Anda di halaman 1dari 21

Catalysis Surveys from Asia, Vol. 9, No. 2, May 2005 ( 2005) DOI: 10.

1007/s10563-005-5997-x

117

Synergism of clay and heteropoly acids as nano-catalysts for the development of green processes with potential industrial applications
Ganapati D. Yadav*
Department of Chemical Engineering, University Institute of Chemical Technology (UICT), University of Mumbai, 400 019, Matunga, Mumbai, India

A large number of inorganic oxides, mixed oxides, including alumina, silica, titania, zirconia, zeolites, carbon, clays and ion exchange resins have been employed as both supports and solid acid catalysts. Clay structure collapses at high temperatures and has to be stablilized. The thermal stability and pore size issues were addressed by pillaring of clays. Natural clays are acid treated or ionexchanged to be used as solid acids, and the acidity and pore structure are dependent on the treatment methodology. Another important class of catalysts is the heteropoly acids (HPA), which are employed as homogeneous or heterogeneous catalysts, having both acid and redox properties. The focus of the current paper is to address the work done in our laboratory on the synergism between clays and heteropoly acids for the development of green processes. A comparison is also provided for the activity of these catalysts with other solid acids. Several alkylation, acylation, isomerization, condensation, dehydration, esterication, nitration and oxidation reactions which are useful in a wide spectrum of industries such as bulk, intermediates, dyes, plasticizer, pharmaceuticals, perfumes, avours and other ne chemicals were investigated to improve the selectivity of the desired products.

1. Introduction Green chemistry and nano-catalysis, the vibrant research areas of today, have been studied by several researchers over decades, without ever using these terminologies. Due to the environmental issues aecting all societies, the emphasis on green aspects of chemical processes will be appreciated. Many solid catalysts have been traditionally synthesized in nano sizes per se or created as nano materials inside a porous matrix. A large number of inorganic oxides, mixed oxides, including alumina, silica, titania, zirconia, zeolites, and clays have been employed as both supports and catalysts [16]. Amongst them, clays have resurfaced to the fore as solid acids. Although clays were the original FCC catalysts, they were replaced by thermally more stable silicaalumina in 1940s, and thereafter in 1960s most protably, by zeolites of dierent hue. Clays structure collapses at high temperatures and must be stabilized. The thermal stability and pore size issues were addressed through pillaring of the clays, which led to a repertoire of new applications. Natural clays are acid treated or ion-exchanged to be used as solid acids, and the acidity and pore structure are dependent on the treatment methodology [57]. Other important class of catalysts are heteropoly acids (HPA), having both redox and acid properties [716]. There is a continuous search for better and cheaper catalytic materials, with precise pore
* To whom correspondence should be addressed. E-mail: gdyadav@yahoo.com, gdyadav@udct.org

structure at micro or meso level. Inorganic solids possessing high-surface areas such as alumina, silica, titania, zirconia, ceria, zeolites, clays, etc. are overwhelmingly used as catalysts or supports for dispersing active centres and reagents [57,1719]. We have been investigating dierent aspects of green processes with benign solid acid catalysts, which have direct industrial relevance in reneries, petrochemicals, pharmaceuticals, rubber chemicals, dyestu, agrochemicals, perfumery and avour chemicals. Clays, heteropoly acids supported on clays, sulphated zirconia, UDCaT series of catalysts (named after University Department of Chemical Technology, UDCT; the former name of authors institute) and ion exchange resins have been evaluated in several processes. However, the focus here is to address the synergism between clays and heteropoly acids in comparison with other solid acids for the development of green processes carried out in this laboratory. The importance of each reaction is also highlighted.

2. Clays Clays are ubiquitous and cheap. Important characteristics of clays responsible for the varied applications are particle size, surface chemistry, particle shape, surface area, and other properties specic to a particular application such as viscosity; colour; plasticity; dry and red strength; absorption and adsorption; abrasion; among others. Clays are also used as binders for
1571-1013/05/05000117/0 2005 Springer Science+Business Media, Inc.

118

G. D. Yadav/Synergism of clay and heteropoly acids

zeolites, which increases the stability of zeolites for processes carried out at higher temperature and pressure. Clay minerals are essentially crystalline materials of very ne particle size (<2 lm), with hydrous layer silicates of the so-called phyllosilicate family, in which extremely stable SiO4 tetrahedral structural units polymerize to form 2-D sheets, which is a result of the sharing of three of the oxygens at the corners of the tetrahedra. The oxygen, which can be imagined as being at the base of an equilateral triangle (basal oxygens), forms repeating patterns of regular hexagon. Clays are thermally unstable and can be stabilized through special treatment, for instance, through pillaring. Smectite clays are very frequently used in the synthesis of pillared clays. The smectite group, containing all low charge 2:1 minerals, is divided into saponites (which are trioctahedral) and montmorillonite (which are dioctahedral). The commonest example of the subgroup of smectite is saponite, which has the ideal formula : Mx+ (Si8)x Alx) IV (Mg)6VI O20(OH)4, in which M+ is the charge balancing cation at the surface and x is the layer charge constrained by 0.6< x < 1.2. Hectorite clay is also in this subgroup wherein Mg is substituted by Li in the octahedral layer. Montmorillonite is an important end member of the species of dioctahedral smectite in which Al is replaced by Mg in the octahedral layer to produce a charge decit (x) of about 0.6 and 1.2 (Figure 1). Montmorillonite clays cover a wide range of minerals within this subgroup and are represented by two extreme compositions in which the charge is either completely tetrahedral in origin (i.e., bidelite, Mx+ (Si8)x Alx)IV (Al4)VI O20(OH)4. ) or completely octahedral (i.e., montmorillonite ,Mx+ (Si8)IV (Al4)x Mgx)VI O20(OH)4). Mg2+ is necessary for the formation of montmorillonite. Mg2+ increases the size and the charge of the octahedral sheet and tends to decrease the layer strain. 2.1. Acid Treatment of Clays Montmorillonite clays have very poor thermal stability. When the clay is heated above 200 C, water from

Figure 1. Structure of montmorillonite clay.

its interlayers is removed which results in collapsing of the solid structure. Surface area and activity of clay is also reduced due to such heat treatment. To increase the thermal stability and improve surface properties, montmorillonite clays are often modied. Clay in its natural form does not show much catalytic activity. It needs to be activated to convert it to the desired form. Al3+ is the source of acidic nature of the surface in aluminosilicates. The Al3+ ions are electron pair acceptors and function as aprotic acids of Lewis acid type. Various modications of clay such as acid activation, pillaring etc. alter the structure of clay and thereby aect the acidity. During acid activation, clay is treated with an acid to replace interlamellar cations with protons and the acid leaches Al3+ from the octahedral layers. Dealumination develops mesopores in the layers and contributes to high surface area. The protons exchanged for interlamellar ions and the leached hydrated alumina occupying the cation exchange sites result in enhanced acidity [6,2025]. It is now accepted that acid attack on the clay structure progresses inward from both the edge and the basal surface of the clay platelets, leaching cations, particularly Mg if present, from the octahedral sheet. Acid activation causes little damage to the silicate layer, and consequently, the structure in the centre of platelet, at the limit of acid attack remains intact. The rate of dissolution of octahedral sheet is a rst order process that increases not only with increasing concentration of acid, temperature, and the contact time but also with increasing Mg content in the octahedral sheet. Clays with organic cations (possessing long-chain alkyl groups) on the surface impart a hydrophobic quality to the mineral surface and are able to sorb hydrophobic molecules. Acid treatment occurs even under mild conditions. When montmorillonite from Wyoming, USA was treated with 0.01 N HCl at room temperature, 60% of exchanged sites were occupied by Al3+ [26]. Acid treatment also reveals the unexposed parts of the clay thereby increasing the clay surface area and porosity. Clay is converted into NH+ 4 form via ion exchange followed by heat treatment at 550 C. During heat + treatment, NH+ ion on the 4 decomposes to leave H clay and imparts good catalytic activity [27]. After acid treatment clay structure becomes rigid and the amorphous character becomes predominant. Its property of swelling is also reduced. Eect of acid activation with organic acids on the structure and properties of layered aluminosilicates has received only scant attention. Carboxylic acid cause mild activation of clays without the loss of crystallinity which usually occurs on acid activation of hectorite with dierent mineral acids [28]. In the acid activation of hectorite with dierent mineral acids, the rate of dissolution increased with the concentration of the acid and for the specic molar concentration, the

G. D. Yadav/Synergism of clay and heteropoly acids

119

rate of dissolution varied in the order HCl (max.) > HNO3 > H2SO4 (min). The conditions maintained during acid activation such as concentration of acid and temperature inuence the nal acidity [29]. Montmorillonite KSF, a commercially available proton exchanged clay is prepared by treatment with cold dilute acid [30]. The ions associated with octahedral sheet are removed during activation with hot concentrated acid. The extent of leaching depends on the duration of activation and this in turn results in dierent products, for example, montmorillonite K-10, KP-10 [11].

2.2. Clay catalysis Clays are widely used as catalysts in synthetic organic chemistry for alkylation (C, O, N, S) and acylation of aromatics, esterication, isomerization, cracking, hydration and dehydration, cyclization, condensation, nitration, NOx reduction and hydrodesulphurization [5,7]. In comparison with ion exchange resins which can be operated upto 120 C, clays possess higher thermal stability and thus can be safely used up to 280 C. Also clays are inexpensive compared to other new generation solid acid heterogeneous catalysts. K10, KSF, KS, Filtrol-24 and KP10 are some examples of commercially available acid treated clays of monmorillonite type. They have proved to be good catalysts in modied forms. For example, ion exchanged and salt impregnated forms of monmorillonite are used in numerous industrially important reactions such as alkylation [17], polymerization [18] acylation [19], etherication [20], Diels Alder reaction [21], etc.

them most preferred structure for catalysis than any other form of structure [1016]. HPAs are reported as the most suitable catalysts in acid catalysed and oxidation reactions [721]. There are three dierent types of catalysis by unsupported HPAs, which make them versatile in chemical reactions: (i) surface catalysis, (ii) bulk type I catalysis (pseudo-liquid catalysis), and (iii) bulk type II catalysis. Dodecatungstophosphoric acid (DTP) is the most stable among all HPAs and is commonly used for acid catalysis since it possesses the highest Bronsted acidity [810]. The Keggin unit is composed of a central PO4 tetrahedron, surrounded by 12 WO6 octahedra (Figure 2). However, low surface area, rapid deactivation and relatively poor stability are some of the major problems associated with HPAs in bulk form. The structure has a net ()3) charge, which requires three cations to satisfy the electroneutrality. If these cations are proton, the material functions as a Bronsted acid catalyst. One of the unique properties of HPAs is their ability to tune catalytic activity by substitution of other metals in place of tungsten atom in the Keggin structure. Substitution of Mo for W in the structure provides a good catalyst for oxidation chemistry. Replacement of protons with cesium increases the surface area and thermal stability of HPA. Acid strength of HPA varies with compositions. HPAs composed of W are more acidic than those composed of Mo. The thermal stability of hydrogen forms of HPAs changes with heteroatom, polyatom and polyanion structure as follows: H3PW12O40 (dodecatungstophoshoric acid, DTP) > H4SiW12O40 (dodecatungstosilicic acid, DTS) > H3PMo12O40 (dodecamolybdophosphoric acid, DMP) > H4SiMo12O40 (dodecamolybdosilicic acid, DMS) Thus, DTP and DTS are strong acids in which protons are dissociated completely from the structures in aqueous solution [821]. 3.1. Catalysis by heteropoly acids HPAs can be employed in a number of states as catalysts (Table 1). In homogeneous liquid phase catalysis, the advantages of the HPAs are more distinctive due to their low volatility, low corrosiveness, high acidity, activity and exibility. Upon comparing HPAs ` -vis their reactivity in Bronsted with sulfuric acid, vis-a acid-catalysed liquid-phase reactions, it has been found that the activity/mole proton of HPA is higher by a factor of 3100, depending on the reaction conditions. HPAs are most commonly used in the hydration of olens, the reaction rate being dependent on the concentration of HPA. Other reactions such as dehydration of 1,4-butanediol, polymerisation of benzyl alcohol, isomerization, alkylation, esterication of alcohols and olens, synthesis of ethers, etc. have been successfully catalysed [5, 7, 921]. They are also used in conjunction with phase transfer catalysts.

