Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Article

pubs.acs.org/JAFC

Preparation and Properties Evaluation of Biolubricants Derived from


Canola Oil and Canola Biodiesel
Rajesh V. Sharma, Asish K. R. Somidi, and Ajay K. Dalai*
Catalysis and Chemical Reaction Engineering Laboratories, Department of Chemical and Biological Engineering, University of
Saskatchewan, Saskatoon, SK S7N 5A9, Canada

ABSTRACT: This study demonstrates the evaluation and comparison of the lubricity properties of the biolubricants prepared
from the feed stocks such as canola oil and canola biodiesel. Biolubricant from canola biodiesel has a low cloud and pour point
properties, better friction and antiwear properties, low phase transition temperature, is less viscous, and has the potential to
substitute petroleum-based automotive lubricants. Biolubricant from canola oil has high thermal stability and is more viscous and
more effective at higher temperature conditions. This study elucidates that both the biolubricants are attractive, renewable, and
ecofriendly substitutes for the petroleum-based lubricants.
KEYWORDS: Canola oil, Canola biodiesel, biolubricant, epoxy ring-opening and esterification, lubricity

■ INTRODUCTION
Biomass-based materials such as biodiesel, biolubricants,
bioplastics, surfactants, etc. are capable of substituting the
fossil-based products and fuels and, hence, have the promising
future market. Diesel fuel is being replaced by the biodiesel
obtained from vegetable oils due to its engine compatibility.1,2
Biodiesel has a carbon foot print and basically originated from
plants. Biodiesel is produced from edible and nonedible
vegetable oils, and this technology is already commercial-
ized.3−5 The lubricants obtained from vegetable oils are not
much explored and has attracted more attention from
researchers.
Industry and automobile sectors extensively use lubricants
for lubricating their machineries and materials. The demand for
the lubricants is growing by 1.6% per year.6 The petroleum-
based lubricants generate a significant amount of wastes, which
increases the concern for pollution and are environmentally
unsafe.7 Biolubricants are green lubricant in comparison with
the crude oil derived lubricants because they are eco-friendly,
renewable, and biodegradable.8−10 Also, they possess the
characteristic properties required by the ideal lubricants.10,11
Biolubricants prepared from vegetable oils which have high
oleic content can substitute conventional mineral oil-based
lubricants.12 In this context, soyabean oil has been explored
effectively for making a variety of lubricating base stocks, which
find applications in chemical and automobile industries.
However, lubricants obtained from vegetable oils have the
drawback of poor oxidative stability, which can be improved by
the removal of polyunsaturation present in the oil.13−15 Figures
1 and 2 explain the schematic representation of preparation of Figure 1. Reaction scheme for the synthesis of canola oil biolubricant.
biolubricants from vegetable oil and biodiesel.
Several reports are available on the conversion of the CC
double bond to epoxy linkage by using homogeneous catalysts
and ion-exchange resins. The epoxidation of mahua oil
(Madhuca longifolia) and cotton seed oil was carried out Received: December 2, 2014
using hydrogen peroxide as oxidant in the presence of mineral Revised: March 13, 2015
acids such as H2SO4/HNO3 as catalysts.16,17 The epoxidation Accepted: March 15, 2015
of unsaturated triglycerides in the wild safflower oil was carried

© XXXX American Chemical Society A DOI: 10.1021/jf505825k


J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

out using Amberlite IR-122 resin as catalyst.18 Also, Amberlite


IR-120H resin was found to be active catalyst for the
epoxidation of polyunsaturated canola triglycerides.12
A few studies have been reported on the vicinal diacylation of
epoxidized vegetable oils to obtain the product biolubricant.
Hwang and Erhan19 studied epoxy ring-opening by various
linear and branched alcohols. Adhvaryu et al.20 synthesized
acylated soybean oil derivatives (biolubricants) using different
carboxylic anhydrides and pyridine as catalysts. The synthesis
method reported by Salimon et al.21 for the acylation of
ricinoleic acid includes sulfuric acid as a catalyst. Limited
literature reports are available on the application of
heterogeneous catalysts for the simultaneous oxirane ring-
opening and diacylation of epoxidized triglycerides in vegetable
oils to obtain biolubricant (i.e., esterified product). Application
of solid metal oxide catalysts for the preparation of
biolubricants from vegetable oils not only promote green
Figure 2. Reaction scheme for the synthesis of biodiesel derived
technology but also minimize workup procedures to separate
biolubricant. the product and the catalyst can be reused several times.22,23
The present investigation deals with the synthesis of
biolubricants from canola oil and canola biodiesel by using
heterogeneous catalysts (Figures 1 and 2). As per our
knowledge, this report finds novelty with the preparation of

