Open access

Valorization of Waste Cooking Oil into Biodiesel over Heteropolyacids Immobilized on Mesoporous Silica — A Kinetic Study

Written By

M. Caiado, A. Tropecêlo and J.E. Castanheiro

Submitted: 30 September 2014 Published: 30 September 2015

DOI: 10.5772/59584

From the Edited Volume

Biofuels - Status and Perspective

Edited by Krzysztof Biernat

Chapter metrics overview

1,697 Chapter Downloads

View Full Metrics

1. Introduction

Biodiesel is a source of energy derived from renewable sources, which can be a substitute for diesel fuel. It is biodegradable, and it has limited greenhouse gas emissions because of the closed CO2 cycle and a lower combustion emission profile (especially SOx). Biodiesel consists of fatty acid alkyl esters (usually methyl esters, FAME) derived from either the transesterification of triglycerides and/or the esterification of free fatty acids (FFAs) with low molecular weight alcohols [1-3].

Traditionally, the production of biodiesel is carried out in the presence of a homogeneous base catalyst; however, effluent disposal leads to environmental problems and economical inconveniences. These problems can be overcome by the use of heterogeneous catalysts. Another problem with commercial production of biodiesel is the high cost of raw materials. In order to overcome this problem, waste oils and fats can be used as feedstocks [4-7]. However, waste cooking oils contain a high amount of free fatty acids, which is a problem for biodiesel production by the traditional process (homogeneous alkali-catalyzed transesterification) [8,9]. Different solid acids such as resins with sulfonic acid groups [10-14], zeolites [15-17], solid super acid catalysts [18] and carbonaceous catalysts [19] have been used as catalysts in the conversion of waste cooking oil containing a high amount of free fatty acids into biodiesel.

Heteropolyacids (HPAs) have several advantages as catalysts, which make them economically and environmentally attractive. On the one hand, HPAs have a very strong Bronsted acidity, approaching the superacid region; on the other, they are efficient oxidants, exhibiting fast reversible multielectron redox transformations under rather mild conditions. Their acid-base and redox properties can be varied over a wide range by changing the chemical composition. Solid HPAs possess a discrete ionic structure, comprising fairly mobile basic structural units, e.g., heteropolyanions and counter cations (H+, H3O+, H5O2 +, etc.) unlike the network structure of, e.g., zeolites and metal oxides [20-22]. HPAs have low specific surface areas (1–10 m2/g). In order to increase the specific area of HPAs or, even better, to increase the number of accessible acid sites of the HPAs, a variety of supports like activated carbon [23-26], silica [27-32], zeolite [33-38] and polymeric matrix [39-43] have been used as support to immobilize HPAs.

HPAs and HPAs supported on different supports have been used as acid catalysts for biodiesel production [44-48].

Transesterification of waste cooking oil with high acid value and high water content was performed using heteropolyacid H3PW12O40.6H2O (PW) as homogeneous catalyst. PW was found to be the most promising catalyst, which exhibited the highest ester yield (87%) for transesterification of waste cooking oil and an ester yield of 97% for esterification of long-chain palmitic acid. The PW acid catalyst showed higher activity under the optimized reaction conditions compared with the conventional homogeneous catalyst sulfuric acid, and it can easily be separated from the products by distillation of the excess methanol and can be reused several times [45].

Biodiesel production in the presence of 20 wt% myristic acid from soybean oil was carried out over heteropolyacid immobilized on mesoporous Ta2O5 materials. Different catalysts were prepared. The network structures of the hybrid materials and the functions of the incorporated alkyl groups on the catalytic activity of the materials have been put forward [46].

Biodiesel synthesis from waste cooking oil was carried out over 12-tungstosilicic acid on SBA-15. The heteropolyacid was prepared using impregnation method. The effect of different reaction parameters like percentage loading, catalyst amount, mole ratio, time and temperature were studied for the supreme conversion. The catalyst was recycled up to four times after simple workup without notable loss in the activity [48].

In this work, we report the transesterification of waste cooking oil with different alcohols over heteropolyacids immobilized on SBA-15. We also studied the effect of free fatty acid addition into waste cooking oil. A kinetic model is proposed.

