Вы находитесь на странице: 1из 7

Bioresource Technology 99 (2008) 87528758

Contents lists available at ScienceDirect

Bioresource Technology
journal homepage: www.elsevier.com/locate/biortech

Efcient production of biodiesel from high free fatty acid-containing waste oils using various carbohydrate-derived solid acid catalysts
Wen-Yong Lou, Min-Hua Zong *, Zhang-Qun Duan
Laboratory of Applied Biocatalysis, South China University of Technology, Guangzhou 510640, China

a r t i c l e

i n f o

a b s t r a c t
In the present study, such carbohydrate-derived catalysts have been prepared from various carbohydrates such as D-glucose, sucrose, cellulose and starch. The catalytic and textural properties of the prepared catalysts have been investigated in detail and it was found that the starch-derived catalyst had the best catalytic performance. The carbohydrate-derived catalysts exhibited substantially higher catalytic activities for both esterication and transesterication compared to the two typical solid acid catalysts (sulphated zirconia and Niobic acid), and gave markedly enhanced yield of methyl esters in converting waste cooking oils containing 27.8 wt% high free fatty acids (FFAs) to biodiesel. In addition, under the optimized reaction conditions, the starch-derived catalyst retained a remarkably high proportion (about 93%) of its original catalytic activity even after 50 cycles of successive re-use and thus displayed very excellent operational stability. Our results clearly indicate that the carbohydrate-derived catalysts, especially the starch-derived catalyst, are highly effective, recyclable, eco-friendly and promising solid acid catalysts that are highly suited to the production of biodiesel from waste oils containing high FFAs. 2008 Elsevier Ltd. All rights reserved.

Article history: Received 21 January 2008 Received in revised form 11 April 2008 Accepted 11 April 2008 Available online 27 May 2008 Keywords: Carbohydrate-derived catalysts Starch-derived solid acid catalyst Biodiesel production Waste oils High FFAs

1. Introduction Biodiesel is dened as the simple alkyl esters of fatty acids produced from vegetable oils and animal fats. There has been an increasing interest in biodiesel as a green and alternative fuel as a result of recent legislations that require a major reduction of vehicle emissions, as well as the soaring price of petroleum (Ma and Hanna, 1999; Vicente et al., 2004). However, biodiesel has currently not been commercialized all over the world. The major bottleneck is the high cost of feedstock used for biodiesel production, which prohibits its widespread application (Kulkarni and Dalai, 2006). One way of reducing the cost of biodiesel production is to employ low quality feedstocks such as waste cooking oils which are readily available and inexpensive, instead of neat vegetable oil (Watanabe et al., 2005; Zaropoulos et al., 2007; Canakci, 2007). However, such a process is challenging due to the presence of considerable undesirable components especially free fatty acids (FFAs) and water. Use of alkaline catalysts for transesterication of such feedstock is problematic because the alkali reacts with the FFAs to form large amounts of unwanted soap by-products which create serious problem of product separation and ultimately lower the yield substantially (Veljkovic et al., 2006). Homogeneous acid catalysts do not exhibit measurable susceptibility to FFAs, but are difcult to recycle and operate at high temperatures, and give rise
* Corresponding author. Tel.: +86 20 87111452; fax: +86 20 22236669. E-mail address: btmhzong@scut.edu.cn (M.-H. Zong). 0960-8524/$ - see front matter 2008 Elsevier Ltd. All rights reserved. doi:10.1016/j.biortech.2008.04.038

to serious environmental and corrosion problems (Lotero et al., 2005; Canakci and Van Gerpen, 2003). Lipases are generally effective catalysts and are non-polluting (Selmi and Thomas, 1998), but they are expensive and there exist problems associated with their usage in the presence of FFAs and short chain alcohols (such as methanol and ethanol), which denature the enzyme to some extent. Glycerol, which is one of the products of the reaction, manifests a serious negative effect on the enzyme (Dossat et al., 1999; Watanabe et al., 2000; Kose et al., 2002). Solid acid catalysts offer signicant advantages of eliminating separation, corrosion, toxicity and environmental problems (Clark, 2002; Okuhara, 2002), and, therefore, have recently attracted considerable attention. A few reports have dwelt on the importance of solid acids for biodiesel production (Lotero et al., 2005; Kiss et al., 2006; Kulkarni et al., 2006). Apart from recyclability and reusability, an ideal solid acid catalyst for biodiesel preparation should have high stability, numerous strong acid sites, large pores, hydrophobic surface and low cost (Lotero et al., 2005). Inorganic-oxide solid acids such as zeolite and Niobic acid have low densities of effective acid sites and readily lose their activities under harsh conditions (Van Rhijn et al., 1998; Harmer et al., 1998). In particular, these catalysts have small pore and thus are not suitable for biodiesel production because of the diffusion limitation of the large fatty acid molecules. Although strongly acidic ion-exchange resins such as Amberlyst-15 and Naon-NR50 have abundant sulfonic acid groups, these resins are expensive and show bad stability (Kiss et al., 2006; Kulkarni et al., 2006; Harmer and Sun, 2001; Harmer