3. Heteropoly acids (HPA) Heteropoly compounds are known as condensates of dierent oxoacids [716]. In a typical heteropoly compound, a heteroatom, which is located in the centre of the HPA molecule, can be chosen from among a wide variety of group I to VII elements in the periodic table. The atoms which are bound to the central heteroatom through oxygen atoms are called polyatoms or addenda atoms. They are represented by transition metal such as Mo, W, and V. Amongst all these known structures for a HPA molecule, the catalyst material is limited exclusively to the Keggin-type and Dawson-type structures owing to their availability and chemical stability [8]. The Keggintype HPAs have received the most attention due to the ease of preparation and strong acidity. Not only water but also certain polar organic solvents can enter and leave the bulk of the solid. Hydration of HPA leads to increase in proton activity in it. Further, the mobility of Keggin anion even in dehydrated HPA crystal, makes

120

G. D. Yadav/Synergism of clay and heteropoly acids

Figure 2. Keggin structure of dodecatungstophosphoric acid.

3.2. Pseudo-liquid phase HPA catalysis Pseudo-liquid phase HPA catalysis has characteristics of both dissolved and solid HPA [11,16]. Pseudo-liquid phase HPA is found under conditions intermediate to those for fully heterogeneous or homogeneous catalysis. Pseudo-liquid phase behaviour occurs most often with basic molecules like ammonia and pyridine, and with alcohols at lower temperatures. The HPA is present as a solid, sometimes as crystals on a carrier, containing a certain number of molecules of water of hydration. If the catalyst is fully dehydrated the pseudo liquid phase behaviour is lost for most compounds. However, pyridine and ammonia are still taken up by the bulk of the catalyst even at 300 C. Application of this typical behaviour is possible in reactions where the reactant(s) and product(s) have dierent hydrophobicities. A reaction well suited to the pseudo-liquid HPA catalysis is dehydration of alcohols to alkenes, wherein this selectivity is obtained because of the bypassing of the bulk of the catalyst by the hydrophobic alkane, leading to the possibility of attaining higher conversions per pass without losing selectivity [916].

HPAs catalyse oxidation reactions by several mechanistic pathways. Reactions such as dehydrogenation of isobutyric acid, oxidation of aldehydes are reported [5, 7, 916]. 3.3. Role of water of hydration in solid HPA catalysis The crystal form of HPA depends on the amount of water of crystallisation. The water molecules in HPA crystal are arranged as protonated aquo-cations, e.g., as 3+ H3(H2O)29 , in the face centered cubic lattice of H3PW12O40.29H2O. The water of crystallisation can be eliminated on heating as easily as the water in zeolite, whereby the proton activity in HPA crystal is increased, resulting in higher activity in acid catalysed reactions. However, unlike the rigid anionic zeolite framework, the Keggin anions are still mobile even in dehydrated HPA crystal, and not only water but also certain kinds of polar organic molecules can enter and leave the bulk of the solid, being accompanied by a volume change of the HPA crystal. Such a structural exibility of HPA crystal should be taken into account when employing HPA as a heterogeneous catalyst. HPA may be a kind of mixed

Table 1 Catalysis by heteropolyacids (Acid and Redox) Catalysis type Homogeneous catalysis Catalyst form pseudo-liquid HPA bulkHPA Dissolved HPA Dissolved HPA and phase transfer catalyst Dissolved HPA and supported phase transfer catalyst Supported on clays, alumina, silica, MCM-41, HMS, Carbon, etc. Locations phase(s) of reactants Gas Gas Liquid Liquid Liquidliquid Liquidliquid Gas Liquid Liquidliquid Gasliquid Gas

Heterogeneous catalysis

Supported pseudo phase

G. D. Yadav/Synergism of clay and heteropoly acids

121

metal oxide in terms of chemical composition but it must be strictly distinguished from ordinary mixed metal oxides, in that, HPA crystal is composed of Keggin anions which are arranged discretely. This structural features of HPA is deeply associated with its unique properties and catalytic functions in contrast to other mixed metal oxides and isopoly compounds such as paramolybdate. HPA used in aqueous or organic media retains its Keggin structure in the course of reactions. It has also been reported that the carbenium ions formed in acid catalyzed reactions are stabilized by heterpoly anion. Although HPAs are protonic acids in nature, they may exhibit Lewis acid like catalytic function through the stabilizing eect of heterpoly anion [30]. Heteropoly acids as catalysts in organic reactions reveal that in crystalline HPAs there are always number of molecules of water of hydration present (normally 6 per Keggin unit). The protons are localized between two water molecules, forming cations, or larger aggregates if more water molecules are present. So protons do not bind directly to the anion until the HPA is fully dehydrated, which is the case after heating to about 300 C. At higher temperatures beyond 300 C, two protons combine with an oxygen atom from the HPA to detach a water molecule and HPA molecule is left with oxygen vacancies and thus destabilizing it. The positions of protons in dehydrated HPAs are not well known and in all probability the protons are still quite mobile [30].

4. Heteropoly acids supported on inorganic oxides The disadvantage of using HPAs in the bulk form is that they form a homogeneous phase with the reaction medium and create problems related to homogeneous catalysis. Such type of problems can be solved by supporting the HPA on an inorganic support. The process gives many advantages such as easy separation from the reaction mixture, no euent problems, low consumption of reagents and above all, product selectivity can be advantageously exploited. Supported HPA catalysts are better because unsupported bulk HPAs have a low specic surface (15 m2 g)1) [7, 8, 21]. Procedures for supporting HPAs on silica, titania, alumina, etc. and characterization are covered in literature and reviewed [30]. Heterogeneous catalysis with HPAs is mostly performed at elevated temperatures. In many cases, the catalyst is present as a partial salt, the catalytic activity resulting from the non-exchanged protons or from the Lewis acidity of the added counter cations. The catalytic activity of supported HPAs depends on the type of carrier, the HPA loading, conditions of pretreatment etc. Acidic or neutral supports such as SiO2, active carbon, acidic ion-exchange resin, etc., are suitable as supports, the most often used being SiO2. Under these conditions the catalyst is quite stable, especially if water or another possible oxygen

donor is constantly present or when the reactants do not diuse easily into the bulk of the catalyst. Basic solids such as Al2O3 and MgO tend to decompose the HPA. Liquid-phase as well as vapour-phase reactions (dehydration of alcohols, Friedel-Crafts alkylations and de-alkylation) have been reported over supported HPAs. HPAs have been supported on dierent types of silica : MCM-41 [3033], Aerosil 380 [33], amorphous silica powder (Davison 62, 300 m2/g) [34] and SiO2 (Grace, 311 m2/g) [35]. Amongst all these catalysts, there is an evidence for the retention of Keggin structure on the SiO2 surface, with loadings ranging from 13 to 87 wt.% as conrmed by the IR and XRD studies. Attempts to improve the eciency and stability of HPAs have been made by using dierent supports including mesoporous silica, mesoporous aluminosilicates, alumina, carbon and zirconia [30,36]. Alumina and zirconia tend to decompose HPAs because of the basic nature of these supports resulting in deformation of the parent Keggin structure and thus reduction in overall activity [36]. The use of various supports such as silica, titania, calumina and zirconia for DTP in the preparation of dioctyl phthalate was investigated by us to observe that silica is the best support since it was neutral and retained the delity of DTP, whereas zirconia and titania were the least eective due to their lower surface areas and partly due to decomposition of DTP (Table 2) [36]. The use of hexagonal mesoporous silica (HMS) as a support has also been explored in detail and its merit is brought out in the synthesis of acetoveratrone [37]. Acetroveratrone (3,4-dimethoxyacetophenone) is an industrially important aromatic ketone, derived from veratrole (1,2-dimethoxybenzene) using polluting acids as catalysts. It is used for the synthesis of papaverine (1-(3,4-dimethoxybenzyl)-6,7-dimethoxyisoquinoline), an opium- alkaloid, which is a useful antispasmodics. Of late, several researchers have focused their attention to Hb and HY zeolites for acylation of aromatics and substituted aromatics, including anisole and veratrole [3135]. However, zeolites have been found to deactivate [3840] and there are three important parameters which govern the choice of a suitable catalyst, namely, activity, accessibility of the active sites without intra-particle diusion and reusability. Since acetoveratrone is bigger molecule, catalysts with pores in mesoporous region were thought to be primary candidate and the ecacy of 20% w/w DTP/HMS was studied [37]. The textural characteristics are given in Table 3. The initial rates were extremely high in the absence of a solvent, whereas the overall conversions were severely limited. There appeared to be an equilibrium or poisoning of the catalyst in the absence of solvent. In the presence of ethylene dichloride as a solvent, the nal conversion was much higher and the reaction progressed smoothly to give a conversion of 70% in 2 h at a mol ratio of 1:3 of veratrole to acetic anhydride at

122

G. D. Yadav/Synergism of clay and heteropoly acids

Table 2 Eect of support for heteropoly acids on the esterication of phthalic anhydride monoester with 2-ethylhexanol to prepare dioctyl phthalate [36] No. 1 Catalyst DTP/SiO2 Loading (%) w/w 30 Time (h) 1.0 2.0 3.0 1.0 2.5 1.0 1.5 1.0 2.0 Monoester conversion (%) 67.90 70.69 75.58 54.43 65.42 48.78 57.56 48..07 54.51

2 3 4

DTP/c-Al2O3 DTP /TiO2 DTP/ZrO2

30 30 30

Mole ratio monoester: 2-ethylhexanol=1:1; reaction temperature: 195 C; calcinations temperature: 650 C, 30% (w/w) DTP/support, catalyst loading: 2.5% (w/w).

60 C at the catalyst loading 0.015 g/cm3, which is far ` -vis other reported catalysts [37]. better vis-a

5. Clay supported reagents and Lewis acids Many reagents have been supported on clays to intensify rates of reactions, selectivities and easy workup [4150]. The unique property of swelling allows the intercalation of larger molecules, which is not possible to be exchanged either in the synthetic zeolites or in the conventional silica and alumina catalysts. Furthermore, the reasonable amount of negative charge that montmorillonite carries in the range 0.6 < x < 1.2 helps in the manipulation of the amount of exchange that has to be done to achieve the desirable results. Another important aspect in montmorillonite clays is the amount of hydration in the interlayers. The extent of swelling is related to the amount of water adsorbed in the interlayers, which in turn depends on the type of the exchangeable cation present. Besides water, organic molecules can also be intercalated and it too depends on the type of the exchangeable cation. Clays have been employed as supports with oxidants such as potassium permanganate, thallium (III) trinitrate, potassium dichromate and ferric nitrate. Potassium permanganatesupported clay is reportedly very eective in the oxidation of secondary alcohols to ketones [44]. Supported thallium (III) trinitrate is a useful reagent for the oxidative rearrangement of olens and enols. Clay supported copper nitrate and ferric nitrate, nicknamed claycop and clayfen are very useful for the oxidation of alcohols to aldehydes or ketones, benzoins to

benzils, and the oxidative coupling of thiols to disuldes [42,43]. Of late, transition metal halides supported on clays have been used as Friedel-Crafts alkylation catalysts [41]. These catalysts in addition to being heterogeneous require very mild reaction conditions. Zinc chloride, nickel chloride, magnesium chloride, and copper chloride supported on clay have been eectively used as catalysts for the alkylation [3844].

