+ All Categories
Home > Documents > 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to...

4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to...

Date post: 22-Jul-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
29
O Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts Craig E. Barnes 137 Modern Surface Organometallic Chemistry. Edited by Jean-Marie Basset, Rinaldo Psaro, Dominique Roberto, and Renato Ugo Copyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim ISBN: 978-3-527-31972-5 4 4.1 Introduction The contributions that catalysts make to almost every facet of our daily lives cannot be overstated. Virtually every natural resource (crude oil, coal, biomass, minerals) and every source of energy (petrochemical fuels, nuclear, natural gas, solar) require the use of many catalysts before finished products (fine chemicals, pharmaceuti- cals, polymers, composites) arrive in our homes, offices and industries [1]. Cata- lysts also play increasingly important roles in solving some of the most challenging environmental problems that we currently face (global warming, the greenhouse effect, limited natural resources and pollution) [2]. While the use of heterogeneous catalysts is pervasive in science and technology, it rarely appears in the “credits” that herald the arrival of the next, new high strength material or medicine. Furthermore, the historical development of cataly- sis, the so-called “art and science” of the discipline has at times led to a picture of incompletely understood chemical systems that accomplish extraordinary chemi- cal transformations [3, 4]. This view of catalysis has changed in the past 50 years as new techniques and instrumentation have been developed to address some of the challenges posed in understanding heterogeneous catalysts. Recently, however, a new opportunity to quicken the pace of this change has presented itself that is having a unifying effect on several disciplines of science and technology involved in catalysis research: nanoscience (Figure 4.1) [5]. While many applications of nanoscience are outside the scope of catalysis (con- ducting nanowires, devices, electronic and magnetic nanomaterials) the nanome- ter size regime is located precisely where the disciplines of chemistry, materials science and heterogeneous catalysis meet. The synthesis and tailoring of discrete molecules has long been the goal and focus of the modern chemist [6] while developing strategies to prepare tailored materials is of great current interest in materials science [7]. Chemical engineers now spend much time and effort on understanding exactly what the active catalyst is and in developing methodologies by which more finely tailored catalysts can be prepared. Within the context of Druckfreigabe/approval for printing Without corrections/ ` ohne Korrekturen After corrections/ nach Ausfçhrung ` der Korrekturen Date/Datum: ................................... Signature/Zeichen: ............................ c04.indd 137 3/27/2009 4:47:43 PM
Transcript
Page 1: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

Building Block Approaches to Nanostructured, Single Site, Heterogeneous CatalystsCraig E. Barnes

137

Modern Surface Organometallic Chemistry. Edited by Jean-Marie Basset, Rinaldo Psaro, Dominique Roberto, and Renato UgoCopyright © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, WeinheimISBN: 978-3-527-31972-5

4

4.1Introduction

The contributions that catalysts make to almost every facet of our daily lives cannot be overstated. Virtually every natural resource (crude oil, coal, biomass, minerals) and every source of energy (petrochemical fuels, nuclear, natural gas, solar) require the use of many catalysts before finished products (fine chemicals, pharmaceuti-cals, polymers, composites) arrive in our homes, offices and industries [1]. Cata-lysts also play increasingly important roles in solving some of the most challenging environmental problems that we currently face (global warming, the greenhouse effect, limited natural resources and pollution) [2].

While the use of heterogeneous catalysts is pervasive in science and technology, it rarely appears in the “credits” that herald the arrival of the next, new high strength material or medicine. Furthermore, the historical development of cataly-sis, the so-called “art and science” of the discipline has at times led to a picture of incompletely understood chemical systems that accomplish extraordinary chemi-cal transformations [3, 4]. This view of catalysis has changed in the past 50 years as new techniques and instrumentation have been developed to address some of the challenges posed in understanding heterogeneous catalysts. Recently, however, a new opportunity to quicken the pace of this change has presented itself that is having a unifying effect on several disciplines of science and technology involved in catalysis research: nanoscience (Figure 4.1) [5].

While many applications of nanoscience are outside the scope of catalysis (con-ducting nanowires, devices, electronic and magnetic nanomaterials) the nanome-ter size regime is located precisely where the disciplines of chemistry, materials science and heterogeneous catalysis meet. The synthesis and tailoring of discrete molecules has long been the goal and focus of the modern chemist [6] while developing strategies to prepare tailored materials is of great current interest in materials science [7]. Chemical engineers now spend much time and effort on understanding exactly what the active catalyst is and in developing methodologies by which more finely tailored catalysts can be prepared. Within the context of

���������������������� ��� ��������

������ �������������

���� ����������

���� ���������������� ��������� �

��� ����������

����������� � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � � �

������������������ � � � � � � � � � � � � � � � � � � � � � � � � � � � �

c04.indd 137 3/27/2009 4:47:43 PM

Page 2: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

138 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

nanoscience, chemists are taking on the challenges involved in creating large, organized assemblies of molecules while catalysis scientists are focusing in on exactly what constitutes the optimum arrangements of metal and support to maxi-mize catalytic activity. The final piece of the puzzle is supplied by materials scien-tists who are now able to prepare materials that can be exquisitely tailored to the chemical and physical properties needed to optimize the support material for application in catalysis. The fact that the efforts of all three disciplines focus in the nanometer size regime represents an opportunity to bring heterogeneous catalysis to a greater level of recognition for the role it plays in science and tech-nological advancement.

4.2Current Challenges in Catalysis

The science of catalysis may be divided into two major parts: catalysts that function in the same phase as reactants and products (homogeneous) and those that func-tion in a different phase (heterogeneous). Both share many of the same goals and challenges: identify the components that make up the actual catalyst, understand mechanisms and optimize activity with respect to rate and selectivity.

But there are also significant differences. Supported catalysts frequently involve reactions that occur over a much broader range of temperatures (room tempera-

(a) (b)

2.20

1.80

1.40

1.00

0.600.0 2.0 4.0 6.0

Cluster Diameter (nm)

Act

ivit

y

30.0 nm

CO + ½ O2 Æ CO2

30.0

nm

8.0 10.0

(c)

Figure 4.1 (a) STM image of gold nanoparticles on a titania crystal surface; (b) plot of the catalytic activity of gold nanoparticles versus size in the selective oxidation of carbon monoxide with oxygen to carbon dioxide; (c) an artist’s rendition of the raft-like shapes of the nanoparticles. (From Reference [5] with permission.)

c04.indd 138 3/27/2009 4:47:44 PM

Page 3: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.3 What is a Nanostructured Catalyst? 139

ture to 1000 °C) such as hydrocarbon activation [8] and oxidations [9, 10] and therefore are much more thermally stable than homogeneous analogues. Most homogeneous catalysts are stable at below 200 °C [11, 12]. Homogeneous catalysts are in general better characterized than heterogeneous analogues, having bene-fited greatly from several characterization tools (e.g., NMR spectroscopy and X-ray diffraction) that allow scientists to precisely define at least the stable precursors to active species. Furthermore, the ligands found in homogeneous catalysts can be tailored to a much higher degree than is typically possible in the case of hetero-geneous analogues [12]. The most common type of heterogeneous catalyst involves a catalyst “ensemble” bound to the surface of a thermally and chemically robust support material. When metal atoms are involved, the ligands in their coordination spheres are a mixture of functionality on the support and substrates involved in catalysis. One of the current challenges in the development of new, “next-genera-tion” heterogeneous catalysts is to articulate new synthetic methodologies that allow one to design and tailor targeted catalyst ensembles on support surfaces with 100% fidelity [13]. A second challenge involves making sure that no other catalyti-cally active species is present in the system. Achieving these goals should open doors to whole new families of “ultraselective” catalysts [14, 15] that will be more efficient, selective, cost effective and more widely applicable than current catalysts. It is precisely in the area of catalyst design that new nanostructured heterogeneous catalysts are expected significantly impact future research in the area of catalysis[16].

4.3What is a Nanostructured Catalyst?

What is a “nanostructured” catalyst and how does it differ from the myriad of currently known catalysts? First, a nanostructured heterogeneous catalyst involves creating a local environment around the catalyst ensemble that is tailored to its function and application. The ensemble in this context may be interpreted as involving one or more metal atoms and two types of ligands: those that fix the ensemble onto the support surface and those that reside on the exposed, outer surface of the ensemble. Tailoring the catalyst and its immediate environment involves collecting together the targeted number of metal atoms and placing ligands around them in such a fashion that activity and attachment to the support are optimized for catalysis. Chemical groups that terminate the surface of the support have a dual role in this context. They serve as ligands to the metals, thus effecting catalytic activity, and they bind the metals to the support and prevent leaching or aggregation [17, 18]. In this context, organization or “nanostructuring” of surface functionality near the catalyst site is critical to obtaining single site surface bound catalysts.

A nanostructured heterogeneous catalyst also involves “organization” in several other contexts and at several other length scales. The distribution of active sites must be arranged, more often than not, to prevent aggregation under the

c04.indd 139 3/27/2009 4:47:44 PM

barnes
Sticky Note
Marked set by barnes
Page 4: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

140 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

conditions of catalysis. Controlling site “dispersion” in the initial preparation of a supported catalyst and maintaining dispersion in application are two additional challenges that must also be addressed in developing nanostructuring methodologies.

Two ways of achieving site isolated catalysts are to go to extremely low catalyst loadings or to “prepare” the support surface such that there are relatively few, widely spaced groups that will chemically bind a metal precursor from solution. However, since catalytic activity is directly related to the number of sites, a com-promise must be made with these strategies so that useful activities are obtained. A goal of next-generation nanostructured catalysts is not to settle for this compro-mise but to develop methodologies that achieve high site densities while maintain-ing site isolation and site fidelity.