Figure 3. 1H NMR spectra: (A) canola oil, (B) epoxidized canola oil, (C) biolubricant from canola oil, (D) D2O exchanged biolubricant from canola
oil, (E) canola biodiesel, (F) epoxidized canola biodiesel, (G) biolubricant from canola biodiesel, and (H) D2O exchanged biolubricant from canola
biodiesel.

B DOI: 10.1021/jf505825k
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Figure 4. 13C NMR spectra: (A) canola oil, (B) epoxidized canola oil, (C) biolubricant from canola oil, (D) canola biodiesel, (E) epoxidized canola
biodiesel, and (F) biolubricant from canola biodiesel.

new value-added product biolubricant from canola biodiesel Where, N stands for HBr solution normality.
and the characteristic lubricity properties associated with it in
comparison to canola oil biolubricant. epoxy conversion (%)

■ EXPERIMENTAL SECTION
Chemicals and Reagents. Edible grade canola oil
=
oxirane value at O h − oxirane value at time (t )
oxirane value at O h
× 100
(2)
(Loblaws Inc. Montreal, Canada), canola biodiesel (Milligan
Biofuels Inc., Saskatoon, Canada), glacial acetic acid (reagent Kinematic viscosity, cloud point, and pour point were
grade), 30 wt % hydrogen peroxide (GR grade) (EMD measured as per the ASTM methods. The oxidative stability
Chemicals Inc., Darmstadt, Germany), Amberlite IR-120H measurements were carried out by AOCS methods. Lubricity of
(Sigma-Aldrich, St. Louis, MO), Amberlyst-15 (Sigma-Aldrich, the products was performed according to ASTM methods.
St. Louis, MO), Wijs’ solution (VWR, San Diego, CA), ethyl Thermogravimetric analysis was carried to determine the
acetate, and 33 wt % hydrobromic acid in acetic acid (EMD weight loss of materials. The details of the analysis procedure
Chemicals Inc., Montreal, Canada) were purchased. and equipment used for the determination of tribological
properties of biolubricants are described in the literature.6,23


Analytical Methods. The iodine value (% unsaturation)
analysis was carried out as per the AOCS Cd 1−25 standard
method. The amount of epoxy content (oxirane oxygen) in the
EXPERIMENTAL SETUP AND PROCEDURE
oil was calculated using AOCS Cd 9-57 volumetric titration Epoxidation. The unsaturated fatty acids present in the feed stocks
method. Each experiment was repeated thrice and found to such as canola oil and canola biodiesel were epoxidized by using
hydrogen peroxide and acetic acid.6 A sample of 22.6 g of canola oil/
have ±3% error in the conversion. Equations 1 and 2 give the biodiesel, acetic acid, and Amberlite IR −120H (22 wt % w.r.t to g of
appropriate formulas employed for determining the change in feed stock taken) were placed in a three neck round-bottom flask and
the percentage unsaturation and the epoxy content in the oils maintained at 65 ± 2 °C. Then, aqueous hydrogen peroxide (30 wt %)
during the course of the reaction. was added with a molar ratio of 1.5:1 (unsaturation in canola oil/
canola biodiesel: H2O2). The reaction was continued at 65 ± 2 °C for
oxirane oxygen content 8 h with rapid stirring. The oil phase was extracted using ethyl acetate
mL of HBr consumed × N × 1.60 as solvent and was separated by evaporation. The conversion was
= assessed by the iodine value (7.5 and 7.6 g of I2/100 g for canola oil
mass of sample, g (1) and biodiesel, respectively), oxirane content (5.6 and 5.8%), and glycol

C DOI: 10.1021/jf505825k
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Figure 5. FTIR spectrum: (A) canola oil, (B) epoxidized canola oil, (C) biolubricant from canola oil, (D) canola biodiesel, (E) epoxidized canola
biodiesel, and (F) biolubricant from canola biodiesel.

