Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Ordered Porous Solids: Recent Advances and Prospects
Ordered Porous Solids: Recent Advances and Prospects
Ordered Porous Solids: Recent Advances and Prospects
Ebook1,758 pages16 hours

Ordered Porous Solids: Recent Advances and Prospects

Rating: 0 out of 5 stars

()

Read preview

About this ebook

The developments in the area of ordered nanoporous solids have moved beyond the traditional catalytic and separation uses and given rise to a wide variety of new applications in different branches of chemistry, physics, material science, etc. The activity in this area is due to the outstanding properties of nanoporous materials that have attracted the attention of researchers from different communities. However, recent achievements in a specific field often remain out of the focus of collaborating communities. This work summarizes the latest developments and prospects in the area of ordered porous solids, including synthetic layered materials (clays), microporous zeolite-type materials, ordered mesoporous solids, metal-organic-framework compounds (MOFs), carbon, etc. All aspects, from synthesis via comprehensive characterization to the advanced applications of ordered porous materials, are presented. The chapters are written by leading experts in their respective fields with an emphasis on recent progress and the state of the art.

  • Summarizes the latest developments in the field of ordered nanoporous solids
  • Presents state-of-the-art coverage of applications related to porous solids
  • Incorporates 28 contributions from experts across the disciplines
LanguageEnglish
Release dateAug 11, 2011
ISBN9780080932453
Ordered Porous Solids: Recent Advances and Prospects

Related to Ordered Porous Solids

Related ebooks

Materials Science For You

View More

Related articles

Reviews for Ordered Porous Solids

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Ordered Porous Solids - Valentin Valtchev

    Tsapatsis

    Brief Table of Contents

    Copyright

    Contributors

    Index

    Table of Contents

    Copyright

    Contributors

    Index

    Copyright

    Elsevier

    The Boulevard, Langford lane, Kidlington, Oxford, OX5 1GB, UK

    Radarweg 29, PO Box 211, 1000 AE Amsterdam, The Netherlands

    First edition 2009

    Copyright © 2009 Elsevier B.V. All rights reserved

    No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means electronic, mechanical, photocopying, recording or otherwise without the prior written permission of the publisher

    Permissions may be sought directly from Elsevier's Science & Technology Rights Department in Oxford, UK: phone (+44) (0) 1865 843830; fax (+44) (0) 1865 853333; email: permissions@elsevier.com. Alternatively you can submit your request online by visiting the Elsevier web site at http://elsevier.com/locate/permissions, and selecting Obtaining permission to use Elsevier material

    Notice

    No responsibility is assumed by the publisher for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions or ideas contained in the material herein. Because of rapid advances in the medical sciences, in particular, independent verification of diagnoses and drug dosages should be made

    British Library Cataloguing in Publication Data

    A catalogue record for this book is available from the British Library

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is available from the Library of Congress

    ISBN: 978-0-444-53189-6

    For information on all Elsevier publications visit our web site at books.elsevier.com

    Printed and bound in Great Britain 08 09 10 10 9 8 7 6 5 4 3 2 1

    Contributors

    Section I. Synthesis, Modification and Characterization of Ordered Porous Materials

    Chapter 1. A New Family of Mesoporous Oxides—Synthesis, Characterisation and Applications of TUD-1

    Contents

    Introduction 4

    MCM-41 and FSM-16 5

    TUD-1 6

    Redox Metal Incorporation into Siliceous TUD-1 Framework 10

    Ti-TUD-1 10

    Fe-TUD-1 11

    Co- and Cr-TUD-1 14

    Ce-TUD-1 17

    Zr-TUD-1 18

    Al² O³ -TUD-1 and Al-TUD-1 19

    TUD-1 as Potential Drug Carriers 21

    Particle Incorporation 22

    Zeolite Beta 22

    ITQ-2—Delaminated zeolite 26

    Conclusion 28

    References 29

    Keywords

    Mesoporous oxides

    Nanoparticles

    Composite materials

    Synthesis

    Catalysis

    Abbreviations

    CTMA

    Cetyltrimethylammonium

    FTIR

    Fourier Transform Infrared

    MPV

    Meerwein-Ponndorf-Verley

    NMR

    Nuclear Magnetic Resonance

    SMPO

    Styrene Monomer Propylene Oxide

    TBHP

    tert-Butylhydroperoxide

    TEA

    Triethanolamine

    TEAOH

    Tetraethylammonium hydroxide

    TEOS

    Tetraethoxysilane

    TOF

    Turnover frequency

    (HR) TEM

    (High Resolution) Transmission Electron Microscopy

    UV-vis

    Ultraviolet-Visible Spectroscopy

    XRD

    X-ray Diffraction

    Abstract

    There exists much work regarding the synthesis of mesoporous materials. TUD-1 is one recent development whereby a non-surfactant organic compound (triethanolamine) templates the formation of mesoporous oxides. Transition metals varying from atomically dispersed isolated centres to nanoparticulate oxides are incorporated easily and controllably in a one-pot synthesis mixture, and have shown considerable catalytic performance when evaluated in various reaction types. Furthermore, composite micro-/mesoporous systems have been synthesised and applied successfully in a range of acid-catalysed reactions. This chapter aims to provide an up to date review of TUD-1 and its derivatives, their synthesis, characterisation and application.

    1. Introduction

    Porous materials of varying chemical characteristics (basic, acidic, redox-active, inert, conducting, semi-conducting, etc.) are of fundamental importance in the areas of science and technology.[¹,²] This is due to the presence of voids of controllable dimensions at the atomic, molecular and nanometre scale.[³] Their use in industry is extensive ranging from areas such as petroleum refining,[⁴] detergents,[⁵] medicinal applications[⁶] and separations.[⁷] The International Union of Pure and Applied Chemistry (IUPAC) divides porous materials into three classes based on their pore diameter (d): microporous d < 2.0 nm, mesoporous 2.0 ≤ d ≤ 50 nm and macroporous d > 50nm.[⁸] Pore architectures (size, shape, connectivity) and the nature of the pore distribution, in combination with the chemical characteristics of the pore walls, determine the properties (and hence applications) of such materials.

    Arguably, the most important group of porous materials so far is the microporous one, in particular zeolites. Although these crystalline aluminosilicates are used extensively in industry, they do suffer from significant drawbacks as a result of their small pore sizes. Zeolites can experience problems in mass transfer, affecting diffusivities of the reactants and products to and from the active sites. In addition, in some reactions, like catalytic cracking, this might lead to coke formation, degrading the catalytic activity of a zeolite. The inability of bulky substituents to make use of the extensive internal surface area restricts their use in important chemical processes. Much work has been carried out in the literature to address the limitations of zeolites, one of the methods being the synthesis of materials with larger sized pores. One of the most significant contributions has sprung from work involving mesoporous material synthesis, initiated by the groups led by Beck[⁹] and Kuroda.[¹⁰]

    2. MCM-41 and FSM-16

    The best-known family of mesoporous materials is MS41, and from that group, MCM-41, first reported in 1992.[⁹] The material was synthesised in an effort to form an improved alternative to zeolites. MCM-41 material exhibits a regular, hexagonal arrangement of pores (P6mm symmetry) with one-dimensional parallel channels, formed as a result of liquid crystal templating. This mechanism is thought to be a result of either the silicate synthesis gel condensing around the pre-arranged micelles formed from the cetyltrimethylammonium (CTMA+) template or by the silicate influencing the formation of the liquid crystal phase (Fig. 1.1),[¹¹] depending on the concentrations used. Pore sizes obtained for this material can range from 1.5 to 10 nm.

    Figure 1.1. Possible mechanistic pathways for the formation of MCM-41: (1) liquid crystal phase initiated and (2) silicate anion initiated.

    Reprinted with permission from Beck et al. [¹¹] Copyright (1992) TheAmericanChemical Society.

    FSM-16, a material with a similar pore arrangement was synthesised by Inagaki et al.[¹⁰] This method consists of interlayer cross-linking of a layered silicate (kanemite, NaHSi²O⁵·3H²O) that has undergone an ion exchange (usually alkyltrimethylammonium ions, such as CTMA+). The silicate layers of the single-layered polysilicate are then able to condense to form silicate networks (Fig. 1.2). For the synthesis of FSM-16, a smaller amount of CTMA+ is employed compared to MCM-41, since it is used as a swelling agent, rather than as a template, to direct the formation of the material.

