Sunteți pe pagina 1din 9

Bioresource Technology 98 (2007) 639647

Biodiesel production using a membrane reactor


M.A. Dube *, A.Y. Tremblay, J. Liu
Department of Chemical Engineering, University of Ottawa, Ottawa, Ont., Canada K1N 6N5 Received 31 May 2005; received in revised form 5 January 2006; accepted 3 February 2006 Available online 31 March 2006

Abstract The immiscibility of canola oil in methanol provides a mass-transfer challenge in the early stages of the transesterication of canola oil in the production of fatty acid methyl esters (FAME or biodiesel). To overcome or rather, exploit this situation, a two-phase membrane reactor was developed to produce FAME from canola oil and methanol. The transesterication of canola oil was performed via both acid- or base-catalysis. Runs were performed in the membrane reactor in semi-batch mode at 60, 65 and 70 C and at dierent catalyst concentrations and feed ow rates. Increases in temperature, catalyst concentration and feedstock (methanol/oil) ow rate signicantly increased the conversion of oil to biodiesel. The novel reactor enabled the separation of reaction products (FAME/glycerol in methanol) from the original canola oil feed. The two-phase membrane reactor was particularly useful in removing unreacted canola oil from the FAME product yielding high purity biodiesel and shifting the reaction equilibrium to the product side. 2006 Elsevier Ltd. All rights reserved.
Keywords: Biodiesel; Methanol; Acid-catalyzed transesterication; Base-catalyzed transesterication; Two-phase membrane reactor

1. Introduction Biodiesel is a clean-burning diesel fuel produced from vegetable oils, animal fats, or grease. Its chemical structure is that of fatty acid alkyl esters (FAAE). Commercially, biodiesel is produced by transesterication. This reaction consists of transforming triglyceride (TG) into FAAE, in the presence of an alcohol (e.g. methanol, ethanol) and a catalyst (e.g. alkali, acid, enzyme) with glycerol as a major by-product. The reaction scheme is shown in Fig. 1. In Fig. 1, X represents the alkyl group of the alcohol (e.g. CH3 for methanol) while R is a carbon chain typically on the order of 1120 carbon atoms long. The conversion of TG to FAAE comprises three consecutive reversible reactions with diglyceride (DG) and monoglyceride (MG) as intermediate products. Biodiesel is being used increasingly in public transportation in Europe, Japan and North America.
Corresponding author. Tel.: +1 613 562 5800x6108; fax: +1 613 562 5172. E-mail address: Marc.Dube@uottawa.ca (M.A. Dube). 0960-8524/$ - see front matter 2006 Elsevier Ltd. All rights reserved. doi:10.1016/j.biortech.2006.02.019
*

Due to diminishing petroleum reserves and the environmental consequences of exhaust gases from petroleumderived fuels, such as gasoline and diesel, biodiesel has attracted attention during the past decade as a renewable and environmentally friendly fuel. Because biodiesel is made entirely from vegetable oil or animal fats, it is renewable, environmentally benign (biodegradable), and does not contain any sulphur, aromatic hydrocarbons, metals or crude oil residues. Like petroleum diesel, biodiesel operates in compressionignition engines such as those used in farm equipment, and private and commercial vehicles. Essentially no engine modications are required, and biodiesel maintains the payload capacity and range of diesel. Because biodiesel is oxygenated, it is a better lubricant than diesel fuel, increasing the life of engines, and is combusted more completely. Indeed, many countries are introducing biodiesel blends to replace the lubricating eect of sulfur compounds in low-sulfur diesel fuels (Anastopoulos et al., 2001; Dmytryshyn et al., 2004). The higher ash point of biodiesel makes it a safer fuel to use, handle and store. With its relatively low emission prole, it is an ideal fuel for use

640

M.A. Dube et al. / Bioresource Technology 98 (2007) 639647

H 2C-O-R Catalyst HC-O-R H 2C-O-R + 3X-OH

H2C-O-H HC-O-H H 2C-O-H + 3RCOOX

Fig. 1. Stoichiometric reaction for conversion of triglyceride to fatty acid alkyl esters.

