Sie sind auf Seite 1von 6

Eur. J. Lipid Sci. Technol.

2008, 110, 341346

341

Research Paper Studies on the kinetics of in situ epoxidation of vegetable oils


Chuanshang Cai, Honghai Dai, Rongsheng Chen, Caixia Su, Xiaoya Xu, Shi Zhang, Liting Yang
School of Chemistry and Environment, South China Normal University, Guangzhou, China

In the present work, the kinetics of the epoxidation of soybean oil (SBO) by peroxyacetic acid (PAA) generated in situ in the presence of sulfuric acid as the catalyst was studied at various temperatures (45, 65 and 75 7C). It was found that epoxidation with almost complete conversion of unsaturated carbon and negligible oxirane cleavage can be attained by the in situ technique. The rate constant for epoxidation of SBO was found to be of the order of 106 mol1s1 and the activation energy of epoxidation is 43.11 kJ/mol. Some thermodynamic parameters: enthalpy, entropy and free activation energy of 40.63 kJ/mol, 208.80 J/mol and 102.88 kJ/mol, respectively, were obtained for the epoxidation of SBO. The kinetic and thermodynamic parameters of epoxidation obtained from this study indicate that an increase in the process temperature would increase the rate of epoxide formation. The epoxidation of corn oil and sunflower oil were also investigated under the same conditions. The results show that the reaction rate is in the order of soybean oil . corn oil . sunflower oil.
Keywords: Epoxidation / Kinetics / Peroxyacetic acid / Vegetable oil

Received: April 12, 2007; accepted: November 22, 2007 DOI 10.1002/ejlt.200700104

Supporting Information for this article is available from the author or on the WWW under http://www.wiley-vch.de/contents/jc_2114/2008/200700104_s.pdf

1 Introduction
Epoxides of all kinds of plant oils are well known commercially because of the many important reactions they undergo. Epoxidation of long-chain olefins and of unsaturated fatty acid derivatives such as soybean oil (SBO) and other plant oils is carried out on the industrial scale [1]. Today, one of the most important epoxidized vegetable oils is epoxidized soybean oil (ESO), and its worldwide production is about 200,000 t/year [2]. Fatty epoxides are used directly as plasticizers which are compatible with polyvinyl chloride (PVC), and as stabilizers for PVC resins to improve flexibility, elasticity, and toughness and to impart the stability of polymers towards heat and UV radiation. Due to high reactivity of the oxirane ring, epoxides also act as a raw materials for a variety of chemicals, such as alcohols, glycols, alkanolamines, carbonyl compounds, olefinic compounds and polymers like polyesters, polyurethanes (PU), and epoxy resins.

Correspondence: Liting Yang, School of Chemistry and Environment, South China Normal University, Guangzhou 510631, China. E-mail: yanglt63@yahoo.com.cn, yanglt@scnu.edu.cn Fax: +86 20 39310187

Especially, the oil was characterized for its hydroxyl values and fatty acid composition. The modified oils had higher hydroxyl values and lower unsaturated acids than regular unmodified oils. The modified soy-based vegetable oil polyols could be incorporated as a replacement for conventional polyols, reacting with isocyanates (TDI and MDI), to produce flexible slabstock PU foams, elastomers, and coatings. SBO is highly hydrophobic; so an excellent weather stability of the derived PU can be expected; then, the thermal and oxidative stabilities of the SBO-based PU are comparable with those of the polypropyleneoxide-based PU [36]. As energy demands increase and fossil fuel reserves are limited, there has been a growing interest in the utilization of renewable resources as an alternative to petroleum-based polymers. Consequently, much attention has been focused on the development of polymeric materials from vegetable oils, a sustainable resource [7, 8]. Vegetable oil, which is readily available and is a comparatively inexpensive material, can be used to synthesize various types of polymers. There are four known technologies to produce epoxides from olefinic types of molecules: (a) epoxidation with percarboxylic acids [9], the most widely used in the industry, which can be catalyzed by acids or by enzymes [1, 10]; (b) epoxidation with organic and inorganic peroxides, which includes
www.ejlst.com

