You are on page 1of 11

CONTINUOUS PROCESS FOR THE PRODUCTION OF FATTY ACID METHYL ESTERS IN A FALLING LIQUID LIQUID FILM REACTOR

Paulo Csar Narvez*, Lady Zamira Castaeda, Sandra Milena Rincn, Francisco Snchez, Jess Torres
Universidad Nacional de Colombia, Faculty of Engineering, Department of Chemical Engineering Abstract. This study introduces the results of an evaluation made to a falling liquid liquid film reactor involved in the continuous production of fatty acid methyl esters (specifically of methyl esters from palm oil). The study analyzed the influence of temperature, percentage of NaOH, and palm oil-methanol ratio on constant values for oil residence time and Reynolds number. Results show that it is possible to perform the palm oil methanolysis in this type of reactor, thus facilitating separation operations and reducing problems arising from the use of sodium hydroxide as a catalyst. Additionally, results for the process simulation employing the ASPEN PLUS TM tool are also included.

Key words: Methyl Esters, Bio-diesel, Falling Liquid Liquid Film Reactor.

1. Introduction
Fatty acid methyl esters (FAME), which are a product of transesterification of oils and fats, are used as raw materials in the production of fatty alcohols, sulphonated methyl esters, metallic soaps, fatty amines and amides, industrial processes in which they have gradually replaced fatty acids (MPOB, 2000; Baileys, 1996). Additionally, these esters are used as solvents and as additives in formulations for adhesives and renewable fuels, and as partial or total replacements of petrochemical diesel (bio-diesel). When comparing production processes involving fatty acids and methyl esters, the latter prove to have advantages, since they require less critical operation conditions during methanolysis and, therefore, generate less energy consumption and lower investments in equipment. Furthermore, processes involving methyl esters generate higher concentrations of glycerin and have the presence of methanol instead of water, meaning less costs relating to purification processes for this sub-product. Commonly, industrial processes involved in the production of fatty acid methyl esters use basic homogeneous catalysts, particularly sodium hydroxide. The catalytic action of bases is due to the presence of methoxide ions, which may be introduced as alcoholate in the reactive medium. Variables which have a significant influence on fatty-compound methanolysis, as catalyzed by alcohol-soluble hydroxides, include the presence of water and free fatty acids (FFA), the alcohol:oil molar ratio, temperature, and mixing. The presence of water and free fatty acids has a negative impact on the production of methyl esters (Freedman et al., 1984). Cvengros and Povazanec (Cvengros, 1996) suggested that this negative effect is the consequence of the formation of catalytically inactive soaps (product of FFA neutralization with the catalyst, or through hydrolysis of triglycerides followed by neutralization), which emulsify the phase rich in glycerin within the organic phase. The stoichiometric molar ratio for alcohol:oil is 3:1. As the ratio increases, the amount of monoglycerides (MG), diglycerides (DG) and triglycerides (TG) present in the final product diminish, and the production of methyl esters increases. Results suggest that at a 6:1 ratio the maximum production of FAME is achieved, and that higher ratios have no further effect on the increase in contents of FAME, but difficult the

Universidad Nacional de Colombia, Faculty of Engineering, Department of Chemical Engineering Ciudad Universitaria, Laboratorio de Ingeniera Qumica, Oficina 212, Bogot D. C., Colombia. Fax 3165617. pcnarvaezr@unal.edu.co