The requirement that one prepare as many isolated catalyst sites as possible on the surface of a solid support can best be met with porous solids. Zeolites are good examples of a family of heterogeneous catalysts that address this requirement [19]. These crystalline, microporous (dpore < 2 nm) solids have seen broad application in catalysis (Figure 4.2) [20]. Limitations inherent with the microporous nature of these materials have led to the development of mesoporous (2 nm < dpore < 50 nm) supports where the combination of sol–gel technologies with surfactant templat-ing have led to a wide range of solids with controlled pore diameters and ordered pores within the material [21, 22]. At the same time, however, the targeted prepara-tion of mixed metal, mesoporous catalysts still faces several significant experimen-tal hurdles (vide infra) and thus remains a challenge confronting the development of next-generation catalysts.

Figure 4.2 Illustration of pores and molecular scaffolding in ITH zeolite. (From Reference [19].)

c04.indd 140 3/27/2009 4:47:45 PM

Page 5: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.5 Current Approaches to Nanostructured Catalysts 141

4.4Benefits of Nanostructuring Catalysts

The goals and benefits of nanostructured catalysts are synonymous with those that catalysis scientists have been working on for several decades. Tailored, single site catalysts should be more selective than systems that contain more than one active species. Atom efficiencies should approach 100% for single site catalysts, thus lowering their effective cost. If high site densities can be achieved while maintain-ing site isolation, then high activities should be attainable. Finally, supported cata-lysts that sit in tailored pockets of surface ligands are expected to be more strongly bound to the support and therefore more stable under the conditions of application. All of these benefits are well known in catalysis science. Simultaneously satisfying all of these requirements in the design and construction of new catalysts is the principle challenge that the science of heterogeneous catalysis currently faces. Nanoscience in the field of catalysis is confronted by the same challenges.

The following sections are organized into two parts. The first presents a brief overview of several approaches to preparing nanostructured supports and catalysts that are under current development. Approaches that incorporate building blocks in either part of the system (catalyst or support) will be highlighted. The second part describes a new approach to preparing nanostructured catalysts that we are currently developing in our laboratories.

4.5Current Approaches to Nanostructured Catalysts

If efforts to achieve and tailor metal dispersion on the surfaces of support materials are included as a form of nanostructuring catalysts then, indeed, catalysis scien-tists were among the first to work in the area of nanostructured materials [23]. One of the simplest and most widely used methods to bind metals to support surfaces is the incipient wetness technique. In most cases this involves little more than dissolving a catalyst precursor in a solvent (frequently water) containing the suspended support and then removing the solvent by evaporation. Ligands origi-nally coordinated to the metal precursor are replaced by hydroxyl or oxide groups on the surface of the support [24]. Much effort has been devoted to influencing the distribution and dispersion of supported catalysts made in this fashion. Thermal pretreatments of the support that remove adsorbed water and then reduce the number of surface hydroxyl groups through condensation reactions may be used to “tailor” the distribution of potential binding sites for a catalyst [25–27]. From the point of view of the catalyst, going to very low catalyst loadings is gener-ally thought to lead to atomic dispersions of the metal centers [28, 29].

However, as mentioned above, a compromise must be made between low loading and reasonable activity. Furthermore, although hydroxyl groups may on average appear to be well separated on a dehydroxylated surface, there is evidence

c04.indd 141 3/27/2009 4:47:45 PM

Page 6: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

142 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

that the distribution is still far from uniform or homogeneous (Figure 4.3) [30, 31]. Many different configurations of hydroxyl groups (geminal, vicinal diols, clusters of hydroxyls) can exist, each of which can lead to different catalyst species. Therefore, tailored dispersions of surface binding sites via thermal pretreatments coupled with the method of incipient wetness may be considered as nanostructur-ing a catalyst, but only in its most rudimentary form [32].

In this context, much effort has also been invested in controlling the nuclearity of the catalyst ensemble through the selection of its precursor. One area in which considerable progress has been made involves the adsorption of small polynuclear clusters onto supports [33]. Examples involving the immobilization of small, preformed polynuclear clusters on supports are the reactions of carbonyl clusters of the late metals [16, 34], the binding of polyoxometalates (POMs) and their neutral alkoxy analogues [35] and heteropolyacids such as the Keggin cluster [36, 37].

The immobilization of carbonyl clusters on the surfaces of metal oxides has been extensively investigated by many groups [38]. TEM evidence for the formation of several discrete clusters such as “Os5” on silica (Figure 4.4) has been described [39]. In most cases, however, the chemistry that leads to formation and binding of the cluster to the surface can be complex, difficult to predict and control, and involves the same heterogeneous distribution of surface functionality alluded to above. Preventing fragmentation or aggregation of precursors on the surface are also constant challenges. The set of ligands at the interface between the metal ensemble and surface usually varies across the spectrum of bound clusters. For small nuclearity clusters such differences can have profound effects on the cata-lytic activities that otherwise identical clusters might exhibit [40].

Supported nanoparticles are related to the idea of starting with polynuclear cluster precursors. While there is no clear line that divides polynuclear clusters from nanoparticles, clusters are generally small, low nuclearity (MnLx; n = 3–20), structurally well-characterized species approximately 1–2 nm in size. Nanoparti-cles are larger (>2 nm) and frequently defined by a size distribution rather than a discrete number of atoms and ligands [41]. In the area of catalysis, gold nanopar-

Si Si Si SiO

OH

OHHOOH OH

Si SiSi Si

O

OOOO

OOMM M

Figure 4.3 Different configurations of surface hydroxyl groups produce different, atomically dispersed surface bound metal species.

c04.indd 142 3/27/2009 4:47:45 PM

Page 7: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.5 Current Approaches to Nanostructured Catalysts 143

ticles supported on titania are one of the most cited examples in which particles in the 1–3 nm size range exhibit unusual catalytic activities relative to either atomic or bulk gold [42]. Although this system and many others currently under study are often pointed to as examples of nanostructured catalysts, they satisfy only some of the criteria above and none rigorously. Distributions in particle diameter that span even 1 nm can cause changes in nuclearity that involve thousands of metal atoms. Furthermore, although model studies under UHV conditions involving clean single-crystal surfaces reduce the interfacial heterogeneity where particles are in contact with the surface, the steps, corners and other types of surface defects lead to questions regarding catalytic activities that are still under active debate [43].

Despite having literally thousands of mononuclear and polynuclear cluster [44–46] formulations to choose from, there are no generally applicable methods available for preparing rigorously defined, single site catalysts on supports using simple adsorption procedures. Furthermore, heterogeneous distributions of surface functionalities present on currently available supports will generally lead to a distribution of surface species that in turn gives rise to multicomponent cata-lyst systems.

Within a discussion of nanostructured catalysts, mention must be made of zeolitic systems. Zeolites are a broad family of natural and synthetic aluminosili-cates that exhibit two important properties that makes them ideal for consideration as heterogeneous catalysts: they are crystalline and porous. Crystallinity brings with it precise definition at the atomic scale that is absent with amorphous or polycrystalline metal oxides. The combination of a well-defined structure and

Figure 4.4 TEM image of the edge of silica showing Os5 particles on the surface. (From Reference [39] with permission.)

c04.indd 143 3/27/2009 4:47:46 PM

Page 8: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

144 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

microporosity is an ideal combination to meet the requirements of high surface area and well-defined catalyst binding sites in nanostructured catalysts. Examples of well-studied single site zeolite catalyst systems thought to involve framework replacement of silicon are aluminosilicates [20], titanium silicalite, TS-1 [47, 48] and more recently Sn-β-zeolite [49].

The four-coordinate, framework aluminium sites in the walls that define the pore structures of zeolites may be converted into strong Brønsted acid centers [50]. The combination of an ordered array of highly active acid sites confined within the micropores of a thermally robust support is, in many respects, a “dream come true” in the science of catalysis. Numerous technologically important applications of heterogeneous catalysts based upon zeolites have been developed [20]. Further-more, the confinement offered by the chambers and pores of zeolites has been used advantageously to bind a wide variety of metals into the framework and on the pore walls, ultimately resulting in a high degree of control of the reactions that they catalyze. In many ways, zeolites are prototypical examples of well-defined, nanostructured catalysts and the benefits that accrue when all the criteria for true nanostructuring are achieved.

There are, however, two limitations associated with preparation and application of zeolite based catalysts. First, hydrothermal syntheses limit the extent to which zeolites can be tailored with respect to intended application. Many recipes involving metals that are interesting in terms of catalysis lead either to disruption of the balance needed for template-directed pore formation rather than phase separation that produces macroscopic domains of zeolite and metal oxide without incorporating the metal into the zeolite. When this happens, the benefits of catalysis in confined chambers are lost. Second, hydrothermal synthesis of zeolitic, silicate based solids is also currently limited to microporous materials. While the wonderfully useful molecular sieving ability is derived precisely from this property, it also limits the sizes of substrates that can access catalyst sites as well as mass transfer rates of substrates and products to and from internal active sites.

The limitation of microporosity in zeolites has been a driving force behind the development of surfactant templating of sol–gel technologies for preparing ordered, mesoporous metal oxide solids [51]. This form of nanostructuring has seen explosive growth following its discovery [52]. A well-understood, broadly applicable methodology for using the phase behavior of surfactants to template metal oxide matrices in aqueous solution has been articulated through the work of many research groups around the world (Figure 4.5) [53]. The most well-developed systems involve the low temperature solvolysis and condensation of orthosilicate precursors in the presence of micellular phases of surfactants. Both the ordering of and size of the mesopores may be tailored through choice of sur-factant and conditions. Pore sizes are continuously variable across virtually the entire mesopore range (2–50 nm).

