You are on page 1of 7

A NEW SOLUBILITY MODEL TO DESCRIBE BIODIESEL FORMATION KINETICS

K. Gunvachai1, M. G. Hassan2, G. Shama2 and K. Hellgardt1,


1 2

Department of Chemical Engineering and Chemical Technology, Imperial College London UK. Department of Chemical Engineering Loughborough University, UK.

Abstract: The development of a mathematical model that accurately describes the formation of fatty acid methyl esters (FAMEs) (otherwise known as biodiesel) from rapeseed oil and methanol in the presence of sodium hydroxide catalyst at temperatures in the range of 293333 K is described. The model accounts for the existence of more than one phase at any given stages of the reaction and for the solubilities of the reacting species in various phases. The latter were obtained experimentally and are presented in the form of ternary phase diagrams. The model appears to give a good t to the experimental data obtained using both batch and continuous ow reactors. Keywords: biodiesel; kinetics; solubility; phase diagram; transesterication.

INTRODUCTION
There is considerable interest in the direct use of vegetable oils as well as vegetable oil or animal fat derived biodiesel as renewable liquid fuels (Meher et al., 2006; Boocock et al., 1998). Whilst apparently a simple option, there are some limitations that mitigate against the use of vegetable oils. For example, the viscosities of plant oils are very much higher than those of diesel fuels and this leads to engine-start difculties in colder climates (Ma and Hanna, 1999). The fuel injection systems of engines running on vegetable oils also display a tendency to become blocked by gummy deposits (Srivastava and Prasad, 2000; Meher et al., 2006). Moreover, vegetable oils do not burn cleanly but produce aldehydes such as acrolein as well as organic acids (Goering et al., 1981). To overcome these problems, vegetable oils can be converted to their methyl esters, so called biodiesel. Biodiesel has a viscosity of typically six to 10 times lower than that of the vegetable oils and does not form engine deposits or generate harmful pollutants. Since biodiesel can be used in diesel engines in neat form or blended with diesel fuel with little or no engine modication, it can be considered as a potential diesel fuel extender (Canakci and Van Gerpen, 2003). Other advantages of biodiesel over conventional diesel fuel have also been identied. For instance, the exhaust
383

Correspondence to: Dr K. Hellgardt, Department of Chemical Engineering and Chemical Technology, Imperial College London, South Kennington Campus, London, SW7 2A2, UK. E-mail: k.hellgardt@ imperial.ac.uk

DOI: 10.1205/psep07033 09575820/07/ $30.00 0.00 Process Safety and Environmental Protection Trans IChemE, Part B, September 2007 # 2007 Institution of Chemical Engineers

emissions from biodiesel, including carbon monoxide, particulate emissions, and unburned hydrocarbon, are lower than those from the conventional diesel fuel (Van Gerpen, 2005) and therefore biodiesel could be an effective means to reduce air pollution in urban areas. In addition, because of its high lubricity, biodiesel could be added to ultra low sulphur diesel fuel, which is generally characterized by poor lubricity, in order to improve its fuel properties (Van Gerpen, 2005). Moreover, biodiesel does not contribute to global warming because of its closed carbon cycle: it was estimated in a life cycle analysis that the overall carbon dioxide emissions resulting from the use of biodiesel were reduced by 78% when compared to mineral diesel (Sheehan, et al., 1998). Apart from the technical view point, the use of fuel derived from natural oils and fats would also create a new market channel for vegetable oils and animal fats and hopefully it would help alleviate the oversupply problems in some areas of the world. Transesterication of vegetable oils with short chain alcohols such as methanol and ethanol in the presence of homogeneous basic catalysts such as sodium hydroxide (NaOH) and potassium hydroxide (KOH) has been widely used as an industrial biodiesel production process. The reaction comprises three consecutive reversible steps: the formation of diglycerides, the formation of monoglycerides and the formation of glycerol.
Vol 85 (B5) 383389

