Anda di halaman 1dari 18

Catalysts Selectivity.

, A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

Catalysts Selectivity: Shape Selective Catalysis Over Zeolites A.V. Ramaswamy Department of Chemistry, University of Pune, Pune 411 007 Email: avram@chem.unipune.ernet.in ________________________________________________________________________ Abstract The catalysts either homogeneous or heterogeneous are more important for shifting selectivity to a desired, value added product rather than turnover, in the present context of environmental safety and by-products disposal. Among the general class of chemo-, regio and stereo selectivity associated with a catalyst, shape-selective catalysis is of particular importance, as the zeolites offer tremendous advantage over amorphous catalysts. The ability to synthesize zeolites and mesoporous materials with precise structural and porous features has proved an impetus to their application in the manufacture of petrochemicals, chemical intermediates and fine chemicals. This presentation highlights some of the recent work on the use of acidic zeolites for the production of alkyl aromatics. The main theme is the alkylation of naphthalene and biphenyl, as these substrates and their alkyl derivatives, both mono- and di-substituted alkyl derivatives provide the challenge for maximum selectivity to either 2,6dialkylnapthalene or 4,4-dialkylbiphenyl, respectively, among many isomers that can possibly form on the acid zeolites both within the pore system and outside through associated reactions of isomerization and trans-alkylation. These isomers are important targets for making advanced heat resistant polymers and liquid crystals. Some of the strategies for improved selectivity through passivation of the external acid sites, and the selectivation of the catalyst through controlled coking and deposition of rare-earth metal oxides are discussed with examples drawn from the recent literature. ________________________________________________________________________ Key words: Catalysts selectivity, shape-selective catalysis, regio-selectivity, alkylation of aromatics, selectivation of catalysts, alkylation of naphthalene, alkylation of biphenyl 1. Introduction The role of catalysts is shifting more and more towards getting better selectivity to a desired product in a chemical reaction that leads to more than one products normally. Even though the search for more active catalysts, novel materials and methodology for exposing more active sites in a catalyst for a particular reaction continues, catalysts selectivity has become a crucial issue recently. The problem of separation of the desired product from other byproducts, means of disposal of byproducts and environmental considerations for the same have contributed to development of catalysts, which are more selective even at the cost of activity for a reaction. Equilibrium constants for the product isomers, and difference in the H between an intermediate and the final product in a consecutive reaction such as A B C are to be considered in the choice of a catalyst for a desired product. How the rate constants k1 and k2, respectively for the two reactions, can be controlled by the use of specific catalysts can be seen

140

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

by high yields of either B or C. Similar considerations will be the guiding factors for the selection of catalysts in simultaneous reactions, which will be discussed in this presentation with some examples. Ethylene to its epoxide and propylene to propylene oxide are some of the outstanding examples of industrial success in oxidation reactions, which otherwise will lead to thermodynamically more stable final products, viz., CO2 and H2O. By the right choice of a catalyst, the yield of ethylene oxide today is close to theoretical value or even in excess of the same under controlled conditions of oxidation and the reactor design. The selectivity can, generally, be classified into chemo-, regio- and stereo selectivities for convenience for a number of chemical reactions. The classical example of chemo-selectivity is the reaction of isopropanol to give, on the one hand, acetone by dehydrogenation on oxides such as ZnO and propene by dehydration with mildly acidic oxides such as alumina. An understanding of the nature of the oxides and supported metal catalysts has provided a powerful tool to design catalysts for specific applications, including the ability to control the rates in a way that reflects catalysts selectivity to a desired product. The chemical nature of the material is important in providing the required surface properties for the selectivity. There are innumerable examples of regio selective reactions in which the design of the catalyst is invariably linked to the spatial configuration of the catalyst. One of the classical examples is the hydroformylation of propylene to nbutyraldehyde (anti-Markovnikov product) in homogeneous catalysis. A selectivity of >95% for the linear

aldehyde has been achieved by the use of a Rh complex catalyst, having the right choice of CO and triphenylphosphine ligands around the metal center. The sterically crowded environment around the metal center favors anti-Markovnikov addition of the olefin to the metal. Further examples can be found in the design of metallocene catalysts for the production of stereo tactic polymers, where the symmetry and chirality of the metallocenes are important considerations. 2. Shape selective catalysis Another aspect of regioselectivity is associated with the use of zeolites as acid catalysts in isomerization and alkylation reactions. Shape selectivity, a term given by Csiscery [1] and further discussed in many publications [2], is in a sense an outcome of macroscopic design of a catalyst. A classification of this into reactant shape selectivity, product shape selectivity and restricted transition state shape selectivity could explain the results observed initially with medium pore pentasil (ZSM-5) zeolite as a solid acid catalyst and is well documented in the literature [2-5]. An understanding of the diffusional characteristics of molecules through well-defined pores of zeolites and a match of the molecular dimensions of the reactants, products or the intermediates with the pore dimensions, including the intersections in a threedimensional pore system have contributed to the right choice of a zeolite for a reaction. Some of the bestknown illustrations of shape selectivity are the erionite (small pore) based post reforming catalyst (M-forming), catalytic dewaxing of distillates based on ZSM-5 (medium pore) (MDDW process for diesel), isomerization of C8 aromatics

