Vous êtes sur la page 1sur 33

Transworld Research Network 37/661 (2), Fort P.O.

, Trivandrum-695 023, Kerala, India

Zeolites: From Model Materials to Industrial Catalysts, 2008: 357-389 ISBN: 978-81-7895-330-4 Editors: J. ejka, J. Perz-Pariente and W. J. Roth

14

Zeolites and zeolite-like materials in industrial catalysis


C. Perego and A. Carati ENI SPA, Via G. Fauser 4, 28100 Novara, Italy

Abstract
This article sets out to survey the remarkable development that has taken place in zeolite catalysis for refining and petrochemical industries, highlighting what appears to be the most significant recent progress. In the first part of the article a general review of the main features of zeolites as catalyst is discussed. Than an overview of the catalytic processes based on zeolites is reported, underlining the peculiar properties of the different zeolites utilized. The overview will cover the main catalytic reactions characterizing refinery and petrochemistry: cracking, hydrocracking, dewaxing, reforming, naphta hydrodesulfurization, oligomerization, aromatic saturation, aromatization, olefin/paraffin alkylation, aromatic alkylation and transalkylation, disproportionation and isomerization.
Correspondence/Reprint request: Dr. C. Perego, Centro Ricerche Energie non Convenzionali Istituto Eni Donegani, Direzione Strategie e Sviluppo, ENI S.p.A., ENI S.p.A., Via Fauser 4, 28100 Novara (NO), Italy. E-mail: carlo.perego@eni.it

358

C. Perego & A. Carati

1. Introduction
Catalysis is truly a well-established technology, thanks to remarkable improvement occurred in the 20th century from the point of view of both fundamental knowledge and application development. Catalyst-based chemical synthesis accounts for over 60% of todays refining and petrochemical production and 90% of current processes [1]. Catalyst industry has a very large impact on the process industry: in 2004 the market of catalysts was of $6.5 billion contributing to a refining and chemical industry production worth over $1.5 trillion. According to a recent published report by Business Communications Company, the catalyst market is expected to grow to nearly $13 billion in 2009, of which 90% is currently related to petroleum refining [2]. The requirement of the refining and petrochemical industry to catalyst companies is the development of high activity catalysts that increase reaction yields, enhance life cycle, selectivity and conversion rate. Catalysis is a key technology for the increase of both the profit and the greening of industry. Indeed, the improvement in catalyst performances plays a main role in the reduction of both wasteful by-products and energy consumption, moving in the direction of a sustainable chemistry. In particular, heterogeneous catalysis using solid catalysts shows several advantages with respect to homogeneous catalysis, because of the reduction of the separation problems and of salt and waste production. Today the majority of solid acid catalysts used in refinery and petrochemical industry are based on zeolites. Zeolites are a component of about 130 catalysts on the total of 840 catalysts used in the world [3].

Zeolite properties: The reason of their success as catalysts


The first steps in zeolite story [4,5], resulting in their first application in refinery as industrial catalysts are summarized in Figure 1. Zeolite or zeolite-like materials are porous crystalline three-dimensional framework based essentially on tetrahedrally coordinated T-atoms characterized by the presence of channels (and eventually cages) with the smallest opening larger than six T-atoms (where T = Si, Ge, Ti, Al, Ga, B, Fe, Be, P, etc.). The T-atoms forming the framework are linked together by the common sharing of oxygen ions to form polynuclear complexes. Different types of zeolites result from differences in the way how the T-atoms may link in the space. The tri-dimensional networks encompass welldefined micropores with uniform pore diameters that are similar to the dimensions of simple organic molecules. Pore diameters depend on the number of T-atoms in the ring around the pore. According to the number of T-atoms forming the pore opening zeolites, are grouped as follows: 1) 2) 3) 4) small-pore zeolites (8- membered rings), e.g., Erionite; medium-pore zeolites (10-membered rings), e.g., ZSM-5, ZSM-11, ferrierite, MCM-22 large-pore zeolites (12-membered rings), e.g. zeolite Y, Beta, mordenite, ZSM-12; extra-large-pore zeolites (14-membered rings), e.g. CIT-5 and UTD-1.

Zeolites and zeolite-like materials in industrial catalysis

359

1962: Introduction of synthetic faujasites (zeolites X and Y) in refinery (FCC of heavy petroleum distillates) 1961: Alkylammonium cations were introduced in the synthesis recipes 1959: Union Carbide classified zeolites as a new class of industrial catalyst for hydrocarbon conversion 1957: J.A. Rabo related the strong acid catalytic activity of X and Y zeolites to their crystallinity 1953: Linde Type A zeolite becames the first synthetic zeolite to be commercialized 1949: Zeolites A, X, and Y were discovered by R.M. Milton and D. Breck at Union Carbide Corporation. 1948: The era of synthetic zeolites starts with Barrers synthesis of small-port mordenite and identification of polycrystalline materials by X-ray powder data. 1945: R.M. Barrer classified zeolite minerals depending on the size of absorbable molecules 1926: The adsorption properties of chabazite were attributed to its porosity 1756: The first zeolite mineral (stilbite) was described in Sweden by A. Cronstedt

Figure 1. Zeolite: from the discovery to the introduction in refinery.

Tridimensional networks can act as reaction channels, whose activity and selectivity will be enhanced by introducing active sites. The presence of Al (or another trivalent element) in the framework produces a deficiency in electrical charge that must be locally neutralized by the presence of an additional cation within the pore structure. By ion exchange it is possible to substitute one cation for another to modify the catalytic properties. When cationic sites are exchanged by H+, acid sites are obtained. The Brnsted acid sites are associated with framework aluminum or other trivalent ions (e.g. B, Ga, Fe). The number of the Brnsted acid sites is directly proportional to the concentration of these framework trivalent ions. Their strength depends on the kind of trivalent ions and on its concentration. In the case of Al, the strength of the Brnsted acid sites is usually inversely proportional to the concentration of framework aluminum, up to about a silica/alumina ratio of 10.

360

C. Perego & A. Carati

Lewis sites in zeolites are usully associated with tricoordinated or extraframework Al. The simultaneous presence of Brnsted and Lewis sites could enhance the acid activity of zeolites. Besides, extraframework Al affects different reactions differently and, therefore, could affect selectivities. When some silica atoms of the framework are isomorphically substituted by Ti, the zeolite shows redox properties. While the active sites present in the channels are related with catalytic activity of zeolites, the diameter, interconnectivity and dimensionality of pore system give rise to the molecular sieving action. Zeolites are able to discriminate among reactants, products, or reaction intermediates according to shape and size of their pores. This ability, called shape selectivity, is a very important property of zeolite catalysts. According to Degnan [6], few concepts have had a greater impact on the design and development of novel catalytic processes for petroleum refining and petrochemical manufacture than that of shape selective catalysis. Firstly proposed by Weisz in 1960 [7] and extended by Csicsery [8] the concept of shape selectivity is today the basis of several commercial processes. Three main types of shape selectivity have been described: 1. Reactant shape selectivity (RSS) distinguishes between competing reactants: only some of them can easily reach the internal active sites due to the geometry at the pore entrance and to their energy barrier for diffusion inside the pores. 2. Product shape selectivity (PSS) refers to the situation where the pore diameter discriminates on the basis of their size and the energy barrier for diffusion among products that may be formed in the larger pore intersections. 3. Restricted transition-state selectivity (TSS) acts when the geometry of the pore around the active sites imposes steric constraints on the transition state inhibiting the formation of too large ones. Other types are subject to vigorous debate: inverse shape selectivity, molecular traffic control, pore mouth-key-lock shape selectivity, the Window Effect, and the Nest Effect [9]. According to ejka and Wichterlov [10] the success of zeolites or zeolite-like catalysts compared with conventional solid acids can be so summarized: (a) well-defined inorganic crystalline structures, with a variety of structures differing in channel diameters, geometry, and dimensionality; (b) a precisely defined inner void volume providing high surface area; (c) the ability to adsorb and transform molecules in the inner volume; (d) isomorphous substitution of some trivalent cations into the silicate framework enabling tuning of the strength and concentration of the acid sites; (e) shape selectivity, given by the ratio of the kinetic diameters of the reactants, intermediates, and products to the dimensions of the channels. (f) environmental tolerance.

Zeolites as catalysts: Markets


In 2001 the overall consumption of synthetic zeolites was around 1.46 million metric tons [11], concerning three major fields of activity: detergents (81%), adsorbents and desiccants (6%) and catalysis (13%).

Zeolites and zeolite-like materials in industrial catalysis

361

The uses as catalysts only accounted for a low part of the total volume. However, from the point of view of the market value the overall consumption of synthetic zeolites as catalysts represents about 2 billions $, over passing the 55% of the total market value, being about the 95% devoted to Fluid Catalytic Cracking (FCC) process.

Zeolites as catalysts: Industrial applications


Almost all the zeolites used as catalysts in the world are applied in refining and petrochemical industry. Today the current 83 million bpd (barrel per day) word crude oil production continues to grow at about 1,8% per year, due to the increasing demand from geographic area of growing economies such as China and India [12]. From a catalysis perspective, the new challenges can be related to three main needs of refinery:

Changes in the product distribution in order to meet the market demand: the future projections show the increase of demand for high quality middle distillates and the decreasing for residue and fuel oil demand. - The increase in the ability to process lower-quality feedstocks: a large part of the estimated 1.2 trillion barrels of crude in the world is in the medium-to-heavy range. Refinery process complexity and catalysts with improved activity and resistance to metals contamination are expected. - The respect of increasingly stringent environmental regulations regarding product quality (Table 1). In refining, zeolites have been introduced since the beginning of the sixties, when Ytype zeolite started to be used in catalytic cracking. From the early 1960s on, the use of synthetic zeolites in catalysis and in related adsorption separation processes has dramatically
Table 1. Time table for clean fuel requirements in EU [13,14].