384
Triglycerides CH3 OH ! Diglycerides
k2 k1

GUNVACHAI et al.
UK). Methanol of 99% purity was purchased from Fisher Scientic Ltd (Loughborough, UK). Sodium hydroxide pellets of 99% purity were purchased from BDH Laboratory Supplies (Poole, UK). For the quantitative analysis of fatty acid methyl esters using a gas chromatograph, methyl pentadecanoate of more than 99.5% purity, methyl palmitate, methyl stearate, methyl oleate and methyl linoleate of at least 99% purity were purchased from Sigma-Aldrich Company Ltd (Gillingham, UK).

CH3 COOR1 Diglycerides CH3 OH CH3 COOR2 Monoglycerides CH3 OH CH3 COOR3 The overall reaction can be written as Triglycerides 3CH3 OH Glycerol
k7 k8 k5 k6 k3 k4

(1) ! Monoglycerides (2) ! Glycerol (3)

Experimental Procedures Generation of phase diagrams


Series of experiments were carried out to determine the solubility of rapeseed oil in methanol at different concentrations of FAMEs at temperatures of 293, 313 and 333 K. The FAMEs were added using a micro pipette to a stirred mixture of methanol and oil of known composition in a sample bottle until the onset of cloudiness was reached. This was taken as indicating incipient phase separation. The mixture was then allowed to stand for an hour and if the clouding disappeared additional FAMEs were added. The volume of FAMEs used and the composition of the resulting mixture were plotted to form a phase diagram. In order to validate the results generated from varying the oil to methanol concentration, a second set of experiments was carried out by varying the FAMEs/methanol concentration and gradually adding oil.

! 3CH3 COOR (4)

Because vegetable oils and alcohols are immiscible, the initial reaction system comprises two phasesan alcohol phase and an oil phase. Under such conditions mass transfer would govern the kinetics of the reaction because transesterication must occur in the methanol phase, in which the catalyst is dissolved. As a result, the initial rate of reaction is relatively slow and mechanical agitation is necessary to promote a more rapid reaction. Upon vigorous mixing, an emulsion is formed, aided by natural surfactants present in the oil, including free fatty acids, and lecithin (Garti, 1999; Novales et al., 2005). Once an emulsion is formed it appears that the rate of reaction is controlled by the solubility of the oil in the catalytically active phase. At high conversions of the oil, the reaction system forms a single phase and methyl esters, which act as a cosolvent, are formed. At this point simple homogeneous batch kinetics will govern the process, and the rate of reaction is rapid (Vincente et al., 2005; Boocock et al., 1996). As a result, the curves of the concentrations of the nal reaction products, i.e., methyl esters and glycerol, usually have a sigmoidal shape, which indicates a slow reaction rate at the early stage of reaction followed by a sudden increase in the reaction rate and nally a plateau as the reaction reaches its equilibrium. This behaviour is typical for the autocatalytic reactions or the reactions with changing mechanisms. While the kinetics of the reaction in the kinetically controlled region (fast region), where the system is a homogeneous solution, can be modelled by conventional methods, the early stage kinetics (slow region), which is controlled by the amount of vegetable oil dissolved in the methanol phase, must be modelled by using the inter-solubility data between vegetable oil, methanol and fatty acid methyl esters (FAMEs). Therefore in this study, by constructing the ternary phase diagrams of rapeseed oil, methanol, and FAMEs, we are able to propose a new semi-empirical solubility model to describe the early stage kinetics of the transesterication of commercial rape seed oil, using sodium hydroxide as a catalyst and a molar ratio of methanol to oil of 5 : 1. We also report here the overall rate constants of the reaction at the reaction temperatures of 293, 313 and 333 K.