6. Synergism between heteropoly acids and acid treated clay K10 and Filtrol-24 are acid treated clay which are used as catalysts in a number of reactions in our laboratory among others [4856]. In our pioneering work, we supported dierent HPAs on K10 clay and studied the eect of various parameters on their activity in dehydration and condensation reactions [47,48]. In the case of K-10 montmorillonite clay, which was dried at 150 C before use, the surface area was 230 m2/g and the pore volume was 3.52 10)3 m3/kg, with pore sizes in the range of 4.515 nm. The HPA was initially dissolved in a suitable solvent, the volume of which was equal to the pore volume of the support material. The quantity of substrate to be supported and the choice of solvent was important for this matter. Once the desired quantity of substrate was loaded on to the support, heat treatment was given to remove water and other volatiles. Then it was calcined to activate the catalyst (Table 3). Eect of calcination temperature plays an important role during the preparation of supported HPAs. The properties of support and the precursor ought to be

Table 3 Surface area, pore volume and pore diameter of catalysts [56,58] Catalyst K-10 20% w/w DTP/K-10 20% w/w Cs2.5H0.5PW12O40/K-10 20% w/w DTP/HMS Sulfated zirconia Surface area (m2/g) 230 107 207 909 100 Pore volume (cm3/g) 0.36 0.32 0.29 0.538 0.115 Pore diameter (nm) 6.4 7.1 5.8 2.6 2.8

G. D. Yadav/Synergism of clay and heteropoly acids

123

taken into consideration during calcinations. The amount of HPA on clay was varied in the range of 5 40% as well as the eect of calcinations temperature from 150650 C was studied systematically on the activity and reusability of the catalysts. Thus, DTP, DTS, and DMP were supported on K-10 clay. HPA loading of 20% w/w on K10 clay with 285 C as the calcinations temperature was found to give the best reusable catalyst without any leaching. The leaching was studied independently by a well established procedure [48]. The catalyst was ltered and the reaction was allowed to continue which showed no further conversions. Also the ascorbic acid test was performed. Clay loses all water when it is heated beyond 200 C resulting in drastic reduction in the surface area. Due to the presence of heteropoly acids in the interlayers, the clay has the capacity to withstand higher temperatures. But at higher temperature, beyond 400 C, heteropoly acids sinter resulting in the loss of surface area and activity. There is also a probability of the formation of the lacunary heteropoly-acids. It has been found that the activity of HPA/K10 attained a maximum at a temperature of 285 C. Beyond this temperature there was observed a gradual loss in activity. 20% w/w DTP/K10 was found to be the best catalyst in a number of reactions of industrial relevance. In one of our pioneering papers, we have discussed the synergism of heteropolyacids with clays to give much higher activity and selectivity and this aspect was covered in a large number of papers subsequently with reference to activities of clays, pillared clays, sulfated zirconia, ion exchange resins and zeolites [48, 56]. The catalyst has been fully characterized and activity trends are explained. 6.1. Characteristics of 20% w/w DTP/K10 The surface area of catalysts measured by nitrogen BET method for K10 montmorillonite and DTP/K10 were found to be 230 and 107 m2/g, respectively. Since both the K10 support and catalysts were pre-treated in a similar way prior to analysis the reduction in surface area of DTP/K10 may be due to the blockage of smaller pores by active species. It appears the active spices are held in few junctions of such dimensions from where the access to smaller pores is denied, thereby leading to the

decrease in accessible surface area. The average particle sizes of both K10 and DTP/K10, as determined by image analysis, were in the range of 60 lm. The XRD studies of DTP/K10 had indicated that in the impregnation process, the clay has lost some of its crystallinity compared to K10. The cation exchange capacity(CEC) of the catalyst were determined by the method to observe that DTP/K10 possesses higher CEC than K10. This may be due to some additional surface protons, which come by way of heteropoly acid impregnation, which may play a role on account of easy availability for the exchange reaction (Table 4). In the case of DTP/ K10, there is a lot of disperse particles on the surface of the support K10 as revealed by the SEM photograph. FT IR studies of K10 showed that there was SiO and SiOAl linkage and the OH groups are bonded to the Al atoms, whereas in DTP/K10 are indicated the presence of H3O+ (Bronsted acidity) and linkage of phosphorous. 6.2. Characterisation of 20% w/w Cs2.5H0.5PW12O40/ K10 A method that could possibly enhance the HPA stability and activity in liquid phase reactions is to prepare catalysts in the form of alkali metal salts of HPA. Especially Cs salts of DTP have been reported to be better catalysts than DTP itself, but the particles of Cs2.5H0.5PW12O40 are very ne and separation of the catalyst from the liquid remains a problem. For example, partial substitution of protons of HPAs with Cs+ renders them with higher surface area and improved thermal stability. Cs2.5H0.5PW12O40 has been found to be the most active catalyst and due to its insolubility in liquids as polar as water; however, its small particle size limits its catalytic applications in commercial xed bed or slurry reactors [5860]. Unfortunately, direct impregnation of Cs2.5H0.5PW12O40 is not possible due to its insolubility in any kind of solvent and is a quite challenging. Thus, we developed a procedure to create nano-particles of Cs2.5H0.5PW12O40 in the pore space of K10 clay. We developed a procedure to prepare the nanosize catalyst [6061]. K-10 clay was rst impregnated with aqueous solution of Cs+ precursor (cesium chloride), dried at 110 C and calcined at 300 C. DTP was then

Table 4 Cation exchange capacity of modied clays [53] No. 1. 2. 3. 4. 5. 6. Catalyst K10 montmorillonite DTP/K10 Al-Fe/K10 Al pillared K10 Cr pillared K10 Zr pillared K10 CEC (meq/100 g) 35 39 20 16 18 18

124

G. D. Yadav/Synergism of clay and heteropoly acids

impregnated on the support, from a methanolic solution, dried at 110 C for 12 h and calcined at 300 C for 3 h. Thus, 20% w/w Cs2.5H0.5PW12O40/K10 was dried at 110 C for 12 h and then calcined at 300 C for 3 h. Bulk Cs2.5H0.5PW12O40 was prepared by adding the Cs precursor solution dropwise to the DTP solution while stirring. The resulting precipitate was dried at 110 C for 12 h and calcined at 300 C for 3 h. All catalysts were powdered and dried at 120 C for 3 h prior to their use. It was necessary to study the stability of Cs2.5H0.5PW12O40/K-10 in order to reuse the catalyst in liquid phase. Any leaching of the active sites from the catalyst would render it commercially unattractive. It was therefore decided to verify the possibility of leaching of Cs2.5H0.5PW12O40 from the support under severe conditions. When supernatant portion of Cs2.5H0.5PW12O40/K-10 catalyzed-reaction mixture was subjected to UV spectrum, it did not exhibit the presence of any absorption band at 265 nm assigned to Keggin 3) type PW12O40 6.3. Characterization of 20% w/w Cs2.5H0.5PW12O40/ K10 clay The catalyst was fully characterized and the details are reported recently by us [5859]. Crystallinity and textural patterns of the catalysts predicted from XRD of Cs2.5H0.5PW12O40/K-10 demonstrate that although the Cs2.5H0.5PW12O40 salt loses some of its crystallinity during deposition on K10 clay. The Keggin structure remains intact, which was also conrmed by the FT-IR analysis. The salt makes nanoparticles. A characteristic split in W=O band of Cs2.5H0.5PW12O40 suggested the existence of direct interaction between the Keggin polyanion and Cs+. The SEM micrographs revealed that both K10 and 20% w/w DTP/K10 samples possess rough and rugged surfaces, whereas 20% w/w Cs2.5H0.5PW12O40/K10 shows a smoother surface due to a layer of (Cs salt of DTP) over the external surface of K10. The BET surface area of 20% w/w Cs2.5H0.5PW12O40/K-10 was measured to be 207 m2g)1 and the pore volume and pore diameter were 0.29 cm3/g , respectively (Table 3). The nitrogen adsorpand 58 a tiondesorption isotherm for Cs2.5H0.5PW12O40/K10 showed that they have the form of Type IV isotherm with the hysteresis loop of type H3, which is a characteristic of mesoporous solid.

readily available on the plant site. Problems associated with transportation and handling of isobutylene in small quantities for such purposes prohibits its economic use. Hence it is of interest to nd other sources of isobutylene which provide safer and easier handling on site. Both methyl tert-butyl ether (MTBE) and tert-butanol were employed for the in situ generation of pure isobutylene over acidic catalysts for the alkylation and etherication reactions. The co-product of cracking of MTBE is methanol whereas that of tert-butanol is water which will have signicant bearing on the rates and selectivities. Thus the synthesis of MTBE from tertbutanol and methanol, dehydration of tert-butanol [51,52], and the alkylation reactions with MTBE and tert-butanol as feedstocks with several aromatics were investigated in our laboratory. 7.2. Synthesis of MTBE from methanol and tert-butanol Methyl-tert-butyl ether (MTBE), ethyl tert-butyl ether (ETBE), tert-amyl methyl ether (TAME) and tertamylethyl ether (TAEE) are important oxygenates. These ethers are increasingly being used as ecologically clean additives to the so called reformulated gasoline, which meets strict environmental anti-pollution measures. Yadav and Kirthivasan [51] studied the synthesis of MTBE from the condensation of methanol and tertbutanol over a number of catalysts including heterpoly acids (HPA) supported on clay (Table 5). 20% w/w DTP/ K-10 was found as a very novel catalyst with unique characteristics that have given a very high conversion and the best selectivity of MTBE, in comparison with 15 other non-zeolitic catalysts. The role of water, generated in-situ, as well as that added externally was investigated in detail [52]. The rate of reaction increases with water concentration up to a critical value, which was necessary to enhance the acidity of the catalyst. Again due to the increased density of Bronsted over Lewis acid sites in DTP/ K10, the enhanced activity in dehydration/condensation of the alcohols leading to MTBE was observed. 7.3. Dehydration of tert-butanol to isobutylene Dehydration of aliphatic alcohols is amongst the most thoroughly investigated reactions catalyzed by metal oxides which exhibit wide ranges of acid and base strengths [62,63]. It is pertinent to understand the role of water on the rate of dehydration of tert-BuOH, when heteropoly acids are used as catalysts. The dehydration of tert-butanol was studied using 20% DTP/K 10 at 85 C in an autoclave. It was seen that the rate increased as the concentration of water was increased by small amounts and with further increase in the water concentrations, resulted in reduction in the rate. That is, at a mole ratio of 1:4 of water to tert-BuOH, there was an appreciable increase in the rate of dehydration. It can be surmised that a small amount of water increases the

7. Applications of 20% w/w DTP/K10 in commercially relevant reactions 7.1. Relevance of MTBE and tert-butanol as sources of isobutylene There are several medium and small scale manufacturing processes where isobutylene is required but not

G. D. Yadav/Synergism of clay and heteropoly acids Table 5 Activities of the catalysts used for the synthesis of MTBE from tert-butanol and methanol [51] # 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
a

125

Catalyst Na+ Montmorillonite Al3+ ILC (5 meq Al3+/g, calcined at 450 C) Al3+ ILC (25 meq Al3+/g, calcined at 450 C) Al3+ ILC (5 meq Al3+/g, dried at 120 C) Zr2+ ILC Cr2+ ILC Sulfated Zirconia K-10 Montmorillonite 30 wt % DTP/silica, calcined at 650 C 30 wt % DTP/silica, calcined at 285 C 20 wt % DTP/K-10 clay, calcined at 285 C Amberlyst-15 20 wt % DTP/carbon, calcined at 285 C HZSM-5 20 wt % DTS/K-10 clay, calcined at 285 C 20 wt % DMP/K-10 clay, calcined at 285 C

Conversion of t-BuOH,% 21 24 34 25 52 44 66 56 08 55 71 74 00 77 74 76

Selectivitya to MTBE,% 33 37 73 84 94 95 93 70 100 76 99 86 0 52 91 86

The byproduct is isobutene; t-BuOH: 82.5 mmol; MeOH: 165 mmol; 1,4-Dioxane: 34 g, Catalyst loading: 2.2% w/w of reaction mixture; reaction temperature: 85 C; speed of agitation: 1000rpm; DTP: dodeca tungstophosphoric acid; DTS: dodeca tungstosilicic acid; DMP: dodeca molybdophosphoric acid; reaction time: 6 h except for Amberlyst-15 for 2 h.