W.-Y. Lou et al. / Bioresource Technology 99 (2008) 87528758

8753

et al., 2000). Sulphated zirconia, on the other hand, is an efcient solid acid catalyst (Kiss et al., 2006; Yadav and Nair, 1999), but is expensive because zirconium is a rare and costly metal and high temperatures are required both for the calcination and for reactivation of the catalyst. These limitations of the currently available solid acids have restricted their practical utility in biodiesel production. Recently, a new strategy of preparing novel carbon-based solid acids has been developed by Haras research group (Hara et al., 2004; Toda et al., 2005; Takagaki et al., 2006; Okamura et al., 2006): sulfonation of incompletely carbonized D-glucose. Incomplete carbonization of D-glucose leads to a rigid carbon material consisting of small polycyclic aromatic carbon sheets in a threedimensional sp3-bonded structure (Okamura et al., 2006). Sulfonation of such carbon material has been demonstrated to afford a highly stable solid with a high density of active SO3H sites, which is physically robust and there is no leaching of SO3H groups during use and so displays remarkable catalytic performance for the esterication of higher fatty acids. Only recently have we presented the preliminary characterization of this novel D-glucose-derived solid acid catalyst and its successful use for biodiesel production from higher fatty acids and especially waste oils with a high acid value (Zong et al., 2007). Except this preliminary work we previously reported, to the best of knowledge, utilization of carbohydrate-derived solid acid catalysts for biodiesel production from low quality waste oils has not been explored in detail so far. In the present work, we describe the development of a series of carbohydrate-derived catalysts prepared from various cheap starting materials, including starch, cellulose, D-glucose and sucrose, and examine their catalytic activities of both esterication and transesterication. Furthermore, such carbohydrate-derived catalysts are here evaluated for biodiesel preparation from waste cooking oils containing 27.8 wt% FFAs by simultaneous esterication and transesterication. Inuences of several crucial variables such as reaction temperature, molar ratio of alcohol to oil, catalyst loading and reaction time on biodiesel production using the most efcient starch-derived catalyst, are also studied. 2. Experimental 2.1. Materials The pre-treated waste cooking oils (FFAs content: 27.8 wt%; acid value: 55.3 mg KOH/g; saponication value: 194 mg KOH/g; water content: 0.03 wt%) was kindly provided by a local company that collects the waste oils from restaurants by the authority of local government. Non-oil components of the waste cooking oils were removed by separation prior to use. Methyl palmitate, methyl stearate, methyl oleate, methyl linoleate, methyl linolenate, methyl heptadecanoate, oleic acid and triolein were purchased from SigmaAldrich (USA) and were all of over 98% purity. All other chemicals were also obtained from commercial sources and were of the highest grade available. 2.2. Catalyst preparation The carbohydrate-derived solid acid catalysts were prepared via a modication of the published methods (Toda et al., 2005; Takagaki et al., 2006; Okamura et al., 2006). In a typical procedure, 10 g of carbohydrate powder (D-glucose, sucrose, cellulose or starch) was heated for a specied time at an appropriate temperature (P300 C) under N2 ow to produce a brown-black solid (an incomplete carbonization). The resulting material was then ground to a powder and heated in 100 mL of concentrated H2SO4 (>96%) under N2 ow to introduce SO3H into the aromatic carbon rings. After sulfonation of incomplete carbonization for a certain time at a given temperature (P100 C), the mixture was cooled to

30 C and diluted with 500 mL of distilled water to form a black precipitate. Then, the precipitate was collected by ltration and washed repeatedly with hot distilled water (>80 C) until impurities such as sulfate ions were no longer detected in the wash water. The resulting black solids (i.e. carbohydrate-derived solid acid catalysts) were dried at 60 C for 48 h in vacuo to remove water absorbed on the surface of the catalyst. Details about raw materials, carbonization temperature and time, and sulfonation temperature and time are specied for each case. The total yield of the prepared catalyst is around 45 wt% based on the raw material weight. These carbohydrate-derived catalysts are insoluble in the tested solvents and liquid reactants (water, methanol, n-hexane, t-butanol, oleic acid, triolein and waste cooking oils). 2.3. Catalyst characterization Scanning electron microscopy (SEM) of the prepared catalysts was performed by using a FEI Quanta 400 FEG electron microscope with an acceleration voltage of 10 kV. The X-ray diffraction (XRD) analysis was conducted on a Rigaku D/MAX b powder X-ray diffractometer using Cu Ka radiation (k = 0.18415 nm) at 30 kV and 30 mA in the scanning angle (2h) of 260 at a scanning speed of 10/min. The SO3H groups were determined using X-ray photoelectron spectroscopy (XPS) with a Kratos Axis Ultra DLD apparatus. A monochromatic Al Ka (hm = 1486.6 eV) source operating at 150 W was used. The photoelectron pass energies for wide and narrow scans were 160 and 40 eV, respectively. The base pressure of the XPS chamber was 1 109 Torr. Elemental composition of the prepared catalysts was determined by elemental analysis (EA) using Elementar vario EL b apparatus. The absolute errors were 60.1% (CHS) and 60.2% (O). NH3-TPD spectra were recorded using a XianQuan TP5000 ow unit apparatus to characterize the acid site adsorption distribution for each solid acid catalyst. Each sample (50 mg) was placed in a quartz reactor and heated to 300 C in helium (30 mL/ min) for 1 h to remove adsorbed impurities. The temperature was then cooled to 40 C and saturated for 2 h with 100 mL/min of 10% (v/v) NH3/He as carrier gas. Subsequently, the system was ushed with 30 mL/min of He for 2 h. The temperature was ramped up to 300 C at a rate of 10 C/min. A thermal conductivity detector (TCD) was used to measure the desorption prole of NH3. A blank (no catalyst) study of the TPD ramp was made in order to account for baseline drift. As COOH and SO3H groups were present in catalyst samples prepared from carbohydrate (Okamura et al., 2006), the acid site densities estimated by NH3-TPD are total amounts of both functional groups. According to XPS analysis, it is expected that all S atoms in the carbohydrate-derived solid acids are contained in SO3H groups (see below). The densities of SO3H groups were thus estimated by EA. The acid strength of catalyst sample was examined by ultravioletvisible diffuse reectance spectroscopy (UVvis DRS) on a Shimadzu UV-2501PC spectrophotometer. A mixture of catalyst sample (0.2 g) and BaSO4 powder (a reference material for DRS measurements, 1.0 g) was evacuated at 150 C for 1 h to remove adsorbed water. In an Ar-lled glove box, the mixture was packed into a sealable quartz cell, and benzene (with or without color-producing reagent) was then added to the cell. The DRS of the mixture in each benzene solution was measured without exposure to air. The DRS of the color-producing reagent in the presence of the carbohydrate-derived catalyst was obtained by subtracting the spectrum for the mixture in pure benzene from that of the mixture in the benzene solution containing the color-producing reagent. The DRS was also similarly measured using BaSO4 in benzene with each color-producing reagent and also without added color-producing reagent. The specic surface area and the pore size of the catalysts were assayed on Micromeritics Flowsorb III 2310 equipment at 196 C using liquid nitrogen. Prior to the analysis the catalyst was pre-treated at 120 C under vacuum for above