362

C. Perego & A. Carati

transformed petroleum refining by vastly increasing the yield of high-quality fuels and reducing capital and operating costs, energy requirements, and environmental impact. Zeolites also played a major role in allowing the efficient reformulation of fuels. An interface role between refinery and petrochemistry can be ascribed to processes which maximize olefins and aromatics production, so providing opportunities both for modern refinery and petrochemical industry [15]. Aromatics, together with olefins, are indeed the main building blocks of petrochemical industry. Besides new legislative limits on aromatics and benzene concentration in fuels switch the application of aromatics from fuels to petrochemicals, especially for benzene and xylenes. In modern refinery, zeolite catalysts are involved in several processes, as reported in Figure 2. In petrochemical industry benzene, toluene and xylenes (BTX) are three basic building blocks for the production of most intermediates of several aromatic derivatives (Figure 3). Zeolite catalysts have been present in petrochemical industry since the seventies when mordenite replaced the amorphous acid support in the isomerization of xylenes [14]. Today zeolite based catalysts find application in several processes, related to the synthesis and the interconversion of alkylbenzenes, as alkylation, transalkylation, isomerization and disproportionation [17]. During the last three decades the main innovation in this field has been represented by the introduction of new zeolitic catalysts with the aim to improve the overall selectivity
Carbons
Aromatization, Oligomerization, Cracking, Olefin/paraffins alkylation, Olefin isomerization

GASES 4 7 LSR NAPHTHA MIDDLE DISTILLATES LUB OILS VGO RESIDUE

Hydroisomerization Reforming, Hydrodesulfurization, Transalkylation of heavy aromatics

Dewaxing, Saturation of heavy aromatics

16 36 44 80

Dewaxing Fluid catalytic cracking, Hydrocracking, Mild pressure hydrocracking

Figure 2. Refinery catalytic processes using zeolite containing catalysts.

Zeolites and zeolite-like materials in industrial catalysis

363

Figure 3. Petrochemical derivatives from BTX (benzene, toluene and xylenes) (adapted from Tsai [16]).

364

C. Perego & A. Carati

of the processes and to comply with the requirement of the more stringent environmental legislation. In fact, most of the oldest technologies, mainly the alkylation/transalkylation ones, use strong mineral acids or Lewis acids (e.g.: HF, H2SO4, AlCl3) as catalysts. The success of zeolites in fine chemicals is more recent and benefits mainly from the zeolitic materials developed mostly for the first two industrial sectors (i.e. refinery and petrochemistry). Significant improvements are still possible, especially by the use of different zeolite structures tailored on the application requirements. Indeed, fine chemistry comprises a large variety of reactions and the catalysts are frequently not completely optimized. Cross-sector technology transfer can greatly accelerate the greening of many highly wasteful chemical processes, by substitution of both soluble acid (such as H2SO4 and AlCl3) and organometallic catalysts [18]. The high added value of the fine chemical products specialties and pharmaceutical intermediates makes the high cost of certain zeolitic material acceptable. However, the quantities of products involved in the fine chemicals sector are small, and the volumes of zeolitic materials used will still be minor compared with those of refining and petrochemical industry. The use of zeolite in fine chemical will not treated in this paper. Excellent reviews on synthetic uses of zeolites over the years are available in the literature [18-21].
Table 2. Structural characteristics for zeolites relevant as industrial catalysts.

Zeolites and zeolite-like materials in industrial catalysis

365

Today 176 zeolite and zeotype structures are recognized by International Zeolite Association [22]. In principle it is possible to select the most appropriate zeolite (acid strength, dimensionality, pore-size, and pore-shape) from a large variety of structures to catalyze almost any reaction. Despite the continued increase in the availability of new molecular sieve materials and the associated research activities into their applications, only a handful are used in industrial processes (MFI, BEA, MWW, MTW, MOR, MAZ, FER, FAU, LTL, CHA, AEL, TON), being these zeolites often used in more than one process application [6,23]. Table 2 reports the main characteristics of the above mentioned zeolites. This trend should be modified by increasing of the application of zeolite-based catalysts in speciality and fine chemical industry where the catalyst cost engraves less on the overall cost of the processes. Playing the catalyst a very important role in industrial processes sometimes the type of catalyst used is not revealed. In Table 3 there are reported refinery and petrochemical applications where the use of zeolite based catalysts is described. These applications will be discussed in this paper, with particular attention to improvements in consolidate processes or to new processes developments.
Table 3. Refinery and petrochemical applications using zeolite based catalysts.
IZA code AEL BEA CHA FAU Zeolite type SAPO-11 Beta SAPO-34 Y, USY Kind of application Dewaxing Dewaxing, Aromatic alkylation and transalkylation Methanol to olefins Fluid catalytic cracking, Hydrocracking, Mild pressure hydrocracking, Saturation of heavy aromatics, Olefin/paraffin alkylation, Aromatic alkylation and transalkylation Skeletal isomerization of n-butenes Light paraffin hydroisomerization, Fluid catalytic cracking, Dewaxing, Reforming, Reformate alkylation process, Hydrodesulfurization of naphtha, Catalytic cracking of olefinic/paraffinic streams to propylene, Aromatization, Oligomerization, Methanol to olefins, Xylenes isomerization, Toluene disproportionation, Toluene alkylation with methanol, Aromatic alkylation, Ammoximation, Beckmann rearrangement Light paraffin hydroisomerization, Oligomerization, Aromatic alkylation and transalkylation Aromatic alkylation Aromatic alkylation Olefin skeletal isomerization

FER MAZ MFI

Ferrierite Omega ZSM-5, TS-1, S-1

MOR MTW MWW TON

Mordenite ZSM-12 MCM-22 ZSM-22

366

C. Perego & A. Carati

2. Refinery processes based on zeolitic catalysts


Fluid Catalytic Cracking (FCC)
The process gives a selective conversion of a wide range of gas oils into high-value products. Typical feedstocks are straight-run or hydrotreated gas oils but also include lube oil extract, coker gas oil and residues. Typical products are high octane gasoline, light olefins and distillate. Flexibility of mode operation allows for maximizing the most desirable products: FCC unit can be operated to maximize gasoline or distillate production, or it can be operated at high severity to maximize petrochemical feedstocks production such as propylene. FCC units use small fluidizable catalyst particles of multifunctional catalysts which are composed of: different amorphous and crystalline mixed oxides, with different pore size distribution; metal passivators; SOx removal components; promoters for total combustion and octane enhancing additives. If all these components are important, zeolites are the heart of the catalyst. Zeolites provide vastly superior combinations of strong acid catalytic sites, uniformity of pore structure, and stability, all of which provide improved selectivity, yield, durability, and cost over non-zeolite alternatives, providing much higher yield of gasoline and other high-quality fuels per barrel of crude oil. Zeolite Y continues to be the primary zeolitic component in FCC catalysts, nearly 45 years after its first commercial introduction. While many research programs have attempted to identify alternative materials [15], zeolite Y continues to provide the greatest gasoline yield at the highest octane number with the greatest degree of catalytic stability. Other zeolites and molecular sieves have failed principally for their low selectivity and stability properties. Hydrothermal stability is a fundamental requirement for FCC catalyst that during the process is subjected to high regeneration temperatures (about 720 C) in presence of steam. From the point of view of catalyst formulation two main components constitute the FCC catalyst: the binder and the zeolites. At equilibrium a typical catalyst has a particle size distribution centred at 50-70 m (Figure 4), being made up of an amorphous matrix (typically a silica alumina) including micro-particle of Y zeolite (e.g. 2-10 m size). The production is usually performed by spray-drying a slurry of silica-alumina sol and of zeolite powder, as represented in Figure 5. The reactions occurring on Y zeolite during the cracking process are C-C cleavage, dealkylation, transalkylation, oligomerization, dehydrocyclization, isomerization, hydrogen transfer, coke formation [25]. Even if the crystal structure is the same, important improvements have been obtained on Y zeolite for FCC process, passing from rare-earth-ion exchanged zeolite Y (REY) to the ultrastable zeolite Y (USY), characterized by lower unit cell parameters and thus by reduced acid site density coupled to extra-framework aluminum or mesoporosity generation in the zeolite crystals. Optimization of framework and extra-framework composition affects the site accessibility and the final catalytic behavior, suppressing coke formation and improving cracking activity, hydrothermal stability, gasoline selectivity and octane number [15] (Figure 6). Since 1980s, ZSM-5 zeolite has been used as shape-selective catalyst additive in the amount of 0.5-3.0 wt% in the global catalyst to gain in octane number. Being ZSM-5 mixed with the catalyst, the gasoline component formed on the FCC catalyst can move from the catalyst to the additive for further upgrading, while, due to its smaller pore dimension,

Zeolites and zeolite-like materials in industrial catalysis

367

Figure 4. SEM micrograph of FCC catalyst.

Figure 5. Schematic representation of the production of a FCC catalyst, including the preparation of zeolite powder and the silica-alumina, followed by the spray drying (adapted from J.A. Moulijn et al. [24]).

ZSM-5 can not crack the large FCC feed molecules. For RSS effect only linear and mono methyl paraffins and olefins would have access to the high active sites, decreasing C7+ olefins and increasing C6- olefins [6, 25]. Besides a restricted TSS prevents the formation of the transition state required for hydrogen transfer and prevents the formation of coke within the zeolite pores [6].