A fundamental difference between zeolites and sol–gel derived metal oxides is found in the “order” exhibited by each of these systems. Zeolites are crystalline systems so that, to first order, all pores, chambers and the atomic scale positioning

1

c04.indd 144 3/27/2009 4:47:46 PM

Page 9: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.5 Current Approaches to Nanostructured Catalysts 145

of every atom in the zeolite scaffolding is well defined and identical throughout the material. Sol–gel reactions are generally conducted at lower temperatures, thus leading to porous glasses in which the walls defining the pores are largely amor-phous. Therefore, while ordered arrangements of large pores are possible via surfactant templating, the pore surfaces are quite similar to the heterogeneously functionalized surfaces of normal silicates and metal oxides.

Another challenge faced by sol–gel technologies involves controlling the disper-sion of different metals within a mixed metal (e.g., silicon and titanium) matrix. The solvolysis and condensation steps for metal alkoxide precursors involved in sol–gel reactions can be quite different from that of orthosilicates, which often leads to the loss of dispersion and formation of separate silica and other metal oxide domains [54].

Imprinting the surface of silicates derived from sol–gel reactions is an interest-ing approach to preparing nanostructured materials that could have a significant impact in catalysis[55–57]. In its simplest form metal ions are added to the sol–gel solution and allowed to bind to the surfaces of the developing pores. In this way, surface functionality is organized according to the electronic and steric signatures of the imprinting metal cation or complex anion. The most common applications of imprinted sol–gel materials is in the areas of ion recognition, separation

Figure 4.5 TEM image of MCM-41, a surfactant templated silica with ordered mesopores. (Kindly provided by S. Dai, Oak Ridge National Laboratory, Oak Ridge, TN.)

c04.indd 145 3/27/2009 4:47:47 PM

Page 10: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

146 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

science and selective ion sorbents [58]. There have also been several efforts to prepare supported catalysts that combine imprinting techniques with surface functionalization [59–61] and grafted ligands, in essence “heterogenizing” homo-geneous catalysts [60, 62, 63].

The picture that emerges from this overview of new and traditional approaches to nanostructuring catalysts is one in which progress has been made, but also where significant challenges remain. All synthetic strategies that rely on adsorp-tion to the surfaces of preexisting supports face the challenge posed by the inher-ent heterogeneity of the surface functionality that binds to metals [13]. This usually leads to a variety of potential catalysts in the system, reducing atom efficiency and frequently leading to lower selectivities in application. While zeolites exemplify many of the properties of nanostructured, single site catalysts they suffer from the limit of microporosity and the requirements of hydrothermal synthesis. Sol–gel technologies open the door to mesoscale porosity but generally at the expense of amorphous surface structure and heterogeneous functionality. New approaches to preparing nanostructured support–catalyst systems will have to be developed before next-generation catalysts can emerge.

4.6Building Block Approaches to Nanostructured Materials and Catalysis

Interest in the use of molecular building blocks (mbbs) to prepare nanostructured solids with applications to catalysis has recently seen significant progress. A broad new class of porous, crystalline, building block solids has recently been described that is expected to have great impact on many technologies, including gas storage, separation, chemical delivery systems as well as catalysis. [64, 65]. Using a com-bination of rigid metal and organic based building blocks, a broad array of new metal-organic framework (MOF), covalent organic frameworks (COF) [66] and most recently zeolitic imidazolate framework (ZIF) materials [67] have been pre-pared (Figure 4.6). Furthermore, a design theory [68] based upon the shapes of rigid building blocks and linking units, and the number of connection points between each, is being developed that should allow researchers to plan syntheses of porous materials with targeted properties such as pore and cavity sizes, overall surface area and thermal stability. In contrast to zeolites, these materials fre-quently contain stoichiometric numbers of metal atoms in their frameworks and thus have great potential for metal-mediated catalysis. At the same time, however, the metals are usually coordinatively saturated and it is therefore not clear how they can participate in catalytic cycles that usually require rather significant changes in coordination geometries and ligands and still participate in holding the framework together. Nonetheless, applications of these zeolite-like materials to catalysis are beginning to appear [69, 70]. While some MOFs exhibit limitations with respect to thermal and chemical stability in application, the metal centers in these materials satisfy many of the requirements necessary to consider them as well-defined, single site, nanostructured catalysts.

c04.indd 146 3/27/2009 4:47:47 PM

Page 11: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.6 Building Block Approaches to Nanostructured Materials and Catalysis 147

Polyoxometalates and oxoalkoxides were mentioned above in the context of potential polynuclear catalysts within nanostructured solids. They can also serve equally well as building blocks that make up the bulk of the solid matrix [71]. Polyhedral silsesquioxanes (POSSs) and silicates and derivatives thereof have also been used to prepare building block solids[72]. Similar to POSSs are the spherosilicates, SinO2.5n (n = 6, 8, 10, 12, 14) [73]. They are ideal candidates for building block materials [74] and some work in this area has already been reported [75, 76]. An approach utilizing solvolysis and condensation reactions related to sol–gel methodologies with the cubic Si8O20 core has been described by Klemperer and coworkers [77]. Finally, Feher and Walker have described the reaction of the trimethyltin substituted cubic silicate, Si8O12(OSnMe3)8, with several metal chlorides under anhydrous conditions [78]. The linking reaction occurs under mild conditions and involves a simple metathesis reaction in which M-O-Sicube links are formed with the loss of trimethyltin chloride. This work also illustrates the potential that non-aqueous based linking reactions have in strategies to construct metal oxide materials from molecular building blocks [79].

The challenges that stand between heterogeneous catalysts prepared by traditional methodologies and “next-generation” heterogeneous catalysts begin with the need for new synthetic methodologies that simultaneously control

2

Figure 4.6 Illustration of a covalent organic framework (COF) material composed of hexahydroxytriphenylene and either tetra(4-dihydroxyborylphenyl)methane or silane crosslinking moieties. (From Reference [66] with permission.)

c04.indd 147 3/27/2009 4:47:48 PM

Page 12: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

148 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

functionality that define the active site as well as structure at multiple length scales. At the atomic level (1–2 nm) it will be necessary to prepare supported catalysts that have only a single type of active site throughout the matrix in which the metal nuclearity and ligands that define the ensemble as well as bind it to the support are uniformly identical. Surface functionality around a catalyst ensemble could be optimized by having it or a suitable precursor present as the support is formed and play a role in templating the surface to which it will be bound. High site densities within the matrix dictate the need for porous solids while mesoporosity will open the door to a broad spectrum of substrates as well as high mass transport rates.

4.7Nanostructured Catalysts via a Non-Aqueous Building Block Methodology

For the past five years, we have investigated strategies by which multifunctional-ized building blocks may be linked together to produce tailored distributions of site isolated, single site metal based supported catalysts. We wished to develop a broadly applicable synthetic strategy for the preparation of catalytically active metals on metal oxide supports in which high densities of identical sites are pro-duced. Initial areas of interest were solid acid and oxidation chemistry such as hydrocarbon functionalization, oxidation, reforming and cracking. Therefore, ther-mally robust catalyst ensembles involving high valent metals and main group elements were sought. Ligands normally associated with grafted organometallic complexes would generally not survive the conditions of catalysis in this context and were therefore not the initial focus of our investigations. Readily available metal halide or mixed alkoxy-halide complexes were chosen as the synthetic pre-cursors of the eventual catalysts in the matrix.

The requirement that surface functionality at the interface with the catalyst ensemble be identical for every site obviates most approaches that simply adsorb a catalyst precursor onto an existing array of surface functionality found on most metal oxides. One way to meet this requirement is through synthetic strategies that involve a precursor to the actual catalyst in the formation of the support material. This strategy is identical to imprinting strategies used in molecu-lar and ion recognition studies in the area of separation science [56, 80, 81]. For supported catalysts, however, the goal is to control and tailor interaction with the surface so that strong bonding of the catalyst ensemble to the surface permanently attaches it to the support. Aqueous based approaches to patterned metal oxides (e.g., zeolite and sol–gel) have been investigated extensively and their advantages and limitations summarized briefly above. Feher’s metathesis reaction involving trialkyl tin as a substitute for the proton on a terminal metal hydroxide circumvents several problems encountered in aqueous solution [78]. Trialkyl tin groups cleanly react with M-Cl groups to produce the desired Si-O-M linkage and the alkyl tin chloride, which can in most cases be easily removed from the system (Equation 4.2):

c04.indd 148 3/27/2009 4:47:48 PM

Page 13: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.7 Nanostructured Catalysts via a Non-Aqueous Building Block Methodology 149

Si M H2O

O

Si

OH

+ M

HO

+

Si Si

O

M M

O

(4.1)

SnR3

O+ M

Cl

Si M

O

ClSnR3+

(4.2)

Recasting the linking reaction in terms of complementary “A” and “B” func-tional groups effectively eliminates undesired homocondensation reactions that can lead to speciation and loss of dispersion in the final matrix.

The second theme we chose to pursue in preparing nanostructured catalysts was the use of rigid, nanometer sized building blocks as the main structure directing entities on which the solid matrix would be built. The spherosilicates are a family of potential building block molecules that are well suited for catalyst synthesis. The most easily synthesized member of the family is the cubic analog, which has the Si8O12 core (Figure 4.7). Either the neutral octahydrido (Si8O12H8) or the octa-anion, [TMA]8[Si8O20]nH2O (TMA = tetramethylammonium), may be converted into the desired octakis(trimethyltin) compound to serve as a materials building block.