acidity of the catalyst, whereas a further increase in the water content poisons it. 20% DTP/K-10 was calcined at 285 C, which could be sucient to drive-away the water of crystallisation. Thus, the lacunary phase of DTP was lost. At temperatures above 300 C, the DTP crystallites form aggregates leading to decrease in surface area. This sort of thermal treatment leads to a permanent deformation. However, at lower temperatures such loss in activity is not observed. In the presence of aqueous media, DTP takes up water to restore its lacunary phase, thereby rendering activity. The increase in activity could be due to the formation of lacunary HPA formation that increases the turnover number by opening itself up. 7.4. Synthesis of TAME from tert-amyl alcohol with methanol Synthesis of alkyl-tert-alkyl ethers by reactions of isoolens with alcohols is an ecient process. tert-Amyl methyl ether (TAME) with an octane number of 106 which is almost the same as that of MTBE (109) can be employed as a possible high-octane additive to motor fuels, since MTBE is caught in a controversy of politics of ethanol producers and reneries, particularly in the US and has gone out of favour in the Western countries. A variety of catalysts were evaluated by us including Amberlyst-15, Amberlyst-36, DTP, Filtrol-24, K-10 and 20% DTP/K10 [64]. DTP/K10 does not give good selectivity to TAME and majority of olens were formed in this reaction although conversions were high. Amberlyst-36 was a better catalyst from the view point of selectivity (Table 6). The parallel reactions of tertamyl alcohol adsorbed on the sites were found to control the overall rate of reaction which led to the

formation of TAME, 2-methyl-butene-1 (2MB1) , and 2-methyl-butene-2 (2MB2). The reaction follows pseudo rst-order kinetics at a xed catalyst loading. The selectivity of TAME over 2MB1 and 2MB2 is also found to be consistent with the model. 7.5. Alkylation reactions with MTBE and tert-butanol Alkylation of phenol and substituted phenols using a variety of olens and alcohols and ethers using acidic catalysts produces a valuable industrial chemicals. 7.5.1. Alkylation of phenol with MTBE and tert-butanol 2-tert-Butylphenol is an intermediate for pesticides, fragrances and other products whereas 4-tert-butylphenol is used to make phosphate esters, fragrances, oil eld chemicals and demulsiers. 2,4-Ditert-butylphenol is an intermediate for antioxidants and 2,6 di-tert-butylphenol is used as an antioxidant intermediate and in pharmaceuticals. tert-Butylated phenols are generally prepared by reacting phenol with pure isobutylene gas or C4 fraction of naphtha, by using a liquid acid catalyst, giving wide product distribution. Both C- and O-alkylation are possible depending on reaction conditions such as temperature, source of isobutylene and type of the catalyst. Yadav and Doshi [65] have studied several solid acids, majority of which were based on montmorillonite K10 clays using both MTBE and tert-butanol as alkylating agents in a pressure reactor under autogenous pressure. 20% w/w DTP/K10 was found to be the most active catalyst with both MTBE and tert-butanol at 150 C (Table 7). The rate of alkylation is higher with MTBE for which the conversion of phenol was 68% compared to 55% with tert-butanol. This can be

126

G. D. Yadav/Synergism of clay and heteropoly acids Table 6 Synthesis of TAME from reaction of tert-amyl alcohol with methanola [64]

Catalyst Amberlyst-15 Amberlyst-36 DTP Filtrol-24 K-10 20% DTP/ K-10

Conversion (%) 82 86.8 74.3 37 51.2 58.3

% yieldb of TAME 44.0 44.7 9.4 25.0 6.1 5.6

Selectivityc 0.644 0.701 0.0762 0.123 0.0325 0.0346

a tert-Amyl alcohol:methanol: 1:2; catalyst loading: 0.02 g/cm3; time: 4 h; temperature: 70 C speed of agitation: 800 rpm; solvent: 1,4-dioxane. b Yield= amount of TAME formed/amount of tert-amyl alcohol reacted. c Selectivity = rate of TAME formation/rate of formation of 2-methyl-butene-1 and 2-methyl-butene-2.

related to the eect of the co-product and its vapour pressure under operating conditions. The co-product water coming from tert-butanol leads to variable activities depending on the amount used or generated and it also changes pore structure. The clay is hydrophilic in nature. The dehydration is more likely to be aected by equilibrium than cracking of MTBE. The 2-/4-TBP ratio was marginally higher with tertbutanol as catalyst because the rate of dialkylation was lower with tert-butanol as alkylating agent. No 2,6-DTBP was formed with tert-butanol. A typical second order kinetics was found to prevail [65]. 7.5.2. Butylated hydroxytoluenes from p-cresol and MTBE Butylated hydroxytoluene (BHT) is a well known antioxidant and the basic raw material for the manufacture of oil-soluble phenol-formaldehyde resins. It is conventionally prepared by the acid catalysed reaction of isobutylene with p-cresol. The reaction is consecutive with formation of 2-tert-butyl-p-cresol followed by 2,6-di-tert-butyl-p-cresol which is popularly called BHT. The alkylation of aromatic substrates is important in many sectors of the chemical industry. Traditional methods of manufacture based on environmentally hazardous catalysts such as

aluminium chloride are increasingly being replaced by more benign alternatives but there remains a great need for new catalysts. Solid acids are especially popular in this context and Yadav et al.[66] reported a new material that brought out the advantages of a mesoporous inorganic solid UDCaT-1 (good thermal stability and molecular diusion rates) with those of a known strong solid acids, in the important alkylation of p-cresol using MTBE. This oers the additional environmental advantage of using a relatively safe in situ source of isobutylene rather than the gas itself. DTP/K10 was also an excellent catalyst (Table 8) [66]. 7.5.3. Alkylation of hydroquinone with MTBE and tertbutanol The alkylation of hydroquinone yields industrially important compounds. Acid catalysis can be employed with appropriate reaction conditions to get either C- or Oalkylated products which are useful in a variety of industries. Among C-alkylation, tert-butylhydroquinones are widely used as antioxidants and stabilisers for oils, fats, plastics and rubbers. 2-tert-Butylhydroquinone is a very important precursor for its use in pharmaceuticals and in developing photographic plates. Generally it is prepared by alkylation of hydroquinone with isobutylene. Twenty percent of DTP/K10 was found to be a very e-

Table 7 Ecacy of dierent catalysts in alkylation of phenol with MTBE [65] # Catalyst Conversion of phenol (%) Selectivity (%) 2-TBP 1 2 3 4 5 6 7 8 DTP(20%)/K10 DTP K10 ZnCl2(20%)/K10 Al exchanged K10 Zr exchanged K10 Cr exchanged K10 S-ZrO2 68 55 52 32 20 14 11 30 38 4 33 38 48 52 59 40 4-TBP 38 5 39 42 45 45 41 48 2,4-DTBP 22 86 25 19 7 3 10 2,6-DTBP 2 5 3 1 2 2-/4-ratio 1.0 0.8 0.84 0.90 1.06 1.15 1.43 0.83

Phenol:MTBE 1:2 mole; catalyst loading 0.04 g/cm3; temperature 150 C; speed 1000 rpm; Time 4 h, autogenous pressure 200 psi.

G. D. Yadav/Synergism of clay and heteropoly acids Table 8 Activity of various catalysts for the alkylation of p-cresol with MTBE [66] Catalyst 1. 2. 3. 4. 5. 6. UDCaT-1 Indion 130 Filtrol-24 Sulfated zirconia K-10 20% w/w DTP/K-10 Conversion (%) 45 39 19 15 12 30

127

% Selectivity to 2-tert-butyl-p-cresol 97 92 96 91 96 96

MTBE (220 mmol); p-cresol (220 mmol); Catalyst loading = 3.5% of reaction mixture; T = 100 C; speed = 700 rpm; reaction time 3 h; Byproducts formed are isobutylene, diisobutylene, triisobutylene and methanol.

cient and novel catalyst in comparison with several others (Table 9) for alkylation of hydroquinone with dierent alkylating agents such as methyl-tert-butyl-ether (MTBE) and tert-butanol at 150 C in an autoclave. The rate of alkylation with MTBE was much faster than that with tert-butanol. The reaction follows a typical second order kinetics at a xed catalyst loading with weak adsorption of both the species. The energy of activation was found to be 19.34 kcal/mol [67]. 7.5.4. Alkylation of catechol with MTBE Alkylation of dihydroxybenzenes is a very important reaction from industrial point of view since tert-Butylated dihydroxybenzenes are used as antioxidants, dye developers and stabilisers for fats, oils, plastic, rubbers etc. Commercial alkylation of aromatic compounds requires severe conditions since it proceeds through via electrophilic substitution. The synthesis of tert-butylated dihydroxybenzenes from catechol and resorcinol with MTBE has been carried out in presence of variety of solid acid catalysts. For instance, 4-tert-butyl catechol is used as a polymerisation inhibitor during the manufacture and storage of monomers such as styrene and butadiene. Also, 4-alkyl catechols are less toxic as compared to biologically active 3-alkyl catechols. Yadav et al. [68] have addressed the use of solid acid catalysts in the synthesis of tert-butyl catechols and tertbutyl resorcinol from alkylation of catechol and resorcinol with MTBE. Among the variety of catalysts

used, 20% w/w DTP/K10 was found to be the best catalyst (Table 10). The eects of various parameters on the rates of reaction were studied systematically, and the reactions were found to be kinetically controlled. The best reaction conditions for catechol and resorcinol were, 20% DTPA/K10, 150 C, mole ratio of catechol: MTBE= 1:3, catalyst loading 0.015 g/cm3. The reaction mechanism involves weak adsorption of MTBE on the catalyst followed by surface reaction with catechol leading to typical second order kinetics at a xed catalyst loading. The value of energy of activation was found as 8.86 kcal/mol for the alkylation of catechol. The catalyst is reusable. This is once again another example of a green process. 7.5.5. Alkylation of resorcinol with MTBE Alkylation of resorcinol was also carried out using MTBE as an alkylating agent over 20% DTP/K10 as a catalyst under otherwise similar conditions as were used in the case of catechol. The conversion of resorcinol was 97% in 4 h with a selectivity of 100% towards 4-tert-butyl resorcinol. It was observed that the alkylation of resorcinol was faster as compared to alkylation of catechol. This is due to the fact that resorcinol is a more active species than catechol [68]. 7.5.6. Alkylation of anisole with MTBE Alkylation of anisole is a very important reaction from industrial point of view since tert-butylated ani-

Table 9 Eect of various catalysts on alkylation of hydroquinone with MTBE [67] No. Catalyst Conversion (%) Selectivity of (%) 2-TBHQ 1. 2. 3. 4. 5. 6. 7. DTP DTP(20%)/K10 DTS(20%)/K10 K10 montmorillonite ZnCl2(20%)/K10 S-ZrO2 DTP/ZnCl2/K10 80 46 42 29 25 26 24 100 72 42 100 100 100 89 2,6-DTBHQ 28 58 11

Hydroquinone:MTBE 1:3 mole; cat.load 0.014 gm/cc; temp 150 C; speed 1000 rpm; solvent 1,4 dioxan; TBHQ: tert-butylhydroquinone; DTBHQ: di-tert-butylhydroquinone.

