Vous êtes sur la page 1sur 6

Ind. Eng. Chem. Res.

XXXX, xxx, 000

Component Distribution between Light and Heavy Phases in Biodiesel Processes


Renzo Di Felice,*, Danilo De Faveri, Paola De Andreis, and Piero Ottonello
Dipartimento di Ingegneria Chimica e di Processo G.B. Bonino, UniVersita ` degli Studi di GenoVa, Via Opera Pia 15, 16145 GenoVa, Italy, and Chemtex Italia srl - Gruppo Mossi & Ghisol, Strada SaVonesa 9, 15050 RiValta ScriVia, Italy

Knowledge of the component distribution in the biodiesel production process is fundamental both during the transesterication reaction, where reactants are partially miscible, and during the recovery of the nal products, which exist in two separate phases, a heavy one containing nearly all of the glycerol and a light one containing nearly all of the biodiesel. In this article, a simple methodology able to predict the product distribution at equilibrium between the heavy and light phases is suggested. Experimental equilibrium data for mixtures of two (biodiesel and glycerol), three (biodiesel, glycerol, and methanol), and four (biodiesel, glycerol, methanol, and water) components are collected and correlated with the Wilson activity coefcient model. The approach is based on thermodynamic data already available in the literature, without the addition of any empirical tting parameter. A small effect brought about by the presence of water at low concentrations on methanol distribution between the two phases is pointed out and qualitatively supported by the model predictions. The inuence of the addition of electrolyte contaminants (soap or catalysts) is also considered.
Introduction Awareness in the industrialized world of the depletion of crude oil reserves is pushing toward the search for renewable alternative energy sources. Painless and comprehensive solutions are not, by any means, readily available at present. However, certain actions that are already available and are planned for the immediate future, as also evidenced by European Directive 2003/30/EC,1 can contribute to alleviate this problem. The use of biodiesel as a transportation fuel is one such action, and this explains the unprecedented interest in this alternative form of energy. Biodiesel, the common name for fatty acid methyl esters (FAMEs), is a liquid fuel obtained from the transesterication of vegetable (or animal) oils. The process has been known for a long time and certainly cannot be dened as high-tech, as practically anyone can do it in his or her own kitchen, without the use of any sophisticated equipment or material. It involves simply the reaction, under very mild conditions, between vegetable oil and typically a large excess of alcohol of methanol, in the presence of an acid or basic catalyst, which produces FAMEs as the main product and glycerol as a byproduct.2 The separation of the transesterication reaction products is then greatly facilitated by the formation of two immiscible liquid layers: a heavy one containing nearly all of the glycerol and a light one containing nearly all of the biodiesel, with the excess alcohol being distributed between the two phases. Further biodiesel cleaning processes are then carried out to obtain a product of the required purity and characteristics, so that it can be used, alone or more commonly mixed with diesel of fossil origin, as a transportation fuel.3 The simplicity of the process seems to have been transferred to the industrial level, without any further renement. Indeed, the industrial process is currently carried out based mostly on the specic knowledge and experience of the plant operators, very much like a craftsman who treasures his know-how and
* To whom correspondence should be addressed. Tel.: +39 0103532924. Fax: +39 0103532586. E-mail: renzo.difelice@unige.it. Universita ` degli Studi di Genova. Chemtex Italia srl - Gruppo Mossi & Ghisol.

