Anda di halaman 1dari 16

Ind. Eng. Chem. Res.

1996, 35, 2075-2090

2075

REVIEWS Advances in Mesoporous Molecular Sieve MCM-41


Xiu S. Zhao, G. Q. (Max) Lu,*, and Graeme J. Millar
Department of Chemical Engineering and Department of Chemistry, The University of Queensland, Brisbane, Queensland 4072, Australia

The discovery of mesoporous molecular sieves, MCM-41, which possesses a regular hexagonal array of uniform pore openings, aroused a worldwide resurgence in this field. This is not only because it has brought about a series of novel mesoporous materials with various compositions which may find applications in catalysis, adsorption, and guest-host chemistry, but also it has opened a new avenue for creating zeotype materials. This paper presents a comprehensive overview of recent advances in the field of MCM-41. Beginning with the chemistry of surfactant/ silicate solutions, progresses made in design and synthesis, characterization, and physicochemical property evaluation of MCM-41 are enumerated. Proposed formation mechanisms are presented, discussed, and identified. Potential applications are reviewed and projected. More than 100 references are cited. Contents
1. Introduction 2. General Background 3. Chemistry of Surfactant/Silicate Aqueous Solution 3.1. Behavior of Surfactant Molecules in an Aqueous Solution 3.2. Chemistry of Silicates/ Aluminosilicates in an Alkaline Aqueous Solution 4. Synthesis of MCM-41 4.1. Expanding Synthesis Conditions 4.2. Synthesis of Aluminum-Rich MCM-41 4.3. Synthesis of Hybrid Atom MCM-41 5. Formation Mechanism 5.1. Liquid Crystal Templating (LCT) Mechanism 5.2. Transformation Mechanism from Lamellar to Hexagonal Phase 5.3. Folded Sheets Mechanism 5.4. Identifying the Formation Mechanism 6. Characterization, Physicochemical Properties, and Structure Model 6.1. Characterization and Structure Model 6.2. Acidity 6.3. Stability 6.4. Interaction with Water 7. Application 7.1. Catalysis 7.2. Ship in a Bottle 7.3. Model Adsorbent 2075 2076 2077 2077 8. Perspective 8.1. Host -Guest Encapsulation 8.2. Modification 8.3. Adsorbent 8.4. Environment 9. Acknowledgment 10. Literature Cited 2086 2086 2087 2087 2087 2087 2087

2077 1. Introduction 2078 2078 2079 2080 2080 2080 2081 Over the past 15 years, there has been a dramatic increase in the literature of design, synthesis, characterization and property evaluation of zeolites and molecular sieves for catalysis, adsorption and separation, environmental pollution control, and intrazeolite fabricating technology. Progress has been made in a variety of disciplines, including inorganic and materials chemistry (Davis and Lobo, 1992; Ozin, 1992), mineralogy and crystallography (Smith, 1988), petrochemistry (Chen et al., 1989; Corma, 1995a), environmental science (Armor, 1992; Iwamoto, 1994; Tabata, 1994), and even biochemistry (Mann, 1993). As excellent examples, bicontinuous phases have been used to orient organic species which are then photopolymerized (Davis, 1993). After the extraction of the unpolymerized components, porous organic monoliths resulted. This process has led to a new concept of self-assembled microstructure serving as structure-directing agents. Again, material scientists have used large organic molecules to intercalate into layered clays to generate novel materials with controllable pore size (Yanagisawa et al., 1990). Biochemists have established a model that makes use of the cooperative organization of inorganic and organic
* Author to whom all correspondence should be addressed at the Department of Chemical Engineering, The University of Queensland, Brisbane, Queensland 4072, Australia. Phone: 61 7 33653708. Fax: 61 7 33654199. E-mail: maxlu@ cheque.uq.edu.au. Department of Chemical Engineering. Department of Chemistry.

2083 2083 2083 2083 2084 2084 2085 2085 2085 2086 2086

S0888-5885(95)00702-0 CCC: $12.00

1996 American Chemical Society

2076

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996

Table 1. Pore Size Definition of Zeolites and Molecular Sieves pore size () >500 20-500 <20 definition macroporous mesoporous microporous ultralarge pore typical material MCM-41 cloverite JDF-20 VPI-5 AlPO4-8 faujasite AlPO4-5 ZSM-12 ZSM-48 ZSM-5 CaA SAPO-34 20 20 18 14 12 12 12 10 10 8 8 ring size pore diameter () 15-100 6.0 13.2 6.2 14.5 12.1 7.9 8.7 7.4 7.3 5.5 5.9 5.3 5.6 5.3 5.6 5.1 5.5 4.2 4.3 reference Beck et al., 1992a Estermann et al., 1991 Jones et al., 1993 Davis et al, 1988 Dessau et al., 1990 Olson, 1970 Bialek et al., 1991 Fyte et al., 1990 Schlenker et al., 1978 van Koningsveid et al., 1990 Meier and Olson, 1987 Lok et al., 1987

large pore medium pore small pore

molecular species into three dimensionally structured arrays for the synthesis of nanocomposite materials (Mann, 1993). Molecular sieve scientists have taken advantage of this concept to generate novel inorganic materials, which resulted in the discovery of the new family of mesoporous molecular sieves M41S (Beck et al., 1992; Kresge et al., 1992). MCM-41, one member of the M41S family, possesses a regular hexagonal array of uniform pore openings with a broad spectrum of pore diameters between 15 and 100 (Beck et al., 1992a; Kresge et al., 1992). The mesoporous structure can be controlled by a sophisticated choice of templates (surfactants), adding auxiliary organic chemicals (e.g., mesitylene) and changing reaction parameters (e.g., temperature, compositions). Hence, numerous studies of this novel materials have been published since its discovery. However, there lacks a multiple review on MCM-41 except the brief report (Casci, 1994) which was aimed at summarizing the work on ultra large pore molecular sieves and related zeotype materials (pillared layered clays). This paper is intended to provide a comprehensive overview on synthesis, formation mechanisms, characterization, modification, and applications of MCM-41 in order to identify significant trends. 2. General Background Zeolite was first discovered in 1756 by the Swedish scientist Cronstedt when an unidentified silicate mineral was heated in a blowpipe flame which fused readily with marked fluorescence. He called, therefore, this mineral zeolite (in Greek, zeo ) boil and litho ) stone). The term molecular sieve was derived from McBain (1932) when he found that chabazite, a mineral, had a property of selective adsorption of molecules smaller than 5 in diameter. Since then the nomenclature of this kind of porous material seems to be ambiguous. The success of synthetic crystalline aluminosilicates, in particular the emergence of the new family of aluminophosphates (Wilson et al., 1982) and silicoaluminophosphates (Lok et al., 1984), made the concept of zeolite and molecular sieves more intricate. In a broad sense, zeolites are molecular sieves. Strictly speaking, zeolites are crystalline aluminosilicates with molecular sieve properties. Hence, MCM-41 is called a molecular sieve. According to the definition of IUPAC, mesoporous materials are those that have pore diameters between 20 and 500 . Table 1 typically lists some zeolites and molecular sieves with variable pore sizes, mostly in the microporous range prior to the discovery of MCM-41. There has been, however, an ever growing interest in expanding the pore sizes of zeotype materials from

the micropore region to mesopore region in response to the increasing demands in both industrial and fundamental studies. Examples are treating heavy feeds (e.g., processing of tar sand from the high distillates of crude oils to valuable low-boiling products), separating and synthesizing large molecules (e.g., protein separation and selective adsorption of large organic molecules from waste water) and intrazeolite fabricating technology (e.g., supramolecular assembly of molecular arrays, metal complexes encapsulated in zeolite frameworks, and introduction of nanometer particles into zeolites and molecular sieves for electronic and optical applications) (Davis, 1992; Mitchell, 1994; Ozin, 1992). Therefore, numerous works to create zeotype materials with pore diameters larger than those of the traditional zeolites were carried out. It was not until 1982 (Wilson et al., 1982) that success was achieved by changing the synthesis gel compositions when the first so-called ultra large pore molecular sieve, AlPO4-8, which contains 14membered rings, was discovered (yet its structure was only solved in 1990 (Dessau et al., 1990)). Indeed, this not only broke the deadlock of the traditional viewpoint that zeolite molecular sieves could not be constructed with more than 12-membered rings, but also stimulated further investigations on other ultra large pore molecular sieves, such as VPI-5 (Davis et al., 1988), cloverite (Estermenn et al., 1991), and JDF-20 (Jones et al., 1993). However, these materials have not found, up till now, any significant applications because of either their inherently poor stability or their weak acidity. As a consequence, they seem to be inferior compared to pillared layered clays. In 1990, Yanagisawa et al. (1990) reported the synthesis of the mesoporous materials characteristic of MCM-41. Through intercalation of long-chain (typically C16) alkyltrimethylammonium cations into the layered silicate kanemite followed by calcination to remove the organic species, a mesoporous material was obtained. The silicate layers condensed to form a three-dimensional structure with nanoscale pores. 29Si solid-state NMR indicated that a large number of Q3 species were converted to Q4 species during the intercalation and calcination process. X-ray powder diffraction gave only an uninformative intensity centered at extreme low angles. Unfortunately, no further characterization data were available, so Yanagisawa et al.s results have been ignored. In 1992, researchers at Mobil Corporation discovered the M41S family of silicate/aluminosilicate mesoporous molecular sieves with exceptionally large uniform pore structures (Beck et al., 1992a; Kresge et al., 1992), which has resulted in a worldwide resurgence in this

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996 2077

Figure 1. Phase sequence of surfactant-water binary system (Myers, 1992; Lawrence, 1995).