2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

342

C. Cai et al.

Eur. J. Lipid Sci. Technol. 2008, 110, 341346

alkaline and nitrile hydrogen peroxide epoxidation as well as transition metal-catalyzed epoxidation [11]; (c) epoxidation with halohydrins, using hypohalous acids (HOX) and their salts as reagents for the epoxidation of olefins with electrondeficient double bonds [9]; and (d) epoxidation with molecular oxygen [9]. But of the ways of epoxidation of vegetable oils, the available technologies that need to be explored are only (a) and (b) as described above, which are clean and efficient, namely epoxidation with percarboxylic acids and epoxdation with organic and inorganic peroxides [12]. These technologies can be rendered cleaner by using heterogeneous catalysts replacing traditional homogeneous ones [13]. Although the cleanest epoxidation process is direct insertion of oxygen from air into the olefinic bond in the absence of any catalyst in the mixture, the conversion of a double bond to an oxirane ring under this condition is far from application in the industry. In the present work, we investigate the kinetics of epoxidatin of three oils (SBO, corn oil and sunflower oil) by PAA generated in situ and the relation between the extent of epoxidation and temperature. The kinetic parameters and some thermodynamic parameters, i.e. enthalpy, entropy and free activation energy of the epoxidation of oils, were obtained. Partial epoxidation of oils will give unsaturated epoxidized oils with different oxirane contents, which can generate polyols with different OH values. These kinds of polyols react with isocyanate and can produce series of PU for different applications.

100. The flow rate of the carrier gas was 1.2 mL/min. The gas chromatograph injector temperature was 200 7C. The oven temperature was held at 50 7C for 6 min and then programmed from 50 to 250 7C at 10 7C/min. The oil reference standards, AOCS, from Sigma Chemicals were used as external standards for quantitative analysis.

2.3 Epoxidation of SBO


Epoxidation of SBO was carried out in a 1000-mL fournecked round-bottom flask equipped with a thermometer sensor, a mechanical stirrer, a condenser, and an isobaric funnel. The whole apparatus was kept in a water bath to maintain the respective reaction temperature of 45, 65 and 75 6 1 7C. Thus, the temperature was measured inside the reaction vessel. Of SBO, 150 g (0.172 mol) was placed in the round-bottom flask. With agitation, PAA, prepared in situ by reacting various mixtures of 35 g (0.58 mol) 99.5% glacial acetic acid and 165 g (1.46 mol) 30% H2O2 in the presence of small quantities of concentrated sulfuric acid for about 12 h, was added slowly through an isobaric funnel. This precaution was taken to prevent overheating of the system due to the exothermic nature of epoxidation reactions. The stirring rate was controlled so that the oil in the mixture was finely dispersed. The reaction was monitored by withdrawing aliquots of the reaction mixture at various time intervals. The samples were quenched inside the refrigerator to stop the epoxidation reaction, and delaminated soon. To purify the samples, the samples were poured into a separator funnel. The aqueous layer was drawn off and the oil layer was washed successively with warm dilute sodium carbonate solution until the pH was neutral, and then washed with saturated sodium chloride solution and distilled water, respectively. The solvent was removed using a rotary evaporator, with the help of a water vacuum pump. The oil phase was further dried above anhydrous magnesium sulfate and then filtered. The epoxy oxygen content was measured according to a standard procedure for oils and fats [15]. The epoxidation reactions of corn oil and sunflower oil were carried out according to the same procedure as describe above. The infrared spectra were recorded on a Perkin-Elmer Spectrum-1000 FT -IR spectrometer. Samples were prepared as thin liquid films on a KBr or NaCl salt plate.

2 Materials and methods


2.1 Materials
The refined SBO, corn oil and sunflower oil were kindly provided by Nanhai Oil Co. Ltd. Glacial acetic acid (99.5%), sodium carbonate, sodium hydroxide, hydrogen peroxide (30%) and sulfuric acid were of analytical reagent grade and purchased from Guangzhou Chemical Reagent Co. They were used as supplied.

2.2 Determination of fatty acid composition


The fatty acid composition of the oils was determined by gas chromatography after methanolysis according to reference [14]. A sample (0.1 g) was placed in a 50-mL reaction bottle with a reflux condenser, and 0.75 g calcium methoxide dispersed in methanol (20 mL) was added. The bottle was heated at 70 7C for 30 min. The products of the transesterification reaction were cooled and then analyzed using a Hewlett Packard 6890 gas chromatograph equipped with a mass selective detector HP 5973. The capillary column (30 m long with 0.25 mm inside diameter) had an INNOWAX phase. Helium was used as the carrier gas with a split ratio of
2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

3 Results and discussion


The main compositions of SBO, corn oil and sunflower oil were analyzed by gas chromatography after methanolysis according to [14] and the results are shown in Table 1. As seen from Table 1, the saturated acid contents of SBO, corn oil and sunflower oil are 13.5, 12.9 and 12.7, respectively, and the unsaturated acid contents are 86.1, 87.1 and 88.9, respecwww.ejlst.com

Eur. J. Lipid Sci. Technol. 2008, 110, 341346

Kinetics of in situ epoxidation of vegetable oils

343

Table 1. Main fatty acid composition of oils.