141

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

to p-xylene (ZSM-5) or the catalytic dewaxing for lubes (Encilite). The ability to synthesize different zeolites, which have well-defined micro pore structure and crystallinity has given a great impetus to take advantage of the phenomenon of shape selectivity in reactions that are controlled more by pore diffusion rather than intrinsic kinetics. Given the big scope of different types of selectivities and the role of catalysts in influencing the selectivity to a desired product, it is impossible to cover all the above aspects in this paper. We shall essentially concentrate on shape-selective catalysis, with particular reference to zeolites and some meso porous materials and draw some general conclusions from the recent literature on the status of shape selectivity. It is pertinent to note here that we will deal with zeolites and acid catalysis for the production of chemicals and intermediates of industrial importance, mainly the alkyl aromatics. As the aromatics are being phased out from the fuel (gasoline and diesel), the efficient ways of converting them into useful and value-added petrochemicals and chemical intermediates assume a lot of importance. While giving examples for the production of the alkyl aromatics, the role of zeolites in shape selective catalysis becomes apparent, as only some of the isomers are industrially important. The question of selectivity becomes the focus of many investigations. 3. Production of p-xylene The isomerization of C8 aromatics (that includes ethylbenzene) to make selectively p-xylene is well known and practiced widely in the industry. This is perhaps the most important

process and for the first time demonstrated the use of pentasil (HZSM-5) zeolite, in terms of product shape selectivity, wherein the diffusion out of the linearly di-substituted benzene (p-xylene) from the pores of ZSM-5 is favorable under the conditions of reaction. Another important reaction to make p-xylene is the dispropor-tionation of toluene, which is again catalyzed by acidic ZSM-5, the advantage of this reaction being the utilization of toluene as the feedstock. Since the introduction of the first commercial process based on ZSM-5 zeolite, there have been considerable improvements in the design of the catalysts for this reaction, taking the selectivity to p-xylene to a very high level. We will briefly discuss the strategy for further improvement in selectivity. Selectivation is a new direction to avoid producing unwanted and lower value by-products in some of these reactions. Although zeolites and clays as such are shape selective, as mentioned earlier, non-selective reactions may take place on the external surface of the zeolites. Selective attenuation of the active sites on the external surface of the zeolite will reduce the formation of these by-products. Passivation of the external surface and further fine-tuning of the dimensions of the pores in zeolite or the pore mouth openings by carefully controlled deposition of non-reactive species have been successfully tried in some of the reactions catalyzed by HZSM-5. Selectivation of the catalyst by silylation or controlled coke deposition has been demonstrated to improve the selectivity in many cases. In the production of p-xylene by toluene disproportionation, a comparison of the two processes of Mobil, both reportedly use the same catalyst, is given in Table

142

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

1. The MSTDP process introduced in the 90s shows an extremely high selectivity to p-xylene. A ZSM-5 catalyst is the heart of the Mobil process in the above reaction. Pre-treatment of the catalyst or selectivation of the catalyst is said to be the key to the improved selectivity to pxylene and is claimed to be through

selective coking of the catalyst which probably passivates the external acid sites as well as modifies the sinusoidal pore dimensions to some extent, particularly the intersections in the MFI pore system.

Table 1 Toluene disproportionation (Mobil, 1995): Comparison of yields at 100% toluene conversion __________________________________________________ Component MTDP-3 MSTDP __________________________________________________ C5 and lighter 3.0 6.0 Benzene 41.4 46.3 Ethylbenzene 1.2 2.0 p-Xylene 11.1 37.0 m-Xylene 24.3 5.3 o-Xylene 11.9 1.3 C9+ Aromatics 7.1 2.0 p-Xylene/total xylene (wt.%) 23.5 84.7 __________________________________________________ 4. Alkylation of aromatics In the manufacture of alkyl aromatics, several processes have been commercialized in the last two decades that use zeolites as acid catalysts, replacing older Friedel-Crafts catalysts and solid phosphoric acid [3,4]. The advantages of acid zeolites have not only been in finding reusable, regenerable alternative solid acid catalysts, but also in improvement in selectivity to desired isomers. Ethylebenzene, cumene and linear alkyl benzene are large volume petrochemicals and alkyl-naphthalenes and alkyl-biphenyls are some of the intermediates to new polymers and resins. Even in the simplest case of alkylation of benzene to ethylbenzene, (Mobil-Badger process, NCL process) where ZSM-5 is used as the catalyst, the selectivity to ethylbenzene, and hence its yield will depend on some of the associated reactions that could occur simultaneously on the same catalyst. One is that normally alkyl benzenes are more reactive than benzene itself in such electrophilic substitution reactions and hence form heavier alkylates in the product stream that can finally lead to coke formation. Secondly, the olefins can oligomerize on the acid catalyst to give some oligomers, and finally, nonselective alkylation and isomerization may take place on the external surface of the zeolite to give many isomers of diand trialkylbenzenes, which are not expected from the shape-selectivity considerations. Further modification of the catalyst is therefore required and there are several ways to circumvent the

