You are on page 1of 22

Friedel-Crafts alkylation of xylenes with tertbutanol over mesoporous superacid UDCaT-5

Document by: Bharadwaj Visit my website

www.engineeringpapers.blogspot.com
More papers and Presentations available on above site
Abstract Friedel-Crafts alkylation of xylenes with tert-butanol in the presence of novel mesoporous superacidic catalysts named as UDCaT-4, UDCaT-5 and UDCaT-6 was investigated. The catalysts are modified versions of zirconia showing high catalytic activity, stability and reusability. The catalytic activity is in the order: UDCaT-5 (most active) > UDCaT-6 > UDCaT-4 > sulfated zirconia (least active). Synergistic effect of very high sulfur content present (9% w/w S) and preservation of tetragonal phase in UDCaT-5, in comparison with sulfated zirconia (4% w/w S), were responsible for higher catalytic activity. The performance of UDCaT-5 in alkylation of xylenes was studied with tert-butanol with reference to selectivity and stability. Alkylation of m-xylene over UDCaT-5 gives 96% conversion of tert-butanol with 82% selectivity towards 5-tertbutyl-m-xylene (5-TBMX) under optimum reaction conditions. The formation of products is correlated with the acidity of the catalyst. The reactions were conducted in liquid phase at relatively low reaction temperatures (130160 C). A systematic investigation of the effects of various operating parameters was done to describe the reaction pathway. The reaction was carried out without any solvent in order to make the process cleaner and greener. An overall second order kinetic equation was used to fit the experimental data, under the assumption that both xylene and tert-butanol are weakly adsorbed. An independent dehydration study of tert-butanol (TBA) was also done. Keywords: Friedel-Crafts alkylation; tert-Butyl xylenes; Mesoporous superacidic catalysts UDCaT-5; Sulfated zirconia; Green chemistry. Introduction Alkylation processes normally require FriedelCrafts acid catalysts such as AlCl3, BF3, TiCl4, liquid HF, and AlCl3 with elemental iodine.1,2 Several problems are associated with these catalysts such as toxicity, corrosiveness, low reaction selectivity, and disposal of effluents.35 Relatively high concentration of catalyst is needed; often the amounts are more than stoichiometric making the reactions inherently polluting. 6 Due to ever increasing societal, environmental, and economic pressure, efforts are devoted to the development of environmentally friendly catalysts for the production of industrially important chemicals and intermediates.79 Alkylation of xylenes with tert-butanol (TBA) is an important reaction both in organic synthesis and chemical manufacturing. Some dimethyl alkylbenzenes have assumed practical significance. In particular, 1,2-dimethyl-4-tert-butylbenzene or 4-tertbutyl-o-xylene (4-TBOX) was proposed as the starting substance for the production of 1

novel phthalocyanine pigments, plasticizers, photographic materials and other valuable products.1013 tert-Butylated xylenes are usually manufactured by reacting xylenes in the presence of liquid acid catalysts, with pure isobutylene or C4 fraction from naphtha crackers containing isobutylene, giving wide product distribution. These processes suffer from problems associated with the use of highly corrosive liquid acids and also the source of isobutylene. The development of a technologically efficient, highly productive and environmentally safe method for the synthesis of dimethyl-tert-butylbenzene (DMTBB) is challenging. There is a tremendous scope for devising a new catalytic process for the synthesis of tert-butylated xylenes to replace conventional homogenously catalyzed, highly polluting processes. Besides, due to the problems associated with unavailability, transportation and handling of isobutylene, particularly for usage in low-tonnage fine and speciality chemical industry (typically 10-100 TPA production), it is advantageous to generate isobutylene in situ. Dehydration of tert-butanol is an attractive source for the same. Further, tert-butanol is available as a by-product in the ARCO process for propylene oxide which could be used effectively for this purpose. We have successfully carried out tert-butylation of several aromatic compounds by using tert-butanol, methyltert-butyl ether (MTBE) and isobutene as alkylating agents using ecofriendly solid acid catalysts.1417 The only problem which needs to be considered was activity and stability of solid acid catalysts in the presence of water, which is evolved during the reaction. For the past few years, our laboratory is engaged vociferously in the synthesis, characterization and application of selective, ecofriendly and active catalysts such as UDCaT series catalysts, sulfated zirconia, heteropolyacids and their modified versions supported on clays. In particular, sulfated zirconia has been extensively studied in a number of reactions1820 and it should be modified to bring in shape selectivity, mesoporosity and better acidity. Recently we prepared novel mesoporous superacids named as UDCaT-4, UDCaT-5 and UDCaT-6, which are modified versions of zirconia and have found a great potential for industrially important reactions.2123 We have reported, for the first time, that a sulfated zirconia, with sulfur content as high as 9% w/w, was produced with preservation of tetragonal phase by using chlorosulfonic acid as a new source for sulfate ion. It was designated as UDCaT-5. The acronym UDCaT symbolizes our research institute, formerly called University Department of Chemical Technology (UDCT), which is now renamed as Institute of Chemical Technology (ICT). The work summarizes the investigation of activity and selectivity of these new breed of catalytic materials in alkylation of xylenes with tert-butanol, including reaction kinetics. The catalyst can be used as the basis for a high performance and environmentally safe method for the synthesis of DMTBB. The model studies were done with m-xylene and then extended to other xylene isomers. Experimental Section Chemicals and Catalysts Pure xylene isomers, tert-butanol, zirconium oxychloride, aluminum nitrate, ammonium persulfate and aqueous ammonia solution were procured from Aldrich, USA. Hexadecyl amine and chlorosulfonic acid was purchased from Spectrochem Ltd. Mumbai, India. Tetraethyl orthosilicate (TEOS) was procured from Fluka, Germany. All chemicals were of analytical reagent (A.R.) grade. These were used as received without any further purification. Preparation of Catalysts