368

C. Perego & A. Carati

Figure 6. Improvements in gasoline selectivity and research octane number (RON) by modifications in zeolite Y (adapted from T.F. Jr. Degnan [15])

The ability of small amounts of ZSM-5 added to the FCC unit to improve gasoline octane number while producing more light olefins has prompted a substantial amount of process and catalyst research into zeolite-based FCC additives [25]. The application of different zeolites as FCC catalyst additives has been extensively studied. Corma et al. [26] evaluated a large variety of zeolites, including Beta, Mordenite, SSZ-24, ZSM-5, Ferrierite, CIT-1, NU-87, NU-86, and MCM-22. Using the cracking of linear paraffin within the range of the gasoline fraction as test reactions, they suggested that the relative rates for cracking and hydrogen transfer and the amount of branched products are correlated with the zeolite pore dimensions and topologies. For example Figure 7 shows the Paraffin /Olefin ratio in the gas products against the void spaces formed within the zeolite structure by pores, cavities, channels crossing or lobes. Since 1990s the demand for incremental gasoline octane has diminished with increased oxygenate addition (e.g. MTBE and ETBE), while the ZSM-5 zeolite has been increasingly used to produce an additional amount of olefins (especially propylenes and butenes) to meet the petrochemical demand for olefins. Today the market is moving towards higher active ZSM-5 or new generation shape-selective additives. By using these ultra-high activity additive (i.e. OlefinsUltra, by W R Grace & Conn, Engelhard Maximum Olefins Additive) FCC units can achieve propylene yield as high as 8-10 %. FCC units revamped to operate in a petrochemicals mode are targeted till to 15-20 % of propylene yield [27, 28]. The optimization work on both Y and ZSM-5 structures in order to increase their catalytic performances have played a key role in the evolution of FCC catalysts, demonstrating that activity and selectivity of a zeolitic catalyst are affected not only by its crystalline structure but also by its compositional and morphological properties.

Zeolites and zeolite-like materials in industrial catalysis

369

Figure 7. Paraffin/Olefin ratio vs. void space dimensions in zeolites [26].

At present the changing needs of refiners present new challenges for FCC catalyst manufacturers: the varieties of crude and refinery processing schemes have resulted in a huge range of needs that require proper catalytic solution. Two of the most significant needs are coke selectivity and metals tolerance [29]. Today by a proper combination of process units either upstream or downstream from the FCC unit and the optimum catalyst formulation it is possible to process cheaper, heavier and more sour crude maximizing profitability. Optimization of FCC catalyst is a very complex field, indeed all the components and not only the zeolitic one are involved. For example to increase the residue processing, Grace Davison commercializes the IMPACT catalyst series that combine the new Z-28 or Z-32 high stability zeolites (structures not revealed), a new generation Integral Vanadium Trap and novel Tunable Reactive Matrix (TRM) technology. Through a combination of materials and processing technology advancements, TRM technology provides control of the matrix acidity and porosity, so that the catalyst matrix has properties specifically calibrated to the requirements of the feed to be processed. Grace Davison Impact catalysts are actually applied with TMR-100 and Z-28 zeolite in 15 units worldwide, 13 of which are processing residue as part of FCC units feedstock and with TRM-400 and Z-28 or Z-32 in 12 units with high level of Ni contamination in the residue feed [29]. Similarly, Engelhard, now BASF, has developed Flex-Tec and DSM technologies for residual fuel oil cracking. Flex-Tec comprises a matrix pore structure (DSM architecture) optimised for the facile transport of heavy feed molecules

370

C. Perego & A. Carati

and the resulting products. Also included are technology for the passivation of nickel and vanadium, as well for maintaining high stability in residual fluid oil FCC operations [30].

Hydrocracking
Traditionally, hydrocracking is used for upgrading of heavy feedstocks in presence of hydrogen, producing increased amount of low-sulfur distillates, including LPG, high octane gasoline, reformer naphtha, diesel and jet fuel. Unconverted products are excellent feedstocks for FCC, coker or lube oil basestock production. Typical feedstocks are heavy and extra heavy crudes, vacuum gas oil (VGO), vacuum residues (VR). Hydrocracking is a capital-intensive process, due to the capital investment, the high level of hydrogen consumption and operating pressures. In spite of its high cost, the hydrocracking of heavy fractions of crude oil, primarily VGO, is one of the fastestgrowing catalytic conversion processes in refinery. Its development is related to the excellent quality of the middle distillates, particularly atmospheric gas oil (AGO), that it produces and for which there is a growing demand in Europe. Hydrocracking is a flexible process: an extensive range of catalysts, various types of processes and process configurations are used to match the refiners objectives. Hydrocracking catalyst systems are more complex than FCC or hydrotreating catalysts, because the process includes the hydrogenation of aromatic hydrocarbons, the opening of naphthenic rings and isomerization. In particular polynuclear aromatics that can not be processed in FCC units, in a hydrocracking unit are first partially hydrogenated and then cracked [25]. The process is based on a bifunctional catalyst, associating hydro-dehydrogenation function, which is generally a metal sulfide phase, with acid function, which is an amorphous or crystalline mixed oxides (e.g zeolites) [1]. The zeolites most frequently used in commercial hydrocracking catalysts are Y-type zeolites. Other zeolites and mixtures of zeolites also are used. The zeolites often are embedded in a high-surface-area amorphous matrix, which serves as a binder. The metals can be supported on both the zeolite domain and the amorphous matrix [4]. The hydrocracking catalyst based on Y-type zeolites is in competition with the amorphous silica-alumina or silica-alumina-titanium oxide type catalysts. These latter are much less active and less stable, but more selective to middle distillates [9]. A role of RSS effect can be hypothesized to explain the differences in stability and selectivity between the two categories of catalysts. Internal acid sites of the zeolite are protected against large coke precursor molecules contained in the feed, even if the coking potential of these precursors is seriously reduced by the hydrogen pressure. This effect of protection can explain why the Y-type zeolite, used at lower temperature than amorphous catalysts, has more stable performances over time. Besides while all the molecules can react on the acid sites of amorphous catalysts, the largest ones can not reach the internal acid sites of the zeolite. Consequently, these molecules are only slightly converted. Furthermore, the zeolite transforms the molecules entering into its micropores, more in-depth. This results in reduced selectivity to middle distillates with a higher production of gasoline and LPG and higher kerosene content in the middle distillates. Another cause of the inadequate middle distillates selectivity of the Y-type zeolite could be a bad balance between the hydrodehydrogenating functions and acidity [9].

Zeolites and zeolite-like materials in industrial catalysis

371

These drawbacks can be solved by using more highly dealuminated USY zeolites: by them it is possible to produce distillate selective hydrocracking catalysts that approach the desired selectivity of amorphous catalysts while enjoying substantially lengthened cycles and milder operating conditions (pressure and temperature) [15]. This has led to the development of several moderate pressure hydrocracking (MPHC) processes. These MPHC processes operate at temperatures in the 320450C range and pressures between 50-100 bar, saving investment and operating costs with respect to traditional HC [31]. The MPHC catalysts generally contain acidic cracking components such as zeolites or silica-alumina associated with a combination of metal sulphides of the VI (Mo, W) and VIII (Co, Ni) groups. MPHC processes take advantage of the inherent coke resistance of large pore, highly dealuminated Y-zeolites, in hydrogen or rare-earth forms. These units operate at pressure ranges that are consistent with the typical design pressures for existing hydrotreating units. Since the early 1980s refiners have added incremental conversion to their refineries by revamping existing VGO hydrotreating units into mild hydrocracking units The revamping usually involves just a catalyst substitution in the hydroprocessing unit along with some changes in the operating parameters like temperature, hydrogen to oil ratio etc. Among commercial processes available for MHC the use of zeolite based catalyst is claimed for MAKFining alliances Moderate pressure Hydrocracking [31] and ExxonMobils EMRE process [32].

Catalytic dewaxing
The aim of the process is to remove waxy components from various grades of lube oils and fuel components in order to meet desiderated cold properties (i.e. properties at low temperatures). Typical feedstocks are middle distillates and lube oils. Long-chain paraffinic (waxy) components, which precipitate out at low temperatures, are removed by cracking to form shorter-chain (non-waxy) molecules or by isomerization to branched paraffins obtaining fuels (diesel and jet fuel) of good stability and low pour point and lubricants of high-viscosity index. For jet fuels and diesel, besides octane and cetane number respectively, other properties are also required. In the case of jet fuels, high flash point to reduce fire hazard during re-fuelling, high heating value and low freezing point for high altitude are needed. In the case of diesel, the flow properties such as viscosity, pour point, freezing point, cloud point, and cold filter plugging point become critical especially in cold climates. Figure 8 shows the melting points of pure paraffins depending on the carbon number and branching. It appears therefore that in order to decrease the pour point of a diesel fraction the n-paraffins concentration should be reduced. In the early 1970s the catalytic dewaxing was introduced in refinery as alternative process to solvent dewaxing: the use of a zeolite-based catalyst permitted to remove the least branched paraffins by cracking [14]. The first zeolite-based bifunctional catalyst proposed was Pt/Hmordenite developed by BP [33], soon followed by ZSM-5 proposed by Mobil [34]. The exclusion of multiply branched paraffins in the selective catalytic dewaxing of waxy distillates and lube fractions over ZSM-5 is a typical example of RSS [6], as illustrated in Figure 9.

372

C. Perego & A. Carati

Today, the process is used worldwide by 27 licensees who have a combined dewaxing capacity in excess of 160,000 bpd. Further improvements in catalyst were realised: in the late 1980s Akzo-Fina added a desulfurization function to the ZSM-5 catalyst and improve the hydrotreating properties to reduce heteroatoms and to saturate olefins and aromatics [15]. In all cases, the improvements in product quality are obtained at the expense of end-product yield. More recently, in the 1990s, a new category of more selective catalysts gradually became established [14]: paraffin cracking is slowly being replaced by paraffin isomerization to reduce product pour point and improve flow properties. Isomerization improves product yield and product quality especially for lubricants where the isoparaffins are critical for achieving high-viscosity indices. The catalysts that permitted to move towards the isodewaxing process was the Chevrons SAPO-11 [35] that limits cracking favouring the isomerization of long straight paraffinic chains.