128

G. D. Yadav/Synergism of clay and heteropoly acids Table 10 Ecacies of various catalysts on the reaction of catechol with MTBE [68]

Catalyst

Conversion catechol (%)

Selectivity of (%) 4-TBC 3,5-DTBC 8 18 25 22 13 12

DTP 20 w/w DTP%/K10 20% w/w DTS/K10 Filtrol-24 K10 S-ZrO2

85 42 38 39 28 22

92 82 75 78 87 88

Catechol:MTBE = 1:3; catalyst loading = 0.015 g/cm3; temperature = 150 C; time = 4 h; speed of agitation = 800 rpm; solvent = 1, 4-dioxane; autogeneous pressure = 0.7 MPa; DTP: Dodecatungsto phosphoric acid; DTS: Tungsto silicic acid;4-TBC: 4-tert-butyl catechol; 3,5DTBC: 3,5-di-tert-butyl catechol.

soles are used as antioxidants, dye developers and stabilisers for fats, oils, plastic, rubbers etc. Commercial alkylation of aromatic compounds requires severe conditions since it proceeds through electrophilic substitution. The conventional acid catalysts like AlCl3, ZnCl2, and FeCl3 etc., are required to be employed in excessive amounts to carry out such reactions. Among the variety of catalysts used 20% DTP/K-10 was found to be the best catalyst (Table 11). The eects of various parameters on the rates of reaction were studied systematically and the reaction was found to be kinetically controlled. The reaction mechanism involves weak adsorption of MTBE on the catalyst followed by surface reaction with anisole leading to typical second order kinetics at a xed catalyst loading. The kinetic constants are determined [68]. 7.5.7. Alkylation of aniline with MTBE and tert-butanol tert-Butylanilines are used in pharmaceuticals, pesticides, plastics, additives and dyes. Among alkylated anilines, tert-butylanilines are used in pharmaceuticals, pesticides, plastics, additives and dyes. They are usually prepared by reacting aniline in presence of liquid acid catalysts with pure isobutylene or C4 fraction from naphtha crackers containing isobutylene. Alkylation of aniline is an acid catalyzed consecutive reaction with formation of N-alkylaniline in the rst step which gets further dialkylated to di-alkylaniline under suitable

conditions. Twenty percent w/w DTP/K10 was found to be the most active for alkylation of aniline with both MTBE and tert-butanol among others (Table 12). Only C-alkylated products were obtained. The yield of monoalkylated products was over 84% with a selectivity of 53% to 2-tert-butylaniline with MTBE at 175 C. With tert-butanol only the mono-alkylated products are obtained at 150 C with equal isomer distribution. A typical second order reaction kinetics was suitable [69] . 7.5.8. Synthesis of butylated hydroxyanisoles by akylation of 4-methoxyphenol with MTBE Butylated hydroxy anisoles, are very valuable and dominant food anti-oxidants and they are produced by highly polluting Friedel-Crafts alkylation of 4-methoxyphenol using homogeneous acids or from hydroquinone in a two-step reaction using alkylation followed by etherication using dimethyl sulphate under alkaline conditions. The activities of Filtrol-24, K-10 montmorillonite clay, 20% w/w DTP/K10, sulphated zirconia (S-ZrO2) and cation exchange resin Deloxane ASP in the alkylation of 4-methoxyphenol with MTBE were evaluated (Table 13). The order of catalytic activity of the solid acid catalysts studies is [70]: Filtrol-24 > 20% w=w DTP=K-10 > Deloxane ASP commercial resin > K 10 Montmorillonite clay > S-ZrO2

Table 11 Ecacies of various catalysts on alkylation of anisole with MTBE [68] Catalyst Conversion (%) Selectivity of (%) 4-TBA DTP 20% DTP/K-10 20% DTS/K-10 Filtrol-24 K-10 S-ZrO2 75 59.5 55.4 57 32.7 17 49 57 53 56 69.4 70.4 2-TBA 20.3 22 24 13.5 18.2 23.7 2,4-DTBA 30.7 21 23 30.4 12.4 5.9

Anisole: MTBE: 1:3; catalyst loading: 0.03 g/cm3; temperature: 170 C; Solvent: 1,4-dioxane; Time: 5 h; speed of agitation: 800 rpm; autogeneous pressure: 1 Mpa; 4-TBA: 4-tert-butyl anisole; 2-TBA: 2-tert-butyl anisole; 2,4-DTBA: 2,4-di-tert-butyl anisole.

G. D. Yadav/Synergism of clay and heteropoly acids Table 12 Ecacies of dierent catalysts in alkylation of aniline with MTBE [68] No. Catalyst Activity 106 mol/g.s Conversion of aniline (%) Product selectivity (%) 2-TBA 1. 2 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 20% w/w DTP/K-10 clay 20% w/w DTP/KSF clay 20% w/w DTP/Filtrol-24 clay 20% w/w DTP/SWy2 clay DTP K-10 clay KSF clay Filtrol-24 clay 20% w/w DTS/K10 DTS 10% AlCl3/ 10% FeCl3/K10 Al pillared clay S-ZrO2 3.506 2.717 2.630 2.410 2.586 1.972 1.314 1.840 2.717 2.454 2.016 1.884 1.402 70 56 51 47 52 40 24 35 56 48 41 37 26 53 56 54 53 63 44 45 43 58 60 54 67 53 4-TBA 31 32 28 36 29 37 45 41 34 31 36 26 36

129

2,4-DTBA 16 12 18 11 8 19 10 16 8 9 10 7 11

Aniline:MTBE 1:4 mole; catalyst loading 0.05 g/cm3; temperature 175 C; time 4 h; speed 1000 rpm; autogenous pressure 350 psig; autoclave 100 ml; TBA: tert-butylaniline; DTBA: di-tert-butylaniline. Table 13 Ecacies of various catalysts in alkylation of 4-Methoxyphenol with MTBE [70] Catalyst Conversion (%) % Selectivity of 2-TBHA S-ZrO2 K10 montmorillonite 20% w/w DTP/K10 Filtrol - 24 Deloxane ASP 22.2 41.7 56.1 75.0 48.2 85 79 74 70 84 2,5-DTBHA 15 21 26 30 16

4-Methoxyphenol: MTBE 1:3 mole; catalyst loading 0.03 g/cm3; temperature 150 C; speed 800 rpm; solvent 1,4-dioxane; reaction time 3 h; autogeneous pressure 150 psi.

Only two products 2-tert-butyl-4-methoxyphenol and 2, 5-di-tert-butyl-4-methoxyphenol were formed with maximum selectivity to the monoalkylated product, both of which are used as antioxidants either as a mixture or alone. The eects of various parameters were studied to arrive at the reaction kinetics. The reaction is intrinsically kinetically controlled with an activation energy of 12.81 kcal/mol.

7.6. Etherication of phenethyl alcohol with alkanols Substituted phenyl and phenethyl ethers such as phenethyl methyl ether (PEME), phenethyl ethyl ether (PEEE), phenethyl isopropyl ether (PEIPE), and phenethyl isoamyl ether (PEIAE), and ethyl o-methoxybenzyl ether are important perfumery ingredient. Twenty percent of w/w DTP/K10 was found to be a very ecient and novel catalyst in comparison with several others such as DTP alone, DTP/C, Indion 830, an ion exchange resin (IER), and K-10 clay, in the etherication of phenethyl alcohol with a variety of

alkanols such as methanol, ethanol, isopropanol and isoamyl alcohol. The typical catalyst loading in the reaction mixture was 10% wt/wt, of the entire mass of the reaction mixture (0.08725 g/cm3 of the liquid phase). The reactions were carried out at 70 C except where the eect of temperature on the rate of reaction was studied. The runs were typically conducted for 5 h. The reactions are 100% selective towards the formation of the ethers. The rates of reaction phenethyl alcohol with dierent alkanols were the same showing zero order dependence on the concentration of alkanol. The experimental data demonstrate that the chemisorption of phenethyl alcohol onto the catalytic site controls the overall rate of reaction for all etherication reactions with all alcohols. An Eley-Rideal type of mechanism prevails [55]. 7.7. Etherication of b-naphthol with alkanols The ethers of b-naphthol are commercially very attractive due to their extensive applications in the ne chemical industry. For instance, naphroxen (6methoxy-R-methyl-2-naphthaleneacetic acid), which is an antiinammatory, analgesic, and antipyretic drug,

130

G. D. Yadav/Synergism of clay and heteropoly acids

is produced from b-naphthyl methyl ether as the starting material. The ecacy of solid acid catalysts such as acid-treated clays (Filtrol-24 and K-10), heteropolyacids (DTP, DTS, DTM), zeolites (X, Y, mordenite, ZSM-5), and sulfated zirconia was tested in the etherication of b-naphthol with methanol. A 2.0% w/w loading of catalyst based on the organic mass of the reaction mixture was employed at a mole ratio of methanol to b-naphthol of 27:1 at 150 C and an agitational speed of 600 rev/min. It was found that the zeolites X, Y, mordenite, ZSM-5 and the raw clay Na+-montmorillonite (Swy2) did not show any activity to this reaction. Twenty percent of DTP/K-10 showed very good activity followed by Filtrol-24 compared to other catalysts. The activity of the various catalysts is evaluated on the basis of the second order rate constants [kR2] that are obtained under the kinetically controlled mechanism. Among these catalysts, 20% DTP/K-10 clay was found to be the best. (Table 14). The rates of etherication with other aliphatic alcohols over the same catalyst at 150 C were in the following order n-BuOH > 2-PrOH > MeOH > EtOH > n-PrOH. The catalyst can be easily recovered and reused without any signicant loss in activity [71]. 7.8. Synthesis of hydroquinone monomethyl ether (HMME) from hydroquinone and methanol Hydroquinone monomethyl ether (HMME) is industrially a very important anti-oxidant having several applications. A variety of clay supported catalysts were employed in this reaction such as 40% w/w DTP/K10, 20% (50:50) w/w AlCl3FeCl3/K10, 20% w/w ZnCl2/ K10. It was found that the addition of benzoquinone as a co-catalyst was necessary for initiation of the reaction. Amongst all catalysts, 40% DTP/ K10 was the only catalyst with benzoquinone as co-catalyst was eective. The main product of formation was monoalkylated hydroquinone, because methanol was taken in far excess molar quantity vis-a-vis hydroquinone. No di-ether of hydroquinone was formed and the selectivity of hydroquinone monomethyl ether was 100%. The reaction mechanism suggests that benzoquinone couples with
Table 14 Rate constants for various catalysts for the etherication of b-naphthol at 150 C [71] Catalyst 20% w/w DTP/K10 Filtrol-24 20% w/w DTS/K10 S-ZrO2 20% w/w DPM/K10 K10 kR2 (cm3/g) (cm3/gmole) s)1 3.23 2.69 2.15 1.61 0.27

protonated methanol on the catalyst site to yield an intermediate, which reacts with hydroquinone adjacent to the site in the absence of external mass transfer and intra-particle diusion limitations. The surface reaction between the intermediate and chemisorbed hydroquinone controls the overall rate. There is 100% selectivity towards HMME. The experimental data are well correlated with the model. Excess of methanol favours the formation of hydroquinone monomethyl ether [72]. 7.9. Acetalization of aldehydes with alcohols Acetals are very well known in organic chemistry as the protective group for aldehydes. Perfumery aldehydes, used in soaps and cosmetics are unstable under the prevalent alkaline conditions and are also prone to chemical attack by oxygen leading to alterations in odour and discolouration of the nished products. Methyl and ethyl acetals, particularly of n-octanal and n-decenal have widespread applications in perfume and avour industry. CH3 CH2 nCHO 2CH3 OH CHOCH3 2 H2 O CH3 CH2 n !

With20% w/w DTP/K10, with 0.5% w/w loading of the catalyst, the reaction was complete within 30 min and was found to be mass transfer controlled for a large molar excess of alcohol to aldehyde. The ecacy of dierent catalysts per unit weight basis, in the kinetic regime, was in the following order for octanal to methanol mole ratio of 1:30 at catalyst loading of 0.5% w/w at 65 C, was, with the corresponding conversions after 6 h as: Filtrol-2493:3% > Amberlyst-1582:5% > 10% DTP=K1072% > 10% DTP=ZSM-5 53% The ecacy of supported heteropoly acids on K-10 clay is little lower than Filtrol-24. It appears that the bulky heteropoly acid crystals can block some of the pores in small ranges and therefore ZSM-5 support gives smaller conversion values. More than 60% conversion is obtained within 1 h for all the catalysts except 10% DTP/ZSM-5. The overall rate of the reaction was controlled by the formation of the acetal from the hemiacetal on the surface of the catalyst, which is formed in the rst step from the aldehdye, giving an overall third order rate equation [73]. 7.10. Synthesis of dodecylbenzenes via alkylation of benzene with 1-dodecene 2-, 3-, and 4-Phenyldodecanes would be preferred for making linear alkyl benzene sulfonates (LABS) from

b-naphthol : 5 g; methanol : 30 g and catalyst : 0.7 g total volume of reaction mixture: 42 cm3; agitation speed : 600 rev/min.