Advertisement

2. Waste cooking oil raw converted to biofuel

The McDonald's Corporation is the world's largest chain of fast-food restaurants, serving around 68 million customers daily in 119 countries across 35,000 outlets [49]. The waste cooking oil in about 90% of McDonald’s restaurants is used for after-market uses, including biodiesel. The used cooking oil from the restaurants is collected, recycled into biodiesel, and put back into distributors’ trucks to fuel their deliveries to McDonald’s restaurants. In a few countries, McDonald’s waste cooking oil are a closed-loop system. Since 2013, McDonald's developmental licensee in the United Arab Emirates operates a 100% closed-loop biodiesel system. More than 25,000 liters of waste cooking oil are being collected from McDonald’s approximately 108 restaurants each month and converted into biodiesel. Waste cooking oil is stored at the restaurants in a specially designed container. 1.75 million liters of biodiesel have been generated from waste cooking oil and used to power the distribution fleet. Excess biodiesel can be sold, helping other companies reduce their transportation impacts. In the United Kingdom (in 2013), McDonald’s replaced more than 3 million liters of diesel fuel with biofuel made from used cooking oil collected at approximately 1,200 restaurants. Across markets in Europe, the waste cooking oil from more than 80% of the McDonald’s Europe restaurants is being converted into biodiesel. In Brazil 19 restaurants are participating in a closed-loop system that has recycled more than 187,000 liters of used cooking oil into 65,000 liters of biodiesel since 2009, for use in five trucks that deliver to the restaurants [50].

Advertisement

3. Experimental

3.1. Catalyst preparation

Mesoporous silica SBA-15 was prepared according to the literature [51] An ethylene oxide (EO)/propylene oxide (PO) triblock copolymer (P123) with composition EO20PO70EO20 and with an average molecular weight of 5800 was used as the template. The synthesis consisted of 2.0 g of triblock P123 dissolved in 60 cm3 of 2 mol.dm-3 aqueous HCl and 15 cm3 of distilled water under stirring. Then, 4.4 g of tetraethyl orthosilicate (TEOS) was added dropwise to the solution at room temperature. The mixture was stirred for 24 h at 313 K, and then the temperature was raised to 373 K and kept at 373 K for another 24 h in a Teflon-lined autoclave. Finally, the resulting precipitate was filtered, washed carefully with distilled water, air-dried, and calcined at 773 K in air for 5 h to remove the template and to obtain the final product SBA-15.

The heteropolyacid was immobilized on SBA-15 by the grafting technique. The grafting of tungstophosphoric acid (PW) was carried out by mixing 2.0 g dried SBA-15 with 3-aminopropyltriethoxysilane (23.4 μL) containing freshly distilled toluene refluxing for 48 h. The obtained solid material was immersed in an aqueous solution of PW with stirring for 5 h. The solid was then dried in vacuum to obtain the heteropolyacid anchored mesoporous catalyst [52].

3.2. Catalyst characterization

The textural characterization of the catalysts was based on the nitrogen adsorption isotherm, determined at 77 K with a Micromeritics ASAP 2010 apparatus.

The FTIR spectra were recorded with a Bio-Rad FTS 155 instrument.

The amount of W in silica catalysts was measured by dissolving the catalyst in H2SO4/HF 1:1 (v/v) and analyzing the obtained solution using inductively coupled plasma analysis (ICP), which was carried out in a Jobin-Yvon ULTIMA instrument.

The X-ray diffraction (XRD) patterns of the heteropolyacid, silica and catalysts were obtained by using a Bruker powder diffractometer with built-in recorder, using Cu Kα radiation, nickel filter, 30 mA and 40 kV in the high voltage source, and scanning angle between 0.7° and 5° of 2θ at a scanning rate of 1°/min.

Transmission electron microscopy (TEM) analyses were performed on a Hitachi S-2400 scanning electron microscope, at a current voltage of 25 kV.

Catalyst acidity was measured by means of potentiometric titration [53].

3.3. Catalytic experiments

The catalytic experiments were carried out in a stirred batch reactor at 60°C. In a typical experiment, the reactor was loaded with 30 mL of methanol and 0.2 g of catalyst. Reactions were started by adding 2.5 ml waste cooking oil (WCO).

Stability tests of the catalyst were carried out by running four consecutive experiments, using the same reaction conditions. Between the catalytic experiments, the catalyst was separated from the reaction mixture by filtration, washed with acetone and methanol and dried at 70° C overnight.

In order to study the reusability, the PW-NH2-SBA-15 catalyst was filtered from the reaction mixture. After this operation, it was soaked in hexane overnight and it was dried overnight.