separation and purification process (Freedman et al., 1984). The rise in temperature has an impact on production of FAME, increasing reaction velocity. After 0,1 hours of reaction in experiments performed at 32C, 45C, and 60C, with a methanol:oil ratio of 6:1, and using NaOH as catalyst, the contents of FAME was 64%, 87%, and 94%, respectively. However, after one hour of reaction, the contents was basically the same for experiments at 60C and 45C, and slightly lower for experiments at 32C (Freedman et al., 1984). Agitation is an important variable when considering the lack of miscibility between methanol and oil. Initially, reactants form a two (non-miscible) liquid system, and the reaction speed is controlled by diffusion. Once the reaction starts, given the emulsifying nature of DG and MG, the reactant mixture transforms itself into a semi-homogeneous system. In lengthier reaction times, in which concentrations of these intermediates are low, the mixture separates into two phases: a phase rich in ester and a phase rich in glycerol (Jordan, 2001; Srivastava, 2000). One of the critical variables of reaction is the mixing intensity, which according to studies by Noureddini and Zhu (Noureddini, 1997) must approach Reynolds numbers above 10.000. Assmann suggests the use of Reynolds numbers above 2300 for continuous processes in a tubular reactor (Assmann, 1996). Recent researches have intended to find a catalyst that may replace NaOH, looking to reduce the notorious influence of present water and free fatty acids, and to avoid the formation of sodium soaps that diminish catalytic activity, increase viscosity, drive formation of gels, and generate difficulties during the separation and purification stage (Fukuda, 2001; Gryglewicz, 1999; Leclercq, 2001, Ma, 1999b). This study evaluates the use of a falling liquid liquid film reactor in the production of fatty acid methyl esters. The reactor is a vertical tube with a structured packing in its interior that enables the creation of an interfacial area without dispersing droplets from one phase into the other. The use of this type of reactor enhances the advantages associated to homogenous alkaline catalysts, diminishing problems related to soaps and emulsions formation; this, while keeping in mind that turbulence required for contacts between reacting phases is eliminated: the flow regime is laminar and the reaction takes place in the interface between oil and methanol films with catalyst. The mass transfer velocity between the phases increases, due to the distance required for molecules to diffuse and react is shorter, and the chemical balance is displaced toward the desired product, as a consequence of the diffusion of glycerine molecules into the alcoholic phase.

2. Methodology
2.1. Materials The following materials were used in the experiments: refined, bleached and deodorized palm oil, edible grade, produced by company PALMALI LTDA, Bogot D. C.; anhydride methanol at 99.8% purity, and sodium hydroxide at analytical grade (MERCK); chromatographic standards: methyl palmitate, methyl oleate, glyceryl 2-monopalmitate, glyceryl dipalmitate (mix of isomers), glyceryl tripalmitate and glyceryl trioleate (Sigma Aldrich Chemical Company); silylant agent: N-Bistrimethylsilyltrifluoroacetamide (BSTFA) (Sigma Aldrich Chemical Company); internal standard: tricaprine (Fluka). 2.2. Equipment Experiments were carried out in a reaction system as shown in Figure 1. The reaction system is made up of two precision scales of 0.1g, which together with a precision chronometer of 0.01s, enable the measurement

of the mass flow of methanol and oil; two metering pumps EMEC HMS EXT 2001with digital flow control and geared with pulsation dampeners at discharge; a stainless-steel 304 reactor with 30 cm of effective length, 1 nominal diameter, and structured packing inside; a two-phase separator; and, temperature controls with a precision of 0.1C.

