Vous êtes sur la page 1sur 12

Journal of Industrial and Engineering Chemistry 36 (2016) 1–12

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


journal homepage: www.elsevier.com/locate/jiec

Review

Development of biolubricants from vegetable oils via chemical


modification
Josiah McNutt, Quan (Sophia) He *
Department of Engineering, Faculty of Agriculture, Dalhousie University, Truro, Nova Scotia, Canada B2N 5E3

A R T I C L E I N F O A B S T R A C T

Article history: In response to the increasing environmental pollution concern and depleting petroleum reserves, bio-
Received 23 November 2015 based lubricants have received a great deal of interest as a substitute for mineral oil-based lubricants.
Received in revised form 7 February 2016 Biolubricants have a number of advantages over mineral lubricants, including the high biodegradability,
Accepted 12 February 2016
low toxicity, excellent lubrication performance, and minimal impact on human health/environment.
Available online 21 February 2016
This paper reviewed the most recent advancements in the synthesis of biolubricants from vegetable oils
through chemical modification methods such as esterification/transesterification, estolide formation,
Keywords:
and epoxidation of vegetable oils.
Biolubricant
Vegetable oils
ß 2016 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
Chemical modification reserved.
Esterification
Estolide
Epoxidation
Oxidative stability
Viscosity

Contents

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Feedstock considerations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
Performance requirements of biolubricants . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
Modification methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
Esterification/transesterification. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
Estolide formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Epoxidation, ring opening and acetylation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
Other efforts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

Introduction the use of vegetable oil-based lubricants in place of the commonly


used petroleum-based lubricants. These products, known as
With concern over global climate change and depleting ‘‘biolubricants’’, carry several environmental, health, and perfor-
petroleum reserves growing, the search for environmentally mance benefits over current petroleum-based lubricants.
sustainable alternatives to current practices continues to intensify It is estimated that 20% of the 5.2 million tons of lubricant
[1]. One area of interest which could serve to reduce both reliance consumed every year in Europe is released into the environment,
on petroleum and anthropogenic impact on the environment, is and a kilogram of said mineral oil is capable of polluting a million
liters of water [2]. Therefore, pollution caused by lubricants is far
from insignificant. The petroleum-based lubricants can also
* Corresponding author. Tel.: +1 9028936180; fax: +1 9028901859. contaminate soil directly, and pollute the air due to its volatility
E-mail address: quan.he@dal.ca (Q.(. He). [3]. This pollution is hazardous to not only plants and animals

http://dx.doi.org/10.1016/j.jiec.2016.02.008
1226-086X/ß 2016 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
2 J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12

inhabiting the contaminated areas, but potentially human trees, cannot be used for edible purposes due to the presence of
residents as well [4]. Several studies have documented the harmful glycoside. However, the tree can grow in a wide range of pH values,
effects of petroleum based lubricants on human health. Chronic meaning it could be produced on land that is largely unproductive
inhalation or dermal exposure to petroleum-based lubricants can [20]. Non-edible crop oils such as jatropha, linseed, karanja, neem,
have inflammatory effects on the respiratory system and location castor, coriander, cuphea, rice bran, milkweed, and many others
of contact, while also being carcinogenic [5–7]. These negative have little to no impact on world food prices or production, as they
effects are even more severe in the used petroleum-based oils, as can be grown on nutrient-deficient land and do not compete with
degradation leads to increased toxicity [8]. Many biolubricants existing agricultural resources [2,19,21]. However, harvesting,
however, are rapidly biodegradable and nontoxic, and therefore cultivation, and processing of these crops are often a challenge, as
pose little or no risk to the environment or operators [4,9,10]. many of these crops had little incentive to be purposely produced
Biolubricants also boast several performance benefits, including until recently [21].
better lubricity, higher flash point, lower volatility, higher viscosity Genetic modification of oil-bearing crops is also a topic to
indices, higher shear stability, lower compressibility, higher consider when examining feedstocks. A majority of the genetic
detergency, higher resistance to humidity, and higher dispersancy research in the field of lubricants involves the creation of high-oleic
[4,9,11–16]. varieties of oil seed crops. As mentioned above, oleic acid is one of
Despite these advantages, biolubricants are still not widely the most favorable fatty acids in biolubricant production due to its
used due to several major challenges and difficulties regarding its good balance of low temperature properties and high temperature
performance and production. Aside from issues regarding feed- thermo-oxidative stability. Sunflower, palm, camelina, rapeseed,
stock reliability and consistency as well as industry acceptance, and soybean all have genetically modified varieties which alter the
biolubricants also have two main negative physical properties: composition of the seed oils [22–26].
poor low temperature performance, and low thermal oxidative Another feedstock which is being investigated is waste cooking
stability [3,9,10,12,17,18]. However, through appropriate chemical oil. One of the biggest barriers for biolubricant production is the
modification processes, these two properties can be improved to high cost of feedstock, which can account for 70–80% of the total
make biolubricants a feasible alternative to mineral lubricants for production cost [27]. Waste cooking oil is significantly cheaper
various applications. Much research has been done in previous than unused edible vegetable oils, and has therefore been
years on the exploration of new feedstocks and modification examined as a possible alternative feedstock [27–30]. Waste
methods, development of more efficient catalysts for modification cooking oil is generally 30–60% cheaper than regular vegetable oil,
of vegetable oils, and optimization of modification. However, there which makes it a potentially promising candidate for profitable
is need for a review paper on this topic. This paper aims to biolubricant production [27].
summarize the most current research focused on chemical
modification methods and their respective advantages and Performance requirements of biolubricants
disadvantages. The paper will also provide information on
common feedstocks and the required properties of biolubricants, The main function of a lubricant is to reduce the friction
as well as the relevant testing methods. It is noteworthy that the between contacting surfaces. Lubricants are used in a variety of
focus is on the development of base oil for biolubricants from industries such as agriculture, forestry, mining, automobile, and
vegetable oils, therefore bio-based additive development is beyond fishing, serving as engine oils, chainsaw oils, transmission oils, and
the scope of this paper. hydraulic oils. With different applications, lubricants may have
specifically required characteristics in terms of viscosity, chemical
Feedstock considerations stability, fluidity, flammability, range of working temperature and
water solubility. Currently, there are not a wide array of
The most commonly used feedstock for developing biolubri- specifications for biolubricants, but generally the properties of
cants is vegetable oils. Molecularly, vegetable oils are triglycerides, biolubricants must be comparable to those of mineral oil based
esters of glycerol and three straight chained fatty acids. The chain lubricants as regulated in USA or European standards. Table 1
length of the fatty acids is usually in the range of C12–C24. The shows some lubricant specifications, as well as some unmodified
three fatty acids vary between feedstocks, and play an important vegetable oil properties. One of the most popular sources for
role in determining the properties of the oil. The two main evaluation methods of lubricants is ASTM International, formerly
variables among fatty acids are the number of double bonds and known as the American Society for Testing and Materials. Their
the chain length. In general, a longer chain length results in a website contains over 1200 different testing methods relevant to
higher melting point and viscosity, and more double bonds lubricants, demonstrating how broad the use of lubricants is. Some
correspond to lower melting points, decreased viscosity, and of the most common ASTM tests include D97, D445, D2270, D2500,
decreased thermo-oxidative stability [2]. Monounsaturated fatty and D4172 which measure the most important properties such as
acids, such as oleic and palmitoleic acid, have been found to have a pour point, viscosity, viscosity index, cloud point, and anti-wear
good balance of low melting point, with good thermo-oxidative characteristics, respectively [38–44].
stability and viscosity [2,3,17]. It is for this reason feedstocks with
high oleic or palmitoleic acid contents are generally preferred and Modification methods
sought after.
Vegetable oil can be extracted from the over 350 different crops Most currently available vegetable oils cannot be used as
with oil-bearing seeds throughout the world [1,19]. Popular lubricants directly due to poor low temperature performance and
feedstocks include palm, canola, soybean, sunflower, coconut, low oxidative and thermal stability. There are a number of methods
safflower, rapeseed, cottonseed, jatropha, karanja, castor, lesquer- to improve these undesired properties, such as genetic modification
ella, pennycress, and peanut oils, with many others being tested for of fatty acid profile of vegetable oils, direct addition of antioxidants,
potential use. While both edible and non-edible crops are currently viscosity modifiers, and pour point depressant to vegetable oils,
being researched, non-edible crops are more desirable for several emulsification of vegetable oils, and chemical modification of
reasons. Non-edible vegetable oils are often derived from plants vegetable oils [3,9,12]. Among these methods, chemical modification
that are not in direct competition with cultivation of edible oil is the most promising one with great potential to improve chemical
crops. For example, rubber seed oil, which is collected from rubber and broad temperature range stability. Chemical modification
J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12 3

Table 1
Lubricant requirements and unmodified vegetable oil properties.

