9781849732772-00253

download 9781849732772-00253

of 31

Transcript of 9781849732772-00253

  • 7/31/2019 9781849732772-00253

    1/31

    Hierarchical zeolites: materials withimproved accessibility and enhancedcatalytic activity

    D. P. Serrano,1, 2 J. Aguado3 and J. M. Escola3

    DOI: 10.1039/9781849732772-00253

    1 Introduction

    Conventional zeolites are crystalline aluminosilicates formed by a network of

    SiO4 and AlO4 tetrahedra linked by shared oxygen atoms, which are usually

    obtained in the form of crystals/particles of micrometer size (W1mm).1 The

    presence of aluminium atoms into their framework confers them acidic

    properties (both Bro nsted and Lewis) although they can also incorporate

    other heteroatoms such as Ti, V, Sn, etc., allowing their application ascatalyst in many different types of chemical reactions.2 However, their major

    feature is likely the occurrence of micropores of molecular dimensions

    (generally 0.40. 75 nm), which turn them into molecular sieves. This unique

    property has been also called shape selectivity and allows the zeolites to

    discriminate among different reactants, products or even transition states

    according to their shape and size. This phenomenon has led to remarkable

    selectivities exhibited by zeolites in a large number of reactions. This is the

    case of toluene disproportionation, wherein enhanced yields and selectivities

    towards the p-xylene isomer have been attained over the ZSM-5 zeolite dueto its improved diffusion through the channel systems with regard to the

    m/o- isomers.3 In fact, many of the current successful applications of zeolites

    are based on their molecular sieves character.

    However, zeolites fail when dealing with bulky substrates, usually

    exhibiting low activities, often even below the values obtained with

    amorphous materials such as silica-alumina. This result is to be expected as

    the bulky substrates cannot access the active sites located inside the zeolite

    micropores due to small size of the latter. Thus, only those sites situated

    over the external surface of the catalyst particles and crystals, or close to the

    micropore openings, are accessible for large molecules, and these represent

    usually a low share (o 5%) of the total content of active sites. A clear-cut

    example of this circumstance can be found in the catalytic cracking of

    polypropylene over ZSM-5 zeolite,4 wherein just 11% of conversion

    was observed at 3751C, despite the high acid strength of this zeolite, while a

    99% conversion was obtained over the mild acid strength Al-MCM-41

    mesoporous material. On the other hand, even if the substrate can enter into

    the zeolite micropores, the diffusion rate is usually too slow bringing about

    the appearance of intracrystalline mass transfer constraints, which limit

    meaningfully the performance of the zeolite catalyst.5

    1Department of Chemical and Energy Technology, ESCET, Universidad Rey Juan Carlos,c/ Tulipan s/n, 28933, Mostoles, Madrid, Spain2IMDEA Energy Institute, c/Tulipan s/n, 28933, Mostoles, Madrid, Spain3Department of Chemical and Environmental Technology, ESCET, Universidad Rey Juan

    Carlos, c/ Tulipan s/n, 28933, Mostoles, Madrid, Spain

    Catalysis, 2011, 23, 253283 | 253

    c The Royal Society of Chemistry 2011

  • 7/31/2019 9781849732772-00253

    2/31

    Seemingly, the increasing need of processing bulky compounds put the

    zeolites into a difficult situation, specially when ordered mesoporous

    materials such as MCM-416 and SBA-157 were discovered in the early and

    late 90s, respectively. These materials possess pore sizes whose dimension

    may be tailored within the 1.5 30.0 nm range by a suitable choice of

    synthesis conditions (mostly type of surfactant, temperature and aging

    time). However, the huge number of publications devoted to these materialssoon realized their main limitations, mainly derived from the amorphous

    nature of their walls. Accordingly, ordered mesoporous aluminosilicates do

    not possess the high hydrothermal stability and strong acidity of zeolites.

    Therefore, zeolites are still the preferred choice in numerous industrial

    applications.

    An alternative solution to deal with bulky substrates has been the syn-

    thesis and application of nanozeolites (crystal size below 100 nm) since they

    show a high share of external surface area whose active sites are accessible

    to large molecules.

    8

    However, nanozeolites are usually produced in lowyields since they are very difficult to separate from the synthesis medium.9,10

    Recently, several new strategies have appeared in order to improve the

    properties of zeolites when processing bulky substrates. Delaminated

    zeolites,11 large pore zeolites12,13 and hierarchical zeolites14,15 are examples

    of these new strategies. Delaminated zeolites, such as ITQ-2, result from the

    delamination of a layered zeolite precursor (e.g. MCM-22) giving rise to

    thin zeolite sheets (B2.5 nm thick) having a huge external surface area

    (Z700m2 g1).11 However, this method of synthesis is inherently bound to

    the occurrence of a zeolite layered precursor. The preparation of large pore

    zeolites, having channels formed by rings with more than 12 members, is

    another strategy that has been pursued in the late years. One of the main

    achievements in this line has been the discovery of ITQ-33, a silicogerma-

    nate having circular pores of 18 member-rings interconnected by 10

    member-ring channels and a crystallographic pore diameter of 1.22 nm.13

    This means a substantial enhancement in pore size considering that

    a conventional zeolite Y shows roughly 0.75 nm of pore dimension.

    Nevertheless, the main disadvantage of this approach is the need of

    employing Ge in the synthesis, as well as the addition of special structure

    directing agents (SDA) which are usually expensive reagents.One of the most successful strategies for improving accessibility is likely

    the case of hierarchical zeolites.14,15 These zeolites are characterized by the

    presence of a bimodal pore size distribution, formed by both micropores

    and mesopores. The microporous structure is the one inherent to the

    classical zeolite topology while the secondary mesoporous structure can be

    generated by a variety of specific synthetic procedures. The presence of a

    secondary porosity in hierarchical zeolites, usually in the mesopore range, is

    responsible for the improved mass transfer properties of these materials,

    which have proved to be advantageous in numerous reactions. On the otherhand, the surface area associated to this secondary porosity implies the

    presence of active sites that are not sterically hindered for interacting with

    bulky molecules. In some cases the mesopore surface area is confused with

    the external surface area, likely due to the application of the t-plot method

    for the joint calculation of both parameters. Although the nature and

    254 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    3/31

    performance of the active sites located in both mesopore and external

    surface area are probably quite similar, they are different concepts. External

    surface refers to the surface area located in the outer part of the zeolite

    particles, whereas mesopore surface corresponds with the surface area in the

    mesopore walls. Therefore, the occurrence of high external surface area

    does not mean the presence of mesopores in the material. Thus, nanozeolites

    show high external surface area but many times mesopores are absent. Inthis regard, hierarchical zeolites cannot be regarded truly as nanozeolites

    (although they usually consist of zeolitic nanounits as building blocks).

    From a catalytic viewpoint, most part of the active sites in hierarchical

    zeolites are placed inside both micropores and mesopores, the latter having

    improved accessibility. Henceforth, the main preparation strategies of

    hierarchical zeolites as well as their applications in different reactions are

    commented.

    2 Methods of preparation of hierarchical zeolitesHierarchical zeolites can be prepared by different procedures which show

    distinct features. Although all of them lead to materials with bimodal pore

    size distributions, the features and contribution of the generated secondary

    mesoporosity depends heavily on the chosen procedure. Some of these

    synthesis strategies present common aspects which allow them to be clas-

    sified as follows:

    Dealumination

    Desilication

    Hard templating by carbon materials

    Hard templating by polymers

    Incorporation of organosilanes

    Other methods

    2.1 Dealumination

    Dealumination methods represent the most classical alternative for de-

    veloping mesoporosity in zeolites. They comprise steaming at elevated

    temperatures and acid leaching.16 Steaming at high temperature of zeolite Y

    has been traditionally performed for its application as catalyst in fluidcatalytic cracking reactors (FCC) since it creates mesopores by removal of a

    certain amount of the aluminium from the zeolite framework. This treat-

    ment causes an enhancement of the hydrothermal stability of the zeolite as

    well as an improvement of the diffusion of bulky molecules inside the zeolite

    pores. The steaming treatment consists of first performing an ion exchange

    of an alkali metal containing zeolite Y with an ammonium salt solution in

    order to reduce the alkali content to 1025% of its starting value, followed

    by washing to remove the excess salt. Subsequently, the zeolite is heated to

    2006001C to enable the migration of sodium ions towards easilyexchangeable sites. Then, a second ion exchange is performed with an

    ammonium salt solution to remove completely the sodium. Finally the

    zeolite is heated under a steam atmosphere at a temperature between

    6008001C.17 The non framework aluminium resulting from the steaming

    treatment can be removed from the zeolite by subsequent washing with

    Catalysis, 2011, 23, 253283 | 255

  • 7/31/2019 9781849732772-00253

    4/31

    dilute acids. In addition, as a result of the steaming treatment, some

    shrinkage of the zeolite framework of around 0.020.03 nm takes place.

    Acid leaching also allows mesopores to be created by a quite similar

    mechanism, i.e. through the extraction of some aluminium from the

    framework positions. A typical procedure is the treatment of the zeolite in

    an slurry with a solution of ethylenediaminetetraacetic acid (EDTA) under

    reflux for 18 h.18

    Afterwards, the zeolite is heated with an inert gas at 8001C.However, this method has the drawback of removing preferentially the

    aluminium present over the external surface leading to non uniform

    distributions within the crystals.19 A related method which employs a strong

    acid treatment has been reported recently by Van Oers et al.20 They attained

    hierarchical Beta from nanoparticles of this zeolite using a procedure

    consisting in the ageing at 1401C of a zeolite Beta nanoparticle solution,

    followed by stepwise cooling or quenching. Then, this mixture is strongly

    acidified with concentrated HCl and subjected to a second hydrothermal

    treatment at 1501

    C for 72 h. The obtained mesopore size depended on theconditions employed in the cooling step, giving rise to roughly 10 nm pores

    with slow cooling and 6.0 nm pores with quenching.