in sensitive environments, such as marine areas, national parks and forests, and heavily polluted cities. Existing technology to produce biodiesel continues to require government subsidies in order to be protable. One reason is the high cost of virgin vegetable oil, the source of TG (Coltrain, 2002). In order to reduce production costs and make it competitive with petroleum diesel, low cost feedstock, such as non-edible oils, waste frying oils and animal fats, could be used as raw materials (Zhang et al., 2003a). However, the relatively higher amounts of free fatty acids (FFA) and water in this feedstock results in the production of soap in the presence of alkali catalyst. Thus, additional steps to remove any water and either the FFA or soap from the reaction mixture are required. In fact, commercial base-catalyzed processes often employ an acid-catalyzed pre-esterication reactor to remove excess FFAs. Alkali-catalyzed transesterication proceeds much faster than that catalyzed by an acid and is the one most used commercially (Freedman et al., 1984). Considerable research has been done on biodiesel made from virgin vegetable oils (e.g. soybean oil, sunower oil, rapeseed oil) using alkali catalysts. The majority of biodiesel today is produced by alkali-catalyzed (e.g. NaOH, KOH) transesterication with methanol, which results in a relatively short reaction time. However, as mentioned above, the vegetable oil and alcohol must be substantially anhydrous and have a low FFA content because the presence of FFA promotes soap formation. The soap formed partially consumes the catalyst, lowers the yield of esters and renders the downstream separation of the products dicult (Freedman et al., 1984). Freedman et al. (1985) reported a 98% yield of biodiesel in 1 h using alkali catalysts such as sodium hydroxide or sodium methoxide with alcohols such as methanol, ethanol, and 1-butanol. Freedman et al. (1984) investigated the eect of the molar ratio of the alcohol to oil, type of catalyst (base vs. acid), temperature and degree of renement of the oil on the yield of biodiesel. For the alkali-catalyzed reaction, the eect of alcohol to oil ratio was found to be the most important variable aecting the yield, while temperature had a signicant eect on the initial reaction rate. It also was mentioned in their study that acid catalysts would be more eective when the degree of renement of oil was low, and for oils that had a high FFA content. Several studies have been done on the production of biodiesel from waste oils or animal fats (Nye et al., 1983; Watanabe et al., 2001) describing the feasibility of making

quality biodiesel from this feedstock while identifying the problems with the FFAs present in the raw materials. Experiments in our laboratory (McBride, 1999; Ripmeester, 1998; Zheng, 2003) showed that greater than 99.5% conversion of waste frying oil to biodiesel could be achieved within a reasonable time period ($3 h at 70 C) through transesterication with an excess of methanol (methanol:oil molar ratio was >50:1) catalyzed by sulphuric acid. Despite the slower reaction rate, this approach has several advantages over the base-catalyzed method (Canakci and Van Gerpen, 1999; Zhang et al., 2003a,b): it employs a one-step process as opposed to a two-step process; it can handle feedstock with a high FFA content; downstream separation of the biodiesel is straightforward; and a high quality glycerol by-product is produced. Aside from the slow reaction rate, another drawback of the acid-catalyzed process is the requirement for the reactor to withstand an acidic environment. Nonetheless, our economic assessment carried out on four dierent continuous processes with dierent types of oil (virgin vs. waste) and catalysts (acid vs. alkali) showed that although the alkalicatalyzed process using virgin oil had the lowest capital investment cost, the cost of using virgin oil led to a higher total manufacturing cost (Zhang et al., 2003b). When waste frying oil was used in the alkali-catalyzed process, a pretreatment unit was required to reduce the content of the FFA. Thus, the cost associated with the pre-treatment unit oset the cost savings due to the use of waste frying oil. In contrast, the acid-catalyzed process using waste frying oil did not require a pre-treatment step and had the lowest total manufacturing cost, and the highest protability. Yet another drawback to the acid-catalyzed process, is that high alcohol to oil ratios are necessary to promote the conversion of oil to FAAE (Freedman et al., 1985). These higher amounts of alcohol increase the reactor size but recycling can mitigate some of the associated increases in cost. Nonetheless, the issue of separating these substantial amounts of alcohol from the FAAE becomes important. Another issue that plagues biodiesel production is the removal of residual TG and glycerol from the biodiesel product. One approach is to drive the reaction as close to complete conversion of the TG as possible. However, the transesterication of TG is an equilibrium reaction, and there are thus, limits to this approach. Other approaches employ multiple water washing steps of the product stream, which can give rise to a challenging waste treatment problem in the wastewater stream. Karaosmanoglu et al. (1996) studied three dierent separation methods and found that hot water washing at 50 C was the best way to separate and purify the biodiesel product. Bam et al. (1995) suggested that glycerine could be added as a solvent to wash impurities. Hayafuji et al. (1999) studied the use of a solid absorbent, such as activated clay, activated carbon, activated carbon bre, etc. to purify the resultant esters. Miscibility is an important factor in biodiesel production. The conventional transesterication method results

