Вы находитесь на странице: 1из 18

Selectivity engineering in isopropylation of mesitylene with isopropyl alcohol over cesium substituted heteropolyacid supported on K-10 clay

Document by: Bharadwaj Visit my website

www.engineeringpapers.blogspot.com
More papers and Presentations available on above site
Abstract Alkylation of aromatics catalyzed by solid acids constitutes a class of reactions of both academic and industrial importance. Among alkylation reactions, isopropylation of aromatic compounds has attracted considerable attention. Use of propylene, as alkylating agent at very high temperatures leads to coke formation which results in deactivation of the catalyst. The use of isopropanol as an alkylating agent is attractive when propylene is not readily available. In-situ dehydration of isopropanol leads to prolonged activity since water of reaction suppresses coke formation. Further, isopropanol dehydration also generates diisopropyl ether (DIPE) which itself is an excellent alkylating agent. Alkylation of mesitylene with propylene or isopropanol (IPA) results in the formation of 2-isopropyl-mesitylene (2-IPMT) which is almost extensively used as a precursor in a number of industrial chemicals. This work covers the evaluation of clay supported heteropolyacids and sulfated zirconia. A variety of solid acid catalysts such as K-10 clay, sulfated zirconia, Filtrol-24, 20% w/w dodecatungstophosphoric acid (H3PW12O40, DTP) supported on K-10 montmorillonite clay and 20% w/w cesium substituted dodecatungstophosphoric acid (Cs2.5H0.5PW12O40, Cs-DTP) supported on K-10 montmorillonite clay were investigated for the liquid phase isopropylation of mesitylene to 2-IPMT using IPA at much milder conditions vis--vis other catalysts reported so far. 20% w/w Cs-DTP/K-10 clay was found to be the best catalyst which gives 98% conversion of limiting component, IPA and 98% selectivity towards the desired product, 2-IPMT after 2 h of total reaction time. This catalyst could be reused without any further chemical treatment, eliminating the effluent disposal problems. The reaction was carried out without using any solvent and the process subscribes to the principles of green chemistry. The catalytic activity is in the following order: 20% w/w Cs-DTP/K-10 clay (most active) > 20% w/w DTP/K-10 clay > Filtrol-24 > sulfated zirconia > K-10 clay (least active). The effect of various operating parameters and catalyst reusability were also systematically investigated. A mathematical model was proposed to probe into the intricate reaction kinetics and mechanism consistent with the experimental results. The reaction is free from any external mass transfer as well as intraparticle diffusion limitations and is intrinsically kinetically controlled. An overall second order kinetic equation was used to fit the experimental data, under the assumption that all the species are weakly adsorbed on the catalytic sites.

Keywords: Isopropylation, Mesitylene, Isopropyl alcohol, 2-Isopropyl-mesitylene, Solid acid catalyst, Heteropolyacids, Sulfated zirconia, Green chemistry, Selectivity. Introduction Alkylation is an interesting industrial organic reaction having extensive commercial utility. Products from every sector of the organic chemical industry make use of this methodology at some stage or other. Alkylation processes normally require FriedelCrafts acid catalysts such as H2SO4, BF3, TiCl4, liquid HF, and AlCl3 with elemental iodine etc. Some of the well established processes still employ homogeneous acid catalysts in batch reactors using large excess of the substrate or solvent causing problem of corrosion and pollution, loss of selectivity of the desired product, etc. In most of the industrial alkylation processes different catalysts such as mineral acids, anhydrous AlCl3 etc. and solvents like nitrobenzene, carbon disulphide and halogenated hydrocarbons are used. Relatively high concentration of catalyst is needed; often the amount being more than stoichiometric and these make most alkylation reactions highly polluting (Olah 1963; Kirk and Othmer 1996; Ullmann 2002; Olah et al. 1991; Franck and Stadelhofer 1988). In addition, since the reagents are mixed with acids, separation of the products from the catalyst is often a difficult and energy consuming process. Since several problems are associated with FriedelCrafts acid catalysts such as toxicity, corrosiveness, low reaction selectivity, and disposal of effluents. In recent years, the demand to replace mineral acids with environment-friendly catalysts such as ion exchange resins (Harmer and Sun 2001; Jayadeokar and Sharma 1993; Chakrabarti and Sharma 1993), clay-based catalysts (Yadav et al. 2003; Yadav and Asthana 2003; Yadav and Kirthivasan 1997), zeolites (Corma 1995) and metal oxides (Yadav and Nair 1999; Yadav and Murkute 2004) has increased in order to make the process cleaner and greener. Hitzler et al. (1998) reported the Friedel-Crafts alkylation of mesitylene and anisole with propene and/or 2-propanol using a heterogeneous polysiloxane-supported solid acid catalyst (Degussas Deloxan) in a small fixed bed continuous reactor (10 ml volume) using supercritical propene or CO2 as the reaction solvent. They have reported that, at temperature 160-180 0C with pressure 200 bar, yield of monoalkylated product (2isopropyl-mesitylene) was only approximately 25% due to the formation of dialkylated product as well as dimers and trimers of propene. Selectivity to the monoalkylated product was significantly higher (40% yield) in case of alkylation with 2-propanol in supercritical CO2. The authors have demonstrated success in tuning the product selectivity through control of temperature, pressure and reactant concentrations. Reduction of catalyst deactivation due to coking was potential advantage for conducting the process in a supercritical fluid. Continuous-flow reactor process operating in the supercritical fluid regime using a heterogeneous catalyst can also affects the FriedelCrafts alkylation or acylation reactions. We have recently reported the vapor phase alkylation of mesitylene with isopropanol (IPA) including dehydration of various alcohols and cracking of diisopropyl ether over UDCaT-4 (Yadav and Murkute 2004). The novel mesoporous solid acid catalyst UDCaT-4 was synthesized by incorporating superacidic centers of persulfated alumina and zirconia into highly ordered and well defined hexagonal mesoporous silica. The results are novel. We found that the conversion of mesitylene was dependent on the temperature and space time. The conversion of mesitylene increases with temperature up to 220 0C and then remains the same up to 250 0C, but above 250 0C the conversion is found to decrease. This is due to coke formation at high temperature which leads to decrease in conversion of mesitylene. The range of space time suggests that 40816 g h mol-1 as the optimum space time for mesitylene which gives 44% conversion of