area. The template agent used is no longer a single, solvated organic molecule or metal ion, but rather a selfassembled surfactant molecular array as suggested initially (Beck et al., 1992a). Three different mesophases in this family have been identified, i.e., lamellar (Dubois et al., 1993), hexagonal (Beck et al., 1992a), and cubic phases (Vartuli et al., 1994), in which the hexagonal mesophase, MCM-41, possesses highly regular arrays of uniform-sized channels whose diameters are in the range of 15-100 depending on the templates used, the addition of auxiliary organic compounds, and the reaction parameters. The pores of this novel material are nearly as regular as, yet considerably larger than those present in crystalline materials such as zeolites, thus offering new opportunities for applications in catalysis (Corma et al., 1995b; Kozhevnikov et al., 1995; Kloetstra and van Beckkum, 1995; Armengol et al., 1995), chemical separation (Thomas, 1994), adsorption media (Rathousky et al., 1994, 1995; Llewellyn et al., 1995; Branton et al., 1993, 1994, 1995; Feuston et al., 1994), and advanced composite materials (Huber et al., 1994; Wu and Bein, 1994a-c; Abe et al., 1995). Accordingly, MCM-41 has been investigated extensively because the other members in this family are either thermally unstable or difficult to obtain (Vartuli et al., 1994). 3. Chemistry of Surfactant/Silicate Aqueous Solution Reaction gel chemistry is considered to play an important role in zeolite synthesis (Feijen et al., 1994). It is, therefore, not surprising that much effort has been made to elucidate the gel chemistry related to the formation mechanisms and eventually to the resultant products for MCM-41 synthesis (Beck et al., 1994; Vartuli et al., 1994; Chen et al., 1993; Huo et al., 1994a,b). In this sense, knowledge of the chemistry of surfactant/silicate solution is a prerequisite for understanding the synthesis and mechanisms responsible for the formation of MCM-41 from its precursors. 3.1. Behavior of Surfactant Molecules in an Aqueous Solution. In a simple binary system of water-surfactant, surfactant molecules manifest themselves as very active components with variable structures in accordance with increasing concentrations, as schematically shown in Figure 1. At low concentrations, they energetically exist as monomolecules. With increasing concentration, surfactant molecules aggregate together to form micelles in order to decrease the system entropy. The initial concentration threshold at which monatomic molecules aggregate to form isotropic micelles is called cmc (critical micellization concentration). As the concentration process continues, hexagonal close

packed arrays appear, producing the hexagonal phases (Lawrence, 1994). The next step in the process is the coalescence of the adjacent, mutually parallel cylinders to produce the lamellar phase. In some cases, the cubic phase also appears prior to the lamellar phase. The cubic phase is generally believed to consist of complex, interwoven networks of rod-shaped aggregates (Fromherz, 1981). According to Myers (1992), the particular phase present in a surfactant aqueous solution at a given concentration depends not only on the concentrations but also on the nature of itself (the length of the hydrophobic carbon chain, hydrophilic head group, and counterion) and the environmental parameters (pH, temperature, the ionic strength, and other additives). This is reflected by the effect of the aforementioned matters on cmc. Generally, the cmc decreases with the increase of the chain length of a surfactant, the valency of the counterions, and the ion strength in a solution, respectively. On the other hand, it increases with increasing counterion radius, pH, and temperature. For example, in an aqueous solution at 25 C, the cmc is about 0.83 mM for surfactant C16H33(CH3)3N+Br-; between the cmc and 11 wt %, small spherical micelles are present; in the concentration range of 11-20.5 wt %, elongated flexible rodlike micelles are formed (Chen et al., 1993b); hexagonal liquid crystal phases appear in the concentration region between 26 and 65 wt %, followed by the formation of cubic, lamellar, and reverse phases with increasing concentration (Lawrence, 1995). At 90 C, the hexagonal phase is observed at a surfactant concentration of more than 65% (Steel et al., 1994). Many characterization techniques can be employed to identify the individual aggregate state, among which the 14N NMR is one of the most effective means. Figure 2 shows the 14N NMR spectra for cetyltrimethylammonium chloride (CTACl) in aqueous solution at 90 C (Steel et al., 1994). It is clear that with the increase in concentrations of surfactant CTACl the phase sequence is isotropic micelles f hexagonal phase f hexagonal + cubic phase f cubic + intermediate phase f lamellar phase. 3.2. Chemistry of Silicates/Aluminosilicates in an Alkaline Aqueous Solution. As many as 25 individual silicate structures in an alkaline solution have been identified so far, as illustrated in Figure 3 (Swaddle et al., 1995), which are deemed as important species for zeolite and molecular sieve synthesis. It has been found that the distribution of these anionic silicate species are sensitive to pH, temperature, cation, and Si concentration. A reduction in silicon concentration or increase in temperature or pH favors the formation of monomer and small oligomers (Kinrade and Swaddle,

2078

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996

Figure 2. 14N NMR spectra of CTACl aqueous solution with different concentrations at 90 C (Steel et al., 1994).

intermediates with the same structures responsible for the formation of zeolites and molecular sieves. That is, the aluminosilicate species are much like those of the numerous silicate oligomers shown in Figure 3. 27Al NMR has revealed that an increase in the SiO2/Al2O3 ratio favors the reaction between these two species (Muller et al., 1981). This means that large oligomers of silicate easily react with aluminate species because increasing silicate concentration results in the large extent of oligomerization. It has also been found that both the solubility and the distribution of aluminosilicate structures can be altered upon adding organic alkylammoniums (Engelhardt and Mickel, 1987). For instance, when tetramethylammonium hydroxide (TMAOH) is used to dissolve SiO2, the predominant species are D3R (Q36), D4R (Q38), and D5R (Q310). Additionally, increasing the length of the alkyl chains of organic species often results in the predominance of the D3R anionic species (McCormick et al., 1989). This reveals that the nature of the amine species exerts a key influence on the distribution of silicate/aluminosilicate species. Undoubtedly, the anion-cation pairing is one of several factors acting on the resulting species distributions (Hendricks et al., 1991), as well as the resultant products when considering zeolite synthesis. 4. Synthesis of MCM-41 As the traditional zeolite and molecular sieve synthesis, preparation of MCM-41 materials is also made by mixing organic amine (surfactants), silica, and/or silica-alumina source to form a supersaturated solution while maintaining the mixture at a temperature between 70 and 150 C for selected periods of time. Since the first report of MCM-41 (Beck et al., 1992a; Kresge et al., 1992), a number of patents (see recent review of Casci (1994)) and publications (see Table 2) on the synthesis of MCM-41 have been reported. Table 2 attempts to list the publications concerning synthesis of MCM-41, and it is shown that the majority of interests in the previous reports were aimed at expanding synthesis conditions, generating acidity, synthesizing hybrid atom MCM-41, and identifying synthesis mechanisms. 4.1. Expanding Synthesis Conditions. It has been demonstrated that the preparation of MCM-41 materials can be achieved by using a broad spectrum of surfactants and a wide range of synthesis conditions (e.g., temperatures, pH, reaction time) (Huo et al., 1994a; Monnier et al., 1993; Tanev et al., 1994). Table 3 lists four possible routes through which the hexagonal MCM-41 materials with various compositions have been obtained. Route I, the initially discovered pathway, and route II, expanded by Huo et al. (1994a), are easily imagined. In the case of route III, the cooperative assembly of S+H-I+ complexes was suggested through the mediation of the counterion H- (Huo et al., 1994a, b). Route IV, postulated by Tanev and Pinnavaia (1995), was proposed to be templated by the neutral ammine liquid crystal. The various routes for creating MCM-41 materials offer the possibility to generate mesophases with various compositions and characteristics. Such materials are anticipated to find main applications in electrochromic or solid electrolyte devices (Lampert and Ganqvist, 1990), high surface area redox active catalysts (Thomas, 1994), as well as biochemical and pharmaceutical separation matrices (Fisher, 1995). Interestingly, the pore sizes or interlayer distances can also be tailored by changing the chain length of

Figure 3. Silicate species that have been identified by 29Si NMR in alkali aqueous solution (Swaddle et al., 1995). Filled circles represent tetrahedral Si atoms; lines represent Si atoms linked through O atoms.

1988). When the ratio of SiO2/Na2O is more than 1, for instance, the major form of Si is Q3 [Si(OSi)3OH] and Q2 [Si(OSi)2(OH)2] connectivity. Small anions, such as the cyclic trimer, are found at high concentrations of Si instead of the larger anions (McCormick et al., 1989). Some other silicate species, such as the hexagonal prism (Q312) and the four-way connectivity Q4 [Si(OSi)4] species, characteristic of many zeolite structures are, on the contrary, generally not observed in an alkaline aqueous solution. The aforementioned silicate species can react with monomeric Al(OH)4- species (the only species in alkaline solution with pH ) 7-13) to produce aluminosilicate

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996 2079

alkylhexadecyldimethylammonium, i.e., (C16H33)(CnH2n+1)(CH3)2N+ (Sayari et al., 1995a). In this system, a hexagonal phase was obtained for n ) 1, 3, 5, and 7, while a lamellar phase was obtained for all other templates (see Figure 4). The d100 spacing of the hexagonal MCM-41 increases with the increasing carbon number (n) at a rate of 2.45 /C atom, which is much higher than the 1 /C atom reported by Beck et al. (1992a) for a series of calcined samples. Surprisingly, the d100 spacing of the lamellar phases was almost constant at ca. 32.5 regardless of the n value, indicating that this distance of lamellar phases is controlled by the C16 carbon chain only. Ryoo et al. (1995) reported the synthesis of highly ordered MCM-41 by adding acetic acid to the reaction system to shift the equilibrium between the reactants and the mesophases formed toward to the desired direction. The authors suggested that MCM-41 phase was in dynamic equilibrium with the reactants. Addition of acid to the reaction mixture was proposed to neutralize the hydroxyl so as to shift the equilibrium toward the positive direction. We have synthesized MCM-41 in an alkali-metal-free system using aqueous ammonia solution to adjust the pH of the reaction gel Zhao et al. (1995a). The resultant product was found to exhibit higher catalytic activity for n-heptane cracking compared to those obtained from the sodium-containing system. This was explained that more mild acidic sites could be generated from the sample synthesized in an alkali-metal-free system than those synthesized in a sodium-containing system. Other modified synthesis methods, such as low temperature synthesis (Elder et al., 1995; Zhao and Goldfard, 1995) and simplifying the reactants (Chen et al., 1993a), are also found in the literature. 4.2. Synthesis of Aluminum-Rich MCM-41. Acid sites, in particular the Brnsted ones (from the tetrahedrally coordinated Al), are the active locus for most hydrocarbon reactions. Therefore, incorporating tetrahedrally coordinated aluminum to the framework of MCM-41 has been the subject of much attention. Schmidt et al. (1994a) reported the synthesis of MCM41 with Si/Al ratios as low as 8.5. No disordered aluminum was found either in the as-synthesized or the calcined sample (540 C for 7 h) as characterized by 27Al MAS NMR. Borade et al. (1995) reported the synthesis of MCM-41 with a Si/Al ratio as low as 2, without observing the presence of octahedral Al in the 27Al MAS NMR using sodium aluminate as the aluminum source. However, no data concerning the acidity were given in the above authors works. In fact, one is most interested in acidity rather than the Si/Al ratios. Luan et al. (1995) have extensively investigated the influence of the aluminum source upon the resulting coordination state by examining the 27Al NMR and 29Si NMR spectra. They found that when Captal aluminum or sodium aluminate was used, virtually all Al in the product was 6-coordinate, whereas 4-coordinate aluminum in the MCM-41 framework with a Si/Al ratio as low as 10 could only be obtained by using aluminum sulfate as the aluminum source. On the other hand, Corma et al. (1994b) and Chen et al. (1993a) have found that the most active aluminum source for the formation of MCM-41 is sodium aluminate. In our laboratory (Zhao et al., 1994), we have also examined the effect of aluminum source for synthesis of MCM-41 in a hydrothermal system of 4.0Na2O:29.2SiO2:Al2O3:6.0CTACl: 900H2O and discovered that highly ordered MCM-41

synthesis using primary amine at room temperature in acidic media synthesis of V-MCM-41 synthesis of Ti-MCM-41 in sodium-free system synthesis of MCM-41 in an alkali-free system synthesis of MCM-41 with highly tetrahedral aluminum (Si/Al ) 8.5) synthesis of MCM-41 with Si/Al ratio as low as 2 influence of aluminum source on coordination state synthesis of highly ordered MCM-41 by adding acetic acid synthesis of (Ti)MCM-41 in a sodium system using surfactant with two alkyl chains synthesis of Fe-MCM-41 synthesis of B-MCM-41 synthesis of Mn-MCM-41 18 144 28 72 24 16-70 48 72 30-168 ambient 100 140 100 100 100 150 100 100-150 150 100 21-100 Yuan et al., 1995 Sayari et al., 1995b Zhao and Goldfarb, 1995 168-240 24 72 Tanev et al., 1994 Reddy et al., 1994 Corma et al., 1994a Zhao et al., 1995a Schmidt et al., 1995a Borade et al., 1995 Luan et al., 1995 Ryoo et al., 1995 Sayari et al., 1995a

synthesis of MCM-41 via liquid crystal mechanism relationship between surfactant chain length and pore diameter; role of auxiliary organized synthesis and characterization influence of surfactant/silica molar ratio on mesophase synthesis of MCM-41 in an acidic medium (pH ) 1-5) influence of temperature and surfactant chain length on mesophases

temperature (C)