Lubricant Viscosity Viscosity Viscosity Pour Flash Oxidative Coefficient of Wear scar Ref.
requirement 40 8C (cSt) 100 8C (cSt) index point (8C) point (8C) stability (min) friction (mm)

ISO VG32 >28.8 >4.1 >90 6 204 – – – [27]


ISO VG46 >41.4 >4.1 >90 6 220 – – – [27]
ISO VG68 >61.4 >4.1 >198 6 226 – – – [27]
ISO VG100 >90.0 >4.1 >216 6 246 1670.26 – – [27,31]
Paraffin VG95 95 10 102 – – – – – [13]
Paraffin VG460 461 31 97 – – – – – [13]
R150 150.04 – – – 195 931.16 – – [31]
SAE20W40 105 13.9 132 21 200 – 0.117 0.549 [4]
AG100 216 19.6 103 18 244 – – – [32]
75W-90 120 15.9 140 48 205 – – – [33]
75W-140 175 24.7 174 54 228 [34]
80W-140 310 31.2 139 36 210 – – – [33]

Vegetable oil
Soybean 28.86 7.55 246 9 325 – – – [15]
Sunflower 40.05 8.65 206 12 252 – – – [15]
Passion Fruit 31.78 – – – 228 7.5 – – [31]
Moringa 44.88 – – – 204 28.27 – – [31]
Castor 220.6 19.72 220 27.00 250 – – – [35]
Rapeseed 45.60 10.07 180 12.00 240 – – – [35]
Jatropha 35.4 7.9 205 6 186 5 – – [36]
Coconut 24.8 5.5 169 21 325 – 0.101 0.601 [4]
Rice bran 40.6 8.7 169 13 318 – 0.073 0.585 [4]
Palm 52.4 10.2 186 5 – – – – [37]
Lesquerella 119.8 14.7 125 21 – – 0.045 0.857 [16]
Pennycress 40.0 9.3 226 21 – – 0.054 0.769 [16]

mainly revolves around modifying the acyl (C5 5O) and alkoxy (O–R) of vegetable oils to obtain lubricants is a multistep process. First,
functional groups and double bonds present in the oil. As shown in unmodified triglycerides are reacted with a short-chain alcohol
Fig. 1, one way is to rearrange the acyl moieties to form new triesters (methanol) with a base catalyst to produce fatty acid methyl esters.
from triglyceride through esterification/transesterification; the The resulting methyl esters are then reacted with various types of
second way is to modify acyl group by the formation of estolides alcohol, in the presence of an acidic or basic catalyst to yield
after the hydrolysis of triglyceride to give a variety of branched esters, triesters. This process is one of the most popular in the literature,
and the third path is to modify double bonds by epoxidation and due to the wide array of reactants that can be used, resulting in
subsequent ring opening to give place to a versatile intermediate for biolubricants with varying properties, thus broader applications.
the synthesis of different diesters. These three modification methods Generally, the transesterification of methyl esters leads to
are comprehensively reviewed in the following sections, summariz- decreased pour points and increased thermo-oxidative stability,
ing the most recent developments and advancements in this field. while maintaining the beneficial viscosity and lubricity character-
istics of the base oils [20,25,37,45,46]. The reaction of methyl
Esterification/transesterification esters (ME) and trimethylolpropane (TMP) with an alkaline or
enzyme catalyst is the most common approach. Much work has
Esterification and transesterification are commonly used to been done in recent years to improve the efficiency of this process,
arrange acyl moieties in vegetable oils to form new esters with and the properties of the lubricant, by examining different aspects
improved physical properties. Esterification or transesterification of the reaction. Vegetable oil methyl esters, waste cooking oil,

Fig. 1. Reaction pathways of the three main chemical modification methods. Red, green and blue solid arrows represent the pathway of esterification/transesterification,
estolide formation, and epoxidation/ring opening, respectively. (For interpretation of the references to color in figure legend, the reader is referred to the web version of the
article.)
4 J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12

biodiesel, estolides, and ring opened products, can all be esterified effect of different alcohols such as 2-ethylhexanol (2-EP), neopentyl
or transesterified. This section will focus on vegetable oil methyl glycol (NPG) and trimethylolpropane (TMP) on lubricant properties
esters, waste cooking oil methyl esters, and biodiesel, as the of rapeseed derived biolubricants in a study done by Gryglewicz
esterification of estolides and ring opened products will be et al. [25]. It was observed that using 2-EP led to an ester product
discussed in Sections ‘Estolide formation’ and ‘Epoxidation, ring with a better low temperature property, the pour point of 31.3 8C
opening and acetylation’, respectively. compared to the products resulting from employing NPG and TMP.
Table 2 outlines the conversion of a number of methyl esters In addition, transesterification of FAME with NPG and TMP required
derived from vegetable oils to triesters. Jatropha oil has recently a longer reaction time.
received significant attention as a feedstock for developing
lubricants. Using sodium methoxide, the oxidative stability of Estolide formation
the resulting TMP triesters has been improved significantly, and
the degradation temperature was higher than 325 8C as demon- Estolides are formed by the bonding of a fatty acid’s carboxylic
strated in studies [37,45]. Palm oil based triester was synthesized acid functionality to the double bond of another fatty acid [12].
under the temperature range of 120–150 8C with sodium Fatty acids are usually obtained from the base triglyceride through
methoxide as catalyst for 45 min, and the pour point of obtained hydrolysis, and the resulting fatty acids react in the presence of
product was as low as 37 8C [23]. Heterogeneous catalyst calcium hydrogen ions to produce estolides. Lubricants based on estolides
methoxide [46] was also used to synthesize palm TMP triesters, can provide improved lubricity, improved oxidation stability, and
and obviously a longer reaction time of 8 h was required compared decreased pour points [58–60]. Another benefit of estolides is the
to 45 min using homogenous catalyst [23]. Koh et al. attempted reaction temperatures in which they are produced are reasonably
introducing a novel oscillatory flow reactor for the synthesis of low (<100 8C), and therefore require less energy.
palm biolubricant from palm methyl esters and trimethylolpro- Estolides can be formed from a variety of different base oils.
pane. The final product had improved thermal and oxidative Franco’s group has been dedicated to the synthesis and property
stability, a lower pour point, and comparable properties to other characterization of lubricants from vegetable oils. They synthe-
biolubricants, and the conversion time was less than half of other sized a variety of very high viscosity estolides from high-oleic
reported reaction times [47]. Biodiesel, another group of alterna- sunflower oil, olive pomace acid oil, oleic acid, and ricinoleic acid
tive feedstocks, is a mixture of fatty acid methyl/ethyl esters, which [58–61]. They also examined the difference the choice of catalyst
was studied to synthesize TMP-based triesters [48–50]. As shown had on the physically properties of the estolides, finding that
in Table 2, the oxidative stabilities of the finished triesters were sulfuric acid provided higher viscosities than perchloric acid or
satisfactory. Similarly, studies have also been done using waste p-Toluensulphonic acid, but also increased the frictional coefficient
cooking oil (WCO) methyl esters to produce TMP triesters leading to increased wear [58,59]. The results are summarized in
[27,29,30]. Waste cooking oil is a good source of methyl esters Table 4. The formation of estolides from pennycress, composed
for making TMP triesters-based lubricants. The major drawback mostly of erucic and linoleic acid, by Cermak’s team yielded
is an additional purification step required prior to performing biolubricant with remarkably high viscosities [62]. Cermak et al.
esterification and/or transesterification reactions, due to the low [64] also synthesized estolides from lesquerella and castor fatty
quality of WCO compared to methyl esters derived from the virgin acid esters with differing degrees of saturation, and multiple
vegetable oils. The properties of the resulting TMP trimesters capping fatty acids. The study found that the estolides with the
presented the satisfactory lubricating properties. best low temperature performance were those formed with oleic
Recently, improved metal-based heterogeneous catalysts (sul- acid or 2-ethylhexanoic acid as a capping material, and that
fated zirconia complexes [51] and Fe–Zn double-metal cyanide saturation of the estolide resulted in higher pour and cloud points.
complexes [52]) as well as solid acid catalysts (Indion-130 [53] and Formation of estolides from esters was also economically
silica-sulfuric acid [54]) have been examined in order to improve beneficial, as it required no catalyst or solvent.
efficiency by reducing waste. The recyclability of catalysts was Unmodified estolides are not always used on their own; often
evaluated in these studies. they are mixed with vegetable oils, esterified, or modified in some
Enzymatic catalyst has advantages over chemical catalysts such other way to further enhance their properties [59,63–65]. Cermak
as high selectivity and low reaction temperature required. The et al. also conducted research on the esterification of estolides,
application of microfluidic reactor, an emerging novel continuous which are presented in Table 5. Petroselinic acid found in coriander
reactor was also explored in the chemical modification of was combined with various capping fatty acids, and the resulting
vegetable oil for lubricant production. Enzymatic reactions estolide was then esterified with 2-ethylhexyl alcohol, to yield
performed in microchannel reactors represents an exciting and estolide 2-EH esters. These petroselinic based estolide esters were
attractive field as the continuous packed bed microreactors are found to be comparable to their oleic based counterparts,
able to provide the extended lifetime of the bio-catalyst, simple possessing good low temperature properties, and viscosities
enzyme reuse and relatively easy product recovery. Madarasz et al. [66]. In a separate study, estolides derived from oleic acid were
[55] found that the oleic fatty acid and isoamyl alcohol could be esterified using a variety of linear and branched alcohols to
reacted with Novozym 435 as an enzymatic catalyst in an evaluate the effect it had on physical properties. Esterification with
H-CubeTM microfluidic reactor continuously for 144 h without branched alcohols yielded significant improvements in the low
any loss in catalytic from the enzyme. Happe et al. [56] created a temperature properties of currently available commercial pro-
novel microwave barrel reactor for the use in lipase catalyzed ducts, and as expected the longer the chain length, the higher the
biolubricant synthesis, which reduced solvent use as well as viscosity [65]. Esterification of estolides derived from pennycress
energy consumption significantly. Immobilized lipase was also was also performed. This esterification improved the low
tested in biolubricant syntheses [57]. Unfortunately, the properties temperature properties, and had much better viscosity properties
of the products derived from the employment of enzymatic than any other estolide esters synthesized [62]. Another method of
catalysts were not tested or not reported. altering estolides, is by sulfur modification. Biresaw et al. created a
Development of other alcohol esters rather than TMP triesters sulfur modified castor 2-EH ester estolide, by first synthesizing an
was exploited for biolubricant applications [20,25,32,51,52]. Re- estolide from castor 2-EH esters and capping lauric acid, and then
search has also been done on how the type of alcohol used impacts reacting the resultant estolide with butanethiol in a photochemical
the physical properties of the lubricant. Table 3 summarizes the reactor. The sulfur modified castor 2-EH ester estolide showed a
Table 2
Physical properties and reaction conditions of lubricants derived from fatty acid methyl esters.