Fatty acids [wt-%] Palmitic (16:0) Soybean oil Corn oil Sunflower oil 10.5 11.6 9.2 Stearic (18:0) 3.0 1.3 3.5 Oleic (18:1) 23.1 30.6 20.4 Linoleic (18:2) 56.5 55.8 68.1 Linolenic (18:3) 6.5 0.7 0.4

tively. But the contents of oleic, linoleic and linolenic acids in the three oils are very different.

3.1 Oxidation of vegetable oils


The oxidation of oils was carried out according to the Materials and methods section at 75 7C and the results are shown in Fig. 1. As can be seen from Fig. 1, the reaction rate of SBO is the fastest and gives the highest oxirane ring content. This may be due to the different acid contents of oleic, linoleic and linolenic in the oil. As is seen from Table 1, the linolenic acid content in SBO is 6.5; linolenic acid has three double bonds in the chain, thus having higher reactivity. The extent of epoxidation of SBO by PAA generated in situ at various temperatures is shown in Fig. 2. The results indicated that increasing the temperature shows a favorable effect on the extent of SBO by PAA generated in situ. The initial increase in the extent of epoxidation of SBO with reaction time reached maximum values (for epoxidation at 75 7C), and then decreased with further increases in reaction time. The decreased extent of epoxidation at 75 7C was very obvious, while the extent of epoxidation at 65 7C hardly

decreased during the initial 8 h. The reaction at lower temperature showed a lower rate, but gave a more stable oxirane ring (for epoxidation at 45 7C). These results suggested that the optimum level of epoxidation could be obtained in a shorter time at moderate reaction temperatures ranging from 65 to 75 7C, at which the rate of epoxide degradation was relatively low. Consequently, only the initial stage of the reaction was taken into account for the kinetic study. The extents of epoxidation of corn oil and sunflower oil by PAA generated in situ at various temperatures are shown in Figs. A-1 and A-2 (see Supporting Information on the WWW). SBO and the purified sample spread on an NaCl plate were analyzed using FT -IR to check whether SBO was converted to ESO. The FT -IR spectra of SBO and ESO are described in Fig. 3. In the FT-IR spectrum of SBO, the characteristic peak at 3009 cm1 was attributed to the CH stretching of the SBO C=CH. The peak at 3009 cm1 disappeared after the epoxidation reaction. As Fig. 3 shows, obviously, the peak at 3009 cm1 had disappeared almost completely at 65 and 75 7C after 6 h, indicating that almost all the C=C had taken part in the epoxidation reaction. The C=O double bonds at 1725 cm1 (C=O stretching vibrations) remained unaltered after the epoxidation reaction. The presence of new peaks in the FT -IR spectrum of ESO at 823 and 833 cm1, attributed to the epoxy group, corroborated the conclusion relative to the success of the epoxidation reaction of SBO. The other new peak at 3463 cm1 was attributed to the hydroxyl OH stretching, indicating that the epoxy group has been opened; then, the intensity of the 3463 cm1 band indicated the extent of hydroxyl of ESO. Consequently, this obviously shows that the reaction temperature was higher, and the epoxy group had been opened partially.

Figure 1. Oxirane content of three oils versus reaction time at 75 7C.

2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.ejlst.com

344

C. Cai et al.

Eur. J. Lipid Sci. Technol. 2008, 110, 341346

Figure 2. Extent of epoxidation of SBO at different temperatures.

Figure 4. Kinetics of in situ epoxidation of SBO by PAA.

where the subscript 0 denotes the initial concentrations and EP denotes the epoxides. Hence
LnfH 2 O2 0 EPg kRCOOH0 t LnH 2 O2 0 (2)

Figure 3. FT-IR spectra: Spectra of SBO and ESO prepared at various temperatures, reacting for 6 h.

