You are on page 1of 10

Fuel Processing Technology 106 (2013) 366–375

Contents lists available at SciVerse ScienceDirect

Fuel Processing Technology


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

Factors affecting oxidation stability of commercially available biodiesel products


Zeyu Yang a, b,⁎, Bruce P. Hollebone b, Zhendi Wang b, Chun Yang b, Mike Landriault b
a
Key Laboratory of Catalysis and Materials Science of the State Ethnic Affairs Commission & Ministry of Education, College of Chemistry and Materials Science,
South-Central University for Nationalities, Wuhan, 430,074, China
b
Emergencies Science and Technology Section (ESTS), Science and Technology Branch, Environment Canada, 335 River Road, Ottawa, ON, Canada K1A 0H3

a r t i c l e i n f o a b s t r a c t

Article history: Factors affecting oxidation stability for several commercially available biodiesels were primarily investigated
Received 15 August 2011 by acid value (AV) and induction period (IP) evaluations in this study. It was found that the measured IP at
Received in revised form 31 August 2012 different storage time points were somewhat dependent on the saturated degree of fatty acid methyl esters
Accepted 1 September 2012
(FAMEs), the corresponding measured AV scattered randomly. Generally, AV increased and IP decreased after
Available online 20 September 2012
one year of storing in a dark cold room in an air‐tight tank. Solvents (methanol, acetone and water) did not
Keywords:
show a contribution to altering IP. Metals (copper and lead) showed the strongest detrimental effects to ox-
Acid value idative stability although somewhat depending on the particle size and oxide coating thickness, however,
Induction period aluminum alloy and steel were not the case. Among the antioxidants, pyrogallol (PY) was the best in enhanc-
Oxidative stability ing IP with a concentration of less than 3000 ppm, however, tert-butylhydroquinone (TBHQ) was the best
Biodiesel after 3000 ppm, followed by propyl gallate (PG), butylated hydroxyanisole (BHA), 3,5-di-tert-butyl-4-
Metal contaminants hydroxyltoluene (BHT), and α-tocopherol. The appropriate dosage of PY was also evaluated to achieve the
Antioxidants specified IP regulated by EN-14112 for samples with copper or lead contamination.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction they reflect the change of fuel properties more directly compared with
the variation of chemical composition.
Biodiesel as an alternative diesel fuel consisting of alkyl mono- Rancimat test is the specified standard method for oxidation sta-
esters of fatty acids prepared from vegetable oils or animal fats, has bility testing for biodiesel sample in accordance to EN-14112. The bio-
attracted more and more attention because it is renewable and diesel standard EN-14214 calls for determining oxidative stability at
environmental-friendly. However, biodiesel can be very sensitive to 110 °C with a minimum induction time of 6 h by the Rancimat meth-
oxidative and thermal degradation due to its ester chemical structure. od (EN-14112) [4]. American Standard Test Method (ASTM) standard
The first problem of biodiesel is that it undergoes oxidative degrada- D-6751-08 recently introduced a minimum induction period of 3 h
tion over time due to the presence of a significant amount of fatty for biodiesel samples. However, biodiesels sourced from vegetable
acids with double bonds. Storage stability of a fuel is its ability to re- oil are abundant with fatty acids with more than one double bond,
sist physical and chemical changes brought about by its action with their IP values are very difficult to meet the EN-14112 and ASTM
the environment. The problems arising from the deterioration of the D-6751 limits [5–7] unless antioxidants are present.
fuel properties of biodiesel during storage are expected to be more There are numerous publications on the storage oxidation stability
severe than diesel fuel. The oxidation can result in the formation of and factors affecting the stability of various biodiesels [5,6,8–15]. The
corrosive short chained acids and deposits that may cause increased storage problems can be caused by the chemical composition of bio-
wearing of engine fuel pumps. diesel and the storage conditions [6,8,16]. Firstly, it is well known
Currently, the deterioration of biodiesels under different storage con- that fatty acid methyl esters (FAMEs) are the main chemical compo-
ditions has been investigated by evaluating several commonly-used phys- sition for biodiesel. The molecular structure of FAMEs can vary with
icochemical parameters, such as, peroxide values (PV), acid value (AV), feedstocks by chain lengths, levels of unsaturation, and conformation.
iodine value (IV) viscosity (ν), and insoluble impurities (II) [1–3], as Some parameters have been evaluated to represent oxidative stability
or physical property of biodiesel based on FAME composition, espe-
cially using unsaturated FAMEs [6]. For example, Knothe and Dunn
[17] reported the good correlation between IP and bis-allylic position
⁎ Corresponding author at: Key Laboratory of Catalysis and Materials Science of the equivalent (BAPE, the number of doubly allylic carbons present in the
State Ethnic Affairs Commission & Ministry of Education, College of Chemistry and Ma-
terials Science, South-Central University for Nationalities, Wuhan, 430074, China.
fatty oil or ester), but not for allylic position equivalent (APE, the
Tel.: + 86 27 59880491. number of single allylic carbons present in the fatty oil or ester).
E-mail address: yangzeyu72@163.com (Z. Yang). Another parameter concerned with FAME structure was expressed

0378-3820/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.fuproc.2012.09.001
Z. Yang et al. / Fuel Processing Technology 106 (2013) 366–375 367

as oxidizability (OX) by Neff et al. [18]. OX was expressed as: OX = 2. Experimental