M.A. Dube et al. / Bioresource Technology 98 (2007) 639647

641

in a two-phase reaction which is, as a result, mass-transfer limited. More specically, the vegetable oils and methanol are not miscible. The approach of many existing commercial enterprises has been largely focused on steps to enhance reaction rate by attempting to overcome this immiscibility. For example, the addition of a co-solvent to generate a homogeneous reaction mixture can greatly enhance the reaction rate (Boocock et al., 1996, 1998). While this enhances reaction rate signicantly, the co-solvent must eventually be separated from the biodiesel and this requires additional processing. Considering that reaction rate may not necessarily restrict process protability (Zhang et al., 2003b), transesterication is an equilibrium reaction and downstream processing of the biodiesel is of utmost concern, another approach can be envisaged. This approach involves the use of membrane separation technologies, which exploit the immiscibility of the oil and methanol. In this paper, we report on the production of biodiesel using a semi-batch two-phase carbon membrane reactor using either an acid- or base-catalyzed transesterication of canola oil. Carbon membranes are of particular interest as they can be stable at high temperatures and resist chemical attack. 2. Background Membrane based separations are well-established technologies in water purication, protein separations and gas separations. To date, commercial applications of membrane technologies are limited to separations involving aqueous solutions and relatively inert gases. The use of membranes to treat non-aqueous uids is an emerging area in membrane technologies (Kim et al., 2002; White and Nitsch, 2000). Synthetic membranes also have been used in reactor systems to remove the products of reactions as they are being formed. Product-inhibited and equilibrium reaction systems are very common in commercially important reactions. Removing reaction products as they are formed drives reactions to very high conversions that cannot be reached in conventional reactors. To date, much eort has been spent on membrane reactors involving easily separated mixtures such as hydrogen and methane (Armor, 1998; Hsieh, 1991, 1996; Saracco et al., 1999; Saracco and Specchia, 1994; Zaman and Chakma, 1994). The membrane can be either organic in nature (i.e., polymeric) or inorganic. Inorganic membranes are more suitable for use with organic solvents and, due to their excellent thermal stability, they can be used at high reaction temperatures (Saracco et al., 1994, 1999). Membrane reactors can have several advantages over conventional reactors (Armor, 1989): (a) an integration of reaction and separation into a single process, thus reducing separation costs and recycle requirements,

(b) an enhancement of thermodynamically limited or product-inhibited reactions resulting in higher conversions per pass, (c) a controlled contact of incompatible reactants and (d) an elimination of undesired side reactions.

