Metal Nanoparticles For Catalysis Advances and Applications
Metal Nanoparticles For Catalysis Advances and Applications
Metal Nanoparticles For Catalysis Advances and Applications
View Online
Advisory Board:
Krijn P de Jong, University of Utrecht, The Netherlands
James A Dumesic, University of Wisconsin-Madison, USA
Chris Hardacre, Queens University Belfast, Northern Ireland
Enrique Iglesia, University of California at Berkeley, USA
Zinfer Ismagilov, Boreskov Institute of Catalysis, Novosibirsk, Russia
nchen, Germany
Johannes Lercher, TU Mu
Umit Ozkan, Ohio State University, USA
Chunshan Song, Penn State University, USA
View Online
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-FP001
Franklin Tao
University of Notre Dame, USA
Email: [email protected]
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-FP001
View Online
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-FP005
Biography
Franklin (Feng) Tao, PhD, is a tenure-track
assistant professor of chemistry. After receiving a PhD from Princeton University
and being a postdoctoral fellow at UCBerkeley and Lawrence Berkeley National
Lab, he started his independent career in
2010. He was elected a Fellow of the Royal
Society of Chemistry in 2013. Currently, he
is leading a research group focusing on
synthesis, evaluation of catalytic performance, and in situ and operando characterization of catalytic materials in catalytic
reactions for chemical and energy transformations toward fundamental understanding of catalytic processes at the
molecular level. He has published about 100 peer-reviewed publications and
three books with Wiley and the Royal Society of Chemistry.
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-FP005
View Online
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-FP007
Acknowledgements
I would like to thank all the contributors who made the publication of this
book possible, the Royal Society of Chemistry books team for their eort in
publishing this book, and support from the Chemical Sciences, Geosciences
and Biosciences Division, Oce of Basic Energy Sciences, Oce of Science,
U. S. Department of Energy under the grant DE-FG02-12ER16353. An acknowledgement must also go to Dr Radha Narayanan for her time at the
early stages of the development of this book.
vii
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-FP007
View Online
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-FP009
Contents
Chapter 1 Introduction: Synthesis and Catalysis on Metal
Nanoparticles
Franklin (Feng) Tao, Luan Nguyen and Shiran Zhang
Chapter 2 Nanocatalysis: Definition and Case Studies
Choumini Balasanthiran and James D. Hoefelmeyer
2.1
2.2
2.3
2.4
Introduction
2.1.1 Flash Synopsis of the History of Catalysis
2.1.2 Reporting Turnover Frequency: the Common
Denominator in Catalysis
Factors Contributing to Structure Sensitivity in
Catalysis
2.2.1 Statistical Shape Analysis of Polyhedra
Crystals and Relation to Catalysis
2.2.2 Equilibrium Shapes of Nanocrystals
2.2.3 Surface Restructuring
2.2.4 Mobility of Surface Adsorbates
2.2.5 Change in the Electronic Structure of Solids
at the Nanometre Scale
2.2.6 Example to Illustrate Anomalous TOF
Behavior: Ethane Hydrogenolysis on
Rhodium
Synthesis and Properties of Well-defined
Nanocrystals
The Dawn of Nanocatalysis and Case Studies
2.4.1 CO Oxidation on Au
2.4.2 TiO2 Nanocrystals with Reactive Facets
2.4.3 Catalysis on Shape-controlled Pt Nanocrystals
ix
6
7
8
9
9
11
12
14
14
15
16
18
18
19
21
View Online
Contents
2.4.4
Introduction
New Route for the Preparation of Supported Metal
Nanoparticle Catalysts
3.2.1 A Photo-assisted Deposition Method Using
a Single-site Photocatalyst
3.2.2 A Microwave-assisted Deposition Method
3.2.3 Deposition of Size-controlled Metal
Nanoparticles as Colloidal Precursors
3.3 Multifunctional Catalysts Based on Magnetic
Nanoparticles
3.3.1 Coreshell Magnetic FePt@Ti-containing
Silica Spherical Nanocatalyst
3.3.2 Water-soluble FePt Magnetic NPs Modified
with Cyclodextrin
3.3.3 FePd Magnetic NPs Modified with a Chiral
BINAP Ligand
3.3.4 Coreshell Nanostructured Catalyst for
One-pot Reactions
References
Chapter 4 Organometallic Approach for the Synthesis of Noble
Metal Nanoparticles: Towards Application in Colloidal
and Supported Nanocatalysis
Solen Kinayyigit and Karine Philippot
4.1
4.2
4.3
Introduction
Organometallic Synthesis of Noble Metal
Nanoparticles
Nanoparticles for Colloidal Catalysis
4.3.1 Hydrogenation Reactions
4.3.1.1 Ligand Stabilized Nanoparticles as
Catalysts
4.3.1.2 Water-soluble Nanoparticles as
Catalysts
23
23
24
25
30
30
31
31
34
35
38
38
39
41
42
44
47
47
48
49
50
50
52
View Online
Contents
xi
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-FP009
4.3.1.3
Introduction
Nanoparticles as Relevant Catalysts for Biphasic
Hydrogenation
57
60
63
63
65
67
69
70
71
71
72
75
75
76
78
78
79
83
83
84
86
88
94
96
96
99
99
100
View Online
xii
Contents
6.3
Introduction
Synthetic Scheme for the Fabrication of Pd
Nanoparticles
7.3 CC Coupling Reaction Mechanism for Pd
Nanocatalysts
7.4 Pd Nanoparticles in the Stille Coupling Reaction
7.5 Pd Nanoparticles in the Suzuki Coupling Reaction
7.6 Pd Nanoparticles in the Heck Coupling Reaction
7.7 Summary and Conclusions
References
Chapter 8 Metal Salt-based Gold Nanocatalysts
Zhen Ma and Franklin (Feng) Tao
8.1
8.2
Introduction
Metal Salt-based Gold Nanocatalysts
8.2.1 Metal Carbonate-based Gold Catalysts
8.2.2 Metal Phosphate-based Gold Catalysts
8.2.3 Hydroxyapatite-based Gold Catalysts
8.2.4 Hydroxylated Fluoride-based Gold Catalysts
8.2.5 Metal Sulfate-based Gold Catalysts
8.2.6 Heteropolyacid Salt-based Gold Catalysts
8.3 Summary
Acknowledgments
References
Chapter 9 Catalysis with Colloidal Metallic Hollow Nanostructures:
Cage Eect
Mahmoud A. Mahmoud
9.1
9.2
9.3
9.4
Introduction
Synthetic Approaches to Hollow Metallic Nanocatalysts
Assembling the Nanocatalysts on Substrates
Hollow Nanostructures are Dierent in Catalysis
9.4.1 Hollow Nanostructures with a Catalytically
Active Inner Surface and an Inactive Outer
Surface
105
106
107
108
109
112
112
115
116
126
134
144
148
149
157
157
158
158
160
164
166
167
168
169
169
169
172
172
174
176
179
180
View Online
Contents
xiii
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-FP009
9.4.2
Introduction
Steric and Structural Eects
10.2.1 Dendrimers
10.2.2 Microgels
10.2.3 Polymer Coreshell Structures
10.2.4 Hydrophobichydrophilic Structures:
Micelle, Emulsion, and Liposome
10.3 Absorbing Nanocatalyst Surface
10.3.1 Micelle and Emulsion
10.3.2 Carbon Nanotubes
10.4 Conclusion
References
11.2
11.3
11.4
History
11.1.1 Iodine Clock Reaction
11.1.2 BZ Reaction
11.1.3 BrayLiebhafsky Reaction
11.1.4 BriggsRauscher Reaction
11.1.5 The Blue Bottle Experiment
Recent Work
11.2.1 Clock Reaction of Methylene Blue
Mechanistic Approach
11.3.1 EleyRideal Mechanism
11.3.2 LangmuirHinshelwood Mechanism
Applications
11.4.1 Water Purification
11.4.2 Memory Facilitation by Methylene Blue
and Brain Oxygen Consumption
182
183
184
185
187
192
192
193
193
194
195
195
198
198
198
199
199
203
204
204
205
205
206
207
207
210
215
215
216
216
216
217
View Online
xiv
Contents
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-FP009
Introduction
Computational Method
Metal Nanocatalysts
12.3.1 Copper Nanocatalysts for WaterGas
Shift Reactions: the Importance of
Low-coordinated Sites
12.3.2 Metal (Core)Platinum Shell Nanocatalysts
for Oxygen Reduction Reactions in Fuel
Cells: the Essential Role of Surface
Contraction
12.4 Supported Metal Nanocatalysts
12.5 Conclusions
Acknowledgements
References
Introduction
Synthesis and Morphology Control
Catalysis
13.3.1 Selective Oxidation and Fine Chemical
Synthesis
13.3.2 CH Activation
13.3.3 CO2 Activation
13.3.4 Environmental and Green Chemistry
13.4 Conclusion
Acknowledgments
References
Subject Index
217
217
217
219
219
220
220
221
223
227
230
231
231
235
235
236
239
239
241
243
244
246
247
247
251
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00001
CHAPTER 1
View Online
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00001
Chapter 1
View Online
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00001
under-coordinated catalyst atoms at corners and edges from the site on the
surface since there is a size-dependent density of these under-coordinated
atoms at corners and/or edges.
Chapter 2 describes structural factors of catalyst nanoparticles, which
could influence catalytic performances including catalytic activity and selectivity. In addition, it reviews correlations between these structural factors
and catalytic performance, and discusses the potential restructuring of the
surfaces of catalyst particles. A few examples are discussed from a structural
point of view.
Chapter 3 reviews a few synthetic routes of metal nanoparticle catalysts. It
presents three methods for the preparation of metal nanoparticles supported on a substrate including photo-assisted deposition methods using
single-site photocatalysts, a microwave-assisted deposition method, and
deposition of size-controlled metal nanoparticles as colloidal precursors. In
addition, the syntheses of multi-functional catalysts are discussed.
In Chapter 4, syntheses of noble metal nanoparticles are reviewed from an
organometallic point of view. It emphasizes the nature of the ligand stabilizing the nanocatalysts in solution and the role of the support for a supported catalyst. Syntheses of Pt, Rh, Ru, Ir and other metals are exemplified.
Factors influencing the synthesis of metal nanoparticles and their performance in catalysis are discussed in detail.
After the review of syntheses in Chapters 24, Chapter 5 discusses the
catalytic transfer hydrogenation of organic compounds. It focuses on the
utilization of nanoparticle catalysts of the earth-abundant metal, Ni, a replacement for noble metal catalysts in hydrogen-transfer reductions of
functional groups.
Chapter 6 reports the recent progress achieved in nanocatalysis, particularly use of quaternary ammonium salts as water-soluble capping agents of
rhodium nanoparticles. Hydrogenation under biphasic liquidliquid conditions and asymmetric catalysis including ethylpyruvate hydrogenation and
prochiral arene hydrogenation are described.
As well as hydrogenations on Ni and Rh nanoparticles in Chapters 5 and 6,
the catalysis of CC coupling on Pd nanoparticles is reviewed in Chapter 7.
The importance of Pd-catalyzed CC coupling pioneered by Heck, Suzuki
and Negishi was recognized by the 2010 Nobel prize in chemistry. Thanks to
the advance in the synthesis of metallic Pd nanocatalysts with specific size
and shape, ecient catalysis on Pd nanoparticles in contrast to traditional
small molecules was revealed. Due to the limited space, only CC coupling
on selected examples are discussed in Chapter 7, though a large number of
applications of Pd nanoparticles in CC coupling have been reported in the
literature.
Gold nanoparticles supported on oxide substrates exhibit exciting catalytic
activities for many inorganic and organic reactions in contrast to the inert
nature of macroscopic gold particles. Tremendous eort has been put into
exploration of the origin of catalytic activity of Au nanoparticles following
the discovery of activity for many reactions on them. Numerous papers on
View Online
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00001
Chapter 1
this topic have been published in the literature. In these reports, most of the
Au catalyst particles were supported on an oxide. Other than metal oxides,
salts have been used as supports for Au nanoparticles. Some salts are solid
acids, which are used in acid catalysis. In fact, integration of the acidic
support of a salt with gold nanoparticles forms a bi-functional catalyst that is
highly active in organic reactions. Chapter 8 reviews the preparation and
catalysis of salt-based Au catalysts. These salts in the Au/salt catalysts are
carbonate, phosphate, hydroxyapatite, hydroxylated fluoride, metal sulfate,
and heteropolyacid. A comparison between the roles of salts and oxides in
gold catalysis is made.
All the metal nanoparticle catalysts discussed so far are metal nanoparticles with a solid core. From a structural point of view, a nanoparticle
with a hollow or porous core could provide dierent catalytic activity.
A porous metal nanoparticle consist of an external surface and an internal
surface. Thus, the total surface area is significantly increased due to the
existence of the internal surface in the hollow or porous core of the particle.
More importantly, the density of catalyst atoms with a low coordination
number in such a metal particle is in fact much higher than that of a metal
nanoparticle with a solid core. In addition, its inner surface could be covered
with much less capping molecules due to the limited space in a pore. Thus,
the density of active sites could be significantly larger. In addition, the
limited space at a scale of nanometres or a sub-nanometre, in fact increases
the rate of collision of reactant molecules with catalyst atoms. These structural features could oer porous metal nanoparticles dierent catalytic activity and selectivity in contrast to a metal nanoparticle with a solid core.
Chapter 9 presents the synthesis of porous metal nanocatalysts and discusses the catalysis of the nanoparticles.
As discussed in Chapter 9, the local structure of a catalyst nanoparticle is
critical for its catalytic activity. Other than the localized structure of a metal
nanoparticle, such as the surface of a catalyst particle with a solid core and
the surface and inner surface of a porous nanoparticle, the external environment of a particle could influence catalytic shape to some extent.
A metal particle together with its surrounding layers or shell can be considered as a single reactor at the nanoscale. The surroundings could be a
scaold of a dendrimer around the loaded metal nanoparticles, a crosslinked linear polymer of microgels, a polymer with a hydrophobic or
hydrophilic substrate around a metal particle, or a shell of a metal particle.
Chapter 10 describes several types of nanoreactors and discusses the eect
of the surroundings on the catalysis of metal nanoparticles.
Other than these catalytic reactions, metal nanoparticles can also catalyze
clock reactions. Chapter 11 reviews the clock reactions on metal and oxide
particles, and discusses the mechanisms.
Computational chemistry is significant for a mechanistic understanding
of chemical reactions. The application of a computational approach to
catalysis studies is certainly the most successful synergy in chemical sciences. A tremendous eort has been made in this field in recent decades and
View Online
.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00001
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
CHAPTER 2
2.1 Introduction
Catalysis is the process of introducing a catalyst to a reaction to increase the
rate of the reaction. The catalyst provides an alternate reaction pathway with
a lower activation energy. It facilitates the reaction without being consumed
and contributes to multiple turnovers of the catalytic reaction cycle at each
catalytic site. Catalysis has enormous technological significance, being
important in energy production, the chemical industry, and environmental
technologies, and is a foundation of our modern way of life as well as of life
itself.1
Early examples of catalysis in the chemical industry include the catalytic
oxidation of sulfur dioxide to sulfur trioxide, which allowed the large-scale
commercial production of sulfuric acid; ammonia synthesis from nitrogen
and hydrogen (Haber process); and ammonia oxidation (Ostwald process).2
Catalysis has been utilized extensively in the petroleum industry, for example in hydrodesulfurization and reforming. These uses arose from an
increasing demand for high-octane gasoline. The long history of the growing
demand for chemicals and fuel resulted in a parallel demand for better
catalysts, and has triggered much research into understanding catalytic
processes.
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
2.1.1
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
Chapter 2
2.1.2
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
2.2.1
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
10
Figure 2.1
Chapter 2
cubic metals composed of five types of surface atoms: corners (C66), atoms
within {100} faces (C84,5), atoms within {111} faces (C93), edges between two
{111} faces (C79), and edges between {100} and {111} faces (C75). The ratio of
the surface atom types to total surface atoms calculated for crystals with
n shells shows the rapidly diminishing importance of corner and edge atoms
and dominance of facial atoms with increasing size (Figure 2.1).
This is an idealized model; whereas, nanoparticle samples have some
distribution of size and shape that complicates the statistical shape analysis.
In fact, the ability to precisely control the dimensions of nanocrystals in
order to achieve minimum size and shape distribution can allow statistical
shape analysis with smaller standard deviations. This seems to be an opportunity that defines the field of nanocatalysis. Ideally, the nanocatalyst
samples would be characterized and reported with quantitative details including statistical shape analysis. Perhaps this is an opportunity for mathematics to merge with chemistry. Some reports that include mechanistic
details for the formation of special non-equilibrium nanocrystal shapes incorporate detailed qualitative shape analysis.13 In special cases where
homogeneous samples can be created, quantitative relations between surface topology and turnover frequency can be determined. This tends to work
well for larger high-quality nanocrystals with small relative size distribution
and uniform shape. For particles that have poorly defined shapes (size o4 nm),
it becomes extremely challenging to develop quantitative relations, which
interestingly is often the regime in which materials exhibit maximum
turnover frequency!
Crystal domain size clearly contributes to structure sensitivity in catalysis.1421 As crystal domains become smaller, the ratio of corner and edge
atoms to facial atoms increases, and the surface topology becomes increasingly disordered, or roughened. The open coordination spheres of
corner and edge atoms should bind substrates actively. A reaction mechanism operating on corner and edge atoms is likely to be very dierent to
one on an extended crystal face.
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
2.2.2
11
It is already apparent from the discussion in section 2.1 that the shape of a
nanocrystal in relation to its crystal structure gives rise to specific crystal
facets. The example of a cubo-octahedron shape composed of atoms in a
face-centered cubic arrangement gives rise to {100} and {111} low-index
facets. In fact, the cubo-octahedron is one of many possible equilibrium
shapes of nanocrystals. Each facet has a measurable surface tension or
surface free energy such that the most stable facet has the lowest surface free
energy. Low-index facets tend to have lower surface free energy and are often
atomically smooth; whereas high-index facets are atomically rough and have
high surface free energy. At high temperatures, the dierences in the surface
free energy of facets are lower than thermal energy, and the nanocrystal
adopts the perfectly isotropic shape of a sphere. At lower temperatures, the
dierences are large enough to influence the equilibrium shape of the
nanocrystal. The ratio of the surface free energy of two facets is proportional
to the relative distance between the nanocrystal center and the face-centroids, which allows the theoretical construction of the equilibrium crystal
shapeknown as a Wul construction.22,23 If a nanocrystal shape deviates
from the Wul construct, then some other influence (for example, surface
stabilizing agent or catalyst) leads to the non-equilibrium shape. This idea is
fundamentally important in the design and synthesis of nanocrystals.
So, an origin of structure sensitivity is the arrangement of atoms on specific facets. For the face-centered cubic cubo-octahedron, the atoms on the
{111} face are found within a hexagonal 2D array of atoms wherein each
atom has six in-plane nearest neighbors and six three-fold hollow sites,
which of course diers significantly from the atoms on the {100} face that
are found within a square 2D array wherein each atom has eight in-plane
nearest neighbors and four four-fold hollow sites. Geometric dierences
between surface facets can have enormous implications for their chemical
properties. Examples include the extremely structure-sensitive nature of CO
dissociation on rhodium,24 or ammonia synthesis catalyzed on iron single
crystal surfaces that dier in turnover frequency depending on the orientation of the surface plane.25,26 Comparison of five iron single crystal surfaces showed the Fe(111) surface was B500 times more active than the
Fe(110) surface. The Fe(111) surface is rough compared with the Fe(110)
surface; however, it also has a high ratio of C7 sites. The Fe(211) surface also
shares these features and exhibits a high turnover frequency for ammonia
synthesis.
One of the principal challenges defining nanocatalysis is the ability to
engineer nanocrystals with uniform size and shape that possess characteristics optimized for specific catalytic reactions. Some recent advances in the
literature include the synthesis of well-defined nanocrystals with high-index
facets that have reactive surfaces.27,28 High-index facets feature surface
atoms at step-edges with lower coordination numbers that contribute to
higher reactivity.
View Online
12
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
2.2.3
Chapter 2
Surface Restructuring
The shapes of nanocrystals are not static. At high temperatures, a nanocrystal will morph into the thermodynamic equilibrium shape. The temperature at which a Wul construct shaped nanocrystal becomes spherical is
often stated as the nanocrystal melting point, and can be determined using
transmission electron microscopy with high temperature stage capability.
The melting point of nanomaterials generally decreases dramatically in
comparison with bulk materials. This is because surface atoms, which make
up a higher atom fraction of nanomaterials than bulk materials, are mobile
and can desorb from the surface. At high temperatures, it is possible for interparticle atom migration to occur as thermal energy mobilizes atoms. On
surfaces, this phenomenon leads to sintering of the nanoparticles characterized by an increase in average particle size. This eect reduces surface area,
leads to loss of control in surface site types, and is detrimental to catalysis.
Surface restructuring and reshaping of nanocatalysts is a common phenomenon in heterogeneous catalysis to optimize bonding between adsorbed
atoms and surface atoms. Dierent atomic termination and crystal planes
revert to thermodynamically more stable structures at higher temperatures
and pressures during catalysis. Surface restructuring has been studied extensively and can occur upon adsorption of molecules on the surface or
formation of thin films. Surface science techniques including LEED, STM,29
and XPS are powerful tools used to characterize such processes, and these
studies have contributed enormously to the understanding of catalysis.3032
Recent studies in the research of surface restructuring via oxide layer
formation in bimetallic, metallic, and metal oxide nanocrystals have been
described. In situ crystal plane controlled surface restructuring during
CO oxidation on Cu2O NPs has been investigated using XPS, TEM, HRTEM,
SEM and H2-TPR.33 The fcc Cu2O nanocrystals with octahedral or cubic
shapes with exposed {111} and {100} facets, respectively, do not show any
oxidation under ambient conditions. During CO oxidation, the Cu2O octahedra and Cu2O cubes show activation energies and conversion rates of
73.4 2.6 kJ mol1, 91.5% and 110 6.4 kJ mol1, 45.1%, respectively,
under similar reaction conditions. During catalysis, XPS data show Cu(II)
in both nanocrystals, indicating complete oxidation of the Cu(I) in the assynthesized nanocrystals. Electron microscopy results show that the nanocrystals retain their original octahedral or cubic shape; however, with
rounder corners and rougher edges on the surface. HRTEM shows that the
inner region maintains the same lattice fringe structures as Cu2O. Together,
the data indicate an in situ formation of CuO thin films during catalytic
oxidation of CO. Theoretical studies show that the faster rate of catalytic
oxidation of Cu2O octahedra over Cu2O cubes is associated with the existence of coordinatively unsaturated Cu atoms in {111} planes that are more
prone to surface oxidation.
High pressure STM experiments on Pt(110) single crystals show the flexible nature of the surface.32 Exposure to 1.6 atm H2 led to a missing row
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
13
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
14
Chapter 2
2.2.4
2.2.5
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
15
within a glassy matrix. The optical spectra were shown to depend on crystallite size approaching the Bohr exciton radius of the semiconductor.
A major advance came with the development of synthesis methods to
produce measurable quantities of uniform, processable nanocrystals.5961
The quantum dots were composed of CdSe and were synthesized via a nonhydrolytic molecular precursor route in which surfactants mediated nanocrystal growth and stabilized the nanocrystal surface. From this seminal
discovery, in addition to the discovery of fullerenes and the ongoing study of
colloids, was born the systematic investigation of uniform, high-quality
nanocrystals from synthesis, characterization, and applications. These developments touched o a firestorm of research activity to develop easy, lowcost, robust methods for uniform, high-quality nanocrystals and other types
of nanomaterials that have given rise to the fields of nanoscience or
nanotechnology. While spherical quantum dots show size-dependent
electronic structures, more recent studies have also shown shape-dependent
electronic structures.62
2.2.6
We end section 2.2 with an example from the catalysis literature that illustrates the size-dependent properties of matter. Yates and Sinfelt reported the
results of ethane hydrogenolysis on Rh/SiO2 catalysts.63 The catalysts were
synthesized by impregnation of RhCl3 onto Cabosil, followed by H2 reduction. Rhodium particle sizes of 11, 12, 20, 23, 41, and 127 were obtained. Hydrogen chemisorption was used to determine the Rh particle
sizes. Unsupported Rh powder was found to have an average particle size of
2560 . The specific activity of catalytic ethane hydrogenolysis varied dramatically over this size range. Rh particle sizes of 11 and 12 exhibited a
specific activity of 4 mmol C2H6 converted/hr/m2Rh. The activity was about
four-fold higher for Rh particle sizes of 20, 23, and 41 , and decreased 20 to
40-fold for large particle sizes. A peak in TOF was observed for Rh crystallite
sizes close to 2 nm.
Schmidt and co-workers64,65 prepared Rh/SiO2 catalysts by wetness
impregnation of RhCl3 on Cabosil. The Rh loading was varied as 1, 5, and
15 wt% on SiO2, and led to Rh particle sizes of 12, 32, and 44 nm, respectively. The pretreatment procedure had a significant influence on the catalyst
structure and activity. Briefly, when Rh/SiO2 samples were calcined in O2, the
particles were oxidized to Rh2O3. The samples were then reduced at dierent
temperatures under hydrogen. Reduction at 300 1C in H2 caused the oxide
particles to fracture into smaller particles of 520 , separated by channels or
cracks 510 wide. Increased reduction temperature caused the small particles to become annealed, and resulted in a reformation of the original Rh
particle size at 600 1C. The particle structure was characterized with TEM.
Intermediate states between fractured particles and fully annealed particles
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
16
Chapter 2
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
17
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
18
Chapter 2
2.4.1
CO Oxidation on Au
Gold clusters can be grown on the surface of rutile TiO2 (110) single crystal
surfaces under ultra-high vacuum conditions. A tungsten filament wrapped
with high purity Au wire provides Au atoms with well-calibrated flux. The Au
atoms are deposited on the TiO2 surface at 300 K followed by annealing at
850 K for 2 minutes. With this method, Au clusters (16 nm diameter) with
relatively narrow size distributions can be formed in which size depends on
the initial Au coverage.96
The clusters were characterized using scanning tunneling microscopy and
scanning tunneling spectroscopy to establish their size and electronic
structure. A quantum eect was clearly observed for the set of Au particles.
Particles with diameter x height greater than 3.5 nm1.0 nm showed
metallic properties. Below this size, the Au clusters exhibit an increasing
band-gap. Clusters with band-gaps in the 0.20.6 eV range were found in the
2.53.0 nm range.
View Online
19
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
2.4.2
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
20
Figure 2.2
Chapter 2
terminated crystal has higher activity due to the slightly larger band-gap;
though, the dierences are only on the order of a few tens of millivolts.
Exposed facets of nanocrystals influence the trapping and interfacial
transfer of photogenerated carriers. UV irradiation of rutile nanocrystals
with {011} and {110} facets in aqueous hexachloroplatinate led to deposition
of platinum on the {110} facets while water oxidation occurred on the {011}
facets. UV irradiation of the Pt-TiO2 powder in aqueous Pb21 led to the
growth of PbO2 on the {011} facets (Figure 2.2).104 Similar behavior was
observed with anatase TiO2; however, platinum deposition required addition
of isopropanol as a sacrificial oxidant. The platinum was preferentially deposited on the {101} facet of anatase, and PbO2 growth occurred on the {001}
facet. The results indicate separation of electrons and holes on dierent
facets of the nanocrystal during irradiation. The eect was stronger for rutile
particles than for the anatase particles.
A similar result was found when anatase TiO2 nanocrystals with {101} and
{001} facets were evaluated for mineralization of aqueous phenol in the
presence of O2. Charge trapping centers on the catalyst surface (Ti31, O,
and O2) that formed upon UV irradiation were quantitatively evaluated
using EPR spectroscopy.105 During exposure to vacuum conditions, the
trapped holes (O centers) increased with higher proportions of {001} facets;
while the Ti31 sites were associated with the {101} surfaces. Such a mechanism can allow for separation of the self-trapped polarons, which could
provide a big advantage in photocatalysis. Otherwise, their proximity on the
same facet could make recombination more competitive. The photooxidation of phenol was found to increase with nanocrystals that had higher
proportions of {001} facets. Because of the correlation between trapped holes
and increased activity with increasing ratio of the {001} facet, the mechanism for photooxidation appeared to be driven by trapped holes. The Ti31
sites act as reduction centers that form superoxide in the presence of O2 and
contribute to the mineralization of phenol. It remains to be elucidated
whether the improved photooxidation of nanocrystals with {100} facets was
due to separation of charge-trapped centers that allow simultaneous
View Online
21
31
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
2.4.3
Electron transfer between hexacyanoferrate(III) and thiosulfate on Pt nanocrystals with tetrahedral, cubic, or spherical shape was studied.106 The tetrahedral particles are bound by {111} facets whilst cubic particles are bound
by {100} facets. The near spherical particles are likely a statistical distribution of truncated octahedra that feature both facets. The reaction was
conducted near room temperature and, consequently, there was little change
in the shape of the nanocrystals during catalysis. The activation energy of
catalytic electron transfer was 14.0 kJ mol1, 22.6 kJ mol1, and 26.4 kJ
mol1 for tetrahedral, spherical, and cubic particles, respectively. The result
shows that electron transfer was facet- (and therefore shape-) dependent,
which is consistent with established structural eects in electrocatalysis on
Pt single crystal electrodes.107
Monodisperse PVP-stabilized platinum nanocrystals with three shapes
cubes, cubo-octahedra, and octahedrawere synthesized.108 Shape control
was achieved with careful addition of an Ag1 ion that enhances the crystal
growth rate along [100]. The PVP stabilizer, however, is dicult to remove
from the Pt surface and blocks adsorbates thereby lowering catalytic activity.
Cetyltrimethylammonium bromide (CTAB) stabilized Pt nanocrystals were
synthesized with cubic, cubo-octahedral, and dendritic shapes.109 CTABPt
particles were an order of magnitude more active than PVPPt particles for
ethylene hydrogenation. Benzene hydrogenation was found to be shapedependent on the CTABPt nanocrystals.110 Cyclohexane and cyclohexene
were formed on cubo-octahedral nanoparticles; whereas, only cyclohexane
was produced on cubic nanoparticles. The results are consistent with data
obtained from single crystal studies.111 These findings implicate nanocrystal
morphology as a way to control selectivity in catalysis.
Turnover frequency of ethane hydrogenolysis on platinum supported on
mesoporous silica was studied. Monodisperse PVP-stabilized platinum
nanoparticles were synthesized with sizes 1.7 7.1 nm and added to
mesoporous silica supports.112,113 This model catalyst system diers from a
conventional catalyst in its synthesis approach. The Pt nanoparticles were
synthesized first using precise chemical methods that yielded monodisperse
particle samples. A conventional approach, such as wetness impregnation,
can yield much larger particle size distributions. This is a distinction between nanocatalysis and a more conventional catalysis study. Two routes
were used for the addition of Pt particles to the support. In the capillary
inclusion method, the PVPPt particles were combined with SBA-15 and
sonicated at room temperature for 3 hours; whereas, in the nanoparticle
encapsulation method, the PVPPt particles were added to the triblock
copolymer surfactant mixture in which mesoporous silica condenses.
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
22
Chapter 2
The latter method was devised to yield a catalyst with more even distribution
of particles throughout the support. The catalysts were calcined at 623723 K
to remove the PVP stabilizer. Chemisorption analysis with H2, O2, CO, and
H2O2 titration were performed to measure dispersion of the catalyst. The
activity for ethylene hydrogenation was measured as a means to calculate
dispersion due to its structure-insensitive nature. The dispersions measured
from chemisorption were compared with data from TEM and powder XRD
measurements, which establish the number of surface sites. The correct
tabulation of surface sites is required for the calculation of turnover frequency that enables a true comparison of the catalysts. Ethane hydrogenolysis, known to be a structure-sensitive reaction, was studied on the Pt/
SBA-15 catalysts. The turnover frequency increased with decreasing particle
size. This was attributed to the higher surface roughness in the smaller
particles; however, an electronic structure eect cannot be ruled out.