3.2 Kinetics of epoxidation


The in situ epoxidation reaction generally takes place in two steps: (i) formation of a peroxy acid and (ii) reaction of the peroxy acid with the unsaturation [16]. (i) Formation of a peroxy acid:
RCOOH H 2 O2 , RCOOOH H 2 O

According to Eq. (2), the plot of {[H2O2]0 [EP]} vs. time should yield straight lines for those reactions with negligible degradation of oxirane. Deviations from linearity were observed for the reaction with substantial ring opening. In such cases, rate constants were obtained with the initial slopes. Figure 4 shows the kinetic plots for in situ epoxidation of SBO at different temperatures, and Figs. A-3 and A-4 show the kinetic plots for in situ epoxidation of corn oil and sunflower oil at different temperatures, respectively. The rate constants obtained for the reaction were of the order 106 mol1s1 (Table 2). The values of the rate constant showed the expected dependence on temperature, as an increase in the reaction temperature from 45 to 75 7C was accompanied by an about fourfold increase in the value of the rate constant of epoxidation. Because the epoxidation reagent (PAA) had been formed and mixed with SBO to prepare ESO, the rate of the epoxidation of SBO by PAA was very fast. Consequently, the rate constant of epoxidation was bigger than the one in references [13, 18].

(ii) Epoxidation
RCOOOH 1 R1CH=CHR2 ? R1CHOCHR2 1 RCOOH

Table 2. Rate constants of SBO epoxidation at various temperatures.


Temperature [7C] Rate constant of epoxidation K6106 [mol1s1] 7.4 11.4 18.0 29.6

If the first step is considered to be rate determining and the concentration of the peroxy acid is assumed to be constant throughout the reaction, then the rate of epoxidation will be given by the following expression [17]:
dEP kfH 2 O2 0 EPg RCOOH0 dt
2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

45 55 65 75

(1)
www.ejlst.com

Eur. J. Lipid Sci. Technol. 2008, 110, 341346

Kinetics of in situ epoxidation of vegetable oils

345

where k is the rate constant; R the gas constant, T is the absolute temperature, N the Avogadro constant, and h is Plancks constant.
DF DH T DS (5)

The average values of the thermodynamic parameters were calculated using the equations above and the results are shown in Table 3.

4 Conclusions
The results from this study show that the epoxidation of SBO by PAA generated in situ can be carried out at moderate temperatures with minimum epoxide degradation. The kinetic and thermodynamic parameters of the epoxidation obtained indicate that an increase in the process temperature would increase the rate of epoxide formation and would be useful for the scaled-up production of ESO using the in situ technique. The activation energy of the reaction follows the sequence SBO , corn oil , sunflower oil.

Figure 5. Activation energy, Ea, for the epoxidation of SBO by PAA.

The kinetic data for the in situ epoxidation of SBO by PAA gave a good Arrhenius plot (Fig. 5), from which the activation energy, E, of epoxidation was determined to be 43.11 kJ/mol. Figures A-5 and A-6 also show the Arrhenius plots of corn oil and sunflower oil. The activation energy values for in situ epoxidation of corn oil and sunflower oil were obtained according to the same process as above, and the results are shown in Table 3. As can be seen from Table 3, the activation energy of the oils follows the sequence ESoybean oil , ECorn oil , ESunflower oil, and this result may be due to the different fatty acid compositions of the oils. A higher linolenic acid content gives a lower activation energy.

Acknowledgment
This work was supported by the Natural Science Foundation of Guangdong Province (No. 06025028).

Conflict of interest statement


3.3 Thermodynamics of epoxidation of vegetable oils
The enthalpy of activation, DH, was calculated using the following equation [19]:
DH E RT (3)

The authors have declared no conflict of interest.