mesitylene with 97% selectivity towards the monoalkylation. The space time was low enough to prohibit any significant dialkylation. We have also concluded that there is no effect of temperature and space time on the selectivity towards monoalkyaltion of mesitylene and remains the same. This report also suggests that dehydration of IPA was very fast in comparison with alkylation of mesitylene. In present study, the alkylation of mesitylene with IPA was chosen as a reaction because the reaction generates water as a co-product and thus the stability of the catalyst in the presence of water can be really tested. Another reason is that the reaction was carried out without using any solvent in order to make the process cleaner and greener. The alkylated product with mesitylene is a promising precursor for a number of industrial chemicals. There is practically no literature available on liquid phase alkylation of mesitylene with IPA over solid acids. Supercritical alkylation of mesitylene with propylene is reported and environmentally acceptable (Hitzler et al. 1998) but it require high pressure, appropriate costly instruments and become uneconomical. The current work covers the use of a variety of ecofriendly solid acid catalysts such as K-10 clay, sulfated zirconia, Filtrol-24, 20% w/w DTP/K-10 clay and 20% w/w Cs-DTP/K-10 clay. The study of dehydration of IPA was also undertaken independently to throw light on mechanism and selectivity. The effects of various parameters on rates and product distribution are used to deduce the kinetics of the reaction. Experimental section Chemicals and catalyst Mesitylene and isopropyl alcohol were obtained from M/s s. d. Fine Chemicals Pvt. Ltd. Mumbai, India. Filtrol-24 which is commercially available clay was obtained from Engelhard, USA and K-10 clay was obtained from Aldrich, USA. All chemicals were of analytical reagent (A.R.) grade. These were used as received without any further purification. Catalysts preparation The following catalysts were prepared by well-developed procedures and characterized in our laboratory: (i) 20% w/w dodecatungstophosphoric acid supported on K-10 clay (20% w/w H3PW12O40/K-10 clay i.e. 20% w/w DTP/K-10 clay) (Yadav and Kirthivasan 1997), (ii) 20% w/w cesium substituted dodecatungstophosphoric acid supported on K-10 clay (20% w/w Cs2.5H0.5PW12O40/K-10 clay i.e. 20% w/w Cs-DTP/K10 clay (Yadav et al. 2004; Yadav and Asthana 2003), and (iii) sulfated zirconia (S-ZrO 2) (Kumbhar and Yadav 1989; Kumbhar et al. 1989; Yadav and Nair 1999). All catalysts were dried in an oven at 120 0C for 1 h before use. Experimental setup The reactions were carried out in a 100 cm3 capacity Parr autoclave reactor with an internal diameter of 5 cm, equipped with four bladed pitched turbine impeller. The temperature was maintained at 1 0C of the desired value with the help of an in-built PID controller. Specific quantities of desired reactants and catalyst were charged into the reactor and the temperature was raised to the desired value. Then, an initial sample was withdrawn and agitation started. Further samples were withdrawn at periodic time intervals up to 2 h to monitor the reaction. Reaction procedure In a typical reaction, 0.316 mol mesitylene was reacted with 0.079 mol isopropanol (IPA) (4:1 mole ratio of mesitylene to IPA) with 2 g of catalyst; this makes