Batch reactor experiment


Experiments were conducted in a stirred, jacket reactor of 1 l total capacity. Reaction temperatures were maintained at 293, 313 or 333 K. A molar ratio of methanol to rapeseed oil of 5 : 1 was employed and the catalyst used was sodium hydroxide at a concentration of 5 wt% based on methanol (approximately 1 wt% of oil). The stirrer speed was kept constant at 500 rpm for all experiments. Initial experiments indicated that increased stirrer speeds did not substantially affect the observed transesterication rates. To ensure that the contents of the reactor were at the desired temperature, the rapeseed oil (0.0005 m3) was introduced into the reactor rst and allowed to reach the desired reaction temperature. Then the solution of sodium hydroxide (0.002 kg) in methanol (0.0001 m3) was added via the addition port in the top of the reactor. Samples were taken after 2, 4, 6, 8, 10, 15, 20 and 25 min by opening the valve on the bottom of the reactor and allowing approximately 0.00003 m3 to ow out before retaining the sample. Each sample was placed in a sealed container and immediately placed on ice in order to stop the reaction. FAMEs were decanted from the lower layer of glycerol, diluted to 10 ppm with hexane and analysed according to the procedure described later.

Continuous ow minireactor experiment


The transesterication reactions at 333 K were repeated in a continuous ow minireactor (0.008 m ID 0.05 m, stainless steel 316), using the same molar ratio of methanol to oil and catalyst concentration as were reported above. The outlet of reactor was connected to a cooling loop which was immersed in an ice bath so that the reaction was quenched immediately once the reaction mixture exited the reactor. An inline lter with 0.5 micron pore size (Swagelok

MATERIALS AND EXPERIMENTAL PROCEDURES Materials


Rapeseed oil of food grade quality (rened and bleached) was purchased from Sainsburys Supermarket Ltd (London,

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B5): 383 389

SOLUBILITY MODEL TO DESCRIBE BIODIESEL


London, Hertfordshire, UK) was placed at the inlet of the reactor to cause formation of an emulsion. The minireactor was placed inside a Carbolite tube furnace (Carbolite Furnaces, Shefeld, UK) to control the reaction temperature (+1 K). Methanol was pumped to the reactor system using a dual-piston reciprocating HPLC pump (Waters Associates, Massachusetts, USA) Whereas methanol could be pumped with a standard chromatography pump, the high viscosity of the oil precluded the use of a reciprocating system (with check valves) and hence a continuous high pressure ISCO 100DM syringe pump was employed (ISCO Inc., Nebraska, USA). The ow rates of methanol and oil were adjusted such that the molar ratio of methanol to oil of 5 : 1 and the space times of 0.5 to 5 min were obtained. Samples collected from the reactor were kept overnight in a refrigerator to allow phase separation and then analysed using the procedure described later.

385

Figure 1. Mass of fatty acid methyl esters measured during rapeseed oil transesterication at 293, 313 and 333 K for molar ratio of methanol to oil of 5 : 1 and 5 wt% NaOH in a batch reactor.

Analytical Methods
To follow the change in the reaction product concentrations over time, the samples collected from the batch reactor and the continuous ow reactor were analysed based on gas chromatography technique which, for the types of the capillary columns selected, allowed the quantication of fatty acid methyl esters. Because different models of gas chromatographs and different types of detectors were used, two sets of standard calibration curves were constructed separately for the quantitative analysis purpose.

Batch reactor experiment


The samples were analysed using a HP 5890 gas chromatograph (Hewlett Packard, Ipswich, UK), equipped with a TRIO-1 mass spectrometer detector (Mooreld Associates, Manchester, UK), a HP 7673 auto-sampler and Mass Lab 1.3 software. The column used was a 30 m and 0.25 mm Supelco polar EC-Wax (Supelco, UK, Poole, UK). Helium was used as a carrier gas at a ow rate of 0.2 m s21. The injector temperature was maintained at 473 K. The oven temperature was initially held at 393 K for 1 min and then increased at a rate of 7 K min21 to 473 K whereupon it was held for 5 min.

