Sie sind auf Seite 1von 14

Reactive & Functional Polymers 61 (2004) 101114

REACTIVE & FUNCTIONAL POLYMERS


www.elsevier.com/locate/react

Kinetics of esterication of lactic acid with ethanol catalyzed by cation-exchange resins


Yang Zhang, Li Ma, Jichu Yang *
Department of Chemical Engineering, Tsinghua University, Beijing 100084, Peoples Republic of China Received 17 March 2004; accepted 26 April 2004 Available online 25 June 2004

Abstract The esterication of lactic acid with ethanol was carried out in the presence of ve dierent cation-exchange resins. The eect of catalyst type, catalyst loading, and temperature on reaction kinetics was evaluated. In order to study which components had the strongest adsorption strength on the resin surface, two simplied mechanisms based on Langmuir Hinselwood model were compared by correlating the experimental data. FTIR method was used to verify the rationality of the mechanism. Nonideality of the liquid phase was taken into account by using activities, which were predicted by UNIFAC method instead of concentrations. The thermal stability and mechanical strength of the resin catalysts were tested by SEM. 2004 Elsevier B.V. All rights reserved.
Keywords: Lactic acid; Esterication; Ethanol; Kinetics; Cation-exchange resin

1. Introduction Ethyl lactate is an important organic ester, which is biodegradable and can be used as food additive, perfumery, avor chemicals and solvent, which can dissolve acetic acid cellulose and many resins [1]. Furthermore, the esterication of lactic acid with ethanol is a step in the purication of lactic acid by reactive distillation [2,3]. The conventional way to produce ethyl lactate is the esterication of lactic acid with ethanol
Corresponding author. Tel.: +86-10-62-788568/785514; fax: +86-10-62-770304. E-mail address: yjc-dce@mail.tsinghua.edu.cn (J. Yang).
*

catalyzed by sulphuric acid. Since this kind of homogeneous catalyst may cause a lot of problems, many heterogeneous solid catalysts were used in the reaction such as ion-exchange resin, clays and clay supported heteropoly acids [4]. Among these catalysts, cation-exchange resin is a perfect substitute which has many advantages such as: (a) corrosion problems could be avoided and it is easier to dispose of the waste liquor from the reaction mixture; (b) continuous operations in columns are possible; (c) the catalyst can be easily removed from the reaction products by decantation or ltration; (d) the purity of the products is higher since side reactions can be eliminated or are less signicant [5,6].

1381-5148/$ - see front matter 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.reactfunctpolym.2004.04.003

102

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

Nomenclature ai ci DA De EA DH Keq k k0 ki0 ki M MRD N nLA;0 R r r0 S SRS activity of i (mol/l) concentration of i at the surface of the catalyst (mol/l) molecular diusion coecient eective diusion coecient apparent activation energy (kJ mol1 ) enthalpy change (kJ mol1 ) reaction equilibrium constant reaction rate constant (mol g1 min1 ) pre-exponential factor (mol g1 min1 ) adsorption coecient at initial temperature adsorption coecient molar mass of the solvent mean relative deviation number of data points initial molar amount of lactic acid (mol) gas constant (J mol1 K1 ) reaction rate (mol g1 min1 ) radius of catalyst particle vacant site on catalyst surface sum of residual squares T t W X VA temperature (K) time (min) dry catalyst weight (g) conversion molar volume at boiling point

Greek symbols c activity coecient e porosity s tortuosity / thiele modulus U0 association factor l viscosity (Pa s) Subscripts calc calculated values exp experimental values LA lactic acid E ethanol EL ethyl lactate W water

The industrial production of esters by esterication of acid and alcohol was rst carried out in a continuous stirred tank reactor (CSTR) and later in a catalytic distillation column over cation-exchange resins. Nowadays, some investigations have focused on the water-permeable membrane reactors applied to liquid-phase reversible reactions [1,7,8]. The conversion of the esterication of lactic acid with ethanol exceeded the equilibrium limit remarkably with the aid of vapor-permeation according to their researches. In order to optimize the design of a CSTR, a reactive distillation column or a membrane reactor, it is necessary to have some information on the reaction kinetics. Although there have been many researches about the esterication of lactic acid with dierent alcohols over cation-exchange resins [49], very few reports concern the synthesis of ethyl lactate with heterogeneous catalysts. In this paper, the emphasis of our work was to use LangmuirHinselwood (LH) model [10] in a study

of the esterication kinetics over dierent cationexchange resins. Two simplied LH mechanisms were compared in order to nd which one described the reaction kinetics better. Fourier transform infrared spectroscopic (FTIR) analysis was also used to verify the rationality of the mechanism. Scanning electron microscopy (SEM) was introduced to evaluate the mechanical strength of the resins and their thermal stability.