Figure 8. Melting point of n-paraffins (curve a), 2-, 3-, 5-methyl paraffins (curves b, c, d, respectively).

Figure 9. RSS of zeolite catalyst in dewaxing reaction.

Zeolites and zeolite-like materials in industrial catalysis

373

ExxonMobil announced a new fuels dewaxing process, commercially proven, in 1996. Mobil Isomerization Dewaxing, or more simply MIDW, lowers the pour point of waxy gas oils by hydroisomerization and selective hydrocracking of paraffins. Premium quality, low pour point distillates are obtained by retaining paraffins in the distillate boiling range. A family of proprietary, dual-functional noble metal-molecular sieve catalysts is used to isomerize. ExxonMobil has developed a pre-activated, distillate selective MIDW catalyst version, which simplifies co-loading with other catalyst types [36]. The type of zeolite has not been disclosed, however according to the patent literature [37,38] it should be a zeolite Beta, able to convert feedstock containing a significant portion of aromatic with high selectivity for isomerization over-cracking. Two isomerization steps should be preferred, the first with zeolite Beta and a second with a medium pore zeolite (ZSM-22, IZA code TON, structural characteristics reported in Table 2, or ZSM-23, IZA code MTT, medium pore zeolite, 1D, 0.45 x 0.52 nm) with a high degree of shape selectivity so that only linear (or almost linear) paraffins can enter in the internal structure of zeolite [38]. Application of MIDW after a moderate pressure hydrocracking (MPHC) process has been recently described by ExxonMobil process. MIDW converts paraffins and hydrocracks MPHC bottoms to selectively produce high quality kerosene and diesel. A commercial plant is in operation at Jurong refinery (Mobil Oil Singapore Ltd), where an existing diesel HDS unit has been revamped [32].

The process converts C6-C10 n-paraffins in aromatics and branched alkanes, obtaining high-octane gasoline blending components with increased Research Octane Number, RON and Motor Octane Number, MON. Combined processes for alkylation of benzene to more benign alkylaromatics or for saturate aromatics can be desired in response to tighter restrictions on aromatic contents in fuels. Hydrogen is a significant by-product of reforming units and it is separated from the reformate for recycling and use in other processes. Typical feedstocks of reforming units are low-octane naphthas. Reforming represents the total effect of numerous reactions such as cracking, polymerization, dehydrogenation, isomerization, cyclization taking place simultaneously. Depending on both the properties of the naphtha feedstock (as paraffin, olefin, naphthene, and aromatic content) and catalysts used, reformates can be produced with very high concentrations of toluene, benzene, xylene (BTX) and other aromatics useful in gasoline blending and petrochemical processing. These blending components could be less used in the future, due restriction on aromatics level in fuels. Processes optimized to produce BTX will be described after, as refinery/petrochemical interface processes. When Pt-Re or Pt-Re-Sn-Al2O3 catalysts are used, the main limitations are the low selectivity to dehydrocyclization of C6-C7 paraffins and the high amount of low octane normal paraffins [25]. The first zeolite introduced in reforming process was erionite (i.e. small pore zeolite with 8 member ring, IZA code ERI, 3D, 0.36 x 0.51 nm), that was placed in the last bed of the reactor. This post-reforming process was called Selectoforming. Only normal paraffins can react inside the small pore (RSS) and the octane number increase is due to their selective cracking producing LPG and to the increase in the concentration of iparaffins and aromatics.

Catalytic reforming and combined processes to upgrade and/or remove aromatics

374

C. Perego & A. Carati

Further improvement was represented by the M-Forming process [39], characterized by the use of the medium pore zeolite ZSM-5. The medium channel size of ZSM-5 catalyst can admit singly branched paraffins as well as simple aromatics such as benzene and toluene. The second lowest octane components, the single branched paraffins, are removed by hydrocracking. Furthermore, the aromatic species are alkylated by the olefinic component of the cracked product. The alkylaromatics contribute to octane number and reduce the loss of cracked products to gas, thus increasing the liquid yield [40]. A combination of M-Forming and reforming has been commercialized by ExxonMobil (BTXtra): the zeolite catalyst is placed near in the last bed of a semiregenerative reformer to improve product quality, increasing the concentration primarily of toluene and xylenes by transalkylation of the methyl groups from C9 aromatics with benzene. Some additional benzene, toluene, and xylenes are produced as a result of dealkylation. This process was first applied at Exxon-Mobils Chalmette, LA refinery in 1996 [15]. Another process, known as ExxonMobil Benzene Reduction (MBR), is a reformate alkylation process converting olefins (C2= and C3=) from FCC off gas and benzene-rich reformate stream to produce alkyl aromatics suitable as gasoline blendstock. RON values of alkylbenzenes is ~112 [41], while it is ~104 for benzene/toluene and ~93 for gasoline obtained by oligomerization of light olefins (Shells Polygasoline process, reported below in the oligomerization section). This fluid-bed process reduces the gasoline pool benzene below 1.0 vol% while increasing the pool octane by up to 1.0 number. The ZSM-5 catalyst permits to produce a higher octane mixture of C8C10 alkylaromatics. The catalyst also cracks some linear paraffins further enhancing gasoline octane. Approximately 6070% benzene reduction can be achieved [15]. In response to tighter restrictions on aromatics in diesel, particularly in Europe, Shell has developed a middle distillate hydrogenation process that permits to saturate heavier aromatics. It uses cation-exchanged Y-zeolite as a support for noble metal (Pt, Pd). The process takes advantage of the reduced sensitivity of ion exchanged Pt and Pd to poisoning by sulfur and nitrogen compounds. The Shell catalyst can tolerate feed sulfur levels of at least 500 ppm with little impact on hydrogenation activity. The zeolite has very low cracking activity, permitting to maximize yield through aromatic saturation [15].

Hydrodesulfurization of FCC products: innovative sulfur removal technologies


Over 90 % of the sulfur in gasoline comes from FCC naphtha, being the 80 % concentrated in the heavy FCC naphtha stream. The sulfur is present as thiophenic sulfur compounds, which are difficult to remove except by severe hydrotreatement [42]. Several processes have been developed and are now commercially proven: Prime G+ (IFP), SCANfining (ExxonMobil), CD Hydro / CD HDS (CD Tech), ISAL (UOP), OCTgain (ExxonMobil), S Zorb SRT (Phillips) [43]. Only some of these processes claim the use of zeolite-components. SCANfining process uses zeolite based catalysts, designed specifically for selective HDS of FCC naphtha. RT-225 catalyst minimizes olefin saturation and octane loss, while RT-229 catalyst breaks-up recombination of mercaptans [32]. OCTgain and ISAL1 processes associate conventional deep HDS to subsequent octane recovery through isomerization of the paraffins and alkylation essentially.

Zeolites and zeolite-like materials in industrial catalysis

375

Both use a fixed bed reactor with a first catalyst bed over which the organic sulfur compounds of the feed are converted into H2S, while hydrocarbons and the olefins are almost completely hydrogenated. The second bed containing a different catalyst leads to octane recovery through cracking, isomerization and alkylation of the paraffins, which compensates the significant olefin saturation. ISAL1 process uses a CoMoP/Al2O3 associated to a GaCr/HZSM-5 zeolite [44], while the OCTgain uses a not revealed shape selective zeolite [15]. One of the drawbacks of these processes can be some yield loss in gasoline due to cracking into light products and the hydrogen consumption needed for the quasi total olefin saturation. S Zorb Sulfur Removal Technology is a sorbent-based sulfur-removal method for extracting sulfur atoms from hydrocarbon product streams till to 10ppmw. Traditionally, sulfur removal has been accomplished through hydrotreating. Treating lighter portions of the Fluid Catalytic Cracker stream can, however, result in hydrogenation and resultant octane loss. S Zorb technology permits overcoming the conflicting goals of sulfur removal and olefin retention. S Zorb Sulfur Removal Technology uses a proprietary sorbent that attracts sulfur-containing molecules and removes the sulfur atom from the molecule. The sulfur atom is retained on the sorbent while the hydrocarbon portion of the molecule is released back into the process stream. H2S is not released into the product stream and therefore prevents recombination reactions of H2S and olefins to make mercaptans, which would otherwise increase the effluent sulfur concentration. An example of the basic sulfur removal pathway is shown in Figure 10 [45, 46]. The first commercial S Zorb unit began operations in April 2001 at ConocoPhillips Refinery in Borger, Texas USA. In November 2003, the second commercial S Zorb unit began operations at ConocoPhillips Refinery in Ferndale, Washington, USA. The benefit of using zeolite catalyst in conjunction with proprietary sorbent has been investigated in pilot plant test. Zeolite, due to RSS, can permit isomerization and cracking of linear alkanes, while larger branched molecules are excluded by its porosity. Tests with sorbent and 5 % of zeolite added show an average increase in RO N and MON of 0.8 and 0.5 respectively in comparison with tests with sorbent only [46].

Figure 10. Example of S Zorb chemistry.

Light paraffin hydroisomerization


The process permits the upgrading of the octane number of C5/C6 refinery streams by conversion of n-paraffins to branched isomers. High RON and MON products so obtained are suitable for addition to the gasoline pool. Typical feedstocks include light straight run naphtha, light reforming and light hydrocracking streams.