Continuous ow minireactor experiment


The samples were analysed using a HP5890 gas chromatograph (Hewlett Packard, Ipswich, UK) equipped with a 0.25 mm 0.22 mm 25 m polar SGE BPX70 column (SGE Europe Ltd, Milton Keynes, UK), a ame ionization detector and JCL 6000 data system. Methyl pentadecanoate was used as an internal standard. The initial oven temperature was kept constant at 448 K for 2 min and then raised at a rate of 5 K min21 to 493 K. Helium was used as a carrier gas and the split ratio of 100 : 1 was used. The injector and detector temperatures were maintained at 523 K.

reaction to reach its highest conversion was decreased as temperature was increased. The increase in the equilibrium ester concentration at higher reaction temperature is due to the fact that the equilibrium constant shifts to higher values. It was observed that at 333 K the mass of FAMEs reached the maximum value of 0.165 1023 kg and then decreased after 8 min. This may be because small amounts of FAMEs dissolved in the glycerol phase, which formed at high conversion. Higher solubility of FAMEs in glycerol at elevated temperature is possibly because, at higher energy state, FAME molecules have a tendency to dissolve more in the glycerol phase. The lag time corresponding to the two-phase reaction mixture as reported by Noureddini and Zhu (1997) was not observed in this batch experiment. This is because the reaction proceeded very rapidly especially at 333 K, where the nal conversion was reached only 2 min (the rst data point) after the reaction was commenced. To study the initial reaction kinetics, it is necessary to follow the progress of the reaction at its early stages, when two phases exist and the reaction is promoted by the FAMEsaided dissolution of the oil in the methanol. However, the transesterication of rapeseed oil is so rapid that, when using a batch reactor, it prevented us from efciently perform the sampling and the quenching of the reaction. Therefore a tubular ow micro-reactor was constructed for carrying out the experiments. This type of reactor not only enabled us to obtain more data points in the solubility-controlled region, but it also allowed an immediate quenching of the reaction once the mixture exits the reactor. Figure 2 shows an initial slow rate of reaction and an overall sigmoidal pattern to the reaction time course. After leaving overnight in a

RESULTS AND DISCUSSION Transesterification Reaction Analysis


Figure 1 shows the mass of esters produced from a batch reactor as a function of time at the constant reaction temperatures of 293, 313 and 333 K. As expected, the time for the
Figure 2. Concentration of fatty acid methyl esters during rapeseed oil transesterication at 333 K for molar ratio of methanol to oil of 5 : 1 and 5 wt% NaOH in a continuous ow minireactor.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B5): 383 389

386

GUNVACHAI et al.

refrigerator for settling, the reaction mixtures in this region (0.5 and 1.0 min residence time) separated into three phases: a top layer of methanol, a middle layer of oil and a small droplet of glycerol at the bottom. This lag is the time needed for the concentration of the esters in the reaction system to increase to a point at which a single phase is formed. When the residence time was increased, a twophase system of biodiesel and glycerol was observed which corresponds to the sudden increase in the ester concentration in Figure 2. As previously observed in the batch experiment, the ester concentration also decreased after it reached its maximum value.

Phase Diagrams
In order to fully follow the transesterication reaction path within the reactor, the intersolubility data of the major reaction components, i.e., rapeseed oil, methanol, FAME and glycerol needed to be identied. For clearer depiction, a tetrahedron is used to represent the reaction system (Figure 3). Each corner is marked with each of the major constituents of the reaction mixture. The planes of the tetrahedron that frame the composition space of the reacting mixture during the reaction are the ternary phase diagrams of oilmethanol FAME, oilmethanol glycerol, oil FAME glycerol and methanol glycerol FAME and by constructing the solubility diagrams (phase diagrams), the composition of the different phases could be extracted and used in the solubility models. Figure 4 shows the phase diagram of oil methanol FAME obtained from the procedure described earlier. It is evident that although both rapeseed oil and methanol are completely soluble in FAME, they are almost immiscible at all temperatures studied because of the difference in polarity. However, their intersolubility gradually increases with increasing concentration of FAME and when the concentration of FAME in the mixture reaches approximately 55 vol%, the two-phase mixture becomes homogeneous. As expected, the intersolubility between rapeseed oil and methanol increases with increasing temperature. The intersolubilities between FAME methanol glycerol, oilglycerol MeOH and oilglycerol FAME were not determined in this study because the purpose is to investigate the kinetics of the transesterication reaction at the initial stage where the concentrations of the reaction products, which