The following catalysts were prepared by well-developed procedures and characterized in our laboratory: (i) sulfated zirconia (S-ZrO2), (ii) UDCaT-4, (iii) UDCaT-5, and (iv) UDCaT-6. Preparation of Sulfated Zirconia (S-ZrO2) Sulfated zirconia (S-ZrO2) was prepared by adding aqueous ammonia solution to zirconium oxychloride solution at a pH of 10, as detailed elsewhere.20 The precipitate was thoroughly washed with distilled water and made free from ammonia and chloride ions. It was dried in an oven at 120 C for 24 h. The sulfation of the zirconia was done using 15 cm3 g-1 of 0.5 M sulfuric acid. It was dried at 110 C and calcined at 650 C for 3 h. Preparation of UDCaT-4 The ordered hexagonal mesoporous silica (HMS) was prepared according to our earlier work.24 Desired quantities of zirconium oxychloride and aluminum nitrate were dissolved in aqueous solution and added to precalcined HMS by incipient wetness technique. After addition the solid was dried in an oven at 110 C for 3 h. The dried material was hydrolyzed by ammonia gas and washed with deionized water until a neutral filtrate was obtained and the absence of chlorine ion in the filtrate was detected by phenolphthalein indicator and silver nitrate tests. It was then dried in an oven for 24 h at 110 C. Persulfation was carried out by immersing the above solid material in to 0.5 M aqueous solution of ammonium persulfate for 30 min. It was dried at 110 C for 24 h and calcined at 650 C for 3 h to get active catalyst called UDCaT-4 with 0.6% w/w of alumina.21 Preparation of UDCaT-5 UDCaT-5 was prepared by adding aqueous ammonia solution to zirconium oxychloride (ZrOCl2.8H2O) solution at pH of 9-10. The precipitated zirconium hydroxide so obtained was washed with deionized water until a neutral filtrate was obtained. The absence of chlorine ion was detected by the AgNO3 test. A material balance on chloride ions before and after precipitation and washing shows no retention of Cl on the solid. Zirconium hydroxide (Zr(OH)4) was dried in an oven for 24 h at 100 C and was crushed to 100 mesh size. Zr(OH)4 then immersed in 15 cm3 g-1 of 0.5 M solution of chlorosulfonic acid and ethylene dichloride. All materials were immersed for 5 min in the solution and then without allowing moisture absorption were kept in an oven and the heating was started slowly to 120 C after about 30 min. These materials were kept in oven at 120 C for 24 h and calcined at 650 C for 3 h to get the active catalysts UDCaT5.22 Preparation of UDCaT-6 UDCaT-6 was prepared by adding an aqueous solution of 2.5 g zirconium oxychloride to 5 g precalcined HMS by incipient wetness technique and it was dried in an oven at 120 C for 3 h. The dried material was hydrolyzed by ammonia gas and washed with distilled water until no chloride ions were detected which was confirmed by the AgNO3 test. It was further dried in an oven for 2 h at 120 C. Zr(OH)4/HMS was immersed in 15 cm3 g-1 of 0.5 M chlorosulfonic acid in ethylene dichloride. It was soaked for 5 min in the solution and then without allowing moisture absorption, it was oven dried to evaporate the solvent at 120 C after about 30 min. The sample was kept in the oven at 120 C for further 24 h and calcined thereafter at 650 C for 3 h to get the active catalyst UDCaT-6.23

Characterization of Catalysts UDCaT-4,21 UDCaT-5,22 UDCaT-6,23 and sulfated zirconia20 were completely characterized by ammonia-temperature programmed desorption (NH3-TPD), X-ray powder diffraction (XRD), Brunauer-Emmett-Teller (BET) surface area and Fourier transform infrared (FTIR) and the details were published recently by us. Only a few salient features which thought be important are reported here. Characterization of UDCaT-4 The XRD, BET surface area and pore size analyses provided an explanation for the entrapment of nanoparticles of persulfated alumina zirconia (PAZ) (<3.6nm) in mesoporous of HMS. The XRD of UDCaT-4 suggested that the structural integrity of HMS was maintained even after converting it into UDCaT-4. The diffractogram of UDCaT-4 further revealed that the introduction of small amount of alumina (0.16% w/w) and sulfate ion (1.17% w/w) stabilized the tetragonal phase of the zirconia, which is an ideal phase conducive for superacidity in sulfated zirconia, into the pores of HMS. Furthermore, the pore volume of UDCaT-4 (0.21 cm3 g-1) is much less than that of pure HMS (0.78 cm3 g-1) indicating that large amount of crystalline zirconia (9.01% w/w), and alumina must be present inside pores of UDCaT-4. FTIR spectroscopy and energy dispersive X-ray (EDAX) analysis further support the assumption drawn on introduction of sulfate ion in UDCaT-4. The sulfur Ka1 and zirconium La1 distribution spectra determined by EDAX analysis shows the incorporation and homogeneous distribution of zirconia and sulfur atoms in UDCaT-4. Temperature programmed desorption (TPD) profiles of PAZ and UDCaT-4 show that although UDCaT-4 possesses both weak and medium acid sites, the total acid sites of UDCaT-4 (0.56 mmol g -1) are greater than those of PAZ (0.09 mmol g-1).21,24 Scanning electron micrographs (SEM) of UDCaT-4 revealed that similar to the morphology of HMS, UDCaT-4 is made up of sub-micrometer sized free standing or aggregated sphere shaped particles. SEM analysis further supports the argument that active centers of the PAZ are successfully embedded in HMS and the structural integrity of HMS is unaltered even after it is converted to UDCaT-4.21 Characterization of UDCaT-5 Ammonia-TPD was used to determine the acid strength of UDCaT-5. AmmoniaTPD analysis shows that apart from intermediate and strong acidic sites present in sulfated zirconia, UDCaT-5 also contains superacidic sites. Elemental analysis shown that UDCaT-5 contain 9% w/w sulfate content which is highest as compared to reported so far in the literature. IR spectroscopy confirms that the chlorosulfonic acid is decomposed during calcination at 650 C and sulfate ions are retained on the surface of UDCaT-5 and thus the sulfur content is higher in UDCaT-5. The XRD study proved that the tetragonal phase in the UDCaT-5 was preserved. From BET surface area and pore size analysis showed that the surface area of UDCaT-5 decreased abruptly at the maximum sulfur content. It was due to the migration of sulfate ions from bulk phase to zirconia matrix. Thus maximum sulfur present on surface decreases its surface area. Thus, UDCaT-5 is superacidic in nature due to the presence of very high sulfur content present on the zirconia matrix with preservation of tetragonal phase of zirconia.22 Characterization of UDCaT-6 FTIR spectroscopy and EDAX analyses support the introduction and retention of sulfate ion in UDCaT-6. XRD, BET surface area and pore size analysis provided an explanation for entrapment of nanoparticles of zirconia in mesoporous of HMS. The XRD of UDCaT-6 suggested that the structural integrity of HMS was retained even after