8754

W.-Y. Lou et al. / Bioresource Technology 99 (2008) 87528758

2 h to remove all adsorbed moistures (mainly water) from the catalyst surface and pores. 2.4. Catalytic activity of carbohydrate-derived solid acid catalyst Esterication activity was evaluated by selecting the esterication of oleic acid with methanol as the model reaction. The reaction was conducted by adding 0.14 g of the carbohydrate-derived catalyst into a methanol-oleic acid mixture (methanol: 100 mmol; oleic acid: 10 mmol) at 80 C and a stirring rate of 500 rpm. Aliquots (200 lL) were withdrawn at specied time intervals from the reaction mixture and centrifuged (10,000 rpm, 10 min) to remove the catalyst, and were then kept for 10 min at 70 C under vacuum to remove methanol from the samples, followed by addition of anhydrous n-hexane (200 lL) containing 1.0 mM methyl heptadecanoate (as an internal standard) for GC analysis. Transesterication activity was evaluated by selecting the transesterication of triolein with methanol as the model reaction. The reaction was carried out by adding 0.14 g of the prepared catalyst into a methanoltriolein mixture (methanol: 100 mmol; triolein: 10 mmol) at 80 C and a stirring rate of 500 rpm. Prior to GC analysis, samples (200 lL) were taken and worked up as described above. The impact of stirring rate on both esterication reaction rate and transesterication reaction rate with each catalyst was examined (data not shown). When the stirring rate was 500 rpm or above, no external mass transfer limitations were observed. 2.5. Biodiesel production from waste cooking oils Simultaneous esterication and transesterication for biodiesel production were conducted by adding a xed amount of the prepared catalyst to a mixture of waste cooking oils (27.8 wt% FFAs, 5 g) and methanol at a stirring rate of 500 rpm and an appropriate temperature. Samples (200 lL) were taken, worked up and analyzed as described in Section 2.3. Details about molar ratio of methanol to oils, catalyst loading, reaction temperature and reaction time are specied for each case. 2.6. Operational stability of starch-derived solid acid catalyst In order to assess the operational stability of the starch-derived solid acid catalyst, the re-use of the prepared catalyst was investigated in the production of biodiesel from waste cooking oils (27.8 wt% FFAs). Initially, 0.5 g (10 wt% catalyst loading) of the catalyst was added into the mixture of waste cooking oils (5 g) and methanol (30:1 methanol to oil molar ratio) and the reaction was conducted at 500 rpm and 80 C. Then, the reaction was repeated over P50 cycles (8 h per cycle) under the same reaction conditions described above. After each cycle of use in the production of biodiesel, the catalyst was recovered by ltration before re-use. The relative activity of the catalyst employed for the rst batch was dened as 100%. 2.7. GC analysis Reaction mixture were assayed with a GC 2010 gas chromatograph (Shimadzu Corp., Kyoto, Japan) with a HP-5 capillary column (0.53 mm 15 m, HewlettPackard, USA) equipped with a ame ionization detector. The column temperature was held at 180 C for 1 min, then raised to 186 C at 0.8 C/min and kept at constant temperature for 1 min. Nitrogen was used as the carrier gas at a ow rate of 2 mL/min. The split ratio was 1:100 (v/v). The injector and the detector temperatures were set at 250 C and 280 C, respectively. The retention times for methyl myristate, methyl palmitate, methyl heptadecanoate, methyl oleate, methyl linoleate,

and methyl stearate were 2.11, 4.12, 5.88, 7.64, 7.68 and 8.45 min, respectively. The average error for this determination was less than 0.5%. All reported data are averages of experiments performed at least in duplicate. 3. Results and discussion 3.1. Catalytic activity of carbohydrate-derived solid acid catalyst Recent studies have shown that a novel carbon-based solid acid catalyst, consisting of small polycyclic aromatic carbon sheets with high densities of SO3H groups, can be readily synthesized by sulfonation of incompletely carbonized D-glucose and exhibits remarkable catalytic performance for esterication reaction as well as hydration reaction (Toda et al., 2005; Takagaki et al., 2006; Okamura et al., 2006). However, such highly active and stable carbon-based solid acid could not be prepared by sulfonation of an incompletely carbonized resin, amorphous glassy carbon, activated carbon, or natural graphite. Catalyst samples prepared from such starting materials showed no signicant activity for esterication, hydration or hydrolysis (Okamura et al., 2006). These clearly suggest that the catalytic performances of the novel catalysts are signifcantly dependent on the starting materials used for their preparation. Therefore, we initially focused on inuences of different starting materials on catalytic and textural properties of novel solid acid catalysts prepared from them. Several cheap carbohydrate compounds such as starch, cellulose, sucrose and D-glucose were selected as starting materials, and were pyrolyzed, accompanied by dehydration and dissociation of COC at 400 C for 15 h, leading to the formation of polycyclic aromatic carbon rings (incomplete carbonization). SO3H groups were then introduced into the aromatic carbon rings by sulfonation with concentrated H2SO4 at 150 C for 15 h, resulting in the four corresponding carbohydrate-derived solid acids. Subsequently, the four catalyst samples were examined through methyl oleate (high-grade biodiesel) formation from oleic acid (Fig. 1), one of main ingredients in various vegetable oils, as an example of the esterication of higher fatty acids. After 3 h reaction, the four corresponding catalysts prepared from starch, cellulose, sucrose and D-glucose gave a yield of 95%, 88%, 80%, and 76%, respectively. It is obvious that different raw materials have shown great impact on catalytic activity of the resulting catalyst samples. Time-dependent curves of methyl oleate yield as depicted in Fig. 1 clearly indicated that the starchderived catalyst was much more active than the other ones, achieving the maximum yield of 95% only within 3 h, compared to 45 h for the other catalysts. The marked difference in catalytic performance of the four catalysts prepared from different raw materials is expected to be closely related to the physical and chemical properties of the catalyst samples. No signicant difference in morphology was observed among the prepared four catalysts by the SEM analysis (data not shown). The particles of all these carbohydrate-derived catalysts reach micrometre dimensions and do not signicantly aggregate. All the XRD diffraction patterns exhibit broad and weak diffraction peaks (2h = 1030; 2h = 3550), which are attributable to amorphous carbon composed of polycylic aromatic carbon sheets oriented in a considerable random fashion (Tsubouchi et al., 2003). Also, no clear difference in the XRD patterns was found among the four catalysts. However, the notably different textural properties of the four catalysts summarized in Table 1 could be responsible for such difference in catalytic activity of methyl oleate formation. Elementary analysis (EA) showed that the catalyst samples had clearly different compositions. The starch-derived catalyst contained a higher content of S element than the three catalysts. In the XPS for the catalysts prepared from starch, cellulose, sucrose and D-glucose (Fig. 2), the single S 2p peak was observed at 168 eV, assigned to SO3H groups. It meant that