resists any change that aims at rationalizing and improving the overall process. Quantication of the basic parameters governing the biodiesel production process is the rst step toward better knowledge, and better operation as a direct consequence, and it represents the aim of the work in which we are currently involved. In this article, we concentrate on a very specic aspect, namely, the separation of the products after the transesterication reaction and the washing with water of the resulting biodiesel-rich streams. These operations are carried out in the liquid phase by exploiting the formation of two immiscible phases of quite different densities. Consequently, the two phases, in terms of component distribution, must conform to the thermodynamic liquid-liquid equilibrium that is established between them. A simple methodology is suggested here that is capable of predicting the product distribution at equilibrium between the heavy and light phases. The components that must be considered for the current problem are essentially biodiesel, methanol, glycerol, water, soap, and catalysts. First, we limit the study to the behavior of a mixture made up of all of the nonelectrolyte products: biodiesel, glycerol, methanol, and water. Then, we consider the effects on the obtained results due to the addition of soap or catalysts. This approach allows a simple predictive model to be put forward for the nonelectrolyte mixture. It has to be stressed that such an approach is based on thermodynamic data already available in the literature, without the introduction of any empirical tting parameter. Theoretical Analysis The situation addressed here considers n components distributed between two liquid phases. The component distribution, at equilibrium conditions, is governed by the chemical potential, so that the following relationship must hold for every component i (ixi)L ) (ixi)H (1)

In this equation, x is the component molar fraction; is its activity coefcient; and H and L represent the light and heavy phases, respectively. The activity coefcient for the ith component is generally a complex function of the molar fraction of

10.1021/ie800510w CCC: $40.75 XXXX American Chemical Society Published on Web 09/25/2008

Ind. Eng. Chem. Res., Vol. xxx, No. xx, XXXX

Table 1. Wilson Paired Energy Parameters ij (J/mol) biodiesel biodiesel methanol glycerol water
a

methanol 33387 46287 21788

glycerol
a

water
a

55507
a a

because Wilsons equation is valid only for nonelectrolytes, the effects of soap and catalyst could not be included in the model predictions. Experimental Procedure Materials. Biodiesel (palm oil methyl esters) was kindly supplied by Mythen SpA, Ferrandina (MT), Italy. Its purity and bound glycerol and free glycerol contents are >98.5%, < 0.8%, and <200 mg/kg, respectively. Methanol (g99.9%) and glycerol (g99.0%) were purchased from Sigma-Aldrich, Milan, Italy. Distilled water and potassium oleate (g87.0%, Sigma-Aldrich, Milan, Italy) were used as contaminants. Liquid-Liquid Equilibrium Experiments. Experimental tests were performed in a separating funnel at different temperatures (20, 40, and 65 C). After addition of exactly known amounts of components, the mixtures were vigorously mixed for 5 min and allowed to stand for 1 h. For tests carried out at 40 and 65 C, after mixing, the separating funnel was placed in an oven for settling. Afterward, the mixtures were separated into two phases, a light phase, rich in biodiesel, and a heavy phase, rich in glycerol, that were analyzed following the analytical procedure described below. Tests were carried out using 50-60-g samples of mixtures of two (biodiesel and glycerol), three (biodiesel, glycerol, and methanol), four (biodiesel, glycerol, methanol, and water), and ve (biodiesel, glycerol, methanol, water, and potassium oleate) components. Analytical Procedures. The contents of glycerol and methanol in the light phase were determined by gas chromatography and headspace gas chromatography following the methods described in UNI EN 1410611 and UNI EN 14110,12 respectively, using a Perkin-Elmer gas chromatograph, model Clarus 500, equipped with a ame-ionization detector (FID) and a headspace sampler (Perkin-Elmer, model HS40). The calibration curves were based on biodiesel standards containing known amounts of glycerol and methanol, respectively. The content of biodiesel in the heavy phase was measured by means of an analytical method based on Fourier transform infrared (FT-IR) spectroscopy, which was developed using method UNI EN 1407813 as the reference norm. Biodiesel was extracted from samples (1 g) by cyclohexane (25 mL): the mixtures were magnetically stirred for 30 min and allowed to stand for a few minutes in sealed bottles. The calibration curve was made with standards containing known amounts of biodiesel, and the FT-IR absorbance of the extracts was recorded at 1748 cm-1. The content of methanol in the heavy phase was determined by headspace gas chromatography using the same gas chromatograph as above. For each analysis, a 20-mL vial containing a 20-mg sample was used. The calibration curve was made with glycerol standards containing known amounts of methanol. The water content in both phases was measured according to the coulometric Karl Fisher titration method (UNI EN ISO 1293714) using a coulometric Karl Fisher titrator (MettlerToledo, DL 37). Finally, the content of potassium oleate was determined by means of a titration method. Specically, samples were diluted with an acetone/water mixture and titrated with a 0.01 M HCl solution in the presence of bromophenol blue. Results and Discussion Biodiesel-Glycerol System. The results listed in Table 2, expressed as mass fractions, show that biodiesel and glycerol are soluble in each other to a very negligible extent. The