    Figure 1.2. Schematic model for the formation of mesoporous material from kanemite.[¹⁰]

    Reproduced by permission of The Royal Society of Chemistry.

    Even though aluminosilicate versions of these materials with essentially amorphous walls were consequently no match for the activity associated with the high acidity of crystalline zeolites, it did open up a new door to mesoporous material synthesis. Early examples of these mesoporous materials exhibited low hydrothermal stability, thin walls, the use of an expensive template and the one-dimensional pore arrangement. Thus, much work has been targeted at either the improvement of these materials or the synthesis of novel mesoporous materials.

    3. TUD-1

    One such example of a novel, next-generation, mesoporous material set out to improve the short-comings of MCM-41 is TUD-1 (Technische Universiteit Delft), first reported in 2001 by one of us (T. Maschmeyer).[¹²] Along with its high surface area, thicker mesopore walls and hydrothermal stability, TUD-1 employs an inexpensive, non-surfactant structure directing agent, triethanolamine (TEA), to direct the formation of mesostructures during the polycondensation of inorganic species. The resulting material is one which exhibits high surface areas (up to 1000 m[²] g-1), tunable mesopore size distribution (25–250 Å) and a three-dimensional arrangement of interconnecting pores. The three-dimensional connectivity of TUD-1 was proven via carbon inverse replication, a method first reported by Ryoo et al.[¹³] to prove the interconnected pores of SBA-15.[¹³] The resulting carbonaceous material showed a similar foam-like mesopore structure to TUD-1 (Fig. 1.3).

    Figure 1.3. Mesoporous carbon templated by TUD-1silica, yielding an inverse TUD-1structure.[¹²]

    Reproduced by permission of The Royal Society of Chemistry.

    For a typical synthesis of TUD-1, TEA and tetraethoxysilane (TEOS) are combined in the following ratios—1TEOS:0.25–2TEA:10–40H²O. The gel is then aged, dried and treated hydrothermally for a certain amount of time. Since there exists a direct relationship between pore sizes and heating time, one can obtain a desired pore size simply by adjusting the heating time, as shown in Fig. 1.4. The template is removed via Soxhlet extraction with ethanol, followed by calcination in air, resulting in TUD-1.

    Figure 1.4. Typical graph obtained when tuning the mesoporosity by variation of the heating timeduring hydrothermal treatment (in this case a homogeneous synthesismixturewith a molar composition of TEOS:0.5TEA:0.1TEAPH:11H²O was, after drying at 98 °Cfor 24 h, heated in an autoclave to 190 °C for different times). D is the mesopore diameter at maximum peak height calculated using the BJH model based on nitrogen desorption branch, S is the mesopore surface area calculated using the t-plot method.[¹²]

    Reproduced by permission of The Royal Society of Chemistry.

    A schematic mechanism of formation is outlined in Fig. 1.5. Firstly, the TEOS is hydrolysed to give Si(OH)⁴ which condenses with TEA. Since C–O–Si bonds are more easily hydrolysed than Si–O–Si bonds, over time the TEA–silica linkages get repeatedly broken and the silica nucleus grows slowly, while being surrounded by TEA. Upon heating, the processes of TEA hydrolysis and silanol condensation are accelerated, leading to a controlled phase separation (i.e. microsyneresis). Subsequently, the TEA aggregates template the mesoporous silica framework as the silica species condense around them. Figure 1.6 is a transmission electron microscopy (TEM) image of the final calcined product where foam-like porosity is clearly evident.

    Figure 1.5. Mechanism of formation of TUD-1: (i) TEA hydrolysis and TEOS condensation from the initial silica nucleus, resulting in phase separation (ii) and pore formation (iii).

    Figure 1.6. TEMimage of TUD-1.

    In addition to directing pore formation, TEA is able to stabilise active metals in the synthesis gel, which then allows their incorporation into the siliceous framework. The synthesis of metal-incorporated mesoporous materials has been a field of active investigation, as these materials offer much better diffusion to and from the active sites than their microporous counterparts. The initial metal concentration for the synthesis of such materials must be taken into careful consideration: TEA can stabilise many active metals via complexation, thus it is possible to generate a heterometal-rich TEA phase that templates the mesopores. During calcination, the organic species are removed and the metal atoms are deposited into/onto the internal mesoporous surface (in situ grafting), depicted in Fig. 1.7a. At higher metal loadings, the possibility for hetero-metal-oxide formation increases and nanosized scale crystals can grow inside the pores, and/or bulky metal oxides are formed as extra framework crystals (Fig. 1.7b). Thus, with materials containing higher metal loadings, more than one type of catalytic site is likely to exist.

    Figure 1.7. (a) Low concentration of the metal precursor allows the metal to be atomically dispersed and (b) higher concentration results in nanoparticles.The smaller sized particles (i) are controlled by concentration (lower end sizes) and the larger sized particles (ii) controlled by the material's pore sizes that can be adjusted via heating time.

    Taking Ti as an example, TEA is known to form complexes more easily with Ti than Si (due to the interaction of the lone electron pair on nitrogen atom with empty d orbital of the Ti). Thus, Ti can be deposited in a controlled manner into the pores and onto the walls of TUD-1, as first reported by Shan et al.[¹⁴] The stabilisation of the active metal during the synthesis is extremely important as the starting materials (Ti(IV) alkoxides) have a tendency to form TiO² upon exposure to H²O in a rapid and not easily controlled manner. Thus, not only would this lower the number of isolated and tetrahedrally co-ordinated metal sites, but also lower the integrity of the mesoporous structure as such.

    4. Redox Metal Incorporation into Siliceous TUD-1 Framework

    4.1. Ti-TUD-1

    The motivation for the synthesis of a Ti-TUD-1 material has its origin in the desire to epoxidise bulky alkenes with peroxides and preferentially with hydrogen peroxide, analogous to titanosilicate-1 (TS-1). Initial interest was based on the SMPO (Styrene Monomer Propylene Oxide) process, which employs a titanium-on-silica (epoxidation) catalyst, first established in 1979,[¹⁵] and is a process still in use by Shell today. The report of epoxidation with hydrogen peroxide using TS-1 ignited the field.[¹⁶] Incorporation of Ti into the crystalline framework largely avoids the problem of leaching that typically occurs when metals are incorporated into zeolites with an ion-exchange procedure, but more importantly the bond angles due to the lattice are thought to prevent Ti-site deactivation by water and/or hydrogen peroxide. TS-1 has shown to exhibit remarkable activity due its structure and selective oxidations of alkanes,[¹⁷] alkenes[¹⁸] and alcohols.[¹⁸] Many other microporous titanosilicates and titanoaluminophosphates have also been synthesised, such as TS-2,[¹⁹] Ti-ZSM-48,[²⁰] Ti-beta,[²¹] TAPO-5[²²] and TAPO-11[²²] all with tetrahedrally co-ordinated Ti centres. The small pore sizes, as with all microporous materials, are by definition the common problem, restricting their use with bulky substrates because these are unable to penetrate into the micropores and only the external surface area can be utilised—wasting everything but a small fraction of the available sites.

    Thus, a number of mesoporous titanosilicates have been synthesised, for example using the M41S family, where Ti can be introduced into the mesoporous wall during synthesis or grafted onto the wall surface after synthesis of the material (e.g. Ti-MCM-41[²³] and Ti-HMS[²⁴]) to offer a solution to these problems. At loading above 0.5 wt.% Ti, a significant amount of Ti(IV) centre lies inaccessibly in the framework and dimmer/oligomer formation can occur, reducing the peroxide selectivity of the titanium centres. This problem was overcome by grafting an organometallic titanium catalytic-site precursor onto the inner walls after synthesis, leading to isolated and dispersed sites, at high loadings.[²⁵]

    Along with the drawbacks of MCM-41 mentioned previously, the one-dimensional pore structure can lead to problems associated with pore blockage and the stabilisation of the isolated, tetrahedral Ti-sites. However, Ti-TUD-1,[²⁶] with its three-dimensional arrangement of pores, offers an alternative to such materials. It was tested as a potential catalyst and compared with Ti-MCM-41, Ti-grafted TUD-1 and Ti-grafted MCM-41 in the epoxidation of cyclohexene with tert-butylhydroperoxide (TBHP). The results are summarised in Table 1.1.