143

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

problem of non-selective reactions. As cited in the case of toluene disproportionation, selective coke deposition is one option. Dealumination of the zeolite is often known to be advantageous. Among other ways of external surface passivation, silylation of the zeolite leads to improvement in selectivity in many reactions [5,6]. In the alkylation of ethylbenzene to paradiethyl benzene, for example, the catalyst is silylated in situ to a level that the para-selectivity improves beyond what normal medium pore zeolites could give. A selectivity of >98% could be achieved at a per pass conversion of 15% of ethylbenzene at 340oC and at

ethylbenzene to ethanol (mole) ratio of 4 to 10 [7]. 4.1 Isopropylation of benzene The solid phosphoric acid based process for the alkylation of benzene with propene is being modified in many units with zeolite-based catalysts. The zeolite catalysts, particularly, the betazeolite, have proved to be advantageous, both in terms of reusability of the catalyst and selectivity to cumene. Further improvements are being introduced to reduce heavier alkylates and products of side reactions, such as oligomerization of propene.

Table 2 Catalyst performance in the alkylation of benzene by propene [9] ______________________________________________________________________ Parameters Catalyst Catalyst MCM-22 MCM-56 ______________________________________________________________________ Temperature, oC 112 113 -1 Propene flow, WHSV, h 1.3 10.0 Propene conversion, % 98.0 95.4 Selectivity, % - Cumene 84.35 84.98 - Diisoprooyl benzene 11.30 13.20 - Triisopropyl benzene 2.06 1.28 - C3 Oligomers 1.8 0.52 - n-Propyl benzene, ppm 70 90 The Mobil-Badger cumene process probably uses a novel zeolite catalyst (believed to be MCM-22) whose acidity is controlled to such an extent that both coke formation and propene oligomerization is kept to a minimum [8]. All unwanted by-products (ethylbenzene, n-propylbenzene and butylbenzenes) are claimed to be less than 250 ppm. MCM-22 has an unusual pore system that is claimed to be responsible for the observed benefits of high selectivity to cumene. A recent Mobil patent claims another catalyst, labeled as MCM-56 to be superior to MCM-22 in cumene production [9]. The data in Table 2 show a high propene conversion even at higher space velocities of propene (almost one order of magnitude higher) on MCM-56, compared to MCM-22. At these conditions, C3 oligomers and triisopropyl benzene are lower while cumene selectivity is unaffected.

144

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

4.2 Alkylation of naphthalene The synthesis of symmetrically substituted polynuclear aromatic hydrocarbons such as 2,6dialkylnaphthalene and 4,4dialkylbiphenyl are industrially important intermediates. 2,6Dialkylnaphthalene can be oxidized to the dicarboxylic acid and then esterified with ethylene glycol to produce PEN polyester, which is a high performance engineering plastic that provides superior mechanical, thermal and chemical resistance and barrier properties relative to polyethylene terephthalate (PET). Substantial market potential and growth have been predicted for PEN (15-20 million kg in 2005 for PEN films from about 8 million kg in 2000). The major challenge lies in the manufacture of 2,6dialkylnaphthalene in a cost effective manner. In the alkylation of naphthalene, there are ten different dialkylnaphthalene isomers. The mono-alkylated product, which can either be - or - isomer undergoes further alkylation, wherein the difference in the molecular dimensions of different, di-substituted naphthalene isomers can be taken advantage of by using a molecular sieve of proper pore structure. The isomers are formed not only by alkylation, but also from isomerization and trans alkylation reactions that can simultaneously take place on the catalyst. It is difficult to distinguish between these reactions that ultimately contribute to the product spectrum. It is to be noted that even in the initial stages of the reaction, the coke formation is rapid and that the coke precursors are essentially polyalkylated naphthalenes. As indicated by Guisnet recently [10], the coke molecules trapped in the micropores of

zeolites can be the real active species in hydrocarbon transformation. The transalkylation reaction between poly alkylaromatics trapped in the micropores of large pore zeolites and aromatic molecules in the liquid phase may also contribute to the product selectivity. This is essentially pore mouth catalysis. 4.2.1 Production of dimethyl naphthalene The simplest case of methylation of naphthalene with methanol over a variety of zeolites has been the subject of several investigations [11-14]. 2,6Dimethylnaphthalene is easier to oxidize than other alkyl naphthalenes to its dicarboxylic acid. However, producing 2,6-DMN is challenging both technically and economically. The reasons are that the positions of the methyl groups on the naphthalene ring make them difficult to control, that only two DMN isomers (1,6- and 1,5-DMN) can be efficiently isomerized to the desired 2,6-DMN via conventional acid catalysts, and that 2,6DMN is difficult to separate from other nine DMN isomers by distillation because their boiling points are very close to each other. Many routes have been developed for producing 2,6-DMN, as reviewed by Lillwitz [15]. The major routes can be grouped into, a) reaction of a monocyclic aromatic and a diolefin through alkenylation, cyclization and dehydrogenation, followed by isomerization (BP-Amoco process, Fig. 1); b) transalkylation of naphthalene with polyalkylbenzenes; c) recovery from appropriate cut of cycle oil produced in fluid catalytic cracking (FCC); and d) methylation (and also isopropylation) of naphthalene. The alkylation of naphthalene with methanol under different conditions of reaction and catalysts has not been very successful. However, using 1,2,4-

