You are on page 1of 72

Accepted Manuscript

Title: A review on deoxygenation of triglycerides for jet fuel


range hydrocarbons

Authors: Saima Khan, Andrew Ng Kay Lup, Khan


Muhammad Qureshi, Faisal Abnisa, Wan Mohd Ashri Wan
Daud, Muhamad Fazly Abdul Patah

PII: S0165-2370(18)31087-8
DOI: https://doi.org/10.1016/j.jaap.2019.03.005
Reference: JAAP 4567

To appear in: J. Anal. Appl. Pyrolysis

Received date: 19 November 2018


Revised date: 30 January 2019
Accepted date: 7 March 2019

Please cite this article as: Khan S, Ng Kay Lup A, Qureshi KM, Abnisa F,
Wan Daud WMA, Patah MFA, A review on deoxygenation of triglycerides for
jet fuel range hydrocarbons, Journal of Analytical and Applied Pyrolysis (2019),
https://doi.org/10.1016/j.jaap.2019.03.005

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. 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.
A review on deoxygenation of triglycerides for jet fuel
range hydrocarbons

Saima Khana, Andrew Ng Kay Lupa,*, Khan Muhammad Qureshia, Faisal Abnisab, Wan
Mohd Ashri Wan Dauda, Muhamad Fazly Abdul Pataha

a
Department of Chemical Engineering, Faculty of Engineering, University of Malaya, 50603,
Kuala Lumpur, Malaysia

T
b
Department of Chemical and Materials Engineering, Faculty of Engineering, King

IP
Abdulaziz University, Rabigh, 21911, Saudi Arabia.

* Corresponding authors e-mail address: drewanyak@hotmail.com

R
SC
Graphical Abstract

U
N
A
M
ED
PT

Highlights:
 HDO of triglyceride (TGs) based oils for jet fuel production.
E

 Study of HDO reaction pathway.


CC

 Effects of active metal, support and promoters in HDO of TGs.


 Effect of operating parameters on HDO processes.
A

Abstract:
Production of bio-jet fuel from triglycerides (TGs) based vegetable oils have recently received
increased attention from research and industries because of its renewability and environmental
benefits. Catalytic deoxygenation (DO) is a suitable way to produce bio-jet fuel from TGs.
Presently, the main challenges faced by DO are the optimized selection of feedstocks,

1
catalysts, reaction pathways and parameters. This review includes discussion on the feedstock
and assessment of several potential catalysts: noble metals, sulphided, non-sulphided for DO
of TGs. This assessment elucidates the model compounds of TGs, effect of operating
parameters, potential catalysts and different DO reaction pathways to attain optimum yield
and selectivity of desired products. In addition, some relevant discussion of TGs derived jet
fuel specification, characteristics and fuel properties are also discussed. Overall, this review
provides a comprehensive discussion on DO of TGs based vegetable oils in enhancing
alternative jet fuel production in all relevant technical aspects.

T
IP
Key words: Hydrodeoxygenation, triglycerides, catalyst, jet fuel.

R
SC
Table of Contents
Table of Contents................................................................................................................................... 2
1.
U
Introduction ................................................................................................................................... 3
N
2. Overview of Triglycerides and Jet Fuel....................................................................................... 5
2.1 Characteristics and properties of triglycerides ..................................................................... 5
A
2.2 Jet fuel specifications ............................................................................................................. 7
2.3 Characteristics and properties of jet fuel............................................................................... 9
M

3. Selection of feedstock/model compound for jet fuel production ............................................. 12


4. Reaction pathway of deoxygenation .......................................................................................... 16
ED

4.1 Hydrogenation ..................................................................................................................... 17


4.2 Decarboxylation (DCO2) ..................................................................................................... 19
4.3 Decarbonylation (DCO) ...................................................................................................... 20
PT

4.4 Hydrodeoxygenation (HDO) ................................................................................................ 21


4.5 Cracking and isomerization reactions ................................................................................. 22
5. Deoxygenation reaction mechanism of model compounds ...................................................... 24
E

5.1 Methyl laurate ...................................................................................................................... 24


CC

5.2 Tristearin and tripalmitin..................................................................................................... 26


5.3 Tristearin and stearic acid ................................................................................................... 27
5.4 Triolein................................................................................................................................. 28
5.5 Palm kernel oil ..................................................................................................................... 29
A

5.6 Karanja oil ........................................................................................................................... 31


6. Deoxygenation catalyst................................................................................................................ 32
6.1 Properties of catalysts.......................................................................................................... 32
6.2 Types of catalysts ................................................................................................................. 34
6.2.1 Noble metals ............................................................................................................... 35
6.2.2 Sulfided/non-sulfided catalysts................................................................................... 37

2
6.3 Choice of support ................................................................................................................. 38
6.4 Active metal or promoter ..................................................................................................... 42
7. Effect of operating conditions on DO of TGs to jet fuel conversion ....................................... 44
7.1 Type of reactor ..................................................................................................................... 44
7.2 Effect of temperature............................................................................................................ 46
7.3 Effect of pressure ................................................................................................................. 48
7.4 Effect of residence time ........................................................................................................ 49
7.5 Amount of catalyst................................................................................................................ 50
8. Industrialization of bio-jet fuel................................................................................................... 51

T
9. Conclusion .................................................................................................................................... 52
References ............................................................................................................................................ 53

R IP
1. Introduction
Aviation or jet fuel is a multi-component fuel with a carbon chain length of C8-C16 [1,2].

SC
Commercial and military jet fuels are mixtures of paraffins, naphthenes or cycloparaffins,
aromatics and olefins compounds [2,3]. Paraffins, iso-paraffins and cycloparaffins
(approximately 70 to 85%) are the major components [2] which are mainly responsible for
U
N
reducing the freezing point of jet fuel [2,4]. Aromatics play a key role in aviation fuel since
A
they enhance the energy density, shrinkage of aged elastomer seals and minimize fuel leakage
issues. The aromatic hydrocarbons in the existing jet fuel are typically present in the range of
M

8-25% [2]. However, high aromatic hydrocarbons may lead to more smoke formation which
can damage the combustion chambers or the turbine blades, causing a reduction in engine
ED

lifespan [4]. The jet fuel utilization worldwide is in excess of 800 million liters/day, which
accounts for 10% of the overall transportation energy [5]. International Air Transport
PT

Association (IATA) estimated that the energy demand of jet fuel will be increased annually
by 5% till 2030 [6]. Due to depletion of petroleum resources and increasing demand of jet
E

fuel, scientists are currently moving towards the use of triglycerides (TGs) as one of the
potential feedstocks for alternative jet fuel production [7].
CC

Triglycrides (TGs) are the main constituents of animal fats and vegetable oils that are
A

composed of one glycerol molecule attached with three free fatty acid molecules. Typically,
TGs differ by their carbon chain length, number of double bonds and the type of oil sources
[8]. Different types of TGs have been widely used for the production of jet fuel from plant-oil
TGs [9], coconut oil [10], castor oil [11], jatropha oil [12], soybean oil [13] and palm kernel
oil [14]. Animal fats and vegetable oils would be hydrotreated to produce high cetane number
and straight chain alkanes ranging from C9 to C18 that can be used in the aviation industry
3
[15]. Furthermore, TGs based vegetable oils can be directly converted into cycloparaffin and
aromatic components for jet fuel production. These components basically meet the
requirements of jet fuel quality based on the H/C mole ratio, heat of combustion and the
average molecular formulation [3]. Therefore, these important properties are making TGs as
potential feedstock for production of renewable and sustainable jet fuel. In addition, TGs
based feedstocks contain lower oxygen content ranging from 8-22% [16,17] while biomass
pyrolysis oil contains higher oxygen content in the range of 26-47% [18]. High oxygen
content of fuel is undesirable as it may lead to some major problems such as low heating

T
value, high corrosivity and instability issues when used as fuel [16,19]. To overcome these

IP
fuel related issues, different routes for production of high quality jet fuel from vegetable oil

R
have been researched. In many cases, hydrodeoxygenation is an essential step to convert TGs
into deoxygenated jet fuel that are highly compatible with conventional petroleum based fuels

SC
[20].

U
Deoxygenation reaction (DO) is a viable process to remove oxygen from TGs based oil to
N
form linear long chain hydrocarbons [21,22]. The DO process of vegetable oil was initially
commercialized by the Finland Company, Neste Oil [23]. Moreover, DO process can be
A
performed via two reaction pathways such as indirect and direct DO pathway. In case of
M

indirect DO pathway, the fatty acids, alcohol and/or fatty acid ester are produced as
intermediate products through deoxygenation reaction conditions by β-elimination in the
ED

presence of pure H2 pressure [24]. Through β-elimination, TGs release fatty acids and become
unsaturated glycol difatty acid ester (UGDE). Subsequently, UGDE releases fatty acid and
PT

transformed to alkanes or normal paraffin [24]. On the other hand, the direct DO pathway
transforms TGs to hydrocarbons without any formation of fatty acid intermediates. The exact
mechanism of direct DO is still not clearly known [24].
E
CC

In regard to the catalyst for DO application, researchers are more intensive to find out the
most suitable catalyst which shows high stability, surface area and extensive range of
A

porosity. From the previous findings, it was proven that the hydrocarbon conversion of TGs
does not merely depend on the catalyst activity but also the selectivity and nature of the
catalysts [9]. Some catalysts with high acidity showed severe hydrocracking and consequently
reduced hydrocarbon yield. However, catalysts with moderate acidity are good candidates for
DO reaction and more alkanes formation with C7-C14 selectivity [23]. Besides that, the
amount of catalyst plays an important role in hydrocracking activity and product selectivity

4
during DO reaction. In addition, low amount of catalyst provides low acidity in DO reaction.
While the high amount of catalyst loading leads to high cracking and more gas formation
which resulting as reduced jet-fuel yield [11]. Additionally, performance of catalyst is also
dependent on high H2 pressure condition. Low H2 pressure is commonly preferred in
decarboxylation (DCO2) and decarbonylation (DCO) reaction, while high H2 pressure is more
desired in the HDO reaction of TGs for O2 removal [25]. Likewise, different temperatures in
catalytic HDO reaction pathways have a great effect on high selectivity. Some researchers
have noted that high temperature is more desired for DCO2 and DCO due to the endothermic

T
nature [26]. However, HDO reaction was found as enormously exothermic in nature [14,26].

IP
From the previous findings, it is proven that the operating parameters: feedstock, temperature,

R
pressure, catalyst type, catalyst quantity and residence time are the main factors that influence
the DO process as well as product selectivity of hydrocarbons.

SC
A descriptive analysis was done in regard to the potential use of several TGs based oils in

U
DO process for jet fuel production. It has been noted that the product yield and quality are
N
heavily dependent on different DO reaction pathways, selection of TGs based oil, operating
parameters, type of catalyst and amount of catalyst. Therefore, this review attempts to
A
highlight the different DO reaction pathways and model compounds of TGs to obtain jet fuel
M

hydrocarbons. Different type of catalysts with the main affecting parameters in DO of TGs
based oils which are instrumental in maximizing the jet fuel production and to improve the
ED

fuel quality were also discussed. The main parameters: temperature, pressure, residence time,
catalyst type, catalyst support type, promotor and catalyst quantity are broadly described here.
PT

Additionally, some relevant discussion on TGs as alternative jet fuel for the aviation industry
with their fuel specifications, characteristics and fuel properties were also presented in this
review.
E
CC

2. Overview of Triglycerides and Jet Fuel


A

2.1 Characteristics and properties of triglycerides


At present, TGs have shown high potential to be used as an alternative source of jet fuel
production due to its simple chemical structure and lower oxygen content in contrast to
biomass pyrolysis oil. The major fatty acids in TGs based oil are palmitic acid (C16:0), oleic
acid (C18:1), linoleic acid (C18:2), palmitic acid (C16:0) and stearic acid (C18:0), while
minor fatty acid in TGs based oil are arachidic (C20:0), eicosenoic, (C20:1), behenic (C22:0)

5
acids (Table 4). In addition, alkanes, alkenes, aromatics, carboxylic acids, aldehyde and
ketone are also the main components found through pyrolysis of TGs based oils [17,27].
Furthermore, TGs based oils may possibly lead to more paraffinic and less aromatic
compounds on catalytic cracking which are preferred for better engine performance [23]. In
fact, TGs derived oil contains paraffins from C8-C24 with main composition of C16 and C18 of
hydrocarbon chains [28] which can be simply converted into jet fuel without the use of an
energy intensive process [29].

T
Literature accounts that, TGs based oils consisting of around 57 wt.% paraffinic straight

IP
chain hydrocarbons which are based on n-alkanes and alkenes contents [27] whereas the

R
percentage of aromatic compounds accounts about 13.56 wt.% [17]. These chemical
properties of TGs are making them as an appropriate option for gasoline or jet fuel production

SC
[8]. Furthermore, Kim et al. also suggested that TGs are showing good potential to be used for
commercial airlines and in jet fuel industry due to the presence of high paraffinic components.

U
Therefore, to produce high yield of light hydrocarbons, some typical physicochemical
N
properties such as water content, acid number, viscosity, density and higher heating value
(HHV) of vegetable oil are considered in Table 1. In general, TGs based oils have lower
A
oxygen content than other organic feedstock such as wood, bagasse and industrial residues.
M

The typical amount of oxygen content in TGs is 10 to 22 wt.% [17] and low acidity of 5.36 to
6.96 mg KOH/g [17,30].
ED

Furthermore, densities of TGs based vegetable oils vary from 0.88 g/mL to 0.97 g/mL
PT

due to different sources [38]. Low density is more suitable for jet fuel production because it
enhances the liquid flow property of the jet fuel product [39]. The viscosities of TGs based
oils are noticeably higher, ranging from 25 cP to 50 cP due to the long carbon chains and
E

larger molecules with oxygen atoms [33]. Another important property of TGs based oil is
CC

higher heating value (HHV). HHV of TGs based oil ranges from 37.1 MJ/kg to 40.6 MJ/kg,
which are significantly higher than the HHV of biomass derived pyrolysis oils which ranges
A

from 16.00 MJ/kg to 20.00 MJ/kg. These biomass derived pyrolysis oils are commonly
produced from lignocellulosic biomass feedstock such as softwoord [17], hardwood [40],
bagasse [41], waste tires and plastic [42]. It is generally observed that biomass pyrolysis oil
has much higher oxygen content which tends to reduce its heating value [43]. From this study,
it can be determined that physicochemical properties: density, viscosity, acidity and higher
heating value of TGs are more favorable as compared to biomass pyrolysis oil, making them

6
to be more suitable as an alternative source for jet fuel, gasoline or transportation fuel
productions [8,33,43].

2.2 Jet fuel specifications


Typically, jet fuel composed of different types of compounds such as 20% of paraffins,
40% of isoparaffin, 20% of naphthenes and 20% of aromatic [44]. The major components in
jet fuel are paraffins, iso-paraffins and cycloparaffins (about 70 to 85%) [2] which are
essentially responsible for decreasing the freezing point in jet fuels [4]. The high hydrogen-to-

T
carbon ratio of normal and iso-paraffins offers high heat-to-weight ratio and cleaner

IP
combustion. However, cycloparaffins may possibly reduce the hydrogen-to-carbon and heat-

R
to-weight ratios. Additionally, it reduces the fuel freezing point which is an important
parameter for safe fuel operation at high altitude flights [2]. However, the presence of

SC
aromatics may lead to more smoke formation which causes damage to combustion chamber or
turbine blades [4]. Furthermore, jet fuel also contains traces of sulfur, nitrogen and oxygen

U
containing compounds which may have an adverse effect on the anti-oxidation property and
N
lubricity of jet fuel [2].
A
Fuel specification is a list of standards through which the producers and users can
M

identify and control the necessary fuel properties for suitable and reliable engine performance.
The first fuel specification was made available in 1943 in the UK and 1944 in US [2]. Two
ED

organizations have taken the leading roles in setting and maintaining the specifications for jet
fuels, namely the American Society for Testing and Material (ASTM) and the United
PT

Kingdom Ministry of Defense [45]. In fact, the ASTM standard specification of D1655
basically defines the specific types of jet fuels preferably used in civil aviation, which are Jet
A and Jet A-1. It also specifies that, prior to its use, jet fuel must be tested for volatility,
E

fluidity, combustion properties, corrosion, thermal stability, contaminants and additives to


CC

fulfill the essential fuel requirements. However, on the behalf of United Kingdom, the
Ministry of Defense Aviation Fuels Committee issued another standard that is Defense
A

Standard 91-91 (DEFSTAN91-91) for aviation turbine fuel that defines Jet A-1 requirements.
According to the DEFSTAN 91-91, Jet A-1 is very similar to ASTMD1655 except for a small
number of areas where DEFSTAN 91-91 is more stringent on the additional synthetic
components which are required for jet fuels [45].

7
The quality of jet fuel depends on its physicochemical properties which can classified as
oxygen content, dynamic viscosity, heating value, density and flash point. Jet fuel
specifications and requirements are mostly defined in terms of the required performance
properties. The specifications required for jet fuels are minimum energy density by mass,
maximum freeze point temperature, maximum allowable viscosity, acidity and minimum
aromatic content [39]. Based on these properties, jet fuel can be classified according to the
standards into Jet A, Jet A-1, JP-5 and JP-8 [2,39]. These fuel specifications are defining the
specific type of jet fuel such as typical Jet A and Jet A-1 which are used in commercial

T
airplanes and Jet A-5 and JP-8 which are commonly used in military aircrafts. Jet A fuel is

IP
generally used in the United States, while Jet A-1 is adopted in the rest of the world. The only

R
difference between these two jet fuels is the freezing point for Jet A and Jet A-1 which are -
40 °C and -47 °C respectively [2]. Furthermore, JP-5 is a high-flash-point jet fuel developed

SC
by the Navy. JP-5 is a specifically refined type of kerosene oil consisting of C9-C16
hydrocarbons. However, JP-8 is comparable to Jet A-1 commercial fuel, it was precisely

U
developed for Air Force application to provide a safe kerosene-based jet fuel that have better
N
reliability and acceptable freezing point [2]. Therefore, some jet fuel specifications and
properties are summarized in Table 2.
A
M
ED
E PT
CC
A

8
2.3 Characteristics and properties of jet fuel
Alternative jet fuel production is one of the key initiatives to mitigate the growth
limitations of the aviation industry. There are several conversion pathways that have been
accepted for jet fuel production. Recently, many researchers have indicated that DO of TGs is
one of the most promising pathways to produce bio-jet fuel which is in compliance with
coventional aviation fuel specifications. The physicochemical properties of alternative jet
fuels: net heat of combustion, viscosity, density, flash point, freezing point, final boiling
point, acidity, aromatics and smoke point were presented in Table 2.

T
IP
Net heat of combustion or higher heating value (HHV) is one of the important properties

R
of jet fuel specification to measure the maximum amount of heat energy released during the
combustion [35]. Some authors reported that the calculated calorific value of alternative jet

SC
fuel was noticed to be comparable with commercial grade gasoline and jet fuel, making it a
suitable alternative for jet fuel production [2,50]. Typically, the minimum limit of heating

U
value of 42.8 MJ/kg and freezing point of -47°C are recommended for aviation fuel
N
specification [51].
A
Kinematic viscosity is an important property which affects the fuel flow property [52].
M

Fuels with high viscosity may have problems in atomization and damages the fuel injector
system. This would cause incomplete combustion, poor engine performance and unburned
ED

solid particle deposition. On the other hand, fuels of lower viscosity may lack in lubrication
for pump and injector systems. According to ASTM D-1655 standard, the optimum viscosity
is prescribed within the range as 8 mm2/s at -20°C for jet A/jet A-1 and 8.5 mm2/s for JP-5
PT

[51]. The viscosity value for production of alternative jet fuel with different vegetable oils
was determined as outlined in Table 2. Based on Table 2, it can be seen that the viscosity
E

values of hydro-processed esters and fatty acids (HEFA) [53], Camelina oil [39], Carinata oil
CC

[39] were comparable to the viscosity of conventional jet fuel at -20°C. Many researchers also
observed the enhancement effect on fuel viscosity in DO reaction over TGs. They indicated
A

that when O2 was removed from TGs, the intermolecular boundaries of hydrogen bonds and
liquid viscosity were noticeably decreased which are desired conditions for fuel property [54].

As for fuel density, density of jet fuel ranges from 775 to 840 kg/m3 [46] and can be
controlled by keeping the fuel specification like temperature. This is because the fuel density
is a temperature dependent property, as the temperature increases, the density of the fuel will

9
also decrease. Tim et al. studied the effect of temperature on product density achieved through
DO of TGs. They informed that the product density was slightly changed from 775 to
785 kg/m3 when the temperature was varied. Therefore, these density values are still in
compliance with the commercial jet fuel specifications [46].