As shown, the cubic Si8O12 core is relatively rigid and the terminal oxygen atoms associated with one building block are well separated, such that the formation of MOx groups is prevented during growth of the matrix [82]. Equally important, the geometry and bonds along the Si–O–Si bridge (cube edge) between two metal

The core of the Si8O20building block

Linking point separations:O1···O2 ~5 ÅO1···O3 ~7 ÅO1···O4 ~9 Å

O1

O2

O4

O3

Si

Figure 4.7 Illustration of the core of the cubic spherosilicate Si8O20 building block, showing the overall metrics and distances separating the terminal oxygen linking points.

c04.indd 149 3/27/2009 4:47:49 PM

Page 14: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

150 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

centers effectively isolate the terminal oxygen atoms from one another in terms of potential chemical interactions important to catalysis (in practice metal ions bound to these oxygens will generally be much further separated from one another in the matrix).

Reaction of the octakis(trimethyltin) silicate cube (hereafter referred to as the “tin cube”) with high valent metal and main group halides at room temperature in aprotic organic solvents leads to crosslinked matrices of Si8O20 cubes connected by the single atoms derived from the halide reagent (Figure 4.8). It is important to note two properties of these matrices from the outset. First, if simple mono-nuclear halide complexes are used in the reaction, then rigorous atomic dispersity of the linking atom will be achieved throughout the matrix. Second, because the linking reaction between trialkyltin groups and metal chloride linking reagents is irreversible, the substitution pattern around any cube or linking atom should be more or less random. This leads to the prediction that these building block matri-ces should be amorphous glasses, which is observed in practice. Finally, it is dif-ficult to imagine that the cubes would link together to form dense solids under these conditions [74]. A highly “defected” glass-like solid with considerable void volume throughout the matrix is expected and in general this is observed. The void spaces that develop between cubes are expected be quite irregular but should scale approximately with the size of the main structure-directing unit, that is the building block. Thus, we predict that matrices composed of larger and larger build-ing blocks will be porous and the size of the “pores” should increase with the size of the building block. Both of these predictions are realized in practice [82].

In one of the first reactions investigated, the tin cube was exposed to several equivalents of SiCl4 in toluene between room temperature and 50 °C. The solid-state 29Si NMR (SS NMR) spectrum of the product (Figure 4.9) shows that a dis-tribution of chlorosiloxane linking groups with different “connectivities” was obtained in the crosslinking reaction. We use the term connectivity to highlight a critical property of these linking groups. 1-connected groups derived from SiCl4 do not actually link cubes together in the matrix but are capping –SiCl3 groups. 2-connected groups are –SiCl2– moieties, 3-connected are ≡SiCl and 4-connected

3

– ClSnR3+ MCl4

Si8O12(SnR3)8

MM

M

M

MCl2

Cl3M

MM

Cl3M

MCl3 MCl3MCl3

Cl

Cl

Cl

Cl

Cl

M = Si or Ti“tin cube”

Figure 4.8 Illustration of the random orientation of Si8O20 cubes and random distribution of linking MClx groups that form when the tin cube is reacted with an arbitrary amount of a metal chloride crosslinking reagent such as SiCl4 or TiCl4.

c04.indd 150 3/27/2009 4:47:49 PM

Page 15: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.8 A Model for the Growth of Building Block Matrices and a Nanostructuring Strategy 151

linking groups are synonymous with embedded Q4 silicate centers found in sili-cates. Note that every time a trialkytin group is replaced by a bond to a linking group a Q4 silicon is also formed at that corner of the Si8O20 cage.

Titanium tetrachloride and vanadyl chloride (VOCl3) [83] give similar distribu-tions of linking groups when reacted with the tin cube. Early investigations showed that simply exposing the tin cube to an arbitrary amount of a linker will not, a priori, lead to the type of nanostructuring we seek. Although atomic dispersity is achieved, a distribution of different catalyst sites (i.e., sites with different connec-tivity) develops within the matrix. To address this problem, we conducted a brief study of how these bb-matrices grow as crosslinking progresses. This led to the development of a simple, straightforward synthetic approach for preparing bb-matrices that have exactly one type of linking group catalyst center in them [84].

4.8A Model for the Growth of Building Block Matrices and a Nanostructuring Strategy

Linking the entire building block matrix together with the active metal is generally not desirable. Therefore, multiple linking reagents – one to insert the active metal into the matrix and others to knit the matrix together into a stable material – are needed to produce viable catalysts. Fortunately, chlorosilane reagents satisfy the second requirement quite well in that they form robust silicate (when SiCl4 is used) or siloxane linkages (e.g., MeSiCl3, Me2SiCl2 are used) between building blocks. The alkyl groups in these eventual siloxane groups play several roles in tailoring the properties of these matrices. First, they control the overall crosslinking within the matrix simply by limiting the maximum connectivity that may be achieved around them (in practice, we find that all chlorides in silane reagents remain active in forming bonds to cubes), which will influence how cubes aggregate and the overall pore structures in the final materials. Second, the polarity of the matrix can be controlled by changing both the identity and number of blocking groups in the matrix. The ability to tailor both the polarity and crosslinking throughout the

-44 ppm

SiCl Cl Cl

-68 -92

Si

Cl

(Sicube + 4-connected

silicon links)

SiCl Cl

-120-100-80-60-40 ppm

4 SiCl4 + Si8O20(SnMe3)8 Si8O20 matrix

-112 Q4

Figure 4.9 29Si SS NMR spectrum (MAS) showing the three types of chlorosiloxane linking groups that develop in the general reaction of SiCl4 with the tin cube.

c04.indd 151 3/27/2009 4:47:50 PM

Page 16: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

152 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

matrix is a very powerful tool for adjusting several parameters known to influence catalytic activities and selectivities in porous solids.

We have studied the reactions of chlorosilanes with the tin cube to determine the effects of solvent, time and temperature on tailoring the properties of the bb-matrices that are formed [84]. Reaction of the first chloride ligand to make 1-connected capping groups is much faster than subsequent reaction of other chlorides. To obtain high initial degrees of crosslinking in the matrix therefore requires that both the stoichiometry and time of reaction be carefully controlled. Alternatively, since the first chloride reacts faster than subsequent ones, inverse addition of the tin cube to excesses of silyl chlorides can give rise to complemen-tary functionalized building blocks such as Si8O20(SiCl3)8 (M.-Y. Lee and C.E. Barnes, unpublished results).

As crosslinking begins, first small oligomers and then colloidal particles form. The size of oligomers and colloids increases until phase separation occurs to form sol–gel like materials. The point of phase separation can be influenced by the solvent. Poorly solvating solvents (hexane) cause early phase separation while more polar solvents (toluene, methylene chloride) can delay its onset. Phase sepa-ration appears to play an important role in the overall crosslinking, porosity and surface area that develop in these matrices. If it occurs too early, the degree of crosslinking needed to develop porous solids is not reached and low surface area solids are obtained. If crosslinking continues long enough, then high surface area (100–600 m2 g–1) materials are obtained. In all of these materials, bimodal distribu-tions of pore sizes containing both micro- and mesoporous components are obtained.

The stoichiometry of linking reagent to cube (given either as M : cube or SnR3 : Cl ratios) is of critical importance in obtaining single site catalysts. Stoichiometric or excess linking reagents lead to distributions in linking group connectivity or lightly crosslinked matrices with a large proportion of capping groups on the surface. Limiting the amounts of linking reagents that are allowed to react completely leads to matrices in which all of the linking groups will have achieved exactly the same connectivity to the matrix. These linking groups are the precursors to the active sites for catalysis in the matrix.

The matrix will stop growing when all linking reagent is consumed and a dis-tribution of oligomers with variable degrees of average crosslinking and residual trimethyl tin groups will have developed. A second, more subtle point about these three-dimensionally crosslinked matrices involves their rigidity. When enough crosslinks are formed between cubes, they and the linking groups become rigidly fixed in space relative to one another. This has the important consequence of spatially isolating unreacted tin and silyl (or metal) chloride groups in the matrix. Thus, there is an intermediate phase of growth of these matrices where both M-Cl and cube-OSnR3 groups can be present simultaneously but are prevented from reacting with one another because of the rigidity of the matrix. Once isolated from one another, trimethyl tin groups can no longer participate in further crosslinking reactions. Addition of other chloride-containing reactants yields only capping groups regardless of the number of tin groups present.

4

c04.indd 152 3/27/2009 4:47:50 PM

Page 17: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.9 A General Procedure for Preparing Nanostructured Catalysts in Silicate Matrices 153

The presence of any residual trialkyl tin groups in the final matrix is not desir-able in the context of catalysis. Tin is a catalytically active metal and thus could complicate and potentially interfere with reactions of other active metals in the matrix. Fortunately we found that all residual tin groups in these matrices can be replaced with trimethyl-silyl (TMS) groups simply by exposure to TMSCl in the gas phase. This observation also indicates that none of these sites becomes inac-cessible to further reaction when crosslinking in the matrix is appropriately adjusted.

4.9A General Procedure for Preparing Nanostructured Catalysts in Silicate Matrices

The observations described above and the qualitative model for the growth phases of these bb-matrices may be recast into a simple, straightforward methodology for preparing nanostructured, single site catalyst systems that we refer to as the “method of sequential additions.” Because of the ease with which individual silox-ane groups may be observed and identified via 29Si solid state NMR (SS NMR), we conducted an initial model study using silyl chloride reagents to illustrate the utility of this methodology in ultimately creating targeted metal sites on the sur-faces of these bb-matrices [84].