Figure 4. The phase diagram of oilmethanol FAME.

are produced in the methanol phase, are relatively low. Therefore it is reasonable to assume that all the products stay in the methanol phase without diffusing to the oil phase.

Solubility Models for a Batch Reactor


The basis of the new model is that the solubility of the oil in the methanol phase increases as the reaction progresses and FAMEs are produced leading at a later stage to a single phase system (at conversions in excess of 60%). At the beginning of the reaction the species present are methanol and rapeseed oil. In order to determine, which phase is dispersed and which is continuous, a correlation factor (X), proposed by Selker and Sleicher (1965), was calculated based on phase volume ratios, density and viscosity of each phase. Table 1 shows the criteria for determining types of dispersion by using the correlation factor (X). x   QL rL mL 0:3 QH rH mH (5)

A correlation factor of approximately 0.06 was obtained, indicating that in our case methanol is dispersed in a continuous phase of rapeseed oil. A number of initial assumptions were made to simplify the model, including: (1) None of the species is soluble in the oil phase at low concentration. This simplies the model considerably as it means that when a tie line is plotted on the diagram between the two phases, the oil phase end remains at 100% oil (the left corner of the diagram). This enabled us to modify the phase diagram of Figure 4 to the one shown in Figure 5. (2) None of the species are soluble in the glycerol phase, however it is known that methanol is soluble in the

Table 1. Relationship between dispersion types and values of the correlation factor. Correlation factor X ,0.3 0.30.5 0.52.0 2.03.3 .3.3 Figure 3. The interaction of all phases within the reaction system. Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B5): 383 389 Results Light phase always dispersed Light phase probably dispersed Phase inversion possible Heavy phase probably dispersed Heavy phase always dispersed

SOLUBILITY MODEL TO DESCRIBE BIODIESEL


Higher order reaction kt NTG=Me0 x
0

387

dXTG n m VCTG=Me CMe=Me

(8)

Figure 5. A solubility phase diagram used for the generation of the reaction kinetics.

(3)

(4)

(5) (6) (7)

(8)

glycerol phase, but in the early stages of the reaction when a two-phase emulsion exists, the concentration of glycerol will be low. The dissolution of methanol in glycerol could be built into an extended version of the model. The solubility of oil in the methanol phase is a function of conversion. The production of FAME acts as a solubilizing agent. A small amount of oil is always present in the methanol phase even at the beginning of the reaction. This is in agreement with the phase diagrams in Figures 4 and 5. The edge point of the triangular diagram in Figure 5 has been forced to be zero. The actual solubility can be inferred by extrapolating the low concentrations of oil in methanol. The reaction only occurs in the methanol phase, where the catalyst is dissolved. Rates are considered with respect to the volume of methanol only. The reaction intermediates, i.e., diglycerides and monoglycerides are neglectedtherefore the three reaction steps [equations (13)] can be simplied to equation (4). This assumption is based on the fact that only FAME and glycerol are the nal products for the overall reaction. As a result, the rate constants calculated in this study are for the overall reaction. Reverse reactions are negligible because methanol is used in excess in order to drive the reaction forward and the reaction products formed at the early stage are at fairly low concentrations.