the catalyst loading as 0.04 g/cm3 of liquid phase. The total volume of the reaction mixture was 50 cm3. The reaction was carried out at 180 0C at a speed of agitation of 1000 rpm under autogenous pressure. The reaction was carried out without any solvent. Propylene formed in-situ was not allowed to escape from the reaction vessel. Method of analysis Clear liquid samples were withdrawn at regular time intervals by reducing the speed of agitation momentarily to zero and allowing the catalyst to settle at the bottom of the reactor. Analysis of the samples or compounds were performed by Gas Chromatograph (Chemito Model 8610 GC) equipped with a 10% SE-30 (liquid stationary phase) stainless steel column (3.175 mm diameter 4 m length) with FID detector. Products were isolated and confirmed through GC-MS and their physical properties and retention times were recorded and compared with authentic samples. Calibrations were done with authentic samples for quantification of data. The conversions were based on the disappearance of isopropanol (IPA), the limiting reactant in the reaction mixture. Results and discussions Catalyst characterization The catalyst was fully characterized, and the details are reported recently by us (Yadav et al. 2003, 2004; Yadav and Asthana 2003). Only a few salient features are reported here. Crystallinity and textural patterns of the catalysts predicted from X-ray diffraction data of 20% w/w Cs-DTP/K-10 clay (Yadav et al. 2003) show that DTP is crystalline while K-10 is amorphous. The diffractogram obtained suggested that, although the Cs-DTP salt loses some of its crystallinity in the process of supporting it on K-10, the Keggin structure of DTP remains intact. The Fourier transform infrared analysis fortified the preservation of the keggin structure of DTP in the catalyst. A characteristic split in the W = O band of Cs-DTP suggested the existence of direct interaction between the Keggin polyanion and Cs+. The scanning electron micrographs reveal that both K-10 and 20% w/w Cs-DTP samples possess rough and rugged surfaces, whereas 20% w/w Cs-DTP/K10 clay shows a smoother surface because of a layer of Cs salt of DTP over the external surface of K-10. The Brunauer-Emmett-Teller surface area of 20% w/w Cs-DTP/K-10 clay (Yadav et al. 2004) was measured to be 207 m2 g-1, and the pore volume and pore diameter were 0.29 cm3 g-1 and 58 , respectively. The adsorption-desorption isotherm for 20% w/w Cs-DTP/K-10 clay showed that they have the form of a type IV isotherm with the hysteresis loop of type H3, which is a characteristic of a mesoporous solid. Effect of different catalysts Different solid acid catalysts were used to assess their efficacy in this reaction. A 0.04 g/cm3 loading of catalyst based on the organic volume of the reaction mixture was employed at 180 0C. The catalysts were 20% w/w DTP/K-10 clay, 20% w/w Cs-DTP/K10 clay, sulfated zirconia (S-ZrO2), Filtrol-24 and K-10 clay. It was found that 20% w/w Cs-DTP/K-10 clay showed higher conversion compared to other catalysts and the order of activity was: 20% w/w Cs-DTP/K-10 clay (most active) > 20% w/w DTP/K-10 clay > Filtrol-24 > sulfated zirconia > K-10 clay (least active) (Fig. 1). The acid strength of inorganic catalysts such as 20% w/w DTP/K-10 clay (Yadav and Kirthivasan 1997), 20% w/w Cs-DTP/K-10 clay (Yadav and Asthana 2003) and sulfated zirconia (Kumbhar and Yadav 1989; Kumbhar et al. 1989) was determined by temperature-programmed desorption of ammonia. This has been already reported by us earlier and hence the details are avoided. The purpose of using several different solid acid catalysts was to study the effect of nature, strength and distribution of acidity, pore size distribution and stability of 4

the catalyst on conversion of isopropanol and selectivity to monoalkylated product, 2isopropyl-mesitylene (2-IPMT). The catalyst properties and final conversion are given in Table 1. Conversion of isopropanol was more (98%) with 20% w/w Cs-DTP/K-10 clay as compared to conventional sulfated zirconia (61%) and also the catalyst gave maximum selectivity (98%) towards the desired product, 2-isopropyl-mesitylene. The comparison of initial activity of catalysts suggested that 20% w/w Cs-DTP/K-10 clay was far superior to others, mainly because of its nano-size, higher surface area and ease of accessibility of the active sites to reacting molecules. Hence further experiments were conducted with 20% w/w Cs-DTP/K-10 clay. The observed concentration profile of different products for this reaction at 180 0C is depicted in Fig. 2, which clearly shows that the selectivity of 98% for 2-isopropyl-mesitylene was achieved. The pores of the catalyst get narrowed when K10 clay is impregnated with Cs-DTP nanoparticles as reported in our earlier work (Yadav and Asthana 2003). Thus, only monoalkylated product, 2-IPMT is formed as the major product with trace amount of dialkylated product, 2,6-diisopropyl-mesitylene (2,6DIPMT) while the trialkylated product, 2,4,6-triisopropyl-mesitylene (2,4,6-TIPMT) does not form during the reaction course. Effect of speed of agitation To assess the role of external mass transfer on reaction rate, the effect of speed of agitation (Fig. 3) was studied. The speed of agitation was varied from 800 to 1200 rpm. It was observed that the conversion of isopropanol was practically the same in all the cases. The external mass transfer effects did not influence the reaction. Hence, all further reactions were carried out at 1000 rpm. The influence of external solidliquid mass transfer resistance must be ascertained before a true kinetic model could be developed. Depending on the relative magnitudes of external resistance to mass transfer and reaction rates, different controlling mechanisms have been put forward (Yadav et al. 2003, 2004; Yadav and Asthana 2003). This reaction is a typical solidliquid slurry reaction involving the transfer of limiting reactant IPA (A), and mesitylene (B) from the bulk liquid phase to the catalyst wherein external mass transfer of reactants to the surface of the catalyst particle, followed by intraparticle diffusion, adsorption, surface reactions, and desorption, take place. Thus experimental and theoretical analyses were also done to establish that there was no effect of external mass transfer limitations. Effect of catalyst loading In the absence of external mass transfer resistance, the rate of reaction is directly proportional to catalyst loading based on the entire liquid phase volume. The catalyst loading was varied over a range of 0.01-0.05 g/cm3 on the basis of total volume of the reaction mixture. Fig. 4 shows the effect of catalyst loading on the conversion of IPA. The conversion increased with an increase in catalyst loading, which was due to the proportional increase in the number of active sites. The final conversion obtained with 0.05 g/cm3 loading was not much different than that of 0.04 g/cm3, which suggested that the number of active sites available were little more than those required. Hence all further experiments were carried out at 0.04 g/cm3 loading. At this loading, the intra-particle diffusion resistance sets in. Proof of absence of intra-particle resistance Because the average particle size of 20% w/w Cs-DTP/K-10 clay was found to be in the range of 2-10 m and the catalyst is amorphous in nature, it was not possible to study the effect of catalyst particle size on the rate of reaction. The average particle

