel. Environmental Progress & Sustainable Energy: An
re
Official Publication of the American Institute of Chemical Engineers, 2009. 28(1): p. 111-
120.
[14] Bezergianni, S. and A. Dimitriadis, Comparison between different types of renewable diesel.
lP

Renewable and Sustainable Energy Reviews, 2013. 21: p. 110-116.


[15] Sanhoob, M.A., et al., Steam cracking of green diesel (C12) to BTX and olefins over silane-
treated hierarchical BEA. Fuel, 2020. 263: p. 116624.
[16] Chu, S. and A. Majumdar, Opportunities and challenges for a sustainable energy future.
na

nature, 2012. 488(7411): p. 294-303.


[17] Tomes, D., P. Lakshmanan, and D. Songstad, Biofuels: global impact on renewable energy,
production agriculture, and technological advancements. 2010: Springer Science &
Business Media.
ur

[18] Munasinghe, P.C. and S.K. Khanal, Biomass-derived syngas fermentation into biofuels:
opportunities and challenges. Bioresource technology, 2010. 101(13): p. 5013-5022.
[19] Almeida, J.R., L.C. Fávaro, and B.F. Quirino, Biodiesel biorefinery: opportunities and
Jo

challenges for microbial production of fuels and chemicals from glycerol waste.
Biotechnology for biofuels, 2012. 5(1): p. 1-16.
[20] Lin, L., et al., Opportunities and challenges for biodiesel fuel. Applied energy, 2011. 88(4): p.
1020-1031.
[21] Shanmugam, S., H.-H. Ngo, and Y.-R. Wu, Advanced CRISPR/Cas-based genome editing
tools for microbial biofuels production: A review. Renewable Energy, 2020. 149: p. 1107-
1119.
[22] Cheah, W.Y., et al., Pretreatment methods for lignocellulosic biofuels production: current
advances, challenges and future prospects. Biofuel Research Journal, 2020. 7(1): p. 1115.
[23] Pablo, G., et al., Recent trends on seaweed fractionation for liquid biofuels production.
Bioresource technology, 2020. 299: p. 122613.
[24] Battista, F., et al., Added-value molecules recovery and biofuels production from spent
coffee grounds. Renewable and Sustainable Energy Reviews, 2020. 131: p. 110007.
[25] Zhang, Y., et al., Fast microwave-assisted pyrolysis of wastes for biofuels production–A
review. Bioresource technology, 2020. 297: p. 122480.
[26] Ali, S., et al., Biofuels production from weed biomass using nanocatalyst technology.
Biomass and Bioenergy, 2020. 139: p. 105595.
[27] Dufey, A., Biofuels production, trade and sustainable development: emerging issues. 2006:
Iied.
[28] Bazargan, A., et al., A calcium oxide-based catalyst derived from palm kernel shell
gasification residues for biodiesel production. Fuel, 2015. 150: p. 519-525.

of
[29] Kostić, M.D., et al., Optimization and kinetics of sunflower oil methanolysis catalyzed by
calcium oxide-based catalyst derived from palm kernel shell biochar. Fuel, 2016. 163: p.
304-313.

ro
[30] Mohammad, M., et al., Overview on the production of paraffin based-biofuels via catalytic
hydrodeoxygenation. Renewable and Sustainable Energy Reviews, 2013. 22: p. 121-132.
[31] Galadima, A. and O. Muraza, Catalytic upgrading of vegetable oils into jet fuels range

Engineering Chemistry, 2015. 29: p. 12-23.


-p
hydrocarbons using heterogeneous catalysts: A review. Journal of Industrial and

[32] Schmidt, J.H., Life cycle assessment of five vegetable oils. Journal of Cleaner Production,
re
2015. 87: p. 130-138.
[33] Sugiyono, A., et al. Prospects for the Development of Green Gasoline and Green Diesel from
Crude Palm Oil in Indonesia. in Materials Science Forum. 2020. Trans Tech Publ.
lP

[34] Norouzi, N. and S. Talebi, An overview on the green petroleum production. Chemical
Review and Letters, 2020. 3(1): p. 38-52.
[35] Nikolopoulos, I., et al., Waste cooking oil transformation into third generation green diesel
catalyzed by nickel–alumina catalysts. Molecular Catalysis, 2020. 482: p. 110697.
na

[36] Jin, M., et al., Coproduction of Value-Added Lube Base Oil and Green Diesel from Natural
Triglycerides via a Simple Two-Step Process. Industrial & Engineering Chemistry
Research, 2020. 59(19): p. 8946-8954.
[37] JCG Filho, P., et al., L.: Biodiesel production from Sterculia striata oil by ethyl
ur

transesterification method. Ind. Crops Prod, 2015. 74: p. 767-772.


[38] Nazário, J., et al., Empirical Evaluation of Stirring Procedures in The Production of Biodiesel
From Castor Oil. Brazilian Journal of Petroleum and Gas, 2016. 10(2).
Jo

[39] Peiter, A.S., et al., Stirring and mixing in ethylic biodiesel production. Journal of King Saud
University-Science, 2020. 32(1): p. 54-59.
[40] Mahlia, T., et al., Patent landscape review on biodiesel production: Technology updates.
Renewable and Sustainable Energy Reviews, 2020. 118: p. 109526.
[41] Abomohra, A.E.-F., et al., Potential of fat, oil and grease (FOG) for biodiesel production: A
critical review on the recent progress and future perspectives. Progress in Energy and
Combustion Science, 2020. 81: p. 100868.
[42] Douvartzides, S.L., et al., Green diesel: Biomass feedstocks, production technologies,
catalytic research, fuel properties and performance in compression ignition internal
combustion engines. Energies, 2019. 12(5): p. 809.
[43] Kim, S.K., et al., Production of renewable diesel via catalytic deoxygenation of natural
triglycerides: Comprehensive understanding of reaction intermediates and
hydrocarbons. Applied energy, 2014. 116: p. 199-205.
[44] Yulia, D. and A. Zulys. Hydroprocessing of kemiri sunan oil (reutealis trisperma (blanco) airy
shaw) over NiMoCe/γ-Al2O3 catalyst to produce green diesel. in IOP Conference Series:
Materials Science and Engineering. 2020. IOP Publishing.
[45] Gamal, M.S., et al., Effective catalytic deoxygenation of palm fatty acid distillate for green
diesel production under hydrogen-free atmosphere over bimetallic catalyst CoMo
supported on activated carbon. Fuel Processing Technology, 2020. 208: p. 106519.

of
[46] Choo, M.-Y., et al., Deposition of NiO Nanoparticles on Nanosized Zeolite NaY for
Production of Biofuel via Hydrogen-Free Deoxygenation. Materials, 2020. 13(14): p.
3104.

ro
[47] Choo, M.-Y., et al., Deoxygenation of triolein to green diesel in the H2-free condition: Effect
of transition metal oxide supported on zeolite Y. Journal of Analytical and Applied
Pyrolysis, 2020. 147: p. 104797.