    Nevertheless, recent experimental evidences have been found that ques-

    tion markedly the goodness of the mesoporous structure created in zeolite Y

    by steaming. 3-D TEM technique has allowed the nature of the obtained

    mesopores to be studied, indicating that their size encompass a broad range

    (2 50 nm).21 But the most important finding has been that the mesopores

    consist chiefly in cavities connected by micropores instead of a network of

    mesopore channels connecting the outer surface and the internal micro-

    pores. This is nicely illustrated in Fig. 1, which shows the 2-D and 3-D TEM

    micrographs of a USY zeolite obtained by steaming of a NH4Y sample.

    This finding has important consequences since this mesopore structure is

    not expected to cause significant enhanced diffusion rates inside the zeolite

    crystals. Likewise, other works have been devoted specifically to elucidate

    this fact by measuring intracrystalline diffusion rates in steamed USY

    Fig. 1 TEM micrographs of USY zeolite: a) 2D-TEM image of a crystal, b) 3D TEMreconstruction of a crystal. (Reprinted with permission from ref. 21, Copyright Wiley-VCHVerlag GmbH & Co KGaA, 2001)

    256 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    5/31

    zeolites using the PFG NMR technique.22 Two probe molecules were used:

    n-octane and 1,3,5 triisopropylbenzene, with molecular diameters smaller

    and larger than the zeolite micropores, respectively. The main conclusion of

    this work was that the intracrystalline diffusion inside zeolite USY is

    practically not affected by the presence of the mesopores generated by

    steaming, due to the absence of an interconnected network structure. This

    result indicates that the classical procedure of steaming is not reallyappropriate for the preparation of a network of mesopores, highlighting the

    importance of developing new methods of synthesis of hierarchical zeolites.

    2.2 Desilication

    The technique of desilication is based on the treatment of zeolites with a

    base (usually sodium hydroxide) under relatively mild conditions

    (To363 K). Mesopores are formed due to the preferential removal of silica

    from the framework. This is a remarkable difference from the acid leachingtreatment wherein aluminium atoms are selectively extracted. Ogura et al.23

    carried out the desilication of ZSM-5 zeolite with 0.2 M NaOH aqueous

    solution at 353 K. After the treatment, the ZSM-5 sample retains its crys-

    tallinity and the majority of their microporous structure, whereas it contains

    additional mesopores. The morphology of the zeolite is meaningfully

    altered by the alkaline treatment leading to the appearance of voids and

    grooves over the zeolite surface. This is clearly appreciated in Fig. 2 which

    shows the SEM micrograph of both the raw sample and the alkaline treated

    ZSM-5 zeolite. These authors also observed that the acidity of the zeolite

    was little affected after the alkaline treatment according to ammonia TPD

    measurements.24 Moreover, it was concluded that the size of the pores

    formed by the treatment with NaOH is around 1.8 nm, being indeed

    supermicropores.25

    A key parameter of desilication by alkaline treatment is the Si/Al atomic

    ratio of the zeolite.26 Thus, a Si/Al atomic ratio in the range 20 50 has been

    found as optimal in the case of ZSM-5 zeolite giving rise to intracrystalline

    mesopores with a size around 10 nm. In contrast, for lower Si/Al atomic

    ratios, the high aluminium concentration in the framework prevents the

    removal of silicon leading to almost no mesopore formation. On the con-trary, if the Si/Al is very high (Si/Al W 200), too much silicon is extracted

    creating large pores. This can be appreciated in Fig. 3, illustrating the

    evolution of the porous structure of the material with the alkali treatment

    for different Si/Al atomic ratios.

    On the other hand, the presence of extraframework aluminium proved to

    inhibit silicon extraction due to realumination during the alkaline treat-

    ment. However, this problem could be solved by previous acid washing in

    order to eliminate the extraframework aluminium species.27 An interesting

    fact that highlights the importance in this method of the Si/Al atomic ratiois the existence of Al concentration gradients in large ZSM-5 crystals. Thus,

    the external zeolite surface is often rich in Al species, while the concen-

    tration of aluminium in the inner part of the crystals is much lower. In this

    case, the application of the desilication procedure inevitably brings about

    the generation of hollow zeolite architectures.28 Therefore, special synthetic

    Catalysis, 2011, 23, 253283 | 257

  • 7/31/2019 9781849732772-00253

    6/31

    methods which lead to more uniform Al concentration into the zeolites are

    required if the desilication procedure is to be applied.

    The alkali treatment can be applied using also organic bases (TPAOH and

    TBAOH) instead of NaOH, which causes a higher extent of the aluminumleaching.29 In this way, the mesopore volume can be optimized by incorpor-

    ating in the alkaline treatment both NaOH and quaternary ammonium cat-

    ions,30 such as TPA or TBA . These cations act as pore growth moderators,

    being necessary for the success of the method that they do not enter the zeolite

    micropores (TMA does not work). The addition of these cations together

    with the alkaline solution allows the mesopore surface area to be increased

    without reducing the micropore volume (the ratio between these two

    parameters has been called hierarchy factor). In addition, the mesopore size

    decreased from 10.0 nm to 4.5 nm on augmenting the TPA

    concentration.Based on FTIR measurements using CO as probe molecule, it has been

    concluded that the acid strength of the Bro nsted sites of the zeolitic

    materials did not change significantly after desilication.31 However, the

    appearance of new strong Lewis acid sites have been observed assigned to

    dislodged Al species. The enhanced accessibility of the Bro nsted acid sites

    Fig. 2 SEM micrographs of the original (a) and alkaline treated ZSM-5 (b). (Reprinted withpermission from ref. 23, Copyright Japan Chemical Society)

    258 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    7/31

    in the desilicated zeolites has been proved by FTIR using substituted

    alkylpyridines such as 2-6-lutidine (size of 0.67 nm) and 2,4,6-collidine (size

    of 0.74 nm).32 Collidine is too bulky to enter the MFI zeolite micropores

    while lutidine can probe certain amount of the zeolite sites (below 50%).

    This fact allowed an accessibility index (ACI) to be established as the ratio

    of the Bro nsted acid sites detected with the probe molecule with regard to

    the total number of Bro nsted sites, measured using a fully accessible probe

    molecule such as pyridine (size of 0.57 nm). This accessibility index showed

    a remarkable increase with the mesopore surface area reaching maximum

    values of 0.4 and 1 with collidine and lutidine, respectively, for a mesopore

    surface area of 277 m2 g 1.

    The desilication procedure has been extended to other zeolites differentfrom ZSM-5, such as mordenite33 and ZSM-12.34 Its main advantage is that

    the procedure to be applied is quite simple and does not involve the use of

    expensive reagents. However, one inherent limitation of the method is the

    need of having a Si/Al atomic ratio within a given range. Thereby, its

    applicability is lower than other reported procedures that do not show this

    bound. In addition, desilication involves a partial destruction of the zeolite

    structure what should be carefully controlled in order to avoid the

    appearance of amorphous material or the generation of extraframework

    aluminium species.

    2.3 Hard templating by carbon materials

    The outburst in the synthesis of mesoporous/hierarchical zeolites has been

    largely marked by the work of Jacobsen and col.35 It was preceded by the

    development of the confined synthesis method for the preparation of

    Fig. 3 Evolution of the porous structure of ZSM-5 zeolite during desilication by alkalitreatment as a function of the Si/Al ratio. (Reprinted with permission from ref. 27, CopyrightWiley-VCH Verlag GmbH & Co KGaA, 2005)

    Catalysis, 2011, 23, 253283 | 259

  • 7/31/2019 9781849732772-00253

    8/31

    nanozeolites, performed also by this research group.36 The latter consists of

    the synthesis of nanozeolites inside a mesoporous inert carbon matrix.

    Subsequently, the carbon is burnt off releasing the nanozeolites with a

    crystal size of around 20 nm that could be tailored by controlling the pore

    size of the used inert matrix. The preparation of mesoporous zeolites could

    be carried out by modifying slightly this method. Thus, if an excess of

    zeolite gel is used in the synthesis, the zeolite grows over and around theparticles of the carbon matrix forming large zeolite crystals which en-

    capsulates the inert carbon matrix. Once the zeolite crystal is formed, the

    carbon matrix is eliminated by combustion in air giving rise to large zeolite

    crystals containing mesopores roughly of the size of the carbon matrix

    particles (see Fig. 4).

    Seemingly, the dimension of the mesopores can be easily controlled by

    adjusting the size of the starting carbon particles. In this sense, in the original

    work of Jacobsen, carbon black pearls 2000, with a mean size of 12 nm, were

    employed leading to a mesopore volume of 1.01 cm

    3

    g

    1

    and a pore sizewithin the range 5 50 nm. The dimensions of the zeolite particles so obtained

    were relatively large (0.3 1.2 mm), being formed by agglomerates of smaller

    crystals. However, these nanocrystals show their zeolite planes aligned

    spanning throughout the entire agglomerate which is indicative of them being

    organized as a single zeolite crystal. In addition, electron diffraction patterns

    also indicate that the apparent agglomerates are actually single zeolite crys-

    tals. Other carbon sources have been also used as matrixes such as multiwall

    carbon nanotubes (MWNT) which led to narrower mesopore distributions

    than with carbon black pearls.37 In addition, the use of carbon fibres also

    allowed cylindrical mesopores to be obtained with low tortuosity.38

    This method was further applied successfully to the synthesis of a rela-

    tively large number of hierarchical zeolites, such as MFI,39 MEL,40 MTW,41

    BEA and CHA.42 However, one of the limitations of the method is the

    availability of adequate carbon templates to obtain the desired mesoporous

    structures. Accordingly, a modification of the carbon template method has

    been proposed also by Jacobsen et al.43 which allows at least partially this

    drawback to be overcome. They devised a procedure consisting of using a

    Fig. 4 Scheme of the synthesis of mesoporous zeolites by carbon templating according to thework of Jacobsen. (Reprinted with permission from ref. 35, Copyright American ChemicalSociety, 2000)

    260 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    9/31

    carbohydrate (sucrose) instead of a carbon template. The method comprises

    the incipient wet impregnation of silica gel with a sucrose solution, followed

    by its thermal decomposition under inert atmosphere (Ar). Thus, a carbon-

    silica composite is formed. Subsequently, a base and the corresponding

    zeolite template are added and the mixture is left crystallizing. Afterwards,

    the carbon template is eliminated by combustion in air and a hierarchical

    zeolite is formed. The resulting mesoporous structure can be tailored tosome extent by controlling the concentration of the sucrose solution. Fig. 5

    shows TEM micrographs of one of the mesoporous zeolites so obtained.

