Sunteți pe pagina 1din 6

Fuel 90 (2011) 773–778

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Evaluation of the oxidative stability of corn biodiesel


M.B. Dantas a, A.R. Albuquerque a, A.K. Barros b, M.G. Rodrigues Filho c, N.R. Antoniosi Filho d,
F.S.M. Sinfrônio e, R. Rosenhaim a, L.E.B. Soledade f, I.M.G. Santos a, A.G. Souza a,⇑
a
Departamento de Química, CCEN, Universidade Federal da Paraíba, João Pessoa, PB, Brazil
b
Departamento de Física, Universidade Federal do Maranhão, São Luís, MA, Brazil
c
Coordenação de Química, CCEN, Universidade Estadual do Piauí, Teresina, PI, Brazil
d
Departamento de Química, Universidade Federal de Goiás, Goiânia,GO, Brazil
e
Pró-Reitoria de Pesquisa e Pós-Graduação, Universidade Federal do Maranhão, São Luís, MA, Brazil
f
Universidade Federal do Maranhão, Campus de Pinheiro, Pinheiro, MA, Brazil

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

Article history: The main factors responsible for the oxidative processes of biodiesel are oxygen, metal traces, high tem-
Received 30 November 2009 perature and the amount of unsaturated fatty acids. The biodiesel quality is harmed by the oxidation
Received in revised form 14 September products, which are corrosive to engine chambers and may lead to clogging of the injection pumps
2010
and filters, besides increasing the biodiesel viscosity. Thus, this work aimed at investigating the oxidative
Accepted 16 September 2010
Available online 10 November 2010
stability of corn biodiesel, obtained by base-catalyzed transesterification reaction, using the ethanol
route. As-synthesized, stored and heated biodiesel samples were characterized by peroxide value, iodine
value and dynamic viscosity. UV/Vis absorption was used to evaluate the oxidative degradation of biodie-
Keywords:
Biodiesel
sel, by means of the 232 and 272 nm absorption peaks, ascribed to double bonds and carbonyl groups,
Oxidative stability respectively. The TG and PDSC thermal analysis techniques were also employed to analyze the oxidative
UV/VIS stability. The oxidative decomposition was confirmed by the increase of the dynamic viscosities. PDSC
Thermal analyses curves showed that the oxidation onset temperature was reduced as the sample was exposed to degra-
dation factors during heating and storage. Such results were in agreement with the peroxide and iodine
values, as well as with dynamic viscosity values.
Ó 2010 Elsevier Ltd. All rights reserved.

1. Introduction petroleum-based fuels [7,8]. However, biodiesel is more suscepti-


ble to auto-oxidation than fossil fuels [7].
Corn oil has been used in the production of biodiesel along with The oxidative stability of biodiesel is affected by numerous fac-
other oil sources, such as soybean, physic nut, cotton and castor tors during its commercialization. As the parent vegetable oil, bio-
oils [1–4]. All of them have been under investigation in order to diesel is vulnerable to the oxidation of its unsaturated fatty acid
improve the productive biodiesel chain. Modifying the alcohol esters [8,9]. Ultraviolet irradiation, metal contamination, humidity,
moiety is another possibility for inherently improving the fuel pigments, some enzymes and heating can alter biodiesel, deterio-
properties of biodiesel [5]. The most frequently employed are the rating its fuel performance when exposed to oxygen [10,11]. Both
short chain alcohols, such as methanol, ethanol, propanol and oxygen content and exposition time play an important role in the
butanol. Ethyl esters are the most likely esters other than methyl formation of undesired compounds, which can corrode the engines
to possess some commercial significance and are the most com- or clog the filters and injection systems of the engines [12]. Thus,
mon form of biodiesel in Brazil [6]. Moreover, the fuel is obtained oxidative stability has to be considered as an essential characteris-
from biomass and, thus the whole process becomes totally inde- tic in the control of the biodiesel properties [10,13,14].
pendent of petroleum. The oxidative stability determination is quite often based on the
A serious problem associated with the use of petroleum fuels methodology of accelerated tests, originally proposed by Hadorn
including diesel oil is the increase in the pollutant emissions. Bio- and Zurcher [15] for the monitoring of the rancidity of edible oils,
diesel appears as a renewable natural fuel that inflicts minimal known as the Rancimat method. This test is based on the increase
harm to the environment and can be applied as an alternative to of the conductivity of deionized water contained in a reservoir,
which retains the volatile acids liberated during the propagative
oxidation of fatty materials. The European standard EN 14112
establishes that the oxidative stability of biodiesel should be deter-
⇑ Corresponding author. Tel./fax: +55 83 3216 7441. mined at 110 °C by the Rancimat method, requiring a minimum
E-mail address: agouveia@quimica.ufpb.br (A.G. Souza). value of 6 h for the induction period [16].