-p
[48] Kamaruzaman, M.F., Y.H. Taufiq-Yap, and D. Derawi, Green diesel production from palm
fatty acid distillate over SBA-15-supported nickel, cobalt, and nickel/cobalt catalysts.
Biomass and Bioenergy, 2020. 134: p. 105476.
re
[49] Baharudin, K.B., et al., Renewable diesel via solventless and hydrogen-free catalytic
deoxygenation of palm fatty acid distillate. Journal of Cleaner Production, 2020. 274: p.
122850.
lP

[50] Baharudin, K.B., et al., Mesoporous NiO/Al-SBA-15 catalysts for solvent-free deoxygenation
of palm fatty acid distillate. Microporous and Mesoporous Materials, 2019. 276: p. 13-
22.
[51] Water, C., Green diesel production by hydrorefining renewable feedstocks. 2008.
na

[52] MEHRA, S., Physiochemical properties characteristics of Jatropha Curcas biodiesel and its
blends with tetrachloroethylene addition. International Journal of Chemistry Research,
2012: p. 30-33.
[53] VERDUZCO, L.F.R., et al., Hydrodeoxigenation process of vegetable oils for obtaining green
ur

diesel. 2020, Google Patents.


[54] Pratas, M.J., et al., Biodiesel density: experimental measurements and prediction models.
Energy & Fuels, 2011. 25(5): p. 2333-2340.
Jo

[55] Kaewtrakulchai, N., et al., Palm Oil Conversion to Bio-Jet and Green Diesel Fuels over Cobalt
Phosphide on Porous Carbons Derived from Palm Male Flowers. Catalysts, 2020. 10(6): p.
694.
[56] Ameen, M., et al., Catalytic hydrodeoxygenation of triglycerides: An approach to clean
diesel fuel production. Renewable and Sustainable Energy Reviews, 2017. 80: p. 1072-
1088.
[57] Harnos, S., G. Onyestyák, and D. Kalló, Hydrocarbons from sunflower oil over partly reduced
catalysts. Reaction Kinetics, Mechanisms and Catalysis, 2012. 106(1): p. 99-111.
[58] Šimáček, P., et al., Fuel properties of hydroprocessed rapeseed oil. Fuel, 2010. 89(3): p. 611-
615.
[59] Loe, R., et al., Effect of Cu and Sn promotion on the catalytic deoxygenation of model and
algal lipids to fuel-like hydrocarbons over supported Ni catalysts. Applied Catalysis B:
Environmental, 2016. 191: p. 147-156.
[60] Emori, E.Y., et al., Catalytic cracking of soybean oil using ZSM5 zeolite. Catalysis Today,
2017. 279: p. 168-176.
[61] Itthibenchapong, V., et al., Deoxygenation of palm kernel oil to jet fuel-like hydrocarbons
using Ni-MoS2/γ-Al2O3 catalysts. Energy conversion and management, 2017. 134: p.
188-196.
[62] Gong, S., et al., Isomerization of n-alkanes derived from jatropha oil over bifunctional
catalysts. Journal of Molecular Catalysis A: Chemical, 2013. 370: p. 14-21.

of
[63] Yasir, M., et al., Hydroprocessing of crude jatropha oil using hierarchical structured TiO2
nanocatalysts. Procedia engineering, 2016. 148: p. 275-281.
[64] Lee, W.-S., et al., Selective vapor-phase hydrodeoxygenation of anisole to benzene on

ro
molybdenum carbide catalysts. Journal of catalysis, 2014. 319: p. 44-53.
[65] da Rocha Filho, G., D. Brodzki, and G. Djéga-Mariadassou, Formation of alkanes,
alkylcycloalkanes and alkylbenzenes during the catalytic hydrocracking of vegetable oils.
Fuel, 1993. 72(4): p. 543-549.

-p
[66] Ameen, M., et al., Parametric Studies on Hydrodeoxygenation of Rubber Seed Oil for Diesel
Range Hydrocarbon Production. Energy & Fuels, 2020. 34(4): p. 4603-4617.
re
[67] Murti, S.S., et al. Synthesis of green diesel through hydrodeoxygenation reaction of used
cooking oil over NiMo/Al2O3 catalyst. in AIP Conference Proceedings. 2020. AIP
Publishing LLC.
lP

[68] de Jesus, S.S., et al., Biodiesel production from microalgae by direct transesterification
using green solvents. Renewable Energy, 2020. 160: p. 1283-1294.
[69] Veriansyah, B., et al., Production of renewable diesel by hydroprocessing of soybean oil:
Effect of catalysts. Fuel, 2012. 94: p. 578-585.
na

[70] Zikri, A. and M. Aznury. Green diesel production from Crude Palm Oil (CPO) using catalytic
hydrogenation method. in IOP Conference Series: Materials Science and Engineering.
2020. IOP Publishing.
[71] Asikin-Mijan, N., et al., Production of renewable diesel from Jatropha curcas oil via
ur

pyrolytic-deoxygenation over various multi-wall carbon nanotube-based catalysts.


Process Safety and Environmental Protection, 2020. 142: p. 336-349.
[72] Baek, H., et al., Production of bio hydrofined diesel, jet fuel, and carbon monoxide from
Jo

fatty acids using a silicon nanowire array-supported rhodium nanoparticle catalyst under
microwave conditions. ACS Catalysis, 2020. 10(3): p. 2148-2156.
[73] Ochoa-Hernández, C., J.M. Coronado, and D.P. Serrano, Hydrotreating of Methyl Esters to
Produce Green Diesel over Co-and Ni-Containing Zr-SBA-15 Catalysts. Catalysts, 2020.
10(2): p. 186.
[74] Phimsen, S., et al., Oil extracted from spent coffee grounds for bio-hydrotreated diesel
production. Energy Conversion and Management, 2016. 126: p. 1028-1036.
[75] Srihanun, N., et al., Biofuels of green diesel–kerosene–gasoline production from palm oil:
effect of palladium cooperated with second metal on hydrocracking reaction. Catalysts,
2020. 10(2): p. 241.
[76] Vázquez-Garrido, I., et al., Synthesis of NiMo catalysts supported on Mn-Al2O3 for
obtaining green diesel from waste soybean oil. Catalysis Today, 2021. 365: p. 327-340.
[77] Scaldaferri, C.A. and V.M.D. Pasa, Production of jet fuel and green diesel range
biohydrocarbons by hydroprocessing of soybean oil over niobium phosphate catalyst.
Fuel, 2019. 245: p. 458-466.
[78] Wada, Y., et al., Physical insight to microwave special effects: nonequilibrium local heating
and acceleration of electron transfer. Journal of the Japan Petroleum Institute, 2018.
61(2): p. 98-105.
[79] Kishimoto, F., et al., Remote Control of Electron Transfer Reaction by Microwave

of
Irradiation: Kinetic Demonstration of Reduction of Bipyridine Derivatives on Surface of
Nickel Particle. The journal of physical chemistry letters, 2019. 10(12): p. 3390-3394.
[80] Vardon, D.R., et al., Hydrothermal catalytic processing of saturated and unsaturated fatty

ro
acids to hydrocarbons with glycerol for in situ hydrogen production. Green Chemistry,
2014. 16(3): p. 1507-1520.
[81] Mizugaki, T., et al., New routes for refinery of biogenic platform chemicals catalyzed by

p. 1-6.
-p
cerium oxide-supported ruthenium nanoparticles in water. Scientific reports, 2017. 7(1):

[82] Yamada, Y.M., et al., A Palladium‐Nanoparticle and Silicon‐Nanowire‐Array Hybrid: A


re
Platform for Catalytic Heterogeneous Reactions. Angewandte Chemie, 2014. 126(1): p.
131-135.
[83] Xu, G., et al., Efficient hydrogenation of various renewable oils over Ru-HAP catalyst in
lP

water. ACS Catalysis, 2017. 7(2): p. 1158-1169.