converting it into UDCaT-6. Furthermore, the pore volume of UDCaT-6 (0.7 cm3 g-1) is much less than that of pure HMS (1.2 cm3 g-1) indicating that large amount of crystalline nano-particles of zirconia must be present inside the pores of UDCaT-6. The sulfur Ka1 and zirconium La1 distribution spectra determined by EDAX analysis shows the incorporation and homogeneous distribution of zirconia and sulfur atoms in UDCaT-6. The SEM of UDCaT-6 revealed that similar to the morphology of HMS, UDCaT-6 is made up of sub-micrometer sized free standing or aggregated sphere shaped particle and that active centers of zirconia are successfully embedded in HMS and the structural integrity of HMS is unaltered even after it is converted to UDCaT-6.23 Apparatus and Procedures 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 C of the desired value with the help of an in-built proportional integral differential (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. All catalysts were dried in an oven at 120 C for 1 h before use. In a typical reaction, 0.342 mol xylene isomer was reacted with 0.0855 mol tertbutanol (TBA) (4:1 mole ratio of xylene to TBA) with 2 g of catalyst; this makes the catalyst loading as 0.04 g cm-3 of liquid phase. The total volume of the reaction mixture was 50 cm3. The reaction was carried out at 150 C at a speed of agitation of 1000 rpm under autogenous pressure. The reaction was carried out without any solvent. Isobutylene 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, India: Model 8610 GC) equipped with a 10% SE-30 (liquid stationary phase) stainless steel column (3.175 mm diameter 4 m length) with flame ionization detector (FID). Products were isolated and confirmed through gas chromatography-mass spectrometry (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 tert-butanol (TBA), which is the limiting reactant in the reaction mixture. Results and Discussion Choice of Catalysts The activity and stability of UDCaT-4, UDCaT-5 and UDCaT-6 catalysts were tested by carrying out separate reactions which produced HCl or water as the co-products. Thus, benzylation of toluene with benzyl chloride and esterification of p-cyclohexanol with acetic acid, where HCl and water are by-products, respectively, were studied systematically, whose details are given elsewhere.2025 The catalysts showed higher activity and good reusability as compared to sulfated zirconia. It indicated that the catalysts were stable in presence of HCl and water. Moreover all catalysts were reusable in all these reactions, without loss of activity. So it was thought desirable to investigate the efficacy of these catalysts in the liquid phase tert-butylation of xylene isomers with

TBA, where water is one of the co-products. Effects of various operating parameters such as effect of speed of agitation, catalyst loading, mole ratio of xylene to TBA and reaction temperature on rates and product distribution are discussed with m-xylene as the model isomer and then extended to other xylene isomers to deduce the kinetics of the reaction. Efficacies of Various Catalysts The activities of UDCaT-4, UDCaT-5, UDCaT-6 and sulfated zirconia were evaluated at 150 C. It was found that UDCaT-5 showed higher conversion of limiting reactant, TBA (96%) as compared to the other solid acid catalysts with selectivity of 82% towards the desired product, 5-tert-butyl-m-xylene (5-TBMX). The order of activity was: UDCaT-5 (most active) > UDCaT-6 > UDCaT-4 > sulfated zirconia (S-ZrO 2) (least active) (Figure 1). All other catalysts contain less number of acidic sites as compared to UDCaT-5 and thus the results are in this order. 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 the catalyst on conversion and selectivity in a complex network of reactions involving dehydration, etherification and alkylation. Since the substrate is bulky and involves generation of water, it is essential that sulfated zirconia which gets deactivated is modified. Besides, our group was the first one to report the highest superacidity in solid superacids. This has been already published by us.22 The catalyst properties, selectivity of 5-TBMX and final conversion of TBA are given in Table 1, which clearly indicated that UDCaT-5 is the most active and selective catalyst in terms of conversion and selectivity. Hence further experiments were conducted with UDCaT-5 due to its excellent activity, reusability and stability. The observed concentration profile of reactants and several products for this reaction at 150 C is depicted in Figure 2, which clearly shows that the selectivity of 82% for 5-tert-butyl-m-xylene was achieved. Effect of Speed of Agitation To ascertain the influence of external resistance to mass transfer of the reactants to the catalyst surface, the speed of agitation was varied over the range of 800 to 1200 rpm under otherwise similar conditions (Figure 3). At all speeds, solid catalyst particles are supposed to be completely in suspension. It was observed that the final conversion of TBA was practically the same in all these cases and thus there was no limitation of external mass transfer of reactants from bulk liquid phase to the outer surface of the catalyst. Also there was no significant effect on the selectivity of the desired product. Theoretical analyses were also done to establish that there was no effect of external mass transfer limitations as delineated in our earlier work.18,25 Hence, all further reactions were carried out at 1000 rpm. Effect of Catalyst Loading The catalyst loading was varied over a range of 0.010.05 g cm-3 on the basis of total volume of reaction mixture under otherwise similar conditions. Figure 4 shows the effect of catalyst loading on the conversion of TBA. The initial rates of reaction, in the absence of external mass transfer resistance, are plotted against catalyst loading in Figure 5. It demonstrates that the rates are directly proportional to the catalyst loading based on the entire liquid phase volume. The conversion of TBA increased with an increase in catalyst loading (Figure 4), which is obviously due to the proportional increase in the number of active sites. The final conversion of TBA was practically the same in case of 0.04 g cm-3 and 0.05 g cm-3 loading. This suggests that the number of sites available were