70-150 100 ambient 100-200

Table 2. Reports on MCM-41 Synthesis

Kresge et al., 1992 Beck et al., 1992a

Chen et al., 1993a Vartuli et al., 1994 Huo et al., 1994a Beck et al., 1994

CTA ) cetyltrimethylammonium[C16H33(CH3)3N+]. b TMA ) (CH3)4N+. c n ) 8, 9, 10, 12, 14, 16.

2.6OH-:30SiO2:Al2O3:6.3CTAa:8.4TMAb:382H2O 2.6OH-:30SiO2:Al2O3:6.3[CnH2n+1(CH3)3N+]c:8.4TMA:382H2O

150 150

reference

30SiO2:0.4Al2O3:2.6CTA:4.5TMAOH:500H2O 30SiO2/CTAOH ) 2 26.9H+:30SiO2:3.5CTA:3800H2O 30SiO2/15[CnH2n+1(CH3)3N+]d ) 2, pH ) 10, 11 wt % surfactant of total mixture 0.6H+:30SiO2:192C2H5OH:8DDAe:1060H2O 5.1Na2O:30SiO2:0.6VO2:15CTMA:900H2O 30SiO2:0.5TiO2:5.2CTA:7.6TMAOH:715H2O 4.0(NH4)2O:30SiO2:Al2O3:6CTA:700H2O 13OH-:30SiO2:xAl2O3:2.1C14H29(CH3)3N+:500H2O 6.5OH-:30SiO2:Al2O3:6.9CTA:2.7TMA:3660H2O 7.8OH-:30SiO2:Al2O3:8CTA:7.8TMA:1755H2O 15OH-:30SiO2:Al2O3:5CTMA:1125H2O 2.3OH-:30SiO2:x(Ti,V)O2:5.8[(C16H33)(CnH2n+1)(CH3)2N+]f:7.6TMAOH:2165H2O 15OH-:30SiO2:0.3Fe2O3:6CTA 2.5Na2O:30SiO2:(6.25-)BO2:4.8CTA:1890H2O (4.7-14.3)OH-:30SiO2:(0.01-2.6)MnO:3.5CTA:4000H2O

molar ratio of gel compositions

n ) 6, 8, 10, 12, 14, 16. e DDA ) dodecylamine [CH3(CH2)11NH2]. f n ) 1-12.

summary

24-240 48 0.5-24

time (h)

48 48

2080

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996

Table 3. Five Possible Routes for Generating Mesophase routea I (S+I-) II (S-I+) III (S+H-I+) IV (S0I0)
a

typical example CTA + silicate species C16H33SO3- + lead oxide C12H25PO42- + ion oxide CTABr + silicate species CTABr + zincophosphate C12H25NH2 + (C2H5O)4Si I-

pH <7 or 10-13 1-5 <2 <3 <7 H-

resulting phase hexagonal, cubic, and lamellar hexagonal lamellar hexagonal lamellar hexagonal ) anionic halides; M+

reference Beck et al., 1992a; Huo et al., 1994a,b Huo et al., 1994a,b Huo et al., 1994a,b Tanev et al., 1995

S+

) surfactant ions;

) anionic inorganic species;

) cationic alkaline ions.

B-MCM-41 (Sayari et al., 1995b) have also been successfully synthesized. 5. Formation Mechanism The most outstanding feature of preparation of MCM41, in contrast to the traditional single organic molecule or metal ion templating preparation, is that the templates used are surfactants, having an alkyl chain length of greater than 6 carbon atoms (in most cases greater than 10 carbon atoms). Therefore, the mechanisms responsible for the formation of MCM-41 from its precursors have attracted much attention. Two typical mechanisms have been proposed so far (Beck et al., 1992a; Monnier et al., 1993), accompanied by other modified routes (Inagaki et al., 1994; Steel et al., 1995). It is well-known that formation of the traditional zeolites and molecular sieves is templated by single molecules or ions. Here we take ZSM-5 as an example as shown in Figure 5. The crystallization is via silicate condensation around a tetrapropylammonium cation. The initial ordered species may consist of an aggregate of water molecules or silicate moieties (Kerr, 1966; Feijen, 1994). Subsequent growth proceeds because of nucleation by this initial structure or assembly of a number of such structures, but crystal growth is the result of the initial silicate organization. In the case of MCM-41 formation, however, Beck et al. (1992a) initially proposed the liquid crystal templating mechanism; i.e., supermolecular arrays preformed are the templates that act as the structure-directing agents (see Figure 6). 5.1. Liquid Crystal Templating (LCT) Mechanism. A key feature of the LCT mechanism is that the liquid crystalline mesophases or micelles act as templates rather than individual single molecules or ions. Accordingly, the final product is a silicate skeleton which contains voids that mimics these mesophases. The silicate condensation is not the dominant factor in the formation of the mesoporous structure. The whole process may be via two possible mechanistic pathways as schematically shown in Figure 6: (1) the liquid crystal mesophases may form prior to the addition of silicate species; (2) the silicate species added to the reaction mixture may influence the ordering of the isotropic rodlike micelles to the desired liquid crystal phase, i.e., hexagonal mesophase. Therefore, the mesophase formed is structurally and morphologically directed by the existing liquid crystal micelles and/or mesophases. The influence of alkyl chain length and the addition of mesitylene on the pore size have been taken as strong evidence for the LCT mechanism, since this phenomenon is consistent with the well-documented surfactant chemistry (Winsor, 1966; Myers, 1992). The auxiliary organic species added to the reaction gel can be solubilized inside the hydrophobic regions of micelles, causing an increase in micelle diameter so as to increase the pore size of MCM-41. The observed pore size

Figure 4. Influence of another alkyl chain length (n) of (C16H33)(CnH2n+1)(CH3)2N+ on the mesophase and the d100 spacing (Sayari et al., 1995a).

can be synthesized from gels with a Si/Al ratio as low as 14 by using sodium aluminate rather than aluminum sulfate. In contrast, when aluminum sulfate is used, highly ordered MCM-41 could only be obtained if the Si/Al ratio was as high as 40, which is in good agreement with the results reported by Corma et al. (1994b) and Chen et al. (1993a). 4.3. Synthesis of Hybrid Atom MCM-41. Isomorphous substitution of T atoms (T ) Si, Al) by other elements can generate a new hybrid atom molecular sieve with interesting properties. A typical example is titanium-containing ZSM-5 zeolite (TS-1), which is an effective catalyst for selective catalytic oxidation of organic compounds in the presence of hydroperoxide (Notari et al., 1988). The range of reactive compounds was greatly limited by the pore sizes of the traditional zeolites and molecular sieves prior to the invention of MCM-41. Therefore, titanium-containing MCM-41 molecular sieves have attracted much attention because their large pore size can permit almost all organic reactants, intermediates, and products to easily pass in and out. Titanium-substituted MCM-41 has been synthesized in an acidic system by using either ionic surfactants (CTA+) or primary amine (DDA) as template (Tanev et al., 1994). The Ti-containing mesophases obtained from the two different templates differ greatly in textural mesoporosity reflected by both the X-ray diffraction (XRD) patterns and the nitrogen adsorptiondesorption isotherms. The catalytic activities of these samples were also remarkably distinct (vide infra). Corma et al. (1994a) have synthesized Ti-MCM-41 in a sodium-free system. A band at 960 cm-1 in the IR spectra and two bands at 210-230 nm in the UV-vis spectra were taken as proof for Ti incorporation into the framework of MCM-41 without isolated Ti atom. V-MCM-41 (Reddy et al., 1994), Mn-MCM-41 (Zhao and Goldfard, 1995), Fe-MCM-41 (Yan et al., 1995), and

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996 2081

Figure 5. Mechanism of the formation model for ZSM-5 templating by single TPA molecule.