-36337 16569

3478 53299 -

Because biodiesel is practically immiscible with glycerol or water, these parameters are not necessary.

each component in the respective phase, as well as the temperature. Numerous efforts have been made to express such a function with workable numerical expressions having some physical basis. The generalized Wilson relationship4 is one such effort, and it is expressed as ln i ) - ln

( )
n j)1

xjij + 1 -

k)1

xkki

x
j j)1

(2)

kj

which contains only binary interaction parameters, ij, given by ij ) j ij exp i RT

( )

(3)

where is the component molar volume and ij is a numerical constant to be determined experimentally. It follows that knowledge of ij is the only input needed to be able, in principle, to quantify the component distribution between two phases in thermodynamic equilibrium for any given temperature, the only difculty being numerical, as it involves the solution of a system of highly nonlinear equations. For the specic case considered here, we are interested in predicting how biodiesel, methanol, glycerol, and water would distribute between the light and heavy phases. This prediction therefore requires knowledge of the pair of binary interaction parameters for all six possible component pairs. To simplify the present approach, in this case, we assumed that no biodiesel was present in the heavy phase and that no water or glycerol was present in the light phase. As discussed later, these assumptions are based on both experimental results from this work and literature evidence.5,6 In this way, the number of required binary interaction parameters was reduced to four pairs. They were all retrievable from the open literature, inferred in all cases from measurements of vapor-liquid equilibrium, and they are summarized in Table 1. It has to be stressed that, as reported by Orye and Prausnitz,10 the Wilson equation has two features that make it particularly useful for engineering applications. First, it has a built-in temperature dependence that has some theoretical signicance, as the quantities i,j are independent of temperature. The values of these parameters, obtained from data at one temperature, can be used to predict activity coefcients at other temperatures, which is an important advantage. Second, Wilsons model for a multicomponent solution requires only parameters that can be obtained from binary mixture data, thus reducing the required amount of experimental work. Even if other local-composition models, such as NRTL and UNIQUAC, have the same features, in this study, the Wilson equation was used because of its greater simplicity as well as the availability in the open literature of the parameters involved in the present investigation. However,

Ind. Eng. Chem. Res., Vol. xxx, No. xx, XXXX C

Figure 1. Methanol distribution between phases for the biodieselglycerol-methanol system. Temperature ) 20 C. Table 2. Experimental Data On Mutual Solubility in the Biodiesel-Glycerol Binary System at Different Temperatures T (C) 20 40 65 light-phase glycerol mass fraction (kg/kg) 0.0008 0.0009 0.0010 heavy-phase biodiesel mass fraction (kg/kg) 0.0016 0.0016 0.0008

Figure 2. Methanol distribution between phases for the biodieselglycerol-methanol system. Temperature ) 65 C.

solubility of biodiesel in the glycerol phase is slightly greater than that of glycerol in the biodiesel phase, which is in good agreement with the results obtained by Zhou et al.5 Moreover, the solubility in the binary mixture seems to change only slightly with temperature in the range from 20 to 65 C. These results therefore support the simplied assumption made in the predictive approach of no reciprocal solubility for the biodiesel-glycerol binary system. Biodiesel-Glycerol-Methanol System. The three-component biodiesel-glycerol-methanol system is the most widely studied, and therefore, an abundance of experimental evidence is available.5,7,15,16 The reason for this is quite simple, as these three components constitute the main bulk of the two-phase exit stream from the transesterication reactor; in fact, the oil feed is essentially completely converted to biodiesel and glycerol, and some methanol is still present as it is fed in large excess. According to the previous hypothesis, biodiesel is assumed to be found exclusively in the light phase and glycerol is assumed to be found exclusively in the heavy phase, whereas methanol is distributed between the two phases. An analysis performed using experimental evidence from this work and other data presented in the literature17 supported this simplication: such an analysis revealed, in fact, that the heavy phase contains over 97% of the overall glycerol (i.e., the amount of glycerol present in the mixture before the separation of the two phases) and, at the same time, no more than 1% of the overall biodiesel (i.e., the amount of biodiesel present in the mixture before the separation of the two phases). Methanol is the only distributed component, and its concentrations in the light and heavy phases have been measured experimentally. The results for ambient temperature are presented in Figure 1, where experiments are compared with predictions from the model outlined in a previous section. As expected, the methanol weight concentration is much higher in the heavy, glycerol-rich phase than in the lighter,