    Table 1.1. Comparison of catalytic activities for cyclohexene epoxidation over Ti-TUD-1, Ti-MCM-41, Ti grafted TUD-1, and Ti grafted MCM-41

    a Titanium loading.

    b Surface area obtained from nitrogen adsorption.

    c Mesopore diameter at maximum peak calculated by using the BJH model.

    d Turnover frequency after 6 h reaction.

    e Selectivity of TBHP after 6 h reaction.

    The turnover frequency (TOF) of Ti-TUD-1 was found to be about 5.6 times higher to that of framework-substituted Ti-MCM-41 and similar to that of Ti-grafted MCM-41. The selectivity of TBHP to cyclohexene epoxide was ∼94% after 6 h of reaction. One advantage of Ti-TUD-1 over Ti-grafted MCM-41 is that the former titanosilicate can be synthesised in a one-pot synthesis. The higher activity of Ti-TUD-1 originates from both the high accessibility of the substrates to the catalytic site provided by the three-dimensional mesopore system and from the formation of isolated titanium centres on the surface of the mesopore wall, to give optimum usage of Ti.

    4.2. Fe-TUD-1

    In a similar vein to Ti-TUD-1, it has been possible to successfully immobilise Fe in siliceous TUD-1 to produce Fe-TUD-1. In addition, by adjusting the Si/Fe concentration one can also obtain nanoparticles of iron oxide, located preferentially inside the pores. The synthesis of Fe-TUD-1 was first reported by Hamdy et al.[²⁷] and materials made with varying Si/Fe ratios were tested as a potential catalyst for the Friedel–Crafts benzylation of benzene (Table 1.2). Generally speaking, Friedel–Crafts alkylations are catalysed by Lewis acids in the liquid phase. As with other processes, there is a push to replace liquid acids by solid acid catalysts, as they are more easily separated from their reaction mixtures and conform more readily to the principles of ‘Green Chemistry’. Fe-TUD-1 provides an effective alternative to the traditional FeCl³ catalyst for the benzylation of benzene, as shown in Table 1.3. The catalytic test shows no reaction with pure siliceous TUD-1 or with pure iron oxide powder (Fe²O³), consistent with the need for dispersed nanoparticles in TUD-1 (for all active samples, the selectivity to diphenyl methane is 100%).

    Table 1.2. Comparison between samples composition and activity (i.e. conversion of benzyl chloride)

    Table 1.3. Asymmetric hydrogenation of methyl 2-acetamidoacrylate using AlTUD-1 as catalyst[a]

    a 5 bar H², 50 mL solvent, [4a] = 0.05 M, 0.1 g catalyst with 1 wt% Rh.

    b 10 bar H², 50 mL solvent, [4a] = 0.2 M, 0.01 g catalyst with 1 wt% Rh.

    c Percentage of total amount of Rh determined by AAS of the filtrate.

    Fe-TUD-1 exhibits different types of catalytic sites; at lower Fe loadings isolated Fe(III) sites are consistent with diffuse reflectance UV–vis, TEM and XRD data, while at higher Fe loadings nanoparticles of iron oxide embedded into the mesopores of the TUD-1 silica matrix are observed by HR-TEM (Fig. 1.8). From the table above, it is clear that the sample with Si/Fe = 10 ratio was the most active.

    Figure 1.8. HR-TEMimages of different Fe-TUD-1 samples.

    Reproduced from Hamdy et al. [²⁷] by permission of Elsevier.

    By examining the time taken to reach for 90% as a function of Fe loading, we can observe that the increase in activity as a function of Fe loading is highly non-linear, pointing to a fundamental change in the nature of the active site (Fig. 1.9).

    Figure 1.9. Time taken to reach 90% conversion of benzyl chloride as a function of Fe loading.

    It has been suggested that one reason for the improved activity with higher loading is that the nanoparticles of iron oxide are less co-ordinatively saturated (due to lattice strain and defects) than isolated iron species or than the iron centres present in larger particles, thereby giving rise to more active centres. On comparing Fe-TUD-1 with other materials from the literature,[²⁸–³⁰] Fe-TUD-1 is more active by more than a factor of 15 when compared to the next best material (Fe-ZSM-5). This is likely due to the combination of improved local environment of active site (strained/unsaturated nanoparticles) as well as the improved global structure of TUD-1, which is thought to be responsible for the higher accessibility of the reactants inside Fe-TUD-1.

    4.3. Co- and Cr-TUD-1

    The ease with which Fe and Ti could be selectively incorporated into the TUD-1 structure type (i.e. varying from atomically dispersed to stabilised nanoparticles inside the mesopores) paved the way for modifications with other transition metals. Co- and Cr-TUD-1 materials can also be synthesised in much the same way as their Fe- and Ti-TUD-1 analogues, as they also from relatively stable complexes with TEA. Thus, Cr- and Co-TUD-1 were investigated as suitable catalysts for the liquid phase oxidation of cyclohexane with TBHP and air as an oxidant.

    Cyclohexane oxidation is an important industrial process, but as the product is more easily oxidised than the substrate, the cyclohexane conversions need to be kept low (∼4%) such as to retain high selectivity for cyclohexanone and cyclohexanol (∼70–85%). This has prompted much research into the development of new and more selective catalytic systems which work under mild conditions. One approach has been to replace the traditional homogeneous cobalt-based catalysts with heterogeneous catalysts, which prompted the synthesis of Co- and Cr-TUD-1 materials, containing a Si/M ratio of 100.[³¹] In the liquid-phase cyclohexane oxidation, at 70 °C, with TBHP as oxidant both the conversion of cyclohexane (Fig. 1.10a) and TBHP (Fig. 1.10b) were catalysed by all M-TUD-1 catalysts. By plotting the conversion of cyclohexane against selectivity for mono-oxygenated products[¹] (Fig. 1.10c), both Co-TUD-1 and Cr-TUD-1 exhibited high activity and very high selectivity to these products (at conversion levels of 5–8%). Co-TUD-1 showed a very slow selectivity decline (even at higher conversion levels) of cyclohexane (compared to Cr-TUD-1). Cr-TUD-1 produced more cyclohexanone than cyclohexanol but Co-TUD-1 produced initially more of the alcohol (A), and then slowly oxidised it to the ketone (K) as shown in Fig. 1.10d. In summary, Cr and Co-TUD-1 showed >90% selectivity towards mono-oxygenated products and conversion levels as high as 8–10%. However, not unexpectedly, leaching was a problem, particularly in the chromium case.

    ¹ Mono-oxygenated products are cyclohexanol, cyclohexanone, cyclohexyl hydroperoxide and minor products such as cyclohexylformate and cyclohexyltertbutylper-ether.

    Figure 1.10. (a) Conversion of cyclohexane and (b) tert-butylhydroperoxide (TBPH) with TBPH as oxidant over M-TUD-1 at 70 °C.The conversion of cyclohexane versus selectivity for monooxygenated products (c) is also shown and (d) the formation of cyclohexanone (K)/cyclohexanol (A) against time.

    Reproduced from Anand et al. [³¹] by permission of Elsevier.

    In a further study, Co-TUD-1 was synthesised with different Si/Co ratios (100, 50, 20 and 10) and analysed under the same catalytic reaction.[³²] Higher catalytic efficiency was observed in samples with a low Co loading (Co atoms in the framework). Figure 1.11 shows that the conversion of both cyclohexane and cyclohexanol decreases as the Co concentration in TUD-1 increases. Agglomeration of Co at higher metal loadings is thought to reduce the accessibility of individual Co atoms and therefore the number of active sites. Thus, it is most likely that the isolated Co(II) sites are responsible for the catalyst's activity highlighting the importance of metal site dispersion and the ability of TUD-1 to achieve this requirement.

    Figure 1.11. (a) Conversion of cyclohexane and (b) tert-butylhydroperoxide (TBHP) over Co-TUD-1at various metal loadings.