The foundation of this methodology rests on controlling the connectivity that a catalyst center has in the matrix through a series of carefully controlled doses of linking reagents to solutions of the tin cube (Figure 4.10). To obtain fully embed-ded, framework catalyst sites an initial, limiting dose of the metal chloride (Me2SiCl2 in the model study) is reacted with the tin cube. To date we have seen that all chloride ligands are active in crosslinking for MCl4 (MCl4, M = Si, Ti, V, Sn) reagents and as high as 5-connected centers may be achieved in the case of WCl6 under similar conditions.

After completion of the reaction, the mixture will consist of n-connected oligo-mers and possibly excess tin cube. Critically, all the crosslinking metal centers in these oligomers will have the same maximum connectivity to cubes. While addi-tional doses of different crosslinking reagents generally serve to knit these n-connected oligomers together into a robust bb-matrix, one can easily imagine broadening this strategy to prepare bifunctional catalysts via additional doses of catalytically active linkers.

Spectra (a) and (b) in Figure 4.11 show the MAS and CPMAS spectra for a Si8O20 matrix that was first exposed to a limiting amount of dimethyldichloro silane (1 Me2SiCl2 : cube or 2Cl : 8SnMe3) followed by one or more doses of silicon tetrachlo-ride to finish crosslinking the matrix and remove all tin from the matrix (TMSCl could also be used to remove the last tin groups). Cross polarization to silicon from the methyl groups on the dimethyl siloxane groups selectively enhances these signals in the matrix and allows for unambiguous assignments to be made. It can be seen from MAS and CPMAS spectra that only 2-connected O–SiMe2–O linking groups are present, exactly as desired.

c04.indd 153 3/27/2009 4:47:50 PM

barnes
Sticky Note
Marked set by barnes
barnes
Underline
Page 18: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

154 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

silicon grease

(∫SiO)SiCl3

(∫SiO)2SiCl2

(∫SiO)3SiCl

(∫SiO)2SiMe2

MAS

ppm

CPMAS

“Surface” (∫SiO)SiClMe2

(∫SiO)SiClMe2

Si(OSi∫)4

“Embedded” Me2SiCl2

MAS

CPMAS

-125-75-2525 0

a

b

c

d

Figure 4.11 29Si SS NMR spectra (MAS+CPMAS) for two samples, showing selective cross polarization enhancement of the 2-connected SiMe2 groups in the “embedded” sample [spectra (a) and (b)] and 1-connected SiMe2 groups in the “surface” sample [(c) and (d)].

SiCl4

MCln MClX

= SiClx links

MCln

catalyst ensemblescolloidal particles

MM

MM

“capping” catalyst sites

SiCl4

rigid building block platform

M

M

M

porous, supportedcatalyst particles

building block

Catalyst Nanostructuring Strategy “embedded”catalyst sites

Figure 4.10 Illustration of the “method of sequential additions” strategy for preparing nanostructured solids containing atomically dispersed metals in building block matrices.

c04.indd 154 3/27/2009 4:47:51 PM

Page 19: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.9 A General Procedure for Preparing Nanostructured Catalysts in Silicate Matrices 155

The procedure to obtain the opposite type of catalyst center in the matrix, 1-con-nected “surface” O–ClSiMe2 groups, simply involves reversing the dosing sequence of reagents described above. Initial treatment of a solution of the tin cube with an amount of SiCl4 calculated to reach the point where all cubes are rigidly held in place but still leaving some unreacted trialkyl tin groups in the matrix produces a Si8O20-bb “platform.” As noted above, once a rigid three-dimensionally crosslinked matrix is obtained, additional metal or silyl chlorides can no long crosslink tin sites in the matrix. The second set of 29Si SS NMR spectra [(c) and (d)] in Figure 4.11 show that a subsequent dose of Cl2SiMe2 gives rise to only the desired 1-con-nected –ClSiMe2 groups.

Catalyst centers with connectivities intermediate between the minimum and maximum can be obtained by replacing chlorides with several blocking groups. Alkyl groups serve well for this purpose in the case of silanes (SiCl4, MeSiCl3, Me2SiCl2, TMSCl) while alkoxy groups can be used in the case of transition metal chlorides [85]. Notably, while matrices prepared in this manner are amorphous glasses they exhibit “nanostructuring” in the tailoring and isolation of catalysts sites from one another, in the control of the crosslinking and, finally, in the ability to adjust the polarity and porosity of the matrix [86].

Having demonstrated the efficacy of this nanostructuring methodology we have turned our attention to preparing metal based catalyst systems. In the model study above, the ability of solid-state NMR data to resolve and identify individual groups throughout the matrix makes it an extremely powerful tool in the context of site identification. However, few catalytically active metals have spin active isotopes that give as clear a picture as we observed in the case of silicon [87]. EXAFS is a technique beloved by catalysis scientists because of its potential to provide struc-tural data about the catalyst sites present on supports that exhibit no long-range order [88]. Only in the case of single site catalysts, however, is the interpretation of EXAFS data relatively straightforward. Distributions of catalyst sites that are more commonly encountered can lead to very complex EXAFS in which a com-posite picture derived from all the different species present is obtained. For this reason, reliance on EXAFS data alone to prove the identity of supported catalysts is seldom advisable. When SS NMR data are not available, defining the sites that we construct is more challenging and we combine as many independent lines of spectroscopic and analytical measurements as possible with EXAFS data to piece together the composition of the catalyst ensemble.

We have recently developed another type of analysis that is quick and convenient but provides extremely important information about linking group connectivities in the matrices we prepare: gravimetric measurements. Simple weighings taken before and after doses of linking reagents have allowed us to ascertain metal con-nectivities quickly and have proven essential in verifying that the linking reaction has occurred as expected and in interpreting EXAFS data.

An example will illustrate the power of gravimetric analysis in the context of this building block strategy. If it is assumed that the crosslinking reaction proceeds as illustrated in Equation 4.2, every time a link is formed a trialkyl tin chloride byproduct is released from the matrix. Given that the precise number of tin groups

c04.indd 155 3/27/2009 4:47:51 PM

Page 20: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

156 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

and amounts of linking reagents are known at the beginning of each dose then, in a stepwise manner, it is possible to follow the progress of crosslinking simply by weighing the reaction mixture after each dosing reaction. However, only average connectivities achieved by each added linking reagent may be determined gravi-metrically. In the two extremes where fully embedded or surface catalysts are targeted, the connectivity of every linking group must be identical and therefore measurement of the average defines the catalyst sites unambiguously. As an example, when a limiting amount of AlCl3 was added to the tin cube in toluene, a weight loss consistent with reaction of all chloride ligands was measured. There-fore, we can conclude that all the aluminium sites in the matrix achieved their maximum connectivity of 3. Subsequent doses of SiCl4 and TMSCl were also fol-lowed in this manner (Table 4.1). Note that in the case of capping TMS groups we can assume a connectivity of one and gravimetric data can then be used to count the number of TMS groups added to the matrix.

With this evidence in hand, we predict that an aluminosilicate matrix has been formed in which all the aluminium sites are bound to exactly three oxygens from different Si8O20 cubes. Stepping back from the details of this reaction, it is of inter-est to compare this bb-aluminosilicate with more traditional zeolites and silicas. Since the bb-solid has never been exposed to protic reagents such as water or alcohols there should be no Brønsted acid sites in the matrix (no bands, broad or sharp, are found in the 3700–3000 cm−1 region of IR spectra). All the aluminium atoms in the material should be identical and uniformly Lewis acidic. This is a unique starting place from which to study the properties of these interesting sites in the context of catalysis. Both 27Al SS NMR and base binding studies are in progress to further define the nature of the sites and probe their reactivity.

4.10Atomically Dispersed Titanium and Vanadium, Single Site Catalysts

The stage has now been set to move into many other systems of interest in the context of catalysis. Initial metal chloride linking reagents studied have all been simple mononuclear chloride reagents such as TiCl4, VOCl3, VCl4, SnCl4 and WCl6,

Table 4.1 Gravimetric analysis of average crosslinking for embedded Al catalysta.

X : cube Cl : Sn Connectivity based on mass change

Residual Sn/cube

Dose 1 AlCl3 1(Al) : 1 3 : 8 3.06 (6) 4.93Dose 2 SiCl4 0.5(Si) : 1 4 : 4.9 2.4 (1) 3.7Dose 3 TMSCl Excess – 1 (theoretical) −0.2

a 1.010 g Si8O20(SnMe3)8; final Al: 2.3 wt%; theoretical molecular formula: Si8O20Al(SiCl2.41)0.5TMS3.73 = Si12.9Al1.0Cl1.2Me11.2; BET surface area: 183 m2 g−1.

c04.indd 156 3/27/2009 4:47:51 PM

barnes
Cross-Out
replace "3" with "three"
Page 21: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.10 Atomically Dispersed Titanium and Vanadium, Single Site Catalysts 157

-1000 ppm-800-600-400-2000

VOCl3

3-connected2-connected1-connected

*

*

** * : spinning side bands

*

**

(a)

(b)

(c)

Figure 4.12 51V SS NMR (MAS) for three atomically dispersed vanadyl containing Si8O20-building block matrices. (a) “Embedded” (3-connected) vanadyl; (b) mainly 2-connected vanadyl; (c) “surface” (1-connected) vanadyl moieties. Variable speed MAS experiments were used to identify the chemical shifts of the isotropic signals within the envelopes of spinning sideband peaks (not shown).

which lead to atomically dispersed metal sites in Si8O20 matrices. Two areas of interest for catalysts of this type are selective oxidation reactions and solid acid catalysis [89, 90]. Given that 51V is a NMR active isotope with quadrupole properties that allow it to be observed relatively easily [83], we initially investigated atomically dispersed vanadyl (V=O) catalysts derived from VOCl3. As chloride is replaced by oxygen in the coordination sphere around the vanadyl group, the chemical shift of the 51V center moves upfield (more negative chemical shift) due to the chloride effect [91]. The expected isotropic shift ranges for vanadyl groups with 0, 1, 2 and 3 chloride ligands are indicated in Figure 4.12. 51V SS NMR spectra corresponding to three samples, one containing embedded (3-connected) vanadyl centers, one primarily 2-connected vanadyl centers and an extremely sensitive material containing surface vanadyl dichloride (≡Si–O–VOCl2) sites are also shown in the figure.