0016-2361/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.fuel.2010.09.014
774 M.B. Dantas et al. / Fuel 90 (2011) 773–778

Alternatively the oxidation process can also be evaluated by D6584), total glycerin (ASTM D6584), water and sediment (ASTM
thermoanalytical techniques, such as Thermogravimetric Analysis D2706), flash point (ASTM D93), specific mass (ASTM D4052), cop-
(TG), Differential Scanning Calorimetry (DSC) and Pressurized Dif- per corrosion (ASTM D130) and oxidative stability (EN 14112) by
ferential Scanning Calorimetry (PDSC) [17,18]. the Rancimat method.
PDSC measures the liberation of energy during the oxidation The physico-chemical analyses performed in the B0, BH1, BH2,
reactions, instead of measuring the oxidation products with the BH6 and BS samples were iodine value, peroxide value, dynamic
analyses being carried out directly in the sample. It is a powerful viscosity, thermogravimetric analyses, PDSC and UV/Vis
tool for the high precision determination of the thermodynamic spectroscopy.
properties, showing significant applicability for oils, lubricants
and polymers. Furthermore, PDSC analyses reach temperature
ranges not often detected by conventional oxidative techniques, 2.2.1. Gas chromatography
such as the Rancimat test. This technique differs from the Ranci- The gas chromatography analyses of as-synthesized (B0) were
mat, as it is a rapid method and contains an extra variable, the made in a CG-FID VARIAN 3800 gas chromatograph, with split/
pressure, having thus the possibility to increase the number of oxy- splitless injector.
gen moles present in the environment and speeding up the reac-
tion, besides using minor amounts of sample, without losing 2.2.2. Oxidative stability by Rancimat
accuracy [1,17–19]. The oxidative stability of B0 was determined according to the
PDSC is efficient in its dynamic mode, in which the oxidation standard EN 14112, using the Rancimat equipment (model 743,
temperature of the sample is determined and in the isothermal Metrohm). In this method, 3 g of sample were heated at 110 °C un-
mode, when the onset oxidation time of the sample is determined. der constant air flow (10 L h 1).
In both modes the sudden energy liberation is measured in relation
to the heat flow baseline, corresponding to the temperature or time
of oxidative induction [20]. 2.2.3. Iodine value
Along with titrimetric analyses, UV/Vis has been applied for the For the determination of the iodine value, the Hübl’s methodol-
study of biodiesel oxidation [21]. It allows measuring the absorp- ogy was employed [23]. The sample was diluted in chloroform
tion due to the formation of dienes, trienes and peroxides in bio- (Merck, 99.4%), with the later addition of solutions of alcoholic io-
diesel, resulting from oxidative processes. dine and mercuric chloride. The iodine excess was titrated with a
Therefore, this work aims at studying the oxidative stability of the sodium thiosulfate (Vetec) solution with starch as indicator, until
corn biodiesel, obtained by the ethanol route, during its long-term the disappearing of the blue color.
storage and heating (150 °C). For this purpose, thermal analysis tech-
niques, such as TG and PDSC were employed in conjunction with UV/
2.2.4. Peroxide value
Vis spectroscopy, physico-chemical parameters and viscosity.