[84] Moreira, J.d.B.D., D.B. de Rezende, and V.M.D. Pasa, Deoxygenation of Macauba acid oil
over Co-based catalyst supported on activated biochar from Macauba endocarp: A
potential and sustainable route for green diesel and biokerosene production. Fuel, 2020.
na

269: p. 117253.
[85] Silva, L.N., et al., Biokerosene and green diesel from macauba oils via catalytic
deoxygenation over Pd/C. Fuel, 2016. 164: p. 329-338.
[86] da Silva César, A., et al., The prospects of using Acrocomia aculeata (macaúba) a non-edible
ur

biodiesel feedstock in Brazil. Renewable and Sustainable Energy Reviews, 2015. 49: p.
1213-1220.
[87] Evaristo, A.B., et al., Actual and putative potentials of macauba palm as feedstock for solid
Jo

biofuel production from residues. Biomass and Bioenergy, 2016. 85: p. 18-24.
[88] Shim, J.-O., et al., Facile production of biofuel via solvent-free deoxygenation of oleic acid
using a CoMo catalyst. Applied Catalysis B: Environmental, 2018. 239: p. 644-653.
[89] Papanikolaou, G., et al., Highly selective bifunctional Ni zeo-type catalysts for
hydroprocessing of methyl palmitate to green diesel. Catalysis Today, 2020. 345: p. 14-
21.
[90] Yenumala, S.R., et al., Hydrodeoxygenation of karanja oil using ordered mesoporous nickel-
alumina composite catalysts. Catalysis Today, 2020. 348: p. 45-54.
[91] Aliana-Nasharuddin, N., et al., Production of green diesel from catalytic deoxygenation of
chicken fat oil over a series binary metal oxide-supported MWCNTs. RSC Advances, 2020.
10(2): p. 626-642.
[92] Ameen, M., et al., Process optimization of green diesel selectivity and understanding of
reaction intermediates. Renewable Energy, 2020. 149: p. 1092-1106.
[93] de Oliveira Camargo, M., et al., Green diesel production by solvent-free deoxygenation of
oleic acid over nickel phosphide bifunctional catalysts: Effect of the support. Fuel, 2020.
281: p. 118719.
[94] Ameen, M., et al., HY zeolite as hydrodeoxygenation catalyst for diesel range hydrocarbon
production from rubber seed oil. Materials Today: Proceedings, 2019. 16: p. 1742-1749.
[95] Scaldaferri, C.A. and V.M.D. Pasa, Hydrogen-free process to convert lipids into bio-jet fuel
and green diesel over niobium phosphate catalyst in one-step. Chemical Engineering

of
Journal, 2019. 370: p. 98-109.
[96] Kochaputi, N., et al., Catalytic behaviors of supported Cu, Ni, and Co phosphide catalysts for
deoxygenation of oleic acid. Catalysts, 2019. 9(9): p. 715.

ro
[97] Jeon, K.-W., et al., Synthesis and characterization of Pt-, Pd-, and Ru-promoted Ni–Ce0.
6Zr0. 4O2 catalysts for efficient biodiesel production by deoxygenation of oleic acid. Fuel,
2019. 236: p. 928-933.

-p
[98] Taromi, A.A. and S. Kaliaguine, Green diesel production via continuous hydrotreatment of
triglycerides over mesostructured γ-alumina supported NiMo/CoMo catalysts. Fuel
processing technology, 2018. 171: p. 20-30.
re
[99] de Sousa, F.P., C.C. Cardoso, and V.M. Pasa, Producing hydrocarbons for green diesel and
jet fuel formulation from palm kernel fat over Pd/C. Fuel processing technology, 2016.
143: p. 35-42.
lP

[100] Liu, Y., et al., The production of diesel-like hydrocarbons from palmitic acid over HZSM-22
supported nickel phosphide catalysts. Applied Catalysis B: Environmental, 2015. 174: p.
504-514.
[101] Kiatkittipong, W., et al., Diesel-like hydrocarbon production from hydroprocessing of
na

relevant refining palm oil. Fuel processing technology, 2013. 116: p. 16-26.
[102] Sági, D., et al., Co-hydrogenation of fatty acid by-products and different gas oil fractions.
Journal of Cleaner Production, 2017. 161: p. 1352-1359.
[103] Szarvas, T., et al., Radioisotopic investigation of the oleic acid-1-14C HDO reaction
ur

pathways on sulfided Mo/P/Al2O3 and NiW/Al2O3 catalysts. Applied Catalysis B:


Environmental, 2015. 165: p. 245-252.
[104] Srifa, A., et al., Roles of monometallic catalysts in hydrodeoxygenation of palm oil to green
Jo

diesel. Chemical Engineering Journal, 2015. 278: p. 249-258.


[105] Wang, C., et al., High quality diesel-range alkanes production via a single-step
hydrotreatment of vegetable oil over Ni/zeolite catalyst. Catalysis Today, 2014. 234: p.
153-160.
[106] Sotelo-Boyás, R., Y. Liu, and T. Minowa, Renewable diesel production from the
hydrotreating of rapeseed oil with Pt/Zeolite and NiMo/Al2O3 catalysts. Industrial &
Engineering Chemistry Research, 2011. 50(5): p. 2791-2799.
[107] Sousa, F.P., et al., Simultaneous deoxygenation, cracking and isomerization of palm kernel
oil and palm olein over beta zeolite to produce biogasoline, green diesel and biojet-fuel.
Fuel, 2018. 223: p. 149-156.
[108] Toor, S.S., L. Rosendahl, and A. Rudolf, Hydrothermal liquefaction of biomass: a review of
subcritical water technologies. Energy, 2011. 36(5): p. 2328-2342.
[109] Nokkosmäki, M., et al., Catalytic conversion of biomass pyrolysis vapours with zinc oxide.
Journal of Analytical and Applied Pyrolysis, 2000. 55(1): p. 119-131.
[110] Graça, I., et al., Catalytic cracking in the presence of guaiacol. Applied Catalysis B:
Environmental, 2011. 101(3-4): p. 613-621.
[111] Chiaramonti, D., et al., Development of emulsions from biomass pyrolysis liquid and diesel
and their use in engines—Part 1: emulsion production. Biomass and bioenergy, 2003.
25(1): p. 85-99.

of
[112] Chiaramonti, D., et al., Development of emulsions from biomass pyrolysis liquid and diesel
and their use in engines—Part 2: tests in diesel engines. Biomass and Bioenergy, 2003.
25(1): p. 101-111.

ro
[113] Junming, X., et al., Bio-oil upgrading by means of ethyl ester production in reactive
distillation to remove water and to improve storage and fuel characteristics. Biomass
and bioenergy, 2008. 32(11): p. 1056-1061.