[0.02(%O) + (%L) + 2(%Ln)] / 100, where O refers to methyl oleate
(cis-9, C18:1, ME), L refers to methyl linoleate (cis-9, 12, C18:2, ME) 2.1. Materials and chemicals
and Ln refers to methyl linolenate (cis-9, 12, 15, C18:3, ME).
Secondly, the presence of antioxidants usually prevents radical Various metals were used to test their impact on induction periods.
formation in oils and subsequent degradation. Based on this theory, They include NIST SRM 8k (Bessemer Steel (Simulated, chip form), 0.1%
some synthetic antioxidants, such as pyrogallol, 3,4,5-trihydroxybenzoic Carbon), NIST SRM 1240c (Aluminum Alloy 3004, chip and disk form),
acid propyl ester (propyl gallate), tert-butyl hydroquinone, tert-butyl lead powder (~325 mesh), and copper granule (−10 to +40 mesh)
hydroxyanisole, tert-butyl methyl-phenol, butylated hydroxytoluene (Bellefonte, PA, USA). Both NIST SRM standards were purchased from
(BHT), and some natural antioxidants (e.g., tocopherols), have been the National Institute of Standard and Technology (NIST, Gaithersburg,
used to significantly slow down, but not entirely prevent the oxidation MD, USA), their elemental composition is known but no size informa-
of various biodiesels [3,10,15,19–21] because these chemicals can only tion is available. Additionally, metals with different particle sizes, in-
inhibit the oxidation process. cluding lead shot (~1 cm), copper shot (2−5 cm), and aluminum
Thirdly, the presence of impurities (e.g., transitional metals: iron, powder (~325 mesh) purchased from Sigma-Aldrich, were utilized to
nickel, manganese, cobalt, and copper) and degradation products compare the variation of oxidative stability of biodiesels with different
can catalyze the formation of radicals in oils, while other impurities metal particle sizes.
can catalyze the degradation pathways of oil once radicals are formed Biodiesel samples were obtained from domestic Canadian manufac-
[5,12]. tures, e.g., Milligans Bio-Tech. Inc. (Calgery, AB, Canada). Three of them
For storage conditions like exposure to air and/or light, tempera- were sourced from tallow fat (Ban), three of them were from soybean
ture can accelerate the degradation of biodiesel [5,20,22]. Water con- oil (Bsoy), and three of them are produced from canola oil (Bca), please
tent in biodiesel will enhance biodiesel degradation due to hydrolysis see Table 3 for detailed information. Where Bsoy-2, Bca-1 and Ban-1 were
but its effect is much less than the above two factors. sampled in December 2006, all other samples were sampled in Oct. to
However, these researches mainly focused on home-made biodiesels. Dec. 2009. These biodiesel products were stored in 10 L air‐tight steel
For commercial biodiesel products, few researches reported the influence tanks and kept in a dark cold room (−4 °C). The initial physico-
of storage parameters (storage time, the presence of water, methanol, ac- chemical properties for five samples sampled in 2009 are shown in
etone, steel, aluminum alloy and lead) on oxidative stability. The correla- Table 1 (data were not available for other four samples). It is obvious
tion between oxidative stability and FAME composition for commercial that the initial quality for these five samples, except for Bsoy-1, satisfied
products was rarely reported either. with ASTM D6751 limits. For Bsoy-1, the initial induction period is
The objective of the present study is to select nine commercial bio- shown to be 2.3 h; the mass of total glycerine is 0.263%, which exceeded
diesel products sourced from three sources (soybean, canola oil, and the maximum value of 0.24%; the mass of monoglycerides, diglycerides,
animal fat) to study the effect of storage parameters on their oxida- and triglycerides is 0.608, 0.382, and 0.395%, respectively, which are far
tive stability. The variation of acid values and induction periods higher than the other four samples. The concentration of antioxidants
with storage time will be investigated. The correlation between the was not available; the chemical composition of FAMEs with a mass
measured IP and FAME composition will be discussed. The impact of ratio higher than 1% for all target samples is discussed in Section 3.1.
several solvents (methanol, water and acetone) and metals (copper, Antioxidants including α-tocopherol, propyl gallate (PG), butylated
lead, aluminum alloy and steel) on IP is to be evaluated. Several hydroxyanisole (BHA), 3,5-di-tert-butyl-4-hydorxytoluene (BHT),
commonly used antioxidants to enhance the oxidative stability will pyrogallol (PY), and tert-butylhydroxyquinone (TBHQ) were obtained
be compared for biodiesel with poor quality and simulated metal from Sigma-Aldrich (Bellefonte, PA, USA). Their chemical structures
contaminated products. and their physico-chemical properties are listed in Fig. 1 and Table 2,

Table 1
Initial physico-chemical properties for several representative samples.

Property (units) Test method ASTM D6751-08 limits Sample number

Bsoy-1 Bsoy-3 Bca-2 Bca-3 Ban-2

Acid value, mg KOH/g ASTM D664 0.31 0.34 0.15 0.09 0.18
Ash, sulfated, mass% ASTM D874 Max. 0.02 0.001 0 0 0 0
Determination of carbon, hydrogen, and nitrogen ASTM D5291
Carbon, mass% 76.86 77.1 77.28 77.15 76.24
Hydrogen, mass% 12.01 12.03 12.14 12.1 12.54
Cetane number ASTM D613 Min. 47 57.2 54 54.6 56.7 63.7
Cloud point, °C ASTM D5773 8.4 6.6 -1.1 -1.3 13.8
Copper corrosion ASTM D130 Max. 3 1a 1a 1a 1a 1a
Flash point, °C ASTM D93, procedure A Min. 130 144 / / 162 174
Free glycerin, mass% ASTM D6584 Max. 0.02 0.007 0.002 0.005 0.004 0.006
Total glycerin, mass% ASTM D6584 Max. 0.24 0.263 0.061 0.027 0.082 0.151
Monoglycerides, mass% ASTM D6584 0.608 0.055 0.055 0.219 0.404
Diglycerides, mass% ASTM D6584 0.382 0.177 0.043 0.075 0.225
Triglycerides, mass% ASTM D6584 0.395 0.183 0.022 0.093 0.065
Density, kg/m3, at 15 °C ASTM D4052 883.1 882.5 883.8 883.8 875.7
Specific gravity, kg/L, 60/60 °F ASTM D4052 0.8836 0.8843 0.8762
Phosphorus (P) EN 14107 modified Max. 0.001 b2.0 b2.0 b2.0 b2.0 b2.0
Carbon residue, mass% ASTM D4530 Max. 0.05 0 0.002 0.001 0.002 0
Total nitrogen, mg/L ASTM D4629 26.46 29 14 12.18 63.49
Oxidation stability, 110 °C, hours EN 14112 Min. 3 2.3 5.6 5.3 5 11.8
Total sulfur by ultraviolet fluorescence, mg/kg ASTM D5453 Max. 50000 5.7 3.2 8.6 7.5 15
Kinematic viscosity, mm2/s (cSt) clear, at 40 °C ASTM D445 1.9–6.0 4.298 / / 4.474 4.758
Water and sediment, volume% ASTM D2709 Max. 0.05 0.005 b0.005 0.005 b0.005 b0.0
368 Z. Yang et al. / Fuel Processing Technology 106 (2013) 366–375

OH
OH
HO OH
HO OH

C=O
OC3H7
Pyrogallol (PY)
Propyl gallate (PG)

OH CH3
H3C OH CH3
CH3
H3C CH3
C C C

H3C
CH3 CH3

OCH3 CH3

Butylated hydroxyanisole (BHA) 3, 5-di-tert-butyl-4-hydroxytoluene (BHT)

OH
H3C
CH3
C

CH3

OH
tert-butylhydroxyquinone (TBHQ)

CH3

HO

CH3 O

CH3
a-tocopherol

Fig. 1. Chemical structures of antioxidants used in the present study.