The catalytic hydrogenation of pyrrole was studied on several Pt catalysts.114 A series of uniform Pt nanocrystals with a cubic shape with sizes of
5, 6, 7, and 9 nm were prepared, as well as rounded polyhedra with sizes of
3.5 and 5 nm. The nanocrystals were supported on MCF-17 large-pore
mesoporous silica. The activity of the catalysts for ethylene hydrogenation
exhibited a turnover frequency of B6 s1 (10 Torr C2H4, 100 Torr H2, and 298
K) that is similar to Pt(111), and demonstrated active catalyst materials and
the structure-insensitive nature of the reaction. Pyrrole hydrogenation can
proceed step-wise to form pyrrolidine, n-butylamine, and butane/ammonia
wherein the latter products arise from CN bond scission. The reaction was
studied in the temperature range 385415 K. Over this temperature range,
cubic Pt nanocrystals showed high selectivity for n-butylamine. Interestingly,
at lower temperatures, the rounded polyhedra Pt catalyzed formation of
n-butylamine/pyrrolidine in approximately 2 : 1 ratio that became increasingly selective to n-butylamine at higher temperatures. SFG investigation of
pyrrole hydrogenation on Pt(111) and Pt(100) suggested a preference for
n-butylamine on the Pt(100) surface. In comparison, hydrogenation of pyrrole on B1 nm Pt particles over the same temperature range exhibited very
dierent selectivity, and pyrrolidine was the major product.115 The B1 nm
particles were formed from a dendrimer template that confines the size of
the metal domain within a very narrow size range. The small particles do not
feature any long range surface facets, so the preferential stabilization of
n-butylamine on Pt(100) was not possible.
Particle shape was found to be important in controlling the cistrans
isomerism of olefins.116 The interactions of cis- and trans-2-butene were
compared for Pt(111), Pt(557), Pt(100), and Pt(110)-(2 1) surfaces. Except in
the case of Pt(110), cis-2-butene desorbs from Pt surfaces at a higher temperature than the trans isomer, which indicates a higher stability of the cis
isomer on the surface. DFT calculations show deformation of the Pt(111)
surface on adsorption of 2-butene. The deformation involves movement of a
Pt atom out of plane toward the olefin, and is more severe (and energetically
unfavorable) for the trans isomer. The dierence in surface deformation is
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
23
likely the reason for the slightly higher stability of the cis isomer. The data
suggest that olefin isomerization on Pt should lead to conversion to the cis
isomer, as long as the catalyst surfaces exhibit the preferential stabilization
of the cis isomer. To this end, tetrahedral Pt nanocrystals were prepared and
deposited on silica xerogel. The Ptxerogel SiO2 catalysts were calcined at 475
K, 525 K, and 575 K. At 475 K, the tetrahedral shapes were retained; however,
at 575 K, the particles had become spherical in shape. Isomerization of
2-butene at 375 K was studied over the three catalysts with tetrahedral,
intermediate, and spherical shapes. The catalyst with tetrahedral Pt particles
exhibited selectivity for the cis isomer that greatly diminished with increasing population of spherical particles.
2.4.4
The use of dendrimers has emerged as a tool for the highly precise preparation of nanoparticles in the 12 nm range.117 Dendrimers are somewhat
like highly pre-organized micelles. They are polymers that grow radially from
a starting molecular block and repeating units that branch. Due to the
geometric expansion of the dendrimer with successive generations, their size
tends to be limited yet uniform. Functional groups, typically amines, on the
repeating units can coordinate metal ions. In this way, the core of a dendrimer can be filled with metal ions, and subsequent reduction leads to formation of a nanoparticle within the dendrimer. Due to the high uniformity of
the dendrimer, the nanoparticles are monodisperse. More importantly, it is
straightforward to synthesize well-defined bimetallic nanoparticles simply by
adjusting the metal ion ratio in the dendrimer.118 Such control is typically not
possible in untemplated nanoparticle synthesis. For example, random alloys
of Pt/Cu, Pd/Cu, Pd/Au, Pt/Au, and Au/Cu DENs were synthesized.119 These
materials were used to catalyze reduction of p-nitrophenol to p-aminophenol
by NaBH4. Theoretically, combinations of metals should exhibit unique
binding energy to p-nitrophenol that governs the reaction rate. Experiments
showed this to be true for metal combinations with matching lattice constants; however, metals with lattice mismatch showed bimodal distribution of
binding energies. This is an excellent example of synergism between theory
and experiment in the design and testing of model nanocatalysts. Whereas
dendrimers are limited in size, star-like block copolymers oer the advantages
of micellular pre-organization over a larger size range.120 Both of these phenomena are made possible through the use of uniform, high-quality polymers, and there is an interesting parallel between monodisperse polymers
and the ability to organize matter at the nanoscale.
2.4.5
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
24
Chapter 2
growing nanocrystal are of dierent compositions and crystal phases. Reactive facets or surface defects on the seed crystal serve as nucleation sites to
form the hybrid nanocrystal. Given the seed is a well-defined nanocrystal, it
is theoretically possible to construct the hybrid nanocrystal through a rational approach upon selective nucleation of the new phase at specific
positions on the seed. Hybrid nanocrystals that consisted of CdSe nanorods
with Au tips were synthesized according to this principle.121 Other examples
have appeared in the literature, and excellent reviews can be found in the
literature.122 An attractive feature of hybrid nanocrystals is the ability to join
completely dierent materials at small length scales. For instance, one may
combine a semiconductor particle with a magnetic particle to build a
multifunctional nanocrystal. From a catalysis perspective, there is an enormous opportunity to capitalize on studying the eect of material combinations in catalytic reactions.
Metalsemiconductor hybrid nanocrystals have attracted attention as
photocatalyst materials.123 Excitation of the semiconductor component results in an exciton that can become separated as the electron becomes
trapped on the metal domain. Careful selection of the metal and semiconductor domain in which they catalyze reduction and oxidation half-reactions, respectively, should give rise to an enhanced photocatalytic eect.
This idea was the impetus for the study of modified titania colloids that
could achieve water splitting.124 New metalsemiconductor hybrid nanocrystals based on semiconductor nanorods can oer the specific advantages
of high longitudinal mobility of charge carriers and well-defined sites for
selective nucleation of catalyst particles. Ultrafast charge separation in
matchstick type AuCdS hybrid nanocrystals was studied using an optical
time-resolved pump probe technique and probing the light-induced change
of their optical response.125 Electronhole pairs photoexcited in the semiconductor part of the nanohybrids are shown to undergo rapid charge separation with the electron transferred to the metal part on a sub-20 fs
time scale.
2.5 Conclusion
Catalysis began with a tradition of empirical work that evolved into a rational
design of materials to attain improved performance. The development of
model systems has been critical to the understanding of catalytic processes.
The study of chemistry on surfaces has had profound implications, such as
identification of surface species, the role of crystal facets in adsorption energy or reaction rates, restructuring processes, and surface mobility. Model
three-dimensional systems were lacking until it became possible to synthesize monodisperse nanocrystals. Advances in nanocrystal synthesis
methods extend the limits of model catalyst systems, and will surely open
the door to new performance possibilities in catalysis. The ability to precisely
control matter at the nanoscale, therefore its electronic structure and surface
features, is at the heart of nanocatalysis.
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
25
References
1. I. Chorkendor and J. W. Niemantsverdriet, Concepts of Modern Catalysis, Wiley-VCH, Weinheim, Germany, 1st edn, 2003.
2. J. H. Sinfelt, Surf. Sci., 2002, 500, 923946.
3. More detailed historic accounts can be found in:(a) A. J. B. Robertson,
Platinum Met. Rev., 2002, 19, 6469; (b) J. K. Smith, History of Catalysis,
in Encyclopedia of Catalysis, John Wiley & Sons, Inc., 2002;
(c) J. N. Armor, Catal. Today, 2011, 163, 39.
4. M. Che and C. O. Bennett, Adv. Catal., 1989, 36, 55172.
5. D. R. Rolison, Science, 2003, 299, 16981701.
6. P. L. J. Gunter, J. W. Niemantsverdriet, F. H. Ribiero and G. A. Somorjai,
Catal. Rev.: Sci. Eng., 1997, 39, 77168.
7. M. Boudart, Chem. Rev., 1995, 95, 661666.
8. It is important to point out that reaction order can be temperature
dependent!.
9. T. Kwan, J. Phys. Chem., 1956, 60, 10331037.
10. H. Song, R. M. Rioux, J. D. Hoefelmeyer, R. Komor, K. Niesz, M. Grass,
P. Yang and G. A. Somorjai, J. Am. Chem. Soc., 2006, 128, 30273037.
11. R. M. Rioux, H. Song, J. D. Hoefelmeyer, P. Yang and G. A. Somorjai,
J. Phys. Chem. B, 2005, 109, 21922202.
12. R. Van Hardeveld and F. Hartog, Surf. Sci., 1969, 15, 189230.
13. S. Maksimuk, X. Teng and H. Yang, J. Phys. Chem. C, 2007, 111,
1431214319.
n, S. Fuentes and J. M. Dominguez, Surf. Sci., 1981, 106,
14. M. J. Yacama
472477.
15. M. Gillet and A. Renou, Surf. Sci., 1979, 90, 91101.
16. B. Coq and F. Figueras, Coord. Chem. Rev., 1998, 178, 17531783.
17. A. S. McLeod and L. F. Gladden, J. Catal., 1998, 173, 4352.
18. L. F. Gracia and E. E. Wolf, Chem. Eng. J., 2001, 82, 291301.
19. S. Ladas, Surf. Sci., 1986, 175, L681L686.
20. R. B. Greegor and F. W. Lytle, J. Catal., 1980, 63, 476486.
21. J. J. Burton, Catal. Rev., 1974, 9, 209222.
22. G. Wul, Z. Kristallogr., 1901, 34, 449530.
23. (a) L. D. Marks, Surf. Sci., 1985, 150, 358366; (b) L. D. Marks, Rep. Prog.
Phys., 1994, 57, 603649.
24. M. Mavrikakis, M. Baumer, H. J. Freund and J. K. Norskov, Catal. Lett.,
2002, 81, 153156.
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
26
Chapter 2
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
27
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
28
Chapter 2
86. Y. W. Jun, J. H. Lee, J. S. Choi and J. Cheon, J. Phys. Chem. B, 2005, 109,
1479514806.
87. Y. W. Jun, J. W. Seo, S. J. Oh and J. Cheon, Coord. Chem. Rev., 2005, 249,
17661775.
88. S. Maksimuk, X. Teng and H. Yang, J. Phys. Chem. C, 2007, 111,
1431214319.
89. G. Viau, R. Brayner, L. Poul, N. Chakroune, E. Lacaze, F. Fievet-Vincent
and F. Fievet, Chem. Mater., 2003, 15, 486494.
90. M. W. Brink, M. A. Peck, K. L. More and J. D. Hoefelmeyer, J. Phys.
Chem. C, 2008, 112, 1212212126.
91. J. Watt, C. Yu, S. L. Y. Chang, S. Cheong and R. D. Tilley, J. Am. Chem.
Soc., 2013, 135, 606609.
92. G. A. Somorjai, F. Tao and J. Y. Park, Top. Catal., 2008, 47, 114.
93. R. Schlogl and S. B. Abd Hamid, Angew. Chem., Int. Ed., 2004, 43,
16281637.
94. G. A. Somorjai and J. Y. Park, Top. Catal., 2008, 49, 126135.
95. This is not a comprehensive account. We note an excellent special issue
Nanoparticles for Catalysis in Acc. Chem. Res., 2013, 46(8), 1671 that
includes recent major developments.
96. M. Valden, X. Lai and D. W. Goodman, Science, 1998, 281, 16471650.
97. G. R. Bamwenda, S. Tsubota, T. Nakamura and M. Haruta, Catal. Lett.,
1997, 44, 8387.
98. U. Diebold, Surf. Sci. Rep., 2003, 48, 53229.
99. H. G. Yang, C. H. Sun, S. Z. Qiao, J. Zou, G. Liu, S. C. Smith,
H. M. Cheng and G. Q. Lu, Nature, 2008, 453, 638641.
100. M. Liu, L. Piao, L. Zhao, S. Ju, Z. Yan, T. He, T. He, C. Zhou and
W. Wang, Chem. Commun., 2010, 46, 16641666.
101. X. Zhao, W. Jin, J. Cai, J. Ye, Z. Li, Y. Ma, J. Xie and L. Qi, Adv. Funct.
Mater., 2011, 21, 35543563.
102. X. Han, B. Zheng, J. Ouyang, X. Wang, Q. Kuang, Y. Jiang, Z. Xie and
L. Zheng, Chem.Asian J., 2012, 7, 25382542.
103. J. Pan, G. Liu, G. Q. M. Lu and H. M. Cheng, Angew. Chem., Int. Ed.,
2011, 50, 21332137.
104. T. Ohno, K. Sarukawa and M. Matsumura, New J. Chem., 2002, 26,
11671170.
105. M. DArienzo, J. Carbajo, A. Bahamonde, M. Crippa, S. Polizzi, R. Scotti,
L. Wahba and F. Morazzoni, J. Am. Chem. Soc., 2011, 133, 1765217661.
106. R. Narayanan and M. A. El-Sayed, Nano Lett., 2004, 4, 13431348.
107. For example:(a) J. Clavilier, R. Parsons, R. Durand, C. Lamy and
J. M. Leger, J. Electroanal. Chem., 1981, 124, 321326; (b) C. Lamy,
J. M. Leger, J. Clavilier and R. Parsons, J. Electroanal. Chem., 1983, 150,
7177.
108. H. Song, F. Kim, S. Connor, G. A. Somorjai and P. D. Yang, J. Phys.
Chem. B, 2005, 109, 188193.
109. H. Lee, S. E. Habas, S. Kweskin, D. Butcher, G. A. Somorjai and
P. D. Yang, Angew. Chem., Int. Ed., 2006, 45, 78247828.
View Online
23/06/2014 08:14:40.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00006
29
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
CHAPTER 3
3.1 Introduction
Among the various catalyst materials, metal NPs (1 to 100 nm) in particular
are gaining increasing attention because of their existence on the borderline
molecular states with discrete quantum energy levels.13 Thus, NP-based
catalysts are considered to bridge the gap between mononuclear metal
complexes and heterogeneous bulk catalysts; this system is often referred to
as a quasihomogeneous system.4 Their large surface area-to-volume ratio
and the high concentration of low coordination sites and surface vacancies
allow eective utilization of expensive catalyst metals.
The general routes to NP synthesis are based on the chemical reduction of
metal precursors with the appropriate reducing reagents in the presence of
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
30
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
31
Single-site photocatalysts, exemplified by Ti-containing zeolites and Ticontaining mesoporous silicas, have been shown to exhibit exceptional
photocatalytic activities compared to bulk TiO2.911 These dierences can be
explained by considering their photo-excitation mechanism. In the case of
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
32
Figure 3.1
Chapter 3
bulk TiO2, electrons and holes in the conduction and valence bands formed
via photoexcitation contribute to the reduction and oxidation reaction, respectively (Figure 3.1A).12 On the other hand, in the case of single-site
photocatalysts, excitation by light at around 230270 nm brings about an
electron transfer from the oxygen to Ti41 ions, resulting in the formation of
pairs of trapped hole centers (O) and electron centers (Ti31).13,14 These
charge transfer excited states, i.e. the electron hole pair state which localize
quite near to each other compared with the electrons and holes generated in
TiO2, play a crucial role in photocatalytic reactions (Figure 3.1B). It can be
expected that the metal precursor species can be easily deposited on the
excited state of a single-site photocatalyst to form well-controlled metal NPs
from the mixture of single-site photocatalysts in the aqueous solution of
metal precursors. Recently, we have successfully utilized single-site photocatalysts as a platform for the synthesis of nano-size metal catalysts by the
advanced photo-assisted deposition (PAD) method.1520
Under UV-light irradiation of a slurry of Ti-containing zeolites (TS-1,
Pd/Ti 0.6) in an aqueous PdCl2 solution, the Pd metal can be successfully
deposited on the TS-1. No Pd deposition was observed both on the silicalite zeolite without Ti-oxide under UV-light irradiation and the TS-1 zeolite
without UV-light irradiation. These results suggest that the Pd precursor
underwent anchoring on the single framework Ti41 center of TS-1 under UVlight irradiation. The subsequent calcination at 723 K for 5 h and reduction
with H2 at 473 K for 1 h generates nano-sized Pd metal particles (PAD-Pd/TS-1).
The TEM images showed that the PAD-Pd/TS-1 exhibits nano-sized Pd
metals with a mean diameter of ca. 2.1 nm and has a narrow size distribution, while the imp-Pd/TS-1, which was prepared by a conventional impregnation method, had Pd metal particles with various sizes within the
range of 215 nm. These results clearly suggest that the size of the Pd metal
particles depends on the preparation method, and that the smaller Pd metal
particles are formed on the photo-deposited catalyst rather than on the
impregnated catalyst.
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
Figure 3.2
33
View Online
34
Chapter 3
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
23
(Ti-HMS). The PAD-PdNi/Ti-HMS catalyst exhibited specifically higher hydrogenation activity than imp-PdNi/Ti-HMS prepared by conventional
methods. In particular, it was proven that the PAD-PdNi/Ti-HMS with
Ni/Pd 1.5 mol% exhibited a strong synergistic eect. These studies
verified a promising approach for the preparation of supported metal catalysts, and we expect that this synthetic strategy will be extendable in the
preparation of other bimetallic nanoparticles as unique catalysts and/or
photocatalysts.
3.2.2
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
Figure 3.3
35
3.2.3
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
36
Figure 3.4
Chapter 3
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
Figure 3.5
37
The reaction rate increased with increasing the pH value of the prepared
colloidal solution, and the Ag/Al2O3 (3.7) with the smallest Ag NPs exhibited
higher activity in the reduction of 4-nitrophenol in the presence of NaBH4,
because the higher dispersion of Ag metal NPs is preferable for the catalytic
reactions.
We also exploited the size controlled deposition of Ag nanoparticles on
Al2O3 with the assistance of a photo-induced chromic reaction.31 2-Hydroxychalcon derivatives are known to be one of the organic photochromic
compounds. By light irradiation to trans-2-hydroxychalcon derivatives under
slightly acidic conditions, flavylium cations are photochemically produced
via cis-2-hydroxychalcon and flav-3-en-2-ol, as shown in Figure 3.5.32 In the
presence of trans-2-hydroxychalcone, the photo-irradiation gradually neutralized the negative electric charge of 3-MPA-stabilized Ag NPs owing to the
formation of flavylium cations, which induced the assembly of Ag NPs. We
demonstrated that such photo-induced assembly-dispersion control of Ag
NPs enables a size selective deposition on an Al2O3 support by varying the
photo-irradiation time while keeping their inherent aggregated form. After
calcination and reduction processes, the assemblage of Ag NPs transforms
into monocrystalline Ag particles with dierent diameters. From the TEM
image, the average diameter gradually increased with increasing the irradiation time in the preparation sequence; the average diameters after 0, 2, 4,
and 8 hours of irradiation times were estimated to be 10.2, 12.3, 15.9,
and 20.5 nm, respectively. It can be concluded that these preparation
methods based on the precise control of the assembly-dispersion state of the
colloidal NPs enable a strong protocol to create supported metal catalysts
with dierent diameters. The present strategy provides great flexibility in
the selection of the kind of metals as well as primary NP sizes. Introduction
of these features into the catalyst design enables the achievement of desired
supported metal catalysts for the target catalytic reactions.
View Online
38
Chapter 3
3.3.1
Figure 3.6
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
Figure 3.7
39
TEM images show that the spherical particles have a 4550 nm thick TiSiO2 shell (Figure 3.7). Each Ti-SiO2 sphere encapsulates one FePt NPs, although a small fraction of the particles possesses either two or zero FePt
NPs. The average diameter of FePt NPs was determined to be ca. 6 nm. Ti
K-edge XAFS spectra also confirm the formation of isolated and tetrahedrally-coordinated Ti-oxide moieties within the silica shell framework.
The catalytic ability of the FePt@Ti-SiO2 was investigated in the epoxidation of cyclooctene using 30% H2O2 as a simple test reaction. The results
of this reaction indicated that the corresponding cyclooctene oxide was
obtained in 499% selectivity and the turnover number (TON) based on
Ti approached 48 after 24 h. A more important advantage of the FePt@TiSiO2 is the facile recovery from the reaction mixture and the high reusability.
Upon completion of the oxidation reaction, the magnetic properties of the
FePt@Ti-SiO2 can aord a straightforward means to isolate the catalyst from
the colloidal solution. By applying an external permanent magnet, the
catalyst was attracted. The recovered catalyst could then be recycled in the
epoxidation of cyclooctene at least three times while maintaining identical
inherent activity to the initial run.
3.3.2
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
40
Figure 3.8
Chapter 3
composed of six (a-CD), seven (b-CD), or eight (g-CD) a(1,4)-linked glucopyranose units. The cavity exhibits a hydrophobic character, which allows
CDs to form inclusion complexes with hydrophobic molecules that fit into
the cavity.37 We present a CD-stabilized FePt magnetic NP exhibiting catalytic activity under aqueous conditions,38 in which the composition of FePt
NPs was controlled to form FecorePtshell by the first decomposition of Fe(CO)5
to form the Fe core, followed by the successive reduction of Pt(acac)2 in the
presence of oleic acid and oleylamine. Elemental analysis confirmed that the
average composition of the NPs was Fe48Pt52.
As shown in Figure 3.8, the as-synthesized FePt NPs can be dispersed in
nonpolar hexane solvent due to the presence of oleic acid and oleylamine.
A mixture of a hexane suspension of as-synthesized FePt NPs and an equal
volume of g-CD aqueous solution was stirred vigorously at room temperature. After stirring for 24 h, the top hexane layer becomes colorless, confirming the existence of water-compatible g-CD as host molecules
(Figure 3.8.). It can be said that the surface properties of the NPs have been
modified through the formation of an inclusion complex between surfacebound organic molecules and g-CD. The surface coverage of g-CD on the
FePt NPs was roughly estimated to be ca. 60 % by CHN elemental analysis.
The aqueous reaction of allyl alcohol using FecorePtshell NPs capped with
g-CD proceeded smoothly to give the corresponding n-propanol. It should be
noted that the reaction using the FeshellPtcore NPs as well as Fe NPs stabilized
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
41
by g-CD hardly occurred because Fe atoms do not accelerate the aforementioned hydrogenation reaction, confirming that our FePt NPs consists of
an Fe-rich core and a Pt-rich shell. More interestingly, the reaction rate
performed with FePt NPs capped with g-CD in water was three times higher
than hydrophobic FePt NPs stabilized by oleic acid/oleylamine in THF.
Similar results were observed in other organic solvents, such as toluene and
hexane. We suppose that the hostguest complexation of organic substrate
(allyl alcohol) with g-CD on FePt NPs is the major driving force for the enhancement of the catalytic eciency under aqueous conditions. Such a
driving force is lost in the case of the hydrophobic FePt NPs stabilized
by oleic acid/oleylamine in an organic solvent as the reaction medium,
resulting in a low catalytic activity.
3.3.3
With a similar strategy for catalyst design, the FePd NP with an Fe-rich core
and a Pd-rich shell have been modified with chiral 2,2 0 -bis(diphenylphosphino)-1,1 0 -binaphthyl (BINAP) through the simple ligand exchange procedure (Figure 3.9).39 In the FT-IR spectrum, the peaks due to the methylene
n(CH) stretch vibration of oleic acid and/or oleylamine disappeared, accompanied with the appearance of the aromatic ring n(CH) stretch vibration of BINAP ligand at around 28003000 cm1. The circular dichroism (CD)
spectra of (S)-BINAP-modified NPs in CHCl3 showed a negative cotton eect,
while (R)-BINAP-modified NPs showed a positive cotton eect. For the
structural model of the FePd NPs, the XAFS results suggested that most of
the Pd atoms were preferentially located in the shell region, while the Fe
atoms were preferentially located in the core region.
It was found that the FePd NPs modified with (S)-BINAP served as a
chiral nanocatalyst in the SuzukiMiyaura coupling reaction of 1-bromo2-methylnaphthylene (1) and naphthylboronic acid (2) aord the
corresponding (S)-binaphthalene (3) in 499% yield with a moderate enantioselectivity of 48% (Figure 3.9). Since Fe atoms do not accelerate these
coupling reactions, our FePd NPs are thought to consist of an Fe-rich core
and a Pd-rich shell. In the case of (R)-BINAP, (R)-binaphthalene was obtained
with the same level of chemical yield and enantioselectivity, revealing
that the absolute configuration of the coupling products is determined
by the employed ligands. Periodic monitoring of the coupling reaction
showed no induction period or degradation: the enantioselectivity appeared
to be independent of the conversion level and remained unchanged
during the reaction, suggesting that racemization of the product did not
take place.
Figure 3.9 Ligand exchange procedure of FePd NP from oleic acid/oleylamine to chiral BINAP and asymmetric SuzukiMiyaura coupling reaction.
The red blue core is Fe rich and the yellow shell is Pt rich.
View Online
Chapter 3
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
42
Figure 3.9
3.3.4
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
43
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
44
Chapter 3
Figure 3.10
active site. With further optimization of the pore diameter and thickness of
the Ti-MS shell, enhancement of the catalytic activity can be attained by a
factor of 6.41 Therefore, the coreshell catalyst is considered to be a promising structure for one-pot reactions.
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
45
View Online
23/06/2014 08:14:41.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00030
46
Chapter 3
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
CHAPTER 4
4.1 Introduction
Catalysis is an important area of chemistry in which metal nanoparticles
(MNPs) are considered as promising substitutes of catalysts; both for
homogeneous (molecular complexes) and heterogeneous (bulk metals on
supports) ones.14 MNPs with their small size are highly interesting systems
due to their high proportion of surface atoms that oer numerous active
sites and to their unique electronic properties at the frontier between the
molecular and metallic states.5 Furthermore, the surface chemistry of MNPs
can be tuned by addition of a stabilizer or by using a support. Polymers,
dendrimers, ionic liquids, surfactants, ligands etc. used as capping agents
can transfer their own properties such as solubility or chirality whereas
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
47
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
48
Chapter 4
widely used supports such as ceria, titania, alumina, silica, or carbon materials can give a synergy to the nanoparticle systems to orientate a catalytic
reaction.6,7 There are some examples in literature that focus on both aspects
by combining the ligands with supports. With intensive studies over the past
two decades, nanocatalysis has emerged as a modern domain of catalysis at
the borderline between homogeneous and heterogeneous ones. However,
from an industrial point of view regarding sustainable and economical
concerns, the use of MNPs as catalysts is still hampered by two important
drawbacks: the stability of the particles and the recovery/recyclability of the
catalyst.8,9 To overcome these problems, several approaches have been developed, such as the design of appropriate stabilizing agents or the immobilization of MNPs in a magnetic solid support, which allows an easy
separation of the nanocatalyst from the reaction mixture. Regardless of the
method followed for the preparation, promising alternatives to increase
stability and recycling of the nanocatalysts have been reported.
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
Figure 4.1
49
Schematic representation of the organometallic approach for the synthesis of MNPs by hydrogen decomposition of a metalloorganic
complex.
chemistry. The decomposition step is performed through reduction or ligand displacement from the metal coordination sphere, most often in tetrahydrofuran (THF) solution and in the presence of a stabilizer (see
Figure 4.1).1517 This allows the control of particle size, shape and surface
state in order to reach a monodisperse assembly of NPs with the desired
properties. This method can also be applied to bimetallic systems.18,19 The
choice of the stabilizing agent is fundamental as it will govern the growth,
the stability and the catalytic performance of the MNPs.20 Various functional
organic molecules such as polymers,21,22 alcohols,23,24 ionic liquids2527 and
various ligands2830 can be used for this purpose. Immobilization of NPs
using supports like alumina, silica31 or carbon materials is also possible and
they allow the easy orientation of the catalytic reaction and facilitate the
recovery of the nanocatalysts. During these eorts, simple spectroscopic
methods derived from molecular chemistry, such as infrared (IR) and nuclear magnetic resonance (NMR), both in solution and in the solid state, aid
in precisely characterizing the metallic surface.3237
In the following parts of this chapter, we will present an overview of the
recent developments in the organometallic synthesis of noble metal nanoparticles for application as catalysts in solution or supported conditions.
This review is not an exhaustive report but aims to show the progress obtained recently. The results are reported taking into account the targeted
catalytic reaction, with the emphasis on either the nature of the stabilizing
ligand (ionic liquids, organic solvent- or water-soluble ligands) for nanocatalysis in solution or of the support for supported catalysis.
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
50
Chapter 4
interaction with the metallic surface. Another interest for us in using ligands
was their potential to orientate the course of a catalytic reaction by their
selectivity, a well-known phenomenon in homogeneous catalysis. Numerous
eorts have been devoted to the use of simple ligands but also to more sophisticated ones as these parameters are major for the development of
nanocatalysts with improved catalytic performances in comparison to
known homogeneous and heterogeneous catalysts.
Studies in the improvement of the NP synthesis mainly involved ligands
containing nitrogen (aminoalcohols, oxazolines), phosphorus (phosphines,
diphosphites) or carbon (carbenes) as coordinating atoms to the metal
surface. Ionic liquids were also proven to be very ecient stabilizers to obtain size-controlled MNPs, and the coaddition of ligands could oer new
perspectives in this way. Some of our NP systems have been applied in
catalysis, mainly in hydrogenation reactions, but also in the dehydrogenation of amine-boranes, carboncarbon coupling and hydroformylation
reactions.
4.3.1
Hydrogenation Reactions
A high number of papers describe the preparation of various MNPs in suspension for their application as catalysts in the hydrogenation of substrates
mainly of olefin and arene derivatives in addition to the carbonyl substrates
such as ketones.38 The asymmetric hydrogenation of prochiral substances is
also a domain explored with nanocatalysts.
4.3.1.1
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
Figure 4.2
51
aminooxazoline-stabilized RuNPs and their catalytic behavior in hydrogenation of ortho- and para-anisoles.39
catalyst was found to be inactive. In both cases, the trans-isomer was favoured, with a transcis ratio up to 19. The very low asymmetric induction
encountered with this colloidal catalyst suggested a fluxional behaviour of
the ligands at the surface of RuNPs, as previously described for Ru/amine
colloids.22
4.3.1.1.2 Diphosphite-stabilized Ru, Rh and Ir Nanoparticles. In a coln, Claver and Roucoux, 1,3-diphoslaboration with the groups of Castillo
phites derived from carbohydrates were used as stabilizers for the
preparation of Ru, Rh and Ir NPs for their investigation in asymmetric hydrogenation reactions.40,41 These ligands allowed the formation of very
small Ru, Rh and Ir NPs in mild conditions (THF; 3 or 6 bar H2; r.t.) from
[Ru(COD)(COT)], [Rh(Z3-C3H5)3] and [Ir(COD)Cl]2 complexes, respectively.
Structural modifications in the diphosphite backbone were found to influence the NP size, dispersion and catalytic activity. In the hydrogenation of
ortho- and meta-methylanisoles, the RhNPs were shown to display the
highest activity while the IrNPs presented the lowest one. In contrary with
the results previously described for the aminooxazoline-stabilized RuNPs,
the hydrogenation of o-anisole gave total selectivity for the cis-product in
all cases. However, the ee of the product was always less than 6%.