2. Materials and catalysts Ethanol (purity >99.7 wt%) was purchased from Yili ne chemical Co., Beijing. Ethyl L -lactate was purchased from SigmaAldrich.L lactic acid (80 wt%) was obtained from PURAC Biochem, Netherlands. The catalysts used in the experiments were commercial strong-acid cationexchange resins. Their physical properties are listed in Table 1. Before use, the fresh catalysts

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114 Table 1 Physical properties of the strong-acid cation-exchange resins Property Shape Size (mm) Internal surface area (m2 /g) Weight capacity (mEq/g) Temperature stability (C) Manufacturer Amberlyst-15 (Mp) Bead 0.5 P 90% 50 4.7 120 Rohm and Haas Co., USA D001 (Mp) Bead 0.321.25 P 95% N/A P 4.35 100 Shandong Dongda Chemical Industry Co., PRC D002 (Mp) Bead 0.41.25 P 95% 3540 P 4.8 120 Jiangyin Organic Chemical Plant, PRC NKC (Mp) Bead 0.41. 25 P 95% 77 P 4.7 100 Chemical Plant of Nankai Univ., PRC 002 (G)

103

Bead 0.51.25 P 95% N/A P 5.0 180 Jiangyin Organic Chemical Plant, PRC

Mp Macroporous; G Gel; N/A Not available.

were dried at 90 C over 12 h to remove moisture completely after being washed with pure water, ethanol, hydrochloric acid and pure water sequentially.

4. GC analysis GC (Shimadzu GC-9AM) was used to analyze the sample, which was equipped with a ame ionization detector (FID). The sample size for GC was 0.2 ll. The injection port and detector temperatures were set to be 240 and 250 C separately. A capillary column (25 m 0.5 mm, SE-30) was used. The column temperature was programmed to rise from an initial value of 100125 C at 2.5 C/min, and then held constant at 125 C for an additional 4 min. High purity helium gas (99.999%) was used as a carrier gas. The ow rate of the carrier gas was 40 ml/min. 5. FTIR and SEM The catalyst samples, which had been used in the kinetics experiments, were tested in FTIR analysis. Before test, they were treated under 100 C with dierent desorption times from 0 to 8 h in an oven, and then sealed in plastic bags. The FTIR analysis was carried out at room temperature using NICOLET 560 equipment. The samples were ground to ne powders using an agate mortar immediately prior to analysis. KBr was used as the embedding medium to make every sample tablets containing around 1.3 mg of ionexchange resin powder [11,12]. A SEM equipment (KYKY-2800, KYKY Technology Development Ltd.) was used to

3. Kinetics experimental apparatus and procedure 3.1. Apparatus The esterication reactor consisted of a threenecked ask of 250 ml capacity tted with a condenser, a sampling device and a thermometer. The temperature was controlled by a thermostating bath, which ensured a temperature constancy of 0.2 C in the reactor. A magnetic stirrer was used to mix the reactants, and the frequency was about 450 rpm. 3.2. Procedure Lactic acid and cation-exchange resins were charged into the reactor, and then heated to the desired temperature. Finally, ethanol was added. This was taken as zero time for a run. The initial molar ratio of lactic and ethanol was 1:3, and the total volume of the reactant was 196.77 ml. About 1 ml of liquid sample was withdrawn from the reactor at regular intervals for gas chromatography (GC) analysis. In a typical run, about 10 samples were taken from the system.

104

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

evaluate the mechanical strength and the thermal stability of the catalysts, and the acceleration voltage is 20 kV.