respectively. All these antioxidants are phenolic types of antioxidants, was programmed with the following settings: initial temperature of
five of them are sourced from chemical synthesis, only α-tocopherol 50 °C, held for 1 min, then increased at a rate of 7 °C/min to 185 °C,
can be sourced from nature. All antioxidants were dissolved in target held for 10 min, then increased at a rate of 15 °C/min to 230 °C, with
biodiesel with about 0.2 mL of acetone. The addition of acetone did a final hold for 5 min. The temperatures of injector, transfer line, ion
not show any alternation of IP values for the target biodiesel samples source, and MS quadrupole analyzer were held at 280, 280, 230, and
(please see Section 3.3). 150 °C, respectively. Samples were injected in splitless mode with
helium as a carrier gas at the flow rate of 1 mL/min. The mass-
2.2. Analyses selective detector (MSD) was operated at an electron impact mode
(70 eV) in selected ion monitoring (SIM) mode.
2.2.1. GC/MS analysis
The fatty acid methyl ester composition of all tested biodiesels was 2.2.2. Determination of oxidative stability
determined by an Agilent 6890 GC system interfaced to an Agilent Oxidative stability of biodiesel in the presence of different impuri-
5973 mass spectrometer by triplication. In brief, a DB-225 MS GC col- ties (solvents, elemental metals, antioxidants, and metals plus antiox-
umn (30 m × 0.25 mm i. d., film thickness: 0.25 μm) obtained from idants) was studied in Rancimat equipment model 743 (Herisau,
Agilent (Mississauga, ON, Canada) was employed to separate FAME Switzerland) according to the biodiesel standard method EN-14214.
compounds. The Supelco 37 component FAME mixture with cis- and In the Rancimat method, a sample of 3 g was heated under a heating
trans-isomers (Bellefonte, PA, USA) had good separation efficiency on block temperature of 110 °C, with a temperature correction factor ΔT
the DB-225 MS column. An internal calibration method was used to de- to be set to 1.5 °C as recommended by the test method. A constant air
termine all target analytes in the present study. The oven temperature flow of 10 L/h passed through the heated samples. The volatile
Z. Yang et al. / Fuel Processing Technology 106 (2013) 366–375 369

Table 2
Physico-chemical properties of used antioxidants in the present study.

Antioxidants CAS number Molecular formula Molecular weight (g/mol) Water solubility Melting point (°C) Boiling point (°C)

α-Tocopherol 59-02-9 C29H50O2 430.71 Insoluble 2.5−3.5 200−220 (0.1 mm Hg)


Propyl gallate 121-79-9 C10H12O5 212.2 / 150 Decompose
Butylated hydroxyanisole 25103-16-5 C11H16O2 180.24 Insoluble 48−55 264−270
3,5-di-tert-butyl-4-hydorxytoluene 128-37-0 C15H24O 220.35 1.1 mg/L (20 °C) 70−73 265
Pyrogallol 87-66-1 C6H6O3 126.11 / 131−134 309
Tert-butylhydroxyquinone 1948-33-0 C10H14O2 166.22 Slightly soluble 127−129 273

carboxylic acid produced during the oxidation process released into indicator for titration. About 20 g of biodiesel sample together with
the vessel containing 60 mL of ultra pure water together with the 0.5 mL of 10 g/L of p-naphtholbenzein indicator in titration solvent
air, and absorbed in the water. The variation of conductivity of was dissolved in a 100 mL titration solvent. The standardized potassi-
water with time was monitored by an electrode, the point that the um hydroxide solution was used for the titration until the appearance
conductivity begins to increase rapidly is designated as the measured of the first permanent green color persisting for at least 15 s. Each
IP. Data for all analytical measurements are means of triplication. sample was analyzed for at least three times for the determination
of average acid values.
2.2.3. Acid value determination
Acid value, as described by ASTM D974-04, measures the number 3. Results and discussion
of acidic functional groups in a sample by titration with potassium
hydroxide. The titration solvent was prepared by mixing toluene, 3.1. Analysis of biodiesel samples
water, and anhydrous isopropyl alcohol in the volume ratio of
100:1:99. Phenolphthalein indicator solution (0.1% phenolphthalein Table 3 lists the fatty acid methyl ester (FAME) composition with
in a 1:1 mixture of water, free of CO2, and ethanol) was used as >1% mass percentage, acid value (expressed as mg KOH/g sample),

Table 3
Fatty acid methyl ester composition (%, mass/mass), acid value and induction periods (n = 3) of tested biodiesel samples.

Biodiesel feedstock Soybean oil Canola oil Tallow fat

Sampling time DEC10,2009 NOV24,2006 OCT11,2009 NOV242006 OCT11,2009 DEC10,2009 NOV24,2006 DEC10,2009 DEC23,2009

Biodiesel number Bsoy-1 Bsoy-2 Bsoy-3 Bca-1 Bca-2 Bca-3 Ban-1 Ban-2 Ban-3

FAME compounds Abbreviationsa Average SD Average SD Average SD Average SD Average SD Average SD Average SD Average SD Average SD

