Sie sind auf Seite 1von 6

Fuel 80 (2001) 693698

www.elsevier.com/locate/fuel

Kinetics of transesterication in rapeseed oil to biodiesel fuel as treated in supercritical methanol


D. Kusdiana, S. Saka*
Department of Socio-Environmental Energy Science, Graduate School of Energy Science, Kyoto University, Yoshida Honmachi, Sakyo-ku, Kyoto 606-8501, Japan Received 28 December 1999; accepted 3 August 2000

Abstract A kinetic study in free catalyst transesterication of rapeseed oil was made in subcritical and supercritical methanol under different reaction conditions of temperatures and reaction times. Runs were made in a bath-type reaction vessel ranging from 2008C in subcritical temperature to 5008C at supercritical state with different molar ratios of methanol to rapeseed oil to determine rate constants by employing a simple method. As a result, the conversion rate of rapeseed oil to its methyl esters was found to increase dramatically in the supercritical state, and reaction temperature of 3508C was considered as the best condition, with the molar ratio of methanol in rapeseed oil being 42. q 2001 Elsevier Science Ltd. All rights reserved.
Keywords: Kinetics of transesterication; Supercritical methanol; Methyl esters; Biodiesel fuel

1. Introduction Transesterication of vegetable oils with simple alcohol has long been a preferred method for producing biodiesel fuel [13]. Generally speaking, there are two methods of transesterication reaction. One is the method using a catalyst and the other is without the help of a catalyst. The former method has a long story of development and now biodiesel fuel produced by this method is in the market in some countries such as North America, Japan and some west European countries. However, there are at least two problems associated with this process; the process is relatively time consuming and purication of the product for catalyst and saponied products are necessary. The rst problem due to the two phase nature of vegetable oil/methanol mixture requires vigorous stirring to proceed in the transesterication reaction. To solve this problem, Boocock et al. reported that the use of a simple ether such as tetrahydrofuran can make this two phase nature into one phase of its mixture and that methyl esters can be produced in less than 15 min depending on the catalyst concentration [4]. Yet, the catalyst problem cannot be solved for purication. Therefore, this conventional process still requires a high production cost and
* Corresponding author. Tel./fax: 181-75-753-4738. E-mail address: saka@energy.kyoto-u.ac.jp (S. Saka).

energy. The overall process, thus, includes transesterication reaction, recovery of unreacted methanol, purication of methyl esters from catalyst and separation of glycerin as a co-product from saponied products. The latter method involves uncatalyzed transesterication of vegetable oil in supercritical methanol as recently reported by Saka and Kusdiana [5]. The supercritical state of methanol is believed to solve the two phase nature of oil/ methanol mixture to form a single phase due to a decrease in dielectric constant of methanol in supercritical state [6]. As a result, the reaction was found to be complete in a very short time within 24 min, as described in their previous work [5]. In addition, because of non-catalytic process, the purication of products after transesterication reaction is much simpler and environmentally friendly, compared with the conventional commercial method in which all the catalyst and saponied products have to be removed for biodiesel fuel. Some researchers have reported kinetics for both acidand alkali-catalyzed transesterication reactions. Dufek and coworkers studied the acid-catalyzed esterication and transesterication of 9(10)-carboxystearic acid and its mono- and di-methyl esters [7]. Freedman et al. reported transesterication reaction of soybean oil and other vegetable oils with alcohols [8], and examined in their study were the effects of the type of alcohol, molar ratio, type and amount of catalyst and reaction temperature on rate

0016-2361/01/$ - see front matter q 2001 Elsevier Science Ltd. All rights reserved. PII: S 0016-236 1(00)00140-X

694

D. Kusdiana, S. Saka / Fuel 80 (2001) 693698

Fig. 1. Relationship between reaction temperature and pressure inside a bath-type reaction vessel. The shadowed zone is in supercritical state of methanol.

constants and kinetic order. Noureddin and Zhu applied the effects of mixing of soybean oil with methanol on its kinetics of transesterication [9]. Recently, Diasakov et al. reported kinetics on uncatalytic transesterication reaction of soybean oil [10]. However, the kinetic study of vegetable oil in supercritical medium without catalyst has not yet been presented. Therefore, in this study, kinetics of transesterication of rapeseed oil to biodiesel fuel was studied as treated in supercritical methanol without using any catalyst. We reported the effects of molar ratio and reaction temperature on methyl ester formation followed by a proposed simple method for the kinetics of the transesterication reaction.

