You are on page 1of 38

Chapter 1

Introduction
1.1 General introduction
Porous materials have been widely used as catalysts, adsorbents, molecular sieves,
ion exchangers and chromatographic agents in fine chemicals industry because of their
particular properties such as uniform pore size, large surface area and flexible
frameworks. Following the definition accepted by the IUPAC [1], porous materials can
be grouped into three classes based on their pore diameter (): microporous, < 2.0 nm;
mesoporous, 2.0 < < 50 nm; and macroporous, > 50 nm (Fig. 1-1). Among the
porous solids, microporous zeolites and mesoporous molecular sieves are mostly used as
acid catalysts, especially in the petrochemical industry. They possess several catalytically
desirable properties including high surface area, adjustable pore size, acidity and high
thermal stability [2-4].
Zeolites are crystalline porous solids containing pores and cavities of molecular
dimensions. Their structures reveal regular arrays of channels and cavities in the
molecular size range of 0.3 to 1.5 nm. In their acid form, zeolites are probably the most
important heterogeneous acid catalysts used in industry and are particularly used in the
fields of oil refining and petrochemistry as fluid cracking catalysts. Their key properties
are size and shape selectivity, together with the potential for strong acidity [5]. Although
the micropores of zeolite have been well described as having excellent potential for
chemical functions, diffusion in the micropores often interferes in chemical process [6].

This limits the application of these pore systems to reactant molecules having kinetic
diameters below 10 .
Microporous
Domain

Mesoporous
Domain

Macroporous
Domain
Porous Glasses

Porous Gels

M41S

Mesoporous Solids (M41S), and other


materials with controlled porosity, obtained
by self-assembly (polymers) or colloidal
(latex bends) templating

Pillared layered solids (ex: clays, LDH)

Zeolites and
related materials

0,5

5
10
Pore Diameter (nm)

50

100

Fig. 1-1 Examples of micro-, meso-, and macroporous materials, showing pore size
domains and typical pore size distributions [1].

In recent years, mesoporous molecular sieves with well-defined pore sizes of 2 to


50 nm have received much attention. In comparison to microporous zeolites, ordered
mesoporous materials with larger pore dimensions overcome the pore size constraint of
zeolites and allow the diffusion of larger molecules which are highly desirable in certain
processes such as fluid catalytic cracking and chemical conversions in condensed media.
2

However, due to the amorphous nature of the pore walls, mesoporous materials generally
exhibit weaker acidity and poorer hydrothermal stability than conventional zeolites.
Therefore, a lot of efforts have been made to improve properties of mesoporous materials.
Among them, the synthesis of composite materials with combined micro- and
mesoporousity has become the challenging target. This kind of material is expected to
offer many significant advantages over single micro- or mesoporous molecular sieves
including improved diffusion rate, high acidity and hydrothermal stability.

1.2 Microporous materials zeolites


Zeolites are three-dimensional, microporous, crystalline solids with well-defined
structures that contain aluminum, silicon, and oxygen in their regular framework. The
elementary building units of zeolites are TO4 tetrahedra (SiO4 and AlO4-). Adjacent
tetrahedral are linked at their corners via a common oxygen atom. A network of SiO4
tetrahedral is neutral while each AlO4 tetrahedron in the framework bears a negative
charge. Thus, zeolite frameworks are typically anionic, and charge compensating cations
(Na+, K+, or NH4+) populate the pores to maintain electrical neutrality [7]. The
framework of a zeolite contains channels, channel intersections and/or cages with
dimensions from ca. 0.3 to 1.5 nm, which coincides with the range of dimensions of a
very large number of molecules. Most of the zeolites can be classified into three
categories (Fig. 1-2): small pore zeolites with eight membered-ring pore apertures having
free diameters of 3.0-4.5 (e.g. zeolite A); medium pore zeolites with ten memberedring apertures, 4.5-6.0 in free diameter (e.g. zeoite ZSM-5) and large pore zeolites with
12 membered-ring apertures, 6.0-8.0 (e.g. zeolite Beta and zeolite Y).

8-Ring, zeolite A

12-Ring, zeolite Beta

10-Ring, zeolite ZSM-5 (MFI)

12-Ring, zeolite Y

Fig. 1-2 Representative structures of zeolites [7].

Zeolites have been the most important solid acid catalysts in the past few decades. A
landmark was the introduction of synthetic faujasites (zeolites X and Y) on an industrial
scale in fluid catalytic cracking (FCC) of petroleum to produce the majority of the
worlds gasoline - one of the most important chemical processes worldwide. ZSM-5 is
another example being of utmost importance in heterogeneous catalysis. It is used

industrially in the synthesis of ethylbenzene, the isomerization of xylenes and the


disproportionation of toluene. Nevertheless, compared to the successful use of zeolites in
hydrocarbon processing, the application of zeolite catalysts in the manufacture of fine
chemicals is still limited. This can be attributed to their small pores which are generally
too narrow to admit the molecules involved in catalytic reactions. A number of large pore
zeolites have been synthesized to date, such as the super-large-pore materials with pores
formed by more than 12 TO4- tetrahedral (e.g. zeolite UTD-1, 14-membered-ring) [8].
Amongst the large-pore zeolites, the three-dimensional pore structure containing 12membered-rings in zeolite beta and zeolite Y make them highly suitable catalyst in
organic reactions.

1.2.1 Structure of zeolite Y and Beta


Zeolite Y exhibits the FAU (faujasite) structure (Fig. 1-3) [9]. Its pore system is
relatively spacious with a pore diameter of 7.4 since the aperture is defined by 12 TO4tetrahedra. It consists of spherical cages, referred to as supercages, with a diameter of
13 . Therefore, zeolite Y is classified to possess a three-dimensional, 12-membered-ring
pore system [7].

Fig. 1-3 Structure of zeolite Y (FAU) [9]. 12-ring viewed along [111]; supercages : 13
.

Fig. 1-4 Projections along principal crystallographic directions of zeolite beta structures:
polymorph A (left) and Polymorph B. The unit cell outlines are indicated by the dashed
lines [10].

The structure of zeolite beta was only recently determined because the structure is
very complex. Zeolite beta consists of an intergrowth of two distinct but closely related
structures termed Polymorphs A and B (Fig. 1-4) which are normally found in a 60:40
ratio, respectively. Both polymorphs have three-dimensional 12-ring pore systems. The
intergrowth of the polymorphs does not significantly affect the pores in two of the
dimensions, where two mutually perpendicular straight channels with a cross section of
7.6 x 6.4 run in the a- and b-directions. However, the pore becomes tortuous in the
direction of the faulting, where a sinusoidal channel of 5.5 x 5.5 runs parallel to the cdirection (Fig. 1-5). The two polymorphs are depicted in references [9-13]. It has been
proposed that a third polymorph C should also exist which has a 3-dimensional pore
topology with all 12-membered-ring channels being linear. Indeed, pure polymorph C of
zeolite beta was synthesized by using framework isomorphous substitution (Si replaced
partially by Ge) as a structure-directing mechanism [14].