content (0.08 and 0.06 mol/100 g) of canola oil and canola biodiesel, canola biodiesel exhibited no signal between 120 and 140 ppm,
respectively. Further, the product formation was also confirmed by indicating the conversion of the olefinic carbon (−CC−) to
spectroscopic techniques such as 1H NMR, 13C NMR, and FT-IR. epoxy linkage, given by the signal appearing between 53 and 58
Epoxy Ring-Opening of Epoxidized Canola Oil/Epoxidized
Biodiesel. The oxirane ring-opening and diacylation reactions were ppm. Biolubricant from canola oil and biolubricant from canola
performed in a single pot. Typically, 30.0 g of epoxidized canola oil/ biodiesel showed no signal between 53 and 58 ppm, which
epoxidized biodiesel, 45 g of acetic anhydride, and Amberlyst-15 (10 indicates the conversion of epoxy linkage to ester linkage. This
wt % of epoxidized product) were mixed in the three neck round- is indicated by a new signal at 170 ppm, which is attributed to
bottom flask and stirred for 15 h at 130 °C. The percentage epoxy carbonyl carbon. The presence of glycerol backbone in canola
conversion was measured periodically by the oxirane content method oil biolubricant is indicated by the shift at 62 and 68 ppm for α
(eq 1), and the product formation was also identified with FTIR, 1H and β carbon atoms of the glycerol backbone, respectively.
NMR, and 13C NMR analysis. Figure 4 also confirmed the presence of −CH3 carbon of