biodiesel-rich phase. This is perfectly understandable given the greater afnity between methanol and glycerol than between methanol and biodiesel. The good predictive power of the present approach should be stressed here, mostly because the predictions reported in Figure 1, and in all subsequent gures, are based on the values of coefcients ij that were obtained from liquid-vapor equilibrium measurements, but they are applied here to liquid-liquid equilibrium. In particular, the model was able to predict the methanol weight concentration in the heavy, glycerol-rich phase with an average percent error18 (APE) of 6.5%, which points out its good predictive power. The temperature does not have a very noticeable effect on the methanol distribution, and it is reassuring that the same result was obtained from calculated values. In fact, a comparison of the model predictions obtained at 20 C (Figure 1) and at 65 C (Figure 2) reveals a small rise in the methanol mass fraction in the light, biodiesel-rich phase as the temperature increases. Nevertheless, such behavior is in good agreement with other literature evidence.7 Biodiesel-Water-Methanol System. The biodieselwater-methanol three-component system is certainly less important than the one previously reported in industrial biodiesel production. However, water washing of the biodiesel-rich stream is a common feature of these processes, and therefore, information on this system is of some practical relevance. Biodiesel and water are practically immiscible, whereas methanol is distributed between the two phases. To the best of our knowledge, only one work has been published reporting the quantication of the methanol distribution in this system.6 As in the previous section, measured methanol concentrations in the biodiesel- and water-rich phases are depicted in Figure 3, together with model predictions. In this case, the model was able to predict the methanol weight concentration in the waterrich phase with an APE of 6.4%, which again points out the good predictive power of the present approach. Not surprisingly, the data are qualitatively very similar to those regarding the biodiesel-glycerol-methanol system, given the similarity between glycerol and water. Again, methanol is found preferentially in the heavy phase, and the predictions behave accordingly. Effect of Water on the Biodiesel-Glycerol-Methanol System. It is well-known that the stream leaving the transesterication reactor contains some water. The presence of water

Ind. Eng. Chem. Res., Vol. xxx, No. xx, XXXX

Figure 3. Methanol distribution between phases for the biodiesel-watermethanol system. Temperature ) 20 C. Table 3. Comparison between Experimental and Predicted Component Distribution in the Water-Biodiesel-Glycerol-Methanol System at 20 C light-phase mass fraction (kg/kg) component experimental predicted heavy-phase mass fraction (kg/kg) experimental nda 0.2060 0.7940 0 0.0014 0.1523 0.7878 0.0585 0.0008 0.1860 0.6902 0.1230 predicted

Figure 4. Biodiesel mass fraction in the heavy phase as a function of soap mass fraction. Temperature ) 20 C.

initial water content ) 0 wt % biodiesel methanol glycerol water biodiesel methanol glycerol water biodiesel methanol glycerol water
a

0.9784 0.0210 0.0006 0 0.9781 0.0208 0.0008 0.0003 0.9589 0.0404 0.0007 nda

0.9928 0.0072 0 0 0.9933 0.0067 0 0 0.9938 0.0062 0 0

0 0.1782 0.8218 0 0 0.1727 0.7739 0.0534 0 0.1678 0.7116 0.1206

initial water content ) 3 wt %

initial water content ) 7 wt %

Not determined.