A maximum of 81% cis-selectivity was obtained for the hydrogenation of
m-anisole, but with no enantioselectivity.
4.3.1.1.3 Carbene-stabilized Ru Nanoparticles. In a collaborative work
with van Leeuwen et al., RuNPs stabilized by N-heterocyclic carbenes
(NHC), namely N,N-di(tert-butyl)imidazole-2-ylidene (ItBu) and 1,3-bis(2,6diisopropylphenyl)imidazole-2-ylidene (IPr),42 were used as catalysts in the
hydrogenation of several substrates under various reaction conditions
(solvent, substrate concentration, substratemetal ratio, temperature). The
Ru/NHC NPs appeared as active catalysts in the hydrogenation of aromatics and showed an interesting ligand eect; Ru/IPr NPs were generally
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
52
Figure 4.3
Chapter 4
Synthesis of Ru/NHC NPs and their catalytic performance in hydrogenation of ortho-methylanisole (0.5 M) in dierent solvents (Ru/IPr0.2
(0.3% Ru); 40 bar H2; 298 K : pentane (black), THF (red), methanol
(blue), no-solvent (green, 0.1% Ru).43
more active than Ru/ItBu NPs.43 The influence of the reaction parameters
(changes in solvent, temperature, etc.) was also studied (see Figure 4.3).
4.3.1.2
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:1
Figure 4.4
Synthesis of PTA-stabilized NPs and TEM images in THF and in water for Ru NPs (left) and Pt NPs (right).44
53
View Online
54
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
Table 4.1
Chapter 4
Hydrogenation of olefins and aromatic derivatives with aqueous Ru/PTA
and Pt/PTA colloidal solutions.a
Catalyst
Substrate
P H2 (bar)
Time (h)
Conversion (%)f
Pt/PTAb
Pt/PTAb
Pt/PTAc
Pt/PTAc
Pt/PTAc
Pt/PTAc
Octene
Octene
Toluene
Toluene
m-Methylanisole
m-Methylanisole
1
1
10
10
10
10
2
2
2
16
2
16
100
68g
8
100
15
100
Ru/PTAd
Ru/PTAd
Ru/PTAe
Ru/PTAe
Dodecene
Octene
Toluene
m-Methylanisole
1
10
10
10
5
1
16
16
100h
100
100
60
with pentane and filtration, and were further dissolved in water without
any change in dispersion and in their mean diameter, with a value of B1.4
nm and B1.1 nm for the RuNPs and the PtNPs, respectively. Aqueous suspensions of these NPs were stable for weeks when kept under an argon atmosphere. 1H, 13C, 31P solution and solid-state NMR studies showed the
strong coordination of PTA at the surface of the particles by the
phosphorous atom.
Biphasic liquidliquid hydrogenation was investigated with model olefins
and aromatic substrates using aqueous colloidal solutions of Ru/PTA and
Pt/PTA NPs as catalysts, with Roucoux et al. (see Table 4.1).45 Octene and
dodecene were totally converted into the corresponding alkanes (r.t.; 1 bar H2),
with moderate activities. An increase in the hydrogen pressure (P(H2) 10 bar)
was not detrimental for the colloidal suspension stability. Complete hydrogenation of toluene into cyclohexane was observed overnight with Ru/PTA NPs
whereas 60% of m-methoxymethylcyclohexane was formed from methoxymethylanisole. In comparison, the reduction with Pt/PTA NPs was achieved
after 16 h even if very low conversions were obtained after 2 h with 8% and 15%
cyclohexyl derivatives, respectively. In summary, these PTA-stabilized Ru and
Pt NPs were active in the hydrogenation of olefinic and aromatic compounds
under mild conditions despite the change of environment that they underwent after their dissolution into water.
4.3.1.2.2 Sulfonated Diphosphine-stabilized Ru Nanoparticles. By applying the same procedure as previously described with the PTA ligand
(see Figure 4.5), we used sulfonated diphosphines to stabilize RuNPs.46 In
this collaborative work with Roucoux et al. and Monflier et al., the RuNPs
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:1
Figure 4.5
Synthesis and TEM images of sulfonated diphosphine-stabilized RuNPs in water with (a) dppb-TS (b) dppp-TS, and (c) dppeTS with [L][Ru] 0.1.46
55
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
56
Chapter 4
were prepared from [Ru(COD)(COT)] and the diphosphines in THF (3 bar H2;
r.t.). Dierent diphosphines, (1,4-bis[(di-m-sulfonatophenyl)phosphine]butane dppb-TS, 1,4-bis[(di-m-sulfonatophenyl)phosphine]propane
dppp-TS, 1,4-bis[(di-m-sulfonatophenyl)phosphine]ethane dppe-TS), and
ligandRu ratios were employed in order to analyze the eect of the backbone
and the diphosphine concentration on the stability and the size of the NPs
and thus on their catalytic properties. Depending on the ligand amount,
well-crystallized RuNPs in a mean size range of 1.21.5 nm were formed. The
coordination of the sulfonated diphosphines at the surface of the RuNPs
allowed their further dispersion in water giving rise to very homogeneous
and stable aqueous colloidal solutions (up to several months) without any
change in their mean sizes.
The catalytic behaviour of these aqueous colloidal solutions showed
promising results in terms of reactivity when tested for the hydrogenation of
unsaturated substrates (tetradecene, styrene and acetophenone) in biphasic
liquidliquid conditions. Interestingly, minor structural dierences in the
diphosphine ligands, such as the alkyl chain length, influenced the catalytic
activity in styrene hydrogenation significantly, in addition to the positive
eect of an increase in temperature (from 20 1C to 50 1C) or pressure (from 1
to 10 bar H2) (see Table 4.2). As the [ligand][Ru] ratio increased, conversion
and selectivity (expressed as ethylbenzene (EB)ethylcyclohexane (EC) ratio)
Table 4.2
Ru NPs
H2
ST
EB
EC
Nanocatalysts
T (1C)
Time (h)
Ru/dppb-TS
20
20
20
1
20
40
25
0
0
75
45
3
0
55
97
Ru/dppp-TS
20
20
20
1
20
40
59
0
0
40
47
2
1
53
98
Ru/dppe-TS
20
20
20
1
20
40
75
0
0
24
41
1
1
59
99
Ru/dppb-TS
50
50
50
1
2
3
10
0
0
90
82
0
0
18
100
Reaction conditions: [ligand][Ru] 0.1; ruthenium (3.9 105 mol), styrene (3.9 103 mol),
1 bar H2, water (10 mL).
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
57
also increased at short reaction times. The best results were obtained with
the dppb-TS ligand with the longest alkyl chain, giving rise to 75% EB after
1 h compared to 40% EB and 20% EB with dppp-TS and dppe-TS, respectively. As all the RuNPs display similar mean sizes, the observed dierences
in their catalytic properties were correlated to the dierence in the flexibility
of the alkyl chain. Due to having the highest number of carbon atoms, the
dppb-TS ligand had the highest flexibity and therefore favoured a better
diusion of the substrate towards the metal surface. Although the mean size
(1.25 nm) of the RuNPs did not alter with the increase of [dppe-TS][Ru] from
0.2 to 0.5, the variation in selectivity was explained by a limited access of the
aromatic substrate to the NP surface due to increased steric hindrance when
more ligands were coordinated. A molar ratio of 0.1 appeared to be a good
compromise between the stabilization and the catalytic activity of the NPs.
Finally, as also observed for Ru/PTA NP systems, preliminary results of recycling were encouraging for the recovery of these water-soluble Ru/sulfonated diphosphine nanocatalysts.
4.3.1.3
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:1
58
Chapter 4
Figure 4.6
Influence of the ionic liquid chain length on the size of the RuNPs.26
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
Figure 4.7
59
was smaller at 0 1C than at 25 1C, and the stirring induced their agglomeration. Additionally, the increase in alkyl chain lengths was found to be linearly proportional to the mean size of the RuNPs.
The influence of added amines on the size control of IL-prepared RuNPs
and on their stability was also studied.51,52 RuNPs are commonly used as
catalysts in hydrogenation reactions but IL-stabilized RuNPs are generally not
stable in the necessary catalytic conditions. The ionic liquids [RMIm][NTf2]
(R CnH2n11, n 2, 4, 6, 8, 10) and two amines, octylamine (OA) and hexadecylamine (HDA), were used as added stabilizers. The synthesis of the
NPs was performed following the previously described conditions, but with
0.2 molar equiv. of the chosen ligand. Regardless of the alkyl chain length of
the IL, well-dispersed and well-crystallized NPs were obtained with a mean
size in the range of 1.11.3 nm (see Figure 4.7). These RuNPs displayed a
better size dispersity than the corresponding pure amine-stabilized RuNPs.
The IL prevents particles from agglomeration through a confinement eect
but amines also play an important role in controlling their size and dispersion. This demonstrates the interest in combining the properties of ionic
liquids to confine nanoparticles in non-polar domains (nanoreactors) with the
presence of a ligand that stabilizes the particles at a very small size.
The hydrogenation of toluene was first performed with the RuNPs synthesized in C1C4ImNTf2 in the presence of octylamine, at dierent temperatures (see Table 4.3). The conversion obtained was low but reproducible.
A temperature of 75 1C was further chosen for studying the reaction with the
other colloids. Since Ru/HDA and Ru/OA NPs have similar size and shape,
the slightly lower conversion obtained by hexadecylamine was explained by
View Online
60
Chapter 4
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
Table 4.3
Hydrogenation
of
toluene
C1C4ImNTf2 octylamine.a
with
RuNPs
synthesized
in
Nanocatalyst
IL/ligand
T
(1C)
PhMe/
RuT
PhMe/
RuS
RuS/
RuT
Conversion
(%)
TONa kinitialb
C4 OA
C4 OA
C4 OA
C4 OA
C4 HAD
C4 OA
C6 OA
C8 OA
C10 OA
30
50
75
100
75
75
75
75
75
38/1
38/1
38/1
38/1
38/1
38/1
38/1
38/1
38/1
46/1
46/1
46/1
46/1
52/1
46/1
49/1
49/1
52/1
0.82
0.82
0.82
0.82
0.73
0.82
0.77
0.77
0.73
4
9
17
17
14
17
16
11
14
2
4.5
8
8
6.5
8
8
5.5
7
12 1
12 1
81
51
71
Reaction conditions: P(H2) 1.2 bars; reaction time 5 h; RuT total amount of Ru atoms;
RuS amount of surface Ru atoms. aTurn over number (moles of product converted per mol
of RuS).
b
Estimated initial rates in 102 mol L1 h1.
the viscosity changes in the reaction mixtures and/or by a higher steric hindrance of HDA compared to that of OA once coordinated on the RuNP surface.
These systems were also tested in the hydrogenation of 1,3-cyclohexadiene
(CYD), styrene (STY) and R-()-limonene (LIM). The results evidenced that
for the hydrogenation of 1,3-CYD, STY and LIM, the activity of the RuNPs
increased with s-donor ligands such as C8H17NH2 and H2O, but decreased
with the bulkier and p-acceptor ligands, PPhH2 PPh2H and CO. This underlined the pseudo-molecular nature of NPs with sizes o3 nm since their activity can be tuned by the s-p-ligand character as in homogeneous catalysis.
In summary, the confinement properties of ionic liquids based on imidazolium cations help to control the growth of RuNPs, and their catalytic
performance can be further improved by the addition of a stabilizing ligand.
4.3.2
Among the hydrogen storage materials, amine-boranes are interesting candidates due to their high hydrogen content (19.6 wt %).54 Recent studies have
shown that the catalytic dehydrogenation of dimethylamine-borane
(Me2NHBH3; DMAB) can be achieved under mild conditions. In a collaborative
zkar, we studied the catalytic performances of 3-amiwork with the team of O
nopropyltriethoxysilane-stabilized Ru nanoclusters (Ru/APTS) with dierent
sizes in this reaction (see Figure 4.8).55,56 The mean particle size of Ru/APTS
NPs decreased with increasing APTS ligand concentration (see Table 4.4).
After the addition of DMAB into a THF solution of Ru/APTS 3, the
hydrogen evolution started immediately with an initial turnover frequency
(TOF) of 53 h1 and continued until 1 equiv. H2 per mol DMAB was released.
In order to dierentiate between homogeneous and heterogeneous catalysis,
Hg(0), a poison for heterogeneous metal(0) catalysts, was added into the
reaction mixture. The suppression of the catalysis by Hg(0) was evidence of
heterogeneous catalysis (see Figure 4.9). The initial TOF value of 53 h1
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
Figure 4.8
(a) Synthesis of Ru/APTS NPs, (b) TEM and HRTEM images of Ru/APTS
NPs (B1.7 nm) immediately after the synthesis.55
Table 4.4
Nanocatalyst
[APTS][Ru]
Ru/APTS
Ru/APTS
Ru/APTS
Ru/APTS
Ru/APTS
0.25
0.50
1.0
2.0
4.0
2.36 0.61
1.96 0.44
1.67 0.41
1.29 0.31
1.16 0.30
61
1
2
3
4
5
Figure 4.9
View Online
Chapter 4
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
62
Figure 4.10
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
63
obtained with this system was comparable to that of the prior best heterogeneous catalyst of rhodium nanoclusters (TOF 60 h1). Moreover, it was
the first example of an isolable, bottleable and reusable transition metal
nanocatalyst for the dehydrogenation of dimethylamine-borane.
An increase of APTS concentration caused a significant decrease in the
catalytic activity as a result of a higher coverage of metallic surface (see
Figure 4.10 (a)). The maximum catalytic activity was achieved with Ru/APTS
3 NPs, which appeared to have the best compromise between the NP mean
size and the surface accessibility. The influence of temperature on the
hydrogen generation rate using Ru/APTS 3 NPs gave high activity even at low
temperature (see Figure 4.10 (b)).
4.3.3
4.3.3.1
mez, Castillo
n and Claver et al., chiral
In a collaborative project with Go
diphosphites were used as ligands for the synthesis of PdNPs and corresponding molecular complexes for a comparative study in catalytic allylic
alkylation (see Figure 4.11).65,66
PdNPs stabilized by the chiral xylofuranoside diphosphite ligand, L1,
were achieved by decomposition of [Pd2(dba)3] (THF; P(H2) 3 bar; r.t.)
and were investigated in the allylic alkylation of rac-3-acetoxy-1,3-diphenyl1-propene (rac-I) with dimethyl malonate (see Table 4.5). With the Pd/L1
nanocatalyst, this reaction proceeded with only one enantiomer of
the racemic substrate (495% ee), while the corresponding molecular species stabilized with the same ligand yielded both enantiomers. These
results demonstrated a very high degree of kinetic resolution with
Pd/L1 NPs.
Modifications in the diphosphite ligands changed the selectivity of the
reaction, drastically. For instance, Pd/L2 NPs (see Figure 4.11) did not
induce any selectivity under the same conditions. When the carbohydrate
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
64
Figure 4.11
Table 4.5
Chapter 4
Catalyst
IPdL1
Moleculara
Colloidalb
Colloidalb
Colloidalb
Colloidalb
Colloidalb
100
100
100
100
100
100
a
b
:
:
:
:
:
:
1
1
1
1
1
1
:
:
:
:
:
:
1.25
0
0.2
1.05
0.2
1.05
t(h)
Conv. (%)
ee II (%)
ee rac-I (%)
1.5
24
24
24
168
168
83
2
56
56
59
61
90 (S)
nd
97 (S)
97 (S)
97 (S)
97 (S)
0
0
89
89
89
89
(S)
(S)
(S)
(S)
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
65
4.3.3.2
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:1
[Pd complex] or
Pd MNPs, tBuOK
B(OH) 2
66
Table 4.6
DMF/H 2O (4:1)
X= I, Br
+
BT
PhMe
TT
Catalyst [PhB(OH)]/[Pd]
Pd/L2
315
Pd/L3 315
Pd/C
315
Pd/L3d
3125
Pd/L2
Pd/L3
Pd/C
XI
Convb (%)
BTc
TTc
X Br
Convb (%)
BTc
TTc
PhMec
100
69
31
69
12
14
86
100
31
61
65
39
35
61
20
37
84
17
53
46
47
37
20
39
100
90
10
90
100
100
18
24
76
100
100
64
70
30
45
100
99
83
41
59
34
PhMec
Reaction conditions: 1102 mmol Pd, 2.5 mmol 4-halogenotoluene, 5.0 mmol tBuOK, 0.5 mmol naphthalene as internal standard, 8.0 ml DMF2.0 mL
H2O, 100 1C.
b
Conversion after 6 h.
c
Product distribution after 6 h.
d
4-Halogenotoluene tBuOKPd 2500 : 5000 : 1.
e
Yield of 4-methylbiphenyl (BT).
f
Yield of 4,4 0 -dimethylbiphenyl (TT).
a
Chapter 4
View Online
67
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
Figure 4.12
4.3.4
Hydroformylation Reactions
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:1
68
Synthesis of RhNPs from [Rh(Z3-C3H5)3] and [Rh(m-OMe)(COD)]2 in the presence of chiral diphosphite ligands 1 and 2.68
Chapter 4
Figure 4.13
View Online
Table 4.7
69
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
H2/CO
CHO
+
cat.
Catalyst
RhLSb
t(h)
% Conversionc
% Regioselectivityd
% ee
Rh3
Rh3
Rh4
Rh4
1
1
1
1
24
24
24
24
28
11
95
80
57
499
54
90
0
40 (S)
13 (S)
24 (S)
:
:
:
:
: 200
0.2 : 200
: 200
0.2 : 200
Reaction conditions: 3 mg RhNPs, 80 1C, p 20 bar pCOH2 1, RhS 1 : 200 styrene 5.8 mmol,
10 ml toluene.
b
Molar ratio between rhodium, excess ligand added to the catalysis and substrate.
c
% Conversion styrene G.C.
d
% 2-Phenylpropanal.
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
70
Chapter 4
4.4.1
In collaboration with Schmid et al., the filling of nanoporous alumina membranes of various pore widths has been carried out from [Ru(COD)(COT)]
decomposed in THFMeOH mixtures without any stabilizer in two dierent
ways.70 The first approach involved the impregnation of an alumina support
with colloidal solutions of RuNPs in dierent sizes, which were dependent on
the ratio of MeOHTHF in the reaction mixture. Colloidal solutions were
transferred into membranes by vacuum induction. Very few agglomerates
were observed outside the pores whereas there were dense areas within
the membrane channels. The second approach consisted of decomposition
Figure 4.14
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
71
4.4.2
Figure 4.15
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
72
Chapter 4
4.4.2.2
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:1
Figure 4.16
Synthesis and TEM images of Ru and RuO2NPs inside the pores of a phosphonate-functionalized mesoporous silica.73
73
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
74
Chapter 4
Figure 4.17
(a) Schematic representation of the gas sensors, (b) sensitivity variations of the gas sensor under dierent gas compositions with (black
line) or without (grey line) an on chip filter layer, and (c) the role of
filter for CO.73,74
Figure 4.18
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
Figure 4.19
4.4.3
75
Today, carbon materials are used often for the immobilization of MNPs
since they oer several advantages as solid supports. These include easy
availability, relatively low cost, high mechanical strength and chemical stability. In addition, their porous structure makes them attractive for surface
chemistry as it enables easy modifications, such as its functionalization in
order to suit the immobilization needs of the NPs.
4.4.3.1
Hydrogenation of Cinnamaldehyde
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
76
Chapter 4
Figure 4.20
NPs through stabilization by the ligand first and then their impregnation on
the CNTs. An alternative route was the co-decomposition of the two organometallic precursors in the presence of both the ligand L and the CNTs.
The best results in terms of confinement of the particles were obtained by
the impregnation method employing amide-functionalized CNTs. The resulting hybrid material had B22.5 nm NPs which were located inside the
CNTs (80% of NPs for a 23 wt% of metal). All the prepared systems, in
addition to the non-supported RuPt/L NPs, were tested as catalysts in the
hydrogenation of cinnamaldehyde (see Table 4.8). The catalytic activity and the
selectivity of CNT-supported RuPt/L NPs were higher than those of the nonsupported ones. In the case of the NPs mainly located inside the CNTs, a remarkable selectivity towards the formation of cinnamyl alcohol was achieved.
4.4.3.2
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
Table 4.8
Catalytic hydrogenation
nanomaterials.a,76
of
cinnamaldehyde
77
with
RuPt/CNT
Catalyst
inside/outside
% NP int.
TOF (h1)
HCAL
HCOL
COL
PtRu/L NPs
PtRu/CNT2
PtRu/CNT1
PtRu/CNT3
2.2
2.2/2.2
1.6/2.2
2/2.5
10
30
80
30
56
75
85
50
33
18
0
15
8
12
5
35
59
69
95
Figure 4.21
are well dispersed on the carbon materials and display mean sizes of 1.21.3
nm and 1.92.2 nm, respectively. These hybrid systems were successfully
used as catalysts in the oxidation of benzyl alcohol in water at 80 1C, giving
rise to excellent conversion and selectivity (499%) towards the aldehyde
for both metals (see Figure 4.21). The recyclability studies showed that carbon-supported PdNPs can be used in up to eight successive reaction cycles
without significant loss in activity. The catalytic activity of these systems was
View Online
78
Chapter 4
Table 4.9
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
Cat.
Cat.
-CO
(decarbonylation)
Nanocatalyst
Reaction
time (min)
Conversion
(mol%)
Selectivity in
furan (mol%)
Selectivity in
THF (mol%)
No catalyst
MB-H2O2
MB-1500
Pd(TOP)/MB-H2O2
Pd(TOP)/MB-1500
Pd(TPP)/MB-H2O2
Pd(TPP)/MB-1500
Pd/MB-H2O2
Pd/MB-1500
60
60
60
30
30
30
30
45
45
490
75
69
65
70
45
15
20
10
o10
30
35
80
72
85
80
70
65
H2
hydrogenation
Reaction conditions: 2 mmol furfural, 1.5 mL formic acid, 1.5 mL water, 0.1 g catalyst, 100 1C,
microwave power 100150 W, P 150200 psi (developed in the systems).
4.4.3.3
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
79
References
1. D. Astruc, F. Lu and J. Ruiz Aranzaes, Angew. Chem., Int. Ed., 2005, 44,
78527872.
2. Nanoparticles and Catalysis, ed. D. Astruc, Wiley-VCH, Weinheim, 2008.
3. J. M. Thomas, Chem. Cat. Chem., 2010, 2, 127132.
4. Z. X. Li, W. Xue, B. T. Guan, F. B. Shi, Z. J. Shi, H. Jiang and C. H. Yan,
Nanoscale, 2013, 5, 12131220.
5. Nanoparticles: From Theory to Application, ed. G. Schmid, Wiley-VCH,
Weinheim, 2012.
6. B. R. Cuenya, Thin Solid Films, 2010, 518, 31273150.
m and M. Ba
umer, Phys. Chem. Chem. Phys., 2011, 13,
7. P. Sonstro
1927019284.
8. T. J. Geldbach and P. J. Dyson, Metal-Catalysed Reactions in Ionic Liquids,
Springer, The Netherlands, 2005.
9. N. Yan, C. Xiao and Y. Kou, Coord. Chem. Rev., 2010, 254, 11791218.
10. B. L. Cushing, V. L. Kolesnichenko and C. J. OConnor, Chem. Rev., 2004,
104, 38933946.
11. N. Semagina and L. Kiwi-Minsker, Catal. Rev.: Sci. Eng., 2009, 51,
147217.
12. Nanomaterials in Catalysis, ed. K. Philippot and P. Serp, Wiley-VCH,
Weinheim, 2013.
13. D. Uzio and G. Berhault, Catal. Rev.: Sci. Eng., 2010, 52, 106131.
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
80
Chapter 4
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
81
View Online
23/06/2014 08:14:50.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00047
82
Chapter 4
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
CHAPTER 5
5.1 Introduction
The catalytic transfer hydrogenation of organic compounds1,2 is an advantageous methodology with respect to other reduction methods for several
reasons: (a) the hydrogen source is easy to handle (no gas containment or
pressure vessels are necessary), (b) possible hazards are minimized, (c) the
mild reaction conditions used can aord enhanced selectivity and, (d)
catalytic asymmetric transfer hydrogenation can be applied in the presence
of chiral ligands. Among the dierent types of hydrogen donors, 2-propanol
is probably the most popular since it is cheap, non-toxic, volatile, possesses
good solvent properties, and it is transformed into acetone, which is environmentally friendly and easy to remove from the reaction medium. Some
other sources of hydrogen, such as formic acid derivatives or hydrazine, are
in less common use.
Transition metal catalysis is dominated by the elements of the three triads,
where ruthenium, rhodium, iridium and, above all, palladium, generally
stand out from the others because of their incomparable catalytic activity and
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
83
View Online
84
Chapter 5
70
62.3
60
/mmol
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
50
43.0
40
34.4
30
20
12.5
8.2
10
0
Rh
Figure 5.1
Au
Ir
Pd
Ru
0.086
0.007
Ni
Fe
eciency. These noble metals have obscured the role of the first triad metals
in catalysis until the recent revival of iron.3 Indeed, methodologies based on
modest transition metals that can promote organic transformations typically reserved for noble metals are challenging and praiseworthy. This assertion gains points when we compare the relative prices of some transition
metals compared to the corresponding anhydrous chlorides. Thus, rhodium,
gold, and iridium are extraordinarily expensive; palladium and ruthenium
are somewhat cheaper, whereas nickel and iron are about 100-fold and 1000fold cheaper than ruthenium, respectively (Figure 5.1).
It is noteworthy that the combination of inexpensive transition metals
with green and safe sources of hydrogen has been practically unexplored
until recently. In this sense, the potential of nickel has been undervalued in
favor of more exotic and attractive catalytic systems containing noble
metals. On the other hand, in recent years, nano-catalysis has emerged as a
sustainable and competitive alternative to conventional catalysis since the
metal nanoparticles possess a high surface-to-volume ratio, which enhances
their activity and selectivity, while at the same time maintaining the intrinsic
features of a heterogeneous catalyst.47
This chapter summarizes the application of catalytic systems based on
nickel nanoparticles as an alternative to noble-metal catalysts in the
hydrogen-transfer reduction of functional groups. In particular, we will deal
with the transfer hydrogenation of olefins, carbonyl compounds, and the
reductive amination of aldehydes.
5.2 Antecedents
In 1996,8,9 due to our incipient interest in active metals,10,11 we discovered
the NiCl2 2H2O-Li-arene(cat.) combination as a useful and versatile mixture
View Online
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
H
C C
Y
R
C C
C C
[Y = CH, N]
Scheme 5.1
CH NH
C O
C N
N O
RH
Ar
R'
R
Ar
R
NH
N NH
N N Ar
Ar
RHal
ROSO2
Y
H
CH OH
85
ArNH2
Ar N N Ar
Ar N N Ar
or ArNH2
View Online
86
Chapter 5
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
18
View Online
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
C C
20 mol% NiNPs
>99%, 3 h
H C C H
i-PrOH, 76 C
Ph
>99%, 2 h
Ph
60%, 1 h
87
>99%, 2 h
Ph
Ph
>99%, 1 h
Ph
>99%, 1 h
Ph
>99%, 4 h
CO2 Et
89%, 2 h
>99%, 2 h
>99%, 72 h
>99%, 5 h
MeO
CO 2Et
>99%, 2 h
HO
MeO
>99%, 3 h
OMe
>99%, 48 h
OH
OH
Ph
O
>99%, 4 h
96%, 2 h
90%, 3 h
OH
H
N
Ph
O
>99%, 3 h
>99%, 2 h
51%, 72 h
Scheme 5.2
View Online
88
Chapter 5
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
C C
H C C H
MeO
82%
68%
74%
88%
O
MeO
79%
52%
91%
CO2 Me
81%
74%
Scheme 5.3
48%
S
79%
View Online
OH
i-PrOH, 76 C
Ar
OH
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
89
R2
OH
OH
OH
Ph
82%, 1 h
(80%, 1 h)
OH
72%, 1 h
(40%, 3 h)
Scheme 5.4
F3 C
MeO
57%, 1 h
(73%, 3 h)
OMe
34%, 24 h
(33%, 3 h)
73%, 1 h
(72%, 4 h)
OH
OH
OH
51%, 1 h
(52%, 5 h)
Ph
73%, 1 h
(52%, 4 h)
Ph
Ph
86%, 1 h
(63%, 3 h)
Ph
OH
42%, 2 h
(30%, 3 h)
achieved with 20 mol% NiNPs being somewhat lower than that achieved with
stoichiometric NiNPs. The NiNPs exhibited a superior performance in
comparison with other forms of nickel under the same reaction conditions.
Moreover, the NiNPs could be easily separated by decantation with an extra
amount of 2-propanol and reutilized up to four times (with 1 equiv. NiNPs, 1
h, 9495%) or five times (20 mol% NiNPs, 1 h, 8777%) in a very simple
reaction medium composed of the NiNPs, 2-propanol and the substrate,
with no base. Deactivation of the catalyst with further repeated reuse was
ascribed to surface oxidation.41 According to some deuteration experiments,
the reaction seemed to proceed through a dihydride-type mechanism, in
which the two hydrogens of the donor are transferred to the metal, thus
losing their original identity and becoming equivalent (Scheme 5.6). After
our work, several reports emerged describing the application of supported
NiNPs, prepared by dierent methods, to the transfer hydrogenation of
carbonyl compounds with isopropanol (see below).
Laboratory-made magnetic nano-ferrites were surface-modified with dopamine followed by the addition of NiCl2 at basic pH and reduction with NaBH4.
The resulting Ni-coated nano-ferrite (1013 nm) was used in the transfer hydrogenation of aromatic ketones in the presence of KOH and isopropanol at
100 1C under microwave irradiation (45 min (Scheme 5.7).42 Although high
yields were generally attained, brominated acetophenones (either on the ring or
a position) underwent hydrodebromination whereas nitroacetophenones were
transformed into the corresponding anilines. The catalyst could be recovered
magnetically and reused, after simple treatment (washing with MeOH and
drying at 60 1C for 30 min), five times with no apparent loss of activity (9895%).
View Online
90
Chapter 5
O
1
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
R1
i-PrOH, 76 C
OH
4
OH
R2
OH
OH
OH
73%, 1 h
(73%, 4 h)
92%, 1 h
(76%, 4 h)
53%, 48 h
83%, 1 h
(90%, 3 h)
OH
HO
OH
58%, 1h, exo/endo 85:15
(81%, 4 h, exo/endo 80:20)
H
64%, 1 h, t rans/cis 90:10
(78%, 3 h, t rans/cis 66:34)
OH
OH
H
H
H
H
HO
HO
84%, 1 h, d.r. 1:1
79%, 2 h
(34%, 1 h)
OH
n-C 9H 19
OH
39%, 1 h, t rans/cis 8:92
Scheme 5.5
62%, 1 ha
O
Ph
OH
O
- acetone
O
Ph
Ph
OH
+
OH
NiNPs, THF, 76 C
(CH3)2CDOH
Ph
45% D
Ph
H
50%
50%
[H-M-H]
Scheme 5.6
OH
40%, 1 h
View Online
OH
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
OH
7 mol% Ni/nano-Fe2 O3
2
91
R2
Ar
OH
OH
OH
Ph
98%
Cl
Br
NH2 OH
NH2
95%
98%
NH2 OH
80%a
O
OH
Br
Ph
NH2
80% a
Scheme 5.7
Br
5%
98%
5%
Transfer hydrogenation of aromatic ketones catalyzed by nanoferriteNi in 2-propanol. aStarting from the nitrocompound.