Figure 6. Schematic model of liquid crystal templating mechanism via two possible pathways (Beck et al., 1992a).

increase of the aluminosilicates compared to the siliceous MCM-41 was due to the replacement of the shorter Si-O bonds (1.6 ) by the longer Al-O bonds (1.75 ) (Beck et al., 1992a). The LCT mechanism has been further confirmed by subsequent reports (Chen et al., 1993a; Beck et al., 1994; Vartuli et al., 1994). Vartuli et al. (1994) have studied the effect of surfactant/silica molar ratio on the resultant phases in a simple system containing alkali metal, tetraethylorthosilicate, water, and CTAOH at 100 C. They found that as the surfactant/silica molar ratio increased from 0.5 to 2, the siliceous products obtained could be classified into four separate groups: MCM-41 (hexagonal), MCM-48 (cubic), thermally unstable lamellar phase, and the cubic octamer [(CTMA)SiO2.5]8. The data are in excellent agreement with the behavior of surfactants in solution as mentioned above (Myers, 1992). The influence of alkyl chain length (C6-C16) and the reaction temperatures (100-200 C) on the resulting products was also extensively investigated by Beck et al. (1994). Figure 7 shows the XRD patterns of the products synthesized with different surfactants and at different temperatures. At 100 C, with the increase of the surfactant alkyl chain length, the products are in the sequence of amorphous materialsf poorly defined MCM-41f well-defined MCM-41. At 150 C, ZSM-5 zeolite with high crystallinity was generated from both the C6 and C8 surfactants. The C10 surfactant preparation produced less well-defined MCM-41 compared with the C12, C14, and C16 preparation. At 200 C, ZSM-5 was synthesized from C6; a mixture of ZSM-5, ZSM-48, and dense phases from C8-14 surfactants; and amorphous phases from C16. These data are also in good agreement with the surfactant solution chemistry (Winsor, 1968; Myers, 1992). Another interesting deduction from their study was that mesoporous and zeolite materials were not coproduced in the investigated temperature range using C6-16 surfactants as templates. This behavior was explained in terms of the

formation mechanism of MCM-41 which is quite distinct from that of the traditional zeolite templating mechanism. Chen et al. (1993b) have investigated the formation mechanism by employing XRD, 29Si NMR, in situ 14N NMR, and thermogravimetric analysis (TGA) techniques. No hexagonal liquid crystalline mesophases were detected either in the synthesis gel or in the surfactant solution used as template in their study. This has also been confirmed by a later study (Steel et al., 1994). It was, therefore, concluded that formation of MCM-41 phase is possibly via pathway 2 (see Figure 6) rather than pathway 1. That is, the randomly ordered rodlike micelles interact with silicate species by Coulombic interactions in the optimal reaction mixture to produce approximately two or three monolayers of silicate encapsulation around the external surfaces of the micelles. These randomly ordered composite species spontaneously pack into a highly ordered mesoporous phase with an energetically favorable hexagonal arrangement, accompanied by silicate condensation. With the increase in heating time, the inorganic wall continues to condense. It must be noted that the majority of reports regarding LCT mechanism have been investigated in a system containing relatively large amounts of surfactant (generally more than 10 wt % of the total mixture). It is clear that liquid crystal micelles, even crystalline mesophases, indeed exist under such circumstances, and therefore, the LCT mechanism seems to be plausible. 5.2. Transformation Mechanism from Lamellar to Hexagonal Phase. Another formation mechanism proposed by Stucky and co-workers (Monnier et al., 1993; Huo et al., 1994b; Firouzi et al., 1995) is the transformation mechanism from lamellar to hexagonal phase. The central tenet of their conclusion is that the mesophase formed is governed by charge density, coordination state, and steric requirements of the inorganic and organic species at the interface and not necessarily

2082

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996

Figure 7. X-ray powder diffraction patterns of calcined products obtained using different surfactants at (A) 100, (B) 150, and (C) 200 C (Beck et al., 1994).

Figure 8. X-ray powder diffraction patterns of (A) the initially lamellar phase from the reaction mixture of 3.4Na2O:29.2SiO2: 0.7Al2O3:8.3CTACl:2.6TMAOH:3650H2O, (B) an intermediate material of A after hydrothermal treatment, and (C) the hexagonal phase transformed from A after hydrothermal treatment for 10 days (Monnier et al., 1993).

by a preformed liquid crystal structure. In contrast, a micellar assembly of organic molecules will be broken up and rearranged upon addition of inorganic species to form a new phase often with lamellar morphologies (Huo et al., 1994b; Firouzi et al., 1995). As reported previously (Monnier et al., 1993), a lamellar phase was indeed isolated between the reaction period of 1-20 min at 75 C, and after 20 min the hexagonal mesostructures were simultaneously detected (see Figure 8). When the lamellar phase was hydrothermally treated at 100 C, it was converted to the hexagonal phase over 10 days with intermediate, and the final X-ray patterns are shown in Figure 8. Therefore, the authors proposed that in a surfactant/silicate aqueous mixture with relatively low pH, a low degree of polymerization of silica species, and low temperatures, small silica oligomers (three to eight silicon atoms, e.g., S3R, D4R) play an important role in the formation of mesoporous phase

(Monnier et al., 1993). In part, this is because they can interact with surfactant cations by Coulombic interactions at the interfaces, leading to a strong interaction by multidentate binding between them. This multidentate binding of silicate oligomers to the cationic surfactant can subsequently further polymerize to form very large ligands, and enhance the cooperative binding between the surfactant and silicate species. It is these multidentate surfactant-silicate ligands that lead to a lamellar biphase governed by the optimal surfactant average head group area (A0) (Monnier et al., 1993). As polymerization of the silicate species proceeds, the average head group area (A) of surfactant assembly increases due to the decrease of the charge density of larger silicate polyanions. This leads to the corrugation of the silicate layers and ultimately results in the hexagonal mesophase precipitation. Figure 9a schematically shows the transformation mechanism from lamellar to hexagonal phase through charge matching. Alternatively, Huo et al. (1994b) and Firouzi et al. (1995) suggested that ion exchange between surfactant anions (OH-, Br-, Cl-) and multiply charged anionic silica oligomer (D4R, D3R) may take place in a surfactant-silicate aqueous solution because the high anionic charge densities of the oligomer enhance Coulombic attractions (Figure 9b). This multidentate bonding can screen the intraaggregate electrostatic head group repulsions so as to reduce the average head group area of the surfactants. The collective results can further decrease the local curvature of the aggregate, which implies that the cooperative assembly of inorganicorganic species is formed with structures different from that of the precursor micelles. This transformation process and the final mesophase with hexagonal pore structures (rather than cylindrical pore structures) was proved by high-resolution transmission electron microscopy (Alfredsson et al., 1994). Interestingly, Steel et al. (1995) suggested that MCM41 materials are indeed transformed from a lamellar mesophase. However, it is the sandwiched micelles around which silicate species assemble into lamellar mesophase rather than the coorganization process.

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996 2083

been performed under various experimental conditions (e.g., temperature, pH, compositions of reaction gel). These parameters are found to exert key influence on both the surfactant behavior and the distribution of silicate species (Myers, 1992; McCormick, 1992). Extrapolation of such results to a detailed mechanism thus seems impossible. However, it is worth noting that a high surfactant concentration, high pH, low temperatures, and low degree of silicate polymerization always favor the formation of cylindrical micelles, as well as the hexagonal mesophases. In fact, the majority of MCM-41 syntheses were carried out using a precursor consisting of a surfactant solution with concentrations of ca. 25 wt %. In this way, it is inevitable that the dominate aggregates of surfactants are cylindrical micelles which are in equilibrium with the monomeric molecules. Upon combining with silicate species, the reaction between them will take place via ion exchange to form ion parings due to Coulombic interactions at the interfaces, leading to the precipitation of inorganicorganic complexes. The equilibrium between micelles and monomeric molecules is shifted toward the micelle direction. This is indeed the case when a reaction gel contains a relatively high surfactant concentration (Chen et al., 1993b). On the other hand, in a reaction gel containing relatively low surfactant concentrations, the isotropic surfactant molecules or micelles may be disassociated and reorganized upon the adding of silicate species to form lamellar inorganic-organic complexes in order to precipitate (Steel et al., 1995; Fironiz et al., 1995; Huo et al., 1994b). 6. Characterization, Physicochemical Properties, and Structure Model 6.1. Characterization and Structure Model. Proper preparations of MCM-41 will always result in products with high surface areas (more than 1000 m2/ g), large adsorption capacity of benzene (more than 60 wt %), narrow pore size distribution (centered at ca. 3538 using CTA+ as template), and X-ray diffraction patterns with a few distinct maxima in the extreme low angle region. X-ray diffraction patterns give only hko reflections, and no reflections at diffraction angles larger than about 6 2 can be observed (Behrens and Stucky, 1993). The position of these peaks approximately fits the positions for the hko reflections from a hexagonal lattice. However, only a single peak at ca. 2 2 can be detected in most cases; other reflections are virtually absent (Zhao et al., 1994). Preparations with only one remarkable peak have also been found to contain substantial amounts of MCM-41 (Borade et al., 1995; Tanve et al., 1994). Nitrogen adsorption-desorption isotherms of MCM-41 materials exhibit a sharp step commencing at ca. p/p0 ) 0.38, as shown in Figure 11 (Zhao et al., 1994), and the reversible type IV isotherms without hysteresis (Kresge et al., 1992; Schmidt et al., 1995). The inflection point is attributed to the commencement of pore filling from which the pore diameter can be roughly estimated. Figure 12 typically shows the relationship between XRD d100 spacing, Ar pore size, wall thickness (t), and the alkyl chain length (n) for siliceous MCM-41 prepared using alkyltrimethylammonium surfactants with different chain lengths (Beck et al., 1992a). It exhibits that the wall thickness for siliceous MCM-41 is approximately constant at about 10 . Considering a Si-O bond length of 1.6 and a Si-O-Si bond angle of 150 (Chen et al., 1993a), a wall thickness of about 10 can

Figure 9. Schematic model for transformation mechanism from lamellar to hexagonal phase via (a) charge matching and (b) ion exchange pathways (Monnier et al., 1993; Huo et al., 1994; Feuston et al., 1994).

Figure 10. Model schematically representing folded sheets mechanism (Ingaki et al., 1994).

Furthermore, the silicate layers pack between two lamellar biphases due to condensation. The sandwiched micelles finally turn back to randomly isotropic surfactant molecules. 5.3. Folded Sheets Mechanism. Although the folded sheets mechanism, proposed by Inagaki et al. (1994), is based on the intercalation of surfactant to layered silicates (kanemite) process, it should be elucidated here in order to give a comparison. Figure 10 is a schematic model representing the folded sheets mechanism. Initially, the surfactant cations intercalate into the bilayers of kanemite via an ion exchange process. As the ion exchange proceeds, the interlayer cross-linking occurs by condensation of silanols. The silicate sheets of kanemite have the required flexibility to be folded due to its single sheet structure. 5.4. Identifying the Formation Mechanism. Studies concerning MCM-41 formation mechanisms have

2084

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996

Figure 11. Nitrogen adsorption-desorption isotherms of MCM41 (Zhao et al., 1994).

Figure 12. Relationship between surfactant alkyl chain length (n) and argon pore size, a0 (a0 ) 2d100/ 3), and wall thickness (t).