376

C. Perego & A. Carati

This application has been prompted by the world drive to remove the lead additives from motor gasoline in order to reduce air pollution. The octane loss caused by the removal of lead antiknock additives can be compensated for by isomerization of pentane/hexane paraffin fraction of the light gasoline fractions. In order to operate at low temperature (where thermodynamic is more favourable) and to obtain an acceptable yield of isomers, the catalyst systems used in the early units were based on superacidic aluminum chloride. This catalyst system, however, had the drawback of being highly corrosive and difficult to handle. From the 1960s and 1970s, modern isomerization units utilize dual- function solid catalysts consisting of a support having an acidic carrier and a hydrogenation function, frequently a noble metal, and operate in vapour phase and in presence of hydrogen. Zeolites can be used as acidic carrier; even if they have a much lower activity and have to be used at a higher temperature (230250 C approximately), they are more convenient to use with respect chlorinated alumina. In fact zeolite catalysts are characterized by their outstanding tolerance of feedstock poisons such as sulfur and water, the chlorinated catalysts suffer from extreme sensitivity to all kinds of feed contaminants [47,48]. The incorporation of a metal, such as Pt, not only increases the selectivity by carrying out the process in a true bi-functional way, but also prevents catalyst deactivation [25]. Zeolite-based catalysts are used in some versions of UOPs hydroisomerization processes; for example UOP I-7 catalyst is a Pt loaded modified synthetic (large-port) mordenite. In the presence of hydrogen at moderate conditions, such catalysts optimize isomerization and minimize hydrocracking [15]. A negative PSS effect can be evidenced because bulky isomers can hardly diffuse in the zeolitic pores. For example comparing omega zeolite and mordenite, both having 1-D large pore system, 2,2-dimethyl butane isomer is present in larger amount in omega probably because this bulky isomer diffuses more easily in its pores which are larger this respect to those of mordenite [9]. Nevertheless the advantage deriving by omega zeolite use is not sufficient to justify the development of a new catalyst. Other platinum promoted zeolite based catalysts are proposed for example by SdChemie. HYSOPAR catalyst, based on a not revealed zeolite structure, is tolerant to feedstock contaminants, for example can process feed of up to 200 ppm sulfur without any significant influence in product pool octane. It shows very low selectivity to light gases (C1-C4) which means higher isomerate yield. The catalyst can be regenerated (3-5 times) and shows long life (10 years on average) [49]. Recently Sd-Chemie has also introduced a new HYSOPAR-SA catalyst (SA stands for Super Acid)a sulphated zirconia catalyst distinguished by outstanding activity along with greatly improved tolerance towards water. Despite the advantages in terms of higher LHSV and lower operating temperature (about 60 C), the HYSOPAR-SA catalyst displayed noticeably higher iso-C5 and 2,2-DMB activities than the zeolite based HYSOPAR catalystbut at slightly lower poison tolerance levels. Therefore, this novel metal oxide catalyst is not a simple replacement for zeolitic isomerization catalysts but a complementary product [47].

Zeolites and zeolite-like materials in industrial catalysis

377

Olefin/paraffin alkylation
The products of olefin/paraffin alkylation are valuable clean fuels as blending component for gasoline pool. In fact the alkylate is characterized by no olefins or aromatic content, a low sulfur content, a limited heavy end, a low vapor pressure and high octane numbers. The main reaction is the alkylation of n-butene with iso-butane to produce iso-octanes (i.e., 2,2,3-, 2,2,4-, 2,3,3- and 2,3,4-trimethylpentane). As MTBE use further diminishes, alkylate will continue its growing importance as a means to compensate for lost octane-barrels and to recapture isobutylene back into the gasoline pool. The olefin/paraffin alkylation is a very difficult reaction and is thermodynamically promoted by low temperatures. Therefore highly acid catalysts are required. Today the liquid catalysts H2SO4 and HF, industrially used since 1940s, are still used in the alkylation units of refineries, in spite of the considerable research carried out in the 1990s to replace them by solid catalysts [48]. Recently a solid acid technology named Alkyclean has been developed and validate on a 10 bpd demonstration unit, in Porvoo, Finland in 2002 [50,51]. The process employs a zeolite-based catalyst so avoiding all the drawbacks of the existing hydrofluoric (HF) and sulfuric (H2SO4) acid based technologies. The AlkyClean process achieves superior performance by coupling its optimized solid acid catalyst (SAC) with an alkylation reactor system that efficiently minimizes the peak olefin concentration in the reaction zone. In the AlkyClean process, multiple reactors are used to allow continuous cyclic operation, alternating reactor between periods of alkylation and rejuvenation, following an inventive procedure established during the process development effort. The core of the process is the regeneration of the catalyst by rejuvenation before there is any substantial decrease in its activity, permitting to produce a high yield of high quality alkylate over a very long period of time [52]. The zeolite claimed in the patent is a Y structure, with a 0.5 % of platinum. Al2O3 can be present as matrix material and different particle size have been used (in the range 0.4-3 mm). The advantages of AlkyClean on H2SO4 and HF processes is summarised in Table 4, using modern sulphuric acid process as comparison base [51].
Table 4. AlkyClean process vs liquid acid technologies.

378

C. Perego & A. Carati

Light olefins, as propylene and mixed butenes (or both) contained in lower value refinery streams, are oligomerized into C6+ iso-olefin fractions that can be used as high octane blending stocks for gasoline pool and high-smoke-point blending stocks for kerosene and jet fuel [53]. Supported phosphoric acid (SPA) on silica was used since 1930s in the UOP Catpoly process oriented to gasoline production [14]. Since 1970s, different amorphous silicaalumina (e.g. Polynaphtha and Selectopol processes) or homogeneous organometallic complexes catalysts (Dimersol process) have been proposed to substitute SPA [53]. Also catalysts based on mesoporous silica alumina have been claimed for olefin oligomerization to gasoline and kero [54]. Several processes using zeolite have been reported. Shells Polygasoline and Kero SGPK process uses a Nimordenite catalyst. The process converts C2C5 olefins to gasoline, kerosene, and distillate at high pressures 15 MPa (1050 bar) and low temperatures (200280 C). The product is primarily isoolefins with very low aromatic content. The gasoline has a high octane (RON = 93) and is low in C5C7 olefins, while the distillate products have very low pour points and very good cetane number (cetane number >50). Distillate and gasoline selectivity can be varied by changing recycle composition [15]. The Mobil Oligomerization to Gasoline and Diesel (MOGD) process uses a modified ZSM-5 catalyst. ZSM-5 converts the oligomers into C5+ components by acid-catalyzed oligomerization, hydrogen transfer, aromatization, and isomerization reactions. In the distillate mode the process is operated at elevated pressure and low temperature to produce iso-olefins; while in the gasoline mode higher reaction temperature are required [55]. Very recently, a new ExxonMobil Oligomerization process (EMOGAS) has been developed, to convert olefins into motor gasoline by a simple fixed bed catalytic process [32,56]. The products are primarily octane isomers, the di-branched ones being preferred [56]. EMOGAS uses a new catalyst with a winning combination between zeolite and metal oxides that is more stable than any traditional solid acid catalyst in the presence of high olefin content and of high exothermic reactions. This catalyst does not deactivate as fast as traditional zeolites, a faster removal of coke precursors is facilitated. The first commercial scale trial started in 2004 at Giant refinery in Yorktown, Virginia, USA. Early results shows that the EMOGAS catalyst gives equivalent or improved performance than SPA catalyst in terms of feed processed, product qualities, pressure drop and ease of handling [56]. First grassroots unit start-up is planned for 2007 [41].

Olefins oligomerization

3. Refinery/petrochemical interface processes based on zeolitic catalysts


Processes which maximize olefins and aromatics provide opportunities both for modern refinery and petrochemical industry [15].

Methanol to olefins (MTO)


The UOP/HYDRO MTO [57] process converts methanol primarily into light olefins (ethylene and propylene) using a fluidized bed reactor with a continuous fluidized bed regenerator.

Zeolites and zeolite-like materials in industrial catalysis

379

The process can provide a broad range of propylene to ethylene product ratios. By simply changing the reactor operating severity, the UOP/HYDRO MTO process user can adjust between operation modes as market demands dictate. The reaction is catalyzed by the MTO-100 silicoaluminophosphate synthetic molecular sieve based catalyst, a SAPO-34 [57]. Long-term methanol conversion of 99.8% and stable product selectivity has been demonstrated at HYDROs large demonstration plant in Norway. This plant circulates and regenerates catalyst continuously and uses crude methanol as a feedstock at a rate of more than 0.75 MT per day. Lurgi Methanol to Propylene technology (MTP) [58], developed by Lurgi and ExxonMobil, is based on a fixed-bed reactor system, using a ZSM-5 catalyst. Methanol fed to the MTP plant is first converted to dimethyl ether (DME) and water in a DME prereactor. Using a highly active and selective catalyst, thermodynamic equilibrium is achieved, resulting in the methanol/water/DME mixture at appropriate operating conditions. The methanol/DME conversion rate exceeds 99%, with propylene as the essential compound. Lurgi is offering the MTP technology on full commercial scale terms [59].

Catalytic cracking of olefinic/paraffinic streams to propylene


Catalytic cracking of olefinic and paraffinic streams (suitable feedstocks include light cracked naphtha LCN, coker naphtha, steam cracker C4/C5, and light pyrolysis gasoline) to produce primarly propylene along with ethylene and butylenes has been described by several companies in recent years, including Mobil (MOI, PCC), Kellogg Brown & Root (SUPERFLEX) and Lurgi (PROPYLUR). These processes are characterized by the use of ZSM-5 catalysts to selectively convert higher molecular weight olefins and paraffins to lighter olefins. ZSM-5 pore size provides RSS by limiting access to the interior of the catalyst to mostly linear (non-branched or mono-methyl) paraffin and olefin molecules. In addition, the use of ZSM-5 zeolite limits the level of polyaromatics in the product. The resulting distribution of C3 and higher olefins approaches equilibrium, with propylene being the olefin product having highest yield. Ethylene is also produced and its yield is largely dependent on reaction conditions [60].