DA

p 7:4 108 U0 MT VA l
0:6

6. Results and discussion 6.1. External and internal diusion In a heterogeneous catalytic reaction, there are several processes that inuence the rate of reaction, which are external and internal diusion of reactants, adsorption, surface reaction, desorption, internal and external diusion of reaction products. In order to study the intrinsic kinetics at the catalyst surface, external and internal diusion should not be the rate limiting steps [8]. According to Sanzs work [9], if the speed of stirring was between 300 and 700 rpm for the similar cationexchange resins, the inuence of external diusion could be neglected. Therefore, the stirring speed was set around 450 rpm in all further experiments to ensure the absence of external mass transfer resistance. To evaluate the eect of internal diusion on the cation-exchange resins, Eq. (1) was used [13]. /
2 r0 k ; 9D e

In Eq. (3), U0 and M denote the association factor and the molar mass of the solvent. VA is the molar volume at boiling point and l is the viscosity of the solution. For solvent mixtures, the association molar mass can be calculated by p p P U0 M xi U0i Mi . Le Bas method was used in order to calculate VA [15]. With the increase of temperature, the viscosity of the reaction solution will decrease while the diusion coecient will increase. As the result, if the inuence of internal diusion could be neglected at lower temperature, the inuence at higher temperature could also be neglected according to Eq. (1). At 25 C, through experiment, the viscosity of the equilibrium reaction solution was 2.3963 103 Pa s. Then the calculation results of Eqs. (1) and (2) were / 0:02 and De 2:2 106 m2 s1 , which indicated that the eect of internal diusion on the reaction rate could be ignored reasonably. Based on the discussion above, the experimental kinetic results could be considered to reect the intrinsic kinetics of the esterication reaction catalyzed by cationexchange resins. 6.2. Assumption of the model and parameters estimation The heterogeneous reaction has been described with many models such as LangmuirHinselwood, EleyRideal (ER) and pseudo-homogeneous model [6,10]. Among them, LH model was found to be more appropriate for this kind of esterication reaction [9,14,16]. The reaction mechanism can be described as follows: LA S ( ) LA S ES( )ES E S LA S ( ) EL S W S EL S ( ) EL S WS( )WS

where r0 and De denote the radius of catalyst particle and the eective diusion coecient respectively. k is the reaction rate constant, and / is the thiele modulus. If the calculated value of / was smaller than 1, the internal diusion could be neglected. The eective diusion coecient was dened as follows: De DA e ; s 2

where DA is the molecular diusion coecient. s is the particle tortuosity and e is the porosity. For most resin catalysts, the values of e=s are between 0.25 and 0.50 [14]. In the calculation, the value of e=s was taken 0.50. The molecular diusion coefcient for liquid phase diusion can be evaluated from the WilkeChang equation (Eq. (3)) [15].

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

105

Therefore, the LH model could be written as:   nLA;0 dX r dt W aW aLA aE aEL Keq k ; 2 1 kW aW kE aE kLA aLA kEL aEL   c i S aEL aW ki ; Keq ; ai c S aLA aE eq i Water, Ethanol, Lactic acid, Ethyl lactate 4 In the equations, nLA;0 is the initial molar concentration of lactic acid and W is the catalyst loading, k represents the reaction rate constant and ki represents the adsorption coecient. cS and ciS denote the concentration of vacant site on catalyst surface and the concentration of component i at the catalyst surface respectively. ai is the activity for each component and Keq is the reaction equilibrium constant. The activities were calculated by the UNIFAC method [15]. The splitting of the groups is shown in Table 2. The volume and area parameters of the groups and their interaction parameters were taken from the book of Fredenslund et al. [16]. Since the denominator of Eq. (4) is a multicomponent complex, the parameters cannot be regressed correctly from the experimental results. It was assumed that the adsorption of the molecules was competitive on the same active site, and only those that had the strongest adsorption were taken into account in the simplied mechanisms. There were some dierent opinions about which components had the strongest adsorption on the catalyst surface [810,14,17]. Two mechanisms were tested in our work to nd which one was
Table 2 UNIFAC group identication of the components Molecule Group identication Group name Ethyl Lactate CH3 CH3 OHCH CH2 COO H2 O CH3 CH2 OH COOH CH3 OHCH Main 1 5 11 7 5 17 5 Secondary 1 15 26 20 17 40 15

more reliable. Mechanism A assumes that ethanol and water adsorbed much stronger than other components in the esterication solution, so the adsorption of ethyl lactate and lactic acid was neglected. We get rk
aW aLA aE aEL Keq

1 kW aW kE aE

In contrast to mechanism A, it was assumed that lactic acid and water had the strongest adsorption in mechanism B (Eq. (6)). rk
aW aLA aE aEL Keq