Undecano was used as the internal standard. Samples were taken periodically and analyzed by GC, using a Hewlett Packard instrument equipped with a 30 m x 0.25 mm HP-5 column.

Advertisement

4. Results and discussion

4.1. Catalyst characterization

SBA-15 and PW-NH2-SBA-15 catalysts showed a typical IV adsorption isotherm with H1 hysteresis loop as defined by IUPAC. It was observed that the specific surface area and total pore volume of SBA-15 decreased with the immobilization of PW on SBA-15 (Table 1).

PW loada
(wt%)
Surface
areab (m2/g)
VTc
(cm3/g)
SBA-15 - 1050 1.38
PW-NH2-SBA-15 5.6 735 0.62

Table 1.

Physicochemical characterization of SBA-15 and PW-NH2-SBA-15

(a)HPA load determined by ICP analysis


(b) BET


(c) (p/p°) = 0.98


X-ray diffraction patterns of the SBA-15 and PW-NH2-SBA-15 catalysts are shown in Figure 1. All three reflections are still detectable after PW immobilization, suggesting that hexagonal pore structure of the support is retained.

Figure 1.

X-ray diffractograms of SBA-15 (A) and PW-NH2-SBA-15 (B).

Figure 2 shows the FT-IR spectra of SBA-15 and PW-NH2-SBA-15 samples. For SBA-15 bands were observed at 1084, 812 and 476 cm-1, which are attributed to asymmetric stretching, symmetric stretching and bending modes of Si–O–Si, respectively. For PW-NH2-SBA-15 catalyst the characteristic bands of SBA-15 are present along with the bands at 1083, 982, 890 and 808 cm-1, which are the fingerprint of Keggin structure of HPW. However, some bands typical of the Keggin-type HPA structures are overlapped or partially overlapped with the bands of the SBA-15 matrix framework, probably due to the low loading of PW. Similar results were also observed by Liu et al. [52].

Figure 2.

FT-IR spectra of catalysts: (A) SBA-15 and (B) PW-NH2-SBA-15.

TEM image of PW-NH2-SBA-15 catalyst was carried out. The morphology of the support remains after immobilization of PW on SBA-15 (Fig. 3).

Figure 3.

Transmission electron microscopy (TEM) image of PW-NH2-SBA-15.

Table 2 shows the initial electrode potential (Ei) of the materials. The Ei indicates the maximum acid strength of the surface sites [53]. It can be observed that Ei increased with the amount of PW immobilized in SBA-15, which can be due to the increase in the amount of protons with PW loading of the SBA-15 support.

Sample Ei (mV)
SBA-15 +110
PW-NH2-SBA-15 +408

Table 2.

Initial electrode potential (mV) of materials

4.2. Catalytic experiments

The biodiesel production from WCO was carried out over PW-NH2-SBA-15 catalyst at 60°C.

Different catalytic experiments were carried out at different stirring speeds to study the influence of external resistances to mass transfer. It was observed that experiments carried out with 700 rpm have got a good mix of the compounds and eliminate possible mass transfer problems.

4.3. Effect of the nature of alcohol

The alcohols most frequently used in biodiesel production are methanol and ethanol. Due to the low cost of methanol, this alcohol is the first choice for the esterification/transesterification reactions of WCO. However, for biodiesel production to be more environment friendly, ethanol is the ideal candidate for the synthesis of a fully biogenerated biodiesel, since ethanol is derived from agricultural products (renewable sources) [3].

Figure 4 shows the effect of the nature of alcohol (methanol, ethanol and 1-propanol) on the fatty acid ester concentrations obtained over PW-NH2-SBA-15 catalyst, at 60°C. It was observed that the ester concentrations obtained with 1-propanol led to lower conversion, when compared with methanol and ethanol. Similar results were also observed by Sreeprasanth et al. [54], on the transesterification of rubber seed with ethanol over Fe-Zn-1 catalyst. This behavior can be explained not due to the reaction rate with ethanol is slower than with methanol, as ethyl nucleophile is less reactive than methyl nucleophile.

Figure 4.

Biodiesel production from WCO with methanol over PW-NH2-SBA-15 catalyst. Effect of alcohol nature. FAME concentration (mol.dm-3) versus time (h): (○) Methanol; (▲) Ethanol; (□) 1-Propanol.