G. D. Yadav/Synergism of clay and heteropoly acids

131

biodegradability view point in comparison with the 5and 6- isomers. The ratio of total concentrations 2-,3and 4- isomers to that of 5- and 6- isomers (with no branched phenyldocanes or didoceylbenzenes) is taken as a yardstick of biodegradability and environmental acceptability and less content of 5- and 6-isomers is benecial for sulphonation and formulation. The liquid phase alkylation of benzene with 1-dodecene was examined by using 13 non-zeolities, based on clays, pillared clays and clay supported heteropolyacids such as dodecatungstophosphoric acid (DTP), dodecatungstosilicic acid (DTS), and dodecamolybdophosphoric acid (DMP). The activities and selectivities of K-10 clay, 20% w/w heteropolyacids (DTP, DMP and DTS) supported on K-10, Filtrol-24, Al pillared clay, 20% w/w DTP/silica,10% AlCl3/10% FeCl3/K10, DTP, Cr exchanged K10, sulfated zirconia, Zr exchanged K10 and 20% DTP/activated carbon. Twenty percent of w/w DTP/K10 clay leads to the best activity and selectivity to 3-phenyldodecane than all other catalysts used. Excess of benzene favours the formation of dodecylbenzenes. However, with decreasing benzene concentration, the formation of didodecylbenzene increases. A mechanistic and kinetic model was developed and validated against experimental data. Benzene alkylation was also carried out with 1-octene, 1-decene and 1-tetradecene under otherwise similar sets of conditions. It was found that the rate of benzene alkylation decreased with an increase in the chain length of alpha olen. With 1-octene and 1-decene the rate of reaction was very fast compared to 1-dodecene and it was very slow with 1-tetradecene (Table 15) [74]. 7.11. Polyalphaolen (PAO) based lubricants Polyalphaolen (PAO) based synthetic lubricants are produced in two stages oligomerization of a-olen and hydrogenation. Currently the synlubes industry

produces low viscosity grade oligomers from 2 cSt to 10 cSt, by using BF3-alcohol catalytic system whereas for the production of higher viscosity grade oligomers (40 cSt-100 cSt), AlCl3 is used as catalyst. The current industrial processes use AlCl3 and BF3 as catalysts for oligomerization, which are polluting, hazardous and give a very wide molecular weight distribution. We reported a green synthesis of a PAO synlube by using 1-decene as the starting material with a variety of solid acidic catalysts, mainly based on clays. The activities for oligomerisation were found to be in the following order (Table 16) [75]: K10 montmorillonite clay > 20% w=wZnCl2=K10 > 5% w=w DTP=K10 > 20%DTP=K10 > 10%AlCl3 10%FeCl3=K10 > S-ZrO2 > Al-pillared clay > DTP > 20%=carbon:

7.12. FriedelCrafts acylation reactions FriedelCrafts electrophilic acylation is a powerful tool for functionalisation of aromatic compounds and is useful in a variety of industries. The acylation of aromatic hydrocarbons requires, in majority of cases, a considerable increase in the electrophilicity of the reagent. An electrophilic catalyst is thus normally required in the addition to the aromatic substrate and the acylating reagent and many of highly polluting homogeneous acids are used in industry. Many acylated derivatives are precursors for important drugs and agrochemicals. The most common method of introducing an acyl group into an organic compound is using either aluminum chloride in nitrobenzene or BF3 in liquid HF as catalyst. Various other Lewis acids such as FeCl3,

Table 15 Eect of dierent catalysts on alkylation of benzene with 1-dodecene [74] No. Catalyst 1 2 3 4 5 6 7 8 9 10 11 12 13 20% w/w DTP/K10 20% w/w DMP/K10 20% w/w DTS/K10 Filtrol-24 Al pillared clay K10 20% w/w DTP/silica 10% w/wAlCl3/ 10% w/w FeCl3/K10 Al exchanged K10 Cr exchanged K10 Sulphated zirconia Zr exchanged K10 20% w/w/ DTP/carbon Conversion of C12 (%) Selectivity of 2-phenyldodecane (%) Selectivity of 3-phenyldodecane (%) 97 92 91 85 80 71 57 32 22 12 12 10 0 32 28 29 22 31 27 35 45 33 30 30 30 0 62 52 50 47 47 30 42 20 42 41 32 35 0

1-dodecene:benzene: 1:10 mol; temperature: 150 C; catalyst loading: 0.05 g/cm3; speed :1000 rpm; autoclaved; DTP: dodecatungstophosphoric acid; DTS: dodecatungstosilicic acid; DMP: dodecamolybdophosphoric acid.

132

G. D. Yadav/Synergism of clay and heteropoly acids Table 16 Ecacies of various catalysts in the oligomerization of 1-Decene [75]

Catalyst

Conversion of 1-decene(%)

Selectivity(%) Dimer Trimer 12 5 10

20% w/w ZnCl2 /K10 K10 montmorillonite clay 5% DTP w/w /K10 20% DTP/K10 10%AlCl3 +10%FeCl3/K10 S-ZrO2 Al pillared clay DTP 20%HPA/carbon Filtrol-24

68 73 48 46 23 23 16 8 5 3

100 88 100 100 100 95 90 100 100 100

1-decene 0.25 mol; n-dodecane 10% w/w of 1-decene; catalyst loading 0.05 g/cm3; temperature 190 C; speed 1000 rpm; reaction time 4 h.

ZnCl2, SnCl2, InCl3, SbCl5, various triates, and Bronsted acids, such as H2SO4 and HCl have been studied extensively. As a consequence, the major drawback of FriedelCrafts acylation on the industrial scale consists of serious euent problems. The traditional homogeneous catalyzed reaction requires at least two mols of Lewis acid to generate one mole of the product [56, 76]. It is thus clear that there is great scope for devising new heterogeneous catalytic processes for regioselective mono-acylation. In our laboratory, several non-zeolitic benign catalysts have been studied. In particular the advantages of supported heteropolyacids (HPA) on K 10 clay with reference to others have been brought out. 7.12.1. Selective mono-acylation of mesitylene Monoacylated mesitylene is industrially an important precursor used in ne chemical and intermediate industries. Selective synthesis of 2,4,6-trimethyl acetophenone (monoacylated mesitylene) from 1,3,5trimethyl benzene (mesitylene) and acetyl chloride was studied using dierent catalysts such as DTP, DTS, 20% DTP/K-10, 20% w/w DTS/K-10 clay, 20% w/w DTP and DTS (50:50)/K-10 and K-10. . K-10 clay and DTP/ K-10 showed high activity followed by DTP, DTP+DTS/K-10 and 20% w/w DTS/K-10 at 50 C. Although K-10 showed higher activity, it had

poor selectivity towards mono-acylmesitylene. DTP/ K-10 was very eective leading to a good conversion of mesitylene and 100% selectivity towards monoacylmesitylene as compared to other catalysts (Table 17). A complete theoretical and experimental analysis was presented. The reaction follows the Eley Rideal type of mechanism with weak adsorption of the acylating species. The energy of activation was found to be 18 kcal mol)1, indicating that the reaction is intrinsically kinetically controlled [56].

7.12.2. Acylation of 2-methoxy-naphthalene Acylation of 2-methoxynaphthalene (2-MON), also known as yarayara, with acetic anhydride, was carried out using dierent solid acid catalysts such as DTP, K10, Filtrol-24, 20% DTP/K10, Al-pillared clay, sulphated zirconia, Amberlyst-15 and ion exchange resin Indion- 130 at 50 C using the 2-MON to acetic anhydride mole ratio of 1:10. [76], in which Indion-130, 20% DTP/K10 and suphated zirconia were active in that order. However, only 1-acetyl-2-methoxy-naphthalene was the sole product which needs to be isomerized for 6acetyl-2-methoxy-naphthalene. The products of the reaction are precursors for many organic and pharmaceutical products. At high molar ratios of acetic anhydride to 2-MON, the reaction is limited by

Table 17 Ecacies of dierent catalysts in acylation of Mesitylene with actetyl chloride [56] No. Catalyst Conversion of Mesitylene (%) Product Selectivity (%) Mono- acetylated 1. 2. 3. 4. 5. DTP (unsupported) 20% w/w DTP/K10 DTP+DTS (10% each)/K10 20% w/wDTS /K10 K-10 (alone) 27 34 23 16 40 100 100 100 87 80 Di- acetylated 13 20

Catalyst loading : 0.05 g/cm3; temperature : 50 C; speed : 800 rpm; reaction time : 4 h; mole ratio of mesitylene to acetyl chloride : 1:10.

G. D. Yadav/Synergism of clay and heteropoly acids Table 18 Ecacies of various catalysts on acylation of resorcinol [77] Catalyst Conversion (%) Selectivity of resoacetophenone (%) 98.0 98.7 42.6 64.8 41.5 71.8

133

Table 19 Isomerization of 1,2-epoxyoctane over various heterogeneous catalysts [10] No. 1 2 3 4 4a 5 6 7 8 9 10 11 12 13 14 15 15a 16 Catalyst H3PO4/SiO2 H3PO4/C Aq. H3PO4 H3PO4/K-10 a H3PO4/SiO2 DTP/K-10 DTP DTP/SiO2 DTP/C ZnCl2/C ZnCl2/K-10 ZnCl2/Filtrol-24 ZnCl2/SiO2 Sulfated ZrO2 Filtrol-24 ZSM-5 ZSM-5a DTP/Y-zeoliteb Conversion (%) 100 76 31 100 100 74 100 66 32 12 29 87 25 24 55 12 80 100

Indion-130 Amberlyst-36 S-ZrO2 Filtrol-24 K 10 DTP/ K 10

56.7 63.0 6.1 7.4 3.8 12.3

Resorcinol: acetic acid :: 1:10; catalyst loading : 0.0515 g/cm3; temperature : 120 C; speed of agitation : 800 rpm; time: 5 h.

intraparticle diusion of the substrate 2-MON. However, this barrier is overcome by using equimolar quantities of the reactants. The reaction proceeds through the Eley Rideal type of mechanism, wherein the chemisorbed acetic anhydride generates a carbocation and acetic acid, and the carbocation reacts with 2-MON from the liquid phase within the pore space to form 1-acetyl-2-methoxy-naphthalene. 7.12.3. Acylation of resorcinol with acetic acid 2,4-Dihydroxyacetophenone, also known as resoacetophenone, is a commercially important intermediate which is generally prepared by the acylation of resorcinol with acetic acid in the presence of a molar excess of zinc chloride, which leads to waste disposal problems. The synthesis of 2,4-dihydroxyacetophenone from resorcinol and acetic acid was carried out in the presence of a variety of solid acid catalysts such as K-10 , 20% w/w DTP/K-10, sulfated zirconia and ion exchange resins (Table 18). Amongst these catalysts, Amberlyst-36, an ion exchange resin, was found to be the most eective although 20% DTP/ K10 showed synergism and better activity in comparison with remaining catalysts. It was possible to deduce the adsorption equilibrium constant and rate constant simultaneously for the reaction including the corresponding energies of activation [77]. 7.13. Isomerisation of 1,2-epoxyalkanes The rearrangement of epoxides over acidic and superacidic catalysts leads to several products with selectivity to the desired aldehyde depending upon the nature of the catalyst surface and the reaction conditions. Yadav and Satoskar [52] reported the studies with 1,2epoxyoctane as a model compound, over several catalysts such as sulphated zirconia and zeolites, and dodecatungstophosphoric acid, phosphoric acid and supported on silica, clay and carbon. The typical ZnCl2 products of rearrangement are aldehyde, allylic alcohols, furan and aldol. H3PO4/SiO2, DTP/ K-10 clay and ZSM-5 were found to be ecient under dierent con-

Epoxide, 0.078 mol; solvent, heptane, 30 cm3; temperature, 50 C; time, 4 h; catalyst loading, 2.5% (w/w) on total mass; Five products were formed; all supported catalysts have a loading of 10% (w/w) of the active species. ZSM-5 = SiAl ratio =80. a Reaction time: 2 h. b Autogenous pressure: 1034.2 kPa, 2 h time, 180 C. For 15a, at 180 C and 1034 kPa.

ditions (Table 19). A kinetic model and mechanism are also presented. 7.14. Nitration of chlorobenzene Industrial aromatic nitrations are carried out by employing a mixture of nitric acid and sulfuric acid predominantly yielding ortho- and para-substituted products from substituted benzenes as substrates. Nitro compounds nd uses in many industries such as drugs and dye intermediates, explosives, and the use of para isomers is found commercially to be the most common. However, this mixed acid process has a major problem of the spent acid disposal which is environmentally unfriendly and can be costly. The formation of water during the reaction causes the dilution of the nitrating mixture thus reducing the rate of nitration as the reaction proceeds. At about 68% sulfuric acid, the nitration becomes extremely slow. Several catalysts have been reported for nitration of chlorobenzene which include Cu(NO3)2/K-10, HNO3/silicaalumina, HNO3 supported on H-ZSM-5, zeolite b and HNO3 supported on SiO2 molecular sieves amongst which zeolite b was reported to give the best para selectivity of about 93%. Yadav and Nair reported a series of supported or modied catalysts such as 20% DTP/K10 and UDCaT2 to get better selectivity in nitration of chlorobenzene and toluene, the latter being better shape selective catalyst (Table 20) [78].