more than the number of molecules. Hence all further experiments were conducted by using 0.04 g cm-3 loading and at this loading, the intraparticle diffusion resistance sets in. Proof of Absence of Intraparticle Resistance Because the average particle size of UDCaT-5 was found to be in the range of 40 50 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 UDCaT-5 used in the reactions was 0.0045 cm, and thus a theoretical calculation was done based on the Weisz-Prater criterion to assess the influence of intraparticle diffusion resistance. According to the Weisz-Prater criterion,26,27 the dimensionless parameter CWP, which represents the ratio of the intrinsic reaction rate to the intraparticle diffusion rate, can be evaluated from the observed rate of reaction (-robs = 3.3410-5 mol gcat-1 s-1), density of catalyst particle ( p = 1.255 g cm-3), the particle radius (Rp = 2.2510-3 cm), the effective diffusivity of the limiting reactant (De = 4.066 10-6 cm2 s-1), and the concentration of the reactant at the external surface of the particle ([As] = 1.7110-3 mol cm-3). The dimensionless parameter CWP is given by: 2 robs p R p CWP = (1) De [ As ] Here, (i) if CWP >> 1, then the reaction is limited by severe internal diffusion resistance, and (ii) if CWP << 1, then the reaction is intrinsically kinetically controlled. In the present case, the value of CWP was calculated as 3.05 10-2 for the initial observed rate (-robs) which is much less than 1 and therefore, the reaction is intrinsically kinetically controlled. A further proof of the absence of the 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 m-xylene to TBA was varied from 1:1 to 5:1 under otherwise similar operating conditions to assess its effect on the conversion of TBA (Figure 6). As the m-xylene:TBA mole ratio was increased, an increase in conversion of TBA and the rate of reaction with increase in selectivity for monoalkylated desired product was observed. There was insignificant increase in conversion beyond mole ratio 4:1. The mole ratio study indicates that using excess TBA results in to fast deactivation of active sites, which may be attributed to passivation of active sites by strong adsorption of either excessive alcohol molecule or water produced,28 but in the present case there is no deactivation of UDCaT-5 over the range studied. There are three more reasons to carry out experiments at 4:1 mole ratio: (i) to avoid the formation of large amounts of secondary products, such as the oligomers of isobutene and dialkylated products, (ii) to diminish the influence of water formed by the dehydration of TBA in situ, and (iii) lower the concentration of TBA, more the adsorption of TBA to produce isobutene on the catalyst sites and therefore, more availability of isobutyl cation to react with m-xylene. Thus all the subsequent reactions were carried out with a mole ratio of 4:1. With an increase in amount of m-xylene, the activity of isobutene produced increases which results in increase in the rate of alkylation. The reaction was also carried out with mxylene to TBA mole ratio 1:4. Even though the conversion of TBA was significant, the rate of alkylation of m-xylene with isobutene formed was very slow under the same reaction conditions. Also the products formed mainly were isobutylene, di-tert-butyl ether (DTBE) 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. Dehydration of TBA
OH CH3 H3 C H 3C O CH3 CH3 CH3 CH3

2
H 3C CH3 CH3

- H 2O UDCaT- 5

- H 2O UDCaT- 5

2
H 3C CH2

TBA

DTBE

Isobutylene

Scheme 1. Dehydration of TBA An independent dehydration study of TBA (Scheme 1) was studied in the temperature range of 130160 C (Figure 7) to ensure no coke formation on the catalyst. Isobutylene and di-tert-butyl ether (DTBE) were the products formed in this reaction. DTBE was initially formed and then cracked to isobutylene. This can be explained as, tert-butyl group is bulky and hence ether which was formed is unstable and breaks into isobutylene. The rate of dehydration increased with increase in temperature. Because isobutylene is difficult to sample and quantify, the concentrations of TBA and DTBE were first quantified by gas chromatography (GC) and then a mass balance was established to calculate the concentration of isobutylene. Alkylation of m-Xylene with TBA
CH3 OH CH3 CH3 CH3 CH3

+
CH3

H3C

CH3

CH3

UDCaT-5 150C

H3C CH3 H3C CH3

UDCaT-5 150C

CH3 H3 C H3 C CH3 CH3

m-Xylene

TBA

5-TBMX

2,5-DTBMX

Scheme 2. Alkylation of m-xylene with TBA The alkylation of m-xylene with TBA (Scheme 2) is highly temperature dependent. The temperature effect was studied at temperatures from 130160 C (Figure 8). With an increase in temperature, both the rate of reaction as well as selectivity for monoalkylated product increased. The final conversion of TBA was increased from 74% at 130 C to 96% at 150 C and the selectivity for 5-tert-butyl-m-xylene (5-TBMX) was 82% at 150 C after 2 h. There is no significant difference in the final conversion of TBA between 150 C and 160 C. Therefore, 150 C is the optimum temperature for this reaction. Spectroscopy evidences proved that UDCaT-5 contains three types of acidic sites, namely, intermediate, strong and very strong. Strong acidic sites present in UDCaT5 are in more number and hence monoalkylation was dominant. The formation of dialkylated product i.e. 2,5-di-tert-butyl-m-xylene (2,5-DTBMX) increases with an increase in temperature. Alkylation of Xylene Isomers with TBA