In order to elucidate the combustion phenomena and fuel properties in jet fuel
application, flash point is one of the important properties which gives off sufficient vapors to
ignite in air under external flame source. Different authors have indicated that the flash points

T
of bio-jet fuels derived from vegetable oils was close to the commercial jet fuel. Based on

IP
Table 3, it can be observed that the flash point of alternative jet fuel derived from vegetable

R
oil was comparable to the light petroleum distillate fuel. According to ASTM D 1655
standard, the minimum flash point for jet fuel is 38oC while the minimum flash point based on

SC
thefuel specification of US Navy would be 60oC. From a safety point of view, higher flash
point and more ignition time are preferred for jet fuel application [51].

U
N
A
M
ED
E PT
CC
A

10
Freezing point of an aviation fuel is the lowest temperature at which the fuel remains free
of solid hydrocarbon crystals that can restrict the flow of fuel through filters [56]. Generally,
the freezing point has great influence on the fuels pump ability or mass transfer rate at low
temperature condition. Based on ASTM D 1655, the freezing point specifications for Jet A
and Jet A-1 are -40 oC and -47 oC [51]. The freezing point is strongly dependent on carbon
number and n-paraffins fractions [45]. The lowest freezing point is due to the lesser amount of
n-paraffins existing in TGs based jet fuel product [54]. Besides that, corrosion problem is
typically related to the total acid number in the fuel. Based on ASTM D 1655, the acid

T
number of jet fuel must be less than 0.1 mg KOH/g [51]. The fuel acid number represents the

IP
content of macromolecular organic acids, which normally contains naphthenic acids. The
extreme quantity of these acids will possibly lead to severe metal corrosion as such levels of

R
acidity must be kept low to prevent any such occurrence [57].

SC
Aromatic contents are important characteristics for jet fuels as they have the greatest
impact on volatile number and mass emission indices (EIs) in all engine powers [58]. The
U
aromatics content significantly varies with the type of oil [13]. Typically, jet fuel contains
N
aromatic content of about 5-25% [59]. Based on Table 2 and 3, it can be seen that all
A
alternative jet fuel derived from vegetable oil contains comparable aromatic contents with the
standard jet fuel aromatic. Some researchers reported that the aromatic content of less than
M

8% in jet fuel may possibly induce seal contraction which leads to leakage and system
depressurization. Therefore, the aromatic content of synthetic fuel blends is currently fixed at
ED

a minimum of 8% [2]. More aromatic contents result in lower hydrogen content and heavier
paraffins formation, leading to more soot formation [2]. Besides that, an increase in aromatic
PT

content also reduces the net heat of combustion and decreases the freezing point [59].
E

Another important property of jet fuel is its smoke point. It is termed as the maximum
CC

flame height at which fuel will burn without smoking, which can be determined by the height
of flame at which soot generation is initiated. The sooting tendencies of the fuels were found
to increase with the increase in the C/H and C/O ratios [60]. Furthermore, Xue et al. described
A

that the soot formation in jet fuels is directly related to the aromatic and hydrogen contents of
the test fuels. They also stated that fuels with higher aromatics content or smaller hydrogen
content produce more soot and thus having lower smoke point. This is because the
combustion of heavier hydrocarbon components has higher sooting propensity [47].

11
Several studies were also conducted on alternative jet fuel quality by DO of TGs based
oil for jet fuel production. Verma et al. performed DO of jatropha oil to investigate on the
liquid hydrocarbon products in regard to their heat of combustion, density, flash point,
freezing point and aromatic content. They observed a maximum quality of jet fuel products
such as net heat of combustion of 43.3MJ/kg, density of 776 kg/m3 at 15°C, flash point of
47°C, maximum freezing point of -55.1°C, acidity of 0.022 mg KOH/g, maximum aromatics
of 8.0 vol. % and smoke point of about 24 mm were measured [12]. Likewise, Anand et al.
studied the physicochemical properties of alternative jet fuel from jatropha oil over Ni-

T
W/SiO2-Al2O3 catalyst system. With continuous operation of 100 hours, they observed several

IP
products such as naphtha, kerosene and diesel in the product and the properties of fuel product
were in comparable range with the Jet A-1 fuel specifications [61].

R
SC
3. Selection of feedstock/model compound for jet fuel production
Selection of feedstock for biofuel production specifically for jet fuel production depends
U
on two sets of criteria. The first important set of criteria would be the number of double
N
bonds, composition of the fatty acids, cold flow properties and cetane number [62]. While, the
A
second set of criteria would be the feedstock availability, transportation cost and economic
feasibility [63].
M

The first set of criteria for selection of feedstock for jet fuel production is dependent on
ED

the number of double bonds and composition of fatty acids in TGs based oils. This is because
the the degree of unsaturation and the composition of fatty acids would affect the chemical
PT

stability and reactivity in undergoing different deoxygenation reaction pathways to form


different deoxygenated products for jet fuel. Typically, TGs comprise of saturated and
E

unsaturated bonds of fatty acid which are typically found in isomeric form. Besides that,
CC

another important property of saturated fatty acids is high oxidative stability. Unsaturated
fatty acids possess poor oxidation stability due to the presence of double bonds. Therefore, to
solve the poor oxidation stability, hydrogenation or HDO is one of the possible techniques to
A

form saturated hydrocarbons [62]. A decrease in degree of unsaturation may enhance the gas
production or light hydrocarbon yield [64].

Chuan et al. also noticed that the decrease in degree of unsaturation of oil would decrease
the hydrocarbon liquid yield and increase the gas yield from 63.6 to 44.3 wt% and 10 to 14

12
wt% respectively. The increased gas yield may be attributed to zeolite acidity, BET surface
area, and pore volume. In addition, HZSM-5 with a higher acidity content dominated the
catalytic cracking performance, in which more monomolecular cracking took place which
resulting lighter hydrocarbons and gas product [65]. Rao et al. also reported similar effects of
decrease in degree of unsaturation of oil over catalytic cracking using zeolite catalyst on
triglycerides (e.g., rapeseed oil) into a wide range of products. They observed several products
after catalytic cracking of TGs such as gasoline, light cycle oil [66], light gaseous
hydrocarbons, and some heavy hydrocarbons containing negligible content of oxygenated

T
hydrocarbons [67].

IP
Additionally, limitation of unsaturated fatty acids of glycerides is essential because on

R
higher temperature, it leads to polymerization reaction which increases the deposits formation

SC
or reduces fuel lubricity. The correlation between the iodine value and the degree of
unsaturation [68] in a mixture of fatty acid or TGs can be specified by iodine value. The
composition of fatty acids in oil/TGs is often determined by the esterification method with
U
strong acids such as Boron trifluoride (BF3) or oxalic acids [69]. Composition of feedstock
N
has strong impact on product distribution, product yield [2] and product properties such as
A
cetane number and cold flow properties [70]. The contents and compositions of fatty acid in
different vegetable oils are presented in Table 4 [69].
M

Cetane number is widely used for fuel quality parameter which linked to the ignition
ED

delay time and burning quality [41]. A higher cetane number would enable better ignition
properties [68]. Besides that, high cetane number would indicate a fuel having good cold fuel
PT

flow property which can reduce the formation of white smoke. A delay in ignition time or a
change in cetane number was reported to be dependent on some physicochemical properties:
E

type of fuel, volatility, viscosity, surface tension and molecular structure [71]. The presence of
CC

highly mono-unsaturated compounds like (C18:1), (C20:1) and (C22:1) are showing a trend of
higher cetane number in contrast to poly-unsaturated compounds such as linoleic (C18:2) and
linolenic (C18:3) acids. Ramos et al. described that the longer the fatty acid carbon chains and
A

the more saturated the molecules, the higher the cetane number [68,72]. Therefore, a proper
analysis on the TG model compounds within feedstock for its selection is important to
produce jet fuel that is of higher cetane number.

13
The second set of criteria for selection of feedstock is the feedstock availability and the
transportation cost. In the past decades, sunflower and rapeseed oil were commonly used for
biofuel production due to their high production rates in European region [9,39]. The major oil
producing countries with high production rates are summarized in Table 5.

According to the department of foreign agricultural service, the worldwide major


sunflower oil producing countries are: Ukraine, Russian Federation, European Union and
other countries with respective production rates of 5110.172, 3934.46, 2990.081 and 2189.944

T
million tons during the year 2017/18. China, Canada, India and Northern Europe are the major

IP
producers of rapeseed or canola seed oil with production rates of about 7 to 10 million

R
tons/year. Canada only exports about 3-4 million tons of canola seed/year and canola oil about
8 million tons/year [74]. In tropical countries, particularly in south-east Asian countries,

SC
Malaysia, Indonesia and Thailand are the major producers of coconut and palm oil. Food and
Agriculture Organization (FAO) reported that Indonesia is one of the largest palm oil

U
producers with a production rate of 29.27 million tons/year. While, Malaysia is the second
N
largest palm oil producer with a production rate of 19.67 million tons/year [75]. As the second
largest palm oil producer, Malaysian government planned to decrease their crude oil imports
A
by using palm oil as alternative feedstock for jet fuel production through DO pathway [74].
M

Similarly, coconut oil also appeared as another potential feedstock for production of bio-fuel
or jet fuel, since it contains substantial amount of saturated fatty acids [77].
ED

There are several sources of TGs based feedstocks which can be used as model
PT

compounds in the DO process since they have similar molecular structures. Therefore, model
compound studies can help to describe the kinetic trend and reaction mechanism in order to
obtain information on hydrocarbon product selectivity [78]. In this regard, several types of
E

feedstocks and their model compounds are presented in Fig. 1 that can be used for selective
CC

hydrocarbon production through DO pathway.


A

14
T
R IP
Fig. 1. Common sources of TGs based feedstocks and their model compounds for DO.

SC
For jet fuel production, a great number of edible and non-edible oils were studied in their
feedstock suitabilities. In this aspect, many researchers have used edible oils [3,13] and non-

U
edible oils [79]. Edible oil as feedstock is not an economically feasible option for jet fuel
N
production because it can possibly cause food security issue in developing countries typically
A
in Asian sub-continent [74]. On the other hand, non-edible oil feedstock such as jatropha oil,
algae oil [39], non-edible sunflower [32] and non-edible green canola seed [74] are more
M

suitable options for jet fuel production, since that cannot be used as human food due to their
toxic property [80]. The advantage of non-edible oil crops are their lower cultivation cost than
ED

edible oil crops [4] which makes them more suitable for jet fuel production [4]. Choi et al.
also described that high quality jet-fuel product of about 31% from non-edible oils can be
PT

produced catalytically, thus making non-edible oils as suitable feedstocks for jet fuel
production [81].
E
CC

Economic feasibility is another important feature for feedstock selection. From this point,
hotels, restaurants and bakeries are the main sources of waste cooking oil [82]. In addition,
one major advantage of using waste cooking oil is its economic feasibility, since it is 2 to 3
A

times cheaper than unused cooking oil and recently appeared as an alternative source of
energy [83]. Furthermore, some researchers described that waste cooking oil and waste fish
oil encouraged the address of environmental problems and waste disposal issues, since these
feedstocks are not in use for human consumption or in agriculture field [8]. Many researchers
have reported that China is the largest producer of waste cooking oil in the world with a

15
production rate of 500 million tons each year, which is typically generated from the catering
industry of big and average cities. If the waste cooking oil is not managed properly, then it
will pollute the land and cities [82,84]. Therefore, utilization of waste cooking oil for jet fuel
application is a potential option in waste cooking oil management [83].

From the biological research, microalgae also appeared as a potential source of bio-jet fuels as
it has a high oil content and fast growth rate [73]. It has an ability to isolate carbon dioxide,
ability to grow in non-arable and low quality agricultural lands or saline water and do not

T
disturb the agronomy, since having no overlap with food supply [85]. It also helps to reduce

IP
the food-fuel competition as compared to other edible oil sources [20,39]. However, there are

R
some drawbacks related to microalgae commercialization for jet fuel production. Those can
be counted as sensitivity to weather condition, large land area requirement, extensive demand

SC
of energy for water circulation, lack of adequate mixing, irregular distribution of solar light,
low productivity and risks of culture contamination [86]. Due to these drawbacks, microalgae
for jet fuel production is still yet to be commercialized.
U
N
A

4. Reaction pathway of deoxygenation


M

Deoxygenation (DO) pathway is a promising approach for the production of jet fuel from
TGs with a suitable catalyst under high H2 pressure condition [23]. Deoxygenation is an
ED

exothermic reaction which produces n-alkanes with the same number of carbons as the
corresponding fatty acids and eliminates water as a byproduct [87]. It is actually a
PT

hydrogenolysis process for removing the oxygen from oxygen containing compounds.
Furthermore, the deoxygenation reaction of TGs typically involves a complex reaction
mechanism.
E
CC

In DO reaction, the unsaturated bonds of TGs can be transformed to saturated bonds under
high H2 pressure over suitable catalyst. After that, saturated TGs bonds could be further
A

converted into fatty acids through intermediate ‘di’ and ‘mono’ glycerides [28]. Subsequently,
the formation of C—C bonds and breaking of glycerol compound would lead to formation of
propane and free fatty acids [53]. Saturated fatty acids are further reduced into hydrocarbons
through three possible pathways: decarboxylation (DCO2), decarbonylation (DCO) and HDO
reactions (Fig. 2) [23,88]. DCO2 and DCO are endothermic reaction in which O2 is removed
from fatty acid and n-alksanes are produced with one less carbon atom as compared to the

16
original fatty acid composition. In addition, DCO2 produces CO2 as by-product and DCO
reaction forms CO and water as by-product [53]. However, HDO pathway is specifically
performed under high H2 pressure to produce alkane with the same number of carbon atoms as
the fatty acid chain and two moles of water.

T
R IP
SC
U
Fig. 2. Possible deoxygenation reaction pathway for TG upgrading
N
A
To meet the jet fuel specification, cloud point and viscosity of linear paraffins obtained
from DO process are generally higher, resulting in poor cold flow properties [89]. Therefore,
M

it is necessary to hydrocrack and hydro-isomerize the normal paraffins produced during DO


reaction. This would result in lighter alkanes formation, ranging from C5 to C15 in order to
ED

improve the cold-flow properties and product selectivity for jet fuel production [39].
PT

4.1 Hydrogenation
Hydrogenation of TGs based feedstock transforms unsaturated hydrocarbons into
saturated hydrocarbons under H2 pressurized condition. The leading pathway
E

hydrogenation/hydrolysis of TGs is combined with either β-elimination or γ-hydrogen transfer


CC

as shown in Fig. 3 [90].


A

17
T
IP
Fig. 3. Dominant pathways in the thermal cracking of TGs determined by (a) β-elimination

R
and (b) γ-hydrogen transfer. Redrawn based on [90].

SC
Fig. 3(a) shows severe β-elimination in which the releasing of one fatty acid of TGs

U
under H2 pressure produces carboxylic acid/fatty acid and unsaturated glycol di-fatty
N
ester/diglycerides within 20 min [62]. Subsequently, hydrogenation of diglycerides through β-
elimination forms a negligible extent of monoglycerides in the form of palmitic acid (C16:0),
A
stearic acid (C18:0) and arachidic acid (C20:0) as intermediate products in 40 min. However,
M

majority of the monoglycerides were converted into free fatty acids (FFAs) and glycerol after
60 min. A complete set of reaction mechanism of TGs to FFAs is given in Eqn. 1 [91]:
ED

Triglyceride + H2O→Diglyceride + FFA


Diglyceride + H2O →Monoglyceride + FFA
PT

Monoglyceride + H2O →Glycerol + FFA (1)


E

Fig. 3(b) represents the cleavage of C—C bond in the acyl group by γ–hydrogen transfer
CC

[90] and formation of terminal olefin with two carbons lesser (Cn-2) than the original fatty acid
chain [24]. From this view, no researcher has reported on the significant formation of (Cn-2)
A

during the deoxygenation process [24,90]. During hydrogenation, fatty acid yield might
gradually decrease and hydrocarbon yield may increase in the form of alkanes or α-olefins. In
this process the gaseous products are not expected and only stable saturated fatty acids were
formed. These results confirmed that the fatty acids are indeed the intermediate product of
TGs and no extra double bonds are present in the final product [62]. A possible
hydrogenation/hydrolysis reaction pathway of vegetable oils (TGs) is shown in Fig. 4 [74].

18
T
IP
Fig. 4. A possible hydrogenation reaction pathway of vegetable oil (TGs).

R
SC
Reasonable amount of unsaturated fatty acids in TGs model compounds are converted to
saturated fatty acids by hydrogenation. In this prospective, Smejkal et al. investigated

U
hydrogenation reaction of tristearate model compound to get hydrocarbon at a temperature of
N
250 and 450°C under 7-70 bars. They attained hydrocarbons with two different chain lengths,
which indicates that in hydrogenation reaction two different reaction pathways (DCO2 and
A
HDO) are involved. They also described that variation in selectivity from C17 and C18
M

products is attributed to the different temperatures and pressure condition, hence favored
DCO2 and HDO reaction [92].
ED

4.2 Decarboxylation (DCO2)


Decarboxylation (DCO2) refers to the removal of carboxyl group in H2 environment in
PT

which oxygen O2 is eliminated in the form of carbon dioxide CO2 from TGs carbon chain
[74]. In DCO2 reaction, two leading reaction pathways are involved: β-elimination and γ-
E

hydrogen transfer mechanisms. Through β-elimination, carboxylic acid and an unsaturated


CC

glycol difatty ester (UGDE) are produced from TGs. Subsequently, hydrogenation of UGDE
releases fatty acid and forms lesser number of Cn-1 hydrocarbons or normal paraffin [24].
A

However, γ‐hydrogen transfer mechanism initiates C-C bond cleavage within the acyl group
and produces a terminal olefin, which had two carbons lesser than the fatty acid chain [90].
The major products of DCO2 reaction are C15 and C17 hydrocarbons which correspond to n-
pentadecene and n-heptadecene respectively and leading to hydrogen generation during bond
scission [90,93]. From previous study, it can be determined that DCO2 reaction is responsible

19
to form straight chain hydrocarbons (alkane) with one carbon lesser [87] [90] or odd number
of carbon atoms than the corresponding TGs [54] as shown in Eqn. 2 [91].

CnH2nO2 → Cn-1 H2n + CO2 (2)

4.3 Decarbonylation (DCO)


Decarbonylation (DCO) reaction refers to the removal of carbonyl group in H2
environment to produce alkanes with one carbon lesser than the original fatty acid and CO as

T
a byproduct [88]. As mentioned earlier that, in most cases, fatty acids were observed as an

IP
intermediate product via β-elimination or hydrolysis of TGs in DO process. During DCO

R
reaction, fatty acid intermediates release formic acid rather than CO2 as a primary product
[24]. Afterwards, the formic acid decomposition can take place via two parallel pathways:

SC
dehydration and hydrogenation. CO and H2O are released via dehydration while CO2 and H2
are released via dehydrogenation to produce alkenes/olefin [93]. Throughout DCO,

U
dehydration route is the major pathway of formic acid decomposition under H2 atmosphere
N
[90]. Santillan et al. investigated DCO reaction pathways using tristearin as a model
compound of TGs to get n-pentadecane, CO and H2O under H2 pressure, as illustrated in Eqn.
A
3. While, complete reaction mechanism of both DCO2 and DCO of tristearin into straight
M

chain hydrocarbons is shown in Fig. 5 [90].


ED

CnH2nO2+ H2 → Cn-1 H2n + CO+ H2O (3)


E PT
CC
A

Fig. 5. DCO2 and DCO of tristearin. Redrawn based on [90].

20
4.4 Hydrodeoxygenation (HDO)
HDO is an oxygen exclusion process which depends on the use of pure hydrogen H2
under certain temperature and high pressure condition for direct scission of C—O bond
[3,19]. During HDO process, unsaturated bonds of fatty acid present in TGs based oils are
converted to saturated bonds and thus oxygen O2 is removed in the form of water [74,93] as
mentioned in Eqn. 4.

CnH2nO2+ 3H2 → CnH2n+2+ 2H2O (4)

T
IP
Consequently, HDO reaction leads to n-alkane with the same number of carbon atoms

R
corresponding to fatty acid bond in the original TGs. For a clear understanding of the HDO
reaction pathway, a reaction mechanism of tristearin as model compound of TGs is presented

SC
in Fig. 6 to obtain straight chain hydrocarbons [94].

U
N
A
M
ED

Fig. 6. HDO of tristearin as model compound of TGs. Redrawn based on [90].