Aromatization of olefinic/paraffinic streams


C3-C8 olefins and paraffins streams can be converted to aromatics, mainly BTX. Light C3+ paraffin aromatization process, also called dehydrocycloligomerization (DHCO), involves a complex network of reactions. Specifically, DHCO includes: paraffin dehydrogenation and cracking; olefin oligomerization; olefin alkylation; oligomers cyclization into naphtenes ; naphtene dehydrogenation into aromatics. According to this mechanism, the most effective catalysts proposed for DHCO are generally bi-functional ones, consisting of a shape-selective acidic zeolite combined with a dehydrogenating metal or oxide. Several processes have been developed, namely Cyclar, Aromax, RZ Platforming, GTC, Asahi Chemicals' Alpha. The Cyclar process is based on Ga/H-ZSM-5 catalyst. This process jointly developed by BP and UOP converts a saturated feedstock (LPG) into BTX aromatics, hydrogen and light hydrocarbons. The process is carried out in a number of vertically-stacked moving-

380

C. Perego & A. Carati

bed radial-flow reactors. Catalyst is withdrawn from the bottom reactor, regenerated and supplied to the top reactor without interrupting the process flow. The shape-selectivity of the ZSM-5 helps promote the cyclization reactions and limits the size of the rings [61]. Chevron Phillips Chemicals Aromax technology [62] is the premier on-purpose benzene production technology. Based on a L-zeolite catalyst, it selectively converts C6C8 hydrocarbons and naphthenes to benzene, toluene and hydrogen utilizing conventional fixed-bed reforming equipment. The larger application is a world scale Aromax technology unit start-up in 1999 in Saudi Arabia; while a second generation Aromax catalyst was developed, having a superior performance with heavier feed-stocks. The UOPs RZ Platforming process is a fixed-bed system that converts C6 and C7 paraffins to aromatics, with high yields of benzene, toluene and hydrogen. The RZ Platforming process uses a not described RZ-100TM catalyst, that respect to conventional reforming catalyst shows a significantly higher selectivity for benzene and toluene. The GTC technologys BTX plus process uses a proprietary ZSM-5 catalyst for olefin dimerization and dehydrocyclization to aromatics in a swing fixed bed reactor. Commercial plant is in operation since 1993 [63]. Presently, the only process tailored for the production of BTX aromatics from olefinrich feeds, is the Alpha process developed by Sanyo Petrochemical Co. Ltd, a subsidiary of Asahi Chemical. It is based on a Zn/-Al2O3/H-ZSM-5 catalyst optimised by using a proprietary hydrothermal treatment [64].

4. Petrochemical processes based on zeolitic catalysts


Xylene isomerization
The xylene isomerization process converts the much less used meta- and ortoxylenes to para-xylene, to meet the market for para-xylene demand [65]. Xylene isomerization feed-stocks usually contain more than 10% ethylbenzene (EB), that has to be removed to avoid its accumulation in the recycle streams. Xylene isomerization is a thermodynamic equilibrium controlled reaction. Separation of para-xylene from mixed C8 aromatics is essential and can be achieved either by crystallization and adsorption processes. Adsorption technologies prevalently utilize modified faujasite zeolites (e.g. Y and X zeolites) as selective adsorbent. Well-known technologies are Isomar (UOP), Aromax (Toray) and Eluxyl (IFP). Three are the main requirements for xylene isomerization process: a) the suppression of xylene loss by disproportionation to toluene and trimethylbenzenes; b) the enhancement of EB conversion; c) the increase in para-xylene selectivity. Introduction of zeolite in this industrial process started in the seventies when mordenite replaced the amorphous acid support in the isomerization of xylenes [14]. In 1975 Mobil introduced the Vapor Phase Isomerization (MVPI) process, based on a ZSM-5 catalyst [66-68]. Shape selective ZSM-5 shows higher selectivity and lower coking than the large pore zeolites. There is a strong increase in the ratio between the rates of isomerization and disproportionation when zeolite pores are smaller. This ratio increases from 20, to 70 up to 1000 for faujasite, mordenite and ZSM-5 respectively [69]. Trialkylbenzenes formation

Zeolites and zeolite-like materials in industrial catalysis

381

and coking are inhibited in ZSM-5 pore system because there is not enough space to accommodate the large, bimolecular transition states of these reactions (TSS effect) [70]. In ZSM-5, at 200C the kinetic reaction controls the rate of isomerization, while over 300C the rate is diffusion controlled. The rate of diffusion of para-xylene is higher than that of ortho-xylene; meta-xylene has the lowest diffusion rate. After MVPI process, Mobil has introduced several technology progressions, relating both to process conditions and catalytic material, in order to increase EB conversion with low xylene losses. The most recent processes developed are: Advanced Mobil High Activity Isomerization (AMHAI) and XyMax, introduced in 1999 and 2000 respectively. AMHAI can operate with low para-xylene concentration in isomerization feeds, offering reduced xylenes losses, a wide range of EB conversion and operating at lower temperatures than its predecessors. The XyMax process works in a fixed bed reactor with two distinct zone catalysts. In the top bed the catalyst is designed to convert EB to benzene and ethylene (then hydrogenated to ethane), and to crack non-aromatics. In the bottom bed the catalyst is optimized for isomerization of the para-xylene depleted feedstock to an equilibrium mixture of xylenes. In 2002 there were 21 units worldwide using ExxonMobil xylene isomerization technologies. These units represent over 30 % of the worlds para-xylene production capacity from xylene isomerization [67]. Other xylene isomerization processes have been developed by Indian Petrochemical Corporation Limited (IPCL), in association with National Chemical Laboratory (NCL, Pune). In the mid-1980s a composite catalyst containing high-silica ZSM-5-type metallosilicate, called Encilite, was developed with unique catalytic activity for mxylene isomerization. This led to the development of Indias first petrochemicals catalyst and process (Xylofining) for the isomerization of m-xylene to o-and para-xylene with conversion of EB to benzene. Subsequently, based on extensive fundamental studies, like changes in the composition of the high-silica zeolite, support modification and impregnation of active metal ingredients, IPCL developed an improved new generation catalyst of its own, known as Encilite-501. The performance of the newly developed Encilite-501 catalyst, a conventional C8 aromatic isomerization catalyst and the Encilite catalyst is presented in Table 5.
Table 5. Xylene isomerization processes: typical process parameters [34].

382

C. Perego & A. Carati

The zeolite-based Encilite-501 has shown excellent stability and life for more than 10 years in commercial operation for the production of para- and o-xylene [71]. UOP Isomar process can be operated with different catalysts, that are classified according to how they treat the EB present in the mixed xylenes feed. I-400 catalyst is an EB isomerization catalyst that converts EB to additional xylenes: ethyl group from the ethylbenzene is converted into methyl groups which are then rearranged to form xylenes. I-300 is an EB dealkylation catalyst that converts EB to ethane and benzene, Both I300 and I-400 catalysts are platinum loaded zeolite base [72].

Toluene disproportionation-transalkylation
The first generation of toluene disproportionation (TDP) process was commercialized in 1975 by Mobil (MTDP process) in Naples, Italy plant [73]. During the late 1980s Mobil developed a selective toluene disproportionation process (MSTDP), converting two moles of toluene to one mole each of xylenes and benzene, that uses shape-selective ZSM-5 catalyst to maximize para-xylene in the produced xylene stream. Application of ZSM-5 in toluene disproportionation process is an excellent example of the application of shape selectivity to vastly exceed equilibrium yields of preferred products by playing off rates of diffusion vs. rates of reaction [6]. Shape selectivity inhibits the secondary isomerization of the primary para-xylene product. As the diffusivity of para-xylene in ZSM-5 is three orders of magnitude larger than for orto- and meta-xylene, by a proper optimization of zeolitic catalyst it is possible to minimize the isomerization so to produce para-xylene in excess of equilibrium. Such optimization includes the increase of the crystal radius, the incorporation of different element and the deposition of coke on the exterior of the MFI molecular sieve, so increasing the diffusion path length and poisoning all of the non-selective external sites without affecting the number of sites internal to the crystal [74]. Later improvements of MSTDP process led to the new PxMax, the state-of-the-art STDP technology by ExxonMobil Chemical, put into commercial operation in 1996. Compared to MSTDP, the PxMax process offers higher purity para-xylene-rich xylenes (90%+), higher total xylenes yield and superior xylenes/benzene ratio [66]. PxMax catalyst has also been modified in order to reduce by-products formation, especially EB. This is done by incorporating a hydrogenating/dehydrogenating function in the catalyst, i.e. platinum. Patent examples show that EB can be reduced by a factor of 3-4 and C9+ aromatics by a factor of 3 adding 0.025 % platinum on a 10 % SiO2/HZSM-5 catalyst formulation [75]. To treat up a high level of C9+ aromatics, with either toluene or benzene (fresh feed composition can be varied from 100% toluene to 100% C9+ aromatics, heavy aromatics with up to 25 % C10 aromatics) a transalkylation process, TransPlus [66,76] has been codeveloped by ExxonMobil and the Chinese Petroleum Corporation of Taiwan and commercialized in 1997 at CPC's Lin-Yuan Petrochemical Plant in Kaohsiung, Taiwan [66]. This technology offers to refiners a low cost solution for upgrading heavy aromatics streams, alone or in conjunction with either toluene or benzene co-feed, to higher value mixed xylenes with either high purity benzene or toluene as a co-product. The TransPlus process is typical for a vapour phase fixed bed reactor system. Toluene or benzene feed, along with C9+ aromatics, are combined with a hydrogen-rich recycle gas, heated and passed through a catalyst bed. The process utilizes an ExxonMobil proprietary zeolite

Zeolites and zeolite-like materials in industrial catalysis

383

based catalyst that has excellent yield performance while exhibiting a high tolerance towards heavy aromatic species which can potentially shorten catalyst life. Mechanistically, the TransPlus process works by first dealkylating ethyl and higher alkyl groups leaving primarily methyl aromatics behind. The remaining feed components are then converted via disproportionation and transalkylation to produce a near-equilibrium product. Similarly UOP developed Tatoray, PX-Plus and PX-Plus XP processes [72]. Tatoray is a disproportionation and transalkylation process, to selectively convert toluene, C9 and C10 aromatics into more valuable benzene and xylenes. Tatoray process can handle up to 100 % C9 aromatics and has operated commercially with up to 10% C10 aromatics in the feed. PX-Plus XP process, licensed in 1999, is an integration of the previous PX-Plus process (1993) and is more selective to para-xylene compared to Tatoray process. Details about formulation and selectivation procedure for the catalysts are not disclosed. A new catalyst (ATA-12) [77], that utilizes a zeolite as the active material, has been developed in the SK Corp and Zeolyst laboratories for conversion of toluene and C9+ aromatics. ATA catalyst was commercialized in 1999 and today there are a total of 9 units in operations. One of the major benefits of the ATA-12 catalyst is its feed flexibility. The ATA-12 catalyst can process feed compositions ranging from benzene or 100% toluene to 100% C9+ aromatics.