Reactants Product Catalyst Reaction Viscosity Viscosity Viscosity Pour Oxidative/Thermal Yield (%) Ref.
conditions 40 8C (cSt) 100 8C (cSt) index point (8C) stability

Jatropha and TMP TMP triesters Sodium 150 8C, 10 mbar, 43.90 8.71 180 6 – >80 [37]
methoxide 3h
Jatropha ME and TMP triesters Sodium 150 8C, 55 min 42.57 9.37 183 6 325 8C – [45]
TMP methoxide Degradation temp
High oleic palm ME TMP triesters Sodium 120–150 8C, 45.5–50.7 9.2–10 183–200 ( 37) to ( 9) – – [23]
and TMP methoxide 0.3 mbar, 45 min
Palm ME and TMP TMP triesters Sodium 140 8C, 25 mbar, 47.1 9.0 176 2 355 8C 94.6 [46]
methoxide 25 min, Degradation temp

J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12


oscillatory flow
reactor at 1.5 Hz
with 20 mm
amplitude
Palm ME and TMP TMP triesters Calcium 180 8C, 50 mbar, – – – – – 92.38 [47]
methoxide 8h
Canola biodiesel TMP triester Sodium 110 8C, 1 mbar, 40.5 7.8 204 66 Induction time: 90.9 [49]
ME and TMP methoxide 5h 0.74 h
Castor biodiesel TMP triester Dibutyltin 170 8C, 0.01 bar 287.2 26.13 119 27 RPVOT: 43 min 89.7 [50]
and TMP dilaurate (Butylated
hydroxytoluene
added)
Castor biodiesel TMP triester Amberlyst 15 120 8C, 0.01 bar 20.94 4.467 127 – – – [50]
and TMP ionic exchange
resin
Castor biodiesel TMP triester Sodium 120 8C, 0.01 bar 11.28 3.100 141 – RPVOT: 150 min – [50]
and TMP methoxide (Butylated
hydroxytoluene
added)
WCO ME and TMP TMP triester KOH 128 8C, 200 Pa, 38.60 8.44 204 8 FP: 240 8C 85.7 [27]
1.5 h
Soybean oil and n-alcohol-esters Sulfated zirconia 140 8C, 4 h 10.3–432.7 3.0–34.4 45–195 – – >80 [51]
various alcohols catalysts
Sunflower oil and FA-n-octyl esters Fe–Zn double- 170 8C, 8 h 7.93 2.74 226 3 23 min (RBOT) 98 [52]
octanol metal cyanide
(DMC) complexes
Pentaerythritol and Pentaerythritol Ion exchange 110 8C, 6 h 63.08 12.00 190 24 – – [53]
oleic acid tetraoleate ester resin, Indion-130 toluene solvent
Valeric acid TMP Valeric acid TMP Silica–sulphuric 70 8C molar ratio 9.5 2.5 80 75 – – [54]
ester acid of 3:1, toluene
Rubber ME and NPG/TMP/PE p- 135–140 8C, until 23.1–62.6 5.9–12.6 206–222 ( 15) to ( 3) 10–15 min (RBOT) 94.5–96.5 [20]
NPG/TMP/PE triesters Toluensulphonic theoretical FP: 266–308 8C
acid reaction
complete
Rapeseed ME and NPG/TMP/PE C Antarctica 150, 200, 50 h 7.8–38.2 2.7–8.4 205–224 ( 31.3) to ( 18) Dv: 90.1–147.1 98 [25]
NPG/TMP/PE triesters lipase DAc: 2.9–7.7
Thumba ME, xylene NPG/TMP/PE p- 135–140 8C, until 20.65–60.26 5.45–11.89 209–220 ( 12) to ( 3) 10–15 min (RBOT) 89–95 [32]
and NPG/TMP/PE triesters Toluensulphonic complete FP: 270–318 8C
acid

Note: RPVOT, rotating pressure vessel oxidation test; RBOT, Rotary Bomb Oxidation Test; RSSOT, Rapid Small Scale Oxidation Test; FP, flash point.

5
6 J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12

Table 3
Physical properties of rapeseed esters using various alcohols.