PT

While through HDO reaction, two possible side reactions could occur, that is a water gas shift
and methanation reactions which lead to increase in hydrogen consumption as shown in Fig. 7
E

[12].
CC
A

21
T
IP
Fig. 7. Methanation and water gas shift reactions

R
SC
Jeczmionek et al. reported that CO2 methanation reaction can be recognized during HDO
reaction as the joining of two reactions: water gas shift reaction and CO methanation [95].
The main gaseous products from HDO of TGs based oil are CO2, CO, H2 and C3H8 [28,96]. It

U
is generally observed that the decarbonylation/decarboxylation pathway consumes lesser H2
N
and produced C15 and C17 hydrocarbons. However, HDO has higher hydrogen consumption
A
and gives C16 and C18 hydrocarbons products [12]. During DCO2 reaction, more CO2 is
formed as byproduct and when it reacts with four molecules of H2, it will initiate methanation
M

and produces CH4 with two water molecules [97]. Likewise, Sudhakara et al. explained that
when CO and three H2 molecules react with each other, they form CH4 and two H2O
ED

molecules. Additionally, this reaction leads to formation of lighter products. Besides that,
when CO and H2O undergo through water gas shift reaction, they form CO2 and H2 as shown
PT

in Fig. 7 [28].
E

4.5 Cracking and isomerization reactions


Cracking and isomerization reactions are used to improve the HDO process. According to
CC

the literature, HDO generally produces paraffinic hydrocarbons with a high cetane number,
but these paraffinic oils may have compatible issues with jet fuels due to their poor cold flow
A

properties. Therefore, it is essential to improve their physiochemical properties by


hydrocracking and isomerization reactions. After hydrocracking, normal paraffins are
converted to C9 to C15 hydrocarbon chains which are more suitable for jet fuel application
[39]. After hydrogen isomerization, the alkanes retain the high cetane number and also
improve the cold flow properties [39,89].

22
In addition, cracking and isomerization reactions can be performed either concurrently or
sequentially for the formation of hydrocarbons [39]. Hydrocracking of alkanes is a well
known industrial process which was developed by petroleum refineries for the production of
gaseous hydrocarbons, gasoline and kerosene from heavy fractions [23]. Cracking reaction
produces straight chain n-alkanes (C15 above) which need further cracking to be a compatible
jet fuel. On the other hand, the isomerization process converts straight-chain hydrocarbons (n-
paraffins) into branched (iso-paraffin) structures to reduce the freezing point in order to meet
the jet fuel specification and standard [39]. Hydro-isomerization and hydrocracking processes

T
are followed by a fractionation process to separate the mixtures to obtain paraffinic kerosene

IP
or jet fuel range hydrocarbons [98]. Wei et al. also suggested that hydrocracking is an

R
essential processe before isomerization reaction to obtain lighter alkanes (C5 to C15) [39].

SC
The hydro-isomerization process normally proceeds over bifunctional supported catalyst
[23,79,97,98]. In a typical isomerization reaction, normal paraffins are dehydrogenated on the

U
metal sites of the catalyst and react on the acid sites to produce a corresponding n-alkane with
N
the same number of carbon chain [23,39]. The interaction of the latter with Brønsted acid sites
generates carbenium ion, which undergo subsequent isomerization into iso-carbenium ion and
A
consequently desorbs as iso-alkane [23]. The carbenium ions are rearranged into mono-
M

branched, di-branched and tri-branched carbenium ions on the acid sites. The branched
carbenium ions are deprotonated and hydrogenated to produce the corresponding paraffins
ED

[39]. Mechanism of n-paraffins/n-alkanes hydro-isomerization over bifunctional catalysts


(e.g. zeolites loaded with metals) using n-decane as the model compound is shown in Scheme
PT

1 [23]. Catalytic performance of some typical zeolite catalysts for isomerization process with
different feedstocks and varied reaction conditions for jet fuel production is shown in Table 6.
E
CC
A

Scheme 1. Mechanism of n-paraffins/n-alkanes hydro-isomerization over bifunctional


catalysts (e.g. zeolites loaded with metals) using n-decane as a model compound [23].

23
It is commonly observed by many researchers that isomerization is more dependent on
moderate acidity of the catalyst and desired hydrocarbon selectivity. In regard to moderate
acidity for jet fuel production, Wang et al. reported that Ni metal with SAPO-11 as catalyst
support had the best moderate acidity and produced the maximum hydrocarbon selectivity
(i.e. 100% selective to n-C7-C14 alkanes). In addition, they also noticed that an increase in Ni
loading from 0 to 8 wt.% reduced the amount of strong acid sites and increased Ni metal
active sites which resulted isomerized alkanes yield increase up to 85% [23]. While, zeolite
attached with metal sites which is necessary to favor hydro-isomerization and avoids from

T
undesired side reactions such as hydrogenolysis and hydrocracking [23]. Some researchers

IP
have also confirmed that mordenite supported metal or high-silica zeolite catalysts favored

R
hydro-isomerization reaction between 250 and 400oC with 20 bars of H2 pressure [23,100-
102].

SC
5. Deoxygenation reaction mechanism of model compounds

U
Several investigations were made based on model compound studies in order to develop
N
further insights on the reaction kinetics of DO pathway with respect to the model compounds
[9]. In this regard, several researchers have investigated DO pathway of TGs based model
A
compounds and liquid yield. Therefore, model compound study is useful in selection and
M

designing of appropriate catalysts for the production of jet fuel hydrocarbon [42]. Several
model compounds such as methyl laurate, tristearin, tripalmitin, palm kernel oil, palm olein,
ED

karanja oil and stearic acid to get hydrocarbons through DO reaction were examined [42].
Typically, conversion of these model compounds to hydrocarbons is accompanied by catalytic
PT

DO reaction via major reaction pathways corresponding on DO and DCO2/DCO to estimate


the hydrocarbon yield [42,103]. Several reaction mechanisms of methyl laurate, tristearin
and tripalmitin, tristearin and stearic acid, triolein, palm kernel oil and karanja oil conversion
E

processes are concisely discussed as below.


CC

5.1 Methyl laurate


A

The deoxygenation (DO) reaction of methyl laurate as a model compound was studied by
Chen et al. in a continuous flow stainless steel fixed bed reactor to obtain the alkane
hydrocarbons. The reaction was performed at temperature of 300°C, 2.0 MPa pressure and
WHSV (weight-hourly space velocity) of 5.2 h-1 with several silica-supported metal (Ni, Co,
Fe, Mo and W) phosphides. The proposed reaction pathway for the deoxygenation of methyl
laurate is shown in Scheme 2 which includes hydrogenolysis, hydrolysis, hydrogenation,

24
decarbonylation, dehydrogenation, dehydration, isomerization as well as cracking process.
Initially, methyl laurate can be converted to: lauric acid by hydrogenolysis or hydrolysis;
dodecanal via hydrogenolysis and undecene via decarbonylation [91].

For hydrogenolysis or hydrolysis pathway, lauric acid is the primary intermediates and
can be further converted to dodecanal (C11H23CHO) via hydrogenation or undecene (C11H22)
via decarbonylation. After that, undecene is transformed to undecane (C11H24) through
hydrogenation process. For the HDO pathway, dodecanal can be hydrogenated to produce

T
dodecanol. While dodecanol (C12H26O) is formerly dehydrated to form dodecene (C12H24) or

IP
undesirably dehydrogenated to form dodecanal. Furthermore, dodecene is either hydrogenated

R
to form dodecane (C12H26) or converted into carbocation on Bronsted acid site. This process is
followed by isomerization and hydrogenation to produce iso-dodecane (C12H26), iso-undecane

SC
(iso-C11H24) and lighter alkanes. While in the case of DCO pathway, dodecanal is converted
to undecene by decarbonylation/dehydrogenation. After that, undecene is feasibly converted

U
to undecane, isoundecane and lighter alkanes over cracking reaction.
N
A
M
ED
E PT
CC
A

Scheme 2. Proposed reaction pathway for deoxygenation of methyl laurate. Redrawn based
on [104].

The authors observed that the oxygen-containing intermediates such as lauric acid,
dodecanal, dodecanol, dodecyl laurate and methanol during DO reaction were formed in the
25
liquid product and subsequently undergone DCO and HDO reactions. The HDO reaction
formed the final product of C12 hydrocarbons over Mo and W based P/SiO2 catalysts.
However, DCO produced C11 hydrocarbons over P/SiO2 supported metal catalysts (Ni, Co,
Fe). Thus, the authors determined that DO reaction mechanism is mainly dependent on the
electron property of metal site and Brønsted acidity. The DCO reaction was promoted by
increasing the electron density of metal Ni and Co phosphide sites and giving rise to the larger
C11/C12 ratio and lesser amount of alkenes respectively. However, Fe, Mo and W phosphides
produced alkenes and iso-hydrocarbons apart from the n-alkanes, which is attributed to their

T
low hydrogenation ability and high Brønsted acidity. Furthermore, Ni and Co phosphide

IP
demonstrated the highest conversion and selectivity for C11/C12 hydrocarbons [104].

R
5.2 Tristearin and tripalmitin

SC
Yenumala et al. developed a mechanistic kinetic model for HDO of tristearin and
tripalmitin as model compounds. The HDO experiments were carried out using γ-Al2O3

U
supported nickel catalyst at a temperature range of 280-390oC with 30 bar H2 pressures as
N
shown in Scheme 3.
A
M
ED
E PT
CC

Scheme 3. Reaction mechanism of HDO of TGs. Redrawn based on [103].


A

Under these experimental conditions, tristearin and tripalmitin TGs were converted to
FAs (palmitic and stearic acid) and appeared as major oxygenated intermediate products with
propane as a coproduct. These intermediates were further reduced and converted to
octadecanal and hexadecanal. Subsequently, these octadecanal and hexadecanal were further
converted to alkanes by two parallel reaction pathways which are DCO and HDO
26
respectively. From the DCO reaction, octadecanal and hexadecanal were converted to
pentadecane (C15H32) and heptadecane (C17H36) and released one mole of CO. In this manner,
DCO appeared as the leading pathway as compared to HDO, in which the former produces
alkanes with one carbon lesser than the corresponding fatty acids. However, during HDO
reaction, octadecanal and hexadecanal were further reduced to octadecanol and hexadecanol,
and produces octadecene and hexadecene olefins through dehydration reaction. Afterward, the
olefins were converted to alkanes with the same number of carbon atoms as present in the
fatty acids via hydrogenation reaction. However, the formation of olefins was not observed

T
during conversion of fatty acids to alkanes. The transformation of the aldehyde to alkane via

IP
decarbonylation reaction typically occurs on active metallic sites of the catalyst. On the other

R
hand, HDO occurs over acidic sites to convert alcohol to alkene via dehydration and metallic
sites for reduction/hydrogenation reaction. The study on its product distribution clearly

SC
showed the decarbonylation of fatty aldehyde was the dominating route as compared to the
reduction of aldehyde to alcohol followed by its dehydration and hydrogenation reaction
[103].
U
N
5.3 Tristearin and stearic acid
Santillan et al. also studied the catalytic deoxygenation reaction pathway of tristearin
A
(95%) and stearic acid (97%) as model compounds of pure TGs and fatty acids to produce
M

high selectivity of C10-C17 hydrocarbons. The HDO experiments were performed over 20%
Ni/Al2O3 catalyst at a temperature of 300 ± 2o C with a pressure of 135 psi for 1.5 h in a
ED

stainless-steel autoclave reactor using semi-batch mode. Possible DO reaction pathway of


stearic acid is shown in Scheme 4. They indicated that the conversion of tristearin produced
three molecules of stearic acid as intermediate products over β-elimination process [94].
E PT
CC
A

27
T
R IP
SC
Scheme 4. Possible DO reaction pathway of stearic acid. Redrawn based on [90,93].

U
Based on Scheme 3, it can be seen that the stearic acid can either be feed or the reaction
N
intermediate which undergoes DCO2 and DCO simultaneously to produce n-heptadecane and
n-pentadecane as the main products with CO2 and H2 as the byproducts. Subsequently, some
A
of the stearic acid was hydrogenated using the hydrogen source to produce minor amount of
M

alkane (C18H38) and H2 as identified in reaction 4. Furthermore, C17H36 and C17H34 are
isomerized and/or hydrogenated/dehydrogenated to produce C17 cyclic and aromatic
ED

molecules as seen in reactions 5-9. Additionally, cracking of stearic acid produces shorter
fatty acids (C10-C17 acid) and C13-C16 hydrocarbons can be seen in reaction 10 [93]. While,
stearic acid is converted into symmetrical ketone and unsaturated hydrocarbons can cause the
PT

formation of heavier products, as shown in reaction 11, 12, and 13. Both C15 and C16 alkanes
possibly will be cracked into the light alkanes over an acidic catalyst [90,93,94].
E

5.4 Triolein
CC

Possible DO reaction pathway of triolein over alumina supported Ni, Co and Mo catalysts
were shown in Scheme 5. In this study, Horacek et al. observed that the triolein was initially
A

hydrogenated by hydrogen source and then completely converted to single chain products in
the form of di- and mono-glycerides with minor amount of alkane like stearic acid
C17H35COOH as intermediates.

28
T
R IP
SC
Scheme 5. Possible DO reaction pathway of triolein. Redrawn based on [105].

U
N
Besides that, the intermediate products were converted to C17 and C18 via two reaction
pathways: (i) HDO producing even-numbered hydrocarbon chains and [106] [106] hydro-
A
decarboxylation yielding odd-numbered hydrocarbon chains. In the HDO pathway,
M

hydrogenation reaction occurs which produces C18H38 along with propane. For the
decarboxylation pathway, it involves decarboxylation and decarbonylation reactions which
ED

remove carboxyl groups to respectively form CO2 and CO as byproducts. The difference in
product distribution in the different model compounds indicated that the catalytic activity
depends on the degree of unsaturation in the feedstock. Meanwhile, in HDO reaction, a
PT

mixture of stearic acid, C17H36 and C18H38 were observed as the major products. The authors
determined that the alumina-supported NiMo bimetallic promoter-active metal combination
E

was found as most active catalyst type in the temperature range of 250 and 270 oC. However,
CC

high selectivity for HDO reaction pathway (total hydrogenation) in the same temperature
range was observed with CoMo type catalysts [107].
A

5.5 Palm kernel oil


Another group of researchers studied the refined palm kernel oil as a TGs model
compound for the production of jet fuel-like hydrocarbons. They used Ni-MoS2/γ-Al2O3
catalysts for deoxygenation of model compound at temperature range of 270-330oC under

29
varied hydrogen pressure conditions from 30-50 bars with liquid hourly space velocity
(LHSV) of 1-5 h-1 as shown in Scheme 6.

T
R IP
SC
U
N
A

Scheme 6. Proposed catalytic conversion of palm kernel oil into jet fuel like hydrocarbon via
M

deoxygenation. Redrawn based on [14].


ED

The palm kernel oil was dissociated through the hydrogenation of C—C bonds in
unsaturated triglycerides, followed by C—O bond cleavage via hydrogenolysis of saturated
PT

triglycerides to produce propane gas, C12, C14 and C18 fatty acid molecules as major
oxygenated intermediates. The formation of fatty acids is relatively fast when compared with
E

deoxygenation reactions (HDO, DCO, and DCO2). The deoxygenation of these fatty acids
intermediates to jet fuel was proposed to occur via three main reaction pathways: HDO, DCO,
CC

and DCO2.
Typically, HDO is the primary pathway and responsible for cleavage of C—O bond to
A

produce alkanes with the same number of carbon atoms such as dodecane (C12H26),
tetradecane (C14H30) and octadecane (C18H38) as well as water as a byproduct. In the case of
DCO reaction, minimum H2 gas was required to break the C—C bond and it produces alkane
of one carbon lesser such as undecane (C11H24), tridecane (C13H28), heptadecane (C17H36) with
CO as the by-product. However, in the case of DCO2, it does not require much amount of H2
gas to produce n-alkanes and CO2 gas. In addition, the side reactions such as cracking and
30
methanation also occur at 330 oC. Therefore, it is assumed that the formation of CH4, C2H6,
and C3H8 species in the gas product might be associated with cracking reaction in the liquid
phase during deoxygenation of palm oil. These effects revealed that HDO pathway of palm
kernel oil essentially produces the major hydrocarbon products of C12 fatty acid, indicating it
as a promising option for jet fuel production [14].

5.6 Karanja oil


Sudhakara et al. also investigated the DO reaction pathway for karanja oil as a model

T
compound for the production of C10 to C14 alkanes over Ni metal catalyst supported over γ-

IP
Al2O3, HZSM-5 and SiO2 at 35 bar pressure with a temperature range of 340-400°C is shown

R
in Scheme 7.

SC
U
N
A
M
ED

Scheme 7. Possible reaction pathway for DO of Karanja oil. Redrawn based on [28].
PT

In this process, unsaturated TGs were initially saturated under high H2 pressure through
E

hydrogenation. The saturated TGs were then converted into fatty acids over intermediate
CC

diglycerides and monoglycerides by liberating equivalent amount of propane. On the other


hand, during HDO reaction, diglycerides were not detected due to their rapid conversion.
During HDO of karanja oil, stearic acid, palmitic acid, octadecanol, monostearate and mono-
A

palmitate were observed as oxygenated intermediates. Furthermore, these intermediate


products were subsequently converted by DCO and HDO routes and form alkanes with the
number of carbon atoms either (i) one less than or [106] equal to corresponding TGs.

31
In the first route, the intermediates undergo through dehydrogenation reaction followed
by DCO over the catalyst metallic center and formed alkane with one carbon atom lesser than
corresponding alcohol intermediate. While, through second route, alcohol intermediates
undergo dehydration followed by hydrogenation or HDO and formed alkane with same
number of carbon atoms. These alkanes ranging from (C10 to C22) such as n-decane, n-
undecane, n-dodecane, n-tetradecane, PD, n-hexadecane (HXD), n-heptadecane (HPD), n-
octadecane (OD), n-nonadecane, n-icosane, n-heneicosane, and n-docosane with HPD. The
heptadecane appeared as the major product, leading to C10–C14 alkanes formation during

T
HDO of karanja oil [28].

R IP
It can be summarized that HDO and DCO2/DCO are the primary reaction pathways for
DO of various model compounds of vegetable oils to produce hydrocarbon fuel. However,

SC
these reaction pathways are also accompanied by other reactions such as hydrogenation/
dehydrogenation, hydrogenolysis, hydrocracking and isomerization for production of

U
hydrocarbons. Undesirable reactions like cracking, coking, and cyclization may also occur
N
during DO thereby reducing the hydrocarbon selectivity and leading to the higher aromatics
formation.
A
M

6. Deoxygenation catalyst
ED

6.1 Properties of catalysts


One of the typical features of the catalyst is to crack the reactant and accelerate the
PT

reaction rate (i.e. ion exchange) and remains unchanged at the end of the process. In this
respect, catalysts are widely used in process industries and in research works to optimize the
product distribution and to increase the product selectivity [52]. Catalyst pore structure, pore
E

volume, surface area and acidity are some of the important properties which have significant
CC

effects on cracking property and product selectivity [28,92,108]. In this regard, zeolites and
alumina-based catalysts were extensively used in the upgrading of vegetable oil for biofuel
A

production. Zeolites are crystalline aluminosilicate sieves structure with open pores and ion
exchange capabilities. In addition, zeolites are composed of different SiO2/Al2O3 ratios which
depend on their acidic properties and shape selectivity. Typically, SiO2/Al2O3 ratio determines
the zeolite reactivity on TGs based oils which has great impact on the final product [109].
Sharuddin et al. described that alumina is an amorphous acid catalyst that encloses Brønsted
acid sites with ionizable hydrogen atoms and Lewis acid sites with an accepting electron site.

32
Due to these chemical properties, alumina-based catalysts are likely used in the upgrading of
oil for fuel application [52]. Some typical properties of zeolite and alumina-based catalysts for
upgrading the vegetable oil are summarized in Table 7.