The peroxide value was determined according to the AOCS Cd 8-
53 standard [24]. The sample was dissolved in an acetic acid–chlo-
2. Experimental roform solution, with the addition of also a saturated solution of
potassium iodide and a 1% starch (Vetec) solution. The liberated io-
2.1. Synthesis and storage of biodiesel dine was titrated with a sodium thiosulfate (Vetec) solution, until
the disappearing of the bluish color.
Refined edible corn oil (Bunge Salada), ethanol (Merck, 99.5%)
and potassium hydroxide (Merck) were used for synthesis of
2.2.5. Dynamic viscosity
biodiesel.
The dynamic viscosities of the samples were determined in a
The biodiesel was obtained by means of the transesterification
Brookfield equipment; model LV-DVII, coupled to a heat controller
reaction, using 1:6 oil:ethanol molar ratio and potassium hydrox-
at the temperature of 25 °C, using a small sample adaptor.
ide (1% mass ratio) as a homogeneous basic catalyst [1]. Initially
potassium ethoxide was prepared by mixing 30 g of ethanol with
1 g of KOH for each 100 g of oil in beaker, under stirring up to 2.2.6. Thermogravimetric analyses
the complete dissolution of KOH. Next potassium ethoxide was The non-isothermal thermogravimetric curves were obtained in
added to corn oil, placing the beaker on a plate featuring magnetic an equipment SDT 2960 (TA Instruments), in the 25–600 °C tem-
stirring, in order that the transesterification reaction occurs at perature range, using a synthetic air atmosphere (100 mL min 1),
room temperature. The reaction reaches equilibrium (steady state) alumina crucibles, heating rate of 10 °C min 1 and a sample mass
within 80 min. Taking into account the concept of chemical equi- of 10.0 mg [1].
librium, the formation of fatty acid esters is favored by increasing
the alcohol/oil molar ratio. After the separation of ester/glycerin
phases, the biodiesel was washed with distilled water and dried 2.2.7. PDSC
with anhydrous Na2SO4 (Vetec) [22]. The PDSC curves (dynamic mode) were recorded in a DSC 2920
After transesterification, the as-synthesized corn biodiesel (B0) (TA Instrument) with pressure cell, using platinum crucible, heat-
was stored in an amber glass container at room temperature for ing rate of 10 °C min 1, 1400 kPa oxygen gas (99.5% purity and
12 months, being denominated BS. Artificial accelerated aging of constant volume), at the 25–150 °C temperature range and a sam-
the biodiesel sample was carried out in a closed glass system under ple mass of 10.0 mg [25].
air atmosphere (30 mL min 1), in which the B0 sample was heat
treated at 150 °C during 1 (BH1), 2 (BH2) and 6 (BH6) h.
2.2.8. UV/Vis absorption spectroscopy
The UV/Vis spectra were obtained in a Shimadzu UV-2550 spec-
2.2. Characterization of the biodiesel samples trophotometer, using 1:1000 (in volume) dilution ratio in dichloro-
methane (Merck). All data were collected in the 200–400 nm
The as-synthesized biodiesel sample (B0) were analyzed by gas range. The deconvolution of UV–Vis spectra was performed using
chromatography, acid number (ASTM D664), free glycerin (ASTM the free Peak Fit software.
M.B. Dantas et al. / Fuel 90 (2011) 773–778 775

Table 1
Chemical composition of the ethanol corn biodiesel (B0) and corn oil (literature data).