20(6): p. 2717-2720.
-p
[114] Zhang, Q., et al., Upgrading bio-oil over different solid catalysts. Energy & Fuels, 2006.

[115] Oh, Y.-K., et al., Recent developments and key barriers to advanced biofuels: a short
re
review. Bioresource technology, 2018. 257: p. 320-333.
[116] Elliott, D.C. and T.R. Hart, Catalytic hydroprocessing of chemical models for bio-oil. Energy
& Fuels, 2009. 23(2): p. 631-637.
lP

[117] Mahfud, F., et al., Acetic acid recovery from fast pyrolysis oil. An exploratory study on
liquid-liquid reactive extraction using aliphatic tertiary amines. Separation Science and
Technology, 2008. 43(11-12): p. 3056-3074.
[118] Luo, J., et al., Two-step hydrothermal conversion of Pubescens to obtain furans and phenol
na

compounds separately. Bioresource technology, 2010. 101(22): p. 8873-8880.


[119] Cheah, K.W., et al., Process simulation and techno economic analysis of renewable diesel
production via catalytic decarboxylation of rubber seed oil–A case study in Malaysia.
Journal of environmental management, 2017. 203: p. 950-961.
ur

[120] Kumar, P., et al., Kinetics of hydrodeoxygenation of stearic acid using supported nickel
catalysts: Effects of supports. Applied Catalysis A: General, 2014. 471: p. 28-38.
[121] Janampelli, S. and S. Darbha, Metal Oxide-Promoted Hydrodeoxygenation Activity of
Jo

Platinum in Pt-MO x/Al2O3 Catalysts for Green Diesel Production. Energy & Fuels, 2018.
32(12): p. 12630-12643.
[122] Ameen, M., et al., Catalytic hydrodeoxygenation of rubber seed oil over sonochemically
synthesized Ni-Mo/γ-Al2O3 catalyst for green diesel production. Ultrasonics
sonochemistry, 2019. 51: p. 90-102.
[123] Li, Z., et al., Effect of incorporation manner of Zr on the Co/SBA-15 catalyst for the Fischer–
Tropsch synthesis. Journal of Molecular Catalysis A: Chemical, 2016. 424: p. 384-392.
[124] Liu, Y., et al., Synthesis of Zr-grafted SBA-15 as an effective support for cobalt catalyst in
Fischer–Tropsch synthesis. Chemistry letters, 2008. 37(9): p. 984-985.
[125] Bach, Q.-V., K.-Q. Tran, and Ø. Skreiberg, Accelerating wet torrefaction rate and ash
removal by carbon dioxide addition. Fuel Processing Technology, 2015. 140: p. 297-303.
[126] Zarchin, R., et al., Hydroprocessing of soybean oil on nickel-phosphide supported catalysts.
Fuel, 2015. 139: p. 684-691.
[127] Phimsen, S., et al., Nickel sulfide, nickel phosphide and nickel carbide catalysts for bio-
hydrotreated fuel production. Energy Conversion and Management, 2017. 151: p. 324-
333.
[128] Ruangudomsakul, M., et al., Influential properties of activated carbon on dispersion of
nickel phosphides and catalytic performance in hydrodeoxygenation of palm oil.
Catalysis Today, 2021. 367: p. 153-164.
[129] Jeong, H., et al., Comparison of activity and stability of supported Ni2P and Pt catalysts in
the hydroprocessing of palm oil into normal paraffins. Journal of Industrial and

of
Engineering Chemistry, 2020. 83: p. 189-199.
[130] Ivars-Barceló, F., et al., 6 Advances in the application of transition metal phosphide
catalysts for hydrodeoxygenation reactions of bio-oil from biomass pyrolysis. Biomass

ro
and Biowaste: New Chemical Products from Old, 2020: p. 145.
[131] Alvarado Rupflin, L., C. Boscagli, and S.A. Schunk, Platinum Group Metal Phosphides as
Efficient Catalysts in Hydroprocessing and Syngas-Related Catalysis. Catalysts, 2018.
8(3): p. 122.

-p
[132] Xin, H., et al., Production of high-grade diesel from palmitic acid over activated carbon-
supported nickel phosphide catalysts. Applied Catalysis B: Environmental, 2016. 187: p.
re
375-385.
[133] Srifa, A., et al., Production of bio-hydrogenated diesel by catalytic hydrotreating of palm
oil over NiMoS2/γ-Al2O3 catalyst. Bioresource technology, 2014. 158: p. 81-90.
lP

[134] Mahdi, H.I. and O. Muraza, Conversion of isobutylene to octane-booster compounds after
methyl tert-butyl ether phaseout: the role of heterogeneous catalysis. Industrial &
Engineering Chemistry Research, 2016. 55(43): p. 11193-11210.
[135] Mahdi, H.I. and O. Muraza, An exciting opportunity for zeolite adsorbent design in
na

separation of C4 olefins through adsorptive separation. Separation and Purification


Technology, 2019. 221: p. 126-151.
[136] Zikri, A., et al. Production of Green Diesel From Crude Palm Oil (CPO) Through
Hydrotreating Process by Using Zeolite Catalyst. in 4th Forum in Research, Science, and
ur

Technology (FIRST-T1-T2-2020). 2021. Atlantis Press.


[137] Suhendi, E., et al., The Effect of Time on the Activation of Bayah Natural Zeolite for
Application of Palm Oil Shell Pyrolysis. Bulletin of Chemical Reaction Engineering &
Jo

Catalysis, 2021: p. 16-3.