W.-Y. Lou et al. / Bioresource Technology 99 (2008) 87528758

8755

essentially all S atoms in the prepared catalyst samples were contained in SO3H groups. Hence, the densities of SO3H sites in the carbohydrate-derived catalysts could be calculated based on the S content in catalyst compositions. Table 1 clearly indicated that the starch-derived catalyst afforded higher densities of SO3H groups than the other three catalysts, which is consistent with the observation that the starch-derived catalyst gave a faster reaction rate as shown in Fig. 1. It was interestingly noted that the acid site densities of all the four catalysts estimated by NH3-TPD were slightly higher (0.14 mmol/g or so) than the SO3H site densities, possibly resulting from small amounts of COOH groups present in the carbohydrate-derived catalysts. It has been conrmed by Haras group using 13C MAS NMR that the COOH groups are present in the catalyst prepared from D-glucose (Takagaki et al., 2006; Okamura et al., 2006). Additionally, the starch-derived catalyst manifested clearly larger BET surface area, pore volume and pore size than the catalysts prepared from the other three starting materials, which could also well explain for higher activity of starch-derived catalyst in esterifying oleic acid to methyl oleate. Surprisingly, for all the four carbohydrate-derived catalysts the densities of SO3H groups as main functional sites were as high as 1.471.83 mmol/g despite the small surface areas (4.17.2 m2/g). These densities were much too high to be attributed to SO3H groups attached to the catalyst surface. This suggests that SO3H groups in the amorphous carbon bulk can also participate in the esterication. Therefore, the incorporation of large reactant molecules into the bulk of amorphous carbon can greatly affect the reaction. Thanks to the relatively large pore volume and pore size, the starch-derived catalyst made reactants more accessible to the SO3H sites in the amorphous carbon bulk than each of the three catalysts prepared from cellulose, sucrose and D-glucose, thus dis100

playing high efciency in catalyzing the esterication of higher fatty acids as illustrated in Fig. 1. Therefore, among the carbohydrate compounds tested, starch was thought to the best starting material for preparing highly active carbohydrate-derived solid acid catalysts. Certainly, further detailed investigations are required to get a deep insight into the observation that the four raw materials resulted in such signicant difference in catalytic and textural properties of the catalysts prepared from them. 3.2. Inuences of the preparation variables on the catalytic activity of carbohydrate-derived catalyst In the course of preparing such carbohydrate-derived catalysts, it was found that carbonization temperature and time, and sulfonation temperature and time also signicantly affected the catalytic performance of the resulting catalysts. Hence, subsequent work was made on the effects of the above-mentioned preparation variables on the esterication of oliec acid catalyzed by the resulting starch-derived catalyst. As can be seen in Table 2, with increasing carbonization temperature up to 400 C, the resulting starch-derived catalyst became more active and led to the improvement in the yield of methyl oleate within 3 h. Further increase in carbonization temperature above 400 C resulted in a clear decline in catalytic activity of the prepared catalyst. This may be because the sample carbonzied at lower temperature (below 400 C) has higher densities of OH groups and more water from the esterication can be adsorbed on the surface of the resulting catalyst, and therefore the formed water layer will prevent the access of relatively hydrophobic oleic acid to the catalyst. Also, the sample carbonized at lower temperature (below 400 C) is soft aggregate of small polycylic aromatic carbon with SO3H rather than rigid carbon material, and

500 80

Starch-derived catalyst

Yield of methyl oleate (%)

450 Cellulose-derived catalyst 60 400 350 40

SO3H

Sucrose-derived catalyst D-glucose-derived catalyst

20

Starch-derived catalyst Cellulose-derived catalyst Sucrose-derived catalyst D-glucose-derived catalyst

CPS

300 250 200

0 0 1 2 3 4 5 6

Reaction time (h)


Fig. 1. Time-dependent curves of methyl oleate formation catalyzed by various carbohydrate-derived catalysts prepared from different starting materials (D-glucose, sucrose, cellulose and starch). Reaction conditions: 10 mmol oleic acid, 100 mmol methanol, 80 C, 500 rpm, 0.14 g catalyst.