4.4. Effect of the initial amount of free fatty acid

In order to study the influence of the initial free fatty acid amount in the WCO, three different amounts of fatty acid (as a model) was added to the reaction mixture. The catalytic experiments were carried out at different amounts of initial palmitic acid in the WCO, over PW-NH2-SBA-15 catalyst, while the initial concentration of WCO (0.088mol.dm-3) and the catalysts loading (m=0.2 g) were kept constant. Figure 5 shows the influence of the initial amount of FFA in WCO on the biodiesel production. It was observed that the initial reaction rate increases with the amount of FFA. When the initial amount of FFA increases, a slight increase on the conversion was observed. Similar results were also observed by Marchetti and Errazu [16]. Therefore, this effect could also be seen on the total FAME production since the final amount of biofuel will be produced from the triglycerides as well as from the fatty acids present in the reaction mixture.

Figure 5.

Biodiesel production from WCO with methanol over PW-NH2-SBA-15 catalyst. Effect of the amount of palmitic acid. FAME concentration (mol.dm-3) versus time (h): (○) 0%; (□) 4%; (▲) 12 % ; (×) 27 %.

4.5. Catalyst stability

In order to study the catalytic stability of the PW-NH2-SBA-15, different batch runs with the same catalyst sample and under the same conditions were carried out. Figure 6 shows the catalytic activity of PW-NH2-SBA-15 at different batch runs. It was observed that the catalytic activity decreases only about 10%, after the fourth use.

Figure 6.

Catalytic stability of PW-NH2-SBA-15 catalyst in biodiesel production with methanol. The initial activities are taken as the maximum observed reaction rate, which was calculated from the maximum slope of the methyl ester kinetic curve.

4.6. Kinetics modeling

A simple kinetic model can be established based on the following assumptions:

  1. Isothermal and isobaric reaction conditions;

  2. First order kinetics with respect to the reactants is assumed. The forward and reverse reactions follow second order overall kinetics;

  3. Due to the excess of methanol used, the reverse reaction could be minimized and it was not considered in the reaction rate.

  4. Triglycerides are consumed according to the consecutive reaction network:

T+Ak1D+E
D+Ak2M+E
M+Ak3G+E

where T represents triglycerides, D represents diglycerides, M represents monoglycerides, A represents alcohol, G represents glycerol and E represents esters of fatty acids;

The reaction rate of these three pseudo elementary reaction are expressed as:

r1=k1CT.CAE1
r2=k2CD.CAE2
r3=k3CM.CAE3

For batch reactor the mole balance equations may be written as

dCTdt=WV(r1)E4
dCAdt=WV(r1+r2+r3)E5
dCDdt=WV.(r1r2)E6
dCMdt=WV.(r2r3)E7
dCGdt=WV(r3)E8
dCEdt=WV(r1+r2+r3)E9

The differential equations system was integrated using the Euler Method. The optimization was carried out by the SOLVER routine in a Microsoft Excel spreadsheet.

Figure 7 shows the concentration of fatty acid methyl ester (FAME) versus time (h) on the transesterification of WCO with methanol. The solid line represents the model fitted to the data points. It was observed that the kinetic model fits experimental concentration data quite well. The model parameters, k1, k2 and k3, have got the value of 0.00979, 0.01348 and 0.01956 dm6.mol-1.h-1.gcat-1, respectively. It was observed that k1<k2<k3, which can be explained due to the molecular size of monoglycerides, diglycerides and triglycerides and due to the textural characteristics of PW-NH2-SBA-15. The size of monoglycerides is smaller than that of diglycerides and triglycerides. Consequently, it is expected that, near the active sites of catalyst, the amount of monoglycerides is higher than the amount of diglycerides and triglycerides. As the reaction rate is dependent on reactant concentration, a high concentration of monoglycerides leads to high reaction rates.

Figure 7.

Concentration of FAME (mol.dm-3) versus time (h).

Advertisement

5. Conclusions

Biodiesel production from WCO with methanol was carried out over tungstophosphoric acid immobilized on SBA-15 by grafting technique, at 60°C. After PW immobilization, the morphology of the support remained.

In order to optimize the conditions, the influence of various reaction parameters, the nature of alcohol and the amount of initial free fatty acid on the transesterification of WCO in the presence of PW-NH2-SBA-15 catalyst were carried out. The catalytic stability of the material was also studied.

The esterification/transesterification of WCO with propanol and ethanol led to lower concentration of fatty acid esters than with methanol.

When different amounts of free fatty acids (palmitic acid) were added to the WCO, a slight increase on the concentration of FAME was observed.