145

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

trimethlbenzene as a solvent or reagent together with methanol on MTW-zeolite stands out to be the best reported so far, in terms of activity and selectivity to 2,6dimethylnaphthalene [16]. Transalkylation seems to be the main reaction and this can be explained on the

basis of a restricted transition state shape selectivity to be the main feature with MTW catalyst.

CH3 CH3 + CH2=CH-CH=CH2 CH3 Zeolite Catalyst Cyclisation Alkali Metal Catalyst NaK Alkenylation

CH3 CH2-CH2-CH=CH-CH3 OPT

1,5-DMT CH3

Pt / Al2O3 Dehydrogenation 1,5-DMN

Zeolite Beta Isomerization

2,6-DMN

146

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

Fig. 1 Synthesis of 2,6-dimethyl naphthalene from o-xylene and 1,3-butadiene (BPAmoco) Pt/Re/chlorided alumina at conditions of A new route, reported to be more high temperatures (450 to 500oC, 1000 efficient for the synthesis of 2,6-DMN, has recently been announced by kPa, WHSV of 3 h-1 and H2 to TP ratio Chevron-Texaco [17,18]. It uses toluene of 5). The conversion of tolylpentanes is and pentenes as less expensive 100% and at least 5 different DMN feedstocks and consists of alkylation, isomers are formed which are reforming, hydroisomerization followed hydroisomerized and dehydrogenated by dehydrogenation and 2,6-DMN over a Pd/Y or Pd/beta catalyst. Here, purification. The sequence reactions and one of the aromatic rings is saturated via the catalysts used in each step are given hydrogenation so that the energy barriers in Fig. 2. for the positional shifting of methyl Tolylpentanes can be synthesized groups on the ring are overcome and the by alkylation of toluene with pentenes well-defined zeolite pore structure from a variety of sources over many provide for isomerization to 2,6 and 2,7 solid acid catalysts. Typically a Ydimethyl tetrahydronaphthalene, the later zeolite is used at 160oC, 1310 kPa then readily isomerize to 2,6-DMT. In -1 the following step, the hydroisomerized pressure, WHSV of 2-6 h and at 2,6-DMT is dehydrogenated over a low toluene to pentenes molar ratio of about acidity catalyst such as Pt/Na-ZSM-5 to 10 in a fixed bed reactor, where the the desirable 2,6-DMN, which is olefin conversion is almost 100%. purified via crystallization and Reforming of the tolylpentanes is carried adsorption. out over a typical bimetallic, bifunctional catalyst such as

147

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

CH3 + Zeolite Y C5H10 Alkylation

CH3

TPs C5H11

CH3 Pt / Re / Al2O3 / Cl Reforming C5H11 Pd / Beta Hydroisomerization DMTs

DMNs

Pd / Beta Hydroisomerization 2,6-DMT

Pt / Na-ZSM-5 Dehydrogenation 2,6-DMN

Fig. 2 Synthesis of 2,6-dimethyl naphthalene from toluene and pentenes (ChevronTexaco) 4.2.2 Isopropylation of naphthalene The shape selectivity effects are more pronounced with the isopropylation or t-butylation of naphthalene using propene (or isopropanol) and t-butanol, respectively. The alkylation can be done in a fixed bed flow reactor or in a batch reactor using a suitable solvent (cyclohexane, for example) either at autogeneous or high pressures. Among the zeolites, large pore zeolites, particularly Hmordenite are suitable for the formation of the least bulky products from polynuclear aromatics such as naphthalene and biphenyl, although the selectivity changes with reactants and zeolites [19-21]. We will briefly discuss the shape selective isopropylation of naphthalene. The general reaction scheme is given in Fig. 3. As mentioned, 2,6-DIPN is one among the ten diisopropylnaphthalenes isomers that can possibly form in this reaction. It may be noted that the ultimate selectivity to 2,6DIPN may be determined even at the monoalkylation stage of the reaction, i.e., the ratio of - to isopropylnaphthalenes (1-IPN vs. 2-IPN) as the primary products. In subsequent alkylation step, the 2,6- and 2,7-DIPN probably have the advantage of shape selective formation due to their linear and almost linear molecular dimensions that can match with the given pore structure of a zeolite, compared to other isomers. The catalyst is normally tuned to be selective to such isomers and characterized by definite silica to alumina ratio, porosity, NMR characteristics and a symmetry index.