Fatty chain This work/wt.% [26]/wt.% [27]/wt.%


Caprylic (C8:0) 0.06 – –
Lauric (C12:0) 0.01 – –
Myristic (C14:0) 0.04 0 –
Palmitic (C16:0) 2.58 12 11.8
Stearic (C18:0) 1.66 2 2.0
Oleic (C18:1) 13.00 25 24.8
Linoleic (C18:2) 77.18 60 61.3
Linolenic (C18:3) 0.66 Traces 0.0
Others 4.80 – 0.3

3. Results and discussion

The biodiesel samples B0 were analyzed by gas chromatogra-


phy, in order to assess the conversion of the triglycerides into their
ethyl esters. Table 1 presents these results, as well as theoretical
biodiesel compositions, based on the fatty acid compositions of
corn oil available in the literature [26,27]. The results, as expected
by the corn oil composition, indicated the prevalence of the ethyl
linoleates (77.18 wt.%), a value higher than those found in the lit-
erature. The second richest ester in the biodiesel is ethyl oleate,
with 13 wt.%.
In Table 2, are shown the physico-chemical properties of biodie-
sel B0, as well as the limits established by the standards EN 14214
[28] and ASTM D6751 [29]. The tests acid number and total glyc-
erin displayed values slightly higher than those set up by EN and
Fig. 1. TG/DTG curves for (a) as-synthesized (B0) and long-term stored (BS) and (b)
ASTM. The oxidative stability was determined by the method EN
samples of thermally treated corn biodiesel, obtained by the ethanol route (BH1, BH2
14112, the method Rancimat, showing a figure of 2.89 h, smaller and BH6).
than the lower limits of 3 h and 6 h from ASTM and EN,
respectively.
The TG/DTG data (Fig. 1 and Table 3) showed two main mass the amount of linoleic acid in natura, the more easily biodiesel is
loss steps for the B0 sample and three for BS, BH1 and BH2, with oxidized.
all the samples possibly yielding keto-aldehyde compounds and However, the oxidation temperature is also very sensitive to the
polymeric products. In all samples the first mass loss step was amount of peroxides existing before the DSC analysis [31].
attributed to volatilization (mainly ethyl linoleate) and peroxide Although in the previously heat treated samples the amount of
decomposition reactions. It was noticed that the more the sample bis-allylic hydrogen is low (high conjugation), the amount of pre-
was oxidized before the Thermogravimetric Analysis; the lower existing peroxides is high (rise of the peroxide value for the previ-
was the onset temperature of the first mass loss step. This was ously heat treated samples), leading to a more rapid mass loss (no-
due to the volatilization of the aldehydes and ketones previously ticed in the TG) and diminishment of the onset oxidation
formed (step 4 of Fig. 2). temperature (noticed in PDSC). Such results are in agreement with
For the BH6 (Fig. 1b), four steps were observed, being the first the peroxide and iodine values, as well as with the dynamic viscos-
one related to oxidation and decomposition reactions and the last ity values presented in Table 4.
three steps related to the subsequent polymerization and combus- According to Table 4 and Fig. 4, the dynamic viscosity of biodie-
tion reactions. That thermal behavior was confirmed by the abrupt sel raised considerably as consequence of the thermal stress at
increase on the BH6 absolute viscosity (Table 4). 150 °C, being barely modified during the long-term storage at
The PDSC curves (Fig. 3) demonstrate that the onset oxidative room temperature. It was reported that viscosities of mineral lubri-
temperatures were reduced from 131 °C (B0) to 115 °C (BS), cant oils increase when submitted to thermal stress, due to the for-
103 °C (BH1) and 95 °C (BH6). According to Knothe et al. [30], the mation of carbonyl compounds, which interact with hydrogen
oxidation temperature of a biodiesel is inversely proportional to bonds [32,33]. Similarly, the biodiesel rheology can also be ex-
the amount of bis-allylic hydrogens; in other words, the higher plained by this kind of interaction.

Table 2
Physico-chemical properties of ethanol corn biodiesel (B0), as compared with the EN and ASTM specification limits.