In order to study the catalytic stability of PW-NH2-SBA-15 catalyst, four consecutive batch runs were carried out. It was observed that the catalytic activity of PW-NH2-SBA-15 tends to stabilize.

Advertisement

Acknowledgments

This work was carried out with support from the Project PTDC/CTM-POL/114579/2009.

References

  1. 1. Ma F., Hanna M.A. Biodiesel production: a review. Bioresource Technology 1999; 70(1) 1–15.
  2. 2. Szybist J.P., Song J., Alam M., Boehman A.L. Biodiesel combustion, emissions and emission control. Fuel Processing Technology 2007; 88 (7) 679–691.
  3. 3. Lotero J. E., Liu Y., Lopez D. E., Suwannakarn K., Bruce D. A., Goodwin J. G. Synthesis of biodiesel via acid catalysis. Industrial and Engineering Chemical Research 2005; 44(14) 5353–5363.
  4. 4. Jacobson K., Gopinath R., Meher L. C., Dalai A. K. Solid acid catalyzed biodiesel production from waste cooking oil. Applied Catalysis B: Environmental 2008; 85(1-2) 86–91.
  5. 5. Hass M.J, McAloon A.J., Yee W.C., Foglia T.A. A process model to estimate biodiesel production cost. Bioresource Technology 2006; 97 (4) 671–678.
  6. 6. Echim C., Verhe R., De Greyt W., Stevens C. Production of biodiesel from side-stream refining products. Energy & Environmental Science 2009; 2 (11)1131–1141.
  7. 7. Atadashi I.M., Aroua M.K., Abdul Aziz A.R., Sulaiman N.M.N. Production of biodiesel using high free fatty acid feedstocks. Renewable and Sustainable Energy Reviews 2012; 16 (5) 3275– 3285.
  8. 8. Kulkarni M.G., Dalai A.K. Waste cooking oil-An economical source for biodiesel:A review. Industrial and Engineering Chemical Research 2006; 45 (9) 2901–2913.
  9. 9. Phan A.N., Phan T.M. Biodiesel production from waste cooking oils. Fuel 2008; 87 (17-18) 3490–3496.
  10. 10. Talukder M.M.R., Wu J.C., Lau S.K., Cui L. C., Shimin G., Lim A. Energy Fuels 2009; 23 (1) 1-4.
  11. 11. Shi W., He B., Li J. Esterification of acidified oil with methanol by SPES/PES catalytic membrane. Bioresource Technology 2011; 102 (9) 5389–5393.
  12. 12. Park J.-Y., Wang Z.-M., Kim D.-K., Lee J.-S. Effects of water on the esterification of free fatty acids by acid catalysts, Renewable Energy 2010; 35 (3) 614–618.
  13. 13. Tesser R., Casale L., Verde D., Di Serio M., Santacesaria E. Kinetics and modeling of fatty acids esterification on acid exchange resins. Chemical Engineering Journal 2010; 157 (2-3) 539-550.
  14. 14. Tesser R., di Serio M., Guida M., Nastasi M., Santacesaria E. Kinetics of Oleic Acid Esterification with Methanol in the Presence of Triglycerides. Industrial and Engineering Chemical Research 2005;44 (21) 7978-7982.
  15. 15. Chung K.-H., Chang D.-R., Park B.-G. Removal of free fatty acid in waste frying oil by esterification with methanol on zeolite catalysts. Bioresource Technology 2008; 99 (16) 7438–7443.
  16. 16. Marchetti J. M., Errazu A. F. Comparison of different heterogeneous catalysts and different alcohols for the esterification reaction of oleic acid. Fuel 2008; 87(15-16) 3477-3480.
  17. 17. Chung K.-H., Park B.-G. Esterification of oleic acid in soybean oil on zeolite catalysts with different acidity. Journal of Industrial and Engineering Chemistry 2009;15 (3) 388–392
  18. 18. Li Y., Zhang X., Sun L. Fatty acid methyl esters from soapstocks with potential use as biodiesel. Energy Conversion and Management 2010; 51 (11) 2307–2311.
  19. 19. Guo F., Xiu Z., Liang Z. Synthesis of biodiesel from acidified soybean soapstock using a lignin-derived carbonaceous catalyst. Applied Energy 2012; 98 (1) 47–52.