1 kW aW kLA aLA

In mechanisms A or B, there are three parameters instead of ve (Eq. (4)) to be estimated at a constant reaction temperature. In mechanism A, they are k , kW and kE . While for mechanism B, the corresponding parameters are k , kW and kLA . A two-stage optimization procedure was adopted for parameter evaluation. Firstly, the regression of these parameters was carried out by minimizing the sum of residual squares (SRS) between the experimental and calculated reaction rates. Then, numerical integration method was used for integrating the calculated reaction rates with previously determined parameters to get the conversions. The calculated conversion values were then compared with experimental values through the mean relative deviation (MRD). X SRS rexp rcalc 2 ; 7 N ! 1 X Xcalc Xexp 100%: 8 MRD N Xexp N

vj

Volume parameter Rj

Area parameter Qj

Water Ethanol Lactic acid

1 1 1 1 1 1 1

0.9011 1.8780 1.6764 0.92 2.1055 1.3013 1.8780

0.8480 1.660 1.420 1.40 1.972 1.224 1.660

106

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

Table 3 Values of parameters for the kinetic equations based on the mechanisms A and B Mechanism Parameters 002 Value A k (mol g kW kE
1

NKC SRS 1.83 109 MRD (%) 4.27 Value 1.701 5.073 2.245 1.449 5.510 2.010 SRS 2.10 109 MRD (%) 4.16

min )

1.704 4.973 2.093 1.650 5.016 1.554

k (mol g1 min1 ) kW kLA

5.88 109

8.61

_ 9 9.06 10

7.30

Regression results are summarized in Table 3, which included the values of the regression parameters, SRS and MRD at 353 K. The dierences of the two mechanisms were compared through the values of SRS and MRD. From the regression results, the rate equation for mechanism A described more accurately the experimental data than mechanism B. Therefore, the former was more reliable for this kind of esterication reaction. Similar mechanisms were applied in Sanz and Wei Songs work [9,17]. In order to verify this conclusion, FTIR analysis was used to test the component absorbing strength on the catalyst surfaces as discussed below. 6.3. Desorption experiment and FTIR analysis In the model assumption section, through the comparison of two dierent mechanisms, it was concluded that ethanol and water absorbed stronger than other components in the reaction solution. In order to examine the previous results, FTIR analysis was introduced. The used resins (002) were treated at high temperature in order to determine the desorption information which was inverse to the reaction of adsorption. The resin samples were withdrawn from the esterication reactor, and then treated at 100 C in an oven for dierent period of time (0, 0.5 and 8 h). In Fig. 1, the overall spectra of three samples are presented (wave numbers from 400 to 4000 cm1 ), in which any component changes on the surfaces of the catalysts could be observed. There is a broad band in the range from 3200 to 3600 cm1 , which is due to the symmetric and asymmetric stretching of the OH groups [18]. Fig. 1 shows that the intensity of
Fig. 1. Absorbance FTIR spectrum of 002 at dierent desorption times: 0 h (), 0.5 h (- - - - -), 8 h ( ), under 100 C, wave number region: 4004000 cm1 .

Fig. 2. Absorbance FTIR spectrum of 002 at dierent desorption times: 0 h (), 0.5 h (- - - - -), 8 h ( ), under 100 C, wave number region: 15001800 cm1 .

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

107

Fig. 3. Absorbance FTIR spectrum of lactic acid, wave number region: 4004000 cm1 .

the band decreases with desorption time, which means that the extent of the desorption reaction took place on the catalyst surface can be measured by FTIR. Fig. 2 shows two important bands, which are located at 15001800 cm1 . One is at around 1640 cm1 , and the other is located at 1740 cm1 . The former band is attributed to the bending vibration of water and the latter to the stretching vibration of the C@O [18]. It can be seen that the absorbance of the band at 1640 cm1 decreases as a function of desorption time, which indicates that water held a large proportion of active sites on the catalyst

Fig. 5. Conversion versus time for the esterication with 002 at dierent temperatures, 333 K ., 343 K d, 353 K N, 361 K j. The continuous lines represent the results of the LH model. All the reactions were carried out with an initial molar ratio 3:1 (ethanol:lactic acid). The catalyst loading was 4% (w/w) in all experiments.

surfaces. The intensity of the band at 1740 cm1 is relatively much smaller in comparison with the band at 1640 cm1 . However, the intensity of the stretching vibration of the C@O group is always very strong and sharp, which can be testied by the

Fig. 4. Absorbance FTIR spectrum of ethyl lactate, wave number region: 4004000 cm1 .