LC 290 2

LC 2901

T-2902
TIC 2901 LG 2901 TE 2906

T-2901

TI 290 1

PI 2902

TE 29 01

HE-2901
PI 2901

HE-2902

TE 2904

TIC 2904

R-2901
TE 2903

TIC 2903

FI 2902

FI 2901 TIC 2902 LC 2903

P-2902

P-2901

TI 2902 FC 2902 FC 2901 HLC 2901 TE 2905 LLC 2901 A TANQUE DE ALMACENAMIENTO PI 2903

SD-2901
TE 2902

LG 2902

A TANQUE DE ALMACENAMIENTO

Figure 1. Experimental liquid liquid film reactor

2.3. Procedure The reaction system was preheated at 50C to avoid oil solidification. Palm oil was loaded with the aid of the metering pump, and the oil flow and the feeding temperatures for oil, the reactor and the separator were adjusted to meet the values required for the experiment. Oil flow in all experiments remained constant at 1,8 kg/min. Once the operating conditions were defined, the methanol solution, previously prepared to the concentration required was fed to the reactor. During the initial 10 minutes, the entry temperature and the methanol feeding flow were adjusted. Then, 10 ml samples were taken every 20 minutes, until completing 200 minutes, counted as of the start of the methanol feed. These samples were decanted for a minute, and a subsequent sample of approximately 10 mg was taken and further silylated adding BSTFA and pyridine, according to the procedure IUPAC Standards methods for the analysis of oils, fats and derivatives, 7th Edition. Then, samples were stored at 4 C, and further analyzed through gas chromatography. Gas chromatography was performed in a Hewlett-Packard 5890 Series II chromatograph equipped with a spli/splitless injection system, a flame-ionization detector, and HP Chemstation software. The column was a SUPELCO SGE HT-5 (aluminum clad, 12m x 0,53mm x 0,15m) with N2 at 7,10 cm/s as the carrier gas and a split ratio of 50:1. Injector temperature was 250 C and detector was 450 C, oven temperature started at 120 C increased at 350 C at a rate of 20C/min, and held at this temperature for 10 minutes.

3. Results and Discussion


Figure 2 shows products and oil concentrations during methanolysis at 50C, with methanol:oil molar ratio of 6:1, 1% of NaOH as catalyst, and residence time for oil of 8 minutes. The reactor reaches a stable status 40 minutes after beginning feed of methanol, a figure that matches the average residence time for methanol (31 minutes). For an average residence time of 8 minutes, using 1% of NaOH as catalyst and a methanol:oil molar ratio of 6:1, a product with a methyl esters weight percentage of 57%, 10% monoglycerides, 15% diglycerides, and 18% triglycerides is achieved. Using a CSTR reactor, Darnoko and Cheryan (Darnoko, 2000) achieved a methyl ester percentage of up to 80%, with a residence time of 60 minutes. Experiments performed in a stirred tank reactor under the same conditions resulted in a product with 89% of methyl esters after 8 minutes of reaction. Considering that in methanolysis of fatty compounds the contact area between reacting phases is a key factor, results show that effects regarding transportation phenomena effectively decrease the percentage of methyl esters in the product, and that with longer residence times in the reactor, better results should be achieved. Furthermore, although in the CSTR reactor one phase disperses into the other, the contact area generated by the seal of the falling film reactor seems to be larger than that achieved by Darnoko and Cheryan (Darnoko, 2000) in their work; this explains the high concentration of methyl esters in the product for a significantly shorter time of residence.

1.00 0.80

M ass fr ac ti o

0.60 0.40 0.20 0.00 0 20 40 60 80 100 120 Time (minutes)


MG DG

140

160

180

200

FAME

TG

Figure 2 Concentrations of products and oil against operation times: falling liquid liquid film reactor, 50C temperature, alcohol:oil molar ratio of 6:1, 1% NaOH.

Results of batch processes show that the influence of temperature on the 50-60C range is insignificant.
Keeping in mind, that the final temperature of the fusion for palm oil is 47C, and the boiling temperature for methanol at the working pressure is 55C, two different levels of temperature in the falling film reactor were evaluated: 50C and 55C.

Results show that as temperature increases, concentrations of methyl esters decrease, as a consequence of methanol evaporation within the reactor. Productivity of methyl esters diminishes from 2,20 to 1,84 mol of ester

per mol of oil. Figure 3 presents the impact of temperature on concentrations of methyl esters.

100 C o n c e n tr ati o n (% w e i 90 80 70 60 50 40 30 20 10 0 0 20 40 60 80 100 120 140 160 180 200 Time (minutes)
50 C 55 C

Figure 3 Impact of temperature on percentage of methyl esters from palm oil: falling liquid liquid film reactor, alcohol:oil molar ratio of 6:1, 1% NaOH.

Figure 4 shows the effect of catalyst percentage variations on palm-oil methanolysis in a falling film reactor, at 50C and methanol:oil molar ratio of 6:1. The positive impact of increasing the percentage of catalyst on the concentration of methyl esters can be clearly seen here.