    It should be remarked that hierarchical zeolites obtained by the carbon

    templated method of Jacobsen et al. contain interconnected micropores

    and mesopores, so enhanced mass transfer rates are to be expected, unlike

    it occurred with the hierarchical USY zeolites obtained by steaming. This

    fact has been proved by 3D-TEM of a hierarchical zeolite synthesized

    using carbon nanotubes as hard-templates. The 3D images showed that

    the mesopores formed defined channels spanning throughout the zeolitecrystal.44

    Carbon aerogels monoliths have been also used as templates for the

    generation of mesoporosity in zeolites.45 In this case, the mechanism of

    formation of mesopores is slightly different from those shown above

    according to the Jacobsen method. Carbon aerogels were obtained from

    resorcinol-formaldehyde gels after drying with CO2 under supercritical

    conditions and pyrolysis under nitrogen atmosphere at 1323 K. The

    resulting carbon aerogel possessed mesopores with 23 nm size. The for-

    mation of the hierarchical zeolites takes place by the crystallization of the

    ZSM-5 zeolite inside the mesopores of the carbon aerogel. The latter is

    Fig. 5 TEM micrographs of hierarchical zeolites obtained using concentrated sucrosesolutions as precursors of the carbon template. (Reprinted with permission from ref. 43,Copyright American Chemical Society, 2007)

    Catalysis, 2011, 23, 253283 | 261

  • 7/31/2019 9781849732772-00253

    10/31

    subsequently removed by burning off under oxygen leaving available the

    space corresponding to the carbon aerogel pore walls. The obtained

    material is a ZSM-5 zeolite with a mesopore size of 11 nm. This method

    presents the added advantage that the mesoporous zeolite has been

    synthesized as a monolith.

    Ordered mesoporous carbons, prepared by nanocasting, have been also

    used successfully as templates for the synthesis of hierarchical zeolites.46,47

    This is the case of CMK-3, an ordered mesoporous carbon attained by

    nanoreplication of pure silica SBA-15. The hierarchical zeolites obtained

    employing CMK-3 as template present mainly supermicropores or small

    mesopores with a size around 2 nm. These are mesopores smaller than those

    reported previously with other carbon sources. The textural properties of

    the hierarchical zeolites can be tuned by changing the type of CMK-3

    carbon used. It should be pointed out that the size of the mesopores in

    CMK-3 carbon can be tailored using SBA-15 templates with different

    mesopore size. A modification of this method consists of impregnatingdirectly the composite SBA-15/carbon or MCM-41/carbon with TPAOH

    and left the mixture crystallizing hydrothermally under steam.47 After cal-

    cination, a mesoporous ZSM-5 is formed. In this case, the size of the

    mesopores is around 3.5 nm and 10 15 nm depending on whether MCM-

    41 or SBA-15 are used as raw templates, respectively.

    In summary, carbon-templating offer many possibilities for synthesizing

    hierarchical zeolites due to the large availability of carbon materials that

    can be employed. In contrast, the main disadvantage of these carbon-based

    strategy is related to the need of burning the carbon template, which

    represents a great amount of material being destroyed and having, in many

    cases, a high cost. Moreover, the carbon combustion may generate high

    temperatures that can damage the zeolite structure.

    2.4 Hard templating by polymers

    Closely related with the use of carbon based templates, is the application of

    polymers as hard templates for the synthesis of hierarchical zeolites. In this

    regard, the work of Xiao et al.48 deserves special mention. These authors

    have employed mesoscale cationic polymers, like polydiallyldimethyl-

    ammonium chloride, for the preparation of hierarchical Beta zeoliteby a one-step hydrothermal route. The size of the mesopores obtained

    varies in the range 5 40 nm, which is in agreement with the molecular

    dimension of the cationic polymer used. According to the authors, the

    method can be extended to the synthesis of other zeolites different from Beta.

    Polystyrene spheres have also been employed for obtaining hierarchical

    zeolites with a microporous/macroporous pore structure.49 Likewise, the

    use of nanospheres of poly(methyl methacrylate) (PMMA) has allowed

    mesoporous microspheres of zeolite ZSM-5 to be prepared by a mechanism

    involving self assembly of the aluminosilicate source, the PMMAnanospheres and TPAOH under basic conditions.50 Subsequently, hydro-

    thermal crystallization takes place and the hierarchical zeolite is obtained

    after calcination. Mesopores with about 13 nm diameter are achieved

    with this method despite that the PMMA nanosphere size is far higher

    (around 80 nm). Mesoporous zeolite A was also obtained using as templates

    262 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    11/31

    aerogels formed by sol-gel polymerization of resorcinol-formaldehyde.

    The hierarchical zeolite A so obtained showed mesopores with a size of

    roughly 15 nm. This method has the advantage that the nanoscale structure

    of the polymer aerogel template can be tuned just by changing the resorcinol -

    catalyst ratio, allowing some modification and control of the mesopore

    size.51

    2.5 Incorporation of organosilanes

    One alternative initially attempted to obtain hierarchical zeolites was based

    on putting together in the synthesis gel the structure-directing agent of the

    zeolite and that of an ordered mesoporous solid in order to obtain a hybrid

    material formed by an array of ordered mesoporous with crystalline zeolitic

    walls. However, this approach has failed to get the desired product as it

    usually led to a physical mixture of segregated phases.52 Pinnavia et al.53

    succeeded in solving this problem by a two-step synthesis strategy involving

    the previous formation of a precursor solution containing the zeolite seedsthat subsequently were assembled into the mesostructure by aggregation

    around surfactant micelles. However, the materials so obtained lack of

    zeolitic crystalline features, being wide-angle X-ray amorphous, which

    makes difficult to establish their real crystalline character.

    In order to avoid the phase separation of the mesoporous material

    and the zeolite, Choi et al.54 designed a specific amphiphilic organosilane

    ((3-trimethoxysilyl) propyl/hexadecyldimethylammonium chloride) of

    formula [(CH3O)3SiC3H6N(CH3)2C16H33]Cl. This amphiphilic organosi-

    lane is a surfactant molecule that contains a hydrophobic alkyl chain, a

    quaternary ammonium group (zeolite structure-directing agent) and a

    hydrolysable trimethoxysilyl moiety. The latter interacts strongly with the

    growing zeolite crystals due to the formation of covalent bonds with the silyl

    moiety avoiding the separation of the phases. The silanization constitutes a

    crucial step for the success of the method. The amount of amphiphilic

    organosilane used is low (4% mol), being first added to a typical syn-

    thesis composition of ZSM-5 zeolite, containing tetrapropylammonium

    hydroxide (TPAOH). The evolution of the crystallinity of the products

    obtained using this method with the synthesis time is shown in Fig. 6,

    which depicts the low and wide-angle XRD patterns. At short synthesistimes (3 h), only a mesoporous phase with amorphous walls is obtained.

    However, after 12 h, the appearance of zeolite MFI begins to be appreciated

    in the high angle XRD pattern, and finally, after 2 days, the presence of a

    mesoporous structure with zeolitic pore walls is appreciated.

    The authors suggested a solution-based mechanism to explain the for-

    mation of the mesoporous zeolite, involving the dissolution of the starting

    mesophase, followed by the crystallization of the mesoporous zeolite crys-

    tals from the dissolved species. The mesopore diameter can be tailored

    within the range 2 20 nm by changing the length of the alkyl chain of theorganosilane or by modifying the hydrothermal conditions of the synthesis.

    The mesoporosity created by this procedure is claimed to be rather uniform.

    On the other hand, the mesopore size can be widened up to 24 nm by using

    triblock copolymers (EO20PO70EO20) as pore expanding agents.55 This

    method for the synthesis of hierarchical zeolites has allowed the preparation

    Catalysis, 2011, 23, 253283 | 263

  • 7/31/2019 9781849732772-00253

    12/31

    of MFI, LTA54,55 and sodalite56 materials. Regarding the acidity of the

    samples, and according to infrared mass spectroscopy/temperature

    programmed desorption measurements, hierarchical ZSM-5 synthesized by

    this method contains Bro nsted acid sites of similar strength as those of

    conventional ZSM-5 but in lower number.57

    Different approachs have been also developed by other authors to syn-thesize hierarchical zeolites by perturbing the crystallization mechanism

    through the addition of organosilanes. Thus, our research group5861

    envisaged a procedure for the preparation of hierarchical zeolites by using

    seed silanization agents. The method is based on the fact that during the early

    stages of crystallization of MFI zeolites from clear synthesis solutions, the

    precursors are nanounits with a particle size of 25 nm.6264 If the

    agglomeration of these nanocrystals into bigger entities is partially hindered

    by the presence of bulky organic substituents anchored over their external

    surface, the obtained zeolite shows, after calcination, the occurrence of

    mesopores whose dimension correspond to the voids occupied by the bulky

    substituents during aggregation. This strategy comprises the following stages:

    a) Precrystallization of the zeolite synthesis gel to promote the formation

    of ZSM-5 protozeolitic nanounits, using tetrapropylammonium hydroxide

    (TPAOH) as structure directing agent.

    Fig. 6 XRD patterns of mesoporous MFI zeolite generated using amphiphilic organosilanealong different synthesis times. (Reprinted with permission from ref. 54, Copyright McMillan

    Publishers Ltd., 2006)

    264 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    13/31

    b) Functionalization of the zeolite seeds by anchoring organosilanes on

    their external surface. These organosilanes form a surface passivating layer

    that avoids or reduces meaningfully the nanounit aggregation.

    c) Crystallization to complete the zeolitization of the functionalized

    protozeolitic units. This stage can be performed hydrothermally under

    standard zeolite crystallization temperatures (e.g. 1701C).

    d) Calcination to remove the silanization and the structure directingagents rendering accessible both micropores and mesopores in the hier-

    archical zeolite.