The integrals on the right hand side of the equation were calculated from the batch experimental data and plotted against the reaction time. It was found that a reaction order of 1.5 with respect to methanol and 1 with respect to oil gave the best t to the experimental data (Figures 6 8). The reaction rate constants (k) were then calculated from the slopes of the plots and the initial concentrations of oil in the methanol phase obtained from the extrapolation of the phase diagram. The k values of 0.00098, 0.00309 and 0.0067 (kmol m23)21.5 s21 were obtained at reaction temperatures of 293, 313 and 333 K, respectively. The 2.5 order rate constant obtained in this study at 333 K was converted to the second order rate constant and compared with the values for the second-order reaction (pseudo homogeneous reaction) reported in the literature (Table 2). The rate constant obtained in this study was found to be within the range of the gures reported in the literature. The discrepancy is probably because the reaction conditions used in each study are even though similar but not identical and therefore the rate constants obtained are different. The explanation for the relatively low rate constants obtained from this batch experiments is that these values could be the rate constants at near equilibrium conditions as the experimental results shows that the reaction was very fast

Figure 6. A 2.5 order (rst order with respect to oil and 1.5 order with respect to methanol) solubility model plot of rapeseed oil at 293 K.

By taking into account the above assumptions, a solubility model for a pseudo rst order reaction with respect to methanol for a batch reactor can be written as kt NTG=Me0 x dXTG VCTG=Me 0 (6)

For second order and higher order reactions, the model equations can be written in a similar form as Second order reaction kt NTG=Me0 x dXTG VCTG=Me CMe=Me 0 (7)
Figure 7. A 2.5 order (rst order with respect to oil and 1.5 order with respect to methanol) solubility model plot of rapeseed oil at 313 K.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B5): 383 389

388

GUNVACHAI et al.

Figure 9. A segregated model for a tubular reactor.

Figure 8. A 2.5 order (rst order with respect to oil and 1.5 order with respect to methanol) solubility model plot of rapeseed oil at 333 K.

of small intervals (Dt). Then the reaction in each time interval can be represented by a batch reactor (Figure 10). For a pseudo rst order reaction (the methanol in the droplets is in great excess), the solubility model can now be written as 1 dNTG=Me KCTG=Me V dt (9)

Table 2. Rate constants of the second order transesterication reaction reported in the literature and in this study. Reported by Noureddini and Zhu (1997) Darnoko and Cheryan (2000) Cheng et al. (2004) Vincente et al. (2005) This study (batch reactor) Rate constant1, m3 kmol21 min21 0.0502 3.382 (originally reported as 0.036 wt%21 min21) 0.097 0.612 0.0673

If we further assume that the volume of the droplet is constant, the equation can then be rearranged and integrated to ln CTG=Me0 kt CTG=Me (10)

Notes: 1 The reaction conditions are similar to those employed here but not identical. 2 The numbers reported are the lowest forward rate constants. 3 The rate constant was converted from 2.5 order to the second order by multiplying by the square root of the initial concentration of oil in the methanol phase.

and it seemed to be complete within the rst 2 min of the reaction course.

A Solubility Model for a Tubular Flow Reactor


To develop a solubility model for a tubular ow reactor, a segregated model was applied (Fogler, 1992). The initial assumptions are: (1) When rapeseed oil and methanol enter the reactor, tiny droplets of methanol with small amounts of dissolved oil are formed and these droplets travel through the reactor without coalescing together until they exit the reactor. (2) Both reactants inside the droplets are perfectly mixed. (3) There is no interchange of the reaction products, i.e., FAMEs and glycerol, between the droplets and the continuous phase (oil phase). As the FAMEs are formed, they promote the solubility of oil in methanol so that the oil keeps dissolving in the methanol phase as time progresses. Therefore each droplet behaves like a tiny semibatch reactor with a continuous feed of oil. (4) Each droplet spends equal time in the reactor (ideal tubular reactor) leading to the same residence time (or batch time for that matter). (5) Assumptions 5 8 from the batch reactor solubility model are also adopted. The model can be further simplied by dividing the total time that each droplet spends in the reactor into a number