CH3 OH

CH3

CH3 CH3 CH3

CH3

CH3 CH3

+
CH3

H3 C

CH3

CH3

UDCaT-5 150C
CH3

UDCaT-5 150C

CH3 H3 C H3 C CH3 CH3

p-Xylene

TBA

2-TBPX

2,5-DTBPX

Scheme 3. Alkylation of p-xylene with TBA


CH3 CH3 OH CH3 CH3 H3 C CH3 CH3 CH3

H 3C

CH3

CH3

UDCaT-5 150C
CH3 CH3

UDCaT-5 150C

H3 C

H3 C

H 3C

CH3

CH3

o-Xylene

TBA

4-TBOX

4,6-DTBOX

Scheme 4. Alkylation of o-xylene with TBA Alkylation of p-xylene (Scheme 3) and o-xylene (Scheme 4) was also carried out using TBA as an alkylating agent over UDCaT-5 as a catalyst under otherwise similar operating conditions as were used for m-xylene. The observed concentration profile of reactants and several products for these reactions at 150 C are depicted in Figure 9 & 10. These both the reactions gives maximum conversion of 96-97% of the limiting reactant, TBA with 80-82% selectivity towards the desired product; 2-tert-butyl-p-xylene (2TBPX) in case of p-xylene with TBA and 4-tert-butyl-o-xylene (4-TBOX) in case of oxylene with TBA reaction. It was also observed that the alkylation of o-xylene was faster relative to the alkylation of p-xylene and m-xylene due to the fact that o-xylene is a more active species than other xylene isomers (Table 2). Reaction Kinetics We have recently reported the alkylation of several substituted aromatics using linear and branched alcohols, and ethers using a variety of solid acid catalysts, wherein a free olefin and water are generated in situ as a co-product, along with the C or/and O alkylated products.25,29,30 In condensation reaction of two alcohols it was found that the symmetrical or asymmetrical ethers are formed (e.g. methyl-tert-butyl ether from methanol and tert-butanol) in which sulfated zirconia, ion exchange resin, 20% w/w Cs2.5H0.5PW12O40/K-10 clay i.e. 20% w/w Cs-DTP/K-10 clay were employed as catalysts. In some cases there was formation of alkenes along with C or/and O alkylated products; whereas isopropylation using isopropanol gave diisopropyl ether (DIPE) and this ether further act as an alkylating agent. The rate of alkylation was higher by using ether than alcohol.30 Thus, the alkylation with alcohols does not follow a simple reaction pathway and our current rate data using a new catalyst needed testing of different hypotheses. Several models were tried to fit the experimental data which were collected in the absence of any external mass transfer and intraparticle diffusion limitations. The result was expected to obey a first order kinetics for weakly adsorbed species, but contrary to this, the model followed second order kinetics. Further, it was found that the alkylation of m-xylene also followed a second order reaction. The analysis of earlier published reports on the dehydration of ethanol and isopropanol31,32 suggests that there is a production of ether at lower temperature and also two types of sites should be involved in the reaction. Solid superacids have both Lewis and Bronsted sites and thus the mechanism involves 9

bifunctional sites S1 and S2. These two species participate in the reaction. Thus mechanistic models were needed to describe TBA dehydration to isobutylene and water and alkylation of m-xylene with TBA. These two cases are considered here: (i) dehydration of TBA, and (ii) alkylation of m-xylene with TBA. Dehydration of TBA In this case, a model based on two catalytic sites was proposed according to which TBA (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 AS 1 and AS2 to form di-tert-butyl ether and water as the surface complexes (ES1) and (WS2) respectively and ES1 being unstable subsequently decomposes instantly into isobutylene (P) in the gas phase as shown below: K A + S1 AS1 (2) K A + S 2 AS2 (3)
1 A 2 A

SR AS 1 +AS 2 ES+ WS 2 1 1

(4) (5) (6) (7) (8) (9) (10)

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


CT S2 = CV S2 + C AS2 + CW S2

k SR2

The following adsorption equilibria for different species hold


K1 W + S1 W WS1 K2 W + S 2 W WS2 K1 E E + S1 ES1

Thus the rate of formation of isobutylene, -rP' (mol gcat-1 s-1) is: k SR1K 1ACAK 2 ACACT S1C S 2 T rP '= (11) + 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 (11) is reduced to 2 (12) rP ' = kwC A Where, k = k SR1 K1A K 2ACT S1 CT S2 (13) Writing in terms of conversion, and further integration results into the following equation: XA = kwC A0 t (14) 1 X A XA Thus a plot of against t (Figure 11) was made to get an excellent fit 1 X A thereby supporting the model. This is an overall second order reaction for weak adsorption of TBA (A). Alkylation of m-Xylene with TBA Various models were tried, including typical first order kinetics (weak adsorption of TBA and strong adsorption of m-xylene) and overall second order kinetics (weak adsorption of TBA and m-xylene). The overall second order dehydration model was taken as a basis. As is validated above, TBA dehydration follows second order kinetics by adsorption of TBA on two adjacent sites S1 and S2 and the product di-tert-butyl ether (E) is formed, which is decomposed instantaneously to isobutylene (P). Thus in the temperature range studied, the rate of alkylation is not controlled by the dehydration rate,