is mainly due to some small amounts originally present in the oil and in the catalysts (at least in the case of KOH), in addition to that obtained from the reaction between free fatty acids and basic catalysts; some of this water can then hydrolyze the product. Although, with the current level of the process knowledge, a precise quantication of water content in the reactor exit stream is difcult, there is little doubt that, depending on the process operating conditions, water can be present at a concentration of a few percent. Even less information is available in the literature about the effect of the presence of water on the component distribution between the heavy and light phases. Stloukal et al.6 seem to imply that water can have quite a strong effect on the component distribution; however, the tests carried out in the present work did not conrm such a suggestion. Table 3 reports measured component distributions between the light and heavy phases for different amounts of water initially present in the mixture. As a comparison, Table 3 also includes a section with 0% water content, corresponding to the biodiesel-glycerol-methanol system. It is clear that the results obtained for the four-component system are very similar to those reported for the water-free system. Moreover, Table 3

indicates that these results are also qualitatively supported by the present model predictions. The Wilson approach, on which the present model is based, is particularly suitable for predicting phase equilibrium in this case as it is based only on binary interaction parameters. The very mild effect brought about by the presence of water at this low level of concentration can be justied by considering that, from a thermodynamic point of view, water has very similar characteristics to glycerol, so that a strong effect is not expected when some glycerol is substituted with the same amount of water. Effect of Soap Formation on Component Distribution between the Two Phases. When an alkali catalyst is utilized, free fatty acids react with the catalyst to form soap and water. Soap formation is one of the main problems in biodiesel production given its strong effect on the component phase distribution. The analysis of the present experimental study revealed that, when different quantities of soap, up to 6% by weight, were added to the original mixture (biodieselmethanol-glycerol), the soap was found almost completely (measured values ranged from 97% to 99.55%) in the heavy, glycerol-rich phase. The very important effect of this segregation was that the soap dragged with it, as expected, a nonnegligible quantity of biodiesel. Given that the Wilson model is valid only for nonelectrolytes, the system with soap cannot be included in the model predictions. Moreover, to the best of our knowledge, in the open literature, no mathematical model for predictions of component distributions in biodiesel processes includes the effect of soaps. As illustrated in Figure 4, the amount of biodiesel present, and therefore lost, in the glycerol-rich phase increases with soap concentration in the original mixture, before phase separation. The results depicted in Figure 4 are quite interesting in this respect. The amount of biodiesel lost in the glycerol-rich phase appears to be very small and roughly constant up to a critical soap concentration of about 1-2%, and only after that critical concentration does it increase proportionally. This nding is qualitatively in line with the well-known phenomenon of critical micelle concentration (cmc), well-documented when dissolution of a component by a surfactant or a soap is investigated.19,20 Similar conclusions can be obtained from the work by Vicente et al.21 Although Vicente et al.21 did not measure soap concentrations directly, their data were obtained from measurements carried out at the end of the actual reaction experiments

Ind. Eng. Chem. Res., Vol. xxx, No. xx, XXXX E


Table 4. Distribution of Catalyst between Light and Heavy Phases: Literature Dataa catalyst temperature (C) 25 75 25 75 ambient
a b

catalyst (wt%) molar ratiob 0:1:3 3:1:3 6:1:3 0:1:3 3:1:3 6:1:3 0:1:3 3:1:3 6:1:3 0:1:3 3:1:3 6:1:3 3:1:3 heavy phase 8.60 5.79 4.29 8.11 5.57 4.13 4.52 2.91 2.27 3.39 1.98 1.60 5.57 light phase 0.088 0.061 0.056 0.172 0.125 0.119 0.075 0.055 0.049 0.108 0.071 0.066 0.05 Kc 98 95 77 47 45 35 60 53 46 31 28 24 111

type KOH KOH H2SO4 H2SO4 NaOCH3

concentration (wt%) 1 1 0.5 0.5 1

Figure 5. Biodiesel mass fraction in the heavy phase as a function of soap mass fraction as estimated from the data of Vicente et al.21