Methyl myristate C14:0, ME 0.6 0.0 0.5 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.5 0.0 1.5 0.0 1.4 0.0
Methyl palmitate C16:0, ME 13.6 0.3 9.1 0.0 8.3 0.0 3.3 0.0 3.3 0.0 3.4 0.0 18.7 0.1 20.4 0.0 19.8 0.0
Methyl palmitoleate cis-9, C16:1, 0.9 0.0 1.0 0.0 0.1 0.0 0.2 0.0 0.2 0.0 0.2 0.0 2.0 0.0 2.4 0.0 2.4 0.0
ME
Methyl stearate C18:0, ME 6.0 0.2 2.7 0.0 2.8 0.0 1.2 0.0 1.3 0.0 1.3 0.0 11.2 0.0 10.9 0.1 11.1 0.0
Methyl oleate cis-9, C18:1, 22.9 0.4 32.2 0.3 24.7 0.1 60.3 0.6 62.8 0.8 64.2 0.7 38.0 0.1 39.9 0.3 41.6 0.1
ME
cis-10, octadecenate cis-10,C18:1, 1.8 0.1 1.5 0.1 1.2 0.0 3.6 0.4 3.6 0.0 3.4 0.4 2.2 0.0 2.2 0.1 1.9 0.0
ME
cis-11, octadecenate cis-11,C18:1, 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.1 0.2
ME
Methyl linoleate cis-9,12 C18:2, 29.4 0.9 41.8 0.2 50.8 0.5 19.4 0.1 19.2 0.0 19.4 0.1 6.9 0.0 8.2 0.1 9.4 0.0
ME
Methyl linolenate cis-9,12,15, 4.4 0.1 5.1 0.1 6.6 0.0 10.5 0.0 7.4 0.1 7.5 0.0 0.6 0.0 0.5 0.0 0.7 0.0
C18:3,ME
Methyl arachidate C20:0, ME 0.1 0.0 0.2 0.0 0.2 0.0 0.4 0.0 0.4 0.0 0.4 0.0 0.1 0.0 0.1 0.0 0.1 0.0
Methyl erucate cis-11,C20:1, 0.2 0.0 0.3 0.0 0.1 0.0 1.1 0.1 1.1 0.0 1.1 0.0 0.3 0.0 0.3 0.0 0.4 0.0
ME
Methyl behenate C22:0, ME 0.2 0.0 0.1 0.0 0.2 0.0 0.2 0.0 0.2 0.0 0.2 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Methyl lignocerate C24:0, ME 0.0 0.0 0.0 0.0 0.1 0.0 0.0 0.0 0.1 0.0 0.1 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Total FAME (%) 82.7 4.2 95.2 0.5 95.2 0.7 100.5 0.5 99.9 0.7 101.6 0.3 83.5 0.0 88.2 0.3 90.4 0.4
∑SFAME 20.4 0.1 12.9 0.2 11.7 0.1 5.1 0.0 5.4 0.0 5.5 0.0 32.4 0.1 33.6 0.1 32.9 0.1
∑MUFAME 28.1 0.2 35.2 0.2 26.0 0.1 65.4 0.5 67.8 0.8 69.0 0.3 43.2 0.1 45.5 0.3 47.1 0.4
∑PUFAME 36.5 0.4 47.1 0.3 57.5 0.5 30.0 0.2 26.7 0.1 27.1 0.1 7.9 0.0 9.1 0.1 10.4 0.0
Acid value 0.31 / 0.3 0.1 0.3 / 0.3 0.0 0.15 / 0.09 / 0.45 0.1 0.18 / 0.3 0.1
(mg KOH/g
sample)b
Acid value 0.23 0.0 0.6 0.1 0.6 0.1 0.4 0.0 0.2 0.0 0.2 0.0 0.6 0.1 0.3 0.0 0.4 0.1
(mg KOH/g
sample)c
IP (hours)b 2.3 / 4.6 0.1 5.6 / 5.1 0.1 5.2 / 5.4 / 11.1 0.2 11.8 / 9.6 0.2
IP (hours)c 0.7 0.1 4.8 0.0 4.7 0.1 5.0 0.0 4.8 0.1 5.2 0.0 10.8 0.1 10.0 0.2 9.2 0.4
Kinematic viscosity, 4.298 / / / / / / / / / / / 4.474 / 4.758 / / /
mm2/s (cSt) at
40 °C

Notes:
a
The abbreviations of FAME compounds based on fatty acid nomenclature. ∑SFAME indicates the sum of all saturated FAMEs; ∑MUFAME indicates the sum of FAMEs with one
double bond; ∑PUFAME indicates the sum of FAMEs with more than one double bond;
b
Data were measured at DEC, 2009;
c
Data were measured at Dec, 2010.
370 Z. Yang et al. / Fuel Processing Technology 106 (2013) 366–375