■ RESULTS AND DISCUSSION


Characterization of Products. NMR Spectroscopy. Figure
−COOCH3 that appeared at 51 ppm, which is characteristic of
canola biodiesel. Therefore, the formation of canola oil
biolubricant and canola biodiesel-derived biolubricant is
3 represents 1H NMR spectra of starting materials and confirmed by the NMR spectra.
products. The olefinic hydrogens (H−CC−H) of canola FT-IR Spectroscopy. FT-IR transmittance spectra of the
oil and canola biodiesel showed a chemical shift between 5.2 products and reactants were recorded to confirm the formation
and 5.4 ppm. These proton shifts were found to be absent in of biolubricant on the basis of functional groups present. FT-IR
the epoxidized canola oil and epoxidized canola biodiesel, spectra are shown in Figure 5. Canola oil and canola biodiesel
indicating that all olefinic groups are converted to the epoxy show the bands at 3007 and 738 cm−1, which are due to the
group. 1H NMR spectra of epoxidized canola oil and epoxidized bending and stretching vibrations of H−CC−H. No bands
canola biodiesel showed a chemical shift between 2.7 and 3.1 appeared at 3007 and 738 cm−1 indicating that −CC− bonds
ppm due to the vicinity of epoxy oxygen.12 The CH2 proton α- in the oil are converted into epoxide. Epoxy band was found to
to epoxy group showed a chemical shift at 1.9 ppm, while the β- appear at 831 cm−1.24 The FT-IR spectra of biolubricant from
CH2 proton has a chemical shift at 1.6 ppm. The terminal canola oil and biolubricant from canola biodiesel have no band
−CH3 proton indicates the chemical shift at 0.8−1.0 ppm. 1H at 831 cm−1, indicating the absence of the epoxy group.
NMR spectra of canola oil biolubricant and biolubricant from However, two new bands appeared at 725 and 604 cm−1, with
canola biodiesel show a chemical shift at 5.0 ppm, which also a corresponding increased intensity of the band at 1750
represents both CH− protons associated with the carbonyl cm−1. This confirmed the ester structure present in the
group. The D2O exchanged 1H NMR spectrum of canola oil products. Hence, FT-IR analysis indicated the formation of
biolubricant, and biolubricant from canola biodiesel confirmed biolubricant from canola oil and canola biodiesel which was also
the absence of hydroxyl proton in the molecule which indicates confirmed by NMR results.
the formation of diacylated product after the epoxy ring- Low Temperature Properties. The cold flow temperature
opening step. Figure 3 confirmed the presence of methine properties of the biolubricants from canola oil and canola
proton of glycerol backbone by the signal appearing between biodiesel were represented by the cloud point and pour point.
5.2 and 5.4 ppm and also the presence of −COOCH3 proton These properties of lubricants derived from vegetable oil are
that appeared at 3.7 ppm, which is characteristic of canola poor, hence, making it difficult to use in cold countries. It can
biodiesel. be improved by reducing the aliphatic carbon chain length.
Figure 4 represents 13C NMR spectra of starting materials Table 1 represents the low-temperature properties of
and products. Canola oil and canola biodiesel showed the signal epoxidized canola oil, epoxidized biodiesel, biolubricant from
appearing between 120 and 140 ppm due to the existence of canola oil, and biolubricant from canola biodiesel. The
olefinic carbon atoms. Epoxidized canola oil and epoxidized epoxidized canola oil and epoxidized canola biodiesel have
D DOI: 10.1021/jf505825k
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Table 1. Tribological Properties of Products biolubricant from canola oil. It might be because of less
number of aliphatic carbon present in the biolubricant derived
low temperature kinematic
property viscosity (cSt) from canola biodiesel, which creates less steric hindrance and
fewer hydrogen bondings as compared with the biolubricant
cloud pour oxidative
point point stability derived from canola oil (Figure 2). Therefore, biolubricant
entry product (°C) (°C) 40 °C 100 °C (OIT) h derived from biodiesel can be used as lubricant for automobiles
1 epoxidized 15 9 − − − at low temperature. Hence, it can be anticipated that the
canola oil biolubricant from biodiesel has a better future prospect for
2 epoxidized 6 0 − − − automobile sector.
canola
biodiesel Friction and Anti-Wear Properties. Lubricity testing was
3 biolubricant −3 −9 0 670 56.1 performed according to the report by Sripada et al.26 All the
from canola samples were tested twice, and the average value is reported.
oil The wear scar diameter of standard diesel sample was found to
4 biolubricant −12 −18 116 19 76.3 be 600 μm. With the addition of 1% of canola oil biolubricant
from canola
biodiesel to the standard diesel sample, the diameter of the wear scar was
found to reduce to 130 μm, while 1% of biolubricant derived
from canola biodiesel integrated in the standard diesel sample
the cloud points of 15 and 6 °C and the pour points of 9 and 0 showed a wear scar diameter of 95 μm. It was noticed that
°C, respectively. It was found that cloud and pour point of canola oil biolubricant resulted in more wear scar on the steel
epoxidized biodiesel are low as compared with epoxidized test ball as compared with biolubricant from canola biodiesel
canola oil, which can be due to the presence of less number of (Figure 6). This can be due to the presence of more steric
aliphatic carbon atoms present in epoxidized biodiesel. Hence, hindrance and intramolecular hydrogen bonding present in
an increase in the aliphatic carbon chain length reduced the canola biolubricant, which increases the viscosity and hence
low-temperature properties. Also, these low-temperature decreases the solubility in the standard diesel sample and
properties can be improved by adding the ester linkage in the resulted in more wear scar diameter (130 μm). Furthermore, a
epoxidized molecules, which promote steric hindrance. The lower number of aliphatic carbon atoms is present in biodiesel
cloud points of biolubricant from canola oil and biolubricant biolubricant as compared with canola biolubricant, which
from canola biodiesel were found to be −3 and −12 °C, while makes it less viscous and less nonpolar and, hence, increases the
pour points were −9 and −18 °C, respectively. Hence, it solubility in diesel fuel and results in less wear scar (95 μm).
elucidates that biolubricant obtained from biodiesel has better The better lubricity property of biodiesel biolubricant can also
low-temperature properties as compared to biolubricant be explained by the presence of ester oxygen moieties in 1, 9,
obtained from canola oil, which can be due to the less steric and 10 positions, which are free of hydrogen bonding and help
hindrance. Hence, biolubricant from canola biodiesel is superior to reduce the friction by developing the antifrictional film on
for low-temperature applications and has better future metal surfaces.27 Table 2 indicates that with the increase in
prospects for the automotive industry as compared to lubricant content in pure diesel fuel, the wear scar on the steel
biolubricant from canola oil. test ball is decreased.
Viscosity. The viscosity of the fluid defines the effectiveness
of the lubricant to lubricate the contact surface of the metals. Table 2. Wear Scar Diameter of Bio-lubricants Diluted in
The kinematic viscosity of canola biolubricant and biodiesel Low Lubricity Diesel
biolubricant were measured at 40 °C and found to be 0 and 116
cSt, whereas at 100 °C, it was found to be 670 and 19 cSt wear scar diameter (μm)
(Table 1). The viscosity index for biolubricant from canola biolubricant added in pure biolubricant from biolubricant from
biodiesel was found to be 185. High viscous nature of canola oil diesel (%) canola oil biodiesel
biolubricant is because of the addition of ester linkage in the oil 1 130 95
at the unsaturation. The increase in the viscosity is also 2 125 90
attributed to the intramolecular hydrogen bonding of the ester 5 120 86
linkages.25 Canola oil biolubricant is highly viscous and can be 10 115 81
used as a lubricant for high temperature, high load, and high-
speed operations like heavy machinery. However, biolubricant Thermogravimetric (TGA) Analysis. Thermal stability of
from canola biodiesel is less viscous as compared with canola oil, epoxdized canola oil, canola biolubricant, biodiesel,