Alkylation

148
1,5-DIPN

1,4-DIPN

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

Isomerization

1,4-DIPN

1,3-DIPN

1,7-DIPN

2,7-DIPN

1,6-DIPN 1,5-DIPN

2,6-DIPN

Transalkylation
2,6-DIPN

1,3-DIPN

2,7-DIPN

Fig. 3. Reaction path of the isopropylation of naphthalene to give different di-substituted isomers through alkylation, isomerization and transalkylation conversion was lower compared to the Table 3 summarizes the results other two zeolites. The distribution of reported by Sugi et al [22] on the use of the isomers was found to be constant three different large pore zeolites, viz., during the reaction. H-Mordenite has H-mordenite, HY and HL zeolites in two types of pores, which are comparison with a medium pore ZSM-5 perpendicular to each other, one is the zeolite, whose pores are inaccessible to 12-membered ring elliptical pores (0.67 naphthalene and IPN isomers. A high x 0.71 nm) and the other 8-membered selectivity for 2,6 and 2,7-DIPN was ring pores (0.29 x 0.57 nm). The active observed with HM, although the

149

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

centers are the Brnsted acid sites, the strength of which is determined by the Si/Al ratio. They are inside the elliptical pores and naphthalene diffuses through these pores preferentially. These pores restrict the transition state of the intermediates to form the least bulky products of alkylation, viz., 2-IPN among the monoalkylaromatics and 2,6DIPN among the di-substituted isomers.

The formation of mono- and disubstitution in the - position is prevented because the corresponding transition states have bulky conformation, which requires larger space than is available at the acid sites inside the pores.

Table 3 Isopropylation of naphthalene over conventional zeolites (from Ref. 22) _______________________________________________________________________ Catalyst SiO2/Al2O3 Conversion Product Distribution of di-isopropyl naphthalene,% Ratio % 1,3- 1,4- 1,5- 1,6- 1,7- 2,6- 2,7_______________________________________________________________________ HZSM-5 70 1.0 HY 7.3 96.1 23.7 HL 6.1 95.1 39.9 HM 25.3 68.3 5.3 Among the 2,6 and 2,7-isomers, the difference in the diffusion rates of the two may contribute to the observed selectivity. Horseley et al have simulated the diffusion of 2,6 and 2,7diisopropylnaphthalene in HM and HL pores [23]. Based on the minimum energy profiles for their diffusion in HM pore, the diffusion of 2,6 isomer with a calculated energy barrier of 4 kcal mole-1 is significantly less hindered than that of 2,7-isomer which carries a value of 18 kcal mole-1. However, there is no significant barrier for either isomer against their diffusion in HL pores. The transition state for the formation of 2,6isomer would be linear whereas that of 2,7-isomer can be assumed to be somewhat bent configuration. On the basis of minimum energy profiles for the di-substituted isomers in HM pores, the transition state for 2,7 isomer would have a higher energy than that for 2,6isomer, resulting in higher activation -0.6 0.2 6.8 4.9 32.6 31.2 7.9 6.7 15.3 16.3 6.7 7.2 3.8 1.9 7.1 6.1 50.8 24.9 energy for the formation of 2,7-isomer. In other large pore zeolites such as HY and HL, there may be no energy difference between the isomers and hence they are not regio-selective for the formation of 2,6 and 2,7 isomers. There is also a suggestion that the formation of cationic intermediate for the 2,6-isomer is favored electronically as compared to that of 2,7 and 2,3-isomers during the isopropylation of 2isopropylnaphthalene inside HM channel [24,25]. An AlCl3 immobilized MCM-41 is described by Zhao et al as an interesting catalyst for the isopropylation of naphthalene [26]. While several functionalized mesoporous materials have been investigated for this reaction, this report illustrates the selectivation of a catalyst. The silanol network in MCM-41 may be responsible for many side reactions. For the modification of the external surface, silylation of MCM-

150

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

41 was carried out with trimethylchlorosilane before incurporation of AlCl3. A comparison of the

activity and selectivity of the two samples is given in Table 4.

Table 4 Isopropylation of naphthalene on AlCl3-MCM-41 (Data from Ref. 26) Catalyst Conversion, wt% DIPN DIPN Selectivity, % yield,% 2,6- 2,6-/2,7- others AlCl3-MCM-41 88.5 65.5 41.0 2.2 40.4 AlCl3-Ext.Surface Modified MCM-41 85.2 72.3 60.9 2.8 17.5 Conditions: Parr reactor; naphthalene = 0.02 mol; iPrOH = 0.04 mol; solvent = cyclohexane, 200 ml; Temp. = 200oC; Duration = 4 h; Autogeneous pressure Even though the conversion of naphthalene is slightly lower on the silylated sample, there is an appreciable increase in DIPN yield, which is more enriched with 2,6- and 2,7-DIPN isomers than in the case of the first catalyst. Silylation has not only passivated the external active sites, but has also narrowed the meso pore dimension from 2.3 to 2.0 nm, which probably favored the formation of less bulky 2,6-isomer in the product stream. 4.2.3 Isopropylation of biphenyl Alkylation of biphenyl is another good model for shape selective catalysis involving bulky organic compounds. The symmetrically substituted dialkyl biphenyl (4,4-diisopropylbiphenyl, 4,4DIPB, for example) is a precursor for the synthesis of advanced materials and polymers. Among the three mono-alkyl biphenyls (IPBs), which are the primary alkylated products in the reaction, only 4-IPB will enter the pores in a reactant shape selective reaction that uses a zeolite of appropriate pore structure (Fig. 4), if this mechanism operates in the isopropylation of BP. Among the DIPB isomers, the least bulky and linear 4,4DIPB will diffuse out preferentially from the DIPB mixture inside the pores. The composition of DIPB isomers will be near equilibrium inside the pores. From a consideration of restricted transition state selectivity, the 4,4-DIPB isomer will be obtained preferentially because the transition state will be restricted least among other isomers inside the pore, if this mechanism operates. A clear-cut discrimination