2. Materials and methods The rapeseed oil from Nacalai Tesque was used in this work as a vegetable oil. In addition, various methyl esters from fatty acids such as palmitic, stearic, oleic, linoleic and

linolenic acids were also obtained from Nacalai Tesque and used as a standard. The supercritical methanol biomass conversion system employed in this work is described in a previous work [5]. In brief, the rapeseed oil and methanol with a different molar ratio was fed to the reaction vessel made of Inconel-625 and it was shaken to mix. Subsequently, the entire reaction vessel was quickly immersed into the tin bath preheated at a designated reaction temperature and kept for a set time interval for subcritical and supercritical treatments of methanol (2005008C). It was, subsequently, moved into the water bath to stop the reaction. The temperature and pressure inside the vessel were monitored during the treatment to conrm that a subcritical or supercritical condition was achieved. For a subcritical treatment, the temperature was too low for tin bath to be melt. Therefore, oil bath was used for the treatments at 200 and 2308C with a temperature controller. The treated liquid was then removed from the reaction vessel and allowed to settle for about 30 min. After the product was separated to be two portions, the upper and lower, each portion was evaporated at 908C for 20 min to remove methanol. The obtained products were analyzed on its composition by using the high performance liquid chromatography (HPLC) (Shimadzu, LC -10AT) which consists of the column (STR ODS-II, 25 cm in length 4.6 mm in inner diameter, Shinwa Chem. Ind. Co.) and refractive index detector (Shimadzu, RID-10A) operated at 408C with 1.0 ml/min ow rate of methanol as a carrier solvent. The sample injection volume was 20 ml and peak identication was made by comparing in the retention time between the sample and the standard compound as in a previous work [5]. 3. Results and discussion The supercritical experiments of methanol with rapeseed oil were carried out with a batch-type of reaction vessel. Therefore, the temperature and the pressure inside the reaction vessel are different in different reaction conditions. Fig. 1 shows a typical relationship between the maximum reaction temperature and pressure inside the reaction vessel during the treatment. Since a critical point of methanol is 2398C (Tc) in temperature and 8.09 MPa (Pc) in pressure, supercritical state of methanol can be achieved in the shadowed zone in Fig. 1. 3.1. Effect of molar ratio of methanol to rapeseed oil The molar ratio of methanol to rapeseed oil is one of the most important variables affecting the yields of methyl esters converted. The stoichiometry of the transesterication of rapeseed oil requires three molecules of methanol to react with one molecule of rapeseed oil. Encinar [11] stated in the conventional commercial process with alkaline catalyst that the yield of methyl esters increases as the molar

Fig. 2. HPLC chromatograms of rapeseed oil as treated in supercritical methanol at 3508C in various molar ratios of methanol to rapeseed oil.

D. Kusdiana, S. Saka / Fuel 80 (2001) 693698

695

Fig. 3. Effect of the molar ratios of methanol to rapeseed oil in transesterication reaction on producing methyl esters, as treated at 3508C.

ratio of methanol to Cynara oil rises and that the optimal ratios for its transesterication result between 4.05 and 5.67. For its molar ratio less than 4.05, the reaction is reported to be incomplete, whereas at higher than 5.67, it becomes difcult to separate glycerin from methanol as a by-product. Another worker [12] further noted that a 98% conversion of vegetable oils could be made to the methyl esters at the molar ratio of 6, but that even higher molar ratio up to 45 was necessary when the oil contained a large amount of free fatty acids. However, as the molar ratio decreased to the theoretical value of 3, its conversion was decreased down to 82%. In this work, therefore, the effect of the molar ratio of methanol to rapeseed oil was studied in the range between 3.5 and 42 on the yield of methyl esters formed for supercritical methanol treatments, assuming that the average molecular weight of rapeseed oil is 806 as triglycerides. Fig. 2 shows the obtained HPLC chromatograms of rapeseed oil as treated in various molar ratios for 4 min under