Toluene alkylation with methanol


GTC Technology Corporation has developed the GT-TolAlk process technology, in which toluene reacts with low-cost methanol across a selective zeolite catalyst to produce para-rich xylenes plus water [78,79]. These xylenes are easily purified to chemical-grade para-xylene in a simple crystallization unit. There is virtually no benzene by-product from this reaction, so that the para-xylene yield from the toluene is very high. Another process has been developed till pilot plant scale, using a ZSM-5 after silanization with tetraethyl orthosilicate (TEOS) for pore-size regulation. Para-xylene selectivity above 85 % with toluene conversion up to 28 % has been obtained, with catalyst stability for more than 500 h on stream [71].

Alkylbenzenes by alkylation-transalkylation reaction


Ethylbenzene (EB), cumene, para-diethylbenzene (p-DEB), paradiisopropylbenzene (p-DIPB), C10-C14 linear alkylbenzenes (LAB), para-ethyltoluene and cymene are some of the chemical intermediates obtained by acid catalyzed alkylation on the aromatic ring of benzene or toluene. Figure 11 summarizes several aromatic alkylations industrially applied for the preparation of important chemical intermediates. These reactions include the most important aromatic substrates, benzene, toluene and xylene, and different olefins. Besides they include two different kind of alkylation: the electrophilic alkylation on the aromatic ring catalyzed by acids and the side-chain alkylation catalyzed by bases. Different types of acids are used as catalysts for the alkylation of aromatic hydrocarbons: a) metal halides, such as aluminium and gallium chloride and borium fluoride; b) mixed oxides and zeolites; c) protonic acids such as sulphuric acid, hydrofluoric acid and phosphoric acid; and d) sulfonic resins. The most active ones are the Brnsted acids, which contain an acidic proton. Zeolites are of high interest for the alkylation of aromatics, permitting green processes.

384

C. Perego & A. Carati

The effort devoted to introduce zeolite catalyst for benzene alkylation with ethylene led to the first industrial demonstration in 1976 by Mobil. Since that several improvements have been introduced as summarized in the Table 6. For the production of cumene many more years were necessary to arrive at a significant result, than with EB. This is due mainly to the fact that the zeolite ZSM-5, the catalyst for Mobils EB process, demonstrated a major limitation in the cumene reaction, represented by the elevated level of co-production of n-propyl benzene. On the other hand, ZSM-5, being of medium pore size, is not sufficiently active in the liquid phase [89]. Also for cumene, a noticeable improvement was thus obtained by operating in the liquid phase with large pore zeolites. Mordenite structure has been applied by Dow-Kellog in the process, named 3-DDM Cumene Process. 3-DDM stands for three dimensional dealuminated mordenite. In fact, considering only the 12-ring channels, mordenite is a one-dimensional zeolite. Dealumination converts the material into a novel catalyst by introducing some mesoporosity, resulting in a pseudo three-dimensional structure, which provides optimal performance and stability. This 3-DDM catalyst has been industrially applied by DowKellog as transalkylation catalyst in the cumene plant of Terneuzen since 1992. The other zeolites used in the current processes for cumene production are similar to that used for EB: i.e. zeolite Beta (EniChem, now Polimeri Europa, and UOP) and MCM22 (Mobil-Raytheon). EniChem, Mobil-Raytheon and UOP, independently started-up industrial zeolitebased cumene plants in 1996. Cumene yields are above 99.5 % in all zeolite-based processes. Product purity is as high as 99.95 % [90].

Figure 11. Industrial aromatic alkylation reactions (adapted from C. Perego et al. [80]).

Zeolites and zeolite-like materials in industrial catalysis

385

According to Chemical Week, in 2004 about 50 % of cumene production uses zeolite catalysts, as producers have switched from older phosphoric acid-based processes [91]. Zeolites find industrial applications also in synthesis of para-ethyltoluene and pDEB. The main goal in all these reaction is the para-selectivity, i.e. the capability of the catalytic system to selectively favour the formation of para-isomer. Mobil developed an ethyltoluene process based on the selective alkylation of toluene with ethylene, which was industrially applied in Baton Rouge, Lousiana in 1982 using a modified H-ZSM-5 with phosphorus and by impregnation with Mn, Mg, and B salts[92,93]. A dramatic increase in the para-selectivity up to 98 % was observed and ascribed mainly to spatial restrictions in the narrowed zeolite pores rather than differences in the zeolite acidity. A process for p-DEB production by EB disproportionation, was commercialized by Taiwan Styrene Monomer Company (TSMC) in 1988. The commercial plant that produced p-DEB with 96% purity went on stream in 1990 with a capacity of ca. 4000 tons/year.
Table 6. Industrial benzene alkylation with ethylene processes.

386

C. Perego & A. Carati

Recently, TSMC further upgraded their p-DEB production purity to 99%. The TSMC process has a cycle length of over six months and its catalyst is fully regenerable [16]. The silanization of ZSM-5 has also been applied in the p-DEB production plant operated by Indian Petrochemical Company since 1994 with a capacity of 1200 tons/year. This process is based on the selective alkylation of EB with ethylene [71]. For this reaction high para-selectivity is also reported with ZSM-12 [94].

-Caprolactam
Caprolactam, which is employed in the production of Nylon-6, may be derived from cyclohexanone through the ammoximation and the acid-catalyzed Beckmann rearrangement reactions. Use of concentrated sulfuric acid (e.g. oleum) and production of high amount of ammonium sulfate as a by-product are important drawbacks of conventional technologies. A new integrated process (Figure 12) is now available that combine EniChem and Sumitomo Chemical developed processes for the ammoximation and the vapour-phase Beckmann rearrangement steps [95]. Titanium silicalite (TS-1) in the presence of ammonia and hydrogen peroxide is the catalyst of ammoximation reaction [96]. An industrial unit (12000 t/y) was operated since 1994 in Porto Marghera, Italy. The reaction is performed in a continuous stirred tank reactor (CSTR) in t-butanol as a solvent. Both conversion and selectivity of cyclohexanone to oxime is over 99% and the yield based on hydrogen peroxide is over 90% [97]. Silicalite-1 (S-1) is used as catalyst for vapour phase Beckmann rearrangement process developed by Sumitomo Chemical Co., Ltd [98]. Sumitomo Chemical Co., Ltd. has industrialized the combined process of vapour phase Beckmann rearrangement and ammoximation in 2003 [99].

Figure 12. New -Caprolactam synthesis steps.

5. Conclusions
Catalyst industry is closely tied into current processing trends and market forces characterizing the evolution of the chemical industry. In particular, refining and petrochemical industry are continuously growing, under the push of the global demand of oil products. Refining and petrochemistry can be considered, by certain aspects, as mature industries, but this is only an appearance for these industries that are really constantly changing. They continue to change to adapt to numerous constraints, which are not likely to decrease in the foreseeable future, and which are due to the need to limit

Zeolites and zeolite-like materials in industrial catalysis

387

consumption of petroleum products (in particular CO2-related issues), to have greater respect for the environment and minimize refining costs. In this context, zeolites have an important role to play. The contribution of zeolitic catalysts to refining and petrochemistry is already substantial. Further achievements in the future through improvements to existing catalysts and the development of new catalysts, can be forecasted. The development of new catalysts for new or improved processes is a pivotal target for catalyst industry, but not lower importance has to be attributed to optimization of catalyst formulation by stretching the performance of primary catalytic-conversion units. Industrial zeolite-based catalysts are complex materials comprehending several building blocks: Zeolite is the heart of the catalyst working as active phase for acidic (or basic or redox, depending by its composition) catalyzed reactions and affecting the selectivity of reactions through the shape selectivity. Activity and selectivity of a zeolite are affected not only by its crystalline structure but also by its compositional and morphological properties. Metal oxide can be present and its role, especially in modern refinery industry, is increasing due to the need of metals and sulfur tolerance, control in coke formation, hydrotreating contribute required to treat low quality oils. Matrix is fundamental to shape the catalyst, besides its acidity and porosity has to be calibrated to the requirements of the feed to be processed. Additives can permit to have specialized catalysts, required to process a wider variety of feed-stocks through processing facilities. Optimized catalysts will result by a winning combination of all their components. A multidisciplinary approach is also essential: a close cooperation among catalyst suppliers, reactor designers, process control experts, driven by a global overview of refinery and petrochemical systems, with attention on synergies among different processes is required.