and induction periods (hours) of the target biodiesel products. Two se- storage, they ranged from 4.8 to 5.4 h for Bca samples, 4.6–5.6 h for
ries of acid values and induction period values were measured at Dec. Bsoy-2 and Bsoy-3 samples. It is clear that IP generally decreased after a
2009, and Dec. 2010, respectively, to investigate the effect of storage one year storage, especially for Bsoy-1, its IP decreased from 2.3 to 0.7.
time on the physico-chemical property of biodiesel products. It can be Ban samples at any measured points meet the EN-14214 minimum
seen that FAME carbon number of 18 was the most abundant one re- limit of 6 h. All other samples, except for Bsoy-1, meet the ASTM
gardless of the varied saturation degree, followed by carbon number D-6751 minimum limit of 3 h. That is to say, the oxidative stability of
of 16, 14 and 20. This is in good agreement with many previous reports Ban is higher than the Bca and Bsoy samples. Similar IP ranges were ob-
[23,24]. In both Ban and Bca samples, methyl oleate (cis-9, C18:1, ME) served among Bsoy (except for Bsoy-1) and Bca samples. It seems that
was the most abundant composition (38−64%, mass/mass), but in the relative higher contents of saturated FAMEs (∑SFAME) counteract
Bsoy samples, methyl linoleate (cis-9, 12, C18:2, ME) was the most abun- with the negative influence of ∑PUFAME on oxidative stability in most
dant one (29−51%). The saturated FAME in animal fat sourced biodiesel Bsoy samples, which makes most Bsoy samples to have similar IP with Bca
was the highest (about 33%), followed by Bsoy (12%−20%), and Bca (7%). samples. It is well known that even small amounts of poly-unsaturated
Biodiesels from same feedstocks showed similar FAME distribution ex- fatty compounds containing bis-allylic carbons are supposed to strongly
cept for Bsoy-1 with a relatively higher saturated FAMEs, and less mono-, decrease the IP values [17]. In conclusion, AV generally increased, and IP
poly-unsaturated, and total FAMEs compared with Bsoy-2 and Bsoy-3 generally decreased after storing for about one year.
samples. The total FAME mass percentages in all these samples were de- The exception of Bsoy-1 can be explained by many assumptions.
termined to be in the range of 83−100%. The discrepancy of the total Firstly, free fatty acids have been shown to have a significant effect on
FAME contents depends on two aspects, the initial quality and the stor- the oxidizability of oil [17]. Miyashita and Takagi [25] have compared
age time. Because all these samples were not fresh biodiesels, it is the oxidation of oleic (cis-9, C18:1), linoleic (cis-9, 12, C18:2) and
possible that some polyunsaturated FAMEs have been degraded or po- linolenic (cis-9, 12, 15, C18:3) acids with their corresponding methyl
lymerized. For Bsoy-1, headspace gas chromatography has been used esters, the free acids were found to be far more vulnerable to be oxi-
to analyze the methanol residue, however, we did not find obvious de- dized. However, the measured acid values for the Bsoy-1 were lower
tectable methanol. As mentioned in Section 2.1, the mass concentra- than the other Bsoy samples, indicating the absence of a high concentra-
tions of glycerin and glycerides are relatively high compared with tion of free fatty acids. Secondly, the effect of the addition of solvents
other four samples with initial properties. This may partly contribute (methanol, acetone, and water) into biodiesels on IP values was tested
to the low FAME contents for Bsoy-1. to evaluate their contribution to the quite low IP for Bsoy-1 sample, how-
ever, the IP values did not show obvious variation (data not shown
3.2. Influence of storage time on acid value and induction periods here) when altering the solvent volume ratio in the range of 1.7 to
16.7%. Thirdly, as indicated in many previous reports [5,16,17], a variety
The acid values measured at Dec. 2009 and Dec. 2010 were ob- of other factors, for example, exposure to air and/or light, temperatures
served to show some random array regardless of the sources of bio- above ambient, as well as presence of extraneous materials with cata-
diesel, they ranged from 0.09 to 0.45, and from 0.2 to 0.6 mg KOH/g lytic and inhibitory effects on oxidation, possess catalytic or inhibitory
sample, respectively. It is obvious that all acid values measured at effects on oxidative stability. All these products were stored in similar
Dec. 2009 did not exceed the ASTM D6751 limit of 0.5 mg KOH/g air-tight tank in a −4 °C cold room. Therefore, the initial IP of Bsoy-1 is
samples, even for samples sampled in 2006 (e.g., Bsoy-2, Bca-1 and 2.3 h indicating that without antioxidants present in the Bsoy-1 sample.
Ban-1). But almost all acid values increased after a one year storage, It can be concluded that Bsoy-1 was not contaminated by metals. If Bsoy-1
three of them (Bsoy-2, Bsoy-3 and Ban-1) exceeded 0.5 mg KOH/g. was contaminated by metals, the initial IP should be far less than 2.3 h.
The random AV profile regardless of biodiesel sources suggests that
feedstocks show little influence on acid values, especially for fresh 3.3. The correlation between FAME composition and IP values
biodiesel. It is true that good processing procedures can convert free
fatty acids to biodiesel either before or during the transesterification To investigate the dominant factor influencing the IP values for
reaction. The increase of AV with the extension suggests that the stor- various biodiesel products, the relationship of the measured IP at
age time is a main factor contributing to the poor quality of some two measured time points and APE/BAPE for all samples except for
samples. After a one year storage, the AVs of most soya feedstock Bsoy-1 was estimated. It was found that the IP values increased with
fuels were found to be higher than the samples with a high abun- the decrease of APE and BAPE values. However, the value of the
dance of cis-9, C18:1 ME (Table 3). This is because FAMEs with square of correlation coefficient (r 2) is quite low (less than 0.2) for
more than one double bond are more susceptible to oxidation [2] dur- both of them, and the impact of APE on IP values is much less than
ing storage. As seen from Table 3, the sum of FAMEs with more than BAPE. Similar correlation has been achieved by some previous reports
one double bond (∑PUFAME) was the most abundant composition [6,17], their explanation for the low r 2 may be that the amount of nat-
(47−58% by mass) in Bsoy-2 and Bsoy-3. The high abundance of unsta- ural antioxidants is different in different biodiesels, which depends
ble ∑PUFAME could result in the significant increase of acid values on the process of production and the process of distillation of biodie-
during storage. Similarly, some studies showed that the quality of bio- sel. Similarly, all biodiesels in the present study are commercially
diesel over a longer period of storage strongly depended on the tank available products, the concentrations of antioxidants were not avail-
material as well as contact to air or light. They found the increases in able, which may partly contribute to the low coefficients.
viscosities and acid values versus the decreases in induction periods For another parameter of OX, the measured IP generally increased
after storing for a longer time [2,14]. The variation of AV for Bsoy-1 with the decrease of OX. However, the correlation coefficient be-
is out of regular expectation; this may partly result from the experi- tween OX and IP values was less than 0.4. Similar to Jain's conclusion
mental errors in measuring acid values. Additionally, acid values for [6], OX is not a very reliable parameter to denote IP values.
some samples at different storage time points showed some random As all above parameters did not accurately describe the relation-
variation regardless of the FAME composition (e.g., Bca-2, Bca-3, ship between FAME composition and IP, the sum of FAMEs based on
Ban-2, Ban-3). The possible presence of antioxidants in samples may their saturated degree were calculated and listed in Table 3. Based
contribute to the variation of acid values with storing extension ex- on the measured IP at different storage time points and the FAME
cept for FAME composition. composition, it can be qualitatively concluded that IP generally posi-
Similarly, induction periods at different storage time points were tively proportional to the concentration of saturated FAMEs and neg-
also obtained with a one year gap (Table 3). The average IP values atively proportional to the concentration of poly-unsaturated FAMEs.
of Ban samples were determined to be 9.2−11.8 h within a one year The plottings of various FAMEs by saturated degree versus two series
Z. Yang et al. / Fuel Processing Technology 106 (2013) 366–375 371

of measured IP values are shown in Fig. 2. It was found that all IP values, Knothe et al. [17] in 2003 when they tested the effect of copper,
except for Bsoy-1, increased with the increasing of ∑SFAME. The Pear- iron, and nickel on IP of methyl oleate at 90 °C.
son Correlation analysis indicated that both variables positively corre- Currently, the commonly used tank materials for storing biodiesel
late with each other with a correlation coefficient of 0.959 and p include steel and aluminum. In this study, steel and aluminum as stan-
values below 0.050. A decreasing tendency between ∑PFAME versus dard reference material from NIST, were blended with varying concen-
IP was observed; the Pearson Correlation analysis suggested a negative trations in biodiesels from three feedstocks to study the catalytic effects
correlation with a correlation coefficient of −0.914, p b 0.05. However, on IP. The influence of copper and lead was investigated as copper
the Pearson Correlation analysis demonstrated no significant relation- showed a strong catalytic effect on oxidative stability expressed as IP
ships between ∑MUFAME and IP (p> 0.05). Additionally, the linear re- values [17] and little was known for lead. Fig. 3 shows the effect of the
lationship of the measured IP values versus ∑SFAME, and ∑PUFAME above metals with varying concentration on IP values for biodiesels
(except for Bsoy-1) was fitted to mathematically describe the correla- sourced from three feedstocks. The IP values decreased significantly
tion. The positive linear correlation with a square of correlation coeffi- with the addition of copper (−10 to +40 mesh) and lead (~325
cient up to as high as 0.90–0.92 for IP versus ∑SFAME, and the mesh) from 0 to 0.5% by mass for all these three biodiesels, then they
negative linear correlation with a square of correlation coefficient of were almost unaltered with the continuous increasing of these two
0.64–0.67 between IP and ∑PUFAME were observed. The good corre- metals. A very slight decrease of IP was found when steel (SRM 8k)
lation between IP and ∑SFAME and ∑PUFAME indicates that the and aluminum alloy (SRM 1240c) were added. It is clear that small
FAME composition based on saturated degree is one of the controlled amounts of copper or lead had nearly the same influence on the induc-
factor affecting IP values, although the information of antioxidants tion period as larger amounts. Therefore, lead and copper showed sig-
was not available for these samples. nificant catalytic effects on the oxidative stability for all biodiesels.
However, the presence of steel and aluminum alloy did not show obvi-
ous catalytic or inhibited effects. t-Test showed that no significant dif-
3.4. Effect of metal contaminants on oxidation stability ference (p> 0.05) was found between copper and lead, and steel and
aluminum alloy, but significant difference (pb 0.05) was observed
Sarin et al. [5] investigated the effect of several transition metal between copper/lead and steel/aluminum alloy. The dependence of
naphthenates (iron, nickel, manganese, cobalt, and copper) on oxida- the induction period on copper and lead in biodiesel resulted in the ac-
tion stability of their own synthesized Jatropha methyl ester. They celeration of free radical oxidation due to a metal-mediated initiation
found that copper showed the strongest catalytic effect on the reaction [5,24]. The effect of these metals on IP is similar among the dif-
depression of induction time. Similar results have been achieved by ferent biodiesel products, which is independent on biodiesel sources. As