supercritical conditions. In the previous study [5], it was demonstrated that the intensive peak in the chromatogram observed in the short retention times (310 min) are methyl esteried compounds, while in the longer retention times, intermediates such as monoglycerides and diglycerides appeared (1020 min). Therefore, from Fig. 2, it is apparent that the conversion state of rapeseed oil is different as various molar ratios of methanol were applied to the transesterication reaction of the rapeseed oil. With a higher molar ratio of methanol applied, the methyl esteried compounds are increased with a decrease in the intermediate compounds. Fig. 3 shows the content of methyl esters produced as different supercritical treatments were carried out at 3508C. For a molar ratio of 42 in methanol, almost complete conversion was achieved in a yield of 95% of methyl esters, whereas for the lower molar ratio of 6 or less, incomplete conversion was apparent with the lower yield of methyl esters. These lines of evidence, therefore, indicate that the higher molar ratios of methanol result in the better transesterication reaction, due perhaps to the increased contact area between methanol and triglycerides. 3.2. Effect of temperature on methyl esters formation To determine the effect of temperature on methyl esters formation, transesterication reactions of rapeseed oil were carried out with a xed molar ratio of 42 in methanol, the best condition found in Fig. 3, at various temperatures ranging from 200 to 5008C. Fig. 4 shows the obtained HPLC chromatograms of rapeseed oil as treated in various conditions of temperatures and reaction times, while the content of methyl esters obtained is shown in Fig. 5, in which the obtained experimental data are shown by the symbols, whereas the simulated curves are shown by the lines as discussed later. At temperatures of 200 and 2308C, the relatively low conversion to methyl esters is evident in Figs. 4 and 5 due to the subcritical state of methanol. In these conditions, methyl esters formed are at most about 68 and 70% at 200 and 2308C, respectively, at 3600 s (1 h) treatment. These results are in good accordance with those already reported [10]. At a temperature of 2708C, the conversion rate is still low which might be related with the stability of supercritical condition. As can be seen in Fig. 1, maximum pressure reached in this treatment is 14 MPa, still in the transition between subcritical and supercritical state of methanol. However, at 3008C, a considerable change in the conversion rate can be seen with about 80% of methyl esters produced in 240 s. As observed in the previous study [5], at 3508C, 240 s treatment resulted in a high conversion of rapeseed oil to methyl esters with its yield of 95%. An important result here is that the composition of methyl esters yielded is very similar with that prepared by the conventional commercial process with alkaline catalyst.

Fig. 4. HPLC chromatograms of rapeseed oil as treated at various conditions of temperatures and reaction times with molar ratio of 42 in methanol.

696

D. Kusdiana, S. Saka / Fuel 80 (2001) 693698

erides react to produce monoglycerides. Finally monoglycerides react with methanol to give glycerin as a by-product. At each reaction step, one molecule of methylated compounds is produced for each molecule of methanol consumed. As a result, six different rate constants of the reaction are reported for the whole reaction. Due to reality that nal products for the whole reaction in the transesterication reaction for biodiesel fuel production are methyl esters with glycerin, we dened a simpler mathematical model for this reaction by ignoring the intermediate reactions of diglycerides and monoglycerides, so the 3 steps can be simplied to be one step as follows:
Fig. 5. Effect the reaction temperature on the methyl esters formation. The experimental data are presented by the symbols, whereas the solid lines are simulated curves based on Eqs. (3) and (6).

At even higher temperature of 4008C, the transesterication reaction is essentially completed for 120 s to convert almost all rapeseed oil to their methyl esters. However, in such a high reaction temperature, new peaks in the shorter retention time (34 min) are dominating in the HPLC chromatograms as shown in Fig. 4 and the previous study [5]. This indicates that decomposition reaction takes place at temperature above 4008C due to the thermal degradation. As a result, the transesterication reactions of rapeseed oil to methyl esters proceed appropriately at temperature of 3508C under supercritical condition of methanol without any catalyst used. 3.3. Kinetics of rapeseed oil to methyl esters To correlate experimental data and to quantify the temperature and reaction time effects observed above, the experimental results were analyzed further in terms of the kinetics of rapeseed oil to methyl esters. As mentioned earlier, the model is based on overall reaction. Since the molar ratio of methanol to rapeseed oil was xed to be 42, the concentration of methanol was not taken into account, as reported by other researchers [10,12]. Diasakov [10] proposed the thermal transesterication reaction to be divided into 3 steps. Triglycerides react with methanol to produce diglycerides, and then diglyc-

This reaction is assumed to proceed in the rst order reaction as a function of the concentration of triglycerides (TG) and reaction temperature. The rate constant of the reaction can be determined based on the increased amount of the product that occurs in some reaction time interval [13,14], or alternatively, based on the decreased amount of one reactant. In this work, the decreased amount of one reactant, that is TG, was chosen. Therefore, the rate constant of the reaction can be given by Eq. (1) Rate 2d TG dt 1

where [TG] refers to the content of vegetable oil used in this study. In this supercritical methanol method, three species were dened as methyl esters (ME), glycerin (GL) and unmethyl esteried compounds (uME) which include triglycerides, diglycerides, monoglycerides and unreacted free fatty acids. Therefore, Eq. (1) can be modied to be Rate 2d or 2duME kuME dt 3 uME dt 2

where [uME] refers to the content of the species, excluding methyl esters and glycerin, that result or remain after the supercritical treatment was carried out. Assuming that the initial concentration of uME, at time t 0; is uME, 0 and that it falls down to uME, t at some later time t, the integration gives
Fig. 6. Semilog plot of unmethyl esters content in rapeseed oil during transesterication reaction. Legends see Fig. 5.