Figure 6. HFRR wear scar images of (A) standard diesel fuel, (B) 1% canola biolubricant, and (C) 1% biodiesel biolubricant.

E DOI: 10.1021/jf505825k
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Figure 7. TGA thermogram of (A) canola oil, (B) epoxidized canola oil, (C) canola biolubricant, (D) biodiesel, (E) epoxidized biodiesel, and (F)
biodiesel biolubricant.

epoxidized biodiesel, and biodiesel biolubricant was studied by can be due to the arrangement of the saturated molecules that
thermo gravimetric differential thermal analyzer (TG/DTA). influence the crystallization of the compound. Hence, bio-
The detailed analysis procedure is described in the literature.26 lubricant from biodiesel has better low-temperature properties.
The graphs of weight loss against temperature are shown in Oxidation Stability Study. The oxidation stability of
Figure 7. It is clear from the graph that canola oil, epoxidized biolubricants depends on the degree of unsaturation present in
canola oil, and canola biolubricant are thermally stable below the molecules which are very prone to oxidation. These
305 °C. Maximum weight loss (90−95% wt) was observed in oxidation processes of lubricants take place during the
all these three materials at the temperatures 460, 461, and 530 combustion process of the fuel inside the engine, hence it is
°C, respectively. However, canola biodiesel, epoxidized canola a critical parameter.28 Canola biolubricant showed a lower value
biodiesel, and biodiesel biolubricant are found to be thermally of 56.1 h, whereas biodiesel biolubricant gave 76.3 h of
stable below the temperatures 305, 160, and 194 °C, oxidative induction time (Table 1). This increase in the
respectively, and about 90−95% weight loss was found for oxidative stability is because of the existence of short chain ester
each sample at the temperatures 462, 251, 370 °C, respectively linkage in canola biodiesel biolubricant as compared with
(Table 3). Hence, it is conclusive that canola biolubricant is canola biolubricant.27 Hence, biodiesel biolubricant is close to
1.5 times more stable when compared with canola biolubricant,
Table 3. Thermal Stability Data of Feed and the Products under similar oxidative conditions, which makes it more
temperature (°C) suitable for automotive applications.
Catalyst Reusability. In the present study, the two-step
product stability 15−20 wt % loss 90−95 wt % loss
process was demonstrated for preparation of biolubricant,
Canola oil 333 375 460 hence two different heterogeneous catalysts were used (i.e.,
epoxidized canola oil 319 370 461 Amberlite IR-120H and Amberlyst-15). The reusability of the
Canola biolubricant 309 364 530
catalysts was tested up to the third run. The reusability
biodiesel 309 360 462
procedure was followed as in the literature.6 The conversion
epoxidized biodiesel 160 204 251
decreased marginally in both the steps, and both the catalysts
biodiesel biolubricant 194 247 370
were used successfully up to three times.
In conclusion, in the present study two different bio-
thermally more stable as compared to biolubricant from canola lubricants were prepared from canola oil and canola biodiesel
biodiesel, which can be due to the presence of intra molecular using Amberlite IR-120H for the epoxidation step and
hydrogen bonding. Thermogravimetric analysis (TGA) also Amberlyst-15 for the simultaneous ring opening and ester-
confirms that canola biolubricant can be more useful for high- ification step. NMR and FT-IR spectral analyses confirmed the
temperature applications, whereas biodiesel biolubricant is formation of products. Biolubricant from canola biodiesel has
useful for low-temperature applications. good low temperatue properties (−12 °C and −18 °C),
Differential Scanning Calorimetry (DSC) Analysis. kinematic viscosity (19 cSt at 100 °C), low friction and
Figure 8 represents a DSC thermogram of epoxidized canola antiwear scar (95 μm), and high oxidative stability (76.3 h).
oil, epoxidized canola biodiesel, canola biolubricant, and Therefore, biolubricant derived from biodiesel can be used for
biodiesel biolubricant. The exothermic crystallization peaks of low temperature and general application like automotive
epoxy canola oil and epoxy canola biodiesel appeared at −7 and lubricants. However, canola oil biolubricant has a cloud point
−12 °C, respectively, because epoxy canola oil is well-stacked of −3 °C and pour point of −9 °C, kinematic viscosity
due to intramolecular forces.25 Epoxy canola biodiesel measured at 373 K is 670 cSt, friction and antiwear scar (130
crystallizes at lower temperature because of the presence of μm), and oxidative stability (56.1 h). As biolubricant from
the cis epoxy linkage.26 Biolubricant from canola oil shows the canola oil has high thermal stability (305 °C), and is highly
peak at −118 °C, and the biolubricant from biodiesel showed viscous, even at high temperature, it can be used at high
no peak for exothermic crystallization up to −140 °C, which temperature, heavy load, and high speed operations such as
F DOI: 10.1021/jf505825k
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