100 Concentration (% weight) 80 60 40 20 0 0 20 40 60 80 100 120 140 160 180 200 Time (minutes)
0, 50 % 0,75 % 1%

Figure 4 Effect of catalyst percentages on methyl esters percentages from palm oil: falling liquid liquid film reactor, 50C temperature, alcohol:oil molar ratio of 6:1.

The phase separation time for the experiment with 1% of catalyst was 3 minutes, and no gelation of the phase rich in glycerin was noticeable until two hours after the sample had been taken. Figure 5 shows the impact of catalyst concentrations on palm-oil conversion, and Table 1 illustrates the influence of catalyst percentages on conversion and productivity. Since residence time and flows for methanol are identical, difference may be attributed almost exclusively to kinetic effects.

90 80 Conversion (% ) 70 60 50 40 30 20 40
1%

60

80

100

120
0,75%

140

160

180
0,50%

200

Time (minutes)

Figure 5. Effect of catalyst percentages on palm-oil conversion: falling liquid liquid film reactor, 50C temperature, alcohol:oil molar ratio of 6:1.

Tabla 1. Influence of catalyst percentage on conversion and productivity of palm-oil methanolysis in a falling liquid liquid film reactor: 50C temperature, methanol:oil molar ratio of 6:1; and, time of residence: 8 minutes for oil and 31 minutes for methanol

Catalyst Percentage (%) 0,50 0,75 1,00

Conversion (%) 43,47 60,58 78,50

Productivity (mol of product /mol de oil that reacts) FAME MG DG 1,45 0,32 0,54 1,61 0,35 0,45 2,20 0,26 0,22

Freedman et al. (Freedman et al., 1984) studied the effect of the methanol:oil ratio on methanolysis for soy, sunflower, peanut, and cottonseed oils, and concluded that as the molar ratio approaches the stoichiometric value, the concentration of methyl esters decreases and that of monoglycerides and diglycerides increases, thus showing that conversion into esters is not complete. Likewise, their results suggest that ratios higher than 6:1 do not increase the productivity of methyl esters and rather difficult separation and purification operations. Three oil:alcohol ratio 3:1, 6:1 and 12:1, was evaluated in the liquid-liquid film reactor. In order to analyze this behavior, it is important to keep in mind that although the residence time for oil is the same in the three experiments, it changes for methanol (62 minutes for 3:1 ratio experiment, and 15.5 minutes for 12:1 ratio experiment). Figure 6 shows the influence of the methanol:oil molar ratio on the concentration of methyl esters.

100 90 80 70 60 50 40 30 20 10 0 0 20 40 60 80 100 120 140 160 180 200

Concentration (% weight)

Time (minutes)
3a1 6a1 12 a 1

Figure 6. Effect of alcohol:oil molar ratio on methyl ester percentages from palm oil: falling liquid liquid film reactor, 50C temperature, and 1% of NaOH.

The increase of the methanol:oil molar ratio up to 12:1 generated a drastic reduction in concentrations of methyl esters in the falling liquid liquid film reactor, as a consequence of shorter times of residence for methanol and increases in film thickness, which involve increases in the resistance to mass transfers and reductions in transfer velocity. Table 2 shows the results for conversion and productivity for these experiments.
Table 2. Influence of the methanol:oil molar ratio on the conversion and productivity levels of palm-oil methanolysis in a falling liquid liquid film reactor: 50C temperature, 1 % of NaOH, and time of residence of 8 minutes for oil

Methanol:oil molar ratio 3:1 6:1 12:1

Conversion (%) 70,62 78,50 36,39

Productivity (mol of product/mol of oil that reacts) FAME MG DG 1,89 0,32 0,33 2,20 0,26 0,22 1,30 0,32 0,62

It is important to point out that residence time for methanol in the 3:1 ratio experiment is twice than that for the 6:1 ratio experiment, lowering conversion and productivity differences and attenuating the effect of alcohol in excess. For an eight minute reaction at the agitation tank reactor, while under the same conditions, conversion amounts to 96% for a 6:1 ratio and to 80% for a 3:1 ratio.