    As it can be appreciated in Fig 7, the hierarchical ZSM-5 so obtained is

    made up by 200 400 nm aggregates formed by ultrasmall ZSM-5 units

    whose size varies within the 5 10 nm range depending on the synthesis

    conditions. Interestingly, the lattice fringes share their orientation among

    the different nanounits which indicates that they are not really independent

    nanocrystals but a significant degree of intergrowth exists between them.

    This is an important aspect as it implies that the actual size of the crystallinedomains is quite larger than the size of the nanounits, which is expected to

    improve the stability of these materials compared to nanozeolites formed

    solely by independent nanocrystals.

    On the other hand,27Al-MAS NMR analyses point out that most of

    the aluminium atoms are incorporated inside the framework of the

    as-synthesized zeolites. Additionally, 1D and 2D NMR analyses provided

    evidences about the location of the TPA and silanization agent (pheny-

    laminopropyltrimethoxysilane, PHAPTMS) moieties bearing out that they

    Fig. 7 TEM images of hierarchical ZSM-5 obtained by crystallization of silanized seeds.(Reprinted with permission from ref. 8, Copyright Royal Society of Chemistry, 2008)

    Catalysis, 2011, 23, 253283 | 265

  • 7/31/2019 9781849732772-00253

    14/31

    are located inside the zeolite nanounits and grafted over their external

    surface, respectively. This can be appreciated clearly in the schematic

    diagram shown in Fig. 8.

    The effects induced by the seed silanization treatment heavily depend

    on the chosen synthesis variables, in particular upon the nature and

    concentration of the seed silanization agent and on the precrystallization

    stage. Different bulky organosilanes have been used as seed silanization

    agents: octadecyltrimethoxysilane, isobutyltriethoxysilane, 3-aminopropyl-

    trimethoxysilane and phenylaminopropyltrimethoxysilane.8 The former

    (octadecyltrimethoxysilane) was only anchored onto the protozeolitic units

    in a reduced extent (1.3 wt %), being not successful in developing meso-

    porosity. In contrast, both isobutyltriethoxysilane and 3-aminopropyl-

    trimethoxysilane led to materials having mesopores with a size of 8.0 nm

    and an external surface area of around 200 m2 g1. But the most interesting

    result has been achieved with phenylaminopropyltrimethoxysilane, leading

    to materials having 2.0 3.0 nm mesopores, their walls being formed by the

    smallest ZSM-5 nanounits (5 10 nm). Regardless of the organosilane used,

    the crystallinity of all the samples has been proved by both XRD patterns,which indicate the absence of amorphous material, and FTIR, that shows

    the appearance of the 550 cm 1 band, typical of the asymmetric stretching

    mode of five membered rings present in ZSM-5 zeolite.65

    The concentration of the seed silanization agent is also an important

    variable. Thus, an optimum of 12 mol%, referred to the total silica content,

    appears to exist with phenylaminopropyltrimethoxysilane as seed silaniza-

    tion agent.61 On the other hand, the precrystallization step for the

    formation of the nanozeolite precursors containing TPA occluded is

    completely necessary, since in its absence the synthesis of an amorphousmaterial took place. In addition, the temperature used in the pre-

    crystallization step influenced deeply the BET surface areas of the samples:

    698m2 g 1 at 901C and 785m2 g1 at 401C. The latter value is specially

    remarkable since it represents the highest reported BET surface area for a

    hierarchical ZSM-5 zeolite.61

    Fig. 8 Location of TPA and PHAPTMS moieties in hierarchical ZSM-5 obtained from sila-nized seeds. (Reprinted with permission from ref. 61, Copyright American Chemical Society, 2009)

    266 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    15/31

    The seed silanization approach can be considered as a general route for

    synthesizing hierarchical zeolites since it is not bound to just ZSM-5 zeolite.

    Hierarchical Beta,59 mordenite66 and TS-160 zeolites have been prepared by

    the seed silanization procedure. In the case of hierarchical Beta, BET

    surface areas as high as 857 m2 g 1 have been reported using phenylami-

    nopropyltrimethoxysilane as seed silanization agent. A great proportion of

    supermicropores (1.7 1.8 nm) were formed in addition to mesopores inhierarchical Beta zeolite.

    Another method based on the use of organosilanes has been devised by

    Pinnavaia et al.67 In this case, the procedure consists of employing a silane

    functionalized polymer, as shown in the schematic diagram depicted in

    Fig. 9. During the nucleation of the zeolitic entities, a silylated polymer

    (e.g. silylated polyethyleneimine oxide) is grafted over their external surface

    hindering the nanozeolite aggregation but not avoiding the zeolite crystal

    formation. This leads to the generation of an intracrystalline polymer

    network inside the zeolite structure. Finally, the polymer is removed bycalcination forming a zeolite with intracrystalline mesopores. Fig. 10 shows

    the N2 adsorption isotherm at 77 K of hierarchical zeolite (MSU-MFI),

    obtained using this approach, compared with a reference ZSM-5 sample

    wherein the differences in the porosity between both samples are evident (see

    the steep adsorption of MSU-MFI in the p/p0 range of 0.15 0.6). The

    polymers employed had molecular weights within the range 600 25000

    leading to mesopore sizes around 2.0 and 3.0 nm, respectively. Silylation has

    Fig. 9 Mechanism of formation of a mesoporous zeolite using a silylated polymer. (Reprintedwith permission from ref. 67, Copyright Wiley-VCH Verlag GmbH & Co KGaA, 2006)

    Catalysis, 2011, 23, 253283 | 267

  • 7/31/2019 9781849732772-00253

    16/31

    probed to be essential for the success of the procedure since the use of a

    non-silylated polyethyleneimine was not effective. In addition, the degree of

    silylation is also important for the generation of mesopores. The procedure

    is not limited to ZSM-5 zeolite but it has also been applied to the synthesis

    of hierarchical zeolite Y.

    These organosilanes based methods have the advantage of the possibility

    of tailoring the mesopore size by choosing the adequate synthesis

    conditions. This is particularly true for the seed silanization and theamphiphilic organosilane strategies. In addition, although organosilanes

    may be expensive reagents, the employed amounts are relatively small and

    their removal by calcination takes place at temperatures (300 5001C) low

    enough for not damaging the structure of the hierarchical zeolite.

    2.6 Other methods

    A variety of other methods, that cannot be classified within any of the

    previous categories, has been reported for the preparation of hierarchical

    zeolites. Just as interesting examples, three of them are described in thissection. Zhang et al.68 employed bacteria (Bacillus subtilis) as templates in

    the synthesis of microporous/macroporous MFI hierarchical zeolites. The

    zeolite nanounits grow occupying the void spaces of the bacterial template,

    so the formation of a microporous/macroporous framework occurs after

    calcination. The final products are fibres formed by 0.5 mm channels and

    Fig. 10 N2 adsorption isotherm at 77 K of a hierarchical ZSM-5 (MSU-MFI), prepared usinga silylated polymer, and of a conventional ZSM-5. (Reprinted with permission from ref. 67,

    Copyright Wiley-VCH Verlag GmbH & Co KGaA, 2006)

    268 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    17/31

    100 nm thick walls of silicalite. Other curious method to prepare

    hierarchical zeolites is the one based on the usage of the leaves of the plant

    Equisetum arvense as template.69 The presence of silica in the plant pro-

    motes zeolite crystallization leading to a microporous/mesoporous zeolite

    (roughly 0.79 cm3 g1 of intracrystalline mesoporosity). On the other hand,

    inorganic compounds have also been used as templates for the preparation

    of hierarchical zeolites. In this regard, nanosized CaCO3 can act as templatefor the preparation of hierarchical silicalite-1. After the synthesis, the car-

    bonate can be easily dissolved by acid treatment, developing the secondary

    porosity (pore size within 50100 nm). In this procedure, the presence of

    hydroxyl groups over the surface of the nanosized CaCO3 is essential to

    interact with the silanol groups of the silica leading to the encapsulation of

    the salt inside the zeolite crystal. Thus, if the hydroxyl groups are protected

    by fatty acids leading to hydrophobic CaCO3, the synthesis is unsuccessful

    yielding conventional zeolites instead of the hierarchical ones.70

    3 Singular features of hierarchical zeolites

    The presence of a bimodal micro/mesoporous structure provides hier-

    archical zeolites with an improved accessibility to the active sites, which in

    many cases influences positively on their catalytic activity compared to

    conventional zeolites with micrometer crystal sizes. Hierarchical zeolites

    possess a collection of singular properties that are commented henceforth

    according to the following order:

    Improved surface area.

    Increase in mass transfer rates.

    Resistance to deactivation.

    High dispersion of active phases.

    3.1 Improved surface area

    For conventional zeolites, having crystal sizes in the micrometer range, the

    proportion of external surface area is usually negligible, i.e. the BET surface

    area corresponds almost completely with the surface area associated to the

    micropores. However, in the case of hierarchical zeolites the presence of

    mesoporosity implies that a great part of the surface area is related to thelatter, while a reduction is usually observed in the micropore surface area

    compared with the standard zeolites. This is an essential aspect since the

    mesopore surface area is not sterically hindered, as it occurs with the surface

    area of the micropores, being capable of adsorbing and interacting with

    bulky compounds.

    Other interesting fact is that in most cases hierarchical zeolites have been

    reported to present enhanced BET surface area, which also occurs for

    nanozeolites. This fact can be interpreted as a result of the strongly

    restricted adsorption that takes place within the zeolite micropores, whereasadsorption on the external/mesopore surface area is not so limited. Thus, in

    the case of nanocrystalline ZSM-5, a good correlation has been found

    between the enhancement of the BET surface area and the reduction in

    the size of the nanocrystals.9 A similar trend is expected to be also valid

    for hierarchical zeolites, with an increase of the BET surface area as the

    Catalysis, 2011, 23, 253283 | 269

  • 7/31/2019 9781849732772-00253

    18/31

    contribution of the secondary mesoporosity is more pronounced. Accord-

    ingly, the BET surface area can be used as a parameter for comparison

    among the different afore commented synthesis procedures of hierarchical

    zeolites.