Viewed along [100]

(a)

(b)

Fig. 1-5 Structure of zeolite beta. (a) 12-ring viewed along [100] and (b) 12-ring viewed
along [001] [9].

1.2.2 Synthesis of zeolites


Zeolites are normally synthesized hydrothermally from basic reaction gels at
temperatures between 60 C and 200 C under an autogeneous pressure and in the
presence of a large excess of water. Zeolite Y was first synthesized in the sodium form in
1964 by Union Carbide [15]. Typically, it is synthesized in a gelling process under
hydrothermal conditions. Sources of alumina (sodium aluminate) and silica (sodium
silicate) are mixed in alkaline media to give a gel. The gel is then usually heated to 70120 C to crystallize the zeolite. The zeolite is present in Na+ form and must be converted
to acid form. To prevent disintegration of the structure from acid attack, it is first
converted to the NH4+ form by ion-exchange with NH4NO3 solution. Zeolite Y is not a
particularly siliceous zeolite with Si/Al ratio in the range of 1 to 25. This high aluminum
content allows zeolite Y to be synthesized easily without using any organic structure
directing agents.
Beta zeolite was first synthesized in the basic medium by Mobil Oil Corporation in
1967 [16]. In comparison to the zeolite Y, the Si/Al ratio of zeolite beta can range from
approximately 10 to and synthesis requires the addition of an organic compound, such
as the tetraethylammonium species, as a structure directing agent. These compounds are
often referred to as templates as the zeolite structure appears to form around them.
Besides the basic medium synthesis, an alternative route is using fluoride as the
mineralizing agent, in which case the pH can be much lower. Caullet et al. [17]
succeeded in preparing zeolite Beta from near neutral aqueous aluminosilicate gels in the
presence of fluoride as mineralizing media and 1,4-diazabicyclo [2,2,2] octane (DABCO)
and methylamine as the organic species. The fluoride approach has the advantages that

(i) nucleation rates are reduced so that larger crystals are formed and (ii) the acid or
neutral pH regime facilitates the structural incorporation of those heteroatoms where the
precursor species would be precipitated at higher pH [18]. In recent years, this approach
comes to great prominence, particularly in the synthesis of large crystals, novel structures
and hetero-substituted materials. For example, incorporation of Ti into the framework of
Al-free zeolite beta in F- medium was achieved by Corma et al. [19]. It was found that
upon contact with ambient humidity, there is no hydrolysis of Si-O-Ti bonds in Ti-Beta
zeolites prepared by the fluoride route, while it is probably a major feature of those
synthesized in OH- medium. The more hydrophobic character of Ti-beta-F (synthesized
in the fluoride media) gives it attractive properties in oxidation reactions. Furthermore,
Al-free Sn-beta zeolite was successfully synthesized in fluoride medium [20] and it is
shown to be a very efficient catalyst for Baeyer-Villiger oxidation. Recently, Al-free Zrbeta zeolite with Si/Zr up to 75 was synthesized in fluoride medium by our group [21].
The incorporation of zirconium into zeolite beta induced the formation of increased
amounts of polymorph B. Moreover, Zr-beta was found to be excellent catalyst in the
MPV reduction of alkyl- and aryl-substituted cyclohexanones, with high selectivity to the
corresponding alcohols.

1.2.3 Acidity of zeolites


Among the most important properties of zeolites with respect to their use as catalysts
is their surface acidity. Both Brnsted and Lewis acid sites occur in zeolites. This is
important as most of the reactions catalyzed by zeolites involve proton transfer, or at least
activation by a Brnsted or Lewis acid.

10

Nature of the acid sites

H
O

Si
O

O
O O

Si

O
O O

Al

Si

O O

O
O O

Si

O
O

Fig. 1-6 Brnsted acid site formation due to Si/Al substitution in a zeolite framework

Brnsted acid sites are formed by the bridging hydroxyl protons which compensate
the excess negative charges generated by isomorphous substitution of silicon with
trivalent metal cations, Al3+ for instance, into the framework. In zeolites, the bridging
hydroxyl groups are attached to framework oxygen linking tetrahedral Si and Al atoms
(Fig. 1-6). Obviously, the number of protonic sites in a zeolite is related to the number of
framework aluminum atoms. While aluminum incorporation in the framework results in
Brnsted acidity, extra framework aluminum (or more accurately aluminum
oxyhydroxides) in the pore system results in Lewis acidity [22]. Extra framework
aluminum species created by mild steaming or calcination of hydrated zeolites were
shown to increase the catalytic activity of zeolites. This increase in activity was attributed
to the creation of sites exhibiting enhanced acidity through interaction of bridging
hydroxyl groups (Brnsted acid sites) and neighboring small extra framework aluminum
species (Lewis acid sites) [23]. However, the exact nature is still a matter of debate.

11

Strength of acid sites


The acid strength of a given acid site in zeolite depends on several factors including
the T-O-T bond angles, the type of T-atom, and Si/T ratio. It was found that a relation
exists between the T-O-T bond angles and the acid strength of the associated proton in
the zeolites: the greater the angle, the stronger the sites [24]. Thus, the protonic sites of
H-MOR (bond angle range between 143-180 ) and H-MFI (133-177 ) are stronger than
those of H-FAU (138-147 ). This explains why H-MOR is while H-FAU is not active in
n-butane and n-hexane isomerization reactions which require very strong acid sites. The
type of T-atom affects the acid strength. From IR spectroscopy (wavenumber of the OH
groups) and ammonia TPD over MFI samples, the acid strength increased in the
following order: B(OH)Si< In(OH)Si << Fe(OH)Si < Ga(OH)Si < Al(OH)Si [25]. In
addition to the T-O-T band angles and type of T-atom, the acid strength of the proton at
the aluminum site is also strongly dependent on framework Si/Al ratio. Theoretical and
experimental studies of the effect of framework Si/Al ratio on the acidity led to the
conclusion that the strength of the protonic sites of zeolites is influenced by the presence
of neighbors [22]. It is well known and supported by quantum chemical calculations that,
due to the higher electronegativity of silicon compared to aluminum, the strongest
Brnsted acid sites in zeolites will occur on completely isolated AlO4- tetrahedral, i.e.,
those which lack AlO4- tetrahedral as next nearest neighbors [8]. This is the reason why
the acid strength of zeolite HY increases with dealumination up to Si/Al ratio of 10. In
this region, the gain in acid strength overcompensates the decrease in the density of
Brnsted acid sites.