  20. 20. Kozhevnikov I.V. Catalysis by heteropoly acids and multicomponent polyoxometalates in liquid-phase reactions. Chemical Reviews 1998; 98(1) 171–198.
  21. 21. Kozhevnikov I.V. Heteropoly acids and related compounds as catalysts for fine chemical synthesis. Catalysis Reviews Science and Engineering 1995; 37 (2) 311–352.
  22. 22. Okuhara T., Mizuno N., Misono M. Catalytic chemistry of heteropoly compounds. Adv. Catal. 1996; 41, 113−252.
  23. 23. Dupont P., Védrine J.C., Paumard E., Hecquet G., Lefebvre F. Heteropolyacids supported on activated carbon as catalysts for the esterification of acrylic acid by butanol. Applied Catalysis A: General 1995;129 (2) 217-227.
  24. 24. Pizzio L.R., Cáceres C.V., Blanco M.N. Adsorption of tungstophosphoric or tungstosilicic acids from ethanol–water solutions on carbon original. Journal of Colloid and Interface Science 1997; 190 (2) 318-326.
  25. 25. Obali Z., Dog˘u T. Activated carbon–tungstophosphoric acid catalysts for the synthesis of tert-amyl ethyl ether (TAEE). Chemical Engineering Journal 2008; 138 (1–3) 548–555.
  26. 26. Ferreira P., Fonseca I.M., Ramos A.M., Vital J., Castanheiro J.E. Acetylation of glycerol over heteropolyacids supported on activated carbon. Catalysis Communications 2011, 12 (7) 573–576.
  27. 27. Chimienti M.E., Pizzio L.R., Cáceres C.V., Blanco M.N., Tungstophosphoric and tungstosilicic acids on carbon as acidic catalysts. Applied Catalysis A: General 2001;208 (1–2) 7–19.
  28. 28. Caetano C.S., Fonseca I.M., Ramos A.M., Vital J., Castanheiro J.E. Esterification of free fatty acids with methanol using heteropolyacids immobilized on silica. Catalysis Communications 2008; 9 (10) 1996–1999.
  29. 29. Schwegler M.A., Vinke P., van der Eijk M., van Bekkum H. Activated carbon as a support for heteropolyanion catalysts. Applied Catalysis A: General 1992; 80 (1) 41–57.
  30. 30. Dias A.S., Pillinger M., Valente A.A., Mesoporous silica-supported 12-tungstophosphoric acid catalysts for the liquid phase dehydration of d-xylose. Microporous and Mesoporous Materials 2006;94 (1–3) 214–225.
  31. 31. Caiado M., Machado A., Santos R.N., Matos I., Fonseca I.M., Ramos A.M., Vital J., Valente A.A., Castanheiro J.E. Alkoxylation of camphene over silica-occluded tungstophosphoric acid. Applied Catalysis A: General 2013;451 (1) 36–42.
  32. 32. Kasztelan S., Payen E., Moffat J.B. The existence and stability of the silica-supported 12-molybdophosphoric acid keggin unit as shown by Raman, XPS, and 31P NMR spectroscopic studies. Journal of Catalysis 1990; 125 (1) 45–53.
  33. 33. Haber J., Pamin K., Matachowski L., Much D. Catalytic performance of the dodecatungstophosphoric acid on different supports. Applied Catalysis A: General 2003;256 (1) 141–152.
  34. 34. Sulikowski B., Haber J., Kubacka A., Pamin K., Olejniczak Z., Ptaszynski J. Novel "ship-in-the-bottle" type catalyst: evidence for encapsulation of 12-tungstophosphoric acid in the supercage of synthetic faujasite. Catalysis Letters 1996;39 (1–2) 27–31.
  35. 35. Mukai S.R., Masuda T., Ogino I., Hashimoto K. Preparation of encaged heteropoly acid catalyst by synthesizing 12-molybdophosphoric acid in the supercages of Y-type zeolite. Applied Catalysis A: General 1997; 165 (1–2) 219–226.
  36. 36. Mukai S.R., Lin L., Masuda T., Hashimoto K. Key factors for the encapsulation of Keggin-type heteropoly acids in the supercages of Y-type zeolite. Chemical Engineering Science 2001; 56 (3) 799–804.
  37. 37. Tran M.H., Ohkita H., Mizushima T., Kakuta N. Hydrothermal synthesis of molybdenum oxide catalyst: Heteropoly acids encaged in US-Y. Applied Catalysis A: General 2005;287 (1) 129–134.