Fig. 6. Conversion versus time for the esterication with NKC at dierent temperatures, 333 K ., 343 K d, 353 K N, 361 K j. The continuous lines represent the results of the LH model. All the reactions were carried out with an initial molar ratio 3:1 (ethanol:lactic acid). The catalyst loading was 4% (w/w) in all experiments.

108

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

Table 4 Regression results of 002 at dierent reaction temperatures Temperature (K) 333 343 361 k (mol g1 min1 ) 0.762 1.015 2.555 kW 5.999 5.475 5.500 kE 3.254 2.601 2.900 SRS 8.01 10 2.84 1010 7.23 1010
10

MRD (%) 4.58 4.79 2.68

The results at 353 K are shown in Table 3.

Table 5 Regression results of NKC at dierent reaction temperatures Temperature (K) 333 343 361 k (mol g1 min1 ) 0.578 1.101 2.505 kW 5.921 5.507 5.501 kE 3.922 2.945 2.902 SRS 3.99 1010 2.58 109 3.60 1010 MRD (%) 3.94 4.05 2.33

The results at 353 K are shown in Table 3.

Fig. 7. (a) Arrhenius plot of esterication catalyzed with 002. The continuous line represents the result of linear regression. (b) vant Ho plot of adsorption coecient of water catalyzed with 002. The continuous line represents the result of linear regression. (c) vant Ho plot of adsorption coecient of ethanol catalyzed with 002. The continuous line represents the result of linear regression.

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

109

Fig. 8. (a) Arrhenius plot of esterication catalyzed with NKC. The continuous line represents the result of linear regression. (b) vant Ho plot of adsorption coecient of water catalyzed with NKC. The continuous line represents the result of linear regression. (c) vant Ho plot of adsorption coecient of ethanol catalyzed with NKC. The continuous line represents the result of linear regression.

FTIR spectra of lactic acid and ethyl lactate in Figs. 3 and 4. Thus it was concluded that lactic acid and ethyl lactate were not absorbed strongly in comparison with water on the resin surfaces. Through the FTIR experiments, it is proved that water absorbs much stronger on the resin surfaces than lactic acid and ethyl lactate do. Although the adsorption intensity of ethanol could not be conrmed in this way, mechanism B is

Table 6 Regression results of DH for 002 and NKC 002 Water DH (kJ/mol) )6.42 Ethanol )21.33 NKC Water )7.47 Ethanol )26.97 Fig. 9. Conversion versus time for the esterication with 002. The catalyst loading was 2% (w/w) j, 4% N and 6% d respectively. All the reactions were carried out with an initial molar ratio 3:1 (ethanol:lactic acid), at temperature 343 K.

110

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

surely unreasonable since the adsorption of lactic acid is thought as one of the strongest. 6.4. Esterication of lactic acid with ethanol 6.4.1. Eect of reaction temperature The eect of temperature on the esterication reaction was studied over a temperature range from 333 to 361 K, under atmospheric pressure, with 002 and NKC as catalysts. The boiling point of this mixture is about 361 K. As mentioned above, the stirring speed was set to around 450 rpm. So the inuence of external and internal diusion could be neglected under the given conditions. The experimental and regression results are displayed in Figs. 5 and 6. From the gures, the reaction rate increases with increasing temperature. The same trend was described in the reaction of lactic acid with methanol by Seo and

Fig. 10. Conversion versus time for the esterication with ve dierent cation-exchange resins. The symbols: 002 j, Amberlyst-15 N, D001 d, D002 ., NKC r. All the reactions were carried out with an initial molar ratio 3:1 (ethanol:lactic acid), at temperature 353 K. The catalyst loading was 4% (w/w) in all experiments.

Fig. 11. SEM photomicrographs of the surface of the resins: (a) fresh catalyst, (b) after esterication experiment and (c) reused in the catalytic distillation column. Catalyst: 002. Magnication: 150.