    Reproduced from Hamdy et al. [³²] by permission of Wiley-VCH.

    4.3.1. Cr-TUD-1 as photocatalysts in the gas phase

    There is an extensive body of work in which Cr has been incorporated into both mesoporous materials (such as HMS) and microporous materials (ZSM-5).[³³–³⁶] In particular, Cr⁶+ has shown a considerable amount of photocatalytic efficacy in the framework of both silica and silica–alumina supports. Therefore, overcoming the problem of leaching, while still making use of the catalytic potential of Cr-TUD-1, the selective photocatalytic oxidation of propane to acetone in the gas phase was investigated. Cr-TUD-1 (Si/Cr ratio of 130) was found to show high activity and high selectivity towards acetone under visible light irradiation.[³⁷] Absorbed species arising from the propane oxidation were monitored using FTIR (Fig. 1.12a). The rate of total product formation is plotted as a function of exposure time (Fig. 1.12b).

    Figure 1.12. (a) The development of infrared absorbance spectroscopy during photocatalytic oxidation of propane over Cr-TUD-1with 435 nm illumination. (b) Propane conversion and acetone selectivity plotted during 200 min of irradiation.

    Reproduced from Hamdy et al. [³⁷] by permission of Elsevier.

    Various Ti-TUD-1 materials (Si/Ti = 100, 20, 2.5, 1.6) were also investigated for the same reaction.[³⁸] As a function of increasing metal loading, either isolated Ti atoms (when Ti loading ∼2.5 wt.%) or combinations of isolated Ti atoms and anatase (TiO²) nanoparticles are obtained. Propane oxidation occurred at an irradiation of λ = 365 nm (which selectively activates anatase nanoparticles). The rate of total product formation is plotted as a function of exposure time (Fig. 1.13). Clearly, Ti-TUD-1 is able to selectively photocatalyse the dehydrogenation of propane under blue light, especially at higher Si/Ti ratios. The selectivity is thought to arise from the ‘de-tuned’ catalytic activity that has been achieved by increasing the bandgap due to quantum confinement in the nanoparticle.

    Figure 1.13. Propane consumption over different Ti-TUD-1 compositions.

    Reproduced from Hamdyet al.[³⁸] by permission of Wiley-VCH.

    4.4. Ce-TUD-1

    Ce-TUD-1 has shown to be an active catalyst for both the oxidation of p-tert-butyl toluene to p-tert-butyl benzaldehyde and peroxidative halogenation of Phenol Red. TUD-1 samples containing different Si:Ce ratios (20, 50, 100) were investigated. Complete incorporation of Ce into the framework of TUD-1 was achieved for Ce-TUD-1 (100) whereas in Ce-TUD-1 (20), a small portion of Ce as nanometre sized CeO² particles was found.[³⁹] Regardless of the concentration of Ce incorporated in the TUD-1 synthesis gel, the conversion of p-tert-butyl toluene exhibits a conversion of 37%, with 57% selectivity towards the desired p-tert-butyl benzaldehyde product. These values are similar to those reported for Ce(III) acetate, suggesting the reaction mechanism over heterogeneous Ce-TUD-1 and homogeneous Ce(III) acetate catalysts is the same. With Ce-TUD-1 (20), leaching occurs (∼7%); however, when the regenerated catalyst is reused, the substrate conversion increased to 90%, with an aldehyde selectivity of 55%. This surprising result is subject to further investigations.

    4.5. Zr-TUD-1

    Tetrahedrally co-ordinated zirconium has shown to be particularly active in the Meerwein–Ponndorf–Verley (MPV) reduction of aldehydes by Ramanathan et al.,[⁴⁰] a mild redox reaction used extensively in steroid chemistry (Fig. 1.14).

    Figure 1.14. Proposedmechanismfor the Meerwein—Ponndorf—Verley (MPV) reduction.

    The MPV reduction is typically catalysed by equimolar amounts of aluminium(III) isopropoxide. The proposed mechanism involves a complex in which both the reactants are coordinated to the Lewis acid metal centre and a hydride transfer from the alcohol to the carbonyl group occurs. Other Lewis acid metals such as Zr[⁴¹] and Ln[⁴²] have shown their potential as homogeneous catalysts for the described reaction. More recently, the introduction of heterogeneous catalysts, in particular Al-free Sn-beta[⁴³] and Al-free Zr-beta,[⁴⁴–⁴⁶] show significant improvements in activity. Al-free Zr-beta was found to be the superior catalyst of the two, as it exhibited higher regioselectivity and was less sensitive to moisture than Al-free Sn-beta. Work has also been conducted on grafting of Zr on MCM-41, MCM-48 and SBA-15, to extend the use to bulkier substituents, but the process is time consuming as grafting is performed post synthesis.[⁴⁷] With Zr-TUD-1, the metal is able to be incorporated into the framework of TUD-1 using the one-pot synthesis approach described previously.[⁴⁰] The catalyst displays high selectivity, in particular towards the sterically hindered steroids, compounds that are unable to be reduced with H-beta or Al-free Zr-beta. One particular example is the reduction of 5α-cholestan-3-one (Fig. 1.15). The keto-group was reduced almost quantitatively (> 95%) to yield 5α-cholestan-3-ol with α/β ratio of 1.6.

    Figure 1.15. Reduction of 5α-cholestan-3-one to 5α-cholestan-3-ol over Zr-TUD-1.

    Thus, along with the Zr to act as an effective Lewis acid to catalyse the MVP reduction, the sponge-like mesoporous structure of TUD-1 allows for the conversion of large, bulky substituents.

    5. Al²O³-TUD-1 and Al-TUD-1

    In 2003, Shan et al.[⁴⁸] published results for the formation of a stable mesoporous alumina denoted Al²O³-TUD. In the same way that all-silica TUD-1 is templated by a small organic compound, Al²O³-TUD is synthesised with a similar compound, tetra(ethylene) glycol (TEG). In addition, due to the high reactivity of aluminium alkoxides with water, a controlled amount of water is used in the initial synthesis to avoid the formation of alumina particles. Surface areas of ∼600 m[²] g-1 can be obtained with Al²O³-TUD.

    Apart from forming a pure Al²O³-TUD, it is possible to synthesise a aluminosilicate TUD-1 (denoted Al-TUD-1), a method first described in Simons et al.[⁴⁹] The charge imbalance, induced by the partial substitution of silicon with aluminium, allows this material to act as an effective anionic carrier onto which cationic metal complexes can be bound non-covalently. In addition, for high Brønsted acidity, a high Al/Si ratio (∼0.25) is required and the aluminium should be preferentially tetrahedrally co-ordinated. It is important to note that in Al-TUD-1, we do not see the same dependence of hydrothermal treatment time and pore size distribution that we observe in TUD-1. It is thought that the Al-TUD-1 system is less dynamic during the early stages of heating because of the faster degree of Al–O–Si bond formation compared to that of Si–O–Si bonds. Thus, this results in a system that is less sensitive towards changes in pore size with temperature. In addition, TEG is not able to completely suppress the formation of hexa-coordinated Al. However, the resulting material does fulfil the requirements for an effective anionic carrier, with a high surface area and a good Altetrahedral/Si ratio (= 0.11, total Al/Si = 0.25).

    The ability of Al-TUD-1 to act as an efficient support was investigated by Simons et al.[⁴⁹,⁵⁰] whereby ([RhI(cod)((R)-MonoPhos)²]BF⁴) (Fig. 1.16), a well-known asymmetric hydrogenation catalyst, was immobilised on Al-TUD-1 with a loading typically ∼1% (Fig. 1.17) and tested with the asymmetric hydrogenation of methyl 2-acetamidoacrylate (Fig. 1.18).

    Figure 1.16. Mono PhosRh catalyst.

    Figure 1.17. Typical heterogenisation of an Rh catalyst on Al-TUD-1.

    Reproduced from Simonset al.[⁵⁰] by permission of Springer.

    Figure 1.18. Asymmetric hydrogenation of methyl 2-acetamidoacrylate.

    Reproduced from Simons et al.[⁵⁰] by permission of Springer.