References
[1] L. A. Rios, P. Weckes, H. Schuster, W. F. Hoelderich: Mesoporous and amorphous Ti-silicas on the epoxidation of vegetables oils. J Catal. 2005, 232, 1926. [2] M. R. Klaas, S. Warwel: Complete and partial epoxidation of plant oils by lipase-catalysed perhydrolysis. Ind Crops Prod. 1999, 9, 125132. [3] A. Guo, I. Javni, Z. S. Petrovic: Rigid polyurethane foams based on soybean oil. J Appl Polym Sci. 2000, 77, 467473. [4] A. Guo, Y. Cho, Z. S. Petrovic: Structure and properties of halogenated and non halogenated soy based polyols. J Polym Sci A Polym Chem. 2000, 38, 39003910. [5] Z. S. Petrovic, A. Guo, W. Zhuang: Structure and properties of polyurethanes based on halogenated and nonhalogenated soy based polyols. J Polym Sci A Polym Chem. 2000, 38, 4062 4069. [6] I. Javni, Z. S. Petrovic, A. Guo, R. Fuller: Thermal stability of polyurethanes based on vegetable oils. J Appl Polym Sci. 2000, 77, 17231734.
www.ejlst.com

The average energy of activation, DS, and the free energy of activation, DF, were obtained using the relationship [20]:
Table 3. Kinetic parameters for in situ epoxidation of vegetable oils.
E [kJ/mol] Soybean oil Corn oil Sunflower oil 43.11 6 0.33 74.22 6 0.50 85.21 6 0.42 R2 DH [kJ/mol] DS [J/mol] DF [kJ/mol]

0.9987 40.63 0.9980 71.74 0.9986 82.73

208.80 102.88 120.01 110.00 83.17 110.01

RT DS=R E=RT e e Nh

(4)

2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

346

C. Cai et al.

Eur. J. Lipid Sci. Technol. 2008, 110, 341346

[7] S. Ahmad, F. Naqvi, K. L. Verma, S. Yadav: Studies on a newly developed linseed oil-based alumina-filled polyesteramide anticorrosive coating. J Appl Polym Sci. 1999, 72, 16791687. [8] S. Ahmad, S. M. Ashraf, F. Naqvi, A. Hasnat, S. Yadav: A polyesteramide from Pongamia glabra oil for biologically safe anticorrosive coating. Prog Org Coat. 2003, 47, 95102. [9] S. Guenter, R. Rieth, K. T. Rowbottom (Eds.): Sixth Ullmanns Encyclopedia of Industrial Chemistry. Vol. 12. WileyVCH Verlag, Weinheim (Germany) 2003, pp. 269284. [10] S. Warwel, M. R. Klaas: Chemo-enzymatic epoxidation of unsaturated carboxylic acids. J Mol Cat B Enzym. 1995, 1, 2935. [11] K. B. Sharpless, S. S. Woodard, M. G. Finn: On the mechanism of titanium-tartrate catalyzed asymmetric epoxidation. Pure Appl Chem. 1983, 55, 18231836. [12] V. V. Goud, A. V. Patwardhan, N. C. Pradhan: Studies on the epoxidation of mahua oil (Madhumica indica) by hydrogen peroxide. Bioresour Technol. 2006, 97, 13651371. [13] A. Campanella, M. A. Baltanas, M. C. Capel-Sanchez, J. M. Campos-Martin, J. L. G. Fierro: Soybean oil epoxidation with hydrogen peroxide using an amorphous Ti/SiO2 catalyst. Green Chem. 2004, 6, 330334.

[14] S. Gryglewicz, K. Grabas, G. Gryglewicz: Use of vegetable oils and fatty acid methyl esters in the production of spherical activated carbons. Bioresour Technol. 2000, 75, 213218. [15] M. He, X. Wang, J. Liu: Epoxidation of vegetable oils catalyzed by heteropoly acids. Chinese J Appl Chem. 1998, 15, 117118. [16] B. Rangaraja, A. Harvey, E. A. Grulke, P. D. Culman: Kinetic parameters of a two-phase model for in situ epoxidation of soybean oil. J Am Oil Chem Soc. 1995, 72, 11611169. [17] L. H. Gan, S. H. Goh, K. S. Ooi: Kinetics studies of epoxidation and oxirane cleavage of palm olein methyl esters. J Am Oil Chem Soc. 1992, 69, 347351. [18] F. E. Okieimen, O. I. Bakare, C. O. Okieimen: Studies on the epoxidation of rubber seed oil. J Ind Crops Prod. 2002, 15, 139144. [19] F. A. Zaher, M. H. El-Mallah, M. A. El-Hefnawy: Kinetics of oxirane cleavage in epoxidised soybean oil. J Am Oil Chem Soc. 1989, 66, 698700. [20] A. A. Frost, R. G. Pearson: Kinetics and Mechanism. 2nd Edn. John Wiley, New York, NY (USA) 1961.

2008 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

www.ejlst.com

Das könnte Ihnen auch gefallen