    Fig. 11 illustrates the ranges, as well as the average values, of BET

    surface areas reported for hierarchical ZSM-5 zeolite synthesized with the

    most important methods previously discussed. It can be observed that thestrongest effect on the BET surface area is obtained with the seed silani-

    zation strategy61 reaching values close to 800 m2 g1. Moreover, this

    method allows the BET surface area to be adjusted in a wide range by

    changing the synthesis conditions. Significant enhancements in the BET

    surface area of ZSM-5 are also obtained with the use of silane-containing

    polymers,67 desilication29 and amphiphilic organosilanes,54 although with

    clearly lower values (up to around 600 m2 g1). This is a remarkable

    achievement as the BET surface area of a conventional ZSM-5 is about

    400m

    2

    g

    1

    . In contrast, lower BET surface areas are exhibited by thehierarchical ZSM-5 samples obtained using different types of carbon

    materials as hard-templates. Thus, the use of carbon blacks71 just gives rise

    at most to a BET surface area of 418m2 g 1, very close to the value

    corresponding to a conventional ZSM-5. These results indicate that the

    1 2 3 4 5 6 7 80

    100

    200

    300

    400

    500

    600

    700

    800

    900

    1000

    Carbonnanotubes

    Carbon

    aerogels

    Carbon

    blacks

    Mesostructured

    carbons

    Amphiphilic

    organosilane

    Desilication

    Silane

    polymer

    Seed

    silanization

    BETsur

    facearea(m2g-1)

    Method of preparation

    Fig. 11 Ranges of BET surface area values reported for hierarchical ZSM-5 samples preparedaccording to different synthesis methods (seed silanization,8,58,61 silane polymer,79

    desilication,26,30 amphiphilic organosilanes,76,77 mesostructured carbons,46,47 carbonblacks,42,70 carbon aerogels45 and carbon nanotubes44).

    270 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    19/31

    degree of modification of the zeolite textural properties depends strongly on

    the type of method employed for the generation of the hierarchical porosity.

    Those methods based on the incorporation of organosilanes into the

    synthesis gel appear to cause a stronger and more effective modification of

    the textural properties of the zeolite in terms of surface area than the route

    using carbon templates.

    The improvement of the BET surface area, as well as the presence ofa high share of non-microporous surface area, opens the possibility for

    the functionalization of hierarchical zeolites with different agents.

    Thus, hierarchical ZSM-5 has been successfully functionalized with

    3-aminopropylsilane, reaching the incorporation of 25% more organic

    groups than in the case of using an ordered mesoporous SBA-15 support.

    The organic-functionalized hierarchical zeolites show the added advantages

    of their high hydrothermal stability and reusability. This has been proved in

    the Pd-catalyzed Sonogashira cross coupling reaction72 of terminal alkynes

    with chlorobenzenes under the presence of Na2CO3. In this reaction,the crystallinity of the hierarchical zeolite made possible their reuse since

    conversion just decreased slightly (from 96 to 91%) after 5 reaction cycles.

    In contrast, for SBA-15 the conversion values dropped from 84 to 38% due

    to the lower hydrothermal stability of this material compared to the

    mesoporous zeolite.

    3.2 Increase in mass transfer rates

    The existence of an interconnected network of mesopores and micropores is

    expected to favour the intracrystalline mass transfer phenomena in hier-

    archical zeolites. The mesopores allows a faster diffusion of the reacting

    molecules towards the active sites leading to enhanced kinetics. In the same

    way, the products may diffuse faster from the active sites reducing the extent

    of secondary reactions which alters the obtained selectivity. This fact has

    been nicely shown by Christensen et al.73 in the catalytic alkylation of

    benzene with ethene. Fig. 12.a) shows the activity so obtained (TOF values)

    versus the temperature, while Fig 12.b) illustrates the selectivity towards

    ethylbenzene versus benzene conversion. These results indicate clearly

    that the activity of the mesoporous zeolite is always higher than that of

    the conventional ZSM-5 in the whole range of temperatures studied(583 643 K). Fig 12.b) evidences that selectivities towards ethylbenzene

    attained over the hierarchical ZSM-5 are higher than the values obtained

    over the conventional ZSM-5 catalyst. The occurrence of mass transport

    constraints when using a conventional ZSM-5 zeolite with regard to the

    hierarchical zeolite was concluded from the lower values of the activation

    energies obtained with the former (59 vs 77 kJ/mol, respectively). Con-

    sequently, this enhanced mass transport rates drives to higher benzene

    conversions for the hierarchical zeolite as well as to larger selectivities

    towards ethylbenzene. The latter effect has been ascribed to the shorterdiffusion path present in the hierarchical zeolite which suppresses the

    secondary ethylation reaction.

    Further analyses confirmed that the mass transport rates of both benzene

    and ethylbenzene are diffusion limited over the conventional zeolite

    since the value of the Thiele modulus was higher than 0.1, while over the

    Catalysis, 2011, 23, 253283 | 271

  • 7/31/2019 9781849732772-00253

    20/31

    hierarchical zeolite was much lower. In addition, experiments of diffusion

    of isobutane point out that the effective diffusivity is three times higherover the hierarchical zeolite.74 These results have been obtained using

    hierarchical zeolites prepared according to the carbon templating route.

    Similar conclusions have been drawn by Groen et al.75 with hierarchical

    zeolites synthesized using the desilication method. Thus, a two-order of

    magnitude improvement has been denoted in the diffusion of neopentane

    inside desilicated ZSM-5, due to the shorter diffusion path length and the

    presence of an accessible network of mesopores.

    3.3 Resistance to deactivationBesides the above described improvement of the intraparticle mass trans-

    port, hierarchical zeolites show another key feature when compared with

    conventional ones: increased catalyst lifetime due to their high resistance to

    deactivation. The large size of the mesopores hinders the occurrence of

    micropore blocking phenomena by coke, typically found in many catalytic

    Fig. 12 Benzene alkylation with ethane over conventional and mesoporous ZSM-5 catalysts:a) TOF; b) selectivity towards ethylbenzene. (Reprinted with permission from ref. 73,Copyright American Chemical Society, 2003)

    272 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    21/31

    reactions when using conventional zeolites. In this regard, coke precursors

    can diffuse out of the hierarchical zeolite easily due the presence of

    mesopores, which avoids or delays their transformation into coke deposits,

    slowing down the zeolite deactivation. This fact has been clearly shown by

    Ryoo et al.76 for a hierarchical MFI zeolite, synthesized using an amphipilic

    organosilane, in three different reactions: isomerisation of 1,2,4

    trimethylbenzene, cumene cracking and esterification of benzyl alcohol withhexanoic acid. Fig. 13 illustrates the deactivation curves obtained in these

    three reactions wherein the superb performance of the hierarchical

    MFI zeolite over conventional MFI zeolite and ordered mesoporous

    Al-MCM-41 can be appreciated.

    In the isomerisation of 1,2,4-trimethylbenzene (Fig. 13.a), the conversion

    obtained over conventional MFI catalyst drops quickly from 30 to 8% after

    30 min, while for Al-MCM-41 falls from 13 to 5%. In contrast, the con-

    version over hierarchical MFI slowly decreases from 25 to 17% after much

    longer reaction times (180 min). In cumene cracking (Fig. 13.b), the con-version obtained over Al-MCM-41 is below 10% along the whole reaction

    time (140 min). Hierarchical MFI and conventional MFI initially show al-

    most complete conversion (B 95%) for this reaction but the evolution of

    the activity with the time shows rather different trends between them.

    Thus, hierarchical MFI keeps its high conversion (W 90%) while for the

    conventional MFI conversion drops to around 50% after 140 min.

    Likewise, as it can be observed in Fig. 13.c), the benzyl alcohol conversion

    over conventional MFI and Al-MCM-41 drops after 5 reaction cycles from

    22% and 81%, respectively, to almost zero. In contrast, over hierarchical

    MFI only a slight decrease from 90 to 80% in conversion is observed after 5

    reaction cycles.

    The slow deactivation undergone by hierarchical zeolites has been mainly

    ascribed to the presence of mesopores that favour a fast diffusion of the

    coke precursors out of the zeolite, avoiding their accumulation inside the

    micropores which would cause pore blocking phenomena. Likewise, the

    difference in performance between the ordered mesoporous Al-MCM-41

    materials and the hierarchical ZSM-5 zeolite (both of them possess

    mesopores) has been assigned to the stronger acidity of the hierarchical

    ZSM-5 as well as to the higher concentration of Al sites in the Al-MCM-41.The latter proposal takes into account that the Al sites are relatively close in

    Al-MCM-41, which can promote the formation of polymeric coke

    precursor species.

    Another example of resistance to deactivation over hierarchical MFI has

    been denoted in the methanol to hydrocarbon reaction (MTH).77 The

    lifetime of the catalyst has been observed to increase more than three

    times over the hierarchical ZSM-5 with regard to conventional ZSM-5.

    This effect has been also related to the deposition of coke on the mesopores

    of the hierarchical zeolite, while for the conventional sample it takes placepreferentially inside the zeolite micropores having a stronger deactivating

    effect by pore blockage. This is nicely shown in Figs 14.a and 14.b, re-

    spectively, wherein the coke content as well as the hydrocarbon production

    obtained over conventional (ZST-12) and mesoporous ZSM-5 (OSD-5) are

    shown.

    Catalysis, 2011, 23, 253283 | 273

  • 7/31/2019 9781849732772-00253

    22/31

    Fig.

    13

    Deactivationperformanceofhierarchical

    MFIindifferentreactions:a)1

    ,2,4,

    TMBisomerisation,

    b)cumenecrackingandc)benzylalc

    oholesterificationwith

    hexanoicacid

    .(Reprintedwithpermissionfromref.76

    ,CopyrightRoyalSocietyo

    fChemistry,

    2006)

    274 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    23/31

    It can be observed that the hydrocarbon production over the con-

    ventional ZSM-5 catalyst drops to practically 0% after a time on stream of

    40 h while over OSD-5 it is still around 50% after 130 h, indicating the

    longer lifetime of the mesoporous ZSM-5. In addition, both Figs 14.a and

    14.b provides information of the coke content, distinguishing between in-ternal coke (inside the micropores) and external coke (mainly within the

    mesopores). It is clear that the internal coke is formed before and in a far

    greater amount over the microporous ZSM-5 than over hierarchical ZSM-5,

    wherein the coke is mostly external (W80%). The internal coke possesses a

    stronger deactivation effect, which explains the fast decrease observed in the

    conversion with the conventional ZSM-5.