    Hydrogenising this traditional homogeneous catalyst allows for an easier and more complete separation of the products from the reaction mixture (Fig. 1.17). Importantly, a significant improvement on the reusability of this traditionally expensive metal complex can be observed. Furthermore, as the complex is non-covalently bound to the aluminosilicate support, behaviour similar to that of the homogeneous catalyst could be observed, that is, the supported metal complex had not undergone significant changes in its catalytic performance.

    Tests were carried out to examine the effect of the solvent used, in terms of conversion, selectivity and leaching.[⁵⁰] The results are summarised in Table 1.3. It was found that once the catalyst was immobilised, the most problematic aspect was its susceptibility to leaching. With increasing solvent polarity, the solvent displays an increasing ability to stabilise charged species, therefore the amount of Rh lost in the system increases. Figure 1.19 shows the loss of Rh as a function of the polarity of the solvent.

    Figure 1.19. The correlation between the polarity of the solvent (ETN) and the loss of Rh for 5-Al-TUD-1.

    Reproduced from Simons et al.[⁵⁰] by permission of Springer.

    However, the problem of leaching could be overcome by using an extremely polar solvent, water, which is simply too polar and the support becomes again the preferential location of the cationic catalyst. By using water as the solvent, the catalyst exhibited excellent enantioselectivity, good activity and a very low loss of Rhodium. In addition, the catalyst can be reused without significant deterioration—proving that the phosphoramidite (MonoPhos) is very stable in aqueous reaction media.

    In further work by Simons et al.,[⁵¹] a comparison of supports for the electrostatic immobilisation of Rhodium-MonoPhos was investigated. The metal complex was immobilised on four different anionic carrier materials: Al-TUD-1,[⁴⁸] Nafion, a Nafion silica composite (SAC-13) and phosphotungstic acid on alumina (PWTUD). The efficiencies of these catalysts were examined, again, for the asymmetric reduction of methyl-2-acetamidoacrylate.

    Although most of the catalysts are highly selective, the activity and loss of Rhodium is strongly dependent on the type of support. The bonding between Rh and the phosphotungstic acid (supposedly covalent) is thought to be the reason behind the catalyst's superior anchoring ability to PWTUD. This then translates to minimal leaching of the catalyst, regardless of which solvent system is employed. In addition, the enantioselectivity remained very high. The other supports did experience some leaching (as a result of a lower binding interaction between the support and catalyst) and in the case of SAC-13, a lower activity is also observed. When using Nafion, the catalysts were encapsulated, leading to transport limitations which resulted in low activity.

    6. TUD-1 as Potential Drug Carriers

    Mesoporous silica-based materials have been studied widely as potential drug delivery systems. Heikkilä et al.[⁵²,⁵³] studied the suitability of TUD-1 as a drug carrier, comparing the drug release kinetics of the drug loaded onto TUD-1, MCM-41 and of the solid drug by itself. When considering the pore diameters of TUD-1 and MCM-41, ibuprofen was selected as the model drug due to its molecular dimensions (1.0 nm×0.5 nm). Loading was accomplished via the ‘soaking method’ (ethanolic solution) where the carrier/drug ratio was 1:48 (w/w). Drug release experiments performed in acidic dissolution medium (mimicking conditions at the start of the small intestine) showed a rapid and almost complete (95%) liberation of ibuprofen from TUD-1 over 210 min. This is an acceptable time frame considering the drug transit time of the small intestine (often cited as 200 min). That the initial 60% of the drug was released quickly (15 min) is a positive result. The more restricted pore characteristic of MCM-41 is thought to be responsible for the slower initial release of ibuprofen (60% over 40 min). What is of interest to note is that both carriers released ibuprofen faster than is the case for the pure form of the drug (25% over 45 min) (Fig. 1.20). The improvement of dissolution is thought to be a result of the mesoporous material changing the solid state properties of the drug (crystal habit and/or amorphotisation, leading to faster dissolution rates). This effect is more pronounced in low pH environments as in neutral conditions the release times and amounts are not as significant since ibuprofen dissolves much more easily. Thus, the application of a siliceous mesoporous carrier diminished the pH dependency of the ibuprofen dissolution and demonstrated the potential for TUD-1 to act as a drug carrier.

    Figure 1.20. Ibuprofen release from TUD-1,MCM-41 and the pure crystalline form at HBSS pH5.5medium.

    Reproduced from Heikkilä et al.[⁵²] by permission of Elsevier.

    7. Particle Incorporation

    The formation of hierarchical materials by the combination of porous systems of different dimensions is of great interest, in particular to catalysis. The potential of TUD-1 to be a host for the stabilisation and dispersion of zeolite/zeotype nanoparticles has led to the design and synthesis of solid catalysts with very low diffusion barriers, enabling a deep level of control of transport phenomena.[⁵⁴] The network of the mesoporous TUD-1 structure minimises diffusion barriers and allows for the incorporation and stabilisation of highly reactive, microporous zeolite and zeotype particles. These particles are very important structures, as they are the principle catalysts used in petrochemical conversions and represent by far the greatest tonnage of catalysts used today; however, they suffer from high intrinsic diffusion barriers due to their intricate and narrow pore systems. While reducing the size of the particles minimises this limitation, very small particles are unstable and tend to aggregate, negating their initial size advantage.

    7.1. Zeolite Beta

    The first reported instance of a composite zeolite Beta-TUD-1 material was reported in 2002:[⁵⁵] Prior to the incorporation of zeolite Beta into the synthesis gel, the particles were homogeneously dispersed in aqueous ammonia to avoid initial aggregation of the particles. To this suspension were added TEOS, TEA, H²O and finally TEAOH. The homogeneity of the dispersion was maintained (during and after gelation) due to the sudden and rapid increase in viscosity brought on by the base in the transition of liquid to thick gel. With a molar ratio of TEAOH/Si < 0.1, one observes gelation. However, with TEAOH/Si > 0.2, no gelation occurs and zeolite particles separate out.[⁵⁴] X-ray Diffraction (XRD), NMR, nitrogen adsorption and HR-TEM can be used to confirm the incorporation of zeolite Beta into the mesoporous matrix and that the zeolites structure is maintained (Fig. 1.21). The XRD pattern (Fig. 1.21b) shows the characteristic peaks of zeolite Beta (7–8.5° and 22.4° in 2θ) are clearly evident, even at low zeolite loadings (20%), along with the characteristic mesoporous periodicity of TUD-1 (1.2° in 2θ).

    Figure 1.21. (a) X-ray powder diffraction (XRD) patterns of Beta-TUD-1 containing different amounts of zeolite Beta and pure zeolite Beta. (b) [²⁷]AlNMR of pure zeolite Beta and Beta-TUD-1 containing 20 and 40 wt.% of zeolite Beta.

    Reproduced from Waller et al.[⁵⁴] by permission of Wiley-VCH.

    [²⁷]Al NMR (Fig. 1.21b) indicates that all aluminium atoms in the composite material are four co-ordinate (as in pure zeolite Beta) thus one can assume that integrity and structure of the zeolite was maintained during the mesopore formation and subsequent calcination.

    The catalytic activity of Beta-TUD-1 with different loadings was tested using n-hexane cracking, shown in Fig. 1.22. Zeolite Beta-TUD-1 with a zeolite loading of 40 wt.% showed the highest activity out of the composite materials, and was twice as active per gram of zeolite than either the pure zeolite Beta or the physical mixture (40 wt.%) of zeolite and TUD-1. This clearly illustrated the use of TUD-1 to prevent aggregation of the nanoparticles, but also points to an additional effect.

    Figure 1.22. Pseudo-first-order reaction rate constants based on the mass of zeolite n-hexane cracking at 538 °C on zeolite Beta-TUD-1 catalysts, the pure zeolite Beta, and a physical mixture of 40wt.%zeolite and TUD-1.

    Reproduced from Waller et al.[⁵⁴] by permission from Wiley-VCH.

    The enhanced activity of the Beta-TUD-1 composite (40 wt.% loading) is a result of the modification in the number and strength of acid sites resulting from the interactions between the amorphous TUD-1 and zeolite Beta. FTIR (employing CO and NH³ adsorption) is able to differentiate between various Brønsted acid sites with overlapping bands. Besides isolated and vicinal silanols, two other types of Brønsted acid sites occur in the composites: (1) hydroxyl groups linked to partially extra-framework aluminium ions, sites of medium acidity and (2) bridged hydroxyl groups that form the strong Brønsted acidity in zeolites. Figure 1.23 depicts NH³ adsorption spectra whereby typical adsorption species and sites for zeolite Beta can be detected.