The rate constant (k) was calculated using a spreadsheet technique (Microsoft Excel). A time interval (Dt) of 10 s was used in the calculation. The total time studied was 60 s as it was observed in the experiments that oil and methanol were still in separate phase during this period. The initial concentration of oil in the methanol phase (CTG/Me0) was determined from the ternary phase diagram and the concentration of FAMEs at the beginning of each time interval (or at the end of the previous time interval). To extrapolate the concentration of oil at a very low concentration of FAMEs, the ternary phase diagram was converted to a 2-D plot by plotting the ratios of the concentrations of oil to that of methanol versus the concentrations of FAMEs. Then CurveExpert1.3 was used to t curves to those data. Several mathematical models were found to give a good t to the experimental data but the most suitable model was selected according to two criteria. The rst criterion is that the model must allow small amounts of oil to present in the system even in the absence of FAMEs which is in accordance with assumption 1. Another criterion is that the curve should be shallow at the beginning and then followed by a steep region which corresponds to the lag time and the sudden surge reported in the literature as well as observed in this study. This means that, at the beginning of the reaction, an increase in FAME concentration has only small effect on the solubility of oil in methanol but when the FAMEs build up to a certain concentration, they greatly enhance the

Figure 10. Segregated model incorporating molar ux based on solubility envelope.

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B5): 383 389

SOLUBILITY MODEL TO DESCRIBE BIODIESEL


Table 3. Rate constants of the rst order transesterication reaction reported in the literature and in this study. Reported by Freedman (1986) This study (batch reactor) This study (continuous ow reactor) Rate constant, min21 0.7922.3731,2 0.018 1.78

389

Notes: 1 The reaction conditions are similar to those employed here but not identical. 2 The numbers reported are the lowest forward rate constants.

rH rL m and n NTG/Me NTG/Me0 QH QL mH mL t V X XTG

density of the heavy component density of the light component orders of the reaction mole of oil in methanol initial mole of oil in methanol volume of the heavy component volume of the light component viscosity of the heavy component viscosity of the light component reaction time volume of the reactor correlation factor conversion of oil

intersolubility of the two components. It was found that Gompertz function satised both criteria and gave a good t to the experimental data. The boundary concentrations of FAMEs at t 0.5 and 1 min, which were known from the experiments, were applied and the k value of 1.78 min21 was nally obtained from the spreadsheet calculation. As can be seen in Table 3, this value is within the same range of the k values for the pseudo homogeneous rst order reaction reported by Freedman et al., (1986). Even though the 2.5 reaction order was found to give the best t to the batch experimental data, an attempt was made to obtain the rst order rate constant for comparison purpose. It was found that, at the same reaction temperature, the k value from the continuous ow reactor is larger than that from the batch experiment by approximately two orders of magnitude. This may be explained based on the hypothesis that the kinetics of the transestercation reaction comprises three regions: the solubility controlled region (slow), the kinetically controlled region (fast) and the slow region as equilibrium is approached. Even though we aimed to calculate the k value of the rst region, by using a batch reactor, we could not collect the data in this region because the reaction was extremely fast. We therefore used the data in rst 4 min to calculate the rate constant. As a result, the value obtained in the batch experiment is actually an average of the rate constants of the three regions.