3. Experimental 3.1. Materials Methanol (99.85% Reag. Grade containing <0.1% water) was supplied by (Commercial Alcohols Inc., Brampton, ON, Canada) and the canola oil by (No Name, Toronto, ON, Canada, purchased at a local foodstore). Fatty acid methyl esters (FAME or biodiesel) used for miscibility testing was produced from the acid-catalyzed transesterication of waste oils from a previous study (Zheng, 2003). Sulfuric acid (9598%, Reagent Grade) and tetrahydrofuran (99.95%, Chromatography Grade) were supplied by (EMD Chemicals Inc., Gibbstown, NJ, USA). 3.2. Membrane reactor design and experimental design A 300 mL membrane reactor system was constructed and is shown schematically in Fig. 2. A carbon membrane (Koch Membrane Systems, Inc., Wilmington, DE, USA) was used in the reactor. The pore size of the membrane was 0.05 lm. The inside and outside diameters of the membrane were 6 and 8 mm, respectively. The length of carbon membrane tube was 1200 mm giving a surface area of 0.022 m2 for the entire membrane. A controlled volume pump (Milton Roy Company, Ivyland, PA, USA) was used to feed the oil and methanol/catalyst mixtures to the system while a seal-less centrifugal canned motor pump (Labcor Inc. Concord, ON) was used to circulate the mixture at a speed of 15.2 mL/min. A heat exchanger (Neslab Instruments, Inc., Portsmouth, NH, USA) coupled with LabViewTM software was used to control the reaction temperature. Experiments were carried out at 60, 65 and 70 C in a 300 mL membrane reactor for 6 h, 0.5, 2, 4 and 6 wt.% concentrations of sulfuric acid catalyst were investigated (see Table 1). Canola oil (100 g) was used in each run. Pressure was controlled at 138 kPa between the permeation side and reaction side of the membrane. All experiments and sample analyses were carried out in random order to minimize any potential experimental errors. Several replicate runs also were performed (see Table 1). Additional experiments were conducted to verify the eect of methanol feed ow rate and the use of a base catalyst. 3.3. Membrane reactor experiments procedure The methanol and sulfuric acid were pre-mixed and charged into the membrane reaction system prior to each reaction. Canola oil (100 g) was charged into the membrane reactor, the membrane reactor was sealed and the

642

M.A. Dube et al. / Bioresource Technology 98 (2007) 639647

Fig. 2. Schematic diagram of separative membrane reactor.

Table 1 Experimental conditions Temperature (C) 60 65 70 60 65 70 60 65 70 60 65 70 Catalyst concentration (wt.%) 0.5 0.5 0.5 2 2 2 4 4 4 6 6 6 # of replicates 2 2 2 3 3 2 2 2 2 4 4 4

circulation pump was started. After a 10 min circulation time, methanol and acid catalyst were charged continuously into the membrane reactor with the feed pump at a ow rate of 6.1 mL/min. The heat exchanger was switched on to achieve the reaction temperature (60, 65 and 70 C). A thermocouple was used to monitor the reaction temperature. A stable reaction temperature (0.1 C) was achieved within 30 min for 60 C, 40 min for 65 C and 45 min for 70 C of starting the heat exchanger. Pressure was controlled at 138 kPa. The permeate product was collected in a 2000 mL ask. All experiments were conducted for 6 h. Additional experiments were conducted to observe the eect of methanol/acid catalyst feed ow rate on conver-

sion for both acid- and base-catalyzed transesterications. These ow rates were 2.5, 3.2 and 6.1 mL/min. The permeate product collected during the entire experiment time was mixed with an equivalent volume of reverse osmosis water (produced from tap water) and shaken by hand for about 5 min. This step served to stop any further reaction in the samples by promoting a phase separation of the glycerol phase containing most of the catalyst from the FAME phase. The mixture was allowed to settle for 24 h and ltered using a 0.5 lm membrane lter (Nalge Company, New York, NY, USA). The upper layer of the resulting two-phase mixture was transferred to a separatory funnel and washed with 1 L of reverse osmosis water. The resulting mixture was allowed to settle for 24 h, after which the upper layer was analyzed using high performance liquid chromatography (HPLC) according to the method used by Dube et al. (2004). Any unreacted oil in the retentate stream was also analyzed by HPLC. The retentate solution was neutralized by sodium hydroxide solution before analysis by HPLC. The HPLC analysis revealed that the purication method was eective and no residual acid was found in the samples. 3.4. High performance liquid chromatography (HPLC) analysis A Waters Corp. HPLC system consisting of an HPLC pump, a controller, a dierential refractometer and autosampler was used to analyze the contents of the permeate