uME;t duME t k dt uME;0 uME 0

D. Kusdiana, S. Saka / Fuel 80 (2001) 693698 Table 1 The rate constant of transesterication reaction Reaction condition Temperature (8C) 200 230 270 300 350 385 431 487 Pressure (MPa) 7 9 12 14 19 65 90 105 0.0002 0.0003 0.0007 0.0071 0.0178 0.0249 0.0503 0.0803 k (s 21)

697

and 2ln or K lnuME;t 2 lnuME; 0 t 6 uME;t kt uME; 0 5

Fig. 6 shows the correlation between the content of unmethyl esterifed compounds and reaction times. As mentioned previously, unmethyl esteried compounds are dened as other compounds obtained from the upper portion excluding ve types of methyl esters, such as methyl palmitate, methyl oleate, methyl stearate, methyl linoleate and methyl linolenate. The straight line was determined to t the data in order to adopt the rst order rate equation. Based on the results in Fig. 6, the rate constant was obtained for each reaction temperature as shown in Table 1 and the corresponding Arrhenius plot for this method is presented in Fig. 7. It is evident that at subcritical temperature below 2398C, the reaction rates are so low but much higher at supercritical state, with the rate constant increased by a factor of about 85 at the temperature of 3508C. Liquid methanol is a polar solvent and has hydrogen

bondings between OH oxygen and OH hydrogen to form methanol clusters. Because the degree of hydrogen bonding decreases with increasing temperature, the polarity of methanol would decrease in supercritical state. This means that supercritical methanol has a hydrophobic nature with the lower dielectric constant. As a result, non-polar triglycerides can be well solvated with supercritical methanol to form a single phase of vegetable oil/methanol mixture. This phenomenon with the high temperature conditions seems to be likely to promote transesterication reaction of rapeseed oil. The simulation was made on a relationship between the formation of methyl esters and reaction times, based on Eqs. (3) and (6) to examine the tness of the experimental results, as shown in Fig. 5. In this gure, the simulated curves are shown by lines and the experimental data are represented by symbols. In the subcritical temperature, simulated curves are somewhat different from those of experimental data. This would be because at the longer treatment, the conversion rate is low due to the equilibrium reaction approached. However, at the supercritical state, the simulated curves t well with the experimental results in all cases. Therefore, a simple method proposed to determine the rate constants in transesterication must be valid.

4. Concluding remarks A highly efcient transesterication process has been described and the proposed kinetics in transesterication of rapeseed oil has been proven to t very well with those of experimental data. A reaction temperature of 3508C with the molar ratio of methanol being 42 were considered as the best condition for a free-catalyst process of biodiesel fuel production. The supercritical methanol method, therefore, offers a potentially low cost method with simpler technology for producing an alternative fuel for compression ignition engines. The considerable yield of methyl esters by the environmentally friendly method renders this technique ideally suited for industrialization.

References
[1] Serdari A, Lois E, Stournas S. Ind Engng Chem Res 1999;38: 3543. [2] Aksoy HA, Karaosmanoglu BF, Yatmaz HC, Civelekoglu H. Fuel 1990;69:600. [3] Schwab AW, Bagby MO, Freedman B. Fuel 1987;66:1372. [4] Boocock DGB, Konar SK, Mao V, Lee C, Buligan S. JAOCS 1998;75:1167. [5] Saka S, Dadan K. Biodiesel fuel from rapeseed oil as prepared in supercritical methanol. Fuel 2001;80:225. [6] Deslandes N, Bellenger V, Jafol F, Verdu J. Appl Polym Sci 1998;69:2663. [7] Dufek EJ, Buttereld RO, Frankel EN. JAOCS 1972;49:302. [8] Freedman B, Buttereld RO, Pryde EH. JAOCS 1986;63:1375. [9] Noureddin H, Zhu D. JAOCS 1997;74:1457.

Fig. 7. First order reaction rate constant in Arrhenius plot of rapeseed oil in methanol during transesterication reaction.

698

D. Kusdiana, S. Saka / Fuel 80 (2001) 693698 [13] Barrow GM. Physical chemistry. Tokyo: McGraw-Hill Kokusha Ltd, 1973 (p. 419). [14] Steinfeld JI, Francisco JS, Hase WL. Chemical kinetics and dynamics. New York: Prentice Hall, 1989 (p. 6).

[10] Diasakov M, Loulodi A, Papayannakos N. Fuel 1998;77:1297. [11] Encinar JM, Gonzalez JF, Sabio E, Ramino MG. Ind Engng Chem Res 1999;38:2927. [12] Freedman B, Pryde CH, Mounts TL. JAOCS 1984;61:1638.

Das könnte Ihnen auch gefallen