12

Location of the acid sites


The accessibility of the protonic sites plays a significant role in the catalytic activity
of zeolites. More and more examples for distribution of the location of acid sites are
emerging in the literature. One example is the acid form of zeolite FAU where the portion
of protonic sites located in the supercages is accessible to many organic molecules
whereas the others, located in the hexagonal prisms, are inaccessible to many potential
reactant organic molecules [22].

1.2.4 Zeolties in catalysis


Zeolites have the ability to act as catalysts for chemical reactions which take place
within the internal cavities. An important class of reactions is that catalyzed by hydrogenexchanged zeolites, whose framework-bound protons give rise to very high acidity. This
is exploited in many organic reactions, including crude oil cracking, isomerisation and
fuel synthesis. Zeolites can also serve as oxidation or reduction catalysts, often after
metals have been introduced into the framework. Essentially, zeolites have two properties
which make them particularly suitable as catalysts, viz. (i) zeolites are crystalline porous
materials with pore dimensions in the same order as the dimensions of simple molecules.
Thus the shape and size of the particular pore system can exert a steric influence on the
reaction, controlling the access of reactants and products; (ii) they are cation exchangers,
hence it is possible to introduce a large variety of cations with different catalytic
properties into their intracrystalline pore system, which in turn offers the opportunity to
create different catalytic properties, e.g., in acid- or metal-catalyzed reactions [26].

13

1.2.4.1 Shape-selective catalysis in zeolites


Molecular shape-selective, or simply shape-selectivity, is the reaction specificity
arising from the presence of a sterically confined environment in which the molecules are
converted [27]. Zeolites can be shape-selective catalysts on the basis of a subtle matching
of size and shape of micropores of the zeolite with the size and shape of reactants,
transition states and products in the reaction system.

Fig. 1-7 Schematic diagram of shape selectivity in zeolites [28].

14

Reactant and product shape selectivity


The simplest forms of shape selectivity come from the impossibility of certain
molecules in a reactant mixture entering the zeolite pores - reactant shape selectivity or of
certain product molecules which forms inside the pore network exiting from these pores
product shape selectivity. An example of reactant shape selectivity is in the
transformation of n-alkanes of light gasoline into propane and n-butane. Being linear, the
n-alkanes can enter the pores of the zeolite catalysts and be transformed, whereas the
branched alkanes are excluded from the pores and do not react (Fig. 1-7) [28]. Product
shape selectivity plays a key role in the process for the selective synthesis of paradialkylbenzenes in toluene alkylation by methanol (Fig. 1-7). During the reaction, the
bulkier ortho- and meta-isomers are also formed in the pores, but their exit is hampered
due to their lower diffusion coefficient relative to that of para. The selective removal of
para-isomer allows the bulkier isomers to be isomerized into the less bulky species [29].
An example of reverse shape selectivity was proposed by Venuto et al. [30] for coke
formation in a Y zeolite at a high temperatures - the bulky molecules trapped in the
supercages (13 ) cannot escape, the pore apertures (7 ) being too narrow and the coke
molecules limits or inhibits the access of reactant molecules to the active sites leading to
a deactivation of the zeolite catalyst.

Transition state shape selectivity


Transition state shape selectivity occurs when the formation of reaction intermediates
(and/or transition states) is sterically limited by the space available near the active sites.
A typical example of this type of selectivity was first proposed by Csicsery [31] for the
15

transformation of dialkyl-benzenes over H-mordenite. Although 1,3,5-trialkylbenzene


can diffuse in the zeolite channels, it is not found amongst the products. This can be
attributed to the insufficient space within the zeolite channels for the diphenylmethane
intermediate which is involved in the formation of 1,3,5-trialkylbenzenes (Fig. 1-7).
Another example of transition state shape selectivity is provided by stereoselective
MPV reduction of 4-tert-butyl-cyclohexanone to cis-4-tert-butyl-cyclohexanol (>95%)
over zeolite beta which was first reported by Creyghton et al. [32, 33]. Other active solid
catalysts, including zeolites, gave predominantly the thermodynamically stable transisomer. The observed high selectivity to the thermodynamically unfavorable cis-isomer
was explained by a restricted transition-state for the formation of the trans-isomer within
the pores of the zeolite beta (Fig. 1-8). In an extension of this work, van der Waal et al.
[34] reported on the application of Al-free Ti-beta in the same reaction. Again, a very
high selectivity of 98% to the cis-isomer was achieved, indicating similar steric
constraints, in this case, around a Lewis-acid titanium site. More recently, the MPV
reduction was performed over Al-free Sn-beta [20] and Al-free Zr-beta [21], which were
both synthesized in fluoride medium, and the high activity and selectivity was attributed
to the introduction of adequate Lewis acidity by Sn or Zr and the pore size constraints.

16

Fig. 1-8 Transition states for the formation of cis-4-tert-butyl-cyclohexanol (left) and
trans-4-tert-butyl-cyclohexanol (right) in the proposed reaction mechanism over zeolite
beta. Illustration from Ref. [32].

Other types of shape selectivity


In most cases, the observed selectivity effects can be rationalized in terms of the
traditional classification outlined above. However, various other types of shape
selectivity have been proposed to account for more recent research results. These include
shape selectivity related to molecular concentration in zeolite micropores [35]; the cage
or window effect [36]; molecular traffic control [37]; shape selectivity at the external
surface-the nest effect [38]; tip-on adsorption of molecules diffusing inside the pore
system [39] and secondary shape selectivity/inverse shape selectivity [40], etc. Some of
these concepts have, in the meantime, been disproved, while others continue to be a
matter of debate.

17

1.2.4.2 Zeolite-supported metal catalysts


Zeolites can be used as a support material because of their large surface area and
cage-like pore system. The metal can be prepared as very small particles with the same
dimensions of the zeolite channel or cage. After the metal has been introduced into the
micopores, the zeolites can act as oxidation or reduction catalysts. Typically, metal
particles can be introduced into zeolite by ion-exchange of amine complexes with the
zeolite or impregnation with a solution using the so-called incipient wetness technique,
followed by calcination and reduction. The metal must have an even distribution over the
available zeolite surface. Careful reduction or oxidation/reduction of the metal cations is
necessary to give finely dispersed, nano-sized metal particles [25].
Two types of zeolite-supported metal catalysts may be distinguished. The first is
represented by monofunctional catalysts, consisting of a metal supported on a neutral
catalytically inactive zeolite. One example is the selective aromatization of n-alkanes (e.g.
n-hexane into benzene) proceeding through Pt-LTL catalyst which presents no protonic
sites [22]. The second type comprises bifunctional catalysts, consisting of an acidic
function and a metallic hydrogenation/dehydrogenation function. Bifunctional catalytic
reactions involve a series of catalytic steps over acidic and hydrogenationdehydrogenation sites with formation of intermediate compounds. For example, n-hexane
isomerization involves the following steps - n-hexane dehydrogenation to n-hexenes
(metal catalyzed), skeletal isomerization of n-hexenes into isohexenes over protonic acid
sites, followed by the hydrogenation of isohexenes into isohexanes (metal catalyzed) [22].
The activity, stability and selectivity of bifunctional zeolite catalysts depend mainly on