M.A. Dube et al. / Bioresource Technology 98 (2007) 639647

643

and retentate streams. Waters Millennium 32TM software (Waters) was utilized for analysis. The columns used were two 300 7.5 mm Phenogel columns of 3 lm particles and 100 A pore size (Phenomenex, Torrance, CA, USA) connected in series. The mobile phase was tetrahydrofuran (THF) at a ow rate of 0.5 mL/min at 23 C. THF was used to make a 20 mg/g solution of the sample. Two grams of the solution was injected into the autosampler vials. Prior to analysis, the solutions were ltered through a 0.5 lm polytetrauoroethylene (PTFE) syringe lter. The HPLC analysis was conducted according to the method shown by Dube et al. (2004) and Darnoko et al. (2000). A calibration curve was generated from ve standards: triolein (TG), diolein (DG), monoolein (MG), methyl oleate (FAME), glycerol. The injection masses were plotted against the peak area. Each standard was injected three times at ve dierent concentrations. The calibration curves of the standard solutions showed good linearity. The retention times of the standards are shown in Table 2. Fig. 3 shows a typical chromatogram of a mixture of standards (note: sample concentrations were 0.548 mg/mL TG, 0.654 mg/mL DG, 0.602 mg/mL MG, 0.642 mg/mL FAME and 0.584 mg/mL glycerol (injection volume was 2 lL)).

The fractional conversion of oil to FAME, based on the amount of oil remaining in the reactor, was taken to represent the actual conversion. The oil to FAME conversion at time t was calculated from X M oilt0 M oiltt M oilt0 1

where X was the fractional conversion, Moil(t = 0) was the original of mass of oil (or TG equivalents in order to account for the presence of any DG or MG) in the reactor. Moil(t = t) was the mass of TG left in the reactor after the 6 h reaction time. 4. Results and discussion 4.1. Principle of membrane reactor operation It is known that methanol is only slightly miscible in canola oil and that temperature has only a slight eect on this miscibility (Liu, 2004). For all practical purposes, one could say that the two phases are immiscible. On the other hand, FAME is entirely miscible in methanol over a broad range of temperatures. At normal reaction temperatures (e.g. 60 C), FAME and methanol are miscible. The immiscibility of oil and methanol is at the root of the mass transfer issues for biodiesel production. However, in the operation of a membrane reactor, the formation of a two-phase (emulsied) system is exactly what is necessary. Thus, the transesterication of TG to FAME is ideally suited to operation in a membrane reactor. The principle of membrane reactor operation is depicted in Fig. 4. Due to the immiscibility of canola oil and methanol, and due to various surface forces, the canola oil will exist in the form of an emulsion; i.e. droplets suspended in methanol. One can thus envisage the transesterication occurring at the surface of the canola oil droplets. In the

Table 2 Retention time of standards Standard Triolein (TG) Diolein (DG) Monoolein (MG) Methyl oleate (FAME) Glycerol Retention time (min) 24.57 25.45 27.12 28.68 30.95 Relative retention time 1 1.04 1.10 1.17 1.26

80

Diolein
70

Monoolein

Triolein
60

Methyl oleate Glycerol

50

mV

40

30

20

10

0 22 23 24 25 26 27 28 29 30 31 32 33

Elution time (min)

Fig. 3. HPLC chromatogram of mixture of standards.

644

M.A. Dube et al. / Bioresource Technology 98 (2007) 639647

was used for the determination of compounds in both the permeate and retentate. A typical HPLC chromatogram of the retentate is shown in Fig. 5. It is seen that the main component in the retentate is TG (retention time = 25 min) or canola oil. Trace amounts of DG (retention time = 26 min) and FAME (retention time = $29 min) also are evident. Fig. 6 shows a typical chromatogram of the permeate. The complete absence of a peak at 25 min indicates that very high purity FAME was produced by the membrane reactor. At the reaction conditions in this study, as mentioned previously, methanol is only slightly miscible in canola oil. At the same time, FAME and methanol are miscible. These physical characteristics are what permit the membrane reactor to separate the FAME from the oil. 4.2. Eect of temperature
Fig. 4. Separation of oil and FAME by a separative membrane.

presence of a permeable membrane, the oil droplets are too large to pass through the pores of the membrane. On the other hand, the FAME is soluble in the methanol and, due to its small molecular size, will pass through the membrane pores along with the methanol, glycerol and catalyst. The main reaction for producing FAME was shown earlier in Fig. 1. The reaction is reversible, although the equilibrium lies towards the production of FAME and glycerol. In order to increase the production of FAME, FAME or glycerol should be removed during the reaction in order to shift the equilibrium to the product side. A microporous carbon membrane reactor can selectively permeate FAME, methanol and glycerol during the transesterication from the reaction zone. The molecule of canola oil is trapped in droplets forming an emulsion. The droplets cannot pass through the pores of the membrane because they are larger than the pore size of the carbon membrane. Results showed that during the reaction, canola oil did not appear in the permeate side. HPLC
1500 1400 1300 1200 1100 1000 900 800 700 600 500 400 300 200 100 0 22 23 24 25 26