O
OH
0.05 mol% NiNPs/SiO2
NaOH, i-PrOH, 100 C, 1 h
93%
Scheme 5.8
The group of Song prepared NiNPs by thermal decomposition of nickeloleylamine complexes, which were successfully coated with silica using the
microemulsion method. The nickel cores inside silica shells formed nickel
phyllosilicate with an urchin-like morphology under hydrothermal conditions, and NiNPs (24 1 nm) were regenerated on the silica spheres by
hydrogen reduction.43 The resulting morphology of the Ni/SiO2 nanostructure can be regarded as active NiNPs embedded in the silica support.
Acetophenone was converted into 1-phenylethanol in a high yield (93%),
using only 0.05 mol% of the catalyst and NaOH in isopropanol at 100 1C for
1 h (Scheme 5.8). The catalyst could be easily separated from the reaction
medium by centrifugation and reused five times without any loss of catalytic
activity (9792%).
The same group reported the synthesis of Ni@SiO2 yolkshell nanocatalysts with nickel cores of ca. 3 nm. The NiNPs were coated with silica through
the microemulsion method, with the resulting Ni@SiO2 yolkshell NPs
being subjected to partial etching by acid treatment. Uniform yolkshell
nanocatalysts were obtained after calcination of the NPs.44 The transfer
hydrogenation of cyclohexanone and a series of aromatic ketones was effected with low catalyst loading (0.03 mol%) and NaOH (20 mol%) at 150 1C
for 30 min (Scheme 5.9). Although the yields and TOF (6000 h1) were high,
the reaction conditions are rather harsh and might be incompatible with the
View Online
92
Chapter 5
O
R1
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
OH
OH
R1
Ph
93%
OH
Ph
83%
94%
80%
OH
86%
OH
OH
MeO
98%
87%
R2
OH
OH
Ph
Scheme 5.9
OH
OH
Cl
75%
94%
OH
1 mol% NiNPs/montmorillonite
NaOH, i-PrOH, 76 C, 4 h
98%
Scheme 5.10
presence of other functional groups. Nonetheless, the catalyst could be reused for six cycles maintaining an excellent performance (9290%).
The impregnation of Ni(OAc)2 into the nanopores of two acid-modified
montmorillonite supports, followed by polyol reduction (with ethylene glycol) provided the corresponding catalysts of NiNPs (o18 nm) diering in the
time of activation of the support.45 The catalyst with the longer activation
time (2 h) reached a higher yield and selectivity in the hydrogen-transfer
reduction of acetophenone, using low catalyst loading (1 mol% Ni) and
NaOH in isopropanol at 76 1C under nitrogen (Scheme 5.10). The higher
catalytic activity of the latter was related to the wider specific pore volume of
the catalyst matrix. The presence of base and absence of oxygen were
mandatory for the reaction to take place. The catalyst could be recovered by
filtration and reused (after washing with water and drying) in three runs with
a decline in the conversion (9864%).
More recently, a highly ecient transfer hydrogenation catalyst was prepared by incorporation of NiNPs into a mesoporous aluminosilicate framework.46 The latter was obtained by the solgel method, followed by calcination
at 800 1C. Reduction of the impregnated support with hydrazine gave rise to
the nickel aluminosilicate nanocomposite. The catalyst was fully characterized, and the powder XRD pattern showed the presence of nickel oxide with a
fcc lattice and a crystallite size of ca. 5 nm. The TPR profile indicated that the
major portion of the original NiNPs is on the surface of the framework and
View Online
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
OH
Ph
OH
85%
OH
O2N
13%
OH
OH
O2N
N
Cl
83%
41%
MeO
OH
OH
OH
N
82%
Cl
Cl
OH
58%
OH
OH
OH
OH
Ph
Ph
>99%
Cl
84%
83%
Ph
94%
MeO
71%
71%
OH
S
Scheme 5.11
NO 2
98%
88%
99%
R2
Br
OH
OH
OH
Cl
98%
H2N
OH
93
OH
46%
N
96%
OH
48%
was oxidized to NiO in the atmosphere. The catalyst exhibited a high performance in the transfer hydrogenation of aldehydes and ketones with 0.026
mol% Ni, KOH (2 equiv.) at 90 1C for 30 min (aldehydes) or 3 h (ketones)
(Scheme 5.11). GC yields 482% were generally recorded for aromatic and
heteroaromatic aldehydes as well as for ketones, though cyclohexanone was
shown to be more reluctant to react. The process was highly chemoselective,
with the reaction conditions being compatible with the presence of halogens
(Cl and Br), nitro, methoxy, or amino groups. Furthermore, the catalyst was
recyclable for four cycles with no significant loss in activity (9492%).
Kidwai et al. prepared NiNPs by reduction of Ni(II) to Ni(0) in a reverse
micellar system containing an aqueous solution of Ni(NO3)2 and an alkaline
solution of NaBH4. The NiNPs (1085 nm) were eciently applied to the
transfer hydrogenation of aldehydes47 and ketones,48 using ammonium
formate as a hydrogen source. The maximum reaction rate was recorded for
an average particle diameter of about 20 nm. All reactions proceeded in
relatively short reaction times at room temperature (for aldehydes,
Scheme 5.12) or 4550 1C (for ketones, Scheme 5.13) in THF under nitrogen
and were high yielding. Moreover, the NiNPs could be recycled, with a
View Online
94
Chapter 5
O
1
10 mol% NiNPs
H
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
OH
85%
NC
OH
O2 N
OH
HO
OH
MeO
93%
72%
MeO
93%
OH
OH
OH
MeO
92%
90%
89%
OH
78%
Scheme 5.12
OH
Br
97%
NO2
OH
OH Br
OH
Ph
R1
OH
O
60%
S
88%
OMe
93%
O
R1
OH
R1
R2
OH
OH
Ph
93%
Br
OH
OH
OH
Ph
Ph
78%
OH
Cl
Br
84%
91%
OH
OH
O
93%
Scheme 5.13
90%
93%
84%
decrease in the catalytic activity of up to 76% after the sixth cycle, which was
attributed to partial particle agglomeration.
View Online
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
95
contrast with other transition metals, nickel has been scarcely used in reductive aminations or in the transfer hydrogenation of imines. Reductive
aminations involving nickel were performed with hydrazine and borohydride exchange resin-nickel acetate56 or with stoichiometric nickel boride in
methanol.57 On the other hand, the transfer hydrogenation of pre-formed
imines was carried out with 2-propanol and an excess of aluminium isopropoxide in the presence of Raney nickel (the reaction failed in the absence
of aluminium isopropoxide),58 or with sodium isopropoxide catalyzed by a
nickel(0)N-heterocyclic carbene complex.59
To the best of our knowledge, however, nickel-mediated reductiveamination reactions by transfer hydrogenation had not been reported
previous to our work. We proved that NiNPs can catalyze the reductive
amination of aldehydes by transfer hydrogenation, using 2-propanol as the
hydrogen source in the absence of any added base (Scheme 5.14).60 The
20 mol% NiNPs
R 1CHO + R2 NH 2
R1
i-PrOH, 76 C
R2
N
H
OMe
Ph
Ph
N
H
73%, 9 h
Ph
Ph
N 7
H
90%, 10 h
N
H
65%, 10 h
Ph
N
H
Ph
N
H
43%, 24 h
Ph
N
H
97%, 4 h
N
H
99%, 12 h
Ph
N
H
N
H
44%, 6 h
MeO
98%, 8 h
Ph
H
N
O
Ph
30%, 48 h
30%, 48 h
99%, 12 h
H
N
Ph
80%, 24 h
MeO
36%, 6 h
Scheme 5.14
N
H
N
H
N
H
MeO
N
Ph
H
92%, 12 h
Ph
MeO
N
H
Ph
iBu
87%, 10 h
Ph
40%, 24 h
N
OMe
H
80%, 8 h
N
Ph
H
77%, 8 h
85%, 6 h
N
H
Ph
Ph
N
H
Ph
70%, 12 h
View Online
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
96
Chapter 5
process was especially eective in the reductive amination of aromatic aldehydes, most of the corresponding secondary amines being obtained in
good-to-excellent yields. The application to aliphatic aldehydes was, however, more limited. Although any attempt to reuse the NiNPs failed, this
methodology can be considered advantageous because (a) the preparation
step of the imine is avoided, aldehydes and amines being used directly as
starting materials, (b) the source of hydrogen is 2-propanol, a cheap and
environmentally friendly solvent, and (c) the NiNPs were shown to be superior to other nickel catalysts in this reaction.
5.6 Conclusions
In view of the above results, we can state that NiNPs are a real alternative to
noble-metal based catalysts in hydrogen-transfer reactions. Simple catalytic
systems, such as that composed of NiNPs and 2-propanol, exhibited a high
performance in the transfer hydrogenation of olefins, carbonyl compounds,
and reductive amination of aldehydes. This represents the first transitionmetal catalyzed transfer hydrogenation in 2-propanol performed in the absence of any base or acid. Other unsupported NiNPs also succeeded in the
transfer hydrogenation of carbonyl compounds with ammonium formate at
room temperature. However, NiNPs immobilized on inorganic supports are
more stable and allow the reactions to proceed with very low metal loadings,
albeit the presence of a base is mandatory. It must be underlined that, due to
the heterogeneous nature of all the catalytic systems shown herein, they can
be easily recovered and reused in 36 cycles maintaining high catalytic activities and displaying a reactivity superior to other nickel catalysts, including Raney nickel. The results presented herein open an array of
possibilities for further research in reactions where the more expensive, and
sometimes sophisticated, noble-metal catalysts might be replaced by the
cheaper and, generally, easy-to-prepare NiNP-based catalysts.
References
1. R. A. W. Johnstone and A. H. Wilby, Chem. Rev., 1985, 85, 129170.
2. R. M. Kellogg, Reduction of CX to CHXH by hydride delivery from
carbon in Comprehensive Organic Synthesis, ed. B. M. Trost and I. Fleming,
Pergamon, Oxford, 1st edn, 1991, vol. 8, ch. 1.3, pp. 79106.
3. Iron Catalysis in Organic Chemistry. Reactions and Applications, ed.
B. Plietker, Wiley-VCH, Weinheim, 2008.
4. Nanoparticles and Catalysis, ed. D. Astruc, Wiley-VCH, Weinheim, 2008.
5. M. Kidwai, Nanoparticles in Green Catalysis in Handbook of Green
Chemistry, ed. P. T. Anastas and R. H. Crabtree, Wiley-VCH, Weinheim,
2009, vol. 2, ch. 4, pp. 8192.
6. N. Yan, C. Xiao and Y. Kou, Coord. Chem. Rev., 2010, 254, 1791218.
7. V. Polshettiwar and R. S. Varma, Green Chem., 2010, 12, 743754.
8. F. Alonso and M. Yus, Tetrahedron Lett., 1996, 37, 69256928.
View Online
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.
38.
39.
97
View Online
23/06/2014 08:14:51.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00083
98
Chapter 5
40. F. Alonso, P. Riente and M. Yus, Tetrahedron Lett., 2008, 49, 19391942.
41. F. Alonso, P. Riente, J. A. Sirvent and M. Yus, Appl. Catal., A: Gen., 2010,
378, 4251.
42. V. Polshettiwar, B. Baruwati and R. S. Varma, Green Chem., 2009, 11,
127131.
43. J. C. Park, H. J. Lee, J. U. Bang, K. H. Park and H. Song, Chem. Commun.,
2009, 73457347.
44. J. C. Park, H. J. Lee, H. Y. Kim, K. H. Park and H. Song, J. Phys. Chem. C,
2010, 114, 63816388.
45. D. Dutta, B. J. Borah, L. Saikia, M. G. Pathak, P. Sengupta and
D. K. Dutta, Appl. Clay Sci., 2011, 53, 650656.
46. N. Neelakandeswari, G. Sangami, P. Emayavaramban, S. G. Babu,
R. Karvembu and N. Dharmaraj, J. Mol. Catal. A: Chem., 2012, 356, 9099.
47. M. Kidwai, V. Bansal, A. Saxena, R. Shankar and S. Mozumdar, Tetrahedron Lett., 2006, 47, 41614165.
48. M. Kidwai, N. Kumar Mishra, V. Bansal, A. Kumar and S. Mozumdar,
Catal. Commun., 2008, 9, 612617.
49. V. A. Tarasevich and N. G. Kozlov, Russ. Chem. Rev., 1999, 68, 5572.
50. S. Gomez, J. A. Peters and T. Maschmeyer, Adv. Synth. Catal., 2002, 344,
10371057.
51. E. W. Baxter and A. B. Reitz, Org. React., 2002, 59, 1714.
52. A. F. Abdel Magid and S. J. Mehrman, Org. Process Res. Dev., 2006, 10,
9711031.
53. V. I. Tararov, R. Kadyrov, T. H. Riermeier, U. Dingerdissen and
A. Boerner, Org. Prep. Proced. Int., 2004, 36, 99120.
54. V. I. Tararov and A. Boerner, Synlett, 2005, 203211.
55. P. Roszkowski and Z. Czarnocki, Mini-Rev. Org. Chem., 2007, 4, 190200.
56. J. H. Nah, S. Y. Kim and N. M. Yoon, Bull. Korean Chem. Soc., 1998, 19,
269270.
57. I. Saxena, R. Borah and J. C. Sarma, J. Chem. Soc., Perkin Trans., 2000, 1,
503504.
58. M. Botta, F. De Angelis, A. Gambacorta, L. Labbiento and R. Nicoletti,
J. Org. Chem., 1985, 50, 19161919.
59. S. Khul, R. Schneider and Y. Fort, Organometallics, 2003, 22, 41844186.
60. F. Alonso, P. Riente and M. Yus, Synlett, 2008, 12891292.
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
CHAPTER 6
Ammonium Surfactant-capped
Rh(0) Nanoparticles for
Biphasic Hydrogenation
AUDREY DENICOURT-NOWICKI*a,b AND ALAIN ROUCOUX*a,b
a
6.1 Introduction
In the drive towards the development of environmentally-friendly and economically viable processes, nanometre-sized metallic species, finely dispersed in water, have appeared as an unavoidable family of active and
reusable catalysts under biphasic conditions.1 Thus, soluble metal nanoparticles have emerged as sustainable alternatives to conventional molecular
catalysts, potentially being easily recycled as heterogeneous systems.24 These
nanocatalysts have proved to be ecient and selective for various reactions,
owing to a great number of potential active sites and an original surface
reactivity.57 In particular, they have showed great promise in arene hydrogenation reactions,8,9 presenting high activities in mild conditions, especially
with Rh(0) nanoparticles. Recently, the preparation of metal particles in
green solvents, such as water, ionic liquids, fluorous medium or supercritical
fluids, has received great attention.10 Among them, water has appeared as a
relevant alternative to conventional organic media for economic and environmental reasons,11,12 as well as the easy catalyst recovery through a
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
99
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
100
Chapter 6
biphasic approach thanks to its absence of miscibility with the usual organic
phases.13,14 Since naked nanoparticles tend to aggregate, protective agents
are usually used and chosen according to the reaction media. Besides the use
of ligands (polymers, cyclodextrins, dendrimers, etc.) as stabilizers,1,15 surfactants, such as ammonium salts, are widely known as ecient protective
agents for electrosteric stabilization of particles within the aqueous phase,
thanks to their amphiphilic character. Generally, nanospecies with wellcontrolled size were obtained, providing pertinent reactivity in catalysis.
In this chapter, we will mainly focus on the use of Rh(0) nanoparticles in
hydrogenation reactions in pure biphasic liquidliquid conditions, namely
when a liquid substrate constitutes the only organic phase (no solvent
added).
Figure 6.1
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
101
(CH3SO3),
Table 6.1
Entry
Surfactant
CMCa
(mmol L1)
1
2
3
4
5
6
7
HEA16Cl
HEA16F
HEA16HCO3
HEA16BF4
HEA16CF3SO3
HPA16Cl
THEA16Cl
1.23
2.12
0.6
0.62e
0.25
2.00
0.63
gCMCb
(mN m1)
Ac (2)
CPPd
Geometry
40.5
43.2
40.3
30.3
32.0
36.9
33.5
33.4
87.1
46.6
37.3
49.6
91
128
0.63
0.24
0.45
0.55
0.42
0.23
0.16
Bilayers
Spheres
Rods
Rods
Rods
Spheres
Spheres
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
102
Chapter 6
Figure 6.2
(a) Rh(0)@THEA16Cl
Figure 6.3
(b) Rh(0)@HEA16HCO3
(c) Rh(0)@HEA16BF4
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
103
Entry
Surfactant
t (h)
TOF
1
2
3
4
5
6
7
8
9
10
HEA18Br
HEA16Br
HEA16Cl
HEA16F
HEA16BF4
HEA16HCO3
HEA16CH3SO3
HEA16CF3SO3
HPA16Cl
THEA16Cl
9.1
5.3
3.6
6
3.7
6
3.7
3
5
1
33
57
83
50
81
50
81
100
60
300
(h1)
View Online
104
Chapter 6
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
Table 6.3
Entry Substrate
t (h)
TOF c
(h1)
1
2
Toluene
Anisole
3.6
4
83
75
3
4
5
6
7
8
9
10
11
12
13
Ethyl benzoate
Aniline
Furan
Pyridine
Quinoline
1,3,5-Triazine
o-Xylene
m-Xylene
p-Xylene
Chorobenzene
2-Chloroanisole
Methylcyclohexane (100)
Methoxycyclohexane (70)
cyclohexanone (30)
Ethylcyclohexanoate (100)
Cyclohexylamine (100)
1,2,3,4-Tetrahydrofuran (100)
Piperidine (100)
1,2,3,4-Tetrahydroquinoline (100)
1,3,5-Hexahydrotriazine (100)
1,2-Dimethylcyclohexane (cistrans: 91 : 9)
1,3-Dimethylcyclohexane (cistrans: 80 : 20)
1,4-Dimethylcyclohexane (cistrans: 65 : 35)
Cyclohexane (100)
Methoxycyclohexane (53)
cyclohexanone (47)
4.7
10
2.1
3.9
9.1
1.7
5.3
6.8
4.1
1.7
11
64
30
95
77
22
176
57
44
73
34
OMe
OMe
Cl
H2
- HCl
Figure 6.4
OMe
+ 2H 2
OMe
+H
+ H2 O
-H
HO
OMe
O
+ MeOH
Proposed mechanism of cyclohexanone formation during the chloroanisole hydrogenation (adapted from ref. 29).
hydrogen atoms to only one face of the arene.31,32 The formation of the
minor trans diastereoisomers could be explained through a roll-over mechanism, in which the partially hydrogenated intermediate dissociates from
the catalysts surface and then re-associates through the opposite face before
further reduction occurs.30,33 In all cases, the nanocatalysts proved to be
easily recycled by simple extraction of the product with the adequate solvent
(alkanes or ether), during several successive reduction reactions with similar
TOFs and no metal leaching. In the case of halogenoarenes (entries 1213), a
tandem dehalogenationhydrogenation reaction occurred with Rh(0) colloids.34 Moreover, the hydrodechlorination of halogenoanisoles as model
substrates of endocrine disruptors (entry 13) yielded the relatively non-toxic
saturated product and added-value cyclohexanone in a significant amount
(nearly 50% GC yield).29 The formation of the ketone could be attributed to
the decomposition of the hemiacetal intermediate, generated from the
partially hydrogenated product in acidic conditions due to the release of HCl
during the dehalogenation step (Figure 6.4). However, in some cases, a
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
105
X
N
N
10
N
13
(1S,2S)-NMeProl16X
HO
(1R,2S)-NMeProl16X
(1S,2S,4S,5R)-QCI16X
HO
13
13
OH
13
OH
(1R,2S)-NMeEph12X
OH
(1S,2R,4S,5R)-QCD16X
Figure 6.5
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
106
Chapter 6
(a) Rh(0)@(-)-NMeEph12Br
Figure 6.6
6.3.1
Ethylpyruvate
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
Table 6.4
107
Entry
Surfactant
t (h)
Conversionb (%)
e.e.b (%)
1
2
3
4
5
6
7
8
9
10
11
12
(1R,2S)-()-NMeEph12Br
(1S,2R)-()-NMeEph12Br
(1R,2S)-()-NMeEph12HCO3
(1R,2S)-()-NMeEph12(R)-lactate
(1R,2S)-()-NMeEph12(S)-lactate
(2S)-()-NMeProl16Br
(1S,2S)-()-NMeProl16Br
(1R,2S)-()-NMeProl16Br
(1S,2S,4S,5R)-()-QCI16Br
(1S,2R,4S,5R)-()-QCD16Br
(1R,2S)-()-NMeEph12Brd
(1R,2S)-()-NMeEph12(S)-lactated
1
1
1.5
1
1
0.5
1
1
2c
2
1
1
100
100
100
100
100
100
100
100
100
60
100
100
12
15
4
9
13
1
3
5
1
0
15
18
(R)
(S)
(R)
(R)
(R)
(S)
(S)
Unsatisfactorily, QCI, QCD and NMeProl derivatives demonstrated poor results with e.e. o5%.
The low asymmetric induction observed could be attributed to a lack of
steric hindrance around the rhodium(0) particles and also to the dynamic
behavior of the protective agent at the metals surface.
6.3.2
Prochiral Arenes
View Online
108
Chapter 6
OMe
OMe
Me
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
Me
Me
MeO
Me
cis
MeO
OMe
Me
trans
Figure 6.7
Table 6.5
Entry Surfactant
1
2
3
4
5
6
t (h)
(1R,2S)-()-NMeEph12Br
3
(1R,2S)-()-NMeEph12(R)-lactate
3
(1R,2S)-()-NMeEph12(S)-lactate
3
(2S)-()-NMeProl16Br
16c
(1S,2S,4S,5R)-()-QCI16Br
1
(1S,2R,4S,5R)-()-QCD16Br
1
62
65
66
60
70
70
(0)
(1)
(0)
(1)
(1)
(1)
:
:
:
:
:
:
38
35
34
40
30
30
(0)
(0)
(0)
(1)
(2)
(2)
6.4 Conclusions
In summary, hydroxylated ammonium surfactants have been widely applied
as protective agents for the ecient electrosteric stabilization of rhodium(0)
nanoparticles in neat water. The easy modulation of the skeleton such as the
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
109
lipophilic chain length, and the nature of the associated counter-ion or the
polar head allow tuning of the stability of the aqueous suspensions and finally the hydrogenation catalytic properties. The obtained micelles are able
to confine metallic species within their cores, thus acting as promising
nanoreactors. In contrast to homogeneous species, these nanocatalysts
proved to be active in the reduction of arene derivatives under mild conditions and easily recyclable under biphasic conditions (water/substrate).
Undoubtedly, the development of asymmetric catalysis with well-defined
nanospecies will constitute an ambitious goal in the field of catalysis, which
remains unexplored in neat water.
References
1. A. Denicourt-Nowicki and A. Roucoux, Metallic Nanoparticles in
Neat Water for Catalytic Applications in Nanomaterials in Catalysis,
ed. P. Serp and K. Philippot, Wiley-VCH Verlag GmbH & Co. KGaA,
Weinheim, 1st edn, 2013, pp. 5595.
2. R. Narayanan and M. El-Sayed, Top. Catal., 2008, 47, 1521.
3. V. Polshettiwar and R. S. Varma, Green Chem., 2010, 112, 743754.
4. D. Astruc, F. Lu and J. R. Aranzaes, Angew. Chem., Int. Ed., 2005, 44,
78527872.
5. J. P. Wilcoxon and B. L. Abrams, Chem. Soc. Rev., 2006, 35, 11621194.
6. Nanoparticles and Catalysis, ed. D. Astruc, Wiley-VCH Verlag GmbH & Co.
KGaA, Weinheim, 2008.
7. A. Roucoux, A. Nowicki and K. Philippot, Rhodium and Ruthenium
nanoparticles in catalysis in Nanoparticles and Catalysis, ed. D. Astruc,
Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, 2008, pp. 349388.
8. A. Roucoux, Stabilized noble metal nanoparticles: An unavoidable family
of catalysts for arene derivatives hydrogenation, in Topics in Organometallic Chemistry: Surface and Interfacial Organometallic Chemistry and
ret and B. Chaudret, Springer GmbH, Heidelberg,
Catalysis, ed. C. Cope
2005, vol. 16 pp. 261279.
9. A. Gual, C. Godard, S. Castillon and C. Claver, Dalton Trans., 2010, 39,
1149911512.
10. N. Yan, C. X. Xiao and Y. Kou, Coord. Chem. Rev., 2010, 254, 11791218.
11. U. M. Lindstrom, Chem. Rev. (Washington, DC, U. S.), 2002, 102, 2751
2772.
12. C.-J. Li and L. Chen, Chem. Soc. Rev., 2006, 35, 6882.
13. Comprehensive Organic Reactions in Aqueous Media, ed. C. J. Li and
T. H. Chan, Wiley, New-York, 1997.
14. Aqueous-Phase Organometallic Catalysis: Concepts and Applications, ed.
B. Cornils and W. A. Hermann, Wiley-VCH Verlag GmbH & Co. KGaA,
Weinheim, 2nd edn, 2005.
15. A. Denicourt-Nowicki and A. Roucoux, Curr. Org. Chem., 2010, 14, 1266
1283.
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
110
Chapter 6
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00099
111
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
CHAPTER 7
Pd Nanoparticles in CC
Coupling Reactions
DENNIS B. PACARDO AND MARC R. KNECHT*
Department of Chemistry, University of Miami, 1301 Memorial Drive,
Coral Gables, Florida 33146
*Email: [email protected]
7.1 Introduction
Transition metals have played an important role in synthetic chemistry as
they have the unique capacity of activating various organic compounds
towards chemical transformation and/or generation of new bonds. As such,
organic ligand-stabilized metal catalysts based on Mg,14 Fe,59 Ni,1016
Au,1723 Ag,2430 Pd,3142 and Pt4349 have been synthesized to aord industrially important transformations such as carboncarbon (CC) coupling
reactions, alkylation, hydrogenation, oxygen and nitro-group reduction, and
carbon dioxide oxidation. These reactions, however, typically involve the use
of organic solvents, high temperatures, and high metal loadings. Furthermore, these processes also produce toxic by-products and are becoming
economically unsustainable and environmentally unviable.50,51 As such, new
processes are being developed to reduce the energy consumption and ecological impact of these reactions that are essential to maintain current
technological advances.52
One reaction that has garnered significant attention over the past decade
is the CC coupling reaction using Pd-based catalysts due to its applications
in pharmaceutical, natural products, and materials synthesis.5356 The
coupling reaction, shown in Scheme 7.1, typically involves the formation of a
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
112
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
Scheme 7.1
113
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
114
Scheme 7.2
Chapter 7
Grignard reagent also restricted their applications to a selection of functionalities. As an alternative to highly toxic tin and highly reactive Grignard
transmetalation agents, Suzuki developed boron-based organometallic reagents for coupling reactions with aryl halides,63,64 generating a non-toxic
and practical alternative that can be used under mild reaction conditions. In
this reaction, activation of organoboron by a base is required to produce the
boronate intermediate that initiates transmetalation, where a wide variety of
functional groups can be employed.
The importance of Pd-catalyzed CC coupling reactions in organic and
material synthesis, as well as industrial applications, was recognized by the
2010 Nobel prize in chemistry for the pioneering works of Heck, Suzuki, and
Negishi.65,66 In more recent times, technological advancements have led to
novel synthetic approaches for producing highly active Pd catalysts by using
dierent types of organic-based stabilizers that can control the reactivity and
selectivity of the metal catalyst. Furthermore, these new Pd materials have
unique characteristics and properties that enable their use in catalytic
transformations for a wide array of applications in material,6769 natural
products,7075 and pharmaceutical synthesis.54,76,77
With the advent of nanotechnology, new synthetic strategies for the generation of Pd nanocatalysts have been developed with particular emphasis
on the control of the particle size, shape, and functionality.7883 Since
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
115
nanomaterials encompass the gap between the atomic level and bulk materials, interesting physical, chemical, electronic, and optical properties can
be obtained,81,84,85 thereby generating a new class of structures for specific
applications. Modern synthetic methods allow for the production of nanosized catalysts with optimized surface-to-volume ratios, allowing for more
ecient catalytic reactivity as compared to traditional small molecule materials. The use of Pd nanoparticles for CC catalysis marks a transition from
using traditional organic solvent-based and high temperature reactions
toward more ambient conditions such as aqueous solutions, mild temperatures, and low catalyst loadings.8688 This is achieved by employing
highly functionalized stabilizers such as dendrimers,33,86,8993 proteins,31
peptides,87,88,94 and other small molecules.9598 These materials allow for
size-specific interactions, as well as control over the nanoparticle shape,
thereby generating solution stability, but at the same time exposing active
sites for interaction with substrates.
This chapter will cover select recent advances regarding the use of Pd
nanoparticles for CC coupling reactions. Due to the large volume of work
done in this area, this report will mostly focus on homogeneous Pd nanoparticles; however, it will not discuss bimetallic catalysts for coupling reactions. Emphasis will be given to Pd nanoparticles used in energy-ecient
and environmentally benign catalytic reactions, as well as information regarding their catalytic mechanisms.
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
116
Chapter 7
View Online
117
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
Pd atoms are abstracted from the surface of the nanoparticles during oxidative addition.
A thorough study of the leaching mechanism of Pd nanoparticles in coupling
reactions was investigated using a 2-compartment reactor, shown in
Figure 7.1(a).168,169 This system separates the two compartments with an
Figure 7.1
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
118
Chapter 7
alumina membrane that only allows for passage of metal species of o5 nm.169
As such, the nanocluster exclusion process was achieved in which only small
Pd clusters, Pd0 atoms, and/or Pd21 ions can traverse the alumina membrane, thereby monitoring and analysis of the leached species was possible.
Pd nanoparticles were prepared by mixing the metal precursor, Pd(OAc)2,
with tetraoctylammonium glycolate (TOAG) acting as both a stabilizer and a
reducing agent to produce B14 nm particles.169 The CC coupling initially
studied using the unique reaction system was the Heck reaction between nbutyl acrylate and iodobenzene using DMF as the solvent and NaOAc as the
base to generate n-butylcinnamate as the product. The two starting materials
were added to both sides of the reaction chamber, except the solid base was
placed only in side B and the Pd nanoparticles were fully contained in side A
at the initiation of the reaction.169 NaOAc base was necessary for the Heck
coupling to occur so no product formation was expected in side A. Control
studies showed that the solubility of NaOAc in DMF was negligible thus base
diusion from side B to side A was not observed.168 In this system, however,
diusion of the active Pd species to side B through the mesoporous membrane will allow the reaction to proceed. To this end, Heck coupling was
observed for the reaction at side B generating 88% of the n-butylcinnamate
product after 120 h, although a small amount of the product (4.9%) was also
formed in the opposite side of the reaction chamber, which was attributed to
the diusion of the product from side B to side A.169
To further verify the leaching of the Pd species through the membrane,
another reaction was performed using the Suzuki coupling between phenylboronic acid and p-iodotoluene with the same conditions as the Heck
coupling. Approximately 50% of the total anticipated product was formed on
side B after 250 h of reaction time, indicating that the catalytic Pd species
was abstracted from the nanoparticle cluster during oxidative addition and
subsequently transferred from side A to side B; however, diusion of the
product from side B to side A was also observed to be similar to the Heck
reaction albeit at a greater yield of B20%.169
The nature of the leached species was then investigated to determine
whether Pd0 species, small Pd0 clusters, or both were released through the
membrane. Initially, Pd nanoparticles were added to the DMF solvent in side
A of the chamber while only DMF was placed in side B.169 The leaching
process was then monitored for 144 h at 100 1C by taking small aliquots of
the reaction mixture in side B and analyzing the particles using inductively
coupled plasma (ICP) and transmission electron microscopy (TEM).169 TEM
results indicated that small Pd clusters do not diuse through the membrane, whereas only 20% of the total Pd was found on side B after 144 h, as
determined by ICP analysis.169 Only the Pd0 species was observed in side B,
forming irregular shaped aggregates stabilized by the presence of TOAG ions
that were released from the nanoparticle surface and diused through the
membrane due to concentration gradient.169 This was verified by a UV-vis
experiment using samples from both sides of the reaction chamber, wherein
only a broad spectrum was observed for side B indicating the presence of Pd
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
119
View Online
Chapter 7
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
120
Figure 7.2
Possible transfer mechanisms of leached Pd species in a twocomponent reactor. (a) Nanoparticle fragmentation and transfer of
small clusters does not occur, (b) leaching of Pd0 atoms and ensuing
transfer, and (c) formation of Pd21 complexes during oxidative addition
and subsequent leaching and transfer.169
Reproduced with permission from A. V. Gaikwad, A. Holuigue, M. B.