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

111

Sanz et al. [6,9], and in the reaction of isopropanol with lactic acid [4]. The reaction constant, adsorption coecients, SRS and MRD are provided in Tables 4 and 5. In both tables, the values of SRS were smaller than 3.0 109 and MRD were smaller than 5%. The adsorption coecient of water is greater than that of ethanol, which suggests that water adsorbs more strongly than ethanol on the active sites of the resins. 6.4.2. Activation energy and adsorption coecient Figs. 7 and 8 show the Arrhenius plots for the esterication reaction with 002 and NKC as catalysts at dierent temperatures. Arrhenius relation was introduced to describe the change of the reaction rate constant with temperature. The linear correlation coecients were higher than 0.995 for both catalysts. From the slopes of the straight lines in Figs. 7(a) and 8(a), the activation energies can be calculated by Eq. (9), which are 51.58 kJ/mol for

002 and 52.26 kJ/mol for NKC. The high values of the activation energy indicate that the reaction is kinetically controlled. From Figs. 7(b) and (c), it can be deduced that the variation of adsorption coecient can be described by vant Ho law (Eq. (10)) [8] in the temperature range from 333 to 353 K, but there was a big deviation at 361 K (the boiling point of the reaction mixture). Since a lot of bubbles appeared in the liquid reaction mixture at the boiling point, the concentrations ai of components in the liquid phase, mainly ethanol and water, became lower because of their higher volatility. Meanwhile, as the surface reaction was the rate-determining step, the concentration of absorbed components on catalyst surface was considered unchanged. In Eq. (4), ciS and cS did not change, while ai got smaller. So the regression value of ki was higher than that estimated by vant Ho law. Figs. 7 and 8 show that the adsorption of water and ethanol on cation-exchange resin surface

Fig. 12. SEM photomicrographs of the surface of the resins: (a) fresh catalyst, (b) after esterication experiment and (c) reused in the catalytic distillation column. Catalyst: 002. Magnication: 5000.

112

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

is an exothermic process, while the whole esterication reaction is an endothermic reaction. The regression results of DH for both catalysts were listed in Table 6. lnk lnk 0 lnki lnki0 EA ; RT DHi : RT 9 10

6.4.4. Eect of catalyst type Dierent types of cation-exchange resins were used to assess their eciency in the esterication reaction. Fig. 10 shows the plots of conversion of lactic acid against time for the various catalysts. It can be concluded that the catalytic activity increased in the order D002 < D001 < Amberlyst15 < NKC < 002 in the experimental conditions. 6.5. SEM study

6.4.3. Eect of catalyst loading Fig. 9 shows the eect of catalyst loading on reaction rates. It was observed that the equilibrium constant nearly did not change with the increase of catalyst loading. The time required to reach the equilibrium was reduced as the catalyst loading increased. The reason was that the more catalysts loaded, the more active sites were available for reaction.

SEM photomicrographs of the surface of two kinds of cation-exchange resins are presented in Figs. 1114. Each gure contains photographs of fresh catalyst (a), catalyst after reaction kinetics experiment as mentioned in the Section 6.4.4 (b), and catalyst after being reused in a catalytic distillation column (c). The column was operated for

Fig. 13. SEM photomicrographs of the surface of the resins: (a) fresh catalyst, (b) after esterication experiment and (c) reused in the catalytic distillation column. Catalyst: NKC. Magnication: 150.

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114

113

Fig. 14. SEM photomicrographs of the surface of the resins: (a) fresh catalyst, (b) after esterication experiment and (c) reused in the catalytic distillation column. Catalyst: NKC. Magnication: 5000.

more than 40 h at around 363 K in the lactic acid ester catalytic distillation process [19]. Figs. 11 and 12 show the SEM photographs of 002 under magnication of 150 and 5000. In these gures, the surface of 002 was very smooth like most gel resins. Figs. 12(b) and (c) show that after being used in the esterication and catalytic distillation experiments, there were few changes on the gel surfaces. Some small defects were observed on their surfaces, which were probably caused by impurities in the column or by a small amount of poly-lactic acid formed in the experiments, just like what happened on other catalysts after being used for a long time. In the same way, SEM micrographs of NKC are shown in Figs. 13 and 14. The surface had obvious dierences in comparison with 002. The pore diameters on the surfaces of fresh NKC ranged from 50 to 1000 nm. The surfaces of NKC had much more cracks and mass loss than 002

after the same intensity of stirring according to Figs. 13(b) and 14(b). In Fig. 14(c), the pore sizes of NKC became larger than the fresh catalysts. It was due to the swelling eect of reaction liquid under high temperature in the catalytic distillation column. All these indicated that 002 had better mechanical strength than NKC. For industrial applications, mechanical strength is an important factor in evaluation of catalyst. Taking lactic acid ester production for example, cation-exchange resins were used in a CSTR or a catalytic distillation column. Under the stirring force, the resins may be destroyed if they do not have enough mechanical strength. And that would cause a lot of problems in further separation and purication processes. In a reactive distillation column, the destroyed resin fragments would plug the packing section and lead to much pressure drop in it. Taking this factor into account, 002 is a better choice for industrial applications.