[138] Mahdi, H.I., et al., Glycerol carbonate production from biodiesel waste over modified
natural clinoptilolite. Waste and biomass valorization, 2016. 7(6): p. 1349-1356.
[139] Vieira, S.S., et al., Use of HZSM-5 modified with citric acid as acid heterogeneous catalyst
for biodiesel production via esterification of oleic acid. Microporous and Mesoporous
Materials, 2015. 201: p. 160-168.
[140] Wang, C., et al., One‐Step Hydrotreatment of Vegetable Oil to Produce High Quality
Diesel‐Range Alkanes. ChemSusChem, 2012. 5(10): p. 1974-1983.
[141] Azreena, I.N., et al., A promoter effect on hydrodeoxygenation reactions of oleic acid by
zeolite beta catalysts. Journal of Analytical and Applied Pyrolysis, 2021. 155: p. 105044.
[142] Chintakanan, P., et al., Bio-jet fuel range in biofuels derived from hydroconversion of palm
olein over Ni/zeolite catalysts and freezing point of biofuels/Jet A-1 blends. Fuel, 2021.
293: p. 120472.
[143] Freitas, L.N.S., et al., Study of direct synthesis of bio-hydrocarbons from macauba oils
using zeolites as catalysts. Fuel, 2021. 287: p. 119472.
[144] Kang, Y.-H., et al., Green and effective catalytic hydroconversion of an extractable portion
from an oil sludge to clean jet and diesel fuels over a mesoporous Y zeolite-supported
nickel catalyst. Fuel, 2021. 287: p. 119396.
[145] Arumugam, M., et al., Hierarchical HZSM-5 for Catalytic Cracking of Oleic Acid to Biofuels.
Nanomaterials, 2021. 11(3): p. 747.

of
[146] Wang, F., et al., Promoting hydrocarbon production from fatty acid pyrolysis using
transition metal or phosphorus modified Al-MCM-41 catalyst. Journal of Analytical and
Applied Pyrolysis, 2021. 156: p. 105146.

ro
[147] Lee, C.-W., et al., Hydrodeoxygenation of palmitic acid over zeolite-supported nickel
catalysts. Catalysis Today, 2020.
[148] Liu, Y., et al., Rapid and green synthesis of SAPO-11 for deoxygenation of stearic acid to

110280.
-p
produce bio-diesel fractions. Microporous and Mesoporous Materials, 2020. 303: p.

[149] Ghaffar, N., et al. Catalytic Cracking of High Density Polyethylene Pyrolysis Vapor over
re
Zeolite ZSM-5 towards Production of Diesel. in IOP Conference Series: Materials Science
and Engineering. 2020. IOP Publishing.
[150] Choo, M.-Y., et al., The role of nanosized zeolite Y in the H 2-free catalytic deoxygenation
lP

of triolein. Catalysis Science & Technology, 2019. 9(3): p. 772-782.


[151] Putra, R., et al., Fe/Indonesian natural zeolite as hydrodeoxygenation catalyst in green
diesel production from palm oil. Bulletin of Chemical Reaction Engineering & Catalysis,
2018. 13(2): p. 245-255.
na

[152] Zandonai, C.H., et al., Production of petroleum-like synthetic fuel by hydrocracking of


crude soybean oil over ZSM5 zeolite–improvement of catalyst lifetime by ion exchange.
Fuel, 2016. 172: p. 228-237.
[153] Gomes, L.C., et al., Hydroisomerization of n-hexadecane using Pt/alumina-Beta zeolite
ur

catalysts for producing renewable diesel with low pour point. Fuel, 2017. 209: p. 521-
528.
[154] Hachemi, I., et al., Sulfur-free Ni catalyst for production of green diesel by
Jo

hydrodeoxygenation. Journal of Catalysis, 2017. 347: p. 205-221.


[155] Nugraha, R.E., et al., The effect of structure directing agents on micro/mesopore
structures of aluminosilicates from Indonesian kaolin as deoxygenation catalysts.
Microporous and Mesoporous Materials, 2021. 315: p. 110917.
[156] Silva, G.C.R. and M.H.C. de Andrade, Simulation of deoxygenation of vegetable oils for
diesel-like fuel production in continuous reactor. Biomass Conversion and Biorefinery,
2021: p. 1-15.
[157] Fereidooni, L., M. Enayati, and A. Abbaspourrad, Purification technology for renewable
production of fuel from methanolysis of waste sunflower oil in the presence of high silica
zeolite beta. Green Chemistry Letters and Reviews, 2021. 14(1): p. 1-13.
[158] Paniagua, M., et al., Understanding the role of Al/Zr ratio in Zr-Al-Beta zeolite: Towards
the one-pot production of GVL from glucose. Catalysis Today, 2021. 367: p. 228-238.
[159] Zhao, Q., et al., Core–shell structured zeolite–zeolite composites comprising Y zeolite cores
and nano-β zeolite shells: Synthesis and application in hydrocracking of VGO oil.
Chemical Engineering Journal, 2014. 257: p. 262-272.
[160] Kurnia, I., et al., In-situ catalytic upgrading of bio-oil derived from fast pyrolysis of lignin
over high aluminum zeolites. Fuel Processing Technology, 2017. 167: p. 730-737.
[161] Shi, Y., et al., Recent progress on upgrading of bio-oil to hydrocarbons over metal/zeolite
bifunctional catalysts. Catalysis Science & Technology, 2017. 7(12): p. 2385-2415.

of
[162] Glisic, S.B., J.M. Pajnik, and A.M. Orlović, Process and techno-economic analysis of green
diesel production from waste vegetable oil and the comparison with ester type biodiesel
production. Applied Energy, 2016. 170: p. 176-185.

ro
[163] Dickinson, J.G., J.T. Poberezny, and P.E. Savage, Deoxygenation of benzofuran in
supercritical water over a platinum catalyst. Applied Catalysis B: Environmental, 2012.
123: p. 357-366.

-p
[164] de Miguel Mercader, F., et al., Production of advanced biofuels: Co-processing of
upgraded pyrolysis oil in standard refinery units. Applied Catalysis B: Environmental,
2010. 96(1-2): p. 57-66.
re
[165] Damartzis, T. and A. Zabaniotou, Thermochemical conversion of biomass to second
generation biofuels through integrated process design—A review. Renewable and
Sustainable Energy Reviews, 2011. 15(1): p. 366-378.
lP

[166] Naik, S., et al., Characterization of Canadian biomass for alternative renewable biofuel.
Renewable energy, 2010. 35(8): p. 1624-1631.
[167] Papageridis, K., et al., Effect of operating parameters on the selective catalytic
deoxygenation of palm oil to produce renewable diesel over Ni supported on Al2O3, ZrO2
na

and SiO2 catalysts. Fuel Processing Technology, 2020. 209: p. 106547.