Figure 14. (A) Pyridine adsorption IR spectra for (a) amorphous silica-alumina (Si/Al ) 2.5) and (b) MCM-41 (Si/Al ) 14) in vacuum at (1) 423, (2) 523, and (3) 623 K (Corma et al., 1994b). (B) NH3-TPD profiles for MCM-41 (Si/Al ) 15), amorphous silicaalumina (Si/Al ) 3), and ZSM-5 (Si/Al ) 25) (Zhao et al., 1995a). Figure 13. Structural model of MCM-41 with cylindrial pore (A) (Feuston et al., 1994) and hexagonal pore (B) (Behrens and Stucky, 1993).

be calculated which is in good agreement with the experimental data. Combining with other characterization techniques, such as transmission electron microscopy (Kresge et al., 1992; Beck et al., 1992a), electron diffraction pattern (Kresge et al., 1992), lattice images (Feuston et al., 1994), 27Al and 29Si MAS NMR (Chen et al., 1993a; Firouzi et al., 1995), Fourier transform infrared spectroscopy (Chen et al. 1993), temperatureprogrammed desorption of ammonia (NH3-TPD), and TGA (Chen et al., 1993a), two structural models for MCM-41 with an amorphous wall have been constructed as shown in Figure 13. Model A, a cylindrical pore structure with lattice constant a ) 44.6 and a wall thickness of 8.4 , based on the LCT mechanism, was proposed by Feuston et al. (1994) through the classical molecular dynamics simulation approach. Model B, a hexagonal pore structure with an interpore distance of ca. 35 , based on the lamellar to hexagonal phase

transformation mechanism, was proposed by Behrens et al. (1993). Interestingly, both the cylindrical and hexagonal pore structures have been visualized through transmission electron microscopy by Chenite and Le Page (1995) and Walker et al. (1995), respectively. 6.2. Acidity. Figure 14 shows the pyridine adsorption IR (A) spectra (Corma et al., 1994b) and NH3-TPD profiles (B) (Zhao et al., 1994) of H-form MCM-41 with other samples for comparison. In contrast to ZSM-5, MCM-41 possesses only some weak- and middlestrength acid sites that are similar to amorphous aluminum-silica. The pyridine adsorption spectra also reveal that both the Brnsted acid sites (band at 1540 cm-1) and the Lewis acid sites (band at 1450 cm-1) are comparable to amorphous silica-alumina (Si/Al ) 2.5) This is in good agreement with NH3-TPD results. Both the NH3-TPD and IR data show that MCM-41 materials are aluminosilicates with weak- and middle-strength acid sites similar to amorphous silica-alumina. 6.3. Stability. It has been found that MCM-41 is thermally stable, but the hydrothermal stability is relatively poor, in particular for the as-synthesized

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996 2085


Table 4. Thermal/Hydrothermal and Acid-Base Proof Stability of MCM-41 (Zhao et al., 1994) samplea a treatment method calcination in N2 at 540 C for 1 h and then in air for 8 h, 1 C/min XRD intensity 100 2/g) 1016 BET surface area (m benzene uptakeb (%) 65.5
a

sample b calcination in air at 540 C for 8 h, 1 C/min 95 940 54.0

sample c hydrothermally treated at 450C for 2 h 55 670 33.3

sample d

sample e

20% nitric acid, 5% KOH solution, room temperature, room temperature, overnight overnight 100 1114 3.5 65.5

Sample a as the batch material for other treatments. b At 25 C, 65 Torr. Table 5. Catalytic Activity of Ti-MCM-41 Compared with TS-1 and TiO2a (Tanev et al., 1994) 2,6-DTBP oxidation catalyst Ti-MCM-41 TS-1 TiO2 benzene oxidation conversion selectivity to conversion selectivity to (%) quinone (%) (%) phenol (%) 83 6.5 14.5 >95 >95 >95 37 31 >95 >95 -

a Conditions: 0.1 g of catalyst, 80 C, 2 h, 5 mmol of 2,6-DTBP or 10 mmol of benzene, 0.13 mol of acetone as solvent, 29 mmol of 30% aqueous H2O2.

Thus, the conclusion could be drawn that siliceous MCM-41 is hydrophobic, while aluminosilicate MCM41 is slightly hydrophilic. 7. Application
Figure 15. TGA curves of MCM-41 with different Si/Al ratios (Zhao et al., 1994).

samples (Chen et al., 1993a; Corma et al., 1994b). We have also studied the various stabilities of MCM-41 (see Table 4). When MCM-41 was hydrothermally treated at 450 C for 2 h, considerable losses of both the BET surface area and the benzene sorption capacity occurred. Calcination of MCM-41 in dry air resulted in slight loss of BET surface area, indicating that the thermal stability was satisfactory. When MCM-41 was impregnated in 20% nitric acid overnight, both the BET surface area and the benzene sorption amount were slightly increased. This observation may be associated with the removal of some blockages in the channels of MCM-41. In sharp contrast with the acid treatment, the basic treatment in 5% potassium hydroxide bought about almost complete destruction of the MCM-41 structure demonstrated by both the nitrogen and benzene adsorption and XRD spectra (only a broad intensity at extremely low angle regions was detected). These data indicate that MCM-41 material is thermally stable with high acid resistance, but hydrothermally unstable and with low base tolerance. 6.4. Interaction with Water. The adsorption of water steam over MCM-41 characterized by a type V isotherm reveals an initially repulsive character followed by a capillary condensation step of water, indicating that MCM-41 possesses both hydrophobic and hydrophilic properties (Llewellyn et al., 1995). Three distinct stages were obtained in the TGA curves (Figure 15) (Zhao et al., 1994). The first stage at 25-150 C is associated with the desorption of physically adsorbed water and other gases; the second stage at 150-380 C is attributed to the decomposition and combustion of organic species, and the third stage at 380-800 C may be related to the water losses due to condensation of silanol groups to form siloxane bonds (Chen et al., 1993a). In addition, with the increase of the aluminum content in MCM-41 framework the amount of water desorbed increases and the organic species decrease.

7.1. Catalysis. A US patent (Pelrine et al., 1992) claimed good catalytic activity of MCM-41 impregnated with Cr for olefin oligomerization to produce lube oil additives. Both the pour points and the viscosity indexes were improved on Cr-MCM-41 catalyst compared with the commercial catalyst (Cr-SiO2). The support (MCM-41, alumina-silica, and USY) effect on the catalytic activity of hydrodesulfurization (HDS), hydrodenitrogenation (HDN), and mild hydrocracking (MHC) was also investigated by Corma et al. (1995b). A very high activity was observed on Ni,Mo-MCM-41 catalyst which was attributed to a combination of high surface area and large pore size that favors a high dispersion of the active species coupled with easy accessibility of the large feedstock molecules. Use of MCM-41 as an acid catalyst for Friedel-Crafts alkylation of 2,4-di-tert-butylphenol (bulky aromatic) with cinnamyl alcohol (Kloetstra et al., 1995a) and for the tetrahydropyranylation of alcohol and phenol (Armengol et al., 1995) was reported where the advantages of MCM-41 were manifested. Besides being an acid catalyst, Na-MCM-41 and Cs-MCM-41 catalysts exhibit satisfactory performance in base catalysis (Kloetstra et al., 1995b). For example, in the Knoevenagel condensation of benzaldehyde with ethyl cyanoacetate, 81% conversion of benzaldehyde and 75% selectivity to the desired product were obtained at 150 C within 7 h, and 90% conversion of benzaldehyde and ca. 100% selectivity were observed at 100 C within 3 h in water solvent. Ti(V,Cr)-MCM-41 materials exhibit excellent catalytic oxidation performance in the presence of hydrogen peroxide or even the bulky oxidant THP (terbutyl hydroperoxide), as reported by Tanev et al. (1994) and Corma et al. (1994a). A typical example from Tanev et al. (1994) is shown in Table 5. Both Ti-MCM-41 and TS-1 are effective catalysts for the hydroxylation of benzene, but for substrate 2,6-DTBP, this is not the case because the medium-sized pore structure of TS-1 (MFI) cannot permit such a large molecule to diffuse into the

2086

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996

Table 6. Catalytic Activity of MCM-41 in SCR of NOx NOx conversion (%) catalyst 6.1%Ti, 2.5%V/MCM-41 3.0%Ti, 2.6%V/SiO2 2.0%V2O5, 8.0WO3/TiO2
a

350 C 81 70 95

450 C 41 2 63

structures of MCM-41, thus, offer an opportunity to investigate this fundamental problem. Reports regarding this issue are readily available (Branton et al., 1993, 1994, 1995; Rathousky et al., 1994, 1995). 8. Perspective As indicated above, the marked characteristics of MCM-41 are as follows: (1) well-defined pore structure with apertures in the range of 15-100 which can be controlled by careful choice of surfactants, auxiliary chemicals, and reaction parameters, (2) high thermal stability, (3) mild acidity, (4) large BET surface area and pore volume, and (5) hydrophobic/hydrophilic property which can be modified by changing Si/Al ratios. Therefore, MCM-41 materials have a promising potential in catalysis, adsorption, and advanced molecular sievebased materials. 8.1. Host-Guest Encapsulation. Semiconductor clusters anchored in molecular sieves have received much attentions as advanced composite materials (Stucky and Mac Dougall, 1990; Ozin, 1992) and photocatalysts (Liu et al., 1993; Tanguay et al., 1989). In this case, the uniform pore structures of molecular sieves can act as solid solvents to control both the size and topology of the particles encapsulated inside. Nevertheless, it was found that introduction of such species, in particular large transition metal species (e.g., MoO3) could result in partial collapse of the host structures (Fierro et al., 1987) and ultimately lead to the decrease in surface area as well as the catalytic activity. From this consideration, MCM-41 should exhibit superior properties compared to other common molecular sieves because of its mesopore structure. As an example, we have successfully encapsulated nanosized TiO2 or MoO3 particles in the MCM-41 channels through either ion exchange or impregnation approaches (Zhao et al., 1995b). X-ray powder diffraction patterns show that no structure collapse occurred. Furthermore, no diffraction peaks attributed to these metal oxides at the external surface of MCM-41 could be observed for carefully purified samples. This result was also confirmed by X-ray photoelectron spectroscopy. A remarkable increase in the band gaps of these composite materials was observed in the UV-vis absorption spectra due to quantum size effect (Brus, 1986), indicating that nanometer TiO2 clusters were indeed formed in the MCM41 channels. Metals with two different, stable oxidation states, such as Co(II)/Co(III), Fe(II)/Fe(III), Cr(II)/Cr(V), and Cu(I)/Cu(II), were found to be effective oxygen carriers for catalysis (Imamura and Lunsford, 1985; Mortier and Schoonheydt, 1985; Thomas, 1994). It was found, however, that the oxygen-carrying ability of these ligands was due to the formation of M-O-M (M ) Cu, Cr, Fe, etc.) bridges in the cubooctahedra. This would lead to diffusion limitations of reactants and products into and out of the cubooctahedra. On the other hand, metal-organic supermolecule complexes encapsulated in zeolite cavities were also found to be effective gas carriers and reaction centers for catalysis (Ozin and Gil, 1989; Mallouk and Lee, 1990; Mitchell, 1991). It was pointed out that the degree of [CoII(bpy)(terpy)]2+ complex formation inside a NaY zeolite was very low, hence the efficiency for the utilization was low. In addition, some complexes are impossible to be encapsulated inside the cages of the traditional zeolites because of their too large dynamic diameters (e.g., the ligand of [Co(phthalo-

Conditions: 2.0 g of catalyst, WHSV ) 400 mL/min, 125 ppm NO, 125 ppm NH3, 0.12% O2, helium balance.