150 176 174 172 170 168 166 164 162

Binding energy (eV)


Fig. 2. S 2p XPS spectra for the catalysts prepared from starch, cellulose, sucrose and D-glucose.

Table 1 Textural properties of various solid acid catalysts Catalyst (composition) S content (wt%) Acid site density (mmol/g) Total
D-Glucose-derived

SO3H 1.47 1.59 1.68 1.83

Surface area (m2/g) 4.1 5.0 5.7 7.2 218 128

Average pore volume (cm3/g) 0.44 0.52 0.65 0.81 0.18

Average pore size (nm) 4.0 5.1 6.4 8.2 2.7

catalyst (CH1.14O0.39S0.030) Sucrose-derived catalyst (CH1.06O0.30S0.029) Cellulose-derived catalyst (CH1.01O0.28S0.031) Starch-derived catalyst (CH0.85O0.23S0.032) Sulphated zirconia Niobic acid (Nb2O5 nH2O)

4.7 5.1 5.4 5.9 2.5

1.60 1.71 1.82 1.97 0.4 0.3

8756

W.-Y. Lou et al. / Bioresource Technology 99 (2008) 87528758

therefore the sulfopolycyclic aromatic compounds can be readily leached from the solid at high reaction temperature or with higher fatty acids as reactants. On the other hand, the sample carbonized at higher temperature (above 400 C) becomes more rigid due to the growth of large polycylic aromatic carbon sheets and the stacking of carbon sheets, and thus the resulting catalyst contains lower densities of SO3H sites as well as less xible structure. 400 C represents a balance between these conicts and so the resulting catalyst sample carbonized at 400 C displays higher activity in methyl oleate formation. At the optimal carbonization temperature of 400 C, the suitable carbonization time was thought to be 15 h on account of higher catalytic activity of the resulting catalyst (Table 2). Similarly, under the preparation conditions tested the most effective catalyst was produced when sulfonation was conducted at 150 C for 15 h (Table 2). There is no doubt that the textural properties of the starch-derived catalysts such as the compositions, the acid site density, and the inner porous structure will be affected when the variables of carbonization and sulfonation are changed. The effect proles of such variables on the textural properties of the prepared catalysts are expected to be capable of further explaining the results described above, and are the subjects of ongoing investigations in our laboratory. 3.3. Comparison of the catalytic activities of various catalysts It is of particular interest to compare the catalytic activities of the carbohydrate-derived catalysts with those of the concentrated H2SO4 (>96%) and the two typical strong solid acid catalysts, sulphated zirconia and Niobic acid (Nb2O5 nH2O) that are widely used in industrial acid-catalyzed processes (Okuhara, 2002; Ecormier et al., 2003), in the esterication and transesterication reactions involved in biodiesel production from waste oils. The esterication and transesterication activities were evaluated through the esterication of oleic acid (10 mmol) and transesterication of triolein (10 mmol) with methanol (100 mmol) at 80 C, respectively. The results for 0.14 g of all the tested catalysts were shown in Fig. 3. Obviously, the four carbohydrate-derived catalysts exhibited much higher esterication and transesterication activities (on a weight basis unless otherwise specied) than those of sulphated zirconia and Niobic acid. Although the two typical solid acid catalysts possess much larger specic surface areas than the four carbohydrate-derived catalysts, the acid site densities of the four carbohydrate-derived catalysts are 46 times lager than the two solid acid catalysts (Table 1). Therefore, the superior catalytic activity of the carbohydrate-derived catalysts can be explained on the basis of their markedly greater densities of active acid sites. Also, the larger pore volume and size of the carbohydrate-derived catalysts, which will result in accessibility of large molecular reactants (oleic acid or triolein) to the active acid sites in the bulk of the catalysts, can well account for their considerably excellent catalytic performance for both reactions compared to Sulphated zirconia. It is well known

that Niobic acid (providing acidic OH groups) is a non-porous solid acid and displays only marginal acidity in the presence of water. So the negative effect of the formed by-product water in the esterication reaction on Niobic acid can be partially responsible for its low esterication activity towards oleic acid. Additionally, Niobic acid showed only slight transesterication activity towards triolein, which might be attributable to the inaccessibility of more hydrophobic triolein to the effective acid sites attached on the relatively hydrophilic surface of the catalyst. As can be clearly seen from the data depicted in Fig. 3, the esterication and transesterication activities of the three carbohydrate-derived catalysts from D-glucose, sucrose and cellulose were much higher than half those of concentrated H2SO4, while the activities of Sulphated zirconia and Niobic were clearly below 19% of those of concentrated H2SO4. Amazingly, the starch-derived catalyst manifested almost comparable esterication and transesterication activities with those of concentrated H2SO4. Also, ultravioletvisible diffuse reectance spectroscopy (UVvis DRS) analysis indicates that the four carbohydrate-derived catalysts have a strong acid strength with pKa being between 11 and 8, which is comparable to that of concentrated H2SO4 (pKa = 11.9). On the other hand, the concentrated H2SO4 as a homogeneous liquid acid catalyst cannot be readily recycled and presents a threat to the environment and the operators health, especially when it is employed on a large scale, whilst the carbohydrate-derived solid acid catalysts themselves are relatively non-toxic, eco-friendly, and can be easily recy-

800

Esterification activity (mol/min g)

160 700 140 600 120 500 400 300 200 100 0 100 80 60 40 20 0
4

Fig. 3. Esterication and transesterication activities of various catalysts tested. Reaction conditions: 10 mmol oleic acid or triolein, 100 mmol methanol, 80 C, 500 rpm, 0.14 g catalyst.