[168] Tran, C.-C., D. Akmach, and S. Kaliaguine, Hydrodeoxygenation of vegetable oils over
biochar supported bimetallic carbides for producing renewable diesel under mild
conditions. Green Chemistry, 2020. 22(19): p. 6424-6436.
ur

[169] Kukushkin, R., et al., Deoxygenation of esters over sulfur-free Ni–W/Al2O3 catalysts for
production of biofuel components. Chemical Engineering Journal, 2020. 396: p. 125202.
[170] Phichitsurathaworn, N., L. Simasatitkul, and S. Assabumrungrat, Techno-economic
Jo

feasibility analysis of the H2 recovery in bio-hydrogenated diesel (BHD) from palm oil
process by methanol synthesis via ASPEN Plus. 2020.
[171] Glisic, S. and D. Skala, The problems in design and detailed analyses of energy
consumption for biodiesel synthesis at supercritical conditions. The journal of
supercritical fluids, 2009. 49(2): p. 293-301.
[172] Glisic, S.B. and A.M. Orlović, Review of biodiesel synthesis from waste oil under elevated
pressure and temperature: Phase equilibrium, reaction kinetics, process design and
techno-economic study. Renewable and Sustainable Energy Reviews, 2014. 31: p. 708-
725.
[173] Glisic, S.B. and A.M. Orlovic, Modelling of non-catalytic biodiesel synthesis under sub and
supercritical conditions: The influence of phase distribution. The Journal of Supercritical
Fluids, 2012. 65: p. 61-70.
[174] West, A.H., D. Posarac, and N. Ellis, Assessment of four biodiesel production processes
using HYSYS. Plant. Bioresource technology, 2008. 99(14): p. 6587-6601.
[175] Glišić, S., I. Lukic, and D. Skala, Biodiesel synthesis at high pressure and temperature:
analysis of energy consumption on industrial scale. Bioresource Technology, 2009.
100(24): p. 6347-6354.
[176] Tan, K.T. and K.T. Lee, A review on supercritical fluids (SCF) technology in sustainable
biodiesel production: Potential and challenges. Renewable and Sustainable Energy
Reviews, 2011. 15(5): p. 2452-2456.
[177] Kusdiana, D. and S. Saka, Kinetics of transesterification in rapeseed oil to biodiesel fuel as

of
treated in supercritical methanol. fuel, 2001. 80(5): p. 693-698.
[178] He, H., et al., Transesterification kinetics of soybean oil for production of biodiesel in
supercritical methanol. Journal of the American Oil Chemists' Society, 2007. 84(4): p.

ro
399-404.
[179] Almagrbi, A.M., et al., Determination of kinetic parameters for complex transesterification
reaction by standard optimisation methods. Hemijska industrija, 2014. 68(2): p. 149-159.

-p
[180] Song, E.-S., et al., Transesterification of RBD palm oil using supercritical methanol. The
Journal of Supercritical Fluids, 2008. 44(3): p. 356-363.
[181] Silva, C., et al., Continuous production of fatty acid ethyl esters from soybean oil in
re
compressed ethanol. Industrial & engineering chemistry research, 2007. 46(16): p. 5304-
5309.
[182] Madras, G., C. Kolluru, and R. Kumar, Synthesis of biodiesel in supercritical fluids. Fuel,
lP

2004. 83(14-15): p. 2029-2033.


[183] Duan, P. and P.E. Savage, Catalytic hydrotreatment of crude algal bio-oil in supercritical
water. Applied Catalysis B: Environmental, 2011. 104(1-2): p. 136-143.
[184] Mittelbach, M. and B. Trathnigg, Kinetics of alkaline catalyzed methanolysis of sunflower
na

oil. Lipid/Fett, 1990. 92(4): p. 145-148.


[185] Gopinath, S., et al., Efficient mesoporous SO42−/Zr-KIT-6 solid acid catalyst for green
diesel production from esterification of oleic acid. Fuel, 2017. 203: p. 488-500.
[186] Abdulkareem-Alsultan, G., et al., A review on thermal conversion of plant oil (edible and
ur

inedible) into green fuel using carbon-based nanocatalyst. Catalysts, 2019. 9(4): p. 350.
[187] Hongloi, N., et al., Nickel catalyst with different supports for green diesel production.
Energy, 2019. 182: p. 306-320.
Jo

[188] Asikin-Mijan, N., et al., Production of green diesel via cleaner catalytic deoxygenation of
Jatropha curcas oil. Journal of Cleaner Production, 2017. 167: p. 1048-1059.
[189] Abdulkareem-Alsultan, G., et al., Pyro-lytic de-oxygenation of waste cooking oil for green
diesel production over Ag2O3-La2O3/AC nano-catalyst. Journal of Analytical and Applied
Pyrolysis, 2019. 137: p. 171-184.
[190] Abdulkareem-Alsultan, G., et al., Efficient deoxygenation of waste cooking oil over Co 3 O
4–La 2 O 3-doped activated carbon for the production of diesel-like fuel. RSC Advances,
2020. 10(9): p. 4996-5009.
[191] Wang, W.-C. and C.-H. Hsieh, Hydro-processing of biomass-derived oil into straight-chain
alkanes. Chemical Engineering Research and Design, 2020. 153: p. 63-74.
[192] Borugadda, V.B., G. Kamath, and A.K. Dalai, Techno-economic and life-cycle assessment of
integrated Fischer-Tropsch process in ethanol industry for bio-diesel and bio-gasoline
production. Energy, 2020. 195: p. 116985.
[193] Hari, T.K., Z. Yaakob, and N.N. Binitha, Aviation biofuel from renewable resources: Routes,
opportunities and challenges. Renewable and Sustainable Energy Reviews, 2015. 42: p.
1234-1244.
[194] Pattanaik, B.P. and R.D. Misra, Effect of reaction pathway and operating parameters on
the deoxygenation of vegetable oils to produce diesel range hydrocarbon fuels: A review.
Renewable and Sustainable Energy Reviews, 2017. 73: p. 545-557.
[195] Ochoa-Hernández, C., et al., Hydrocarbons production through hydrotreating of methyl

of
esters over Ni and Co supported on SBA-15 and Al-SBA-15. Catalysis today, 2013. 210: p.
81-88.
[196] Morgan, T., et al., Conversion of triglycerides to hydrocarbons over supported metal

ro
catalysts. Topics in Catalysis, 2010. 53(11-12): p. 820-829.
[197] Puspawiningtiyas, E., et al. Effect of metal type on basic soap pyrolysis produce bio-
gasoline. in IOP Conference Series: Materials Science and Engineering. 2020. IOP
Publishing.

-p
[198] Sihombing, J.L., et al., Characteristic and catalytic performance of Co and Co-Mo metal
impregnated in sarulla natural zeolite catalyst for hydrocracking of MEFA rubber seed oil
re
into biogasoline fraction. Catalysts, 2020. 10(1): p. 121.
[199] Suiuay, C., et al., Effect of gasoline-like fuel obtained from hard-resin of Yang
(Dipterocarpus alatus) on single cylinder gasoline engine performance and exhaust
lP

emissions. Renewable Energy, 2020. 153: p. 634-645.


[200] Xu, H., U. Lee, and M. Wang, Life-cycle energy use and greenhouse gas emissions of palm
fatty acid distillate derived renewable diesel. Renewable and Sustainable Energy
Reviews, 2020. 134: p. 110144.
na

[201] Makertihartha, I.G.B., et al., Biogasoline production from palm oil: optimization of
catalytic cracking parameters. Arabian Journal for Science and Engineering, 2020. 45(9):
p. 7257-7266.
[202] Taufiqurrahmi, N. and S. Bhatia, Catalytic cracking of edible and non-edible oils for the
ur

production of biofuels. Energy & Environmental Science, 2011. 4(4): p. 1087-1112.