Alcohol used Viscosity 40 8C (cSt) Viscosity 100 8C (cSt) Viscosity index Pour point (8C) Dv (%) DAc
2-EH 7.8 2.7 224 31.3 147.1 7.7
NPG 17.4 4.7 209 19.5 136.5 6.0
TMP 38.2 8.4 205 18.0 90.1 2.9
None (original oil) 35.1 7.9 207 19.1 145.7 7.5

Note: Dv denotes change in viscosity at 40 8C after 12 L/h of air bubbles were passed through the lubricant for 24 h at 100 8C. DAc denotes change in acid number at 40 8C after
12 L/h of air bubbles were passed through the lubricant for 24 h at 100 8C.

much improved oxidative stability over the unmodified estolide, to high quality biolubricants. Early work used acidic ion exchange
and a decreased viscosity index [63]. resin catalyst in the epoxidation of canola oil with hydrogen
peroxide [72]. Recently, a number of high performance catalysts
Epoxidation, ring opening and acetylation were developed and applied in the synthesis of canola oil-based
lubricants. For example, Somidi et al. [68] found that sulfated-SnO2
Epoxidation involves the removal of double bonds between two catalyst provided very good catalytic activity and had a 100%
carbons via an atom of oxygen, which results in an epoxide conversion rate of canola oil to epoxidized canola oil within 6 h,
functional group, a three atom ring composed of two carbon atoms and the induction time of the resulting epoxidized canola oil was
and an oxygen atom. This reaction usually involves a base olefinic 60 h. Catalyst Amberlite IR-120H [75,76] was also tested and
material, reacted with hydrogen peroxide, in the presence of evaluated in epoxidation of canola oil or canola biodiesel. Sharma
formic or acetic acid, and sometimes involving the use of various et al. found that epoxidation followed by simultaneous ring opening
enzymatic or heterogeneous catalysts, such as Amberlite IR-120H, and esterification of canola oil and canola biodiesel made viable
sulfuric acid, sulfated-SnO2 catalyst, Novozym 435, or peracetic biolubricants. Modified canola biodiesel had good low temperature
acid [31,36,67–73]. Epoxidation of vegetable oils generally results properties, viscosity, anti-wear properties, and oxidative stability,
in increased oxidative stability, better acidity value, increased making it a promising candidate for use in general automotive
adsorption to metal surfaces which results in better lubricity, applications. Modified canola oil was highly viscous, and worked
increased viscosity, decreased viscosity index, and increased pour extremely well at high temperatures which gave it a promising
point [26,31,36,67–70,74]. Some physical characteristics and outlook for use in heavy machinery. In the study on the ring opening
reaction conditions of the epoxidation of vegetable oils and their and esterification of epoxidized canola oil, another novel sulfated Ti-
derived fatty acids are shown in Table 6. As can be seen, this SBA-15 catalyst demonstrated higher activity, selectivity, stability,
modification is generally favorable due to the low temperature and reusability than the commonly used catalysts such as
required for the reaction, however the pour points of the products Amberlyst-15, IRA-200 and IRA-400 [81]. In the research carried
are unsatisfactory for many applications. out by Ahn et al. magnesium stearate was an efficient catalyst for
As the epoxidized vegetable oil still has poor low temperature solvent-free ring opening of epoxidized methyl oleate, and had a
properties, it is essential to further modify it. The epoxidized superior efficiency when compared to many other catalysts. This
product is often subjected to a combination of oxirane ring- process increased the viscosity, and was also less wasteful than most
opening, esterification, and/or acetylation. After the vegetable oil is other processes, as it was solvent free [78].
epoxidized, the first modification step performed is oxirane ring Along with improved catalysts, alternative basestocks have also
opening, then esterification, or a reaction that does both been studied. Sankaranarayanan and Srinivasan [89] experimented
simultaneously. Because the ring opening and esterification with ricinoleic derivatives, as opposed to the more popular oleic
reactions of the epoxidized vegetable oils can be performed with derivatives, and found that they were highly tunable and showed
an array of alcohols and other materials, the resulting ring opened promise as for a range of industrial lubricants. The study performed
products and esters can have differing beneficial properties. This by Salih et al. [35] found that longer midchain esters were good for
process is sometimes followed by the acetylation of the produced low temperature operability and anti-wear properties, but resulted
ester. Numerous studies have shown that ring opening, esterifica- in worse thermal stability than shorter midchain esters. Kamalakar
tion and/or acetylation result in improved viscosity index, better et al. [32] observed that epoxidized thumba oil could be used as
low temperature flow properties, increased thermal and oxidative aviation grade lubricant, while esterification of the epoxy oil
stability, lower coefficients of friction, and better lubricity derivatives were suitable for hydraulic and metal working applica-
characteristics [28,35,70,75,76,78–86]. Table 7 summarizes the tions. Hashem et al. [87] produced epoxidized castor, linseed,
properties of some of these modified epoxidized vegetable oils as sunflower, and jatropha oil and then esterified them with oleic acid,
well as modification conditions. which produced several promising lubricants. Li and Wang [28]
Much attention has been given in recent years to further epoxidized waste cooking oil, formed methyl esters, and subse-
improving the efficiency and properties of the products of quently esterified those methyl esters with branched alcohols,
epoxidation of vegetable oils, ring opening reactions, esterification, resulting in increased viscosity index and improved pour points.
and acetylation of the epoxidized vegetable oils. This is being done Aside from improved catalysts or feedstocks, research has also
by finding new catalysts, improving existing processes, and been done on modifying the traditional conversion methods.
experimenting with new feedstocks. For example, Sammaiah Hwang and Erhan found that epoxidized soybean oil reacted with
et al. [36] employed a mineral acid, H2SO4, to accelerate the Guerbet alcohols gave ring opened products which were either
epoxidation of Jatropha oil, and the product yield was 96%. H2SO4 transesterified or not depending on the amount of alcohol used.
was also widely used in the modification of epoxidized vegetable The ring opened products which were simultaneously transester-
oils such as soybeans [80], mustard oil [83] and oleic acid [84]. ified had lower pour points and viscosities, and higher viscosity
However, the mineral acid catalyst is generally not favorable due to indices [80]. Lee et al. [90] experimented with the amidation of
the potential corrosion to reactor vessels as well as waste stream epoxidized vegetable oil derivatives, and successfully formulated
handling problems. Dr. Dalai’s group has conducted considerable cross-linking products, but gave no information on the physical
research on developing advanced catalysts for modifying canola oil properties of these materials.
Table 4
Physical properties and reaction conditions of assorted estolides.

Reactants Products Catalyst Reaction Viscosity Viscosity Viscosity Pour Point (8C) Oxidative/Thermal Yield (%) Ref.
Conditions 40 8C (cSt) 100 8C (cSt) Index stability

Sunflower oil Estolide H2SO4, HClO4, or 50–100 8C 102.4–425.3 17.2–42.7 153–185 – – – [58]

J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12


p-Toluensulphonic depending on
acid catalyst, 3–24 h
Olive oil Estolide H2SO4 100 8C, 3–24 h 271.8–518.6 33.5–60.2 168–188 – – – [58]
Sunflower oil Estolide H2SO4 50 8C, 3–24 h 278.8–430.8 50.1–35.3 – – – – [60]
Oleic acid Estolide H2SO4, HClO4, or 50–100 8C 75.61–415.59 9.51–17.02 – – – – [61]
p-Toluensulphonic depending on
acid catalyst
Ricinoleic acid Estolide H2SO4, HClO4, or 50–100 8C 581.56–6712.98 51.26–232.62 – – – – [61]
p-Toluensulphonic depending on
acid catalyst
Pennycress FA FFA estolide HClO4 60 8C, 7.5–10.9 kPa, 494.4–497.1 42.2–75.3 134–163 ( 15) to ( 6) – – [62]
24 h
Castor 2-EH ester Estolide Tin (II) 2- 130 8C, 12–18 Pa, 51.4 9.9 183 < 54 RPVOT 16 min 73 [63]
and lauric acid ethylhexanoate 24 h
Castor 2-EH ester Estolide – ( 28) to ( 18)8C, 56.0 10.6 144 < 54 RPVOT 224 min 96 [63]
estolide and photochemical
butanethiol reactor, 3 h
Saturated Saturated – 200 8C, 20 Pa, 24 h 37.0–45.7 7.9–9.1 187–196 ( 12) to 6 – – [64]
lesquerella FA estolide
ester and capping
FA
Unsaturated Unsaturated – 200 8C, 20 Pa, 24 h 35.4–51.1 7.8–10.1 189–200 ( 54) to 3 – – [64]
lesquerella FA estolide
ester and capping
FA
Saturated castor FA Saturated – 200 8C, 20 Pa, 24 h 43.6–68.3 8.7–12.2 178–186 ( 36) to 6 RBOT: 403 min – [64]
ester and capping estolide with 3.5% of an
FA Lubrizol oxidative
stability package
Unsaturated castor Unsaturated – 200 8C, 20 Pa, 24 h 29.0–70.6 6.5–11.8 164–196 (< 54) to 23 RBOT: 159 min – [64]
FA ester and estolide with 3.5% of an
capping FA Lubrizol oxidative
stability package

7
8 J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12

Table 5
Physical properties of assorted estolide esters.