diameter of 20% w/w Cs-DTP/K-10 clay used in the reactions was 0.001 cm, and thus a theoretical calculation was done based on the Weisz-Prater criterion (Fogler 1995; Reid et al. 1977) to asses the influence of intraparticle diffusion resistance. According to the Weisz-Prater criterion, the value of {-robs pRp2/De[As]} has to be far less than unity for the reaction to be intrinsically kinetically controlled and which can be evaluated from the observed rate of reaction (-robs), density of catalyst particle ( p), the particle radius (RP), the effective diffusivity of the limiting reactant (De), and the concentration of the reactant at the external surface of the particle ([AS]). The calculated value 4.1810-3 further reveals that the absence of mass transfer limitation at the reaction conditions and therefore, the reaction is intrinsically kinetically controlled. A further proof of the absence of intraparticle diffusion resistance was obtained through the study of the effect of temperature and it will be discussed later. Effect of mole ratio The mole ratio of mesitylene to IPA was varied from 1:1 to 5:1 under otherwise similar operating conditions to assess its effect on the rate and selectivity. The overall reaction rate of IPA increased with an increase in the mole ratio of mesitylene to IPA from 1:1 to 4:1. Further increase in mole ratio did not have any significant effect on conversion of IPA. Thus, all the subsequent reactions were carried out with a mole ratio of 4:1 (Fig. 5). A mole ratio of 4:1 was maintained: (i) to avoid the formation of large amounts of secondary products, such as the oligomers of propylene and dialkylated products, and (ii) to diminish the influence of water formed by the dehydration of IPA in situ. The reaction was also carried out with mesitylene to IPA mole ratio 1:4. Even though the conversion of isopropanol was significant, the rate of alkylation of mesitylene with propylene formed was very slow under the same reaction conditions. Also the products formed mainly were diisopropyl ether (DIPE) and monoalkylated product. Effect of temperature Intrinsically kinetically controlled reactions show significant increase in the conversion profile with temperature. Since almost all mass transfer limitations were eliminated, the effect of temperature was studied on two reaction steps. Case 1: Dehydration of IPA
OH

2
H3 C CH3

- H2O Cs-DTP/K- 10
H3 C

CH3

CH3

- H 2O
CH3

CH2

Cs-DTP/K- 10

2
CH3

IPA

DIPE

Propylene

IPA dehydration reaction was studied in the temperature range of 160190 0C. Propylene and diisopropyl ether (DIPE) were the products formed. The rate of dehydration increased with increase in temperature (Fig. 6). The product distribution is shown in Fig. 7. It was found that the formation of DIPE increased sharply with temperature from 9% at 160 0C to 41% at 190 0C after 2 h. Since propylene is difficult to sample and quantify, the concentrations of isopropanol and diisopropyl ether were first quantified by GC and then a mass balance was established to calculate the concentration of propylene. Case 2: Alkylation of mesitylene with IPA