10

but the alkylation of m-xylene adsorbed on site S2 with TBA adsorbed on adjacent site S1, to give the monoalkylated desired product (D), which is formed due to the surface reaction as shown below: K A + S1 AS1 (15) K B + S 2 BS2 (16) k BS 2 + AS1 DS 2 + WS1 (17) Analogously, the site balance can be written to obtain: k SR2 K1 AC A K 2B C B CT S1 CT S2 rA ' = (18) (1 + K1 AC A + K1W CW + K1E C E )(1 + K 2B C B + K 2DC D ) With weak adsorption of all species, equation (18) is reduced to rA ' = k SR w AC B C (19) Where, k SR = k SR 2 K1A K 2B CT S1 CT S2 (20) Writing in terms of conversion, and further integration results into the following equation: M XA ln (21) = 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 Figures 1214 in the M (1 X A ) alkylation reaction of m-xylene, p-xylene and o-xylene respectively with TBA as an 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 TBA (A) and xylenes (B). The values of rate constants (k or kSR) at different temperatures were calculated and an Arrhenius plots (Figure 15) was used to estimate the frequency factor (k0) and apparent activation energy (E) of each reaction and is tabulated in Table 2. The values 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 of the catalyst. The values of rate constants also shows that o-xylene is most active isomer than p-xylene and m-xylene (Table 2) and their activity is in the order o-xylene (most active) > p-xylene > m-xylene (least active).
Reusability of Catalyst Reusability of UDCaT-5 was verified by conducting three runs. After each run the catalyst was filtered, and then refluxed with 50 cm3 TBA for 30 min in order to remove any adsorbed material from catalyst surface and pores. The catalyst was then dried at 120 C for 2 h and weighed before using in the next batch. There was some attrition of catalyst particle during agitation. In a typical batch reaction, there were inevitably losses of particles during filtration due to attrition. Although the catalyst was washed after filtration to remove all adsorbed reactants and products, there was still a possibility of retention of small amount of 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, which have been taken into account because no fresh catalyst was added into the reaction mixture to maintain the same catalyst loading during reusability tests. The actual amount of catalyst used in the next batch, was almost 5% less than the previous batch. It was observed that there is only a marginal decrease in conversion (Figure 16); even there was no effect on selectivity of the products. When make-up quantity of the catalyst was added, almost similar conversions were found to suggest that the catalyst is stable. Conclusion 11

The liquid phase alkylation of xylenes with TBA as an alkylating agent was examined by using various modified version of zirconia named as UDCaT-4, UDCaT-5 and UDCaT-6. The UDCaT-5 catalyst was found to be the best catalyst in terms of activity and selectivity, which lead to 96% conversion of the limiting reactant, TBA, with 82% selectivity toward the desired product, dimethyl-tert-butylbenzene. No oligomerisaton of isobutylene was formed at standard reaction conditions and monoalkylated products were obtained exclusively. The rate of reaction increased with temperature, catalyst loading and xylene to TBA mole ratio. The reaction should be carried out at 150 C with catalyst loading 0.04 g cm-3 with mole ratio of xylene:TBA, 4:1 to obtain dimethyl-tert-butylbenzene with maximum yield. We believe that this is the novel method to synthesize dimethyl-tert-butylbenzene or in other words tert-butylated xylenes at such a low temperature and with high activity and selectivity using mesoporous superacidic UDCaT-5 catalyst. Comprehensive mathematical models were developed and validated with experimental results. The overall second order kinetic equation fits the data very well and the activation energy was calculated for all tertbutylation reactions including dehydration of TBA. The values 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; it is an intrinsically kinetically controlled reaction on active sites of the catalyst. The reaction is solvent free which could be advantageous as a green process and the conditions are optimized in a way to get dimethyl-tert-butylbenzene selectively at low temperatures. The present study also shows that UDCaT-5 has a potential as a solid superacid in a number of reactions. Acknowledgement Professor G. D. Yadav acknowledges receipt of a research grant and chair from Darbari Seth Professorship Endowment. Dr. S. B. Kamble acknowledges receipt of Senior Research Fellowship (SRF) from University Grants Commission (UGC), Government of India, New Delhi. We also wish to thank anonymous reviewer and editor for their critical and useful comments which refined the manuscript a lot. Nomenclature A B D P E W M Ci CA CB CV CT CA0 K1 k1 k1' KA KB kSR limiting reactant species A, tert-butanol excess reactant species B, m-xylene or xylene isomers monoalkylated desired product, dimethyl-tert-butylbenzene isobutylene di-tert-butyl ether water mole ratio of xylene to tert-butanol concentration of species i, mol cm-3 concentration of A, mol cm-3 concentration of B, mol cm-3 concentration of vacant sites of catalyst, mol cm-3 concentration of total sites of catalyst, mol cm-3 initial concentration of A at solid catalyst surface, mol cm-3 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-1 adsorption equilibrium constant for B, cm3 mol-1 second order rate constant, cm6 gcat-1 mol-1 s-1

12

k k0 -ri' E Si SR i-j T-Si V-Si w t XA

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-1 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 cm-3 of liquid phase reaction time interval, min. fractional conversion of A