Figure 8. DSC thermogram of (A) epoxidized canola oil, (B) epoxidized canola biodiesel, (C) canola biolubricant, and (D) biodiesel biolubricant.

heavy machinery. This study shows that both the biolubricants (11) Battersby, N. S.; Pack, S. E.; Watkinson, R. J. A correlation
are effective, attractive, renewable, and thus have good between the biodegradability of oil products in the CEC L-33-T-82
characteristics as industrial lubricants. and modified sturm tests. Chemosphere 1992, 24, 1989−2000.


(12) Mungroo, R.; Pradhan, N. C.; Goud, V. V.; Dalai, A. K.
AUTHOR INFORMATION Epoxidation of canola oil with hydrogen peroxide catalyzed by acidic
Corresponding Author ion exchange resin. J. Am. Oil Chem. Soc. 2008, 85, 887−896.
(13) Johansson, L. E.; Lundin, S. T. Copper catalysts in the selective
*E-mail: ajay.dalai@usask.ca. Tel: +1 306 966 4771. Fax: +1
hydrogenation of soybean and rapeseed oils: I. the activity of the
306 966 4777.
copper chromite catalyst. J. Am. Oil Chem. Soc. 1979, 56, 974−980.
Funding (14) Sinadinovic-Fiser, S.; Jankovic, M.; Petrovic, Z. S. Kinetics of in
The authors acknowledge Saskatchewan Agricultural Develop- situ epoxidation of soybean oil in bulk catalyzed by ion exchange resin.
ment Fund (ADF) for providing funding for this research. J. Am. Oil Chem. Soc. 2001, 78, 725−731.
Notes (15) Adhvaryu, A.; Erhan, S. Z. Epoxidized soybean oil as a potential
The authors declare no competing financial interest. source of high-temperature lubricants. Ind. Crops Prod. 2002, 15, 247−