Data were obtained from ref except where indicated otherwise. Methanol/biodiesel/glycerol molar ratio. c K ) weight percentage of catalyst in the heavy phase/weight percentage of catalyst in the light phase.

for different initial concentrations of free fatty acids. By estimating soap concentrations from free fatty acid concentrations, we obtained the results reported in Figure 5, where the same qualitative pattern found in the present experimental investigation (Figure 4) is reproduced, with a rapid increase of the biodiesel concentration in the heavy phase after a threshold value has been reached, even though the two series of data are not exactly the same. Catalyst Distribution between the Two Phases. In a chemical reaction, catalysts generally should not undergo any transformation, but in this specic case, some of the catalysts can actually be consumed by reacting with free fatty acids present in the feed. Nevertheless, some catalyst is still present in the exit stream, and it is very important to be able to predict how the catalyst will be distributed between the FAME- and glycerol-rich phases, as it should be removed from the biodiesel and, if possible, recycled from the glycerol. No experimental work has been done specically on this issue. However, there is quite a bit of evidence presented in the literature to provide a rather clear picture of the catalyst distribution.7,17 This evidence is summarized in Table 4, from which it can be seen that the majority of the catalysts, regardless of their chemical composition, are eventually found in the glycerol-rich phase. This result is not at all surprising, as it is obvious that the polar structure of glycerol is much more suitable for solubilizing a strong electrolyte acid or base catalyst than the nonpolar FAME solvent. Moreover, the data listed in Table 4 point out that the higher the excess of methanol, the lower the partition coefcient K. Such an effect is probably due to the fact that, even if the methanol weight concentration is generally higher in the heavy phase, as larger amounts of this component are present, its concentration in the light phase increases as well. Moreover, the methanol contained in the light phase brings with it greater amounts of catalyst, thus reducing the value of the partition coefcient. Conclusions A simple mathematical model to predict the distribution of the nonelectrolyte products (i.e., biodiesel, glycerol, methanol, and water) at equilibrium between the heavy and light phases involved in the biodiesel production process was developed.

The proposed approach was able to qualitatively support the behavior of both the ternary (biodiesel-glycerol-methanol and biodiesel-water-methanol) and the quaternary (waterbiodiesel-glycerol-methanol) systems. When considering the effects of electrolyte contaminants, it was shown that the amount of biodiesel lost in the heavy phase increases with soap concentration, once a threshold level of soap concentration is reached, whereas the minority of the catalyst is always found in the light phase. As a direct application of these results in the biodiesel production process, it can be pointed out that the presence of water in the biodiesel-glycerol-methanol system, at least at low concentrations (e7%), can be neglected from the standpoint of product separation and purication, whereas the presence of methanol in the biodiesel-glycerol system, with or without the catalyst, tends to increase the distribution of methanol and, if present, the catalyst in the biodiesel phase. This phenomenon favors the biodiesel reaction, which is quenched by the formation of greater amounts of glycerol, whose polar structure is particularly suitable for solubilizing a strong electrolyte acid or base catalyst, but increases the difculties in the separation and purication of the nal products. Acknowledgment The continuous nancial support of Chemtex Italia srl Gruppo Mossi & Ghisol is gratefully acknowledged. Nomenclature
K ) distribution coefcient R ) universal gas constant T ) absolute temperature x ) molar fraction Greek Letters ) activity coefcient ) Wilson paired energy parameter ) Wilson binary interaction parameter ) molar volume Subscripts H ) heavy phase i, j, k ) components L ) light phase