4. Process Simulation
Process simulations support the gathering of information on the process behavior under specific operating conditions. This work builds on the use of specialized tools called simulators, which include thermodynamic packages and unit operation models, as well as calculation strategies for their resolution. In this study, the

ASPEN PlusTM tool was used to perform the simulation of a process proposed for the generation of fatty acid methyl esters from palm oil in stable conditions. The procedure needed to perform the simulation implies the definition of substances involved, the selection of a thermodynamic package and process units, including operating conditions such as temperature, pressure, material flow, and alike. In this case, the simulator databases have comprehensive information available on methanol, glycerol, and water. As for palm oil and its methyl ester, such databases do not include the information required for the simulation. Considering that triglycerides with radicals from oleic and palmitic acids feature the highest ratios within the composition of palm oil (Baileys, 1996; Goncalves, 2004), palm oil was represented as a mixture of triolein and tripalmitin, and its methyl ester as a mixture of methyl oleate and methyl palmitate. Parameters were set for the calculation of properties of these compounds through the introduction of properties such as boiling point, combustion heat, and density, specific heat and viscosity against temperature. Once the parameters were calculated, the calculation models and routes were defined. Finally, the values estimated by the simulator were matched against data gathered via the experiments. Considering the process operation characteristics, the presence of two non-polar compounds (palm oil and methyl ester) and two highly-polar compounds (methanol and glycerin), and the fact that the esterification reaction takes place in the liquid phase, the decision was made to use a method based on the calculation of activity coefficients (Aspen Technology Inc., 2001; Carlson, 1996). Within the group of possibilities available in the simulator, the UNIFAC method was chosen, since, in contrast with the UNIQUAC and NRTL models which require molecule-interaction parameters - this is a group-contribution model. Literature has reported that predictions on liquid-liquid balance approximate to experimental data when using models such as UNIQUAC, NRTL and UNIFAC for systems covering compounds similar to those that make up the process (Batista et al., 1999a; Batista et al., 1999b; Goncalves, 2004). The percentage differences between the thermodynamic and transportation properties calculated by the simulator and the properties found in literature for glycerin and methanol, as experimentally defined for the oil and ester, are less than 6%. For calculation of liquid-liquid balance by the UNIFAC model, a set of binary-interaction parameters was found for the groups that make up the molecules. The process simulation mainly includes reaction units, separation, heat exchange and distillation. Operating conditions for the reaction stage were selected according to the optimal values reported in literature: 60C, methanol:oil molar ratio of 6:1, and 1% of NaOH as catalyst (Freedman et al., 1984). Furthermore, a tubular reactor model was used, as well as the information on the global kinetics which was experimentally defined by using an agitation-tank reactor operating in batch mode. The objective of separation and distillation units is to separate and purify the ester attained in the reaction of the other components in the mixture, methanol in excess, oil that did not react, glycerin produced, and water employed in the rinsing stages. Distillation columns are also targeted to recover methanol, which is further recirculated, and to purify glycerin, which is a sub-product of significant industrial importance. The distillation is
favoured by the difference in the points of boil of the methanol (64,7C), glycerine (287,85C) and methyl esters (194C).

However, it should be noted that these operations use vacuum pressures, since both glycerol and the ester may decompose at high temperatures. Figure 7 shows a flowsheet for the process, as well as the simulation results. The simulated process proved to be suitable for the production of high quality, palm-oil methyl esters featuring all the characteristics included in Table 3.