134

G. D. Yadav/Synergism of clay and heteropoly acids Table 20 Mono-nitration of chlorobenzene [78]

Catalyst

Conversion of HNO3 a (%) No conversion 47 45 41 21

p : o ratioc

No catalyst S-ZrO2 UDCaT-2 5%DTP/S-ZrO2 c 5%DTP/S-ZrO2/CMSd

10:6: 1 13:2: 1 8:6: 1 10:2: 1

a In a short while after the addition of nitric acid was complete. Conversion of nitric acid was calculated on the basis of chlorobenzene consumed. b Analysed by GC and by GC-MS. Compared with authentic samples. c 1 g of dodecatungstophosphoric acid (DTP) was dissolved in 10 ml of methanol and supported on 19 g of S-ZrO2 by incipient wetness technique, dried at 120 C in an oven for 24 h and calcined at 285 C for 3 h to obtain 5%DTP/S-ZrO2 . d 5%DTP/S-ZrO2 was coated with polyvinyl alcohol solution using incipient wetness technique and calcined at 285 C for 3h.

8. Applications of 20% w/w Cs2.5H0.5PW12O40/ K10 in commercially relevant reactions 8.1. Dehydration of isopropanol to propylene and diisopropyl ether Isopropanol dehydration reaction was studied at 473 K over 20% w/w Cs2.5H0.5PW12O40/ K10, wherein a conversion of 90% with selectivity of 60% of propylene and 40% of diisopropyl ether (DIPE) was obtained within 1 h with the conversion values rising by 5% with every 10 degree rise in temperature. The formation of DIPE increased signicantly with temperature [79]. In fact, the isopropylation of a variety of aromatics with a number of HPA based and zirconia based UDCaT catalysts suggests interesting kinetic mechanisms [80,81]. 8.2. Isopropylation of benzene to cumene Alkylation of benzene with propylene or isopropanol (IPA) results in the formation of isopropyl benzene (cumene) which is used almost exclusively for the production of phenol and acetone using dierent catalysts. However, use of propylene, as alkylating agent at very high temperatures leads to coke formation which results in deactivation. Industrial processes for the production of cumene from alkylation of benzene with propylene are based on the use of solid phosphoric acid (SPA) or zeolites. A wide variety of zeolites has been explored. Use of propylene, as alkylating agent, is associated with coke formation due its oligomerization, which results in deactivation of the catalysts and loss of propylene. Therefore, use of isopropanol as alkylating agent is attracting a lot of attention. Moreover almost all zeolites give best conversions at temperatures above 600 K. Such energy intensive operating conditions constitute a major portion of the operating costs and need to be cut down to ensure sustenance of any industrial operation in todays competitive world. Use of isopropanol is made

in several laboratories and is supposedly helpful in reduction of coke formation. Isopropanol is dehydrated into propylene and it is the olen which is the alkylating species. It has been reported in a number of papers that high operating temperatures with zeolite catalysts result in coke formation and deactivation. Besides, acidity and porosity also cause deactivation. Zeolites are less acidic than HPAs, and hence require higher temperature for the same acid catalyzed reaction The use of lower reaction temperatures with more acidic mesoporous catalysts will thus be advantageous in maintaining high rates without deactivation and the current work was thus undertaken using clay supported HPA. Also, the micropores of zeolites will be clogged easier than the mesopores of HPA/K-10. The alkylation of benzene to cumene is a very important step in the production of phenol in the Hock process. The alkylation is typically carried out with propylene using high temperatures and a variety of catalysts which get deactivated fast. A novel process using Cs2.5H0.5PW12O40 supported on K-10 clay as a catalyst was developed and engineered with excellent activity and selectivity for the isopropylation of benzene to cumene at 437 K with 9:1 benzene to IPA mole ratio and WHSV of 30 h)1 converting 98.7% of IPA with 85% selectivity towards cumene at much lower temperatures than reported so far. This leads to substantial savings of energy. The experimental data are tted with a pseudo-rst order kinetic model [59]. 8.3. Synthesis of phenol from cumene hydroperoxide Phenol production is expected to continue to grow at 6% per annum to about 7.5 million tons capacity in 2003. This growth is due to the use of phenol in the production of phenolic resins, synthesis of adipic acid, caprolactum, bisphenol, nitro and chlorophenols, phenol sulphonic acids and the like. Phenol is mainly by the three-step Hock process (also called the cumene peroxidation method) wherein in the rst step, benzene and propylene are reacted to yield cumene , which is puried and oxidized in the second step with oxygen to generate cumene hydroperoxide (CHP) (step two). CHP is then isolated and decomposed in the third step with sulfuric acid to obtain phenol along with acetone. The rst step calls for the use of a large quantity of benzene, much of which must be recovered and reused. As the main product produced in step two is an explosive peroxide, only 25% of the cumene is allowed to react in order to keep the concentration of the former low. The unreacted cumene must consequently be recovered and reused. Step three yields equimolar quantities of phenol and acetone through decomposition by sulfuric acid but there are other byproducts which need to be controlled. Both phenol and acetone are valuable bulk chemicals serving their own independent as well as interrelated expanding markets.

G. D. Yadav/Synergism of clay and heteropoly acids

135

The Hock process technology will continue to dominate since acetone, of late, has found many more uses in agrochemical, pharmaceutical and resin industry. Thus, dierent catalysts are being explored in all three steps of the Hock process. In particular, the third step of cumere hydroperoxide decomposition needs to be converted into an ecofriendly process. Various Bronsted and Lewis acids are used to decompose CHP in a temperature range of 075 C to yield phenol, acetone with some dicumyl peroxide (DCP) and sulfuric acid is preferred to get high yields of phenol. DCP in turn gets converted into a mixture of a-methyl styrene (AMS), phenol and acetone. A change in strength of the catalyst as well as the reaction temperature facilitates the change in selectivity of the reaction product during the decomposition of CHP. The decomposition of CHP to phenol and acetone was studied by using K-10, 20% w/ w DTP/K10, 20% w/w Cs2.5H0.5PW12O40/K10, 20% w/ w ZnCl2/K10 and sulphated zirconia at 40 C. Cs2.5H0.5PW12O40/K10 was found to be the most active and selective catalyst. The reaction follows the rst order kinetics. The catalyst is stable and reusable giving 100% conversion with exclusive selectivity to 100% phenol and acetone during all the runs [82]. 8.4. Alkylation of phenol with isopropanol Isopropylphenols are formed by reaction of phenol with propylene or isopropanol (IPA) in the presence of Lewis or Brnsted acids such as alumina. Propofol (2,6diisopropyl phenol), a very important drug, is typically synthesized by the isopropylation of phenol over an acid catalyst. This process consists of two series reactions, namely, the formation of 2-isopropyl phenol and its further isopropylation to 2,6- and 2,4-diisopropyl phenols. The selectivity to 2,6-diisopropyl phenol depends upon a number of parameters and the alkylating agent, type of catalyst and temperature play a major role. Industrially, 2-isopropylphenol is produced from liquid phase reaction of phenol and propylene using c-Al2O3 at 250300 C and 10 MPa. If a phenol to propylene molar ratio of 1.3:1 is used, 2-isopropylphenol is obtained with a selectivity of about 85% at complete propylene conversion. 2,6-Diisopropylphenol is formed as a byproduct (1012%) in this process along with smaller amounts of 4-isopropylphenol (12%) and isopropyl phenyl ether. The proportion of ether formed increases considerably at temperatures below 250 C. Mono-isopropyl phenol mixtures enriched in the 4-isomer can be synthesized from phenol and propylene or IPA over zeolite catalysts. When a 1:1 molar mixture of phenol and IPA is passed over a ZSM-5-catalyst at 250 C and LHSV of 1 h1, a phenol conversion of 20% and a selectivity for 4-isopropylphenol of 63% is achieved. Thus the alkylation of both phenol and monoisopropyl phenol was studied in liquid-phase in an autoclave at 200 C using isopropanol and isopropyl ether independently as alkylating

agents, over 20% w/w Cs2.5H0.5PW12O40/K-10 catalyst which is a synergistic combination of cesium modied heteropoly acid and clay. This catalyst is better than zeolites [83].

8.5. FriedelCrafts acylation reactions 8.5.1. Benzoylation of anisole with benzoyl chloride Functionalized aromatic ketones, produced via acylations, are of great importance in a variety of industries. For example, acylation of anisole (4-methoxy benzene) with benzoyl chloride or benzoic anhydride gives 4-benzoyl anisole (4-methoxy benzophenone), which is a valuable perfumery intermediate and also a precursor for antioxidants used in cosmetics, PVC, unsaturated polyesters, acrylics, rubber, alkylated resins, cellulose lacquers and oil paints. The most common method of introducing an acyl group into an organic compound is via the FriedelCrafts reaction by using either aluminium chloride in nitrobenzene or boron triuoride in liquid HF as catalyst. Zeolites have emerged as an alternative for the traditional polluting catalysts but unfortunately the slow diusion of reactants through their microporous structure makes them relatively poor catalysts in liquid-phase reactions and their stability is susceptible in reactions where acids are generated as co-products. The use of highly reactive carboxylic acid halogenide compounds as acylating agents, as well as anhydrous HCl, which is produced during the reaction, may remove the aluminum species from the zeolite framework. Of late, several researchers have focused their attention to Hb and HY zeolites for acylation of substituted aromatics including anisole. However, zeolites have been found to deactivate [57]. The acylation of anisole with benzoyl chloride was studied systematically by using a variety of catalysts (Table 21). Twenty percent of w/w Cs2.5H0.5PW12O40 / K-10 was the best catalyst which could be reused without any further chemical treatment eliminating the euent disposal problem. This acylation reaction is 100% selective towards 4-methoxybenzophenone, which is the desired perfumery intermediate. The eect of various important kinetic parameters was evaluated systematically to establish that the reaction obeys the EleyRideal type of mechanism with a very weak adsorption of the reactants. 8.5.2. Benzoylation of p-xylene Benzophenones form an important class of ketones employed in ne chemical, intermediate, dyestu, pharmaceutical and agrochemical industries. The most common method of introducing an acyl group into an organic compound is by the FriedelCrafts acylation using either aluminum chloride in nitrobenzene or BF3 in liquid HF as catalysts. 2,5-Dimethylbenzophenone is used extensively as a UV light stabilizer in plastics,

136

G. D. Yadav/Synergism of clay and heteropoly acids

Table 21 Ecacies of various catalysts on acylation of anisole with benzoyl chloride [57] Catalyst Initial activity (mol. g-cat)1.s)1) 16.610)4 10 10)4 10 10)4 2 10)4 1.66 10)4 Conversion (%)

20% w/w ZnCl2/K-10 20% w/w Cs2.5H0.5 PW12O40/K-10 20% w/w DTP (/K-10 S-ZrO2 K-10

38 37 32 26 15

K10 clay and 20% w/w Cs2.5H0.5PW12O40/ K10 are excellent catalysts for a number of reactions such as alkylation, acylation, isomerization, oligomerization, condensation, dehydration, esterication and nitration reactions which are useful in industries such as bulk, intermediates, dyes, plasticizer, pharmaceuticals, perfumes, avours and other ne chemicals. This review primarily dealt with the work carried out in authors laboratory.

Benzoyl chloride: Anisole = 1:7; catalyst loading = 0.03 g/cm3; temperature = 70 C; speed of agitation = 800 rpm; time = 3 h.