Liu (1994) noted that heating was required for faster reaction and the reaction time may vary from a few minutes to several hours for a temperature range of 6090 C for acidcatalyzed transesterication. From our experiments, three dierent reaction temperatures, 60, 65 and 70 C, were selected. Fig. 7 shows the conversion versus temperature data as a function of acid concentration. At each acid concentration, an increase in nal conversion was evident as temperature was increased. 4.3. Eect of catalyst concentration The catalyst concentration was found to aect the conversion of canola oil to FAME. It is evident from Fig. 7 that an increase in acid concentration served to increase the conversion of TG to FAME. Based on Fig. 7, one can see that between 0.5 and 2 wt.% acid concentration the conversion increased substantially at higher temperatures, but the conversions of 2, 4 and 6 wt.% were not very dierent (<10% conversion). Thus, concentrations of acid

mV

27 28 29 Elution time (min)

30

31

32

33

Fig. 5. HPLC chromatogram of the phase-separated oil phase of the emulsion in the reactor after 6 h of operation (reaction at 65 C, 2 wt.% acid catalyst).

M.A. Dube et al. / Bioresource Technology 98 (2007) 639647

645

1500 1400 1300 1200 1100 1000 900 800 700 600 500 400 300 200 100 0

mV

22

23

24

25

26

27 28 29 Elution time (min)

30

31

32

33

Fig. 6. HPLC chromatogram of permeate after washing (reaction at 65 C, 2 wt.% acid catalyst).

1 0.9 0.8 0.7 Conversion at 6 h 0.6 0.5 0.4 0.3 0.2 0.1 0 58
0.50% 2% 4% 6%

Table 3 Eect of ow rate on conversion Expt. 1 2 3 Flow rate (mL/min) 2.5 3.2 6.1 Temperature (C) 65 65 65 Conversion via acid-catalyst (%) 35 48 64 Conversion via base-catalyst (%) 95 96 96

60

62

64 66 Temperature (C)

68

70

72

Fig. 7. Eect of reaction temperature and acid-catalyst concentration (linear t for 0.05, 2 and 4 wt.%, t not plotted for 6 wt.%).

beyond 2 wt.% are not necessary at 70 C. In addition, the reaction was more sensitive to temperature at high acid concentration. 4.4. Eect of ow rate

the acid-catalyzed case shows that the base catalyst provided a much higher conversion, than that of acid catalyst. Freedman et al. (1984) studied the eect of the type of catalyst on the reaction. It was found that 98% conversion was observed at 1 wt.% sodium hydroxide. They also found that greater than 90% of the oil was converted to methyl esters at 1 wt.% sulphuric acid. In our base-catalyzed experiments, small amounts of soap were detected in the wash waters. These were not found in the acid-catalyzed runs. One possible reason was that the canola oil may have contained signicant amounts of FFA that were converted to soaps rather than FAME by the base catalyst. This would have implications for the use of an acid catalyst which, despite the slower reaction rate, may provide both a technological and economic advantage for the use of lower cost waste feedstock, which contain higher levels of FFA (Zhang et al., 2003a,b). 4.6. Membrane material resistance to degradation

The methanol/acid catalyst feed ow rate was set to 2.5, 3.2 and 6.1 mL/min for three separate experiments at 2 wt.% acid concentrations (see Table 3). A signicant increase in conversion was observed as the ow rate was increased. 4.5. Eect of base catalyst The use of a 1 wt.% NaOH catalyst concentration was tested at dierent ow rates (see Table 3). Comparison to

An important consideration when dealing with high acid or base catalyst concentration is the life of the carbon membrane used in the reactor. The carbon membrane was able to resist the high acid and base environments in our experiments. FAME also presents very strong solvent qualities. After 10 months of operation and contact with methanol/acid or methanol/base solution, no tangible evidence of degradation of the membrane was observed.