Figure 8. Process Flow Diagram for Palm Oil Methyl Ester Continuous Production

Table 3. Process Simulation Results

Stream Temperature (C) Pressure (bar) Molar flow (kmol/h) Mass flow (kg/h) Volumetric flow (m3/h) Specific heat (kJ/kg K) Density (kg/m3) Viscosity (cP) Mass fraction Palm oil Methyl ester Metanol Glycerin Water Stream Temperatura (C) Presure (bar) Molar flow (kmol/h) Mass flow (kg/h) Volumetric flow (m3/h) Specific heat (kJ/kg K) Density (kg/m3) Viscosity (cP) Mass Fraction Palm oil Methyl ester Methanol Glycerin Water

1 3 4 5 7 8 9 10 11 45,0 25,0 60,0 60,0 60,0 25,0 25,0 25,0 25,0 1,01 1,01 1,01 1,01 1,01 1,01 1,01 1,01 1,01 1,19 3,77 8,27 8,27 3,73 4,54 3,89 8,43 3,98 1000,00 120,74 1227,00 1227,00 190,49 1036,51 70,00 1106,51 1015,78 1,16 0,15 1,47 1,46 0,20 1,22 0,07 1,29 1,19 1,88 2,53 2,12 2,35 2,67 2,12 4,18 2,25 2,11 865,60 789,62 835,61 842,43 964,97 849,69 994,70 857,60 851,56 34,28 0,54 0,65 1,83 1,97 3,81 0,91 1,97 5,13 1,000 0,000 0,000 0,000 0,000 12 25,0 1,01 4,45 90,73 0,10 3,75 931,54 0,84 0,000 0,000 0,263 0,003 0,734 0,000 0,000 1,000 0,000 0,000 13 58,7 0,20 0,51 54,00 69,34 1,63 0,78 0,01 0,000 0,824 0,128 0,000 0,047 0,815 0,000 0,185 0,000 0,000 14 58,7 0,20 3,11 846,00 1,02 2,28 825,73 2,57 0,000 0,996 0,003 0,000 0,001 0,022 0,797 0,094 0,087 0,000 15 130,8 0,30 0,36 115,78 0,15 2,58 777,97 0,69 0,228 0,772 0,000 0,000 0,000 0,000 0,009 0,433 0,559 0,000 17 28,6 0,20 0,74 23,85 0,03 2,56 786,15 0,51 0,000 0,000 1,000 0,000 0,000 0,026 0,942 0,032 0,000 0,000 18 100,0 1,01 3,70 66,88 0,07 4,21 957,28 0,28 0,000 0,000 0,000 0,003 0,996 0,000 0,024 0,000 0,883 0,000 0,030 0,000 0,000 1,000 0,063 19 20 15,5 232,6 0,10 0,20 2,58 1,16 84,09 106,40 0,11 0,10 2,47 3,49 799,73 1107,76 0,62 1,74 0,000 0,020 0,980 0,000 0,000 0,000 0,000 0,000 1,000 0,000 0,026 0,961 0,009 0,000 0,003 21 18,4 1,01 3,32 106,26 0,13 2,49 795,86 0,59 0,000 0,000 1,000 0,000 0,000

Conclusions
This study evaluated the production of fatty acid methyl esters using a continuous process in a falling liquid liquid film reactor, while using sodium hydroxide as a catalyst; more specifically, the study addressed the production of methyl esters from palm oil. Within the studied interval (50-55C), the temperature effect is not significant, although increases in temperature translate into decreases in productivity of methyl esters, due to methanol evaporation. As for constant times of residence for oil and methanol, increases in percentages of NaOH within the 0.50-1% interval against oil generate increases in concentration of methyl esters, and thus, in the oilconversion and methyl-ester productivity levels. As for the methanol:oil molar ratio, the maximum conversion and productivity values were achieved at a 6:1 ratio; furthermore, using methanol in excess (molar ratio of 12:1) does not generate any increase in the concentration of methyl esters,
transfer speed and not by kinetics effects. since the reactor's behavior is governed by mass

For an 8-minute time of residence for palm oil at 50C temperature, 1% of

NaOH (against oil) as catalyst, and methanol:oil molar ratio of 6:1, the conversion rate for palm oil reached 78.50, while productivity yielded 2.20 mol of methyl esters for each mol of reacting oil. Additionally, the process simulation was performed in the ASPEN PLUSTM equipment, generating parameters and selecting calculation models and routes for physical, thermodynamic and transportation properties of oil and its methyl esters, showing differences of less than 6%, as compared to the values experimentally available. Based on the simulation results, it could be said that the process is suitable for continuous production of methyl esters deriving from palm oil.