Reactant shape Selectivity

151

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

Product shape Selectivity

Restricted Transition State Selectivity

Fig. 4 Mechanisms of shape-selective catalysis in isopropylation of biphenyl within the pores of a zeolite during the between product selectivity and reaction. restricted transition state mechanism is Amorphous silica-alumina and a difficult, but the number of large pore zeolites have been product composition inside the pores, if investigated in this reaction. As it could be determined, will be able expected, HZSM-5 gives a low indicate which of the shape selective conversion of BP and no dialkyl product process is operating with a particular is formed. A comparison of the zeolites. This can be done by careful activities of the zeolites and their analysis of the product encapsulated respective selectivities to mono-

152

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

isopropyl and di-isopropylbiphenyl the work of Sugi et al [27]. isomers is given in Table 5, derived from Table 5 Isopropylation of Biphenyl catalyzed by solid acid catalysts (Data from Ref. 27) __________________________________________________________________ Catalyst Si/Al Conv. Selec. for IPB(%) Selectivity for DIPB (%) Ratio (%) 2344,4 3,4 3,3 others __________________________________________________________________ HM 46 48 5 24 71 78 14 2 6 HY 12 83 7 48 45 11 22 13 54 HL 12 84 29 25 46 10 13 6 71 SilicaAlumina 8-9 84 18 32 50 25 26 8 41 HZSM-5 100* 6 16 30 54 __________________________________________________________________ Conditions: Temp = 250oC; Catalyst = 1 g; BP = 50 mmol; propene = 100 mmol; Solvent = trans decalin, 20 ml; Duration = 4 h. *Temp. = 300oC It is seen that H-mordenite is the most effective catalyst in this reaction, even though the conversion of BP is much lower on HM compared to other large pore zeolites (HY and HL) and amorphous silica-alumina. The selectivity among the mono-isomers (IPB) is very high on HM and hence, further alkylation to disubstitution is expectedly very high, following the shape selectivity mechanism. It is also to be noted that among the mono isomers, 4-IPB is highly reactive compared to 3IPB. There is an order of magnitude difference in the initial rates between the two isomers on HM. When present in large amounts, 4-IPB is isopropylated selectively to 4,4-DIPB and 3-IPB is slowly alkylated to products which contain principally 3,3 and 3,4-DIPBs. The catalysis over HY and HL zeolites, on the other hand, is quite different from that over HM. The isopropylation is non-regio selective in the formation of IPBs and DIPBs. By a careful analysis of the bulk products and encapsulated products in the catalysts (HM) used for this reaction, Sugi et al [27] found a high selectivity for 4,4-DIPB in encapsulated products. This indicates that restricted transitionstate mechanism operates in the catalysis over HM, and that product selectivity mechanism is probably not important. The shape selective catalysis over HM is considered to occur inside pores under moderate reaction conditions. They also reported that the selectivity to 4,4-DIPB was not influenced by external surface area. This observation is probably restricted to HM, for in many other instances the external surface of a zeolite contributes to other consecutive and non-selective reactions. Another parameter is the reaction temperature, which influences the product distribution. At low temperatures, 2and 4- positions in the aromatic nucleus exhibit a higher reactivity for electrophilic substitution than the 3position. At higher temperatures, thermodynamically more stable isomers are formed by the isomerization of IPB and DIPB, in the absence of any steric

153

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

control. That explains why other zeolites do not exhibit such a high selectivity to 4,4-isomer as in the case of HM. The formation of 4,4-DIPB over H-offretite and SAPO-11 was also found to be shape selective in comparison to attainable thermodynamic compositions [28]. The former was less selective than HM, while SAPO exhibited a comparable selectivity for 4,4-DIPB. Many new one-dimensional, large pore zeolites have been examined by Sugi et al [27] in this reaction.
100

Typical selectivity for 4,4-DIPB is given in Fig. 5 for a number of zeolites under identical reaction conditions. It is seen that shape-selective catalysis occurs only over one- and twodimensional 12- or 14-membered pores including HM, which exhibited the highest selectivity for 4,4-DIPB. These differences of the selectivity for DIPBs reflect the difference of microporous environments of catalytically active sites.