114

Y. Zhang et al. / Reactive & Functional Polymers 61 (2004) 101114 [2] R.A. Troupe, E. DiMilla, Ind. Eng. Chem. 49 (1957) 847. [3] J. Yang, L. Ma, Chinese Patent CN 1401623A, 03/12/2003. [4] G.D. Yadav, H.B. Kulkarni, React. Funct. Polym. 44 (2000) 153. [5] S. Dassy, H. Wiame, F.C. Thyrion, J. Chem. Tech. Biotechnol. 59 (1994) 149. [6] Y. Seo, W.H. Hong, J. Chem. Eng. Jpn. 33 (2000) 128. [7] J.J. Jafar, P.M. Budd, R. Hughes, J. Membr. Sci. 199 (2002) 117. [8] D.J. Benedict, S.J. Parulekar, S.P. Tsai, Ind. Eng. Chem. Res. 42 (2003) 2282. [9] M.T. Sanz, R. Murga, S. Beltran, J.L. Cabezas, J. Coca, Ind. Eng. Chem. Res. 41 (2002) 512. [10] L.K. Rihko, A.O. Krause, Ind. Eng. Chem. Res. 34 (1995) 1172. [11] M. Quiroga, M.R. Capeletti, N. Figoli, U. Sedran, Appl. Catal. A 177 (1999) 37. [12] G.C. Lukey, J.S.J. van Deventer, R.L. Chowdhury, D.C. Shallcross, S.T. Huntington, C.J. Morton, React. Funct. Polym. 44 (2000) 121. [13] P.B. Weisz, J.S. Hicks, Chem. Eng. Sci. 17 (1967) 265. [14] J. Lilja, J. Aumo, T. Salmi, D.Yu. Murzin, P. MakiArvela, M. Sundell, K. Ekman, R. Peltonen, H. Vainio, Appl. Catal. A 228 (2002) 253. [15] R.C. Reid, J.M. Prausnitz, B.E. Poling, The Properties of Gases and Liquids, fourth ed., McGraw-Hill, New York, 1987. [16] A. Fredenslund, J. Gmehling, P. Rasmussen, Vapor-liquid equilibria using UNIFAC: a group contribution method, Elsevier Scientic Pub. Co., Amsterdam, 1977. [17] W. Song, G. Venimadhavan, J.M. Manning, M.F. Malone, M.F. Doherty, Ind. Eng. Chem. Res. 37 (1998) 1917. [18] K. Nakanishi, P.H. Solomon, Infrared Absorption Spectroscopy, second ed., Holden-Day Inc., Oakland, CA, 1977. [19] L. Ma, Y. Zhang, J. Yang, Chin. J. Chem. Eng., in press.

7. Conclusions Esterication of lactic acid with ethanol over dierent kinds of cation-exchange resins was studied in this paper. The order of catalytic activity was found to be: D002 < D001 < Amberlyst-15 < NKC < 002. Water and ethanol have the strongest adsorption strength on catalyst surface according to the comparison of the regression results of two simplied LH mechanisms and testied by FTIR analysis. The esterication activation energies were found to be 51.58 kJ/mol (002) and 52.26 kJ/mol (NKC), which suggested that this reaction was kinetically controlled. 002 was found to be better than NKC since its catalytic activity, mechanical strength and temperature resistant ability according to kinetics experiments and SEM results. Acknowledgements The nancial support for this work is from National Key Technologies R&D Program of China (NKTRDP 2001BA708B01-01). We also acknowledge Mrs. Rui Yang for her help in the FTIR analysis. References
[1] K. Tanaka, R. Yoshikawa, C. Ying, H. Kita, K. Okamoto, Chem. Eng. Sci. 57 (2002) 1577.

Das könnte Ihnen auch gefallen