[203] Abul-Hamayel, M.A., Kinetic modeling of high-severity fluidized catalytic cracking☆. Fuel,
2003. 82(9): p. 1113-1118.
Jo

[204] Wicke, B., et al., Exploring land use changes and the role of palm oil production in
Indonesia and Malaysia. Land use policy, 2011. 28(1): p. 193-206.
[205] Backhouse, M. and R. Lehmann, New ‘renewable’frontiers: contested palm oil plantations
and wind energy projects in Brazil and Mexico. Journal of Land Use Science, 2020. 15(2-
3): p. 373-388.
[206] Quispe, C.A., C.J. Coronado, and J.A. Carvalho Jr, Glycerol: Production, consumption,
prices, characterization and new trends in combustion. Renewable and sustainable
energy reviews, 2013. 27: p. 475-493.
[207] Fangkoch, S., et al., Solvent-Free Hydrodeoxygenation of Triglycerides to Diesel-like
Hydrocarbons over Pt-Decorated MoO2 Catalysts. ACS omega, 2020. 5(12): p. 6956-
6966.
[208] Rakmae, S., et al., Defining nickel phosphides supported on sodium mordenite for
hydrodeoxygenation of palm oil. Fuel Processing Technology, 2020. 198: p. 106236.
[209] Jin, W., et al., Catalytic conversion of palm oil to bio-hydrogenated diesel over novel N-
doped activated carbon supported Pt nanoparticles. Energies, 2020. 13(1): p. 132.
[210] Srifa, A., et al., NiAl2O4 spinel-type catalysts for deoxygenation of palm oil to green diesel.
Chemical Engineering Journal, 2018. 345: p. 107-113.
[211] Ma, B. and C. Zhao, High-grade diesel production by hydrodeoxygenation of palm oil over
a hierarchically structured Ni/HBEA catalyst. Green Chemistry, 2015. 17(3): p. 1692-
1701.

of
[212] Srifa, A., et al., Catalytic behaviors of Ni/γ-Al 2 O 3 and Co/γ-Al 2 O 3 during the
hydrodeoxygenation of palm oil. Catalysis Science & Technology, 2015. 5(7): p. 3693-
3705.

ro
[213] Liu, Q., et al., One-step hydrodeoxygenation of palm oil to isomerized hydrocarbon fuels
over Ni supported on nano-sized SAPO-11 catalysts. Applied Catalysis A: General, 2013.
468: p. 68-74.

-p
[214] Asikin-Mijan, N., et al., Pyrolytic-deoxygenation of triglycerides model compound and non-
edible oil to hydrocarbons over SiO2-Al2O3 supported NiO-CaO catalysts. Journal of
Analytical and Applied Pyrolysis, 2018. 129: p. 221-230.
re
[215] Pimenta, J.L., et al., Effects of reaction parameters on the deoxygenation of soybean oil for
the sustainable production of hydrocarbons. Environmental Progress & Sustainable
Energy, 2020. 39(5): p. e13450.
lP

[216] Lycourghiotis, S., et al., Nickel catalysts supported on palygorskite for transformation of
waste cooking oils into green diesel. Applied Catalysis B: Environmental, 2019. 259: p.
118059.
[217] Gousi, M., et al., Green Diesel Production over Nickel-Alumina Nanostructured Catalysts
na

Promoted by Copper. Energies, 2020. 13(14): p. 3707.


[218] Kwon, K.C., et al., Catalytic deoxygenation of liquid biomass for hydrocarbon fuels.
Renewable Energy, 2011. 36(3): p. 907-915.
[219] Mortensen, P.M., et al., A review of catalytic upgrading of bio-oil to engine fuels. Applied
ur

Catalysis A: General, 2011. 407(1-2): p. 1-19.


[220] Asikin-Mijan, N., et al., Free-H2 deoxygenation of Jatropha curcas oil into cleaner diesel-
grade biofuel over coconut residue-derived activated carbon catalyst. Journal of Cleaner
Jo

Production, 2020. 249: p. 119381.


[221] Arend, M., et al., Catalytic deoxygenation of oleic acid in continuous gas flow for the
production of diesel-like hydrocarbons. Applied Catalysis A: General, 2011. 399(1-2): p.
198-204.
[222] Asikin-Mijan, N., et al., Catalytic deoxygenation of triglycerides to green diesel over
modified CaO-based catalysts. RSC advances, 2017. 7(73): p. 46445-46460.
[223] Bezergianni, S., et al., Toward hydrotreating of waste cooking oil for biodiesel production.
Effect of pressure, H2/oil ratio, and liquid hourly space velocity. Industrial & engineering
chemistry research, 2011. 50(7): p. 3874-3879.
[224] Liu, Y., et al., Hydrotreatment of vegetable oils to produce bio-hydrogenated diesel and
liquefied petroleum gas fuel over catalysts containing sulfided Ni–Mo and solid acids.
Energy & Fuels, 2011. 25(10): p. 4675-4685.
[225] Chen, N., et al., Effects of Si/Al ratio and Pt loading on Pt/SAPO-11 catalysts in
hydroconversion of Jatropha oil. Applied Catalysis A: General, 2013. 466: p. 105-115.
[226] Kubička, D., P. Šimáček, and N. Žilková, Transformation of vegetable oils into
hydrocarbons over mesoporous-alumina-supported CoMo catalysts. Topics in Catalysis,
2009. 52(1): p. 161-168.
[227] Anand, M. and A.K. Sinha, Temperature-dependent reaction pathways for the anomalous
hydrocracking of triglycerides in the presence of sulfided Co–Mo-catalyst. Bioresource
technology, 2012. 126: p. 148-155.
[228] Patil, S.J. and P.D. Vaidya, On the production of bio-hydrogenated diesel over hydrotalcite-

of
like supported palladium and ruthenium catalysts. Fuel Processing Technology, 2018.
169: p. 142-149.
[229] Huber, G.W., P. O’Connor, and A. Corma, Processing biomass in conventional oil refineries:

ro
Production of high quality diesel by hydrotreating vegetable oils in heavy vacuum oil
mixtures. Applied Catalysis A: General, 2007. 329: p. 120-129.
[230] Hermida, L., et al., Selective acid-functionalized mesoporous silica catalyst for conversion

Engineering, 2018. 34(2): p. 239-265.


-p
of glycerol to monoglycerides: state of the art and future prospects. Reviews in Chemical

[231] Hensley, A.J., et al., Enhanced Fe2O3 reducibility via surface modification with Pd:
re
Characterizing the synergy within Pd/Fe catalysts for hydrodeoxygenation reactions. Acs
Catalysis, 2014. 4(10): p. 3381-3392.
[232] Zheng, H.-Y., et al., An environmentally benign process for the efficient synthesis of
lP

cyclohexanone and 2-methylfuran. Green Chemistry, 2006. 8(1): p. 107-109.