Table 2 Inuence of the preparation variables on the catalytic activity of the starch-derived catalysta Carbonization temperature Temperature (C) 300 350 400 450 500 Yield (%) 70.1 81.3 94.9 85.4 80.2
b

Carbonization time Time (h) 10.0 12.5 15.0 17.5 20.0 Yield (%) 71.2 81.3 94.9 73.3 61.3
b

Sulfonation temperature Temperature (C) 100 125 150 175 200 Yield (%) 83.8 89.8 94.9 86.6 78.8
b

lu co se -d er iv Su ed cr ca os ta ely de C st ri el ve lu lo d se ca -d ta er ly iv st St e d ar ca ch ta -d ly er st iv ed c at Su al lp ys ha t te d zi rc on ia N io bi C ca on ci ce d nt ra te d H 2 S O

-g

Sulfonation time Time (h) 10.0 12.5 15.0 17.5 20.0 Yieldb (%) 76.3 84.4 94.9 86.6 80.5

a Catalyst activity was measured in the model reaction, esterication of oleic acid with methanol. The reaction parameters that were not specically changed as indicated in the table were set as follows: carbonization temperature, 400 C; carbonization time, 15 h; sulfonation temperature, 150 C; sulfonation time, 15 h; oleic acid, 10 mmol; methanol, 100 mmol; starch-derived carbohydrate catalyst, 0.14 g; stirring rate, reaction temperature and time for the esterication, 500 rpm, 80 C and 3 h, respectively. b Yield of methyl oleate.

Transesterification activity (mol/min g)

Esterification activity

Transesterification activity

180

W.-Y. Lou et al. / Bioresource Technology 99 (2008) 87528758

8757

cled and re-used many times though concentrated H2SO4 is employed during their preparation. Therefore, the carbohydrate-derived catalysts in this study, especially the starch-derived catalyst, have clear potential for use as a replacement for H2SO4 in both esterication and transesterication involved in biodiesel production from waste oils. As can be clearly seen in Fig. 3, for all the tested catalysts, the esterication activity towards oleic acid was much higher than the transesterication activity towards triolein under identical reaction conditions, which might be due to the following three possible reasons. First, fatty acids, particularly unsaturated ones such as oleic acid, are more soluble in methanol than triglycerides such as triolein, and, therefore, the esterication reaction proceeded faster. Second, acid-catalyzed esterication reaction requires relatively low activation energy compared to transesterication reaction, thus resulting in a faster reaction rate under identical reaction conditions. The third reason is probably related to the reaction mechanism of esterication and transesterication. Methanolysis of fatty acid proceeds through simple esterication, while methanolysis of triglyceride proceeds through complicated transesterication which consists of three consecutive and reversible reaction steps. It is well known that transesterication of triglyceride with methanol is a stepwise reaction, and in the reaction sequence triglyceride is converted stepwise to di- and monoglyceride and nally glycerol. When tri-, di- and monoglyceride come in contact with the acid sites, respectively, transesterication takes place and generates one mole of fatty acid methyl ester in each step. The transesterication in each step is reversible. Owing to these reasons, the rate of triolein transesterication was markedly slower than that of oleic acid esterication, which is in accordance with much lower activity of transesterication than esterication as illustrated in Fig. 3. Similar observations were also described by other research groups (Kulkarni et al., 2006; Warabi et al., 2004). Thus, the feedstocks with a higher content of FFAs appears preferable to that with a lower FFAs content for biodiesel production by acid-catalyzed simultaneous esterication and transesterication on account of substantially faster reaction rate. 3.4. Biodiesel production from waste cooking oils with carbohydratederived solid acid catalyst In view of the excellent catalytic performance, the carbohydrate-derived catalysts were further tested for the conversion of waste cooking oils containing high free fatty acids (27.8 wt% FFAs), in the presence of methanol, to fatty acid methyl esters that

100

80
Starch-derived catalyst Cellulose-derived catalyst Sucrose-derived catalyst D-glucose-derived catalyst Sulfated zirconia Niobic acid

60

40

20

0 0 2 4 6 8 10 12 14

Reaction time (h)


Fig. 4. The biodiesel production from waste cooking oils containing 27.8 wt% FFAs using various solid acid catalysts. Reaction conditions: methanol to oil molar ratio, 20:1; catalyst loading, 10 wt% based on the weight of waste oils; stirring rate, 500 rpm; reaction temperature, 80 C.