References (1) Olah, G. A. Friedel-Crafts and Related Reactions; Wiley-Interscience: New York, 1963; Vol. 1-4. (2) Olah, G. A.; Krishnamuri, R.; Suryaprakash, G. K. Comprehensive Organic Synthesis; Pergamon: Oxford, 1991; Vol. 3, Chapter 1.8. (3) Clark, J. H.; Macquarrie, D. J. Org. Proc. Res. Develop. 1997, 1, 149. (4) Corma, C.; Garcia, H. Chem. Rev. 2003, 103, 4307. (5) Corma, A. Chem. Rev. 1995, 95, 559. (6) Lorenc, J. F.; Lambeth, G.; Scheffer, W. KirkOthmer Encyclopedia of Chemical Technology; WileyInterscience: New York, 1992; Vol. 2. (7) Sheldon, R. A.; Downing, R. S. Appl. Catal. A: Gen. 1999, 189, 163. (8) Sheldon, R. A.; van Bekkum, H. Fine Chemicals through Heterogeneous Catalysis; Wiley-Interscience, Weinheim: Toronto, 2001. (9) van Bekkum, H.; Hoefnagel, A. J.; Vankoten, M. A.; Gunnewegh, E. A.; Vogt, A. H. G.; Kouwenhoven, H. W. Stud. Surf. Sci. Catal. 1994, 83, 379. (10) Isakov, Y. L.; Minachev, K. M.; Kalinin, V. P.; Isakova, T. A. Russian Chemical Bulletin 1996, 45, 2763. (11) Derfer, J. M.; Derfer, M. M. KirkOthmer Encyclopedia of Chemical Technology; WileyInterscience: New York, 1978; Vol. 22. (12) Fiege, H.; Bayer, A.; Leverkusen, G. Ullmanns Encyclopedia of Industrial Chemistry; 5th ed.; Wiley-VCH Verlag GmbH, Weinheim: Germany, 1991. (13) Franck, H. G.; Stadelhofer, J. W. Industrial Aromatic Chemistry; Springer: Berlin, 1988. (14) Yadav, G. D.; Doshi, N. S. J. Mol. Catal. A: Chem. 2003, 194, 195. (15) Yadav, G. D.; Doshi, N. S. Appl. Catal. A: Gen. 2002, 236, 129. (16) Yadav, G. D.; Thorat, T. S. Tetrahedron Lett. 1996, 37, 5405. (17) Yadav, G. D.; Pujari, A. A.; Joshi, A. V. Green Chem. 1999, 1, 269. (18) Kumbhar, P. S.; Yadav, G. D. Chem. Eng. Sci. 1989, 44, 2535. (19) Yadav, G. D.; Krishnan, M. S.; Doshi, N. S.; Pujari, A. A.; Mujeebur Rahuman, M. S. M. Highly Acidic Mesoporous Synergistic Solid Catalyst and its Applications; US Patent 6,204,424, 2001. (20) Yadav, G. D.; Nair, J. J. Microporous Mesoporous Mater. 1999, 33, 1. (21) Yadav, G. D.; Murkute, A. D. Langmuir 2004, 20, 11607. (22) Yadav, G. D.; Murkute, A. D. J. Catal. 2004, 224, 218. (23) Yadav, G. D.; Murkute, A. D. J. Phy. Chem. A 2004, 108, 9557. (24) Yadav, G. D.; Manyar, H. G. Microporous Mesoporous Mater. 2003, 63, 85. (25) Yadav, G. D.; Salgaonkar, S. S. Ind. Eng. Chem. Res. 2005, 44, 1706.

13

(26) (27) (28) (29) (30) (31) (32)

Fogler, H. S. Elements of Chemical Reaction Engineering; 2nd ed.; Prentice-Hall, New Delhi: India, 1995. Reid, R. C.; Prausnitz, M. J.; Sherwood, T. K. The Properties of Gases and Liquids; 3rd ed.; McGraw-Hill: New York, 1977. Song, C.; Kirby, S. Microporous Mater. 1994, 2, 467. Yadav, G. D.; Bokade, V. V. Appl. Catal. A: Gen. 1996, 147, 299. Yadav, G. D.; Salgaonkar, S. S. Microporous Mesoporous Mater. 2005, 80, 129. Rouge, A.; Spoetzl, B.; Gebauer, K.; Schenk, R.; Renken, A. Chem. Engg. Sci. 2001, 56, 1419. Golay, S.; Wolfrath, O.; Doepper, R.; Renken, A. Dynamics of Surface and Reaction Kinetics in Heterogeneous Catalysis; Elsevier: Amsterdam, 1997; Vol. 109, pp. 295. Table 1. Properties of catalysts and conversion of TBAa

catalyst

source

average pore diameter () 30 40 32 40

surface area (m2 g-1) 233 83 877 100

acidity by NH3-TPD (mmol g-1) 0.562 0.584 0.508 0.433

selectivity of 5-TBMX after 2 h (%) 73 82 66 78

conversion of TBA after 2 h (%) 60 96 74 36

UDCaT-4 UDCaT-5 UDCaT-6 sulfated zirconia (S-ZrO2)


a

this work this work this work this work

Reaction conditions: speed of agitation 1000 rpm, catalyst loading 0.04 g cm-3, mole ratio of m-xylene:TBA 4:1, temperature 150 C, total reaction volume 50 cm3, autogenous pressure. Table 2. Kinetic parameters of dehydration of TBA and alkylation reactions reaction dehydration of TBA alkylation of m-xylene with TBA alkylation of p-xylene with TBA alkylation of o-xylene with TBA rate constant, k or kSR (cm6 gcat-1 mol-1 s-1) at 150 C 1.16 2.11 2.53 3.17 frequency factor, k0 (cm6 gcat-1 mol-1 s-1) 35.7 1015 45.8 108 10.3 1010 15.4 1011 activation energy, E (kcal mol-1) 33.8 19.3 22.4 24.3

14

100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

Figure 1. Effect of different catalysts on conversion of TBA Reaction conditions: speed of agitation 1000 rpm, catalyst loading 0.04 g cm-3, mole ratio of m-xylene:TBA 4:1, temperature 150 C, total reaction volume 50 cm3, autogenous pressure. () UDCaT-5, () UDCaT-6, () UDCaT-4, () sulfated zirconia (S-ZrO2).
1.8 1.6

Concentration (mol cm ) 10

3 -3

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

Time (min)

Figure 2. Concentration profile of various products in alkylation of m-xylene with TBA Reaction conditions: catalyst UDCaT-5, speed of agitation 1000 rpm, catalyst loading 0.04 g cm-3, mole ratio of m-xylene:TBA 4:1, temperature 150 C, total reaction volume 50 cm3, autogenous pressure. () TBA, () 5-tert-butyl-m-xylene, () isobutylene, () di-tert-butyl ether, () 2,5-di-tert-butyl-m-xylene. 15