18

two parameters: the balance between hydrogenating and acid functions [41] and the
zeolite pore structure [42].
Apart from its functions as a metal support and an acidic catalyst, the zeolite
component may also induce shape selectivity, i.e., employing a zeolite as carrier for a
metal enables the shape-selective properties of the microporous structure to be conferred
on the catalytic reactions [43]. Dessau [44] studied the competitive hydrogenation of
branched 4,4-dimethyl-1-hexene and linear 1-hexene over Pt/Cs-ZSM-5. He showed that
over platinum clusters located inside the pores of zeolite ZSM-5, the linear olefins are
preferentially reduced over the branched olefins. However, when the metal was located in
the external surface, no shape selectivity was observed. Therefore, this finding was also
exploited as a method to assess the location of platinum in Pt/Cs-ZSM-5. Moreover, it
was found that the shape selectivity could be obtained only by reduction of the catalyst in
the presence of olefins and hydrogen but not for catalysts reduced in pure hydrogen. This
was explained by metal migration from the interior of the zeolite to the external surface
during the reduction process in hydrogen while the olefins reacted with the platinum
hydride species, thus preventing their migration and agglomeration. Another example for
shape-selective, metal-catalyzed reaction is the selective oxidization of para-xylene in
the presence of the bulkier ortho-xylene over copper-loaded ZSM-5 [45]. Creyghton et al.
[46] reported that 1-decene reacted 18 times faster than trans-5-decene over Pt/Na-beta,
whereas the ratio was only about 2 for platinum on non-microporous supports. This
regioselectivity was attributed to the steric constraints imposed by the microporous
structure of the zeolite. Moreover, regioselectivity was only obtained when the nonselective, external-surface located platinum clusters were selectively poisoned with

19

triphenylphosphine. Furthermore, it was shown that catalytic activity and regioselectivity


were only obtained when non-polar or bulky solvents were applied. This can be explained
by the presence of strong competition between the solvent and the reactant adsorption in
the zeolite where the polar solvent prevented the reactants from attaining a significant
concentration in the pores. All these examples clearly demonstrate the potential of metalloaded zeolites as highly selective hydrogenation catalysts.
As reviewed above, zeolites are important solid acid catalysts but their use has been
limited to molecules that are small enough to enter the zeolite pores. Therefore, when the
M41S family of ordered mesoporous molecular sieves was first reported a decade ago,
considerable interest was generated due to their potential use as catalysts for the
conversion of large molecules.

1.3 Microporous/mesoporous materials


1.3.1 Ordered mesoporous materials
Mesoporous materials can be generally defined as a class of porous materials with
pore diameters in the range of 2-50 nanometers and composition includes silicates and
aluminosilicates. The first successful attempt to synthesize mesoporous materials by
using a surfactant as a template was made by Yanagisawa et al. [47] in 1990. The
mesoporous sieves with pore diameter ranging from 1.8 to 3.2 nm were prepared by
heating

layered

polysilicate

kaenemite

in

an

aqueous

solution

of

alkyltrimethylammonium with different alkyl-chain lengths. The introduction of


supramolecular assemblies (micellar aggregates, rather than molecular species) as

20

templating agents was first developed by a research group at Mobil Oil in 1992 [48]. This
led to a new family of mesoporous silica and aluminosilicate compounds (M41S). These
solid phases are characterized by ordered mesopores presenting narrow pore size
dispersions. The M41S family includes a bidimensional hexagonal phase (MCM-41, for
Mobil Composition of Matter), a cubic phase, MCM-48, and several lamellar phases.
MCM-41 is synthesized by using amphiphilic templating materials together with
inorganic precursors to self assemble into a hexagonal array of uniform pores. The most
outstanding feature in the preparation of MCM-41, 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 10 carbon atoms. Therefore, the
mechanisms responsible for the formation of MCM-41 have attracted much attention.
Several models for the formation of MCM-41 have been proposed [49], e.g., liquid
crystal templating (LCT) approach [48]; transformation mechanism from lamellar to
hexagonal phase [50-53]; and silicate-template interaction hybrid polyelectrolyte model
[54]. All these models are based on one principle: surfactant molecules play a central role
in directing the formation of the inorganic mesostructure from soluble mineral species
[55].
In comparison to microporous zeolites, ordered mesoporous materials with large pore
dimensions overcome the pore size constraint and allow the diffusion of larger molecules.
Moreover, mesoporous molecular sieves present very high surface areas with very regular
pore size dimensions which make them good candidates as support for catalytically active
phases. The introduction of transition metals in the walls gives them catalytic redox
properties which are useful in selective oxidation reactions. However, it was immediately

21

recognized that mesostructured materials have relatively low acidity and poor stability, in
contrast to requirements of many catalytic applications. The instability of these structures
can be attributed in part to the thinness and incomplete crosslinking of the pore walls.
Recent effort has been focused on improving both the acidity and the hydrothermal
stability of mesoporous materials.

1.3.2 Modification of mesoporous materials


1.3.2.1 Thick-walled mesostructures
By using small, ionic hydrophilic head groups and hydrophobic alkyl tails surfactant
molecules as template (such as C16TMABr), ordered arrays of mesoporous silica were
formed with pore diameters around 2.0 - 4.0 nm and wall thickness between 1 and 2 nm.
The poor hydrothermal stability of this mesostructure was quickly recognized and
attributed to the thin, amorphous wall structure. Thick-walled mesostructures were made
by changing the surfactant. In 1998, Stucky et al. [56] successfully synthesized a thickwalled mesoporous material denoted as SBA-15 by using triblock copolymer surfactant polyethylene oxide - polypropylene oxide - polyethylene oxide (PEO-PPO-PEO). SBA15 exhibited a wall thickness between 3-7 nm and large pore sizes of 7-12 nm. Soon
afterwards, Pinnavaia et al. [57] used neutral Gemini amine surfactants to make thickwalled vesicle-like lamellar framework (MSU-G) with improved hydrothermal stability.
The improved hydrothermal stability of these materials is attributable in large part to the
thicker framework walls.