CH3 OH

CH3

CH3

CH3

CH3

+
H 3C CH3

H 3C

CH3

Cs-DTP/K-10 180C
H 3C CH3

Cs-DTP/K-10 180C

H3 C

CH3

H 3C

CH3

H3 C

CH3

Mesitylene

IPA

2-IPMT

2,6-DIPMT

The alkylation of mesitylene with IPA is highly temperature dependent. The temperature effect was studied from 160190 0C to investigate the influence of temperature on the rate of reaction and the selectivity of product. Fig. 8 shows the effect of temperature on the conversion of limiting reactant, IPA. With an increase in temperature from 160 0C to 190 0C, both the rate of reaction as well as the selectivity towards 2-IPMT was increased. The overall reaction rate of IPA increased with an increase in temperature from 160 0C to 180 0C. Further increase in temperature did not have any wide effect on conversion of IPA. At the typical operating condition, the conversion of IPA was 98% with 98% selectivity towards 2-IPMT. No oligomerisation of propylene was occurred in this temperature range. The formation of dialkylated product i.e. 2,6-diisopropyl-mesitylene (2,6-DIPMT) increases with an increase in temperature while the trialkylated product, 2,4,6-triisopropyl-mesitylene (2,4,6-TIPMT) does not form during the reaction course. Reaction kinetics Case 1: Dehydration of IPA The dehydration of IPA over solid acids leads to propylene and also diisopropyl ether which we have earlier studied independently over a number of catalysts including sulfated zirconia and heteropolyacids supported on clay. Independent study of IPA dehydration have been studied over 20% w/w Cs-DTP/K-10 clay and reported in our recent paper (Yadav and Kamble 2009). Solid superacids have both Lewis and Bronsted sites and thus the mechanism involves bifunctional sites S1 and S2. These two species participate in the reaction. Furthermore when mesitylene was reacted with IPA over 20% w/w Cs-DTP/K-10 clay, the reaction was found to follow second order. Thus, it is seen that the alkylation with alcohols does not follow a simple reaction. Thus, a model based on two catalytic sites was proposed according to which IPA (A) gets adsorbed on to two different sites S1 and S2. These two adsorbed species participate in the reaction. Assuming that the rate determining step is the reaction of AS1 and AS2 to form diisopropyl ether and water as the surface complexes (ES1) and (WS2) respectively and ES1 subsequently decomposes instantly into propylene (P) in the gas phase. K A + S1 AS1 (1) K A + S 2 AS2 (2)
1 A 2 A

SR AS 1 +AS 2 ES+ WS 2 1 1 2 ES 1 SR 2 + WS1 P The site balance in this case is CT S1 = CV S1 + C AS1 +C E S1 +CW S1

(3) (4) (5) (6) (7) (8) (9) 7

CT S2 = CV S2 + C AS2 + CW S2

The following adsorption equilibria for different species hold


K1 W + S1 W WS1 K 2 W W + S 2 WS2 K1 E E + S1 ES1

Thus the rate of formation of propylene, -rP' (mol gcat-1 s-1) is: k SR1K 1ACAK 2 ACACT S1C S 2 T rP '= (10) + 1 K AC ( 1 + K1 AC A K WC+ K1 ECE +1) ( 2 + A K2 WCW ) W When the adsorption of all species are very weak, equation (10) is reduced to 2 (11) rP ' = kwC A Where, k = k SR1 K1A K 2ACT S1 CT S2 (12) Writing in terms of conversion, and further integration results into the following equation: XA = kwC A0 t (13) 1 X A XA Thus a plot of against t (Fig. 9) was made to get an excellent fit thereby 1 X A supporting the model. This is an overall second order reaction for weak adsorption of IPA (A). Case 2: Alkylation of mesitylene with IPA Various models were tried, including typical first order kinetics (weak adsorption of IPA and strong adsorption of mesitylene) and overall second order kinetics (weak adsorption of IPA and mesitylene). The overall second order dehydration model was taken as a basis. As is validated above, IPA dehydration follows second order kinetics by adsorption of IPA on two adjacent sites S1 and S2 and the product diisopropyl ether (E) is formed, which is decomposed instantaneously to propylene (P). We have also reported earlier (Yadav and Murkute 2004), the dehydration of IPA over a broad range of temperatures (110150 0C and 180220 0C), which showed that DIPE, although formed at lower temperature, cracks faster than IPA and also the rate of alkylation is not controlled by the dehydration rate. Thus in the temperature range studied, the rate of alkylation is not controlled by the dehydration rate, but the alkylation of mesitylene adsorbed on site S2 with IPA adsorbed on adjacent site S1, to give the monoalkylated product i.e. 2-isopropylmesitylene (D), which is formed due to the surface reaction as shown below. K A + S1 AS1 (14) K B + S 2 BS2 (15) k BS 2 + AS1 DS 2 + WS1 (16) Analogously, the site balance can be written to obtain: k SR2 K1 AC A K 2B C B CT S1 CT S2 rA ' = (17) (1 + K1 AC A + K1W CW + K1E C E )(1 + K 2B C B + K 2DC D ) With weak adsorption of all species, equation (17) is reduced to rA ' = k SR w AC B C (18) Where, k SR = k SR K1A K 2B CT S CT S (19) Writing in terms of conversion, and further integration results into the following equation: M XA ln (20) = k SR wC A0 ( M 1) t M (1 X A )
1 A 2 B
SR2