Acknowledgment Financial support from the Darabari Seth Professor endowment for the personal chair is greatefully acknowledged including DST, CSIR, NIMTLI and AICTE.

cosmetics and lms and produced by the benzoylation of p-xylene using homogeneous catalysts which pose several problems. The benzoylation of p-xylene with benzoyl chloride was carried out in a batch reactor using clay supported catalysts such as 20% w/w DTP/K-10, 20% w/w Cs2.5H0.5PW12O40/K10, 20% w/w ZnCl2/K-10 and K-10 itself and sulfated zirconia. Amongst these catalysts, 20% w/w Cs2.5H0.5PW12O40/K10 was found to be a better catalyst which could be reused without any further chemical treatment eliminating the euent disposal problem. This catalyst was fully characterized. The reaction obeys the EleyRideal type of mechanism with a weak adsorption of the benzoylating species [57]. 8.5.3. Regio-selective benzoylation of xylenes Benzoylation of xylenes leads to dierent dimethyl benzophenone isomers which are important intermediates for a variety of industries including pharmaceuticals, ne chemicals, agrochemicals, explosives, polymers, antioxidants, etc. 3,4-Dimethylbenzophenone, 2,4-dimethylbenzophenone and 2,5-dimethylbenzophenone are all commercially valuable ketones derived from the benzoylation of dierent xylenes using homogeneous catalysts. It is possible to prepare these ketones selectively and economically from commercial mixed xylenes under dierent sets of conditions because of the preferential reactivities of xylene isomers in the presence of solid catalysts. The regioselective benzoylation of pure xylene isomers and of mixed xylenes was studied with benzoyl chloride over Cs2.5H0.5PW12O40/K 10 as catalyst which could be reused without any further chemical treatment eliminating the euent disposal problems. This also forms the basis for the separation of individual xylene components [59].

References
[1] A. Cybulski, J. Moulijn, M.M. Sharma and R.A. Sheldon, Fine Chemicals Manufacture Technology, Engineering rst ed. (Elsevier, Amsterdam, The Netherlands, 2001). [2] R.A. Sheldon, H. Bekkum,van (eds.), Fine Chemicals through Heterogeneous Catalysis (Wiley, Weinheim, 2001)Toronto. [3] E.G. Deroune, G. Crehan, C.J. Dillon, D. Bethell, H. He and S.B.D. Hamid, J. Catal. 194 (2000) 410. [4] G.D. Yadav and J.J. Nair, Micropor. Mesopor. Mater 33 (1999) 1. [5] S.R. Chitnis and M.M. Sharma, React. Func. Polym. 33 (1997) 93. [6] I. Tkac, P. Komadel and D. Muller, Clay Miner. 29 (1991) 101. [7] Y. Izumi, K. Urabe and M. Onaka, Zeolites, Clays and Heteropoly Acids (VCH Publishers Inc, London, 1992). [8] M.T. Pope, Heteropoly and Isopoly Oxometalates (SpringerVerlag, Berlin, 1993). [9] T. Okuhara and T. Nakato, Catal. Survey Jpn. 2 (1998) 31. [10] T. Okuhara, N. Mizuno and M. Misono, Adv. Catal. 41 (1996) 113. [11] N. Mizuno and M. Misono, Chem. Rev. 98 (1998) 199. [12] M. Misono, Catal. Rev.- Sci. Eng. 29 (1987) 269. [13] M. Misono, Catal. Rev.- Sci. Eng. 30 (1988) 339. [14] M. Misono and T. Okuhara, CHEMTECH, Nov. (1993) 23. [15] K.Y. Lee, S. Oishi, H. Igarashi and M. Misono, Catal.Today 33 (1997) 183. [16] J.B. Moat, Metal-Oxygen Clusters: The Surface and Catalytic Properties of Heteropolyoxometalates (Kluwer Academic/Plenum Publishers, New York, 2001). [17] C. Cativiela, J.I. Garcia, M.G. Matres and B. Chiche, Appl. Catal. 123 (1995) 273. [18] J.R. Sohn and M.Y. Park, Appl. Catal. 101 (1993) 129. [19] A. Cornelis, A. Gertsman, P. Laszlo and I. Zieba, Catal. Lett. 6 (1990) 103. [20] B. Labiad and D. Villemin, Synthesis 2 (1989) 143. [21] C. Cativiela, J.M. Fraile, J.I. Garcia, J.A. Mayorah and P.J. Alonso, J. Catal. 137 (1992) 394. [22] C.N. Rhodes and D.R. Brown, J. Chem. Soc., Faraday Trans. 89 (1993) 1387. [23] M. Zyla and M. Olszar, Minorologia Polonica 15 (1984) 67. [24] C.R. Theocaris, K.J. Jacob and A.C. Gray, J. Chem. Soc. Faraday Trans 1 84 (1988) 1509. [25] V. Olphen, An Introduction to Clay Colloid Chemistry (Wiley, New York, 1977).

9. Conclusion In this review, synergism between clays and heteropoly acids was discussed for the development of green processes and activities were compared several other solid acids. 20% w/w dodecatungstophosphoric acid/

G. D. Yadav/Synergism of clay and heteropoly acids [26] A.C. Wright, J.V. Kennedy and W.T. Granquist, J. Catal. 25 (1972) 65. [27] C.R. Theocaris, K.J. Jacob and A.C. Gray, J. Chem. Soc. Faraday Trans. 1 84 (1988) 1509. [28] P. Komadel, J. Madejova, M. Janek and J.W. Stucki, Clays Clay Miner 44 (1996) 228. [29] T. Cseri, S. Bekassy, F. Figueras, E. Cseke and R. Dutartre, Appl. Catal. A Gen. 132 (1995) 141. [30] G.D. Yadav and N.S. Asthana, Catal Rev.-Sci. Eng in preparation (2005). [31] B.M. Choudhary, M. Sateesh, M.L. Kantam and K.V.R. Prasad, Appl. Catal. A Gen 171 (1998) 155. [32] S.Bekassy , K. Biro, T. Cseri, B. Agai, and F. Figueras (Eds) : Stud. Surf. Sci. Catal. Vol 108 : Heterogeneous Catalysis and Fine Chemicals IV (Elsevier, Amsterdam, 1997) p 83. [33] I.V. Kozhevnikov, K.R. Kloetstra, A. Sinnema, H.W. Zandbergen and H. Bekkum, J. Mol. Catal. A 114 (1996) 287. [34] S. Soled, S. Miseo, G. McVicker, W.E. Gates, A. Gutierrez and J. Paes, Catal. Today 36 (1997) 441. [35] L.R. Pizzio, C.V. Caceres and M.N. Blanco, Appl. Catal. A: Gen. 167 (1998) 283. [36] T.S. Thorat, V.M. Yadav and G.D. Yadav, Appl. Catal. 90 (1992) 73. [37] G.D. Yadav and H.G. Manyar, Micro. Meso. Mater. 63 (2003) 85. [38] M. Spagnol, L. Gilbert and D. Alby, Ind. Chem. Libr. 8 (1996) 29. [39] Metivier P., In: Fine Chemicals through Heterogeneous Catalysis, (eds). Sheldon R.A., H. van Bekkum (Wiley, Weinheim, Toronto,2001) p161. [40] P. Moreau, A. Finiels and P. Meric, J. Mol. Catal. 154 (2000) 185. [41] G.D. Yadav, T.S. Thorat and P.S. Kumbhar, Tetrahedron Lett. 34 (1993) 529. [42] P. Lazlo and M.T. Montauer, Tetrahedron Lett. 32 (1991) 1561. [43] P. Lazlo, A. Cornelis, C. Dony and K.M. Nsunda, Tetrahedron Lett. 32 (1991) 2903. [44] J.H Clark, A.P. Kybett, D.J. Macquarrie, S.J. Barlow and P.J. Landon, J. Chem. Soc., Chem. Commun. (1989) 1353. [45] J.H. Clark, A.P. Kybett and D.J. Macquarrie, Supported Reagents; Preparation, Analysis and Applications (VCH, New York, 1992). [46] S.J. Barlow , J.H. Clark, M.R. Darby, A.P. Kybett, P. Landon and K. Martin, J. Chem. Res. (S) (1991) 74. [47] D.R. Brown, J. Massam and H.M.G. Edwards, J. Chem. Soc., Faraday Trans. 92 (1996) 1029. [48] G.D. Yadav and P.H. Mehta, Ind. Eng. Chem. Res. 33 (1994) 2198. [49] G.D. Yadav and N.S. Doshi, Org. Proc. Res. Dev. 6 (2002) 263. [50] G.D. Yadav and D.V. Satoskar, J. Chem. Tech. Biotech. 69 (1997) 438. [51] G.D. Yadav and N. Kirthivasan, J. Chem. Soc. Chem. Commun. (1995) 203. [52] G.D Yadav and N. Kirthivasan, In : Fundamental and Applied Aspects of Chemically Modied Surfaces, (eds). Jonathan P. Blitz

137

[53] [54] [55] [56] [57] [58] [59] [60] [61] [62] [63] [64] [65] [66] [67] [68] [69] [70] [71] [72] [73] [74] [75] [76] [77] [78] [79] [80] [81] [82] [83]

and Charles B. Little (Royal Society of Chemistry, UK, 1999) p. 254. G.D. Yadav and M.S. Krishnan, Ind. Eng. Chem. Res. 37 (1998) 3358. G.D. Yadav and N. Kirthivasan, Appl. Catal. A: Gen. 154 (1997) 29. G.D. Yadav and V.V. Bokade, Appl. Catal. A. Gen. 147 (1996) 299. G.D. Yadav and N.S. Asthana, Ind. Eng. Chem. Res. 41 (2002) 5565. G.D. Yadav, N.S. Asthana and V.S. Kamble, J. Catal. 217 (2003) 88. G.D. Yadav, N.S. Asthana and V.S. Kamble, Appl. Catal. A: Gen. 240 (2003) 53. G.D. Yadav, N.S. Asthana and S.S. Salgaonkar, Clean Tech. Environ. Pol. 6 (2003) 105. G.D. Yadav, S.S. Salgaonkar and N.S. Asthana, Appl. Catal. A Gen. 265 (2004) 153. G.J Hutchings, C.P. Nicolaides and M.S. Scurrel, Catal Today (1994) 23. R. van Le Mao, R. Carli, H. Ahla and V. Ragaini, Catal. Lett. 6 (1990) 321. B.C. Gates, Catalytic Chemistry (Wiley, New York, 1992). G.D. Yadav and A.V. Joshi, Org. Proc. Res. Dev. 5 (2001) 408. G.D. Yadav and N.S. Doshi, Appl. Catal. A Gen. 236 (2002) 129. G.D. Yadav, A.A. Pujari and A.V. Joshi, Green Chem. 1 (1999) 269. G.D. Yadav and N.S. Doshi, Catal. Today 60 (2000) 263. G.D. Yadav, P.K. Goel and A.V. Joshi, Green Chem. 3 (2001) 92. G.D. Yadav and N.S. Doshi, J. Mol. Catal. A: Chem. 194 (2003) 195. G.D. Yadav and M.S.M.M. Rahuman, Appl. Catal. A Gen. 253 (2003) 113. G.D. Yadav and M.S. Krishnan, Ind. Eng. Chem. Res. 17 (1998) 3358. G.D. Yadav , S.A.K. Deshmukh and N.S. Asthana, Ind. Eng. Chem. Res. (2005) (under review). G.D. Yadav and A.A. Pujari, Can. J. Chem. Eng. 77 (1999) 489. G.D. Yadav and N.S. Doshi, Org. Proc. Res. Dev. 6 (2002) 263. G.D. Yadav and N.S. Doshi, Green Chem. 4 (2002) 528. G.D. Yadav and M.S. Krishnan, Chem. Eng. Sci. 54 (1999) 4189. G.D. Yadav and A.V. Joshi, Clean Tech. Environ. Pol. 4 (2002) 157. G.D. Yadav and J.J. Nair, Catal. Lett. 62 (1999) 49. G.D. Yadav, and S.S. Salgaonkar, Ind. Eng. Chem. Res. 44 (2005) 1706. G.D. Yadav and A.D. Murkute, J. Phys. Chem. A 108 (2004) 9557. G.D. Yadav and A.D. Murkute, Langmuir 20 (2004) 11607. G.D. Yadav and N.S. Asthana, Appl.Catal. A Gen. 244 (2003) 341. G.D. Yadav and S.S. Salgaonkar, Micropor. Mesopor. Mater. 80 (2005) 129.

Anda mungkin juga menyukai