22

1.3.2.2 Incorporation of Al into the framework


Modification by introducing tetrahedral AlO4 units into the electrically neutral
siliceous framework can generate acid sites in the material. It is well known that there are
two typical methods used for the incorporation of aluminum. One is direct-synthesis by
homogeneous self-assembly process of silica and aluminum species with the aid of
structure-directing agent; the other is post-synthesis via treatment (impregnation, grafting
or ion-exchange) of preformed mesoporous materials. The procedure of direct-synthesis
is relatively simple. Yue et al. [58] have synthesized Al-SBA-15 at the pH 1.5 and Yang
et al. [59] have prepared mesoporous aluminosilicates in non-aqueous media. However,
the efficiency of direct-synthesis is always low only a small part of the heteroatoms can
be introduced into the mesoporous products. On the other hand, post-synthesis has been
reported to be advantageous with respect to incorporation of more heteroatoms. For
example, Mokaya employed calcined small crystallite MCM-41 as seeds for the
secondary synthesis of the same mesostructure [60]. With post-synthesis grafting of Al
centers onto the framework walls of the secondary silica MCM-41, both the hydrothermal
stability and acidity were significantly improved [61]. It was concluded that the increased
acidity of hydrothermally treated samples was due to an increase in the Al/Si ratio in the
framework and the improved hydrothermal stability was attributed to the combination of
thicker pore walls and improved framework crosslinking upon Al grafting. However, one
disadvantage of post-synthesis is that the uniform mesostructures are sometimes severely
destroyed [62]. Recently, Wu et al. [63] reported a simple and effective pH-adjusting
method in which a large amount of Al could be introduced into the SBA-15 mesophase
and the Al atoms were located at mainly tetrahedrally coordinated sites.

23

1.3.3 Synthesis of microporous/mesoporous materials


Due to the amorphous nature of their frameworks, modified mesoporous materials
still exhibit low stability and acidity compared to zeolites. If micropores could be
incorporated into the mesoporous material, the resulting bimodal pore-structured material
would combine the benefits of both pore-size regimes. For example, micropores in
zeolites provide size- or shape selectivity for guest molecules, while mesopores provide
easier access to the active sites in micropores. This kind of material is expected to offer
many significant advantages over single micro- or mesoporous molecular sieves with
improved diffusion rate, high acidity and hydrothermal stability.
Therefore, synthesis of composite materials with combined micro- and mesoporosity
has gained more attention than any other approaches. One approach is to form the
mesoporous materials followed by transformation of the preassembled walls into zeolitic
structures by post treatment with a microporous zeolite structure director. Another
approach is to form zeolite seeds first followed by assembly of these seeds to form the
mesostructures.

1.3.3.1 Transformation of mesoporous walls into zeolitic structure


The first successful attempt to synthesize combined micro- and mesoporous
materials was made by van Bekkum and co-workers in 1997 [64]. They used the zeolite
structure-directing agent tetrapropylammonium cation to partially recrystallize the pore
walls of Al-MCM-41 mesostructures into ZSM-5 zeolitic structures. The treated product
exhibited a substantial increase in acidity which was attributed to the partial

24

transformation of the amorphous aluminosilicate wall of MCM-41 into an embryonic


crystalline zeolite phase. However, it was found that MCM-41 cannot maintain the
hexagonal mesostructure when its framework wall was transformed to a crystalline
zeolite phase. Once a zeolite phase is formed from MCM-41, it tends to appear as a
separated zeolite phase [65]. Following van Bekkums approach, Kaliaguine et al. [66,
67] investigated the possibility of transforming a thicker wall mesostructured cellular
foams (MCF) aluminosillicate into crystalline zeolitic framework. The acidity of this
material was determined by pyridine adsorption and the results showed that the acidic
strength was much stronger than the parent amorphous mesoporous material. Huang et al.
[68] also prepared mesoporous zeolitic materials by using a secondary templated
crystallization of zeolites starting from the amorphous mesoporous materials. They
transformed the amorphous mesopore walls into crystalline ones by additional
hydrothermal treatment of the amorphous mesoporous precursor in the presence of a
structure-directing agent for the required zeolite phase. Koegler et al. [69] transformed a
mesoporous amorphous material to a composite structure containing the original
mesopores where some of the porous inorganic oxide material was converted to
nanocrystalline zeolite.

1.3.3.2 Assembly from zeolite seeds


Recently Pinnavaia et al. [70, 71] reported an alternative approach based on the use
of nanoclustered zeolite seeds as framework precursors, yielding a product with both
micro- and mesoporosity. These protozeolitic species, known as zeolite seeds, promote
zeolite nucleation by adopting AlO4 and SiO4 tetrahedral connectivity that resembles the

25

secondary structural subunits of a crystalline zeolite. The first steam-stable hexagonal


mesoporous aluminosilicates were successfully assembled from faujasitic-type Y zeolite
seeds denoted as Al-MSU-S [71]. XRD patterns of the seed solution showed no Bragg
peaks, indicating the absence of a well-crystallized zeolite Y phase. It was reported that
Al-MSU-S assembled from nanoclustered zeolite Y seeds retained a well-ordered
hexagonal structure upon steam treatment: the 10 % Al-MSU-S could retain 90 % and
75 % of its surface area and pore volume, respectively. In addition to the high steam
stability, 10 % Al-MSU-S was a far more active acid catalyst for cumene cracking, in
comparison to the conventional 10 % Al-MCM-41. The presence of zeolite connectivities
in Al-MSU-S was deduced by the observation of a single tetrahedral aluminum peak at ca.
61 ppm in the 27Al NMR spectrum of the calcined mesostructure. This is consistent with
the

27

Al chemical shift of the zeolite Y seeds solution and the Al environment of

crystalline zeolite Y. Normally, a single 27Al NMR peak at a chemical shift near 55 ppm
is observed for conventional Al-MCM-41 [72]. The higher chemical shift of the
tetrahedral aluminum sites, together with the stronger acidity of Al-MSU-S for cumene
cracking, implies that zeolite-like connectivities were retained in the Al-MSU-S material.
Besides HY seeds, pentasil zeolite seeds, such as BEA and MFI type seeds have been
used to make micro/mesoporous materials [70, 73-74]. In comparison to the faujasitic
zeolite seeds, which are nucleated by sodium ions, pentasil zeolite seeds are nucleated by
specific tetraalkylammonium ions. Pinnavaia et al. [75] prepared hydrothermally stable
and strongly acidic MCM-41 analogs from zeolite ZSM-5 and Beta seeds, which are
nucleated by tetrapropylammonium and tetraethylammonium cations, respectively. They
proposed that the hydrothermal stability and catalytic activity of MSU-S (MFI) and