Thathagar, J. E. ten Elshof and G. Rothenberg, Chem.Eur. J., 2007, 13,
69086913. Copyright r 2007 WILEY-VCH Verlag GmbH & Co. KGaA,
Weinheim.
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
Scheme 7.3
121
View Online
Chapter 7
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
122
Figure 7.3
(a, b) TEM images, (c) SAED pattern, (d) HRTEM image, and (e) XRD
analysis of the Pd nanoparticles before catalysis. (f) TEM image and
SAED pattern (inset), (g and h) HRTEM images, (i) size distribution
histogram, and (j) EDX spectrum of Pd nanoparticles after catalysis.170
Reproduced with permission from Z. Niu, Q. Peng, Z. Zhuang, W. He
and Y. Li, Chem.Eur. J., 2012, 18, 98139817. Copyright r 2012 WILEYVCH Verlag GmbH & Co. KGaA, Weinheim.
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
123
results suggest that oxidative addition in the presence of aryl halides causes
abstraction of Pd species from the nanoparticle surface, thereby increasing
the amount of Pd released to solution. On the other hand, only minimal
leaching was observed without aryl halides, which can be attributed to the
inherent instability of the nanoparticles.
While the crystallinity results suggest leaching of Pd atoms during oxidative addition, further evidence was obtained using a bimetallic Pd-Au
nanoparticle system. In this case, co-reduction of Pd and Au metal precursors was achieved by adding oleylamine at 230 1C to produce spherical
nanoparticles with a B9 nm average diameter size.170 HRTEM and
EDX mapping analysis revealed that the particles were composed of a 31 : 69
PdAu ratio with lattice spacings between the expected value for face-centered cubic (fcc) Pd and Au. Catalytic reactions between 4-iodoanisole and
phenylboronic acid were used to probe the leaching process driven by the
Pd-Au nanoparticles. TEM analysis of the bimetallic nanocatalyst after the
reaction revealed that the shape and size of the particles remained the same;
however, EDX results showed the PdAu atomic ratio to be 19 : 81, indicating
a decreased amount of Pd atoms in the nanoparticles. Interestingly, highly
crystalline Pd nanoparticles B5 nm in size were formed in this system, as
confirmed by TEM and EDX, suggesting that the Pd atoms were selectively
abstracted from the bimetallic nanoparticle during the coupling reaction.170
To verify the role of oxidative addition by the aryl halide in the leaching
process, control reactions were done without 4-iodoanisole, wherein the EDX
analysis of the materials after the process displayed a PdAu ratio of 27 : 73,
which was similar to the original bimetallic ratio.170 These results indicate
that the aryl halides leached the Pd0 species during the oxidative addition
process in the catalytic reaction.
Overall, the results of these studies indicated that the atom-leaching
mechanism promoted by oxidative addition of the aryl halide was the likely
catalytic process for Pd nanoparticles in the Suzuki coupling reaction. These
results oer direct evidence using various analytical techniques and dierent types of Pd nanoparticles wherein atom abstraction during the catalytic
cycle was both qualitatively and quantitatively analyzed. Aside from indirect
processes such as monitoring the TOF values, these studies provided physical validation of an atom-leaching process for Pd nanoparticles.
The other side of the debate regarding the catalytic mechanism of Pd
nanoparticles centers on the argument that the entire reaction occurs directly on the surface of the particles in a straightforward heterogeneous reaction. To this end, evidence of surface-based Suzuki coupling catalyzed by
Pd nanoparticles was examined by monitoring the local coordination environment of the metal in real time using a recirculating reactor set up for
operando fluorescence X-ray absorption spectroscopy (XAS) illustrated in
Figure 7.4(a).171 Polymer-stabilized Pd nanoparticles were prepared by
refluxing H2PdCl4 and polyvinylpyrrolidone (PVP) with water and ethanol to
generate 1.84.0 nm Pd0 particles.171 Suzuki coupling between iodoanisole
and phenylboronic acid was monitored for these studies wherein the
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
124
Figure 7.4
Chapter 7
(a) Schematic diagram of the recirculating reactor setup for XAS, (b)
normalized TOF data for Suzuki coupling reaction showing rate dependence on correct active sites and not on particle size.171 Reproduced
with permission from P. J. Ellis, I. J. S. Fairlamb, S. F. J. Hackett K.
Wilson and A. F. Lee, Angew. Chem., Int. Ed., 2010, 49, 18201824.
Copyright r 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
catalytic activity was expressed as turnover frequency (TOF) values. When the
reactivity of the nanocatalyst was monitored with respect to particle diameter, the TOF values decreased as the size of the nanoparticles increased as
expected; however, when the TOF values were normalized with respect to the
edge and vertex atoms on the surface of the nanoparticles, statistically
similar TOF values were obtained regardless of particle size
(Figure 7.4(b)).171 This suggests that the reactivity of Pd nanoparticles is
associated with the defect atoms on the particle surface.
As such, monitoring of the possible leached species was performed using
XAS, which can sensitively measure the average PdPd coordination number
(CN), as well as provide information concerning the particle size and
morphology. From the extended X-ray absorption fine structure (EXAFS)
intensities for simulated 1.8 nm Pd nanoparticles, a PdPd CN value of 9.46
was calculated, which is in agreement with the observed value of 9.58.171
Leached Pd species, however, were not observed as there was neither sintering nor dissolution of Pd atoms since there is no change in the CN, which
supported the stability of the nanoparticles in the reaction. HRTEM analysis
before and after the reaction also illustrated no change in the size and shape
of the nanoparticles.171 Simulation of analogous 1.8 nm particles with 48 Pd
defect atoms removed into the solution would lead to a drop of CN to 7.23;
however, this phenomenon was not observed, which indicated that there was
no solubilization and homogeneous contribution of catalytically active Pd
species.171
To further probe the system with regards to leached Pd atoms, trace
amounts of dissolved Pd(OAc)2 were added to the reaction, which was expected to enhance the catalytic activity of the particles; however, no such
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
171
125
View Online
126
Chapter 7
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
Understanding the actual reaction mechanism can provide valuable information for the design of new synthetic routes for the production of more
ecient nanocatalysts.
Figure 7.5
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
127
reactivity of each substrate. To this end, the aryl halide substrates with
electron-withdrawing (COOH) and iodide groups were more reactive than
the corresponding bromide-substituted and phenolic substrates. TOF values
of 2000 200 mol product (mol Pd h)1 for 4-iodobenzoic acid were obtained, suggesting a highly ecient nanocatalyst.86
An interesting control reaction was performed using Pd nanoparticles
generated without the DENs to evaluate the role of the dendritic cage in
catalytic reactivity. As such, K2PdCl4 was reduced with NaBH4 in the absence
of the dendrimers, thus producing polydisperse particles that generated
a TOF value slightly lower than that of Pd DENs (1800 mol product
(mol Pd h)1).86 This result suggests that the size control provided by the
DENs enables the catalyst to have a higher surface area than the particles
without DENs to provide better catalytic reactivity.86 TEM analyses of the Pd
DENs before and after the catalytic reaction were performed, which showed a
moderate increase in particle size and dispersity from 1.7 0.4 nm to
2.7 1.1 nm.86 These results indicated partial aggregation of the Pd nanoparticles as a consequence of the catalytic reaction that was attributed to
leaching of Pd atoms during the catalytic cycle. To confirm this, the Pd DENs
were added to an identical reaction solution but without the substrates and
allowed to proceed for 24 h. TEM analyses of the nanoparticles before and
after the reaction showed statistically similar particle diameter indicating
that the catalytic reaction was responsible for an increase in nanoparticle
size.86
When the Pd DENs were employed in Stille couplings at elevated temperatures, catalyst aggregation was observed by the appearance of a black
precipitate.86 As such, product yields under these reaction conditions were
lower than the reactions at room temperature, which could arise from
dendrimer degradation. Finally, the recyclability of the Pd DENs was also
evaluated using sequential Stille coupling between two dierent substrates.86 For this, the second substrate was added after the reaction of the
initial substrate was concluded. Specifically, a reaction solution containing
4-iodophenol and phenyltin trichloride was added to the Pd DENs and
allowed to proceed for 15 h. After reaching completion, another reaction
solution containing 4-iodobenzoic acid and phenyltin trichloride was added
to the first mixture and allowed to proceed for 15 h such that in the end,
two possible products can be observed if the Pd DENs remain active. Gas
chromatography-mass spectrometry (GC-MS) analysis of the reaction demonstrated the quantitative formation of the two anticipated products
(4-phenylphenol and 4-biphenylcarboxylic acid), indicating that the Pd DENs
were still catalytically active after the initial cycle.86 These studies showed
that Pd DENs catalyzed Stille coupling under ambient reaction conditions,
which could serve as a model for designing catalytic systems.
In a separate study, the eect that the dendrimers possessed over the
catalyst reactivity and selectivity was studied by comparing dierent precatalysts, namely: Pd DENs, Pd(OAc)2, and Pd21-PAMAM complexes.90
In Stille coupling reactions employing the same system, which Crooks
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
128
Chapter 7
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
Figure 7.6
129
(a) Scheme for the synthesis of GnDenP-Pd nanoparticles and (b) their
use in the Stille coupling reaction between methyl-4-bromobenzoate
and tributyl(phenyl)stannane.93
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
130
Chapter 7
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
Figure 7.7
131
The catalytic activity of these nanocatalysts was initially tested using the
Stille coupling reaction between 4-iodobenzoic acid and phenyltin trichloride. Quantitative yields were observed after 24 h for reactions using low
catalyst amounts (Z0.01 mol% Pd for all Pdpeptide ratios), suggesting
that these materials were highly ecient under ambient conditions.87 TOF
analysis of the dierent peptide-templated structures generated modest
values of B450 mol product (mol Pdh)1 for both Pd60 and Pd120, while a
value of B350 mol product (mol Pdh)1 was obtained for Pd90.87 Although
these TOF values are lower than those reported by Crooks and colleagues86
using the same reaction, it is worth noting that for the R5-templated Pd
nanoparticles, lower catalyst amounts were required to generate quantitative
product yields.
Employing biological molecules as tools for the biomimetic synthesis of
metal nanoparticles could generate a model system for green nanocatalysis.
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
132
Chapter 7
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
133
addition and abstract the active metal from the particle surface, from which
the catalytic cycle was completed, indicating an atom-leaching mechanism.
After reductive elimination when the product was released, the free, highly
active Pd0 generated in solution could proceed along three dierent paths, as
illustrated in Figure 7.8(a): (1) reused catalytically until all starting materials
are consumed, (2) quenched by the remaining nanoparticles, or (3) aggregated with other leached Pd atoms to generate Pd black.184 To probe this
catalytic mechanism, TOF analyses using dierent Pd loadings were performed. Based on the results shown in Figure 7.8(b), the TOF values are
constant at B2400 mol product (mol Pdh)1 when using r0.05 mol% Pd;
however, higher Pd loadings generated a decreasing trend in TOF values.184
This indicates decreased catalyst eciency due to the formation of catalytically-inactive Pd black at the higher reaction concentrations.
The fate of the Pd0 atoms after the individual Stille catalytic cycle was
explored using competition experiments containing combinations of two
dierent aryl halides in a single Stille reaction. For this, 4-iodobenzoic acid
and 4-bromobenzoic acid were co-dissolved in water and added with phenyltin trichloride and the peptide-capped Pd nanoparticles at room temperature.185 The results indicated that only the aryl iodide substrate reacts at
room temperature even though the highly reactive Pd0 atoms were present in
solution after the initial cycle. When the reaction temperature was increased
to 40 1C, consumption of the aryl bromide was observed along with an increase in the amount of product generated. Furthermore, when the same
bicomponent reaction was heated to 60 1C after the addition of nanocatalyst
(i.e. no substrate conversion had begun), coupling for both iodo- and bromosubstrates was observed, demonstrating a loss of catalyst selectivity.185 These
results suggest that substrate selectivity can be thermally controlled at rather
lower temperatures wherein the iodo-components reacted at room temperature while the bromo-component required thermal activation to
Figure 7.8
(a) Scheme demonstrating the atom-leaching mechanism for peptidecapped Pd nanoparticles and (b) TOF analysis for Stille coupling using
dierent Pd loadings.184 D. B. Pacardo, J. M. Slocik, K. C. Kirk, R. R.
Naik and M. R. Knecht, Nanoscale, 2011, 3, 21942201.
Reproduced by permission of The Royal Society of Chemistry.
View Online
134
Chapter 7
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
proceed. Unfortunately, no reactivity was observed for comparable chlorohalides. Similar degrees of selectivity were observed for Stille coupling using
aryl dihalides; however, the attachment of electron-donating or withdrawing
groups modulated the catalyst selectivity.185
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
135
acid, 95% yield of the biphenyl product was generated. On the other hand,
using sulfur-containing starting materials such as 2-iodothiophene or
2-thiopheneboronic acid, resulted in decreased product yields ranging from
26% to 92%. The decrease in product yield can be attributed to poisoning
of the nanocatalyst by the thiol groups in the reactants. Although these
reactions showed decent product yields for most of the substrates used,
precipitation of Pd black was noted during the reaction, which was attributed to the high reaction temperatures. In another experiment, the initial
rate of the coupling reaction between phenylboronic acid and iodobenzene
in acetonitrilewater solvent was determined.96 The results indicated that
the reaction rate was directly proportional to the concentration of PVP-Pd
nanocatalyst, from which the authors suggested that CC bond formation
occurred at the catalyst surface. While the PVP-Pd nanoparticles were
eective Suzuki coupling catalysts, the formation of Pd black decreased the
overall eciency of the catalyst.
Another polymer ligand used for the synthesis of Pd nanoparticles is the
block copolymer polystyrene-b-poly(sodium acrylate) (PS-b-PANa).187 The
polymer structure was composed of a hydrophobic polystyrene component
and a hydrophilic poly(sodium acrylate) component. This amphiphilic
polymer generates a micellar structure in solution that provides steric stabilization of nanoparticles. The block copolymer-stabilized Pd nanoparticles
were synthesized by refluxing PS-b-PANa with H2PdCl4 in EtOH, producing a
clear dark brown solution containing 3.0 0.7 nm particles based upon TEM
analysis.187 Suzuki coupling reactions between 2-thiopheneboronic acid and
iodobenzene were performed using a 0.6 mol% Pd catalyst loading with NEt3
or NaOAc as the base, generating yields of 85% or 88%, respectively after
12.0 h under reflux. Incidentally, the base influenced the stability of the Pd
nanoparticles; NEt3 provided additional nanoparticle stabilization, while
NaOAc resulted in particle precipitation. As such, NEt3 was used as a base in
subsequent Suzuki couplings between phenylboronic acid and bromothiophene using 3 : 1 acetonitrilewater as the solvent. In this reaction, competitive formation of the product (2-phenylthiophene) and homocoupling
product (bithiophene) was observed, generating similar yields of 16% with
no precipitation of Pd black.187 In contrast, when PVP-Pd nanoparticles were
employed for a similar reaction between bromothiophene and phenylboronic acid, only homocoupling products were generated; however,
comparing the catalyst stability as a function of ligands, the PS-b-PANabased nanoparticles were more stable than the PVP at these conditions as Pd
black was not produced in the reaction.187
PEG-stabilized Pd nanoparticles have also been employed for Suzuki
coupling reactions in an aqueous medium between aryl halides and phenylboronic acid using K2CO3 as the base.97 Initially, the reaction between
p-bromoacetophenone and phenylboronic acid in a 1 : 5 dimethoxyethane
(DME)water solvent at room temperature generated the expected product
with 95% isolated yield after 3.0 h. Subsequent reactions were performed
in pure water using aryl halides with dierent electron-donating and
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
136
Chapter 7
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
Figure 7.9
137
(a) SEM image of PANI nanofibers, TEM images of Pd-PANI nanoparticles (b) before the reaction, (c) halfway through the reaction, and
(d) after two reaction cycles. (e) Suzuki coupling reaction between
1,4-difluorobenzene and phenylboronic acid using Pd-PANI nanoparticles.188
Reproduced with permission from B. J. Gallon, R. W. Kojima, R. B.
Kaner and P. L. Diaconescu, Angew. Chem., Int. Ed., 2007, 46, 7251
7254. Copyright r 2007 WILEY-VCH Verlag GmbH & Co. KGaA,
Weinheim.
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
138
Chapter 7
chloride and phenylboronic acid. For this, minimal loss of activity was observed after ten reaction cycles, generating Z89% product yields with no
noted catalyst aggregation as observed in the TEM analysis of the particles
after two reaction cycles.188 The decreased product yield was attributed to
the degradation of the nanofiber support after several catalytic cycles.
While synthetic PANI was an excellent support for the catalytic nanoparticles, biologically derived nanofibers, such as bacterial cellulose (BC),
can also be used.189 BC was produced by fermentation of Acetobacter xylinum
with glucose as the carbon source. The three-dimensional nanofibers
interacted with metal ions to serve as a support for nanoparticle production.
The one-pot synthesis involved mixing of metal precursors with BC nanofibers under an N2 atmosphere at 140 1C.189 Once complete, NaBH4 was
added, producing polydispersed, B20 nm Pd nanoparticles deposited on the
nanofibers (termed Pd-BC). Initial catalytic analysis of the materials was
performed using the coupling of iodobenzene and phenylboronic acid,
generating almost quantitative yields of 98% after 3.5 h at 85 1C. Recycling
experiments using the same reactants showed no loss of activity after two
reaction cycles with no visible Pd aggregation as observed by TEM; however,
similar to the Pd-PANI nanoparticles, a decline in the quality of the nanofiber support was observed after extended recycling runs that resulted in
lower product yields.189
Suzuki reactions using aryl iodide and substituted phenylboronic acids
were also studied with the Pd-BC nanocatalyst, which generated excellent
yields between 8899% of the expected coupling products depending
on the substituent attached on the nucleophile.189 To this end, when
3-methoxyphenylboronic acid was used in the reaction with iodobenzene, a
99% yield was obtained, whereas using 4-cyanophenylbenzoic acid generated a yield of only 88%. Recycling experiments were again performed for
the Pd-BC nanoparticles using the substituted nucleophiles wherein product yields 486% were obtained after five catalytic cycles. To further explore
the reactivity of Pd-BC nanoparticles, the least reactive substrate, aryl
chloride, was employed for Suzuki coupling with dierent substituted
phenylboronic acids. Surprisingly, product yields between 7592% were
recorded, indicating that an extremely ecient nanocatalyst was developed.189 Finally, leaching of the active catalyst was also studied using
inductively coupled plasma-atomic emission spectroscopy (ICP-AES) analysis wherein the amount of Pd species before and after the five reaction
cycles was quantitated. The results showed that the Pd concentration before
the reaction was 5.29% and 5.26% after the reaction, indicating a negligible
amount of Pd was leached during the reaction.189 As such, the results of
elemental analysis suggested that the BC nanofibers are good protecting
ligands for the nanoparticles.
The Pd-substituted Keggin-type polyoxometallate-based nanoparticles,
described earlier, were also used for the Suzuki reaction.176 For this, aryl
bromides and phenylboronic acid were reacted in an EtOHwater solvent
with diisopropylamine as the base. Nearly quantitative yields of the expected
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
139
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
140
Chapter 7
ecient catalytic system was obtained using the nearly identical dendrimerbased materials, which may arise from the dierent reagents employed.
Interestingly, no precipitation of Pd black was noted during the reaction.
Aryl bromides were also studied using this system; however, they were not as
reactive as the iodo-based substrates, requiring higher temperatures and
longer reaction times. Aryl chlorides were also shown to be completely unreactive.190 Finally, the recyclability of the nanocatalyst was studied using
the coupling of ethyl-4-iodobenzoate and p-tolylboronic acid wherein formation of Pd black during the subsequent reactions was observed resulting
in decreased product yield, from 99% to 80%, after the third reaction
cycle.190
The Astruc group has developed a unique dendrimer-stabilized Pd nanoparticle system using 1,2,3-triazolylsulfonate dendrimers, synthesized
through click chemistry.191 In this system, Pd21 was introduced to the triazole ligands producing two absorption bands in the UV-vis spectrum at 208
and 235 nm. Upon addition of NaBH4, Pd nanoparticles of B23 nm in
diameter were generated using three dierent dendrimer generations
(Pd-DSN-G0, Pd-DSN-G1, Pd-DSN-G2).191 TEM analysis showed that the
nanoparticles sizes were larger than the size of the ligands, thus leading to a
large degree of polydispersity from the materials as the generation of the
dendrimers increases. Room temperature Suzuki coupling reactions between iodobenzene and phenylboronic acid in a waterethanol solvent were
used to examine the reactivity of the dendrimer-stabilized materials. Using a
0.1 mol% Pd loading, product yields of 96%, 95%, and 70% were obtained
using Pd-DSN-G0, Pd-DSN-G1 and Pd-DSN-G2, respectively, suggestive of a
decrease in catalytic eciency with respect to increasing dendrimer generation.191 Similar trends were observed when lower Pd loadings (0.01 mol%)
were used in the reaction. TOF analysis employing the Pd-DSNs indicated
that there was no significant dierence in the values based upon dendrimer
generation; however, significantly higher TOFs of 11671533 mol PhI
(mol Pdh)1 were obtained for the reaction employing 0.01 mol% Pd, while
TOF values of 117167 mol PhI (mol Pdh)1 were generated when using
0.1 mol% Pd, demonstrating an eect based upon the Pd loading.191 From
these results, the catalytic mechanism of Pd-DSNs was suggested to follow
an atom-leaching process wherein the TOF values of the reaction decrease as
the amount of Pd loading increases. Quantitative yields were also obtained
in the coupling reaction using bromobenzene, but the reaction needed to be
heated to 100 1C.191
The phosphine-based dendrimers, GnDenP-Pd, whose synthesis was discussed earlier, have also been used for Suzuki coupling.93 The reaction between bromobenzene and phenylboronic acid was initially tested with
dierent bases such as NEt3, NaOAc, K2CO3, and K3PO4 to optimize the reaction conditions. As such, K3PO4 was determined to be the best base for the
Suzuki reaction using GnDenP-Pd in dioxane as the solvent under refluxing
conditions. Aryl iodides and aryl bromides with electron-withdrawing or
electron-donating groups were reacted with dierent phenylboronic acids,
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
141
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:1
142
Figure 7.10
Chapter 7
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
Figure 7.11
143
View Online
144
Chapter 7
specific binding between the receptor and substrates controlled the reactivity, leading to high degrees of specificity.
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
145
compared with the third generation PPI-Pd nanoparticles even though the
diameters of the particles were similar. These results suggested that the fifth
generation dendrimers provided larger encapsulated space wherein the
coupling reaction could occur leading to the observed regioselectivity and
higher product yields. In contrast, the third generation dendrimers contained
less interior room such that slower reactions occurred, lowering the product
yields. Nevertheless, product regioselectivity was maintained. As anticipated,
lower yields were obtained when using aryl bromides compared with aryl
iodides and no activation of aryl chlorides was noted. Finally, para-nitrosubstituted aryl bromides generated lower product yields compared with
unsubstituted aryl reagents, which was attributed to the repulsive forces between the electron-donating nitro-group and the perfluorinated polyether
dendrimer groups, making substrate incorporation dicult.193 Taken together, these results suggested that Heck coupling using the PPI-Pd nanoparticles occurred at the dendrimer interior.
Since the reaction using the dendrimer nanocatalysts was performed in a
fluorocarbonhydrocarbon solvent, separation and recovery of the PPI-Pd
nanoparticles was easily achieved.193 To this end, the nanocatalysts remained dissolved in the fluorous phase of the solvent, as shown by the retention of the dark color of the catalyst, while the organic phase that
contained the product was colorless. Recyclability of the nanocatalyst was
performed using the separated particles in the fluorous phase wherein significant loss of activity was observed with each reaction cycle. This loss in
reactivity was attributed to changes in the morphology of the Pd nanoparticles due to the redox cycling of Pd atoms in the Heck reaction. Control
reactions were also performed to determine the eect of base identity on the
reaction. These results showed that even in the absence of Et3N in the reaction, Heck coupling still occurred due to the interior tertiary amines in the
dendrimers that acted as the base. Overall, the results of these initial studies
of Pd DENs for Heck coupling showed highly regioselective intradendrimer
catalytic reactivity under relatively ambient reaction conditions.
Further DEN eects for Pd nanoparticles were performed by Kaneda et al.,
wherein the Pd complex was immobilized inside the dendrimer through
ionic bonds (Figure 7.12).91 Initially, the peripheral amino groups of PPI
dendrimers were functionalized with decanoyl chloride and 3,4,5-triethoxybenzoyl chloride to produce alkylated and arylated dendritic materials,
respectively. On the other hand, Pd nanoparticles were exposed to 4-diphenylphosphinobenzoic acid as ligands, which can then form ionic bonds
between the carboxylic acid moiety and the internal amino groups of the
dendrimers. To this end, the phosphine-stabilized Pd nanoparticles were
encapsulated at the interior of the functionalized PPI dendrimers and employed as nanocatalysts in the Heck reaction.91 Initially, the coupling between iodobenzene and n-butyl acrylate in toluene with KOAc was performed
with the PPI modified Pd DENs. The results of the reaction showed that
the dendrimer generation directly aected the reactivity where the reaction
rate increased with higher generations; however, when the molar ratio of
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
146
Figure 7.12
Chapter 7
phosphine ligandPd in the 5th-generation PPI Pd nanoparticles was increased from 1 to 4, product yields decreased from 92% to 19%.91 These
results suggested that the large amounts of amino groups in higher generation dendrimers provided complexation with the active species and as such
led to high stability and catalytic reactivity.
To further probe the reactivity, as well as the selectivity of the nanocatalyst, a second Heck coupling reaction was employed using p-diiodobenzene
and n-butylacrylate with two possible products, mono- and di-substituted
benzene.91 The results showed that in the presence of the nanocatalyst, only
n-butyl-p-iodocinnamate, the mono-substituted product, was formed, indicating a highly reactive (86% yield) and selective (92%) nanocatalyst. Furthermore, the allylic amination of cinnamyl methyl carbonate was also
catalyzed by the Pd DENs, demonstrating that the catalytic materials were
quite versatile. Finally, the recyclability of the catalyst was established
wherein minimal loss of activity was observed after four reaction cycles.
Expanding the scope of reactivity of the PPI-Pd nanoparticles for Heck
coupling under environmentally friendly conditions was achieved using
supercritical CO2 (scCO2) as the solvent.194 The PPI dendrimer end groups
were functionalized with a perfluoro-2,5,8,11-tetramethyl-3,6,9,12-tetraoxapentadecanoyl perfluoropolyether chain to enhance solubility with
scCO2, generating 12 nm Pd particles after reduction.194 The modified PPIPd nanoparticles were then used as the coupling catalyst between iodobenzene and methylacrylate with Et3N as the base at 75 1C and 5000 psi.194
Formation of the possible products, methyl-2-phenylacrylate and cis/transcinnamate, was monitored using GC wherein 57% of the aryl iodide was
consumed after 24 h.194 Characterization of the reaction solution by 1H NMR
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
147
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
148
Chapter 7
The results displayed that quantitative yields (499%) of the expected products were obtained from activated aryl halides after only 2 h. Iodobenzene,
on the other hand, only managed to generate 49% of the product after 1.5 h,
while increasing the reaction time to 4 h showed a corresponding yield increase to 81%.135 These results indicated that the ionic liquid was a great
solvent for the Heck reaction and the use of Pd-IP-IL nanoparticles shows
excellent reactivity under ambient reaction conditions.
Finally, the Keggin-type polyoxometallate-stabilized Pd nanoparticles were
also employed for the Heck coupling reaction.176 Initially, the reaction between bromotoluene and styrene was performed employing diisopropylamine as the base in a waterethanol solvent. Interestingly, excellent yields of
95% were obtained for the Heck coupling product, stilbene, at 8085 1C after
16 h. When activated aryl chloride, 1-chloro-4-nitrobenzene, was used under
similar reaction conditions, a 91% product yield was obtained.176 Similarly,
for the Heck reaction between aryl halides and methyl acrylate, an almost
quantitative yield (97%) of methyl cinnamate was generated from aryl
bromide, while a 94% product yield was obtained using aryl chloride.176 The
mild reaction conditions employed in the coupling reaction suggested a
highly ecient nanocatalyst that was operational under environmentallyfriendly conditions.
View Online
149
catalytic reactions, which may serve as model systems for the development of
sustainable materials for future chemical transformations.
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
References
1. M. R. Kember and C. K. Williams, J. Am. Chem. Soc., 2012, 134,
1567615679.
2. N. Hanada, T. Ichikawa and H. Fujii, J. Phys. Chem. B, 2005, 109,
71887194.
3. A. Ganguly, P. Trinh, K. V. Ramanujachary, T. Ahmad, A. Mugweru and
A. K. Ganguli, J. Colloid Interface Sci., 2011, 353, 137142.
4. A. B. Patil and B. M. Bhanage, Catal. Commun., 2013, 36, 7983.
pez and A. Corma, ChemCatChem, 2012, 4, 751752.
5. C. Lo
6. J. F. Sonnenberg, N. Coombs, P. A. Dube and R. H. Morris, J. Am. Chem.
Soc., 2012, 134, 58935899.
7. H. M. Torres Galvis, J. H. Bitter, C. B. Khare, M. Ruitenbeek,
A. I. Dugulan and K. P. de Jong, Science, 2012, 335, 835838.
8. S. Lastella, Y. J. Jung, H. Yang, R. Vajtai, P. M. Ajayan, C. Y. Ryu,
D. A. Rider and I. Manners, J. Mater. Chem., 2004, 14, 17911794.
9. Y. Moglie, C. Vitale and G. Radivoy, Tetrahedron Lett., 2008, 49,
18281831.
10. K. Shimizu, K. Kon, W. Onodera, H. Yamazaki and J. N. Kondo, ACS
Catal., 2012, 3, 112117.
11. H. Y. Wang and A. C. Lua, J. Phys. Chem. C, 2012, 116, 2676526775.
12. P.-Z. Li, A. Aijaz and Q. Xu, Angew. Chem., Int. Ed., 2012, 51, 67536756.
13. F. Alonso, I. Osante and M. Yus, Adv. Synth. Catal., 2006, 348, 305308.
14. A. Wang, H. Yin, H. Lu, J. Xue, M. Ren and T. Jiang, Catal. Commun.,
2009, 10, 20602064.
15. W. Zhang, H. Qi, L. Li, X. Wang, J. Chen, K. Peng and Z. Wang, Green
Chem., 2009, 11, 11941200.
16. I. Geukens, J. Fransaer and D. E. De Vos, ChemCatChem, 2011, 3,
14311434.