ZS M-12

SS Z-24

HM

SAPO -5

SS Z-31

80

Selectivity of 4,4'-DIP N (% )

60

40

CIT -5 UT D-1

ZS M-22

20

Fig. 5. Selectivity for 4,4-DIPB over various zeolites in the isopropylation of biphenyl. Reaction conditions: Temp.= 250oC; C3= pressure = 0.8 Mpa [from Ref. 27] 5. Effects of modification of zeolites for enhanced selectivity We have already considered several methods by which some of the reactions that lead to non-selective products can be controlled. Catalytic activity is not always proportional to acid density of zeolites. Major part of the dense acid sites inside the pores of zeolites with a low Si/Al ratio promotes coke deposition, which is rapid in the initial stages of a reaction over large pore zeolites. Severe deactivation of the catalyst often follows as in the case of HY. Dealumination of a zeolite serves the purpose of reduction of the coke deposits and enhance the catalytic performance, in spite of decrease of acid density. A dealuminated mordenite with a SiO2 to Al2O3 ratio of about 200 is

HY

CIT -1

HL

154

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

reported to have a very high selectivity for 4,4-DIPB [28]. Apart from the earlier techniques of passivation of the external surface by silylation or selective coking, ion exchange or impregnation of the zeolite with alkaline earth or rare-earth metal cations (an excess of the same as oxide) is reported to improve the selectivity of the product in some alkylation reactions. The cation exchange property of a HSelectivity for DIPB isomers
80

base property of a zeolite. Many instances are known when a partially exchanged Na form of a zeolite is used to provide weak acidity. In the alkylation reaction cited above, the external surface modification is reported to be done through incorporation of rare earth metal oxides of which ceria modified HM effectively deactivates the external acid sites with minimum pore blockage.
100

100

Selectivity for DIPN isomers (% )

Sel ec tivity of DIPB isomers (% )

80

80

60

(3,3'-)

60

40

(3,4'-)
40

2,7-DIPN
20

(4,4'-)

40

(Conv.)
20

20

Others
0 0 10 20 30 40

1,6-DIPN
50

0 0 5 10 15 20 25 30

Ceria Loading (wt-% as Ce)

Ceria Loading (wt-% as Ce)

zeolite can be used to control the acidFig. 6. Ceria modified H-mordenite catalyst. Influence of cerium content on the conversion and selectivity in isopropylation of naphthalene (left) and biphenyl (right) (From Ref. 27) Fig. 6 Summarizes the effects of ceria incorporation into HM zeolite on the selectivity for 2,6-DIPN in the isopropylation of naphthalene on the one hand and selectivity for 4,4-DIPB in the isopropylation of biphenyl on the one other. The selectivity increases from about 50% to 80% for 4,4-DIPB with about 10 wt.% cerium on HM. The isomerization of 4,4-DIPB was effectively prevented although the catalytic activity decreased to some extent. Similar improvement of selectivity for 2,6-DIPN has been noticed by ceria modification on HM [29]. Although it is not clear as to how the external acid sites are passivated by the deposition of fairly large concentrations of such oxides on zeolites, this technique seems to be more effective in many instances. A HY zeolite modified by 1% MgO and 7% La2O3 is reported to provide better activity, selectivity and catalyst stability than unmodified HY in the liquid phase alkylation of naphthalene with long chain (C11-C12) olefins [30].

155

Conversion (% )

60

2,6-DIPN

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

6. Concluding Remarks Some of the shape-selective catalysis of importance in the production of petrochemicals and chemical intermediates is based on the proper design of the catalysts for which zeolites have been of great significance. Examples of disproportionation (for the production of p-xylene) and alkylation of aromatics, both mono nuclear and dinuclear have been cited wherein the selectivity to the desired regio isomer could be maximized by the right choice of the zeoites and the modifications thereof. While the well-defined pore structure of zeolites enables one to have reasonable control over shape catalysis, the non-selective reactions taking place on the external surface cannot be avoided. Most of the efforts are, therefore, aimed at surface passivation and fine-tuning of the pores to a level that the selectivity to the desired product could be enhanced. These modifications include, dealumination that decreases the rate of coke deposition in the initial stage of the reaction, silylation, selective coking and ion-exchange with alkali, alkaline earth or rare earth metal ions with the zeolites. These have been particularly effective with large pore zeolites, such as H-mordenite, H-Y and H-beta zeolites, where rapid coking of the catalysis under the reaction conditions has been a serious issue. Production of cumene, 2,6-dialkyl naphthalene and 4,4-dialkyl biphenyl are some of the examples that have been discussed based the reports available in the recent literature on this subject. Improvement in selectivity to the desired product even at the cost of conversion of the substrate (aromatics) makes sense because of the problems of separation from other isomers and environmental issues connected with the disposal of