F Ind. Eng. Chem. Res., Vol. xxx, No. xx, XXXX

Literature Cited
(1) Directive 2003/30/EC of the European Parliament and of the Council of 8 May 2003 on the promotion of the use of biofuels or other renewable fuels for transport. (2) Meher, L. C.; Vidya Sagar, D.; Naik, S. N. Technical aspects of biodiesel production by transestericationsA review. Renewable Sustainable Energy ReV. 2006, 10 (3), 248268. (3) Van Gerpen J., Shanks B., Pruszko R., Clements D., Knothe G., Biodiesel production technology; Report NREL/SR-510-36244; National Renewable Energy Laboratory: Golden, CO, 2004. (4) Wilson, G. M. Vapor-liquid equilibrium. XI. A new expression for the excess free energy of mixing. J. Am. Chem. Soc. 1964, 86 (2), 127 130. (5) Zhou, H.; Lu, H.; Liang, B. Solubility of multicomponent systems in the biodiesel production by transesterication of Jatropha curcas L. oil with methanol. J. Chem. Eng. Data 2006, 51 (3), 11301135. (6) Stloukal, R.; Komers, K.; Machek, J. Ternary phase diagram biodiesel fuel-methanol-water. J. Prakt. Chem./Chem.-Zeitung 1997, 339 (5), 485 487. (7) Chiu, C.-W.; Goff, M. J.; Suppes, G. J. Distribution of methanol and catalysts between biodiesel and glycerin phases. AIChE J. 2005, 51 (4), 12741278. (8) Perry R. H.; Green D. Perrys Chemical Engineers Handbook, 7th ed.; McGraw-Hill: New York, 1997; pp 13-20. (9) Celere, M.; Gostoli, C. Osmotic distillation with propylene glycol, glycerol and glycerol-salt mixtures. J. Membr. Sci. 2004, 229 (1-2), 159 170. (10) Orye, R. V.; Prausnitz, J. M. Multicomponent equilibria with the Wilson equation. Ind. Eng. Chem. 1965, 57 (5), 1826. (11) UNI EN 14106. Derivati di grassi e oli, Esteri metilici di acidi grassi (FAME), Determinazione del contenuto di glicerolo libero, Italia, 2003. (12) UNI EN 14110. Derivati di grassi e oli, Esteri metilici di acidi grassi (FAME), Determinazione del contenuto di metanolo, Italia, 2003.

(13) UNI EN 14078. Determinazione del contenuto di esteri metilici di acidi grassi (FAME) nei distillati medi - Metodo spettrometico a infrarossi, Italia, 2004. (14) UNI EN ISO 12937. Prodotti petroliferi, Determinazione del contenuto di acqua, Metodo Karl Fisher mediante titolazione coulometrica, Italia, 2001. (15) Komers, K.; Tichy , J.; Skopal, F. Terna res phasendiagramm biodiesel-methanol-glyzerin. J. Prakt. Chem./Chem.-Zeitung 1995, 337 (1), 328331. (16) Ma, F.; Clements, L. D.; Hanna, M. A. Biodiesel fuel from animal fat. Ancillary studies on transesterication of beef tallow. Ind. Eng. Chem. Res. 1998, 37 (9), 37683771. (17) Zhou, W.; Boocock, D. G. B. Phase distributions of alcohol, glycerol, and catalyst in the transesterication of soybean oil. J. Am. Oil Chem. Soc. 2006, 83 (12), 10471052. (18) Aziz, C. E.; Georgiou, G.; Speitel, G. E., Jr. Cometabolism of chlorinated solvents and binary chlorinated solvent mixtures using M. trichosporium OB3b PP358. Biotechnol. Bioeng. 1999, 65 (1), 100107. (19) An, Y.-J.; Jeong, S.-W. Interactions of peruorinated surfactant with polycyclic aromatic hydrocarbons: critical micelle concentration and solubility enhancement measurements. J. Colloid Interface Sci. 2001, 242 (2), 419 424. (20) Huang, H.-L.; Lee, W.-M. G. Enhanced naphthalene solubility in the presence of sodium dodecyl sulphate: effect of critical micelle concentration. Chemosphere 2001, 44 (5), 963972. (21) Vicente, G.; Mart nez, M.; Aracil, J. A comparative study of vegetable oils for biodiesel production in Spain. Energy Fuels 2006, 20 (1), 394398.

ReceiVed for reView April 1, 2008 ReVised manuscript receiVed July 25, 2008 Accepted July 28, 2008 IE800510W

Vous aimerez peut-être aussi