646

M.A. Dube et al. / Bioresource Technology 98 (2007) 639647 Armor, J.N., 1998. Applications of catalytic inorganic membrane reactors to renery products. J. Membr. Sci. 147, 217233. Bam, N., Drown, D.C., Korous, R., Homan, D.S., Johnson, T.G., Washam, J.M., 1995. Method for Purifying Alcohol Esters. US Patent No. 5,424,467, June 13. Boocock, D.G.B., Konar, S.K., Mao, V., Lee, C., Buligan, S., 1998. Fast formation of high-purity methyl esters from vegetable oils. J. Am. Oil Chem. Soc. 75, 11671172. Boocock, D.G.B., Konar, S.K., Mao, V., Sidi, H., 1996. Fast one-phase oil-rich processes for preparation of vegetable oil methyl esters. Biomass Bioenergy 11, 4350. Canakci, K., Van Gerpen, J., 1999. Biodiesel production via acid catalysis. Trans. ASAE 42, 12101230. Coltrain, D., 2002. Biodiesel: is it worth considering. In: 2002 Risk and Prot Conference, Manhattan, Kansas. Darnoko, D., Cheryan, M., Perkins, E.G., 2000. Analysis of vegetable oil transesterication products by gel permeation chromatography. J. Liq. Chrom. Rel. Technol. 23, 23272335. Dmytryshyn, S.L., Dalai, A.K., Chaudhari, S.T., Mishra, H.K., Reaney, M.J., 2004. Synthesis and characterization of vegetable oil derived esters: evaluation for their diesel additive properties. Biores. Tech. 92, 5564. Dube, M.A., Zheng, S., McLean, D.D., Kates, M., 2004. A comparison of attenuated total reectance-FTIR spectroscopy and GPC for monitoring biodiesel production. J. Am. Oil Soc. Chem. 81, 599603. Freedman, B., Pryde, E.H., Mounts, T.L., 1984. Variables aecting the yields of fatty esters from transesteried vegetable oils. J. Am. Oil Chem. Soc. 61, 16381643. Freedman, B., Buttereld, R.O., Pryde, E.H., 1985. Transesterication kinetics of soybean oil. J. Am. Oil Chem. Soc. 63, 13751380. Hayafuji, S., Shimidzu, T., Oh, S., Zaima, H., 1999. Method and Apparatus for Producing Diesel Fuel Oil from Waste Edible Oil. US Patent No. 5,972,057, October 26. Hsieh, H.P., 1991. Inorganic membrane reactors. Cat. Rev.: Sci. Eng. 33, 170. Hsieh, H.P., 1996. Inorganic Membranes for Separation and Reaction. Elsevier Science, Amsterdam. Karaosmanoglu, F., Cigizoglu, K.B., Tuter, M., Ertekin, S., 1996. Investigation of the rening step of biodiesel production. Energy Fuels 10, 890895. Kim, I.-C., Kim, J.-H., Lee, K.-H., Tak, T.-M., 2002. Preparation and properties of soluble copolysulfoneimide and performance of solvent-resistant ultraltration membrane. J. Appl. Polym. Sci. 85, 10241030. Liu, J., 2004. Biodiesel Production from Canola Oil Using a Membrane Reactor. M.A.Sc. Thesis, Department of Chemical Engineering, University of Ottawa, Canada. Liu, K., 1994. Preparation of fatty acid methyl esters for gas-chromatographic analysis of lipids in biological materials. J. Am. Oil Chem. Soc. 71, 11791187. McBride, N., 1999. Modeling the Production of Biodiesel from Waste Frying Oil. B.A.Sc. Thesis, Department of Chemical Engineering, University of Ottawa, Canada. Noureddini, H., Zhu, D., 1997. Kinetics of transesterication of soybean oil. J. Am. Oil Chem. Soc. 74, 14571463. Nye, M.J., Williamson, T.W., Deshpande, S., Schrader, J.H., Snively, W.H., 1983. Conversion of used frying oil to diesel fuel by transesterication: preliminary test. J. Am. Oil Chem. Soc. 60, 1598 1601. Ripmeester, W., 1998. Modeling the Production of Biodiesel from Waste Frying Oil. B.A.Sc. Thesis, Department of Chemical Engineering, University of Ottawa, Canada. Saracco, G., Neomagus, H.W.J.P., Versteeg, G.F., van Swaaij, W.P.M., 1999. High-temperature membrane reactors: potential and problems. Chem. Eng. Sci. 54, 19972017. Saracco, G., Specchia, V., 1994. Catalytic inorganic-membrane reactors: present experience and future opportunities. Cat. Rev.: Sci. Eng. 36, 305384.