17. Y. Chen, P. Crawford and P. Hu, Catal. Lett., 2007, 119, 2128.
th, Z. Pa
szti, L. To
th, Z. E. Horva
th, A. Karacs and
18. L. Guczi, D. Horva
G. Peto, J. Phys. Chem. B, 2000, 104, 31833193.
19. M. Kidwai, V. Bansal, A. Kumar and S. Mozumdar, Green Chem., 2007,
9, 742745.
20. M. Kidwai and S. Bhardwaj, Appl. Catal., A, 2010, 387, 14.
21. T. Mitsudome, A. Noujima, T. Mizugaki, K. Jitsukawa and K. Kaneda,
Chem. Commun., 2009, 53025304.
22. L. Pasquato, F. Rancan, P. Scrimin, F. Mancin and C. Frigeri, Chem.
Commun., 2000, 22532254.
23. L. Wang, X. Meng, B. Wang, W. Chi and F.-S. Xiao, Chem. Commun.,
2010, 46, 50035005.
24. X. Jiang, Y. Xie, J. Lu, L. Zhu, W. He and Y. Qian, Langmuir, 2001, 17,
37953799.
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
150
Chapter 7
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
151
49. H. Wu, R. Tang, Q. He, X. Liao and B. Shi, J. Chem. Technol. Biotechnol.,
2009, 84, 17021711.
50. J. N. Armor, Appl. Catal., A, 2000, 194195, 311.
51. G. Centi, P. Ciambelli, S. Perathoner and P. Russo, Catal. Today, 2002,
75, 315.
52. D. Astruc, Inorg. Chem., 2007, 46, 18841894.
53. C. Barnard, Platinum Met. Rev., 2008, 52, 3845.
54. T. J. Colacot, Platinum Met. Rev., 2009, 53, 183188.
55. R. Jin, Nanotechnol. Rev., 2012, 1, 3156.
56. J. Yin, Applications of Transition Metal Catalysis in Drug Discovery and
Development, ed. M. L. Crawley and B. M. Trost, John Wiley & Sons, Inc.,
Hoboken, NJ, 1st edn, 2012, pp. 97163.
57. R. F. Heck and J. P. Nolley, J. Org. Chem., 1972, 37, 23202322.
58. A. Balanta, C. Godard and C. Claver, Chem. Soc. Rev., 2011, 40,
49734985.
59. K. Sonogashira, Y. Tohda and N. Hagihara, Tetrahedron Lett., 1975, 16,
44674470.
60. D. Milstein and J. K. Stille, J. Am. Chem. Soc., 1978, 100, 36363638.
61. E. Negishi, A. O. King and N. Okukado, J. Org. Chem., 1977, 42,
18211823.
62. T. Hayashi, M. Konishi and M. Kumada, Tetrahedron Lett., 1979, 20,
18711874.
63. N. Miyaura, K. Yamada and A. Suzuki, Tetrahedron Lett., 1979, 20,
34373440.
64. N. Miyaura and A. Suzuki, J. Chem. Soc., Chem. Commun., 1979,
866867.
65. T. J. Colacot, Platinum Met. Rev., 2011, 55, 8490.
66. Nat. Mater., 2011, 10, 333.
67. F. Qing, Y. Sun, X. Wang, N. Li, Y. Li, X. Li and H. Wang, Polym. Chem.,
2011, 2, 21022106.
68. N. C. Thanh, M. Ikai, T. Kajioka, H. Fujikawa, Y. Taga, Y. Zhang,
S. Ogawa, H. Shimada, Y. Miyahara, S. Kuroda and M. Oda, Tetrahedron, 2006, 62, 1122711239.
69. H. A. M. van Mullekom, J. A. J. M. Vekemans and E. W. Meijer, Chem.
Eur. J., 1998, 4, 12351243.
70. P. D. Thornton, N. Brown, D. Hill, B. Neuenswander, G. H. Lushington,
C. Santini and K. R. Buszek, ACS Comb. Sci., 2011, 13, 443448.
71. P. S. Deore and N. P. Argade, J. Org. Chem., 2011, 77, 739746.
and Y. Ramondenc, Tetrahedron, 2009,
72. W. Bentoumi, J. Helhaik, G. Ple
65, 19671970.
73. F. Zeng and E. Negishi, Org. Lett., 2001, 3, 719722.
74. S. J. Danishefsky, J. J. Masters, W. B. Young, J. T. Link, L. B. Snyder,
T. V. Magee, D. K. Jung, R. C. A. Isaacs, W. G. Bornmann, C. A. Alaimo,
C. A. Coburn and M. J. Di Grandi, J. Am. Chem. Soc., 1996, 118, 28432859.
75. C. Y. Hong, N. Kado and L. E. Overman, J. Am. Chem. Soc., 1993, 115,
1102811029.
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
152
Chapter 7
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
153
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
154
Chapter 7
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
155
View Online
23/06/2014 08:14:56.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00112
156
Chapter 7
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
CHAPTER 8
8.1 Introduction
Gold was traditionally regarded as non-reactive in chemistry and useless in
catalysis. Indeed, the chemical industry relies heavily on other metal catalysts, such as supported platinum, palladium, and silver catalysts. The fact
that bulk gold is stable in air does not mean that gold is not reactive in other
reaction media. Likewise, the fact that gold catalysts prepared inappropriately (with large gold particles dispersed on certain supports) are inactive in
catalysis does not mean that gold catalysts prepared properly (with small and
finely divided gold nanoparticles) are useless. In the 1980s, Haruta and
co-workers found that gold catalysts prepared by co-precipitation or
depositionprecipitation can be highly active for CO oxidation below room
temperature.1,2 Many gold nanocatalysts have been developed ever since,37
and some can catalyze CO oxidation below 80 1C.8,9
Catalytic CO oxidation is an important reaction in saving lives and for
purifying indoor air. The fact that a catalyst can catalyze CO oxidation below
80 1C does not mean we have to use a cooler in practical applications.
Rather, a high activity of a catalyst at a very low temperature usually dilates
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
157
View Online
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
158
Chapter 8
Metal carbonates have dierent solubilities in water and dierent decomposition temperatures. For instance, MgCO3 decomposes at 300 1C to
form MgO and CO2, whereas CaCO3 decomposes at 900 1C. It is possible to
prepare metal carbonate-supported gold catalysts, if the support is insoluble
in water and the preparation and operation temperatures of the gold catalysts are lower than the decomposition temperatures of selected metal carbonate supports.
View Online
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
159
Figure 8.1
View Online
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
160
Chapter 8
We can see from these examples that certain metal carbonate-based gold
catalysts can catalyze CO oxidation and aerobic oxidation of alcohols.
However, the scope of application is very narrow, and the application in a
wide range of catalytic applications has not been demonstrated. Besides, in
the second example,28 the Au/Cs2CO3 was formed in situ, and the sizes of the
formed gold nanoparticles were quite large. It would be interesting to prepare a series of metal carbonate-supported gold nanocatalysts and test these
materials in the aerobic oxidation of alcohols in future research. Finally,
there has been some interest in the synthesis of nanosized metal carbonates
such as CaCO3, but the application of these materials in making gold catalysts has not been demonstrated.
8.2.2
Metal phosphates are generally insoluble in water and have good thermal
stability. Some of them are considered to be solid acids but they also have
varied basic and redox properties.26 Therefore, metal phosphates themselves
can be used as solid catalysts. Alternatively, they can be used as supports for
loading metal nanoparticles. That way, the prepared catalysts may have both
the functionality originated from the support and the functionality bestowed
by the supported gold. The combination of dual functionalities is amazing,
considering that metal particles can catalyze oxidation and hydrogenation
reactions.
Dai and co-workers prepared Au/LaPO4 catalysts.29 LaPO4 nanoparticles
(6.8 nm, 111 m2 g1) were prepared by a sonication method, and gold was
loaded onto the support via depositionprecipitation. For comparison, a
commercial LaPO4 (10 nm, 55 m2 g1) was also used for loading gold. Both
catalysts, in the as-synthesized forms, showed high CO conversions below
0 1C, whereas the Au/LaPO4-nanoparticle catalyst exhibited better thermal
stability and activity after calcination at 500 1C. This work is interesting
because it showed that LaPO4, a metal phosphate, can also be used as a
support for making active gold catalysts.
Ma et al. developed an array of Au/metal phosphate (denoted as Au/M-P-O)
catalysts.30 Zirconium phosphate (denoted as Zr-P-O) was prepared by precipitation; other metal phosphates were purchased from Aldrich. Gold was
then loaded onto these supports via depositionprecipitation. It was found
that 200 1C-pretreated Au/M-P-O (M Fe, Co, Y, La, Pr, Nd, Sm, Eu, Ho, Er)
showed high CO conversions below 50 1C (Figure 8.2),30 and 500 1Cpretreated Au/M-P-O (M Ca, Y, La, Pr, Nd, Sm, Eu, Ho, Er) showed high CO
conversions below 100 1C. On the other hand, other Au-M-P-O (M Mg, Al,
Zn, Zr) catalysts were not particularly active in CO oxidation (Figure 8.2).30
This paper showed that metal phosphates other than LaPO4 can be used to
load gold.29 Nevertheless, the gold loadings on dierent supports were different, and further optimization of catalysts is still needed.
Ma et al. studied the influence of preparation methods on the performance of metal phosphate-supported gold catalysts in CO oxidation.31 Several
View Online
161
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
Figure 8.2
metal phosphates, either regarded as active supports (M-P-O, M Fe, La, Eu,
Ho) or inactive supports (M-P-O, M Al, Zn) were picked as typical metal
phosphate supports. Two methods were selected to load gold. In one,
dodecanethiol-capped gold nanoparticles were loaded onto these supports
View Online
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
162
Chapter 8
View Online
163
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
Figure 8.3
Light-o curves (a) and Arrhenius plots (b) for CO oxidation over
Au/LnPO4-MCFs pretreated at 300 1C.32
Reproduced with permission of the Royal Society of Chemistry.
View Online
164
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
8.2.3
Chapter 8
Figure 8.4
View Online
165
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
41
Figure 8.5
View Online
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
166
Chapter 8
proposed by the authors to account for the tandem reaction that involves
alcohol oxidation and a subsequent condensation step (Figure 8.6).45
We can learn from this summary that the research on hydroxyapatitesupported gold catalysts is quite interesting. This kind of catalyst not only
catalyzes CO oxidation, it can also catalyze the conversion of various organic
substrates. The influences of pretreatment conditions, gold cluster sizes,
and organic substrates were investigated in detail, and some reaction
mechanisms were proposed. It is worthwhile to expand the scope of the
research and demonstrate the application of this kind of catalyst in other
reactions.
8.2.4
Metal fluorides are barely used as supports for making gold catalysts.
Tomska-Foralewska and co-workers found that Au/MgF2 was not active for
CO oxidation between 30 and 300 1C.46 Coman and co-workers developed
Au/MgF2x(OH)x catalysts for the one-pot synthesis of menthol.47 The synthesis of menthol from citronellal usually involves two steps: (1) the cyclization of citronellal to isopulegol, and (2) the hydrogenation of isopulegol to
menthol (Figure 8.7).47 Here, MgF2x(OH)x was prepared by the reaction of
Figure 8.6
Figure 8.7
Scheme for the synthesis of menthol from citronellal in two steps (route
1A 2A) and in one pot (route B).47
Reproduced with permission of Wiley-VCH.
View Online
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
167
8.2.5
Some metal sulfates can be used as solid acids, but they are seldom used as
supports to prepare gold catalysts. Dai and co-workers prepared BaSO4-MCFs
supports by dispersing BaO onto MCFs by impregnation, followed by sulfation of BaO using sodium dodecyl-benzenesulfonate (SDBS) as the sulfate
source.48 Highly dispersed BaSO4 nanocrystals were formed during calcination in air at 500 1C. For comparison, BaSO4-MCFs were also prepared using
KHSO4 as the sulfate source. The BaSO4 formed that way is in the form of
isolated rod-like particles with bigger sizes. Gold nanoparticles with small
sizes were highly dispersed on BaSO4-MCFs (SDBS), whereas the sintering of
gold nanoparticles on BaSO4-MCFs (KHSO4) was more obvious. The former
gold catalyst showed higher activity in CO oxidation than the latter
(Figure 8.8).48
Figure 8.8
View Online
168
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
8.2.6
Chapter 8
Figure 8.9
View Online
169
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
8.3 Summary
Catalysis by supported gold nanoparticle has been a hot topic recently. Most
supported gold catalysts have been prepared by loading gold on oxides or
carbon supports, but metal salts have been seldom used as supports to load
gold. Here, we have highlighted the development and applications of gold
nanoparticles supported on metal salts, including metal carbonates, metal
phosphates, hydroxyapatite, hydroxylated fluorides, metal sulfates, and
heteropolyacid salts. Compared with metal oxides, these metal salts are used
less frequently to make supported gold catalysts. Nevertheless, the results
summarized in this chapter have demonstrated that some metal salts can be
good supports to make active gold catalysts. Not only do they catalyze CO
oxidation, they can also catalyze some organic reactions. The latter aspect is
particularly interesting, considering that metal salt-supported gold catalysts
may be bifunctional, i.e. they not only possess the functionality furnished by
gold, they also possess the functionality furnished by supports. Thorough
experiments are still needed in the future to fine tune the functionalities via
changes in the composition of catalysts and preparation details, and the
potential of these catalysts in catalyzing a variety of organic reactions should
be studied in more detail. Reaction mechanisms can also been understood
based on thorough experiments. This will certainly entail a lot of work in the
future, and good advances are sure to come.
Acknowledgments
Z. Ma thanks the National Natural Science Foundation of China (Grant Nos.
21007011 and 21177028) and the overseas returnees start-up research
fund of the Ministry of Education in China for financial support. F. Tao
acknowledges the University of Notre Dame, the U. S. Department of Energy
Basic Energy Sciences Catalysis Science Program and the ACS Petroleum
Research Fund for funding support.
References
1. M. Haruta, T. Kobayashi, H. Sano and N. Yamada, Chem. Lett., 1987, 405.
2. M. Haruta, N. Yamada, T. Kobayashi and S. Iijima, J. Catal., 1989,
115, 301.
3. G. C. Bond, C. Louis and D. T. Thompson, Catalysis by Gold, Imperial
College Press, London, 2006.
4. Z. Ma and S. Dai, Nano Res., 2011, 4, 3.
5. Z. Ma and S. Dai, ACS Catal., 2011, 1, 805.
rrez, S. Hamoudi and K. Belkacemi, Catalysts, 2011, 1, 97.
6. L.-F. Gutie
7. L. Parti and A. Villa, Catalysts, 2012, 2, 24.
View Online
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
170
Chapter 8
View Online
23/06/2014 08:14:59.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00157
171
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
CHAPTER 9
9.1 Introduction
Noble metallic nanoparticles have been used in many applications such as
catalysis,1 sensing,24 drug delivery,5,6 optical switching,7,8 magnetization
switching in magneto-plasmonics,9,10 and the potential diagnosis and in vivo
treatment of cancer.1113 The great versatility of these nanoparticles comes
from the ability to fine-tune their optical,1416 catalytic,17 photothermal,15,1822 photoelectromagnetic,23 and photoacoustic24 properties
with changing their shape and size.
Nanoparticles of various shapes, sizes, and structures have been prepared
to be used in many applications. These shapes are isotropic such as
spheres,17 cubes,25 triangles,26 shells,27 hollow nanospheres,28 nanocages,25
and frames,29 while the anisotropic shapes are rods,30 wires,31,32 and stars.33
Unlike regular synthetic techniques for the synthesis of solid, shaped,
nanoparticles, which are based on the direct reduction of ions into metallic
nanoparticles, hollow nanostructures have been prepared by special
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
172
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
34
173
35
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
174
Chapter 9
View Online
175
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
25
replacement approach, which was introduced by Sun and Xia. This technique involves the galvanic oxidation of two or more atoms of template
nanoparticles by the metal ions of the cage materials, which will be reduced
to metal on the surface of the template. A typical galvanic replacement example is the synthesis of gold nanocages (AuNCs) using silver nanocubes
(AgNCs) as a template. AgNCs are mixed with gold salt in a boiling solution.
The gold ions oxidize the metallic silver and deposit on the nanoparticles
B
A
C
Figure 9.1
(A) SEM image of gold nanocages. (B) TEM image of platinum nanocages. (C) TEM image of palladium nanocages. (D) SEM image of gold
palladium double shell nanocages. (E) Magnified TEM image of
platinumpalladium double shell nanocages. (F) High-resolution
TEM mage of platinumpalladium nanocages. (G) EDX-SEM mapping
of goldplatinum double shell hollow nanoparticles, gold (green) and
platinum (red).
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
176
Chapter 9
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
177
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
178
Figure 9.2
Chapter 9
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
Figure 9.3
179
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
180
Chapter 9
9.4.1
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
181
Ag on their interior walls, which are oxidized to silver oxide (Ag2O) upon
exposure to dissolved O2 in water. Figure 9.4(A) shows the schematic diagram of AuNCs after oxidation of the residual silver into a Ag2O layer.76 Upon
excitation of Ag2O, electrons are excited from the valence band to the conduction band, generating a hole in the valence band. Water molecules in the
solution are oxidized into hydroxyl radicals and protons by the holes in the
valence band, while the electrons in the conduction band reduce water
molecules into hydroxyl ions and peroxide radicals (Figure 9.4(B)). The
generated radicals attack MO molecules, resulting in MO radicals that
undergo a series of intramolecular fragmentations. The hydroxyl radical
generation process is depicted in Figure 9.4(B) and (C).76
There are a number of observations that strongly suggest the nanoreactor
cage eect. The Ag2O molecules that photocatalyze the reaction are present
only on the inner walls of the nanoreactor. The rate of the photocatalytic
reactions was found to be greatly dependent on the pore size in the wall of
the nanocage, due to diusion of reactant in and product out through the
pores of the cage, and the fact that solid Ag cubes have low significant
catalytic activity after exposure to O2 gas. Therefore, the optimum reaction
rate requires: a high surface area of Ag (which is oxidized to Ag2O) on the
inner wall of the AuNC; pores on the surface of the AuNC that are large
enough to allow the reactants and products to diuse in and out of the cavity
but small enough to keep the radical steady state concentration high; and
the cavity inside the cage being an appropriate size to allow an optimum
collision rate between the reactants. When the pore sizes and surface area of
silver oxide are in balance, the cavity can displaying a cage eect.76 This
means that the concentration of the dye molecules, or the reaction rate
determining species formed from them, has built up its concentration to
drive the observed kinetics of the reaction (Figure 9.4(C)). The hollow
structure nanocatalysts and the hybrid semiconductor metallic structures
showed high catalytic activity i.e. rutheniumcuprous sulfide hybrid inorganic nanocages proved to have ecient electrocatalytic properties due to
their high surface area and exciting electronic properties.75 In fact, the
electronic and catalytic properties of the hybrid material nanocatalysts are
Figure 9.4
Schematic diagram of: (A) AuNC with a remaining layer of silver oxide
at the inner wall, (B) the mechanism of photo-formation of the radicals
by silver oxide inside the AuNC nanoreactor, (C) the reaction and the
cage eect inside the AuNCsAg2O nanoreactor.
View Online
182
Chapter 9
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
9.4.2
View Online
183
0.85
AuHSs
AuNSs
10
15
20
25
30
Concentration of nanocatalyst (pM)
6
0
.00
0.75
44
00
0.60
pe=
0.0
0.65
0.
01
0.55
0.
Slo
EY3-/EY2-
op
e=
0.50
0.45
Sl
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
0.70
0.80
0.40
0.35
Figure 9.5
9.4.3
35
View Online
184
Chapter 9
Table 9.1
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
Rate constant at
Nanocatalyst 25 1C (min1)
PdNCs
PtPdNCs
PdPtNCs
PtNCs
0.0190 8.7104
0.0035 1.3104
0.0190 2.0104
0.0036 2.0104
Entropy of
Activation energy activation
Frequency
(cal/mol K1) factor (min1)
(kcal mole1)
22.6 1.5
20.7 1.8
18.5 1.3
16.2 1.1
67.8 5.0
61.4 6.0
50.4 4.2
43.2 3.6
5.101014
2.131013
8.80109
2.31109
and the reactants diuse inside the catalyst. These results prove that catalysis using a hollow nanocatalyst takes place inside the nanocatalyst and the
enhanced activity is due to the inner surface.67
9.4.4
View Online
626 nm
650 nm
0.7
gold nanocages
(AuNCs)
0.6
B
654 nm
185
0.6
0.5
Extinction
Extinction
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
0.5
0.4
0.3
0.2
0.4
0.3
0.2
Pure AuNCs
Immediately after mixing
30 min after mixing
0.1
0.1
0.0
550
Figure 9.6
600
650
700
Wavelength (nm)
750
800
0.0
500
Pure AuPtNCs
Immediately after mixing
30 min after mixing
600
700
800
900
Wavelength (nm)
1000
1100
increases when the second blue shift occurs due to the reaction on the inner
surface and the production of hydrogen gas.74 A single rapid LSPR blue shift
is observed when using a AuPtNC hollow nanocatalyst and the reaction begins at a slow rate, becoming faster after some time. This is consistent with
the presence of one plasmonic surface for AuPtNCs. Based on these results,
it can be deduced that the catalysis reaction on the inner surface is more
ecient than the catalysis on the outer surface of the hollow nanocatalyst
due to the cage nanoreactor eect.74
View Online
186
Chapter 9
2114
10 to 1
5 to 1
1 to 1
1 to 5
1 to 10
0.40
0.20
2038
2140
0.25
2076
0.30
1117
2052
0.35
996
0.45
1046
A 0.50
Absorbance
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
mixture (solutions of HCFIII and PtNCs-TS flowed separately and were mixed
inside the ATR cell). The products of the HCFIII and TS reaction are tetrathionate and hexacyanoferrate(II) (HCFII);102 the bands corresponding to the
0.15
0.10
0.05
2200
2100
2000
1900
1200
Wavenumber
B O3S
S
PtNSs
1100
1000
(cm1)
O S
3
SSO3
+ 3 K3[Fe(CN)6]
Pt[Fe(CN)6]
PtNSs
+ 2K4[Fe(CN)6]
SSO3
- S4O62
+ 2 K3[Fe(CN)6]
PtNSs
SSO3
Pt
O S
3
Figure 9.7
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
187
(1)
(2)
(3)
References
1. P. L. Freund and M. Spiro, J. Phys. Chem., 1985, 89, 1074.
2. R. Jin, G. Wu, Z. Li, C. A. Mirkin and G. C. Schatz, J. Am. Chem. Soc.,
2003, 125, 1643.
3. A. J. Haes, L. Chang, W. L. Klein and R. P. Van Duyne, J. Am. Chem. Soc.,
2005, 127, 2264.
4. C. E. H. Berger, T. A. M. Beumer, R. P. H. Kooyman and J. Greve, Anal.
Chem., 1998, 70, 703.
5. I. Brigger, C. Dubernet and P. Couvreur, Adv. Drug Delivery Rev., 2002,
54, 631.
6. C. Aymonier, U. Schlotterbeck, L. Antonietti, P. Zacharias, R. Thomann,
J. C. Tiller and S. Mecking, Chem. Commun., 2002, 3018.
7. K. F. MacDonald, Z. L. Samson, M. I. Stockman and N. I. Zheludev, Nat.
Photonics, 2009, 3, 55.
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
188
Chapter 9
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
189
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
190
Chapter 9
View Online
23/06/2014 08:15:00.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00172
191
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
CHAPTER 10
Nanoreactor Catalysis
KYU BUM HAN, CURTIS TAKAGI AND AGNES OSTAFIN*
University of Utah Nano Institute, Department of Chemical Engineering,
36 S Wasatch Drive, Salt Lake City UT 84112
*Email: [email protected]
10.1 Introduction
A catalyst is a substrate that accelerates chemical reactions by reducing their
activation energy. The nanoscale catalyst has expanded the eciencies of
dierent industries via its greater active surface area, controllable physical
morphology and properties (via chemical approaches), and improved stability. Examples of successful nanocatalysts include dendrimers, polymer
microspheres and coreshell structures, microgels, liposomes, emulsions,
micelles and carbon nanotubes. Many of these materials contain protected
catalytic environments and so fall into the category of nanoreactor catalysts.
The nanocatalyst market has grown by an average 6.3% since 2004; the
global market has risen from US$ 3.7 billion in 2004 to US$ 5.0 billion in
2009.1 Petrochemicals, one of the major markets, is a US$ 742 million2
global industry that focuses on obtaining premium quality fuel and improving selectivity control. Refineries previously used lead and benzene to
increase the octane number of gasoline, but these materials were environmentally unfriendly. By adding nanoplatinum as a reforming catalyst with
other metals, such as tin or rhenium, higher octane can be achieved without
these additives. Zeolites are an early form of nanoreactor catalyst with customized cage-like structures B1 nm in size. The method for customizing the
structure was developed at Mobil in 1977. Methanol is reacted on aluminosilicate-based zeolite HZSM-5 to produce a mixture of hydrocarbons for
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
192
View Online
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
Nanoreactor Catalysis
193
Dendrimers
View Online
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
194
Chapter 10
between the p-electrons of the aromatic rings in the polymer, and vacant
orbitals of the metal species.17 Choosing the correct polymer functional
groups, with ionic or hydrogen bonds, for the interior regions of the
dendrimer, attracts and stabilizes the metal cation complexes in a relatively
high concentration in the interior of the dendrimer. For example, Ooe et al.
constructed a dendrimer system with interior amino groups that create a
polar environment around a Pd complex, which leads to higher reactivity
rates.10 Careful manipulation of the dendrimer nanoreactors has often
yielded higher catalytic eciencies than the normal monomeric catalysts.
This is because the catalytic site can force the reagents interacting with the
catalysts into a specific configuration that is similar to their reaction transition state, in a phenomenon which is considered to be similar to biological
enzyme mechanisms. This allows for increased selectivity in reactions
and higher eciencies, especially in reactions with two enantiomeric
products.12,18,19
One of the most promising features of the dendrimeric catalyst nanoreactor systems is the customizability of the dendrimer molecules and, in
eect, the properties of the reaction space and exterior periphery. The first
factor that must be considered is the number of generations, or layers, that
have been instilled in the structure. Dendrimers of high generations can be
so sterically crowded at the edges that they act as filter membranes. Inside,
the large open pores remain, but the exterior layers of the polymer itself
can easily control access to these ports.5,13,20,21 Control over the diusion
of species can be accomplished by keeping the dendrimer surface open
with dierent functional groups. Simple steps, such as protonation or
deprotonation of terminal groups, can open and close the pores at the
surface, and these can be tuned to only allow the diusion of specific reactants or products.5,20,21 Very large pores allow for greater mass transfer, but at
the same time the catalytic pockets can become saturated with reactants and
products, and the diusion kinetics are hindered.22,23 However, if the pores
are too small, then of course, there is not enough opportunity to bring in new
reagents while releasing the product. This is why the pore sizes on the surface
of the dendrimer need to be optimized, to optimize the catalytic properties.
Along with pore size, the peripheral molecules can also control the
solubility properties of the particle. For example, it is possible to have a polar
interior with a non-polar surface; this allows the system to incorporate
itself into an organic environment but catalyze reactions with polar
properties.24,25 This property allows for unconventional designs and ability,
and also it has improved the recoverability of the catalysts, leading to a
greener industry with higher recyclability
10.2.2
Microgels
View Online
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
Nanoreactor Catalysis
195
dendrimer counterparts. Microgels are very stable and highly customizable.26 Like the dendrimer systems, metallic nanoparticles are often embedded in the structures,2730 and access to these sites is controlled by other
external factors. Microgels can expand and contract with temperature
changes, or as they gain or lose water. This opens and closes the pores,
allowing for selectively and responsiveness. For example, poly(N-isopropylacrylamide) (PNIPA) is soluble in water below 32 1C, due to the formation of hydrogen bonds; however, above this temperature the hydrogen
bonds are disrupted, the water escapes and the gel collapses.3135 Pich et al.
have demonstrated the capability of the microgel-supported catalyst by
embedding gold nanoparticles into a microgel system. They found that the
confined catalyst increased eciency to 100% by preventing aggregation of
the particles, and allowing for greater mass transport.30 Because of the high
water content that is possible in microgels, enzymes can also be placed inside rather than the metallic nanoparticle.36,37 This has been demonstrated
with glucose oxidase, which did not lose its catalytic activity when placed in a
water-saturated microgel system. This discovery could have huge industrial
implications where catalysis is needed, because natural enzymes are often
the most ecient and selective catalysts, but their complexity makes it difficult to synthetically duplicate.
10.2.3
10.2.4
View Online
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
196
Chapter 10
View Online
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
Nanoreactor Catalysis
197
View Online
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
198
Chapter 10
complex. Attempts have been made to copy them by taking a molecule in the
transition state and stamping it into a polymer; the idea being that the most
energetic parts will be identified, and leave behind a stabilizing molecule on
the mold polymer. However, this seems very dicult and isnt as eective as
some metal nanoparticle based catalysts.55 In order to fix this problem,
channel proteins or proton pumps were placed in the bilayer so that substrate molecules could easily diuse in and out.56,57 Further, an easier synthesis method has been developed by incorporating an additional block
copolymer, which intrinsically provided the porous bilayer. Later, researchers improved the polymersomes as multifunctional nanocatalysts,
which allowed biochemical reactions not only inside the polymersomes but
also on the surface.5860
10.3.2
Carbon Nanotubes
View Online
Nanoreactor Catalysis
199
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
71,72
between the particle and interior surface than when on the outer surface.
For the control of metal nanoparticle catalysts, the chemical and catalytic
properties of atoms at corners or edges is dierent. So the way the CNT
(also applies to polymer-based structures) attaches may change the catalytic
properties.73 The walls of the CNT are polarized, which may seem odd for a
network of carboncarbon single bonds, but it is this dipolar interaction
between the CNT interior and the embedded catalytic species that can stabilize a transition state and promote more ecient product formation. The
Menshutkin reaction, an SN2 mechanism, showed significant increase in
reaction rate when conducted within a CNT, evidenced by a reduction in the
endothermicity and reaction energy barrier.74
10.4 Conclusion
The nanoscale catalyst has been developed within the last couple of decades,
in the hopes of making industrial processes cheaper, greener, and more
ecient. The nanospace inside nanoreactor structures accelerated chemical
and biochemical reactions due to increased concentrations; it also improved
the selectivity of reactions due to the dierent mechanisms of reagent isolation and selection. However, it is not necessary for these catalysts to be
confined to produce higher eciency rates; this can also be achieved by
increasing the surface area or by stabilizing the catalytic species, such as in
dendrimers and other stabilization procedures. More complex systems,
where a cascading eect that utilizes two or more catalysts, are currently
being developed, mainly for biological scenarios where an enzymatic chain
is desirable. The scale at which it is now possible to isolate and control
environments allows for greater eciency with less chemical waste, thus
allowing industries to expand their capability while maintaining the state of
the environment. Carbon nanotubes, dendrimers, polymers, and emulsion
systems, all work to exploit the dierences that occur at small scales. If
successful, the way certain industries operate can change dramatically, and
the desire to control various chemical reactions can be realized.
References
1. Nanotechnology: nanocatalysts, BCC Inc. Research Market Forecasting,
2009.
2. Global Nanocatalysts Market to Reach $6.0 Billion by 2015, According to
New Report by Global Industry Analysts, Inc., in Nanotechnology Now,
2009 (cited 2011 17, Dec.); available from: http://www.nanotech-now.
com/news.cgi?story_id 32763.