One can observe that the zeolite catalysts show higher surface area as compared to Al 2O3,
indicating the former to have higher catalytic activity. Several metal elements such as Ni and
Mo were commonly used to modify activity of zeolite while Co, Ni, Pd and Pt were widely
used to modify the activity of Al2O3. According to Table 7, it can be seen that the modified

T
zeolite had a higher surface area in the range of 85.6 m2/g to 711.5 m2/g than modified Al2O3

IP
with the surface area range of 148 m2/g to 222 m2/g. Sudhakara et al. described that the

R
modified Ni-ZSM revealed higher performance due to their higher surface area and stronger
acidity than modified Ni-Al [28]. Galadima et al. also reported that modified Al2O3 with

SC
metals has poor performance due to the low surface area and stability issues [23]. Typically,
these stability issues are linked with sintering and agglomeration property of metal particles

U
which decrease the active surface area of catalysts [28]. In addition, high surface area and
N
strong catalyst acidity lead to high conversion of light alkanes (C5 to C15) hence that is
appropriate for jet fuel production [23,28].
A
M

The activity of zeolite and Al2O3 catalysts also depends on the pore size and pore
volume, which have important effects on product selectivity [53]. The pore size and pore
ED

volume of zeolite-based catalyst was found to be between 3-90 A° and 0.10-0.73 cm3/g, while
pore size and pore volume of Al2O3 based catalysts are in the range of 148-172 A° and 0.21-
0.63 cm3/g, respectively as mentioned in Table 7. Pure zeolite ZSM-5 is superior than the
PT

metal modified zeolite (Ni-ZSM-5) due to sintering and agglomeration of nickel particles
during the reduction of the latter [28]. The pore volume and pore size inside the ring structure
E

of different zeolite catalysts vary significantly and are limited due to cation present in the
CC

system. The variation in zeolitic pore structure considerably occurs by changing the
exchangeable cations location typically located near the pore openings [110]. The pore size
A

and pore volume of pure Al2O3 are also greater than the metal modified Al2O3 due to the
addition of metals which may lead to partial blockage in catalytic channels. In addition,
excessive amount of metal particles with modified catalyst might cause blockage in Al2O3
pores which will decrease the catalyst pore volume [28]. Generally, large pore size of
modified catalysts may enable facile transportation of vegetable oil molecules towards the
catalyst active sites. Zhao et al. explained that if large molecules are not converted to smaller

33
ones, then the chances of obstruction might rise within the catalyst pores and cause a decrease
in yield [33]. Typically, vegetable oil with large molecules are quite hard to penetrate deeply
into the catalytic pores (as small as 1-2 nm) thus, limiting the cracking efficiency of the
reaction sites on the catalyst surface [33]. According to Liu et al., they suggested that 2 nm
pore sizes play an important role in HDO due to it creates enough space for the optimum
feedstock and product diffusion [108].

Catalyst acidity is possibly the most important property of zeolites which has significant

T
influence on catalyst activity and product selectivity on TGs based oils. This is due to their

IP
exclusive composite structure and shape selectivity [113]. The strength of catalyst acidity is

R
more attributed to the high silica aluminosilicate zeolites framework. In addition, strong
interaction between Al3+ ions and dissociated silicate species were attributed to increase the

SC
framework stabilization and strengthen their acidity. Acid sites can be classified as weak,
medium and strong acid sites [12]. Weak acidity property is manifested through adsorption of

U
NH3 on Lewis acid sites [12]. The low acidity of weak acid sites of catalyst will potentially
N
encourage the coke formation on the catalyst surface [33] which will lead to lower catalytic
activity [114]. However, the medium and strong acid sites will positively increase the acidic
A
strength by adsorption of NH3 on the Brønsted acid sites [12]. The Brønsted acid sites will
M

then feasibly protonate or transfer a proton (a hydrogen atom) to another ion or molecule of
intermediates [19]. Generally, the acidity of Al2O3 based catalysts are within the range of
ED

0.14-0.21 mmol/g which is a moderate acidity range. However, aluminum Al2O3 with metal
loading works as a Lewis acid site because it increases the electron accepting capacity of
PT

neighboring metal sites [33]. On the other hand, zeolites have both Lewis and Brønsted acid
sites which are useful for the conversion of vegetable oils to lighter and jet fuel range
hydrocarbons [108]. Muraoka et al. reported that silica-rich protonic zeolites showed strong
E

Brønsted acidity relating to alumina-rich protonic zeolites [115]. In addition, the acid sites
CC

density of zeolite-based catalysts were 4.55 mmol/g higher as compared to acid sites in other
catalysts as shown in Table 7. Therefore, high acidity of catalyst may improve the cracking
A

processes of olefin intermediates to enhance the jet fuel production [39].

6.2 Types of catalysts


There are several types of catalyst which can be categorized according to their chemical
nature and type of reactions they catalyze. Generally, three types of catalysts that are widely
used in the DO reaction of TGs and model compounds are noble metal, sulfided and non-

34
sulfided catalysts. There are some noble metal catalysts: Pd/γ-Al2O3, Pt/γ-Al2O3, Ru/γ-Al2O3
that were extensively used for oxygen removal through hydrotreatment of TGs. Beside this,
sulfided and non-sulfided catalysts: NiMo/Al2O3, CoMo/Al2O3, and NiW/Al2O3 and some
zeolites have received considerable attention for high jet fuel production [12]. In addition, the
performance and product selectivity through noble metals, sulfided and non-sulfided catalysts,
supports and promoters were exclusively discussed for DO of several feedstocks and their
model compounds in the subsequent section.
6.2.1 Noble metals

T
Noble metals are the versatile class of catalysts that is commercially applied in different

IP
industrial processes. The most important industrial process includes: HDO, hydrogenation,

R
DCO2, DCO, hydrocracking, isomerization and dehydrogenation for the production of
hydrocarbons as renewable fuel [21]. For instance, noble metals such as Pt, Pd, Ir, Rh, Ru and

SC
Ag were widely used for the removal of oxygen from TGs based oils [116-118]. Some of
them such as Pd, Pt and Ru were extensively used in DO of fatty acids, vegetable oil, TGs and
their model compounds to get liquid hydrocarbons [13,54].
U
N
Li et al. investigated Ru catalysts and observed excellent catalytic activity in DO
A
reactions [89]. The composition of Ru with C support as Ru/C showed good catalytic activity
M

in DO, with 90 % yield in the range of C5 to C16 hydrocarbon [23]. However, Pd and Pt
catalysts are effective for the DCO2 and DCO process and formed hydrocarbons which are
ED

one carbon lesser than the original fatty acids of TGs based oil. Noble metals Pd and Pt are
used in order to produce deoxygenated product with C17 hydrocarbon selectivity [21,54,89].
PT

Noble metals are typically used with supported catalysts such as zeolites or oxides [23].
Noble metal with suitable support might exhibit maximum catalytic performances in the
production of biofuels and jet fuel due to the synergistic effect of different components and
E

favoring the catalytic cracking activity [89]. Apart from that, isomerization reaction is
CC

typically carried out over Pt catalysts. Pt metal with suitable supports such as SAPO-11 and
Al2O3, promotes the isomerization reaction and converts saturated alkanes to branched
A

alkanes in order to improve the fuel cold flow properties and to enhance the jet fuel
production [13].

Different noble metal catalyst were also investigated for DO of TGs based oil for high
conversion of hydrocarbons and C8-C14 product selectivity. Srifa et al. studied the effect of
various catalysts (Co, Ni, Pd, and Pt) combined with γ-Al2O3 support on palm oil in DO

35
process. They observed three main reactions and good conversion tendency with product
selectivity from C8-C14 of jet fuel range hydrocarbons. They described that the metallic sites
of the Co catalyst were responsible for all three major reaction: HDO, DCO2 and DCO [112].
However, metallic sites of Ni, Pd, and Pt were reported to strongly promote DCO2 and DCO
reactions in which the Ni and Co catalysts directly cracked palm oil and formed light
hydrocarbons C8-C14 as the main product. The activities of all catalysts for C8-C14
hydrocarbon yields were reported to be in the following order: 10CoAl= (5.1%) > 5CoAl=
(3.5%) > 10NiAl= (1.6%) > 5NiAl= (0.8%) > 5PtAl= (0.7%) > 5PdAl= (0.6%). These results

T
proved that all catalysts showed a higher conversion of hydrocarbons product ~ 99-100%

IP
[112]. From these insights, it can be determined that Pt as noble metal with Al2O3/SAPO-11

R
showed good activity and appeared as suitable choice in terms of high desired product
selectivity.

SC
Rabaev et al. studied Pt with Al2O3/SAPO-11 at LHSV of 1.0 h-1, 370-385°C and 30 bar

U
pressure for hydrocracking and HDO activity using soybean oil. They observed that several
N
hydrocarbon liquid products were obtained such as aromatic, naphthenic, iso and normal
paraffins compounds in organic liquid with yields of 16.5 wt.%, 5 wt.%, 66.5 wt.% and 17.6
A
wt.% respectively. Hence, this showed that noble metal with appropriate support
M

Pt/Al2O3/SAPO-11 enhanced the jet fuel properties such as specific freezing point and
aromatics content by fractionation and isomerization. In addition, after fractionation, they
ED

obtained 42-48 wt.% of jet fuel with >12% aromatics compounds. The other lighter products
in gas phase were propane, ethane, CO and CO2 with yields of 3, 1, 2.5 and 1.5 wt.%,
PT

respectively.

Veriansyah et al. investigated three different noble metal catalyst: Pd, Pt and Ru
E

supported on γ-Al2O3 catalysts used for conversion of soybean oil by HDO at 400°C with
CC

92 bars H2 pressure for 2 h. Pd and Pt catalysts were found to be more superior than Ru
catalysts for oxygen elimination and hydrocarbon yield. The high oxygen removal was found
A

with Pd/γ-Al2O3 (90.0%) as compared to Pt/γ-Al2O3 (56.4%) and Ru/γ-Al2O3 (15.5%).


Among these noble catalysts, Pd/γ-Al2O3 showed good catalytic activity with n-alkane
product selectivity from C8 to C20 hydrocarbons. The higher oxygen removal with Pd/γ-Al2O3
catalyst is due to its moderate acidity, making it more suitable for HDO of soybean oil while
Pt/γ-Al2O3 and Ru/γ-Al2O3 catalyst showed lower hydroprocessing activity due to the
deactivation by unsaturated deoxygenation and cracking species [34].

36
6.2.2 Sulfided/non-sulfided catalysts
Sulfided/non-sulfided catalyst have been widely used in HDO of plant oils for jet fuel
production due to their higher activity [100]. Sulfided catalysts are found to be highly active
and selective for conversion of TGs based feedstocks [74]. Many authors reported that
vegetable oil derived TGs can be converted into jet fuel with sulfided Ni-Mo/γ-Al2O3 and Co-
Mo/γ-Al2O3 as non-sulfided catalyst by HDO process [119]. However, there are some
drawbacks linked with sulfided catalysts which are responsible for the final product

T
contamination with sulfur content and corrosion problem because the sulfide catalysts cause

IP
sulfidation (oxides to sulfides) in the active sites of the catalyst surface [74].

R
Similarly, Rogelio et al. investigated the effect of three different types of bifunctional

SC
catalysts: non-sulfided Pt/H-Y, Pt/H-ZSM-5 and sulfided NiMo/γ-Al2O3 catalysts for
hydrocracking/HDO of rapeseed oil. Both non-sulfided zeolitic catalysts in HDO of rapeseed

U
oil formed more iso-paraffin than n-paraffin in the range of C5 to C22. The acid sites of Pt/H-
N
ZSM-5 have 10-ring zeolites (having micropores with windows circumscribed by 10 oxygen
atoms) which are stronger than Pt/HY and produced higher cracking products. However, the
A
low acidity of Pt-HY catalyst led to higher production of C17-C18. Among both zeolitic
M

catalysts, Pt/H-Y catalyst was reported to have higher diesel yield while Pt/H-ZSM-5 seems
to be more suitable for production of green gasoline C5-C12. This is because both zeolites had
ED

strong cracking ability that favors the higher isoparaffin formation and providing low pour
point which is a required condition for jet fuel application. Many researchers have reported
PT

that sulfided catalyst NiMo/Al2O3 effectively promoted the hydrocracking of TGs and
carboxylic acids at lower pressure of 8 MPa at 350°C in 3 h [120].
E

Itthibenchapong et al. described that the bimetallic sulfided NiS2, MoS2 and Ni-MoS2
CC

with γ-Al2O3 supported catalyst had the ability to cleave the C—O bond of palm kernel oil by
HDO for the production of jet fuel-like hydrocarbons. The reaction conditions were set as
A

300oC, 30-50 bar of H2 pressure for 1-5 hrs. The acidity was modified by Al2O3 support (via
Brønsted sites) which showed a positive effect on the hydrogenation activity with MoS2 and
Ni-MoS2 sulfided catalysts, due to direct modification in their electronic properties. The
maximum conversion rate of TGs feedstock was about 89-95%, while Ni-MoS2/γ-Al2O3
provided the maximum conversion rate with 92 wt% and enhanced the HDO selectivity from
C10-C12 due to the doping of Ni species on MoS2 structure. Besides that, Ni-MoS2/alumina

37
support was observed to have a reasonably low tendency of isomerization and cracking
activities [14].

There is currently no doubt that noble metals, mainly catalyze DCO2 and DCO reactions
of TGs. However, there are some limitations with their usage, such as their rarity and high
prices. Noble metal catalysts with sulfided support are very sensitive to catalyst poisoning
and catalyst coking due to impurities: sulfur, heavy metals and oxygenated compounds in the
feedstock [121] which induce partial or complete deactivation of the catalyst [122]. Several

T
sulfides/non-sulfides catalyst with different supports are capable in converting TGs to HDO

IP
products. In addition, MoS2 and Co-MoS2 modified catalysts showed a positive effect on

R
hydrogenation activity due to direct alteration in their electronic properties, good stability and
higher hydrocarbon yield [108].

SC
6.3 Choice of support

U
Choice of catalyst support is of special importance because it is attached with active
N
material surface and enhances the performance of several active materials with different
proportions [123]. In addition, support plays an important role in dispersing the active phases
A
of the catalysts. Therefore, the choice of support is one of the key factors for determination of
M

HDO activity and hydrocarbon selectivity with different supported catalysts. In HDO process
many supports were studied such as metal oxides (TiO2, Al2O3, SiO2, CeO2, ZrO2) [123],
ED

zeolites (ZSM-5, HY, H-Beta) [23,33,124], mesoporous material (MCM-41, SAPO-11, SBA-
15, Al-SBA-15, Al-MCM-41) [23] and activated carbon (AC) [54].
PT

Several studies have shown the capability of metal oxide supports in HDO process such
as (TiO2, Al2O3, SiO2, CeO2, ZrO2) which had moderate acidity to catalyze C=O bond in
E

hydrogenation process [123] and C—O bond in hydrogenolysis without rapid coking [33].
CC

TiO2 supported metal catalysts have attracted much interest due to its high activity in several
reduction and oxidation reactions at low pressures and temperature conditions. Additionally,
A

TiO2 as support material was also used with heterogeneous catalyst due to its high surface
area, strong metal support interaction, chemical stability and acid-base property [123].

Nevertheless, γ-Al2O3 appeared as one of the most common and conventional supports
which had many useful properties like high stability and moderate acidity. Some researchers
reported that γ-Al2O3 support showed moderate acidity due to the Al-octahedral structure

38
which makes it a good candidate for HDO of carboxyl groups [14]. The acidity adjustment of
Al2O3 support (via Brønsted sites) has been stated to have a positive effect on the
hydrogenation activity of the MoS2 and Co-MoS2 catalysts because of the direct change in
their electronic properties [14]. Additionally, supports with moderate acidity are always
preferred due to high productivity and selectivity whereas supports of high acidity were
known to cause more cracking and reduce the product selectivity [23]. Tiwari et al.
investigated the effect of mesoporous silica-alumina (SiO2-Al2O3) and alumina oxide
(Al2O3) as a support with sulfided Ni-W and Ni-Mo catalysts effect on hydrocracking and

T
hydrotreating of waste soya oil mixture with refinery oil.

R IP
They also reported that the acidic mesoporous SiO2-Al2O3 with sulfided Ni-W catalyst
was found to be a suitable catalyst for hydrocracking and is more selective for kerosene range

SC
hydrocarbons from C8-C18 in the temperature range of 140-250°C. While the SiO2-Al2O3
catalyst with sulfided Ni-Mo was found to be less acidic and more selective for diesel range

U
hydrocarbons from C15-C18 in the temperature range 250-380°C. The author also observed
N
that Ni-W/SiO2-Al2O3 catalyst favored more on DCO2 and DCO reaction and formed
paraffinic hydrocarbon fuel. In addition, all mesoporous catalyst supports produced liquid
A
fuels with a high cetane number with acceptable density. However, Ni-Mo/Al2O3 catalyst is
M

more preferred in HDO pathway for removal of oxygen from TGs to produce a diesel range
hydrocarbon fuels. [70].
ED

Over the past decades, about 200 different zeolite structures have been discovered, but
PT

only a few structures have been used in industrial applications [124]. Several investigations
were published in the past few years which dealt with the application of zeolites like ZSM-5,
SAPO-11, Meso-Y, Meso-H-beta and Meso-HZSM-5 as catalyst supports for jet fuel
E

production [125]. Zeolites mainly consist of acid sites with high surface area and adsorption
CC

capacity [33]. Additionally, zeolites also have high catalytic activity, remarkable product
selectivity and strong resistance to deactivation [126]. Therefore, zeolites were mainly studied
A

as a support material for the conversion of TGs to jet fuel production [125]. Li et al.
designated the use of zeolites due to uniform pore size, highly ordered structure and molecular
sieving properties, making it as a suitable option to produce biofuels from TGs-based oils.
ZSM-5 appeared as a bifunctional catalyst consisting both cracking and hydrogenation
activities because it holds strong acid sites and good product selectivity in HDO pathway
[125]. In addition, zeolite supported catalysts likely favoring the formation of gasoline range

39
hydrocarbons like jet fuel and gaseous products [11]. Moreover, SAPO-11 is also one of the
best example of supports with moderate acidity which showed maximum selectivity (i.e.
100% selective to n-C7-C14 alkanes) [23].

Mesoporous support materials typically exhibit high surface area which enables high
dispersions and strong interactions with active metal species on their surface, thus
significantly improving the catalytic performance. The high surface area and regular channel
structures could also promote the conversion and selectivity of hydrocarbons as the structures

T
of mesoporous supports facilitate the diffusion of large feedstock molecules, thus leading to a

IP
higher conversion rate [89]. In this regard, Liu et al. investigated the single and two step HDO

R
process over Ni catalysts supported on MCM-41 mesoporous material and several zeolites for
the hydrocracking of castor oil. In the first step, HDO and hydrocracking of castor oil over

SC
Ni/SAPO-11, Pt/SAPO-11, Ni2P/SAPO-11 and Ni-Ag/SAPO-11 produced higher yields of
C17-C18 due to weak acid sites, while the second step was more focused on hydrocracking of

U
C17-C18 to C8-C15. From these findings, the authors found that the level of cracking is
N
dependent on the acid strength and the Ni/Ag molar ratios. They observed that Ni/Ag
produced cracking and strong isomerization activity which leads to the maximum C8-C15
A
alkanes selectivity and low lighter hydrocarbon yield of C5-C7. On the other hand, in the
M

single step processing, mesoporous catalyst Ni/MCM-41-USY was investigated for the
deoxygenation of castor oil and observed that low yield of bio-aviation fuel. In addition, the
ED

authors incorporated APTES with Ni/MCM-41-USY to increase the hydrocarbon yield. They
observed that Ni/MCM-41-APTES-USY showed strong acidity and produced 80.3% yield of
PT

alkanes C8-C15 with high isomerization selectivity. In addition, the mesoporous-microporous


system improved the mass transfer ability of TGs molecule in the pore channels of zeolite
which leads to improved catalytic efficiency. They concluded that both NiAg/SAPO-11 and
E

Ni/USY-APTES-MCM-41catalysts showed excellent bifunctional catalytic activity with a


CC

higher kerosene hydrocarbon yield under similar operating conditions [11].


A

In context of mesoporous zeolite development, Cheng et al. investigated three different


Ni supported Meso-Y, Meso-H-beta and Meso-HZSM-5 catalysts for the production of jet
fuel-like hydrocarbons. They observed that out of the three zeolite catalyst, Ni/Meso-Y
catalyst resulted in higher jet fuel range alkane yield of 55.32% and aromatic hydrocarbon
yield of 9.47%. The change in jet fuel yield was attributed to the Meso-Y zeolites acid density
and surface area which are noticeably higher than other zeolites. However, Ni/Meso-H-beta

40
and Ni/Meso-HZSM-5 catalysts produced lower jet fuel range alkanes yields (29.95% and
4.96%) and higher aromatic hydrocarbons yields (40.3% and 89.43%) respectively [111].
They concluded that higher productivity is more influenced by the mesoporous support
properties such as surface area and acid density.