Reactants Products Catalyst Reaction Viscosity Viscosity Viscosity Pour point (8C) Oxidative/thermal Yield (%) Ref.
conditions 40 8C (cSt) 100 8C (cSt) index stability

Oleic acid estolide Estolide ester BF3 60–80 8C, until 55.2–108.9 10.2–15.3 163–184 ( 33) to ( 9) – – [65]
and linear 99% complete
alcohols
Oleic acid estolide Estolide ester BF3 60–80 8C, until 62.5–209.3 11.1–24.9 149–192 ( 39) to ( 24) – – [65]
and branched 99% complete
alcohols
Coriander FA and 2- Estolide ester HClO4 Estolide: 60 8C, 53.7–92 9.1–14.6 151–165 ( 33) to ( 12) RPVOT: 16– 65–76 [66]
EH, capped with 7.5–10.9 kPa, 24 h 273 min depending
various FA Ester: additional on how much
3–4 h after 2-EH Lubrizol1 7652
added additive
Castor 2-EH ester Sulfide modified – ( 28)-( 18)8C, 56.0 10.6 144 < 54 RPVOT: 224 min 96 [63]
estolide and estolide photochemical
butanethiol reactor, 3 h
Pennycress estolide Estolide ester BF3 80 8C, 7.5– 116.3–245.8 18.2–33.6 169–183 ( 24) to ( 12) – – [62]
and 2-EH, capped 10.9 kPa, 8 h,
with various FA

Table 6
Physical properties and reaction conditions of the epoxidation of vegetable oils and their derived fatty acids.

Reactants Products Catalyst Reaction Viscosity Viscosity Viscosity Pour Oxidative/thermal Yield (%) Ref.
conditions 40 8C (cSt) 100 8C (cSt) index point (8C) stability

Oleic acid, formic Epoxidized oleic – 4 8C, 2 h – – 45.44 0 FP: 113.11 8C – [35]
acid, and H2O2 acid
Thumba oil, formic Epoxidized thumba – 5–10 8C, before 216.9 22.7 128 3 RBOT: 20 min – [32]
acid, and H2O2 oil adding H2O2 and
heating to 60 8C
for 7 h
Passion fruit oil, Epoxidized passion – 30 8C, 3 h 185.65 – – – RSSOT: 16.89 min – [31]
formic acid, and fruit oil
H2O2
Moringa oil, formic Epoxidized – 30 8C, 3 h 80.37 – – – RSSOT: 24.57 min – [31]
acid, and H2O2 moringa oil
Jatropha oil, formic Epoxidized H2SO4 10 8C for 2 h while 146.5 18.2 139 0 RBOT: 20 min 96 [36]
acid, and H2O2 jatropha oil H2O2 added, then FP: 288 8C
60 8C until
complete
Canola oil, acetic Epoxidized canola Sulfated-SnO2 70 8C, 6.5 h 114 19 141 9 Ox induction time: – [68]
acid, and H2O2 oil Catalyst 60 h
Canola oil, acetic Epoxidized canola Amberlite 65 8C, 8 h – – – 9 Therm stab: 319 8C – [75]
acid, H2O2 oil IR-120H
Canola biodiesel, Epoxidized Amberlite 65 8C, 8 h – – – 0 Therm stab: 160 8C – [75]
acetic acid, H2O2 biodiesel IR-120H
Canola oil, acetic Epoxidized canola Amberlite 65 8C, 8 h 151 – – 10 Therm stab: 320 8C – [76]
acid, and H2O2 oil IR-120H
Methyl oleate, Epoxidized methyl – – 8.0 2.5 151 0 Ox onset temp: 97 [77]
formic acid, and oleate 189.75 8C
H2O2
Methyl linoleate, Epoxidized methyl – – 14.3 3.5 132 1.5 Ox onset temp: 95 [77]
formic acid, and linoleate 180.3 8C
H2O2
Methyl linolenate, Epoxidized methyl – – 308 19.3 63 7.5 Ox onset temp: 85 [77]
formic acid and linolenate 131.2 8C
H2O2

Lee et al. [93] hydrogenated soybean oils and observed that the
Other efforts more it was hydrogenated, the higher its viscosity became. They
also experimented with hydrogenation with various catalysts,
Various other methods that do not fall into the above categories such as Ni/SiO2, Ra-N, and Pt/C catalysts and observed that 5% Pt/C
have been examined, but also offer promising advancements in the catalyst gave the highest degree of reduction. Nohaira et al. [94]
study of biolubricants. One of these methods is hydrogenation. used palladium to catalyze the hydrogenation of sunflower oil
Ting and Chen found that various mineral oils can be replaced by ethyl esters with ethanol as a solvent, and found that the addition
blends of hydrogenated, epoxidized, and unmodified soybean oil in of lead to the palladium catalyst, and adding amines to the reaction
different ratios [91]. Shomchoam and Yoosuk hydrogenated palm medium, improved the selectivity of the catalyst.
oil using Pd/g-Al2O3 as a catalyst, and found the partial Examination and production of new feedstocks has also been
hydrogenation improved the oxidative and thermal stability of studied. Eller et al. extracted decanoic acid, which is good for
the oil, while leaving other properties such as viscosity, viscosity estolide formation, from cuphea seed oil using subcritical water
index, and low temperature properties relatively unchanged [92]. without any catalyst in a continuous flow tubular reactor. They
Table 7
Physical properties and reaction conditions of various modified epoxidized vegetable oils.

Reactants Products Catalyst Reaction Viscosity Viscosity Viscosity Pour Oxidative/thermal Yield (%) Ref.
conditions 40 8C (cSt) 100 8C (cSt) index point (8C) stability

Epoxidized mustard oil, Epoxidized mustard oil H2SO4, H3NSO3, 120 8C, until – – 105–159 ( 35) to ( 5) – 92–95 [83]
2-EH ester or CH4O3S complete
Epoxidized canola oil and Epoxidized canola oil Sulfated Ti-SBA- 130 8C, 5 h – 670 – 9 Ox induct time: 56.1 h 100 [81]
acetic anhydride ester 15
Epoxidized soybean oil Transesterified, ring H2SO4 120 8C, 20 h, 59.6–74.5 8.3–10.3 96–135 ( 36) to ( 27) – – [80]
and Guerbet alcohols opened product 0.61 mol alcohol
(1)
Epoxidized soybean oil Ring opened product H2SO4 110 8C, 20 h, 195.6–23.4 16.4–20.9 86–113 18 – – [80]
and Guerbet alcohols 0.47 mol alcohol
(2)
Epoxidized oleic acid, and 9-Hydroxy-10- p- Oleic acid added – – 71.21–232.15 ( 43.25) to FP: 123.09–305.08 8C – [35]