undesired products. Use of new acid catalysts that will be active at moderate conditions of reaction (low temperatures), employing suitable solvents that can reduce coke formation or carrying out reactions at super critical conditions are some of the directions in which current investigations are focused. Acknowledgement The author wishes to thank JSPS, and Prof. Y. Sugi for the opportunity to discuss some of the examples of shape selective catalysis described in this paper. Thanks are also due to Mr. B. Murugan for his help in the preparation of the manuscript. References [1] S.M. Csiscery, Zeolites, 4 (1984) 202. [2] P.B. Venuto, Micropor. Mater., 2 (1994) 297. [3] K. Tanabe and W.F. Holderich, Appl. Catal. A: General, 181 (1999) 399 [4] A.V. Ramaswamy, in Petrotech-99, Third.Intl.Petrol.Conf., New Delhi, 1999, vol.III, p.1. [5] C. Song, in Shape Selective Catalysis: Chemicals Synthesis and Hydrocarbon Processing ACS Symposium series 738, {Eds. C. Song, J.M. Garces and Y. Sugi} Am.Chem.Soc., 1999, p. 248. [6] B.S. Rao, R.A. Shaikh and A.V. Ramaswamy, in Shape Selective Catalysis-Chemicals Synthesis and Hydrocarbon Processing ACS Symposium series 738, {Eds. C. Song, J.M. Garces and Y. Sugi], Am. Chem.Soc., 1999, p.225 [7] B.S. Rao and R.A. Sahikh, Indian Patent Appl. 503/DEL/1994 [8] Chemical Week, 8 (1994) 34. [9] J.C. Cheng et al, US Patent 5,453,554 (1995).

156

Catalysts Selectivity., A V Ramaswamy, Bulletin of the Catalysis Society of India,2(2003) 140-156

[10] M. Guisnet, J. Mol. Catal. A: Chemical, 182-183 (2002) 367. [11] J. Weitkamp et al, Stud..Surf .Sci. Catal., 60 (1991) 291. [12] M. Neuber, H.G. Karge and J. Weitkamp, Catal. Today, 3 (1988) 11. [13] D. Fraenkel, M. Cherniavsky, B. Ittah and M. Levy, J. Catal., 101 (1986) 273. [14] J. Aguilar-P, A. Corma, J.A. de los Reyes, L. Norena, G. Munoz, J.M. Sanchez, A.,Torales and I. Hernandez, 13 IZC, Montpellier, Stud. Surf .Sci. Catal., 135 (2001) 282. [15] L.D. Lillwitz, Appl. Catal. A: General, 221 (2001) 337. [16] G. Pazzucini, G. Terzoni, C. Perego and G. Bellussi, 13 IZC, Montpellier, Stud. Surf. Sci. Catal., 135 (2001) 152. [17] C.Y. Chen, W.L. Schinski, D.J. ORear and T.V. Harris, U.S. Patent, 5,955,641 (1999). [18] C.L. Muson, P.C. Bigot and Z.A. He, U.S. Patent, 6,057,487 (2000). [19] Y. Sugi and M. Toba, Catal. Today, 19 (1994) 187. [20] Z. Liu, P. Moreau and F. Fajula, Chem.Commun., (1996) 2653 [21] Z. Liu, P. Moreau and F. Fajula, Appl. Catal. A : General, 159 (1997) 305. [22] Y. Sugi, Y. Kubota, T. Hanaoka and T. Matsuzaki, Recent Res. Devel. Mat. Sci. Engg., 1 (2002) 395; K. Nakajima, T. Hanaoka, Y. Sugi, T.Matsuzaki, Y.

Kubota, A.Igarashi and K. Kunimori, in Shape Selective Catalysis Chemicals, Synthesis and Hydrocarbon Processing ACS Symposium series 738 {Eds. C. Song, J.M. Garces and Y. Sugi} Am.Chem.Soc., 1999, p. 271... [23] J.A. Horesley, J.D. Fellmann, E.G. Derouane and C.M. Freeman, J. Catal., 147 (1994) 231. [24] C. Song, X. Ma, A.D. Schimitz and H. Scholert, Appl. Catal.,A: General, 182 (1999)175. [25] T. Kamatsu, J.H. Kim and T. Yashima, in Shape Selective CatalysisChemicals Synthesis and Hydrocarbon Processing ACS Symposium series 738, {Eds. C. Song, J.M. Garces and Y. Sugi], Am. Chem.Soc., 1999, p.162. [26] X.S. Zhao, G.Q. Lu and C. Song, Chem. Commun., (2001) 2306 [27] Y. Sugi, Y. Kubota, T. Hanaoka and T. Matsuzaki, Catal. Surveys Japan., 5 (2001) 43. [28] T. Matsuda, T.Kimura, E. Herawati, C. Kobayashi and E. Kikuchi, Appl. Catal., A: General, 136 (1996) 19. [29] J-H. Kim, Y. Sugi, T. Matsuzaki, T. Hanaoka, Y. Kubota, X. Tu, M. Matsumoto, A. Kato, G. Seo and C. Park, Appl. Catal. A: General, 131 (1995) 15. [30] H. Guo, Y. Liang, W. Qiao, G. Wang and Z. Li., Stud. Surf. Sci. Catal., 142 (2002) 999.

157

Anda mungkin juga menyukai