    3.4 High dispersion of active phases

    The presence of mesoporosity in hierarchical zeolites offers great oppor-tunities for the preparation of bifunctional catalysts through the in-

    corporation of other active phases, such as metals and metal oxides, in close

    contact and with an improved interaction with the zeolite support. The

    presence of mesopores is expected to lead towards better dispersions of the

    active phase. This has been shown by Christensen et al.78 when impreg-

    nating both mesoporous and conventional zeolites with Pt (B2 wt %), PtSn

    alloy (B1 wt %) and a b- MoC2 carbide (B10 wt %). For the conventional

    zeolites the metals are deposited mainly over the outer surface, frequently as

    large particles according to EDS scan measurements associated to the TEMimages. In contrast, the same technique indicates that for the mesoporous

    zeolite the metal nanocrystals are evenly placed within the mesopores.

    Moreover, in this case, large metal particles are not observed. This result is

    not only ascribed to the higher BET surface area and pore volume of the

    hierarchical zeolite but to the occurrence of a much higher amount of lattice

    Fig. 14 Evolution along the time on stream of the coke and the hydrocarbon production in themethanol conversion: a) microporous ZSM-5 (ZST-12) and b) mesoporous ZSM-5 (OSD-5).(Reprinted with permission from ref. 77, Copyright Elsevier, 2010)

    Catalysis, 2011, 23, 253283 | 275

  • 7/31/2019 9781849732772-00253

    24/31

    defect sites than in the case of conventional zeolites. These defect sites are

    assumed to act as preferred deposition points for the active phase, leading to

    improved dispersions. Additional examples of the benefits derived from

    using hierarchical zeolites as supports of different active phases are de-

    scribed in the next section dealing with the catalytic applications of these

    materials.

    4 Catalytic applications

    The remarkable and singular properties shown by hierarchical zeolites have

    brought about the potential catalytic applications of these materials in

    numerous reactions, specially those wherein steric or diffusion limitations

    are encountered. The next paragraphs review the literature works dealing

    with the application of hierarchical zeolites in a variety of reactions, which

    have been classified into three groups (oil refining and petrochemical re-

    actions, fine chemicals reactions and environmental catalysis).

    4.1 Oil refining and petrochemical reactions

    A potentially interesting application of hierarchical zeolites is in petroleum

    refining processes since conventional zeolite catalysts cannot refine about

    20% of a petroleum barrel due to the steric hindrances posed by bulky

    molecules. The application of hierarchical zeolites could diminish this

    amount increasing meaningfully the profitability of the refining. In addition,

    it is expected that the share of gasoline and light alkenes might also be

    enhanced by the application of hierarchical zeolite catalysts.

    This approach has been tested by Pinnavaia et al.79 that studied the

    cracking of gas-oil over a mesoporous zeolite (MSU-MFI) and compared it

    with a conventional ZSM-5. These authors observed increased conversions

    over mesoporous MSU-MFI, accompanied by higher yields of gaseous

    products (LPG), gasoline and light cycle oil (LCO) and lower amounts of

    coke. In addition, much more light olefins were also detected.

    Another petroleum-related application of hierarchical zeolites is the

    desulfuration of gasoline and diesel fuels, which is a very important process

    in order to comply with the target of reducing their sulphur content below

    10 ppm.80 Thereby, 4,6, dimethyldibenzothiophene, a highly refractorysulphur-containing organic compound, has been hydrodesulfurized (HDS)

    over Pt, Pd and Pt-Pd mesoporous ZSM-5 zeolite (total metal content of 0.5

    wt %) leading to higher sulphur removal efficiency than metal/microporous

    zeolite or metal/g-Al2O3. Fig. 15 shows the HDS conversion attained over

    the different Pd containing catalysts: mesoporous Na(90%)/H (10%)-

    ZSM-5 (MNZ-5), conventional Na-ZSM-5 (NZ-5) and g-Al2O3. It can be

    appreciated the remarkable performance of Pd/mesoporous Na/H-ZSM-5

    reaching 86% conversion with only 3% for conventional Pd/Na-ZSM-5 and

    21% for Pd/g-Al2O3. This result has been ascribed to a proper combinationof acidity and mesoporosity in the hierarchical zeolite. In line with this

    result, Pd/mesoporous Beta hydrodesulfurized 4,6 dimethyldibenzothio-

    phene at 2501C under 62 bar of hydrogen better than Pd/Al-MCM-41 (51 vs

    35%) due to the higher acidity of the zeolite.81 Additionally, Pd/mesopor-

    ous Beta has shown to be more active in the hydrogenation, at 2501C and 40

    276 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    25/31

    bar of hydrogen, of the bulky aromatic pyrene than Pd/conventional Beta,

    Pd/Al-MCM-41 and Pd/g-Al2O3. The difference in performance with the

    Pd/conventional Beta lies in its larger mesopore volume, whereas regarding

    the other two catalysts this result has been also related to its greater

    acidity.82

    Other reactions of interest are the aromatization and isomerisation of

    1-hexene wherein hierarchical zeolites, obtained by desilication, showed

    enhanced stability. Thus, after a time on stream of 14 h at 3501C, the

    hierarchical zeolite prepared by desilication with 0.5 M NaOH shows

    selectivity towards aromatics of 19.1% while the selectivity values obtained

    with the conventional ZSM-5 drop to 5.1%.83 Likewise, in butene aroma-

    tization at 3501C it has been found that after a time on stream of 34 h, the

    conversion over hierarchical ZSM-5 remains rather stable at 99% while the

    conversion over conventional HZSM-5 drops to 93%.84 This performance

    has been ascribed to a lower deposition of coke inside the micropores,reducing the extent of micropore blocking.

    Hierarchical Mo/HZSM-5 also shows enhanced selectivity to aromatics

    due to a larger tolerance to coke in the catalytic dehydroaromatization

    of methane.85 Hence, after 720 min of reaction at 1003 K, the methane

    conversion was 11% with hierarchical Mo/HZSM-5 while the conversion

    over conventional Mo/HZSM-5 dropped to 3.9%. In addition, the select-

    ivity towards benzene was 68% over the hierarchical zeolite instead of 37%

    over the conventional catalyst.

    In the case of 1-butene isomerisation at 701C, an enhanced activity hasbeen found over H3PW12O40 supported on mesoporous silicalite-1 com-

    pared with H3PW12O40 supported on a conventional silicalite-1.71 These

    differences are specially remarkable since the initial conversion was 72% for

    the former and less than 1% for the latter. Alkylation of benzene with

    ethene over microporous ZSM-5 has also been improved over mesoporous

    Fig. 15 HDS conversion of 4,6, dimethyldibenzothiophene obtained over different Pd-containingcatalysts. (Reprinted with permission from ref. 80, Copyright Wiley-VCH Verlag GmbH & CoKGaA, 2008)

    Catalysis, 2011, 23, 253283 | 277

  • 7/31/2019 9781849732772-00253

    26/31

    ZSM-5, as it was previously commented in point 3.b).73 Mesoporous

    mordenite33 have also been tested for the same reaction at 438 K. This

    catalyst brings about a 5-6 fold increased production of ethylbenzene

    compared to conventional mordenite. The explanation to this behaviour has

    been the presence of mesopores which reduces the deactivation of the

    catalyst.

    On the other hand, confocal fluorescence microscopy86

    has been used tostudy the oligomerization of 4-methoxystyrene. The bimodal porous

    structure of the mesoporous zeolite gives rise to the preferential formation

    of dimeric carbocations due to the shortening of the micropore diffusion

    path, precluding the appearance of higher oligomers. This result is clearly

    bound up to the above indicated resistance to deactivation shown by

    hierarchical zeolites.

    4.2 Fine chemistry reactions

    Hierarchical ZSM-5 has been tested in various reactions involving bulkymolecules with potential application in Fine Chemistry, like the protection

    of benzaldehyde with pentaerythritol, condensation of benzaldehyde with

    2-hydroxyacetophenone or the esterification of benzyl alcohol with

    hexanoic acid.87 High activities were observed for these three reactions over

    hierarchical ZSM-5. In addition, by means of experiments of dealumination

    of the mesopore walls with tartaric acid, it was concluded that bulky mol-

    ecules react at the Al sites placed over the mesopore walls.

    Hierarchical ZSM-5 has been evaluated in the Friedel-Crafts acylation of

    anisole at 1201C with either acetyl chloride or acetic anhydride as acylating

    agent and compared with nanocrystalline HZSM-5 and Al-MCM-41.88

    Superior conversions were achieved over the hierarchical ZSM-5 due to the

    right combination of improved accessibility provided by the mesopores and

    the high acidity and crystallinity of the zeolite.

    Mesoporous sodalite has been also successfully applied in base catalyzed

    reactions such as Knoevenagel condensation, Claisen-Schmidt conden-

    sation and acetonyl acetone cyclization.56 This material shows enhanced

    activity with regards to CsNaX and KAl-MCM-41 catalyst. Thus, in

    Knoevenagel condensation of 4-isopropylbenzaldehyde with ethylcyanoa-

    cetate at 353 K, the conversion obtained with K-mesoporous sodalite was78%, while KAl-MCM-41 and CsNaX gave 45 and 36% conversion values,

    respectively. This result has been related to the basic sites present in the

    mesopores of the hierarchical zeolite.

    Pd exchanged mesoporous sodalite and NaA zeolite have been also

    applied for different aryl coupling reactions (Suzuki, Heck and Sonoga-

    shira) involving bulky substrates.89 These hierarchical catalysts have been

    reported to show high activity and reusability avoiding the usual problem of

    Pd leaching and agglomeration provided that the reactions are carried out

    under air. In this regard, the mesoporous zeolite stabilizes the Pd2

    specieseliminating the formation of unwanted agglomerates. Mesoporous MFI

    zeolite show much higher activity (98%) than conventional ZSM-5 (3.9%),

    Al-MCM-41 (25%) and ZSM-5 seed assembled mesoporous materials

    (SAM, 64%) in the synthesis of jasminaldehyde (all of them were prepared

    with Si/Al=20). In addition, the selectivity obtained with the mesoporous

    278 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    27/31

    zeolite was outstanding (98%), since the other materials presented values

    below 80%. These results have been ascribed to the highly mesoporous

    structure and strong acidity of the hierarchical MFI.54 These same authors

    also tested the synthesis of vesidryl with the same catalysts obtaining again

    superb results with the mesoporous MFI zeolite.