  38. 38. Ferreira P., Fonseca I.M., Ramos A.M., Vital J., Castanheiro J.E. Esterification of glycerol with acetic acid over dodecamolybdophosphoric acid encaged in USY zeolite. Catalysis Communications 2009;10 (5) 481–484.
  39. 39. Choi J.S., Song I.K., Lee W.Y. Simulation and experimental study on the polymer membrane reactors for the vapor-phase MTBE (methyl tert-butyl ether) decomposition. Catalysis Today 2000; 56 (1–3) 275–282.
  40. 40. Choi J.S., Song I.K., Lee W.Y. Performance of shell and tube-type membrane reactors equipped with heteropolyacid-polymer composite catalytic membranes. Catalysis Today 2001; 67 (1–3) 237–245.
  41. 41. Song I.K., Lee W.Y., Kim J.-J. Methyl t-butyl ether decomposition in an inert membrane reactor composed of 12-tungstophosphoric acid catalyst and polyphenylene oxide membrane. Applied Catalysis A: General 1993; 96 (1) 53–-63.
  42. 42. Castanheiro J.E., Ramos A.M., Fonseca I., Vital J. The acid-catalysed reaction of α-pinene over molybdophosphoric acid immobilised in dense polymeric membranes. Catalysis Today 2003; 82 (1–4) 187–193.
  43. 43. Castanheiro J.E., Fonseca I.M., Ramos A.M., Oliveira R., Vital J. Hydration of α-pinene over molybdophosphoric acid immobilized in hydrophobically modified PVA membranes. Catalysis Today 2005; 104 (2–4) 296–304.
  44. 44. Tropecêlo A.I., Casimiro M.H., Fonseca I.M., Ramos A.M., Vital J., Castanheiro J.E. Esterification of free fatty acids to biodiesel over heteropolyacids immobilized on mesoporous silica. Applied Catalysis A: General 2010;390(1–2) 183–189.
  45. 45. Cao F., Chen Y., Zhai F., Li J., Wang J., Wang X., Wang S., Zhu W. Biodiesel production from high acid value waste frying oil catalyzed by superacid heteropolyacid. Biotechnology and Bioengineering 2008;101 (1) 93–100.
  46. 46. Xu L., Li W., Hu J., Yang X., Guo Y. Biodiesel production from soybean oil catalyzed by multifunctionalized Ta2O5/SiO2-[H3PW12O40/R] (R = Me or Ph) hybrid catalyst, Applied Catalysis B: Environmental 2009; 90 (3–4) 587–594.
  47. 47. Baig A., Ng T.T., A single-step solid acid-catalyzed process for the production of biodiesel from high free fatty acid feedstocks. Energy Fuel 2010; 24 (9) 4712– 4720.
  48. 48. Nilesh N., Brahmkhatri V., Patel A. Efficient synthesis of biodiesel from waste cooking oil using solid acid catalyst comprising 12-tungstosilicic acid and SBA-15. Fuel 2014;135 (1) 253–261.
  49. 49. http://en.wikipedia.org/wiki/McDonald's (accessed 14 October 2014).
  50. 50. http://www.aboutmcdonalds.com/mcd/sustainability/planet/minimizing-waste/biodiesel.html (accessed 14 October 2014).
  51. 51. Gao L., Wang Y., Wang J., Huang L., Shi L., Fan X., Zou Z., Yu T., Zhu M., Li Z. A novel ZnII-sensitive fluorescent chemosensor assembled within aminopropyl-functionalized mesoporous SBA-15. Inorganic Chemistry 2006; 45(17) 6844–6850.
  52. 52. Liu H., Xue N., Peng L., Guo X., Ding W., Chen Y. The hydrophilic/hydrophobic effect of porous solid acid catalysts on mixed liquid phase reaction of esterification. Catalysis Communications 2009; 10 (13) 1734–1737.
  53. 53. Pizzio L.R., Vásquez P.G., Cáceres C.V., Blanco M.N. Supported Keggin type heteropolycompounds for ecofriendly reactions. Applied Catalysis A: General 2003;256 (1–2) 125–139.
  54. 54. Screeprasanth P.S., Srivastava R., Scrinivas D., Ratnasamy P. Hydrophobic, solid acid catalysts for production of biofuels and lubricants. Applied Catalysis A: General 2006;314(2) 148–159.

Written By

M. Caiado, A. Tropecêlo and J.E. Castanheiro

Submitted: 30 September 2014 Published: 30 September 2015