Figure 4.13 shows R-space plots (not corrected for phase shift) for the 1-con-nected and 3-connected catalysts just described. We are fortunate that the V=O, V–O and V–Cl features for these species occur at different distances and are easily resolved in R-space plots of the data. Both NMR and gravimetric data provide well-defined constraints that are used in building structural models from EXAFS data. Coordination numbers, V–O, V–Cl, and V=O separations derived from EXAFS data are within the expected ranges from NMR, gravimetric and XANES information. EXAFS data for the sample containing capping VOCl2 groups

c04.indd 157 3/27/2009 4:48:08 PM

Page 22: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

158 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

compare quite well with a previous description of this group on silica surfaces by Scott and coworkers [92].

Turning now to titanium based catalysts, 47Ti/49Ti solid-state NMR data are quite challenging to obtain and interpret [93, 94]. However, gravimetric data allow us to again derive constraints for structural models from EXAFS data. Figure 4.14 shows the R-space plots obtained from X-ray fluorescence data for a sample containing only embedded 4-connected Ti centers in a Si8O20-bb matrix. Consistent with the gravimetric data, no Ti-Cl feature is seen in the R-space plots. Also shown in the figure are the results of an EXAFS analysis for a sample in which the surface –O-

R (Å)

— data- - - FEFF 8 fit

5

0

10

0 1 2 3 4 5 6

TiO

OO

OTi–O

Ti···Si

0 1 2 3 4 5 6

5

0

10

15

(a) (b)

|χ(R

)| (

Å-4

)

Ti Cl

OO

Cl

Ti–O

Ti–Cl

— data- - - FEFF 8 fit

Ti···Si

RTi-O: 1.79 s2: 0.01 CN: 2 (4 - CN(Cl))RTi-Cl: 2.19 s2: 0.0004 CN: 2 (refined)RTi-Si: 3.49 s2: 0.007 CN: 2 (= CN(O))

R (Å)

RTi-O: 1.80 s2: 0.003 CN: 4.0 (refined)RrTi-Si: 3.49 s2: 0.01 CN: 4

Figure 4.14 Titanium EXAFS for (a) 2-connected (average) and (b) embedded titanium centers in Si8O20-building block matrices.

R (Å)

0

5

10

0 1 2 3 4 5 6

VO

OO

OV=O

V–O

V···Si

c(R)- - - FEFF 8 fit

0 1 2 3 4 5 60

0.5

1

1.5

R (Å)

|c(R

)| (

Å-4

)

c(R)- - - FEFF6L Fit

V

O

OlC lC

V=O

V–O

V–Cl

RV=O: 1.61; s2: 0.001RV-O: 1.79; s2: 0.002 CNV-O: 3.0 ±0.1RV-Si: 3.37; s2: 0.01Nind: 22 Npar: 9

RV=O:1.68; s2: 0.002RV-O: 1.83; s2: -0.003 CNV-O: 1 (set)RV-Cl:2.10; s2: 0.002 CNV-Cl: 2 (set)Nind: 14 Npar: 8

(a) (b)

Figure 4.13 Vanadium EXAFS for (a) surface and (b) embedded vanadyl centers in Si8O20-building block matrices.12

c04.indd 158 3/27/2009 4:48:10 PM

Page 23: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.11 Bridge between Nanostructuring and Catalysis 159

TiCl3 species was targeted. The structural model derived from EXAFS data indi-cates that a connectivity slightly greater than one was obtained for the titanium centers in this sample. Both gravimetric and EXAFS coordination numbers rep-resent averages for the distribution of Ti sites in the matrix and thus a single site catalyst was probably not obtained. Even so, we quickly know that 1-connected sites were not achieved so that the initial dosing stoichiometry can be adjusted to produce a larger, more rigid platform to prevent the titanium centers from assum-ing crosslinking positions in the matrix.

4.11Bridge between Nanostructuring and Catalysis

There is a cost associated with the effort needed to prepare these nanostructured materials: building blocks must be found and suitably functionalized; tailored linking reagents are required; air sensitive handling procedures developed and byproducts carefully removed. It is therefore appropriate to ask, “has the effort and expense lead to new and better catalysts?”

In a “philosophical” sense, one can answer yes to this question immediately. Reducing the complexity of a heterogeneous catalyst, and having an exact knowledge of what the initial catalyst sites are, are long held goals in catalysis science. But this statement must be quickly qualified by noting that we generally know only what the sites are before catalysis. Going through a catalytic cycle will, in many cases, change sites irreversibly. Understanding the mechanism by which stable catalyst precursors are transformed into unstable but active catalyst ensembles is a fundamental challenge in catalysis science. Nonetheless we can state that we begin from a nanostructured support matrix containing a homogeneous distribution of a single type of well-defined precursor to the active catalyst.

More pragmatically, one can attempt to address the question of the benefits of nanostructured catalysts by comparison against the properties of similar catalysts prepared by traditional methods. There are a couple of well-known reactions that are thought to be catalyzed by atomically dispersed metals on supports [95]. Epoxi-dation of olefins by titanium on silica is one [96, 97].

Table 4.2 summarizes recent results for epoxidation studies for the Ti-catalysts described above compared to a catalyst, described in the literature [98], that was prepared by exposing predried silica to titanium(IV) isopropoxide followed by calcination. All of the titanium building block catalysts were pretreated with meth-anol to react any remaining silicon or titanium chloride groups, presumably generating methoxy groups in their places. The intermediate connected titanium catalysts are averaged values based upon Ti-cube stoichiometries used in their syntheses. One immediate observation from the data is the very high epoxide TONs exhibited by 4-connected titanium building block catalyst – nearly five times that reported in the literature. However, the selectivity for the epoxide product is not nearly as high as has been reported previously. When this catalyst was

c04.indd 159 3/27/2009 4:48:10 PM

Page 24: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

160 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

prepared and tested in our laboratories it exhibited a selectivity that was similar to the other catalysts tested.

Also of interest is the observation that lower connectivities for titanium in these building block matrices lead to a significant reduction in epoxidation TONs. This observation leads to a preliminary conclusion that 4-connected titanium centers are the most active epoxidation catalysts that we have produced. Currently, we are preparing Ti-bb samples, with rigorously defined 2- and 3-connectivities to better define the activities, as well as other literature catalysts for comparison.

We have also investigated the properties of several of our nanostructured cata-lysts as solid acids in reactions such as the dehydration of alcohols and transesteri-fication reactions [99]. One of the best examples of atomically dispersed solid acid catalysts is aluminosilicates [100]. When aluminium is substituted into silicate frameworks and remains isolated from other Al centers it can behave as a strong acid site [101].

Although there are some important differences between what we describe as 3-connected aluminium sites in our bb-matrices and what the active sites are thought to be in zeolites, we have begun a preliminary study of the activities of the Al, Ti and V-containing bb-catalysts as solid acid catalysts in the dehydration of alcohols. For this type of bench marking reaction, there are two parameters that can be used as preliminary indicators of catalytic activity: lightoff temperatures and product selectivity. A plot of conversion versus temperature produces what is known as a lightoff curve. The temperature at which 50% of the maximum

Table 4.2 Catalytic epoxidation of cyclohexene with anhydrous TBHP (tert-butylhydroperoxide).

Catalysts for epoxidationa

Ti (wt%) (pretreatment)b

Cyclohexene conversion (%)c

Epoxide selectivityd

Epoxide TON

4-Connected Ti 0.9 (MeOH; 140 °C) 52 44 492

⟨3-Connected Ti⟩ 3.45 (MeOH; 140 °C) 44 35 62

⟨2-Connected Ti⟩ 6.08 (MeOH; 140 °C) 58 47 72

Ti on silica 60e 3.85 (calcined: 500 °C, dried at 140 °C)

68 45 125

Ti on silica 60f 3.85 (140 °C) 96 92 131

a Epoxidation conditions: ∼50 mg solid catalyst, 18 mmol cyclohexene; 16 mmol anhydrous TBHP in toluene at 60 °C for 6 h.

b Building block catalysts were washed with MeOH to remove chloride then dried at given temperature under vacuum.

c Based on t-butyl peroxide as limiting reagent.d Other free radical reaction products observed: 2-cyclohexene-1-ol and 2-cyclohexen-1-one.e Prepared by the method in Reference [98].f Data reported in Reference [98].Data reported in Reference [98]

13

c04.indd 160 3/27/2009 4:48:10 PM

Page 25: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

4.11 Bridge between Nanostructuring and Catalysis 161

conversion occurs is referred to as the light off temperature, a common figure of merit for a catalyst. Figure 4.15 shows the lightoff curve for the three-connected aluminium bb-catalyst described earlier in the dehydration of isopropanol to propene. The lightoff temperature we observe for this material is about 175 °C, which is comparable with other aluminosilicates that we have tested in our labo-ratories. Also of interest, is that all of the atomically dispersed metal catalysts that we have tested thus far (Ti, V, Sn and Al) all act as acidic, dehydration catalysts and are quite selective in producing olefin products.