26

MSU-S (BEA) arise from the presence of zeolitic subunits of AlO4 and SiO4 tetrahedra in
the framework walls of the mesostructures. Evidence for the retention of a protozeolitic
connectivity of tetrahedra was provided by IR spectroscopy. Zhang et al. [74] also
reported a hydrothermally stable MCM-41 analog, MAS-5, which was assembled from
zeolite beta seeds. The workers claimed that MAS-5 showed stronger acidity than
conventional Al-MCM-41 for 1,3,5-triisopropylbenzene cracking and higher catalytic
activity than zeolite beta for the alkylation of 2-butene with isobutene. Although the
acidity of MAS-5 reported to be very similar to zeolite beta, the higher catalytic activity
for the alkylation was attributed to the easier diffusion of products in the mesoporous
channels of MAS-5 than in microporous zeolite beta. Xia and Mokaya [76, 77]
synthesized ZSM-5/MCM-48 using a simple two step crystallization process which
involved the assembly of precursor zeolite species, containing ZSM-5 units at various
stages of crystallization, into a mesostructured material. It was found that the acidity and
hydrothermal stability of the composite materials was strongly dependent on the extent of
zeolite phase, i.e., ZSM-5/MCM-48 ratio.
Usually,

the

production

of

these

micro/mesoporous

materials

use

cetyltrimethylammonium bromide (CTAB) as template for the hexagonal structure while


a recent report by Pinnavaia used starch as a porogen [78]. Under appropriate reaction
conditions, these reagents afford nanoparticles that contain mesopores with exceptionally
thick walls between the pores. The new nanoparticles showed remarkable hydrothermal
stability and catalytic reactivity for the cumene cracking reaction which can be attributed
to a unique combination of two factors - the presence of protozeolitic nanoclusters in the
pore walls and the unprecedented pore wall thickness provided through the use of starch.

27

The unique combination of mesoporosity, hydrothermal stability, acidity, and longevity


made it desirable for the efficient cracking of high molecular weight hydrocarbons. More
recently, aluminosilicates with well-ordered MCM-48 mesostructures were prepared via
the assembly of dried zeolite precursors with Gemini surfactant by Kao et al. [79].
They claimed that the use of the Gemini surfactant and the dried zeolite precursor
facilitated the incorporation of more Al into the MCM-48 framework without
compromising the integrity of the cubic mesostructure. The materials assembled from
zeolite precursors exhibited higher hydrothermal stability and better cumene cracking
activities than the Al-MCM-48 sample prepared by direct synthesis.
Larger pore SBA-15 and MCF (with substantially larger pore size of 7 - 35 nm)
analogs can be synthesized under mild acidic condition when sodium silicate was used as
the silica source. Nanoclustered faujasite, ZSM-5 and beta seeds have been successfully
incorporated [73]. The resulting large-pore hexagonal and foam-like materials (denoted
MSU-S/FFAU, MSU-S/FMFI and MSU-S/FBeta) showed higher stability in boiling water
and under steaming conditions. All these materials exhibited higher cumene cracking
activities than the corresponding Al-MCF which do not contain nanoclustered zeolitic
walls. Han et al. [80] reported a hydrothermally stable SBA-15 analog (denoted MAS-9)
assembled from ZSM-5 seeds. Kaliaguine et al. [81] described the production of unusual
ZSM-5-coated mesoporous aluminosilicate SBA-15 using a diluted clear solution
containing primary zeolite units. The resultant SBA-15 analog, ZC MesoAS, showed
higher Brnsted acid sites compared to the parent Al-SBA-15 due to the nanocrystalline
zeolitic nature of their pore wall surface. Furthermore, they prepared a new type of
zeolite-coated mesocellular alumosilicate foams (zeolite-coated MCF) using clear ZSM-5

28

(MFI) and faujasitic zeolite type Y (FAU) gel solutions [82]. With steam treatment, the
coated samples showed much more hydrothermal stability than the parent MCF sample.
This was attributed to the zeolite seeds coated on the mesopore surface, which create
valence bonds with the precursor and heal defect sites, reducing consequently the
concentration of silanol groups. Therefore, it was concluded that zeolite-coated MCF are
promising as new acid catalysts for the conversion of bulky molecules at high
temperature.
The use of protozeolitic nanoclusters for the preparation of improved mesostructures
is not limited to aluminosilicates. Metal-substituted seed compositions, such as those
used to nucleate TS-1, Ga-MFI, and AlPO, may be used as precursors for the preparation
of stable mesostructures [83].

1.3.3.3 Other synthetic methods


Landry et al. [84] described the synthesis of MCM-41/MFI materials by a two-step,
two-template process. The mesoporous material was first synthesized and then
crystallized to form the microporous phase. Tetrapropylammonium bromide was added to
the weakly organized MCM-41 to form the microporous phase. The extent of
transformation was controlled by the crystallization time. Bein et al. [85] prepared
nanosized micro-mesoporous composites via simultaneous hydrothermal treatment of a
MCM-48 precursor solution and a colloidal solution containing X-ray amorphous
aluminosilicate precursors. The formation of composites containing both micro- and
mesoporous phases was confirmed with XRD, IR spectroscopy, TG measurements and
nitrogen sorption data.
29

In summary, micro-mesostructured materials have certainly brought a new dimension


to the design of catalysts. Their potential is currently under intense study in catalytic
applications. One key issue for the applicability of these materials in catalysis is
associated with the hydrothermal stability and acidity. Another key issue is associated
with reactants and products mass transfer-diffusion in the pores during catalytic reaction.
Indeed, structured solids with bimodal pore size distributions are of special interest in
permitting the diffusion of larger molecules. Although great interest has been focused on
the synthesis of composite materials, there have been very few cases of the usage of these
materials as catalysts for reactions producing fine chemicals. Moreover, some of the assynthesized materials do not have any microporosity from nitrogen sorption studies while
others comprised only mixtures of two different phases. In addition, most of the methods
do not demonstrate any control over the textural aspects of the materials.

30

1.4 Aims of the study


We have found that zirconium incorporation into beta zeolite, i.e. Zr-beta, is an
efficient Lewis acid catalyst for the MPV reduction of a number of aldehydes and ketones
with very high steroselectivity to the alcohols. Moreover, by loading metals onto the
zeolite, selective hydrogenation catalysts can be formed that have additional properties of
acidity and shape-selectivity.
Micro-mesostructured materials should have advantages of both microporous and
mesoporous molecular sieves, making them promising catalysts for organic reactions.
However, their use in the synthesis of organic intermediates and fine chemicals is in a
relatively early stage and subtle changes in synthetic methods can make profound effects
on the outcome. One of the goals in this study is to synthesize micro-mesoporous
materials by different processes and to test them as catalysts for organic reactions. The
following are the aims of the project:
1. Synthesis of a variety of materials with varying microporous/mesoporous phases and
investigation of the physical and chemical parameters that may influence the synthesis
of these materials.
2. Synthesis of bifunctional catalysts containing metal functional groups and acidic sites.
3. Characterization of the synthesized materials by employing various characterization
techniques such as powder X-ray diffraction (XRD), infrared spectroscopy (IR),
scanning and transmission electron microscopy (SEM and TEM, respectively), solid
state nuclear magnetic resonance (NMR) and N2 adsorption measurements.