3. C. Chang and A. Silvestri, J. Catal., 1977, 47, 249259.
4. C. Chang, J. Catal. Rev., 1983, 25, 1118.
5. M. Zhao and R. M. Crooks, Angew. Chem., Int. Ed., 1999, 38(3), 364366.
6. M. Zhao and R. M. Crooks, Adv. Mater., 1999, 11(3), 217220.
7. Y. Niu, L. Yeung and R. Crooks, J. Am. Chem. Soc., 2001, 123, 68406846.
View Online
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
200
Chapter 10
View Online
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
Nanoreactor Catalysis
201
View Online
23/06/2014 08:15:01.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00192
202
Chapter 10
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
CHAPTER 11
203
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
204
Chapter 11
11.1 History
There are many reports of clock reactions in the field of chemistry describing
the reversible pathway of a process that takes place periodically. The following are popular demonstrations.
11.1.1
In 1886, Hans Heinrich Landolt reported the clock reaction between bisulfite
and excess iodate known as the iodine clock reaction. Here, two colorless
solutions are mixed and left for some time. After a short time lag, the colorless solution gradually turns dark blue. The reaction starts with an easily
available common reagent such as potassium iodate and an acidic solution
of sodium bisulfite (acidified with sulfuric acid) in the presence of starch.
In this protocol, an iodide ion is generated by the following slow reaction
between the iodate and bisulfite:
IO3 3HSO3-I 3HSO4
This is the rate determining step. Excess iodate will oxidize the iodide
generated in the last step to form iodine:
IO3 5I 6H1-3I2 3H2O
However, the iodine is reduced immediately back to iodide by the
bisulfite:
I2 HSO3 H2O-2I HSO4 2H1
When the bisulfite is fully consumed, the iodine will survive to form the
dark blue complex with starch.
The solution will repeatedly cycle from colorless to blue and back to colorless, until the reagents are depleted.1
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
11.1.2
205
BZ Reaction
11.1.3
BrayLiebhafsky Reaction
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
206
Chapter 11
Figure 11.1
11.1.4
BriggsRauscher Reaction
In 1972, the BriggsRauscher reaction was discovered by replacing the bromate in the BZ reaction with iodate and hydrogen peroxide. Optionally,
starch is used as an indicator for the abrupt change in iodide ion concentration. This reaction is very important in demonstrating the purpose of an
oscillating chemical reaction in which a colorless solution is gradually
transformed to an amber solution and is suddenly changed into a dark blue
solution.
Initially, the aqueous solution consists of hydrogen peroxide, an iodate,
divalent manganese (Mn21) as a catalyst, a strong chemically unreactive acid
such as sulfuric acid (H2SO4) or perchloric acid (HClO4), and an organic
compound such as malonic acid with an active hydrogen atom attached to
carbon which will slowly reduce free iodine (I2) to iodide (I). The reaction
shows habitual cyclic changes, both gradual and sudden, which are visibly
slow changes in the intensity of color, interrupted by abrupt changes in
color. This demonstrates that a dicult arrangement of slow and fast reactions is taking place simultaneously. In the slow process, free iodine is
consumed by malonic acid involving an intermediate species, iodate.
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
207
11.1.5
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
208
Chapter 11
absorption spectra in the visible region. Thus, a crowd-pleasing demonstration can easily be achieved from coinage metals, and the best of these is
the silver nanoparticle redox reaction. The Plasmon absorption, i.e. the coherent oscillation of the valence electron with the impinging electromagnetic radiation interacting with metal nanoparticle surfaces, is easily
demonstrated taking silver into consideration. The absorption maximum
due to stabilized metal nanoparticles varies with particle size, shape and also
with the dielectric medium that stabilizes the nanoparticles. The metal
nanoparticles become progressively reactive as the size decreases. Henglein
described a pushpull reduction method for the preparation of plausibly
monodisperse silver particles in aqueous solution.13 Slow particle growth is
achieved in this method by simultaneously reducing Ag1 ions and partially
oxidizing Ag particles. In particular, a narrow surface Plasmon band at
382 nm was generated for the formation of Ag(0) particles in the solution.
In this investigation, the particles used were prepared by the fast reduction
of Ag1 ions and they had a broader Plasmon absorption band. Figure 11.2
shows the absorption spectrum before and after addition of two concentrations of iodide to the silver sol. It can be seen that the absorption band is
broadened and slightly red-shifted.
Figure 11.2
Absorption spectrum before and after the addition of various concentrations of KI. At [KI] 10 pM, full coverage of the surface of the
particles was reached, as a further increase in the KI concentration
did not lead to additional changes in the shape of the absorption
band; the CTTS absorption band of free I- appeared at [KI]410 pM.
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
209
Below a critical size, the coinage metal nanoparticles also show featureless
absorption curves in the visible region like platinum, palladium etc. These
particles (sub-nanometre particles) are called metallic clusters.14 It is very
dicult to deal with them in solution because of their pronounced reactivity,
identification and stabilization problems. It is also a dicult task to obtain
catalytic activity from these sub-nanometre particles because the easy access
to their surfaces is lost once they are stabilized by a capping agent.
Metal hydrosol bears a negative charge, which is the manifestation of the
capping agent present on the nanoparticles surface. A capping agent may be
replaced by a place exchange reaction with another suitable capping agent.
Still, they may remain as charged particles. Interestingly, the capping agent
may be replaced by neutral long chain amines, thiols, proteins etc. The metal
nanoparticles remain stable and retain their identity with these capping
agents. Then the particles are said to be sterically stabilized and the phenomenon is known as steric stabilization (Figure 11.3).
To elaborate on the fascinating reversible reaction, dilute AgNO3 solution
(104105 M) may be placed in a test tube as a representative coinage metal
salt. Dilute ice cold NaBH4 (0.01 M) has to be introduced into the test tube
with vigorous stirring. A transparent and stable yellow silver hydrosol would
be obtained.15 The sol particles as usual would bear a negative charge, which
makes the silver particles stable due to Coulombs interaction (Figure 11.4).
The maximum absorption of the solution would appear at B400 nm.
Excess BH4 would remain as adsorbed ions, i.e. the adsorbate BH4 on the
silver nanoparticles. Before the decomposition of BH4 ions, if some surfactant molecules are introduced, a dramatic reversible redox reaction may
be observed. Upon shaking the yellow solution, the color completely disappears and then reappears, keeping the solution for some time. The whole
operation may be summarized as a crowd-pleasing demonstration: NaBH4
reduces AgNO3 into small silver nanoparticles. The particles are in the
nanodimension (o50 nm) and they impart a yellow color (Plasmon absorption) to the solution. Excess BH4, more diusive than NO3 or OH,
would adhere onto the silver nanoparticle surfaces. This would shift the
Fermi level towards a more negative region (Figure 11.5).
Thus oxygen in the air would easily oxidize the silver particles surface.
Surfactant in the reaction medium would dissolve out the oxide layer
Figure 11.3
Steric stabilization.
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
210
Chapter 11
Figure 11.4
Electrostatic stabilization.
Figure 11.5
successively from the silver particles surface. The oxidation and surfactantassisted removal of the oxide layer would take place after the oxidation of the
silver surface by air, but in the presence of a strong nucleophile, such as
BH4 (in the present case) or a cyanide ion (in the other). As a result of
oxidation, the silver nanoparticles would be smaller (more reactive) and finally would dissolve/disappear upon shaking. Oxygen in the air assists the
oxidation. Now, the unstirred colorless solution with argento cyanide, still
with excess NaBH4, would again evolve silver nanoparticles. This reversible
cycle would go on and on, just shaking and standing for oxidation (disappearance of the yellow color) and reduction (re-appearance of the yellow
color), respectively, for a crowd-pleasing demonstration (Figure 11.6). This
can even be demonstrated with Au(III) chloride.16
11.2.1
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
Figure 11.6
211
and reduction processes makes the clock reaction of methylene blue important in chemistry and biological systems.
Snehalatha et al.17 demonstrated the clock reaction of methylene blue
using ascorbic acid as a reducing agent. Here, a reduction was performed in
the presence of dilute sulfuric acid. It was observed that with an increase in
acid concentration, the time of reaction decreased, that is, the rate of the
reduction gradually increased. The clear and visible color change made the
reaction simple and so the mechanism of the reaction is easy to explain.
But in the absence of a catalyst, the reduction process becomes time
consuming. So, a methylene blue-ascorbate ion redox reaction has been
catalyzed by a micellar system in which catalysis occurred by the binding of
the substrates onto the micellar surface by hydrophobic and electrostatic
interactions. The rate of the reaction increases with an increase in the
possibility of encounters. In this case, the eect of various salting-in and
salting-out agents, size of reverse micellar water pool, and hydrophobicity of
the dye were studied and no change in reaction rate was found. The reversible electron-transfer reactions for the MB-AA- and MB-BH4- systems
become facile in the presence of coinage metal sol particles as catalysts
(Figure 11.7).18
Additionally, it has been shown that Au nanocrystals act as an eective
catalyst for the reduction of nitrobenzene to aniline and methylene blue to
leucomethylene blue in the presence of NaBH4 and N2H4, respectively, as
reducing agents (Figure 11.8).19
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
212
Figure 11.7
Chapter 11
Only a few noble metals in the elemental form are present in our Earths
crust. However, most metals combine with oxygen and form a binary
compoundan oxide. Oxides are robust and have proved to be excellent
catalysts. Proper tuning of an oxide structure might develop a good semiconductor. Then the semiconductor may be useful for scavenging solar light.
The bottle neck with most oxide materials is their tendency to be activated
under ultraviolet light. Thus, there is a necessity for tuning the size, shape
and/or doping the oxides to make them visible light photocatalysts. The
other aspect of oxide materials is their usefulness in chemical reactions. In
the present context, it may be shown that metallic oxides also exhibit reversible redox reactions.
The popular Fehling test for aldehyde detection and quantification results
in precipitating cuprous oxide, Cu2O. An orecuprite or red copper oxide,
Cu2Ooccurs in nature. This is an unique monovalent simple compound of
copper. In the bulk stage, red Cu2O gets stabilized as a robust crystal.
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
Figure 11.8
213
One can restrict the growth of Cu2O in solution and in the nanodimension,
Cu2O remains as a yellow solid, which has been exploited as a catalyst to
exhibit reversible clock reactions.3 One challenge is to take bulk Cu2O to a
nanoregime and to stabilize it in solution. Cu2O in its nanoregime would act
as a photocatalyst. We report a new synthetic protocol to obtain monodisperse, stable, exclusively cubic Cu2O nanoparticles under surfactant-free
conditions and its catalytic action in the methylene blue (MB)-hydrazine
reaction in an aqueous medium. The blue color of the dye, MB, faded away
upon the addition of hydrazine, producing colorless leucomethylene blue
(LMB), indicating the progress of the redox reaction. The rate of this redox
reaction has been found to be enhanced in the presence of the nanocatalyst,
Cu2O. The success of the reaction demonstrates a simple clock reaction. An
oscillation between a blue MB color and colorless solution due to the formation of LMB is observed on periodic shaking. This oscillation continues
for over 15 cycles. Studies on the eect of bulk Cu2O and nanoparticles of
CuO and Cu(0) have not been successful for the demonstration of the clock
reaction. Thus, the importance of Cu2O nanoparticles in the clock reaction
is established beyond doubt. The electrochemical studies indicate that nanoCu2O shows a couple of redox peaks, which correspond to the redox Cu(II)
Cu(I) system. Kinetic studies authenticate a first-order reaction mechanism.
Further, quantum chemical calculations reveal that the nanoparticles reduce
the activation energy by B17 kcal mol1, thereby making the reaction
2.4107 times faster compared to the gas phase (Figure 11.9).20
Truncated Cu2O cubes have been successfully synthesized using a stable
Cu(II)-EDTA precursor and glucose as the reducing agent under alkaline
conditions and five minutes of microwave irradiation. Interestingly, in the
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
214
Figure 11.9
Chapter 11
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
Figure 11.10
215
11.3.1
EleyRideal Mechanism
Here, one of the reactants (A), generally a reducing agent, is adsorbed on the
catalysts surface and reacts with another reactant (B) present in the system.
Then the newly formed product desorbs from the catalysts surface
(Figure 11.11).22
A S (support)"AS
AS B-Product (P)
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
216
Chapter 11
Figure 11.11
Figure 11.12
11.3.2
LangmuirHinshelwood Mechanism
In this case, reducing agent (A) adsorbed on the catalystss surface, reacts
with the surface and transfers surface-hydrogen species to the catalysts
surface. Again, another reactant (B) is also adsorbed on the catalysts surface.
Both steps are assumed to be fast and also termed as Langmuir isotherm.
Now, reactant B reacts with the adsorbed surface-hydrogen species and is
reduced. After completion of the reaction, the new product apart from
catalyst and again the vacant catalyst surface is ready for repetitive reduction
cycles (Figure 11.12).23
A Support (S)"AS
B Support (S)"BS
AS BS-Product (P)
11.4 Applications
The clock reaction of methylene blue is used in various applications such as:
11.4.1
Water Purification
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
11.4.2
217
11.4.3
11.5 Conclusion
In this chapter, we have described various types of clock reaction, which is
mainly controlled by a reversible redox mediated reaction. In most cases, the
oxidizing agent is a dye or colored molecule, which is reduced by a general
reducing agent; in this context, two reaction mechanisms are discussed in
the preceding section. To make the reduction process facile, various nanocatalysts are used. An investigation into the clock reaction with new catalysts
or dye molecules will be very fascinating.
References
1. H. Landolt, Ber. Dtsch. Chem. Ges., 1886, 19, 1317.
2. B. P. Belousov, Collection of Abstracts on Radiation Medicine, 1959,
147, 145.
3. A. M. Zhabotinsky, Biophysics, 1964, 9, 306.
4. R. J. Field, E. Koros and R. M. Noyes, J. Am. Chem. Soc., 1972, 94, 8649.
5. R. J. Field and R. M. Noyes, J. Am. Chem. Soc., 1974, 96, 2001.
6. I. R. Epstein and K. Showalter, J. Phys. Chem., 1996, 100, 13132.
7. Z. Nagy-Ungvarai, J. J. Tyson and B. Hess, J. Phys. Chem., 1989, 93, 707.
8. V. K. Vanag, L. F. Yang, M. Dolnik, A. M. Zhabotinsky and I. R. Epstein,
Nature, 2000, 406, 389.
9. W. C. Bray, J. Am. Chem. Soc., 1921, 43, 1262.
10. W. C. Bray and H. A. Liebhafsky, J. Am. Chem. Soc., 1931, 53, 38.
11. T. S. Briggs and W. C. Rauscher, J. Chem. Educ., 1973, 50, 496.
View Online
23/06/2014 08:15:04.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00203
218
Chapter 11
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
CHAPTER 12
12.1 Introduction
Catalysis has been essential to life and as a frontier in science. Transportation
fuels, lubricants, chlorine-free refrigerants, high-strength polymers, stainresistant fibers, drugs, and many thousands of other products required by
modern societies would not be possible without the existence of catalysts.1
Most catalysts consist of nanometre-sized particles dispersed on a highsurface-area support. Metal nanocatalysts have a high surface-to-volume ratio
and very active surface atoms, and are very attractive for a wide variety of
reactions, compared to bulk catalysts.26 The nature of the active phase(s) in
these metal/support catalysts and the reaction mechanism are still unclear.
What are the eects of size, shape and support on the catalytic properties? It
is extremely dicult to obtain such necessary details from current experimental techniques, which has been an obstacle in understanding their
structure-dependent catalytic activity and selectivity and therefore hinders
further catalyst development.
Over the past few decades, advances in methodology, software and the
power of supercomputers have made computational approaches, specifically
density functional theory (DFT), capable of providing qualitative, and in
many cases quantitative, insights into catalysis.2,3,718 In this chapter, we will
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
219
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
220
Chapter 12
report our recent DFT studies on metal and supported metal nanocatalysts.
In section 2, DFT and DFT codes that we used will briefly be introduced.
Section 3 will address the DFT studies of catalysis on unsupported metal
nanocatalysts. The eect of support on the catalytic behavior of metal
nanocatalysts will be discussed in section 4. As you will see in this chapter,
theoretical studies play a very important role. It describes catalytic reactions
on nanocatalysts with the detail and accuracy required for computational
results to compare with experiments in a meaningful way. More importantly,
it gives insights into the size and shape eects of nanocatalysts on catalyst
activity and stability. Such understanding can lead to the rational design of
better catalysts.
View Online
221
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
12.3.1
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
222
Chapter 12
Figure 12.1
DFT calculated energy profiles for the WGS on Cu29 (red) and Cu(100)
(black). Thin bar: intermediate; thick bar: transition state. Inset: CO,
H and OH adsorbed on Cu29. Brown: Cu; red: O; grey: C; white: H.
Adsorbed species are denoted by asterisks (*).
The figure is reproduced from ref. 3 and ref. 37.
Figure 12.2
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
223
contributes to strengthening the bindings of the intermediates and transition states with the catalyst (Figure 12.1). Given that, the bottleneck water
dissociation becomes more facile when going from extended surfaces to
nanoparticles, which facilitates the overall WGS reaction.
The importance of low-coordinated sites of metal nanocatalysts was also
observed for other catalytic processes: methanol synthesis,39 CO oxidation,40
ammonia synthesis,41 methanation42 etc. Turnover rates and selectivity of
catalytic reactions can be influenced strongly by the size and shape of
nanoparticles, as their surface structures and active sites can be tailored at
the molecular level. For these reactions, the bond breaking of reactants or
reaction intermediates is the most dicult step that hinders the overall
conversion. Metal nanocatalysts with high surface areas can provide the sites
with the enhanced ability to chemisorb and activate bond cleavage as the
coordination number is reduced at corners and step edges. Consequently,
the reaction pathways leading to product yield and selectivity are in turn
aected. In this case, the catalytic activity is promoted when the size of
catalysts decreases to the nanoscale.
However, for the reaction that is hindered by the bond formation to
produce products or the removal of the products from the surface, the
strengthened interaction associated with the low-coordinated sites of metal
nanoparticles is likely to slow down the overall conversion. As you will see in
the following, the adsorbates may either stay on the surface as spectators
and poison the low-coordinated sites, or destabilize these sites and therefore
the nanocatalysts. In both cases, catalyst degradation with time and decreased activity are expected. That is, nano-sized catalysts may not always be
the choice of catalyst to achieve high performance.
12.3.2
Fuel cells are now widely accepted as the next step in power generation for
many applications including commercial power, automotive transportation
and portable electronic devices. Pt undoubtedly is the best elemental catalyst
in fuel cells. However, there are two main drawbacks that hinder the widespread use of fuel cells. One is the high cost of Pt; the other is the sluggish
kinetics, in particular for the cathodic reaction at low temperatures, i.e. the
oxygen reduction reaction (ORR), even on Pt.43,44 Large eorts have been
spent searching for non-Pt electrocatalysts to lower the cost while maintaining high ORR activity; however, the inadequate activity and stability remain a major obstacle.45,46 Recent studies show that coreshell
nanocatalysts consisting of a single-layered shell of Pt atoms on cores of
metal- or alloy- nanoparticles are able to achieve higher activity and stability
for ORR compared to pure Pt.6,4752 The monolayer concept facilitates
an ultra-low Pt content, high Pt utilization, and the possibility of tuning
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
224
Chapter 12
the activity of the PtML shell through interactions with the supporting
nanoparticle cores.
To understand the enhanced activity and stability of the coreshell
nanocatalysts under the ORR condition, we developed a simplified model to
describe the nanoparticles in real size, which allows us to gain a better
understanding of the eects of size and shape on the stability and activity
under the ORR conditions. In our calculations, assuming that the shape of
the nanoparticle is symmetric, we constructed a semi-sphere-like nanoparticle model (Figure 12.3) to reduce the computational cost, considering
half of the nanoparticle should be representative for the whole. Dierent
from the slab model,5359 relaxations were allowed in both out-of-plane and
in-plane directions for the atoms, while only the Z directions of the atoms on
the bottom layer were fixed to keep the symmetry of the particle. As shown,
this simplified model is able to capture the surface contraction associated
with the particle size and shape with an aordable cost, which was not included in the previous slab model,5357 and aects the ORR activity and
stability significantly. As you will see, using this particle model, our DFT
calculations are capable of describing the experimental results on welldefined coreshell nanoparticles well. Note that for the ORR, the lowcoordinated sites of Pt-based nanoparticles are not active and are blocked by
the O-containing species due to the strong OPt interaction. Accordingly,
only the sites on the terrace, which bind O moderately, are considered for
describing the ORR activity.
(A)
Size Eect
Figure 12.3
Schematic illustration of surface models for (a) pure Pt, (b) solid Pd
core and PtML shell (PdcPt1), and (c) partially hollow Pd core and PtML
shell (hPd1Pt1). c, s, and 1 correspond to core, shell, and
1 monolayer (ML), respectively.
The figure is reproduced from ref. 47.
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
225
in-plane contraction decreases with increasing particle size, the corresponding eect on the binding energy for oxygen (BE-O), a descriptor for the
ORR activity,56,60,61 cannot be ignored for particles of at least up to 5 nm. For
Pd(core)Pt(shell) nanocatalysts, the size and lattice mismatch (0.9%)
jointly determine the strain that aects the BE-O on the (111) facets.6 Considering the fact that the intrinsic ORR activity is largely determined by the
BE-O,5355 one can see that the particle size aects not only the fraction of
active sites at the surface, but also the turnover frequency (TOF) per active
site. The product of the two determines the specific activity. A previous study
of the ORR on transition metal alloys suggested that the best ORR electrocatalysts may bind oxygen more weakly, by about 0.2 eV,55 than Pt(111) does,
i.e. corresponding to 3.89 eV in our calculations (dashed line in
Figure 12.4). This value is reached on 1.8-nm Pt particles solely by the
nanosize-induced surface contraction. Larger particle sizes for optimal BE-O
were found using the 3.4-nm PdCPt3 and 4.4-nm PdCPt2 models, where a
reduced nanosize eect is compensated by the increased influence of the Pd
core. Therefore, the enhancements in specific activity are largely attributed
to the eect of nanosize-induced surface contraction on facet-dependent
BE-O. In accordance with the experiment, the DFT calculations suggest that
the moderately compressed (111) facet is most conducive to the ORR on
small nanoparticles. Our results indicate the importance of concerted
structure and component optimization for enhancing coreshell nanocatalysts activity and stability.
DFT calculations using the particle model were also performed to elucidate the observed high stability of Pd(core)Pt(shell) nanoparticles.47 Three
nanoparticle systems with dierent sizes were considered in our calculations
(Figure 12.3): pure Pt, a Pt monolayer shell with a solid Pd core (PdCPt1), and
a Pt monolayer shell with only a 1 ML Pd inner core (hPd1Pt1). The extended
Figure 12.4
Binding energy for oxygen on dierent sites calculated using the slab
model for extended surfaces (as a reference), and the nanoparticle
model for various coreshell nanoparticles.
The figure is reproduced from ref. 6.
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
226
Chapter 12
Figure 12.5
DFT-predicted dissolution potentials of a 1 ML Pt shell from nanoparticles and extended surfaces of Pt, PdCPt1, and hPd1Pt1 as a
function of particle sizes.
The figure is reproduced from ref. 47.
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
Figure 12.6
227
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
228
Chapter 12
support. Now, the question is: what role does the support play in the catalytic
performance of a catalyst? Tauster et al. used the term strong metal/support
interaction (SMSI) to describe the drastic changes that occur in the
chemisorption properties of Pt and other group VIII (810) metals when they
are dispersed on the surfaces of titanium oxide.63,64 The spreading of TiOx
aggregates on the surfaces of the supported metals could be responsible for
the reduction in their chemical and catalytic activity.64,65 By combining DFT
and kinetic modelling, our theoretical studies show a completely dierent
type of strong metal/support interaction, where supports substantially
enhance the catalytic activity of metal nanoparticles for many reactions,
e.g. the WGS reaction,10,35,66 desulfurization,8,67 CO oxidation68 and CO2
activation.69 In the following, we will show two examples to illustrate the
promoting eect of supports.
Associated with SMSI, supports can not only stablize the small metal
nanoparticles that are unstable in the gas phase, they can also produce large
electronic perturbations for the supported small particles and significantly
enhance the catalytic activity. This is the case for the WGS reaction on
Pt/CeO2.66 It is observed experimentally that upon adding Pt to CeO2(111),
there is a continuous increase in the catalytic activity until a maximum is
reached at a coverage of B0.2 ML. Comparing with the STM results reported
for the Pt/CeO2(111),70 one can conclude that the maximum activity corresponds to catalysts that contain Pt particles with a diameter smaller than
1.7 nm and a height below 0.4 nm. According to our DFT calculations, Pt(111)
has problems in adsorbing and dissociating the water molecule (Figure 12.7).
Figure 12.7
DFT calculated energy profile and the geometries for the dissociation
on Pt(111), Pt79 and Pt8 clusters, and Pt8 on CeO2. Blue: Pt; white: H;
red: O; cream: Ce.
The figure is reproduced from ref. 66.
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
229
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
230
Figure 12.8
Chapter 12
reaction. Our results show that the relatively inert gold nanoparticles can
become very good catalysts when combined with an over-reactive TiO2
surface.
Overall, supports can play an essential role in promoting the catalytic
performance of metal nanoparticles via SMSI. The support is able to stabilize
the small metal nanoparticles, which are catalytically active, by hindering
diusion and coalescence under the reaction conditions. In addition, it can
modify the electronic structure of the supported small particles and therefore promote their catalytic activities. Finally, it can participate in the reaction directly by facilitating the steps that are activated on the supported
particles and therefore facilitate the overall reaction. Such understanding is
the key for the further development of metal/support catalysts.
12.5 Conclusions
The powerful new computing infrastructures are starting to close the gap
between theoretical and experimental studies in the field of nanocatalysis.
Our theoretical studies on unsupported and supported metal nanocatalysts
provide a molecular-scale picture of the mechanisms underlying catalytic
activity. Firstly, the low-coordinated sites at step edges are essential for metal
nanocatalysts, which behave dierently from those on terraces. These lowcoordinated sites can be the only active sites for the reaction and influence
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
231
Acknowledgements
The research is supported by the US Department of Energy, Division of
Chemical Sciences, under contract DEAC02-98CH10886.
References
1. A. T. Bell, Science, 2003, 299, 1688.
2. J. A. Rodriguez, P. Liu, J. Hrbek, J. Evans and M. Perez, Angew. Chem., Int.
Ed., 2007, 46, 1329.
3. P. Liu and J. A. Rodriguez, J. Chem. Phys., 2007, 126, 164705.
4. L. Barrio, P. Liu, J. A. Rodriguez, J. M. Campos-Martin and J. L. G. Fierro,
J. Phys. Chem. C, 2007, 111, 19001.
5. L. Barrio, P. Liu, J. A. Rodriguez, J. M. Campos-Martin and J. L. G. Fierro,
J. Chem. Phys., 2006, 125, 164715.
6. J. X. Wang, H. Inada, L. Wu, Y. Zhu, Y. Choi, P. Liu, W.-P. Zhou and
R. R. Adzic, J. Am. Chem. Soc., 2009, 131, 17298.
rez, Science,
7. J. A. Rodriguez, S. Ma, P. Liu, J. Hrbek, J. Evans and M. Pe
2007, 318, 1757.
es, F. Illas, Y. Takahashi and
8. J. A. Rodriguez, P. Liu, J. F. Vin
K. Nakamura, Angew. Chem., Int. Ed., 2008, 47, 6685.
9. A. Kowal, M. Li, M. Shao, K. Sasaki, M. B. Vukmirovic, J. Zhang,
N. S. Marinkovic, P. Liu, A. I. Frenkel and R. R. Adzic, Nat. Mater., 2009,
8, 325.
10. J. A. Rodrguez, J. Evans, J. Graciani, J. Park, P. Liu, J. Hrbek and
J. F. Sanz, J. Phys. Chem. C, 2009, 113, 7364.
11. J. B. Park, J. Graciani, J. Evans, D. Stacchiola, S. Ma, P. Liu, A. Nambu,
J. F. Sanz, J. Hrbek and J. A. Rodriguez, Proc. Natl. Acad. Sci., 2009,
106, 4975.
es and F. Illas,
12. J. A. Rodriguez, P. Liu, Y. Takahashi, K. Nakamura, F. Vin
J. Am. Chem. Soc., 2009, 131, 8595.
13. J. X. Wang, H. Inada, L. Wu, Y. Zhu, Y. Choi, P. Liu, W. P. Zhou and
R. R. Adzic, J. Am. Chem. Soc., 2009, 131, 17298.
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
232
Chapter 12
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
233
View Online
23/06/2014 08:15:05.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00219
234
Chapter 12
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
CHAPTER 13
Porous Cryptomelane-type
Manganese Oxide Octahedral
Molecular Sieves (OMS-2);
Synthesis, Characterization
and Applications in Catalysis
SAMINDA DHARMARATHNA AND S. L. SUIB*
Department of Chemistry, University of Connecticut, 55, North Eagleville
Road, Storrs, Connecticut, 06269, USA
*Email: [email protected]
13.1 Introduction
Manganese oxide octahedral molecular sieves (OMS) have attracted a large
amount of attention from researchers in recent years as they are excellent
catalysts. Natural OMS, or cryptomelane-type manganese oxide, occurs in
deep sea beds as nodules.1 The motivation to synthesize this mineral was
triggered by the synthesis of a todorokite-type 33 manganese oxide molecular sieve by Shen et al. in 1993, which was named OMS-1.2 Then, in 1994,
DeGuzman et al. synthesized a cryptomelane-type material with a 22 tunnel
structure, as shown in Figure 13.1, using MnO4 and Mn21; the material was
highly porous and showed semiconducting properties.3 After this discovery,
many researchers around the world became very interested in this material.
Many new avenues of research have opened up since then. As an example,
RSC Catalysis Series No. 17
Metal Nanoparticles for Catalysis: Advances and Applications
Edited by Franklin Tao
r The Royal Society of Chemistry 2014
Published by the Royal Society of Chemistry, www.rsc.org
235
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
236
Figure 13.1
Chapter 13
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
Hydrothermal processing of materials has been used as a facile and environmentally benign synthetic approach for years. This is one of the most
versatile solution-based strategies to synthesize nanomaterials, due to the
ease of control of the reaction time, temperature, precursor, product separation and simple setup.10,11
Typical fabrication of OMS-2 using any synthetic technique involves oxidation of Mn21 with a suitable oxidant such as MnO4 , Cr2O72 , S2O82 ,
H2O2, or ozone.1216 Hydrothermal techniques in particular have been
identified in the literature as an eective way to control the morphology and
particle size of the OMS-2 materials. Figure 13.2 shows field emission
scanning electron micrographs (FESEM) of OMS-2 nanomaterials, fabricated
using hydrothermal techniques. Typical fiber-like morphology was obtained
by using a MnO4 /Mn21 redox couple under acidic conditions
(Figure 13.2(a)).3,12 However, when the oxidant was changed to Cr2O72 ,
a self-assembled three-dimensional dendritic structure was formed
(Figure 13.2(b,c)).13 A spontaneous free-standing membrane was formed
using S2O82 as the oxidant. When the fibers grew to sucient length
(B100 mm), they were arranged in an interlocking fashion, Figure 13.2(d), to
yield a free-standing membrane, which was used in sustainable energy and
environmental applications. A recent study by Yuan et al. showed that these
membranes can be used in oil spill cleanup by selective adsorption of
oil.17,18 Compared to fibers of several hundreds of micrometres obtained
using conventional hydrothermal methods, a microwave-assisted
Figure 13.2
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
238
Chapter 13
Figure 13.3
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
13.3 Catalysis
13.3.1
Selective catalysis has gained a great deal of research interest due to the
many economical as well as environmental advantages of the process. OMS-2
is an excellent catalyst that has the proven ability to catalyze both product
and substrate selective reactions.6,7,38 OMS-2 has shown the ability to selectively oxidize many organic functional groups including alcohols, alkenes,
thiols, and amines.7,9,3343 Son et al. presented a selective alcohol oxidation
to an aldehyde without any over oxidation to carboxylic acids, Scheme 13.1.38
The mixed valent Mn active sites for the selective oxidation were located
inside the tunnels, due to the lower activity observed towards larger
substrates.