The results we have gathered thus far are not sufficient to say conclusively what the benefits are from the single site catalysts that we have prepared. They behave more selectively as acidic sites rather than engage in oxidation reactions. They appear to be at least as active as other well-known solid acid catalysts. When we view these results from within the context of atomic dispersion of metals on sili-cate supports, a hypothesis at this stage of our work is that the activities of metal centers that have only M–O–Si groups present in their first coordination spheres is dominated by their acidic properties. The critical functional groups that we believe are absent during catalysis by these sites are hydroxide or oxide ligands that bridge two active metal centers (M–O–M). The absence of the M–O–M group appears to raise the energies associated with having lattice oxygen participate in

0

20

40

60

80

100

0 50 100 150 200 250 300Temperature (∞C)

% C

on

vers

ion

iPrOH

Propene

Other

Figure 4.15 Light-off curve for the dehydration of 2-propanol by a Si8O20-building block catalyst initially containing 3-connected, atomically dispersed aluminium atoms. Conditions: 55 mg catalyst, 95 cc min−1 total flow across catalyst, WHSV: 0.4 h−1.

c04.indd 161 3/27/2009 4:48:10 PM

Page 26: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

162 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

oxidation catalysis. This question points to a straightforward extension of our studies toward double building block syntheses in which polynuclear linking agents containing M–O–M moieties replace the mononuclear ones presently investigation.

4.12Summary

In this chapter I have tried to identify some of the important issues surrounding the notion of what nanostructured catalysts are and approaches to their prepara-tion. The pursuit of nanostructured catalysts goes hand in hand with the goal of preparing single site catalysts. However, nanostructuring catalysts, taken in its broadest context, involves a type of hierarchical control of structure at several dif-ferent length scales: definition of the site at the atomic scale, tailoring access to the site, maintaining separation of sites and tailoring pore distributions are all properties that must be considered. In this context, we can see how tailoring the atomic, meso and macroscopic properties of catalysts all come together in the quest for next-generation supported catalysts.

In our own work, we have articulated a general methodology for preparing nanostructured, single site catalysts in silicate matrices. This methodology enjoys broad applicability to high valent metal and main group halides. The two funda-mental requirements of this approach are a ready source of a rigid, polyfunctional-ized molecular building block and a selective crosslinking chemical reaction that produces bonds only between linkers and building blocks. Once these two require-ments are met the method of sequential additions provides a simple way for a wide range of dispersed, single site catalyst ensembles to be created. Furthermore, both access to the sites and the polarity of the surface of the matrix may be tailored toward application. Using this approach to the synthesis of nanostructured cata-lysts, a broad new range of materials may be rapidly prepared, characterized and their properties as catalysts explored.

Acknowledgments

The generous support of the US Department of Energy (DE-FG02-01ER15259) from the beginning of this project eight years ago is gratefully acknowledged. Support from the Petroleum Research Fund administered by the ACS (PRF 42634-AC5) for our investigations into the development of new solid acids is also acknowledged. I acknowledge the dedicated work of my former (Jason Clark, Ming-Yung Lee, Richard Mayes, Geoff Eldridge) and current (Joshua Abbott, Michael Peretich, Nan Chen) students. I would also like to thank my collaborators, Drs. Edward Hagaman and Jian Jiao (Chemical Sciences Division, Oak Ridge National Laboratory) for obtaining and analyzing SS NMR data and Dr. James Goodwin (Department of Chemical Engineering; Clemson University) and his group for their advice on constructing a gas phase microcatalysis flow reactor.

5

c04.indd 162 3/27/2009 4:48:10 PM

Page 27: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

References 163

References

1 Ying, J.Y. (1997) Nanostructure Processing of Advanced Catalytic materials, ITR Institute, Arlington, VA, pp. 96–9.

2 Dubois, J.-L. (2005) Catal. Today, 99, 5–14.

3 Schlögl, R. (1993) Angew. Chem. Int. Ed. Engl., 32, 381–83.

4 Centi, G., Perathoner, S. and Trifirò, F. (1997) Appl. Catal. A, 157, 143–72.

5 Bell, A.T. (2003) Science, 299, 1688–91.6 Pimentel, G. (1985) Opportunities in

Chemistry, National Academy Press, Washington DC, p. 344.

7 Anderson, M.L., Stroud, R.M., Morris, C.A., Merzbacher, C.I. and Rolison, D.R. (2000) Adv. Eng. Mater., 2, 481–8.

8 Védrine, J.C. (2002) Top. Catal., 21, 97–106.

9 Grasselli, R.K., Oyama, S.T., Gaffney, A.M. and Lyons, J.E. (1997) Third World Congress on Oxidation Catalysis, Vol. 110, Elsevier, New York, p. 1256.

10 Hodnett, B.K. (2000) Heterogeneous Catalytic Oxidation, John Wiley & Sons, Inc., New York, p. 348.

11 Mason, R. and Thomas, J.M. (2003) Angew. Chem. Int. Ed., 42, 18–19.

12 McKnight, A.L. and Waymouth, R.M. (1998) Chem. Rev., 98, 2587–98.

13 Thomas, J.M., Raja, R. and Lewis, D.W. (2005) Angew. Chem. Int. Ed., 44, 6456–82.

14 Derouane, E.C. (1998) A Molecular View of Heterogeneous Catalysis, De Boeck Université, Paris, p. 216.

15 Centi, G., Cavani, F. and Trifiro, F. (2001) Selective Oxidation by Heterogeneous Catalysts, Springer, New York, p. 514.

16 Fierro-Gonzalez, J.C., Kuba, S., Hao, Y. and Gates, B.C. (2006) J. Phys. Chem. B, 110, 13326–51.

17 Henrich, V.E. (1985) Rep. Prog. Phys., 48, 1481–541.

18 Herrero, J., Blanco, C. and Oro, L.A. (1989) Appl. Organomet. Chem., 3, 553–55.

19 Meier, W.M., Olson, D.H. and Baerlocher, C. (••••) Atlas of Zeolite Structure Types, Structure Commission of the International Zeolite Association.6

20 Corma, A. (1997) Chem. Rev., 97, 2373–419.

21 Stein, A., Melde, B.J. and Schroden, R.C. (2000) Adv. Mater., 12, 1403–19.

22 Lee, J., Kim, J. and Hyeon, T. (2006) Adv. Mat., 18, 2073–94.

23 Soleda, S.L., Iglesiaa, E., Fiatoa, R.A., Baumgartnera, J.E., Vromana, H. and Miseo, A.S. (2003) Top. Catal., 26, 101–9.

24 Brown, J., Gordon, E., Henrich, V.E., Casey, W.H., Clark, D.L., Eggleston, C., Felmy, A., Goodman, D.W., Grätzel, M., Maciel, G., McCarthy, M.I., Nealson, K.H., Sverjensky, D.A., Toney, M.F. and Zachara, J.M. (1999) Chem. Rev., 99, 77–174.

25 Lowena, W.K. and Broge, E.C. (1961) J. Phys. Chem., 65, 16–19.

26 Maniar, P. D., Navrotsky, A., Rabinovich, E.M., Ying, J.Y. and Benziger, J.B. (1990) J. Non-Cryst. Solids, 124, 101–11.

27 Thomas, J.M. (1988) Angew. Chem. Int. Ed. Engl., 27, 1673–91.

28 Bronkema, J.L. and Bell, A.T. (2007) J. Phys. Chem. C, 111, 420–30.

29 Lacheen, H.S. and Iglesia, E. (2006) J. Phys. Chem. B, 110, 5462–72.

30 McCool, B. and Tripp, C.P. (2005) J. Phys. Chem. B, 109, 8914–19.

31 Jal, P.K., Patel, S. and Mishra, B.K. (2004) Talanta, 62, 1005–28.

32 Schwarz, J.A., Contescu, C. and Contescu, A. (1995) Chem. Rev., 95, 477–510.

33 Cariati, E., Roberto, D., Ugo, R. and Lucenti, E. (2003) Chem. Rev., 103, 3707–32.

34 Johnson, B.F.G., Raynor, S.A., Brown, D.B., Shephard, D.S., Mashmeyer, T., Thomas, J.M., Hermans, S., Raja, R. and Sankar, G. (2002) J. Mol. Catal. A, 182-183, 89–97.

35 Fornasieri, G., Rozes, L., Calvé, S.L., Alonso, B., Massiot, D., Rager, M.N., Evain, M., Boubekeur, K. and Sanchez, C. (2005) J. Am. Chem. Soc., 127, 4869–78.

36 Dolbecq, A., Mellot-Draznieks, C., Mialane, P., Marrot, J., Férey, G. and Sécheresse, F. (2005) Eur. J. Inorg. Chem., 3009–18.

c04.indd 163 3/27/2009 4:48:11 PM

barnes
Cross-Out
barnes
Replacement Text
change to: Ch. Baerlocher and L. B. McCuster, Database of Zeolite Structures: http://www.iza-structure.org/databases/
Page 28: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

164 4 Building Block Approaches to Nanostructured, Single Site, Heterogeneous Catalysts

37 Mizuno, N. and Misono, M. (1998) Chem. Rev., 98, 199–217.

38 Gates, B.C. (2001) Top. Catal., 14, 173–80.

39 Allard, L.F., Panjabi, G.A., Salvi, S.N. and Gates, B.C. (2002) Nano Lett., 2, 381–84.

40 Homs, N. and d. l. Piscina, P.R. (••••)Carbonyl compounds as metallic presursors of tailored supported catalysts, in Modern Surface Organometallic Chemistry (eds J.M. Bassett, R. Psaro, D. Boberto and R. Ugo), Chapter 8, Elsevier.