    Other interesting application of hierarchical zeolites as catalysts is in

    epoxidation reactions using Ti-containing hierarchical zeolites. Thus, in theepoxidation of cyclohexene39 with hydrogen peroxide, a yield of products

    ten times higher has been detected over mesoporous TS-1 compared to

    conventional TS-1. Likewise, hierarchical TS-1, obtained through the seed

    silanization route, has been employed as catalyst in the epoxidation of

    1-octene60 at 1001C with a bulky oxidant (TBHP, hindered to enter the

    zeolite micropores), leading to much higher conversion (42% vs 5% of

    conventional TS-1) with 100% epoxide selectivity and TBHP efficiency

    higher than 90%. This result is specially relevant as it has opened the

    possibility of using organic hydroperoxides as oxidants, instead of solelyhydrogen peroxide, in combination with TS-1 zeolite.

    4.3 Environmental catalysis

    Hierarchical zeolites have been also investigated in a number of reactions

    within the field of environmental catalysis. Among them, the decomposition

    of NO over hierarchical Cu-ZSM-11 and Cu-ZSM-5 deserves special

    mention.90 In this case, the improved accessibility causes an enhanced

    activity of the hierarchical zeolites because of the formation of dimeric and

    oligomeric Cu species within the mesopores instead of the preferential

    formation of monomeric Cu species over conventional Cu-ZSM-11 and

    Cu-ZSM-5. In addition, hierarchical Cu-ZSM-11 was two-fold more active

    than mesoporous Cu-ZSM-5 due to the occurrence of solely straight

    microporous channels, wherein the active sites are preferentially located.

    Likewise, desilicated Fe-ZSM-5 samples have shown enhanced N2O

    decomposition activity.91 This has been assigned to the fully exchange of

    iron in desilicated ZSM-5 samples due to its enhanced accessibility without

    formation of iron oxides. In contrast, for large zeolite crystals, iron

    exchange is diffusion controlled and leads to the deposition of inactive iron

    species formed by hydrolysis.One less successful application of hierarchical zeolites has been the

    catalytic pyrolysis of lignocellulose92 from beech wood at 5001C. Hier-

    archical Beta zeolite yields less liquid bio-oil and more coke and char than

    Al-MCM-41, leading to increased production of aromatics and PAH due to

    the stronger acidity of their acid sites.

    On the other hand, hierarchical zeolites have shown to be remarkable

    catalysts for the cracking of polyolefins. The latter are bulky substrates

    wherein an easy access to the acid sites leads towards higher activities. Thus,

    a bulky polyolefin such as polypropylene has been cracked at 3601C overhierarchical Beta and ZSM-5 zeolites, obtained both by a seed silanization

    procedure, being compared to the conventional catalysts.58 The improved

    accessibility of these hierarchical catalysts for bulky polypropylene

    molecules caused a four-fold greater conversion than conventional zeolites.

    A similar result has been obtained in the catalytic cracking of low density

    Catalysis, 2011, 23, 253283 | 279

  • 7/31/2019 9781849732772-00253

    28/31

    polyethylene at 3401C using a plastic/catalyst mass ratio=100 with differ-ent hierarchical ZSM-5 samples prepared also by a seed silanization

    method. In this case, while a TOF value of just 0.018 s1 is achieved over

    conventional ZSM-5, the hierarchical ZSM-5 samples leads towards TOF

    values far higher, within the 0.3840.901 range, depending on the combin-

    ation of mesoporosity/acidity of the hierarchical sample.61

    5 Concluding remarks

    The field of synthesis and applications of hierarchical zeolites has been

    extremely fruitful in the last decade. A wide range of synthesis strategies

    have been developed that successfully allowed these zeolites to be prepared

    with bimodal microporous/mesoporous structure. In addition, the number

    of potential applications is growing every year, mainly as catalysts in a large

    variety of reactions. In this respect, a bright future can be envisaged for

    hierarchical zeolites, specially in transformations dealing with bulky sub-strates or suffering of strong deactivation by pore blockage.

    The development of hierarchical zeolites has obliged to change the con-

    ventional vision of zeolites as just pure microporous materials with shape

    selectivity properties. At present, the occurrence of mesoporosity is an

    added feature to a zeolitic material which markedly improves its properties:

    enhanced textural properties, faster intraparticle transport, reduction of

    steric and diffusion constraints, improved dispersion of active phases, better

    resistance to deactivation and higher catalytic performance in many re-

    actions. Interestingly, these improvements in many cases do not imply ne-cessarily a decrease in the selectivity exhibited by the zeolite.

    The availability of a large variety of synthesis methods, which for sure

    will be optimized and enlarged in the future, makes possible to tailor and

    define the contribution and features of the mesoporosity present in hier-

    archical zeolites. Consequently, the research in the field of hierarchical

    zeolites is expected to grow, specially when dealing with the following goals:

    Better control of the mesopore size and distribution.

    Optimization of the ratio mesopore/micropore surface to achieve the

    best catalyst performance.

    Deeper characterization of the nature and strength of the acid sites

    present in the mesopores.

    Increase in the number of catalytic applications, with a strong focus on

    reactions requiring bifunctional catalysts, prepared by incorporation of

    different active phases to hierarchical zeolites.

    References

    1 E. M. Flanigen, Stud. Surf. Sci. Catal., 2001, 137, 11.

    2 G. Bellusi and M. S. Rigutto, Stud. Surf. Sci. Catal., 2001, 137, 911.3 M. A. Uguina, D. P. Serrano, R. Van Grieken and S. Venes, Appl. Catal. A,

    1993, 99(2), 97.

    4 J. Aguado, J. L. Sotelo, D. P. Serrano, J. A. Calles and J. M. Escola, Energy &

    Fuels, 1997, 11, 1225.

    5 M. Hartmann, Angew. Chem. Int. Ed., 2004, 43, 5880.

    280 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    29/31

    6 C. T. Kresge, M. E. Leonowicz, W. J. Roth, J. C. Vartuli and J. S. Beck, Nature,

    1992, 359, 710.

    7 D. Zhao, J. Feng, Q. Huo, N. Melosh, G. H. Fredrickson, B. F. Chmelka and

    G. D. Stucky, Science, 1998, 279, 548.

    8 D. P. Serrano, J. Aguado, J. M. Escola, J. M. Rodriguez and A. Peral, J. Mat.

    Chem., 2008, 18, 4210.

    9 J. Aguado, D. P. Serrano, J. M. Escola and J. M. Rodriguez, Micropor.

    Mesopor. Mater., 2004, 75(1-2), 41.

    10 L. Tosheva and V. P. Valtchev, Chem. Mater., 2005, 17, 2494.

    11 A. Corma, V. Fornes, S. B. Pergher, Th. L. M. Maesen and J. G. Buglass,

    Nature, 1998, 396, 353.

    12 K. G. Strohmaier and D. E. W. Vaughan, J. Am. Chem. Soc., 2003, 125, 16035.

    13 A. Corma, M. J. Diaz-Caban as, J. L. Jorda , C. Martinez and M. Moliner,

    Nature, 2006, 443, 842.

    14 K. Egeblad, C. H. Christensen, M. Kustova and C. H. Christensen, Chem.

    Mater., 2008, 20, 946.

    15 J. Perez Ramirez, C. H. Christensen, K. Egeblad, C. H. Christensen and J. C.

    Groen, Chem. Soc. Rev., 2008, 37, 2530.16 Fluid catalytic cracking V: Materials and technological innovations, ed. M. L.

    Ocelli, P. OConnor, Elsevier, Amsterdam 2001.

    17 R. Szostak, Stud. Surf. Sci. Catal., 2001, 137, 261.

    18 G. T. Kerr, J. Phys. Chem., 1968, 72, 2594.

    19 B. Sulikowski, J. Phys. Chem., 1993, 97, 1420.

    20 C. J. Van Oers, W. J. J. Stevens, E. Bruijn, M. Mertens, O. I. Lebedev, G. Van

    Tendeloo, V. Meynen and P. Cool, Micropor. Mesopor. Mater., 2009, 120, 29.

    21 A. H. Janssen, A. J. Koster and K. P. de Jong, Angew. Chem. Int. Ed., 2001,

    40(6), 1103.

    22 P. Kortunov, S. Vasenkov, J. Karger, R. Valiullin, P. Gottschalk, M. F. Elia,M. Perez, M. Stocker, B. Drescher, G. McElhiney, C. Berger, R. Glaser and J.

    Weitkamp, J. Am. Chem. Soc., 2005, 127, 13055.

    23 M. Ogura, S. Shinomiya, J. Tateno, Y. Nara, E. Kikuchi and M. Matsukata,

    Chem. Lett., 2000, 882.

    24 M. Ogura, S. Shinomiya, J. Tateno, Y. Nora, M. Nomura, E. Kikuchi and M.

    Matsukata, Appl. Catal. A, 2001, 219, 33.

    25 T. Suzuki and T. Okuhara, Micropor. Mesopor. Mater., 2001, 43, 83.

    26 J. C. Groen, J. C. Jansen, J. A. Moulijn and J. Perez Ramirez, J. Phys. Chem. B,

    2004, 108, 13062.

    27 J. C. Groen, L. A. A. Peffer, J. A. Moulijn and J. Perez Ramirez, Chem Eur. J.,2005, 11, 4983.

    28 J. C. Groen, T. Bach, U. Ziese, A. M. Paulaime-van Donk, K. P. De Jong, J. A.

    Moulijn and J. Perez Ramirez, J. Am. Chem. Soc., 2005, 127, 10792.