    Figure 1.23. NH³ adsorption spectra of zeolite Beta at roomtemperature (varying NH³ dosage), showing typical adsorption sites/species. (See color insert.)

    Additionally, NH³ adsorption can also be used to identify strained surface siloxanes, due to its reaction with them leading to a surface group with a distinctive IR band at 1550cm-1. In Fig. 1.24, NH³ adsorption reveals the presence of distorted reactive siloxane bridges formed by the condensation of two vicinal silanols. It is of interest to note that the sample with the highest catalytic activity shows the most abundant presence of these distorted reactive siloxane bridges. It is thought that this synergistic effect of combining TUD-1 and zeolite Beta in such a way is responsible for the higher activity of the composite material.

    Figure 1.24. FTIR spectra in vacuum at room temperature after NH³ adsorption: (a) TUD-1, (b) 20%Beta-TUD-1, (c) 40%Beta-TUD-1, (d) 60%Beta-TUD-1, and (e) zeolite Beta.

    7.2. ITQ-2—Delaminated zeolite

    Delamination is a common phenomenon in clays and layered materials. It occurs when the layers of a clay or layered material are swollen by the incorporation of a solvent or an organic compound, followed by the separation of the layers. Corma et al.[⁵⁶] first utilised this property to separate layered sheets of a zeolitic precursor, MCM-22(P), by the cation exchange of one cation (HMI+) for a larger, bulkier cation, CTAB+, followed by ultra-sonication in a basic medium (pH 13.5), depicted in Fig. 1.25. Once the material is calcined, ITQ-2 is obtained, which exists as single sheets of the respective zeolite organised in a ‘house-of-cards’ arrangement. ITQ-2 exhibits well-structured accessible external surface area (> 700 m[²] g-1), formed by 12-membered ring open cups with a microporous system made up of a sinusoidal 10-membered ring channel (∼0.55 nm diameter) parallel to the main plain of the layer (Fig. 1.26).[⁵⁷]

    Figure 1.25. Structural schematic representation of MWW-type zeolites and ITQ-2 delaminated zeolite.

    Reproduced from Díaz et al.[⁵⁸] by permission from Elsevier.

    Figure 1.26. Crystal structure of ITQ-2 showing the 2.5 nmlayered structure encompassing open cups and the 10-membered ring channels running between the major planes of the layer.

    Reproduced from Corma et al.[⁵⁷] by permission from The Royal Society of Chemistry.

    The ITQ-2 material proved to be a breakthrough due to the ease of accessibility of the material's active sites. The open cups (Fig. 1.26) provide easy entrance and exit points for the reactants and products. The delamination process, though severe, does not compromise the structural integrity of the active sites and thus the activity of the material.[⁵⁶] One important reaction that demonstrates the benefits of the greater site accessibility and smaller diffusion path, among many, is the cracking of larger molecules such as di-isopropylbenzene and vacuum gasoil. Results show that along with ITQ-2 being much more active than the MCM-22, it is also more selective towards the desired gasoline and diesel products with a lower production of gas and coke.

    Like the previous example of zeolite Beta incorporation into a TUD-1, a similar method can be used for ITQ-2. By introducing the material into the synthesis gel, one is able to disperse the sheets inside TUD-1, improving the handling of the material, much in the same way as with zeolite Beta.[⁵⁹] However, an alternative method for ITQ-2 incorporation inside TUD-1 involves the in situ delamination of the layered MCM-22(P) in the synthesis gel during TUD-1 gel formation (see Fig. 1.27). This occurs with the addition of TEA to a suspension of MCM-22(P) in water. The MCM-22(P) is made following a static synthesis procedure.[⁶⁰] It is suspected that TEA disrupts the layered material (assisted with sonication), which is held together by the electrostatic forces between the layers themselves and the CTAB+ used to increase the interlayer spacing. Since this methodology can be applied during the synthesis of TUD-1 in a one-pot mixture, it avoids the use of the harsher conditions employed by Corma et al.[⁵⁶] for the delamination, that is, a highly basic environment (pH ∼13.5). TEM indicates partial, if not fully, delaminated species surrounded by TUD-1 (see Fig. 1.27). Though the XRD pattern of the composite is not sharp (due to the low loading of the initial MCM-22(P) and curved nature of the sheets), it is still possible to confirm the presence of ITQ-2 and TUD-1 in the same composite material (Fig. 1.28).

    Figure 1.27. Schematic for the in situ delamination of MCM-22(P) in the synthesis gel of TUD-1 and the resulting TEM of the compositematerial.

    Figure 1.28. XRDpattern comparing ITQ-2 and ITQ-2/TUD-1.

    8. Conclusion

    In this short review, we have endeavoured to showcase the versatility of the TUD-1 mesoporous oxide family as a catalyst/drug carrier, anion-exchanger and donor material to micro-mesoporous composites.

    In particular, TUD-1 offers an alternative to the regularly employed MCM-41. In addition to the advantage of higher hydrothermal stability and three-dimensional arrangement of pores, the structure directing agent, TEA, is inexpensive and the material allows for a tunable pore size.

    Various metal-incorporated TUD-1 materials are easily synthesised in a one-pot synthesis gel, and the synthesis can be tuned to result in atomic dispersion of the desired metal in the framework, or the formation of stable metal oxide nanoparticles. In addition, pre-formed particles, such as zeolites, can be readily included during the synthesis for adequate dispersion.

    The potential for TUD-1 to act as an effective drug carrier has also been established, to give adequate release times for the model drug, ibuprofen.

    TUD-1's ease of synthesis (even complex systems can be made in a ‘one-pot approach’) and the controllability of its compositional and structural details mean that it may serve as a very useful material for many current and future applications.

    Bibliography

    References

    [1] M.E. Davis, Nature 417 2002813-821

    [2] X.S. Zhao, F. Su, Q. Yan, W. Guo, X.Y. Bao, L. Lv, Z. Zhou, J. Mater. Chem. 16 2006637-

    [3] P. Yang, T. Deng, D. Zhao, P. Feng, D. Pine, B.F. Chmelka, G.M. Whitesides, G.D. Stucky, Science 282 19982244-2246

    [4] C.J. Plank, E.J. Rosinski, W.P. Hawthorne, Ind. Eng. Chem. Prod. Res. Dev. 3 1964165-169

    [5] H.G. Smolka, M.J. Schwuger, Colloid Polym. Sci. 256 1978270-277

    [6] M. Vallet-Regí, F. Balas, D. Arcos, Angew. Chem. Int. Ed. 46 20077548-7558

    [7] H. Verweij, J. Mater. Sci. 38 20034677-4695

    [8] J. Rouquérol, D. Avnir, C.W. Fairbridge, D.H. Everett, J.H. Haynes, N. Pericone, J.D.F. Ramsay, K.S.W. Sing, K.K. Unger, Pure Appl. Chem. 66 19941739-1758

    [9] C.T. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vartuli, J.S. Beck, Nature 359 1992710-712

    [10] S. Inagaki, Y. Fukushima, K. Kuroda, J. Chem. Soc. Chem. Commun. 1993680-682

    [11] J.S. Beck, J.C. Vartuli, W.J. Roth, M.E. Leonowicz, C.T. Kresge, K.D. Schmitt, C.T.-W. Chu, D.H. Olson, E.W. Sheppard, S.B. McCullen, J.B. Higgins, J.L. Schlenker, J. Am. Chem. Soc. 114 199210834-10843

    [12] J.C. Jansen, Z. Shan, L. Marchese, W. Zhou, N. van der Puil, T. Maschmeyer, Chem. Commun. 2001713-714

    [13] R. Ryoo, S.H. Joo, S. Jun, J. Phys. Chem. B. 103 19997743-7746

    [14] Z. Shan, J.C. Jansen, L. Marchese, T. Maschmeyer, Microporous Mesoporous Mater 48 2001181-187