5. Conclusions The purication of FAME from a canola oil/methanol/ catalyst reaction mixture poses some important challenges in the production of biodiesel. This work has demonstrated that a membrane reactor can be used to alleviate many of these diculties and successfully carry out the transesterication of triglycerides to FAME. In the present semi-batch process, a large methanol:oil ratio was employed. From visual observations, the concentration of FAME in the permeate was not constant as the reaction progressed. Initially, the permeate was quite concentrated but as the reaction proceeded, the permeate concentration decreased. In a continuous process, oil and methanol would be fed to the reactor at a xed ratio resulting in the continuous production of a concentrated permeate. The present work illustrates that oil and methanol can readily co-exist in the reactor at a volume ratio of 1:2 without plugging the membrane pores. This allows for the reaction to be carried out in an emulsion where oil and reacted products can be continuously separated in order to produce a TG-free FAME. In most commercial processes, as the reaction progresses, the formed FAME will eventually behave as a mutual solvent for the TG and alcohol phases. Noureddini and Zhu (1997) have discussed the benets of the formation of a homogeneous alcohol/TG/FAME phase as FAME is formed in the reaction. As discussed in Section 4.1, maintaining a two-phase system in the membrane reactor inhibits the transfer of TG and non-reacting lipids to the product stream. One of the benets of producing a TG-free FAME is a simplication of the often onerous downstream purication of FAME. This, of course, leads to the production of high quality FAME. The membrane reactor allows a phase barrier which limits the presence of TG and non-reacting lipids in the product. This is highly desirable in maintaining quality assurance in the production of biodiesel. Maintaining a phase barrier prohibits the transfer of highly hydrophobic molecules to the product. This provides a limiting barrier in the production of biodiesel. This parallels the advantages of using distillation in maintaining product quality in the petroleum processing industries. Acknowledgements The authors thank the support from the Premiers Research Excellence Award of Ontario and the University of Ottawa. References
Anastopoulos, G., Lois, E., Karonis, D., Zanikos, F., Kalligeros, S., 2001. A preliminary evaluation of esters of monocarboxylic fatty acid on the lubrication properties of diesel fuel. Ind. Eng. Chem. Res. 40, 452456. Armor, J.N., 1989. Catalysis with permselective inorganic membranes. Appl. Cat. 49, 125.

M.A. Dube et al. / Bioresource Technology 98 (2007) 639647 Saracco, G., Versteeg, G.F., Van Swaajj, W.P.M., 1994. Current hurdles to the success of high temperature membrane reactors. J. Membr. Sci. 95, 105123. Watanabe, Y., Shimada, Y., Sugihara, A., Tominaga, Y., 2001. Enzymatic conversion of waste edible oil to biodiesel fuel in a xed-bed bioreactor. J. Am. Oil Chem. Soc. 78, 703707. White, L.S., Nitsch, A.R., 2000. Solvent recovery from lube oil ltrates with a polyimide membrane. J. Membr. Sci. 179, 267274. Zaman, J., Chakma, A., 1994. Inorganic membrane reactors. J. Membr. Sci. 92, 128.

647

Zhang, Y., Dube, M.A., McLean, D.D., Kates, M., 2003a. Biodiesel production from waste cooking oil. 1: Process design and technological assessment. Biores. Tech. 89, 116. Zhang, Y., Dube, M.A., McLean, D.D., Kates, M., 2003b. Biodiesel production from waste cooking oil. 2: Economic assessment and sensitivity analysis. Biores. Tech. 90, 229240. Zheng, S., 2003. Biodiesel Production from Waste Frying Oil: Conversion Monitoring and Modeling. M.A.Sc. Thesis, Department of Chemical Engineering, University of Ottawa, Canada.

S-ar putea să vă placă și