internal surface where the active sites (isolated Ti species) are located. Molecular sieves have drawn much interest in environmental catalysis in the past few years (Dartt and Davis, 1994; Iwamoto, 1994), in particular in decomposition of nitrogen oxides (Iwamoto, 1994; Tabata, 1994). The activity of selective catalytic reduction of NO on Ti,V/MCM-41, Ti,V/SiO2, and a commercial catalyst V,W/TiO2 has been compared (Beck et al., 1992b) as shown in Table 6. Ti,V/MCM-41 catalyst exhibits a higher NOx conversion than the silica-based catalyst but is less active than the commercial one. This was explained in terms of the fact that although MCM-41 is capable of supporting more active components than silica due to its high BET surface area it has a low redox performance. In a recent study, we have demonstrated that MCM-41 can be a potential NO decomposition catalyst support with considerably higher activity than ZSM-5 (Zhao and Lu, 1995). 7.2. Ship in a Bottle. The large cavities and uniformed pore structures provide zeolites and molecular sieves with many interesting characteristics. For example, they can act just as molecular factories (bottles) where quantum-sized particles (ships) can be manufactured inside (Ozin, 1992; Dag and Ozin, 1995; Mitchell, 1991). The huge workshop and broad channels (without traffic block) for MCM-41 are expected to manifest itself with specified characteristics in the area of ship in a bottle as Fisher (1995) pointed out. Nanosized catalysts, such as low-valent transition metal moieties, e.g., Me3SnMo(CO)3(-C5H5), have been introduced into the channels of MCM-41 by combining calcined MCM-41 with Me3SnMo(CO)3(-C5H5) in hexane solvent for 18 h under stirring, and importantly these larger ligands seem to be very stable (Huber et al., 1994). When the attached complexes were heated, they were converted into nanometer bimetallic clusters in MCM-41 channels, which displayed high catalytic activity. The fabrication of stable carbon wires in the channels of MCM-41 was reported by Wu et al. (1994). The MCM-41 host was contacted with acrylonitrile vapor at room temperature for 4 h and then briefly evacuated, resulting in a large amount of adsorbed acrylonitrile. Polymerization of acrylonitrile in the channels of MCM-41 was carried out through a freeradical reaction process initiated by adding K2S2O8 and NaHSO3 at 40 C. Finally, the sample was thermally treated in a nitrogen flow between 350 and 1000 C for 24 h. The final nanosized carbon wires exhibited a high thermal stability (more than 800 C) and low-field conductivity. 7.3. Model Adsorbent. As shown in Figure 11, MCM-41 materials exhibit an interesting nitrogen adsorption-desorption isotherms without hysteresis loop. However, it was found that this is not the case for other adsorptives (Branton et al., 1995) and at different temperatures (Rathousky et al., 1995). This reflects the importance of the pore shape and the network of adsorbent. The well-controlled uniform pore

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996 2087

Figure 16. Schematic representation of [Co(phthalocyanine)]2+ metal complex entrapped in MCM-41 channels.

Figure 17. Comparison of p-nitroaniline molecule alignment in the channel of (A) AlPO4-5 and (B) MCM-41.

cyanine)]2+ with a diameter of 15 cannot be encapsulated by NaY zeolite). All of these drawbacks can now be overcome by taking advantage of MCM-41 materials, as shown in Figure 16, because the advantages of MCM41 (high surface area and mesoporous structure) can offer the possibility to support a large amount of chemical ligands with a large dynamic diameter. Attempts have also been made to make zeolite-based solids capable of generating nonlinear optical properties which are dramatically different from those of either host or guest in recent years. For example, p-nitroaniline (pNA), a centrosymmetric crystal, does not exhibit second harmonic generation (SHG), nor does AlPO4-5 molecular sieve. However, when pNA molecules were introduced into the one-dimensional channels of AlPO4-5, a high SHG signal will be detected because the polar host crystal forces alignment of the guest molecules so as to enhance the SHG (Marlow et al., 1994). If one imagines the p-nitroaniline molecules being aligned in the MCM-41 channels with a 40 pore diameter as indicated in Figure 17, the SHG signal generated might be expected to be 5-7 factors more than that for AlPO4-5. 8.2. Modification. From the consideration of the uniform mesoporous structure, mild acid properties, and acid proof stability, introduction of superacid (e.g., SO42-, F-) or heteropoly acid into the MCM-41 channels to modify the acidity seems to be both practical and promising. Kozhernikov et al. (1995) recently reported results for the introduction of tungstophosphoric acid (H3PW12O40) into the MCM-41 pores. By shaking MCM41 with H3PW12O40 solution overnight, finely dispersed H3PW12O40 species on the surface of MCM-41 were obtained, which exhibited strong Brnsted acid sites similar to those of H3PW12O40 supported on amorphous silica. The catalytic activity of this composite material was tested for the alkylation of TBP with isobutene as a model reaction. H3PW12O40/MCM-41 shows a high catalytic activity which is 3-4 times higher than that

for the bulk H3PW12O40 catalyst. Supported SO4- and SO4--ZrO2 over MCM-41 may be a promising approach to get highly dispersed superacids. The pore sizes of MCM-41 materials were claimed to be distributed in the range of 15-100 . It generally ranged from 20 to 40 . The gap between 13 and 20 makes it inconvenient to use this material. Although mesoporous materials with pore sizes in the range of 14-22 have been reported through the gallerytemplated route (Galarneau et al., 1995), this is not satisfactory to zeolite and molecular sieve workers. Therefore, precise control of the pore structures by either hydrothermal synthesis or postmodified approach (Beck et al., 1993) in order to make this novel material a true molecular recognizer is an urgent need. 8.3. Adsorbent. MCM-41 may be expected to find application as an adsorbent since it exhibits both hydrophobic and hydrophilic character depending upon the exact composition and/or postmodification. Removal of hydrocarbons from water, storage of gases (e.g., H2, O2, CH4), adsorptive xylene separation, and separation of biological and pharmaceutical compounds now seems to be potential areas for developing MCM-41 applications. 8.4. Environment. MCM-41 may find wide applications in environmentally safe processes, i.e., replacement of environmentally hazardous catalysts in existing processes. Titanium-containing MCM-41, Ru(II), Co(II), and Ni(II) complexes and Mo(CO)6 ligands encapsulated within the intrasurface of MCM-41, etc. are very promising heterogeneous catalysts to replace traditional homogeneous catalysts because of their ready regenerability, shape selectivity, and easy separation and recovery. 9. Acknowledgment The authors are grateful to Dr. D. Y. Zhao, the Weizmann Institute of Science, Israel, for some valuable discussions. We also thank Mr. David Page, Center for Microscopy and Microanalysis at the University of Queensland, for his help in XRD characterization. X.S.Z. wishes to thank Dalian Institute of Chemical Physics, Chinese Academy of Sciences, where part of the work on MCM-41 was conducted, for permitting a sabbatical leave. 10. Literature Cited
Abe, T.; Tachibana, Y.; Uematsu, T.; Iwamoto, M. Preparation and Characterisation of Fe2O3 Nanoparticles in Mesoporous Silicate. J. Chem. Soc., Chem. Commun. 1995, 1617-1618. Alfredsson, V.; Keung, M.; Monnier, A.; Stucky, G. D.; Unger, K. K.; Schuth, F. High-Resolution Transmission Electron Micros copy of Mesoporous MCM-41 Type Materials. J. Chem. Soc., Chem. Commun. 1994, 921-922. Armengol, E.; Cano, M. L.; Corma, A.; Garca, H.; Navarro, M. T. Mesoporous Aluminosilicate MCM-41 as A Convenient Acid Catalyst for Friedel-Crafts Alkylation of a Bulky Aromatic Compound with Cinnamyl Alcohol. J. Chem. Soc., Chem. Commun. 1995, 519-520. Armor, J. N. Environmental Catalysis. Appl. Catal. 1992, B1, 221256. Beck, J. S.; Vartuli, C.; Roth, W. J.; Leonowicz, M. E.; Kresge, C. T.; Schmitt, K. D.; Chu, C. T-W.; Olson, D. H.; Sheppard, E. W.; McCullen, S. B.; Higgins, J. B.; Schlenker, J. L. A New Family of Mesoporous Molecular Sieves Prepared with Liquid Crystal Templates. J. Am. Chem. Soc. 1992a, 114, 1083410843.