41 Ott, L.S. and Finke, R.G. (2007) Coord. Chem. Rev., 251, 1075–100.

42 Haruta, M., Tsubota, S., Kobayashi, T., Kegeyama, K., Genet, M.J. and Delmon, B. (1993) J. Cat., 144, 175–92.

43 Klabunde, K.J. and Mohs, C. (1998) Nanoparticles and nanostructural materials, in Chemistry of Advanced Materials (eds L.V. Interrante and M.J. Hampden-Smith), John Wiley & Sons, Inc., New York, pp. 271–327.

44 Gilje, J.W. and Roesky, H.W. (1994) Chem. Rev., 94, 895–910.

45 Rozes, L., Steunou, N., Fornasieri, G. and Sanchez, C. (2006) Monatsh. Chem., 137, 501–28.

46 Corbett, J.D. (2000) Inorg. Chem., 39, 5178–91.

47 Bordiga, S., Damin, A., Bonino, F. and Lamberti, C. (2005) Top. Organomet. Chem., 16, 37–68.

48 Bordiga, S., Bonino, F., Damin, A. and Lamberti, C. (2007) Phys. Chem. Chem. Phys., 9, 4854–78.

49 Bare, S.R., Kelly, S.D., Sinkler, W., Low, J.J., Modica, F.S., Valencia, S., Corma, A. and Nemeth, L.T. (2005) J. Am. Chem. Soc., 127, 12924–32.

50 Farneth, W.E. and Gorte, R.J. (1995) Chem. Rev., 95, 615–35.

51 Hüsing, N. and Schubert, U. (1998) Angew. Chem. Int. Ed., 37, 22-45.

52 Kresge, C.T., Leonowicz, M.E., Roth, W.J., Vartuli, J.C. and Beck, J.S. (1992) Nature, 359, 710–12.

53 Dai, S. and Barnes, C.E. (2003) Mesoporous materials, in Encyclopedia of Supramolecular Chemistry (eds J. Atwood and J. Steed), Elsevier, New York.

7

8

54 Raymond, H.G. and Garth, L.W. (1989) Polym. Bull., 22, 527–32.

55 Asanuma, H., Hishiya, T. and Komiyama, M. (2000) Adv. Mater., 12, 1019–30.

56 Marty, J.D. and Mauzac, M. (2005) Molecular imprinting: state of the art and perspectives, in Microlithography – Molecular Imprinting, Vol. 172, Springer, Berlin, pp. 1–35.

57 Dai, S., Burleigh, M.C., Ju, Y.H., Gao, H.J., Lin, J.S., Pennycook, S.J., Barnes, C.E. and Xue, Z.L. (2000) J. Am. Chem. Soc., 122, 992–93.

58 Sharygin, L.M. and Muromskii, A.Y. (2000) At. Energy, 89, 658–62.

59 Liu, J., Feng, X., Fryxell, G.E., Wang, L.-Q., Kim, A.Y. and Gong, A.M. (1998) Adv. Mater., 10, 151–65.

60 Copéret, C., Chabanas, M., Saint-Arroman, R.P. and Basset, J.-M. (2003) Angew. Chem. Int. Ed., 42, pp. 157-81.

61 Basset, J.-M., Baudouin, A., Bayard, F., Candy, J.-P., Copéret, C., Mallmann, A.D., Godard, G., Kuntz, E., Lefebvre, F., Lucas, C., Norsic, S., Pelzer, K., Quadrelli, A., Thieuleux, C., Thivolle-Cazat, J. and Veyre, L. (••••) Preparation of single site catalysts prepared via surface organometallic chemistry, in Modern Surface Organometallic Chemistry (eds J.M. Bassett, R. Psaro, D. Boberto and R. Ugo), Chapter 2, Elsevier.

62 Miller, C.J. and O’Hare, D. (2004) Chem. Commun., 1710–11.

63 (••••) Modern Surface Organometallic Chemistry (eds J.M. Bassett, R. Psaro, D. Boberto and R. Ugo), Chapters 1–3, Elsevier.

64 Ockwig, N.W., Delgado-Friedrichs, O., O’Keefe, M. and Yaghi, O.M. (2005) Acc. Chem. Res., 38, 176–82.

65 Yaghi, O.M., Li, H., Davis, C., Richardson, D. and Groy, T.L. (1998) Acc. Chem. Res., 31, 474–84.

66 El-Kaderi, H.M., Hunt, J.R., Mendoza-Cortés, J.L., Côté, A.P., Taylor, R.E., O’Keeffe, M. and Yaghi, O.M. (2007) Science, 316, 268–72.

67 Banerjee, R., Phan, A., Wang, B., Knobler, C., Furukawa, H., O’Keeffe, M. and Yaghi, O.M. (2008) Science, 319, 939–43.

9

10

c04.indd 164 3/27/2009 4:48:11 PM

Page 29: 4 Building Block Approaches to Nanostructured, Single Site ... · 4.5 Current Approaches to Nanostructured Catalysts . 141. 4.4 Benefits of Nanostructuring Catalysts. The goals and

O

References 165

68 Ockwig, N.W., Delgado-Friedrichs, O., O’Keeffe, M. and Yaghi, O.M. (2005) Acc. Chem. Res., 2005, 176–82.

69 Hwang, Y.K., Hong, D.-Y., Chang, J.-S., Jhung, S.H., Seo, Y.-K., Kim, J., Vimont, A., Daturi, M., Serre, C. and Férey, G.R. (2008) Angew. Chem. Int. Ed., 47, 4144–8.

70 Xamena, F.X.L.I., Casanova, O., Tailleur, R.G., Garcia, H. and Corma, A. (2008) J. Catal., 255, 220–27.

71 Hill, C.L. (1998) Chem. Rev., 98, 1–2.72 Harrison, P.G. (1997) J. Organomet.

Chem., 542, 141–83.73 Agaskar, P. A. (1991) Inorg. Chem., 30,

2707–08.74 Morris, R.E. (2005) J. Mater. Chem., 15,

931–38.75 Laine, R.M. (2005) J. Mater. Chem., 15,

3725–44.76 Lamm, M.H., Chen, T. and Glotzer, S.C.

(2003) Nano Lett., 3, 989–94.77 Klemperer, W.G., Mainz, V.V. and

Millar, D.M. (1986) Mater. Res. Soc. Symp. Proc., 73, 3–13.

78 Feher, F.J. and Weller, K.J. (1994) Chem. Mater., 6, 7–9.

79 Vioux, P. A. (1997) Chem. Mater., 9, 2292–99.

80 Davis, M.E., Katz, A. and Ahmad, W. (1996) Chem. Mater., 8, 1820–39.

81 Sanders, J.K.M. (2000) Pure Appl. Chem., 72, 2265–74.

82 Clark, J.C., Saengkerdsub, S., Eldridge, G.T., Campana, C. and Barnes, C.E. (2006) J. Organomet. Chem., 691, 3213–22.

83 Ghosh, N.N., Clark, J.C., Eldridge, G.T. and Barnes, C.E. (2004) Chem. Commun., 856–7.

84 Clark, J.C. and Barnes, C.E. (2007) Chem. Mater., 19, 3212–18.

85 Finnie, K.S., Luca, V., Moran, P.D., Bartletta, J.R. and Woolfrey, J.L. (2000) J. Mater. Chem., 10, 409–18.

86 Rolison, D.R. (2003) Science, 299, 1698–701.

87 Lambert, J.B. and Riddell, F.G. (1983) Transition Metal NMR Spectroscopy, NATO ASI Series, Vol. 103, D. Reidel: Boston, MA, pp. 445–56.

88 Conradson, S.D. (1998) Appl. Spectrosc., 52, 252A–79A.

89 Weckhuysen, B.M. and Keller, D.E. (2003) Catal. Today, 78, 25–46.

90 Wachs, I.E. (1996) Catal. Today, 27, 437–55.

91 Kidd, R.G. (1983) The multinuclear approach to NMR spectroscopy, in Transition Metal NMR Spectroscopy, NATO ASI Series, Vol. 103 (eds J.B. Lambert and F.G. Riddell), D. Reidel, Boston, MA, pp. 445–56.

92 Deguns, E.W., Taha, Z., Meitzner, G.D. and Scott, S.L. (2005) J. Phys. Chem. B, 109, 5005–11.

93 Pickup, D.M., Sowrey, F.E., Newport, R.J., Gunawidjaja, P.N., Drake, K.O. and Smith, M.E. (2004) J. Phys. Chem. B, 108, 10872–80.

94 Padro, D., Jennings, V., Smith, M.E., Hoppe, R., Thomas, P.A. and Dupree, R. (2002) J. Phys. Chem. B, 106, 13176–85.

95 Thomas, J.M. and Sankar, G. (2001) Acc. Chem. Res., 34, 571–81.

96 Dusi, M., Mallat, T. and Baiker, A. (2000) Catal. Rev. Sci. Eng., 42, 213–78.

97 Vos, D.E.D., Sels, B.F. and Jacobs, P.A. (2003) Adv. Synth. Catal., 345, 457–73.

98 Cativela, C., Fraile, J.M., Garcia, J.I. and Mayoral, J.A. (1996) J. Mol. Catal. A, 112, 259–67.

99 Corma, A. (1995) Chem. Rev., 95, 559–614.

100 Haw, J.F. (2002) Phys. Chem. Chem. Phys., 22, 5431–41.

101 Farneth, W.E. and Gorte, R.J. (1995) Chem. Rev., 95, 615–35.

11

c04.indd 165 3/27/2009 4:48:11 PM


Recommended