J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12


fatty acids acyloxyoctadecanoic Toluensulphonic over 1.5 h at 70– ( 20.11)
acid acid and toluene 80 8C, then
heated to 90–
100 8C for 3 h
9-Hydroxy-10- Octyl 9-hydroxy-10- – 60 8C, 10 h – – 85.32–183.15 ( 44.17) to FP: 123.38–256.34 8C – [35]
acyloxyoctadecanoic acyloxyoctadecanoate ( 21.14)
acid, octanol, and
hexanes
Epoxidized oleic acid, Alkyl 9-alkyloxy-10- H2SO4 60 8C, 20–22 h – – 95–215 ( 28) to ( 11) Onset temp: 77–156 8C – [84]
various alcohols hydroxyoctadecanoate FP: 113–233
Monoepoxide linoleic 9(12)-hydroxy-10(13)- p- 70–80 8C, while – – 153 51 Ox stab: 180.94 8C – [79]
acid, oleic acid oleoxy-12(9)- Toluensulphonic oleic acid added
octadecanoic acid acid over 1.5 h, then
heated to 90–
110 8C for 3–6 h
Epoxidized canola oil, Diacetylated ring Amberlyst-15 130 8C, 15 h 0 670 – 9 Ox stab: 56.1 h – [75]
acetic anhydride opened product Therm stab: 309 8C
Epoxidized canola Diacetylated ring Amberlyst-15 130 8C, 15 h 116 19 – 18 Ox stab: 76.3 h – [75]
biodiesel, acetic opened product therm stab:
anhydride 194 8C
Epoxidized canola oil and Ring opened product Amberlyst-15 100 8C, 15 h 251.7 – – 5 Therm stab: 355 8C – [76]
n-butanol
Epoxidized canola oil and Ring opened product Amberlyst-15 100 8C, 15 h 190.5 – – 8 Therm stab: 361 8C – [76]
amyl alcohol
Epoxidized canola oil and Ring opened product Amberlyst-15 100 8C, 15 h 85.5 – – 15 Therm stab: 405 8C – [76]
2-EH
Epoxidized castor oil, Polyoleate ester p- 150 8C, 4–5 h 95.15 16.53 189 < 36 FP (open/close): – [87]
oleic acid, and xylene Toluensulphonic 227/198 8C
acid Ox stab: 272 min
Therm stab: 325 8C
Epoxidized linseed oil, Polyoleate ester p- 150 8C, 4–5 h 102.88 16.84 179 15 FP (open/close): – [87]
oleic acid, and xylene Toluensulphonic 238/209 8C
acid Ox stab: 181 min
Therm stab: 330 8C
Epoxidized sunflower oil, Polyoleate ester p- 150 8C, 4–5 h 44.79 8.78 180 9 FP (open/close): – [87]
oleic acid, and xylene Toluensulphonic 153/72 8C
acid Ox stab: 278 min
Therm stab: 290 8C
Epoxidized fatty acid Epoxidized branched CaO 90–140 8C 15.9 3.4 157 15 RPVOT: 127.4 min – [28]
waste cooking oil ester PDSC:
methyl esters, 79.2 min
methanol, and
isooctanol

9
10
Table 7 (Continued )

Reactants Products Catalyst Reaction Viscosity Viscosity Viscosity Pour Oxidative/thermal Yield (%) Ref.
conditions 40 8C (cSt) 100 8C (cSt) index point (8C) stability

Epoxidized fatty acid Epoxidized branched CaO 90–140 8C 24.7 5.1 142 20 RPVOT: 92.9 min – [28]
waste cooking oil ester PDSC:
methyl esters, 50.4 min
methanol, and
isotridecanol
Epoxidized fatty acid Epoxidized branched CaO 90–140 8C 43.4 7.4 135 24 RPVOT: 69.8 min – [28]
waste cooking oi ester PDSC:
methyl esters, 35 min
methanol, and
isooctadecanol
Ring opened epoxidized Acetylated, Pyridine 80 8C, 2 h, 35.6–41.5 6.5–7.5 137–149 ( 42) to ( 27) – – [80]
soybean oil (1) and transesterified, ring nitrogen

J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12


acetic anhydride opened product atmosphere
Ring opened epoxidized Acetylated, ring opened Pyridine 80 8C, 2 h, 74.3–103.4 11.7–14.6 136–152 ( 27) to ( 33) – – [80]
soybean oil (2) and product nitrogen
acetic anhydride atmosphere
9,10-hydroxy- Triester derived from Pyridine 50 8C, 5 h 34.634–46.154 6.473–14.925 127–171 ( 20) to ( 60) Onset temp: 176.29– 66–88 [88]
acyloxystearic acid 9,10-hydroxy- 188.5 8C
methyl esters, CCl4 and acyloxystearic acid
acylchlorides methyl esters
Octyl 9-hydroxy-10- Octyl 9-lauroyloxy-10- Pyridine 4 8C, 2 h – – 93.37–185.36 ( 45.34) to FP: 135.12–211.29 8C – [35]
acyloxyoctadecanoate, acyloxyoctadecanoate ( 23.41)
CCl4, and lauroyl
chloride
Epoxidized karanja oil, Acylated karanja oil
alkanoic anhydride,
and xylene
Dimethylaminopyridine 140–150 8C, 7–9 h – 36.5–63.7 111–128 – FP: 228– – [86]
288 8C
J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12 11