[233] Nikul’shin, P., et al., Co-hydrotreating of straight-run diesel fraction and vegetable oil on
Co (Ni)-PMo/Al 2 O 3 catalysts. Petroleum Chemistry, 2016. 56(1): p. 56-61.
[234] Tran, N.T., et al., Vapor-phase hydrodeoxygenation of guaiacol on Al-MCM-41 supported
na

Ni and Co catalysts. Applied Catalysis A: General, 2016. 512: p. 93-100.


[235] Alharbi, K., E. Kozhevnikova, and I. Kozhevnikov, Hydrogenation of ketones over
bifunctional Pt-heteropoly acid catalyst in the gas phase. Applied Catalysis A: General,
2015. 504: p. 457-462.
ur

[236] Yang, Y., et al., Hydrotreating of C18 fatty acids to hydrocarbons on sulphided NiW/SiO2–
Al2O3. Fuel Processing Technology, 2013. 116: p. 165-174.
[237] Alvarez-Galvan, M.C., et al., Metal phosphide catalysts for the hydrotreatment of non-
Jo

edible vegetable oils. Catalysis Today, 2018. 302: p. 242-249.


[238] Asikin-Mijan, N., et al., Waste clamshell-derived CaO supported Co and W catalysts for
renewable fuels production via cracking-deoxygenation of triolein. Journal of Analytical
and Applied Pyrolysis, 2016. 120: p. 110-120.
[239] Janampelli, S. and S. Darbha, Highly efficient Pt-MoOx/ZrO2 catalyst for green diesel
production. Catalysis Communications, 2019. 125: p. 70-76.
[240] Alvarez-Galvan, M.C., J.M. Campos-Martin, and J.L. Fierro, Transition metal phosphides for
the catalytic hydrodeoxygenation of waste oils into green diesel. Catalysts, 2019. 9(3): p.
293.
[241] Wang, H., Biofuels production from hydrotreating of vegetable oil using supported noble
metals, and transition metal carbide and nitride. 2012.
[242] Liang, J., et al., Chemoselective hydrodeoxygenation of palmitic acid to diesel-like
hydrocarbons over Ni/MoO2@ Mo2CTx catalyst with extraordinary synergic effect.
Chemical Engineering Journal, 2020. 391: p. 123472.
[243] Ameen, M., et al., The effect of metal loading over Ni/γ-Al2O3 and Mo/γ-Al2O3 catalysts
on reaction routes of hydrodeoxygenation of rubber seed oil for green diesel production.
Catalysis Today, 2019.
[244] Sweeny, N.P., C.S. Rohrer, and O. Brown, Dinickel phosphide as a heterogeneous catalyst
for the vapor phase reduction of nitrobenzene with hydrogen to aniline and water.
Journal of the American Chemical Society, 1958. 80(4): p. 799-800.
[245] Shin, M., J. Kim, and Y.-W. Suh, Etherification of biomass-derived furanyl alcohols with

of
aliphatic alcohols over silica-supported nickel phosphide catalysts: Effect of surplus P
species on the acidity. Applied Catalysis A: General, 2020. 603: p. 117763.
[246] Fujita, S., et al., Unique catalysis of nickel phosphide nanoparticles to promote the

ro
selective transformation of biofuranic aldehydes into diketones in water. ACS Catalysis,
2020. 10(7): p. 4261-4267.
[247] Wagner, J.L., et al., Zeolite Y supported nickel phosphide catalysts for the

-p
hydrodenitrogenation of quinoline as a proxy for crude bio-oils from hydrothermal
liquefaction of microalgae. Dalton Transactions, 2018. 47(4): p. 1189-1201.
[248] Landau, M.V., et al., Ultradeep hydrodesulfurization and adsorptive desulfurization of
re
diesel fuel on metal-rich nickel phosphides. Industrial & engineering chemistry research,
2009. 48(11): p. 5239-5249.
[249] Cecilia, J.A., et al., Oxygen-removal of dibenzofuran as a model compound in biomass
lP

derived bio-oil on nickel phosphide catalysts: Role of phosphorus. Applied Catalysis B:


Environmental, 2013. 136: p. 140-149.
[250] Nowak, I. and M. Ziolek, Niobium compounds: preparation, characterization, and
application in heterogeneous catalysis. Chemical Reviews, 1999. 99(12): p. 3603-3624.
na

[251] Ziolek, M. and I. Sobczak, The role of niobium component in heterogeneous catalysts.
Catalysis Today, 2017. 285: p. 211-225.
[252] Nico, C., T. Monteiro, and M.P. Graça, Niobium oxides and niobates physical properties:
Review and prospects. Progress in Materials Science, 2016. 80: p. 1-37.
ur

[253] Shao, Y., et al., Selective production of arenes via direct lignin upgrading over a niobium-
based catalyst. Nature communications, 2017. 8(1): p. 1-9.
[254] Xia, Q.N., et al., Pd/NbOPO4 multifunctional catalyst for the direct production of liquid
Jo

alkanes from aldol adducts of furans. Angewandte Chemie, 2014. 126(37): p. 9913-9918.
[255] Reguera, F.M., et al., The use of niobium based catalysts for liquid fuel production.
Materials Research, 2004. 7(2): p. 343-348.
[256] Brandão, R.F., et al., Synthesis, characterization and use of Nb2O5 based catalysts in
producing biofuels by transesterification, esterification and pyrolysis. Journal of the
Brazilian Chemical Society, 2009. 20(5): p. 954-966.
[257] Rade, L.L., et al., Optimization of esterification reaction over niobium phosphate in a
packed bed tubular reactor. Renewable Energy, 2019. 131: p. 348-355.
[258] Bassan, I.A., et al., Esterification of fatty acids with alcohols over niobium phosphate. Fuel
processing technology, 2013. 106: p. 619-624.
[259] da Conceição, L.R.V., et al., Solid acid as catalyst for biodiesel production via simultaneous
esterification and transesterification of macaw palm oil. Industrial Crops and Products,
2016. 89: p. 416-424.
[260] Tago, T., et al., Size-controlled synthesis of nano-zeolites and their application to light
olefin synthesis. Catalysis Surveys from Asia, 2012. 16(3): p. 148-163.
[261] Santillan-Jimenez, E., et al., Catalytic deoxygenation of triglycerides and fatty acids to
hydrocarbons over carbon-supported nickel. Fuel, 2013. 103: p. 1010-1017.
[262] Santillan-Jimenez, E., et al., Catalytic deoxygenation of triglycerides and fatty acids to
hydrocarbons over Ni–Al layered double hydroxide. Catalysis Today, 2014. 237: p. 136-
144.

of
[263] Adams, P., et al., Biomass conversion technologies, in Greenhouse Gas Balances of
Bioenergy Systems. 2018, Elsevier. p. 107-139.
[264] Kim, S., et al., Recent advances in hydrodeoxygenation of biomass-derived oxygenates

ro
over heterogeneous catalysts. Green Chemistry, 2019. 21(14): p. 3715-3743.

-p
re
lP
na
ur
Jo

You might also like