a b
12 12

8 8

4 4
0 5 1 15 20 25 30 35 40 0 5 10 15 20 25 30 35 40
Sum fo SFAME (%, m/m) Sum fo SFAME (%, m/m)
Induction periods (hours)

12 12

8 8

4 4
20 30 40 50 60 70 80 20 30 40 50 60 70 80
Sum of MUFAME (%, m/m) Sum of MUFAME (%, m/m)
12 12

8 8

4 4
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Sum of PUFAME (%, m/m) Sum of PUFAME (%, m/m)

Fig. 2. Correlation of FAME composition vs. IP values. (a) Data obtained at Dec. 2009; (b) data obtained at Dec. 2010.
372 Z. Yang et al. / Fuel Processing Technology 106 (2013) 366–375

6 when traces of water cause hydrolysis of the esters into alcohol and
Bsoy-2 Pb
Cu acids [2], the spiking of water into biodiesel alone and biodiesel with
5 Al lead or copper did not change IP values clearly. It is well known that
Steel
4 the high temperature (110 °C), and the abundant oxygen (10 L/h of
air flow) in the Accelerated Oxidative System accelerate the oxidative
3 rate of FAMEs. Beside oxidation caused by exposure to air (oxygen)
and high temperature, biodiesel is also potentially subject to hydrolytic
2 degradation due to the presence of water. However, the abiotic hydro-
lysis rates of FAME were found to be slower than the microbial degrada-
1 tion rates in seawater matrix, the base-catalyzed hydrolysis half-lives of
FAMEs were estimated to be 7 years at pH 7, and 70 years at pH 6
0
0 2 4 6 8 10 12 14 (25 °C) [26]. This can be used to explain the almost unaltered IP values
12 with the spiking of water into pure biodiesel. Also, the contribution of
Ban-1
water was not beneficial to dissolving metals, or the amount of
Induction period (hours)

10 dissolved metal ions was enough to catalytically oxidize FAMEs even


without the contribution from water.
8
Knothe and Dunn [17] reported that particle size and the thick-
6
ness of oxide coating contributed somewhat to the finally deter-
mined IP values. Herein, some currently produced and long term
4 stored copper and lead with different particle sizes were utilized to
test the impact of particle size and thickness of oxide coating of
2 metal on the oxidative stability of biodiesel. The measured IP values
increased slightly when particle sizes and oxide coating thickness in-
0 creased for copper/lead. However, the catalytic effects for both of
0 2 4 6 8 10 12
6 them were still more significant than the currently produced steel
Bca-1 and aluminum alloy.
5

4 3.6. Improvement of oxidation stability

3 Two types of antioxidants were generally known: chain breakers


and hydroperoxide decomposers [27]. The most commonly used
2
chain breaking antioxidants in improving oxidation stability of bio-
1
diesel are phenolic and amine-types of antioxidants, which can be
sourced from both nature and chemical synthesis [5,6,12]. These anti-
0 oxidants contain a highly labile hydrogen that is more easily abstract-
0 2 4 6 8 10 12 14 ed by a peroxy radical than fatty oil or ester hydrogen. The resulting
Metal concentration (%, mass/mass) antioxidant free radical is either stable or further reacts to form a sta-
ble molecule which is further resistant to a chain oxidation process.
Fig. 3. Effect of various metals on induction periods of biodiesels sourced from three
Thus the chain breaking antioxidants interrupt the oxidation chain
feedstocks.
reaction in order to enhance stability [6].
As mentioned in previous sections, the measured IP of Bsoy-1 was
mentioned in previous sections, the steel and aluminum alloy used in close to 0.7 h in Dec. 2010, which is far less than the EN-14112 and
this study are NIST SRM standards, therefore, steel or aluminum tanks ASTM D-6751 limits of 6 h and 3 h. Therefore, some antioxidants
with similar elemental composition to SRM 8k or 1240c can be used were tested to improve the IP to meet the corresponding limits. The
to store biodiesel without obvious catalytic degradation of biodiesel enhanced efficiency of various antioxidants, including TBHQ, PY, PG,
products. BHA, BHT and α-tocopherol, was evaluated and compared by altering
their concentration ranged from 0 to 8000 ppm. The measured IP for
3.5. Other factors on oxidation stability Bsoy-1 with different antioxidants and spiking concentrations is com-
pared in Fig. 5(a–b).
If lead or copper ions play a major role on the catalysis of the oxida- It can be seen that the IP of Bsoy-1 increased with the increase in dos-
tion of biodiesel, the existence of water is supposed to result in more age of all antioxidants except for α-tocopherol. PY was found to result in
dissolved metal ions due to their high water solubility. In this scenario, the greatest enhancement in IP when the concentration was less than
the synergetic effect of water plus lead or copper in decreasing IP values 3000 ppm, however, TBHQ was the best when the concentration was
should be observed. To verify this assumption, the IP values for higher than 3000 ppm (Fig. 5a), followed by PG, BHA, and BHT, ranked
biodiesels spiked with various concentrations of H2O, lead (~325 from high to low (Fig. 5b). The enhancement performance in IP for nat-
mesh) or copper (−10 to +40 mesh), and H2O plus lead/copper ural antioxidant (α-tocopherol) was not obvious, as α-tocopherol is rel-
were compared (Fig. 4). Similar to the results in former sections, the atively weak in increasing the induction times of biodiesel compared
presence of H2O did not alter the IP values significantly from the spiking with β- and γ-tocopherols [28]. The IP values improved from 0.7 h to
concentration (mass/mass) of 0.05% to 10%, but a slight increase of IP above 6 h as required by the EN-14112 specification when the concen-
values was observed for all samples. The co-contribution of copper/ tration of PY and TBHQ was about 1500 ppm and 3000 ppm, respec-
lead with water did not significantly enhance or decrease the oxidation tively. For PG, the spiking concentration should be higher than
of different biodiesels compared with metals alone, slightly lower or 6000 ppm; for BHA, BHT, and α-tocopherol, this cannot be satisfied
higher IP values were observed in these tests. These slight differences even when their concentrations were as high as 8000 ppm. The weaker
can be a result of experimental errors because t-test showed no signifi- enhancement effectiveness of TBHQ with a lower concentration com-
cant difference between them (p> 0.05). Although acid can be formed pared with that of PY was evaluated by a repeated test. Similar results
Z. Yang et al. / Fuel Processing Technology 106 (2013) 366–375 373