254.
REFERENCES (16) Goud, V. V.; Patwardhan, A. V.; Pradhan, N. C. Studies on the
epoxidation of mahua oil (Madhumicaindica) by hydrogen peroxide.
(1) Vasiliadou, E.; Lemonidou, A. Parameters affecting the formation
of 1,2-propanediol from glycerol over Ru/SiO2 catalyst. Org. Process Bioresour. Technol. 2006, 97, 1365−1371.
Res. Dev. 2011, 15, 925−931. (17) Dinda, S.; Patwardhan, A. V.; Goud, V. V.; Pradhan, N. C.
(2) Mang, T.; Dresel, W. Lubricants and Lubrication, 2nd completely Epoxidation of cottonseed oil by aqueous hydrogen peroxide catalysed
revised and extended ed.; Wiley-VCH: Weinheim, Germany, 2007. by liquid inorganic acids. Bioresour. Technol. 2008, 99, 3737−3744.
(3) Amani, M. A.; Davoudi, M. S.; Tahvildari, K.; Nabavi, S. M.; (18) Meshram, P. D.; Puri, R. G.; Patil, H. V. Epoxidation of wild
Davoudi, M. S. Biodiesel production from Phoenix dactylifera as a new safflower (Carthamusoxyacantha) oil with peroxy acid in presence of
feedstock. Ind. Crops Prod. 2013, 43, 40−43. strongly acidic cation exchange resin IR-122 as catalyst. Int. J.
(4) Baroi, C.; Dalai, A. K. Simultaneous esterification, trans- ChemTech Res. 2011, 3, 1152−1163.
esterification and chlorophyll removal from green seed canola oil (19) Hwang, H. S.; Erhan, S. Z. Modification of epoxidized soybean
using solid acid catalysts. Catal. Today 2013, 207, 74−85. oil for lubricant formulations with improved oxidative stability and low
(5) Peterson, C. L. Vegetable oil as a diesel fuel: Status and research pour point. J. Am. Oil Chem. Soc. 2001, 78, 1179−1184.
priorities. Trans. ASAE 1986, 29, 1413−1422. (20) Adhvaryu, A.; Liu, Z. S.; Erhan, S. Z. Synthesis of novel
(6) Sharma, R. V.; Dalai, A. K. Synthesis of bio-lubricant from epoxy alkoxylated triacyl glycerols and their lubricant base oil properties. Ind.
canola oil using sulfated Ti-SBA-15 catalyst. Appl. Catal., B 2013, 142, Crops Prod. 2005, 21, 113−119.
604−614. (21) Salimon, J.; Salih, N.; Yousif, E. Biolubricants: Raw materials,
(7) Kulkarni, M. G.; Dalai, A. K.; Bakhshi, N. N. Utilization of green chemical modifications and environmental benefits. Eur. J. Lipid Sci.
seed canola oil for biodiesel production. J. Chem. Technol. Biotechnol.
Technol. 2010, 112, 519−530.
2006, 81, 1886−1893.
(22) Fadhel, A. Z.; Pollet, P.; Liotta, C. L.; Eckert, C. A. Combining
(8) Lathi, P. S.; Mattiasson, B. Green approach for the preparation of
biodegradable lubricant base stock from epoxidized oil. Appl. Catal., B the benefits of homogeneous and heterogeneous catalysis with tunable
2007, 69, 207−212. solvents and near critical water. Molecules 2010, 15, 8400−8424.
(9) Matthew, T. S.; Nader, S.; Bigyan, A.; Lambert, A. D. Influence of (23) Sharma, R. V.; Soni, K. K.; Dalai, A. K. Preparation,
fatty acid composition on the tribological performance of two characterization and application of sulfated Ti-SBA-15 catalyst for
vegetable based lubricants. J. Synth. Lubr. 2007, 24, 101−110. oxidation of benzyl alcohol to benzaldehyde. Catal. Commun. 2012, 29,
(10) Campanella, A.; Rustoy, E.; Baldessari, A.; Baltanas, M. A. 87−91.
Lubricants from chemically modified vegetable oils. Bioresour. Technol. (24) Vlcek, T.; Petrovic, Z. S. Optimization of the chemoenzymatic
2010, 101, 245−254. epoxidation of soybean oil. J. Am. Oil Chem. Soc. 2006, 83, 247−252.

G DOI: 10.1021/jf505825k
J. Agric. Food Chem. XXXX, XXX, XXX−XXX
Journal of Agricultural and Food Chemistry Article

(25) Zaher, F. A.; El-Kinawy, O. S.; Abdullah, R. The esterification of


Jatropha oil using different short chain alcohols to produce esters to be
used as biodiesel fuel. Energy Source 2012, 34, 2214−2219.
(26) Sripada, P. K.; Sharma, R. V.; Dalai, A. K. Comparative study of
tribological properties of trimethylopropane-based biolubricants
derived from methyl oleate and canola biodiesel. Ind. Crop. Prod.
2013, 50, 95−103.
(27) Dixit, S.; Kanakraj, S.; Rehman, A. Linseed oil as a potential
resource for bio-diesel: A review. Renew. Sust. Energy Rev. 2012, 16,
4415−4421.
(28) Pedersen, J. R.; Ingemarsson, A.; Olsson, J. O. Oxidation of
rapeseed oil, rapeseed methyl ester (RME) and diesel fuel studied with
GS/MS. Chemosphere 1999, 38, 2467.

H DOI: 10.1021/jf505825k
J. Agric. Food Chem. XXXX, XXX, XXX−XXX

You might also like