100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

Figure 3. Effect of speed of agitation on conversion of TBA Reaction conditions: catalyst UDCaT-5, catalyst loading 0.04 g cm-3, mole ratio of mxylene:TBA 4:1, temperature 150 C, total reaction volume 50 cm3, autogenous pressure. () 800 rpm, () 1000 rpm, () 1200 rpm.
100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

Figure 4. Effect of catalyst loading on conversion of TBA Reaction conditions: catalyst UDCaT-5, speed of agitation 1000 rpm, mole ratio of mxylene:TBA 4:1, temperature 150 C, total reaction volume 50 cm3, autogenous pressure. () 0.01 g cm-3, () 0.02 g cm-3, () 0.04 g cm-3, () 0.05 g cm-3. 16

18 16 y = 335.14x R2 = 0.9963

Initial rates (mol cm s ) 10

7 -3 -1

14 12 10 8 6 4 2 0 0 0.01 0.02 0.03 0.04


-3

0.05

0.06

Catalyst loading (g cm )

Figure 5. Plot of initial rate of reaction as a function of catalyst loading in liquid phase Reaction conditions: catalyst UDCaT-5, speed of agitation 1000 rpm, mole ratio of mxylene:TBA 4:1, temperature 150 C, total reaction volume 50 cm3, autogenous pressure.
100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

Figure 6. Effect of mole ratio of m-xylene:TBA on conversion of TBA Reaction conditions: catalyst UDCaT-5, speed of agitation 1000 rpm, catalyst loading 0.04 g cm-3, temperature 150 C, total reaction volume 50 cm3, autogenous pressure. () 1:1, () 3:1, () 4:1, () 5:1.

17

100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

Figure 7. Effect of temperature on the cracking of TBA Reaction conditions: catalyst UDCaT-5, speed of agitation 1000 rpm, catalyst loading 0.04 g cm-3, total reaction volume 50 cm3, autogenous pressure. () 130 C, () 140 C, () 150 C, () 160 C.
100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

Figure 8. Effect of temperature on alkylation of m-xylene with TBA Reaction conditions: catalyst UDCaT-5, speed of agitation 1000 rpm, catalyst loading 0.04 g cm-3, mole ratio of m-xylene:TBA 4:1, total reaction volume 50 cm3, autogenous pressure. () 130 C, () 140 C, () 150 C, () 160 C. 18

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

-3

Time (min)

Figure 9. Concentration profile of various products in alkylation of p-xylene with TBA Reaction conditions: catalyst UDCaT-5, speed of agitation 1000 rpm, catalyst loading 0.04 g cm-3, mole ratio of p-xylene:TBA 4:1, temperature 150 C, total reaction volume 50 cm3, autogenous pressure. () TBA, () 2-tert-butyl-p-xylene, () isobutylene, () di-tert-butyl ether, () 2,5-di-tert-butyl-p-xylene.
1.8 1.6

Concentration (mol cm ) 10

3 -3

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

Time (min)

Figure 10. Concentration profile of various products in alkylation of o-xylene with TBA Reaction conditions: catalyst UDCaT-5, speed of agitation 1000 rpm, catalyst loading 0.04 g cm-3, mole ratio of o-xylene:TBA 4:1, temperature 150 C, total reaction volume 50 cm3, autogenous pressure. () TBA, () 4-tert-butyl-o-xylene, () isobutylene, () di-tert-butyl ether, () 4,6-di-tert-butyl-o-xylene. 19

12

10 y = 0.2261x R2 = 0.9866 y = 0.0544x 6 R2 = 0.9925 y = 0.0356x R2 = 0.9897

XA/1-XA

y = 0.0152x R2 = 0.9798

0 0 20 40 60 80 100 120 140

Time (min)

Figure 11. Validation of mathematical model for dehydration of TBA () 130 C, () 140 C, () 150 C, () 160 C.
3.0

2.5

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

2.0 y = 0.0805x R2 = 0.9914 y = 0.0437x 1.0 R2 = 0.9916 y = 0.0277x R2 = 0.9972 y = 0.0186x R2 = 0.9948 0.0 0 5 10 15 20 25

1.5

0.5

Time (min)

Figure 12. Validation of mathematical model for alkylation of m-xylene with TBA () 130 C, () 140 C, () 150 C, () 160 C.

20

3.0

2.5

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

2.0 y = 0.0937x 1.5 R = 0.9847


2

1.0

y = 0.0525x R = 0.9801 y = 0.0281x R = 0.9839 y = 0.0183x R = 0.9948


2 2 2

0.5

0.0 0 5 10 15 20 25

Time (min)

Figure 13. Validation of mathematical model for alkylation of p-xylene with TBA () 130 C, () 140 C, () 150 C, () 160 C.
3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 0 5 10 15 20 25 y = 0.1045x R = 0.9698 y = 0.0657x R = 0.9594 y = 0.0334x R = 0.9715 y = 0.0172x R = 0.9839
2 2 2 2

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

Time (min)

Figure 14. Validation of mathematical model for alkylation of o-xylene with TBA () 130 C, () 140 C, () 150 C, () 160 C.

21

2.00

1.00

0.00

ln k

-1.00

-2.00

-3.00

-4.00 2.12E-03 2.16E-03 2.20E-03 2.24E-03


-1

2.28E-03

2.32E-03

1/T (K )

Figure 15. Arrhenius plots () dehydration of TBA, () alkylation of m-xylene with TBA, () alkylation of pxylene with TBA, () alkylation of o-xylene with TBA.
100

80

Conversion (%)

60

40

20

0 0 20 40 60 80 100 120 140

Time (min)

Figure 16. Reusability of the catalyst Reaction conditions: catalyst UDCaT-5, speed of agitation 1000 rpm, catalyst loading 0.04 g cm-3, mole ratio of m-xylene:TBA 4:1, temperature 150 C, total reaction volume 50 cm3, autogenous pressure. () fresh catalyst, () first reuse, () second reuse.

22

You might also like