6 6
Bsoy-2
Bsoy-2
5 5

4 H2O 4 H2O
Cu
Pb
5%H2O+Cu
3 3 5%H2O+Pb

2 2

1 1

0 0

6 6
Bca-1 Bca-1
5 5
Induction period (hours)

4 4

3 3

2 2

1 1

0 0

12 12
Ban-1 Ban-1
10 10

8 8

6 6

4 4

2 2

0 0
0 2 4 6 8 10 0 2 4 6 8 10 12
Mass content of spiked water and metals (%, mass/mass)
Fig. 4. Effect of water plus copper/lead on the oxidative stability of different biodiesels.

were obtained (data not shown here), which indicated the good accura- the natural ones [10,29]. Synergistic effects are possible in the case of
cy of the results. co-contribution of different antioxidants [16].
Numerous studies have evaluated the effect of various antioxidants
on the oxidative stability of biodiesel [10,11,15,29]. The effectiveness 3.7. Improving the oxidation stability for metal contaminated biodiesels
of different antioxidants varied in different situations. For example, PY
was reported to be the best sometimes [15], TBHQ [12], PG [29], or Just as the description in Section 3.4, the addition of metals (copper
BHA [10] was also found to be the best performer in other situations, and lead) resulted in the decreasing oxidation stability for all biodiesels
but α-tocopherol was always found to exhibit the least enhancement from different sources. To find the appropriate antioxidant concentra-
effectiveness regardless of biodiesel sources and storage condi- tion for metal contaminated samples meeting the EN-14214 minimum
tions [10,20]. In all, the effectiveness of an antioxidant can depend on limits, metal contaminated samples (Bsoy-2, Bca-1, and Ban-1) were sim-
a variety of factors, including the fatty acid profile of the oil or fat, the ulated by spiking with 5% (mass/mass) of copper or lead. Then the IP for
amount of naturally occurring antioxidants, and storage or other condi- these samples was measured by altering the PY concentration. The data
tions. The performance of synthetic antioxidants is generally better than are plotted in Fig. 6 versus the PY loading amount. It is noted that the IP
374 Z. Yang et al. / Fuel Processing Technology 106 (2013) 366–375

60 16
TBH a Bsoy-2 Bsoy-2
PY Bsoy-2 with 5% Pb
12 Bsoy-2 with 5%Cu
40
8

20 4

0
0
Induction period (hours)

16
Bca-1 Bca-1
Bca-1 with 5% Pb
0 2000 4000 6000 8000 10000

Induction period (hours)


12 Bca-1 with 5% Cu
8
PG
BHA
b
8
BHT
6 a-tocopherol

4
0

16
Ban-1 5%Cu
0 5% Pb

12

-2
0 2000 4000 6000 8000 10000
Antioxidants concentration (ppm) 8

Fig. 5. Influence of antioxidant concentration on the oxidation stability of Bsoy-1.


4
of Ban was higher than 9 (Table 3), so no data is available for pure Ban
with PY. Only Bca-1 was selected as the representative of Bca samples
0
as their IP values and chemical composition are similar to each other. 0 100 200 300 400 500 600
Clearly, IP values increased with the increment of the dosage of PY. Concentration of PY (ppm)
For Bsoy-2, the original IP was 4.8 h; it reached to 6 h when PY concen-
tration was increased to 50 ppm. For Bsoy-2 mixed with 5% Cu or Pb, the Fig. 6. Influence of antioxidant concentration on the oxidation stability of copper and
IP decreased to less than 1 h, they increased with the increasing dosage lead contaminated biodiesels sourced from three feedstocks.
of PY, and reached to 6 h when PY concentration increased to 500 ppm.
For Bca-1, the original IP value was 5.0 h, it reached to 10 h when the PY were investigated. The results showed that the absolute acid values
concentration increased to 50 ppm, so less than 50 ppm of PY can satis- were independent on FAME composition, but the variation of acid
fy the specification of oxidation stability. Similarly, the IP values for Bca values was kind of dependent on the feedstock of biodiesel with the
spiked with 5% Cu or Pb decreased significantly due to the catalytic ef- extension of storage. When oxidative stability was expressed as in-
fects of both metals. When the PY concentration was about 100− duction periods, oxidative stability was positively correlated with
200 ppm, their IP values could satisfy the EN-14112 limit. For Ban spiked ∑SFAME and negatively correlated with ∑ PUFAME in spite of the
with 5% Cu or Pb, their IP decreased from 10 h to close to 4 h. However, possible interference of antioxidants. The acid values generally in-
they increased to above 6 h when the PY concentration was about creased, and induction periods generally decreased after a one year
50 ppm. storage in a dark cold room (− 4 °C) in an air‐tight tank. The presence
The enhancement effectiveness of PY on copper or lead contaminat- of pyrogallol or tert-butylhydroxyquinone showed the strongest en-
ed biodiesels varied with the source of biodiesels. In detail, for copper hancement of induction periods. Copper and lead showed a strong
and lead contaminated Bsoy-2, t-test showed no significant difference catalytic effect on decreasing oxidative stability, but not for steel
(p> 0.05) between two series of data, therefore, almost similar effec- and aluminum alloy. The solvents (methanol, acetone and water)
tiveness was observed. Although visual stronger effectiveness was ob- did not show contribution to altering the IP values. This study
served for lead contaminated Bca-1 and Ban-1 compared with copper shows a promising theoretical guidance for the transportation, stor-
contaminated ones, t-test showed no significant difference (p>0.05). age and usage of biodiesel products.
This visual difference can be partly attributable to experimental errors.
Acknowledgments
4. Conclusions
This work was funded by the Natural Resources Canada, under the
The factors, including solvents, metals, FAME chemical composi- PERD program through POL 2.1.2 Advanced Fuels & Transportation
tion, and some antioxidants, affecting oxidation stability during stor- Emissions Reduction (AFTER), supported by the Fundamental
age for several commercially available biodiesel products in Canada Research Funds for the Central Universities, South-Central University
Z. Yang et al. / Fuel Processing Technology 106 (2013) 366–375 375