2088

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996


Dubois, M.; Gulik-krzywicki, Th.; Cabane, B. Growth of Silica Polymer in a Lamelar Mesophase. Langmuir 1993, 9, 673-680. Elder, K. J.; White, J. W. Room-Temperature Formation of Molecular Sieve MCM-41. J. Chem. Soc., Chem. Commun. 1995, 155-156. Engelhand, G.; Michel, D. High Resolution Solid-State NMR of Silicates and Zeolites; John Wiley: Chichester, 1987. Estermann, M.; Mccusker, L. B.; Baerlocher, Ch.; Merrouche, A.; Kessler, H. A Synthetic Gallophosphate Molecular Sieves with a 20-Tetrahedral-Atom Pore Opening. Nature 1991, 352, 320323. Feijen, E. J. P.; Martens, J. A.; Jacobs, P. A. Zeolites and their Mechanism of Synthesis. Stud. Surf. Sci. Catal. 1994, 84, 3-21. Feuston, B. P.; Higgins, J. B. Physical Sorption of Argon, Nitrogen and Oxygen by MCM-41, a Model Mesoporous Adsorbent. J. Phys. Chem. 1994, 98, 4459. Fierro, J. C. G.; Conesa, J. C.; Agudo, A. L. Migration of Molybdenum into Intracrystalline Cavities in MolybdateImpregnated NaY Zeolite. J. Catal. 1987, 108, 334-345. Firouzi, A.; Kumar, D.; Bull, L. M.; Besier, T.; Sieger, P.; Huo, Q.; Walker, S. A.; Zasadzinski, J. A.; Glinka, C.; Nicol, J.; Margolese, D.; Stucky, G. D.; Chmelka, B. F. Cooperative Organization of Inorganic-Surfactant and Biomimetic Assemblies. Science 1995, 267, 1138-1143. Fisher, C. Making the Most of Molecular Sieves. Chem. Ind. 1995, 461-462. Frank, O.; Schulz-Ekloff, G.; Rathousky, J.; Starek, J.; Zurcai, A. Unusual Type of Adsorption Isotherm Describing Capillary Condensation without Hysteresis. J. Chem. Soc., Chem. Commun. 1993, 724-726. Fromherz, P. Micelle Structure: A Surfactant-Block Model. Chem. Phys. Lett. 1981, 77, 460-465. Fyte, C. A.; Gies, H.; Kokotailo, G. T.; Marler, B.; Cox, D. E. Crystal Structure of Silica-ZSM-12 by the Combined Use of HighResolution Solid-State MAS NMR Spectroscopy and Synchrotron X-Ray Powder Diffraction. J. Phys. Chem. 1990, 94, 37183721. Galarneau, A.; Barodawalla, A.; Pinnavala, T. J. Porous Clay Heterostructures Formed by Gallery-Templated Synthesis. Nature 1995, 374, 529-531. Hendricks, W. M.; Bell, A. T.; Radle, C. J. Effect of Organic and Alkali Metal Cations on the Distribution of Silicate Anions in Aqueous Solution. J. Phys. Chem. 1991, 95, 9513-9519. Herron, N. Toward Si-based Life: Zeolites as Enzyme Mimics. CHEMTECH 1989, 543-548. Huber, C.; Moller, K.; Bein, T. Reactivity of A Trimethylstannyl Molybdenum Complex in Mesoporous MCM-41. J. Chem. Soc., Chem. Commun. 1994, 2619-2620. Huo, Q.; Margolese, D. L.; Ciesla, U.; Feng, P.; Gier, T. E.; Sieger, P.; Leon, R.; Retroff, P. M.; Schuth, F.; Stucky, G. D. Generalized Synthesis of Periodic Surfactant/Inorganic Composite Materials. Nature 1994a, 368, 317-321. Huo, Q.; Margolese, D. I.; Ciesla, U.; Demuth, D. G.; Feng, P.; Gier, T. E.; Sieger, P.; Firouzi, A.; Chmelka, B. F.; Schuth, F.; Stucky, G. D. Organization of Organic Molecules with Inorganic Molecular Species into Nanocomposite Biphase Arrays. Chem. Mater. 1994b, 6, 1176-1191. Inagaki, S.; Fukushima, Y.; Kuroda, K. Synthesis of Highly Ordered Mesoporous Materials from a Layered Polysilicate. J. Chem. Soc., Chem. Commun. 1993, 680-682. Inagaki, S.; Fukushima, Y.; Kuroda, K. Synthesis and Characterisation of Highly Ordered Mesoporous Material, FSM-16, from a Layered Polysilicate. Stud. Surf. Sci. Catal. 1994, 84, 125-132. Iwamura, S.; Lunsford, J. H. Separation of Oxygen from Air by [CoII(bpy)(terpy)]2+ Complexes in Zeolite Y. Langmuir 1985, 1, 326-330. Jones, R. H.; Thomas, J. M.; Chen, J.; Xu, R.; Huo, Q.; Li, S.; Ma, Z.; Chippindale, A. M. Structure of an Unusual Aluminium Phosphate ([ Al5P6O24H]2-2[N(C2H5)3H]+ 2H2O) JDF-20 with Large Elliptical Apertures. J. Solid State Chem. 1993, 102, 204-208. Kerr, G. T. J. Chemistry of Crystalline Aluminosilicates. I. Factors Affecting the Formation of Zeolite A. J. Phys. Chem. 1966, 70, 1047-1058. Kinrade, S. D.; Swaddle, T. W. Silicon-29 NMR Studies of Aqueous Silicate Solutions. I. Chemical Shifts and Equilibria. Inorg. Chem. 1988, 27, 4253-4259.

Beck, J. S.; Socha, R. F.; Shihabi, D. S.; Vartuli, J. C. Selective Catalytic Reduction (SCR) of Nitrogen Oxides. U.S. Patent, 5,143,707, 1992b. Beck, J. S.; Calabro, D. C.; Mc Cullen, S. B.; Pelrine, B. P.; Schmitt, K. D.; Vartuli, J. C. Sorption Separation over Modified Synthetic Mesoporous Crystalline Material. U.S. Patent, 5,220,101, 1993. Beck, J. S.; Vartuli, J. C.; Kennedy, G. J.; Kresge, C. T.; Roth, W. J.; Schramm, S. E. Molecular or Supramolecular Templating: Defining the Role of Surfactant Chemistry in the Formation of Microporous and Mesoporous Molecular Sieves. Chem. Mater. 1994, 6, 1816-1821. Behrens, G.; Stucky, G. D. Order Molecular Arrays as Templates: A New Approach to Synthesis Mesoporous Materials. Angew. Chem., Int. Ed. Engl. 1993, 32, 696-669. Bialek, R.; Meier, W. M.; Davis, M. E. The Synthesis and Structure of SSZ-24, the Silica Anaology of AlPO4-5. Zeolite 1991, 11, 438444. Boade, R. B.; Clearfield, A. Synthesis of Aluminium Rich MCM41. Catal. Lett. 1995, 31, 267-272. Branton, P. J.; Hall, P. G.; Sing, K. S. W. Physisorption of Nitrogen and Oxygen by MCM-41, a Model Mesoporous Adsorption. J. Chem. Soc., Chem. Commun. 1993, 1257-1258. Branton, P. J.; Hall, P. G.; Sing, K. S. W.; Reichert, H.; Schuth, F.; Unger, K. K. Physisorption of Argon, Nitrogen and Oxygen by MCM-41, a Model Mesoporous Adsorbent. J. Chem. Soc., Faraday Trans. 1994, 90, 2965-2967. Branton, P. J.; Hall, P. G.; Treguer, M.; Sing, K. S. W. Adsorption of Carbon Dioxide, Sulfur Dioxide and Water Vapor by MCM41, a Model Mesoporous Adsorbent. J. Chem. Soc., Faraday Trans. 1995, 91, 2041-2043. Brus, L. E. Electronic Wave Functions in Semiconductor Clusters: Experiment and Theory. J. Phys. Chem. 1986, 90, 25552560. Casci, J. L. The Preparation and Potential Applications of UltraLarge Pore Molecular Sieves: A Review. Stud. Surf. Sci. Catal. 1994, 85, 329-356. Chen, C. Y.; Li, H. Y.; Davis, M. E. Studies on Mesoporous Materials. I. Synthesis and Characterisation of MCM-41. Microporous Mater. 1993a, 2, 17-26. Chen, C. Y.; Li, H. Y.; Davis, M. E. Studies on Mesoporous Materials. II. Synthesis Mechanism of MCM-41. Microporous Mater. 1993b, 2, 27. Chen, N. Y.; Garwood, W. E.; Dwyer, F. G. Shape Selective Catalysis in Industrial Applications; Marcel Dekker: New York, 1989. Chenite, A.; Le Page, Y. Direct TEM Imaging of Tubules in Calcined MCM-41 Type Mesoporous Materials. Chem. Mater. 1995, 7, 1015-1019. Corma, A.; Iglesias, M.; del Oinao, C.; Sanchez, F. Optically Active Complexes of Transition Metals (Rh, Co, Ni) with 2-(aminocarbonyl)pyrrolidine Ligands. Selective Catalysts for Hydrogenation of Prochiral Olefines. J. Organomet. Chem. 1992, 431, 233246. Corma, A.; Navarro, M. T.; Pariente, J. P. Synthesis of an Ultralarge Pore Titanium Silicate Isomorphous to MCM-41 and its Application as a Catalyst for Selective Oxidation of Hydrocarbons. J. Chem. Soc., Chem. Commun. 1994a, 147-148. Corma, A.; Fornes, V.; Navarro, M. T.; Perez-pariente, J. Acidity and Stability of MCM-41 Crystalline Aluminosilicates. J. Catal. 1994b, 148, 569-574. Corma, A.; Martinez, A. Zeolites and Zeotypes as Catalysts. Adv. Mater. 1995a, 137-144. Corma, A.; Martnez, A.; Martnez-Soria, V.; Monton, J. B. Hydrocracking of Vacuum Gasoil on the Novel Mesoporous MCM-41 Aluminosilicates Catalysts. J. Catal. 1995b, 153, 2531. Dag, O. K. A.; Ozin, G. A. Nanostructures: New Forms of Luminescent Silicon. Adv. Mater. 1995, 7, 72-77. Dartt, C. B.; Davis, M. E. Catalysis for Environmentally Benign Processing. Ind. Eng. Chem. Res. 1994, 33, 2887-2899. Davis, M. E. Organizing for Better Synthesis. Nature 1993, 364, 391-393. Davis, M. E.; Lobo, R. F. Zeolite and Molecular Sieve Synthesis. Chem. Mater. 1992, 4, 756-768. Davis, M. E.; Saldarriaga, C.; Montes, C.; Garces, J.; Crowder, C. A Molecular Sieve with Eighteen-Membered Rings. Nature 1988, 331, 698-702. Dessau, R. M.; Schlenker, J. L.; Higgins, J. B. Framework Topology of AlPO4-8: The First 14-Ring Molecular Sieve. Zeolites 1990, 10, 522-527.