propose the method could be used for the extraction of various overcome before large-scale production is viable. Homogeneity of
fatty acids from triglyceride-based fats and oils [95]. Arad et al. the feedstock is a major issue which can lead to inconsistency in
examined the properties of a sulfated polysaccharide from the final biolubricant. While not as much of an issue in small-scale
porphyridium sp., a red microalga. The polysaccharide biolubricant labs tests, variation in oil content during large-scale manufacturing
showed promise as a medical lubricant for use in artificial joints, as could lead to unsatisfactory properties. Similarly, the reliability
well as various other applications. The viscosity was found to be and price of current feedstocks can fluctuate significantly, making
stable over a wide range of temperatures, pH values, and salinities, production of biolubricants economically risky. In terms of
and its wear scar and coefficient of friction were significantly lower reaction processes, mineral acid catalysts and multiple steps are
than the currently used lubricant [96]. Liu et al. modified a high- generally involved, decreasing the economic and environmental
oleic camelina variety by adding the EaDAcT gene from Euonymus viabilities. Another issue is lack of acceptance by many machine
alatus so that it would produce more sn-3 acetyl triacylglycerols, manufacturers and operators, which makes large-scale production
which have reduced viscosity, improved cold temperature pointless.
properties, and increased oxidative stability compared to most Focus moving forward should be on continually improving the
vegetable oils [24]. cost-effectiveness of production methods. Developing cheaper
Several more novel methods of modifying vegetable oils have feedstocks, higher performance catalysts, and optimized reaction
also been proposed. Wang et al. [97] formulated microemulsions processes will be necessary in furthering the study of biolubricants,
from vegetable oil (continuous phase), ionic liquid (IL) 1-butyl-3- as incentivizing the switch from mineral to vegetable based oils
methyl-imidazolium tetrafluoroborate (polar phase), TritonX-100 through economic benefits will be essential to attract support from
(surfactant), and 1-butanol (cosurfactant), which had excellent industry partners. More thorough testing of the use of vegetable
viscosity and frictional wear properties. Doll and Sharma emulsi- oils in the place of mineral oils will also be needed in order to
fied epoxidized vegetable oils and vegetable oil derivatives, and convince manufacturers and operators of vegetable oils’ merits.
found that even 1% emulsification of the vegetable oil can undo the These efforts will lead to a wide use of biolubricants in the future,
negative effects of epoxidation on the coefficient of friction, and offering enormous benefits including excellent biodegradability,
improve frictional properties [98]. Biswas et al. heat-bodied and reduced the reliance on petroleum, less negative impact to human
microwave-irradiated soybean oil and found that while both health and minimum harm to the environment.
increased viscosity, decreased pour point, and increased oxidative
stability, microwave-irradiation improved all these qualities more
than heat-bodying did. However, microwave-irradiation hindered References
the lubricity properties of the soybean oil [99]. Biresaw and
[1] H.M. Mobarak, E. Niza Mohamad, H.H. Masjuki, M.A. Kalam, K.A.H. Al Mahmud,
Bantchev produced phosphonate derivatives of methyl oleate via a
M. Habibullah, A.M. Ashraful, Renew. Sustain. Energy Rev. 33 (2014) 34.
radical chain reaction with varying alkoxy groups. The resulting [2] R. Garcés, E. Martı́nez-Force, J.J. Salas, Grasas Aceites 62 (2011) 21.
phosphonates had increased viscosities, as well as improved [3] P. Nagendramma, S. Kaul, Renew. Sustain. Energy Rev. 16 (2012) 764.
oxidative stability, low temperature properties, and deceased [4] S. Rani, M.L. Joy, K.P. Nair, Ind. Crops Prod. 65 (2015) 328.
[5] W. Dalbey, R. Biles, Appl. Occup. Environ. Hyg. 18 (2003) 921.
coefficients of friction and wear scar [100]. [6] J. Urbanus, R. Lobo, A. Riley, Appl. Occup. Environ. Hyg. 18 (2003) 815.
[7] W.E. Dalbey, R.H. McKee, K.O. Goyak, R.W. Biles, J. Murray, R. White, Int. J. Toxicol.
Conclusions 33 (2014) 110.
[8] S.T. Ivo, M.L. Paulo Renato, N.M. Renato, D.B. Ederio, Braz. Arch. Biol. Technol. 55
(2012) 951.
Chemical modification of vegetable oils is a promising option [9] J. Salimon, N. Salih, E. Yousif, Eur. J. Lipid Sci. Technol. 112 (2010) 519.
for producing biolubricants from vegetable oils. These methods [10] F.M. Luna, B.S. Rocha, E.M. Rola, M. Albuquerque, D.C.S. Azevedo, C. Cavalcante,
Ind. Crops Prod. 33 (2011) 579.
have been found to significantly increase the physical properties of [11] S. Fernando, M. Hanna, S. Adhikari, Appl. Eng. Agric. 23 (2007) 5.
the base vegetable oils, and produce biolubricants that meet or [12] S. Soni, M. Agarwal, Green Chem. Lett. Rev. 7 (2014) 359.
exceed requirements. Esterification/transesterification of vegeta- [13] S. Syahrullail, B.M. Zubil, C.S.N. Azwadi, M.J.M. Ridzuan, Int. J. Mech. Sci. 53
(2011) 549.
ble oils improves their low temperature properties and increases [14] T. Regueira, L. Lugo, O. Fandiño, E.R. López, J. Fernández, Green Chem. 13 (2011)
their stability, but generally requires crops rich in oleic acid, and 1293.
the process requires relatively high reaction temperatures and [15] M.T. Siniawski, N. Saniei, P. Stoyanov, Lubr. Sci. 23 (2011) 301.
[16] S.C. Cermak, G. Biresaw, T.A. Isbell, R.L. Evangelista, S.F. Vaughn, R. Murray, Ind.
negative pressure. Estolide formation provides improved lubricity,
Crops Prod. 44 (2013) 232.
oxidative stability, and low temperature properties and generally [17] H. Wagner, R. Luther, T. Mang, Appl. Catal. A: Gen. 221 (2001) 429.
requires lower reaction temperatures. It can also make use of a [18] R. Kreivaitis, J. Padgurskas, M. Gumbytė, V. Makarevičienė, B. Spruogis, Transport
variety of fatty acids, producing a wide array of lubricants with 26 (2011) 121.
[19] A.K. Jain, A. Suhane, Adv. Eng. Appl. Sci. 1 (2012) 23.
dramatically different properties, and the resulting estolides can be [20] K. Kamalakar, A.K. Rajak, R.B.N. Prasad, M.S.L. Karuna, Ind. Crops Prod. 51 (2013)
esterified to further improve low temperature properties and 249.
stability. However, the initial estolide formation reaction often [21] A.E. Atabani, A.S. Silitonga, H.C. Ong, T.M.I. Mahlia, H.H. Masjuki, I.A. Badruddin,
H. Fayaz, Renew. Sustain. Energy Rev. 18 (2013) 211.
requires the use of capping fatty acids, which are an expensive [22] S.A. Smith, R.E. King, D.B. Min, Food Chem. 102 (2007) 1028.
reactant. Epoxidation improves the lubricity characteristics and [23] R. Yunus, A. Fakhru’l-Razi, T.L. Ooi, R. Omar, A. Idris, Ind. Eng. Chem. Res. 44
stability of the vegetable oil, and has relatively low reaction (2005) 8178.
[24] J. Liu, H. Tjellström, K. Mcglew, V. Shaw, A. Rice, J. Simpson, D. Kosma, W. Ma, W.
temperatures due to the exothermic nature of the reaction, but Yang, M. Strawsine, E. Cahoon, T.P. Durrett, J. Ohlrogge, Ind. Crops Prod. 65
significantly increases the pour point and decreased the viscosity (2015) 259.
index. Subsequent ring opening, esterification, and/or acetylation [25] S. Gryglewicz, M. Muszyński, J. Nowicki, Ind. Crops Prod. 45 (2013) 25.
[26] W. Castro, J. Perez, S. Erhan, F. Caputo, J. Am. Oil Chem. Soc. 83 (2006) 47.
of the epoxidized vegetable oils can provide a final product with [27] E. Wang, X. Ma, S. Tang, R. Yan, Y. Wang, W.W. Riley, M.J.T. Reaney, Biomass
good lubricity, stability, low temperature properties, and viscosity Bioenergy 66 (2014) 371.
index, while requires two or three additional reactions after the [28] W. Li, X. Wang, J. Oleo Sci. 64 (2015) 367.
[29] A. Chowdhury, D. Mitra, D. Biswas, J. Chem. Technol. Biotechnol. 88 (2013) 139.
epoxidation reaction which is economically unfavorable.
[30] A. Chowdhury, D. Mitra, D. Biswas, Environ. Prog. Sustain. Energy 33 (2014) 933.
Despite much progress being made in technical knowledge, [31] M.S. Silva, E.L. Foletto, S.M. Alves, C.D. de, A.A. Dantas Neto, Ind. Crops Prod. 69
large-scale production of biolubricants using the methods (2015) 362.
described in this paper has not been performed. The production [32] K. Kamalakar, G.N.V.T. Sai Manoj, R.B.N. Prasad, M.S.L. Karuna, Grasas Aceites 66
(2015).
and use of biolubricants still faces many challenges that must be
12 J. McNutt, Q.S. He / Journal of Industrial and Engineering Chemistry 36 (2016) 1–12