Test Units Value EN 14214 limit [28] ASTM D6751 limit [29]
1
Acid number mg KOH g 0.8 max. 0.5 max. 0.5
Free glycerin % 0.015 max. 0.02 max. 0.02
Total glycerin % mass 0.27 max. 0.25 max. 0.24
Water and sediment % vol 0.044 – max. 0.050
Flash point °C 169 min. 120 min. 130
Specific mass at 20 °C kg m 3 876.1 860–900 at 15 °C –
Copper corrosion 3 h at 50 °C – 1 max. 1 max. 3
Oxidative stability at 110 °C h 2.89 min. 6 min. 3
776 M.B. Dantas et al. / Fuel 90 (2011) 773–778

Table 3 center, full width at half maximum (FWHM) and analytic area of
Thermogravimetry data for corn biodiesel, obtained by the ethanol route, after storing each band, are summarized in Table 5.
and thermal treatment at 150 °C.
According to literature, for linoleic acid the conjugated dienes
Sample Steps Temperature range (°C) Mass loss (%) absorb at 232 nm, due to p ? p* transitions. The secondary oxida-
B0 1 159–376 95.1 tion products, particularly a-diketones or unsaturated ketones,
2 411–519 3.5 display an absorption peak at 272 nm, also attributed to p ? p*
BS 1 122–294 92.7 transitions [35–38].
2 294–363 4.8
3 364–544 1.9
In the present work, such absorption bands of linoleic acid ester
BH1 1 114–307 87.7 were detected, although minor shifts were noticed, moreover for
2 307–391 8.7 the higher wavelength band. The first band was called band 1
3 411–577 4.0 and the higher wavelength one, band 2. The intensities of both
BH2 1 118–301 91.2
bands grew as the heat treatment time increased. The different
2 301–390 7.3
3 453–560 2.1
BH6 1 116–383 89.7
2 383–432 9.8 Table 4
3 432–471 3.3 Physico-chemical properties of as-synthesized, stored and heated biodiesels.
4 471–565 3.8
Parameter B0 Bs BH1 BH2 BH6
Peroxide value (meq kg 1 oil) 15.30 89.92 90.05 29.65 22.83
Dynamic viscosity (mPa s) 6.03 6.18 6.64 7.12 12.10
The iodine values of the biodiesel samples decreased with the Iodine value (mg I2 100 g 1) 128.00 92.35 113.00 86.00 75.00
heat treatment time at 150 °C, Table 4 and Fig. 4. The value for
BS was lower than B0, indicating that the biodiesel was degraded
during the long room temperature storage. Comparing room tem-
perature storage (BS sample) with heat treatment at 150 °C (BH
samples), iodine values indicated that the 12 month storage at
room temperature corresponded to 1–2 h of heat stress at 150 °C.
As a general rule, the peroxide values considerably increased
during the long-term storage and after the heat treatment process,
for short periods of time (BH1), indicating the prevalence of the
propagation step. In addition, the reduction of the peroxide value
as a function of the time of heat treatment can be attributed to
the termination reaction step, resulting in peroxide decomposition
[34]. Correspondingly, the iodine values reduced in magnitude for
both storage systems, evidencing the oxidative degradation of the
samples. The values presented in Table 4 and Fig. 4 were in agree-
ment with such general rule.
Fig. 5 displays the UV/Vis spectra of all the samples. The spec-
trum of each sample was deconvolved, as exemplified in Fig. 6 Fig. 3. PDSC curves for the as-synthesized corn biodiesel, obtained by the ethanol
for the sample B0. The results obtained, in terms of amplitude, route (B0) and after storage (BS) and thermal treatment (BH1 and BH6).

Fig. 2. Reaction scheme for the double bond conjugation in linoleic acid [36].
M.B. Dantas et al. / Fuel 90 (2011) 773–778 777

Fig. 6. Deconvolution of the UV/Vis spectrum of the B0 sample, related to the


untreated sample.
Fig. 4. Influence of the heat treatment time on the dynamic viscosity and iodine
values of the biodiesel samples.