constitute biodiesel by simultaneous esterication and transesterication. As transesterication of oil (triglyceride) using heterogeneous acid catalysts is well known for its reversibility and slow reaction rate, use of excess methanol can improve the reaction rate and favour forward reaction, maximizing the yield of fatty acid methyl esters (Furuta et al., 2004). Thus, the production of biodiesel using various carbohydrate-derived catalysts was initially conducted at a relatively high molar ratio of methanol to oil (20:1, mol/mol), combined with reaction temperature of 80 C, stirring rate of 500 rpm and catalyst loading of 10 wt% based on the weight of the used waste oils. For comparison, the results for Sulphated zirconia and Niobic acid were also shown in Fig. 4. The four carbohydrate-derived catalysts gave much faster reaction rates than Sulphated zirconia and Niobic acid, achieving the maximum yield (above 80%) of fatty acid methyl esters within 812 h, compared to the yield of 44% for Sulphate zirconia and the yield of only 16% for Niobic acid within 14 h or a longer time. Of all the tested catalysts, the starch-derived catalyst proved to be the most effective in catalyzing the conversion of the waste cooking oils to biodiesel by simultaneous esterication and transesterication, and afforded the yield of around 83% only within 8 h, consistent with the higher catalytic activity of the starch-derived catalyst in the model esterication and transesterication reactions. Since it was the best-performing catalyst for biodiesel production, the starchderived catalyst was selected further to investigate in detail the effect of various variables on the methyl ester yield for optimization of biodiesel production process. Theoretically, one mole of triglyceride requires three moles of methanol to convert it to the corresponding fatty acid methyl esters. In addition, transesterication of triglycerides present in waste oils is a reversible reaction. Hence, the excess of methanol relative to oil can shift the equilibrium towards methyl ester formation. The yield of methyl esters (reaction conditions: 10 wt% catalyst loading; 80 C) markedly increased from 50% to 92% after 8 h reaction with the increase of molar ratio of methanol to oil from 5:1 to 30:1, however, further increase in methanol to oil molar ratio to 40:1 showed only slight enhancement (<2%) in the methyl ester yield. Therefore, the optimal molar ratio of methanol to oil was shown to be 30:1. The excess of methanol used in the process can be collected and reused. The catalyst loading is one of the important parameters that affect the yield of methyl esters and the cost of biodiesel production. As expected, the yield of methyl esters after 8 h (reaction conditions: 30:1 methanol/oil molar ratio; 80 C) greatly increased with increasing catalyst loading up to 10 wt% (based on the oil weight unless otherwise specied). However, when the catalyst loading was further increased to 14 wt%, the yield of methyl esters (about 93%) was similar to that achieved with 10 wt% catalyst loading. Taking into account the yield and the cost, 10 wt% catalyst loading was thought to be suitable for the conversion of waste oils to biodiesel. Reaction temperature shows the signicant effect on the yield of methyl esters after 8 h (reaction conditions: 30:1 methanol/oil molar ratio; 10 wt% catalyst loading). The yield remarkably increased with increasing reaction temperature within the range of 6580 C. However, when reaction temperature was further increased from 80 C to 100 C, the initial reaction rate increased to some extent but the nal yield of methyl esters after 8 h displayed no appreciable improvement. Thus, 80 C was selected as the optimal reaction temperature for the process. Under the optimized conditions described above (30:1 methanol to oil molar ratio, 10 wt% catalyst loading, 80 C reaction temperature), the reaction time-dependent prole of the methyl ester yield with the most effective starch-derived catalyst indicated that the yield clearly increased with increasing reaction time up to 8 h, and thereafter remained almost constant at about 92%, which probably represents a near-equilibrium yield of methyl esters. The FFA content in the reaction system

Yield of fatty acid methyl esters (%)

8758

W.-Y. Lou et al. / Bioresource Technology 99 (2008) 87528758

was substantially reduced to around 0.5 wt%. In the production of biodiesel from waste cooking oils containing high FFAs, the transesterication activity of the starch-derived solid acid catalyst was much lower than its esterication activity (Fig. 3), and consequently it required a relatively long time to get a near-equilibrium yield of methyl esters. It is well known that the alkali catalyst for the transesterication is much more effective than the acid catalyst. Therefore, the use of the alkali catalyst for the transesterication after the use of the starch-derived solid acid catalyst for the esterication can further improve the yield and the efciency of biodiesel production from waste cooking oil containing high FFAs, and this is now under investigation in our laboratory. 3.5. Operational stability of starch-derived solid acid catalyst The catalyst recycling is a crucial step as it reduces the cost of biodiesel production. The efciency of the catalyst also depends on its reusability. In order to evaluate the reusability, the starchderived catalyst was recovered for further conversion of waste oils under the optimized conditions by simple ltering. It was surprisingly found that the starch-derived catalyst still retained about 93% of its original catalytic activity, even after fty cycles of successive re-use, indicating the excellent operational stability. Generally, the activities of most solid acid catalysts with hydrophilic surface are seriously hindered by water produced during the esterication reaction because of the formation of water layer on the hydrophilic surface preventing the access of relatively hydrophobic substrate (Kiss et al., 2006). In the case of the starch-derived catalyst, the XRD diffraction pattern mentioned above, which exhibits broad and weak diffraction peaks (2h = 1030; 2h = 3550), showed that similar to the previously described observation (Tsubouchi et al., 2003; Takagaki et al., 2006), the catalyst was composed of polycylic aromatic carbon sheets. Therefore, the surface of the starch-derived catalyst is relatively hydrophobic and it is less likely to form a water layer on its surface. So the more hydrophobic fatty acids or glycerides could readily access the catalyst. This might partially explain for the observation that the starch-derived catalyst was especially robust and was less hindered by the by-product water during biodiesel production from waste cooking oils containing high FFAs and the in-depth understanding of the observation is on the way in our laboratory. 4. Conclusions The present study clearly showed that the carbohydrate-derived solid acids, prepared from D-glucose, sucrose, cellulose and starch through sulfonation of incompletely carbonized carbohydrate, are non-toxic, inexpensive, recyclable and promising eco-friendly catalysts. The starting materials, carbonization temperature and time, and sulfonation temperature and time for catalyst preparation all had signicant impact on the catalytic and textural properties of the prepared catalysts. The carbohydratederived catalysts prepared at the optimized conditions displayed much higher activities than typical Sulfated zirconia and Niobic acid for both esterication and transesterication. More importantly, various carbohydrate-derived catalysts, especially starchderived catalyst, were shown to be highly effective in converting high FFA-containing waste oils to biodiesel by simultaneous esterication and transesterication. Under the optimized reaction conditions, usage of the most effective starch-derived catalyst for biodiesel production from waste cooking oils containing 27.8 wt% FFAs afforded the methyl ester yield of about 92% after 8 h. This catalyst also manifested very excellent operational stability. As well as being potentially useful for biodiesel production, the carbohydrate-derived catalysts may nd wide applications as a heterogeneous strong acid catalyst.