31

4. Catalytic testing of the Zr-beta zeolite and micro-mesoporous materials in the


cyclisation of ()-citronellal to isopulegol.
5. Investigation of the catalytic activity and regio-selectivity of bifunctional Ni/Zr-beta
and Zr-beta/Ni-MCM-41 catalysts in the domino-cyclisation and hydrogenation of
citronellal and citral to menthol.
6. Application of supported Rh/Zr-beta catalyst in the steroselective hydrogenation of 4tert-butylphenol and p-cresol.
7. Development of Beta/SBA-15 micro-mesoporous catalytic materials and application
in the alkoxylation of limonene.

32

References
1.

K.S.W. Sing, D.H. Everett, W.R.A. Haul, L. Moscou, J. Pierotti, J. Ruquerol and T.
Siemieniewska, Pure Appl. Chem. 57 (1985) 603.

2.

J. Cejka and B. Wichterlova. Catal. Rev. 44 (2002) 375.

3.

M.J. Verhoef and H. van Bekkum, Chem. Mater. 13(2001) 683.

4.

J. Medina-Valtierra, Appl. Catal., A 158(1997) 317.

5.

J.M. Kim, S.K. Kim and R. Ryoo, Chem. Commun., (1998) 259.

6.

C. Herrmann, J. Haas and F. Fetting, Appl. Catal. 35 (1987) 299.

7.

http://www.personal.utulsa.edu/~geoffrey-price/index.html, TU Chemical
Engineering, zeolite page, by Professor Geoffrey L. Price.

8.

J. Weitkamp, Solid State Ionics, 131 (2000) 175.

9.

www.iza-structure.org/databases/

10. web.chemistry.gatech.edu/~wilkinson/Class_notes/CHEM_3111_6170/Introduction
_to_zeolites.pdf
11. J.M. Newsam, M.M.J. Treacy, W.T. Koetsier and C.B. de Gruyter, Proc. R. Soc.
London, Ser. A 420 (1988) 375.
12. M.M.J. Treacy and J.M. Newsam, Nature 332 (1988) 249.
13. J.B. Higgins, R.B. La Pierre, J.L. Schlenker, A.C. Rohrmann, J.D. Wood, G.T. Kerr
and W.J. Rohrbaugh, Zeolites, 8 (6) (1988) 446.
14. A. Corma, M.T. Navarro, F. Rey, J. Rius and S. Valencia, Angew. Chem. Int. Ed. 40
(2001) 2277.
15. D.W. Breck, U.S. Patent 3,130,007 (1964).
16. R.L. Wadlinger, G.T. Kerr, E.J. Rosinski, U.S. Patent 3,308,069 (1967).

33

17. P. Caullet, J. Hazm, J.L. Guth, J.F. Joly, J. Lynch and F. Raatz, Zeolites, 12 (1992)
240.
18. C.S. Cundy and P.A. Cox, Chem. Rev., 103 (2003) 663.
19. T. Blasco, M.A. Camblor, A. Corma, P. Esteve, J.M. Guil, A. Martnez, J.A.
Perdign-Meln, and S. Valencia, J. Phys. Chem. B 102 (1998) 75.
20. A. Corma, L.T. Nemeth, M. Renz, and S. Valencia, Nature, 412 (2001) 423.
21. Y.Z. Zhu, G.K. Chuah and S. Jaenicke, J. Catal., 227 (2004) 1.
22. M. Guisnet and J.P. Gilson, Introduction to Zeolite Science and Technology:
Zeolites for Cleaner Technologies, Eds. M. Guisnet and J.-P. Gilson, Imperial
College Press (2003), p 11.
23. C. Mirodatos and D. Barthomeuf, Chem. Commun. 1981, 39.
24. J. Rabo and G.J. Gajda, Guidelines for Mastering the Properties of Molecular
Sieves, Eds. D. Barthomeuf et al., Plenum Press, New York, 221 (1990) p. 273.
25. W.F. Hlderich and H. van Bekkum, in Stud. Surf. Sci. Catal.: Introduction to
Zeolite Science and Practice, Eds. H. van Bekkum, E.M. Flanigen and J.C. Jansen,
Vol. 137, Elsevier, Amsterdam, (2001) p. 821.
26. J. Weitkamp, S. Ernst and L. Puppe, Catalysis and Zeolite : Shape-Selective
Catalysis in Zeolites, Eds. J. Weitkamp and L. Puppe, Springer (1999) p. 327.
27. J. A. Martens and P. A. Jacobs, Stud. Surf. Sci. Catal. : Introduction to Acid
Catalysis with Zeolites in Hydrocarbon Reactions, Eds. H. van Bekkum, E.M.
Flanigen and J.C. Jansen, Vol. 137, Elsevier, Amsterdam, (2001) p. 663.
28. Website: http://www.che.caltech.edu/groups/med/catmat.htm
29. P.H. Espeel, B. Janssens and P.A. Jacobs, J. Org. Chem., 58 (1993) 7688.

34

30. P.B. Venuto and L.A. Hamilton, Ind. Eng. Chem. Prod. Res. Dev. 6 (1967) 190.
31. S.M. Csicsery, J. Catal. 23 (1971) 124.
32. E.J. Creyghton, S.D. Ganeshie, R.S. Downing, and H. van Bekkum, Chem.
Commun. (1995) 1849.
33. E.J. Creyghton, S.D. Ganeshie, R.S. Downing, and H. van Bekkum, J. Mol. Catal.
115 (1997) 457.
34. J.V. van der Waal, K. Tan, and H. van Bekkum, J. Catal. 173 (1998) 74.
35. E.G. Derouane, J. Mol. Catal. A: Chem, 134 (1998) 29.
36. R.L. Gorring and R.H. Danills, ACS Symp. Ser. 248 (1984) 51.
37. E.G. Derouane and Z. Gabelica, J. Catal. 65 (1980) 486.
38. E.G. Derouane, J- M. Andre and A.A. Lucas, J. Catal. 110 (1988) 58.
39. P. Gallezot, B. Blanc, D. Barthomeuf and Mi. Pas da Silva, Stud. Surf. Sci. Catal. :
Zeolites and related microporous materials: State of the art, Eds. J. Weitkamp,
H.G. Karge, H Pfeifer and W. Hlderich, Vol. 84, Elsevier, Amsterdam, (1994) p.
1433.
40. D.S. Santilli and S.I. Zones, Catal. Lett. 7 (1990) 383.
41. J. Weitkamp, Ind. Eng. Chem. Prod. Res. Dev. 21 (1992) 550.
42. J.A. Martens, M. Thielen and P.A. Jacobs, Stud. Surf. Sci. Catal. 46 (1989) 49.
43. E.J. Creyghton and R.S. Downing, J. Mol. Catal. A: Chem., 134 (1998) 47.
44. R.M. Dessau, J. Catal. 89 (1984) 520.
45. R.M. Dessau, J. Catal. 77 (1982) 304.
46. E.J. Creyghton, R.A.W. Grotenbreg, R.S. Downing and H. van Bekkum, J. Chem.
Soc., Faraday Trans. 92 (1996) 871.