Fig. 7. Evolution of the amplitude of the absorbance peaks, as a function of the heat
treatment time.

Fig. 5. UV/Vis spectra of as-synthesized, stored and heated samples of corn


biodiesel, obtained by the ethanol route.
auto-oxidation mechanism occurred after 2 h. The increase of the
absorbance of the 232 nm band apparently did not agree with
Table 5 the concomitant decrease of iodine value, with the decrease of
Deconvolution of the UV/Vis spectra of the samples. the heat treatment time (Fig. 5). However, this might be due to
Samples B0 BS BH1 BH2 BH6
the conjugation between the two double bonds of linoleic acid,
which took place upon oxidation [36], as depicted in Fig. 2. It
Peak 1
Center (nm) 226 230 229 230 232
should be stressed that the conjugated dienes did not react with io-
Amplitude (a.u.) 0.55 1.87 1.20 2.15 2.93 dine under the conditions prevailing in the iodine test.
Area (a.u.) 29.2 72.4 62.4 89.5 118.2
FWHM (nm) 50.1 36.4 48.7 39.1 38.0
4. Conclusions
Peak 2
Center (nm) 279 280 282 279 279
Corn biodiesel, obtained by the ethanol route, was submitted to
Amplitude (a.u.) 0.16 0.15 0.29 0.49 0.80
Area (a.u.) 4.4 3.1 8.0 15.1 23.3 oxidative degradation, by long time storage at room temperature
FWHM (nm) 25.9 19.7 26.0 28.8 27.3 and by heat treatment at 150 °C. The degradation was found to
r2 0.9970 0.9928 0.9954 0.9918 0.9910 be more pronounced under heating conditions.
Amplitude 3.44 12.47 4.14 4.39 3.67 The profile of TG curve of non-oxidized biodiesel (B0) showed
(Peak1/Peak2) the same tendency observed in literature, while for the oxidized
samples the initial temperature was reduced and the number of
mass loss steps increased. The increase of the number of steps is
consistent with the formation of secondary oxidation products.
intensities of these bands were particularly interesting, once they Both calorimetric data and physico-chemical parameters demon-
allowed evaluating the oxidation stages, based on the amplitude strated that biodiesel stability is reduced with long-term storage
ratios of these two peaks. Initially the amplitude ratio between and more drastically when it is submitted to a heat stress. This
peak 1 and peak 2 was relatively low, corresponding to the onset work demonstrated that UV–Vis spectrometry technique is an
of the oxidation process. As the oxidation process continued, such important tool for monitoring the oxidative stability of biofuels,
ratio increased along with the peroxide content, reaching a maxi- since the amplitude of the absorption bands at 232 and 272 nm
mum at about 2 h of heat treatment. At high oxidation times the respond to the oxidative degradation of this class of products.
ratio decreased corresponding to higher amounts of secondary oxi- The physico-chemical parameters (viscosity, iodine value and
dation products, Fig. 2. peroxide value) were successfully related to the different oxidation
The results of the deconvolved bands, as a function of heat steps and supported the quantitative analysis performed by
treatment time, Fig. 7, pointed out that a step change in the lipid UV–Vis spectrometry.
778 M.B. Dantas et al. / Fuel 90 (2011) 773–778