References
Canakci, M., 2007. The potential of restaurant waste lipids as biodiesel feedstocks. Bioresource Technology 98, 183190. Canakci, M., Van Gerpen, J., 2003. A pilot plant to produce biodiesel from high free fatty acid feedstocks. Transactions of the ASAE 46, 945954. Clark, J.H., 2002. Solid acids for green chemistry. Accounts of Chemical Research 35, 791797. Dossat, V., Combes, D., Marty, A., 1999. Continuous enzymatic transesterication of high oleic sunower oil in a packed bed reactor: inuence of the glycerol production. Enzyme and Microbial Technology 25, 194200. Ecormier, M.A., Wilson, K., Lee, A.F., 2003. Structure-reactivity correlations in sulphated-zirconia catalysts for the isomerisation of alpha-pinene. Journal of Catalysis 215, 5765. Furuta, S., Matsuhashi, H., Arata, K., 2004. Biodiesel fuel production with solid superacid catalysis in xed bed reactor under atmospheric pressure. Catalysis Communications, 721723. Hara, M., Yoshida, T., Takagaki, A., Takata, T., Kondo, J.N., Hayashi, S., Domen, K., 2004. A carbon material as a strong protonic acid. Angewandte Chemie International Edition 43, 29552958. Harmer, M.A., Farneth, W.E., Sun, Q., 1998. Towards the sulfuric acid of solids. Advanced Materials 10, 1255. Harmer, M.A., Sun, Q., 2001. Solid acid catalysis using ion-exchange resins. Applied Catalysis A: General 221, 4562. Harmer, M.A., Sun, Q., Vega, A.J., Farneth, W.E., Heidekum, A., Hoelderich, W.F., 2000. Naon resin-silica nanocomposite solid acid catalysts. Microstructure processing property correlations. Green Chemistry 2, 714. Kiss, A.A., Dimian, A.C., Rothenberg, G., 2006. Solid acid catalysts for biodiesel production towards sustainable energy. Advanced Synthesis and Catalysis 348, 7581. Kose, O., Tuter, M., Aksoy, H.A., 2002. Immobilized Candida antarctica lipasecatalyzed alcoholysis of cotton seed oil in a solvent-free medium. Bioresource Technology 83, 125129. Kulkarni, M.G., Dalai, A.K., 2006. Waste cooking oil an economical source for biodiesel: A review. Industrial and Engineering Chemistry Research 45, 2901 2913. Kulkarni, M.G., Gopinath, R., Meher, L.C., Dalai, A.K., 2006. Solid acid catalyzed biodiesel production by simultaneous esterication and transesterication. Green Chemistry 8, 10561062. Lotero, E., Liu, Y.J., Lopez, D.E., Suwannakaran, K., Bruce, D.A., Goodwin, J.G., 2005. Synthesis of biodiesel via acid catalysis. Industrial & Engineering Chemistry Research 44, 53535363. Ma, F.R., Hanna, M.A., 1999. Biodiesel production: a review. Bioresource Technology 70, 115. Okuhara, T., 2002. Water-tolerant solid acid catalysts. Chemical Reviews 102, 3641 3665. Okamura, M., Takagaki, A., Toda, M., Kondo, J.N., Domen, K., Tatsumi, T., Hara, M., Hayashi, S., 2006. Acid-catalyzed reactions on exible polycyclic aromatic carbon in amorphous carbon. Chemistry of Materials 18, 30393045. Selmi, B., Thomas, D., 1998. Immobilized lipase-catalyzed ethanolysis of sunower oil in a solvent-free medium. Journal of the American Oil Chemists Society 75, 691695. Takagaki, A., Toda, M., Okamura, M., Kondo, J.N., Hayashi, S., Domen, K., Hara, M., 2006. Esterication of higher fatty acids by a novel strong solid acid. Catalysis Today 116, 157167. Toda, M., Takagaki, A., Okamura, M., Kondo, J.N., Hayashi, S., Domen, K., Hara, M., 2005. Green chemistry biodiesel made with sugar catalyst. Nature 438, 178. Tsubouchi, N., Xu, C.B., Ohtsuka, Y., 2003. Carbon crystallization during hightemperature pyrolysis of coals and the enhancement by calcium. Energy and Fuels 17, 11191125. Van Rhijn, W.M., De Vos, D.E., Sels, B.F., Bossaert, W.D., Jacobs, P.A., 1998. Sulfonic acid functionalised ordered mesoporous materials as catalysts for condensation and esterication reactions. Chemical Communications, 317318. Veljkovic, V.B., Lakicevic, S.H., Stamenkovic, O.S., Todorovic, Z.B., Lazic, M.L., 2006. Biodiesel production from tobacco (Nicotiana tabacum L.) seed oil with a high content of free fatty acids. Fuel 85, 26712675. Vicente, G., Martinez, M., Aracil, J., 2004. Integrated biodiesel production: a comparison of different homogeneous catalysts systems. Bioresource Technology 92, 297305. Warabi, Y., Kusdiana, D., Saka, S., 2004. Reactivity of triglycerides and fatty acids of rapeseed oil in supercritical alcohols. Bioresource Technology 91, 283287. Watanabe, Y., Pinsirodom, P., Nagao, T., Kobayashi, T., Nishida, Y., Takagi, Y., Shimada, Y., 2005. Production of FAME from acid oil model using immobilized Candida antarctica lipase. Journal of the American Oil Chemists Society 82, 825831. Watanabe, Y., Shimada, Y., Sugihara, A., Noda, H., Fukuda, H., Tominaga, Y., 2000. Production of biodiesel fuel from vegetable oil using immobilized Candida antarctica lipase. Journal of the American Oil Chemists Society 77, 355360. Yadav, G.D., Nair, J.J., 1999. Sulfated zirconia and its modied versions as promising catalysts for industrial processes. Microporous and Mesoporous Materials 33, 148. Zaropoulos, N.A., Ngo, H.L., Foglia, T.A., Samulski, E.T., Lin, W., 2007. Catalytic synthesis of biodiesel from high free fatty acid-containing feedstocks. Chemical Communications, 36703672. Zong, M.H., Duan, Z.Q., Lou, W.Y., Smith, T.J., Wu, H., 2007. Preparation of a sugar catalyst and its use for highly efcient production of biodiesel. Green Chemistry 9, 434437.

Вам также может понравиться