    [15] J.K.F. Buijink, J.J.M. van Vlaanderen, M. Crocker, F.G.M. Niele, Catal. Today 93–95 2004199-204

    [16] M. Taramasso, G. Perego, B. Notari, US Patent 4,410,501 1983

    [17] D.R.C. Huybrechts, L. Debruycker, P.A. Jacobs, Nature 345 1990240-242

    [18] B. Notari, Stud. Surf. Sci. Catal. 37 1988413-425

    [19] J.S. Reddy, R. Kumar, J. Catal. 130 1991440-446

    [20] D.P. Serrano, H.X. Li, M.E. Davis, J. Chem. Soc. Chem. Commun. 1992745-747

    [21] M.A. Camblor, A. Corma, A. Martinez, J. Perez-Pariente, J. Chem. Soc. Chem. Commun. 1992589-590

    [22] N. Ulagappan, V. Krishnasamy, J. Chem. Soc., Chem. Commun. 1995373-374

    [23] A. Corma, M.T. Navarro, J. Perez-Pariente, J. Chem. Soc. Chem. Commun. 1994147-148

    [24] P.T. Tanev, M. Chibwe, T.J. Pinnavaia, Nature 368 1994321-323

    [25] T. Maschmeyer, F. Rey, G. Sankar, J.M. Thomas, Nature 378 1995159-162

    [26] Z. Shan, E. Gianotti, J.C. Jansen, J.A. Peters, L. Marchese, T. Maschmeyer, Chem. Eur. J. 7 20011437-1443

    [27] M.S. Hamdy, G. Mul, J.C. Jansen, A. Ebaid, Z. Shan, A.R. Overweg, T. Maschmeyer, Catal. Today 100 2005255-260

    [28] K. Bachari, J.M.M. Millet, B. Benaichouba, O. Cherifi, F. Figueras, J. Catal. 221 200455-61

    [29] J. Cao, N. He, C. Li, J. Dong, Q. Xu, Stud. Surf. Sci. Catal. 117 1998461-467

    [30] V.R. Choudhary, S.K. Jana, B.P. Kiran, Catal. Lett. 59 1999217-219

    [31] R. Anand, M.S. Hamdy, P. Gkourgkoulas, T. Maschmeyer, J.C. Jansen, U. Hanefeld, Catal. Today 117 2006279-283

    [32] M.S. Hamdy, A. Ramanathan, T. Maschmeyer, U. Hanefeld, J.C. Jansen, Chem. Eur. J. 12 20061782-1789

    [33] H. Yamashita, M. Ariyuki, S. Higashimoto, S.G. Zhang, J.S. Chang, S.E. Park, J.M. Lee, Y. Matsumura, M. Anpo, J. Synchrotron Radiat. 6 1999453-454

    [34] H. Yamashita, M. Ariyuki, K. Yoshizawa, K. Kida, S. Ohshiro, M. Anpo, Res. Chem. Intermed. 30 2004235-245

    [35] H. Yamashita, S. Ohshiro, K. Kida, K. Yoshizawa, M. Anpo, Res. Chem. Intermed. 29 2003881-890

    [36] H. Yamashita, K. Yoshizawa, M. Ariyuki, S. Higashimoto, M. Anpo, M. Che, Chem. Commun. 2001435-436

    [37] M.S. Hamdy, O. Berg, J.C. Jansen, T. Maschmeyer, A. Arafat, J.A. Moulijn, G. Mul, Catal. Today 117 2006337-342

    [38] M.S. Hamdy, O. Berg, J.C. Jansen, T. Maschmeyer, J.A. Moulijn, G. Mul, Chem. Eur. J. 12 2006620-628

    [39] L.G.A. van de Water, S. Bulcock, A.F. Masters, T. Maschmeyer, Ind. Eng. Chem. Res. 46 20074221-4225

    [40] A. Ramanathan, D. Klomp, J.A. Peters, U. Hanefeld, J. Mol. Catal. A: Chem. 260 200662-69

    [41] B. Knauer, K. Krohn, Liebigs Annalen 1995677-683

    [42] J.L. Namy, J. Souppe, J. Collin, H.B. Kagan, J. Org. Chem. 49 19842045-2049

    [43] A. Corma, E. Domine Marcelo, L. Nemeth, S. Valencia, J. Am. Chem. Soc. 124 20023194-3195

    [44] Y. Zhu, G. Chuah, S. Jaenicke, Chem. Commun. 20032734-2735

    [45] Y. Zhu, G. Chuah, S. Jaenicke, J. Catal. 227 20041-10

    [46] Y. Zhu, G.-K. Chuah, S. Jaenicke, J. Catal. 241 200625-33

    [47] Y. Zhu, S. Jaenicke, G.K. Chuah, J. Catal. 218 2003396-404

    [48] Z. Shan, J.C. Jansen, W. Zhou, T. Maschmeyer, Appl. Catal., A 254 2003339-343

    [49] C. Simons, U. Hanefeld, I.W.C.E. Arends, R.A. Sheldon, T. Maschmeyer, Chem. Eur. J. 10 20045829-5835

    [50] C. Simons, U. Hanefeld, I.W.C.E. Arends, T. Maschmeyer, R.A. Sheldon, Top. Catal. 40 200635-44

    [51] C. Simons, U. Hanefeld, I.W.C.E. Arends, T. Maschmeyer, R.A. Sheldon, J. Catal. 239 2006212-219

    [52] T. Heikkilä, J. Salonen, J. Tuura, M.S. Hamdy, G. Mul, N. Kumar, T. Salmi, D.Y. Murzin, L. Laitinen, A.M. Kaukonen, J. Hirvonen, V.-P. Lehto, Int. J. Pharm. 331 2007133-138

    [53] T. Heikkilä, J. Salonen, J. Tuura, N. Kumar, T. Salmi, D.Y. Murzin, M.S. Hamdy, G. Mul, L. Laitinen, A.M. Kaukonen, J. Hirvonen, V.-P. Lehto, Drug Deliv. 14 2007337-347

    [54] P. Waller, Z. Shan, L. Marchese, G. Tartaglione, W. Zhou, J.C. Jansen, T. Maschmeyer, Chem. Eur. J. 10 20044970-4976

    [55] Z. Shan, W. Zhou, J.C. Jansen, C.Y. Yeh, J.H. Koegler, T. Maschmeyer, Stud. Surf. Sci. Catal. 141 2002635-640

    [56] A. Corma, V. Fornes, S.B. Pergher, T.L.M. Maesen, J.G. Buglass, Nature 396 1998353-356

    [57] A. Corma, V. Fornes, M. Sales Galletero, H. Garcia, C.J. Gomez-Garcia, Phys. Chem. Chem. Phys. 3 20011218-1222

    [58] U. Díaz, V. Fornés, A. Corma, Microporous Mesoporous Mater. 90 200673-78

    [59] Z. Shan, W.P.W. Gerhard, B.G. Maingay, P.J. Angevine, J.C. Jansen, C.Y. Yeh, T. Maschmeyer, F.M. Dautzenberg, L. Marchese, H.D.O. Pastore, Novel zeolite composite, method for making and catalytic application thereof July 15, 2004 US Patent 2004138051 A1

    [60] A.L. Santos Marques, J.L.J.L. Fontes Monteiro, H.O. Pastore, Microporous Mesoporous Mater. 32 1999131-145

    Chapter 2. Organoclays - Preparation, Properties and Applications

    Contents

    Introduction 31

    Preparation and Applications of Organoclays 33

    Organoclays prepared by intercalation of organic moieties in the interlayer space 34

    Organoclays prepared by grafting of organic moieties 34

    One step synthesis of organoclays 43

    Outlook 45

    References 46

    Keywords

    Clays

    Grafting

    Intercalation

    Ion exchange

    One step synthesis

    Organoclays

    Abstract

    The present chapter is devoted to recent developments in the area of layered inorganic–organic hybrid materials. The different approaches used for the preparation of organic nanoclays are summarized. These include intercalation, grafting and in situ incorporation of the organic moieties. Detail description of the methods is presented and their advantages/disadvantages are commented upon. Properties of the obtained hybrid materials are discussed. Examples demonstrating the potential of organoclays for various applications are given.