The utilization of the techniques focused in the present work [17] Gamelin CD, Dutta NK, Choudhury NR, Kehoe D, Matisons J. Evoluation of
kinetic parameters of thermal and oxidative decomposition of base oils by
aimed at an integrated analysis of the influence of the initial
conventional, isothermal and modulated TGA and pressure DSC. Thermochim
oxidation stage (different values of peroxide reached by the ther- Acta 2002;392–393:357–69.
mal pre-treatment) of biodiesel on the thermogravimetric (TG), [18] Tan CP, Che MYB, Selamat J, Yusoff MSA. Comparative studies of oxidative
thermo-oxidative (PDSC), spectroscopic (UV–Vis) and rheological stability of edible oils by differential scanning calorimetry and oxidative
stability index methods. Food Chem 2002;76:385–9.
(dynamic viscosity) behaviors. Each different technique supplies [19] Velasco J, Andersen ML, Skibsted LH. Evaluation of oxidative stability of
information of different oxidative events in the oxidation process, vegetable oils by monitoring the tendency to radical formation. A comparison
differently of the volatile formation, which only occurs in the last of electron spin resonance spectroscopy with the Rancimat method and
differential scanning calorimetry. Food Chem 2004;85:623–32.
oxidation stages, as is the case of the Rancimat method. Further- [20] Kodali DR. Oxidative stability measurement of high-stability oils by pressure
more, all the techniques that were applied in the present work differential scanning calorimeter (PDSC). J Agric Food Chem 2005;53:7649–53.
are of low operational cost and rapid execution. [21] Rodrigues MGF, Souza AG, Santos IMG, Bicudo TC, Silva MCD, Sinfrônio FSM,
et al. Antioxidative properties of hydrogenated cardanol for cotton biodiesel by
PDSC and UV/Vis. J Therm Anal Cal 2009;97:605–9.
Acknowledgements [22] Vasconcelos AFF, Dantas MB, Filho MGR, Rosenhaim R, Cavalcanti EHS,
Antoniosi Filho NR, et al. Influence of drying processes on oxidative stability
of ethyl corn biodiesel by differential scanning calorimetry. J Therm Anal Cal
The authors acknowledge the Brazilian agencies CNPq, CAPES 2009;97:657–60.
and FINEP the financial support. [23] Zenebon O, Pascuet NS, Tiglea P. Métodos químicos e físicos para análise de
alimentos. 4th ed. São Paulo: Instituto Adolfo Lutz; 2008.
[24] AOCS. Method Cd 8-53. Official and tentative methods of the american oil
chemists society. third ed., Champagne, Illinois, USA: American Oil Chemists
References
Society; 1980.
[25] Conceição MM, Dantas MB, Rosenhaim R, Fenandes Jr VJ, Santos IMG, Souza
[1] Dantas MB, Conceição MM, Fernandes Jr VJ, Santos NA, Rosenhaim R, Marques AG. Evaluation of the oxidative induction time of the ethilic castor biodiesel. J
ALB, et al. Thermal and kinetic study of corn biodiesel obtained by the Therm Anal Cal 2009;97:643–6.
methanol and ethanol routes. J Therm Anal Cal 2007;87(3):835–9. [26] Goering CE, Schwab AW, Daugherty MJ, Pryde EH, Heakin AJ. Fuel properties of
[2] Alptekin E, Canakci M. Determination of the density and the viscosities of eleven oils. Trans ASAE 1982;25:1472–83.
biodiesel–diesel fuel blends. Renew Energy 2008;33:2623–30. [27] Demirbas A. Biodiesel fuels from vegetable oils via catalytic and non-catalytic
[3] Ranade SA, Srivastava AP, Rana TS, Srivastava J, Tuli R. Easy assessment of supercritical alcohol transesterifications and other methods: a survey. Energy
diversity in Jatropha curcas L. plants using two single-primer amplification Convers Manage 2003;44:2093–109.
reaction (SPAR) methods. Biomass Bioenerg 2008;32:533–40. [28] European Committee for Standardization. Standard EN 14214. Automotive