M XA Thus, a plot of ln against t is shown in Fig. 10 in the isopropylation M (1 X A ) reaction of mesitylene with IPA as the alkylating agent. It is seen that the data fit very

well, thereby supporting the model. This is an overall second order reaction for weak adsorption of IPA (A) and mesitylene (B). From Fig. 10, the slopes obtained at 160, 170, 180 and 190 0C were found to be 2.83 10-4, 3.65 10-4, 4.97 10-4 and 9.22 10-4 s-1 respectively. Hence the value of rate constants (k or kSR) at different temperature for alkylation reaction can be calculated as kSR (160 0C) = 0.82 cm6 gcat-1 mol-1 s-1 kSR (170 0C) = 1.06 cm6 gcat-1 mol-1 s-1 kSR (180 0C) = 1.44 cm6 gcat-1 mol-1 s-1 kSR (190 0C) = 2.67 cm6 gcat-1 mol-1 s-1 Similarly, in case of IPA dehydration (Fig. 9), the slopes obtained at 160, 170, 180 and 190 0C were found to be 3.02 10-4, 4.88 10-4, 9.10 10-4 and 3.77 10-3 s-1 respectively. Hence the rate constants (k or kSR) at different temperature can be calculated as kSR (160 0C) = 0.38 cm6 gcat-1 mol-1 s-1 kSR (170 0C) = 0.62 cm6 gcat-1 mol-1 s-1 kSR (180 0C) = 1.16 cm6 gcat-1 mol-1 s-1 kSR (190 0C) = 4.79 cm6 gcat-1 mol-1 s-1 Arrhenius plot (Fig. 11) was used to estimate the frequency factor (k0) and activation energy (E). The value of frequency factor and activation energy for alkylation reaction was calculated as 3.78 107 cm6 gcat-1 mol-1 s-1 and 15.3 kcal/mol respectively. In case of IPA dehydration, the value of frequency factor and activation energy was 76.92 1014 cm6 gcat-1 mol-1 s-1 and 32.5 kcal/mol respectively. The value of activation energy also supported the fact that the overall rate of reaction is not influenced by either external mass transfer or intraparticle diffusion resistance and it is an intrinsically kinetically controlled reaction on active sites. Reusability of catalyst The change in the texture of the catalyst from white to gray suggested its deactivation due to coking. After each reaction, reactivation of the catalyst was done by maintaining the catalyst in a solution of IPA at reflux temperature for 4 h in order to remove any adsorbed material from catalyst surface and pores and dried at 120 0C for 2 h and weighed before using in the next batch. There were small losses during filtration. The actual amount of catalyst used in the next batch was almost 5% less than the previous batch. The catalyst was reused with a make-up quantity and the experiments on reusability were repeated. Although the catalyst was washed after filtration to remove all adsorbed reactants and products, there was still a possibility of retention of small amount adsorbed reactants and products species which might cause the blockage of active sites of the catalyst. These are apparent factors for the loss in activity. It was observed that there was only a marginal decrease in conversion, but there was no change in the selectivity of the product to suggest that the catalyst is stable. Fig. 12 depicts the observed conversion profiles. Conclusion The liquid phase isopropylation of mesitylene with IPA was studied with different solid acid catalysts such as K-10 clay, sulfated zirconia, Filtrol-24, 20% w/w DTP/K-10 clay and 20% w/w Cs-DTP/K-10 clay. 20% w/w Cs-DTP/K-10 clay catalyst was the most active and selective catalyst, which lead to 98% conversion of the limiting reactant, IPA with 98% selectivity towards 2-isopropyl-mesitylene. The kinetics of the reaction is also reported. The reactions were found to be intrinsically kinetically controlled. The overall second order kinetic equation fits the data very well and the activation energy was found

to be 15.3 kcal/mol, which suggested that the reaction was an intrinsically kinetically controlled on active sites of the catalyst. The reaction is solvent free which could be advantageous as a green and clean process. Acknowledgement GDY acknowledges receipt of a research grant and chair from Darbari Seth Professorship Endowment. SBK acknowledges receipt of Senior Research Fellowship (SRF) from University Grants Commission (UGC), Government of India, New Delhi. Nomenclature A B D P E W M Ci CA CB CV CT CA0 K1 k1 k1' KA KB kSR k k0 -ri' E Si SR i-j T-Si V-Si w t XA Limiting reactant species A, IPA Excess reactant species B, Mesitylene Monoalkylated desired product i.e. 2-isopropyl-mesitylene Propylene Diisopropyl ether (DIPE) Water Mole ratio of mesitylene to isopropyl alcohol Concentration of species i, mol/cm3 Concentration of A, mol/cm3 Concentration of B, mol/cm3 Concentration of vacant sites of catalyst, mol/cm3 Concentration of total sites of catalyst, mol/cm3 Initial concentration of A at solid catalyst surface, mol/cm3 Surface reaction equilibrium constant, k1/k1' Surface reaction rate constant for forward reaction Surface reaction rate constant for reverse reaction Adsorption equilibrium constant for A, cm3/mol Adsorption equilibrium constant for B, cm3/mol Second order rate constant, cm6 gcat-1 mol-1 s-1 Second order rate constant, cm6 gcat-1 mol-1 s-1 Frequency factor, cm6 gcat-1 mol-1 s-1 Rate of reaction of species i, mol gcat-1 s-1 Apparent activation energy, kcal/mol Site of type i Surface reaction Species j adsorbed on site i Total sites S of type i Vacant sites S of type i Catalyst loading, g/cm3 of liquid phase Reaction time interval, min. Fractional conversion of A