A group of researchers investigated the effect of mesoporous zeolite as support for the
production of jet fuel from waste cooking oil at 400°C. The activity of nickel based catalyst
with three different types of zeolites (Meso-Y, SAPO-34, and HY) was investigated to

T
convert waste cooking oil into jet fuel range aromatic hydrocarbons. They observed that the

IP
jet range alkane selectivity (C8-C16) over Ni/Meso-Y, Ni/SAPO-34 and Ni/HY were 53%,

R
58% and 53%, respectively. Furthermore, among these three catalysts, Ni/HY showed
considerably higher aromatic hydrocarbon yield to 23% of jet-fuel hydrocarbons selectivity

SC
(C8-C16) and lower heating value whereas Ni/SAPO-34 showed the lowest aromatic
hydrocarbon of jet-fuel selectivity of 6.1% which leads to poor lubricating property. It was

U
also concluded that the Ni/Meso-Y catalyst had average aromatic hydrocarbon selectivity for
N
jet fuel hydrocarbons (13.4%) [124].
A
Activated carbon (AC) is also considered to be one of the promising and economically
M

feasible supports in hydroprocessing of vegetable oil with notable properties: large surface
area, weak polarity, textural properties and low thermal stability [54]. High surface area
ED

provides much more anchoring sites for the active component. However, the weak polarity of
surface decreases the contact between the active phase and support material which improves
PT

the dispersion of catalytic phase over metal/sulfide [127]. Furthermore, AC has an acidic
hydrophilic surface at low temperature condition, while at high temperature it tends to form a
basic and hydrophobic surface [63]. In addition, one good property of AC is its low catalyst
E

deactivation property at low temperature due to its neutral nature [33]. These distinct
CC

properties of AC support material, making it as suitable choice in HDO reaction [63].


A

The use of basic support materials has not gained much attention in jet fuel production.
Even though basic support offers higher resistance towards the sintering and deactivation,
basic support materials seems to be less effective than acidic support materials of TGs
transformation and HDO reaction. Hence, previous research directed a limited use of MgO-
supported catalysts for HDO reaction. It can be established from the discussion that acid-base
interaction between the basic support and the metal solution is responsible to promote

41
dispersion of active metal species on catalyst support. Additionally, the coke formation can be
inhibited at some levels by using the support with basic nature [63].

Based on the aforementioned discussion, it can be established that supported metal


catalysts with smaller particle size are effective in HDO of fatty acids for the production of
paraffinic hydrocarbons. Surface area, pore volume and acidity are the major factors that have
important effects on the catalyst behavior and on HDO of fatty acids for the production of
paraffinic hydrocarbons. Support materials such as mesoporous zeolite, silica and activated

T
carbon are more stable in acidic environment than conventional alumina. The variations in

IP
support pore size and porosity had an influence on different reaction pathways such as

R
isomerization. Additionally, supports with moderate acidity is always preferred to avoid the
reduction from product selectivity because excessive high acidity is not favorable for HDO

SC
because of over cracking to get lighter hydrocarbons. Zeolites as support material, favor the
formation of gasoline-range hydrocarbons like jet fuel and gaseous products [11]. However,

U
zeolites as a support material also had a higher degree of coke formation due to its acidic
N
nature [33]. The use of basic supported catalysts are limited in HDO reaction for the
production of jet fuel. The subsequent supports (i.e. alumina, zeolites, SBA-15 and MCM-41,
A
and silica-alumina) are more suitable for vegetable oil to jet fuel production and less favorable
M

for hydrotreating process due to high cracking activity and low acidic nature.
ED

6.4 Active metal or promoter


Active metals or promoters are the important features of catalyst properties. Promoters
PT

are largely defined as materials added throughout the preparation of catalysts that increases
the activity, selectivity or stabilize the catalytic agents [63]. Over the years, several types of
promoters and active metals received more attention in the petroleum industry for refining,
E

HDO and catalytic cracking. Co and Ni catalysts are known as the state-of-the-art catalyst
CC

promoters due to excellent catalytic performance and lower cost for large scale jet fuel
production [128]. Cobalt as a promoter may lead to higher catalyst selectivity (>90%) which
A

is related to oxygen free products [63]. Veriansyah et al. investigated the activity and
selectivity of Ni and Co with Mo during HDO reaction and observed that both promoters
improved the content of acid sites. CoMo catalyst was found to have higher hydrocracking
activity compared to NiMo and favored more cracking, isomerization and formation of iso-
alkane.

42
Various promoters and their precise roles in selective HDO reaction are still yet to be
rigorously determined due to their complicated nature resulting in numerous simultaneous
reactions occurring during the HDO process [42]. Several catalysts were systematically
studied to investigate the effect of promoters in HDO reaction [9]. Zhang et al. studied the
impact of Ni and Co promoters on the performance of MoS2 based catalyst for HDO of canola
oil. They observed that Ni promoted catalyst such as NiMoS2 showed higher selectivity for
HDO reaction over DCO2 and DCO whereas Co promoted MoS2 catalyst favored DCO2 and
DCO reactions. Additionally, NiMoS2 catalyst also improved the hydrogenation ability, since

T
it has provided sufficient sulfur vacancies. However, CoMoS2 catalyst had saturated edge

IP
sites in hydrogen atmosphere which enabled the hydrocracking (C-C hydrogenolysis). The

R
DCO2 reaction was enabled by Co-promoted MoS2 catalyst over adsorption of C atoms on the
S edge that was close to Co-promoter [48]. Thus, in general Ni and Co particles encouraged

SC
the synergistic effect and formed a range of n-C7 to n-C18 alkanes [23]. Ni promoted
bimetallic catalysts with different supports are highly favorable for hydrogenation reaction.

U
Shi et al. reported that the Ni based promoters with HZSM-5 catalysts have substantial impact
N
on HDO activity to transform vegetable oil (Methyl hexadecanoate as the model compound)
to achieve jet fuel range hydrocarbons. It was also observed that Ni merely promotes
A
hydrogenation reaction whereas Co promoted both hydrogenation and HDO reaction. From
M

these insights it can be recognized that the addition of active metal and promoters feasibly
increased the surface area of support material and in turn, enhanced its catalytic activity [128].
ED

An overview of reactant, catalysts, preparation, characterization and reaction conditions for


HDO of TGs based oils over different catalyst and supports is presented in Table 8.
E PT
CC
A

43
7. Effect of operating conditions on DO of TGs to jet fuel conversion
Operating conditions of DO process have significant effects on hydrocarbon selectivity
and yield. Some of the key operating parameters include: temperature, pressure, reaction time,
active metal, catalyst support, promoter and amount of catalyst. Effect of active metal and
catalyst support on selectivity and higher fuel yield were aforementioned in Section 6.

7.1 Type of reactor


Reactor type has an important effect on process control, reaction rate, residence time,

T
heat transfer rate, product selectivity and product yield [136,137]. Therefore, suitable choice

IP
of reactor is an essential feature to perform DO of TGs based feedstocks for jet fuel

R
production. In this regard, several types of reactor were investigated such as trickle bed
reactor [138], batch reactor and semi-batch reactors [136], autoclave with and without stirring

SC
batch reactor [139].

U
Trickle bed reactors is one of the most widely used type of three-phase reactors on an
N
industrial scale. In trickle-bed reactors, the gas and liquid flow concurrently descending over a
fixed bed of catalyst particles. Liquid trickles down, while the gas phase remains continuous.
A
In a trickle-bed reactor, several flow regimes can be distinguished which are dependent on gas
M

and liquid flow rates, fluid properties and packing characteristics [139] Bench scale and
microscale down-flow and up-flow trickle bed reactor were specifically used for process
ED

development and parameter optimization [139]. Several studies showed that both bench and
microscale trickle bed reactors deliver comparable performance. The up-flow mode, trickle
PT

bed reactor has non-ideal flow and dead zone formation inside the catalyst bed, causing
wetting of catalyst and alter the rate of reactions. Thus, down-flow trickle bed reactor was
more preferred over up-flow fixed bed reactor due to its ideal flow [82].
E
CC

Batch and semibatch autoclave reactors are mostly used in deoxygenation of TGs based
oil like waste cooking oil and non-edible oil for the production of liquid hydrocarbon biofuels
A

[136]. The batch autoclave reactor is similar to a pressure cooker, but varied in size, shape and
functionality. It is typically known as a high-pressure batch autoclave for laboratory scale,
having reactor volumes in the range of 50-100 mL. The autoclave batch reactor is basically a
closed system with no inflow or outflow of reactants or products [139]. Batch reactor enables
a high proportion of hydrocarbons production through extended reaction. In contrast, a
semibatch reactor allows reactant addition and product removal at the same time which

44
provides greater flexibility in control of reactant and product concentrations to optimize
product yield [136]. Romero et al performed the deoxygenation of waste cooking vegetable
oil and Jatropha curcas oil under nitrogen atmosphere in batch and semibatch experiments
using CaO and treated hydrotalcite (MG70) as catalysts at 400 oC. In batch conditions, a
single liquid fraction (with yields greater than 80 wt.%) was produced containing a high
proportion of hydrocarbons (83%). In semi-batch conditions two liquid fractions were
obtained: a light fraction and an intermediate fraction containing amounts of hydrocarbons
between 72-80% and 85-88% respectively [136].

T
IP
Some of the batch and semibatch reactors were also equipped with stirrers that run at

R
different speeds depending on the required setting. Low conversion is reported by many
researchers on without stirring and leaving the reactants in the reactor for an extended time,

SC
which is one of its disadvantages [52]. Many researchers have investigated stirring effect of
autoclave reactor under high hydrogen pressure or high temperature on the HDO of TGs [82].

U
Autoclave batch reactor without stirrer at high temperature produced high conversion of TGs
N
to hydrocarbon and high paraffinic fuels. From this aspect, Gusmao et al. used vegetable oil
for HDO in autoclave batch reactor with Ni/SiO2 and sulfide Ni-Mo/γ-Al2O3 catalysts at H2
A
pressure from 10-200 bars at 623-673 K temperature for 2 h. The results revealed that, the
M

maximum conversion of 100% molar yield of hydrocarbon was achieved [140]. Zhao et al.
used sunflower oil as feedstock for cracking in the presence of ZSM-5 without stirring at a
ED

liquid hourly space velocity (LHSV) of 3 h-1 and at three different reaction temperatures (450,
500 and 550oC) in fixed-bed reactor. The liquid yield from 450 °C temperature setting
contained 19.93% oxygenated hydrocarbon fuels and from 500oC setting, it contained pure
PT

hydrocarbons ranging from C18 to C43. However, hydrocarbon fuels produced at 550 °C
setting contained 77.31% hydrocarbons (C7-C12). The results indicated that the high reaction
E

temperature affected the distribution of hydrocarbon fuels in batch reactor [32].


CC

Several researchers have used the autoclave batch reactor with stirring for the conversion
A

of TGs. Guzman et al. used crude palm oil as feedstock for HDO in the presence of Ni-
Mo/Al2O3 with stirring in the pressure range between 40 and 90 bars and 533-613 K in the
autoclave batch reactor. They observed diesel range, high paraffinic fuel with excellent cetane
index. In addition, there was only slight deactivation of the catalyst [141]. Santillan et al. used
a 100 mL autoclave batch reactor with mechanical stirring feature for the deoxygenation of
tristearin over 20 wt.% Ni/C catalyst. The use of stirring conditions led to better results than

45
those previously obtained without using stirring in an autoclave batch reactor, as mixing
improves the heat and mass transfers within the reaction system. Furthermore, remarkable
results were attained over 20 wt.% Ni/C catalyst with stirring, with conversion and selectivity
of C10-C17 and C17 to be at 77% and 55% respectively [142].

7.2 Effect of temperature


Temperature is one of the key parameters in affecting the hydrogenation of TGs,
conversion of TGs based oil, product distribution and product selectivity. Besides that, the

T
reaction temperature also plays an important role in ensuring optimum catalyst activation and

IP
minimization of catalyst sintering [143]. Many researchers have described that the reaction

R
temperature of less than 200 oC showed significant effect of hydrogenation of double bond
and produced corresponding saturated TGs. Additionally, as the temperature increased from

SC
230 to 280 oC, the hydrogenated TGs degraded into several intermediates including:
monoglycerides, diglycerides and free fatty acids. On further increasing the reaction

U
temperature to about 300 oC hydrocarbon yield increased [144]. Reaction temperatures of 350
N
to 375oC were reported to be optimum for the C=O bond scission of fatty acids with minimum
degree of coke formation during HDO reaction [63]. On further increasing to 400oC and
A
above, some adverse effects such as increased coke formation were reported, [63] which was
M

attributed to complex oil composition and large molecules [89].


ED

Verma et al. performed HDO and hydrocracking reaction of jatropha oil in a temperature
range of 375 to 450oC using SAPO-11 supported NiMo and NiW sulfided catalysts. They
observed that lower reaction temperature (<375oC) was more favorable to obtain higher (85-
PT

96%) diesel yield from (C15-C18) selectivity. An increase in temperature to 425oC resulted an
improved the jet fuel range (C9-C14) hydrocarbon yield from 22-37.5 %. However, a further
E

increase in temperature (> 425oC) showed a decrease in jet fuel yield and an increase in the
CC

gas formation which were attributed to excess oil cracking [12]. Based on some earlier works,
Liu et al. used different temperatures (320oC and 400oC) for isomerization of palm oil over 7
A

wt.% Ni/SAPO-11 catalyst under 40 bar hydrogen pressure during HDO reaction [99]. They
observed that gaseous alkanes C1-C4, liquid alkane C5-C14 and C15-C18 were the main
products. The lower reaction temperature of 320oC showed C15-C18 hydrocarbons selectivity
and consequently suppressing heavy cracking of long chain alkanes [99,145]. With further
increasing the reaction temperature from 380oC to 400oC, isomerization was significantly
increased. As the reaction temperature further increased, the liquid product yield from C5-C14

46
was decreased, whereas gas formation was increased due to the significant cracking of long
chain alkanes.

Many studies were conducted to find out the effect of temperature on the selectivity of
deoxygenation pathways. Galadima et al. studied the effect of temperature on deoxygenation
pathways. They found that the DCO and DCO2 reactions were favored at higher temperatures.
[23]. A similar trend was also reported by Chen et al. [26] Cheng et al. investigated the effect
of temperature from 330 to 410°C on jet fuel quality with less than 20 vol.% aromatic

T
hydrocarbons and high number of alkanes. When the temperature was increased from 330 to

IP
390oC, jet fuel range aromatic hydrocarbon, alkanes and total hydrocarbons (C8-C23) contents

R
were significantly increased from 0% to 17.6%, 29.9% and 49.1%, respectively. This is due to
shifting of the reaction pathway from oligomerization to cracking reaction. On further

SC
increase in reaction temperature from 390 to 410oC, jet range aromatic hydrocarbon was
siginificantely increased from 17.6% to 28.7%, which obviously decreased the jet fuel quality
[146].
U
N
Similarly, Li et al. selected different temperatures to optimize the reaction temperature
A
from 370-430°C for the formation of higher aromatic hydrocarbon. They observed an increase
M

in aromatic content from 0.5-24.4% which is responsible for lower heating value and jet fuel
quality. Furthermore, they also observed low hydrocarbons yield of 15% at 370°C whereas,
ED

jet range alkanes and aromatic yields were significantly increased to 40.5% and 11.3%,
respectively when the reaction temperature was further increased to 400°C. These features
PT

indicate the change in reaction pathway from oligomerization to cracking reaction when
reaction temperature increases. On further increasing the reaction temperature to 410 oC, the
yield of jet fuel range alkanes was decreased to 31% and jet fuel range aromatic hydrocarbons
E

increased to 21.6% due to the severe cracking reaction of liquid hydrocarbons which formed
CC

gas products. They concluded that 400°C temperature is the optimum condition to obtain high
jet fuel range alkane yield to 40.5% with low jet fuel range aromatic hydrocarbon yield of
A

11.3% [124].

In the light of these findings, it can be summarized that different temperatures have great
influence on reaction pathway and on product selectivity [82]. It was also noticed that the
reaction temperature of less than 200°C had a significant effect on the saturation of double
bonds [144]. A moderate temperature range from 330 to 390 oC showed low aromatic

47
hydrocarbon yield with high number of alkanes which are desired qualities of jet fuel [146].
However, high temperature (>500 oC) increases the chances of coke formation tendency [63],
light hydrocarbons and high gas formation [82].

7.3 Effect of pressure


The effect of pressure in DO reaction is an important parameter which appreciably
impacts the reaction pathway. Pure H2 or combinations with other inert gases such as H2-Ar,
H2-N2, H2-He etc. were also used in DO reactions. Usually, the H2 reaction atmosphere was

T
noted to be more favoring for the reduction of aromatic hydrocarbon formation [147].

IP
However, different proportions of He, H2-Ar and N2 could lead to maximum aromatics

R
formation which in turn increase fuel density and reduce cold fuel flow properties. Likewise,
the HDO reaction of TGs (unsaturated feedstock) under inert atmosphere (N2, He and Ar) also

SC
led to undesired reactions like cracking and formation of heavier products while H2
atmosphere tends to minimize coke formation. Therefore, Chen et al. investigated the effect of

U
different H2 pressures (0.4, 0.8 and 1.2 MPa) on the conversion of fatty acid methyl esters
N
(FAMEs) to alkane. They observed that conversion of FAMEs to alkane was increased from
78.6% to 83.1% at 280°C and from 85.7% to 91.8% at 300°C when H2 pressure was increased
A
from 0.4 to 0.8 MPa. However, on further increase in H2 pressure to 1.2 MPa, conversion of
M

FAMEs was decreased to 80.2% at 280°C and 87.4% at 300°C. Rogelio et al. also
investigated the effect of H2 pressure (5-11 MPa) in HDO reaction for product distribution.
ED

They observed a high content of saturated carboxylic acids like stearic acid when 8 MPa
pressure was used. Nevertheless, between 8-10 MPa, green diesel range hydrocarbon fuel
PT

(C15-C18) with high yield of n-C17 was observed via DCO2 and DCO reactions. On further
increasing H2 pressure to 11 MPa, high catalytic cracking activity and improved lighter
hydrocarbons (C5 to C12) yield were observed [120].
E
CC

In the catalytic hydroprocessing, Kimura et al. investigated HDO of TGs for jet fuel
hydrocarbon selectivity under low hydrogen pressure from 4-8 bars. They observed that high
A

and low H2 pressure conditions influenced the two reaction pathways. Under low hydrogen
pressure, TGs was converted into hydrocarbons to odd carbon numbers through subsequent
DCO2. However, under high hydrogen pressure condition, it enhanced the reaction of fatty
acids with activated hydrogen and produced hydrocarbons with even carbon numbers. Low
hydrogen pressure condition resulted in the production of 71% jet fuel fraction (C8-C16)
hydrocarbon yield [10]. According to Anand et al., HDO of vegetable oil at higher H2

48
pressure (> 60 bar) and temperature of 420°C was able to improve the C—O bond scission
which led to a higher yield of naphtha and kerosene like jet fuel hydrocarbon whereas lower
H2 pressure (< 60 bar) showed more C—C bond cleavage and dehydrogenation instead.
Furthermore, by increasing the pressure to 90 bars, formation of naphtha and kerosene (2% to
10%) were gradually increased. Additionally, an increase in hydrogen partial pressure also
favored cracking and formation of lower range hydrocarbons [61]. From these findings, it can
be determined that the high H2 pressure is more suitable to minimize the formation of heavy
products in order to improve the lower range hydrocarbon selectivity in HDO process.

T
IP
7.4 Effect of residence time

R
Residence time is the ratio of mass feed rate of liquid over the catalyst mass in the
reaction system [52,82]. In addition, residence time is inversely proportional to the feed rate

SC
and can be expressed as space velocity (SV) as h-1 [25,82]. Itthibenchapong et al. described
that shorter residence time in solid-liquid reaction system is an indication of higher liquid

U
hourly space velocity (LHSV) between the reactant and the catalyst which could lead to
N
decrease in HDO product and increase in DCO2 and DCO product formation. Moreover, high
LHSV with high reaction temperature also promote CO gas product formation by favoring
A
DCO2 pathway [14]. Anand et al. studied the effect of residence time and temperature on
M

product selectivity by catalytic HDO of plant oil (TGs) to hydrocarbons in fix bed reactor.
They found that higher liquid yield C9-C14 of hydrocarbons was achieved (6% to 35%) by
ED

reducing the space velocity from 12 h-1 to 0.5 h−1 when the temperature was set below 420oC
[22]. Similar effects of longer residence time on increasing yield of the lower range of
PT

cracked products (< C15) were also reported using jatropha, soybean, sunflower and waste
cooking oil lipids over different sulfided hydrocracking catalysts.
E

Charusiri et al. investigated the catalytic conversion of 20 grams used oil over 0.2 grams
CC

of HZSM-5 and sulfated zirconia catalyst for liquid fuel production in a stainless steel micro
batch reactor with temperatures of 380-430oC, residence time of 45-90 min, pressure of 10-20
A

bars and different catalyst ratios. It has been shown that the increase in residence time from
45-60 min increased the gasoline yield from 8.54 to 22.15 wt.% and 14.48 to 18.61 wt.% for
HZSM-5 and sulfated zirconia respectively. Furthermore, a further increase in residence time
from 75 to 90 min resulted in a decrease in product distribution of gasoline, kerosene, light
gas oil and gas oil while gaseous product formation was considerably increased. Therefore, it
can be determined that residence time from 45-60 min is an adequate condition of formation

49
of gasoline, kerosene and light gas oil products. Suwannasom et al. also studied the effect of
different residence times from 30-180 min in a prototype autoclave reactor to achieve the
optimum percentage of jet fuel from palm oil TGs over Ni-Mo/HY zeolites. They observed
that the product selectivity of jet fuel increased from 24.8% to 30.15% when residence time
was increased from 30-120 min while a further increase in residence time from 120-180 min
resulted in the decrease in jet fuel yield (30.2 to 19.3%) and product selectivity: gasoline,
kerosene, light gas oil and gas oil. Thus, an optimum residence time is required to ensure the
optimized yield of jet fuel and selectivity of desired hydrocarbon products.