Bibliografy
1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. R.G. Gonzales, ptq catalysis, 9 (2) (2004)13. R.G. Gonzales, ptq catalysis, 10 (2) (2005) 3. Catalyst Report, published by Oil &Gas Journal, Oct. 6, 2003. J.D. Sherman, Proc. Natl. Acad. Sci. USA, Vol. 96, p. 3471, March 1999 Colloquium Paper. D.W. Breck, Zeolite Molecular Sieves, J. Wiley & Sons, Inc. (1974). T.F. Jr Degnan, J. Catal., 216 (2003) 32. P.B. Weisz, V.J. Frilette, J. Phys. Chem., 64(1960) 382. S.M. Csicsery, in: Zeolite Chemistry and Catalysis, ACS Monograph edited by J.A. Rabo, Vol. 171, Am. Chem. Society, Washington, DC, (1976) p. 680. C. R. Marcilly, Top. Catal., 13 (2000) 357. J. ejka, B. Wichterlova, Catal. Rev.-Sci.Eng., 44 (2002) 375. Chemical Economics Handbook Marketing Research Report , August 2002. R.G. Gonzalez, ptq catalysis, 10 (2) (2005) 5. Hydrocarbon processing, February 2006, 41. C. Marcilly, J. Catal., 216 (2003) 47. T.F. Degnan, Top. Catal., 13 (2000) 349.

388

C. Perego & A. Carati

16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57.

Tsai T.C., Liu S.B., Wang I., Appl. Catal. A, 181 (1999) 355. Bellussi G., Stud. Surf. Sci. Catal., 154A (2004) 53. M.G. Clerici, Top. Catal., 13 (2000) 373. A. Corma and H. Garcia, Catal. Today, 38 (1997) 257. R.A. Sheldon and R.S. Downing, Appl. Catal. A, 189 (1999) 163. Roberts, S.M., Catalysts for Fine Chemical Synthesis, Volume 4. Microporous and Mesoporous Solid Catalysts Catalysts For Fine Chemicals Synthesis, John Wiley & Sons (2006). www.iza.-structure.org M. W. Schoonover, M.J. Cohn et al., Top. Catal., 13 (2000) 367. J.A. Moulijn and M. Makkee, Hydrocarbon processes in the oil refinery, TUDelft, (2003). A. Corma, NATO ASI series C, Vol.352, edited by E.G. Derouane et al.,(1992) 373. A. Corma, V. Gonzlez-Alfaro, A.V. Orchills, Appl. Catal. A, 187 (1999) 245. J.R.D. Nee, ptq catalysis, 9 (2) (2004) 6. S. Ismail, ptq Q2 (2005) 129. N. Petti, L. Hunt, G. Yaluris, ptq catalysis, 11 (4) (2006) 43. M. Xu, R.J. Madon, ptq Q2 (2005) 63. G. Valavarasu, M. Bhaskar, and K. S. Balaraman, Petroleum Science And Technology, 21 (78) (2003) 1185. A. Sapre, J. Poturovic, International Downstream Technology Conference & Exhibition (IDTC) (2007). P.A. Lawrance, R.W. Aitken, R.J.K. Harris, Gb1, 134,015 assigned to British Petroleum Co (1968). N.Y. Chen, S.J. Lucki; W.E. Garwood, US 3,700,585 assigned to Mobil Oil Corp (1972). Miller, S. J., Stud. Surf. Sci. Catal., 84C (1994) 2319. W.J. Tracy, G.K. Chitnis, W.J. Novak, T.E. Helton, A. Macris, D.A. Pappal, U. Nagel, 3rd European Catalyst Technology Conference Amsterdam, The Netherlands (2002). R.B. Lapierre, R.D. Partridge, N.Y. Chen, S.S. Wong, US 4,419,220 assigned to Mobil (1983). N.Y. Chen, W.E. Garwood, T.J. Huang, Q.N. Le, S.S. Wong, US 4,919,788 assigned to Mobil (1988). J.C. Bonacci, J.R. Patterson, U.S. 4, 292,167 assigned to Mobil (1981). A. Corma, Catal. Lett., 22 (1-2) (1993) 33. R. D. McGihon, ARTC Annual Refining and Petrochemicals Meeting (2006). Hydrocarbon processing, February 2001, 33. www.kfupm.edu.sa S. Brunet, D. Mey, G. Perot, C. Bouchy, F. Diehl, Appl. Catal. A, 278 (2005) 143. www. conocophillips.com J. Vander laan, E.L. Sughrue, G. Dodwell, P.F. Meier, Hydrocarbon processing, February 2006, 49. H. Weyda, E. Khler, Catal. Today, 81 (2003) 51. J. Weitkamp, Y. Traa, Catal. Today, 49 (1999) 193. www.sud-chemie.com www.albemarle.com V. DAmico, J. Gieseman, E. Van Broekhoven, E. Van Rooijen, H. Nousiainen, Hydrocarbon Processing, February 2006, 65. E. Van Broekhoven, F. Rene Mas Cabre, P. Bogaard, G. Klave, M. Vonhof, US 5,986,158 assigned to Akzo (1997). Hydrocarbon processing - Refining Processes 2004 S. Peratello, M. Molinari, G. Bellussi, C. Perego, Catal. Today, 52 (1999) 271. N.Y. Chen, W.E. Garwood, Catal. Rev. Sci. Eng., 28 (1986) 185. www.exxonmobil.com J.Q. Chen, A. Bozzano, B. Glover, T. Fuglerud, S. Kvisle, Catal. Today, 106 (2005) 103.

Zeolites and zeolite-like materials in industrial catalysis

389

58. www.lurgi.de 59. H. Koempel, W. Liebner, M. Wagner, Gastech 2005, Bilbao, Spain (2005). 60. M.W. Bedell, P.A. Ruziska, T.R. Steffens, On-Purpose Propylene From Olefinic Streams, Tulane Engineering Forum (2003). 61. P.M.M. Blauwoff, J. Gosselink, E.P. Kieffer, S.T. Sie, W.K.J. Stork, edited by J.Weitkamp, L. Puppe, Catalysis and Zeolites: Fundamental and Applications, Springer Verlag, Berlin, Heidelberg (1999) p.512. 62. www.cpchem.com 63. J.C. Gentry, 8th International Downstream Technology Conference & Exhibition (IDTC) (2007). 64. N. Nagamori, M. Kawase, Micropor. Mesopor. Mater., 21 (1998) 439. 65. H.G. Franck, J.W. Stadelhofer, Industrial Aromatic Chemistry, Springer-Verlag, Berlin Heidelberg (1988). 66. www.exxonmobilchemical.com 67. G. Mohr, AIChE 2002 Spring National Meeting New Orleans, Louisiana, U.S.A. (2002). 68. G. Burress, US Patent 3,856,873 assigned to Mobil (1974). 69. H. Heinemann, Catal. Reviews Sci. Eng. 23 (1981). 70. S.M. Csicsery, Pure Appl. Chem., 58 (1986) 841. 71. J. Das, A.B. Halgeri , Catal. Surveys from Asia, 7 (1) (2003) 3. 72. www.uop.com 73. K.W. Smith, W.C. Starr, N.Y. Chen, Oil Gas J., 78 (1980) 75. 74. D.H. Olson, W.O. Haag, ACS Symposium Series, 248 (1984) 275. 75. G.D. Mohr, J.P. Verduijn, US Patent 6,039,864 assigned to Exxon (2000). 76. Xylenes, ChemSystems PERP Report01/02-7 (June 2002). 77. ww.zeolyst.com 78. European Chemical News / April 10-16, 2000. 79. www.gtchouston.com 80. C. Perego, P. Ingallina, Green Chem., 6 (2004) 274. 81. S.H. Wang, Styrene, PEP Report 33C, Supplement C, Process Economics Program, Menlo Park, California, SRI International (1993). 82. P. Ratnasamy, R. Kumar, Catal. Today, 9 (1991) 329. 83. C. G. Wight, US Patent 4,169,111 assigned to Union Oil Company of California (1979). 84. F. Narsolis, G. Woodle, G. Gajda, D. Gandhi, Petroleum Technology Quarterly, Summer (1997) 77. 85. J.C. Cheng, T.F. Degnan, J.S. Beck, Y.Y. Huang, M. Kalyanaraman, J.A. Kowalasky, A: Loehr, D.N. Mazzone, Stud. Surf. Sci. Catal., 121 (1999) 53. 86. J.W.J. Roth, B.J Ratigan, D.N. Mazzone, US 6,984,764 assigned to ExxonMobil Oil Corp (2006). 87. G. Girotti, F. Rivetti, F. Assandri, Hydrocarbon Engineering, 9 (11) (2004) 69. 88. E. Bencini, G. Girotti, WO 2004/056475 assigned to Polimeri Europa (2004). 89. G. Bellussi, G. Pazzuconi, C. Perego, G. Girotti, G. Terzoni, J. Catal., 157 (1995) 227. 90. C. Perego, P. Ingallina, Catal. Today, 73 (2002) 3. 91. Chemical Week, April 7/14 (2004) 41. 92. W.W. Kaeding, L.B. Young, C. Chu, J. Catal., 89 (1984) 267. 93. W.W. Kaeding, L.B. Young, A.G. Prapas, Chemtech, 12 (1982) 556. 94. W. W. Kaeding, J. Catal., 120 (1989) 409 95. G. Bellussi, C. Perego, Cattech, 4 (2001) 4. 96. P. Roffia, G. Leofanti,.A. Cesana, M. Mantegazza, M. Padovan, G. Petrini, S. Tonti, P. Gervasutti, Chim. Ind. (Milan), 72 (1990) 598. 97. G. Petrini, G. Leofanti, M.A. Mantegazza, F. Pignataro, ACS Symposium Series, 626 (1996) 33. 98. H. Ichihashi, M. Kitamura, Catal. Today, 73 (2002) 23. 99. www.sumitomo-chem.co.jp

Vous aimerez peut-être aussi