    29 S. Abello, A. Bonilla and J. Perez-Ramirez, Appl. Catal A, 2009, 364(1-2), 191.

    30 J. Perez Ramirez, D. Verboekend, A. Bonilla and S. Abello , Adv. Funct. Mater.,

    2009, 19, 3972.

    31 M. S. Holm, S. Svelle, F. Joenssen, P. Beato, C. H. Christensen, S. Bordiga and

    M. Bjorgen, Appl. Catal. A, 2009, 356, 23.

    32 F. T. Starzyk, I. Stan, S. Abello , A. Bonilla, K. Thomas, C. Fernandez, J. P.

    Gilson and J. Perez Ramirez, J. Catal., 2009, 264, 11.33 J. C. Groen, T. Sano, J. A. Moulijn and J. Perez Ramirez, J. Catal., 2007, 251, 21.

    34 X. Wei and P. G. Smirniotis, Micropor. Mesopor. Mater., 2006, 97, 97.

    35 C. J. H. Jacobsen, C. Madsen, J. Houzvicka, J. Schmidt and A. Carlsson, J. Am.

    Chem. Soc., 2000, 122, 7116.

    36 C. Madsen and C. J. H. Jacobsen, Chem Commun., 1999, 673.

    Catalysis, 2011, 23, 253283 | 281

  • 7/31/2019 9781849732772-00253

    30/31

    37 I. Schmidt, A. Boisen, E. Gustavsson, K. Stahl, S. Pehrson, S. Dahl, A.

    Carlsson and C. J. H. Jacobsen, Chem. Mater., 2001, 13, 4416.

    38 A. H. Janssen, I. Scmidt, C. J. H. Jacobsen, A. J. Koster and K. P. De Jong,

    Micropor. Mesopor. Mater., 2003, 65, 59.

    39 I. Schmidt, A. Krogh, K. Wienberg, A. Carlsson, M. Brorson and C. J. H.

    Jacobsen, Chem. Commun., 2000, 2157.

    40 M. Y. Kustova, P. Hasselriis and C. H. Christensen, Catal. Lett., 2004, 96, 205.

    41 X. Wei and P. G. Smirniotis, Micropor. Mesopor. Mater., 2005, 89, 170.

    42 K. Egeblad, M. Kustova, S. K. Klitgaard, K. Zhu and C. H. Christensen,

    Micropor. Mesopor. Mater., 2007, 101, 214.

    43 M. Kustova, K. Egeblad, K. Zhu and C. H. Christensen, Chem. Mater., 2007,

    19(12), 2915.

    44 A. Boisen, I. Schmidt, A. Carlsson, S. Dahl, M. Brorson and C. J. H. Jacobsen,

    Chem. Commun., 2003, 958.

    45 Y. Tao, H. Kanoh and K. Kaneko, J. Am. Chem. Soc., 2003, 125, 6044.

    46 Z. Yang, Y. Xia and R. Mokaya, Adv. Mater., 2004, 16(8), 727.

    47 S. I. Cho, S. D. Choi, J. H. Kim and G. J. Kim, Adv. Funct. Mater., 2004

    14, 49.48 F. S. Xiao, L. Wang, C. Yin, K. Lin, Y. Di, J. Li, R. Xu, D. S. Su, R. Schlogl, T.

    Yokoi and T. Tatsumi, Angew. Chem. Int. Ed., 2006, 45, 3090.

    49 B. T. Holland, L. Abrams and A. Stein, J. Am. Chem. Soc., 1999, 121, 4308.

    50 J. J. Zhao, J. Zhou, Y. Chen, Q. J. He, M. L. Ruan, L. M. Guo, J. L. Shi and

    H. R. Chen, J. Mat. Chem., 2009, 19(41), 7614.

    51 Y. Tao, H. Kanoh and K. Kaneko, Langmuir, 2005, 21, 504.

    52 A. Karlsson, M. Stocker and R. Schmidt, Micropor. Mesopor. Mater., 1999, 27,

    181.

    53 Y. Liu, W. Z. Zhang and T. J. Pinnavia, Angew. Chem. Int. Ed., 2001, 40, 1255.

    54 M. Choi, H. S. Cho, R. Srivastava, C. Venkatesan, D. H. Choi and R. Ryoo,Nature Mater., 2006, 5, 718.

    55 K. Choo, H. S. Cho, L. C. De Menorval and R. Ryoo, Chem. Mater., 2009, 21,

    5664.

    56 G. V. Shanbhag, M. Choi, J. Kim and R. Ryoo, J. Catal., 2009, 264, 88.

    57 K. Suzuki, Y. Aoyagi, N. Katada, M. Choi, R. Ryoo and M. Niwa, Catal.

    Today, 2008, 132, 38.

    58 D. P. Serrano, J. Aguado, J. M. Escola, J. M. Rodriguez and A. Peral, Chem.

    Mater., 2006, 18, 2462.

    59 J. Aguado, D. P. Serrano and J. M. Rodriguez, Micropor. Mesopor. Mater.,

    2008, 115, 504.60 D. P. Serrano, R. Sanz, P. Pizarro and I. Moreno, Chem. Commun., 2009, 11,

    1407.

    61 D. P. Serrano, J. Aguado, G. Morales, J. M. Rodriguez, A. Peral, M.

    Thommes, J. D. Epping and B. F. Chmelka, Chem. Mater., 2009, 21, 641.

    62 A. E. Persson, B. J. Schoeman, J. Sterte and J. E. Otterstedt, Zeolites, 1994, 14,

    557.

    63 B. J. Schoeman and O. Regev, Zeolites, 1996, 17, 447.

    64 V. Nikolakis, E. Kokkoli, M. Tirrell, M. Tsapatsis and D. G. Vlachos, Chem.

    Mater., 2000, 12, 845.

    65 D. T. On and S. Kaliaguine, Angew. Chem. Int. Ed., 2001, 40(17), 3248.66 J. Aguado, D. P. Serrano, J. M. Escola and A. Peral, J. Anal. Appl. Pyrol., 2009,

    85, 352.

    67 H. Wang and T. J. Pinnavaia, Angew. Chem. Int. Ed., 2006, 45, 7603.

    68 B. Zhang, S. A. Davis, N. H. Mendelson and S. Mann, Chem. Commun., 2000,

    781.

    282 | Catalysis, 2011, 23, 253283

  • 7/31/2019 9781849732772-00253

    31/31

    69 V. Valtchev, M. Smaihi, A. C. Faust and L. Vidal, Angew. Chem. Int. Ed., 2003,

    43, 2782.

    70 H. Zhu, Z. Liu, Y. Wang, D. Kong, X. Yuan and Z. Xie, Chem. Mater., 2008,

    20, 1134.

    71 K. Zhu, J. Hu, X. She, J. Liu, Z. Nie, Y. Wang, C. H. F. Peden and J. H. Kwak,

    J. Am. Chem. Soc., 2009, 131, 9715.

    72 D. H. Lee, M. Choi, B. W. Yu and R. Ryoo, Chem. Commun., 2009, 74.

    73 C. H. Christensen, K. Johansen, I. Schmidt and C. H. Christensen, J. Am.

    Chem. Soc., 2003, 125, 13370.

    74 C. H. Christensen, K. Johansen, E. Tornqvist, I. Schmidt, H. Topsoe and C. H.

    Christensen, Catal. Today, 2007, 128, 117.

    75 J. C. Groen, W. Zhu Browuer, S. T. Huynink, F. Kapteijn, J. A. Moulijn and J.

    Perez Ramirez, J. Am. Chem. Soc., 2007, 129, 355.

    76 R. Srivastava, M. Choi and R. Ryoo, Chem. Commun., 2006, 4489.

    77 J. Kim, M. Choi and R. Ryoo, J. Catal., 2010, 269, 219.

    78 C. H. Christensen, I. Schmidt, A. Carlsson, K. Johansen and K. Herbst, J. Am.

    Chem. Soc., 2005, 127, 8098.

    79 D. H. Park, S. S. Kim, H. Wang, T. J. Pinnavaia, M. C. Papapetrou, A. A.Lappas and K. S. Triantafyllidis, Angew. Chem. Int. Ed., 2009, 48, 7645.

    80 Y. Sun and R. Prins, Angew. Chem. Int. Ed., 2008, 47, 8478.

    81 T. Tang, C. Yin, L. Wang, Y. Ji and F. S. Xiao, J. Catal., 2008, 257, 125.

    82 T. D. Tang, C. Y. Yin, L. F. Wang, Y. Y. Ji and F. S. Xiao, J. Catal., 2007,

    249(1), 111.

    83 Y. Li, S. Liu, Z. Zhang, S. Xie, X. Zhu and L. Xu, Appl. Catal. A, 2008, 338,

    100.

    84 Y. Song, X. Zhu, Y. Song, Q. Wang and L. Xu, Appl. Catal. A, 2006, 302, 69.

    85 L. Su, L. Liu, J. Zhuang, H. Wang, Y. Li, W. Shen, Y. Xu and X. Bao, Catal.

    Lett., 2003, 91(3-4), 155.86 M. H. F. Kox, E. Stavitski, J. C. Groen, J. Perez Ramirez, F. Kapteijn and

    B. M. Weckhuysen, Chem. Eur. J., 2008, 14, 1718.

    87 V. N. Shetti, J. Kim, R. Srivastava, M. Choi and R. Ryoo, J. Catal., 2008, 254,

    296.

    88 D. P. Serrano, R. A. Garcia and D. Otero, Appl. Catal. A, 2009, 359, 69.

    89 M. Choi, D. H. Lee, K. Na, B. W. Yu and R. Ryoo, Angew. Chem. Int. Ed.,

    2009, 48, 3673.

    90 M. Y. Kustova, S. B. Rasmussen, A. L. Kustov and C. H. Christensen, Appl.

    Catal. B, 2006, 67, 60.

    91 I. Melian-Cabrera, S. Espinosa, J. C. Groen, B. v/d Lenden, F. Kapteijn andJ. A. Moulijn, J. Catal., 2006, 238, 250.

    92 K. S. Triantafyllidis, E. F. Iliopoulou, E. V. Antonakou, A. A. Lappas, H.

    Wang and T. J. Pinnavaia, Micropor. Mesopor. Mater., 2007, 99, 132.