T
IP
7.5 Amount of catalyst

R
The amount of catalyst also has an important role in affecting the HDO reaction rate and
product selectivity [25]. By increasing the amount of catalyst, it is commonly reported to

SC
result in better catalytic activity as more catalytic sites were made available for the reactants
[148]. Liu et al. performed HDO of palm oil over Ni/SAPO11 catalysts with different Ni

U
loadings: 2, 5, 7 and 9 wt.% at 473 K, 1.0 h-1 LHSV and 4.0 MPa H2 pressure. With the
N
increase in Ni loading from 2 to 7 wt.%, liquid alkane yield and isomerization selectivity were
significantly increased from 60% to 67.4% and 46% to 61.5%, respectively. With further
A
increasing Ni loading from 7 to 9 wt.%, liquid alkane yield was retained at the same level.
M

However, C1-C14 alkane selectivity was considerably increased. The high isomerization
selectivity (>83%) was achieved with 9 wt.% of Ni/SAPO-11 catalyst. Consequently, the
ED

correct amount of catalyst loading showed great importance in ensuring optimum product
selectivity and product yield [99]. Sudhakara also reported of similar effects on different Ni
loadings (0-30 wt.%) of Ni/γ-Al2O3 catalyst for HDO of karanja oil at 35 bar H2 pressure and
PT

temperature 340-400 °C. At the lower amount of nickel loading (<20 wt.%), formation of
fatty acid intermediates was not prominent due to catalyst deactivation and undesired product
E

formation. Likewise, by increasing nickel loading to about 20-25 wt.%, HDO route begins to
CC

be more favored over thermal/catalytic cracking to facilitate the conversion of fatty acids to
alkanes.
A

The catalytic activity could also be affected by metal loading due to the changes in
catalyst morphologies. From the literature, it was shown that Ni/HZSM-5 of 7 wt.% Ni
loading had the optimum capacity to convert more than 90% of methyl hexadecanoate into
hydrocarbons with selectivity of 83% for C5-C16. This was due to the synergistic effect of Ni
metal site and HZSM-5 acid site. However, at lower Ni loadings (1 wt.% and 3 wt.%), the

50
NiO was exceedingly distributed on the surface of HZSM-5 and was hardly seen.
Furthermore, with the increase in Ni loading to about 7 wt.%, NiO particle sizes increased to
150 nm and experienced severe agglomeration on the catalyst surface [149]. Thus, the
optimum metal loading on catalyst support is important to ensure the morphologies of catalyst
are conducive to catalyze the HDO of TGs.

8. Industrialization of bio-jet fuel

T
Presently a great attention is made for aviation applications with large amount of

IP
hydrocarbon bio-fuels utilization compatible with existing fossil fuels [150]. Therefore, for

R
bio-jet fuel commercialization, the following major countries have studied and developed bio-
jet fuel industries as mentioned in the Table 9 and airlines that have conducted the bio-jet fuel

SC
tests as summarized in Table 10 [151]. As can be seen from Table 10, different oils with 1:1
ratio of jet fuel and bio-jet fuel were tested by different airlines to ensure the bio-jet fuel

U
specifications. It is estimated that by 2020, nine refineries will be able to deliver bio-
N
hydrocarbons of 2 mega tons, with an estimated total investment cost of about 3 billion Euro
[150].
A
M

A common description of fuel used in the aviation sector, is a fuel that meets the
specifications of fossil or crude oil derived jet fuel with same inclusive performance. A
ED

typical case would be hydrotreated vegetable oil or FAME, which have been industrially
established by several corporations for the production of bio-jet fuel. This type of bio jet-fuel
PT

is presently leading an important part with 2.2 million tonnes/year of production. In addition,
the feedstock used remains as conventional lipids. Therefore, it is important to confirm that
sustainable grown crops such as Camelina or waste oils (as used cooking oils, or other oily
E

residues) are used as feedstocks for jet-fuel production on industrial or commercial scale
CC

[150].
A

51
9. Conclusion
This present study has provided new trends on the DO of TGs for the alternative jet fuel
production based on recent studies. The corresponding selection of appropriate feedstocks,
reaction pathways, selection and synthesis of suitable catalysts and parameter optimization
were discussed. The nature of the vegetable oil as feedstock is a critical factor. From this
aspect, physicochemical properties of feedstock were identified as a key factor for certain
reaction conditions. The saturated feedstock is more useful as compared to unsaturated
feedstock for higher alkane formation during DO. In addition, DO of TGs with longer carbon

T
chain structures is of lower reaction rate and lesser conversion while fatty acids with shorter

IP
carbon chains had better conversion and hydrocarbon selectivity. Noble metals like Pd and Pt

R
with carbon support were found as suitable catalysts in DCO2 and DCO reaction, but their
high costs remain a challenge in their commerciality. On the other hand, non-sulfided catalyst

SC
with high surface area, pore volume and moderate acidity tends to be more active in HDO and
hydrocracking for a jet fuel range hydrocarbons with iso-paraffins. Zeolite and Al2O3

U
supported catalysts are highly preferred for the HDO of vegetable oil to achieve high
N
selectivity. Therefore, the non-sulphided catalysts had been modified with different precursors
and promoters to enhance the catalyst activity and product selectivity. Several studies
A
explicated the influence of different operating parameters like pressure, temperature,
M

residence time and catalyst amount on hydrocarbon conversion and selectivity. By using high
H2 pressure during HDO, the process can significantly remove oxygen from TGs and
ED

decreases the production of aromatic hydrocarbons thereby increasing the hydrocarbon


selectivity. An increase in reaction temperature resulted in higher formation rate of odd
PT

number of carbon product, whereas an increase in residence time decreases the conversion
rate and shortens catalyst lifespan. Likewise, the increase in catalyst amount during HDO also
improved the conversion rate, selectivity of hydrocarbons and reduces the catalyst
E

deactivation. Some relevant discussions about standard and TGs derived alternative jet fuel
CC

specification and fuel properties are also presented in this review. Information about
combustion characteristics, including ignition characteristics for alternative fuels are also
A

discussed. As it is established that HDO process is an optional solution to fulfill the increasing
energy demands and to reduce fossil fuel dependency. With the appropriate selection of
feedstock, supported catalysts and optimization of reaction conditions, HDO could be one of
the promising techniques for production of alternative jet fuel. Based on this critical
assessment, this research is focused on the development of new low-cost catalyst and more
efficient HDO process for the facile production of jet fuel from feedstock.

52
Acknowledgement:

The authors thank GSP-MOHE, University of Malaya for fully funding this study through the
project number UM. 0000066/HME.OM

References

T
[1] P. Kallio, A. Pásztor, M.K. Akhtar, P.R. Jones, Renewable jet fuel, Current Opinion in
Biotechnology 26 (2014) 50-55.

IP
[2] S. Blakey, L. Rye, C.W. Wilson, Aviation gas turbine alternative fuels: A review, Proceedings
of the Combustion Institute 33 (2011) 2863-2885.
[3] X. Wu, P. Jiang, F. Jin, J. Liu, Y. Zhang, L. Zhu, T. Xia, K. Shao, T. Wang, Q. Li, Production

R
of jet fuel range biofuels by catalytic transformation of triglycerides based oils, Fuel 188
(2017) 205-211.

SC
[4] A.M. Ashraful, H.H. Masjuki, M.A. Kalam, I.M. Rizwanul Fattah, S. Imtenan, S.A. Shahir,
H.M. Mobarak, Production and comparison of fuel properties, engine performance, and
emission characteristics of biodiesel from various non-edible vegetable oils: A review, Energy

[5]
Conversion and Management 80 (2014) 202-228.
U
D. Kubička, J. Horáček, M. Setnička, R. Bulánek, A. Zukal, I. Kubičková, Effect of support-
active phase interactions on the catalyst activity and selectivity in deoxygenation of
N
triglycerides, Applied Catalysis B: Environmental 145 (2014) 101-107.
[6] M.Y. Kim, J.-K. Kim, M.-E. Lee, S. Lee, M. Choi, Maximizing Biojet Fuel Production from
A
Triglyceride: Importance of the Hydrocracking Catalyst and Separate
Deoxygenation/Hydrocracking Steps, catalysis 9 (2017) 6526-6267.
M

[7] L. Li, E. Coppola, J. Rine, J.L. Miller, D. Walker, Catalytic hydrothermal conversion of
triglycerides to non-ester biofuels, Energy and Fuels 24 (2010) 1305-1315.
[8] J. Xu, J. Jiang, J. Zhao, Thermochemical conversion of triglycerides for production of drop in
ED

liquid fuels, Renewable and Sustainable Energy Reviews 58 (2016) 331-340.


[9] M. Ameen, M.T. Azizan, S. Yusup, A. Ramli, M. Yasir, Catalytic hydrodeoxygenation of
triglycerides: An approach to clean diesel fuel production, Renewable and Sustainable Energy
Reviews 80 (2017) 1072-1088.
PT

[10] T. Kimura, H. Imai, X. Li, K. Sakashita, S. Asaoka, S.S. Al-Khattaf, "Hydroconversion of


triglycerides to hydrocarbons over Mo-Ni/γ-Al2O3 catalyst under low hydrogen pressure,
Catalyst Letter 143 (2013) 1175-1181.
[11] S. Liu, Q. Zhu, Q. Guan, L. He, W. Li, Bio-aviation fuel production from hydroprocessing
E

castor oil promoted by the nickel-based bifunctional catalysts, Bioresource Technology 183
(2015) 93-100.
CC

[12] D. Verma, B.S. Rana, R. Kumar, M.G. Sibi, A.K. Sinha, Diesel and aviation kerosene with
desired aromatics from hydroprocessing of jatropha oil over hydrogenation catalysts supported
on hierarchical mesoporous SAPO-11, Applied Catalysis A: General 490 (2015) 108-116.
[13] M. Rabaev, M.V. Landau, R. Vidruk-Nehemya, V. Koukouliev, R. Zarchin, M. Herskowitz,
A

Conversion of vegetable oils on Pt/Al2O3/SAPO-11 to diesel and jet fuels containing


aromatics, Fuel 161 (2015) 287-294.
[14] V. Itthibenchapong, A. Srifa, R. Kaewmeesri, P. Kidkhunthod, K. Faungnawakij,
Deoxygenation of palm kernel oil to jet fuel like hydrocarbons using Ni-MoS2/γ-Al2O3
catalysts, Energy Conversion and Management 134 (2017) 188-196.
[15] G. Liu, B. Yan, G. Chen, Technical review on jet fuel production, Renewable and Sustainable
Energy Reviews 25 (2013) 59-70.
[16] H. Wang, H. Lin, P. Feng, X. Han, Y. Zheng, Integration of catalytic cracking and
hydrotreating technology for triglyceride deoxygenation, Catalysis Today (2016)

53
[17] T. Stedile, L. Ender, H.F. Meier, E.L. Simionatto, V.R. Wiggers, Comparison between
physical properties and chemical composition of bio-oils derived from lignocellulose and
triglyceride sources, Renewable and Sustainable Energy Reviews 50 (2015) 92-108.
[18] M. Ferrari, C. Lahousse, A. Centeno, R. Maggi, P. Grange, B. Delmon Influence of the
impregnation order of molybdenum and cobalt in carbon supported catalysts for
hydrodeoxygenation reactions. In B. Delmon et al. (Ed.), Studies in Surface Science and
Catalysis (pp. 505-515). Elsevier; 1998.
[19] A.N. Kay Lup, F. Abnisa, W.M.A.W. Daud, M.K. Aroua, A review on reaction mechanisms
of metal-catalyzed deoxygenation process in bio-oil model compounds, Applied Catalysis A:
General 541 (2017) 87-106.
[20] T. Kandaramath Hari, Z. Yaakob, N.N. Binitha, Aviation biofuel from renewable resources:
Routes, opportunities and challenges, Renewable and Sustainable Energy Reviews 42 (2015)

T
1234-1244.
[21] I.K. kova, D. Kubicka, Utilization of triglycerides and related feedstocks for production of

IP
clean hydrocarbon fuels and petrochemicals: a review, Waste Biomass Valor 1 (2010) 293-
308.
[22] M. Anand, S.A. Farooqui, R. Kumar, R. Joshi, R. Kumar, M.G. Sibi, H. Singh, A.K. Sinha,

R
Kinetics, thermodynamics and mechanisms for hydroprocessing of renewable oils, Applied
Catalysis A: General 516 (2016) 144-152.

SC
[23] A. Galadima, O. Muraza, Catalytic upgrading of vegetable oils into jet fuels range
hydrocarbons using heterogeneous catalysts: a review, Journal of Industrial and Engineering
Chemistry 29 (2015) 12-23.
[24]
U
R.W. Gosselink, S.A.W. Hollak, S.W. Chang, J.v. Haveren, a. Krijn P. de Jong, J.H. Bitter,
D.S.v. Es, Reaction pathways for the deoxygenation of vegetable oils and related model
compounds, ChemSusChem 0 (2013) 1-20.
N
[25] B.P. Pattanaik, R.D. Misra, Effect of reaction pathway and operating parameters on the
deoxygenation of vegetable oils to produce diesel range hydrocarbon fuels: A review,
A
Renewable and Sustainable Energy Reviews 73 (2017) 545-557.
[26] L. Chen, H. Li, J. Fu, C. Miao, P. Lv, Z. Yuan, Catalytic hydroprocessing of fatty acid methyl
M

esters to renewable alkane fuels over Ni/HZSM-5 catalyst, Catalysis Today 259 (2016) 266-
276.
[27] S. Palanisamy, B.S. Gevert, Study of non-catalytic thermal decomposition of triglyceride at
ED

hydroprocessing condition, Applied Thermal Engineering 107 (2016) 301-310.


[28] R.Y. Sudhakara, M. Sunil K., S. Debaprasad, Hydrodeoxygenation of karanja oil over
supported nickel catalysts: Influence of support and nickel loading, Catalysis Science and
Technology 1 (2015) 1-10.
PT

[29] S.K. Kim, J.Y. Han, H.-s. Lee, T. Yum, Y. Kim, J. Kim, Production of renewable diesel via
catalytic deoxygenation of natural triglycerides: Comprehensive understanding of reaction
intermediates and hydrocarbons, Applied Energy 116 (2014) 199-205.
[30] A. Sagiroglu, H.M. Ozcan, S.S. Isbilir, H. Paluzar, N.M. Toprakkiran, Alkali catalysis of
E

different vegetable oils for comparisons of their biodiesel productivity, Journal of Sustainable
Bioenergy Systems 3 (2013) 79-85.
CC

[31] Y. Liu, R.S. Boyas, K. Murata, T. Minowa, K. Sakanishi, 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 25 (2011) 4675-4685.
[32] X. Zhao, L. Wei, J. Julson, Q. Qiao, A. Dubey, G. Anderson, Catalytic cracking of non-edible
A

sunflower oil over ZSM-5 for hydrocarbon bio jet fuel, New Biotechnology 32 (2015) 300-
312.
[33] X. Zhao, L. Wei, S. Cheng, J. Julson, Review of heterogeneous catalysts for catalytically
upgrading vegetable oils into hydrocarbon biofuels, Catalysts 7 (2017) 1-25.
[34] B. Veriansyah, J.Y. Han, S.K. Kim, S.-A. Hong, Y.J. Kim, J.S. Lim, Y.-W. Shu, S.-G. Oh, J.
Kim, Production of renewable diesel by hydroprocessing of soybean oil: Effect of catalysts,
Fuel 94 (2012) 578-585.
[35] O.L. E, D.S.M.L.C. P, Comparative study of calorific value of rapeseed, soybean, jatropha
curcas and crambe biodiesel, Renewable energies and power quality journal (2013) 1-5.

54
[36] S.A.P. da Mota, A.A. Mancio, D.E.L. Lhamas, D.H. de Abreu, M.S. da Silva, W.G. dos
Santos, D.A.R. de Castro, R.M. de Oliveira, M.E. Araújo, L.E.P. Borges, N.T. Machado,
Production of green diesel by thermal catalytic cracking of crude palm oil (Elaeis guineensis
Jacq) in a pilot plant, Journal of Analytical and Applied Pyrolysis 110 (2014) 1-11.
[37] P. Lovas, P. Hudec, M. Hadvinova, A. Haz, Conversion of rapeseed oil via catalytic cracking:
Effect of the ZSM-5 catalyst on the deoxygenation process, Fuel Processing Technology 134
(2015) 223-230.
[38] H. Noureddini, B.C. Teoh, L.D. Clements, Densities of vegetable oils and fatty acids,
Chemical and Biomolecular Engineering Research and Publications (1992) 1184-1188.
[39] C.W. Wei, T. Ling, M. Jennifer, Z. Yanan, T. Eric, B. Liaw, W. Ethan, B. Mary, Review of
biojet fuel conversion technologies. 2016, National Renewable Energy Laboratory: U.S.
[40] C. Telmo, J. Lousada, Heating values of wood pellets from different species, Biomass and

T
Bioenergy 35 (2011) 2634-2639.
[41] M. Carrier, T. Hugo, J. Gorgens, H. Knoetze, Comparison of slow and vacuum pyrolysis of

IP
sugar cane bagasse, Journal of Analytical and Applied Pyrolysis 90 (2011) 18-26.
[42] Z. Si, X. Zhang, C. Wang, L. Ma, R. Dong, An overview on catalytic hydrodeoxygenation of
pyrolysis oil and its model compounds, Catalysts 7 (2017) 1-47.

R
[43] L. Li, K. Quan, J. Xu, F. Liu, S. Liu, S. Yu, C. Xie, B. Zhang, X. Ge, Liquid hydrocarbon
fuels from catalytic cracking of rubber seed oil using USY as catalyst, Fuel 123 (2014) 189-

SC
193.
[44] C. Gutierrez-Antonio, F.I. Gomez Castro, J.A. de Lira-Flores, S. Hernandez, A review on the
production processes of renewable jet fuel, Renewable and Sustainable Energy Reviews 79

[45]
(2017) 709-729.
U
C. Zhang, X. Hui, Y. Lin, C.-J. Sung, Recent development in studies of alternative jet fuel
combustion: Progress, challenges, and opportunities, Renewable and Sustainable Energy
N
Reviews 54 (2016) 120-138.
[46] E. Tim, M. Jeffrey, Stricker, E. William, Harrison, Handbook of aviation fuel properties:
A
Coordinating support of fuels and lubricant research and development (R&D) 2. 2004,
Coordinating Research Council, Inc. 3650 Mansell Road, Suite 140 Alpharetta, GA 30022
M

U.S.A.
[47] X. Xue, X. Hui, P. Singh, C.-J. Sung, Soot formation in non-premixed counterflow flames of
conventional and alternative jet fuels, Fuel 210 (2017) 343-351.
ED

[48] H. Zhang, H. Lin, Y. Zheng, The role of cobalt and nickel in deoxygenation of vegetable oils,
Applied Catalysis B: Environmental 160-161 (2014) 415-422.
[49] D.M. Korres, D. Karonis, E. Lois, M.B. Linck, A.K. Gupta, Aviation fuel JP-5 and biodiesel
on a diesel engine, Fuel 87 (2008) 70-78.
PT

[50] K. Dilip, Adhikari Bio-jet Fuel. In (Ed.), Biofuel and Biorefinery Technologies book series
(BBT) (pp. 187-201). Springer, Cham; 2018.
[51] A.P.P. Pires, J.K. Yinglei Han, M.G. Perez, Chemical composition and fuel properties of
alternative jet fuels, Bio Resources 13 (2018) 2632-2657.
E

[52] S.D. Anuar Sharuddin, F. Abnisa, W.M.A. Wan Daud, M.K. Aroua, A review on pyrolysis of
plastic wastes, Energy Conversion and Management 115 (2016) 308-326.
CC

[53] M.C. Vasquez, E.E. Silva, E.F. Castillo, Hydrotreatment of vegetable oils: A review of the
technologies and its developments for jet biofuel production, Biomass and Bioenergy 105
(2017) 197-206.
[54] F.P. de Sousa, C.C. Cardoso, V.M.D. Pasa, Producing hydrocarbons for green diesel and jet
A

fuel formulation from palm kernel fat over Pd/C, Fuel Processing Technology 143 (2016) 35-
42.
[55] A.a.H. Al-Muhtaseb, F. Jamil, L. Al-Haj, M.A. Al-Hinai, M. Baawain, M.T.Z. Myint, D.
Rooney, Efficient utilization of waste date pits for the synthesis of green diesel and jet fuel
fractions, Energy Conversion and Management 127 (2016) 226-232.
[56] G.E. Totten. Fuels and Lubricants Handbook: Technology, Properties, Performance, and
Testing Printed in Glen Bumie, New York: 2003.
[57] A. Donkor, S. Nyarko, K.O. Asemani, J.-C. Bonzongo, K. Kyeremeh, C. Ziwu, A novel
approach for reduction of total acidity in kerosene based on alkaline rich materials readily

55
available in tropical and sub-tropical countries, Egyptian Journal of Petroleum 25 (2016) 473-
480.
[58] R.H. Moore, M. Shook, A. Beyersdorf, C. Corr, S. Herndon, W.B. Knighton, R. Miake-Lye,
K.L. Thornhill, E.L. Winstead, Z. Yu, L.D. Ziemba, B.E. Anderson, Influence of Jet Fuel
Composition on Aircraft Engine Emissions: A Synthesis of Aerosol Emissions Data from the
NASA APEX, AAFEX, and ACCESS Missions, Energy & Fuels 29 (2015) 2591-2600.
[59] I.A. Al-Nuaimi, M. Bohra, M. Selam, H.A. Choudhury, M.M. El-Halwagi, N.O. Elbashir,
Optimization of synthetic jet fuels aromatic/paraffinic composition via experimental &
property integration methods, Chemical Engineering & Technology (2016) 1-23.
[60] A.A. AKERE, Investigating the links between smoke points, sooting thresholds, and particle
number and size, in Department of Chemical Engineering 2009, UNIVERSITY OF
CAMBRIDGE UNIVERSITY OF CAMBRIDGE

T
[61] M. Anand, S.A. Farooqui, R. Kumar, R. Joshi, R. Kumar, M.G. Sibi, H. Singh, A.K. Sinha,
Optimizing renewable oil hydrocracking conditions for aviation bio-kerosene production, Fuel

IP
Processing Technology 151 (2016) 50-58.
[62] Y. Sugami, E. Minami, S. Saka, Renewable diesel production from rapeseed oil with
hydrothermal hydrogenation and subsequent decarboxylation, Fuel 166 (2016) 376-381.