[4] Conceição MM, Candeia RA, Dantas HJ, Soledade LEB, Fernandes Jr VJ, Souza fuels – fatty acid methyl esters (FAME) for diesel engines – requirements and
AG. Rheological behavior of castor oil biodiesel. Energy Fuel 2005;19:2185–8. test methods.
[5] Knothe G. Dependence of biodiesel fuel properties on the structure of fatty acid [29] American Society for Testing and Materials (ASTM) standard D6751. Standard
alkyl esters. Fuel Process Technol 2005;86:1059–70. specification for biodiesel fuel blend stock (B100) for middle distillate fuels.
[6] Knothe G. ‘‘Designer” biodiesel: optimizing fatty ester composition to improve West Conshohocken (PA): ASTM.
fuel properties. Energy Fuels 2008;22:1358–64. [30] Knothe G, Gerpen JV, Krah J, Ramos LP. The biodiesel handbook. 1st ed. São
[7] Candeia RA, da Silva MCD, Carvalho Filho JR, Brasilino MGA, Bicudo TC, Santos Paulo: Edgard Blucher; 2006.
IMG, et al. Influence of soybean biodiesel content on basic properties of [31] Litwinienko G. Autooxidation of unsaturated fatty acids and their esters. J
biodiesel–diesel blends. Fuel 2009;88:738–43. Therm Anal Cal 2001;65:639–46.
[8] Dinkov R, Hristov G, Stratiev D, Aldary VB. Effect of commercially available [32] Santos JCO, Santos IMG, Souza AG, Sobrinho EV, Fernandes Jr VJ, Silva AJN.
antioxidants over biodiesel/diesel blends stability. Fuel 2009;88:732–7. Thermoanalytical and rheological characterization of automotive mineral
[9] Dunn RO. Effect of antioxidants on the oxidative stability of methyl soyate lubricants after thermal degradation. Fuel 2004;83:2393–9.
(biodiesel). Fuel Process Technol 2005;86:1071–85. [33] Perez JM. Oxidative properties of lubricants using thermal analysis.
[10] Mittelbach M, Gangl S. Long storage stability of biodiesel made from rapeseed Thermochim Acta 2000;357:47–56.
and used frying oil. J Am Oil Chem Soc 2001;78:573–7. [34] Morita M, Tokita M. The real radical generator other main-product
[11] Knothe G, Dunn RO. Dependence of oil stability index of fatty compounds on hydroperoxide in lipid autoxidation. Lipids 2006;41:91–5.
their structure and concentration and presence of metals. J Am Oil Chem Soc [35] Srinivasan S, Xiong YL, Decker EA. Inhibition of protein and lipid oxidation in
2003;80:1021–6. beef heart surimi-like material by antioxidants and combinations of pH, NaCl,
[12] Terry B, McCormick RL, Natarajan M. Impact of biodiesel blends on fuel system and buffer type in the washing media. J Agric Food Chem 1996;44:119–25.
component durability, SAE Techn Paper No. 2006-01-3279. [36] Schneider C, Porter NA, Brash AR. Routes to 4-hydroxynonenal: fundamental
[13] Bouaid A, Martinez M, Aracil J. Long storage stability of biodiesel from issues in the mechanisms of lipid peroxidation. J Biol Chem
vegetable and used frying oils. Fuel 2007;86:2596–602. 2008;283:15539–43.
[14] Knothe G. Some aspects of biodiesel oxidative stability. Fuel Process Technol [37] Frankel EM, Huang S-W, Kanner J, German JB. Interfacial phenomena in the
2007;88:669–77. evaluation of antioxidants: bulk oils versus emulsions. J Agric Food Chem
[15] Hadorn H, Zurcher K. Zurbestimmung der oxydationsstabilitat von olen und 1994;42:1054–9.
fetten. Deut Lebensm-Rundsch 1974;70:57. [38] Huang S-W, Hopia A, Schwarz K, Frankel EM, German JB. Antioxidant activity
[16] European Committee for Standardization, fatty acid methyl esters (fame) – of a-tocopherol and trolox in different lipid substrates: bulk oils versus soil-in-
determination of oxidation stability (accelerated oxidation test). method water emulsions. J Agric Food Chem 1996;44:444–52.
EN14112; 2003.

S-ar putea să vă placă și