10

References Chakrabarti A, Sharma MM (1993) Cationic ion exchange resins as catalyst. React Polym 20(1-2): 1-45. Corma A (1995) Inorganic Solid Acids and Their Use in Acid-Catalyzed Hydrocarbon Reactions. Chem Rev 95: 559-614. Fogler HS (1995) Elements of Chemical Reaction Engineering. 2nd edn. Prentice-Hall, New Delhi, India. Franck HG, Stadelhofer JW (1988) Industrial Aromatic Chemistry. Springer, Berlin. Harmer MA, Sun Q (2001) Solid acid catalysis using ion-exchange resins. Appl Catal A: Gen 221(1-2): 45-62. Hitzler MG, Smail FR, Ross SK, Poliakoff M (1998) Friedel-Crafts alkylation in supercritical fluids: continuous, selective and clean. Chem Commun 3: 359-360. Jayadeokar SS, Sharma MM (1993) Ion exchange resin catalysed etherification of ethylene and propylene glycols with isobutylene. React Polym 20(1-2): 57-67. Kirk and Othmer (1996) Encyclopedia of Chemical Technology. 4th edn. WileyInterscience, New York. Kumbhar PS, Yadav GD (1989) Catalysis by sulfur-promoted superacidic zirconia: Condensation reactions of hydroquinone with aniline and substituted anilines. Chem Eng Sci 44(11): 2535-2544. Kumbhar PS, Yadav VM, Yadav GD in: D. E. Layden (Ed.) (1989) Chemically Modified Oxide Surfaces. Gordon and Breach, New York. Olah GA (1963) Friedel-Crafts and Related Reactions. Wiley-Interscience, New York, vol. 1-4. Olah GA, Krishnamuri R, Suryaprakash GK (1991) Comprehensive Organic Synthesis. Pergamon, Oxford, vol. 3, chapter 1.8. Reid RC, Prausnitz MJ, Sherwood TK (1977) The Properties of Gases and Liquids. 3rd edn. McGraw-Hill, New York. Ullmann F (2002) Encyclopedia of Industrial Chemistry. 6th edn. Wiley-VCH Verlag GmbH, Weinheim, Germany. Yadav GD, Asthana NS (2003) Selective decomposition of cumene hydroperoxide into phenol and acetone by a novel cesium substituted heteropolyacid on clay. Appl Catal A: Gen 244(2): 341-357. Yadav GD, Asthana NS, Kamble VS (2003) Cesium-substituted dodecatungstophosphoric acid on K-10 clay for benzoylation of anisole with benzoyl chloride. J Catal 217(1): 88-99. Yadav GD, Asthana NS, Salgaonkar SS (2004) Regio-selective benzoylation of xylenes over caesium modified heteropolyacid supported on K-10 clay. Clean Tech Environ Policy 6: 105-113. Yadav GD, Kamble SB (2009) Alkylation of xylenes with isopropyl alcohol over acidic clay supported catalysts: Efficacy of 20% w/w Cs2.5H0.5PW12O40/K-10 clay. Ind Eng Chem Res (In Press). Yadav GD, Kirthivasan N (1997) Synthesis of bisphenol-A: Comparison of efficacy of ion exchange resin catalysts vis--vis heteropolyacid supported on clay and kinetic modeling. Appl Catal A: Gen 154(1-2): 29-53. Yadav GD, Murkute AD (2004) Novel Efficient Mesoporous Solid Acid Catalyst UDCaT-4: Dehydration of 2-Propanol and Alkylation of Mesitylene. Langmuir 20(26): 11607-11619. Yadav GD, Nair JJ (1999) Sulfated zirconia and its modified versions as promising catalysts for industrial processes. Micro Meso Mater 33(1-3): 1-48.

11

Table 1. Properties of catalysts and conversion of IPAa Catalyst Source Particle size (m) Surface area (m2/g) Pore volume (cm3/g) Average pore diameter (nm) Conversion of IPA after 2h (%)

Filtrol-24 Engelhard 30-400 350 0.42 7.5 76 Sulfated zirconia This work 50-300 100 0.115 2.8 61 K-10 Aldrich 50-200 230 0.36 6.4 52 b 20% w/w DTP/K-10 This work 50-200 107 0.32 7.1 81 20% w/w Cs-DTP/K-10 This work 200-300c 207 0.29 5.8 98 a Reaction conditions: speed of agitation1000 rpm; catalyst loading0.04 g/cm3; mole ratio of mesitylene:IPA4:1; temperature180 0C; total reaction volume50 cm3; autogenous pressure. b The particle size of the HPA on the K-10 support was 150 nm and c The particle size of the HPA on the K-10 support was 5 nm.