35

47. T. Yanagisawa, T. Shimizu, K. Kuroda and C. Kato, Bull. Chem. Soc. Jpn., 63
(1990) 988.
48. J.S. Beck, J.C. Vartuli, W.J. Roth, M.E. Leonowicz, C.T. Kresge, K.D. Schmitt and
C.T.W. Chu, J. Am. Chem. Soc. 114 (1992) 10834.
49. J.Y. Ying, C.P. Mehnert and M.S. Wong, Angew. Chem. Int. Ed., 38 (1999) 57.
50. A. Monnier, F. Schth, Q. Huo, D. Kumar, D. Margolese, R.S. Maxwell, G.D.
Stucky, M. Krishnamurty, P. Petroff, A. Firouzi, M. Janicke, and B.F. Chmelka,
Science, 261 (1993) 1299.
51. G.D. Stucky, A. Monnier, F. Schth, Q. Huo, D. Margolese, D. Kumar, M.
Krishnamurty, P. Petroff, A. Firouzi, M. Janicke and B.F. Chmelka, Mol. Cryst. Liq.
Cryst. 240 (1994) 187.
52. A. Firouzi, D. Kumar, L.M. Bull, T. Besier, P. Sieger, Q. Huo, S.A. Walker, J.A.
Zasadzinski, C. Glinka, J. Nicol, D. Margolesse, G.D. Stucky and B.F. Chmelka,
Science, 267 (1995) 1138.
53. A. Corma, Chem. Rev. 97 (1997) 2373.
54. J. Frasch, B. Lebeau, M. Soulard, J. Patarin and R. Zana, Langmuir, 16 (2000) 9049.
55. J. Galo, A.A. Soler-lllia, C. Sanchez, B. Lebeau and J. Patarin, Chem. Rev. 102
(2002) 4093.
56. D. Zhao, J. Feng, Q. Huo, N. Mclosh, G.H. Frederickson, B.F. Chmelka and G.D.
Stucky, Science, 279 (1998) 548-552.
57. S.S. Kim, W.Z. Zhang and T.J. Pinnavaia, Science, 282 (1998) 1302-1305.
58. Y. Yue, A. Cedeon, J.L. Bonardet, N. Melosh, J.B. DEsinose and J. Fraissard,
Chem. Commun. (1999) 1697.

36

59. P. Yang, D. Zhao, D. Margoles, B. Chmelka and G.D. Stucky, Nature. 396 (1998)
152.
60. R. Mokaya, W. Zhou and W. Jones, Chem. Commun. (1999) 51.
61. R. Mokaya, Angew. Chem. Int. Ed., 38 (1999) 2930.
62. Z. Luan, M. Hartmann, D. Zhao, W. Zhou and L. Kevan, Chem. Mater., 11 (1999)
1621.
63. S. Wu, Y. Han, Y.C. Zou, J.W. Song, L. Zhao, Y. Di, S.Z. Liu and F.S. Xiao, Chem.
Mater. 16 (2004) 486.
64. K. Richard Kloctstra, J.C. Jansen and H. van Bekkum, Chem. Commun., (1997)
2281
65. M.J. Verhoef, P.J. Kooyman, J.C. van der Waal, M.S. Rigutto, J.A. Peters and H.
van Bekkum, Chem. Mater. 13 (2001) 683.
66. D.T. On, P. Reinert, L. Bonneviot and S. Kaliaguine, Stud. Surf. Sci. Catal. 135
(2001) 929.
67. D.T. On and S. Kaliaguine, Angew. Chem. Int. Ed. Engl., 40 (2001) 3248.
68. L. Huang, W. Guo, P. Deng, Z. Xue and Q. Li, J. Phys. Chem. B 104 (2000) 2817.
69. J.H. Koegler, Heidelberg, Y.Yeh Chuen, Edison, NJ, Philip J. Angevien and
Woodbury, NJ, United States Patent, 2004, Patent No.: US 6,793,911 B2.
70. Y. Liu, W. Zhang, and T.J. Pinnavaia, Angew. Chem. Int. Ed., 40 (2001) 1255.
71. Y. Liu, W. Zhang, and T.J. Pinnavia, J. Am. Chem. Soc. 122 (2000) 8791.
72. R.B. Borade and A. Clearfield, Catal. Lett., 31 (1995) 267272.
73. Y. Liu and T.J. Pinnavaia, Chem. Mater., 14 (2002) 35.

37

74. Z.T. Zhang, Y. Han, F.S. Xiao, S.L. Qiu, L. Zhu, R.W. Wang, Y. Yu, Z. Zhang, B.S.
Zou, Y.Q. Wang, H.P. Sun, D.Y. Zhao and Y. Wei, J. Am. Chem. Soc., 123 (2001)
50145021.
75. S.S. Kim, T.R. Pauly, and T.J. Pinnavaia, Chem. Commun., (2000) 661.
76. Y. D. Xia and R. Mokaya, J. Mater. Chem., 14 (2004) 863.
77. Y. D. Xia and R. Mokaya, J. Phys. Chem. B, 107 (2003) 6954.
78. Y. Liu and T.J. Pinnavaia, J. Am. Chem. Soc. 125 (2003) 2376.
79. H.M. Kao, H.M. Wu, Y.W, Liao and A.S.T. Chiang, Microporous Mesoporous
Mater., 86 (2005) 256.
80. Y. Han, S. Wu, Y. Sun, D. Li, F.S. Xiao, J. Liu and X. Zhang, Chem. Mater., 14
(2002) 11441148.
81. D.T. On and S. Kaliaguine, Angew. Chem., Int. Ed., 41 (2002) 10361040.
82. D. T. On and S. Kaliaguine, J. Am. Chem. Soc., 125 (2003) 618.
83. F.S. Xiao, Y. Han, Y. Yu, X. Zhang, M. Yang and S. Wu, J. Am. Chem. Soc., 124
(2002) 888889.
84. H.R. Poladi and C. Landry, J. Solid State Chem., 167 (2002) 363.
85. P. Prokeov, S. Mintova, J. ejka and T. Bein, Microporous Mesoporous Mater.,
64 (2003) 165.

38

You might also like