for Nationalities (CZZ11006) the Natural Science Foundation of [14] M. Mittelbach, S. Gangl, Long storage stability of biodiesel made from rapeseed and
used frying oil, Journal of the American Oil Chemists' Society 78 (2001) 573–577.
South-Central University for Nationalities (YCZZ11006). Fuel samples [15] M. Mittelbach, R. Schaeffer, The influence of antioxidants on the oxidation stabil-
were generously provided by Debbie Rosenblatt of the Emissions ity of biodiesel, Journal of the American Oil Chemists' Society 80 (2003) 817–823.
Research and Measurement Division of Environment Canada from [16] G. Knothe, Some aspects of biodiesel oxidative stability, Fuel Processing Technol-
ogy 88 (2007) 669–677.
commercial sources. [17] G. Knothe, R.O. Dunn, Dependence of oil stability index of fatty compounds on
their structure and concentration and presence of metals, Journal of the American
References Oil Chemists' Society 80 (2003) 1021–1026.
[18] W.E. Neff, E. Selke, T.L. Mounts, E.N. Rinsch, M.A.M. Zeitoun, Effect of
[1] A. Bouaid, M. Martinez, J. Aracil, Production of biodiesel from bioethanol and triacylglycerol composition and structures on oxidative stability of oils from se-
Brassica carinata oil: oxidation stability study, Bioresource Technology 100 lected soybean germplasm, Journal of the American Oil Chemists' Society 69
(2009) 2234–2239. (1992) 111–118.
[2] A. Bouaid, M. Martinez, J. Aracil, Long storage stability of biodiesel from vegetable [19] P. Bondioli, A. Gasparoli, L.D. Bella, S. Tagliabue, Evaluation of biodiesel storage
and used frying oils, Fuel 86 (2007) 2596–2602. stability using reference methods, European Journal of Lipid Science and Technol-
[3] L.M. Das, D.K. Bora, S. Pradhan, M.K. Naik, S.N. Naik, Long-term storage stability of ogy 104 (2002) 777–784.
biodiesel produced from Karanja oil, Fuel 88 (2009) 2315–2318. [20] Y.C. Liang, C.Y. May, C.S. Foon, M.A. Ngan, C.C. Hock, Y. Basiron, The effect of nat-
[4] EN 14112, Fatty acid methyl esters (FAME)—determination of oxidation stability ural and synthetic antioxidants on the oxidative stability of palm diesel, Fuel 85
(accelerated oxidation test), 2003. (2006) 867–870.
[5] A. Sarin, R. Arora, N.P. Singh, M. Sharma, R.K. Malhotra, Influence of metal contami- [21] L.S. Gang, M. Mittelbach, The impact of antioxidants on biodiesel oxidation stabil-
nants on oxidation stability of Jatropha biodiesel, Energy 34 (2009) 1271–1275. ity, European Journal of Lipid Science and Technology 106 (2004) 382–389.
[6] S. Jain, M.P. Sharma, Stability of biodiesel and its blends: a review, Renewable and [22] D.Y.C. Leung, B.C.P. Koo, Y. Guo, Degradation of biodiesel under different storage
Sustainable Energy Reviews 14 (2010) 667–678. conditions, Bioresource Technology 97 (2006) 250–256.
[7] W. Li, W. Du, D. Liu, Y. Yao, Study on factors influencing stability of whole cell dur- [23] R.M. Balabin, R.Z. Safieva, Biodiesel classification by base stock type (vegetable
ing biodiesel production in solvent-free and tert-butanol system, Biochemical En- oil) using near infrared (NIR) spectroscopy data, Analytica Chimica Acta 689
gineering Journal 41 (2008) 111–115. (2011) 190–197.
[8] G. Karavalakis, S. Stournas, D. Karonis, Evaluation of the oxidation stability of [24] G. Knothe, Dependence of biodiesel fuel properties on the structure of fatty acid
diesel/biodiesel blends, Fuel 89 (2010) 2483–2489. alkyl esters, Fuel Processing Technology 86 (2005) 1059–1070.
[9] R.O. Dunn, Effect of oxidation under accelerated conditions on fuel properties of [25] K. Miyashita, T. Takagi, Study of the oxidative rate and prooxidant activity of free
methyl soyate (biodiesel), Journal of the American Oil Chemists' Society 79 fatty acids, Journal of the American Oil Chemists' Society 63 (1986) 1380–1384.
(2002) 915–920. [26] J.A. DeMello, C.A. Carmichael, E.E. Peacock, R.K. Nelson, J. Samuel Arey, C.M.
[10] R.O. Dunn, Effect of antioxidants on the oxidative stability of methyl soyate Reddy, Biodegradation and environmental behavior of biodiesel mixtures in the
(biodiesel), Fuel Processing Technology 86 (2005) 1071–1085. sea: an initial study, Marine Pollution Bulletin 54 (2007) 894–904.
[11] R.O. Dunn, Oxidative stability of biodiesel by dynamic mode pressurized-differential [27] J. Pospisil, P.P. Klemchuk, Oxidation Inhibition in Organic Materials, CRC Press,
scanning calorimetry (P-DSC), Transactions of the ASABE 49 (2006) 1633–1641. 1990.
[12] A. Sarin, R. Arora, N.P. Singh, R. Sarin, R.K. Malhotra, Oxidation stability of palm [28] A. Fröhlich, S. Schober, The influence of tocopherols on the oxidation stability of
methyl ester: effect of metal contaminants and antioxidants, Energy & Fuels 24 methyl esters, Journal of the American Oil Chemists' Society 84 (2007) 579–585.
(2010) 2652–2656. [29] S.-K. Loh, S.-M. Chew, Y.-M. Choo, Oxidative stability and storage behavior of fatty
[13] R.L. McCormick, M. Ratcliff, L. Moens, R. Lawrence, Several factors affecting the acid methyl esters derived from used palm oil, Journal of the American Oil Chemists'
stability of biodiesel in standard accelerated tests, Fuel Processing Technology Society 83 (2006) 947–952.
88 (2007) 651–657.

You might also like