    1. Introduction

    A nature's remarkable feature is its ability to combine at the nanoscale level organic and inorganic components, thus allowing the construction of smart materials with a compromise between different properties or functions (mechanical, density, permeability, colour, hydrophobia, etc.). Current examples of natural organic-inorganic composites are crustacean carapaces or mollusc shells and bone or tooth tissues in vertebrates.

    The possibility to combine properties of organic and inorganic components in material design and processing was investigated since the early ages of humanity (Egyptian inks, green bodies of china ceramics, prehistoric frescos, etc.). Maya blue is a good example of man made material that combines the colour of organic pigment (indigo) and the resistance of an inorganic host (clay minerals such as palygorskite) leading to a synergic solid with properties and performances beyond those of a simple mixtures.[¹]

    Lately, a lot of attention is paid to a new class of layered materials comprising an organic component. In this chapter, we will provide basic information about hybrid layered materials (organoclays). First structure particularities will be presented, followed by the synthesis methods and the potential applications.

    Layered materials are built by the stacking of ‘two-dimensional’ units known as layers that are bond to each other through weak forces.[²,³] Layered structures can be classified in different categories depending on the type of the layer:

    neutral layers, e.g. brucite (Mg(OH)²) and other hydroxides, phosphates and chalcogenides, and various metal oxides such as V²O⁵,

    negatively charged layers with compensating cations in the interlayer space, e.g. widespread lamellar compounds in nature such as cationic clays (montmorillonite, hectorite, beidellite, etc.)

    positively charged layers with compensating anions in the interlayer space,[⁴] the most common of which are the layered double hydroxides (LDH), also called anionic clays.

    Layered materials from the second category are commonly found in nature (several minerals belong to the montmorillonite group-smectites), while LDH [hydrotalcite (HT)—anionic clays] are typically synthesized in laboratory conditions.[⁵]

    Another interesting group of layered materials is the phyllosilicates whose members can be from natural or synthetic origin.[⁶,⁷] In general, the members from this group have a structure that consists of one layer stacking in which planes of oxygen atoms coordinate to cations such as Si⁴+, A1³+, Mg²+, Fe³+ to form two-dimensional ‘sheets’. The coordination of cations in adjacent sheets typically alternates between tetrahedral and octahedral. Tetrahedral sheets, which commonly contain Si⁴+, consist of hexagonal or ditrigonal rings of oxygen tetrahedral linked by shared basal oxygens. The apical oxygens of these tetrahedra form the base of octahedral sheets having brucite-like or gibbsite-like structures that comprises Mg²+, A1³+, Li+, Fe²+ or Fe³+ cations. A regular repeating assemblage of sheets (e.g., tetrahedral—octahedral or tetrahedral—octahedral—tetrahedral) is referred to as a layer.

    Smectites, a subgroup of the phyllosilicates attract a lot of interest due to their swelling properties. Other subgroups include micas, kaolins, vermiculites, chlorites, talc, and pyrophyllite. These minerals are characterized by a 2:1 layer structure in which two tetrahedral sheets are attached on each side of an octahedral sheet via sharing of apical oxygens. As the apical oxygens from the tetrahedral sheet form ditrigonal or hexagonal rings, one oxygen from the octahedral sheet is located in the centre of each ring and is protonated, thus giving a structural hydroxyl (Fig. 2.1). In 2:1 phyllosilicates, isomorphous substitution of cations having different valences can lead to charge imbalances within a sheet. These may be partly balanced by the opposite charge of the adjacent sheet (e.g. a positively charged octahedral sheet may offset some of the negative charge associated with a tetrahedral sheet). The net charge imbalance on a 2:1 layer, if it occurs, is negative. This charge is referred as layer charge of the mineral and is balanced by larger cations (e.g. Na+, K+, Ca²+ and Mg²+) that coordinate to the basal surfaces of the tetrahedral sheets from adjacent layers. Because these charge-balancing cations are located between adjacent 2:1 layers they are referred to as ‘interlayer cations’, whose hydration is the reason for the high hydrophilicity of the phylosilicates.

    Figure 2.1. Schematic presentation of tetrahedral and octahedral sheets in layered materials (left) and a 2:1 phyllosilicate structure (right).

    Probably, the most widely exploited mineral from the phyllosilicate group is the montmorillonite due to its interesting physicochemical properties:

    particles of colloidal size;

    high specific surface area;

    presence of both Lewis and Brønsted acid sites;

    cation exchange capacity;

    presence of hydroxyl groups at the surface and at the edge of the layers.

    The last two properties render clay minerals hydrophilic, the latter being highly desired for a number of applications. These are indeed necessary when the clay minerals are dispersed in less polar solvents, incorporated in polymers to form clay-polymer nanocomposites or to improve gas and liquid adsorption properties. Finding the way to perform this hydrophobization and, thus, to form organoclays is a challenge, which leads to a growing interest in the field of clay minerals.[⁸–¹⁰]

    2. Preparation and Applications of Organoclays

    Different ways of making organoclays have been reported, namely:

    intercalation of organic molecules or organic cations by adsorption or ion exchange;

    grafting of organic functionalities to the hydroxyl groups presence at the edge of the clay mineral layers;

    direct synthesis of hybrid inorganic-organic materials.

    2.1. Organoclays prepared by intercalation of organic moieties in the interlayer space

    The term ‘intercalation’ refers to a process whereby a guest molecule or ion is inserted into a host lattice. The reaction is usually reversible. The structure of the host or intercalated compound is only slightly changed by the guest species. In three-dimensional systems, the sizes of the guest species are constrained by the host lattice dimensions. In lower dimensional systems, no such restriction exists and the layers may adjust their separation freely to accommodate guest species of different sizes.[⁵] Since the number of appropriate organic guest species and layered materials is vast, the possibilities to combine organic matter with layered hosts are enormous.

    Organoclays are usually prepared by modifying clays with cationic surfactants via ion exchange.[¹¹,¹²] This is accomplished through the replacement of inorganic exchangeable cations, for instance Na+, K+ and Ca²+, within the clay mineral structure with organic cations, for example with quaternary ammonium cations, such as hexadecyltrimethylammonium bromide,[¹³] hexadecylpyridine bromide[¹⁴,¹⁵] and 2-mercaptobenzimidazole.[¹⁶] Cation exchange reactions are most commonly performed by mixing aqueous dispersions of clay minerals and solutions containing surfactants such as organoalkylammonium salts.[¹⁷,¹⁸] Reaction rates depend on many parameters such as the guest compound, temperature and concentration, type of smectite and particle size.[¹⁹–²¹] Products are separated either by centrifugation or filtration and washed repeatedly. Depending on the solubility of the guest species, water-alcohol solution mixtures can be used as solvents. Whatever the method used, the arrangement of the intercalated surfactant cations strongly depends on the layer charge and the alkyl chain length resulting in materials with different interlayer spacing.[²¹] Organoclays are used to remove toxic compounds from the environment and reduce the dispersion of pollutants in soil, water and air. However, as cationic surfactants are weakly bonded to the clay mineral layers through electrostactic forces, they can easily be released in aqueous medium[²²] resulting in secondary pollution. This drawback can be circumvented by having a covalent linkage between organic moieties and the clay minerals (Fig. 2.2).

    Figure 2.2. Schematic presentation of an alkylammonium compound intercalated between two clay layers.

    2.2. Organoclays prepared by grafting of organic moieties

    Substantial efforts of research have been directed to functionalize clays in a durable way[²³–²⁵] to obtain organic-inorganic hybrids where the inorganic and organic components are linked via strong bonds (i.e. covalent or iono-covalent). This approach enables a durable immobilization of the reactive organic groups, preventing their leaching when they are used in liquid media. Such materials are expected to show superior properties in respect to unmodified clays when applied as adsorbents of heavy metals (low loading capacity, relatively weak binding strength and low selectivity).[²⁶,²⁷–³⁴]

    A covalent linkage can be achieved through the reaction between the layers reactive groups (hydroxyl groups) and an adequate molecule, which ensure higher chemical, structural and thermal stability for the compound. Alkylchlorosilane[³⁵] or organoalcoxysilane[³⁶]

    Enjoying the preview?
    Page 1 of 1