REFERENCES
Boocock, D.G.B., Konar, S.K., Mao, V. and Sidi, H., 1996, Fast one-phase oil-rich processes for the preparation of vegetable oil methyl esters, Biomass and Bioenergy, 11(1): 43 50. Boocock, D.G.B., Konar, S.K., Mao, V., Lee, C. and Buligan, S., 1998, Fast formation of high-purity methyl esters from vegetable oils, Journal of the American Oil Chemists Society, 75: 1167 11172. Canakci, M. and Van Gerpen, J.H., 2003, Comparison of engine performance and emissions for petroleum diesel fuel, yellow grease biodiesel, and soybean oil biodiesel, Transactions of the ASAE, 46: 937944. Cheng, S.-F., Choo, Y.-M., Ma, A.-N. and Chuah, C.-H., 2004, Kinetics study on transesterication of palm oil, Journal of Oil Palm Research, 16: 1929. Darnoko, D. and Cheryan, M., 2000, Kinetics of palm oil transesterication in a batch reactor, Journal of the American Oil Chemists Society, 77: 12631267. Fogler, S.C., 1992, Segregation model, in Amundson, N.R. (ed.). Elements of Chemical Reaction Engineering, 38 744 (PrenticeHall, New Jersey, USA). Freedman, B., Buttereld, R.O. and Pryde, E.H., 1986, Transesterication kinetics of soybean oil, Journal of the American Oil Chemists Society, 63: 13751380. Garti, N., 1999, What can nature offer from an emulsier point of view: trends and progress?, Colloids and Surfaces A: Physicochemical and Engineering Aspects, 152: 125146. Goering, C.E., Schwab, A.W., Daugherty, M.J., Pryde, E.H. and Heakin, A.J., 1981, Fuel properties of eleven vegetable oils, ASAE 1981, St Joseph, MI, Paper Number 81-3579. Ma, F. and Hanna, A., 1999, Biodiesel production: a review, Bioresource Technology, 70: 1 15. Meher, L.C., Sagar, D.V. and Naik, S.N., 2006, Technical aspects of biodiesel production by transestericationa review, Renewable and Sustainable Energy Reviews, 10: 248 268. Noureddini, H. and Zhu, D., 1997, Kinetics of transesterication of soybean oil, Journal of the American Oil Chemists Society, 74: 1457 1463. Novales, B., Ropers, M.H. and Douliez, J.-P., 2005, Use of fatty acid/ monoglyceride vesicle dispersions for stabilizing O/W emulsions, Colloids and Surfaces A: Physicochemical and Engineering Aspects, 269: 8086. Selker, A.H. and Sleicher, C.A., 1965, Factor affecting which phase will disperse when immiscible liquids are stirred together, Canadian Journal of Chemical Engineering, 43: 298 301. Sheehan, J., Camobreco, J.D., Graboski, M. and Shapouri, H., 1998, Life cycle inventory of biodiesel and petroleum diesel for use in an urban bus, Final report for U.S. Dept. of Energys Ofce of Fuel Development and the U.S. Dept. of Agricultures Ofce of Energy, by the National Renewable Energy Laboratory, NERL/ SR-580-24089. Srivastava, A. and Prasad, R., 2000, Triglycerides-based diesel fuels, Renewable and Sustainable Energy Reviews, 4: 111 133. Van Gerpen, J., 2005, Biodiesel processing and production, Fuel Processing Technology, 86: 10971107. Vincente, G., Martinez, M., Aracil, J. and Esteban, A., 2005, Kinetics of sunower oil methanolysis, Ind Eng Chem Res, 44: 54475454. The manuscript was received 6 March 2007 and accepted for publication after revision 31 March 2007.

CONCLUSION
It is quite clear form the above discussion that the use of batch reactors to obtain kinetic data for fast transesterication reactions is awed due to the timescale of sampling and possibly spatial variation of concentration. Continuous mini rectors give tighter control and allow for fast intrinsic kinetics to be determined. The new solubility model described above offers a good representation of the kinetics of vegetable oil transesterication at moderate temperatures. The model could be developed further, for example, by considering the reaction in the oil phase and the existence of the glycerol phase, where methanol and catalyst are removed from the reaction. In addition, soap formation and mass transfer effects may also be taken into account. Although additional work needs to be done to better model the complexities of phase effects on reaction kinetics, our model goes some way towards providing a better description of reaction kinetics than was hitherto possible.

NOMENCLATURE
CMe/Me CTG/Me k concentration of methanol in the methanol phase concentration of oil in the methanol phase reaction rate constant

Trans IChemE, Part B, Process Safety and Environmental Protection, 2007, 85(B5): 383 389

You might also like