A more elaborate mechanistic study on alcohol oxidation was done using
isotope labeled oxygen (18O) by Makwana et al.9 Selective aerobic benzyl alcohol oxidation was carried out as the model reaction, in which 100%
conversion as well as 100% selectivity was achieved towards the aldehyde.
A further study confirmed the active participation of structural oxygen during oxidation on OMS-2 catalysts, which were later replenished by atmospheric oxygen. The kinetic data also suggested the presence of a redox cycle
between Mn21 and Mn41 that drives the oxidation of the substrate,
Scheme 13.2. All of the above observations led to the conclusion that the
OH
R2
Scheme 13.1
R1
H
(l)
O
OMS-2, Air
R2
R1
(l)
View Online
240
Chapter 13
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
H 2O2
Mn2+
H2O + 1/2 O2
2e
O
2H+
Mn4+
H
H
O2
Scheme 13.2
OMS-2
OH
R2
R1
OMS-2, Air
R2
R3NH2
R1
OH
R2
R1
NR 3H
R2
R1
NR 3
+ OMS-2 + H 2O
Scheme 13.3
Mars van Krevelen mechanism is active on OMS-2 materials during
oxidations.1,9,20
Sithambaram et al. further utilized this excellent catalytic activity and
selectivity of OMS-2 towards alcohol oxidation in tandem catalysis.20 The
team synthesized imines directly from alcohols and amines. The reaction
progresses through a concerted fashion, Scheme 13.3, where selective oxidation of alcohols to aldehydes through a Mars van Krevelen mechanism
occurs followed by nucleophilic attack from the amine to the carbonyl carbon producing an imine bond. Conversion and selectivity towards the imine
of up to 100% was observed and also the reaction was found to be valid for a
wide variety of alkyl and aryl substrates.
Advancement of this tandem imine formation led to another study by
Sithambaram et al. on the synthesis of quinoxalines, Scheme 13.4, which are
nitrogen-containing heterocyclic compounds, essential intermediates in the
fine chemical industry.44
This reported process was highly ecient (100% yield in 1 h), had good
substrate generality, and used mild reaction conditions. The catalyst, OMS-2,
was also recycled after reaction without major structural changes or
deactivation.
Another transformation important in fine chemical synthesis catalyzed by
OMS-2 is hydrocarbon oxidation, including both alkanes and alkenes. Ghosh
et al. reported the liquid-phase epoxidation of olefins using porous OMS-2
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
Scheme 13.4
O
Homolitic
pathway
Oxo-metal
formation
OMS-2 n+2
+
t-butOH
OMS-2
+
t-butOOH
n+
OH
OMS-2 n+2
+
t-butO
Scheme 13.5
and TBHP (tertiary-butyl hydroperoxide), Scheme 13.5.45 The reaction gave
good conversion with up to 100% selectivity towards epoxide at low temperature (60 1C). Significant conversion was obtained at temperatures as low
as room temperature and catalyst amounts as low as 0.3 mol%. In order to
understand the reaction mechanism, the reaction was carried out in the
presence of radical scavengers/inhibitors. Two dierent radical scavengers/
inhibitors were used in two separate reactions. When inhibitor 4-Oxo
TEMPO (4-oxo-2,2,6,6-tetramethylpiperidine-1-oxyl) was used, the conversion decreased from 59 to 37% and in the presence of radical inhibitor
IONOL (2,6-di-tert-butyl-4-methylphenol), the conversion of cyclooctene was
reduced from 59% to 22.5%.
The results confirmed that the epoxidation occurs on OMS-2 surfaces via
two co-operative mechanisms: oxo-metal formation, and a homolytic pathway
due to the presence of Mn21, Mn31, and Mn41 oxidation states
(Scheme 13.5).45 A follow up to this study was done by Ghosh et al. for a deeper
understanding of the involvement of the Lewis and Brnsted acid sites and
also the eect of the preparation method of OMS-2 catalysts, which we discussed previously in this chapter. OMS-2 materials synthesized by five different synthetic techniques were used in the study, namely reflux (OMS-2R),
hydrothermal (OMS-2HY), high temperature (OMS-2HT), solvent free (OMS-2SF),
and microwave (OMS-2MW). The results of this study show that the acidity of the
material together with porosity play essential roles. The mesopore volume of
OMS-2 synthesized using dierent methods ranged from 0.480.027 cm3 g 1.
The results also showed that the optimum mesopore volume for the styrene
epoxidation was 0.29 cm3 g 1, which corresponds to OMS-2R.
13.3.2
CH Activation
The petroleum industry serves as the primary source of raw materials for the
fine chemical industry; functionalization of these raw chemicals, especially
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
242
Chapter 13
alkanes, requires very high energy due to the inherent inertness of the CH
bond.46 Hence, a catalytic process for the activation of the CH bond is
highly desirable. There are many examples of OMS-2 materials in the literature for the activation of CH bonds and functionalized compounds such
as 9-fluorene,42 cyclohexane,34 and, in some instances, toxic chemicals such
as BTX (benzene, toluene, and xylene) were totally oxidized to CO2.7 These
functionalized compounds have applications in the energy, medical, and
environmental fields.34,42 Scheme 13.6 shows a study by Opembe et al.,
where 9H-fluorene was oxidized to 9-fluorenone.42
The process developed by Opembe et al. used air as an oxidant, unlike the
other processes, which used peroxide, oxygen, and other activation methods,
which are harmful.4749 Excellent conversions (up to 99%) were observed
with 100% selectivity towards 9-fluorenone. Furthermore, the reaction
mechanism and kinetics were studied by isotope-labeled [D10]-fluorene. The
results show the reaction occurs via a late transition state in which CH
bond cleavage is the rate-determining step. Further, the mechanism was
proposed to be a free radical mechanism, as shown in Scheme 13.6.
Recently, Kumar et al. developed a catalytic route for the oxidation of
cyclohexane using OMS-2 as a catalyst.34 This process is highly important in
industry because the products from cyclohexane oxidationcyclohexanol and
cyclohexanoneare starting materials for Nylon and other value-added
chemicals.5054 In this study, the acidity of the OMS-2 catalyst was altered by
using an ion-exchange procedure involving concentrated nitric acid. The high
Scheme 13.6
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
13.3.3
CO2 Activation
Utilization of CO2the most abundant C source in natureas a raw material for fine chemical synthesis has gained immense research attention.5558 However, due to the high inertness or the stability of CO2
molecules, a great amount of energy input is required to activate CO2 in
order to be utilized in a chemical reaction.56,59
Hu et al. developed a OMS-2 catalyzed electrochemical and thermal hybrid
activation of CO2 (Scheme 13.7) to produce paraformaldehyde, a useful
chemical in the polymer and pesticide industry.56 In this work, OMS-2 was
used as a support for Pt nanoclusters, deposited using chemical vapor deposition (CVD). The metal/metal oxide interface was activated by means of a
DC current, and the reaction proceeds through electron and ionic conduction from metal to metal oxide.
Since OMS-2 is a good conductor for electrons as well as ions and also
possesses mixed valency,1,5,9,60 when OMS-2 was used as a support, the reaction temperature was significantly reduced (350450 1C) as compared to
calcia-stabilized zirconia (600900 1C).
Another catalytic transformation on OMS-2 utilizing CO2 is ethane dehydrogenation to ethylene as reported by Jin et al.61 In this study, OMS-2
yields a higher amount of ethane (up to 66%) compared to a conventional
Cr-ZSM-5-type catalyst (36%). CO2 utilization was also very eective when
O
C
Pt
H
+ O2
Support
O
C H H
O O
O
C
O
H2O
Scheme 13.7
CO2
View Online
244
Chapter 13
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
C 2H6(g)
H2(g)
+ CO 2(g)
C 2H6(g) + CO 2(g)
C 2H4(g) + H2(g)
(1)
H2O (g)
(2)
+ CO (g)
Since CO2(g) utilizes H2(g) produced from reaction (1) in a RWGS reaction,
that prevents further reaction between H2(g) and C2H6(g) to give unwanted
byproduct CH4(g). CO2(g) also prevents catalyst deactivation by coke formation, by acting as an oxidizing agent on the OMS-2 surface.
13.3.4
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
and ecology, spillage of any extent needs immediate attention and cleanup.
Yuan et al. have reported oil removal via adsorption onto a modified super
hydrophobic OMS-2 membrane.17 A porous OMS-2 membrane was synthesized according to literature reported methods81 and then the membrane
was further modified using a polydimethyl siloxane (PDMS) polymer.
The resulting material showed an exceptional wettability profile, which
can be altered to be super hydrophobic or super hydrophilic by changing
the temperature. The removal eciency of OMS-2 was 20 the weight of the
membrane, which is much higher than the conventional adsorbents
(Figure 13.4). This OMS-2 membrane can be produced easily to treat a
Figure 13.4
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
246
Chapter 13
large-scale oil spill and also easily recovered. Therefore, OMS-2 materials
are highly desirable in large-scale industrial applications for wastewater
treatment.
Air quality has declined at a rapid rate in recent years as a result of heavy
use of fossil fuels, petroleum-based chemicals, and global industrialization.82,83 VOCs (volatile organic compounds), carbon monoxide, nitrogen
oxides, and particulate matter generation in the lower atmosphere are the
primary cause of smog, which relates to high rates in cardiovascular disease
in human beings.8284 Genuino et al.7 have reported a catalytic route to
oxidize BTX (benzene, toluene, and xylene), a group of VOCs that are highly
toxic, to CO2.7 The reactivity of the substrates studied was xyleneso
ethylbenzeneobenzeneotoluene. The process developed was highly selective;
other byproducts, such as CO or partially oxidized VOC, were not observed.
OMS-2 showed higher activity than commercial MnO2. Temperature programed desorption (TPD) studies of xylenes (ortho, meta, and para-xylene)
showed peak maxima in temperature in the order of paraometaoortho-xylene
suggesting a shape selectivity in the OMS-2 catalysts. The higher catalytic
activity of the OMS-2 catalyst was due to the higher hydrophobicity (0.91),
pore volume (0.46 cm3 g 1), pore diameter (20.1 nm), and higher oxygen
mobility. This study further suggests the potential application of OMS-2 in air
and soil purification from soilvapor extraction, automobile exhaust treatment, and gasoline-stripping of soil. Another example of OMS-2 in air purification is low temperature CO(g) oxidation. Xia et al. reported a catalytic
carbon monoxide oxidation route using Ag1, Co21, and Cu21 doped OMS-2.69
Doped catalysts were very stable under a reducing atmosphere for extended
time on stream.
13.4 Conclusion
This review summarizes the synthesis, structure, morphology, and catalytic
activity of a unique class of OMS materialsOMS-2 systems. From their
initial synthesis in 1993 until now, OMS-2 materials have undergone a large
amount of research throughout the world. Novel synthetic processes have
been developed, which gave this material new perspectives in adsorption
and catalytic applications. The dierent synthetic routes yield similar crystalline phases (cryptomelane) with dierent surface areas, pore volumes,
pore diameters, surface defects, porosity (both micro and mesoporosity),
particle sizes, and catalytic activities. Metal dopants in the structure of
OMS-2 increased the catalytic activities as well as the adsorption capacity.
OMS-2 materials can be fabricated as powders as well as free-standing
membranes for dierent applications such as oil spill cleanup. The OMS-2
catalyst is a non-toxic environmentally benign catalyst for fine chemical
synthesis, oxidation, environmental remediation, and CH and CO2
activation.
Future studies of OMS-2 materials will involve desulfurization, tandem
processes, and their use in more environmental applications such as active
View Online
materials in particulate filters. Research will also continue along the avenues
of formation of novel morphologies in thin OMS-2 films for electrocatalysis.
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
Acknowledgments
We acknowledge the Geosciences and Biosciences Division, Oce of the
Basic Energy Sciences, Oce of Science, and U.S. Department of Energy for
support of our research on manganese octahedral molecular sieves and
other porous materials, under grant DE-FG02-86ER13622-A000.
References
1. S. L. Suib, J. Mater. Chem., 2008, 18, 16231631.
2. Y. F. Shen, R. P. Zerger, R. N. DeGuzman, S. L. Suib, D. I. McCurdy and
C. L. OYoung, Science, 1993, 260, 511515.
3. R. N. DeGuzman, Y. Shen, E. J. Neth, S. L. Suib, C. OYoung, S. Levine
and J. M. Newsam, Chem. Mater., 1994, 6, 815821.
pez, E. Ortiz, M. Alvarez, J. Manjarrez, M. Montes, P. Navarro and
4. T. Lo
J. A. Odriozola, Mater. Chem. Phys., 2010, 120, 518525.
5. R. N. De Guzman, Y. F. Shen, B. R. Shaw, S. L. Suib and C. L. OYoung,
Chem. Mater., 1993, 5, 13951400.
6. L. Espinal, W. Wong-Ng, J. A. Kaduk, A. J. Allen, C. R. Snyder, C. Chiu,
D. W. Siderius, L. Li, E. Cockayne, A. E. Espinal and S. L. Suib, J. Am.
Chem. Soc., 2012, 134, 79447951.
7. H. C. Genuino, S. Dharmarathna, E. C. Njagi, M. C. Mei and S. L. Suib,
J. Phys. Chem. C, 2012, 116, 1206612078.
8. J. Li, R. Wang and J. Hao, J. Phys. Chem. C, 2010, 114, 1054410550.
9. V. D. Makwana, S. Young-Chan, A. R. Howell and S. L. Suib, J. Catal.,
2002, 210, 4652.
10. M. Yoshimura and K. Byrappa, J. Mater. Sci., 2008, 43, 20852103.
11. D. Wang, T. Xie and Y. Li, Nano Res., 2009, 2, 3046.
12. G. Qiu, H. Huang, S. Dharmarathna, E. Benbow, L. Staord and
S. L. Suib, Chem. Mater., 2011, 23, 38923901.
13. J. Yuan, W. Li, S. Gomez and S. L. Suib, J. Am. Chem. Soc., 2005, 127,
1418414185.
14. J. Yuan, K. Laubernds, J. Villegas, S. Gomez and S. L. Suib, Adv. Mater.,
2004, 16, 17291732.
15. J. C. Villegas, L. J. Garces, S. Gomez, J. P. Durand and S. L. Suib, Chem.
Mater., 2005, 17, 19101918.
16. H. M. Galindo, Y. Carvajal, C. Njagi, R. A. Ristau and S. L. Suib, Langmuir, 2010, 26, 1367713683.
17. J. Yuan, X. Liu, O. Akbulut, J. Hu, S. L. Suib, J. Kong and F. Stellacci, Nat.
Nanotechnol., 2008, 3, 332336.
18. X. Li, B. Hu, S. L. Suib, Y. Lei and B. Li, J. Power Sources, 2010, 195, 2586
2591.
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
248
Chapter 13
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
View Online
23/06/2014 08:15:07.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00235
250
Chapter 13
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
Subject Index
p-electrons 194
z potential 102
ABA copolymer 197
ABTA (2,2 0 -azinobis(3-ethylbenzothiazoline-6-sulfonic acid)) 198
Acetobacter xylinum 138
acetonitrilewater solvent 143
acetophenones, brominated 89, 91,
92
4-acetylphenylboronic acid 141,
142
4-acetylphenylchloride 137138
acid chlorides 126
acrylic acid n-butyl ester 144
active site characteristics 7
active supports 161, 162
aerobic oxidation 159, 160, 165
air 207, 242, 246
alcohols 159, 160, 239
aldehydes 7576, 93, 94, 96, 243
aliphatic aldehydes 93, 94, 96
alkenes see olefins (alkenes)
alkenyl halides 134
alkenylboranes 134
alkylamine ligands 19
alloys for templating synthesis 23
alumina membranes in twocompartment reactors 117, 118
alumina supports 7071
aluminium oxides 36, 37,
161, 162
aluminosilicate frameworks 92, 93
HZSM-5 192
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
252
arenes
asymmetric nanocatalysis 105
derivatives biphasic
hydrogenation with
HEA16Cl-capped rhodium
colloids 103, 104
hydrogenation reactions 99
prochiral 107108
aryl halides
palladium-catalysed
carboncarbon coupling
reactions 119, 121, 141
Heck coupling
reaction 147
Stille coupling
reaction 126130, 132,
133
Suzuki coupling
reaction 134, 136, 137,
138, 140, 141, 143
arylboronic acid 143
ascorbic acid 211
assembly-dispersion mechanism 36
asymmetric nanocatalysis see
enantiocontrol/selectivity
atom transfer radical polymerisation
(ATRP) 197
atom-leaching mechanism 116126,
133, 141
ATRP (atom transfer radical
polymerisation) 197
attenuated total reflectance (ATR)
1857
2,2 0 -azinobis(3-ethylbenzothiazoline-6-sulfonic acid) (ABTA)
198
bacterial cellulose (BC) 138
barium salts 159, 167
BE-O (binding energy for oxygen)
225, 226
bead templates 136, 174, 177, 178
BelousovZhabotinsky (BZ)
reaction 205
benzene, biphasic
hydrogenation 103
Subject Index
View Online
Subject Index
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
253
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
254
Subject Index
behaviour in
catalysis 179185
catalytically active inner
surface 180182
hollow/solid
nanocatalysts
comparison 182183
optical properties of
plasmonic
nanocatalysts 184185
proposed mechanism for
nanocatalysis
185187
single/double shell
hollow nanocatalysts
comparison 183184
synthetic
approaches 174176
noble metal nanoparticles
for 4669
packing on crystallographic
faces 23
precursors for size-controlled
metal nanoparticloes
3537
colorimetric oxygen indicators 217
computational approaches
copper nanocatalysts for WGS
reactions 221223
density functional
theory 219231
metal (core)platinum shell
nanocatalysts for ORR in
fuel cells 223227
metal nanocatalysts 220227
method 220
overview 45
shape eects 226227
size eects 224226
supported metal
nanocatalysts 227230
confinement properties 5760, 182,
183, 184, 185
see also cage eect
coordination numbers (CNs) 124
copolymers 135, 196, 197, 198
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
Subject Index
copper
copper oxide nanoparticle
surface restructuring 12
copper/zinc oxide 229
nanocatalysts for water-gas
shift reactions 221223
nanocrystal formation 17
nanoparticles in clock
reactions 207, 209
photo-assisted deposition 35
coreshell nanocatalysts
hydrogen peroxide
formation 4244
magnetic FePt@Ti-SiO2
synthesis 3839
model 224
nanoreactor catalysts 195
corner atoms 9, 174, 179
cost of transition metals 84
Cotton eect 41
Coulombs interaction 209, 210
coupling reactions see
carboncarbon coupling reactions
critical micellar concentration
(CMC) 101
cryptomelane-type manganese oxide
octahedral porous molecular
sieves 235247
crystallinity
copper-29 catalyst crystal
structure 221, 223
domain size/structure
sensitivity 10
exposed planes 173
packing on crystallographic
faces 23
palladium
nanoparticles 119123
surface free energy of facets 11
CTAB (cetyl trimethylammonium
bromide)-Pt nanocrystals 21
CTACI (cetyltrimethylammonium
chloride) 198
cubo-octahedron shape
analysis 910, 11
cuprous oxide 212, 213, 214, 215
255
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
256
N,N-dimethyl-N-alkyl-N(hydroxyalkyl)ammonium
salts 100, 101, 103
N,N-dimethyl-N-cetyl-N-(3hydroxypropyl)ammonium
chloride (HPA16Cl) 100
dimethylamine-borane
(DMAB) 6063
dimethylphenylsilane 165
4-diphenylphosphinobenzoic
acid 145
diphosphate-stabilised noble metal
nanoparticles 51
diphosphines 5457
dipping method of LB film
transfer 178
dissolution potential 226
2,4-di(trifluoromethyl)phenylboronic
acid 137
DMAB (dimethylamineborane) 6063
DMF (dimethylformamide)
solvent 117, 118, 119
DMol3 numerical basis sets 220
double shell nanoparticles 180,
183184
DSN (dendrimer-stabilised
nanoparticles) 128, 140, 144145
DTBB (4,4 0 -di-tert-butylbiphenyl) 85,
86
dual hydrogenationoxidation
reactions 78
dyes 180, 181, 203217, 244
edge atoms 9, 174, 179
EDX-SEM (energy-dispersive X-ray
spectroscopy-scanning electron
microscopy) mapping 175, 176
electrochemical nanoparticle
synthesis method 144
electron beam lithography (EBL) 177
electron microscopy 8, 175, 176, 237
electronic structure 1415, 198199
electrostatic stabilisation 210
EleyRideal mechanism 162,
215216
Subject Index
emulsions 195198
enantiocontrol/selectivity
ammonium surfactant-capped
rhodium nanoparticles for
biphasic
hydrogenation 105108
chiral diphosphite
ligands 6365, 68
cis/trans isomerism 22, 23,
146, 147
hydrogenation
reactions 5051, 8889, 90,
106107
optically-active ammonium
salts 106, 107, 108
plasmonic
nanocatalysts 184185
roll-over trans diastereoisomer
formation 104
stereoselective arylated alkene
products 134
steric hindrance 106, 107, 108,
141
encapsulation 21, 196
endocrine disruptors 104
energy-dispersive X-ray
spectroscopy-scanning electron
microscopy (EDX-SEM)
mapping 175, 176
environmentally-friendly processes
carboncarbon coupling
reactions 112
environmental
remediation 244246
green solvents 99
oil spill cleanup 237, 244246
enzymes 7, 194, 195, 197198
eosin Y 180, 182, 183, 184, 185
epoxidation 39
equilibrium constant K 203204
equilibrium shapes of
nanocrystals 11
Et3N (triethylamine) base 144
ethane 1516, 2122, 243, 244
ethyl-4-iodobenzoate 140
ethylene 14, 22, 243
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
Subject Index
ethylpyruvate
hydrogenation 106107
europium (Eu) 161, 162, 163
ex situ conditions 12
excitons 14, 15
extended X-ray absorption fine
structure (EXAFS) 124
Fehling test 212
Fermi energy levels 173, 210
field emission scanning electron
micrographs (FESEM) 237
finely-divided material structure
sensitivity factors 916
Fisher carbene complexes 129
flavylium cations 37
9-fluorene 242
9-fluorenone 242
fluorine removal 19
fluorocarbon-hydrocarbon
solvent 145
fluorohydrocarbon solvent 144
Fourier transform infrared (FTIR)
spectra 185187
free radical dye photodegradation
mechanism 180181
free-standing membranes 237
fuel cells 223227
functional group transfer
hydrogenation 8396
aldehydes reductive
amination 9496
alkenes 8688
carbonyl compounds 8894
history 8486
functionalised silica materials 125
furfural hydrogenation 78
galvanic replacement
approach 174175, 176
Gibbs law 101
gluconic acid 168, 169
glucose 168, 207
glucose oxidase 195
GnDenP-Pd nanoparticles 140, 141,
142, 143
257
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
258
Subject Index
customised dendrimers
housing metallic
nanocatalysts 193
palladium nanoparticle
catalysis mechanism 125
homolytic pathways 241
horseradish peroxidase (HRP) 198
HPA (hydroxypropylammonium)
100, 101, 103
hybrid nanocrystal catalysis
2324
hydroformylation reactions 6769
hydrogen
see also hydrogenation
reactions
adsorption on platinum or
rhodium 14
hydrogenolysis, ethane on
rhodium 1516
production
semi-conducting
anastase-type titanium
dioxide
photocatalysis 345
watergas shift
reaction 221223,
228230
transfer reduction
aldehydes 9496
alkenes 8688
carbonyl
compounds 8894
hydrogen peroxide
BrayLiebhafsky
reaction 205206
BriggsRauscher
reaction 206207
formation
coreshell
nanostructured
catalyst 4244
cyclooctene
epoxidation 39
PAD-PdAu/TS-1 33
hydrogenation reactions 5060
alumina supports 7071
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
Subject Index
ammonium surfactant-capped
rhodium nanoparticles for
biphasic
hydrogenation 99109
carbon materials as
supports 7576
dual hydrogenationoxidation
reactions 78
ethylpyruvate asymmetric
hydrogenation 106107
ligand-stabilised nanoparticles
as colloidal catalysts 5052
peptide-capped palladium
nanoparticles 132133
silica supports 7172
surfactant-supported
nanoreactor structures 196
transfer hydrogenation of
functional groups 8396
hydrophobichydrophilic
structures 195198
hydrothermal synthesis 19, 237238
hydroxyapatite-based gold
catalysts 164166
2-hydroxychalcon derivatives 37
hydroxyethylammonium (HEA) 100,
101, 103
hydroxyl-terminated PAMAM
dendrimers 126, 127, 139140,
147
hydroxylated fluoride-based gold
catalysts 166167
4-hydroxyphenyl iodide 141
hydroxypropylammonium
(HPA) 100, 101, 103
ICP-AES (inductively coupled
plasma-atomic emission
spectroscopy) 138
imidazolinium-based ionic
polymers 129, 136
imines 95, 165, 240
immobilisation inside
dendrimers 145, 146
in situ conditions 12, 193
inactive outer surfaces 180182
259
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
260
iron (Fe)
active supports for gold
nanocatalysts 161, 162
cost 84
cyclodextrin-stabilised metal
nanoparticles 3941
FePd magnetic nanoparticle
modified with chiral BINAP
ligand 4142
FePt@Ti-SiO2 synthesis 3839
metal salt-based gold
catalysts 162
single crystal surfaces 11
isopropanol 89
isopulegol 166, 167
isotropic shapes 172
Keggin-type polyoxometallates 130,
138139, 148
ketones 93, 9496, 126
kinetics of nanocrystal formation 17
Kirkendall eect 174
Kumada reactions 113, 114
Landolt clock reaction 204
Langmuir isotherm 216
LangmuirBlodgett (LB)
technique 173, 178179
LangmuirHinshelwood
mechanism 162, 216
lanthanides (Ln) 161, 162, 163
lanthanum (La) 160, 161, 162, 163
leaching mechanisms 116126, 133,
141
lead dioxide 20
leucomethylene blue (LMB) 210,
211, 213, 214
Lewis acid sites 241
ligand stabilisers 5052, 115, 116,
144
see also individual ligand
stabilisers
limonene hydrogenation 60
lipophilic chain lengths 100, 101
liposomes 195198
lithography techniques 177
Subject Index
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
Subject Index
261
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
262
Subject Index
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
Subject Index
263
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
264
Subject Index
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
Subject Index
265
recyclability
bacterial cellulose-stabilised
palladium nanoparticles 138
challenge 176177
colloidal metallic hollow
nanostructures 178179
dendrimer-based
materials 127, 140, 141, 145
imidazolinium-based ionic
polymer-stabilised
palladium
nanoparticles 130
palladium-PANI
nanoparticles 137138
redox reactions 203217
reducing agents 203217
reduction pathways 182183, 184,
185
reductive amination of
aldehydes 9496
reporting standards 89
reverse micelles 93, 197
reverse water-gas shift reaction
(RWGS) 244
reversible redox reactions 203217
rhodium (Rh) nanocatalysts
ammonium surfactant-capped
for biphasic
hydrogenation 99109
carbonyl compounds hydrogentransfer reduction 88
cost 84
diphosphate-stabilised
nanoparticles 51
ethane hydrogenolysis 1516
hydroformylation
reactions 6769
nanocrystal formation 17
particle atomic
restructuring 14
polymer nanotubes with PdRh
nanoparticles 196
protective agents 100, 101, 103
Rh/TPPTS catalyst 196
rhodium/palladium
systems 13, 196
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
266
Subject Index
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
Subject Index
267
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
268
Subject Index
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
Subject Index
polymer coreshell
structures 195
single/double shell hollow
nanocatalysts
comparison 183184
SuzukiMiyaura coupling
reaction 41, 65
SXRD (UHV-high-pressure surface
X-ray diraction) 13
synthesis methods
colloidal metallic hollow
nanostructures 174176
micelle-based 196197
multifunctional magnetic
nanoparticle-based
catalysts 3844
noble metal nanoparticles,
organometallic
approach 4779
overview 3
porous cryptomelane-type
manganese oxide
octahedral molecular
sieves 236239
supported gold
nanocatalysts 158
supported metal nanoparticle
catalysts 3137
templating methods 23
titanium dioxide nanocrystals
with reactive facets 19
well-defined
nanocrystals 1618
templating methods
hollow metallic nanocatalyst
synthesis 174
nanocrystal synthesis 23
R5 peptide (Cylindrotheca
fusiformis) 130, 131, 132
TEMPO (4-oxo-2,2,6,6tetramethylpiperidine-1-oxyl) 241
tetraalkylammonium salts 144
tetrabutylammonium salts 129, 134
tetrabutylphenylstannane 130
tetraheptylammonium bromide 130
tetrahydrofuran solution 49
269
tetrakis(triphenylphosphine) 134
tetraphenyltin 130
THEA (trishydroxyethylammonium)
100, 101, 103
theoretical aspects of metal
nanocatalysts 219231
computational
method 219231
copper nanocatalysts for
water-gas shift
reactions 221223
metal (core)platinum shell
nanocatalysts for oxygen
reduction reactions 223227
supported
nanocatalysts 227230
types of metal
nanocatalysts 220227
thermolysis 144
thiazine groupd 210
thiols 239
2-thiopheneboronic acid 135, 139
thiosulfate 21, 185187
time-of-flight (TOF) 1516
tin reagents
CC coupling reactions 113,
114
organostannane
reagents 126
phenyltin
trichloride 126, 127,
131, 132, 133
tetrabutylphenylstannane
130
tetraphenyltin 130
tributyl(phenyl)stannane
128, 129
titanium (Ti)
FePt@Ti-SiO2 synthesis 3839
mesoporous silicas 3134
PAD-PdNi/Ti-HMS catalyst 343
Pd/SiO2@TiMSS 43, 44
titanium dioxide 1921, 31, 32,
22930
zeolites 3134
toluene 5960, 145, 159
p-tolylboronic acid 139140
View Online
23/06/2014 08:15:09.
Published on 12 June 2014 on http://pubs.rsc.org | doi:10.1039/9781782621034-00251
270
transfer hydrogenation
aldehydes 9496
alkenes 8688
carbonyl compounds 8894
overview 3
transition metals 84
see also individual transition
metals
transmission electron microscopy
(TEM) 102, 118, 175, 176
trapped holes 2021, 32
1,3,5-triaza-7-phosphaadamantane
(PTA) 5254
1,2,3-triazolylsulfonate
dendrimers 140
tributyl(phenyl)stannane 128, 129
3,4,5-triethoxybenzoyl chloride 145
triethylamine (Et3N) base 144
trimethyltetradecylammonium
bromide (TTAB) 177
trishydroxyethylammonium
(THEA) 100, 101, 103
TTAB (trimethyltetradecylammonium
bromide) 177
tuning 194, 195, 212
turnover frequency (TOF)
anomalous behaviour of
ethane hydrogenolysis on
rhodium 1516
carboncarbon coupling using
palladium nanoparticles 133
dendrimer-stabilised palladium
nanoparticle system 140
palladium nanoparticlecatalysed Suzuki reaction 124
Pd(core)Pt(shell)
nanoparticles 225
reporting standards 89
turnover number (TON) 147
two-compartment reactors 117126
two-electron reduction 182, 183,
184, 185
Subject Index