R
[63] N. Arun, R.V. Sharma, A.K. Dalai, Green diesel synthesis by hydrodeoxygenation of bio-
based feedstocks: Strategies for catalyst design and development, Renewable and Sustainable

SC
Energy Reviews 48 (2015) 240-255.
[64] T.V. Choudhary, C.B. Phillips, Renewable fuels via catalytic hydrodeoxygenation, Applied
Catalysis A: General 397 (2011) 1-12.
[65]
U
C. Ma, J. Yu, B. Wang, Z. Song, F. Zhou, J. Xiang, S. Hu, L. Sun, Influence of Zeolites and
Mesoporous Catalysts on Catalytic Pyrolysis of Brominated Acrylonitrile–Butadiene–Styrene
(Br-ABS), Energy & Fuels 30 (2016) 4635-4643.
N
[66] D.J. Valco, K. Min, A. Oldani, T. Edwards, T. Lee, Low temperature autoignition of
conventional jet fuels and surrogate jet fuels with targeted properties in a rapid compression
A
machine, Proceedings of the Combustion Institute 36 (2017) 3687-3694.
[67] M.M.C. T. V. Malleswara Rao , Michiel Makkee, Effective gasoline production strategies by
M

catalytic cracking of rapeseed vegetable oil in refinery conditions, ChemSusChem 3 (2010)


807 - 810
[68] M.J. Ramos, C.M. Fernandez, A. Casas, L. Rodriguez, A. Perez, Influence of fatty acid
ED

composition of raw materials on biodiesel properties, Bioresource Technology 100 (2009)


261-268.
[69] D.-S. Lee, B.-S. Noh, S.-Y. Bae, K. Kim, Characterization of fatty acids composition in
vegetable oils by gas chromatography and chemometrics, Analytica Chimica Acta 358 (1998)
PT

163-175.
[70] R. Tiwari, B.S. Rana, R. Kumar, D. Verma, R. Kumar, R.K. Joshi, M.O. Garg, A.K. Sinha,
Hydrotreating and hydrocracking catalysts for processing of waste soya-oil and refinery-oil
mixtures, Catalysis Communications 12 (2011) 559-562.
E

[71] R. Cataluna, R. da Silva, Effect of cetane number on specific fuel consumption and particulate
matter and unburned hydrocarbon emissions from diesel engines, Journal of Combustion 2012
CC

(2012) 1-6.
[72] D. Bajpai, V.K. Tyagi, Biodiesel: Source, production, composition, properties and its benefits,
Journal of Oleo Science 55 (2006) 487-502.
[73] C. Zhao, T. Bruckb, J. A, Catalytic deoxygenation of microalgae oil to green hydrocarbons,
A

Green Chemistry (2013) 1720-1739.


[74] M. Mohammad, T. Kandaramath Hari, Z. Yaakob, Y. Chandra Sharma, K. Sopian, Overview
on the production of paraffin based biofuels via catalytic hydrodeoxygenation, Renewable and
Sustainable Energy Reviews 22 (2013) 121-132.
[75] P. Papilo, Marimin, E. Hambali, I.S. Sitanggang, Sustainability index assessment of palm oil-
based bioenergy in Indonesia, Journal of Cleaner Production 196 (2018) 808-820.
[76] M. Willems, Sustainable palm oil production in Thailand, in Environmental Sciences. 2015,
Wageningen University: Netherland. p. 1-88.

56
[77] Z. Eller, Z. Varga, J. Hancsok, Advanced production process of jet fuel components from
technical grade coconut oil with special hydrocracking, Fuel 182 (2016) 713-720.
[78] K. Jenistova, I. Hachemi, P. Makia Arvela, N. Kumar, M. Peurla, L. Capek, J. Warna, D.Y.
Murzin, Hydrodeoxygenation of stearic acid and tall oil fatty acids over Ni-alumina catalysts:
Influence of reaction parameters and kinetic modelling, Chemical Engineering Journal 316
(2017) 401-409.
[79] A. Galadima, J.A. Anderson, R.P.K. Wells, Solid acid catalysts in heterogeneous n-alkanes
hydroisomerisation for increasing octane number of gasoline, Sci. World J. 4 (2009)
[80] C. Gutierrez Antonio, F.I. Gomez Castro, J.A. de Lira Flores, S. Hernandez, A review on the
production processes of renewable jet fuel, Renewable and Sustainable Energy Reviews 79
(2017) 709-729.
[81] I.H. Choi, K.R. Hwang, J.S. Han, K.H. Lee, J.S. Yun, J.S. Lee, The direct production of jet-

T
fuel from non-edible oil in a single-step process, Fuel 158 (2015) 98-104.
[82] A. Sonthalia, N. Kumar, Hydroprocessed vegetable oil as a fuel for transportation sector:

IP
A review, Journal of the Energy Institute (2017)
[83] Z. Zhang, Q. Wang, H. Chen, X. Zhang, Hydroconversion of Waste Cooking Oil into Green
Biofuel over Hierarchical USY-Supported NiMo Catalyst: A Comparative Study of

R
Desilication and Dealumination," Catalysts 7 (2017) 1-13.
[84] H. Zhang, Q. Wang, S.R. Mortimer, Waste cooking oil as an energy resource: Review of

SC
Chinese policies, Renewable and Sustainable Energy Reviews 16 (2012) 5225-5231.
[85] M. Saber, B. Nakhshiniev, K. Yoshikawa, A review of production and upgrading of algal bio-
oil, Renewable and Sustainable Energy Reviews 58 (2016) 918-930.
R.P. Rastogi, A. Pandey, C. Larroche, D. Madamwar, Algal Green Energy – R&D and
[86]
U
technological perspectives for biodiesel production, Renewable and Sustainable Energy
Reviews 82 (2018) 2946-2969.
N
[87] L.N. Silva, I.C.P. Fortes, F.P. de Sousa, V.M.D. Pasa, Biokerosene and green diesel from
macauba oils via catalytic deoxygenation over Pd/C, Fuel 164 (2016) 329-338.
A
[88] P.L. Chu, C. Vanderghem, H.L. MacLean, B.A. Saville, Process modeling of
hydrodeoxygenation to produce renewable jet fuel and other hydrocarbon fuels, Fuel 196
M

(2017) 298-305.
[89] X. Li, X. Luo, Y. Jin, J. Li, H. Zhang, A. Zhang, J. Xie, Heterogeneous sulfur-free
hydrodeoxygenation catalysts for selectively upgrading the renewable bio-oils to second
ED

generation biofuels, Renewable and Sustainable Energy Reviews (2017)


[90] J.E. Santillan, C. Mark, Catalytic deoxygenation of fatty acids and their derivatives to
hydrocarbon fuels via decarboxylation/decarbonylation, Journal of Chemical Technology and
Biotechnology 87 (2012) 1041-1050.
PT

[91] R.o.a.f. I. A. T. A. Association, " Switzerland 2014 December.


[92] Q. Smejkal, L. Smejkalova, D. Kubicka, Thermodynamic balance in reaction system of total
vegetable oil hydrogenation, Chemical Engineering Journal 146 (2009) 155-160.
[93] L. Hermida, A.Z. Abdullah, A.R. Mohamed, Deoxygenation of fatty acid to produce diesel
E

like hydrocarbons: A review of process conditions, reaction kinetics and mechanism,


Renewable and Sustainable Energy Reviews 42 (2015) 1223-1233.
CC

[94] J.E. Santillan, T. Morgan, J. Shoup, W.A.E. Harman, M. Crocker, Catalytic deoxygenation of
triglycerides and fatty acids to hydrocarbons over Ni-Al layered double hydroxide, Catalysis
Today 237 (2014) 136-144.
[95] L. Jeczmionek, K. Porzycka-Semczuk, Hydrodeoxygenation, decarboxylation and
A

decarbonylation reactions while co-processing vegetable oils over a NiMo hydrotreatment


catalyst. Part I: Thermal effects – Theoretical considerations, Fuel 131 (2014) 1–5.
[96] C. Gonzalez, P. Marin, F.V. Diez, S. Ordonez, Gas-phase hydrodeoxygenation of
benzaldehyde, benzyl alcohol, phenyl acetate, and anisole over precious metal catalysts,
Industrial & Engineering Chemistry Research 55 (2016) 2319-2327.
[97] A. Galadima, R.P.K. Wells, J.A. Anderson, n-Alkane hydroconversion over carbided
molybdena supported on sulfated zirconia, Appl. Petrochem. Res. 1 (2012) 35-43.
[98] A. Galadima, A. Ismail, I. Abdullahi, H. Isah, Supported molybdenum carbide as n-hexane
upgrading catalyst, Chem. Mater. Res. 3 (2013) 73-78.

57
[99] Q. Liu, H. Zuo, Q. Zhang, T. Wang, L. Ma, Hydrodeoxygenation of palm oil to hydrocarbon
fuels over Ni/SAPO-11 catalysts, Chinese Journal of Catalysis 35 (2014) 748-756.
[100] T. Blasco, A. Chica, A. Corma, W.J. Murphy, J. Agúndez-Rodríguez, J. Pérez-Pariente,
Changing the Si distribution in SAPO-11 by synthesis with surfactants improves the
hydroisomerization/dewaxing properties, Journal of Catalysis 242 (2006) 153-161.
[101] F. Alvarez, F.R. Ribeiro, G. Perot, C. Thomazeau, M. Guisnet, Hydroisomerization and
Hydrocracking of Alkanes: 7. Influence of the Balance between Acid and Hydrogenating
Functions on the Transformation ofn-Decane on PtHY Catalysts, Journal of Catalysis 162
(1996) 179-189.
[102] C. Wang, Q. Liu, X. Liu, L. Yan, C. Luo, L. Wang, B. Wang, Z. Tian, Influence of reaction
conditions on one-step hydrotreatment of lipids in the production of iso-alkanes over
Pt/SAPO-11, Chinese Journal of Catalysis 34 (2013) 1128-1138.

T
[103] S.R. Yenumala, S.K. Maity, Reaction mechanism and kinetic modeling for the
hydrodeoxygenation of triglycerides over alumina supported nickel catalyst, Reaction

IP
Kinetics, Mechanisms and Catalysis 120 (2017) 109-128.
[104] J. Chen, H. Shi, L. Li, K. Li, Deoxygenation of methyl laurate as a model compound to
hydrocarbons on transition metal phosphide catalysts, Applied Catalysis B: Environmental

R
144 (2014) 870-884.
[105] M.T. Azizana, K.A. Jais, M.H. Sa’aida, M. Ameena, A.F. Shahudina, M. Yasir, S. Yusupa, A.

SC
Ramli, Thermodynamic equilibrium analysis of triolein hydrodeoxygenation for green diesel
production, Procedia Engineering 148 (2016) 1369-1376
[106] R. Bartosz, M.A. Paivi, T. Anton, R.L. Anne, E. Kari, Y.M. Dmitry, Influence of hydrogen in

[107]
Engineering Chemistry Research 51 (2012) 8922-8927.
U
catalytic deoxygenation of fatty acids and their derivatives over Pd/C, Industrial and

J. Horacek, Z. Tisler, V. Rubas, D. Kubicka, HDO catalysts for triglycerides conversion into
N
pyrolysis and isomerization feedstock, Fuel 121 (2014) 57-64.
[108] Q. Liu, H. Zuo, T. Wang, L. Ma, Q. Zhang, One-step hydrodeoxygenation of palm oil to
A
isomerized hydrocarbon fuels over Ni supported on nanosized SAPO-11 catalysts, Applied
Catalysis A: General 468 (2013) 68-74.
M

[109] N.T.T. Truong, A. Boontawan, Development of bio-jet fuel production using palm kernel oil
and ethanol, in International Journal of Chemical Engineering and Applications. 2017:
Thailand. p. 97-105.
ED

[110] E.F. Vansant Pore Size Engineering in Zeolites. In P.J. Grobet et al. (Ed.), Studies in Surface
Science and Catalysis (pp. 143-153). Elsevier; 1988.
[111] T.L.J. Cheng, R.H.J. Zhou, K. Cen, Conversion pathways of palm oil into jet biofuel catalyzed
by mesoporous zeolites, Royal Society of Chemistry 6 (2016) 103965-103972.
PT

[112] A. Srifa, K. Faungnawakij, V. Itthibenchapong, S. Assabumrungrat, Roles of monometallic


catalysts in hydrodeoxygenation of palm oil to green diesel, Chemical Engineering Journal
278 (2015) 249-258.
[113] Z. Xie, Z. Liu, Y. Wang, Q. Yang, L. Xu, W. Ding, An overview of recent development in
E

composite catalysts from porous materials for various reactions and processes, International
Journal of Molecular Sciences 11 (2010) 2152-2187.
CC

[114] G. Busca, Acidity and basicity of zeolites: A fundamental approach, Microporous and
Mesoporous Materials 254 (2017) 3-16.
[115] K. Muraoka, W. Chaikittisilp, T. Okubo, Energy analysis of aluminosilicate zeolites with
comprehensive ranges of framework topologies, chemical compositions, and aluminum
A

distributions, Journal of American chemical society 138 (2016) 6184−6193.


[116] A.N. Kay Lup, F. Abnisa, W.M.A. Wan Daud, M.K. Aroua, A review on reactivity and
stability of heterogeneous metal catalysts for deoxygenation of bio-oil model compounds,
Journal of Industrial and Engineering Chemistry 56 (2017) 1-34.
[117] J. Yang, Z. Xin, Q. He, K. Corscadden, H. Niu, An overview on performance characteristics
of bio-jet fuels, Fuel 237 (2019) 916-936.
[118] A.N. Kay Lup, F. Abnisa, W.M.A. Wan Daud, M.K. Aroua, Synergistic interaction of metal-
acid sites for phenol hydrodeoxygenation over bifunctional Ag/TiO2 nanocatalyst, Chinese
journal of Chemical Engineering (2018)

58
[119] I.V. Deliy, E.N. Vlasova, A.L. Nuzhdin, G.A. Bukhtiyarova, The comparison of sulfide
CoMo/γ-Al2O3 and NiMo/γ-Al2O3 catalysts in methyl palmitate and methyl heptanoate
hydrodeoxygenation in Recent Researches in Engineering and Automatic Control, V.M. Nikos
Mastorakis, Editor. 2011, WSEAS Puerto De La Cruz, Tenerife, Spain p. 24-29.
[120] S.B. Rogelio, L. Yanyong, M. Tomoaki, Renewable diesel production from the hydrotreating
of rapeseed oil with Pt/Zeolite and NiMo/Al2O3 catalysts, Industrial and Engineering
Chemistry Research 50 (2011) 2791–2799.
[121] M. Shahinuzzaman, Z. Yaakob, Y. Ahmed, Non-sulphide zeolite catalyst for bio-jet-fuel
conversion, Renewable and Sustainable Energy Reviews 77 (2017) 1375-1384.
[122] L.L. Hegedus, R.W. McCabe Catalyst Poisoning. In B. Delmon & G.F. Froment (Ed.), Studies
in Surface Science and Catalysis (pp. 471-505). Elsevier; 1980.
[123] S. Bagheri, N. Muhd Julkapli, S. Bee Abd Hamid, Titanium Dioxide as a Catalyst Support in

T
Heterogeneous Catalysis, The Scientific World Journal 2014 (2014) 21.
[124] T. Li, J. Cheng, R. Huang, J. Zhou, K. Cen, Conversion of waste cooking oil to jet biofuel

IP
with nickel-based mesoporous zeolite Y catalyst, Bioresource Technology 197 (2015) 289-
294.
[125] R.Y. Sudhakara, M. Sunil K., S. Debaprasad, Hydrodeoxygenation of karanja oil over

R
supported nickel catalysts: Influence of support and nickel loading, Catalysis Science &
Technology 00 (2015) 1-10.

SC
[126] D. Li, H. Xin, X. Du, X. Hao, Q. Liu, C. Hu, Recent advances for the production of
hydrocarbon biofuel via deoxygenation progress, Science Bulletin 60 (2015) 2096-2106.
[127] J. Tapia, N.Y. Acelas, D. Lopez, A. Moreno, NiMo-sulfide supported on activated carbon to

[128]
U
produce renewable diesel, Universitas Scientiarum 22 (2017) 71-85.
Y. Romero, F. Richard, S. Brunet, Hydrodeoxygenation of 2-ethylphenol as a model
compound of bio-crude over sulfided Mo-based catalysts: Promoting effect and reaction
N
mechanism, Applied Catalysis B: Environmental 98 (2010) 213-223.
[129] Z. Pan, R. Wang, Z. Nie, J. Chen, Effect of a second metal (Co, Fe, Mo and W) on
A
performance of Ni2P/SiO2 for hydrodeoxygenation of methyl laurate, Journal of Energy
Chemistry 25 (2016) 418-426.
M

[130] F. Jamil, A.a.H. Al-Muhatseb, L. Al-Haj, M.A. Al-Hinai, M. Baawain, Phoenix dactylifera
kernel oil used as potential source for synthesizing jet fuel and green diesel, Energy Procedia
118 (2017) 35-39.
ED

[131] T. Kimura, H. Imai, X. Li, K. Sakashita, S. Asaoka, S.S. Al-Khattaf, Hydroconversion of


triglycerides to hydrocarbons over Mo–Ni/c-Al2O3 catalyst under low hydrogen pressure,
Catalyst Letter 143 (2013) 1175–1181.
[132] E.Y. Emori, F.H. Hirashima, C.H. Zandonai, C.A. Ortiz Bravo, N.R.C. Fernandes Machado,
PT

M.H.N. Olsen Scaliante, Catalytic cracking of soybean oil using ZSM5 zeolite, Catalysis
Today 279 (2017) 168-176.
[133] H. Chen, Q. Wang, X. Zhang, L. Wang, Hydroconversion of jatropha oil to alternative fuel
over hierarchical ZSM-5, Industrial and Engineering Chemistry Research 2 (2014) 1-35.
E

[134] H. Chen, Q. Wang, X. Zhang, L. Wang, Effect of support on the NiMo phase and its catalytic
hydrodeoxygenation of triglycerides, Fuel 159 (2015) 430-435.
CC

[135] A.K. Sinha, M.G. Sibi, N. Naidu, S.A. Farooqui, M. Anand, R. Kumar, Process Intensification
for Hydroprocessing of Vegetable Oils: An Experimental study, Industrial & Engineering
Chemistry Research 0 (2014) 1-36.
[136] M.J.A. Romero, A. Pizzi, G. Toscano, G. Busca, B. Bosio, E. Arato, Deoxygenation of waste
A

cooking oil and non-edible oil for the production of liquid hydrocarbon biofuels, Waste
Management 47, Part A (2016) 62-68.
[137] K.M. Qureshi, A.N. Kay Lup, S. Khan, F. Abnisa, W.M.A. Wan Daud, A technical review on
semi-continuous and continuous pyrolysis process of biomass to bio-oil, Journal of Analytical
and Applied Pyrolysis 131 (2018) 52-75.
[138] T. Nimmanwudipong, R.C. Runnebaum, K. Brodwater, J. Heelan, D.E. Block, B.C. Gates,
Design of a high-pressure flow-reactor system for catalytic hydrodeoxygenation: guaiacol
conversion catalyzed by platinum supported on MgO, Energy & Fuels 28 (2014) 1090-1096.