References
ASPEN Technology Inc. ASPEN Engineering suite V. 11.1 documentation. 2001 Assmann, G., et al. Continuous process for the production of lower alkyl esters, U. S. Patent 5,514,820, May 1996. Batista, E. Monnerat, Kato, K., S. Stragevitch, C., Meirelles, A. Liquid Liquid Equilibrium for Systems of Canola Oil, Oleic Acid, and Short Chain Alcohols. J. Chem. Eng. Data, Vol. 44, 1999a, pp. 1360 - 1364. Batista, E. Monnerat, S. Stragevitch, L, Pina, C., Goncalves, C. Prediction of Liquid Liquid Equilibrium for Systems of vegetable Oils, Fatty Acids, and Ethanol. J. Chem. Eng. Data, Vol. 44, 1999b, pp. 1365 - 1369. Cvengros, J., Povazanec, F., Production and treatment of rapeseed oil methyl esters as alternative fuels for diesel engines. Bioresour. Technol., 55, 1996, pp.145-152. Bailey, A., Bailey's Industrial Oil and Fat Products, Volume 5, Edited by Hui, Y. H., John Wiley & Sons, USA, 1996. Carlson, E. Don`t Gamble with physical Properties for simulations. Chem. Eng. Prog., October, 1996, pp. 35-46. Darnoko D., Cheryan, M. Kinetics of palm oil transsterification in a batch reactor. J. Am. Oil Chem. Soc., 77, 1262-1266, 2000. Freedman, B., Pryde, E. H., Mounts, T. L. Variables affecting the yields of fatty esters from transesterified vegetables oils. J. Am. Oil Chem. Soc., Vol. 61, No. 10, 1984, pp. 1638-1643. Fukuda, H., Kondo, A., Noda, H. Biodiesel fuel production by transesterification or oils. J. Biosc. Bioeng., Vol 92 No. 5, 405-416 2001. Goncalves, C., Meirrelles, A. Liquid-liquid equilibrium data for the system palm oil + fatty acids + ethanol + water at 318,2 K. Fluid Phase Eq., Vol 221, 2004, pp. 139-150 Gryglewicz, S. Rapeseed oil methyl esters preparation using heterogeneous catalyst. Bioresour. Tecnol., 70, 1999, pp. 249253.

IUPAC, Standards methods for the analysis of oils, fats and derivatives, 7th Edition. Jordan, V., Gutshe, B. Development of an environmental benign process for the production of fatty acid methyl esters. Chemosphere, Vol. 43, 2001, pp. 99-105. Leclercq, E., Finiels, A., Moreau, C. Trasesterification of rapeseed oil in the presence of basic zeolites an related solid catalyst. J. Am. Oil Chem. Soc., Vol. 78, No. 11, 2001, pp. 1161-1165. Ma, F., Hanna, M. A., Biodiesel production: a review. Bioresour. Technol., Vol. 70, 1999b, pp. 1-15. Merichem Company, Technical Bulletin FFC-RI-1182, An improved method of mass transfer for refinery treating process, 2004, in www.merichem.com MPBO, Advances in Oil Palm Research, Volume II, Malasya, 2000. Noureddini,H., Zhu, D. Kinetics of transesterification of soybean oil. J. Am. Oil Chem. Soc., Vol. 74, No 11, 1997, pp. 1457-1462. Srivastava, A., Prasad, R. Triglycerides-based diesel fuels. Renew. Sust. E., Vol. 4, 2000, pp. 111-133.

Acknowledgments
This work is part of the project: Production of sulphonate methyl esters from Palm Oil, which is supported by COLCIENCIAS (Contrat RC 389) and the Universidad Nacional de Colombia. The authors acknowledge Laboratorio de Ingeniera Qumica, Universidad Nacional de Colombia and Instituto de Tecnologa Qumica de la Universidad Politcnica de Valencia, in particular Dr. Avelino Corma, Sara Iborra and Alexandra Velty.

You might also like