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996 2089


Kloetstra, R. K.; van Bekkum, H. Catalysis of the Tetrahydropyranylation of Alcohols and Phenols by the H-MCM-41 Mesoporous Molecular Sieve. J. Chem. Res. (Symp.) 1995a, 2627. Kloetstra, R. K.; van Bekkum, H. Base and Acid Catalysis by the Alkali-Containing MCM-41 Mesoporous Molecular Sieves. J. Chem. Soc., Chem. Commun. 1995b, 1005-1006. Kozhevnikov, I. V.; Sinnema, A.; Jansen, R. J. J.; Pamin, K.; van Bekkum, H. New Acid Catalyst Comprising Heteropoly Acid on a Mesoporous Molecular Sieve MCM-41. Catal. Lett. 1995, 30, 241-252. Kresge, C. T.; Leonowicz, M. E.; Roth, W. J.; Vartuli, J. C.; Beck, J. S. Ordered Mesoporous Molecular Sieves Synthesisd by a Liquid-Crystal Template Mechanism. Nature 1992, 359, 710712. Lampert, C. M.; Granquist, C. G. Larger-Area Chromogenics: Materials and Devices for Transmittance Control; SAIE Optical Engineering Press: Washington, DC, 1990. Lawrence, M. J. Surfactant Systems: Their Use in Drug Delivery. Chem. Soc. Rev. 1994, 417-424. Liu, X.; Lu, K. K.; Thomas, J. K. Preparation, Characterization and Photoreactivity of Titanium (IV) Oxide Encapsulated in Zeolites. J. Chem. Soc., Faraday Trans. 1993, 89, 1861-1865. Llewellyn, P. C.; Schuth, F.; Grillet, Y.; Rouqoerol, F.; Rouquerol, J.; Unger, K. K. Water Sorption on Mesoporous Aluminosilicate MCM-41. Langmuir 1995, 11, 574-577. Lok, B. M.; Messina, C. A.; Lyle Patton, R.; Gajek, R. T.; Cannan, T. R.; Flanigen, E. M. Silicoaluminophosphate Molecular Sieves: Another New Class of Microporous Crystalline Inorganic Solids. J. Am. Chem. Soc. 1984, 106, 6092-6093. Luan, Z.; Cheng, C.-F.; Zhou, W.; Klinowski, J. Mesoporous Molecular Sieve MCM-41 Containing Framework Aluminium. J. Phys. Chem. 1995, 99, 1018-1024. Mallouk, T. E.; Lee, H. Designer Solids and Surfaces. J. Chem. Educ. 1990, 67, 829-843. Mann, S. Molecular Tectonics in Biomineralization and Biomimetic Materials Chemistry. Nature 1993, 365, 499-505. Marlow, F.; Hoffmann, K.; Hill, W.; Kornatowski, J.; Caro, J. Optical Properties of Self-Assembly Dipole Chains in Zeolites. Stud. Surf. Sci. Catal. 1994, 84, 2277-2284. McBain, J. W. The Sorption of Gases and Vapors by Solids; Routledge and Sons: London, 1932. McCormick, A. V.; Bell, A. T. The Solution Chemistry of Zeolite Precursors. Catal. Rev.sSci. Eng. 1989, 31, 97-127. Meier, W. M.; Olson, D. H. Atlas of Zeolite Structure Types; Butterworth-Heinemann: London, 1987. Mitchell, P. C. H. Zeolite-Encapsulated Metal Complexes: Biomimetic Catalysts. Chem. Ind. 1991, 308-310. Monnier, A.; Schuth, F.; Huo, Q.; Kumar, D.; Margolese, D.; Maxwell, R. S.; Stucky, G. D.; Krishnamurthy, M.; Petroff, P.; Firouzi, A.; Janicke, M.; Chmelka, B. F. Cooperative Formation of Inorganic-Organic Interfaces in the Synthesis of Silicate Mesostructures. Science 1993, 261, 1299-1303. Mortier, W. J.; Schoonheydt, R. A. Surface and Solid State Chemistry of Zeolites. Prog. Solid State. Chem. 1995, 16, 1-125. Muller, D.; Hoebbel, D.; Gessner, W. 27Al NMR Studies of Aluminosilicate Solutions. Influence of the Second Cordination Sphere on the Shielding of Aluminium. Chem. Phys. Lett. 1981, 84, 2529. Myers, D. Surfactant Science and Technology; VCH: New York, 1992. Notari, B. Synthesis and Catalytic Properties of Titanium Containing Zeolites. Stud. Surf. Sci. Catal. 1988, 37, 413-425. Olson, D. H. A Reinvestigation of the Crystal Structure of the Zeolite Hydrated NaX. J. Phys. Chem. 1970, 74, 2758-2764. Ozin, G. A.; Gil, C. Intrazeolite Organometallics and Coordination Complexes: Internal versus External Confinement of Metal Guests. Chem. Rev. 1989, 1749-1764. Ozin, G. A. Nanochemistry: Synthesis in Diminishing Dimensions. Adv. Mater. 1992, 10, 612-649. Pelrine, B. P.; Schmitt, K. D.; Vartuli, J. C.; Production of Olefin Oligomer. U.S. Patent, 5,105,051, 1992. Rathousky, J.; Zukai, A.; Franke, O.; Schulz-Ekloff, G. Adsorption on MCM-41 Mesoporous Molecular Sieve. Part 1.-Nitrogen Isotherms and Parameters of the Porous Structure. J. Chem. Soc., Faraday Trans. 1994, 90, 2821-2826. Rathousky, J.; Zukai, A.; Franke, O.; Schulz-Ekloff, G. Adsorption on MCM-41 Mesoporous Molecular Sieve. Part 2.-Cyclopentane Isotherms and their Temperature Dependence. J. Chem. Soc., Faraday Trans. 1995, 91, 937-940. Reddy, R. M.; Moudrakovski, I.; Sayari, A. Synthesis of Mesoporous Vanadium Silicate Molecular Sieves. J. Chem. Soc., Chem. Commun. 1994, 1059-1060. Ryoo, R.; Kim, J. M. Structural Order in MCM-41 Controlled by Shifting Silicate Polymerization Equilibrium. J. Chem. Soc., Chem. Commun. 1995, 711-712. Sayari, A.; Danumah, C.; Moudrakovski, I. L. Synthesis, Characterisation and Modification of MCM-41 Molecular Sieves. Mater. Res. Soc. Symp. Proc. 1995a, 371, 81-92. Sayari, A.; Danumah, C.; Moudrakovski, I. L. Boron-Modified MCM-41 Mesoporous Molecular Sieves. Chem. Mater. 1995b, 7, 813-815. Schlenker, J. L.; Rohrbaugh, W. J.; Chu, P.; Valyocsik, E. W.; Kokotailo, G. T. The Framework Topology of ZSM-48: A High Silica Zeolite. Zeolites 1985, 5, 355-356. Schmidt, R.; Akporiaye, D.; Stocker, M.; Ellestad, O. H. Synthesis of Mesoporous MCM-41 Materials with High Levels of Tetrahedral Aluminium. J. Chem. Soc., Chem. Commun. 1995a, 1493-1494. Schmidt, R.; Hansen, E. W.; Stocker, M.; Akporiaye, D.; Ellestad, O. H. Pore Size Determination of MCM-41 Mesoporous Materials by Means of 1H NMR Spectroscopy, N2 Adsorption, and HREM. A Preliminary Study. J. Am. Chem. Soc. 1995b, 117, 4049-4056. Smith, J. V. Topochemistry of Zeolite and Related Materials. 1. Topology and Geometry. Chem. Rev. 1988, 88, 149-182. Steel, A.; Carr, S. W.; Anderson, M. W. 14N NMR Study of Surfactant Mesophases in the Synthesis of Mesoporous Silicates. J. Chem. Soc., Chem. Commun. 1994, 1571-1572. Stucky, G. D.; MacDougall, J. E. Quantum Confinement and Host/ Guest Chemistry: Probing a New Dimension. Science 1990, 247, 669-678. Swaddle, T. W.; Salerno, J.; Tregloan, P. A. Aqueous Aluminates, Silicates, and Aluminosilicates. Chem. Soc. Rev. 1995, 319325. Tabata, T.; Kokisu, M.; Okada, O. Study on Patent Literature of Catalysts for a New NOx Removal Process. Catal. Today 1994, 22, 147-169. Tanev, P. T.; Pinnavaia, T. J. A Neutral Templating Route to Mesoporous Molecular Sieves. Science 1995, 267, 865-867. Tanev, P. T.; Chibwe, M.; Pinnavaia, T. J. Titanium-Containing Mesoporous Molecular Sieves for Catalytic Oxidation of Aromatic Compounds. Nature 1994, 368, 321-323. Tanguay, J. F.; Suib, S. L.; Conhlin, R. W. Dichloromethane Photodegradation Using Titanium Catalysts. J. Catal. 1989, 117, 335-347. Thomas, J. M. The Chemistry of Crystalline Sponges. Nature 1994, 368, 289-290. Treacy, M. M. J.; Newsaw, J. M. Two New Three-Dimensional Twelve Ring Zeolite Frameworks of Which Zeolite is a Disordered Intergrowth. Nature 1988, 332, 249-251. van Koningsveid, H.; Jansen, J. C.; van Bekkum, H. The Monoclinic Framework Structure of Zeolite H-ZSM-5 Comparison with the Orthorhombic Framework of as-Synthesised ZSM-5. Zeolites 1990, 10, 235-241. Vartuli, J. C.; Schmitt, K. D.; Kresge, C. T.; Roth, W. J.; Leonowicz, M. E.; McCullen, S. B.; Hellring, S. D.; Beck, J. S.; Schlenker, J. L.; Olson, D. H.; Sheppard, E. W. Effects of Surfactant/Silica Molar Ratios on the Formation of Mesoporous Molecular Sieves: Inorganic Mimicry of Surfactant Liquid-Crystal Phases and Mechanistic Implications. Chem. Mater. 1994, 6, 23172326. Walker, S. A.; Zasadzinski, J. A. Self-Assembly of Silicate/ Surfactant Mesoporous Materials. Mater. Res. Symp. Proc. 1995, 371, 93-98. Wilson, S. T.; Lok, B. M.; Messina, C. A.; Cannan, T. R.; Flanigen, E. M. Aluminophosphate Molecular Sieves: A New Class of Microporous Crystalline Inorganic Solids. J. Am. Chem. Soc. 1982, 104, 1146-1147. Winsor, P. A. Binary and Multicomponent Solution of Amphiphilic Compounds. Solubilization and the Formation, Structure, and Theoretical Significance of Liquid Crystalline Solutions. Chem. Rev. 1968, 68, 1-40. Wu, C. G.; Bein, T. Conducting Carbon Wires in Ordered, Nanometer-Sized Channels. Science 1994a, 266, 1013-1014. Wu, C. G.; Bein, T. Conducting Polyaniline Filaments in a Mesoporous Channel Host. Science 1994b, 264, 1757-1758. Wu, C. G.; Bein, T. Polyaniline Wires in Oxidant-Containing Mesoporous Channel Hosts. Chem. Mater. 1994c, 6, 1109-1112.

2090

Ind. Eng. Chem. Res., Vol. 35, No. 7, 1996


Zhao, X. S.; Lu, G. Q.; Millar, G. J.; Li, X. S. Synthesis of Mesoporous Molecular Sieves MCM-41 from an Alkali-Free System and its Catalytic Activity. Catal. Lett. 1995a, in press. Zhao, X. S.; Lu, G. Q.; Millar, G. J. Encapsulation of Transition Metal Species into Zeolites and Molecular Sieves as Redox Catalysts. Part 2. J. Porous Mater. 1995b, in preparation.

Yan, J. Y.; Liu, S. U.; Chen, T. H.; Wang, J. Z.; Li, H. X. Synthesis of Iron-Containing MCM-41. J. Chem. Soc., Chem. Commun. 1995, 973-974. Yanagisawa, T.; Schimizu, T.; Kiroda, K.; Kato, C. The Preparation of Alkyltrimethylammonium-Kanemite Complexes and their Conversion to Mesoporous Materials. Bull. Chem. Soc. Jpn. 1990, 63, 988-992. Zhao, D. Y.; Goldfarb, D. Synthesis of Manganosilicates: MnMCM-41, Mn-MCM-48 and Mn-MCM-L. J. Chem. Soc., Chem. Commun. 1995, 875-876. Zhao, X. S.; Wang, Q. X.; Xu, L. Y.; Li, X. S. Direct Synthesis of Mesoporous Molecular Sieve MCM-41. Unpublished results, 1994. Zhao, X. S.; Lu, G. Q. Zeolite-Based Catalysts for NOx Decomposition. Proc. Aust. Combust. Symp. 1995, C1, 1-8.

Received for review November 21, 1995 Revised manuscript received March 25, 1996 Accepted March 27, 1996X IE950702A
X Abstract published in Advance ACS Abstracts, June 1, 1996.

Anda mungkin juga menyukai