12

100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

() 20% w/w Cs-DTP/K-10 clay, () 20% w/w DTP/K-10 clay, () Filtrol-24, () Sulfated zirconia (S-ZrO2), () K-10 clay Fig. 1. Effect of different catalysts on conversion of IPA: speed of agitation1000 rpm; catalyst loading0.04 g/cm3; mole ratio of mesitylene:IPA4:1; temperature180 0C; total reaction volume50 cm3; autogenous pressure.
1.8 1.6
3

Concentration (mol/cm) 10

1.4 1.2 1 0.8 0.6 0.4 0.2 0 0 20 40 60 80 100 120 140

Time (min)

() IPA, () 2-Isopropyl-mesitylene, () Propylene, () Diisopropyl ether, () 2,6Diisopropyl-mesitylene Fig. 2. Concentration profile of various products in isopropylation of mesitylene with IPA: catalyst20% w/w Cs-DTP/K-10 clay; speed of agitation1000 rpm; catalyst loading0.04 g/cm3; mole ratio of mesitylene:IPA4:1; temperature180 0C; total reaction volume50 cm3; autogenous pressure. 13

100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

() 800 rpm, () 1000 rpm, () 1200 rpm Fig. 3. Effect of speed of agitation on conversion of IPA: catalyst20% w/w Cs-DTP/K10 clay; catalyst loading0.04 g/cm3; mole ratio of mesitylene:IPA4:1; temperature180 0 C; total reaction volume50 cm3; autogenous pressure.
100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

() 0.01 g/cm3, () 0.02 g/cm3, () 0.04 g/cm3, () 0.05 g/cm3 Fig. 4. Effect of catalyst loading on conversion of IPA: catalyst20% w/w Cs-DTP/K-10 clay; speed of agitation1000 rpm; mole ratio of mesitylene:IPA4:1; temperature180 0 C; total reaction volume50 cm3; autogenous pressure.

14

100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

() 1:1, () 3:1, () 4:1, () 5:1 Fig. 5. Effect of mole ratio of mesitylene:IPA on conversion of IPA: catalyst20% w/w Cs-DTP/K-10 clay; speed of agitation1000 rpm; catalyst loading0.04 g/cm3; temperature180 0C; total reaction volume50 cm3; autogenous pressure.
100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

() 160 0C, () 170 0C, () 180 0C, () 190 0C Fig. 6. Effect of temperature on the cracking of IPA: catalyst20% w/w Cs-DTP/K-10 clay; speed of agitation1000 rpm; catalyst loading0.04 g/cm3; temperature180 0C; total reaction volume50 cm3; autogenous pressure. 15

100

80

Selectivity (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

() 160 0C, () 170 0C, () 180 0C, () 190 0C (______) Propylene, (______) Diisopropyl ether Fig. 7. Effect of temperature on selectivity of products in dehydration of IPA: catalyst 20% w/w Cs-DTP/K-10 clay; speed of agitation1000 rpm; catalyst loading0.04 g/cm3; temperature180 0C; total reaction volume50 cm3; autogenous pressure.
100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

() 160 0C, () 170 0C, () 180 0C, () 190 0C Fig. 8. Effect of temperature on isopropylation of mesitylene with IPA: catalyst20% w/w Cs-DTP/K-10 clay; speed of agitation1000 rpm; catalyst loading0.04 g/cm3; mole ratio of mesitylene:IPA4:1; total reaction volume50 cm3; autogenous pressure. 16

16 14 12 10 y = 0.2261x R = 0.9866
2

XA/1-XA

8 y = 0.0546x 6 4 2 0 0 20 40 60 80 100 120 140 R = 0.9939 y = 0.0293x R = 0.9933 y = 0.0181x R = 0.9733


2 2 2

Time (min)

() 160 0C, () 170 0C, () 180 0C, () 190 0C Fig. 9. Validation of mathematical model for dehydration of IPA
3.0

2.5

ln[(M-XA)/M(1-XA)]

2.0

1.5

y = 0.0553x R2 = 0.9992 y = 0.0298x R2 = 0.9856 y = 0.0219x R2 = 0.9841 y = 0.0170x R2 = 0.9862

1.0

0.5

0.0 0 5 10 15 20 25

Time (min)

() 160 0C, () 170 0C, () 180 0C, () 190 0C Fig. 10. Validation of mathematical model for isopropylation of mesitylene with IPA

17

2.50

1.50

0.50

y = -7677x + 17.447 R2 = 0.9445

ln k

-0.50 y = -16346x + 36.579 R2 = 0.9229

-1.50

-2.50

-3.50 2.12E-03 2.16E-03 2.20E-03 2.24E-03


-1

2.28E-03

2.32E-03

1/T (K )

() Dehydration of IPA, () Alkylation of mesitylene with IPA Fig. 11. Arrhenius Plot
100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

() Fresh catalyst, () First reuse, () Second reuse Fig. 12. Reusability of catalyst: catalyst20% w/w Cs-DTP/K-10 clay; speed of agitation1000 rpm; catalyst loading0.04 g/cm3; mole ratio of mesitylene:IPA4:1; temperature180 0C; total reaction volume50 cm3; autogenous pressure.

18

Вам также может понравиться