59
[139] S.K. Bej, Performance evaluation of hydroprocessing catalysts- A review of experimental
techniques, Energy & Fuels 16 (2002) 774-784.
[140] B.D. Gusmao J, Djéga-Mariadassou G, Frety R. , Utilization of vegetable oils as an alternative
source for diesel-type fuel: hydrocracking on reduced Ni/SiO2 and sulphided Ni-Mo/γ-
Al2O3., Catalysis Today 5 (1989) 533-44.
[141] T.J. Guzman A, Prada LP, Nunez ML. , Hydroprocessing of crude palm oil at pilot plant scale.
, Catalysis Today 156 (2010) 38-43.
[142] J.E. Santillan, T. Morgan, J. Lacny, S. Mohapatra, M. Crocker, Catalytic deoxygenation of
triglycerides and fatty acids to hydrocarbons over carbon supported nickel, Fuel 103 (2013)
1010-1017.
[143] A.N. Kay Lup, F. Abnisa, W.M.A. Wan Daud, M.K. Aroua, Delayed volatiles release
phenomenon at higher temperature in TGA via sample encapsulation technique, Fuel 234

T
(2018) 422-429.
[144] J.K. Satyarthi, T. Chiranjeevi, D.T. Gokak, P.S. Viswanathan, An overview of catalytic

IP
conversion of vegetable oils/fats into middle distillates, Catalysis Science and Technology 3
(2013) 70-80.
[145] E. Fumoto, Y. Sugimoto, S. Sato, T. Takanohashi, Catalytic cracking of heavy oil with iron

R
oxide-based catalysts using hydrogen and oxygen species from steam, Journal of the Japan
Petroleum Institute 58 (2015) 329-335.

SC
[146] J. Cheng, T. Li, R. Huang, J. Zhou, K. Cen, Optimizing catalysis conditions to decrease
aromatic hydrocarbons and increase alkanes for improving jet biofuel quality, Bioresource
Technology 158 (2014) 378-382.
[147]

[148]
a Unit Operations Laboratory Course 2018.
U
E. Vasquez, Z. J. West, M. DeWitt, R. Wilkens, M. Elsass. Effective Teamwork Dynamics in

A.N. Kay Lup, F. Abnisa, W.M.A. Wan Daud, M.K. Aroua, Acidity, oxophilicity and
N
hydrogen sticking probability of supported metal catalysts for hydrodeoxygenation process,
IOP Conference Series: Materials Science and Engineering 334 (2018) 1-6.
A
[149] N. Shi, Q.-y. Liu, T. Jiang, T.-j. Wang, L.-l. Ma, Q. Zhang, X.-h. Zhang, Hydrodeoxygenation
of vegetable oils to liquid alkane fuels over Ni/HZSM-5 catalysts: Methyl hexadecanoate as
M

the model compound, Catalysis Communications 20 (2012) 80-84.


[150] D. Chiaramonti, M. Prussi, M. Buffi, D. Tacconi, Sustainable bio kerosene: Process routes and
industrial demonstration activities in aviation biofuels, Applied Energy 136 (2014) 767-774.
ED

[151] Development Status for Bio Jet Fuel. 2017, Asia Biomass Energy Cooperation: Immeuble
Kojima 2F, 3-13-2, Higashi-Ikebukuro, Toshima-ku, Tokyo 170-0013 JAPAN.
E PT
CC
A

60
Table 1
Typical physicochemical properties of vegetable oil.

Oxygen Viscosity at Density


Acid number HHV
Feedstock content 30-40°C (g/mL) at Ref.
(mg KOH/g) (MJ/kg)
(wt.%) (cP or mm2/s) (25°C)

Jatropha oil 22.05 27.2 49.40 0.92 37.61 [31]


Palm oil NM 11.5 48.71 0.92 NM [31]
[28]

T
Karanja oil NM 1.23 25.03 0.97 38.41
Sunflower oil 10.22 3.57 NM 0.88 39.52 [32]

IP
Cottonseed oil NM 0.30 58.20 0.92 39.41 [33]

R
Soybean oil NM 0.20 32.61 0.84 39.53 [34]
[35]

SC
Crambe oil NM 0.33 6.64 0.89 40.62
Mustard oil NM 0.55 73.01 0.92 39.01 [33]

Palm oil NM 4.80 48.00 0.90 39.01 [36]

Canola oil NM 1.80 87.20


U 0.91 38.82 [31]
N
Rapeseed oil NM 1.14 35.21 0.91 39.52 [37]
NM= Not Mentioned.
A
M
ED
E PT
CC
A

61
Table 2
Jet fuel specifications and propertiesa
Jet A- Jet A-
1 1
Composition Jet A Jet B JP-4 JP-5 JP-7 JP-8
(AST (IAT
M) A)
Acidity, total (mg 0.01 0.01 0.01
0.1 0.015 0.10 0.10 NM
KOH/g) 5 5 5
300
Final BP (°C) 300 300 300 300 270 – 288 300
330

T
Flashpoint (°C) 60–
38 38 38 38 68 60 38

IP
Min: 79
779
Density at 15°C 775– 775– 775– 775– 602– 775–
814 –

R
(kg/m3) 840 840 840 840 751 840
806
-

SC
Freezing point(°C) -47 -47 -40 - 47 -58 -50 -47
43.5
Viscosity at -20 to -
8.0 8.0 8.0 8.0 1.75 5.39 15 8.0
35.5°C (mm2/s)
Net heat of
combustion 42.8 42.8
43.2
42.8 U
42.8 43 43.5 42.8
N
8
(MJ/kg)
Smoke point (mm) 25 25 25 25 20 19 NM 25
A
Distillation (°C),
170 205 NM 205 145 240 182 152
M

Initial boiling point


18.5 20.3
Aromatics (wt. %) 18.0 25 8–25 25 1.5 13.5
3 6
ED

19.9 13.3
n-paraffins (wt.%) NM NM NM NM NM 19
8 5
Cycloparaffins (wt. 31.8 47.3
NM NM NM NM NM 24.1
%) 0 8
PT

iso-Paraffins (wt. 29.6 18.9


NM NM NM NM NM 38.2
%) 9 1
a
References of work: Jet A-1 (ASTM) [39,46], Jet A-1 (IATA) [39,46], Jet A [44,46,47], Jet B [13,46,48], JP-4
E

[46], JP-5[46,47,49], JP-7 [46] and JP-8 [39,45]


CC
A

62
Table 3
Specifications and properties of various alternative jet fuel derived from vegetable oilsa
Dat Hydroprocess
Compositio e Jatroph ed esters and Soybea Camelin Carinat Karanj Tallo
n pits a oil fatty acids n oil a oil a oil a oil w oil
oil (HEFA)
Net heat of
combustion 43.8 44.3 42.8 NM 42.9 43.2 NM 44.1
(MJ/kg)
Viscosity
(mm2/s) at - 3.71 3.66 8.0 4.6 7.4 6.5 25 5.3

T
20 °C
Density

IP
(kg/m3) at NM 751–840 730–770 780 818 802 970 758
15°C
Flash point

R
NM 46.5 38 44.0 48 46 NM 55
(°C)
Freezing

SC
point (°C), NM -57 -40 – -47 -50 -54 -57 NM -62
max
Final BP
NM 172–243 25 289 NM NM NM 255
(°C)
Acidity,
total (mg NM NM 0.015 0.010 U 0.011 0.012 NM NM
N
KOH/g)
Aromatics
A
NM 8.0 26.5 12.5 24.2 16.8 NM <0.4
(vol%)
Smoke
M

NM NM NM NM 22 26 NM >40
point (mm)
ED

a
References of work: Date pits oil [55], Jatropha oil [44], HEFA [53], Soybean oil [13],
Camelina oil [39], Carinata oil [39], Karanja oil [28] and Tallow oil [45].
E PT
CC
A

63
Table 4
The fatty acid composition of various TGs based oils sources [8,9,68,73].
Typical fatty acid composition in (wt.%)
Fatt Stru Was
Al
y ctur Ca Rub te
Pal Oli Pea Ra Soyb Sunfl Alm Co Jatro ga
acid e nol ber cook
m ve nut pe ean ower ond rn pha l
a seed ing
oil
oil
Capr N N N N
C10:0 NM NM NM NM NM NM NM NM NM

T
ic M M M M
Lauri 0. NM NM

IP
C12:0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
c 1
Myri 0. 2.67 0.22

R
C14:0 0.0 0.1 0.0 0.0 0.0 0.0 0.0 0.0 0.1 0.0
stic 7

SC
Palm 36 11. 10. 13.2 9.28
C16:0 8.0 4.9 11.3 6.2 10.4 6.5 15.9 5.1
itic .7 6 2 5
Palm 17.1 0.49
itilei C16:1
0.
1
1.0 0.0 0.0 0.1 0.1 0.5
U0.6 0.9 0.0 0.0 2
N
c
Stear 6. 0.45 3.95
A
C18:0 3.1 1.8 1.6 3.6 3.7 2.9 1.4 6.9 20.1 8.7
ic 6
46 75. 65. 24. 0.59 54.5
M

Oleic C18:1 53.3 33.0 24.9 25.2 77.1 41.1 57.9


.1 0 6 6 5
Linol 8. 25. 39. 0.43 29.6
C18:2 7.8 28.4 20.4 53.0 63.1 7.6 34.7 24.7
ED

eic 6 2 6 7
Linol 0. 16. 0.37 0.25
C18:3 0.6 0.3 7.9 6.1 0.2 0.8 0.1 0.3 7.9
enic 3 3
PT

Arac 0. NM NM
C20:0 0.3 0.9 0.0 0.3 0.3 0.3 0.1 0.0 0.2 0.0
hidic 4
Eico NM NM
E

0.
senoi C20:1 0.0 2.4 9.3 0.3 0.2 0.0 0.1 0.2 1.0 0.0
CC

2
c
Behe 0. NM NM
C22:0 0.1 3.0 0.0 0.0 0.7 0.1 0.0 0.0 0.2 0.0
nic 1
A

Eruci 0. NM NM
C22:1 0.0 0.0 23.0 0.3 0.1 0.0 0.1 0.0 0.2 0.0
c 0
Lign NM NM N NM NM
0.
oceri C24:0 0.0 1.8 0.0 0.1 0.2 0.2 0.1 M
1
c
Nerv C24:1 0. 0.5 0.0 0.0 0.0 0.0 0.4 0.0 NM NM N NM NM

64
onic
A 0

CC
EPT
ED

65
M
A
N
U
SC
M

RIP
T
Table 5
The world major oil producing countries.

Production in 2017/18
Feedstock Country Reference
(million tons)
Sunflower oil Ukraine 5110 [74]
Sunflower oil Russian Federation 3934 [74]
Sunflower oil European Union 2990 [74]
Sunflower oil Other countries 2190 [74]
Canola seed oil China 7 - 10 [74]
Canola seed oil Canada 7 - 10 [74]

T
Canola seed oil India 7 - 10 [74]
Canola seed oil Northern Europe 7 - 10 [74]

IP
Palm oil Malaysia 19.7 [75]
Palm oil Indonesia 29.3 [75]
Palm oil Thailand 2.3 [76]

R
SC
U
N
A
M
ED
E PT
CC
A

66
v Table 6
Isomerization selectivities of modified zeolites and reaction conditions of oils.

Reaction conditions
Isomerization
Material Catalyst Ref.
T P LHSV selectivity/yield
(°C) (bar) (h-1)

Soybean oil 0.5% Pt/Al2O3/SAPO-11 370-385 30.3 1.0 42–48 wt. % of jet fuel. [13]

C5–C14 and C15-C18 chain


Palm oil 2 wt.% Ni/ SAPO-11 200 40 1.0 [99]
were 18.7% and 70%.

T
C5–C14 and C15-C18 chain
Palm oil 7 wt.% Ni/ SAPO-11 200 40 1.0 [99]
were 67.4% and 61.5%.

IP
Coconut oil Pt/SAPO-11 360 45 1.01 77.6 % jet fuel. [77]

R
SC
U
N
A
M
ED
E PT
CC
A

67
Table 7
Catalyst properties used for upgrading vegetable oil to biofuels.

BET surface Pore volume Pore size Acid sites


Catalyst Types of catalysts Ref.
area (m2/g) (cm3/g) (A°) (mmol/g)

HZSM-5 419 0.28 NM 0.58 [28]


Ni/HZSM-5 386–451 0.36–0.44 39 0.27 [28]
SAPO-11 85.6 0.096 10.2 NM [108]
Ni/SAPO-11 169–224 0.10–0.14 20–23 NM [108]
HMCM-41 925 0.73 NM 0.99 [3]

T
Zeolite HY 620 0.40 NM 1.71 [3]

IP
Zeolite Pd/zeolite 27.3 0.0139 9.9 NM [44]
NiMoC/ZSM-5 446.8 0.13 3–90 NM [92]

R
NiMoC/zeolite β 466.7 0.09 3–90 NM [92]

SC
NiMoC/USY 475.6 0.25 3–90 NM [92]
Ni/Mesoporous-Y 518 NM 39 4.55 [111]
Ni/Mesoporous-HBeta 368 NM 37 2.58 [111]
NiMoC/Al-SBA-15
Al2O3
711.5
181-272 U
0.71
0.510
3–90
7.5
NM
NM
[92]
[14,78]
N
Co/ Al2O3 171–187 0.47–0.50 75 NM [112]
A
Ni/ Al2O3 148–222 0.48–0.63 75 0.14–0.21 [112]
Alumina Pd/ Al2O3 194–214 0.50–0.58 75 NM [112]
M

Pt/ Al2O3 196–217 0.51–0.58 75 NM [112]


NiMoC/γ-Al2O3 216.0 0.21 NM NM [92]
ED

Ni-MoS2/γ-Al2O3 156 0.306 7.0 NM [14]


E PT
CC
A

68
Table 8
An overview of the amount of jet fuel yields from TGs with different active-supported catalysts
under different reaction conditions.
Catalyst Catalyst P Conver
Materi Reactor LHSV Yield R
Active Supp preparation (ba T (°C) sion
al Type (h-1) (%) ef.
metal ort method r) (%)
Refined
γ-
palm Ni- Wetness Trickle- [1
Al2O 50 330 1 ∼92 C10–C12= 60
kernel MoS2 impregnation bade 4]
3
oil
Refined γ-
Ni- Trickle [1

T
palm Al2O Wetness 50 330 1 100 NM
MoS2 bade 4]
oil 3 impregnation

IP
Ni2P, C11-C12 =
14
(Co, Fe, Incipiently < 0.7 [1
Methyl Fixed (W
Mo and SiO2 impregnated 30 340 97.8 C6–C10 = 29
laurate bed HS

R
W) < 0.1 ]
V)
Soya-

SC
oil and
Ni– Al2O Fixed Kerosene = [7
refinery Sol–Gel method 50 380 2 > 99
W/SiO2 3 bed 21.8 0]
-oil

Soya-
oil and Ni–
Al2O Fixed U Kerosene = [7
N
refinery Mo/SiO Sol–Gel method 50 380 2 > 99
3 bed 3.0 0]
-oil 2
A

Phoeni
M

x
Activ
dactylif Continu [1
ated Incipiently Jet fuel =
era Pd ous 10 300oC NM 91.1 30
carbo impregnated 30.4
kernel Stirring ]
ED

n
oil

Bench-
C9–C14 =25–
top
38,
PT

micro-
22–38 and 40
Jatroph NiMo, reactor
Al2O Wet co- for NiMo, [1
a oil NiW, with 60 450 1 84
3 impregnation NiW and 2]
CoMo single
E

CoMo
zone
catalysts,
tubular
CC

respectively.
furnace
Continu
Coconu γ- [1
ous-flow 1-20
t oil Mo-Ni Al2O Co- 4-8 350 97 C8–C16 = 71 31
fixed
3, impregnation ]
A

bed
Refined
and Silic
Micro Kerosene = [1
unrefin ZSM-5 a, Wetness 40.95
fixed- NM 500 4 3.97 and 32
ed zeolite alum impregnation and 41
bed 14.54 ]
soybea ina
n oil

69
Catalyst Catalyst React P LHS Conve R
T Yield
Material Active Sup preparation or (b V (h- rsion ef
(°C) 1 (%)
metal port method Type ar) ) (%) .
C5–C14
Iso-
SAP alkane= [1
Fixed
Palm oil Ni O‐ Incipient wetness 40 500 2 79.3 20.3 and 0
‐bed
11 impregnation Normal 8]
alkane=
7.4

T
[1
Jatropha Ni ZS Incipient wetness Fixed C9-C15=

IP
30 653 3.8 70 3
oil Mo M-5 co-impregnation -bed 32.7 %
3]
SAP

R
O- Incipient wetness Fixed 94.4 C9-C15 = [1
Jatropha Ni
11, co-impregnation -bed 30 380 4 and 24.5 and 3

SC
oil Mo
Al2 100 12.3 4]
O3
γ- Micro 380 46 [1
Jatropha Ni- C6−C14=
Al2 Sol−Gel method chann 50 – (WH > 99 3
oil Mo
O3 el 440
U
SV)
44.5
5]
N
Crude
macauba
A
pulp oil
(CPO),
M

crude CPO= jet


almond o 23, fuel=CP
il (CAO activ CAO= O=96,
pall Fixed 10
), ated Impregnation 25, CAO= [8
ED

adiu bed - 300 5


macauba char method PFA= 97, 7]
m 19
pulp coal 50, PFA=10
fatty acid AFA= 0,
(PFA), 55 AFA=97
PT

almond
fatty acid
from
macauba,
E

(AFA)
γ- 380 53 [1
Jatropha Ni- Mono C6−C14=
CC

Al2 50 – (WH > 99 3


oil Mo Sol−Gel method lithic 36.2
O3 440 SV) 5]
Down
Coconut Ni Al2 NM flow 64.03- 57.85 [7
30 360 1-3
oil Mo O3 tubula 67.67 (Jet fuel) 7]
A

70
Table 9
Different countries studies and developed the bio-jet fuel for aviation industry [151].

Country/region Details
Europe The Alternative Fuels and Biofuels for Aircraft Development (ALFA-
BIRD) project performed studies and development on the quantity of
petroleum-derived fuel that could be replaced with FT synthetic oil,
hydrogenated vegetable oil, and so forth (2008 - 2012).

United States The Department of Energy (DOE) is performing research and


development on FT synthetic oil and fuel from cellulosic biomass for

T
military jets and ships in (APRIL 22, 2013).

IP
China The Chinese Academy of Sciences (CAS) and the Boeing Company
concluded a memorandum of understanding [59] in 2009, and have been

R
engaged in joint research on the manufacture of bio-fuels from algae in (
2010).

SC
Japan The New Energy and Industrial Technology Development Organization
(NEDO) has been conducting research on the manufacture of bio-jet fuel
derived from microalgae (2010 - 2011, 2013 - 2016).

U
N
Table 10
Bio-jet fuel derived from different oils and tested by different airlines [151].
A
Type of
Airline Feedstock Details
fuel
M

Air New Zealand Jatropha oil Tests were performed with flights Bio-jet fuel
equipped with bio-fuel in (December
2008).
ED

Japan Airlines Camelina oil The tests performed with flights using a Bio-jet fuel
fuel containing a 1:1 mixture of jetfuel
and bio jet-fuel in (January 2009).
PT

Thomson Airways Purified waste Used fuel containing a one to one mixture Bio-jet fuel
cooking oil of jet-fuel and bio-fuel on a commercial
flight (October 2011).
E

Lufthansa Airlines Jatropha and Used fuel with 1:1 mixture of jet-fuel and Bio-jet fuel
CC

Camelina oil bio-fuel used between Frankfurt and


Hamburg for six months starting from
(July 2011).
A

China Eastern Palm oil and A teste made through flights with bio-fuel Bio-jet fuel
Airlines waste cooking by SINOPEC in (April 2013).
oil
KLM Royal Dutch Purified waste Flew bio-jet fuel flights from Amsterdam Bio-jet fuel
Airlines cooking oil to Paris (September 2011) and
Amsterdam to New York flight (March
2013).

71

You might also like