[33] Mobil Delvac Synthetic Gear Oil 75W-90, 80W-140, 2014, December Available: [62] S.C. Cermak, A.L. Durham, T.A. Isbell, R.L. Evangelista, R.E. Murray, Ind. Crops
http://www.mobil.com/USA-English/Lubes/PDS/NAXXENCVLMOMobil_ Prod. 67 (2015) 179.
Delvac_Synthetic_Gear_Oil_75W-90_80W-140.aspx. [63] G. Biresaw, G. Bantchev, S. Cermak, Tribol. Lett. 43 (2011) 17.
[34] Castrol Syntrax Limited Slip 75W-140, 2015, July Available: http://msdspds. [64] S.C. Cermak, K.B. Brandon, T.A. Isbell, Ind. Crops Prod. 23 (2006) 54.
castrol.com/bpglis/FusionPDS.nsf/Files/ [65] S.C. Cermak, J.W. Bredsguard, B.L. John, J.S. McCalvin, T. Thompson, K.N. Isbell,
37F61A222EAA409F80257E75002DD9C5/$File/BPXE-9YYLDZ.pdf. K.A. Feken, T.A. Isbell, R.E. Murray, Ind. Crops Prod. 46 (2013) 386.
[35] N. Salih, J. Salimon, E. Yousif, Ind. Crops Prod. 34 (2011) 1089. [66] S.C. Cermak, T.A. Isbell, R.L. Evangelista, B.L. Johnson, Ind. Crops Prod. 33 (2011)
[36] A. Sammaiah, K.V. Padmaja, R.B. Narayna Prasad, J. Oleo Sci. 5 (2014). 132.
[37] T.I. Mohd., Ghazi Gunam Resul, F.M. Mohamad, A. Idris, IJCRE 7 (2009). [67] K.M. Doll, B.K. Sharma, S.Z. Erhan, Clean 36 (2008) 700.
[38] N. Salih, J. Salimon, E. Yousif, J. King Saud Univ.: Sci. 24 (2012) 221. [68] A.K.R. Somidi, R.V. Sharma, A.K. Dalai, Ind. Eng. Chem. Res. 53 (2014) 18668.
[39] J. Salimon, B. Abdullah, R. Yusop, N. Salih, Chem. Cent. J. 8 (2014) 1. [69] G. Gorla, S.M. Kour, K.V. Padmaja, M.S.L. Karuna, R.B.N. Prasad, Ind. Eng. Chem.
[40] Anon., ASTM D97-12, Standard Test Method for Pour Point of Petroleum Res. 52 (2013) 16598.
Products, ASTM International, 2012. [70] S.Z. Erhan, B.K. Sharma, Z. Liu, A. Adhvaryu, J. Agric. Food Chem. 56 (2008) 8919.
[41] Anon., ASTM D445-15, Standard Test Method for Kinematic Viscosity of Trans- [71] E. Milchert, A. Smagowicz, G. Lewandowski, J. Chem. Technol. Biotechnol. 85
parent and Opaque Liquids (and Calculation of Dynamic Viscosity), ASTM (2010) 1099.
International, 2015. [72] R. Mungroo, N. Pradhan, V. Goud, A. Dalai, J. Am. Oil Chem. Soc. 85 (2008) 887.
[42] Anon., ASTM D2270-10e1, Standard Practice for Calculating Viscosity Index [73] S. Sun, G. Yang, Y. Bi, H. Liang, J. Am. Oil Chem. Soc. 88 (2011) 1567.
From Kinematic Viscosity at 40 and 100 8C, ASTM International, 2010. [74] R.D.V.V. Lopes, J.R. Zamian, I.S. Resck, M.J.A. Sales, M.L. Dos Santos, F.R. Da Cunha,
[43] Anon., ASTM D2500-11, Standard Test Method for Cloud Point of Petroleum Eur. J. Lipid Sci. Technol. 112 (2010) 1253.
Products, ASTM International, 2011. [75] R.V. Sharma, A.K.R. Somidi, A.K. Dalai, J. Agric. Food Chem. 63 (2015) 3235.
[44] Anon., ASTM D4172-94(2010), Standard Test Method for Wear Preventive [76] C.S. Madankar, A.K. Dalai, S.N. Naik, Ind. Crops Prod. 44 (2013) 139.
Characteristics of Lubricating Fluid (Four-Ball Method), ASTM International, [77] B.K. Sharma, K.M. Doll, S.Z. Erhan, Green Chem. 9 (2007) 469.
2010. [78] B.K. Ahn, S. Kraft, X.S. Sun, J. Agric. Food Chem. 60 (2012) 2179.
[45] M. Gunam Resul Faiz Mukhtar, T.I. Mohd. Ghazi, A. Idris, Ind. Crops Prod. 38 [79] J. Salimon, N. Salih, B.M. Abdullah, J. Biomed. Biotechnol. (2011) 196565.
(2012) 87. [80] H. Hwang, S.Z. Erhan, Ind. Crops Prod. 23 (2006) 311.
[46] M.Y. Koh, T.I. Mohd. Ghazi, A. Idris, Ind. Crops Prod. 52 (2014) 567. [81] R.V. Sharma, A.K. Dalai, Appl. Catal. B: Environ. 142 (2013) 604.
[47] H. Masood, R. Yunus, T.S.Y. Choong, U. Rashid, Y.H. Taufiq, S. Yap, Appl. Catal. A: [82] S. Arumugam, G. Sriram, Proc. Inst. Mech. Eng. Part J 227 (2013) 3.
Gen. 425 (2012) 184. [83] R.D. Kulkarni, P.S. Deshpande, S.U. Mahajan, P.P. Mahulikar, Ind. Crops Prod. 49
[48] E. Kleinaitė, V. Jaška, B. Tvaska, I. Matijošytė, J. Clean. Prod. (2014). (2013) 586.
[49] P.K. Sripada, R.V. Sharma, A.K. Dalai, Ind. Crops Prod. 50 (2013) 95. [84] J. Salimon, N. Salih, E. Yousif, J. Oleo Sci. 60 (2011) 613.
[50] J.A.C. Silva, A.C. Habert, D.M.G. Freire, Lubr. Sci. 25 (2013) 53. [85] J. Salimon, N. Salih, E. Yousif, Arab. J. Chem. 5 (2012) 193.
[51] J. Oh, S. Yang, C. Kim, I. Choi, J.H. Kim, H. Lee, Appl. Catal. A: Gen. 455 (2013) 164. [86] G. Gorla, S.M. Kour, K.V. Padmaja, M.S.L. Karuna, R.B.N. Prasad, Ind. Eng. Chem.
[52] P.S. Sreeprasanth, R. Srivastava, D. Srinivas, P. Ratnasamy, Appl. Catal. A: Gen. Res. 53 (2014) 8685.
314 (2006) 148. [87] A.I. Hashem, W.S.I. Abou Elmagd, A.E. Salem, M. El-Kasaby, A. El-Nahas, Energy
[53] P. Nagendramma, Lubr. Sci. 23 (2011) 355. Sources Part A: Recov. Util. Environ. Effects 35 (2013) 397.
[54] C.O. Åkerman, Y. Gaber, N.A. Ghani, M. Lämsä, R. Hatti-Kaul, J. Mol. Catal. B: [88] A. Kleinová, P. Fodran, L. Brnčalová, J. Cvengroš, Biomass Bioenergy 32 (2008)
Enzym. 72 (2011) 263. 366.
[55] J. Madarász, D. Németh, J. Bakos, L. Gubicza, P. Bakonyi, J. Clean. Prod. 93 (2015) [89] S. Sankaranarayanan, K. Srinivasan, RSC Adv. 5 (2015) 50289.
140. [90] K. Lee, C. Hailan, J. Yinhua, Y. Kim, K. Chung, Korean J. Chem. Eng. 25 (2008) 474.
[56] M. Happe, P. Grand, S. Farquet, S. Aeby, J. Hritier, F. Corthay, E. Mabillard, R. Marti, [91] C. Ting, C. Chen, Measurement 44 (2011) 1337.
E. Vanoli, A. Grogg, S. Nussbaum, A. Roduit, F. Tiche, S. Salem, C. Constantin, E. [92] B. Shomchoam, B. Yoosuk, Ind. Crops Prod. 62 (2014) 359.
Schmitt, S. Zahno, C. Ellert, A. Habib, J. Wyss, F. Fischer, Green Chem. 14 (2012) [93] K.W. Lee, B.X. Mei, Q. Bo, Y.W. Kim, K.W. Chung, Y. Han, J. Ind. Eng. Chem. 13
2337. (2007) 530.
[57] C.O. Åkerman, A.E.V. Hagström, M.A. Mollaahmad, S. Karlsson, R. Hatti-Kaul, [94] B. Nohair, C. Especel, G. Lafaye, P. Marécot, L.C. Hoang, J. Barbier, J. Mol. Catal. A:
Process Biochem. 46 (2011) 2225. Chem. 229 (2005) 117.
[58] L.A. Garcı́a-Zapateiro, J.M. Franco, C. Valencia, M.A. Delgado, C. Gallegos, M.V. [95] F. Eller, J. Teel, D. Palmquist, J. Am. Oil Chem. Soc. 88 (2011) 1455.
Ruiz-Méndez, Grasas Aceites 64 (2013) 497. [96] S.M. Arad, L. Rapoport, A. Moshkovich, D. van Moppes, M. Karpasas, R. Golan, Y.
[59] L.A. Garcı́a-zapateiro, J.M. Franco, C. Valencia, M.A. Delgado, C. Gallegos, M.V. Golan, Langmuir 22 (2006) 7313.
Ruiz-méndez, Eur. J. Lipid Sci. Technol. 115 (2013) 1173. [97] A. Wang, L. Chen, D. Jiang, Z. Yan, Ind. Crops Prod. 51 (2013) 425.
[60] L.A. Garcı́a-Zapateiro, M.A. Delgado, J.M. Franco, C. Valencia, M.V. Ruiz-Méndez, [98] K. Doll, B. Sharma, J. Surfactants Deterg. 14 (2011) 131.
R. Garcés, C. Gallegos, Grasas Aceites 61 (2010) 171. [99] A. Biswas, A. Adhvaryu, D.G. Stevenson, B.K. Sharma, J.L. Willet, S.Z. Erhan, Ind.
[61] L.A. Garcı́a-Zapateiro, J.M. Franco, C. Valencia, M.A. Delgado, C. Gallegos, J. Ind. Crops Prod. 25 (2007) 1.
Eng. Chem. 19 (2013) 1289. [100] G. Biresaw, G. Bantchev, J. Am. Oil Chem. Soc. 90 (2013) 891.

Vous aimerez peut-être aussi